[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

Search Results (111)

Search Parameters:
Keywords = Taklimakan Desert

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
28 pages, 3751 KiB  
Article
Urbanization Effect on Local Summer Climate in Arid Region City of Urumqi: A Numerical Case Study
by Aerzuna Abulimiti, Yongqiang Liu, Qing He, Ali Mamtimin, Junqiang Yao, Yong Zeng and Abuduwaili Abulikemu
Remote Sens. 2025, 17(3), 476; https://doi.org/10.3390/rs17030476 - 30 Jan 2025
Viewed by 420
Abstract
The urbanization effect (UE) on local or regional climate is a prominent research topic in the research field of urban climates. However, there is little research on the UE of Urumqi, a typical arid region city, concerning various climatic factors and their spatio–temporal [...] Read more.
The urbanization effect (UE) on local or regional climate is a prominent research topic in the research field of urban climates. However, there is little research on the UE of Urumqi, a typical arid region city, concerning various climatic factors and their spatio–temporal characteristics. This study quantitatively investigates the UE of Urumqi on multiple climatic factors in summer based on a decade-long period of WRF–UCM (Weather Research and Forecasting model coupled with the Urban Canopy Model) simulation data. The findings reveal that the UE of Urumqi has resulted in a reduction in the diurnal temperature range (DTR) within the urban area by causing an increase in night-time minimum temperatures, with the maximum decrease reaching −2.5 °C. Additionally, the UE has also led to a decrease in the water vapor mixing ratio (WVMR) and relative humidity (RH) at 2 m, with the maximum reductions being 0.45 g kg−1 and −6.5%, respectively. Furthermore, the UE of Urumqi has led to an increase in planetary boundary layer height (PBLH), with a more pronounced effect in the central part of the city than in its surroundings, reaching a maximum increase of over 750 m at 19:00 Local Solar Time (LST, i.e., UTC + 6). The UE has also resulted in an increase in precipitation in the northern part of the city by up to 7.5 mm while inhibiting precipitation in the southern part by more than 6 mm. Moreover, the UE of Urumqi has enhanced precipitation both upstream and downstream of the city, with a maximum increase of 7.9 mm. The UE of Urumqi has also suppressed precipitation during summer mornings while enhancing it in summer afternoons. The UE has exerted certain influences on the aforementioned climatic factors, with the UE varying across different directions for each factor. Except for precipitation and PBLH, the UE on the remaining factors exhibit a greater magnitude in the northern region compared to the southern region of Urumqi. Full article
18 pages, 6529 KiB  
Article
A Novel Algorithm for Estimating the Sand Dune Density of the Taklimakan Desert Based on Remote Sensing Data
by Mingyu Wang, Yongqiang Liu, Huoqing Li, Minzhong Wang, Wen Huo and Zonghui Liu
Remote Sens. 2025, 17(2), 297; https://doi.org/10.3390/rs17020297 - 16 Jan 2025
Viewed by 541
Abstract
The dune density is an important parameter for representing the characteristics of desert geomorphology, providing a precise depiction of the undulating topography of the desert. Owing to the limitations of estimation methods and data availability, accurately quantifying dune density has posed a significant [...] Read more.
The dune density is an important parameter for representing the characteristics of desert geomorphology, providing a precise depiction of the undulating topography of the desert. Owing to the limitations of estimation methods and data availability, accurately quantifying dune density has posed a significant challenge; in response to this issue, we propose an innovative model to estimate dune density using a dune vertex search combined with four-directional orographic spectral decomposition. This study reveals several key insights: (1) Taklimakan Desert distributes approximately 5.31 × 107 dunes, with a linear regression fit R2 of 0.79 between the estimated and observed values. The average absolute error and root mean square error are calculated as 25.61 n/km2 and 30.48 n/km2, respectively. (2) The distribution of dune density across the eastern, northeastern, southern, and western parts of the Taklimakan Desert is relatively lower, while there is higher dune density in the central and northern areas. (3) The observation data constructed using the improved YOLOv8s algorithm and remote sensing imagery effectively validate the estimation results of dune density. The new algorithm demonstrates a high level of accuracy in estimating sand dune density, thereby providing crucial parameters for sub-grid orographic parameterization in desert regions. Additionally, its application potential in dust modeling appears promising. Full article
Show Figures

Figure 1

Figure 1
<p>Research area overview, (<b>a</b>) research area location, (<b>b</b>) types of dunes and locations of sampling areas, and (<b>c1</b>–<b>c13</b>) sample area Google Earth image. The desert map is provided by the National Cryosphere Desert Data Center. (<a href="http://www.ncdc.ac.cn" target="_blank">http://www.ncdc.ac.cn</a> (accessed on 20 March 2024)).</p>
Full article ">Figure 2
<p>Dunes in the TD. (<b>a</b>,<b>b</b>) Dunes in the Hade (83°42′E, 40°45′N), date: 04/2024. (<b>c</b>–<b>e</b>) Dunes in the Xiaotang (84°18′E, 40°49′N), date: 05/2024. (<b>f</b>,<b>g</b>) Dunes on both sides of the desert highway, (<b>f</b>): (84°19′E, 40°34′N), date: 10/2023 (<b>g</b>): (83°44′E, 39°15′N), date: 05/2024. (<b>h</b>–<b>j</b>) Dunes in the Tazhong (83°38′E, 38°59′N), date: 05/2024. (<b>k</b>,<b>l</b>) Yutian oasis–desert ecotone (81°28′E, 36°56′N), date: 05/2023.</p>
Full article ">Figure 3
<p>The workflow of this study.</p>
Full article ">Figure 4
<p>Extraction algorithm based on dune vertices. (<b>a</b>) DEM. (<b>b</b>) Distribution of dune vertices and areas.</p>
Full article ">Figure 5
<p>The orographic spectral decomposition of the four-directional method. (<b>a</b>) 0°, (<b>b</b>) 45°, (<b>c</b>) 90°, and (<b>d</b>) 135°. Different colors represent different terrain spectral lines.</p>
Full article ">Figure 6
<p>The framework for the acquisition of dune density observation data.</p>
Full article ">Figure 7
<p>Identification results of some sample areas. (<b>a</b>–<b>f</b>) Six randomly selected sample areas; the part circled by the orange yellow box represents the sand dunes recognized by the model.</p>
Full article ">Figure 8
<p>Dune density distribution in TD. (<b>a</b>) Dune density distribution; (<b>b</b>–<b>i</b>) Google Earth images of typical sample regions. (The display of (<b>b</b>–<b>i</b>) is not the actual size of the sample area but only a part of it).</p>
Full article ">Figure 9
<p>Accuracy of the typical sample area.</p>
Full article ">Figure 10
<p>Linear fitting results between estimated and observed values of dune density.</p>
Full article ">
25 pages, 27385 KiB  
Article
Response of Natural Forests and Grasslands in Xinjiang to Climate Change Based on Sun-Induced Chlorophyll Fluorescence
by Jinrun He, Jinglong Fan, Zhentao Lv and Shengyu Li
Remote Sens. 2025, 17(1), 152; https://doi.org/10.3390/rs17010152 - 4 Jan 2025
Viewed by 704
Abstract
In arid regions, climatic fluctuations significantly affect vegetation structure and function. Sun-induced chlorophyll fluorescence (SIF) can quantify certain physiological parameters of vegetation but has limitations in characterizing responses to climate change. This study analyzed the spatiotemporal differences in response to climate change across [...] Read more.
In arid regions, climatic fluctuations significantly affect vegetation structure and function. Sun-induced chlorophyll fluorescence (SIF) can quantify certain physiological parameters of vegetation but has limitations in characterizing responses to climate change. This study analyzed the spatiotemporal differences in response to climate change across various ecological regions and vegetation types from 2000 to 2020 in Xinjiang. According to China’s ecological zoning, R1 (Altai Mountains-Western Junggar Mountains forest-steppe) and R5 (Pamir-Kunlun Mountains-Altyn Tagh high-altitude desert grasslands) represent two ecological extremes, while R2–R4 span desert and forest-steppe ecosystems. We employed the standardized precipitation evapotranspiration index (SPEI) at different timescales to represent drought intensity and frequency in conjunction with global OCO-2 SIF products (GOSIF) and the normalized difference vegetation index (NDVI) to assess vegetation growth conditions. The results show that (1) between 2000 and 2020, the overall drought severity in Xinjiang exhibited a slight deterioration, particularly in northern regions (R1 and R2), with a gradual transition from short-term to long-term drought conditions. The R4 and R5 ecological regions in southern Xinjiang also displayed a slight deterioration trend; however, R5 remained relatively stable on the SPEI24 timescale. (2) The NDVI and SIF values across Xinjiang exhibited an upward trend. However, in densely vegetated areas (R1–R3), both NDVI and SIF declined, with a more pronounced decrease in SIF observed in natural forests. (3) Vegetation in northern Xinjiang showed a significantly stronger response to climate change than that in southern Xinjiang, with physiological parameters (SIF) being more sensitive than structural parameters (NDVI). The R1, R2, and R3 ecological regions were primarily influenced by long-term climate change, whereas the R4 and R5 regions were more affected by short-term climate change. Natural grasslands showed a significantly stronger response than forests, particularly in areas with lower vegetation cover that are more structurally impacted. This study provides an important scientific basis for ecological management and climate adaptation in Xinjiang, emphasizing the need for differentiated strategies across ecological regions to support sustainable development. Full article
Show Figures

