Next Article in Journal
Diurnal Variation Characteristics of Clouds and Precipitation during the Summer Season in Two Typical Climate Regions of the Tibetan Plateau
Next Article in Special Issue
Using Atmospheric Inverse Modelling of Methane Budgets with Copernicus Land Water and Wetness Data to Detect Land Use-Related Emissions
Previous Article in Journal
Using InSAR and GPR Techniques to Detect Subsidence: Application to the Coastal Area of “A Xunqueira” (NW Spain)
Previous Article in Special Issue
Random Forest Classification of Land Use, Land-Use Change and Forestry (LULUCF) Using Sentinel-2 Data—A Case Study of Czechia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Spatiotemporal Characteristics and Driving Factors of Land-Cover Change in the Heilongjiang (Amur) River Basin

1
School of Geography and Tourism, Shaanxi Normal University, Xi’an 710119, China
2
School of Geographical Sciences, Shanxi Normal University, Taiyuan 030031, China
3
State Key Laboratory of Resources and Environmental Information System, Institute of Geographic Sciences and Natural Resources Research, Beijing 100101, China
4
Jiangsu Center for Collaborative Innovation in Geographical Information Resource Development and Application, Nanjing 210023, China
*
Author to whom correspondence should be addressed.
Remote Sens. 2023, 15(15), 3730; https://doi.org/10.3390/rs15153730
Submission received: 11 April 2023 / Revised: 21 July 2023 / Accepted: 23 July 2023 / Published: 26 July 2023

Abstract

:
Monitoring land-use and land-cover change (LUCC) is extremely important in the sustainable development and management of terrestrial ecosystems. Taking the Heilongjiang (Amur) River Basin as the study area, this study aimed to identify the spatiotemporal characteristics of land cover at the entire basin and at the country levels of the three countries using the land-use change index, frequency statistics and land-cover change degree. The results showed that: (1) The LULC types were mainly forest land and grassland (accounting for nearly 83% in total) from 2001 to 2019. The main land-cover types in China, Russia and Mongolia were forest land, forest land and grassland, respectively. (2) The area of urban land, cropland and wetland increased significantly from 2001 to 2019, while the area of forest land and bare land decreased during this time. In general, the ecological environment has greatly improved over the last 19 years, although the relevant ecological restoration still needs to be further implemented. The findings can provide a scientific basis for ecological protection and sustainable development in the Heilongjiang (Amur) River Basin. The approaches here are also applicable to land-cover change research in other similar regions.

Graphical Abstract

1. Introduction

Global environmental change and sustainability are being driven by land-cover change science [1,2]. By utilizing natural resources related to land, humans alter the surface cover of the Earth [3]. Changes in land cover can adversely affect natural elements such as the atmosphere, soil, vegetation, water resources and biological diversity, and then have a wide range of influences on the chemical cycles of the biological earth, energy exchange, water cycles and ecological processes. In the ecosystems of Earth, land-use and land-cover changes reflect human interaction with the environment and are increasingly regarded as factors contributing to global environmental change [4]. Ecosystem services and biodiversity are the results of the interaction between climate change and land-cover change [5]. Hence, understanding the driving factors of land-cover change and their impacts on ecological and anthropogenic processes is of great importance. At the end of the 20th century, under the strong promotion of the International Geosphere and Biosphere Program (IGBP) and the International Human Dimensions Programme on Global Environmental Change (IHDP), land use/land cover made significant progress in various aspects, such as spatiotemporal exploration, driving mechanism analysis, process characterization and simulation and macro-ecological effect evaluation, making effective contributions to understanding and improving global environmental change and achieving sustainable development [6].
Land-cover change has always been a central component in land-change science and plays a central role in current strategies for managing natural resources and monitoring environmental changes [7,8]. Earth observation services guarantee continuous land-cover mapping and are becoming of great interest worldwide [9]. Remote-sensing earth observation technology provides abundant, free, long-term, medium-resolution and high-resolution satellite imagery resources for land-use/land-cover research due to its rapid advancement [10]. With the development of remote sensing technology and classification algorithms, a series of land-use/land-cover products on regional, national and global scales have been developed at different spatial resolutions, such as GlobeLand30 [11], GLC_FCS30 [12], FROM-GLC [13], GLS_LC100 [14] and CCI_LC [15], among others.
The concepts of land use and land cover are fundamentally different but closely related. Land use refers to the purposes for which humans exploit the land cover and includes land management practices such as agricultural land, industrial land, transportation land, residential land, etc. [6,8]. According to the definitions of IGBP and IHDP, land cover refers to the layer of soils and biomass, including natural vegetation, crops and human structures that cover the land surface [8,16], such as various crops, forests, grasslands, houses, cement and asphalt pavements. On the one hand, land use is a direct reason for changing land cover, and on the other hand, its utilization is often constrained by land cover [6]. Land cover is thus directly observable, both in the field and in remote sensing images, while land use is not always easily observable [8]. However, from the perspective of spatial information product classification, land use and land cover can often be mutually transformed [6]. Since the late 20th century, global change issues have attracted a large amount of attention. As a key component of global changes, land-cover and land-use information have become increasingly important for an improved understanding of global environmental changes and feedbacks between social and environmental systems [2,17].
In the analysis of LULC-type changes, the application of the transfer matrix is extensive [18,19,20,21,22], but there are drawbacks such as less intuitive data expression and unfavorable trend analysis. For this reason, some scholars use a chord diagram to visualize the transition matrix in order to identify the flow, direction and diversity of LULC [23,24] more directly. However, chord diagrams cannot intuitively describe the comparative relationships between multiple time periods and have the drawbacks of complex graphics and difficulty in reading. The Sankey Diagram can effectively compensate for the shortcomings of chord diagrams and is very suitable for describing the transformation trajectories of LULC types over multiple research periods [25,26,27]. In terms of analyzing the intensity of LULC changes, comprehensive land-use dynamics [20,21,28], single land-use dynamics [20,28] and LULC change rate [24] are often used to describe the overall changes and severity of LULC. The land-use-type flow index [28] and the land-cover-type transfer rate [22] can further describe the conversion intensity between different LULC types. In addition, LULC has strong spatial differentiation, and analysis methods based on spatial features can better reveal the spatial pattern of LULC changes, such as the landscape pattern index [20], graph analysis [29,30], the grid unit method [30], etc. Among them, the grid element method is an effective means of expressing the spatial change pattern of LULC types [31], which can comprehensively reflect the spatial distribution of LULC change areas.
Affected by the natural environment and local economic policies, the characteristics of distribution and spatiotemporal changes of land-cover types in different countries and regions show obvious differences, especially in cross-border areas [32,33]. The Heilongjiang (Amur) River Basin (HARB) has a unique physical geography environment, rich and diverse species and natural resources, spanning China, Russia and Mongolia, with its southwestern side located within Mongolia, its southern side located within China and its northern side and eastern side located within Russia. In recent years, with various factors such as global population growth, climate change and frequent natural disasters, the ecosystem of the HARB has also been affected to a certain extent, and its function is gradually weakening.
Due to the function that land use and land cover (LULC) serve at a watershed scale, land-use and land-cover change (LULCC) are of paramount importance [34]. We used various methods to monitor land-use/land-cover change to analyze the status of land-use/land-cover change within the HARB. We also elucidated land-cover change and its driving factors in the past 19 years and achieved a comparative analysis of land-cover change characteristics in cross-border areas of China, Russia and Mongolia. The main reasons for land-cover change were also explored. The aim is to provide a reference for the study of human–land relations and land resource management in the HARB, lay the foundation for the sustainable use of land resources and ecological environment protection in China and Russia and provide basic data and a scientific basis for relevant research in the northeastern border area of China.

