ArticlePDF Available

MODELING PROCESS OF THE SPATIAL-TEMPORAL CHANGES OF VEGETATION COVER AND ITS RELATIONSHIP WITH DRIVERS IN DRYLANDS AND WETLANDS IN XIANJIANG (CHINA)

Authors:
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
27
MODELING PROCESS OF THE SPATIAL-TEMPORAL CHANGES OF
VEGETATION COVER AND ITS RELATIONSHIP WITH DRIVERS IN
DRYLANDS AND WETLANDS IN XIANJIANG (CHINA)
Seyed Omid Reza SHOBAIRI *, **, Lingxiao SUN *, ** (C.A.), Haiyan ZHANG *, **,
Chunlan LI *, **, Jing HE *,**, Behnam Asghari BEIRAMI ***,
Samira Hemmati ROUDBARI **** and Qorghizbek AYOMBEKOV *
* Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences, 818 South Beijing Road,
Urumuqi, China, CN-830011, omidshobeyri214@gmail.com, ORCID: 0000-0002-6528-8653 (S. O. R.
S.); sunlx@ms.xjb.ac.cn, ORCID: 0000-0002-1702-5445 (L. S.); hyzhang@ms.xjb.ac.cn, ORCID:
0000-0002-5250-1865 (H. Z.); lichunlan@ms.xjb.ac.cn, ORCID: 0009-0001-9837-7556 (C. L.);
hejing@ms.xjb.ac.cn, ORCID: 0001-0001-7013-1266 (J. H.); ayombekqarghizbek@gmail.com,
ORCID: 0009-0002-2400-1481 (Q. A.).
** University of Chinese Academy of Sciences, 19 Yuquan Road, Beijing, CN-100049, China, CN-
100049, omidshobeyri214@gmail.com, ORCID: 0000-0002-6528-8653 (S. O. R. S.);
sunlx@ms.xjb.ac.cn, ORCID: 0000-0002-1702-5445 (L. S.); hyzhang@ms.xjb.ac.cn, ORCID: 0000-
0002-5250-1865 (H. Z.); lichunlan@ms.xjb.ac.cn, ORCID: 0009-0001-9837-7556 (C. L.);
hejing@ms.xjb.ac.cn, ORCID: 0001-0001-7013-1266 (J. H.).
*** K. N. Toosi University of Technology, 470 Mirdamad Ave. West, Tehran, Iran, IR-19697,
behnam.asghari1370@gmail.com, ORCID: 0000-0002-0314-1912 (B. A. B.).
**** University of Zanjan, Faculty of Agriculture, University Blvd., 45371-38791, Iran, IR-45371-
38791, ss.hemmati82.sh@gmail.com, ORCID: 0009-0000-0002-1702-0907, (S. H. R.).
DOI: 10.2478/trser-2024-0003
KEYWORDS: vegetation, NDVI, drivers, ecoregions, spatial distribution, OLS.
ABSTRACT
Findings reveal that the majority of studied areas are classified as bare lands, while the
lowest amount is covered by lichens and mosses. Grassland and cropland occupy major areas
of the region, with highest normalized difference vegetation index (NDVI) value saw in 2020,
showing dense vegetation in the western, northwestern and northern regions. Afforestation
efforts shown positive results, with a 4% increase in forested area between 2000 and 2022.
RÉSUMÉ: Processus de modélisation des changements spatio-temporels de la
couverture végétale et de sa relation avec les facteurs de changement dans les zones arides et
humides du Xianjiang (Chine).
Les résultats révèlent que la majorité des zones étudiées sont classées comme des
terres nues, tandis que la plus petite partie est couverte de lichens et de mousses. Les prairies et
les terres cultivées occupent une grande partie de la région, la valeur la plus élevée de lindice
de végétation par différence normalisée (NDVI) étant observée en 2020, ce qui indique une
végétation dense dans les régions de louest, du nord-ouest et du nord. Les efforts de
reboisement ont donné des résultats positifs, avec une augmentation de 4% de la superficie
forestière entre 2000 et 2022.
REZUMAT: Procesul de modelare a schimbărilor spațio-temporale ale acoperirii
vegetale și relația acesteia cu factori ai zonelor uscate și zonelor umede din Xianjiang (China).
Rezultatele arată majoritatea ariilor studiate sunt clasificate ca terenuri goale, cea
mai mică proporție este acoperită de licheni și mușchi. Pajiștile și terenurile cultivate ocupă
arii mari ale regiunii, cu o valoare mare a indicelui de vegetație (NDVI) în 2020, indicând
vegetație densă în regiunile de vest, nord-vest și nord. Eforturile de împădurire au arătat
rezultate positive, cu o creștere cu 4% a suprafeței împădurite între 2000 și 2022.
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
28
INTRODUCTION
Vegetation plays a key role in Earths biosphere, influencing biogeochemical cycles,
carbon balance regulation, and climate stability (He et al., 2023; Zhuang et al., 2019; Jiang et
al., 2017). It also provides environmental protection, defence of soil from erosion, wind and
sand damage, and preventing desertification (Wang et al., 2023; Zhuang et al., 2019).
Additionally, vegetation contributes to the energy cycle by reducing greenhouse gas,
supporting global carbon balance, and ensuring climate stability (Zhuang et al., 2019). Thus, it
is vital to monitor changes in vegetation over time and understand the underlying mechanisms
to effectively manage ecological processes and protect the environment (Li et al., 2021).
Many studies conducted worldwide saw significant changes in vegetation across
various scales (Gadiga, 2015). These studies have highlighted the influence of both natural and
human factors, such as climate change, land use modifications, and ecological interventions on
the dynamics of vegetation patterns (Wang et al., 2023; Wang et al., 2022; Gadiga, 2015).
Natural factors exert long-term control, for instance, weather conditions determine the type of
vegetation, and alterations in rainfall deeply impact vegetation growth and health. On the other
hand, social factors play a significant role in short-term scales. Growth of urban economies and
rise in urban population contribute to the expansion of urban areas (Hu and Hu, 2019).
When a large-scale change in vegetation occurs in an area, it affects the energy
balance of the surface, which may cause heating or cooling effects (Han et al., 2022), land use
change is one of the main challenges that affects natural landscapes and has raised concerns
about sustainable development and food security (Hu and Nacun, 2018; Li et al., 2015). That is
why, countries and organizations around the world have accepted the risks of reducing
vegetation cover and considered measures for protection (Wang et al, 2023). IGBP
(International Geosphere-Biosphere Program) and IHDP (Global Change Human Factors
Program) proposed the research program land use/change of cover (LULC) in 1995 (Liu et
al., 2022). As one of the most active economies in the world, China has undergone many
changes in land use over the past decades, increasingly contributing to environmental crises
(Zhu et al, 2022; Yin et al, 2018). In the last fifty years, China has implemented some
measures to protect the Loess plateau, aiming to mitigate soil erosion and enhance ecological
conditions. Among these initiatives, the project of restoring cultivated land to forests and
grasses has emerged as a prominent ecological revitalization scheme. This project, initiated in
1999, has yielded ample improvements in vegetation quantity and quality on the Loess plateau
(Zhang et al., 2020). However previous studies have shown that between 1700 and 1950 there
was an increasing trend in arable land in China, although they show diverse magnitudes and
rates (Miao et al., 2016). Consequently, vegetation is a comprehensive indicator of
environmental change, and therefore, spatial and temporal changes in vegetation and its
response to climate change have become key issues in global change research (Xianfeng et al.,
2014). Xiao and Weng (2007) examined land use changes in southern China from 1991 to
2001, showing that agricultural land decreased and forest land increased slightly. In the study
of spatial-temporal dynamic changes in vegetation, the plant Normalized Difference index
(NDVI), is an important factor for assessing plant growth status and vegetation distribution
pattern (Duan et al., 2021). Gadiga (2015) studied the state and dynamics of vegetation in the
Mubi Region in Nigeria using the NDVI index, and the vegetation in the region has suffered
record degradation due to the growth of urbanization. Zhuang et al. (2020) conducted a study
investigating vegetation dynamics in Chinas Xinjiang province from 1981 to 2018 using the
normalized vegetation difference index (NDVI). Their findings revealed that multiple factors,
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
29
including precipitation, influenced vegetation changes in the region. Areas with notable
vegetation dynamics were primarily concentrated on the northern and southern slopes of the
Tianshan Mountains, the Ely River Valley, and the Altai region. Meng et al. (2020) assessed
the temporal and spatial variations in the NDVI index in Mongolia and examined natural and
human factors influencing these changes. The results indicated major degradation in desert
steppe and the Gobi Desert within arid regions, while grassland steppe and Alpine steppe
exhibited significant upward trends. The researchers also observed that climatic parameters,
such as precipitation, had a positive impact on vegetation distribution. Overall, these studies
highlight the influence of factors such as precipitation on vegetation dynamics and underscore
the specific regions affected by these changes in Chinas Xinjiang province and Mongolia.
Xinjiang Province, in northwestern China, has been susceptible to the effects of
climate change, experiencing warming over the last four decades. The region has witnessed
notable transformations in vegetation patterns during last years (Han et al., 2022; Zhuang et al.,
2019). Understanding the changes in vegetation cover is crucial in realising the temporal and
spatial dynamics of this region and the feedback between vegetation and the atmosphere.
This study aimed to analyze and model the detection of vegetation cover changes in
various areas of Xinjiang Province over a 23-year period. By harnessing the connectivity of
sensor datasets from remote sensing platforms, time series data was utilized to track these
changes. Examining the relationship between vegetation dynamics and climate and
environmental stimuli is vital in the context of global warming and climate change research, as
it provides insights into the dynamic responses of terrestrial ecosystems to these phenomena.
The study employed the ArcGIS platform to predict the impact of climate and environmental
factors on plant indicators, leveraging open-source data and coding techniques.
MATERIAL AND METHODS
Study area
Xinjiang Province is situated in northwestern China, spanning from 73°20E to
96°25E and 34°15N to 49°10N (Jiapaer et al., 2015). With a total land area of 1.66 million
km2, it accounts for approximately one-sixth of Chinas land area (Yu et al., 2020). The
province encompasses a delicate ecological zone characterized by a complex arid environment,
with mountainous areas comprising 51.4% and plain areas comprising 48.6% of the total land
area (Luo et al., 2019; Jiapaer et al., 2015). Xinjiang exhibits diverse landforms, including the
Altai Mountains in the north, the Kunlun Mountains and A-erh-chin Mountains in the south,
and the Tian Shan Mountains spanning the central part of the province. These major mountain
ranges encircle the Junggar Basin and the Tarim Basin (Zhao et al., 2022), which serve as the
sources of numerous rivers (Luo et al., 2019). The Tian Shan Mountains, Altai Mountains, and
Kunlun Mountains harbor extensive forest and grassland vegetation, while oases and cities are
located in the valley plains. As of 2018, forest land and grassland accounted for 31.29% of the
land area, with grassland being the predominant vegetation type (Zhuang et al., 2020). The
Junggar and Tarim Basins are characterized by typical temperate desert vegetation. Despite
belonging to a temperate continental climate, Xinjiang exhibits distinct differences between its
northern and southern regions due to its extensive north-south span. The annual precipitation in
northern Xinjiang ranges from 100 to 500 mm, with average temperatures ranging from 4 to
8°C. In southern Xinjiang, annual precipitation ranges from 20 to 100 mm, while average
temperatures range from 10 to 13°C (Cao and Gao, 2022). Figure 1 provides a visual
representation of Xinjiang Provinces geographical location and its ecoregions.
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
30
Preparation of vegetation data from 2000 to 2023
Modeling the study is based on calculating satellite data. Calculations were completed
with the help of the Google Earth Engine (GEE) platform and catalogs of raster files on the
platform.Other open sources of data were used such as MODIS/061/myd13a1-NDVI (16 daily-
2000 to 2023 with a spatial resolution of 250 meters global scale, and data sets of drivers
such as daily LST (Kelvin), precipitation (mm/d), global land cover extracted from Sentinel-1
and 2 data and Global 30 Arc-Second Elevation to extract DEM and slope. Data through
ArcGIS platform was extracted to track vegetation cover changes, areas, and calculation of the
drivers affecting the vegetation. The main sources and characteristics of the data collected are
displayed in table 1.
Table 1: Main sources of data with the satellite products specifications.
Sattelite/
Model
Product
name
Date
name
Unit
Spatial
resolution
Temporal
resolution
Time
extent
Modis/
Terra
MOD11A1.061
Terra Land
Surface
Temperature
and Emissivity
Daily Global
1 km
LST_Day 1
km
LST_Night
1 km
Kelvin
1,000 m
Daily
2000-02-24
Modis/
Terra
Vegetation
Indices
NDVI
500 m
Daily
2000-02-18
Modis/
Terra
Vegetation
Indices
NDVI
500 m
Daily
2000-02-18
CHIRPS
CSB-
CHG/CHIRPS/
DAILY
Precipitation
mm/d
Daily
1981-01-
01T00:00:00
Z2023-06-
30T00:00:00
Sentinel-1
and 2 data
ESA World
Cover 2020
Global
Land
cover
product
10 m
Annually
2020-2021
GTOPO30
Global 30 Arc-
Second
Elevation
DEM
m
1 km
NaN
1996
Ecoregion data
Ecoregions, defined as regions with similar ecosystems, have become increasingly
significant in the evaluation and administration of environmental matters. They offer a
comprehensive framework that allows for flexible and comparative analysis of intricate
environmental issues (Lovland and Merchant, 2004). We downloaded Ecoregion data from
https://ecoregions.appspot.com and cropped it for Xinjiang Province. There are 14 terrestrial
biomes and 846 ecoregions worldwide. Plant communities within a given biome may exhibit
comparable structural characteristics while harboring distinct species compositions. Table 2,
list the 17 ecoregions and five biomes in Xinjiang Province. Figure 1 shows the positions of
each of the ecoregions of Xianjian Province.