Figure 1

Figure 1
<p>Study area of the Xinjiang arid region in northwest China (Vegetation is classified as follows: Forest (red), high coverage grassland (HCG, &gt;50%, dark green), moderate coverage grassland (MCG, 20–50%, medium green), and low coverage grassland (LCG, 5–20%, light green). The ecological regions (R1–R5) are delineated with different hatching patterns, and meteorological stations are marked with red dots. The map was created using the standard map approved by the Ministry of Natural Resources of China (review number GS (2024) 0650). The base map provided by the Ministry of Natural Resources was used without any modifications. Similarly, all other maps in this study were created using standardized methods and remain unaltered.</p>
Full article ">Figure 2
<p>Technical roadmap for research of vegetation responses to climate change.</p>
Full article ">Figure 3
<p>Temporal variation of SPEI in Xinjiang from 2000 to 2020 (<b>a</b>) Temporal variation of SPEI at a 3-month timescale (SPEI-03). (<b>b</b>) Temporal variation of SPEI at a 6-month timescale (SPEI-06). (<b>c</b>) Temporal variation of SPEI at a 12-month timescale (SPEI-12). (<b>d</b>) Temporal variation of SPEI at a 24-month timescale (SPEI-24). Each panel shows the SPEI data series (blue) and trend line (red). Statistical values, including Z-score, <span class="html-italic">p</span>-value, and slope, are provided for each timescale to indicate the trend significance and direction.</p>
Full article ">Figure 4
<p>Spatial distribution of multi-year average SPEI in Xinjiang from 2000 to 2020 across different timescales. (<b>a</b>) Spatial distribution of SPEI at a 3-month timescale (SPEI-03). (<b>b</b>) Spatial distribution of SPEI at a 6-month timescale (SPEI-06). (<b>c</b>) Spatial distribution of SPEI at a 12-month timescale (SPEI-12). (<b>d</b>) Spatial distribution of SPEI at a 24-month timescale (SPEI-24). Red areas indicate drier conditions, whereas blue areas represent wetter conditions.</p>
Full article ">Figure 5
<p>Inter-annual variation analysis of vegetation NDVI and SIF. (<b>a</b>) Annual mean NDVI with trend line and confidence interval. The trend line (red) represents the linear trend, with the equation y = 0.0012x + 0.1276y = 0.0012x + 0.1276y = 0.0012x + 0.1276 and a correlation coefficient of 0.8929. (<b>b</b>) Annual mean SIF with trend line and confidence interval. The trend line (red) shows the linear trend, with the equation y = 0.0005x + 0.0741y = 0.0005x + 0.0741y = 0.0005x + 0.0741 and a correlation coefficient of 0.7521. Blue triangles represent observed data, and the shaded area indicates the confidence interval around the trend line.</p>
Full article ">Figure 6
<p>Spatial distribution of mean NDVI and SIF in Xinjiang from 2000 to 2020. (<b>a</b>) Spatial distribution of mean NDVI, representing vegetation structural conditions across Xinjiang. The color bar indicates NDVI values, with yellow to dark green representing increasing vegetation coverage from 0.0 to 1.0. (<b>b</b>) Spatial distribution of mean SIF, indicating vegetation physiological activity levels. The color bar reflects SIF values, ranging from 0.00 to 0.20 W·m<sup>−</sup><sup>2</sup>·μm<sup>−</sup><sup>1</sup>·sr<sup>−</sup><sup>1</sup>, with dark green areas showing higher fluorescence.</p>
Full article ">Figure 7
<p>Spatial distribution of correlation coefficients for NDVI and SIF from 2000 to 2020 in Xinjiang. (<b>a</b>) Spatial distribution of the correlation coefficient between NDVI and SPEI across pixels over the study period. (<b>b</b>) Spatial distribution of the correlation coefficient between SIF and SPEI across pixels over the study period. The color scale represents correlation values from −1 to 1, where blue areas indicate a strong positive correlation and red areas indicate a strong negative correlation. The inset bar chart shows the proportion of pixels with positive and negative correlation trends.</p>
Full article ">Figure 8
<p>Spatial distribution and changes in vegetation types from 2000 to 2020 in Xinjiang. (<b>a</b>) Spatial distribution of vegetation types, including forest, high-coverage grassland (HCG), moderate-coverage grassland (MCG), and low-coverage grassland (LCG). (<b>b</b>) Spatial distribution of vegetation change over the study period, identifying common, increased, and decreased areas. The bar chart inserted shows the percentage of each type.</p>
Full article ">Figure 9
<p>Trends in NDVI and SIF for grassland and forested areas (2000–2020). For forest areas, NDVI shows a statistically significant increasing trend (y = 0.0017x + 0.5584, Corr. = 0.569), whereas SIF displays a slight decreasing trend (y = −0.0004x + 0.1943, Corr. = −0.2389). For grassland areas, NDVI exhibits a positive trend (y = 0.0017x + 0.194, Corr. = 0.8019), with a minimal increase in SIF (y = 0.0001x + 0.0779, Corr. = 0.0712). The shaded regions indicate the confidence intervals for each regression line.</p>
Full article ">Figure 10
<p>Spatial distribution of maximum correlation coefficients (R<sub>max</sub>) between vegetation indices and SPEI in Xinjiang. (<b>a</b>) The spatial distribution of the maximum correlation coefficients (R<sub>max</sub>) between NDVI and SPEI across Xinjiang, with values ranging from −0.372 to 0.745, indicating varying vegetation responses to climatic changes in different ecological regions. (<b>b</b>) The spatial distribution of the maximum correlation coefficients (R<sub>max</sub>) between SIF and SPEI, with values from −0.8 to 0.9. This figure highlights the spatial variation in vegetation sensitivity to drought stress across different regions.</p>
Full article ">Figure 11
<p>Maximum correlation coefficients between vegetation indices and SPEI. Box plots depict the distribution of maximum correlation coefficients (R<sub>max</sub>) between NDVI and SIF with SPEI, separated by ecological regions (R1–R5) and vegetation types (Forest, HCG, MCG, and LCG). (<b>a</b>) shows NDVI correlations across ecological regions, whereas (<b>b</b>) displays SIF correlations. (<b>c</b>,<b>d</b>) present NDVI and SIF correlations, respectively, by vegetation type.</p>
Full article ">Figure 12
<p>Spatial distribution of the SPEI time scales corresponding to the maximum correlation coefficients between vegetation conditions represented by NDVI (<b>a</b>) and SIF (<b>b</b>) and SPEI. (<b>a</b>) NDVI and (<b>b</b>) SIF are shown with SPEI03, SPEI06, SPEI12, and SPEI24, representing the dominant timescales of vegetation response to drought. This figure illustrates the temporal dynamics of vegetation response to drought stress.</p>
Full article ">Figure 13
<p>Area proportion of SPEI time scales corresponding to maximum correlation coefficients between vegetation indices and SPEI across different ecological regions and vegetation types. (<b>a</b>) displays NDVI correlations by region, whereas (<b>b</b>) shows SIF correlations by region. (<b>c</b>,<b>d</b>) present NDVI and SIF correlations, respectively, by vegetation type.</p>
Full article ">
19 pages, 6466 KiB  
Article
Increases in Temperature and Precipitation in the Different Regions of the Tarim River Basin Between 1961 and 2021 Show Spatial and Temporal Heterogeneity
by Siqi Wang, Ailiyaer Aihaiti, Ali Mamtimin, Hajigul Sayit, Jian Peng, Yongqiang Liu, Yu Wang, Jiacheng Gao, Meiqi Song, Cong Wen, Fan Yang, Chenglong Zhou, Wen Huo and Yisilamu Wulayin
Remote Sens. 2024, 16(23), 4612; https://doi.org/10.3390/rs16234612 - 9 Dec 2024
Viewed by 653
Abstract
The Tarim River Basin (TRB) faces significant ecological challenges due to global warming, making it essential to understand the changes in the climates of its sub-basins for effective management. With this aim, data from national meteorological stations, ERA5_Land, and climate indices from 1961 [...] Read more.
The Tarim River Basin (TRB) faces significant ecological challenges due to global warming, making it essential to understand the changes in the climates of its sub-basins for effective management. With this aim, data from national meteorological stations, ERA5_Land, and climate indices from 1961 to 2021 were used to analyze the temperature and precipitation variations in the TRB and its sub-basins and to assess their climate sensitivity. Our results showed that (1) the annual mean temperature increased by 0.2 °C/10a and precipitation increased by 7.1 mm/10a between 1961 and 2021. Moreover, precipitation trends varied significantly among the sub-basins, with that in the Aksu River Basin increasing the most (12.9 mm/10a) and that in the Cherchen River Basin increasing the least (1.9 mm/10a). Moreover, ERA5_Land data accurately reproduced the spatiotemporal patterns of temperature (correlation 0.92) and precipitation (correlation 0.72) in the TRB. (2) Empirical Orthogonal Function analysis identified the northern sections of the Kaidu, Weigan, and Yerqiang river basins as centers of temperature sensitivity and the western part of the Kaidu and Cherchen River Basin as the center of precipitation sensitivity. (3) Global warming is closely correlated with sub-basin temperature (correlation above 0.5) but weakly correlated with precipitation (correlation 0.2~0.5). TRB temperatures were found to have a positive correlation with AMO, especially in the Hotan, Kashgar, and Aksu river basins, and a negative correlation with AO and NAO, particularly in the Keriya and Hotan river basins. Precipitation correlations between the climate indices were complex and varied across the different basins. Full article
(This article belongs to the Section Atmospheric Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Location map of the study area. The shaded color indicates the elevation of the TRB (m); red color symbols represent the distribution of stations in the sub-basins of the TRB.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) Time series of the annual mean temperature and precipitation anomalies, red represents a positive anomaly, blue represents a negative anomaly; solid lines represent the detrended 10-year running average; (<b>c</b>,<b>d</b>) linear trends of annual mean temperature (°C) and precipitation (mm) in the TRB from 1961 to 2021; (<b>e</b>,<b>f</b>) spatial distribution of annual mean temperature and mean PRCPTOT in the TRB from 1961 to 2021.</p>
Full article ">Figure 3
<p>Spatial distribution of (<b>a</b>,<b>c</b>) the first eigenvector and (<b>b</b>,<b>d</b>) second eigenvector of the EOF analysis of the annual mean temperature and PRCTOP in the TRB from 1961 to 2021.</p>
Full article ">Figure 4
<p>Temporal trends of (<b>a</b>,<b>c</b>) EOF1 and (<b>b</b>,<b>d</b>) EOF2 for annual mean temperature and PRCPTOT in the TRB from 1961 to 2021, where the solid line is the PC value and the dashed line is the linearly fitted trend.</p>
Full article ">Figure 5
<p>Trends in (<b>a</b>) annual mean temperature and (<b>b</b>) precipitation in the sub-basins of the TRB from 1961 to 2021.</p>
Full article ">Figure 6
<p>The annual mean temperature anomaly after detrending in the TRB sub-basins from 1961 to 2021, where the reference period was 1981–2010, red represents a positive anomaly, blue represents a negative anomaly.</p>
Full article ">Figure 7