2. Materials and Methods

2.1. Study Area

The Heilongjiang (Amur) River Basin (HARB) is located in the middle- and high-latitude parts of eastern Eurasia [35] and is one of the ten biggest river basins in the world. It is situated between 107.051°E–141.128°E and 41.72°N–55.903°N. The total area of the HARB is 208.33 × 104 km2, and the flow path is 1.68 × 104 km. The HARB is conceived as a whole in terms of its natural environment, but it is a cross-border region in the administrative division, which includes parts of China (41%), Russia (50%) and Mongolia (9%). The climate of the HARB varies widely because it is deeply influenced by the high- and low-pressure centers of the inland and seas and the alternation of the monsoon. In winter, the Siberian cold and dry monsoon from the west controls the HARB, while in summer, the warm and humid monsoon from the east controls the basin [36]. At the same time, the basin climate is sensitive to global climate change and is becoming more complicated due to the variable terrain. The terrain of the HARB is higher in the west and lower in the east. The western part of the basin is mainly composed of mountainous and plateau areas, such as the Greater Khingan Mountains and the Mongolian Plateau, and the southern and eastern parts are mainly flat plains, that is, the Songnen Plain and the Sanjiang Plain (see Figure 1). Because of the changeable topography and complex climate, the land-cover types in the basin are varied.

2.2. Data Sources

The datasets used in this paper mainly included land-cover data, the Digital Evaluation Model (DEM), meteorological data, field observation data and socioeconomic data.
(1) Land-cover datasets. The Collection 6 land-cover data (2001–2019) used in this study were the latest version of the Moderate Resolution Imaging Spectroradiometer (MODIS)-based Land-Cover-Type product (MCD12Q1) [37], available at https://earthexplorer.usgs.gov/ (Accessed on 10 May 2021). The MODIS Land-Cover-Type product contains multiple classification schemes that describe land-cover properties derived from observations spanning a year’s input of Terra and Aqua data. The MCD12Q1 product was derived through a supervised decision-tree classification method with five different land-cover classification schemes, including the International Geosphere Biosphere Programme (IGBP) legend [38,39], the University of Maryland (UMD) legend [40], the Leaf Area Index and Fraction of Photosynthetically Active Radiation (LAI/FPAR) biome legend [41], the Biome-Biogeochemical Cycles BGC model legend [42] and the Plant Functional Types (PFTs) legend [43]. The Collection 6 algorithm uses a hierarchical classification model, where the classes contained in each level of the hierarchy reflect structured distinctions between land-cover properties [37]. The Collection 6 algorithm also incorporates a state-space multitemporal modeling framework based on hidden Markov models, which can reduce the spurious land-cover changes caused by classification uncertainty in individual years [37]. The amount of spurious land-cover change has been significantly decreased in Collection 6, accounting for only 1.6%. The accuracy of the IGBP layer of the MCD12Q1 was estimated to be 73.6% globally. A Chinese scholar discussed the accuracy analysis and applicability evaluation of the MODIS land-cover product in the Yellow River source region from two aspects: quantity accuracy and shape consistency. The results showed that, in terms of shape, the overall shape consistency with weights is above 69%, with the consistency of the main grassland types reaching over 88%. In terms of quantity, the relative error of the overall area with weights is within 26%, and the error mainly occurs in unused land and other land types. The MODIS land-cover data product is still of importance and has application value in large-scale land-cover monitoring [44]. An evaluation of the consistency of five land-cover classification schemes of MCD12Q1 was performed based on the 30 m Chinese GlobeLand30-2010 GLC product in Anhui Province. The overall accuracy of IGBP, UMD and PFT in the three sub-regions of Anhui Province was more than 71%, while it was less than 57% for LAI/FPAR and NPP [45]. We also compared five classification schemes based on spatial consistency and area consistency and found that the consistency of the IGBP classification scheme was the best. Thence, the IGBP classification scheme was chosen to analyze the temporal and spatial changes of land cover in the study area.
The MODIS Land-Cover-Type product (MCD12Q1) supplies global maps of land cover at annual time steps and 500 m spatial resolution. MCD12Q1 data are provided as tiles that are approximately 10° × 10° at the Equator using a Sinusoidal grid in HDF4 file format [46]. The HARB contains a total of seven tiles, including H24V04, H25V03, H25V04, H26V03, H26V04, H27V03 and H27V04. We collected 133 tiles of MCD12Q1 products from 1 January 2001 to 31 December 2019. The MODIS Reprojection Tool (MRT) was used to finish the imaging mosaicking, format conversion, reprojection and resampling [47]. Here, the original HDF4 was converted into Geotiff, while the projection was converted to WGS_1984_Albers to ensure the smallest area deformation. In addition, we extracted the land-cover types of the study area by the boundary of the HARB and reclassified them to seven types, that is, Forest (F), Grassland (GL), Wetland (WL), Farmland (FL), Urban (UB), Barren (B) and Water (W) in ArcGIS 10.2.
(2) Boundary data. The boundary data for the Heilongjiang (Amur) River Basin are available at http://www.geodoi.ac.cn/WebCn/Default.aspx (Accessed on 12 May 2021). These data are based on ASTER-DEM data; preliminary results were extracted using watershed area extraction algorithms and then generated by matching them with Google Earth remote sensing images.
(3) DEM. The DEM data were obtained from the geospatial data cloud and had a resolution of 90 × 90 m (http://www.gscloud.cn/ (Accessed on 12 May 2021)). The DEM dataset was resampled from approximately 90 m to 500 m and reprojected to WGS_1984_Albers. In addition, we finished the imaging mosaicking and the image extraction by the boundary of the HARB and calculated the slope of the HARB in ArcGIS10.2.
(4) Meteorological data. The meteorological data used in this study were the monthly gridded Climatic Research Unit Time-Series data version 4.05 (CRU TS4.05), acquired from https://data.ceda.ac.uk/badc/cru/data/cru_ts/cru_ts_4.05/data (Accessed on 15 May 2021). CRU TS4.05 datasets with spatial resolution of 0.5° and long-time span from 1901 to 2020 are stored as NC files, produced by CRU at the University of East Anglia [35]. We chose factors ‘pre’ and ‘tmp’ from CRU TS4.05 as precipitation and temperature to analyze climate change and its relationship with land-cover change. The NC files were converted into Geotiff format by a Python script file. In ArcGIS10.2, the precipitation and temperature datasets were resampled to 500 m by using the cubic convolution method, and annual precipitation and annual average temperature were calculated from 2001 to 2019.
(5) Field observation data. In May 2019, we carried out a week-long field survey in Primorsky Krai, Russia, mainly on the west, east and south banks of Lake Khanka, Amur Bay, Ussuri Bay, the upper reaches of Ussuri and the Razdolnaya basin. We collected relevant data, including surface coverage and vegetation species.
(6) Socioeconomic data. The economic, population and agricultural data of the study area in China were acquired from the 2001–2019 statistical yearbooks of Heilongjiang Province, Jilin Province and Inner Mongolia Autonomous Region Liaoning Province; the relevant data in Russia and Mongolia were downloaded from the World Bank data platform (https://data.worldbank.org.cn/ (Accessed on 20 May 2021)) and the 2001–2019 Russian statistical yearbook, available at https://rosstat.gov.ru/ (Accessed on 15 May 2019).

2.3. Method

2.3.1. Spatiotemporal Changes of Land Cover and Ecological Value Judgment

In this study, the spatiotemporal changes of seven land-cover types were analyzed from four regions: the entire HARB, the Chinese region of the HARB, the Russian region and the Mongolian region. It was found that the ecosystem service value per unit area of each ecosystem type is different, and the order from high to low is wetland, water, forest, grassland, farmland, barren and urban [47,48,49,50]. In the analysis of land-cover dynamic change, we defined the conversion of land-cover types with increasing ecosystem service value as landscape improvement and the conversion with decreasing ecosystem service value as landscape degradation.