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
31
Figure 1: Positions of each of the ecoregions in the Xianjiang, China.
Taklimakan Desert ecoregion with 742,657 km2 accounts for 18.1% of the province
and Tian Shan montane conifer forests ecoregion with 27,568 km2 covers 0.67% of the
province, representing the largest and smallest areas among the Xinjiang-Tian Shan ecoregions
(Tab. 2).
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
32
Table 2: Xinjiang ecoregions, classified by name, area, percentage.
Area km2
%
674,352
16.498
90,434
2.212
142,875
3.495
83,192
2.035
629,190
15.393
65,135
1.594
304,938
7.460
143,265
3.50
374,494
9.162
118,072
2.889
192,147
4.701
34,830
0.852
742,657
18.169
54,533
1.334
129,231
3.162
275,68
0.674
280,611
6.865
Implementation phases
Tracking and analysis of the models took place in several phases: Identifying the
current state of the study border; Calculation and zoning of vegetation at the study boundary
(time series 2000 to 2023) and determination of maximum and minimum density; Analysis of
spatial-temporal distribution of vegetation since 2000 by classification method; Areas where
the value of NDVI is greater than 0.5 and vice versa areas where the value of NDVI is less than
0.5.; Estimation of vegetation density using kernel density and point density: 1) Monitoring and
evaluation of vegetation cover from the past 23 years using land surface measurements
(Divisions by municipalities), 2) Monitoring and evaluation of vegetation cover from 2000-
2023 using land surface measurements (Divisions by ecoregions); Estimation of the area and
percentage of afforestation and deforestation at the beginning and end of the desired time
period (2000-2022); Identifying the pattern of spatial distribution of vegetation in the study
border with spatial autocorrelation with a confidence level above 90%, Average Nearest
Neighbor; Estimation of hot spots and cold spots of vegetation in the border of the study with a
confidence level above 90% using hot spot analysis, clusters and outliers; Identification of
factors affecting the distribution of vegetation in the study border (land use, land surface
temperature, rainfall, slope, elevation); Spatial modeling of factors affecting vegetation (OLS).
RESULTS
Identifying the current state of the study border
Figure 2 shows patterns for the digital elevation model (A), slope (B), precipitation
parameter (C) and LULC variations calculated. According to the digital elevation model, the
area studied has expanded between 152 m and 7,553 m, which will create a wide range of plant
species, each plant specific to an elevation surface.
As much as altitude increase, precipitation increase, temperature decrease, dry stress
decrease, and the vegetation develops, this trend continues till higher altitudes where low
temperatures reduces the vegetation. The gradient variation pattern (Fig. 2B) also represents
gradient fluctuations of 0 to 356 degrees in the study area, which is very diverse. There is a
question of which slopes in the Xinjiang ecoregions have more vegetation?
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
33
For the northern hemisphere of the Earth, including Xinjiang, which is wide and rich
in diversity, the slopes facing south receive more sunlight and become drier and warmer,
support drought-resistant vegetation and are less favorable for trees, while the slopes facing
north retain moisture and are cold and humid and are suitable for humid plants.
The rain pattern is displayed in figure 2C. The highest rainfall was 704 mm and the
lowest was 13 mm. Precipitation is one of the most important factors affecting the growth and
development of vegetation. According to studies conducted in many Xinjiang ecosystems,
especially grasslands and arable land, the growth of seasonal plants occurs entirely within the
rainy season. In times of drought, vegetation grows poorly in these ecosystems. If it rains, the
sudden growth of vegetable plants actually turns. Unsure what is being communicated.
By manipulating the Sentinel-1 and 2 satellite data sets, tracking land use and land
cover changes, for 2020, LULC was considered as one of the drivers that reflected vegetation
in the ecoregions of Xinjiang. Figure 2D shows this classification visually, and table 3
examines the classification of LULC and their area and percentage.
A B
C D
Figure 2: Digital elevation model (A), slope (B), precipitation (C) and LULC (D).
Table 3 shows, the largest area is the Bare area class, (139,433.7 km2, or ying
61.49%), and the smallest area is the Lichens and Mosses class (41.8 km2, or 0.018%). Bare
land refers to areas where dominant vegetation covers under 90%. This category contains areas
without artificial cover and includes regions with less than 4% vegetative cover. Examples
include bare rock regions, sandy areas, and deserts. Table 3 shows that mosses and lichens has
the smallest area, however it is argued that it plays a very important role in study area. Both
mosses and lichens have a significant role in maintaining a healthy environment as they
contribute to the absorption of carbon dioxide and other air pollutants. Lichens, in particular,
possess valuable detoxifying properties that make them highly beneficial to various organisms.
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
34
Table 3: Classification of LULC alignments, area and percentage.
LULC
Area
km2
%
Crop land
16,263.74
7.17
Herbaceous cover
13,615.37
6
Tree cover
2,038.06
0.89
Tree and shrub
1,989.67
0.87
Grassland
48,112
21.2
Lichens and mosses
41.8
0.018
Urban areas
261.83
0.11
Bare areas
139,433.7
61.49
Water bodies
1,306.74
0.57
Snow and ice
3,692.84
1.62
Table 3 is divided into valuable LULC classes and their respective area and percent
coverage of the study area. These classes together create an incredibly diverse and complex
BioCycle and any changes in urban area, cropland and bare area classes will greatly affect the
quality and quantity of valuable classes.
Calculation and zoning of vegetation at the study boundary in time series 2000 to
2023 and determination of maximum and minimum density
Satellites carry numerous sensors that measure the reflection of red and near-infrared
light waves from land surfaces. By utilizing mathematical algorithms, scientists convert the
raw satellite data related to these light waves into vegetation indices. These indices serve as
indicators of vegetation greenness, representing the relative density and health of vegetation
for each pixel in a satellite image. Among various vegetation indices, the Normalized
Difference Vegetation Index (NDVI) is widely employed. NDVI values range from +1.0 to
1.0. Barren areas such as rocks, sand, or snow typically exhibit very low NDVI values, often
0.1 or less. Regions with sparse vegetation like shrubs, grasslands, or aging crops show
moderate NDVI values ranging from approximately 0.2 to 0.5. High NDVI values, around 0.6
to 0.9, correspond to dense vegetation found in temperate and tropical forests or crops during
their peak growth stage (Tahir et al., 2019; Elsu et al., 2017). Researchers can utilize raw
satellite data to convert it into NDVI values, which enables the creation of images and other
products that provide a general assessment of vegetation characteristics, quantity, and
condition on a global scale. NDVI is particularly valuable for monitoring vegetation at
continental to global levels, as it can account for variations in lighting conditions, surface
slope, and viewing angles. However, it should be noted that NDVI may become saturated in
areas with dense vegetation and is influenced by the color of the underlying soil. By averaging
NDVI values over time, researchers can establish a baseline for normal growing conditions in
a specific region during a particular time of year. Subsequent analysis can then evaluate the
relative health of vegetation compared to this baseline. Over time, NDVI analysis can identify
areas where vegetation is flourishing or experiencing stress, as well as detect changes in
vegetation patterns resulting from human activities like deforestation, natural disturbances
such as wildfires, or shifts in plants’ phenological stages (*).
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
35
Figure 3 shows the visual patterns of vegetation classification calculated with
the average NDVI value for the time series 2000 to 2023. The highest value of NDVI is for
2020 (Fig. 3E), with a value of 0.933 predicted, and according to the classification of NDVI
value balances; it represents areas with dense vegetation. The lowest value of -0.199
corresponds to 2010 (Fig. 3C), representing areas with the presence of abundant sand, rocks,
and snow.
Based on the NDVI value patterns depicted in figure 3, the western, northwestern, and
northern areas of Xinjiang Province exhibit the highest NDVI values. These regions
encompass various ecoregions, including the arid steppe of the Tian Shan foothills, conifer
forests of the Tian Shan Mountains, steppe and meadows of the Tian Shan montane region,
steppe in the Emin Valley, steppe and semi-desert in Altai, forest and forest steppe in Altai
Mountains, and alpine meadow and tundra in Altai.
The average value of NDVI is estimated to be approximately between 0.2 and 0.5
which are spread in the central, western and northwestern and southwestern regions, are also
broadcasted in the form of spots in the eastern regions of Xinjiang Province, which this type of
classification represents scattered vegetation, savanna, shrubs, meadows, and agricultural
lands. The classification of certain areas within the Tian Shan montane steppe and meadow
ecoregion serves as a notable illustration of this categorization.
Ecoregions like Taklimakan Desert, north Tibetan Plateau-Kunlun mountains
alpine desert and junggar basin semi desert positioned in the central, central south,
southern and northern regions of Xinjiang Province respectively experienced the lowest
value of NDVI, and are covered with very scattered vegetation, sand, barren rock and rarely
snow (Fig. 3).
Analysis of spatial-temporal distribution of vegetation 2000-2023 by classification
method (Areas with NDVI > 0.5 and NDVI < 0.5).
The temporal and spatial distribution of the average annual NDVI value
was examined. This classification specifically identifies regions where the NDVI value
exceeds 0.5, and regions where the NDVI value is below 0.5, based on data from 2000 to 2023
(Fig. 5). 2015 had the highest area of the regions, with a NDVI value greater than 0.5 in
Xinjiang Province, which has an area of 149,450.6 km2 (Tab. 4, Fig. 5D). 2000 also
experienced the largest area with a NDVI less than 0.5, this value includes 1,614,503.6 km2
(Tab. 4, Fig. 5A). In summary, the range of NDVI values spans from 1.0 to 1.0. Negative
values represent clouds and water, values close to zero indicate bare soil, and higher positive
values of NDVI correspond to varying levels of vegetation density, ranging from sparse
vegetation (0.1-0.5) to dense green vegetation (0.6 and above). 2015 has experienced the
highest level of dense vegetation, and 2000 is also covered with a sporadic and vast vegetation
community.
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
36
A B
C D
E F
Figure 3: Visual patterns of temporal and spatial changes of vegetation
calculated by the mean value of NDVI for the time series 2000 to 2023.
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
37
Table 4: Statistics on area in km2 from 2000 to 2023.
Year
2000
2005
2010
2015
2020
2023
Area (km2) of arenas where
the value of the NDVI < 0.5
1614503.61
1600279.06
1587131.76
1574324.40
1583306.92
1578566.28
Area (km2) of arenas where
the value of the NDVI > 0.5
109396.39
123345.94
136518.24
149450.60
142168.08
146908.72
Figure 4 describes changes in boundary areas from 2000-2023. Figure 4A clearly
shows a decrease in area for arenas with a NDVI of less than 0.5. On the other hand figure 4B
shows a unique phenomenon, where area of areas with NDVI more than 0.5 is increasing,
indicating the existence of suitable conditions for vegetation development as well as correct
management of the arenas. The increasing quantity of positive NDVI values signifies a rise in
the presence of green vegetation. What are the factors that affect NDVI? NDVI is an important
indicator that reflects the state of the regional climate and the environment, which is affected
by precipitation, temperature, the water content of the soil, humans, etc. For instance, Yao et
al. (2023) conducted a study utilizing GeoDetector to investigate the impact of various factors
on NDVI, such as climate, soil, topography, and human activities. Their findings revealed that
the primary contributing factors were night light brightness (51.9%), annual average air
temperature (47%), and annual average atmospheric pressure (45.8%). Additional research has
demonstrated that NDVI exhibits a stronger correlation with changes in rainfall (R = 0.75)
compared to temperature and elevation. Generally, NDVI values tend to decrease under
extremely high and low temperature conditions. Furthermore, the growth pattern of NDVI
varies with altitude, with a more pronounced response observed between 70 m and 1,500 m
above sea level. These aspects will be further examined in the subsequent sections of this
article, exploring the relationship between NDVI and rainfall, land use and land cover changes,
elevation, slope, and LST, as well as modeling the interconnectedness of these factors.
Figure 4: Changes in the area of studied boundaries (km2) where the NDVI value is less
than 0.5.(A) and NDVI value is more than 0.5 (B).
Figure 5 reports visual patterns of NDVI value, for the northern regions of Xinjiang
such as Altai Mountains forest and forest steppe, Altai alpine meadow and tundra, parts of the
Emin Valley steppe and large parts of the Tian Shan ecoregions. Likewise it reports the
montane steppe meadows, the montane conifer forest and the foothill arid steppe and the
central parts of this province are seen as spots that have experienced NDVI values >5 from
2000 to 2023. This value reached its highest value in 2015 (Tab. 4, Fig. 4B), After a hiatus in
2020, it increased again until 2023. Areas that had NDVI values <5 and have scattered
vegetation, bare land, sand and agricultural land (displayed in red), with the exception of 2020,
have experienced a downward trend (Fig. 4A). Overall the very positive ecological event is
taking shape and every year we see an increase in conservation, restoration, and development
of vegetation coverage in all parts of the Xinjiang Province. This constructive process is more
prominent in central and western parts such as the Tian Shan Mountains and northern areas
such as the Altai Mountains and the Emin Valley (Fig. 5).