<p>The annual precipitation anomaly after detrending in the TRB sub-basins from 1961 to 2021, where the reference period was 1981–2010, red represents a positive anomaly, blue represents a negative anomaly.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>d</b>) The spatial distribution of seasonal mean temperature (°C) in the sub-basins from 1961 to 2021; (<b>f</b>–<b>i</b>) the spatial distribution of seasonal precipitation (mm) in the sub-basins from 1961 to 2021; (<b>e</b>,<b>j</b>) thermal maps of seasonal mean temperature (°C) and seasonal precipitation (mm) in the sub-basins.</p>
Full article ">Figure 9
<p>Comparison of observed quantiles for (<b>a</b>) temperature and (<b>b</b>) precipitation with ERA5_Land data. The solid red line represents the 1:1 line.</p>
Full article ">Figure 10
<p>Spatial distribution of (<b>a</b>) annual mean temperature and (<b>b</b>) annual precipitation and (<b>c</b>,<b>d</b>) their spatial variation trends in the TRB and its sub-basins based on ERA5_Land data.</p>
Full article ">Figure 11
<p>The correlation between the annual mean temperature (°C) and annual precipitation (mm) in the sub-basins and global warming (* represents significance <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 12
<p>The correlations between the 10-year running average of (<b>a</b>) annual mean temperature and (<b>b</b>) precipitation in the TRB and the 10-year running average of each climate index after the detrend (* represents significance <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
22 pages, 25759 KiB  
Article
Characteristics of Atmospheric Circulation Patterns and the Associated Diurnal Variation Characteristics of Precipitation in Summer over the Complex Terrain in Northern Xinjiang, Northwest China
by Abuduwaili Abulikemu, Abidan Abuduaini, Zhiyi Li, Kefeng Zhu, Ali Mamtimin, Junqiang Yao, Yong Zeng and Dawei An
Remote Sens. 2024, 16(23), 4520; https://doi.org/10.3390/rs16234520 - 2 Dec 2024
Viewed by 727
Abstract
Statistical characteristics of atmospheric circulation patterns (ACPs) and associated diurnal variation characteristics (DVCs) of precipitation in summer (June–August) from 2015 to 2019 over the complex terrain in northern Xinjiang (NX), northwestern arid region of China, were investigated based on NCEP FNL reanalysis data [...] Read more.
Statistical characteristics of atmospheric circulation patterns (ACPs) and associated diurnal variation characteristics (DVCs) of precipitation in summer (June–August) from 2015 to 2019 over the complex terrain in northern Xinjiang (NX), northwestern arid region of China, were investigated based on NCEP FNL reanalysis data and Weather Research and Forecasting model simulation data from Nanjing University (WRF-NJU). The results show that six different ACPs (Type 1–6) were identified based on the Simulated ANealing and Diversified RAndomization (SANDRA), exhibiting significant differences in major-influencing synoptic systems and basic meteorological environments. Types 5, 3, and 2 were the most prevalent three patterns, accounting for 21.6%, 19.7%, and 17.7%, respectively. Type 5 mainly occurred in June and July, while Types 3 and 2 mainly occurred in August and July, respectively. From the perspective of DVCs, Type 1 reached its peak at midnight, while Type 5 was most frequent in the afternoon and morning. The overall DVCs of hourly precipitation intensity and frequency demonstrated a unimodal structure, with a peak occurring at around 16 Local Solar Time (LST). Basic meteorological elements in various terrain regions exhibit significant diurnal variation, with marked differences between mountainous and basin areas under different ACPs. In Types 3 and 6, meteorological elements significantly influence precipitation enhancement by promoting the convergence and uplift of low-level wind fields and maintaining high relative humidity (RH). The Altay Mountains region and Western Mountainous regions experience dominant westerly winds under these conditions, while the Junggar Basin and Ili River Valley regions benefit from counterclockwise water vapor transport associated with the Iranian Subtropical High in Type 6, which increases RH. Collectively, these factors facilitate the formation and development of precipitation. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Geographical location and terrain altitude (shading, unit: m) of the study area, the area within the red line represents Northern Xinjiang (NX, excluding Turpan and Hami), water-blue areas show lakes, and the water-blue lines represent rivers (the filled red area in the small globe in the upper left corner shows the location of the study area from a broad scope). (<b>b</b>) Spatial distribution of annually averaged accumulated summer (June–August) precipitation (shading, unit: mm) during the period of 2015–2019 and terrain altitude (white contours with intervals of 1000 m, the numbers in black font represent the isoheight lines of 1000 m, 2000 m, and 3000 m). The small illustration figure in the upper left corner shows the different regions divided according to terrain altitude in the study area (AM, TM, WM, JB, and RV represent for the Altay Mountains region, Tishan Mountains region, Western Mountainous region, Junggar Basin, and Ili River Valley, respectively).</p>
Full article ">Figure 2
<p>Six atmospheric circulation patterns (ACPs) were classified using the SANDRA method during the summers of 2015–2019 in Northern Xinjiang (NX) based on the NCEP FNL data. (<b>a</b>–<b>f</b>) Geopotential height (black lines, unit: gpm) and horizontal wind velocity (shading, unit: m s<sup>−1</sup>) at 200 hPa. (<b>g</b>–<b>l</b>) Geopotential height (shading, unit: gpm) at 500 hPa; and (<b>m</b>–<b>r</b>) the water vapor flux divergence (shading, unit: 10<sup>−3</sup> g cm<sup>−1</sup> hPa s<sup>−1</sup>) and horizontal wind velocity (vector arrows, unit: m s<sup>−1</sup>) at the 700 hPa. The bold red closed line in each panel represents the study area (i.e., NX).</p>
Full article ">Figure 3
<p>(<b>a</b>) Annual variations in hourly occurrence frequencies of six types of atmospheric circulation patterns (ACPs) in northern Xinjiang (NX) during the summer of 2015–2019. (<b>b</b>) is the same as (<b>a</b>) but for the temporal variations on a 10-day time scale. (<b>c</b>) is the same as (<b>a</b>) but for the temporal variations on a diurnal scale.</p>
Full article ">Figure 4
<p>Diurnal variation characteristics (DVCs) of (<b>a</b>) hourly averaged precipitation amount, (<b>b</b>) hourly averaged precipitation frequency, and (<b>c</b>) hourly averaged precipitation intensity in six types of ACPs in northern Xinjiang (NX) during the summer of 2015–2019 (the colorful lines represent the values in each type of six ACPs, and the grey bars show the averaged values for the whole summer).</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>f</b>) Spatial distributions of precipitation anomaly percentages in six types of atmospheric circulation patterns (ACPs) in northern Xinjiang (NX) during the summer of 2015–2019 (shading, units: %, grey contours with an interval of 1000 m represent topography, and the numbers in black font represent the isoheight lines of 1000 m, 2000 m, and 3000 m).</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>e</b>) Diurnal variation characteristics (DVCs) of horizontal wind (vector arrows, unit: m s<sup>−1</sup>) at the model’s lowest eta level (eta = 1.00000, represents the near-surface level) of WRF-NJU data in the different regions divided according to terrain altitude in the study area during summer from 2015 to 2019 in six atmospheric circulation patterns (ACPs). The abbreviations in the upper left corner of each panel represent the different regions, i.e., AM, TM, WM, JB, and RV represent the Altay Mountains region, Tishan Mountains region, Western Mountainous region, Junggar Basin, and Ili River Valley, respectively (also shown in <a href="#remotesensing-16-04520-f001" class="html-fig">Figure 1</a>b).</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>e</b>) Diurnal variation characteristics (DVCs) of air temperature (colored contours, unit: K) at the model’s lowest eta level (eta = 1.00000, represents the near-surface level) of WRF-NJU data in the different regions divided according to terrain altitude in the study area during summer from 2015 to 2019 in six atmospheric circulation patterns (ACPs). The abbreviations in the upper left corner of each panel represent the different regions, i.e., AM, TM, WM, JB, and RV represent the Altay Mountains region, Tishan Mountains region, Western Mountainous region, Junggar Basin, and Ili River Valley, respectively (also shown in <a href="#remotesensing-16-04520-f001" class="html-fig">Figure 1</a>b).</p>
Full article ">Figure 8
<p>Diurnal variation characteristics (DVCs) of relative humidity (shading, unit: %) at the model’s lowest eta level (eta = 1.00000, represents the near-surface level) of WRF-NJU data in the different regions divided according to terrain altitude in the study area during summer from 2015 to 2019 in six atmospheric circulation patterns (ACPs). The abbreviations in the vertical coordinates represent the different regions, i.e., AM, TM, WM, JB, and RV represent the Altay Mountains region, Tishan Mountains region, West-ern Mountainous region, Junggar Basin, and Ili River Valley, respectively (also shown in <a href="#remotesensing-16-04520-f001" class="html-fig">Figure 1</a>b).</p>
Full article ">
23 pages, 21957 KiB  
Article
Terrain Analysis According to Multiscale Surface Roughness in the Taklimakan Desert
by Sebastiano Trevisani and Peter L. Guth
Land 2024, 13(11), 1843; https://doi.org/10.3390/land13111843 - 5 Nov 2024
Viewed by 873
Abstract
Surface roughness, interpreted in the wide sense of surface texture, is a generic term referring to a variety of aspects and scales of spatial variability of surfaces. The analysis of solid earth surface roughness is useful for understanding, characterizing, and monitoring geomorphic factors [...] Read more.
Surface roughness, interpreted in the wide sense of surface texture, is a generic term referring to a variety of aspects and scales of spatial variability of surfaces. The analysis of solid earth surface roughness is useful for understanding, characterizing, and monitoring geomorphic factors at multiple spatiotemporal scales. The different geomorphic features characterizing a landscape exhibit specific characteristics and scales of surface texture. The capability to selectively analyze specific roughness metrics at multiple spatial scales represents a key tool in geomorphometric analysis. This research presents a simplified geostatistical approach for the multiscale analysis of surface roughness, or of image texture in the case of images, that is highly informative and interpretable. The implemented approach is able to describe two main aspects of short-range surface roughness: omnidirectional roughness and roughness anisotropy. Adopting simple upscaling approaches, it is possible to perform a multiscale analysis of roughness. An overview of the information extraction potential of the approach is shown for the analysis of a portion of the Taklimakan desert (China) using a 30 m resolution DEM derived from the Copernicus Glo-30 DSM. The multiscale roughness indexes are used as input features for unsupervised and supervised learning tasks. The approach can be refined both from the perspective of the multiscale analysis as well as in relation to the surface roughness indexes considered. However, even in its present, simplified form, it can find direct applications in relation to multiple contexts and research topics. Full article