2.3.2. Dynamic Analysis of Land Cover

Land-use dynamic was used in the quantitative description of land-use change rate in different study periods, including the single land-use dynamic index and the comprehensive land-use dynamic index [51]. In this paper, these two indicators were used to represent a certain land-cover type and the rate of regional land-cover change during the study period [52] and then to analyze the rate of land-cover change in the HARB and explore the impact of different national policies on the land-cover change rate.
The single land-use dynamic index can be defined as:
D s = A t 2 A t 1 A t 1 × 1 t 2 t 1 × 100 %
where A t 1 and   A t 2 are the areas of the land-cover type at times t 1 and t 2 .
The comprehensive land-use dynamic index can be defined as:
D c = i = 1 n Δ A i n i Δ A o u t i 2 i = 1 n A i × 1 t 2 t 1 × 100 %
where A i is the area of the land-cover type i in time t 1 ; Δ A i n i is the total area converted from other types to type i from t 1 to t 2 ; and Δ A o u t i is the total area converted from type i to other types from t 1 to t 2 .
The dynamic degree of land use quantitatively describes the speed of land-use change, which plays a positive role in predicting the trend in land-use change in the future [51].

2.3.3. Land-Use Integrated Index

The change in extent of land use mainly reflects the breadth and depth of land use, not only in its natural attributes but also in the comprehensive effects of human factors and natural environmental factors. According to the land-use integrated index, which was put forward by Liu Ji-yuan [53], the change extent of land cover can be divided into four levels. The highest level is the use of land that completely destroys the natural environment, such as towns, residential areas and industrial development land; the land resources that have not been utilized by humans or cannot be utilized from the perspective of economy and technology are defined as the lowest level. Between these two levels, due to the different degrees of human intervention in the natural environment of the land, two intermediate levels are defined. Assign the value of its own category to the four land-use levels, and then calculate the grading index of the four land-use levels (Table 1).
The degree of land-cover change can be used to better understand the trend and its driving forces [51,52,54]. The equation is shown as follows:
L d = i = 1 n A i C i × 100 ,     L j 100 , 400
where A i is the given index of level i, C i is the percentage of the area of level i in the study area and n is the number of levels.
Furthermore, the change in quantity and the rate of land-use intensity can quantitatively reveal the comprehensive level and development trend in land use in this region.
Δ L b a = L b L a = i = 1 n A i C i b i = 1 n A i C i a × 100
R = i = 1 n A i C i b i = 1 n A i C i a i = 1 n A i C i a
where L b , L a denote the change extent of land use at periods a and b and C i a and C i b are the area ratio of the ith level at periods a and b. When Δ L b a > 0 or R > 0 , land use is at the development stage or at the adjustment or recession stage [47,50].

3. Results

3.1. Spatiotemporal Changes of Land Cover

3.1.1. Spatiotemporal Changes of Land Cover in HARB

The spatial distribution maps of land cover in the HARB from 2001 to 2019 are given in Figure 2. The main land-cover type in the HARB during the study period was forest, accounting for about 50% of the total area and being mainly distributed in the north and east of the HARB. The second largest land-cover type was grassland, accounting for about 33% of the total area, distributed in the Mongolian plateau and the west of the Greater Khingan. Farmland accounted for about 15% of the total area, mainly distributed in the Sonnen Plain and the Sanjiang Plain, and a small part was distributed near Xingkai Lake. Other land-cover types, including wetland, urban, barren and water areas, accounted for less than 2% of the total area.
The area of different land-cover types from 2001 to 2019 is presented in Figure 3. The best trend-fitting equations of each land-cover type’s change with time were all in cubic form, and all passed the significance test (Figure 3, Table 2). The data show that the area change in land-cover type in the study area was affected by many driving factors. We found that the speed of change in wetland areas was the highest, with an increase of 71.23% from 2001 to 2019. The change trend in the wetland area was decreasing first and then increasing, and the increased area of wetland was 1591 km2. The change trend in the water area had the same trend as the wetland area; the area increased by 692.5 km2, and the change rate was 3.86%. The area of urban land continued to increase, from 10,418.25 km2 in 2001 to 10,973.75 km2 in 2019, and the change rate was 5.33%. The change in farmland area also showed a trend of first decreasing and then increasing. On the whole, farmland increased by 14,031.5 km2 during the study period. The change trend in forest and grassland areas was the opposite, but the change extent was the same. The area of forest was 1,070,843.75 km2 in 2010, and it was the largest area in the study period. While the area of grassland was the smallest in 2010, it was 681,478.75 km2. The barren area fluctuated around 1150 km2 and decreased by 200.5 km2 from 2001 to 2019.
During the study period, there were 36 types of land-cover-type transformation processes in the HARB, and the area where land-cover types changed accounted for 9.18% of the total area of the entire basin. The trend that the total area of land experienced a change in land-cover types every year was slow before 2016, but after that, the change intensified. The major land-cover-type transformation processes with the top 10 largest areas of change were selected to analyze the converted processes (Figure 4). The area of forest that converted to grassland was decreasing before 2010, but it increased afterwards, while there was an opposite trend of grassland converting to forest at the same time. The area of grassland converted to other land-cover types has been roughly increasing since 2001.
In order to analyze the macro-trends in landscape dynamic changes throughout the years, we further selected 17 land-cover-type transformation processes that were greater than 0.01% of the entire basin area for landscape quality change analysis (Table 2, Figure 5). The result showed that the area of landscape improvement accounted for 4.04% of the whole basin, and the main processes of landscape improvement included: (1) the transformation from grassland to forest, mainly distributed in the Greater Khingan and the north of the Russian region; and (2) the conversion of farmland to grassland or forest, mainly distributed in the low-altitude area of the Greater Khingan. However, landscape degradation occurred in 5.14% of the whole basin area, and the main processes were as follows: (1) forest degenerated to grassland or farmland, mainly distributed in Zabaykalsky Krai, both sides of the Heilongjiang River and the Greater Khingan; and (2) grassland was reclaimed into farmland, mainly distributed in the low-altitude areas of the Greater Khingan and around Xingkai Lake.

3.1.2. Spatiotemporal Changes of China’s Land Cover in HARB

The total area of the region that is located in the territory of China is 88.18 × 104 km2, accounting for about 41% of the HARB. The main land-cover type in this region in 2019 was forest, followed by farmland and grassland. Forests are distributed in the Greater Khingan, the Little Khingan and Changbai Mountain. The area of farmland in the region was 29.9 × 104 km2 in 2019, which accounted for about 95% of the farmland area in the whole basin. Grassland was mainly distributed in the north of Inner Mongolia, and its area was 22.88 × 104 km2 in 2019. Two-thirds of the urban area in the HARB was located in the Chinese region, and the area was 7421.5 km2 in 2019.
The change trend in land-cover types in the Chinese region was consistent with the change trend in land-cover types in the entire basin, but the change rate of all land-cover types in the Chinese region was greater than that of the whole basin, which indicated that land-cover changes in the Chinese region were more intense. The change trend in forest, wetland and water was increasing first and then decreasing. In the study period, the areas of forest, wetland and water increased by 4708.25 km2, 1178 km2 and 655.25 km2, respectively. The change rate of the wetland area, at 112.51%, is the largest among all the land-cover types. The farmland area and the urban area also showed an increasing trend and increased by 12,636 km2 and 550.25 km2, respectively, while the grassland area and the barren area decreased by 19,508 km2 and 219.75 km2, respectively (Table 3).

3.1.3. Spatiotemporal Changes of Russia’s Land Cover in HARB

Half of the total area of the HARB is located in Russia, covering an area of 101.02 × 104 km2. In the Russian region, forest was the most important land-cover type, accounting for about 70% of the total area of the Russian region and being evenly distributed in the region. Grassland was the second largest land-cover type, with a total area of 28.9209 × 104 km2 in 2019, accounting for about 27% of the total area of the Russian region. The total area of other land-cover types, that is, wetland, farmland, urban, barren and water, accounted for less than 3% (Table 4).
The change trend in land-cover types in the Russian region was consistent with the change trend in land-cover types in the Chinese region. However, except for the fact that the change rate of barren land cover was greater than that of the Chinese region, the change rate of other land-cover types was less than that of the Chinese region. The intensity of land-cover-type change in the Russian region was weaker than in the Chinese region. In the Russian region, only the forest area decreased by 13,005.5 km2 from 2001 to 2019, while other land-cover types increased.