A
B
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
38
A B
C D
E F
Figure 5: Visual patterns of NDVI temporal and spatial distribution along with area of arenas
with a value greater than 0.5 and less than 0.5 for 2000 to 2023.
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
39
Estimation of vegetation density using kernel density and point density
monitoring and evaluation of vegetation cover from 2000-2023 using land surface
measurements (Divisions by municipalities)
Figure 6 shows temporal and spatial patterns of monitoring and evaluating vegetation
changes using land surface measurements, which are borders of counties, and have modeled
changes since 2000. The NDVI value is an average of the total months of the year. The
modeling process relies on assessing vegetation density and leverages the kernel density
method. This approach involves calculating the density of point features surrounding each
output raster cell. Essentially, a smooth curved surface is fitted over each point in the model.
Kernel density estimation, a statistical technique, is utilized for estimating the probability
density function of a random variable using kernel smoothing. In this case, the variable of
interest is the annual average NDVI value.
Kernel estimator results argue that the maximum values close to NDVI >0.55 with
green, and the minimal values are close to NDVI <0.07, highlighted in red. The range between
maximum and minimum is aligned with yellow and orange colors (Fig. 6). In general, the
northern and western counties of Xinjiang Province have experienced the development of
vegetation, in other words the southern and eastern counties have lost the harmony of warm
red and orange colors and turned into pale yellow and green colors (Fig. 6).
According to table 5, the highest average NDVI value is related to Ili Kazakh County,
which experienced a value of 0.54 in 2000. The lowest average NDVI value of 0.07 also
belongs to Khotan County.
According to NASA Earth Observatory website (***), extremely low NDVI values
(0.1 and below) indicate barren areas such as rock, sand, or snow. Moderate values (0.2 to 0.3)
correspond to shrub and grassland, while high values (0.6 to 0.8) signify temperate and tropical
rainforests. Applying this principle, it can be observed that Khotan County predominantly
consists of very sparse vegetation, desert, sand, and rock, whereas Ili Kazakh County and
Shihezi County exhibited a relatively moderate vegetation situation in 2000, characterized by
grasslands, shrubs, as well as broadleaf and coniferous species (Fig. 7).
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
40
A B
C D
E F
Figure 6: Examination of spatial and temporal patterns of vegetation density
using Kernel density estimation.
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
41
Table 5: Classification of vegetation cover (NDVI) based on county divisions in two.
No.
Name
Minimum
Maximum
Mean
1
Aksu
0.052921
0.691787
0.136181
2
Altay
0.123229
0.859978
0.270487
3
Bayingholin
0.191100
0.878285
0.104023
4
Bortala
0.171962
0.874815
0.290748
5
Changji Hui
0.033705
0.803515
0.233974
6
Hami
0.027253
0.678575
0.088444
7
Ili Kazakh
0.043728
0.885946
0.54512
8
Karamay
0.03571
0.541906
0.124138
9
Kashgar
0.075600
0.536531
0.118364
10
Khotan
0.117503
0.677344
0.070696
11
Kizilsu
0.054392
0.547
0.137629
12
Shihezi
0.201837
0.640216
0.484435
13
Tacheng
0.156277
0.819484
0.253239
14
Turfan
0.038627
0.58874
0.082853
15
Urümqi
0.038954
0.724495
0.307904
Figure 7: Statistics of average annual NDVI value
based on the division of county boundaries in 2000.
Tables 5-10 as well as figures 7-12 show the NDVI values of the cities of Xinjiang
Province 2000-2003. The results showed that Ili Kazakh city and then Shihezi city had
the highest NDVI value during the period studied. The lowest NDVI value was for Khotan
city. Examining the NDVI index values in Ili Kazakh city 2000 to 2023 showed that the
amount of vegetation has grown significantly and improved. In this city, the NDVI value in
2000 was 0.54, which gradually increased in 2005, 2010, and 2015 (0.55, 0.57, and 0.57,
respectively).
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
42
Table 6: Classification of vegetation cover (NDVI) based on couty divisions in 2005.
No.
Name
Minimum
Maximum
Mean
1
Aksu
0.122748
0.742965
0.148695
2
Altay
0.192430
0.927757
0.279236
3
Bayingholin
0.184302
0.869611
0.10498
4
Bortala
0.155703
0.875777
0.307521
5
Changji Hui
0.027991
0.773594
0.229612
6
Hami
0.073582
0.669499
0.088761
7
Ili Kazakh
0.046839
0.88735
0.555844
8
Karamay
0.042529
0.670178
0.169616
9
Kashgar
0.175655
0.563468
0.135556
10
Khotan
0.074800
0.60319
0.079338
11
Kizilsu Kirghiz
0.055451
0.5445
0.1639
12
Shihezi
0.170395
0.704664
0.521317
13
Tacheng
0.048387
0.808383
0.279653
14
Turfan
0.011559
0.578973
0.084733
15
Urümqi
0.030313
0.695617
0.269192
Figure 8: Statistics of average annual NDVI value
based on the division of county boundaries in 2005.
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
43
Table 7: Classification of vegetation cover (NDVI) based on couty divisions in 2010.
No.
Name
Minimum
Maximum
Mean
1
Aksu
0.065500
0.687213
0.160299
2
Altay
0.116237
0.864199
0.286228
3
Bayingholin Mongol
0.187300
0.869814
0.111299
4
Bortala Mongol
0.190533
0.820894
0.315226
5
Changji Hui
0.023896
0.81202
0.274984
6
Hami
0.072722
0.68649
0.099629
7
Ili Kazakh
0.033036
0.878573
0.570325
8
Karamay
0.025764
0.712739
0.204186
9
Kashgar
0.080600
0.684593
0.143876
10
Khotan
0.074623
0.694637
0.081276
11
Kizilsu Kirghiz
0.064005
0.662137
0.156734
12
Shihezi
0.261811
0.7112
0.56347
13
Tacheng
0.128363
0.807561
0.305837
14
Turfan
0.011221
0.593036
0.086799
15
Urümqi
0.033983
0.75133
0.304136
Figure 9: Statistics of average annual NDVI value based on the division of county boundaries in 2010.
Table 8: Classification of vegetation cover (NDVI) based on couty divisions in 2015.
No.
Name
Minimum
Maximum
Mean
1
Aksu
0.060800
0.666599
0.167056
2
Altay
‒0.199042
0.866951
0.279071
3
Bayingholin Mongol
‒0.199100
0.857465
0.111562
4
Bortala Mongol
‒0.150868
0.83892
0.337688
5
Changji Hui
‒0.038100
0.841333
0.271983
6
Hami
‒0.085294
0.737637
0.11500
7
Ili Kazakh
0.041731
0.877921
0.579145
8
Karamay
0.007897
0.780512
0.210299
9
Kashgar
‒0.086180
0.682347
0.151257
10
Khotan
‒0.080100
0.592517
0.079768
11
Kizilsu Kirghiz
‒0.056029
0.521963
0.152814
12
Shihezi
0.19053
0.760908
0.510827
13
Tacheng
‒0.189183
0.830494
0.30230
14
Turfan
0.005873
0.629756
0.096969
15
Urümqi
‒0.049939
0.726139
0.30828
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
44
Figure 10: Statistics of average annual NDVI value
based on the division of county boundaries in 2015.
Table 9: Classification of vegetation cover (NDVI) based on couty divisions in 2020.
No.
Name
Minimum
Maximum
Mean
1
Aksu
0.069200
0.779886
0.18251
2
Altay
0.141897
0.848825
0.27150
3
Bayingholin Mongol
0.127306
0.822712
0.115179
4
Bortala Mongol
0.135692
0.933632
0.330082
5
Changji Hui
0.019654
0.797883
0.246454
6
Hami
0.030933
0.652469
0.091052
7
Ili Kazakh
0.039898
0.865697
0.541766
8
Karamay
0.036021
0.781246
0.211123
9
Kashgar
0.057445
0.646838
0.160216
10
Khotan
0.076941
0.614853
0.08908
11
Kizilsu Kirghiz
0.051024
0.52605
0.166212
12
Shihezi
0.134233
0.817762
0.5278
13
Tacheng
0.043740
0.825483
0.30307
14
Turfan
0.013841
0.592871
0.0888
15
Urümqi
0.033376
0.67819
0.267829
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
45
Figure 11: Statistics of average annual NDVI value
based on the division of county boundaries in 2020.
Table 10: Classification of vegetation cover (NDVI) based on couty divisions in 2023.
No.
Name
Minimum
Maximum
Mean
1
Aksu
0.060800
0.736774
0.178739
2
Altay
0.182885
0.83864
0.271577
3
Bayingholin Mongol
0.162868
0.833813
0.113792
4
Bortala Mongol
0.177081
0.847629
0.3529
5
Changji Hui
0.009055
0.800372
0.24254
6
Hami
0.058832
0.640684
0.08512
7
Ili Kazakh
0.037388
0.87388
0.579248
8
Karamay
0.038863
0.790834
0.210868
9
Kashgar
0.076800
0.651084
0.16133
10
Khotan
0.076644
0.558047
0.08468
11
Kizilsu Kirghiz
0.052250
0.661522
0.166611
12
Shihezi
0.136324
0.814858
0.5312
13
Tacheng
0.039781
0.841734
0.29065
14
Turfan
0.019744
0.576473
0.088044
15
Urümqi
0.011569
0.687454
0.264764
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
46
Figure 12: Statistics of average annual NDVI value
based on the division of county boundaries in 2023.
However, in 2020, the NDVI value in Ili Kazakh city area decreased slightly
(0.54), then increased again in 2023 (0.58). In Shihezi city, this significant trend of
vegetation growth was observed, with average NDVI value reaching from 0.48 in 2000 to 0.53
in 2023.
This growth trend was seen NDVI index values for Khotan County, which had the
lowest NDVI rate among the different counties of Xinjiang Province. According to tables 5-10,
as well as figures 7-12, the NDVI value in Khotan County in 2000 and 2005 averaged 0.07,
which gradually reached 0.08 in 2010, 2015, 2020 and 2023.
According to the information provided on the NASA Earth Observatory website (***),
it is stated that low NDVI values (0.1 and below) indicate barren areas consisting of rock,
sand, or snow. Moderate values (0.2 to 0.3) suggest the presence of shrubs and meadows,
while high values (0.6 to 0.8) are associated with temperate and tropical rainforests. Applying
this principle, it can be argued that the region of Khotan has minimal vegetation,
predominantly comprising desert, sand, and rock. On the other hand, the cities of Ili Kazakh
and Shihezi exhibit a relatively moderate level of vegetation, including meadows, shrubs, as
well as broadleaf and coniferous species.
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
47
Monitoring and evaluation of vegetation cover from 2000-2023 using land surface
measurements (Divisions by ecoregions)
Figure 13 shows the pattern of spatial and temporal changes in vegetation cover in the
ecoregions of Xinjiang Province during 2000-2023 using the kernel density function.
A B
C D
E F
Figure 13: Patterns of temporal and spatial changes in vegetation cover
using kernel density estimators in the borders of the ecorigens Xinjiang Province.
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
48
Most of the Taklimakan Desert is completely devoid of vegetation, but hardy plants
such as Halogeton, Halostachys, and camelthorn grow in its foothills. Although the value of
NDVI in this ecoregion is low, it has had a growing trend during the studied years, the value of
NDVI increased from 0.08 in 2000 to 0.09 in 2005 and 2010 and gradually increased to 0.1 in
2015, 2020 and 2023 (Tabs. 11-16, Fig. 13).
This growing trend in the amount of vegetation was also observed in other ecoregions
that have thin and scattered vegetation. Among them are the ecoregions of North Tibetan
Plateau-Kunlun Mountains alpine, Junggar Basin semi-desert, Tarim Basin deciduous forests
and steppe, and Pamir alpine desert and tundra, respectively in the southern, northern, central
and southwest of Xinjiang Province. Therefore, in these ecoregions, the vegetation is sparse
and the NDVI is low. However, in these ecoregions, the amount of vegetation has improved
since 2000, with mean NDVI levels going from 0.08, 0.15, 0.08 and 0.11 in 2000 to 0.10, 0.18,
0.12 and 0.13 respectively in 2023.
Table 11: Classification of vegetation cover (NDVI) based on ecoregions divisions in
2000.