(This article belongs to the Section Land, Soil and Water)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Reprojected COP DEM (30 m resolution, UTM F44) of the area of interest overlaid on the hillshade.</p>
Full article ">Figure 2
<p>Sentinel-2 true color RGB image (bands 4, 3, and 2) of the study area, with the main dune morphologies labeled.</p>
Full article ">Figure 3
<p>Main dune morphologies in the study area, visualized using Sentinel-2 imagery (<b>a</b>), hillshade (<b>b</b>), and residual DEM (<b>c</b>). From top to bottom: network/transverse dunes, longitudinal and transverse dunes, and dome-shaped dunes.</p>
Full article ">Figure 4
<p>Mixed morphologies in the area of interest, visualized using Sentinel-2 imagery (<b>a</b>), hillshade (<b>b</b>), and residual DEM (<b>c</b>). From top to bottom: outcropping bedrock with shadow and linear dunes, fluvial morphology, and a flat area with minor dune morphologies.</p>
Full article ">Figure 5
<p>RA direction, where the RA strength is higher than 0.3, overlaid on the hillshade (<b>a</b>) and the residual DEM (<b>b</b>) calculated for level L2.</p>
Full article ">Figure 6
<p>Omnidirectional short-range roughness (m) for the different resolutions. Different color scales for each diagram.</p>
Full article ">Figure 7
<p>Roughness anisotropy strength at different resolutions. Different color scales for each diagram.</p>
Full article ">Figure 8
<p>RGB image (each band normalized) of 3 omnidirectional roughness indexes computed at different resolutions (B = L1; G = L2; R = L4). Despite the high correlation of the three indexes, they differentiate very well the morphological features of the area. For example, they markedly highlight the characteristic smoothness of interdune areas of the longitudinal dunes south of the mountain ridge.</p>
Full article ">Figure 9
<p>RGB image (each band normalized) of 3 anisotropy strength roughness indexes computed at different resolutions (B = L1; G = L4; R = L16). In the dune fields north of the mountains, long-wavelength anisotropic features prevail; in contrast, for the southern longitudinal dunes, shorter anisotropic features (L4) are highlighted.</p>
Full article ">Figure 10
<p>Landscape clustered according to multiscale surface roughness indexes. The cluster centers in terms of OR and RA are described in <a href="#land-13-01843-f011" class="html-fig">Figure 11</a>.</p>
Full article ">Figure 11
<p>Cluster centers of the 7 classes resulting from K-means clustering for OR and RA at the different levels.</p>
Full article ">Figure 12
<p>MRI clustering results in the area of the northern dune field, characterized by network and transverse dunes. Clustering results (<b>d</b>), Sentinel-2 imagery (<b>a</b>), hillshade (<b>b</b>), and residual DEM (<b>c</b>).</p>
Full article ">Figure 13
<p>MRI clustering results in the area of the southern longitudinal dune fields. Clustering results (<b>d</b>), Sentinel-2 imagery (<b>a</b>), hillshade (<b>b</b>), and residual DEM (<b>c</b>).</p>
Full article ">Figure 14
<p>MRI clustering results in the area with fluvial morphology, outcropping bedrock, and dome dune fields. Clustering results (<b>d</b>), Sentinel-2 imagery (<b>a</b>), hillshade (<b>b</b>), and residual DEM (<b>c</b>).</p>
Full article ">Figure 15
<p>Manual classification of crest lines (<b>a</b>) for large dunes using visual analysis of slope (<b>b</b>), profile curvature (<b>c</b>), and residual DEM (<b>d</b>). Crest lines are associated with high positive profile curvature, strongly positive residual DEM, and low slope. These locations are then located in areas in which the neighborhood is characterized by an abrupt variation in the selected geomorphometric derivatives.</p>
Full article ">Figure 16
<p>Probability of observing a crest obtained by means of RF considering the GDs integrated with the MRIs (<b>a</b>) and only the five GDs (<b>b</b>) to obtain details of the study area, which is located on the western mountain ridge. The RF model integrating the MRIs provides a more focused prediction of crest lines of large dunes. In (<b>c</b>), the prediction of the crest lines of the two RF models is compared. Pixels with a probability higher than 0.8 have been classified as crests. The transparent color is where both models predicted a not-crest pixel, green is where both models predicted a crest, and red and blue are where, respectively, only RF GDs and RF GDs + MRIs predicted a crest.</p>
Full article ">Figure 17
<p>Variables’ importance in the two RF models according to the mean decrease in the Gini index ((<b>a</b>), RF based on GDs; (<b>b</b>), RF based on GDs integrated with MRIs).</p>
Full article ">Figure 18
<p>Prediction of crest lines with the RF model based on GDs and MRIs of an unseen area ((<b>c</b>), green box) external to the one with reference data used for training and testing ((<b>c</b>), red box). The reference crest lines (<b>a</b>) have been manually digitized by means of visual analysis of the profile curvature, the residual DEM, and the slope; the predicted crest lines have been derived as crests of all of the pixels with a probability above 0.8. The predicted crest lines are compared with the reference data (<b>b</b>). Green pixels are correctly classified as crests; red and blue pixels are incorrectly classified, respectively, as crests and not crests.</p>
Full article ">
20 pages, 9093 KiB  
Article
The Role of Subsurface Changes and Environmental Factors in Shaping Urban Heat Islands in Southern Xinjiang
by Cong Wen, Hajigul Sayit, Ali Mamtimin, Yu Wang, Jian Peng, Ailiyaer Aihaiti, Meiqi Song, Jiacheng Gao, Junjian Liu, Yisilamu Wulayin, Fan Yang, Wen Huo and Chenglong Zhou
Remote Sens. 2024, 16(21), 4089; https://doi.org/10.3390/rs16214089 - 1 Nov 2024
Viewed by 642
Abstract
The urban heat island (UHI) effect is one of the most prominent surface climate changes driven by human activities. This study examines the UHI characteristics and influencing factors in the Southern Xinjiang urban agglomeration using MODIS satellite data combined with observational datasets. Our [...] Read more.
The urban heat island (UHI) effect is one of the most prominent surface climate changes driven by human activities. This study examines the UHI characteristics and influencing factors in the Southern Xinjiang urban agglomeration using MODIS satellite data combined with observational datasets. Our results reveal a significant increase in impervious surfaces in the region between 1995 and 2015, with the most rapid expansion occurring from 2010 to 2015. This urban expansion is the primary driver of changes in UHI intensity. The analysis from 2000 to 2015 shows substantial spatial variation in UHI effects across cities. Hotan recorded the highest annual average daytime UHI intensity of 3.7 °C, while Aksu exhibited the lowest at approximately 1.6 °C. Daytime UHI intensity generally increased during the study period, with the highest intensities observed in the summer. However, nighttime UHI trends varied across cities, with most showing an increase in intensity. Temperature, precipitation, and aerosol optical depth (AOD) were identified as the main factors influencing annual average daytime UHI intensity, while PM10 concentration showed a weak and inconsistent correlation with UHI intensity, varying by city and season. Full article
(This article belongs to the Section Atmospheric Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Map of urban agglomeration distribution in South Xinjiang (the number under the city name represents the population in that city in 2015).</p>
Full article ">Figure 2
<p>Trends in impervious surface changes in the South Xinjiang urban agglomeration, 1995–2015.</p>
Full article ">Figure 3
<p>Spatial changes in impervious surface in the Southern Xinjiang urban agglomeration, 1995–2015.</p>
Full article ">Figure 4
<p>Characteristics of daytime land surface temperature changes in urban agglomerations in South Xinjiang.</p>
Full article ">Figure 5
<p>Characteristics of nighttime land surface temperature changes in urban agglomerations in South Xinjiang.</p>
Full article ">Figure 6
<p>Spatial distribution of urban heat island intensity in urban agglomerations in South Xinjiang during daytime.</p>
Full article ">Figure 7
<p>Spatial distribution of urban heat island intensity in urban agglomerations in South Xinjiang during nighttime.</p>
Full article ">Figure 8
<p>Annual variation in urban heat island intensity in the South Xinjiang urban agglomeration, 2000–2015 ((<b>a</b>). Aksu, (<b>b</b>). Atushi, (<b>c</b>). Hotan, (<b>d</b>). Kashgar, (<b>e</b>). Korla).</p>
Full article ">Figure 9
<p>Seasonal variation in urban heat island intensity in the South Xinjiang urban agglomeration, 2000–2015 ((<b>a</b>). Spring, (<b>b</b>). Summer, (<b>c</b>). Autumn, (<b>d</b>). Winter).</p>
Full article ">
17 pages, 7168 KiB  
Article
Evaluating the Prediction Performance of the WRF-CUACE Model in Xinjiang, China
by Yisilamu Wulayin, Huoqing Li, Lei Zhang, Ali Mamtimin, Junjian Liu, Wen Huo and Hongli Liu
Remote Sens. 2024, 16(19), 3747; https://doi.org/10.3390/rs16193747 - 9 Oct 2024
Viewed by 941
Abstract
Dust and air pollution events are increasingly occurring around the Taklimakan Desert in southern Xinjiang and in the urban areas of northern Xinjiang. Predicting such events is crucial for the advancement, growth, and prosperity of communities. This study evaluated a dust and air [...] Read more.
Dust and air pollution events are increasingly occurring around the Taklimakan Desert in southern Xinjiang and in the urban areas of northern Xinjiang. Predicting such events is crucial for the advancement, growth, and prosperity of communities. This study evaluated a dust and air pollution forecasting system based on the Weather Research and Forecasting model coupled with the China Meteorological Administration Chemistry Environment (WRF-CUACE) model using ground and satellite observations. The results showed that the forecasting system accurately predicted the formation, development, and termination of dust events. It demonstrated good capability for predicting the evolution and spatial distribution of dust storms, although it overestimated dust intensity. Specifically, the correlation coefficient (R) between simulated and observed PM10 was up to 0.85 with a mean absolute error (MAE) of 721.36 µg·m−3 during dust storm periods. During air pollution events, the forecasting system displayed notable variations in predictive accuracy across various urban areas. The simulated trends of PM2.5 and the Air Quality Index (AQI) closely aligned with the actual observations in Ürümqi. The R for simulated and observed PM2.5 concentrations at 24 and 48 h intervals were 0.60 and 0.54, respectively, with MAEs of 28.92 µg·m−3 and 29.10 µg·m−3, respectively. The correlation coefficients for simulated and observed AQIs at 24 and 48 h intervals were 0.79 and 0.70, respectively, with MAEs of 24.21 and 27.56, respectively. The evolution of the simulated PM10 was consistent with observations despite relatively high concentrations. The simulated PM2.5 concentrations in Changji and Shihezi were notably lower than those observed, resulting in a lower AQI. For PM10, the simulation–observation error was relatively small; however, the trends were inconsistent. Future research should focus on optimizing model parameterization schemes and emission source data. Full article
Show Figures