3.1.4. Spatiotemporal Changes of Mongolia’s Land Cover in HARB

The area of the Mongolian region in the HARB is 19.1 × 104 km2, accounting for 9% of the total area of the entire basin. In the Mongolian region, the grassland land-cover type, with an area of 18.7 × 104 km2, occupied a dominant position and accounted for more than 98% of the total area in 2019. The other land-cover types, that is, forest, farmland, wetland, urban, barren and water, only accounted for less than 2% and had a scattered distribution in the region.
In the study period, the areas of grassland and wetland in the region increased by 651 km2 and 4 km2, respectively. The urban area did not change, while the area of the other land-cover types decreased by 501.75 km2, 35.25 km2, 3 km2 and 115 km2, respectively (Table 5). The change intensity of all the land-cover types did not vary greatly.

3.2. Dynamic Degree Analysis of Land Cover

The results of the single land-use dynamic index, which was calculated by Equation (1), are shown in Table 6. It can be found that during the study period, the single land-use dynamic index of each land-cover type in the Chinese region was roughly consistent with that of the HARB. In the above two regions, wetland, water, urban and farmland increased at a high speed. Barren and grassland reduced at a high speed and forest increased at a slow speed in the Chinese region, whereas they decreased at a slow speed in the HARB. For forest in the Russian region, the reduced speed was higher than that in the HARB. Except for forest, the other land-cover types increased. In the Mongolian region, the reduced speed of forest was the highest in the four regions, and the area of reduced forest was converted to grassland.
Using Equation (2) combined with the land-cover transformation matrix, the comprehensive land-use dynamic index was calculated from 2001 to 2019 (see Figure 6). From Figure 6, we can see that the change rate of land-cover types in the Chinese region was the highest of the four regions. The comprehensive dynamics of land cover in the Russian and Mongolian regions were lower than the average level of the whole basin. The land-cover changes in the Chinese region play a major role in the change in land cover in the HARB. During the research period, the land-cover change within China’s watershed was more significant, and the intensity of human activities was greater.

3.3. Analysis of the Land-Cover Integrated Index

The integrated index of land use can reflect the overall land-use intensity in a region, which indicates the extent to which land is exploited by humans [55]. The land-use synthetic indices of the HARB, the Chinese region, the Russian region and the Mongolian region are shown in Figure 7. It can be found that, during the study period, the integrated index of land cover in the Chinese region was roughly consistent with that of the HARB. Before 2018, the integrated index of land use in the Russian region remained consistent with that of the HARB but suddenly declined thereafter. While the integrated index of land use in the Mongolian region remained consistent with that of the HARB before 2016, it suddenly declined thereafter.
Furthermore, we used Formulas (4) and (5) to calculate the change quantity and the rate of land-use intensity to reveal the comprehensive level and development trend in land use in the basin. The results (Figure 8 and Figure 9) showed that the development trend in land use in the other regions was a development stage from 2001 to 2019, except for Inner Mongolia. Throughout the entire research period, the state of land-use development went through five stages: (1) 2001–2002, development stage; (2) 2002–2007, recession stage; (3) 2007–2011, development stage; (4) 2011–2013, recession stage; (5) 2013–2019, development stage.

4. Discussion

4.1. The Validation of the Land Cover

Land-cover remote sensing products may have significant differences in product accuracy due to differences in data sources, classification techniques and producers [56,57]. Due to different application objectives, there are significant differences in spatial resolution, classification systems and other aspects of land-cover remote sensing products [57,58,59,60].
We validated the land-cover data using a total of 500 validation samples collected from Google Earth images and field investigations in 2019. The spatial distribution of the validation sample points is shown in Figure 10. In May 2019, we carried out a week-long field survey in Primorsky Krai, Russia, mainly on the west, east and south banks of Lake Khanka, Amur Bay, Ussuri Bay, the upper reaches of Ussuri, the Razdolnaya basin, the Ussuriysk Taijialin Forest Monitoring Station, etc. The obtained sampling points were used to verify the MCD12Q1 land-cover product using the Confusion matrix. The overall accuracy was 78.8%, and the Kappa coefficient was 0.78, which met the research requirements. The error in wetland was the largest.
From the verification results, it can be seen that the accuracy of the MODIS land-cover data within the HARB is relatively low, which will have a certain impact on the analysis of land-cover changes. In order to better study land-use/land-cover changes in the HARB, we need more accurate land-cover products. With the development of remote sensing and image recognition, we will obtain higher spatial resolution of remote sensing images and higher-precision land-cover products.

4.2. The Driving Forces of Land Cover

Land-use/land-cover change is a complex process, and the driving factors that affect land-use/land-cover change include natural and human factors. Understanding the correlation between land-cover change and its various driving factors is the basis for constructing a land-use/land-cover simulation model.

4.2.1. Natural Factors

In recent years, the gradual warming of the temperature and the fluctuation of precipitation in the HARB have been important physical geography factors of land-cover change in the sparsely populated region. The average annual temperature in the HARB from 2001 to 2019 showed a gradual upward trend (see Figure 11). The interannual precipitation fluctuates greatly. With the continuous development of agriculture, the demand for cropland will increase, resulting in the conversion of forest, grassland and wetland into cropland.
On the other hand, the precipitation in the basin grew steadily, and the interannual precipitation fluctuation was small (see Figure 12). By comparing the land-cover changes in the regions of China, Russia, Mongolia and the entire basin under nearly the same temperature and precipitation conditions, we found that the land-cover types in China changed significantly, while those in Russia and Mongolia changed slightly. It can be seen that, although physical geography factors have a certain impact on land-cover changes, they are not the main driving factor in the watershed [22].
Terrain factors are also one of the important factors affecting land-cover change [61]. With the increase in slopes, living costs continue to increase, water flow slows down and soil erosion intensifies, leading to difficulties in agricultural cultivation. The areas where land-cover types undergo transformation are mainly characterized by low altitude, low slopes, etc. Forest land, due to its unique characteristics, often appears in high-altitude and large slope areas. The transformation of farmland and wetlands mostly occurs in areas with slopes below 6° and elevations below 150 m, accounting for more than 90% of the total transformation area. From this, it can be seen that low-altitude and low-slope areas are the main areas where land-cover transformation occurs.

4.2.2. Anthropogenic Factor

Land-cover change is the result of a combination of various related factors. Natural factors mainly affect the spatial pattern of land cover, while human activities act on the spatiotemporal changes of land cover [62]. Population growth is the main driving force behind the intensification of human activities [63]. According to the statistical yearbook data of Heilongjiang Province from 2001 to 2019, the population change in Heilongjiang Province showed a slight downward trend from 38.11 million in 2001 to 37.51 million in 2019, indicating a decrease in the total population (Figure 13). However, the proportion of the agricultural population decreased from 47.62% in 2001 to 39.1% in 2019. While the proportion of the agricultural population decreased, urban migration intensified, resulting in an increase in artificial surfaces in China. This indicates that changes in population structure lead to the expansion of urban land use, and population factors have a certain impact on land-cover change but are not the main driving force of land-cover change in the HARB.
Describing the intensity of land-cover change from a spatial perspective is a challenging task. Widely used methods such as the landscape pattern index and graph analysis in existing research can describe the spatial distribution characteristics of land-cover change intensity. However, for areas with a staggered distribution of multiple land-cover types, the expression effect of these methods is not good, and they cannot effectively highlight the intensity and distribution patterns of their changes. The research method adopted in this article provides an effective means to analyze land-cover change in the Heilongjiang (Amur) River Basin during the past 20 years from multiple dimensions such as basic structure, temporal changes, spatial changes and driving factors. This can not only provide a useful reference for the ecological environment construction of the basin but also be extended to other regional land-cover change studies.