Name
Minimum
Maximum
Mean
Area of
arenas
with NDVI
classify-
cation
(km2)
Alashan Plateau semi-desert
0.020716
0.167403
0.066595
1.549515
Altai alpine meadow and tundra
0.041297
0.8356
0.447685
1.847604
Altai montane forest and forest steppe
0.110255
0.859978
0.586297
2.038968
Altai steppe and semi-desert
0.24145
0.799791
0.521414
0.254039
Central Tibetan Plateau alpine steppe
0.0532
0.0789
0.069438
0.015808
Emin Valley steppe
0.055365
0.819484
0.290983
5.343731
Junggar Basin semi-desert
0.171962
0.874815
0.152806
27.366403
Karakoram-West Tibetan Plateau alpine steppe
0.062345
0.218932
0.031641
0.779973
North Tibetan Plateau-Kunlun Mountains alpine desert
0.164422
0.513398
0.086163
19.851451
Pamir alpine desert and tundra
0.054019
0.424546
0.119441
3.417256
Qaidam Basin semi-desert
0.191100
0.278864
0.080643
2.209025
Rock and Ice
0.075600
0.269455
0.012968
1.099327
Taklimakan Desert
0.181898
0.793929
0.086815
76.751566
Tarim Basin deciduous forests and steppe
0.016457
0.61935
0.089074
5.763608
Tian Shan foothill arid steppe
0.10338
0.789768
0.439169
0.972174
Tian Shan montane conifer forests
0.086819
0.885946
0.530611
1.425375
Tian Shan montane steppe and meadows
0.093751
0.875281
0.353765
21.228148
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
49
Table 12: Classification of vegetation cover (NDVI) based on ecoregions divisions in
2005.
Name
Minimum
Maximum
Mean
Area of
arenas
with NDVI
classify-
cation
(km2)
Alashan Plateau semi-desert
0.024269
0.137483
0.063685
1.549515
Altai alpine meadow and tundra
0.035260
0.83096
0.451796
1.847604
Altai montane forest and forest steppe
0.10491
0.863939
0.57805
2.038968
Altai steppe and semi-desert
0.267149
0.808383
0.544267
0.254039
Central Tibetan Plateau alpine steppe
0.0578
0.0776
0.069378
0.015808
Emin Valley steppe
0.064729
0.80623
0.317591
5.343731
Junggar Basin semi-desert
0.192430
0.927757
0.168551
27.365448
Karakoram-West Tibetan Plateau alpine steppe
0.064121
0.26982
0.026307
0.779973
North Tibetan Plateau-Kunlun Mountains alpine desert
0.184302
0.550939
0.094414
19.846451
Pamir alpine desert and tundra
0.054584
0.422466
0.13886
3.417256
Qaidam Basin semi-desert
0.171300
0.26967
0.082013
2.201525
Rock and Ice
0.074800
0.344393
0.021671
1.099327
Taklimakan Desert
0.175655
0.78328
0.095594
76.756566
Tarim Basin deciduous forests and steppe
0.122748
0.742965
0.101872
5.763608
Tian Shan foothill arid steppe
0.128404
0.761067
0.499164
0.972174
Tian Shan montane conifer forests
0.042952
0.872852
0.492354
1.425375
Tian Shan montane steppe and meadows
0.108059
0.88735
0.342893
21.209103
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
50
Table 13: Classification of vegetation cover (NDVI) based on ecoregions divisions in
2010.
Name
Minimum
Maximum
Mean
Area of
arenas with
NDVI
classify-
cation
(km2)
Alashan Plateau semi-desert
0.0251
0.156013
0.069917
1.549515
Altai alpine meadow and tundra
0.030100
0.8483
0.447748
1.847604
Altai montane forest and forest steppe
0.153305
0.864199
0.567695
2.038968
Altai steppe and semi-desert
0.238124
0.807561
0.587318
0.254039
Central Tibetan Plateau alpine steppe
0.066
0.0851
0.077617
0.015808
Emin Valley steppe
0.067906
0.798935
0.335793
5.343731
Junggar Basin semi-desert
0.128363
0.820894
0.188803
27.378903
Karakoram-West Tibetan Plateau alpine steppe
0.065062
0.195633
0.025916
0.779973
North Tibetan Plateau-Kunlun Mountains alpine
desert
0.163778
0.537248
0.102822
19.848951
Pamir alpine desert and tundra
0.064005
0.415137
0.12624
3.417256
Qaidam Basin semi-desert
0.187300
0.348592
0.098393
2.186525
Rock and Ice
0.080600
0.261311
0.019348
1.099327
Taklimakan Desert
0.125400
0.7671
0.099849
76.726566
Tarim Basin deciduous forests and steppe
0.011591
0.65088
0.101618
5.763608
Tian Shan foothill arid steppe
0.132332
0.795914
0.516423
0.972174
Tian Shan montane conifer forests
0.075842
0.876967
0.548598
1.425375
Tian Shan montane steppe and meadows
0.190533
0.878573
0.365406
21.240648
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
51
Table 14: Classification of vegetation cover (NDVI) based on ecoregions divisions in
2015.
Name
Minimum
Maximum
Mean
Area of
arenas
with
NDVI
classify-
cation
(km2)
Alashan Plateau semi-desert
0.02805
0.152062
0.073175
1.549515
Altai alpine meadow and tundra
0.014300
0.849298
0.46039
1.847604
Altai montane forest and forest steppe
0.1477
0.866951
0.571273
2.038968
Altai steppe and semi-desert
0.2203
0.808246
0.576432
0.254039
Central Tibetan Plateau alpine steppe
0.0643
0.0867
0.074508
0.015808
Emin Valley steppe
0.065934
0.830494
0.333953
5.343731
Junggar Basin semi-desert
0.199042
0.83892
0.187073
27.371403
Karakoram-West Tibetan Plateau alpine steppe
0.064099
0.218998
0.031506
0.779973
North Tibetan Plateau-Kunlun Mountains alpine desert
0.199100
0.544989
0.098162
19.856451
Pamir alpine desert and tundra
0.045398
0.469946
0.137724
3.417256
Qaidam Basin semi-desert
0.185452
0.296632
0.091435
2.201525
Rock and Ice
0.086180
0.290036
0.017878
1.099327
Taklimakan Desert
0.167808
0.7579
0.10416
76.734066
Tarim Basin deciduous forests and steppe
0.030569
0.655555
0.111113
5.763608
Tian Shan foothill arid steppe
0.109939
0.825393
0.53483
0.972174
Tian Shan montane conifer forests
0.04382
0.858642
0.5565
1.425375
Tian Shan montane steppe and meadows
0.085294
0.877921
0.368744
21.230648
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
52
The distinct topographical features found in the Tianshan and Altai mountains have
resulted in the presence of abundant grassland and forest vegetation. Consequently, the
ecoregions situated in this region, namely Tian Shan foothill arid steppe, Altai montane forest
and forest steppe, Altai steppe and semi-desert, Tian Shan montane conifer forests, Altai alpine
meadow and tundra, and Emin Valley steppe, exhibit substantial vegetation with high NDVI
values. As can be seen, the revitalization and development of vegetation in these ecoregions is
more prominent during the period under study. Only in 2020, NDVI decreased in these
ecoregions, but it again has an upward trend and normal vegetation has increased. This
revitalization and development of vegetation is more prominent in the Tian Shan foothill arid
steppe ecoregion, so that its NDVI has increased from 0.48 in 2020 to 0.56 in 2023 (Fig. 13,
Tabs. 15 and 16).
Table 15: Vegetation cover classification (NDVI) based on ecoregions divisions (2020).
Name
Minimum
Maximum
Mean
Area of
arenas
with
NDVI
classify-
cation
(km2)
Alashan Plateau semi-desert
0.031
0.129193
0.066389
1.549515
Altai alpine meadow and tundra
‒0.020200
0.8201
0.425986
1.847604
Altai montane forest and forest steppe
0.11874
0.848825
0.54862
2.038968
Altai steppe and semi-desert
0.20505
0.76781
0.560979
0.254039
Central Tibetan Plateau alpine steppe
0.0695
0.0887
0.080397
0.015808
Emin Valley steppe
0.068533
0.819078
0.324058
5.343731
Junggar Basin semi-desert
‒0.141897
0.825483
0.184406
27.421403
Karakoram-West Tibetan Plateau alpine steppe
‒0.042049
0.190949
0.036133
0.779973
North Tibetan Plateau-Kunlun Mountains alpine desert
‒0.074712
0.614853
0.11043
19.873951
Pamir alpine desert and tundra
‒0.051024
0.405781
0.13389
3.417256
Qaidam Basin semi-desert
‒0.092845
0.535508
0.102987
2.241525
Rock and Ice
‒0.076941
0.259547
‒0.009148
1.099327
Taklimakan Desert
‒0.127306
0.779886
0.107867
76.786566
Tarim Basin deciduous forests and steppe
0.034547
0.728055
0.124142
5.763608
Tian Shan foothill arid steppe
0.142023
0.859359
0.483168
0.972174
Tian Shan montane conifer forests
0.044904
0.845019
0.488253
1.425375
Tian Shan montane steppe and meadows
‒0.102500
0.933632
0.34722
21.240648
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
53
Table 16: Vegetation cover classification (NDVI) based on ecoregions divisions
(2023).
Name
Minimum
Maximum
Mean
Area of
arenas
with
NDVI
classify-
cation
(km2)
Alashan Plateau semi-desert
0.031885
0.112891
0.06282
1.549515
Altai alpine meadow and tundra
‒0.017536
0.813509
0.423099
1.847604
Altai montane forest and forest steppe
0.122865
0.83864
0.54210
2.038968
Altai steppe and semi-desert
0.21795
0.778232
0.53419
0.254039
Central Tibetan Plateau alpine steppe
0.0655
0.0856
0.078207
0.015808
Emin Valley steppe
0.060161
0.828862
0.301536
5.343731
Junggar Basin semi-desert
0.182885
0.841734
0.18347
27.421403
Karakoram-West Tibetan Plateau alpine steppe
0.047465
0.218858
0.046283
0.779973
North Tibetan Plateau-Kunlun Mountains alpine desert
0.135032
0.557652
0.106581
19.873951
Pamir alpine desert and tundra
0.057781
0.411547
0.13789
3.417256
Qaidam Basin semi-desert
0.162868
0.353611
0.094669
2.241525
Rock and Ice
0.076800
0.356202
0.007643
1.099327
Taklimakan Desert
0.145283
0.760485
0.106296
76.786566
Tarim Basin deciduous forests and steppe
0.031389
0.733905
0.12035
5.763608
Tian Shan foothill arid steppe
0.171138
0.870284
0.563711
0.972174
Tian Shan montane conifer forests
0.049584
0.864204
0.48319
1.425375
Tian Shan montane steppe and meadows
0.060800
0.87388
0.353977
21.240648
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
54
Figure 14: Column chart of average NDVI values (annual)
By ecoregion boundaries from 2000, 2005, 2010, 2015, 2020 and 2023.
Tables 11-16 as well as figure 14 show the NDVI values of different eco-origins
during the studied years. As can be seen, the ecoregions located in the range of Tianshan and
Altai mountains have the highest NDVI. According to the NDVI values in the mentioned
tables, it can be said that there have been many changes in the vegetation of the region during
different years. In 2000 and 2005, the Altai montane forest and forest steppe ecoregions had
the highest NDVI values of 0.58 and 0.57 among the Xinjiang ecoregions, respectively. The
highest NDVI value in 2010, 2015 and 2020 with values of 0.58, 0.57 and 0.56 belonged to
Altai steppe and semi-desert ecoregion and in 2023 with value of 0.56 belonged to Tian Shan
foothill arid steppe ecoregion.
Estimation of the area and percentage of afforestation and deforestation (2000 to
2022)
Changes in land use and land cover have significant impacts on ecology, the
environment, and politics at both global and regional levels. Therefore, land cover monitoring
and modeling is important in environmental planning and management. Figure 15 shows the
spatial distribution of deforestation and afforestation in Xinjiang Province (Figure 15A),
ecoregions (Fig. 15B) and counties of Xinjiang Province (Fig. 15C) from 2000 to 2022.
Additionally, the table show the changes in vegetation cover in Xinjiang Province
between 2000 and 2022. Specifically, when examining the vegetation in the northern regions
of Xinjiang Province, such as Altai montane forest and forest steppe, Altai steppe and semi-
desert, Emin Valley steppe, and Tian Shan montane steppe and meadows ecoregions, no
significant alterations have been observed. These regions collectively encompass an average
area of approximately 1.665 million km² within Xinjiang Province. According to figure 15 and
tables 17-18, most parts of Xinjiang Province lack vegetation (1,403,680.527 km2). In figure
15, the areas without vegetation are highlighted with a yellow color, mainly representing the
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
55
southern and central regions of Xinjiang Province. These areas include the Taklimakan Desert,
North Tibetan Plateau-Kunlun Mountains alpine desert, Central Tibetan Plateau alpine steppe,
Karakoram-West Tibetan Plateau, as well as alpine steppe and Pamir alpine desert and tundra.
It is worth noting that there has been no significant change in the vegetation status of these
regions. From 2000 to 2022, approximately one percent of the land area in Xinjiang Province
has experienced vegetation degradation.
Figure 15: Visual patterns of forestry and deforestation, their area and percentage from 2000 to
2022 (Classification based on border of Xinjiang Province (A), the borders of eco-regions ( B),
the borders of the counties (C), the pie chart of the percentage changes in the area of
afforestation and deforestation activities and other changes (D).
D
B
B
B
C
B
A
B
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
56
Deforestation and vegetation destruction have primarily occurred within the
ecoregions of Tian Shan montane steppe and meadows, Emin Valley steppe, Tian Shan
montane conifer forests, and Altai alpine meadow and tundra, covering an area of
21,749.34084 km2. However, due to protective measures and efforts towards vegetation
revitalization and development, significant positive ecological changes have taken place in
various parts of Xinjiang Province. Notably, in an area of 59,743.718885 km2, approximately
4% of the total area of Xinjiang Province, predominantly in the Junggar Basin semi-desert,
Tian Shan foothill arid steppe, Tian Shan montane steppe and meadows, Taklimakan Desert,
and North Tibetan Plateau-Kunlun Mountains ecoregions, there has been substantial
restoration of alpine desert vegetation (Fig. 15).