Figure 1

Figure 1
<p>Domains and terrain height of the WRF-CUACE model. Black dots denote the location of environmental monitoring stations.</p>
Full article ">Figure 2
<p>Comparison between simulated and observed 2 m temperature data at (<b>a</b>) Hotan, (<b>b</b>) Kashgar, (<b>c</b>) Aksu, (<b>d</b>) Atush, (<b>e</b>) Korla, and (<b>f</b>) Turpan stations.</p>
Full article ">Figure 3
<p>Comparison between simulated and observed 10 m wind speed data at (<b>a</b>) Hotan, (<b>b</b>) Kashgar, (<b>c</b>) Aksu, (<b>d</b>) Atush, (<b>e</b>) Korla, and (<b>f</b>) Turpan stations.</p>
Full article ">Figure 4
<p>Comparison between modeled and observed particulate matter (PM)10 data at (<b>a</b>) Hotan, (<b>b</b>) Kashgar, (<b>c</b>) Aksu, (<b>d</b>) Atushi, (<b>e</b>) Korla, and (<b>f</b>) Turpan stations.</p>
Full article ">Figure 5
<p>Comparison between simulated aerosol optical depth (AOD) and MODIS satellite observations from 24 to 26 March 2022.</p>
Full article ">Figure 6
<p>Spatial distribution of averaged dust emission fluxes coupled with 10 m wind speed (<b>a</b>,<b>b</b>), dust transport fluxes and divergence (<b>c</b>,<b>d</b>), and dry dust deposition fluxes (<b>e</b>,<b>f</b>) from 24 to 25 March 2022.</p>
Full article ">Figure 7
<p>Vertical distribution of averaged dust concentrations and wind vectors along line AB [(92.0°E, 40.5°N) to (76.0°E, 38.0°N)] from 24 to 26 March 2022. Brown field represents terrain height (units: m), and blue lines represents boundary layer height (units: m).</p>
Full article ">Figure 8
<p>Comparison between simulated and observed PM2.5 and PM10 concentrations in (<b>a</b>,<b>b</b>) Ürümqi, (<b>c</b>,<b>d</b>) Changji, and (<b>e</b>,<b>f</b>) Shihezi from 14 to 26 December 2023.</p>
Full article ">Figure 9
<p>Comparison between simulated and observed AQIs in (<b>a</b>) Ürümqi, (<b>b</b>) Changji, and (<b>c</b>) Shihezi from 14 to 26 December 2023.</p>
Full article ">
19 pages, 7674 KiB  
Article
Microbial Community Structure in the Taklimakan Desert: The Importance of Nutrient Levels in Medium and Culture Methods
by Feng Wen, Siyuan Wu, Xiaoxia Luo, Linquan Bai and Zhanfeng Xia
Biology 2024, 13(10), 797; https://doi.org/10.3390/biology13100797 - 6 Oct 2024
Viewed by 988
Abstract
Although the Taklimakan Desert lacks the necessary nutrients and conditions to support an extensive ecosystem, it is a treasure trove of extremophile resources with special structures and functions. We analyzed the bacterial communities using oligotrophic medium and velvet cloth replicate combined with an [...] Read more.
Although the Taklimakan Desert lacks the necessary nutrients and conditions to support an extensive ecosystem, it is a treasure trove of extremophile resources with special structures and functions. We analyzed the bacterial communities using oligotrophic medium and velvet cloth replicate combined with an extended culture duration. We isolated numerous uncultured microorganisms and rare microorganisms belonging to genera not often isolated or recently described, such as Aliihoeflea, Halodurantibacterium, and Indioceanicola. A total of 669 strains were isolated from the soil of the Taklimakan Desert, which were classified into 5 phyla, 7 classes, 25 orders, 42 families, 83 genera, and 379 species. Among them, 148 strains were potential new species. Our data show that even when working with samples from extreme environments, simple approaches are still useful for cultivating stubborn microbes. Through comparing the isolation effects of different nutrient levels on microbial diversity and abundance, the results show that reducing the nutrient level of the medium was more conducive to improving the culturability of microorganisms in low-nutrient environments, while the high-nutrient medium was more suitable for the isolation of dominant fast-growing strains. This study helps to better reflect the diversity of microbial resources and lays a foundation for the further research and utilization of soil microbial resources in the Taklimakan Desert. Full article
(This article belongs to the Section Conservation Biology and Biodiversity)
Show Figures

Figure 1

Figure 1
<p>Sample collection points in Taklimakan Desert. Note: A: South, B: North, C: West, D: East, E: Central.</p>
Full article ">Figure 2
<p>PCoA analysis of five sand samples in the Taklimakan Desert. Note: A: South, B: North, C: West, D: East, E: Central.</p>
Full article ">Figure 3
<p>The top ten categories of bacteria at different taxonomic levels: (<b>a</b>) phylum, (<b>b</b>) class, (<b>c</b>) order, (<b>d</b>) family, and (<b>e</b>) genus. Note: A: South, B: North, C: West, D: East, E: Central.</p>
Full article ">Figure 4
<p>Total number of soil microbial plate colonies isolated on media of different nutrient levels. Note: A: South, B: North, C: West, D: East, E: Central.</p>
Full article ">Figure 5
<p>Microorganisms isolated from five soil samples from the Taklimakan Desert. The chord plot (<b>a</b>) of the bacterial community in the desert samples represents the difference in the distribution of microorganisms obtained from the five Taklimakan Desert soil samples. The Venn diagram of the bacterial community in the desert sample represents the quantitative distribution of genus (<b>b</b>). These samples are represented by different circles: a Venn diagram representing the number of genera detected in each sample and the overlap of genera in the sample.</p>
Full article ">Figure 5 Cont.
<p>Microorganisms isolated from five soil samples from the Taklimakan Desert. The chord plot (<b>a</b>) of the bacterial community in the desert samples represents the difference in the distribution of microorganisms obtained from the five Taklimakan Desert soil samples. The Venn diagram of the bacterial community in the desert sample represents the quantitative distribution of genus (<b>b</b>). These samples are represented by different circles: a Venn diagram representing the number of genera detected in each sample and the overlap of genera in the sample.</p>
Full article ">Figure 6
<p>Microorganisms isolated from the Taklimakan Desert at different medium nutrient levels.</p>
Full article ">Figure 7
<p>Proportion of rare bacteria in each nutrient level medium.</p>
Full article ">
20 pages, 4749 KiB  
Article
Variation in Root-Associated Microbial Communities among Three Different Plant Species in Natural Desert Ecosystem
by Yulin Zhang, Yi Du, Zhihao Zhang, Waqar Islam and Fanjiang Zeng
Plants 2024, 13(17), 2468; https://doi.org/10.3390/plants13172468 - 3 Sep 2024
Viewed by 1413
Abstract
The process and function that underlie the assembly of root-associated microbiomes may be strongly linked to the survival strategy of plants. However, the assembly and functional changes of root-associated microbial communities in different desert plants in natural desert ecosystems are still unclear. Thus, [...] Read more.
The process and function that underlie the assembly of root-associated microbiomes may be strongly linked to the survival strategy of plants. However, the assembly and functional changes of root-associated microbial communities in different desert plants in natural desert ecosystems are still unclear. Thus, we studied the microbial communities and diversity of root endosphere (RE), rhizosphere soil (RS), and bulk soil (BS) among three representative desert plants (Alhagi sparsifolia, Tamarix ramosissima, and Calligonum caput-medusae) in three Xinjiang desert regions {Taklimakan (CL), Gurbantünggüt (MSW), and Kumtag (TLF)} in China. This study found that the soil properties {electrical conductivity (EC), soil organic carbon (SOC), total nitrogen (TN) and phosphorus (TP), available nitrogen (AN) and phosphorus (AP)} of C. caput-medusae were significantly lower than those of A. sparsifolia and T. ramosissima, while the root nutrients (TN and TP) of A. sparsifolia were significantly higher compared to C. caput-medusae and T. ramosissima. The beta diversity of bacteria and fungi (RE) among the three desert plants was significantly different. The common OTU numbers of bacteria and fungi in three compartments (RE, RS, and BS) of the three desert plants were ranked as RS > BS > RE. The bacterial and fungal (RE) Shannon and Simpson indexes of C. caput-medusae were significantly lower as compared to those of A. sparsifolia and T. ramosissima. Additionally, bacterial and fungal (RE and RS) node numbers and average degree of C. caput-medusae were lower than those found in A. sparsifolia and T. ramosissima. Root and soil nutrients collectively contributed to the composition of root-associated bacterial (RE, 12.4%; RS, 10.6%; BS, 16.6%) and fungal communities (RE, 34.3%; RS, 1.5%; BS, 17.7%). These findings demonstrate variations in the bacterial and fungal populations across different plant species with distinct compartments (RE, RS, and BS) in arid environments. More importantly, the study highlights how much soil and plant nutrients contribute to root-associated microbial communities. Full article
(This article belongs to the Special Issue Plant-Microbiome Interactions)
Show Figures

Figure 1

Figure 1
<p>The OTU number (bacteria and fungi) and relative abundance {dominant bacteria and fungi taxa (top 10 phyla)} of root endosphere (RE), rhizosphere soil (RS), and bulk soil (BS) in three desert plants. Different lowercase letters (a and b) indicate significant differences among species at the <span class="html-italic">p</span> &lt; 0.05 level (ANOVA and Duncan’s test). (<b>A</b>–<b>C</b>) OTUs number of the bacteria, (<b>D</b>–<b>F</b>) OTUs number of the fungi, and (<b>G</b>) relative abundance of dominant bacteria and (<b>H</b>) relative abundance of dominant fungi.</p>
Full article ">Figure 2
<p>Alpha diversity {(<b>A</b>–<b>C</b>) Chao1, (<b>D</b>–<b>F</b>) Shannon, (<b>G</b>–<b>I</b>) Pielou_e, and (<b>J</b>–<b>L</b>) Simpson indexes} of root endosphere (RE), rhizosphere soil (RS), and bulk soil (BS) bacteria in three desert plants. Different lowercase letters (a–c) indicate significant differences among species at the <span class="html-italic">p</span> &lt; 0.05 level and the ns indicate no significant differences among species at the <span class="html-italic">p</span> &gt; 0.05 level (ANOVA and Duncan’s test).</p>
Full article ">Figure 3
<p>Beta diversity {(<b>A</b>–<b>C</b>) Bray–Curtis and (<b>D</b>–<b>F</b>) nonmetric multidimensional scaling} of root endosphere (RE), rhizosphere soil (RS), and bulk soil (BS) bacteria of three desert plants. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Core and differential microbiota {(<b>A</b>–<b>C</b>) OTUs number of the bacteria and (<b>D</b>–<b>F</b>) OTUs number of the fungi} of root endosphere (RE), rhizosphere soil (RS), and bulk soil (BS) bacteria and fungi among three desert plants.</p>
Full article ">Figure 5
<p>Linear discriminant analysis effect size (LEfSe) {(<b>A</b>–<b>C</b>) LEfSe analysis of the bacteria and (<b>D</b>–<b>F</b>) LEfSe analysis of the fungi} of root endosphere (RE), rhizosphere soil (RS), and bulk soil (BS) bacteria and fungi among three desert plants.</p>
Full article ">Figure 6
<p>Co-occurrence network {(<b>A</b>,<b>D</b>,<b>G</b>) Network characteristics of the <span class="html-italic">A. sparsifolia</span>, (<b>B</b>,<b>E</b>,<b>H</b>) Network characteristics of the <span class="html-italic">T. ramosissima</span>, and (<b>C</b>,<b>F</b>,<b>I</b>) Network characteristics of the <span class="html-italic">C. caput-medusae</span>} of root endosphere (RE), rhizosphere soil (RS), and bulk soil (BS) bacteria of three desert plants.</p>
Full article ">Figure 7
<p>The main drivers of different bacterial and fungal communities {root endosphere (RE), rhizosphere soil (RS), and bulk soil (BS)} at the OTU level. The RDA plots show soil and root nutrients that significantly affect bacterial and fungal communities, according to the reduced model with 999 permutations. The results of HP analysis indicated the relative importance of environmental factors (soil and root) on bacterial and fungal communities. The column diagram shows the individual effect of each environmental factor (from hierarchical partitioning). SOC, soil organic carbon (g·kg<sup>−1</sup>); ROC, root organic carbon (g·kg<sup>−1</sup>); EC, electrical conductivity (mS·cm<sup>−1</sup>); TN, total nitrogen (g·kg<sup>−1</sup>); TP, total phosphorus (g·kg<sup>−1</sup>); TK, total potassium (g·kg<sup>−1</sup>); AN, available nitrogen (mg·kg<sup>−1</sup>); AP, available phosphorus (mg·kg<sup>−1</sup>); AK, available potassium (mg·kg<sup>−1</sup>). {(<b>A</b>) Redundancy analysis and (<b>B</b>) HP analysis of the root endosphere, (<b>E</b>) Redundancy analysis and (<b>F</b>) HP analysis of the rhizosphere soil, and (<b>I</b>) Redundancy analysis and (<b>J</b>) HP analysis of the bulk soil} of the bacteria and {(<b>C</b>) Redundancy analysis and (<b>D</b>) HP analysis of the root endosphere, (<b>G</b>) Redundancy analysis and (<b>H</b>) HP analysis of the rhizosphere soil, and (<b>K</b>) Redundancy analysis and (<b>L</b>) HP analysis of the bulk soil} of the fungi. Note: Significance codes, “*” <span class="html-italic">p</span> &lt; 0.05; “**” <span class="html-italic">p</span> &lt; 0.01; “***” <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>The three sampling sites at Cele, Turpan, and Mosuowan are located in Tarim Basin, Turpan Basin, and Junggar Basin, respectively.</p>
Full article ">
14 pages, 3542 KiB  
Technical Note
Study on Daytime Atmospheric Mixing Layer Height Based on 2-Year Coherent Doppler Wind Lidar Observations at the Southern Edge of the Taklimakan Desert
by Lian Su, Haiyun Xia, Jinlong Yuan, Yue Wang, Amina Maituerdi and Qing He
Remote Sens. 2024, 16(16), 3005; https://doi.org/10.3390/rs16163005 - 16 Aug 2024
Cited by 1 | Viewed by 810
Abstract
The long-term atmospheric mixing layer height (MLH) information plays an important role in air quality and weather forecasting. However, it is not sufficient to study the characteristics of MLH using long-term high spatial and temporal resolution data in the desert. In this paper, [...] Read more.
The long-term atmospheric mixing layer height (MLH) information plays an important role in air quality and weather forecasting. However, it is not sufficient to study the characteristics of MLH using long-term high spatial and temporal resolution data in the desert. In this paper, over the southern edge of the Taklimakan Desert, the diurnal, monthly, and seasonal variations in the daytime MLH (retrieved by coherent Doppler wind lidar) and surface meteorological elements (provided by the local meteorological station) in a two-year period (from July 2021 to July 2023) were statistically analyzed, and the relationship between the two kinds of data was summarized. It was found that the diurnal average MLH exhibits a unimodal distribution, and the decrease rate in the MLH in the afternoon is much higher than the increase rate before noon. From the seasonal and monthly perspective, the most frequent deep mixing layer (>4 km) was formed in June, and the MLH is the highest in spring and summer. Finally, in terms of their mutual relationship, it was observed that the east-pathway wind has a greater impact on the formation of the deep mixing layer than the west-pathway wind; the dust weather with visibility of 1–10 km contributes significantly to the formation of the mixing layer; the temperature and relative humidity also exhibit a clear trend of a concentrated distribution at about the height of 3 km. The statistical analysis of the MLH deepens the understanding of the characteristics of dust pollution in this area, which is of great significance for the treatment of local dust pollution. Full article
Show Figures