5. Conclusions

This article comprehensively analyzes the changes in land cover in the Heilongjiang (Amur) River Basin during the past 20 years from the perspectives of basic structure, temporal changes and spatial changes in order to provide useful references for the sustainable development of grassland ecological environments. The main conclusions are as follows:
From 2001 to 2019, the dominant type of LULC in the entire basin was forest land, covering the vast majority of the northern and eastern regions of the study area. The second was grassland, mainly distributed in the western region. The total area ratio between the two was about 83%. The main land-cover type within the Chinese region was forest, followed by farmland and grassland. The area of farmland accounted for about 95% of the farmland area in the whole basin. In the Russian region, forest was the most important land-cover type, accounting for about 70% of the total area of the Russian region and being evenly distributed in the region. In the Mongolian region, the grassland land-cover type, with an area of 18.7 × 104 km2, occupied a dominant position and accounted for more than 98% of the total area.
The area of the different land-cover types from 2001 to 2019 showed that the area change in land-cover type in the study area was affected by many driving factors. We found that the speed of change in wetland is the highest, with an increase of 71.23% from 2001 to 2019. The trend of changes in water areas and urban land use increased by 3.86% and 5.33%, respectively. The change in farmland area also showed a trend of first decreasing and then increasing. On the whole, farmland increased by 14,031.5 km2 during the study period. The change trend in the forest and grassland areas was the opposite, but the change extent was the same.
It can be found that during the study period, the single land-use dynamic index of each land-cover type in the Chinese region was roughly consistent with that of the HARB. The comprehensive dynamics of land cover in the Russian and Mongolian regions were lower than the average level of the whole basin. The land-cover changes in the Chinese region played a major role in the change in land cover in the HARB.
However, while this study achieved an analysis of annual-scale land-cover changes and driving forces in the watershed, there are still some areas that need improvement. For example, the land-cover product used in this article has a spatial resolution of 500 m and low accuracy. In our next research work, we will consider introducing some of the latest time-series land-cover classification algorithms [37,63,64] to obtain higher-accuracy land-cover data.

Author Contributions

Conceptualization, Y.Y. and S.J.; methodology, Y.Y. and S.J.; software, S.J.; validation, S.J.; formal analysis, S.J.; investigation, Y.Y. and S.J.; resources, Y.Y. and S.J.; data curation, S.J.; writing—original draft preparation, S.J.; writing—review and editing, Y.Y.; visualization, S.J. supervision, Y.Y.; project administration, Y.Y.; funding acquisition, Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by National Earth System Science Data Sharing Infrastructure (2005DKA32300).

Data Availability Statement

All the data used in this study are publicly available. The boundary data for the HARB are available at http://www.geodoi.ac.cn/WebCn/Default.aspx (Accessed on 12 May 2021). The climate dataset used in this study was the monthly gridded Climatic Research Unit Time-Series data version 4.05 (CRU TS4.05), acquired from https://data.ceda.ac.uk/badc/cru/data/cruts/cru_ts_4.05/data (Accessed on 15 May 2021). The digital elevation model (DEM) datasets were obtained from the Shuttle Radar Topography Mission (SRTM), available at http://www.gscloud.cn/ (Accessed on 12 May 2021). The economic, population and agricultural data of the study area in China were acquired from the 2001–2019 statistical yearbooks of Heilongjiang Province (http://tjj.hlj.gov.cn/ (Accessed on 20 March 2022)), Jilin Province (http://tjj.jl.gov.cn/ (Accessed on 20 March 2022)), Inner Mongolia Autonomous Region (http://tj.nmg.gov.cn/ (Accessed on 20 March 2022)) and Liaoning Province (http://tjj.ln.gov.cn/ (Accessed on 20 March 2022)); the relevant data in Russia and Mongolia were downloaded from the World Bank data platform (https://data.worldbank.org.cn/ (Accessed on 20 May 2021)) and the 2001–2019 Russian statistical yearbook, available at https://rosstat.gov.ru/ (Accessed on 15 May 2019).