Table 17: Area of revival and development activities and destruction of vegetation
coverage in Xinjiang province from 2000 to 2022.
Vegetation changes
Area (km2)
No Vegetation to No Vegetation
1,403,680.527142
Vegetation to No Vegetation (Deforestation)
21,749.340839
No Vegetation to Vegetation (Afforestation)
59,743.718885
Vegetation to Vegetation
223,681.424438
Figures 16 and 17 as well as table 18 show the changes in vegetation in 2022
compared to 2000. Figure 16 shows the vegetation cover has had a significant growing trend,
with, 37988.61 km2 of the target areas without vegetation cover being reduced. Also, vegetated
areas in 2022 was 283,462.35 km2, which has increased by an average of 37,988.467 km2
compared to 2000.
Figure 16: Visual patterns of vegetation changes based on the presence or absence of
vegetation from 2000 (A) to 2022 (B).
Table 18: The area of vegetation changes in Xinjiang province from 2000 to 2022.
Vegetation changes in 2000
Area (km2)
No Vegetation
1,463,463.10104
Vegetation
245,473.884643
Vegetation Changes in 2022
Area (km2)
No Vegetation
1,425,474.48557
Vegetation
28,3462.351692
A
C
B
B
B
C
B
B
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
57
Figure 17: Models of changes in vegetation area based on the presence or absence of
vegetation in square kilometers from 2000 (A) to 2022 (B).
Figure 18: Visual patterns of afforestation and deforestation based on county boundaries
and ecoregions in square kilometers from 2000 (A) to 2022 (B).
In contrast, deforestation has been observed in the northern and southeastern parts of
Xinjiang Province, which account for only 27% of the total area. These regions correspond to
the ecoregions of Altai montane forest and forest steppe, Altai alpine meadow and tundra,
Emin Valley steppe, Tian Shan montane steppe and meadows, and Tian Shan montane conifer
forests. In general, it can be said that vegetation has grown significantly and environmental
protection measures and policies have had positive ecological effects. The area of afforestation
and deforestation was 59,743.72 and 21,749.34 km2, respectively, which means that the area of
restored areas is on average 37,994.38 km2 more than the destroyed areas (Tab. 19).
A
C
B
B
B
C
B
B
A
C
B
B
B
C
B
B
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
58
In contrast, deforestation has been observed in the northern and southeastern parts of
Xinjiang Province, which account for only 27% of the total area. These regions correspond to
the ecoregions of Altai montane forest and forest steppe, Altai alpine meadow and tundra,
Emin Valley steppe, Tian Shan montane steppe and meadows, and Tian Shan montane conifer
forests. In general, it can be said that vegetation has grown significantly and environmental
protection measures and policies have had positive ecological effects. The area of afforestation
and deforestation was 59,743.72 and 21,749.34 km2, respectively, which means that the area of
restored areas is on average 37,994.38 km2 more than the destroyed areas (Tab. 19).
Table 19: Area of afforestation and deforestation in Xinjiang Province from 2000 to
2022.
Figure 19. Circular graph percentage of afforestation and deforestation areas
by percentage from 2000 to 2022.
Identifying the pattern of spatial distribution of vegetation within the study
border using spatial autocorrelation (confidence level above 90% (Spatial)
autocorrelation, Average Nearest Neighbor)
Tables 20-25 show clustering of the average distribution of NDVI in the ecoregions of
Xinjiang Province during the years 2000-2023 respectively.
All outputs represent the same result. Hence the clusteriness of the mean distribution
of NDVI at the border of the ecoregions of Xinjiang Province for 2000 which is as follows:
Based on the z-score of 4.125841, it is highly improbable that this clustered pattern is due to
random chance, with a probability of less than 1%.
Vegetation changes from 2000 to 2022
Area (km2)
Deforestation
21,749.34084
Afforestation
59,743.71889
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
59
Table 20: Clusteriness of the mean distribution of NDVI of the ecoregions of Xinjiang
Province in 2000.
Moran’s Index:
0/604016
Expected Index:
‒0/062500
Variance:
0/026097
z-score:
4/125841
p-value:
0/000037
Global Moran’s I Summary
Input Feature Class:
Ecoregions of Xinjiang
Input Field:
Export of 2000 (Mean NDVI)
Conceptualization:
Invers Distance
Distance Method:
Euclidean
Row Standardization:
True
Distance Threshold:
628514/2532 m
Weights Matrix File:
None
Selection Set:
False
Figure 20. Morans index for the year 2000 based on the mean NDVI distribution.
How can we interpret Morans index? When the z-score or p-value indicates statistical
significance, a positive value of Morans I index suggests a tendency towards clustering, while
a negative value of Morans I index suggests a tendency towards dispersion. This tool
calculates a z-score and p-value to determine whether the null hypotheses can be rejected. The
expected value of Morans I is 1/(N-1). Values of I that surpass 1/(N-1) indicate positive
spatial autocorrelation, meaning that similar values, whether high or low, tend to cluster
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
60
together in space. Values of I below 1/(N-1) indicate negative spatial autocorrelation, where
neighboring values are dissimilar (Figs. 20-24). Regarding the results in table 21 for the year
2005, the z-score of 4.202093 suggests that there is a less than 1% chance that the observed
clustered pattern is due to random chance. Therefore, the result can be considered reliable. (**)
Table 21: Clusteriness of the mean distribution of NDVI of the ecoregions of Xinjiang
Province in 2005.
Moran’s Index:
0.617828
Expected Index:
‒0.062500
Variance:
0.026212
z-score:
4.202093
p-value:
0.000026
Global Moran’s I Summary
Input Feature Class:
Ecoregions of Xinjiang
Input Field:
Export of 2005 (Mean NDVI)
Conceptualization:
Invers distance
Distance Method:
Euclidean
Row Standardization:
True
Distance Threshold:
628514.2532 m
Weights Matrix File:
None
Selection Set:
False
Figure 21: Moran's index for the year 2005 based on the mean NDVI distribution.
Table 22, which presents the clustering analysis of the average NDVI distribution in
the ecological regions of Xinjiang Province in 2010, indicate the following: With a z-score of
4.176516, the probability of observing such a clustered pattern by random chance is less than
1%.
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
61
Table 22: Clusteriness of the mean distribution of NDVI of the ecoregions of Xinjiang
Province in 2010.
Moran's Index:
0.614116
Expected Index:
-0.062500
Variance:
0.026246
z-score:
4.176516
p-value:
0.000030
Global Moran’s I Summary
Input Feature Class:
Ecoregions of Xinjiang
Input Field:
Export of 2010 (Mean NDVI)
Conceptualization:
Invers Distance
Distance Method:
Euclidean
Row Standardization:
True
Distance Threshold:
628514.2532 Meters
Weights Matrix File:
None
Selection Set:
False
Figure 22: Morans index for the year 2010 based on the mean NDVI distribution.
Table 23, which depicts the clustering analysis of the average NDVI distribution in the
ecological regions of Xinjiang Province in 2015, reveal the following: With a z-score of
4.165062, the probability of this clustered pattern occurring by random chance is less than 1%.
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
62
Table 23: Clusteriness of the mean distribution of NDVI of the ecoregions of Xinjiang
Province in 2015.
Moran’s Index:
0.612921
Expected Index:
-0.062500
Variance:
0.026297
z-score:
4.165062
p-value:
0.000031
Global Moran’s I Summary
Input Feature Class:
Ecoregions of Xinjiang
Input Field:
Export of 2015 (Mean NDVI)
Conceptualization:
Invers Distance
Distance Method:
Euclidean
Row Standardization:
True
Distance Threshold:
628514.2532 Meters
Weights Matrix File:
None
Selection Set:
False
Figure 23: Morans index for the year 2015 based on the mean NDVI distribution.
Table 24, depicting the clustering analysis of the average NDVI distribution in the
ecological regions of Xinjiang Province in 2020, indicate the following: With a z-score of
4.232639, the probability of observing such a clustered pattern by random chance is less than
1%.
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
63
Table 24: Clusteriness of the mean distribution of NDVI of the ecoregions of Xinjiang
Province in 2020
Moran’s Index:
0.622168
Expected Index:
-0.062500
Variance:
0.026166
z-score:
4.232639
p-value:
0.000023
Global Moran's I Summary
Input Feature Class:
Ecoregions of Xinjiang
Input Field:
Export of 2020 (Mean NDVI)
Conceptualization:
Invers Distance
Distance Method:
Euclidean
Row Standardization:
True
Distance Threshold:
628514.2532 Meters
Weights Matrix File:
None
Selection Set:
False
Figure 24: Morans index for the year 2020 based on the mean NDVI distribution.
Lastly, table 25, illustrating the clustering analysis of the average NDVI distribution in
the ecological regions of Xinjiang Province in 2023, suggest the following: With a z-score of
4.001364, the probability of observing such a clustered pattern due to random chance is less
than 1%.
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
64
Table 25: Clusteriness of the mean distribution of NDVI of the ecoregions of Xinjiang
Province in 2023.
Morans Index:
0.584470
Expected Index:
-0.062500
Variance:
0.026143
z-score:
4.001364
p-value:
0.000063
Global Moran's I Summary
Input Feature Class:
Ecoregions of Xinjiang
Input Field:
Export of 2023 (Mean NDVI)
Conceptualization:
Invers Distance
Distance Method:
Euclidean
Row Standardization:
True
Distance Threshold:
628514.2532 Meters
Weights Matrix File:
None
Selection Set:
False
Figure 25: Morans index for the year 2023 based on the mean NDVI distribution.
We found that z-score or p-value expressed statistical significance, as well as the
Moran i-positive index value indicating a clustering tendency, while the Moran I-negative
index value indicates a dispersion tendency. The tool utilizes a z-score and p-value to
determine the possibility of rejecting null assumptions.
The expected value of Morans I is -1/(N-1). I values exceeding 1/(N-1) indicate
positive spatial correlation, where similar values, whether high or low, tend to cluster together
in space. Conversely, I values below 1/(N-1) represent negative spatial correlation, indicating
that neighboring values differ from each other (Fig. 25).
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
65
Estimation of hot spots and cold spots of vegetation within the border of the study
area (confidence level > 90%) using hot spot analysis, clusters and outliers.
According to Moran outputs, for all the years studied 2000 to 2023, NDVI spatial
distribution patterns were calculated in clusters (Figs. 26A,B).
One point that has caused this distribution to be clustered is how the ecoregions are
distributed. If we look closely, the highest NDVI average is more common in the northern and
northwestern parts, and its actually more concentrated, and that shows the cluster distribution
pattern, and its been going the same way all year. Itis the same locally.
Hot spots and cold spots of NDVI, were calculated with confidence percentages of 90,
95 and 99 percent, which certainly increases the percentage of confidence. The following
patterns show the north and northwest parts are definitely considered hot spots because there is
the highest percentage of average NDVI in these areas (Figs. 26A,B).
Ecoregions include cold spots as well as not significant are also clearly marked which
make up the southern and central parts of the ecoregions of Xinjiang Province.
The high high cluster relates to the hot spots that are available in the Tian Shan
ecoregions and northwestern Xinjiang Province, and the high low outlier, which is red, and the
low low outlier, which are blue, which are described as inconveniences. Low low cluster are
clusters that express cold spots that spread in the southern parts, and clusters that have a very
low confidence percentage are not significant, which are shown white (Figs. 26A,B).
Figure 26: Spatial patterns of hot and cold spots changes in vegetation based on the boundaries
of the ecoregions of Xinjiang Province (Clusters outliers is A, Hot Spot and Cold Spot is B).
Identification of factors affecting the distribution of vegetation within the study
border (land use land cover, LST, rainfall, slope, elevation) by OLS
Ordinary Least Squares regression (OLS) is a widely used method for estimating the
coefficient of linear regression equations. It helps describe the association between a
dependent variable and one or more quantitative independent variables, whether it is a simple
linear regression or multiple linear regression. In this calculation, NDVI is the dependent
variable of the research and LULC, LST, Slope, Rainfall and Elevation are considered
independent variables. The results are shown in figure 27.
A
C
B
B
B
C
B
B
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
66
Figure 27: Spatial patterns of effective land use land cover (A), rainfall (B), LST (C),
slope (D), elevation (E) drivers on vegetation distribution in Xinjiang Province.
Table 26: Summary of OLS results between NDVI and LULC model.
Variable
Coefficient
[A]
Stderror
T-Statistic
Probability
[B]
Robust_SE
Robust _T
Robust _Pr
[B]
Intercept
0/613850
0/129693
4/733110
0/000268*
0/113681
5/399769
0/000074*
LULC
model
0/544243
0/080852
6/731360
0/000007*
0/069100
7/876157
0/000001*
A
C
B
B
B
C
B
B
C
C
B
B
D
C
B
B
F
C
B
B
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
67
In table 26, the column for Coefficient [a] shows the extent and manner of the
relationship between LULC and NDVI. This relationship is strong (0.54) and it should have
been predicted like this, and the relationship is also positive. For this connection to be
meaningful, refer to the Probability [b] columns and Robust_Pr [b], we see that the numbers
are less than 0.03 and are especially starred, the stars of the results indicate the significance of
this relationship and we can calculate that LULC is a driver that has a complete impact on
NDVI and has a positive relationship.
Table 27: Summary of OLS results between NDVI and LST model.
Table 27 shows the summary of OLS results between NDVI and LST models dedicate
that they have a negative relationship with each other (Coefficient [a]) and this value is equal
to 0.003002, in fact, areas covered with dense vegetation have low LST values and this
relationship also argues this point. Unfortunately, the relationship between LST and NDVI is
not very significant according to the Probability column [b] and Robust_Pr [b].
Table 28: Summary of OLS results between NDVI and rainfall model.
Table 28 shows there is a statistically significant relationship between NDVI and
rainfall. However, the impact of this relationship is not as strong as the influence of the LULC
factor.
Table 29: Summary of OLS results between NDVI and elevation model.
Focusing on the column Coefficient [a] and Probability [b] in table 29, the results
show a meaningful and inverse relationship between NDVI and DEM with confidence level
and value of 0.00006, although this relationship is not very strong, but this relationship can be
trusted. It is interpreted as having more vegetation in areas where there is less height.
Variable
Coefficient
[A]
Stderror
T-Statistic
Probability
[B]
Robust_SE
Robust _T
Robust _Pr
[B]
Intercept
1/104029
1/661837
0/664342
0/516554
1/564522
0/705665
0/491210
LST
model
‒0/003002
0/005791
‒0/518356
0/611774
0/005384
‒0/557562
0/585370
Variable
Coefficient
[A]
Stderror
T-Statistic
Probability
[B]
Robust_SE
Robust _T
Robust _Pr
[B]
Intercept
‒0/033441
0/107761
‒0/310326
0/760585
0/052291
‒0/639516
0/532133
Rainfall
model
0/001301
0/000470
2/768490
0/014340*
0/000370
3/518273
0/003106*
Variable
Coefficient
[A]
Stderror
T-Statistic
Probability
[B]
Robust_SE
Robust _T
Robust _Pr
[B]
Intercept
0/414286
0/078734
5/261874
0/000096*
0/085896
4/823133
0/000225*
Eleva-
tion
model
(DEM)
‒0/000064
0/000025
‒2/561612
0/021689*
0/000019
‒3/423192
0/003775*
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
68
Table 30: Summary of OLS results between NDVI and slope model.
The relationship between NDVI and slope is inverse but not a significant relationship
(Tab. 30). Among the results, LULC factor is the most effective factor affecting NDVI and
after LULC factor, the LST factor is not very significant, although it has indirect relationship
and after LULC it can be the most effective factor on vegetation.
The dominant factor influencing NDVI is land use and land cover (LULC) driver,
which has the strongest impact. Following the LULC driver, land surface temperature (LST)
factor can also be considered significant, although its influence is not as pronounced. Despite
having an indirect relationship, the LST factor becomes the most influential factor in the
distribution and growth of vegetation within the counties and boundaries ecoregions of
Xinjiang Province after the LULC driver.
DISCUSION
Many research studies claims that vegetation is an important part of the Earths
ecosystem and biosphere, which, in addition to affecting biogeochemical cycles, carbon
balance regulation and maintaining climate stability, water and soil protection, is responsible
for an important part of completing the energy cycle as well. On the other hand, monitoring
and modeling the time and place changes in vegetation cover and predicting the processes
affecting it are considered an integral part of ecological resource and reserve management.
Recently, understanding dynamics of vegetation patterns, mainly caused by various natural and
human factors such as climate and soil changes, land use changes, wildfires, and mining
measures, has become widespread. Therefore, this article attempts to revise a model of changes
in vegetation and natural and human drivers affecting the temporal spatial distribution process
and dynamics of vegetation at the scale of the counties and ecoregions in Xinjiang Province
from 2000 to present. Satellite platforms such as Modis/Terra, CHIRPS, Sentinel 1-2 and
GTOPO30 platforms were used to extract the NDVI, LST, precipitation (rainfall), global land
cover product and DEM data respectively. NDVI was inserted into the dependent variable and
effective drivers were also listed as LULC, LST, precipitation (rainfall), elevation and slope.
Methods of calculating and zoning vegetation, predicting mean, minimal and maximum
values, temporal and spatial patterns of vegetation, distribution and changes in their area in
km2, kernel density and point density, estimation of afforestation and deforestation, spatial
distribution patterns with confident level, identification of hot and cold spots and prediction of
drivers affecting the development and distribution of vegetation were considered the most
important methodological elements of the study. Terrain factors, such as altitude, play a role in
influencing vegetation coverage. Changes in effective accumulated temperature and soil
moisture caused by altitude variations can impact vegetation growth (Liu et al., 2023). The
output findings from this research argue that; interpretation of the digital elevation model
(DEM) and slope argues the study boundaries have expanded between altitude levels 152 m
and 7,553 m and fluctuations of 0 to 356 degrees that, in conjunction with patterns of average
annual precipitation 0 to 704 mm and average annual LST 250 to 311 K, have created a wide
range of herbaceous plant species, shrubs, and broadleaf and coniferous trees. In a study
Variable
Coefficient
[A]
Stderror
T-Statistic
Probability
[B]
Robust_SE
Robust _T
Robust _Pr
[B]
Intercept
0/253870
0/083998
3/022338
0/008573*
0/069781
3/638121
0/002430*
Slope
model
‒0/001246
0/007745
‒0/160889
0/874325
0/006483
‒0/192223
0/850144
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
69
conducted by Wang et al. (2022) in Southeast Tibet, it was observed that vegetation patterns
exhibit noticeable spatial variations with changes in elevation. The impact of slope and aspect
on NDVI change was found to be relatively less significant compared to elevation. Similarly,
Xianghong et al. (2015) found in their research in China that elevation plays a crucial role in
determining local-scale land use spatial patterns. Areas below 400 m in elevation tend to have
higher land use intensity, and it decreases significantly as the elevation increases. LULC driver
alignment also assessed that the highest amount of area is related to the bare area class, which
is 139,433.7 km2, which occupies almost 61.49% of the study area, and the lowest amount of
area is related to the lichens and mosses class, which accounts for 41.8 km2, or 0.018% of the
total studied area. The grassland and crop land occupy 21.2% and 7.17%, and were calibrated
in second and third place after bare land respectively.
When studying the spatio-temporal dynamics of vegetation, the Normalized Difference
Vegetation Index (NDVI) serves as a vital indicator to assess the state and distribution patterns
of plant growth at both global and regional scales. It also enables the observation of long-term
vegetation changes over time (Zhuang et al., 2020; Gadiga, 2015). We utilized the NDVI of
natural vegetation to track vegetation dynamics. Our analysis revealed that the highest NDVI
value, reaching 0.933, was observed in July for 2020, indicating vigorous vegetation growth.
This predicted value represents areas with dense vegetation which were occupied western,
northwestern and north regions likewise Tian Shan and Altai ecoregions, whereas findings
represented the lowest NDVI value of 0.199 corresponded to 2010, which has the presence of
abundant sand, rocks such as Taklimakan Desert in the central, east and southern ecoregions in
fact.
2015 had the highest area of the regions, with NDVI value of more than 0.5,
(149,450.6 km2), and 2000 also experienced the largest area corresponding to areas with a
NDVI value ofless than 0.5 (1614503.6 km2). Therefore, 2015 has experienced the highest
level of dense vegetation, and 2000 is also covered with sporadic and vast vegetation
community.
Overall, the findings argued that areas where NDVI value has decreased to less than
0.5, their area also declined from 2000 to 2023, or the area NDVI value more than 0.5
increased, which indicates the existence of suitable conditions for the development of
vegetation as well as the correct management of the areas studied by humans. The rising
number of positive NDVI values indicates a growth in the extent of green vegetation.
Interpretation of visual patterns and robust statistical parameters proved that
afforestation activities have increased by 4% from 2000 to 2022 that covers an area of
59,743.71 km2 in all the studied boundaries, which is very commendable. Areas that had a
diversity of vegetation cover have also benefited from the process of restoration, development
and proper management of vegetation cover during the same period, with the area reaching
13%, which occupies approximately 223,681.42 km2 of the total studied area in Xinjiang
Province.
Countries and international organizations worldwide are slowly but surely
acknowledging the threats and risks accompanying the deteriorating and declining vegetation
coverage and are implementing various complex measures to safeguard vegetation (Wang et
al., 2023). Over the past five decades, the Chinese government has actively promoted the
implementation of multiple environmental protection initiatives, resulting in significant
positive changes in vegetation cover. Noteworthy projects include the Grain for Green Project,
the establishment of the Three-North Shelterbelt Program, and the conservation of natural
forest resources.
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
70
Amendments and enhancements to the Forest Law of the Peoples Republic of
China have been implemented alongside ecological programs aimed at curbing deforestation
and increasing forest area. Since 2000, one such ecological restoration initiative is the
Returning Cultivated Land to Forests and Grasses Project (Wang et al., 2023; He et al.,
2023; Han et al., 2022; Yin et al., 2018). Based on the results of various studies, this set of
measures has had an effective role in the growth of vegetation, in this regard, the results of the
studies of Zhuang et al. (2020), in Xinjiang and Zhong Bao et al. (2008), Xuemei et al. (2016),
Karnieli et al. (2014), Shan et al. (2022), He et al. (2023) in other parts of China, indicates that
the vegetation cover has an upward trend.
Also, cluster distribution and hot and cold spot processors calculated at 90, 95 and
99 percent confidence levels that the borders of the ecoregions in the northern
and northwestern regions of Xinjiang Province certainly formed the high percentage of
the average value of the NDVI. The central and southern ecoregions were found to have
minor values of NDVI. The results of the OLS estimate show a strong link between NDVI
and LULC valued at 0.54, which is significant. Vegetation is very sensitive to climate
change (Wang et al., 2023) hence, it becomes vital to examine the spatio-temporal
vegetation changes and how it responds to long-term climate changes. The findings reveal
a weak relationship between NDVI and LST, with a value of 0.003. This lack of
significance is evident in both the probability and robust probability analyses. Notably,
areas characterized by dense vegetation exhibit lower LST values, which aligns with
this relationship. Similarly, a study of Gang et al. (2016) in Central Asia yielded similar
results, indicating a weak negative correlation between annual NDVI and temperature.
The relationship between precipitation and NDVI was significant with a value of 0.001,
although precipitation drivers are not as effective as LULC. In this particular context,
various studies conducted by Liu et al. (2006) and Sun et al. (2001) in the Yellow River
Basin, China, Zhong Bao et al. (2008) in the Loess Plateau, China, and Xuemei et al. (2016)
in the Hexi region of Northwest China have revealed a positive correlation between NDVI
and precipitation. Precipitation is identified as the primary driver of vegetation growth,
while temperature plays a moderating role in this process (Li et al., 2022). Additionally, it
has been observed that there are time lags in the response of NDVI to changes in precipitation
(Gang et al., 2016).
There is a meaningful and inverse relationship between NDVI and DEM
with confidence level and value of 0.00006, showing that vegetation is more present at
lower altitudes. The OLS model also showed a negative relationship between NDVI and
slope with the confidence level 0.0012 which was not very noticeable. Henceforth, the
LULC driver is the most effective factor affecting NDVI, and the LST driver can be
practically emphasized after the LULC, but it is not strongly meaningful, although it has
an indirect relationship, and after the LULC driver, LST can be the most effective driver in
the distribution and development of vegetation within the counties and ecoregions
boundaries of Xinjiang Province. As a consequence, NDVI is an important vegetation
indicator that reflects the state of the regional climate and the environment drivers, which
are significantly affected by precipitation, temperature, topographic complications, and
human activities.
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
71
CONCLUSIONS
Vegetation holds great significance within the ecosystem, and any form of destruction
can have detrimental consequences for the environment. Hence, it is crucial to actively
monitor, conserve, and sustainably utilize this vital resource. The dynamics of vegetation
patterns involve complex processes influenced by a range of biological and human-induced
factors, including climate change, changes in land use, and ecological engineering practices.
The findings of this study indicate that the dynamics of vegetation cover are impacted by
various environmental factors, notably land use and land cover (LULC), land surface
temperature (LST), rainfall, elevation, and slope. Moreover, the implementation of diverse
vegetation protection policies has proven effective in rejuvenating and fostering vegetation
growth. Utilizing the Normalized Difference Vegetation Index (NDVI) to examine vegetation
dynamics facilitates a better comprehension of how vegetation responds to different natural
and human variables.
ACKNOWLEDGEMENTS
This research was supported by Xinjiang Institute of Ecology and Geography, Chinese
Academy of Sciences, Urumqi 830011, China, Tianchi Talents Project of Xinjiang
(E3350107); National Natural Science Foundation of ChinaNo.42107084; Key Research
and Development Program of Xinjiang 2022B01032-4. Professor YU Ruide is also given
special gratitude that provided insight, support and expertise that greatly assisted the research.
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
72
REFERENCES
1.
Cao K., Gao J., 2022 Assessment of climatic conditions for tourism in Xinjiang, China, Open
Geosciences, 14, 382-392.
2.
Duan H., Xue X., Wang T., Kang W., Liao J and Liu, S., 2021 Spatial and temporal
differences in alpine meadow, alpine steppe and all vegetation of the Qinghai-Tibetan Plateau
and their responses to climate change, Remote Sensing, 13, 669, DOI: 10.3390/rs13040669.
3.
Elsu C. A., Ramesh K. V. and Sridevi H., 2017 Quantification and understanding the observed
changes in land cover patterns in Bangalore, International Journal of Civil Engineering and
Technology (IJCIET), 8, 4, 597-603.
4.
Gadiga B. L., 2015 Monitoring the spatio-temporal dynamics of vegetation cover in Mubi
Region, Adamawa State, Nigeria, Journal of Geographic Information System, 7, 598-606, DOI:
10.4236/jgis.2015.76048.
5.
Gang Y., Zengyun H., Xi C. and Tashpolat T., 2016 Vegetation dynamics and its response to
climate change in Central Asia, Journal of Arid Land, 8, 3, 375-388, DOI: 10.1007/s40333-016-
0043-6.
6
Han Z., Wu Q., Lai R, Soomro S., Hou D and Hu C., 2022 Spatio-temporal variations of
vegetation cover and its influence on surface air temperature change over the Yellow River
Basin China, Journal of Water and Climate Change, 13, 9, 3239, DOI: 10.2166/wcc.2022.037.
7.
He Z., Yue T., Chen Y., Mu W., Xi M. and Qin F., 2023 Analysis of spatial and temporal
changes in vegetation cover and driving forces in the Yan River Basin, Loess Plateau, Remote
Sensing, 15, 4240, DOI: 10.3390/rs15174240.
8.
Hu Y. and Nacun, B., 2018 An analysis of land-use change and grassland degradation from a
policy perspective in inner Mongolia, 1990-2015, Sustainability, 10, 4048, DPO:
10.3390/su10114048.
9.
Hu, Y and Hu, Y., 2019 Land cover changes and their driving mechanisms in Central Asia
from 2001 to 2017 supported by Google Earth Engine, Remote Sensing, 11, 554, DOI:
10.3390/rs11050554.
10.
Jiang M., Tian S., Zheng Z., Zhan Q and He, Y., 2017 Human activity influences on
vegetation cover changes in Beijing, China, from 2000 to 2015, Remote Sensing, 9, 271, DOI:
10.3390/rs9030271.
11.
Jiapaer G., Liang S., Yi Q. and Liu J., 2015 Vegetation dynamics and responses to recent
climate change in Xinjiang using leaf area index as an indicator, Ecological indicators, 58, 64-
76.
12.
Karnieli A., Qin Z., Wu B., Panov N and Yan F., 2014 Spatio-temporal dynamics of land-use
and land-cover in the Mu Us Sandy Land, China, Using the change vector analysis technique,
Remote Sensing, 6, 9316-9339; DOI: 10.3390/rs6109316.
13.
Li M., Yan Q., Li G., Yi M and Li J., 2022 Spatio-temporal changes of vegetation cover and
its influencing factors in northeast China from 2000 to 2021, Remote Sensing, 14, 5720, DOI:
10.3390/rs14225720.
14.
Li Z., Deng X., Yin F and Yang C., 2015 Analysis of climate and land use changes impacts
on land degradation in the North China Plain, Hindawi Publishing Corporation Advances in
Meteorology, 2015, 976370, 1-11, DOI: 10.1155/2015/976370.
15.
Liu Y., Haohong H. H., Meng L., Liu M., Wu Z., Liu T and Labatt D., 2023 Spatial temporal
evolution of vegetation coverage and its relationship with terrain and human factors in the
upper reaches of Ganjiang River Basin, China, Frontiers in Earth Science, 10:1043403, DOI:
10.3389/feart.2022.1043403.
16.
Liu X., Hong Z., Zhou L., Liu J., Guo X. and Li S., 2022 Past and future of land use change
in the middle reaches of the Yellow River Basin in China, Research Square, DOI: DOI:
10.21203/rs.3.rs-2144585/v1.
17.
Liu L. Z. and Xiao J. F., 2006 Spatial-temporal correlations of NDVI with precipitation and
temperature in Yellow river basin, Chinese Journal of Ecology, 25, 5, 477-481. (in
Chinese)
Transylv. Rev. Syst. Ecol. Res. 26.1 (2024), "The Wetlands Diversity"
73
18.
Lovland T. and Merchant J., 2004 Ecoregions and ecoregionalization: geographical and
ecological perspectives, Environmental Management, 34, 1-13, DOI: 10.1007/s00267-003-
5181-x.
19.
Luo M., Liu T., Meng F., Duan Y., Bao A., Xing W., Feng X., De Maeyer P. and Frankl A.,
2019 Identifying climate change impacts on water resources in Xinjiang, China, Science of
the Total Environment, 676, 613-626.
20.
Meng X., Gao X., Li S and Lei J., 2020 Spatial and temporal characteristics of vegetation
NDVI changes and the driving forces in Mongolia during 1982-2015, Remote Sensing, 12, 603,
DOI: 10.3390/rs12040603.
21.
Miao L., Feng Z., Sun Z., Moore J and Cui X., 2016 China’s land-use changes during the past
300 years: a historical perspective, International Journal of Environtal Research and Public
Health, 13, 847, DOI: 10.3390/ijerph13090847, Posted on Authorea 17 Aug (2020): The
copyright holder is the author/funder, All rights reserved, No reuse without permission, DOI:
10.22541/au.159769328.83108985, This a preprint and has not been peer reviewed, Data may
be preliminary.
22.
Sun R., Liu C. M. and Zhu Q. J., 2001 Relationship between the fractional vegetation cover
change and rainfall in the Yellow River basin, Acta Geografica Sinica, 56, (6), 667-672. (in
Chinese)
23.
Shan Y., Dai X., Li W., Yang Z., Wang Y., Qu G., Liu W., Ren J., Li C., Liang S. and Zeng B.,
2022 Detecting spatial-temporal changes of urban environment quality by remote sensing-
based ecological indices: a case study in Panzhihua City, Sichuan Province, China, Remote
Sensing, 14, 4137, DOI: 10.3390/rs14174137.
24.
Tahir A. A., Quazi K. H., Sana I., Maleeha B., Hira Jannat B. and Hira J., 2019 Investigative
spatial distribution and modelling of existing and future urban land changes and its impact on
urbanization and economy, Remote Sensing, 11, 2, 105.
25.
Wang C., Wang J., Naudiyal N., Wu N., Cui X., Wei Y. and Chen Q., 2022 Multiple effects
of topographic factors on spatio-temporal variations of vegetation patterns in the three parallel
rivers region, Southeast Tibet, Remote Sensining, 14, 1, 151, DOI: 10.3390/rs14010151.
26.
Wang M., Wang Y., Li Z and Zhang H., 2023 Analysis of spatial-temporal changes and
driving factors of vegetation coverage in Jiamusi City, Forests, 14, 1902, DOI:
10.3390/f14091902
27.
Xianfeng L., Jinshui Z., Xiufang Z., Pan Yaozhong P., Yanxu L., Donghai Z. and Lin Z. L.,
2014 Spatiotemporal changes in vegetation coverage and its driving factors in the Three-
River headwaters region during 2000-2011, Journal of Geographical Sciences, 2014, 24, 2,
288-302. DOI: 10.1007/s11442-014-1088-0.
28.
Xianghong D., Xiyong H., Yuandong W. and Li W., 2015 Spatial-temporal characteristics of
land use intensity of coastal zone in China during 2000-2010, China Geographcal Sciences, 25,
1, 51-61, DOI: 10.1007/s11769-014-0707-0.
29.
Xuemei Y., Shizeng L., Taibao Y., Xianying X., Caizhou K., Jinnian T., Huaidong W.,
Ghebrezgabher M. and Zhiqi L., 2016 Spatial-temporal dynamics of desert vegetation and its
responses to climatic variations over the last three decades: a case study of Hexi region in
Northwest China, Journal Arid Land, 8, 4, 556-568, DOI: 10.1007/s40333-016-0046.
30.
Xiaoa H. and Weng Q., 2007 The impact of land use and land cover changes on land surface
temperature in a karst area of China, Journal of Environmental Management, 85, 245-257,
DOI: 10.1016/j.jenvman.2006.07.016.
31.
Xue J., Wang Y., Teng H., Wang N., Li D., Peng J., Biswas A. and Shi Z., 2021 Dynamics of
vegetation greenness and its response to climate change in Xinjiang over the past two decades,
Remote Sensing, 13, 4063, DOI: 10.3390/rs13204063.
32.
Yao B., Ma L., Si H., Li S., Gong X. and Wang X., 2023 Spatial pattern of changing
vegetation dynamics and its driving factors across the Yangtze River basin in chongqing: a
geodetector-based study, Land, 12, 269.
Shobairi S. O. R. et al. Modeling process of the changes vegetation cover in in Xianjiang (China) (27 ~ 74)
74
33.
Yin H., Pflugmacher D., Lib A., Lic Z. and Hostert P., 2018 Land use and land cover change
in Inner Mongolia understanding the effects of Chinas re-vegetation programs, Remote
Sensing of Environment, 204, 918-930, DOI: 10.1016/j.rse.2017.08.030.
34.
Yu H., Bian Z., Mu S., Yuan J. and Chen F., 2020 Effects of climate change on land cover
change and vegetation dynamics in Xinjiang, China, International Journal of Environmental
Research and Public Health, 17, 1-24.
35.
Zhang X., She D., Huang X. and Wang G., 2020 Identifying the effects of land use changes
and check dams on sediment yield in a watershed of the Loess Plateau, China, Authorea, 17,
DOI: 10.22541/au.159769328.83108985.
36.
Zhao H., Li X., Ezizi M. and Yao J., 2022 Changes in the characteristics of dry-wet periods
in Xinjiang, China based on the SPEI index, Atmósfera, 35, 3, 483-496.
37.
Zhu Z., Zhang Z ., Zhao X., Zuo L and Wang X., 2022 Characteristics of land use change in
China before and after 2000, Sustainability, 14, 14623, DOI: 10.3390/su142114623.
38.
Zhuang Q., Wu S., Feng X. and Niu Y., 2020 Analysis and prediction of vegetation dynamics
under the background of climate change in Xinjiang, China, PeerJ, 8:e8282, DOI:
10.7717/peerj.8282.
39.
Zhong B. X., Jiong X. X. and Zheng W. Z., 2008 Spatiotemporal variations of vegetation
cover on the Chinese Loess Plateau (1981-2006): Impacts of climate changes and human
activities, Science in China Series D: Earth Sciences, 51, 1, 67-78.
40.
* https://www.usgs.gov/special-topics/remote-sensing-phenology/science/ndvi-foundation-
remote-sensing-phenology
41.
** https://pro.arcgis.com/en/pro-app/latest/tool-reference/spatial-statistics/spatial-
autocorrelation.htm#:~:text=When%20the%20z%2Dscore%20or,
can%20reject%20the%20null%20hypotheses
42.
*** https://earthobservatory.nasa.gov/features/MeasuringVegetation
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
This study of vegetation coverage in Jiamusi City provides theoretical support for local urban development, land use, and ecological environmental protection. Based on the land cover data and Landsat remote sensing image data from 2000 to 2020, the vegetation cover and land use data of Jiamusi City were extracted. The study includes the following aspects: (1) an analysis of the spatio-temporal changes in vegetation coverage; (2) analysis of the land use situation in Jiamusi City; and (3) investigation of the impact of natural and human factors on vegetation coverage in Jiamusi City using the Geodetector model. The results show that (1) over the past 20 years, the vegetation coverage of Jiamusi has shown a decreasing trend, declining from 25.22% in 2000 to 17.13% in 2020, representing a decrease of 32%. In terms of spatial distribution, the areas of Fuyuan City and Tongjiang City have experienced more significant decreases in vegetation coverage, decreasing by 73.6% and 54.0%, respectively. (2) The land use pattern of Jiamusi City has undergone significant changes during the study period; except for paddy fields (PF), unused land (UL), and construction land (CL), the areas of all the land categories have decreased. The ranking of the single land use dynamic degree in terms of magnitude is as follows: PF, UL, CL, dry farmland (DF), vegetation coverage land (VCL), and wetland (WET). (3) The changes in vegetation coverage were influenced by both natural and human activities and, according to the Geodetector results, the main influencing factors were CL and DF. The key findings of this study emphasize the need for comprehensive land use planning and ecological environmental protection that focus on sustainable development and conservation practices, and lay the groundwork for future ecosystem management and urban planning efforts in Jiamusi City.
Article
Full-text available
The Yan River Basin of the Loess Plateau is a key region for ensuring the environmental protection and sustainable development of the Yellow River Basin. Therefore, it is essential to identify how vegetation cover has changed and determine the factors that have driven these changes. In this study, we applied a three-dimensional vegetation cover model to examine the spatiotemporal variation characteristics of vegetation cover at the watershed scale in the Yan River Basin from 2001 to 2020 and forecast future trends. Subsequently, the driving forces of fractional vegetation cover (FVC) change were quantified based on meteorological, surface, and anthropogenic factors to explore the common driving relationships among these factors. (1) The accuracy of 3DFVC is better than that of FVC in the Yanhe River Basin, where the terrain is complex. (2) The temporal change trends indicated that the vegetation cover in the Yan River Basin significantly recovered and the basin FVC increased rapidly from 2001 to 2013 (S = 0.0152/a, p < 0.01) and increased gradually from 2013 to 2020 (S = 0.0015/a). The main reason for the increase in vegetation cover was the enhanced growth of medium FVC. (3) The vegetation spatial distribution showed that the FVC values varied substantially from north to south, indicating spatial heterogeneity, and 83.9% of the area presented a trend of increasing vegetation. Furthermore, vegetation cover was predicted to improve in the future. (4) The spatial heterogeneity of FVC was mainly influenced by relative humidity and rainfall, and the spatial variations in FVC were mainly determined by climate factors. Land use and cover change variations, which are influenced by human activities, represent major factors underlying the observed spatial heterogeneity. Most interactions between driving factors showed two-way enhancement or non-linear enhancement, with relative humidity and land use patterns presenting the strongest explanatory power. This study provides a scientific basis for vegetation conservation in the Yan River Basin and contributes theoretical support for decision-making regarding ecological environmental protection in the Loess Plateau and sustainable development in the Yellow River Basin.
Article
Full-text available
Revealing the spatial dynamics of vegetation change in Chongqing and their driving mechanisms is of major value to regional ecological management and conservation. Using several data sets, including the SPOT Normalized Difference Vegetation Index (NDVI), meteorological, soil, digital elevation model (DEM), human population density and others, combined with trend analysis, stability analysis, and geographic detectors, we studied the pattern of temporal and spatial variation in the NDVI and its stability across Chongqing from 2000 to 2019, and quantitatively analyzed the relative contribution of 18 drivers (natural or human variables) that could influence vegetation dynamics. Over the 20-year period, we found that Chongqing region’s NDVI had an annual average value of 0.78, and is greater than 0.7 for 93.52% of its total area. Overall, the NDVI increased at a rate of 0.05/10 yr, with 81.67% of the areas undergoing significant expansion, primarily in the metropolitan areas of Chongqing’s Three Gorges Reservoir Area (TGR) and Wuling Mountain Area (WMA). The main factors influencing vegetation change were human activities, climate, and topography, for which the most influential variables respectively were night light brightness (NLB, 51.9%), annual average air temperature (TEM, 47%), and elevation (ELE, 44.4%). Furthermore, we found that interactions between differing types of factors were stronger than those arising between similar ones; of all pairwise interaction types tested, 92.9% of them were characterized by two-factor enhancement. The three most powerful interactions detected were those for NLB ∩ TEM (62.7%), NLB ∩ annual average atmospheric pressure (PRS, 62.7%), and NLB ∩ ELE (61.9%). Further, we identified the most appropriate kind or range of key elements shaping vegetation development and dynamics. Altogether, our findings can serve as a timely scientific foundation for developing a vegetative resource management strategy for the Yangtze River basin that duly takes into account local climate, terrain, and human activity.
Article
Full-text available
Vegetation coverage is an important indicator for evaluating regional environmental quality. Based on MODIS NDVI and DEM data collected for the upper reaches of the Ganjiang River Basin, China, this study used trend analysis, coefficient of variation, Hurst index, and linear regression to analyze the temporal and spatial evolution of vegetation coverage and its relationship with terrain factors in the basin during the years 2000–2020. The vegetation coverage in the study area showed a fluctuating increasing trend at a rate of 5%/10y, and an increasing trend with increasing elevation. The maximum vegetation coverage was identified in the elevation zone of 750–1,000 m, with an average of 83.54%. Vegetation coverage also showed an increasing trend with increasing slope. The maximum vegetation coverage was up to 82.22% in the slope zone of ≥25°. There were no significant differences among the distributions of vegetation coverage in different aspects because the terrain in the study area is not rugged enough to form barriers against sunlight. The vegetation coverage was relatively stable in the study area, with an average coefficient of variation of 14.8%. Hurst analysis showed that the anti-sustainability effect of vegetation change was stronger than that of sustainability, and weak anti-sustainability was dominant. The effects of human activities mainly concentrated in the areas of low elevation and small slopes less than 2°where cities and towns are located. The findings can provide a scientific basis for the management of regional ecosystems in the future.
Article
Full-text available
The foundation of study on regional environmental carrying capacity is the detection of vegetation changes. A case of Northeast China, we, with the support of normalized difference vegetation index (NDVI) of MOD13A3 (MOD13A3-NDVI), use a three-dimensional vegetation cover model (3DFVC) to acquire vegetation cover from 2000 to 2021. Vegetation trends are then monitored by the spatio-temporal analysis models including the empirical orthogonal function (EOF), the Sen’s slope (Sen), the Mann-Kendall test (MK) and the Hurst index (Hurst). Additionally, we, through the multi-scale geographically weighted regression model (MGWR), explore the spatial heterogeneity of vegetation response to its influencing factors. On the basis of this, it is by introducing the structural equation model (SEM) that we figure out the mechanisms of vegetation response to climate and human activity. The main results are as follows: (1) Compared with the dimidiate pixel model (FVC), 3DFVC, to some extent, weaken the influence of terrain on vegetation cover extraction with a good applicability. (2) From 2000 to 2021, the average annual vegetation cover has a fluctuating upward trend (0.03 · 22a−1, p < 0.05), and spatially vegetation cover is lower in the west and higher in the east with a strong climatic zoning feature. In general, vegetation cover is relatively stable, only 7.08% of the vegetation area with a trend of significant change. (3) In terms of EOF (EOF1+EOF2), EOF1 has a strong spatial heterogeneity but EOF2 has a strong temporal heterogeneity. As for the Hurst index, its mean value, with an anti-persistence feature, is 0.451, illustrating that vegetation is at some risk of degradation in future. (4) MGWR is slightly better than GWR. Vegetation growth is more influenced by the climate (precipitation and temperature) or human activity and less by the terrain or soil. Besides, precipitation plays a leading role on vegetation growth, while temperature plays a moderating role on vegetation growth. What is more, precipitation, on different temperature conditions, shows a different effect on vegetation growth.
Article
Full-text available
China, with notable population blooming and economic development in the last decades, has experienced profound land-use changes, which, in turn, dramatically impacted the regional, even global and environment system. However, characteristics of land-use changes in China have not yet been well addressed, especially around the year 2000 when a series of land policies were put forward, such as the project of “returning farmland to forest”. To fill this gap, this paper investigated the temporal and spatial patterns of land use changes in China for the period from 1987 to 2010, by taking advantage of the continually updated China Land Use Database developed from remote sensing images. The land-use dynamic matrix, zonal model, and transition matrix were employed to characterize land-use change patterns for four time intervals (1987–1995, 1995–2000, 2000–2005, and 2005–2010) on the dimensions of conversion and modification. Results showed that land-use change affected 4 × 105 km2 (4.5%) of the total landscape in China for more than the past twenty years. Of the six land-use types, built-up land experienced the largest net increase by almost 30% (52,434 km2), with the rate of expansion accelerating after 1995. The area of cropland increased before 2000 and declined afterwards, ending with a net increase in 14,280 km2, approximately 1% of its original area. The loss in the eastern coastal region is attributed mainly to built-up land expansion, while the gain in northern China, with the price of grassland and woodland shrinking, reshaped the cropland distribution in China. The area of woodland decreased slightly by 7880 km2 without a clear pattern over time. The modification of woodland indicated an intensive forest management in terms of planting fast-growing trees in the south of China. Grassland continues to shrink at a decreasing rate, and the modification of grassland shows a tendency of transformation from sparse grassland into a dense one in the 21st century. Trade-offs among demands on food security, economic development, and environment protection forced and shaped the contemporary land-use change in China. These results contribute to understanding the trends and causes of land use change in China, which could provide underpinning knowledge for assessing environmental change, and provide insights on future land planning.
Preprint
Full-text available
When looking at maps of land use land cover (LULC) change, often derived from satellite imagery, China's increasing urban construction and decreasing cultivated land area can be identified. Using the middle reaches of the Yellow River Basin in China as the case study, the present work focuses on (i) identifying historical LULC change in the period 1995–2020; (ii) estimating LULC in the next ten years, combining Geographical Information Systems (GIS) with Markov model. Landsat 5/8 images were combined with field evidence to map LULC in three reference years (1995, 2005, 2015and2020), while the Markov model was applied to forecast LULC in 2025, and 2030.The LULC of the middle reaches of the Yellow River Basin was classified into six classes: forest, grassland, wetland, cultivated land, construction land and unused land. The results show that over the past 25 years, cultivated land and forest land have decreased by 8,600 km ² and 6,400 km ² , respectively, while construction land has increased by 7,500 km ² , mainly due to increased urbanization and industrialization. In fact, during the simulated 2020–2030 period, cultivated land and forest land will continue to decrease, while construction land and grassland will increase. It is worth noting that the reduction of forest land will lead to increased soil erosion, while the reduction of cultivated land will lead to food security problems. The study points out that there is an urgent need for response measures in the region to ensure the sustainability of people and nature.
Article
Full-text available
Northwest China has experienced dramatic changes in vegetation cover over the past few decades with the Yellow River Basin (YRB) being the most representative area. As the major climate-sensitive area in China, vegetation cover change is one reason for its impact on surface air temperature (SAT). This study uses the observation minus reanalysis (OMR) method to reveal the spatio-temporal variations of vegetation cover and its impact on SAT change over non-urban areas of the YRB from 1982 to 2015. The Global Inventory Modeling and Mapping Studies dataset, SAT derived from meteorological stations, and ERA-interim reanalysis temperature data were used to analyze the relationship between normalized difference vegetation index (NDVI) and temperature variation caused by vegetation change. The NDVI trend (SlopeNDVI) of the entire YRB reached 1.11×10−2 decade−1, which indicated the recovery of vegetation in general. The impact of variation in vegetation conditions on SAT change during 1982–2015 was estimated to be 0.037 °C decade−1, which contributed 7.62% to the temperature change. The mean annual NDVI (MNDVI) and SlopeNDVI in the YRB were significantly negatively correlated (P<0.001) with OMR temperature variation. A negative correlation was exhibited in semi-arid and semi-humid regions, whereas a positive correlation was found in the arid region. The observed changes in vegetation and SAT in the YRB support the theory of the impact of vegetation variation on SAT in China.
Article
Full-text available
Panzhihua City is a typical agricultural-forestry-pastoral and ecologically sensitive city in China. It is also an important ecological defense in the upper Yangtze River. It has abundant mineral resources, including vanadium, titanium, and water supplies. However, ecological and environmental problems emerge due to the excessive development of mining, agriculture, animal husbandry, and other non-natural urban economies. Therefore, a scientific understanding of the spatio-temporal changes of the eco-environment of Panzhihua is critical for environmental protection, city planning, and construction. To objectively evaluate the eco-environmental status of Panzhihua, the remote sensing-based ecological index (RSEI) was first applied to Panzhihua, a typical resource-based city, and its ecological environmental quality (EEQ) was quantitatively assessed from 1990 to 2020. This study explored the effects of mining activities and policies on EEQ and used change detection to reveal the spatial-temporal changes of EEQ in Panzhihua City over the past three decades. In addition, this study also verified the suitability of RSEI for evaluating EEQ in resource-based city using spatial autocorrelation, revealed the spatial heterogeneity of EEQ in Panzhihua City using optimized hot spot analysis, and showed different ecological clustering by hot spot analysis at two scales of urban and mining areas. According to the results: (1) From 1990 to 2020, the general eco-environmental condition of Panzhihua is improving, but there are still regional differences. (2) The Moran’s I value ranges from 0.436 (1990) to 0.700 (2020), indicating that there is autocorrelation in the distribution of eco-environmental quality. (3) At the mine, the mean value of RSEI dropped by 20–40%, and the EEQ decreased significantly due to mining activities. (4) A series of ecological restoration policies can buffer the negative impact of mining activities on the ecosystem, resulting in a slight improvement in the quality of the ecological environment. This study evaluates the EEQ of resource-based city and its spatial-temporal changes using RSEI constructed by the Google Earth Engine (GEE) platform, which can provide theoretical support for ecological and environmental conditions monitoring, development planning, and environmental protection policy-making of a resource-based city.
Article
Full-text available
Tourism is a highly climate-sensitive sector, and the suitability of climate and weather can be a central motivation for travel. Through the tourism climate index (TCI), this study uses daily data from the period of 1980 to 2014 to identify the spatial and temporal distribution of climatic conditions in Xinjiang of China with regard to its climatic attractiveness for general tourism. The analytical results indicate that the climatic conditions of the most suitable months for tourism in Xinjiang are April, May, September, and October. The best climatic condition (TCI > 80) for tourism activities is found in September in most parts of Xinjiang, except for central Tianshan and Turpan Basins. The climate is not attractive for tourism in Xinjiang during the winter months. The annual average number of climatically suitable days (TCI > 70) are the highest in northwestern Aksu, most parts of Kashgar, and western Hotan. Most parts of northern Xinjiang are favorable for summer travel, with high TCI scores and a high average number of climatically suitable days. By contrast, tourist areas in southern and eastern Xinjiang have favorable climatic conditions during Spring and Autumn. These results can provide useful information to both tourists and the tourism industry.