Figure 1

Figure 1
<p>The 3D topographic map of Taklimakan Desert. The red circle represents the study site of MinFeng.</p>
Full article ">Figure 2
<p>The typical atmospheric mixing layer height results calculated by using the TKEDR threshold method.</p>
Full article ">Figure 3
<p>The seasonal and annual wind frequency rose diagrams during the daytime. (<b>a</b>) Spring. (<b>b</b>) Summer. (<b>c</b>) Fall. (<b>d</b>) Winter. (<b>e</b>) Annual.</p>
Full article ">Figure 4
<p>During the daytime, the average hourly change in the mixing layer height and various surface meteorological elements in different months. (<b>a</b>) Mixing layer height. (<b>b</b>) Atmospheric temperature. (<b>c</b>) Relative humidity. (<b>d</b>) Horizontal visibility. (<b>e</b>) Near-surface wind speed. The data at 8:00 BJT represent the monthly average of the whole hour, and the data of other hours are the same.</p>
Full article ">Figure 5
<p>During the daytime, the average hourly change in the mixing layer height and various surface meteorological elements across different seasons. (<b>a</b>) Mixing layer height. (<b>b</b>) Atmospheric temperature. (<b>c</b>) Relative humidity. (<b>d</b>) Horizontal visibility. (<b>e</b>) Near-surface wind speed.</p>
Full article ">Figure 6
<p>During the daytime, the probability distribution of the mixing layer height and various surface meteorological elements in each month and season. (<b>a</b>) Mixing layer height. (<b>b</b>) Atmospheric temperature. (<b>c</b>) Relative humidity. (<b>d</b>) Horizontal visibility. (<b>e</b>) Near-surface wind speed. (<b>f</b>) Near-surface wind direction.</p>
Full article ">Figure 7
<p>During the daytime, the contribution degree of surface meteorological elements to the mixing layer height at various height intervals. (<b>a</b>) The MLH and its corresponding atmospheric temperature (T) and relative humidity (RH). (<b>b</b>) The MLH and its corresponding horizontal visibility (VIS), near-surface wind speed (WS) and near-surface wind direction (WD). The beginning and end of the arrow indicate two related variables. The width of the arrow trunk signifies the degree of contribution to the mixing layer height, where a wider arrow trunk indicates a greater contribution.</p>
Full article ">Figure 8
<p>During the daytime, the probability density distribution between the hourly mean mixed layer height and the ground meteorological elements in spring and summer. (<b>a</b>) Atmospheric temperature versus MLH. (<b>b</b>) Relative humidity versus MLH. (<b>c</b>) Near-surface wind speed versus MLH. (<b>d</b>) Horizontal visibility versus MLH.</p>
Full article ">
22 pages, 9842 KiB  
Article
Urbanization Effect on Regional Thermal Environment and Its Mechanisms in Arid Zone Cities: A Case Study of Urumqi
by Aerzuna Abulimiti, Yongqiang Liu, Jianping Tang, Ali Mamtimin, Junqiang Yao, Yong Zeng and Abuduwaili Abulikemu
Remote Sens. 2024, 16(16), 2939; https://doi.org/10.3390/rs16162939 - 10 Aug 2024
Cited by 1 | Viewed by 1481
Abstract
Urumqi is located in the arid region of northwestern China, known for being one of the most delicate ecological environments and an area susceptible to climate change. The urbanization of Urumqi has progressed rapidly, yet there is a lack of research on the [...] Read more.
Urumqi is located in the arid region of northwestern China, known for being one of the most delicate ecological environments and an area susceptible to climate change. The urbanization of Urumqi has progressed rapidly, yet there is a lack of research on the urbanization effect (UE) in Urumqi in terms of the regional climate. This study investigates the UE of Urumqi (urban built-up area) on the regional thermal environment and its mechanisms for the first time, based on the WRF (Weather Research and Forecasting) model (combined with the Urban Canopy Model, UCM) simulation data of 10 consecutive years (2012–2021). The results show that the UE on surface temperature (Ts) and air temperature at 2 m (T2m) is strong (weak) during the night (daytime) in all seasons, and the UE on these is largest (smallest) in spring (winter). In addition, the maximum UE on both Ts and T2m is present over southern Urumqi in winter, whereas the maximum UE is identified over the northern Urumqi in other seasons. The maximum UE on Ts occurred in northwestern Urumqi at 18 LST (Local Standard Time, i.e., UTC+6) in autumn (reaching 5.2 °C), and the maximum UE on T2m occurred in northern Urumqi at 4 LST in summer (reaching 2.6 °C). Urbanization showed a weak cooling effect during daytime in summer and winter, reflecting the unique characteristics of the UE in arid regions, which are different from those in humid regions. The maximum cooling of Ts occurred in northern Urumqi at 11 LST in summer (reaching −0.4 °C), while that of T2m occurred at 10 LST in northern and northwestern Urumqi in winter (reaching −0.25 °C), and the cooling effect lasted for a longer period of time in summer than in winter. The UE of Urumqi causes the increase of Ts mainly through the influence of net short-wave radiation and geothermal flux and causes the increase of T2m through the influence of sensible heat flux and net long-wave radiation. The UE on the land surface energy balance in Urumqi can be used to explain the seasonal variation and spatial differences of the UEs on the regional thermal environment and the underlying mechanism. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical location and terrain altitude (shading, units: m) of the study area; the sky-blue thick solid line represents the boundary of the built-up area of Urumqi, and the black fine solid lines represent the administrative boundaries of Urumqi and its districts. The small red box in the small globe in the upper left corner shows the location of the study area from a broader perspective.</p>
Full article ">Figure 2
<p>Terrain altitude (shading, DEM, units: m) and locations of the 20 meteorological stations (indicated by red dots) selected in this study to verify the simulation results. The blue line represents the outline of the built-up area of Urumqi, and the thin black lines indicate administrative boundaries of the districts in Urumqi.</p>
Full article ">Figure 3
<p>(<b>a</b>) The geographic locations and terrain elevation (shading, DEM, units: m) of the WRF model domains, where d01 represents the outer domain, d02 represents the inner domain, and the thin black lines represent administrative boundaries of Xinjiang. (<b>b</b>) The location of the d01 domain on a map of China; the shading represents the terrain elevation (units: m).</p>
Full article ">Figure 4
<p>(<b>a</b>) Land use categories (shading) derived from CLCD dataset in the d02 domain of the Urban (control) experiment of numerical simulation. The land use category of “urban and build-up” is highlighted with red underline in the color bar information. (<b>b</b>) The same as (<b>a</b>) but for the enlarged area (main study area) centered around the built-up area of Urumqi. The small black dot in the central area (indicated by letter C) shows the location of the urban centroid of the built-up area of Urumqi. The black boxes indicate the proximity areas in different directions of the urban centroid of the built-up area of Urumqi, and the abbreviated letters NW, N, NE, W, C, E, S, and SE represent the corresponding northwestern, northern, northeastern, western, central, eastern, southern, and southeastern areas of Urumqi.</p>
Full article ">Figure 5
<p>(<b>a</b>) Land use categories (shading) derived from CLCD dataset in the d02 domain of the NoUrban (sensitivity) experiment of numerical simulation. (<b>b</b>) The same as (<b>a</b>) but for the enlarged area, which shows the same area of <a href="#remotesensing-16-02939-f004" class="html-fig">Figure 4</a>b. The black boxes and corresponding abbreviated letters (NW, N, NE, W, C, E, S, SE) represent the same locations of Urumqi, which are shown in <a href="#remotesensing-16-02939-f004" class="html-fig">Figure 4</a>b. All of the original built-up areas were replaced by grasslands in the NoUrban (sensitivity) experiment of the numerical simulation. The small black dot in the central area (indicated by letter C) shows the location of the urban centroid of the built-up area of Urumqi.</p>
Full article ">Figure 6
<p>Scatter plots and linear regression fitting lines (small blue circles and red lines), with regression equations, Pearson’s correlation coefficient (r), and root mean square error (RMSE) shown in the top of each panel, showing the correspondence of the WRF numerical simulation results (monthly average value) with corresponding observational data from 2012 to 2021. (<b>a</b>–<b>c</b>) indicate the Tmax, Tmean, and Tmin at 2 m in spring, respectively; (<b>d</b>–<b>f</b>) represent the Tmax, Tmean, and Tmin at 2 m in summer, respectively; (<b>g</b>–<b>i</b>) present the Tmax, Tmean, and Tmin at 2 m in autumn, respectively; (<b>j</b>–<b>l</b>) show the Tmax, Tmean, and Tmin at 2 m in winter, respectively.</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>d</b>) Spatial distribution of the UE on the average surface temperature (Ts, unit: °C) in Urumqi in spring, summer, autumn, and winter, respectively. The black boxes indicate the proximity areas in different directions of the urban centroid of the built-up area of Urumqi, which are also shown in <a href="#remotesensing-16-02939-f004" class="html-fig">Figure 4</a>b and <a href="#remotesensing-16-02939-f005" class="html-fig">Figure 5</a>b; The blue line represents the outline of the built-up area of Urumqi, and the thin black lines indicate administrative boundaries of districts in Urumqi. The small black dot in the central area (indicated by letter C) shows the location of the urban centroid of the built-up area of Urumqi. (<b>e</b>–<b>h</b>) The average values of Ts over the eight proximity areas (only calculated values for built-up area) in different directions of the urban centroid of the built-up area of Urumqi in spring, summer, autumn, and winter, respectively. The abbreviated letters NW, N, NE, W, C, E, S, and SE in horizontal axis represent the corresponding eight areas shown in <a href="#remotesensing-16-02939-f004" class="html-fig">Figure 4</a>b and <a href="#remotesensing-16-02939-f005" class="html-fig">Figure 5</a>b, and ALL represents the average value of all built-up areas of Urumqi (i.e., the averaged value of all eight areas).</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>d</b>) Diurnal variation characteristics of UE on the average surface temperature (Ts, unit: °C) in different areas in different directions of the urban centroid of the built-up area of Urumqi in spring, summer, autumn, and winter, respectively. The NW, N, NE, W, C, E, S, and SE represent the corresponding values of Ts calculated in eight areas (only calculated values for built-up area) shown in <a href="#remotesensing-16-02939-f004" class="html-fig">Figure 4</a>b and <a href="#remotesensing-16-02939-f005" class="html-fig">Figure 5</a>b, and ALL represents the average value of all built-up areas of Urumqi (i.e., the averaged value of all eight areas).</p>
Full article ">Figure 9
<p>(<b>a</b>–<b>d</b>) Diurnal variation characteristics of average surface temperature (Ts, unit: °C) over the built-up area of Urumqi in Urban experiment, NoUrban experiment, and the UE on Ts in spring, summer, autumn, and winter, respectively.</p>
Full article ">Figure 10
<p>(<b>a</b>–<b>d</b>) Spatial distribution of the UE on the average air temperature at 2 m (T2m, unit: °C) in Urumqi in spring, summer, autumn, and winter, respectively. The black boxes indicate the proximity areas in different directions of the urban centroid of the built-up area of Urumqi, which are also shown in <a href="#remotesensing-16-02939-f004" class="html-fig">Figure 4</a>b and <a href="#remotesensing-16-02939-f005" class="html-fig">Figure 5</a>b, and the small black dot in the central area (indicated by letter C) shows the location of the urban centroid of the built-up area of Urumqi. (<b>e</b>–<b>h</b>) The average values over of T2m in the eight proximity areas (only calculated values on built-up area) in different directions of the urban centroid of built-up area of Urumqi in spring, summer, autumn, and winter, respectively. The abbreviated letters NW, N, NE, W, C, E, S, SE in horizontal axis represent the corresponding eight areas shown in <a href="#remotesensing-16-02939-f004" class="html-fig">Figure 4</a>b and <a href="#remotesensing-16-02939-f005" class="html-fig">Figure 5</a>b, and ALL represents the average value of all built-up area of Urumqi (i.e., the averaged value of all eight areas).</p>