Acknowledgments

Data support from the National Earth System Science Data Center, National Science and Technology Infrastructure of China (http://www.geodata.cn) is acknowledged. We are grateful to the academic editor and reviewers for their insightful comments, which greatly helped us improve the quality of this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Turner, B.L., II; Lambin, E.F.; Reenberg, A. The emergence of land change science for global environmental change and sustainability. Proc. Natl. Acad. Sci. USA 2007, 104, 20666–20671. [Google Scholar] [CrossRef]
  2. Dong, J.; Kuang, W.; Liu, J. Continuous land cover change monitoring in the remote sensing big data era. Sci. China Earth Sci. 2017, 60, 2223–2224. [Google Scholar] [CrossRef]
  3. Turner, B.L., II; Skole, D.L.; Sanderson, S.; Fischer, G.; Fresco, L.; Leemans, R. Land-Use and Land-Cover Change. Science/Research Plan; Global Change Report (Sweden); The International Geosphere Biosphere Programme: A Study of Global Change (IGBP) of the International Council of Scientific Unions (ICSU) and The Human Dimensions of Global Environmental Change Programme (HDP) of the International Social Science Council (ISSC): Stockholm, Sweden; Geneva, Switzerland, 1995. [Google Scholar] [CrossRef]
  4. Pimnon, S.; Bhumiratana, A. Adaptation of Anopheles Vectors to Anthropogenic Malaria-Associated Rubber Plantations and Indoor Residual Spraying: Establishing Population Dynamics and Insecticide Susceptibility. Can. J. Infect. Dis. Med. Microbiol. 2018, 2018, 9853409. [Google Scholar] [CrossRef] [Green Version]
  5. He, X.; Liang, J.; Zeng, G.; Yuan, Y.; Li, X. The Effects of Interaction between Climate Change and Land-Use/Cover Change on Biodiversity-Related Ecosystem Services. Glob. Chall. 2019, 3, 1800095. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Wang, Z.; Gao, K.; Yang, X.; Su, F.; Huang, C.; Shi, T.; Yan, F.; Li, H.; Zhang, H.; Lv, N.; et al. Land use/land cover classification development from a geographical perspective. Geogr. Res. 2022, 41, 2946–2962. (In Chinese) [Google Scholar] [CrossRef]
  7. Attri, P.; Chaudhry, S.; Sharma, S. Remote Sensing & GIS based Approaches for LULC Change Detection—A Review. Int. J. Curr. Eng. Technol. 2015, 5, 3126–3137. [Google Scholar]
  8. Verburg, P.H.; van de Steeg, J.; Veldkamp, A.; Willemen, L. From land cover change to land function dynamics: A major challenge to improve land characterization. J. Environ. Manag. 2009, 90, 1327–1335. [Google Scholar] [CrossRef]
  9. Orusa, T.; Cammareri, D.; Borgogno Mondino, E.B. A Scalable Earth Observation Service to Map Land Cover in Geomorphological Complex Areas beyond the Dynamic World: An Application in Aosta Valley (NW Italy). Appl. Sci. 2023, 13, 390. [Google Scholar] [CrossRef]
  10. Hu, Y.; Zhang, Y. Spatial–temporal dynamics and driving factor analysis of urban ecological land in Zhuhai city, China. Sci. Rep. 2020, 10, 16174. [Google Scholar] [CrossRef]
  11. Chen, J.; Chen, J.; Liao, A.; Cao, X.; Chen, L.; Chen, X.; He, C.; Han, G.; Peng, S.; Lu, M.; et al. Global land cover mapping at 30 m resolution: A POK-based operational approach. Isprs J. Photogramm. Remote Sens. 2015, 103, 7–27. [Google Scholar] [CrossRef] [Green Version]
  12. Zhang, X.; Liu, L.; Chen, X.; Gao, Y.; Mi, J. GLC_FCS30: Global land-cover product with fine classification system at 30 m using time-series Landsat imagery. Earth Syst. Sci. Data Discuss. 2021, 13, 2753–2776. [Google Scholar] [CrossRef]
  13. Gong, P.; Liu, H.; Zhang, M.; Li, C.; Wang, J.; Huang, H.; Clinton, N.; Ji, L.; Li, W.; Bai, Y.; et al. Stable classification with limited sample: Transferring a 30-m resolution sample set collected in 2015 to mapping 10-m resolution global land cover in 2017. Sci. Bull. 2019, 64, 370–373. [Google Scholar] [CrossRef] [PubMed]
  14. Buchhorn, M.; Lesiv, M.; Tsendbazar, N.-E.; Herold, M.; Bertels, L.; Smets, B. Copernicus Global Land Cover Layers—Collection 2. Remote Sens. 2020, 12, 1044. [Google Scholar] [CrossRef] [Green Version]
  15. Bontemps, S.; Boettcher, M.; Brockmann, C.; Kirches, G.; Lamarche, C.; Radoux, J.; Santoro, M.; Vanbogaert, E.; Wegmüller, U.; Herold, M.; et al. Multi-year global land cover mapping at 300 m and characterization for climate modelling: Achievements of the Land Cover component of the ESA Climate Change Initiative. Int. Arch. Photogramm. Remote Sens. Spat. Inf. Sci. 2015, XL-7/W3, 323–328. [Google Scholar] [CrossRef] [Green Version]
  16. Turner, B.L., II; Meyer, W.B.; Skole, D.L. Global land-use/land-cover change: Towards an integrated study. AMBIO—J. Hum. Environ. 1994, 23, 91–95. [Google Scholar]
  17. Verburg, P.H.; Crossman, N.; Ellis, E.C.; Heinimann, A.; Hostert, P.; Mertz, O.; Nagendra, H.; Sikor, T.; Erb, K.-H.; Golubiewski, N.; et al. Land system science and sustainable development of the earth system: A global land project perspective. Anthropocene 2015, 12, 29–41. [Google Scholar] [CrossRef] [Green Version]
  18. Hao, J.; Liu, H.; Li, Y.; Hu, H.; An, J. Spatio-temporal Variation and Driving Forces of the Coastal Wetland Resources Based on the Transition Matrix in Jiangsu Province. J. Nat. Resour. 2010, 25, 1918–1929. (In Chinese) [Google Scholar] [CrossRef]
  19. Zhang, F.; Tiyip, T.; Feng, Z.D.; Kung, H.; Johnson, V.C.; Ding, J.L.; Tashpolat, N.; Sawut, M.; Gui, D.W. Spatio-Temporal Patterns of Land Use/Cover Changes Over the Past 20 Years in the Middle Reaches of the Tarim River, Xinjiang, China. Land Degrad. Dev. 2015, 26, 284–299. [Google Scholar] [CrossRef]
  20. Zhao, Y.; Yu, X.; Jia, J.; Liu, X. Analysis on dynamic evolution and driving force of land-use landscape in Hongmenchuan basin. Trans. Chin. Soc. Agric. Eng. 2013, 29, 239–248. [Google Scholar] [CrossRef]
  21. Yang, X.; Wei, Z.; Yun, X. The changes of land use and vegetation during recent 30 years in steppe grassland of Xilinhot, Inner Mogolia. J. China Agric. Univ. 2015, 20, 196–204. (In Chinese) [Google Scholar] [CrossRef]
  22. Zhou, D.; Zhao, S.; Zhu, C. The Grain for Green Project induced land cover change in the Loess Plateau: A case study with Ansai County, Shanxi Province, China. Ecol. Indic. 2012, 23, 88–94. [Google Scholar] [CrossRef]
  23. Fyfe, R.M.; Woodbridge, J.; Roberts, C.N. Trajectories of change in Mediterranean Holocene vegetation through classification of pollen data. Veg. Hist. Archaeobotany 2018, 27, 351–364. [Google Scholar] [CrossRef]
  24. Xiao, D.; Niu, H.; Yan, H.; Fan, L.; Zhao, S. Spatiotemperal evolution of land use pattern in the Yellow River Basin (Henan section) from 1990 to 2018. Trans. Chin. Soc. Agric. Eng. 2020, 36, 271–281. (In Chinese) [Google Scholar] [CrossRef]
  25. Perez, G.J.; Comiso, J.C.; Aragones, L.V.; Merida, H.C.; Ong, P.S. Reforestation and Deforestation in Northern Luzon, Philippines: Critical Issues as Observed from Space. Forests 2020, 11, 1071. [Google Scholar] [CrossRef]
  26. Yu, S.; Wang, F.; Qu, M.; Yu, B.; Zhao, Z. The Effect of Land Use/Cover Change on Soil Erosion Change by Spatial Regression in Changwu County on the Loess Plateau in China. Forests 2021, 12, 1209. [Google Scholar] [CrossRef]
  27. Zhang, H.; Song, J.; Li, M.; Han, W.H. Eco-environmental quality assessment and cause analysis of Qilian Mountain National Park based on GEE. Chin. J. Ecol. 2021, 40, 1883–1894. (In Chinese) [Google Scholar] [CrossRef]
  28. Wu, Y.; Liu, T.; Tong, X.; Luo, Y.; Duan, L.; Wang, G. Dynamic evolution analysis of land use (land cover) in Horqin Sandy Land based on long time series landsat data. Acta Ecol. Sin. 2020, 40, 8672–8682. (In Chinese) [Google Scholar]
  29. Zhang, C.; Zhao, Z. Temporal and spatial change of land use/cover and quantitative analysis on the driving forces in the yellow river delta. Beijing Daxue Xuebao Ziran Kexue Ban/Acta Entiarum Nat. Univ. Pekin. 2015, 51, 151–158. (In Chinese) [Google Scholar] [CrossRef]
  30. Li, Z.; Bagan, H.; Yamagata, Y. Analysis of spatiotemporal land cover changes in Inner Mongolia using self-organizing map neural network and grid cells method. Sci. Total Environ. 2018, 636, 1180–1191. [Google Scholar] [CrossRef]
  31. Bagan, H.; Yamagata, Y. Analysis of urban growth and estimating population density using satellite images of nighttime lights and land-use and population data. GIS Sci. Remote Sens. 2015, 52, 765–780. [Google Scholar] [CrossRef]
  32. Kang, Q.; Wang, Z. Land cover change in Wusuli River basin from 1990 to 2015. Bull. Soil Water Conserv. 2020, 40, 312–320. (In Chinese) [Google Scholar]
  33. Ma, E.; Cai, J.; Lin, J.; Han, Y.; Liao, L.; Han, W. Explanation of land use/cover change from the perspective of tele-coupling. Acta Geogr. Sin. 2019, 74, 421–431. (In Chinese) [Google Scholar] [CrossRef]
  34. Tankpa, V.; Wang, L.; Atanga, R.A.; Awotwi, A.; Guo, X. Evidence and impact of map error on land use and land cover dynamics in Ashi River watershed using intensity analysis. PLoS ONE 2020, 15, e0229298. [Google Scholar] [CrossRef] [Green Version]
  35. Chu, H.; Venevsky, S.; Wu, C.; Wang, M. NDVI-based vegetation dynamics and its response to climate changes at Amur-Heilongjiang River Basin from 1982 to 2015. Sci. Total Environ. 2018, 650 Pt 2, 2051–2062. [Google Scholar] [CrossRef]
  36. Wang, X.; Zhu, Y.; Chen, Y.; Zheng, H.; Liu, H.; Huang, H.; Liu, K.; Liu, L. Influences of forest on MODIS snow cover mapping and snow variations in the Amur River Basin in Northeast Asia during 2000–2014. Hydrol. Process. 2017, 31, 3225–3241. [Google Scholar] [CrossRef]
  37. Damien, S.-M.; Josh, M.G.; Abercrombie, S.P.; Friedl, M.A. Hierarchical mapping of annual global land cover 2001 to present: The MODIS Collection 6 Land Cover product. Remote Sens. Environ. Interdiscip. J. 2019, 222, 183–194. [Google Scholar] [CrossRef]
  38. Belward, A.S.; Estes, J.E.; Kline, K.D. The IGBP-DIS global 1-km land-cover data set DISCover: A project overview. Photogramm. Eng. Remote Sens. 1999, 65, 1013–1020. [Google Scholar]
  39. Loveland, T.R.; Belward, A.S. The International Geosphere Biosphere Programme Data and Information System global land cover data set (DISCover). Acta Astronaut. 1997, 41, 681–689. [Google Scholar] [CrossRef]
  40. Hansen, M.C.; DeFries, R.S.; Townshend, J.R.G.; Sohlberg, R. Global land cover classification at 1 km spatial resolution using a classification tree approach. Int. J. Remote Sens. 2000, 21, 1331–1364. [Google Scholar] [CrossRef]
  41. Myneni, R.B.; Hoffman, S.; Knyazikhin, Y.; Privette, J.L.; Glassy, J.; Tian, Y.; Wang, Y.; Song, X.; Zhang, Y.; Smith, G.R.; et al. Global products of vegetation leaf area and fraction absorbed PAR from year one of MODIS data. Remote Sens. Environ. 2002, 83, 214–231. [Google Scholar] [CrossRef] [Green Version]
  42. Running, S.W.; Nemani, R.R.; Heinsch, F.A.; Zhao, M.; Reeves, M.; Hashimoto, H. A Continuous Satellite-Derived Measure of Global Terrestrial Primary Production. BioScience 2004, 54, 547–560. [Google Scholar] [CrossRef] [Green Version]
  43. Bonan, G.B.; Levis, S.; Kergoat, L.; Oleson, K.W. Landscapes as patches of plant functional types: An integrating concept for climate and ecosystem models. Glob. Biogeochem. Cycles 2002, 16, 5-1–5-23. [Google Scholar] [CrossRef] [Green Version]
  44. Cao, M.; Zhang, Y.; Zheng, S.; Chen, L. Accuracy analysis of MODIS land cover data product: A case study of Yellow river source region. Remote Sens. Inf. 2012, 27, 22–27. [Google Scholar]
  45. Liang, D.; Zuo, Y.; Huang, L.; Zhao, J.; Teng, L.; Yang, F. Evaluation of the Consistency of MODIS Land Cover Product (MCD12Q1) Based on Chinese 30 m GlobeLand30 Datasets: A Case Study in Anhui Province, China. ISPRS Int. J. Geo-Inf. 2015, 4, 2519–2541. [Google Scholar] [CrossRef] [Green Version]
  46. Damien, S.-M.; Friedl, M.A. User Guide to Collection 6 MODIS Land Cover (MCD12Q1 and MCD12C1) Product. 2018. Available online: https://lpdaac.usgs.gov/documents/101/MCD12_User_Guide_V6.pdf2018 (accessed on 12 January 2019).
  47. Zhao, J.; Dong, Y.; Zhang, M.; Huang, L. Comparison of identifying land cover tempo-spatial changes using GlobCover and MCD12Q1 global land cover products. Arab. J. Geosci. 2020, 13, 1–12. [Google Scholar] [CrossRef]
  48. Costanza, R.; D’Arge, R.; De Groot, R.; Farber, S.; Grasso, M.; Hannon, B.; Limburg, K.; Naeem, S.; O’Neill, R.V.; Paruelo, J.; et al. The value of the world’s ecosystem services and natural capital. Ecol. Econ. 1997, 25, 3–15. [Google Scholar] [CrossRef]
  49. Chen, Z.; Zhang, X. Value of ecosystem services in China. Chin. Sci. Bull. 2000, 45, 870–876. [Google Scholar] [CrossRef]
  50. Xie, G.D.; Zhen, L.; Lu, C.X.; Xiao, Y.; Chen, C. Expert Knowledge Based Valuation Method of Ecosystem Services in China. J. Nat. Resour. 2008, 23, 911–919. (In Chinese) [Google Scholar] [CrossRef]
  51. Wang, X.L.; Bao, Y.H. Study on the methods of land use dynamic change research. Prog. Geogr. 1999, 18, 81–87. (In Chinese) [Google Scholar] [CrossRef]
  52. Zhang, J.; Liu, Z.; Sun, X. Changing landscape in the Three Gorges Reservoir Area of Yangtze River from 1977 to 2005: Land use/land cover, vegetation cover changes estimated using multi-source satellite data. Int. J. Appl. Earth Obs. Geoinf. 2009, 11, 403–412. [Google Scholar] [CrossRef]
  53. Liu, J.Y. The Land Use in Xizang Autonomous Region; Science Press: Beijing, China, 1992. (In Chinese) [Google Scholar]
  54. Chen, W.; Zhao, X.; Cui, L. Temporal-spatial characteristics of land use in Xinjian County, Jiangxi Province. Chin. J. Appl. Ecol. 2006, 17, 873. [Google Scholar]
  55. Che, Y.; Yang, K.; Chen, T.; Xu, Q. Assessing a riverfront rehabilitation project using the comprehensive index of public accessibility. Ecol. Eng. 2012, 40, 80–87. [Google Scholar] [CrossRef]
  56. Verburg, P.H.; Neumann, K.; Nol, L. Challenges in using land use and land cover data for global change studies. Glob. Chang. Biol. 2011, 17, 974–989. [Google Scholar] [CrossRef] [Green Version]
  57. Gong, P.; Wang, J.; Yu, L.; Zhao, Y.; Zhao, Y.; Liang, L.; Niu, Z.; Huang, X.; Fu, H.; Liu, S.; et al. Finer resolution observation and monitoring of global land cover: First mapping results with Landsat TM and ETM+ data. Int. J. Remote Sens. 2013, 34, 48. [Google Scholar] [CrossRef] [Green Version]
  58. Congalton, R.G.; Gu, J.; Yadav, K.; Thenkabail, P.; Ozdogan, M. Global Land Cover Mapping: A Review and Uncertainty Analysis. Remote Sens. 2014, 6, 12070–12093. [Google Scholar] [CrossRef] [Green Version]
  59. Foody, G.M. Status of land cover classification accuracy assessment. Remote Sens. Environ. 2002, 80, 185–201. [Google Scholar] [CrossRef]
  60. Morisette, J.T.; Privette, J.L.; Justice, C.O. A framework for the validation of MODIS Land products. Remote Sens. Environ. 2002, 83, 77–96. [Google Scholar] [CrossRef]