Full article ">Figure 11
<p>(<b>a</b>–<b>d</b>) Diurnal variation characteristics of the UE on average air temperature at 2 m (T2m, unit: °C) in different areas in different directions of the urban centroid of built-up area of Urumqi in spring, summer, autumn, and winter, respectively. The NW, N, NE, W, C, E, S, SE represent the corresponding values of T2m calculated in eight areas (only calculated values on built-up area) shown in <a href="#remotesensing-16-02939-f004" class="html-fig">Figure 4</a>b and <a href="#remotesensing-16-02939-f005" class="html-fig">Figure 5</a>b, and ALL represents the average value of all built-up area of Urumqi (i.e., the averaged value of all eight areas).</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>d</b>) Diurnal variation characteristics of average air temperature at 2 m (T2m, unit: °C) over the built-up area of Urumqi in the Urban experiment, NoUrban experiment, and the UE of Urumqi on T2m in spring, summer, autumn, and winter, respectively.</p>
Full article ">Figure 13
<p>The UE on surface energy budget (SEB, unit: W m<sup>−2</sup>) in all built-up areas of Urumqi in four seasons. The SWn, LWn, SH, LH, GH, and SEB represent net shortwave radiation, net longwave radiation, sensible heat flux, latent heat flux, ground heat flux, and surface energy budget, respectively.</p>
Full article ">Figure 14
<p>The UE on surface energy budget (SEB, unit: W m<sup>−2</sup>) in the northern part of the built-up area of Urumqi in four seasons. The SWn, LWn, SH, LH, GH, and SEB represent net shortwave radiation, net longwave radiation, sensible heat flux, latent heat flux, ground heat flux, and surface energy budget, respectively.</p>
Full article ">Figure 15
<p>The UE on surface energy budget (SEB, unit: W m<sup>−2</sup>) in the southern part of the built-up area of Urumqi in four seasons. The SWn, LWn, SH, LH, GH, and SEB represent net shortwave radiation, net longwave radiation, sensible heat flux, latent heat flux, ground heat flux, and surface energy budget, respectively.</p>
Full article ">Figure 16
<p>Difference in the UE on surface energy budget (SEB, unit: W m<sup>−2</sup>) between the northern and southern parts of the built-up area of Urumqi in four seasons. The SWn, LWn, SH, LH, GH, and SEB represent net shortwave radiation, net longwave radiation, sensible heat flux, latent heat flux, ground heat flux, and surface energy budget, respectively.</p>
Full article ">
25 pages, 6036 KiB  
Article
Research on Improving the Accuracy of SIF Data in Estimating Gross Primary Productivity in Arid Regions
by Wei Liu, Yu Wang, Ali Mamtimin, Yongqiang Liu, Jiacheng Gao, Meiqi Song, Ailiyaer Aihaiti, Cong Wen, Fan Yang, Wen Huo, Chenglong Zhou, Jian Peng and Hajigul Sayit
Land 2024, 13(8), 1222; https://doi.org/10.3390/land13081222 - 7 Aug 2024
Cited by 2 | Viewed by 1051
Abstract
Coupling solar-induced chlorophyll fluorescence (SIF) with gross primary productivity (GPP) for ecological function integration research presents numerous uncertainties, especially in ecologically fragile and climate-sensitive arid regions. Therefore, evaluating the suitability of SIF data for estimating GPP and the feasibility of improving its accuracy [...] Read more.
Coupling solar-induced chlorophyll fluorescence (SIF) with gross primary productivity (GPP) for ecological function integration research presents numerous uncertainties, especially in ecologically fragile and climate-sensitive arid regions. Therefore, evaluating the suitability of SIF data for estimating GPP and the feasibility of improving its accuracy in the northern region of Xinjiang is of profound significance for revealing the spatial distribution patterns of GPP and the strong coupling relationship between GPP and SIF in arid regions, achieving the goal of “carbon neutrality” in arid regions. This study is based on multisource SIF satellite data and GPP observation data from sites in three typical ecosystems (cultivated and farmlands, pasture grasslands, and desert vegetation). Two precision improvement methods (canopy and linear) are used to couple multiple indicators to determine the suitability of multisource SIF data for GPP estimation and the operability of accuracy improvement methods in arid regions reveal the spatial characteristics of SIF (GPP). The results indicate the following. (1) The interannual variation of GPP shows an inverted “U” shape, with peaks values in June and July. The cultivated and farmland areas have the highest peak value among the sites (0.35 gC/m2/month). (2) The overall suitability ranking of multisource SIF satellite products for GPP estimation in arid regions is RTSIF > CSIF > SIF_OCO2_005 > GOSIF. RTSIF shows better suitability in the pasture grassland and cultivated and farmland areas (R2 values of 0.85 and 0.84, respectively). (3) The canopy method is suitable for areas with a high leaf area proportion (R2 improvement range: 0.05–0.06), while the linear method is applicable across different surface types (R2 improvement range: 0.01–0.13). However, the improvement effect of the linear method is relatively weaker in areas with high vegetation cover. (4) Combining land use data, the overall improvement of SIF (GPP) is approximately 0.11%, and the peak values of its are mainly distributed in the northern and southern slopes of the Tianshan Mountains, while the low values are primarily found in the Gurbantunggut Desert. The annual mean value of SIF (GPP) is about 0.13 mW/m2/nm/sr. This paper elucidates the applicability of SIF for GPP estimation and the feasibility of improving its accuracy, laying the theoretical foundation for the spatiotemporal coupling study of GPP and SIF in an arid region, and providing practical evidence for achieving carbon neutrality goals. Full article
(This article belongs to the Special Issue Land-Based Greenhouse Gas Mitigation for Carbon Neutrality)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Specific locations of the Tianshan Mountains, Ulan Usu Station, Ulastai Station, and Kelameili Station in Xinjiang. (<b>b</b>) Schematic representation of the elevations of the study area. (<b>c</b>) Schematic representation of the land use types at the study area.</p>
Full article ">Figure 2
<p>Interannual variation of monthly average GPP at each site in 2020 (excluding nighttime values).</p>
Full article ">Figure 3
<p>(<b>a</b>) The linear regression fitting of 2020 GPP data from three site with corresponding site data of CSIF satellite products. (<b>b</b>) The linear regression fitting of 2020 GPP data from three site with corresponding site data of RTSIF satellite products. (<b>c</b>) The linear regression fitting of 2020 GPP data from three site with corresponding site data of SIF-OCO-005 satellite products. (<b>d</b>) The linear regression fitting of 2020 GPP data from three site with corresponding site data of GOSIF satellite products.</p>
Full article ">Figure 4
<p>The spatial distribution characteristics of annual mean values of multisource SIF satellite products.</p>
Full article ">Figure 5
<p>Responsiveness of multisource SIF satellite products to major influencing factors of GPP (** indicates significance at the 0.5 level). (<b>a</b>) Kelameili Station, desert vegetation area. (<b>b</b>) Ulastai Station, pasture and grassland area. (<b>c</b>) Ulan Usu Staion, cultivate land and farmland area.</p>
Full article ">Figure 6
<p>Differences in GPP/SIF values under different weather conditions. (<b>a</b>) Kelameili Station, desert vegetation area. (<b>b</b>) Ulastai Station, pasture and grassland area. (<b>c</b>) Ulan Usu Staion, cultivate land and farmland area.</p>
Full article ">Figure 7
<p>Linear fitting graph of 2020 GPP data and RTSIF corresponding station data for each station after improving based on the canopy method. (<b>a</b>) Ulastai Station, pasture and grassland area. (<b>b</b>) Ulan Usu Staion, cultivate land and farmland area.</p>
Full article ">Figure 8
<p>Linear fitting diagram between the 2020 GPP data of each station and the corresponding RTSIF station data after improving based on the linear method. (<b>a</b>) Kelameili Station, desert vegetation area. (<b>b</b>) Ulastai Station, pasture and grassland area. (<b>c</b>) Ulan Usu Staion, cultivate land and farmland area.</p>
Full article ">Figure 9
<p>The R<sup>2</sup> fitting values for various sites based on two accuracy improvement methods: canopy and linear.</p>
Full article ">Figure 10
<p>Changes in spatial characteristics of quarterly average values before and after the improvement of SIF satellite product data. (<b>a1</b>–<b>d1</b>) The spatial variation characteristics of the mean values of each season before improvement, (<b>a1</b>) for spring, and so on. (<b>a2</b>–<b>d2</b>) The spatial variation characteristics of the mean values of each season after improvement, (<b>a2</b>) for spring, and so on.</p>
Full article ">
16 pages, 6207 KiB  
Article
An Evaluation of the Dust Emission Characteristics of Typical Underlying Surfaces in an Aeolian Region in the Middle Reaches of the Yarlung Zangbo River on the Qinghai–Tibet Plateau
by Mingjie Ma, Duo Zha, Qing He, Xinghua Yang, Fan Yang, Ali Mamtimin, Xiannian Zheng and Han Sun
Land 2024, 13(8), 1168; https://doi.org/10.3390/land13081168 - 30 Jul 2024
Viewed by 621
Abstract
Some of the most severe aeolian damage occurs along the middle reaches of the Yarlung Zangbo River in Tibet. Dust emission amounts (DEAs) are often used to assess aeolian damage; however, the research on DEAs in this area is currently almost blank. This [...] Read more.
Some of the most severe aeolian damage occurs along the middle reaches of the Yarlung Zangbo River in Tibet. Dust emission amounts (DEAs) are often used to assess aeolian damage; however, the research on DEAs in this area is currently almost blank. This article uses field-measured wind speed data from 2021 to 2022 in the Shannan wide valley area, combined with the Gillette dust emission estimation model to quantitatively determine the contributions of three surface types (riverbank quicksand area, foothill sand dunes, and the river floodplain vegetation area) to DEAs in the research area. The influence of surface characteristics on DEAs is analyzed and discussed. The results show the following: (1) The threshold friction velocity (u*t) in the riverbank quicksand area, foothill sand dunes, and the river floodplain vegetation area is 30.6 cm/s, 71.2 cm/s, and 85.6 cm/s, respectively, the threshold velocity (ut) is 6.1 m/s, 7.0 m/s, and 7.5 m/s, respectively, and the vegetation area is 2.8 times and 1.3 times that of the quicksand area, respectively. (2) The DEAs were in the following order: the riverbank quicksand area (652.9 t/km2) > foothill sand dunes (326.5 t/km2) > the river floodplain vegetation area (107.8 t/km2), the riverbank quicksand area is about 6.1 times that of the river floodplain vegetation area, and DEAs are a significant seasonal distribution: winter (44.7%) > spring (28.3%) > autumn (15.7%) > summer (11.3%). (3) The DEAs from the dusty weather were in the following order: blowing sand (60.2%) > sandstorms (28.6%) > gusty winds (11.2%). (4) The DEAs increase with the increase in the average wind speed greater than 6.1 m/s, but the increase rate is obviously different, which showed that Changguo and Azha are greater than Sangyesi, Duopazhang, Sangri, and Senburi. At approximately the same average wind speed greater than 6.1 m/s, the DEAs in the quicksand area are much greater than in the vegetation area. Full article
(This article belongs to the Section Land Environmental and Policy Impact Assessment)
Show Figures

Figure 1

Figure 1
<p>Location of the study area in in the middle reaches of the Yarlung Zangbo River.</p>
Full article ">Figure 2
<p>Frequency distribution of logarithms of surface roughness <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>z</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> in the study area.</p>
Full article ">Figure 3
<p>Daily total DEAs variations.</p>
Full article ">Figure 4
<p>Monthly total DEAs variations.</p>
Full article ">Figure 5
<p>The four-season daily total DEAs variation in the riverbank quicksand areas, foothill sand dunes, and river floodplain vegetation areas.</p>
Full article ">Figure 6
<p>Daily DEAs variations by dusty weather in the riverbank quicksand areas, foothill sand dunes, and river floodplain vegetation areas.</p>
Full article ">Figure 7
<p>Monthly DEAs variations by dusty weather.</p>
Full article ">Figure 8
<p>The relationship between the DEAs of the riverbank quicksand areas, foothill sand dunes, and river floodplain vegetation areas and average wind speed &gt;6.1 m/s, &gt;7.0 m/s, and &gt;7.5 m/s, respectively.</p>
Full article ">Figure 9
<p>The relationship between the DEAs of the riverbank quicksand areas, foothill sand dunes, and river floodplain vegetation areas and cumulative wind speed &gt;6.1 m/s, &gt;7.0 m/s, and &gt;7.5 m/s, respectively.</p>
Full article ">Figure 10
<p>Impact of meteorological factors on DEAs.</p>
Full article ">
20 pages, 20652 KiB  
Article
Three-Dimensional Structure and Transport Properties of Dust Aerosols in Central Asia—New Insights from CALIOP Observations, 2007–2022
by Jinglong Li, Qing He, Yonghui Wang, Xiaofei Ma, Xueqi Zhang and Yongkang Li
Remote Sens. 2024, 16(12), 2049; https://doi.org/10.3390/rs16122049 - 7 Jun 2024
Cited by 1 | Viewed by 1128
Abstract
Central Asia (CA) is one of the major sources of global dust aerosols. They pose a serious threat to regional climate change and environmental health and also make a significant contribution to the global dust load. However, there is still a gap in [...] Read more.
Central Asia (CA) is one of the major sources of global dust aerosols. They pose a serious threat to regional climate change and environmental health and also make a significant contribution to the global dust load. However, there is still a gap in our understanding of dust transport in this region. Therefore, this study utilizes Cloud–Aerosol LiDAR with Orthogonal Polarization (CALIOP) data from 2007 to 2022 to depict the three-dimensional spatiotemporal distribution of dust aerosols over CA and to analyze their transport processes. In addition, the Tropospheric Monitoring Instrument (TROPOMI) was employed to assist in monitoring the movement of typical dust events, and the trajectory model was utilized to simulate the forward and backward trajectories of a dust incident. Additionally, a random forest (RF) model was employed to rank the contributions of various environmental factors. The findings demonstrate that high extinction values (0.6 km−1) are mostly concentrated within the Tarim Basin of Xinjiang, China, maintaining high values up to 2 km in altitude, with a noticeable decrease as the altitude increases. The frequency of dust occurrences is especially pronounced in the spring and summer seasons, with dust frequencies in the Tarim Basin and the Karakum and Kyzylkum deserts exceeding 80%, indicating significant seasonal and regional differences. The high values of dust optical depth (DOD) in CA are primarily concentrated in the summer, concurrent with the presence of a stable aerosol layer of dust in the atmosphere with a thickness of 0.62 km. Furthermore, dust from CA can traverse the Tianshan mountains via the westerlies, transporting it eastward. Additionally, skin temperature can mitigate regional air pollution. Our results contribute to a deeper understanding of the dynamic processes of dust in CA and provide scientific support for the development of regional climate regulation strategies. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Overview of the study area.</p>
Full article ">Figure 2
<p>Seasonal distribution of extinction coefficients in CA (km<b><sup>−</sup></b><sup>1</sup>).</p>
Full article ">Figure 3
<p>Seasonal distribution of backscatter coefficients in CA (km<sup>−</sup><sup>1</sup>·sr<sup>−</sup><sup>1</sup>).</p>
Full article ">Figure 4
<p>Seasonal distribution of depolarization ratios in CA.</p>
Full article ">Figure 5
<p>Seasonal distribution of dust occurrence frequency in CA.</p>
Full article ">Figure 6
<p>Seasonal and annual average spatial distribution of DOD in CA.</p>
Full article ">Figure 7
<p>Variation in seasonal mean DOD in five central Asian countries and Xinjiang, China.</p>
Full article ">Figure 8
<p>Monthly average spatial distribution variation in DOD in CA.</p>
Full article ">Figure 9
<p>Variation in monthly mean DOD in five central Asian countries and Xinjiang, China.</p>
Full article ">Figure 10
<p>Seasonal and annual average spatial distribution of dust layers in CA.</p>
Full article ">Figure 11
<p>(<b>a</b>) shows the multi-year trends in dust layer thickness for the five Central Asian countries and Xinjiang, China. (<b>b</b>–<b>d</b>) show the trends in dust layer thickness for the two regions, respectively.</p>
Full article ">Figure 12
<p>Panels (<b>a</b>), (<b>d</b>), and (<b>g</b>) display the CALIPSO satellite trajectories over the CA region on 18, 19 and 20 April 2022, respectively. Panels (<b>b</b>,<b>e</b>,<b>h</b>) show the extinction coefficients observed by CALIPSO (corresponding to the satellite’s overpass times), while panels (<b>c</b>,<b>f</b>,<b>i</b>) present the altitude–satellite orbit profiles of aerosol type classification results (corresponding to the satellite overpass times). Note: The red solid line represents the ground level, white represents the terrain, and the red dots indicate locations in the mountainous regions of Kyrgyzstan.</p>
Full article ">Figure 13
<p>UVAI distribution in the study area observed by TROPOMI from 18 to 20 April 2022.</p>
Full article ">Figure 14
<p>Trajectories of the study site simulated by the HYSPLIT trajectory model for 72 h backward (<b>a</b>) and 72 h forward (<b>b</b>). Note: The red, blue, and green lines, respectively, represent the forward and backward trajectory lines at elevations of 500 m, 1500 m, and 3000 m above the location point.</p>
Full article ">Figure 15
<p>(<b>a</b>) RF simulation results and (<b>b</b>) correlation results. ** represents <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
Back to TopTop