  61. Song, F.; Yang, Z.; Gao, X.; Dan, T.; Yang, Y.; Zhao, W.; Yu, R. Multi-Scale Feature based Land Cover Change Detection in Mountainous Terrain using Multi-Temporal and Multi-Sensor Remote Sensing Images. IEEE Access 2018, 6, 77494–77508. [Google Scholar] [CrossRef]
  62. Wang, C.; Wang, Y.; Wang, R.; Zheng, P. Modeling and evaluating land-use/land-cover change for urban planning and sustainability: A case study of Dongying city, China. J. Clean. Prod. 2018, 172, 1529–1534. [Google Scholar] [CrossRef]
  63. Zhang, H.; Qi, Z.-F.; Ye, X.-Y.; Cai, Y.-B.; Ma, W.-C.; Chen, M.-N. Analysis of land use/land cover change, population shift, and their effects on spatiotemporal patterns of urban heat islands in metropolitan Shanghai, China. Appl. Geogr. 2013, 44, 121–133. [Google Scholar] [CrossRef]
  64. Yang, J.; Huang, X. The 30m annual land cover dataset and its dynamics in China from 1990 to 2019. Earth Syst. Sci. Data 2021, 13, 3907–3925. [Google Scholar] [CrossRef]
Figure 1. Location and elevation of Heilongjiang (Amur) River Basin (ESRI. 102025, WGS-1984 Albers).
Figure 1. Location and elevation of Heilongjiang (Amur) River Basin (ESRI. 102025, WGS-1984 Albers).
Remotesensing 15 03730 g001
Figure 2. The annual land-cover map series of HARB from 2001 to 2019. Maps for 2001 and 2019 are zoomed in.
Figure 2. The annual land-cover map series of HARB from 2001 to 2019. Maps for 2001 and 2019 are zoomed in.
Remotesensing 15 03730 g002
Figure 3. Change trends in different land-cover types from 2001 to 2019 in HARB.
Figure 3. Change trends in different land-cover types from 2001 to 2019 in HARB.
Remotesensing 15 03730 g003
Figure 4. Estimated area changed by the major land-cover change processes by year in the HARB.
Figure 4. Estimated area changed by the major land-cover change processes by year in the HARB.
Remotesensing 15 03730 g004
Figure 5. The change in land-cover type (a) and landscape (b) during 2001–2019.
Figure 5. The change in land-cover type (a) and landscape (b) during 2001–2019.
Remotesensing 15 03730 g005
Figure 6. The comprehensive land-use dynamic index during 2001–2019.
Figure 6. The comprehensive land-use dynamic index during 2001–2019.
Remotesensing 15 03730 g006
Figure 7. The integrated index of land-use extent during 2001–2019.
Figure 7. The integrated index of land-use extent during 2001–2019.
Remotesensing 15 03730 g007
Figure 8. The quantity of land-use intensity during 2001–2019.
Figure 8. The quantity of land-use intensity during 2001–2019.
Remotesensing 15 03730 g008
Figure 9. The rate of land-use intensity during 2001–2019.
Figure 9. The rate of land-use intensity during 2001–2019.
Remotesensing 15 03730 g009
Figure 10. Spatial distribution of sampling points.
Figure 10. Spatial distribution of sampling points.
Remotesensing 15 03730 g010
Figure 11. Trends in temperature changes in the HARB during 2001–2019.
Figure 11. Trends in temperature changes in the HARB during 2001–2019.
Remotesensing 15 03730 g011
Figure 12. Trends in precipitation changes in the HARB during 2001–2019.
Figure 12. Trends in precipitation changes in the HARB during 2001–2019.
Remotesensing 15 03730 g012
Figure 13. Trends in population and agricultural population ratio changes in the HARB during 2001–2019.
Figure 13. Trends in population and agricultural population ratio changes in the HARB during 2001–2019.
Remotesensing 15 03730 g013
Table 1. Land-use degree classification assignment table.
Table 1. Land-use degree classification assignment table.
TypeUnused LandForest, Grass, WaterFarmlandUrban
Land-cover typeBarrenForest, Grassland,
Water, Wetland
FarmlandUrban
Graded index1234
Table 2. Land-cover types’ change from 2001–2019.
Table 2. Land-cover types’ change from 2001–2019.
Change TypesArea (km2)Ratio (%)Change TypesArea (km2)Ratio (%)
F–GL66,891.253.21B–GL395.750.02
GL–F57,156.752.74WL–GL3810.02
GL–FL35,7411.72W–GL340.50.02
FL–GL19,4700.93FL–W275.50.01
FL–F36930.18GL–B2440.01
F–FL23580.11F–W237.50.01
GL–WL1642.50.08FL–WL218.250.01
GL–W4480.02F–WL209.50.01
FL–U4320.02others10690.05
Table 3. The area of land-cover changes in the Chinese region in 2001, 2010 and 2019.
Table 3. The area of land-cover changes in the Chinese region in 2001, 2010 and 2019.
Land-Cover Type2001201020192001–20102010–20192001–2019
Area/km2Area/km2Area/km2Change Area/km2Change Rate/%Change Area/km2Change Rate/%Change Area/km2Change Rate/%
Forest332,612.75353,607337,32120,994.256.31−16,286−4.614708.251.42
Grassland248,360226,667.75228,852−21,692.3−8.732184.250.96−19,508−7.85
Wetland10471599222555252.7262639.151178112.51
Farmland286,430.5286,120.25299,066.5−310.25−0.1112,946.254.5212,6364.41
Urban6871.2570727421.5200.752.92349.54.94550.258.01
Barren793.25769573.5−24.25−3.06−195.5−25.42−219.75−27.70
Water5699.505979.256354.75279.754.91375.506.28655.2511.50
Table 4. The area of land-cover changes in the Russian region in 2001, 2010 and 2019.
Table 4. The area of land-cover changes in the Russian region in 2001, 2010 and 2019.
Land-Cover Type2001201020192001–20102010–20192001–2019
Area/km2Area/km2Area/km2Change Area/km2Change Rate/%Change Area/km2Change Rate/%Change Area/km2Change Rate/%
Forest702,521.75714,935.75689,516.2512,4141.77−25,419.5−3.56−13,005.5−1.85
Grassland278,223.25267,250.75289,209.75−10,972.5−3.9421,9598.2210,986.53.95
Wetland1167.001159.501575.50−7.5−0.6441635.88408.535.00
Farmland13,369.5011,822.2514,800.25−1547.25−11.57297825.191430.7510.70
Urban3396.503398.753401.752.250.0730.095.250.15
Barren173.00236.25196.0063.2536.56−40.25−17.042313.29
Water11,364.2511,412.0011,515.7547.750.42103.750.91151.501.33
Table 5. The area of land-cover changes in the Mongolian region in 2001, 2010 and 2019.
Table 5. The area of land-cover changes in the Mongolian region in 2001, 2010 and 2019.
Land-Cover Type2001201020192001–20102010–20192001–2019
Area/km2Area/km2Area/km2Change Area/km2Change Rate/%Change Area/km2Change Rate/%Change Area/km2Change Rate/%
Forest2626.752251.502125−375.25−14.29−126.5−5.62−501.75−19.10
Grassland187,139.25187,546187,790.25406.750.22244.250.136510.35
Wetland5.259.59.254.2580.95−0.25−2.63476.19
Farmland236.5288.75201.2552.2522.09−87.5−30.30−35.25−14.90
Urban150.5150.5150.5000000
Barren192.25197.75189.255.52.86−8.5−4.30−3−1.56
Water674.75581.25559.75−93.5−13.86−21.50−3.70−115−17.04
Table 6. Dynamic index of single land cover during 2001–2019.
Table 6. Dynamic index of single land cover during 2001–2019.
F_DsGL_DsWL_DsFL_DsU_DsB_DsW_Ds
Chinese region0.079−0.4366.2510.2450.445−1.5390.639
Russian region−0.1030.2191.9450.5950.0090.7390.074
Mongolian region−1.0610.0194.233−0.8280.000−0.087−0.947
HARB−0.047−0.0613.9570.2600.296−0. 9610. 214
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jia, S.; Yang, Y. Spatiotemporal Characteristics and Driving Factors of Land-Cover Change in the Heilongjiang (Amur) River Basin. Remote Sens. 2023, 15, 3730. https://doi.org/10.3390/rs15153730

AMA Style

Jia S, Yang Y. Spatiotemporal Characteristics and Driving Factors of Land-Cover Change in the Heilongjiang (Amur) River Basin. Remote Sensing. 2023; 15(15):3730. https://doi.org/10.3390/rs15153730

Chicago/Turabian Style

Jia, Shuzhen, and Yaping Yang. 2023. "Spatiotemporal Characteristics and Driving Factors of Land-Cover Change in the Heilongjiang (Amur) River Basin" Remote Sensing 15, no. 15: 3730. https://doi.org/10.3390/rs15153730

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop