Next Article in Journal
Insecticidal Potential of Matricaria chamomilla’s Essential Oil and Its Components (E)-β-Farnesene, Germacrene D, and α-Bisabolol Oxide A against Agricultural Pests, Malaria, and Zika Virus Vectors
Previous Article in Journal
Agronomic Impacts on the Performance of Field Crop Seed (Introduction)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Agricultural Land Use Changes as a Driving Force of Soil Erosion in the Velika Morava River Basin, Serbia

1
Faculty of Geography, University of Belgrade, 11000 Belgrade, Serbia
2
Faculty of Civil Engineering, University of Belgrade, 11000 Belgrade, Serbia
3
Geographical Institute “Jovan Cvijić” of the Serbian Academy of Sciences and Arts, 11000 Belgrade, Serbia
4
Faculty of Forestry, University of Belgrade, 11000 Belgrade, Serbia
*
Author to whom correspondence should be addressed.
Agriculture 2023, 13(4), 778; https://doi.org/10.3390/agriculture13040778
Submission received: 10 February 2023 / Revised: 21 March 2023 / Accepted: 22 March 2023 / Published: 27 March 2023
(This article belongs to the Section Agricultural Soils)

Abstract

:
The erosion potential model was applied to estimate the soil erosion status of rural settlements during the years 1971 and 2011. We used univariate and bivariate local Moran’s I indices to detect and visualize the spatial clustering of settlements with respect to changes in erosion intensity and agricultural land use, as well as their mutual spatial correlation. The study area was differentiated into four statistically significant clusters using the calculated bivariate local Moran’s I indices. The statistical analysis examined the two largest clusters, i.e., the high–high and low–low clusters, and the results of the research indicate that the first four principal components explained 70.50% and 73.47% of the total variance, respectively. In the high–high cluster, the low rates of erosion reduction (average Index Z = 98) in the most significant types of rural settlements were determined according to demographic indicators (i.e., the higher population vitality and population density, the smaller share of the old population and the lower average age of the population) and the large proportion of arable land and Neogene sediments. In the low–low cluster, high erosion reduction rates were detected (average index Z = 64). In this cluster, the more statistically significant influence of natural conditions in combination with demographic–agrarian processes (i.e., the larger share of the old population, the higher average age of the population, the lower vitality index and deagrarization) were decisive factors in changing erosion intensity.

1. Introduction

Soil is comprehensively considered to be an essential resource for human security. Ever since humankind began to engage in agriculture, the main threat to soils has been soil erosion by wind and water [1,2]. The latest United Nations (UN) report on the status of global soil resources highlighted that soil erosion is still a major environmental and agricultural threat worldwide [3,4,5]. Additionally, accelerated forms of soil erosion have become a widespread phenomenon that represent a major barrier to achieving the United Nations’ sustainable development goals [6]. The study of erosion processes at different spatial and temporal scales is not only of socioeconomic and political importance, but also has the obvious scientific outcome of understanding the processes and general behavior of ecosystem dynamics (e.g., erosion rates in the climate change modeling of carbon dynamics, and the prediction of soil nutrient balance or pollution dissemination) [7,8,9,10]. In the early 1990s, it was already estimated that 56% of global land on all continents and under all climatic conditions was degraded and suffered from severe forms of soil erosion by water [11]. The latest research has shown that soil erosion is a global phenomenon [12]. It is widespread in less developed, tropical and subtropical countries, where modeling results have indicated greater exposure to erosive processes [13,14,15,16,17,18,19,20]. It is also widespread in industrialized and highly developed countries [21,22,23,24,25,26,27]. One of the leading causes of excessive soil erosion is the conversion of natural vegetation into agricultural land [28,29], which is important considering that almost 40% of the Earth’s land is currently used for agricultural production [30]. Despite both the social and economic benefits that are brought about by land use transformation [22], it can cause ecological changes [31], effects on people’s health and global climate changes [32,33] and changes in the production, transport and quality of agricultural products [34]. Therefore, the conversion of natural vegetation into agricultural land represents an important factor in the degradation of soil and water quality [14]. For that reason, it is necessary to carry out the spatial–temporal analysis, quantification and identification of the main causes. Yet, in certain regions of Europe, there is also the phenomenon of abandoned agricultural land [35,36,37,38,39,40,41].
Serbia’s rural spaces are heterogenic and devastated to different extents, which makes the planning of multifunctional development extremely complicated [42]. Until now, study areas in Serbia have mostly been focused in mountainous basins [43] and peripheral rural areas with distinct evidence of depopulation and deagrarization [44,45]. Unlike those areas, the Velika Morava River Basin has a high population density with numerous economic and administrative centers and significant agricultural production. The Velika Morava Basin is a “pivot of development”, as well as a part of the pan-European Corridor X. Bearing in mind the geostrategic, transportation, economic and social significance of the Velika Morava River Basin for the whole of Serbia, the estimation of vulnerability to water erosion, as the dominant erosion type in the area, plays an essential role in implementing adequate strategies for sustainable development in Serbia. Since soil erosion is directly linked to agriculture, the subject of this study was primarily a rural area comprising a total of 438 rural settlements.
The main initial hypothesis is based on the view that deagrarization and different demographical processes were essential for understanding the ongoing changes in soil erosion intensity in the Velika Morava River Basin. We have divided the initial hypothesis into two secondary hypotheses: (1) in the areas with deagrarization and depopulation processes, the erosion intensity tends to decrease more; (2) agricultural growth and a higher population concentration are found in the peri-urban areas of the larger urban centres of the Velika Morava River Basin, resulting in a slower decrease in erosion intensity.
The main goals of the study were as follows: (1) the spatio–temporal analysis of changes in erosion intensity; (2) the spatial differentiation of rural settlements, determined by the relationships between agricultural land use changes and changes in soil erosion and the identification of the areas with the most significant changes in erosion intensity caused by changes in agricultural land use; (3) the determination of the influence and order of the dominant factors affecting changes in erosion intensity in the two largest clusters (i.e., the high–high and low–low clusters), based on the most dominant variables controlling sediment and soil erosion (i.e., physical–geographical, agrarian–geographical and demographic variables).
In this study, compared to previous studies, we performed a spatial autocorrelation analysis to determine the relationship between the deagrarization processes and changes in erosion intensity at the rural settlement level. We identified the areas with the most significant changes in soil erosion intensity. The results of this study can be implemented in various projects and national strategies (e.g., rural development, spatial planning, regional development policy, environmental protection, etc.).

2. Materials and Methods

2.1. Study Area

The study area is located in the basin of the Velika Morava River (Figure 1). The basin covers an area of 6734 km2. The basin is located between the Carpathian–Balkan Mountains to the east and southeast and the Serbian–Macedonian mass occupies the central part of the basin, along the valley sides [46]. The topography of the basin is characterized by a hilly and mountainous relief in the southern part of the basin and a lowland relief in the northern part of the basin. The altitude ranges from 67 m to 1334 m, with an average altitude of 260 m. The geological structure is dominated by Holocene alluvium and Neogene lake sediments [47]. According to the Köppen climate classification, the north part of the basin belongs to the Cfa climate group, while the south part of the basin belongs to the Cfb group [48]. The study area has a continental climate [49], with an average annual temperature of 11 °C and a positive significant trend of average annual temperature across meteorological stations [50,51]. The basin has between 590 mm and 670 mm of annual precipitation and has a semi-humid and humid climate, according to the De Martonne aridity index [52]. Early studies showed similar results (600–650 mm) [53]. The monthly cloudiness values are below average for Serbia [54]. Across the majority of the hydrological stations in the basin, a negative non-significant trend of mean annual discharge has been recorded [55]. However, the basin can experience torrential rain, and 226 floods were registered over 105 years, resulting in 13 casualties [56]. The maximal discharges have a seasonal character (from May to the first half of June) [57]. The study area has a specific runoff that is below the national level (q < 5 l/s/km2) [58]. There are potential flood zones within the valley, while the edges are at risk of seismic movements and landslides [59,60,61]. According to the WRB classification (World Reference Base), the main soil types in the basin are Eutric Cambisol and Vertisol. Fluvisol and Fluvisol humic are also represented in the alluvial plains of Velika Morava. Leptosol can be found in the highest eastern parts of the basin, while in the mountainous areas to the southwest, Eutric Leptosol is widespread [62].
The population in the Velika Morava River Basin lives in 17 urban and municipal centers and 438 rural settlements. The urban centers are mainly located in the valleys of the Velika Morava River (along the main traffic route, i.e., the pan-European Corridor X) and its larger tributaries. In terms of numbers, small urban settlements prevail: eight settlements have fewer than 10,000 inhabitants and two have between 10,000 and 20,000 inhabitants. There are only six medium-sized towns, with populations ranging between 20,000 and 100,000 inhabitants, while the largest urban center is Kragujevac, with approximately 150,000 inhabitants. The total population has increased from 770,698 people in 1961 to 801,471 people in 2011. However, the total rural population has decreased by 29%, from 559,836 people to 397,730 people [63]. According to the 2012 census, agricultural land covers 286,866 ha in the rural settlements. In terms of the agricultural land, the share of arable land is 79% (227,609 ha). The largest part of arable land is used to grow cereals (68%), with dominant shares of corn (49%) and wheat (38%) and a smaller share of other cereals, e.g., barley, rye and oats (13%) [63].

2.2. Erosion Potential Model

The Gavrilović model, also known as the erosion potential model (EPM), was developed by Slobodan Gavrilović following erosion field research in different catchment areas in Serbia in the middle of the 20th century. Gavrilović proposed an analytical equation for determining the total annual sediment production [64]. In the Balkans, the erosion potential model (EPM) is the preferred method for mapping soil erosion intensity at different scales (i.e., national, regional, local, etc.) [65,66,67,68,69,70,71]. According to an evaluation of different methods, the EPM model is the most suitable at the watershed level for engineering purposes and watershed management in southeastern European countries [72]. Research in other countries has shown that the results obtained using this method are in agreement with field observations. Therefore, the method has been widely accepted in regions all over the world [73,74,75,76,77,78].
The erosion potential model was used to calculate the soil erosion intensity, and is expressed by Equation (1):
W = T H π Z 3 F
where W is the total annual gross erosion (m3/yr), T is the temperature coefficient (T = √(t/10 + 0.1)), t is the mean annual air temperature (°C), H is the mean annual precipitation (mm), Z is the erosion coefficient and F is the watershed area (km2).
The EPM model uses a scoring approach for three descriptive variables to calculate the coefficient of erosion (Table 1): the coefficient of soil resistance (Y), the coefficient of soil protection (X) and the coefficient of the type and extent of erosion (φ). The EPM erosion categorization is shown in Table 2. The erosion coefficient (Z) was calculated using Equation (2):
Z = Y X ( φ + I )
where Y is the coefficient of soil resistance, X is the soil protection coefficient, φ is the erosion and stream network development coefficient and I is the average slope (%).
The Y, X and φ coefficients are dimensionless parameters. Y is the coefficient of soil erodibility, which depends on the lithological characteristics of the watershed and indicates the resistance of soil to erosion. The values of this coefficient range between 0.1 and 1; values close to 0.1 indicate low erodibility, whereas values close to 1 represent strong erodibility. In this study, we used basic 1:100.000 geological maps of the study area to evaluate the Y coefficient. X represents the coefficient of soil protection, which is based on vegetation cover and land use within the catchment area. It varies from 0.05 to 1; values close to 0 indicate low soil protection, while values close to 1 indicate high soil protection. φ is a coefficient that depends on active erosion and the degree of extension of linear erosion. It includes the geomorphological features of the terrain and various erosive–accumulative processes. This dimensionless coefficient has values ranging from 0.1 to 1.
Previous studies [44] on the importance of certain variables in the EPM model have shown that climatic parameters (e.g., precipitation and temperature) do not significantly affect soil erosion. Specifically, correlation matrix results at the statistical significance level of α = 0.05 for function W = f(H, T, Z, X, Y, φ, I, F) showed that the erosion coefficient Z had a major impact on soil erosion intensity (W = f(Z); r = 0.967). Out of the four variables that determine erosion, the erosion coefficient Z and the coefficient of soil protection X were found to be the primary controlling factors in erosion intensity. According to mathematical interactions between the different variables, the correlation coefficients for the functions Z = f(X), Z = f(φ) and Z = f(Y) are r = 0.820, r = 0.767 and r = 0.392, respectively. Given that results have shown that there were no significant changes in the trend of annual precipitation for the period of 1961–2009 in Serbia [79], the focus of this study was on determining the factors that control soil erosion from the perspective of anthropogenic influences, as well as physical–geographical variables that largely determine erosion intensity and are included in the calculation of the erosion coefficient Z.

2.3. Spatial Autocorrelation Analysis

The effects of land use/land cover changes on soil erosion processes have been documented in many studies [41,80,81,82,83,84,85,86]. In our research, we additionally focused on the spatial heterogeneity and spatial distribution of agricultural land use changes, erosion intensity changes and their mutual spatial relationships.
Moran’s I [87,88] is a commonly used indicator for measuring the spatial patterns of geographical phenomena:
I = n Σ i = 1 n Σ j = 1   n W i j   ( X i X ¯ ) ( X j X ¯ ) Σ i = 1 n Σ j = 1 n   W i j   Σ i = 1 n ( X i X ¯ ) 2  
where n represents the number of spatial units (i.e., settlements in our case) in the study area, Xi and Xj are the measured or calculated values of a certain phenomenon or variable at location i and j, respectively, X ¯ is the calculated average value and Wij represents a particular element of a spatial weighted matrix, which is calculated according to the spatial proximity between two locations (in our case, this was calculated using the inverse distance method).
In addition, the local indicator of spatial autocorrelation (LISA) is commonly used to determine spatial heterogeneity and distinguishing local clusters [86]:
I i = ( n 1 ) ( X i X ¯ ) j = 1 , j i n ( X j X ¯ ) 2 j = 1 , j i n W i j ( X j X ¯ )
where Ii is the local Moran’s I index of a spatial unit in location i, which indicates the correlation between that unit and its neighborhood, thereby offering insights into how the spatial units with high and low attributed values are clustered. A high–high cluster indicates that spatial units with high values of the considered variable are surrounded by other units with high values. On the contrary, a low–low cluster designates that units with low values are surrounded by neighbors with low values.
We used bivariate local Moran’s I [89], which recognizes whether the values of one variable (X) in a given spatial unit are correlated with the values of a second variable (Y) in the neighboring spatial units. The formula for bivariate local Moran’s I for the unit k is similar to the univariate version:
I X Y k = Z X k l = 1 n W k l Z Y l
where I X Y k is the standardized value of X at unit k, Z Y l is the standardized value of Y at unit l and Wkl is the spatial weight between units k and l.
In our study, bivariate spatial autocorrelation analysis was used to determine and visualize the rural areas in the Velika Morava River Basin that have experienced the most significant changes in erosion intensity (Index Z) due to changes in agricultural land use (Index AgL). Both the univariate and bivariate local Moran’s I indices were calculated and visualized using open-source GeoDa software that was designed for exploratory spatial data analysis (ESDA) [90].

2.4. Control Variables as Determinants of Changes in Erosion Intensity

After establishing the spatial differentiation of rural settlements using the function Index Z = f (Index AgL), the next aim of this research was to determine the influence and order of the dominant factors affecting changes in erosion intensity. In this context, the two largest clusters were analyzed, i.e., the high–high and low–low clusters. This study focused on examining the combined effects of the selected physical–geographical variables and anthropogenic activities as the most dominant factors controlling sediment and soil erosion.
As presented in Table 3, 17 variables that could potentially affect changes in erosion intensity were taken into consideration. The variables were grouped into three classes: physical–geographical, agrarian–geographical and demographic.
The index was classified according to the following scale: high index < 10; medium–high index = 10–30; medium index = 30–50; medium–low index = 50–70; low index > 70–100 and growth index > 100.
The data for the Z coefficient for 1971 were obtained from the 1:500,000 erosion map of Serbia [91]. To obtain this data for 2011, we used the CORINE land cover database (2012), which was published by the European Environment Agency (EEA) [92]. The coefficient of the erosion type was determined using satellite images from the Landsat 8 satellite, which belongs to the Geological Topographic Institute of the United States (USGS) [93,94,95,96,97,98]. Our results were obtained from all of the gathered data and maps. The data for the mean altitude (Aav) and terrain slope (I) were obtained from the 25–m digital elevation model over Europe (EU–DEM) [99]. The lithological data were obtained from the digitalization of a 1:100,000 basic geological map [100]. The vegetation data were obtained from the CORINE land cover database (2012), which was published by the European Environment Agency (EEA) (Copernicus Land Monitoring Service). The demographic and agrarian–geographic indicators were identified from data analyses by the Statistical Office of the Republic of Serbia for 1961 and 2011 [63], respectively. QGIS 3.8.0. was used for the data analysis, synthesis and visualization.
The determination of the controlling factors among the considered variables that contribute to changes in erosion intensity was carried out using two statistical tools: correlation analysis and principal component analysis (PCA). All the statistical calculations were implemented using the open-source R statistical computing environment with the stats package [101].
In this study, we identified 17 indicators that could provide insights into mutually conditioned relationships, including erosion intensity, natural conditions, land use, deagrarization intensity, human pressure on the land, the degree of depopulation and the demographic vitality of the population. However, the main issue with identifying the typological features is the huge number of variables that determine the attributes, characteristics and features of the area being studied. Based on previous multivariate analysis studies, the principal component analysis algorithm has proven to be the most convenient method for the typologization of the relationship between erosion intensity and agricultural production [43,102,103,104,105]. According to the above-mentioned studies, factor analysis starts from the assumption that a high level of correlation between two or more variables leads to their replacement by a common indicator, thereby providing insights into the unambiguous quantitative formulation of hidden structures via the definition of new indicators.
Correlation coefficient matrices should not be singular, which is why it is necessary to calculate the Kaiser–Meyer–Olkin (KMO) indicator of a sample’s adequacy and also carry out Bartlett’s sphericity test. KMO values range from 0.0 to 1.0; however, a score of 0.50 is the suggested minimum value for a good PCA [102,106,107,108]. In this study, the KMO values for the high–high and low–low clusters were 0.64 and 0.67, respectively. Likewise, to perform the factor analysis successfully, Bartlett’s sphericity test must be significant, as well (i.e., p < 0.05). The statistical significance obtained in this study was p < 0.0001. We also applied the Ward method for hierarchical grouping using the squared Euclidean distance [43,103].

3. Results

3.1. Spatio–Temporal Analysis of Changes in Erosion Intensity

Our comparative quantitative analysis of the two-time series showed a decrease in erosion intensity in the Velika Morava River Basin (Table 4). The results of the EPM model showed that the specific annual gross erosion was W1 = 1013 m3/km2/yr in 1971 and W2 = 747 m3/km2/yr in 2011. The average erosion coefficient in 1971 was Z1 = 0.529, while it was Z2 = 0.420 in 2011. This meant that during the period of 1971–2011, the intensity of soil erosion decreased by 20.6%. According to Table 4, 2219 km2 (33%) of the study area was classified as erosion category V and IV in 1971, while 3055 km2 was classified as these two categories in 2011 (45.4%). The results also indicate that 2093 km2 (31.1%) was classified as erosion category III in 1971, while 2580.5 km2 was classified as this category in 2011 (38.3%). Finally, 2422 km2 (36%) of the study area was classified as erosion category I and II in 1971, while 1098 km2 was classified as these two categories in 2011 (16.3%). It could be concluded that the biggest change lies in the reduction in areas with an excessive and intensive risk of erosion (54%). On the other hand, there was a 38% increase in areas with a weak and very weak risk of erosion, as well as a 19% increase in areas with a medium risk of erosion.

3.2. Erosion Intensity Characteristics in Rural Settlements

The changes in erosion intensity in rural areas were analyzed for 1971 and 2011. For each rural settlement, the average erosion coefficient Z was determined for 1971 (Z1) and 2011 (Z2) (Figure 2A,B). As presented in Figure 2C, only 4% of the total number of rural settlements in the Velika Morava River Basin had a medium Index Z, whereas 37% and 47% of rural settlements had a medium–low and low Index Z, respectively. Finally, 13% of the rural settlements recorded growth in terms of soil erosion intensity. The local Moran’s I statistics for Index Z identified two clusters that represented two distinctively different areas (Figure 2D).
The univariate LISA cluster map singled out a large compact high–high cluster of rural settlements that gravitated toward larger urban centers in the north of the Velika Morava River Basin. Meanwhile, a large low–low cluster in the southern part of the basin indicated a decrease in soil erosion intensity over the period of 1971–2011.

3.3. Agricultural Land Use Characteristics in Rural Settlements

The basic agrarian characteristic of the Velika Morava River Basin is the rapid reduction in agricultural land. During the research period of 1961–2011, the total area of agricultural land decreased by 38% (1961 = 462,515 ha; 2011 = 286,866 ha). Out of 438 rural settlements, the largest proportion of settlements (38%) comprised 500–1000 ha of agricultural land in 1961 (Figure 3A), whereas the largest proportion of rural settlements (37%) consisted of 200–500 ha of agricultural land in 2011 (Figure 3B).
According to Figure 3C, 7% of the rural settlements had a high and medium–high deagrarization index, 26% had a medium deagrarization index, 33% had a medium–low deagrarization index and 30% had a low deagrarization index, while only 4% experienced an increase in agricultural land. The mapped agrarian univariate LISA indicators (Figure 3D) indicated clusters of strong deagrarization areas (low–low), i.e., areas with low deagrarization index values, in the southern part of the Velika Morava River Basin. The cluster of high deagrarization index values and agricultural growth were expressed as a high–high cluster, which were dominant in the northern and central parts of the basin. The clusters primarily reflected a relatively small increase in agricultural land in the area between 1961 and 2011.

3.4. Spatial Differentiation of Rural Settlements: Impact of Agricultural Land Use Changes on Changes in Erosion Intensity

The impact of deagrarization on changes in the intensity of soil erosion in the rural area of the Velika Morava River Basin was demonstrated by the bivariate LISA indicators (Figure 4). Based on the bivariate LISA values of Index AgL and Index Z, the spatial patterns of four clusters were detected. In general, there was a difference between the northern part of the basin (with high–high and low–high clusters) and the southern part of the basin (with low–low and high–low clusters).
The high–high cluster is comprised of rural areas characterized by small changes in erosion and deagrarization intensity (Figure 4). Out of the 88 rural settlements within the Velika Morava River Basin, around 20% belong to this cluster. In terms of space, these rural settlements represent single units of urban centers in the northern and central parts of the basin (Mladenovac, Požarevac, Sopot, Aranđelovac, Topola, Smederevska Palanka and Velika Plana). During the analyzed period, there were no changes in the soil erosion category within this cluster. To be more precise, the erosion coefficient remained in the same category of medium erosion (1971: Z1 = 0.550; 2011: Z2 = 0.529). The average value of the soil erosion index was high (Index Z = 98) and ranged from 49 to 135. Actually, in 42 rural settlements (48%), the value of this index was above 100, which implies an increase in erosion intensity. Generally speaking, the increase in soil erosion intensity was not large and in most of the settlements, it was up to 15% compared to the values from 1971.
There was not a distinct deagrarization process. The mean value of the deagrarization index in the settlements in this cluster was Index AgL = 83. The rural population in this part of the basin owned a total of 130,741 ha of agricultural land in 1971. The biggest number of settlements (37%) comprised 1000–2000 ha of agricultural land (Figure 3A). On the other hand, the population cultivated 106,113 ha of agricultural land in 2011. The largest proportion of settlements comprised 1000–2000 ha of agricultural land (36%) (Figure 3B). According to Figure 3C, 23% and 69% of rural settlements had a medium–low and low deagrarization index, respectively. In the other settlements, there was a slight increase in the area of agricultural land.
The studied area is characterized by intensive agriculture and the use conventional tillage which further increases the intensity of erosion. In addition, a high percentage of the working age population engaged in agriculture in the area allows the improvement of the production process with intensive and frequent tillage interventions and performing work-intensive jobs, which has a direct impact on increasing the intensity of erosion [45,109,110].
The low–high cluster was characterized by rural settlements that experienced higher deagrarization intensity, followed by small changes in erosion intensity. This cluster included 38 settlements, which was 9% of the total number of settlements in the Velika Morava River Basin. These settlements did not have a unique distribution and were mostly surrounded by settlements in the high–high cluster (Figure 4). Just as in the previous cluster, only relatively small changes in soil erosion intensity were detected among the settlements in this cluster. The mean erosion coefficients in 1971 and 2011 were Z1 = 0.529 and Z2 = 0.456, respectively, with an average Index Z = 87. Therefore, over time, the erosion intensity of settlements in this cluster remained within the same category.
However, the deagrarization process demonstrated a more severe decline. Agricultural land reduced by over 50% (Index AgL = 46) over the study period. The total area of agricultural land in all of the settlements covered 43,593 ha in 1971. All in all, 58% of rural settlements comprised 1000–2000 ha of agricultural land (Figure 3A). The relatively high intensity of soil erosion compared to the deagrarization intensity could be explained by the positions of the settlements. The rural settlements in this cluster were in the territory around the peri-urban belt (i.e., rural areas near larger urban settlements), which was a direct consequence of the intensification of agricultural production under the influence of the agrarian markets in the surrounding urban centers [110,111,112].
The low–low cluster was the largest cluster among the rural areas in the Velika Morava River Basin and was characterized by the largest changes in both erosion and deagrarization intensity (Figure 4). There were 132 rural settlements in this cluster, which was 30% of the total number of settlements in the Velika Morava River Basin. This cluster covered the southern part of the basin, which primarily includes mountainous areas and rural settlements in gravitational zone of urban centers, such as Kragujevac, Paraćin, Ćuprija, Jagodina and Despotovac to the north and Ćićevac to the south. The average erosion coefficient in 1971 was Z1 = 0.516, whereas it was Z2 = 0.327 in 2011. This meant that during the study period, the medium erosion level was replaced by low/weak erosion. The average erosion index value was Index Z = 64. Half of the rural settlements had Index Z below the average value, which meant that in these rural areas, the erosion intensity reduced by as much as 75% compared to that in 1971.
The reduction in soil erosion intensity was a consequence of the rather intense deagrarization process in the rural areas. The average deagrarization index (Index AgL = 41) indicates that there was an observable trend of abandoning agricultural land within this cluster. Specifically, most of the rural settlements (55%) had a medium deagrarization intensity, whereas 18% of the settlements had a medium–high and high deagrarization index (Figure 3C). The losses of available agricultural land were more than evident. In total, all the settlements within this cluster comprised a total of 122,154 ha of agricultural land in 1971. The highest number of settlements (45%) consisted of 500–1000 ha of agricultural land (Figure 3A). Over the period of 40 years, the area of agricultural land reduced to a total of 50,794 ha, with a demonstration of the very prominent fragmentation of agricultural land. According to Figure 3B, half of the rural settlements comprised 200–500 ha of agricultural land.
The studied area is characterized by the process of deagrarization, in which, in addition to the absence of intensive practices, we can observe agricultural land use change, which is consistent with the prevailing representation of the autarkic type of agricultural production. Agricultural practices are less intensive, which has a direct impact on the reduction of production productivity and, consequently, on the reduction of erosion intensity.
The high–low cluster was characterized by lower deagrarization intensity, followed by larger changes in erosion intensity. In terms of space, the settlements within this cluster did not represent a single whole, but were instead aerially grouped and were generally positioned around the settlements in the low–low cluster (Figure 4). Out of the total number of rural settlements in the Velika Morava River Basin, this cluster consisted of 79 settlements (11%). The spatial distribution of the settlements indicated that they mostly included urban centers in the central part of the basin (e.g., Kragujevac, Jagodina, Ćuprija and Paraćin). This cluster experienced a significant reduction in soil erosion intensity (Index Z = 69). The medium erosion coefficient in 1971 was Z1 = 0.696, whereas it was Z2 = 0.483 in 2011. The distribution of the Z index was in the range of 35–102 and the predominant erosion index was below the average value.
Concerning the reduction in soil erosion, the deagrarization process was also of lower intensity (Index AgL = 79). Low deagrarization dominated the settlements in this cluster (Figure 3C). The relatively low reduction in agricultural holdings also affected the relatively large reduction in soil erosion intensity. These settlements were mainly located at lower altitudes where the dominant Neogene sediment was vulnerable to the forceful processes of erosion and denudation. On the other hand, the economic transformation caused an increase in demand for non-agricultural occupations around the urban centers [106], which led to the fragmentation of agricultural holdings [110,112,113,114,115,116]. This was reflected in the unfavorable demographic features of the population (i.e., the population decrease of 20%, followed by the process of population aging) [117].

3.5. Geographic Indicators of Changes in Soil Erosion Intensity

In this study, we applied the statistical analysis method to determine the effects and sequences of the dominant factors affecting changes in soil erosion intensity. With this in mind, two bivariate clusters were analyzed: the high–high and low–low clusters. These were the most dominant clusters, according to their features and the catchment of rural settlements. The mathematical interactions between various variables were described using a correlation matrix (Table 5). Generally, the results indicated high correlations between the variables at a significance level of α = 0.05.
In the high–high cluster, the correlation values ranged from r = −0.951 to r = 0.980. The analysis of the results showed that the effects of the physical–geographical indicators on erosion intensity were lower in this cluster than in the low–low cluster. Out of the physical–geographical factors, Z2 only had a relatively high negative correlation with I (r = −0.445) and FC (r = −0.405). This meant that with an increase in the inclination angle, there was an increase in the forest cover and a decrease in erosion intensity. The very strong ratio of NSA= f(Aav) to NSA = f(I) showed the direct correlation between the lithological complex NSA and the altitude and inclination angle. Additionally, the results showed that there were positive correlations between Z2 and Index ArL and between Z2 and ArL (r = 0.473 and r = 0.583, respectively). By the same token, it could be concluded that erosion intensity was affected by changes in arable land with a particular influence on the arable land share within the total area of agricultural land. Along with the deagrarization of agricultural land, there was simultaneously an ongoing deagrarization process of the arable land, as there was a positive correlation (r = 0.711) for Index ArL = f(Index AgL). Additionally, there were noticeable effects of physical–geographical factors on the share of arable land within the total area of agricultural land. This further implied the higher prevalence of arable land on the lithological complex of NSA, located at lower altitudes and on terrain with lower inclination angles. There was a similar situation in this cluster as in the low–low cluster: SDArL was directly linked to GDAgL and showed a high level of correlation (r = 0.932). However, the agrarian pressure on the land (SDAgL and GDAgL) was exclusively determined by demographic indicators (i.e., the Index RP, DRP, Index V and Index H). The balanced coefficients of correlation pointed out that these indicators had almost identical effects on SDAgL and GDAgL, which was not the case with the low–low cluster. The deagrarization process was also defined by demographic indicators (i.e., DRP, Index V, RPav, ORP and Index H), as indicated by the relatively high correlation coefficient values (r > 0.546). The highest correlation was between Index P and Index H (r = 0.980), which meant that Index H was the most significant of the demographic indicators. This was also not the case with the low–low cluster. The main characteristic of this cluster was the direct correlations between the agrarian land pressure indicators and almost all the demographic indicators, as well as the statistically weaker effect of the physical–geographical indicators in comparison to the agrarian–geographical indicators. This meant that the erosion intensity was exclusively determined by agrarian–geographical and demographic changes.
The order of significance of the variables was determined by the magnitude of their eigenvalues, as presented in Figure 5. Eigenvalues explain the percentage of variance and the cumulative variance of the principal components. It is evident that the first point at which it is possible to carry out the reduction of the number of factors is between the third and fifth factor. In this study, the first four principal components explained 70.50% of the total variance.
The data from the PCA were subjected to varimax rotation. After the varimax rotation, four factors affecting changes in the intensity of soil erosion were identified (Table 6). The Factor 1 component explained 26.23% of the soil erosion variance within the dataset. Factors 2 and 3 explained 47.51% and 58.90% of the soil erosion variance, respectively. The addition of Factor 4 increased the model-explained variance to 70.50%.
Based on the obtained results for the factor scores following the varimax rotation (Table 7), the typology of the rural settlements was established according to the dominant variables for changes in soil erosion intensity. There were four types of rural settlements: Factor 1, demographic type; Factor 2, agricultural type of specific physical–geographical characteristics; Factor 3, agrarian type; Factor 4, anthropopression type. The order of significance of the basic indicators in the high–high cluster indicated that the demographic component was the most significant for rural settlements and was essential for understanding ongoing changes in soil erosion intensity.
Factor 1: Demographic type. This type included 30% of the rural settlements in the high–high cluster. This component consisted of the Index P, DRP, Index V, RPav, ORP and Index H variables. The demographic type comprised settlements that experienced positive changes in population numbers and household number, followed by a higher rural population density. That is why the settlements with the highest increase in rural population (Index RP = 125–289) and households number (Index H = 165–371) exhibited the highest anthropogenic pressure on the rural area (DRP = 166–249 rural population/km2). Likewise, this component had the highest positive correlation with Index V (r = 0.951). In some of the settlements, the values of Index V were twice as large as the economically active population in comparison with the older population. On the other hand, this settlement type was inversely proportional to RPav and ORP, which was proven by their high negative correlations (r = −0.831 and r = −0.909, respectively). The average population age ranged from 39.5 to 47.5 years old. In the settlements with the highest demographic growth, the share of the elderly population was 15%. With this in mind, it could be concluded that the rural areas were determined by positive demographic changes, with high population vitality. Index V had a negative correlation with the average population age (RPav) and the share of elderly people within the overall rural population (ORP).
Factor 2: Agricultural type of specific physical–geographical characteristics. This was the most common settlement type, as it comprises 42% of all rural settlements within the cluster. This type of rural settlement consisted of the Aav, I, NSA and ArLs variables. This factor had the highest negative correlations with altitude and inclination angle, whereas its highest positive correlations were with the lithological complex NSA and ArL. This meant that the changes in the altitude and inclination angle of the terrain were inversely proportional to the presence of the lithological complex NSA. The share of the lithological complex NSA was calculated from the share of arable land within the overall area of agricultural land. Therefore, this type of rural settlement was characterized by a higher arable land share within the overall area of agricultural land and was naturally predisposed to lower altitudes and shallower inclinations angles, as well as higher shares of the lithological complex NSA. The dominance of arable land within the overall area of agricultural land (ArL > 90%) was only present in settlements with an NSA > 90%. Additionally, these settlements were located at altitudes Aav < 100 m and on slightly inclined terrains (I ≤ 2°). On the other hand, the settlements at higher altitudes (Aav > 300 m) and on steeper inclinations (I > 80) were characterized by the almost complete absence of the lithological complex NSA (NSA < 2%).
Factor 3: Agrarian type. This type included 16% of all rural settlements in this cluster. This component consisted of Index AgL, Index ArL, Z2 and FC variables. This component only had a negative correlation with the share of forest cover, whereas it had positive correlations with all of the other variables. This meant that erosion intensity Z2 was higher in rural settlements with lower shares of forest cover and with a higher index of agricultural and arable land. Therefore, this type included rural settlements where the intensity of erosion was primarily determined by the deagrarization intensity of agricultural and arable land, as well as forest cover. For instance, these rural areas were characterized by low average shares of forest cover (FC = 7%) and the prevalence of settlements with below average forest cover. Therefore, these were the settlements where deagrarization was not expressed, as indicated by the average values of Index AgL = 95 and Index ArL = 106. Positive changes were more evident in the use of arable land, as half of the rural settlements of this type exhibited an 36% increase in Index ArL.
Factor 4: Anthropopression type. This type included 13% of the settlements in this cluster. This component consisted of the GDAgL and SDArL variables. This component had a highly positive correlation with agrarian densities, meaning that agrarian pressure on agricultural land (GDAgL = 286 rural population/100 ha) was followed by agrarian pressure on arable land (SDArL = 370 rural population/100 ha).
In the low–low cluster, the correlation values ranged from r = −0.951 to r = 0.998 (Table 8). The results of the correlation matrix showed that Z2 had the highest negative correlation with F (r = −0.789). Likewise, Z2 had rather high negative correlations with I (r = −0.782) and Aav (r = −0.725). The very strong ratios between Aav = f(I), Aav = f(FC) and I = f(FC) indicated that with the increase in altitude, the inclination angle of the terrain also increased, as well as forest cover. At the same time, this meant that Z2 decreased. Z2 had a relatively high positive correlation with NSA (r = 0.498). The results also show that the ratios between NSA and I and between NSA and Aav had negative values (r = −0.552 and r = −0.428, respectively). This further implied that Z2 was higher on the lithological complex NSA, which was exclusively located at lower altitudes and on terrains with smaller inclinations angles within the basin. This conformed with the results of previous research [43]. Additionally, the results showed that the correlations between Z2 and Index AgL and between Z2 and ArLs had positive values (r = 0.409 and r = 0.602, respectively). With that in mind, it could be concluded that the erosion intensity was primarily affected by the share of arable land within the total area of agricultural land. Along with the deagrarization of agricultural land, the deagrarization of arable land also occurred, since Index AgL had a positive correlation with Index ArL (r = 0.685). SDArL was directly linked to GDAgL, which was shown by the high correlation level (r = 0.998) for SDArL = f(GDAgL). The similar negative correlation coefficient values between DRP and Aav, I and F indicated that these physical–geographical factors had an almost identical effect on DRP. On the other hand, the results analysis showed that other demographic indicators had a high mutual interdependence, which was why depopulation was directly linked to changes in household number (r = 0.987) and less linked to changes in the average size of households (r = 0.455). There was also a direct connection between RPav and ORP since there was a high positive correlation (r = 0.936) for RPav = f(ORP). Additionally, with the increase in the average population age and the share of the elderly population, there was a decrease in both the demographic vitality index and the average size of households. These results conformed with the results of previous research.
The order of significance of these ”aria’les was determined by the magnitude of their eigenvalues, as presented in Figure 6. Eigenvalues explain the percentage of variance and the cumulative variance of the principal components. It is evident that the first point at which it is possible to carry out the reduction of the number of factors is between the third and fifth factor. In this study, the first four principal components explained 73.47% of the total variance.
The data from the PCA were subjected to varimax rotation. After the varimax rotation, four components affecting changes in the intensity of soil erosion were identified (Table 9). The first component explained 25.23% of the soil erosion variance within the dataset. The second and third components explained 47.47% and 60.72% of the soil erosion variance, respectively. The addition of the fourth component increased the model-explained variance to 73.47%. The square cosines of the variables indicated the best-described variables for each principal component.
Based on the obtained results for the factor scores after the varimax rotation (Table 10), the typology of rural settlements was established, according to the dominant variables for changes in soil erosion intensity. There were four types of rural settlements: Factor 1, agricultural type of specific physical–geographic characteristics; Factor 2, demographic–agrarian type; Factor 3, anthropopression type; and Factor 4, population type. The order of significance of the basic indicators in the low–low cluster indicated that changes in agricultural land use and soil erosion were exclusively the result of the common effects of the specific physical–geographical indicators.
Factor 1: Agricultural type of specific physical–geographic characteristics. This type was the most common and included 52% of the rural settlements in this cluster. This component consisted of the Z2, Aav, I, NSA, FC, Index AgL, ArLs and DRP variables. Based on the variable values for this type, it could be concluded that the erosion intensity was primarily affected by physical–geographical conditions (i.e., the Aav, I, F and NSA indices). This component had positive correlations with Index AgL, ArL and DRP, which meant that the interdependence of these variables also affected soil erosion intensity. This was why the erosion intensity was higher in places with a lower deagrarization intensity of agricultural land, whereas the share of arable land within the area of agricultural land was higher, along with the pressure of the rural population in the settlements. This correlation was best illustrated by the following example. The lowest erosion intensity (Z2 = 0.133) was characteristic of rural settlements located in mountainous areas at the edge of the Velika Morava River Basin. These rural settlements were at an average altitude Aav = 630 m, with an average terrain inclination angle I = 16°. Forest cover occupied 80% of the settlement area and the share of the lithological complex NSA was 3%. In rural areas with these kinds of natural conditions, there was an intense deagrarization process with a 64% reduction in agricultural land.
Factor 2: Demographic–agrarian type. This type of rural settlement consisted of the Index ArL, Index V, RPav, ORP and Index Hs variables. The demographic–agrarian type comprised 36% of the rural settlements in this cluster. There was a high negative correlation between RPav and ORP and positive correlations between Index V, Index Hs and Index ArL. The fall in the average size of households was accompanied by lower demographic vitality and a higher average age of the population and a higher share of elderly people. The ratios between the demographic variables conformed with the deagrarization intensity of arable land. By the same token, the deagrarization of arable land was higher in settlements of this type. Generally speaking, negative demographic tendencies were a main feature of most rural settlements of this type. Within the analyzed period, there was a decrease in the average household size (Index Hs = 62). Additionally, the average population age was RPav = 50.5 years and 46% of the settlements had a population that was older than the average. A third of the rural population consisted of people older than 65 years of age and the average vitality index demonstrated the dominance of the older population in comparison to the economically active population.
Factor 3: Anthropopression type. Out of the total number of rural settlements in this cluster, only 4% belonged to this type. The agrarian pressure on the land variable characterized these rural settlements, as shown by the high positive correlation between GDAgL and SDArL. This was a consequence of the extreme deagrarization of agricultural and arable land in some of the settlements of this type, which was an indicator of the quite small area of agricultural land and the underdeveloped agricultural production. Still, some settlements were determined by these variables and exerted a real agrarian pressure on the land due to the sustainability of both the rural population and the area of agricultural land.
Factor 4: Population type. This type of rural settlement consisted of the Index P and Index H variables. The population type comprised 9% of the rural settlements in this cluster. The high positive correlations between these variables categorized the rural settlements according to depopulation intensity and the change in household numbers within the cluster. This meant that depopulation intensity was lower in places where the household number decrease was smaller. The most prominent rural area in this cluster was a part of the peri-urban belt around Kragujevac. A positive demographic trend was followed by a high multifold increase in the population and household numbers (Index RP = 358, Index H = 379, respectively). Additionally, the main feature of these settlements was the intense and balanced deagrarization of both agricultural and arable land (Index AgL = 33 and Index ArL = 33, respectively).

4. Discussion

The clear differentiation of the study area along the north–south direction undoubtedly indicated the dynamic interactions between natural conditions, demographic and settlement indicators and changes in land use. We observed the broader context of the historical, political and socio–economic patterns during the analyzed period. The historical conditions of agricultural development during the 1960s were determined by the appearance of agricultural cooperatives, which enabled the application of modern technologies via cooperative relationships and contributed to the increase in agricultural production and productivity [118]. However, after that period, the further evolution of rural settlements was directed towards the planned centralized industrialization of socialist Serbia. Economic phenomena and processes were founded on the paradigms of classical industrialization and Marxism, specifically industrialization and the marginalization of agriculture [118]. The politics of this accelerated industrialization were motivated by the large agrarian population density and the differences in economic productivity between the agriculture and non-agricultural fields of human activity [119,120]. Industry hired a lot of workers, causing dramatic decreases in the rural population and agricultural production. These were the most intense during the following period of 1961–1981, when the urban population increased by up to 50% in certain cities (e.g., Aranđelovac, Kragujevac and Topola) [121]. These migration movements followed the same direction for decades (south–north and village–town/city). The tendencies toward the deepening disparity between the north and the south and between the rural and urban [122] were evident in this area.
Structural changes in agriculture as a whole, but primarily those in agricultural land use brought about during the 1990s, were the result of the socio–economic conditions in that period, which was characterized by political instability, social system crisis, sanctions, hyper-inflation and war [123]. The overall reduction in the prices of agricultural products and the increase in market uncertainty became spontaneously (without a pre-planned restructure) oriented towards financially less demanding agricultural production. This was also reflected in the structural changes in arable land [103], which led to the phenomena of excessive land fragmentation and small average farm sizes that were common in former socialist republics [42,85,124,125]. Individual production could not provide commercial profits, so survival relied on agricultural activities alone [42].
The second transition phase ensued after the political changes at the beginning of the 21st century. The main characteristic of this period in post-socialist countries was the reduction in total agricultural production, as well as changes in the share of market-oriented agriculture [113,115,126]. These changes in Serbia mostly caused reductions in the land use categories that were the most intensely cultivated [103].
Urbanization flows were highlighted in the northern parts of the Velika Morava River Basin, where two of the identified clusters (i.e., the high–high and low–high clusters) were a part of a medium sized peri-urban belt, including Smederevo, Požarevac, Smederevska Palanka, Mladenovac and Aranđelovac [127]. These were sustainable rural areas and their development and dynamics were directly related to the urban centers in terms of their production and socio–economic context. Similarly, settlements located to the south of the cluster were in agricultural areas [128]. The relatively high soil erosion intensity in the northern part of the basin and the small changes in erosion intensity over the years were indicators of this agricultural sustainability. Yet, in the low–high cluster, there were huge losses in agricultural land. In this part of the basin, more than any other, there was a rural–urban conflict. This phenomenon was accompanied by dynamic changes in land use [129]. The intensification of migration flows due to the need to find work was the cause of the decrease in the agricultural labor force and as a result, there was also a reduction in intense labor within agricultural production [123]. The higher degree of mechanization partly compensated for the active participation of the labor force in agriculture and helped to increase the productivity of agricultural production. On the other hand, the balance of natural conditions explained the statistical significance of anthropogenic factors as the basic determinants of erosion intensity.
Unlike the northern part of the basin, where the rural–urban conflicts lean toward sustaining erosion intensity, this was not the case in the southern part of the basin (i.e., the low–low cluster). These settlements were oriented toward local agricultural production, whereas the peripheral parts of the cluster included economically weaker rural areas [128]. The sociohistorical dynamics during the study period harmed the demographic, settlement and agrarian features of the area. Here, the urbanization process did not have a positive effect. Regardless of the favorable geographic positions of certain rural settlements, most of them were spatially located in hilly and mountainous parts of the basin, where the different natural conditions (i.e., higher altitudes) affected the flows of population migrations and the intensification of the deagrarization process. The other demographic features also highlighted the differences between the high–high and low–low clusters in terms of the spatial distribution of the demographic aging process and vitality index (Figure 7A,B).
The process of agricultural land abandonment has occurred the most intensively in post-socialist Central and Eastern European countries and European Russia, which began in the 1950s and was mainly due to technological, political and demographic processes [130,131]. The significant reduction in soil erosion identified in the Czech Republic (8.5%) [132], in some Russian river basins (2.0–3.8 fold) [130,133], in the Republic of Srpska (20%) [134] and in Poland (75%) [24]. A similar tendency of soil erosion intensity reduction was recorded in Serbian river basins: the Trgoviški Timok River Basin (59%), the Temštica River Basin (30%), the Nišava River Basin (30%) [43], the Ljig River Basin (43%) [93], the Vranjskobanjska River Basin (83%), the Rasina River Basin (33%) and the Jablanica River Basin (21%) [93,135,136].
Apart from its scientific and practical importance, the study also has some limitations.
(1) This study does not include a temporal component within the 1971–2011 period due to the lack of data on population and agricultural land for ten-year census periods at the settlement level. In this context, it was not possible to determine the soil erosion intensity and identify the key drivers for shorter time periods. Due to the limitations of the EPM model, the study was also unable to determine the intensity of soil erosion at different intra-annual levels.
(2) The spatial aspect of the study did not include smaller regional units within rural settlements, where the zones most at risk of soil erosion are found.
(3) Previous research in Serbia has shown that erosion control measures have an impact on reducing soil erosion [137]. However, this research did not consider the effect of anti-erosion studies.

5. Conclusions

In this study, we found that land use changes had direct effects on soil erosion intensity. The use of spatial statistical techniques and analyses have shown that changes were more pronounced in the southern part of the basin (i.e., the low–low cluster), while a slower decrease of erosion intensity was observed in the northern part of the basin in areas surrounding the peri-urban belts (i.e., the high-high).
As the Velika Morava River Basin is significant for maintaining national cohesion on different levels, this research could further support agricultural development and the development of the rural areas as a whole. Becoming familiar with the dynamics of changes in agricultural land use and soil erosion intensity could enable appropriate planning for land uses and define the economic adequacy of certain types of agricultural production. Likewise, the results of this study could form a basis for projecting and constructing facilities that require larger geographical spaces and an understanding of the causal relationships between physical–geographical, demographic and agrarian–geographical factors (e.g., roads, urban settlements, industrial facilities, melioration systems, etc.). The observed disparities between the northern and southern parts of the basin could be used to identify the main advantages and drawbacks of each area and contribute to obtaining spatial–demographic balance among the settlement network.

Author Contributions

Conceptualization, T.S. and S.M.; methodology, T.S., S.M. and B.B.; software, B.B., I.N., M.V.M. and T.S.; validation, T.S., S.M. and M.S.; formal analysis, T.S., M.S. and M.G.S.; investigation, T.S., M.T. and I.C.; resources, M.G.S.; data curation, T.S., M.S. and M.G.S.; writing—original draft preparation, T.S., S.M., B.B. and M.S.; writing—review and editing, T.S., S.M. and M.S.; visualization, M.V.M., T.S. and B.B.; supervision, S.M.; project administration, M.G.S.; funding acquisition, T.S., M.T. and I.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by projects from the Ministry of Science, Technological Development and Innovation of Republic of Serbia: No. 200091, No. 200092 and No. 200169.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lal, R. Restoring Soil Quality to Mitigate Soil Degradation. Sustainability 2015, 7, 5875–5895. [Google Scholar] [CrossRef] [Green Version]
  2. Amundson, R.; Berhe, A.A.; Hopmans, J.W.; Olson, C.; Sztein, A.E.; Sparks, D.L. Soil and human security in the 21st century. Science 2015, 348, 1261071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Borrelli, P.; Robinson, D.A.; Panagos, P.; Lugato, E.; Yang, J.E.; Alewell, C.; Wuepper, D.; Montanarella, L.; Ballabio, C. Land use and climate change impacts on global soil erosion by water (2015–2070). Proc. Natl. Acad. Sci. USA 2020, 117, 21994–22001. [Google Scholar] [CrossRef] [PubMed]
  4. Kucher, A.; Kucher, L.; Sysoieva, I.; Pohrishchuk, B. Economics of soil erosion: Case study of Ukraine. Agric. Econ. 2021, 7, 27–41. [Google Scholar] [CrossRef]
  5. Di Bene, C.; Francaviglia, R.; Farina, R.; Álvaro-Fuentes, J.; Zornoza, R. Agricultural Diversification. Agriculture 2022, 12, 369. [Google Scholar] [CrossRef]
  6. Keesstra, S.D.; Bouma, J.; Wallinga, J.; Tittonell, P.; Smith, P.; Cerdà, A.; Montanarella, L.; Quinton, J.; Pachepsky, Y.; Van der Puten, W.; et al. The significance of soils and soil science towards realization of the United Nations Sustainable Development Goals. Soil 2016, 2, 111–128. [Google Scholar] [CrossRef] [Green Version]
  7. Verheijen, F.G.; Jones, R.J.; Rickson, R.J.; Smith, C.J. Tolerable versus actual soil erosion rates in Europe. Earth Sci. Rev. 2009, 94, 23–38. [Google Scholar] [CrossRef] [Green Version]
  8. Alewell, C.; Borrelli, P.; Meusburger, K.; Panagos, P. Using the USLE: Chances, challenges and limitations of soil erosion modelling. Int. Soil Water Conserv. Res. 2019, 7, 203–225. [Google Scholar] [CrossRef]
  9. Petlušová, V.; Petluš, P.; Ševčík, M.; Hreško, J. The Importance of Environmental Factors for the Development of Water Erosion of Soil in Agricultural Land: The Southern Part of Hronská Pahorkatina Hill Land, Slovakia. Agronomy 2021, 11, 1234. [Google Scholar] [CrossRef]
  10. Tripolskaja, L.; Kazlauskaite-Jadzevice, A.; Baksiene, E.; Razukas, A. Changes in Organic Carbon in Mineral Topsoil of a Formerly Cultivated Arenosol under Different Land Uses in Lithuania. Agriculture 2022, 12, 488. [Google Scholar] [CrossRef]
  11. Oldeman, L.R. Global Extent of Soil Degradation; Bi-Annual Report 1991–1992/ISRIC; ISRIC: Wageningen, The Netherlands, 1992; pp. 19–36. [Google Scholar]
  12. Borrelli, P.; Alewell, C.; Alvarez, P.; Anache, J.; Baartman, J.; Ballabio, C.; Bezak, N.; Biddoccu, M.; Cerda, A.; Chalise, D.; et al. Soil erosion modelling: A global review and statistical analysis. Sci. Total Environ. 2021, 780, 146494. [Google Scholar] [CrossRef]
  13. Pepin, E.; Guyot, J.L.; Armijos, E.; Bazan, H.; Fraizy, P.; Moquet, J.S.; Noriega, L.; Lavado, W.; Pombosa, R.; Vauchel, P. Climatic control on eastern Andean denudation rates (Central Cordillera from Ecuador to Bolivia). J. S. Am. Earth Sci. 2013, 44, 85–93. [Google Scholar] [CrossRef]
  14. Padonou, E.A.; Anne, M.L.; Yvonne, B.; Rodrigue, I.; Brice, S. Mapping changes in land use/land cover and prediction of future extension of bowé in Benin, West Africa. Land Use Policy 2017, 69, 85–92. [Google Scholar] [CrossRef]
  15. Tadesse, L.; Suryabhagavan, K.V.; Sridhar, G.; Legesse, G. Land use and land cover changes and Soil erosion in Yezat Watershed, North Western Ethiopia. Int. Soil Water Conserv. Res. 2017, 5, 85–94. [Google Scholar] [CrossRef] [Green Version]
  16. Tsegaye, B. Effect of land use and land cover changes on soil erosion in Ethiopia. Int. J. Agric. Sci. 2019, 5, 26–34. [Google Scholar] [CrossRef] [Green Version]
  17. Kogo, B.K.; Kumar, L.; Koech, R. Impact of land use/cover changes on soil erosion in western Kenya. Sustainability 2020, 12, 9740. [Google Scholar] [CrossRef]
  18. Negese, A. Impacts of land use and land cover change on soil erosion and hydrological responses in Ethiopia. Appl. Environ. Soil Sci. 2021, 2021, 6669438. [Google Scholar] [CrossRef]
  19. Milazzo, F.; Fernández, P.; Peña, A.; Vanwalleghem, T. The resilience of soil erosion rates under historical land use change in agroecosystems of Southern Spain. Sci. Total Environ. 2022, 822, 153672. [Google Scholar] [CrossRef]
  20. Abdelsamie, E.A.; Abdellatif, M.A.; Hassan, F.O.; El Baroudy, A.A.; Mohamed, E.S.; Kucher, D.E.; Shokr, M.S. Integration of RUSLE Model, Remote Sensing and GIS Techniques for Assessing Soil Erosion Hazards in Arid Zones. Agriculture 2023, 13, 35. [Google Scholar] [CrossRef]
  21. Erskine, W.D.; Mahmoudzadeh, A.H.M.A.D.; Myers, C. Land use effects on sediment yields and soil loss rates in small basins of Triassic sandstone near Sydney, NSW, Australia. Catena 2002, 49, 271–287. [Google Scholar] [CrossRef]
  22. Leh, M.; Bajwa, S.; Chaubey, I. Impact of land use change on erosion risk: An integrated remote sensing, geographic information system and modeling methodology. Land Degrad. Dev. 2013, 24, 409–421. [Google Scholar] [CrossRef]
  23. Panagos, P.; Borrelli, P.; Meusburger, K.; Van Der Zanden, E.H.; Poesen, J.; Alewell, C. Modelling the effect of support practices (P-factor) on the reduction of soil erosion by water at European scale. Environ Sci. Policy 2015, 51, 23–34. [Google Scholar] [CrossRef]
  24. Latocha, A.; Szymanowski, M.; Jeziorska, J.; Stec, M.; Roszczewska, M. Effects of land abandonment and climate change on soil erosion—An example from depopulated agricultural lands in the Sudetes Mts., SW Poland. Catena 2016, 145, 128–141. [Google Scholar] [CrossRef]
  25. Özşahin, E.; Eroğlu, İ. Soil Erosion Risk Assessment due to Land Use/Land Cover Changes (LULCC) in Bulgaria from 1990 to 2015. Alinteri J. Agric. Sci. 2019, 34, 1–8. [Google Scholar] [CrossRef] [Green Version]
  26. Seeger, M.; Rodrigo-Comino, J.; Iserloh, T.; Brings, C.; Ries, J.B. Dynamics of runoff and soil erosion on abandoned steep vineyards in the Mosel Area, Germany. Water 2019, 11, 2596. [Google Scholar] [CrossRef] [Green Version]
  27. Niacsu, L.; Bucur, D.; Ionita, I.; Codru, I.C. Soil conservation measures on degraded land in the Hilly Region of Eastern Romania: A case study from Puriceni-Bahnari Catchment. Water 2022, 14, 525. [Google Scholar] [CrossRef]
  28. Vanwalleghem, T.; Gómez, J.A.; Amate, J.I.; De Molina, M.G.; Vanderlinden, K.; Guzmán, G.; Laguna, A.; Giráldez, J.V. Impact of historical land use and soil management change on soil erosion and agricultural sustainability during the Anthropocene. Anthropocene 2017, 17, 13–29. [Google Scholar] [CrossRef]
  29. Panagos, P.; Ballabio, C.; Himics, M.; Scarpa, S.; Matthews, F.; Bogonos, M.; Poeson, J.; Borrelli, P. Projections of soil loss by water erosion in Europe by 2050. Environ. Sci. Policy 2021, 124, 380–392. [Google Scholar] [CrossRef]
  30. Foley, J. Living by the lessons of the planet. Science 2017, 356, 251–252. [Google Scholar] [CrossRef]
  31. Wang, L.; Lyons, J.; Kanehl, P. Impacts of urbanization on stream habitat and fish across multiple spatial scales. Environ. Manag. 2001, 28, 255–266. [Google Scholar] [CrossRef]
  32. Falcucci, A.; Maiorano, L.; Boitani, L. Changes in land–use/land–cover patterns in Italy and their implications for biodiversity conservation. Landsc. Ecol. 2007, 22, 617–631. [Google Scholar] [CrossRef]
  33. Molotoks, A.; Smith, P.; Dawson, T.P. Impacts of land use, population, and climate change on global food security. Food Energy Secur. 2021, 10, e261. [Google Scholar] [CrossRef]
  34. Soranno, P.A.; Hubler, S.L.; Carpenter, S.R.; Lathrop, R.C. Phosphorus loads to surface waters: A simple model to account for spatial pattern of land use. Ecol. Appl. 1996, 6, 865–878. [Google Scholar] [CrossRef]
  35. Muller, D.; Kuemmerle, T.; Rusu, M.; Griffiths, P. Lost in Transition: Determinants of post-socialist Cropland Abandonment in Romania. J. Land Use Sci. 2009, 4, 109–129. [Google Scholar] [CrossRef]
  36. Griffiths, P.; Müller, D.; Kuemmerle, T.; Hostert, P. Agricultural Land Change in the Carpathian Ecoregion after the Breakdown of Socialism and Expansion of the European Union. Environ. Res. Lett. 2013, 8, 045024. [Google Scholar] [CrossRef]
  37. Kuškova, P.G. A Case Study of the Czech Agriculture since 1918 in a Socio-Metabolic Perspective– from Land Reform through Nationalisation to Privatisation. Land Use Policy 2013, 30, 592–603. [Google Scholar] [CrossRef]
  38. Pasakarnis, G.; Morley, D.; Maliene, V. Rural Development and Challenges Establishing Sustainable Land Use in Eastern European Countries. Land Use Policy 2013, 30, 703–710. [Google Scholar] [CrossRef]
  39. Prishchepov, A.V.; Müller, D.; Dubinin, M.; Baumann, M.; Radeloff, V.C. Determinants of Agricultural Land Abandonment in post-Soviet European Russia. Land Use Policy 2013, 30, 873–884. [Google Scholar] [CrossRef] [Green Version]
  40. Pazúr, R.; Lieskovský, J.; Feranec, J.; Oťaheľ, J. Spatial Determinants of Abandonment of Large-Scale Arable Lands and Managed Grasslands in Slovakia during the Periods of post-socialist Transition and European Union Accession. Appl. Geogr. 2014, 54, 118–128. [Google Scholar] [CrossRef]
  41. Gusarov, A.V. Land–Use/Cover Changes and Their Effect on Soil Erosion and River Suspended Sediment Load in Different Landscape Zones of European Russia during 1970–2017. Water 2021, 13, 1631. [Google Scholar] [CrossRef]
  42. Bogdanov, N.; Vasiljević, Z. Role of agriculture and multifunctional rural development in Serbia. Appl. Stud. Agribus. Commer. 2011, 5, 47–55. [Google Scholar] [CrossRef]
  43. Manojlović, S.; Sibinović, M.; Srejić, T.; Novković, I.; Milošević, M.V.; Gatarić, D.; Carević, I.; Batoćanin, N. Factors Controlling the Change of Soil Erosion Intensity in Mountain Watersheds in Serbia. Front. Environ. Sci. 2022, 10, 88901. [Google Scholar] [CrossRef]
  44. Manojlović, S.; Antić, M.; Šantić, D.; Sibinović, M.; Carević, I.; Srejić, T. Anthropogenic Impact on Erosion Intensity: Case Study of Rural Areas of Pirot and Dimitrovgrad Municipalities, Serbia. Sustainability 2018, 10, 826. [Google Scholar] [CrossRef] [Green Version]
  45. Manojlović, S.; Sibinović, M.; Srejić, T.; Hadud, A.; Sabri, I. Agriculture land use change and demographic change in response to decline suspended sediment in Južna Morava River basin (Serbia). Sustainability 2021, 13, 3130. [Google Scholar] [CrossRef]
  46. Menković, L.; Košćal, M.; Milivojević, M.; Đokić, M. Morphostructure relations on the territory of the Republic of Serbia. Bull. Serb. Geogr. Soc. 2018, 98, 1–28. [Google Scholar] [CrossRef] [Green Version]
  47. Manojlović, S.; Manojlović, P.; Đokić, M. Dynamics of suspended sediment load in the Morava River (Serbia) in the period 1967-2007. Rev. Geomorfol. 2016, 18, 47–58. [Google Scholar] [CrossRef]
  48. Milovanović, B.; Ducić, V.; Radovanović, M.; Milivojević, M. Climate regionalization of Serbia according to Köppen climate classification. J. Geograph. Inst. “Jovan Cvijić” SASA 2017, 67, 103–114. [Google Scholar] [CrossRef] [Green Version]
  49. Blagojević, B.; Langović, M.; Novković, I.; Dragićević, S.; Živković, N. Water Resources of Serbia and Its Utilization. In Water Resources Management in Balkan Countries; Springer: Berlin/Heidelberg, Germany, 2020; pp. 213–247. [Google Scholar]
  50. Bačević, N.R.; Milentijević, N.M.; Valjarević, A.; Gicić, A.; Kićović, D.; Radaković, M.G.; Nikolić, M.; Pantelić, M. Spatiotemporal variability of air temperatures in Central Serbia from 1949 to 2018. Időjárás/Q. J. Hung. Meteorol. Serv. 2021, 125, 229–253. [Google Scholar] [CrossRef]
  51. Milovanović, B.; Stanojević, G.; Radovanović, M. Climate of Serbia. In The Geography of Serbia; Springer: Berlin/Heidelberg, Germany, 2022; pp. 57–68. [Google Scholar]
  52. Radaković, M.G.; Tošić, I.; Bačević, N.; Mladjan, D.; Gavrilov, M.B.; Marković, S.B. The analysis of aridity in Central Serbia from 1949 to 2015. Theor. Appl. Climatol. 2018, 133, 887–898. [Google Scholar] [CrossRef]
  53. Gocić, M.; Trajković, S. Spatio-temporal patterns of precipitation in Serbia. Theor. Appl. Climatol. 2014, 117, 419–431. [Google Scholar] [CrossRef]
  54. Valjarević, A.; Morar, C.; Živković, J.; Niemets, L.; Kićović, D.; Golijanin, J.; Gocić, M.; Martić Bursać, N.; Stričević, L.J.; Žiberna, I.; et al. Long term monitoring and connection between topography and cloud cover distribution in Serbia. Atmosphere 2021, 12, 964. [Google Scholar] [CrossRef]
  55. Pešić, A. Water regime and discharges trends of the rivers in the Šumadija region (Serbia). In Proceedings of the International Scientific Symposium “New trends in geography”, Ohrid, North Macedonia, 3–4 October 2019; pp. 3–13. [Google Scholar]
  56. Petrović, A. Torrential Floods in Serbia; Special Issues 73; Serbian Geographical Society: Belgrade, Serbia, 2021. [Google Scholar]
  57. Ristić, R.; Kostadinov, S.; Abolmasov, B.; Dragićević, S.; Trivan, G.; Radić, B.; Trifunović, M.; Radosavljević, Z. Torrential floods and town and country planning in Serbia. Nat. Hazards Earth Syst. Sci. 2012, 12, 23–35. [Google Scholar] [CrossRef]
  58. Urošev, M.; Milanović Pešić, A.; Kovačević–Majkić, J.; Štrbac, D. Hydrological Characteristics of Serbia. In The Geography of Serbia, Nature, People, Economy; Manić, E., Nikitović, V., Đurović, P., Eds.; Springer: Berlin/Heidelberg, Germany, 2022; pp. 69–84. [Google Scholar]
  59. Gavrilović, L.; Milanović Pešić, A.; Urošev, M. A Hydrological Analysis of the Greatest Floods in Serbia in the 1960–2010 Period. Carpathian J. Earth Environ. Sci. 2012, 7, 107–116. [Google Scholar]
  60. Dragićević, S.; Filipović, D.; Kostadinov, S.; Ristić, R.; Novković, I.; Živković, N.; Anđelković, G.; Abolmasov, B.; Šećerov, V.; Đurdjic, S. Natural hazard assessment for land-use planning in Serbia. Int. J. Environ. Res. 2011, 5, 371–380. [Google Scholar]
  61. Novković, I.; Dragićević, S.; Đurović, M. Geohazard and Geoheritage. In The Geography of Serbia, Nature, People, Economy; Manić, E., Nikitović, V., Đurović, P., Eds.; Springer: Berlin/Heidelberg, Germany, 2022; pp. 119–131. [Google Scholar]
  62. FAO Soil Portal (World Reference Base). Available online: https://www.fao.org/soils-portal/data-hub/soil-classification/world-reference-base/en/ (accessed on 20 March 2023).
  63. Statistical Office of the Republic of Serbia 1961–2012. Available online: http://www.stat.gov.rs (accessed on 10 September 2022).
  64. Gavrilović, S. Engineering of Torrential Flows and Erosion; Izgradnja: Belgrad, Serbia, 1972; 272p. [Google Scholar]
  65. Zorn, M.; Komac, B. Response of Soil Erosion to Land Use Change with Particular Reference to the Last 200 Years (Julian Alps, Western Slovenia). In Proceedings of the IAG Regional Conference on Geomorphology: Landslides, Floods and Global Environmental Change in Mountain Regions, Brasov, Romania, 15–25 September 2008; pp. 39–47. [Google Scholar]
  66. Tošić, R.; Dragićević, S.; Lovrić, N. Assessment of soil erosion and sediment yield changes using erosion potential model–case study: Republic of Srpska (BiH). Carpathian J. Earth Environ. Sci. 2012, 7, 147–154. [Google Scholar]
  67. Blinkov, I.; Marinov, I.; Kostadinov, S. Comparison of Erosion and Erosion Control Works in Macedonia, Serbia and Bulgaria. Int. Soil Water Conserv. Res. 2013, 1, 15–28. [Google Scholar] [CrossRef] [Green Version]
  68. Efthimiou, N.; Lykoudi, E.; Panagoulia, D.; Karavitis, C. Assessment of Soil Susceptibility to Erosion Using the EPM and RUSLE MODELS: The Case of Venetikos River Catchment. Glob. NEST J. 2016, 18, 164–179. [Google Scholar]
  69. Manojlović, S.; Antić, M.; Sibinović, M.; Dragićević, S.; Novković, I. Soil erosion response to demographic and land use changes in the Nišava River basin, Serbia. Fresenius Environ. Bull. 2017, 26, 7547–7560. [Google Scholar]
  70. Dragičević, N.; Karleuša, B.; Ožanić, N. Modification of erosion potential method using climate and land cover parameters. Geomatics, Nat. Hazards Risk. 2018, 9, 1085–1105. [Google Scholar] [CrossRef] [Green Version]
  71. Lovrić, N.; Tošić, R. Assessment of soil erosion and sediment yield using erosion potential method: Case study–Vrbas River basin (B&H). Bull. Serb. Geogr. Soc. 2018, 98, 1–14. [Google Scholar]
  72. Blinkov, I.; Kostadinov, S. Applicability of various erosion risk assessment methods for engineering purposes. In Proceedings of the BALWOIS Conference, Ohrid, North Macedonia, 25–29 May 2010. [Google Scholar]
  73. Bazzoffi, P. Methods for net erosion measurement in watersheds as a tool for the validation of models in central Italy. In Workshop on Soil Erosion and Hillslope Hydrology with Emphasis on Higher Magnitude Events; K.U.: Leuven, Belgium, 1985. [Google Scholar]
  74. Beyer, P.N. Erosion des Bassins Versant Alpins Suisses par Ruissellement de Surface. Ph.D. Thesis, University of Switzerland, Laussanne, Switzerland, 1998; 438p. [Google Scholar]
  75. Fanetti, D.; Vezzoli, L. Sediment input and evolution of lacustrine deltas: The Breggia and Greggio rivers case study (Lake Como, Italy). Quat. Int. 2007, 173, 113–124. [Google Scholar] [CrossRef]
  76. Poggetti, E.; Cencetti, C.; De Rosa, P.; Fredduzzi, A.; Rivelli, F.R. Sediment Supply and Hydrogeological Hazard in the Quebrada De Humahuaca (Province of Jujuy, Northwestern Argentina)—Rio Huasamayo and Tilcara Area. Geosciences 2019, 9, 483. [Google Scholar] [CrossRef] [Green Version]
  77. Ouallali, A.; Aassoumi, H.; Moukhchane, M.; Moumou, A.; Houssni, M.; Spalevic, V.; Keesstra, S. Sediment mobilization study on Cretaceous, Tertiary and Quaternary lithological formations of an external Rif catchment, Morocco. Hydrol. Sci. J. 2020, 65, 1568–1582. [Google Scholar] [CrossRef]
  78. Mohammadi, M.; Khaledi Darvishan, A.; Spalević, V.; Dudić, B.; Billi, P. Analysis of the Impact of Land Use Changes on Soil Erosion Intensity and Sediment Yield Using the IntErO Model in the Talar Watershed of Iran. Water 2021, 13, 881. [Google Scholar] [CrossRef]
  79. Luković, J.; Bajat, B.; Blagojević, D.; Kilibarda, M. Spatial pattern of recent rainfall trends in Serbia (1961–2009). Reg. Environ. Chang. 2014, 14, 1789–1799. [Google Scholar] [CrossRef]
  80. Solaimani, K.; Modallaldoust, S.; Lotfi, S. Investigation of land use changes on soil erosion process using geographical information system. Int. J. Environ. Sci. Technol. 2009, 6, 415–424. [Google Scholar] [CrossRef] [Green Version]
  81. Sharma, A.; Tiwari, K.N.; Bhadoria, P.B. Effect of land use land cover change on soil erosion potential in an agricultural watershed. Environ. Monit. Assess. 2011, 173, 789–801. [Google Scholar] [CrossRef]
  82. Serpa, D.; Nunes, J.P.; Santos, J.; Sampaio, E.; Jacinto, R.; Veiga, S.; Lima, J.C.; Moreira, M.; Corte–Real, J.; Keizer, J.J. Impacts of climate and land use changes on the hydrological and erosion processes of two contrasting Mediterranean catchments. Sci. Total Environ. 2015, 538, 64–77. [Google Scholar] [CrossRef] [Green Version]
  83. Perović, V.; Kadović, R.; Ðurdević, V.; Braunović, S.; Čakmak, D.; Mitrović, M.; Pavlović, P. Effects of changes in climate and land use on soil erosion: Case study of the Vranjska Valley, Serbia. Reg. Environ. Chang. 2019, 19, 1035–1046. [Google Scholar] [CrossRef]
  84. Uddin, K.; Abdul Matin, M.; Maharjan, S. Assessment of land cover change and its impact on changes in soil erosion risk in Nepal. Sustainability 2018, 10, 4715. [Google Scholar] [CrossRef] [Green Version]
  85. Devátý, J.; Dostál, T.; Hösl, R.; Krása, J.; Strauss, P. Effects of historical land use and land pattern changes on soil erosion–Case studies from Lower Austria and Central Bohemia. Land Use Policy 2019, 82, 674–685. [Google Scholar] [CrossRef]
  86. Moisa, M.B.; Negash, D.A.; Merga, B.B.; Gemeda, D.O. Impact of land-use and land-cover change on soil erosion using the RUSLE model and the geographic information system: A case of Temeji watershed, Western Ethiopia. J. Water Clim. Chang. 2021, 12, 3404–3420. [Google Scholar] [CrossRef]
  87. Anselin, L. Local indicators of spatial association-LISA. Geogr. Anal. 1995, 27, 93–115. [Google Scholar] [CrossRef]
  88. O’Sullivan, D.; Unwin, D. Geographic Information Analysis; John Wiley & Sons: Hoboken, NJ, USA, 2010; 432p. [Google Scholar]
  89. Anselin, L.; Syabri, I.; Smirnov, O. Visualizing Multivariate Spatial Correlation with Dynamically Linked Windows. In New Tools for Spatial Data Analysis: Proceedings of the Specialist Meeting; Anselin, L., Rey, S., Eds.; Center for Spatially Integrated Social Science (CSISS), University of California: Santa Barbara, CA, USA, 2002. [Google Scholar]
  90. Anselin, L.; Syabri, I.; Youngihn, K. GeoDa: An Introduction to Spatial Data Analysis. Geogr. Anal. 2006, 38, 5–22. [Google Scholar] [CrossRef]
  91. Lazarević, R. The Erosion Map of Serbia. Scale 1:500000; Institute of Forestry: Belgrade, Serbia, 1983. [Google Scholar]
  92. Copernicus Land Monitoring Service (2012). CORINE Land Cover (CLC). Available online: https://land.copernicus.eu/pan-european/corine-land-cover/clc2012 (accessed on 15 September 2021).
  93. Novković, I. Natural Conditions as Determinants of Geohazards on the Example of Ljig, Jošanička and Vranjskobanjska River Basins. Ph.D. Thesis, University of Belgrade, Faculty of Geography, Belgrade, Serbia, 2016. [Google Scholar]
  94. Kostadinov, S.; Dragićević, S.; Stefanović, T.; Novković, I.; Petrović, A.M. Torrential flood prevention in the Kolubara River basin. J. Mt. Sci. 2017, 14, 2230–2245. [Google Scholar] [CrossRef]
  95. Durlević, U.; Momčilović, A.; Ćurić, V.; Dragojević, M. Gis application in analysis of erosion intensity in the Vlasina River Basin. Bull. Serb. Geogr. Soc. 2019, 99, 17–36. [Google Scholar] [CrossRef]
  96. Durlević, U.; Novković, I.; Lukić, T.; Valjarević, A.; Samardžić, I.; Krstić, F.; Batoćanin, N.; Mijatov, M.; Ćurić, V. Multihazard susceptibility assessment: A case study–Municipality of Štrpce (Southern Serbia). Open Geosci. 2021, 13, 1414–1431. [Google Scholar] [CrossRef]
  97. Dragićević, S.; Kostadinov, S.; Novković, I.; Momirović, N.; Langović, M.; Stefanović, T.; Radović, M.; Tošić, R. Assessment of Soil Erosion and Torrential Flood Susceptibility: Case Study—Timok River Basin, Serbia. In The Lower Danube River: Hydro-Environmental Issues and Sustainability; Negm, A., Zaharia, L., Ioana–Toroimac, G., Eds.; Springer: Berlin/Heidelberg, Germany, 2022; pp. 357–380. [Google Scholar]
  98. Krstić, F.; Paunović, S. Changes in Soil Erosion Intensity in Jablanica Region; Collection of Papers; University of Belgrade, Faculty of Geography: Belgrade, Serbia, 2022; pp. 83–93. [Google Scholar] [CrossRef]
  99. Copernicus Land Monitoring Service (2020). Imagery and Reference Data, EU-DEM v1.1. Available online: https://land.copernicus.eu/imagery-in-situ/eudem/eu-dem-v1.1 (accessed on 25 September 2021).
  100. Geološki zavod Srbije. Basic Geological Map of Former Yugoslavia 1:100 000; Federal Geological Survey: Belgrade, Serbia, 1978. [Google Scholar]
  101. R Core Team. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing; R Core Team: Vienna, Austria, 2012; Available online: http://www.R-project.org/ (accessed on 7 March 2022)ISBN 3-900051-07-0.
  102. Wuttichaikitcharoen, P.; Babel, M.S. Principal component and multiple regression analyses for the estimation of suspended sediment yield in ungauged basins of northern Thailand. Water 2014, 6, 2412–2435. [Google Scholar] [CrossRef] [Green Version]
  103. Sibinović, M. Typology of Agriculture in Conditions of Transitional Crisis: The Case of the Belgrade Region; Collection of Papers; University of Belgrade, Faculty of Geography: Belgrade, Serbia, 2015; pp. 81–118. [Google Scholar]
  104. Matejka, K. Multivariate Analysis for Assessment of the Tree Populations Based on Dendrometric Data with an Example of Similarity Among Norway spruce Subpopulations. J. For. Sci. 2017, 63, 449–456. [Google Scholar] [CrossRef] [Green Version]
  105. Mizuta, K.; Grunwald, S.; Phillips, M.A. New Soil Index Development and Integration with Econometric Theory. Soil Sci. Soc. Am. J. 2018, 82, 1017–1032. [Google Scholar] [CrossRef] [Green Version]
  106. Kaiser, H. An index of factorial simplicity. Psychometrika 1974, 39, 31–36. [Google Scholar] [CrossRef]
  107. Hair, J.F.; Black, B.; Babin, B.; Anderson, R.E.; Tatham, R.L. Multivariate Data Analysis, 6th ed.; Pearson Prentice Hall: Upper Saddle River, NJ, USA, 2006; pp. 1–816. [Google Scholar]
  108. Williams, B.; Onsman, A.; Brett, A.; Brown, T. Exploratory factor analysis: A five-step guide for novices. Australas. J. Paramed. 2012, 8, 1–13. [Google Scholar] [CrossRef] [Green Version]
  109. Ševarlić, M.; Tomić, D. Serbian agriculture in crisis condition. JUMTO 2009, 14, 157–164. [Google Scholar]
  110. Sibinović, M. How Did Agricultural Patterns Change in Serbia After the Fall of Yugoslavia? Geogr. Teach. 2018, 15, 33–35. [Google Scholar] [CrossRef]
  111. Gatarić, D. The daily urban system of the Knjaževac. In Region of Knjaževac-Potential, Current State and Prospects; Sibinović, M., Stojadinović, V., Popović Nikolić, D., Eds.; National library Njegoš: Knjaževac, Serbia, 2019; pp. 78–85. [Google Scholar]
  112. Sibinović, M. Structural Changes and Spatial Differentiation of Agriculture in Rural Settlements of Belgrade Region. Ph.D. Thesis, University of Belgrade, Faculty of Geography, Belgrade, Serbia, 2014. [Google Scholar]
  113. Bański, J. Agriculture of Central Europe in the Period of Economic Transformation. Rural Stud. 2008, 15, 9–22. [Google Scholar]
  114. Sibinović, M.; Winkler, A.; Grčić, M. Agriculture in a transitional crisis period: Crop production in the administrative region of Belgrade from 1991 to 2002. Mitt. Osterr. Geogr. Ges. 2014, 156, 293–310. [Google Scholar] [CrossRef] [Green Version]
  115. Kostov, P.; Lingard, J. Subsistence agriculture in transition economies: Its roles and determinants. J. Agric. Econ. 2004, 55, 565–579. [Google Scholar] [CrossRef]
  116. Mc Donagh, J. Rural Geography I: Changing Expectations and Contradictions in the Rural. Prog. Hum. Geogr. 2012, 37, 712–720. [Google Scholar] [CrossRef]
  117. Arsenović, D.; Nikitović, V. Demographic Profile of Serbia at the Turn of the Millennia. In The Geography of Serbia, Nature, People, Economy; Manić, E., Nikitović, V., Đurović, P., Eds.; Springer: Berlin/Heidelberg, Germany, 2022; pp. 135–141. [Google Scholar]
  118. Todorović, M. The Base of Typology and Regionalization of Agriculture of Serbia; Serbian Geographical Society: Belgrade, Serbia, 2002; pp. 1–164. [Google Scholar]
  119. Čobeljić, N. Položaj Poljoprivrede u Procesu Privrednog Razvoja, Prilozi sa Skupa Razvojni i Sistemski Problemi Poljoprivrede Jugoslavije; Ekonomski Zbornik: Beograd, Serbia, 1986. [Google Scholar]
  120. Martinović, M.; Ratkaj, I. Sustainable rural development in Serbia: Towards a quantitative typology of rural areas. Carpathian J. Earth Environ. Sci. 2015, 10, 37–48. [Google Scholar]
  121. Drobnjaković, M.; Petrović, G.; Karabašević, D.; Vukotić, S.; Mirčetić, V.; Popović, V. Socio–economic transformation of Šumadija district (Serbia). J. Geograph. Inst. “Jovan Cvijić” SASA 2021, 71, 163–180. [Google Scholar] [CrossRef]
  122. Lukić, V. Migration and Mobility Paterns in Serbia. In The Geography of Serbia, Nature, People, Economy; Manić, E., Nikitović, V., Đurović, P., Eds.; Springer: Berlin/Heidelberg, Germany, 2022; pp. 157–168. [Google Scholar]
  123. Sibinović, M. Structural changes in the rural planting areas of Belgrade region. Bull. Serb. Geogr. Soc. 2012, 92, 112–132. [Google Scholar] [CrossRef]
  124. Pagáč Mokrá, A.; Pagáč, J.; Muchová, Z.; Petrovič, F. Analysis of ownership data from consolidated land threatened by water erosion in the Vlára basin, Slovakia. Sustainability 2020, 13, 51. [Google Scholar] [CrossRef]
  125. Gorgan, M.; Hartvigsen, M. Development of agricultural land markets in countries in Eastern Europe and Central Asia. Land Use Policy 2022, 120, 106257. [Google Scholar] [CrossRef]
  126. Jordan, P. Development of rural space in post-communist Southeast Europe after 1989: A comparative analysis. Revija Za Geografijo-Geogr. J. 2009, 4, 89–102. [Google Scholar]
  127. Živanović, Z.; Tošić, B.; Nikolić, T.; Gatarić, D. Urban System in Serbia—The Factor in the Planning of Balanced Regional Development. Sustainability 2019, 11, 4168. [Google Scholar] [CrossRef] [Green Version]
  128. Gajić, A.; Krunić, N.; Protić, B. Classification of rural areas in Serbia: Framework and implications for spatial planning. Sustainability 2021, 13, 1596. [Google Scholar] [CrossRef]
  129. Antrop, M.; Van Eetvelde, V. Holistic aspects of suburban landscapes: Visual image interpretation and landscape metrics. Landsc. Urban Plan. 2000, 50, 43–58. [Google Scholar] [CrossRef]
  130. Golosov, V.; Yermolaev, O.; Litvin, L.; Chizhikova, N.; Kiryukhina, Z.; Safina, G. Influence of climate and land use changes on recent trends of soil erosion rates within the Russian Plain. Land Degrad. Dev. 2018, 29, 2658–2667. [Google Scholar] [CrossRef]
  131. Mal’tsev, K.A.; Ivanov, M.A.; Sharifullin, A.G.; Golosov, V.N. Changes in the rate of soil loss in river basins within the Southern Part of European Russia. Eurasian Soil Sci. 2019, 52, 718–727. [Google Scholar] [CrossRef]
  132. Van Rompaey, A.; Govers, G.; Verstraeten, G.; van Oost, K.; Poesen, J. Modelling the geomorphic response to land use changes. In Long Term Hillslope and Fluvial System Modelling–Concepts and Case Studies from the Rhine River Catchment; Lang, A., Dikau, R., Hennnrich, K., Eds.; Springer: Berlin/Heidelberg, Germany, 2003; pp. 73–100. [Google Scholar]
  133. Gusarov, A.A. The impact of contemporary changes in climate and land use/cover on tendencies in water flow, suspended sediment yield and erosion intensity in the northeastern part of the Don river basin, SW European Russia. Environ. Res. 2019, 175, 468–488. [Google Scholar] [CrossRef]
  134. Tošić, R.; Lovrić, N.; Dragićević, S. Assesment of the impact of depopulation on soil erosion: Case study Republika Srpska (Bosnia and Herzegovina). Carpathian J. Earth Environ. Sci. 2019, 14, 505–518. [Google Scholar] [CrossRef]
  135. Kostadinov, S.; Zlatić, M.; Dragićević, S.; Novković, I.; Košanin, O.; Borisavljević, A.; Lakićević, M.; Mlađan, D. Anthropogenic influence on erosion intensity changes in the Rasina river watershed-Central Serbia. Fresenius Environ. Bull. 2014, 23, 254–263. [Google Scholar]
  136. Gocić, M.; Dragićević, S.; Radivojević, A.; Martić Bursać, N.; Stričević, L.; Đorđević, M. Changes in soil erosion intensity caused by land use and demographic changes in the Jablanica River Basin, Serbia. Agriculture 2020, 10, 345. [Google Scholar] [CrossRef]
  137. Kostadinov, S.; Braunović, S.; Dragićević, S.; Zlatić, M.; Dragović, N.; Rakonjac, N. Effects of erosion control works: Case study—Grdelica Gorge, the South Morava River (Serbia). Water 2018, 10, 1094. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Geographical position of the Velika Morava River Basin.
Figure 1. Geographical position of the Velika Morava River Basin.
Agriculture 13 00778 g001
Figure 2. The average coefficient of erosion Z in rural settlements of the Velika Morava River Basin in 1971 (A) and 2011 (B); Index Z for the period 1971–2011 (C); LISA univariate cluster map of Index Z for the period of 1971–2011 (D).
Figure 2. The average coefficient of erosion Z in rural settlements of the Velika Morava River Basin in 1971 (A) and 2011 (B); Index Z for the period 1971–2011 (C); LISA univariate cluster map of Index Z for the period of 1971–2011 (D).
Agriculture 13 00778 g002
Figure 3. Agriculture land (AgL) in rural settlements of the Velika Morava River Basin in 1961 (A) and 2011 (B); Index AgL for the period of 1961–2011 (C); LISA univariate cluster map of Index AgL for the period of 1961–2011 (D).
Figure 3. Agriculture land (AgL) in rural settlements of the Velika Morava River Basin in 1961 (A) and 2011 (B); Index AgL for the period of 1961–2011 (C); LISA univariate cluster map of Index AgL for the period of 1961–2011 (D).
Agriculture 13 00778 g003
Figure 4. Bivariate LISA cluster map: Index AgL/Index Z.
Figure 4. Bivariate LISA cluster map: Index AgL/Index Z.
Agriculture 13 00778 g004
Figure 5. Eigenvalues and cumulative variance of principal components (high–high).
Figure 5. Eigenvalues and cumulative variance of principal components (high–high).
Agriculture 13 00778 g005
Figure 6. Eigenvalues and cumulative variance of principal components (low–low).
Figure 6. Eigenvalues and cumulative variance of principal components (low–low).
Agriculture 13 00778 g006
Figure 7. Histogram of demographic variables in rural settlements: (A) vitality index (Index V) and (B) old rural population (ORP) according to the high–high cluster and the low–low cluster.
Figure 7. Histogram of demographic variables in rural settlements: (A) vitality index (Index V) and (B) old rural population (ORP) according to the high–high cluster and the low–low cluster.
Agriculture 13 00778 g007
Table 1. Values of the descriptive variables (Y, X and φ) used in calculating the erosion coefficient (Z).
Table 1. Values of the descriptive variables (Y, X and φ) used in calculating the erosion coefficient (Z).
Coefficient of Soil ResistanceY Value
Fine sediments and soils without erosion resistance0.80–1.00
Sediments, moraines, clay and other rock with little resistance0.60–0.80
Weak rock, schistose, stabilized0.50–0.60
Rock with moderate erosion resistance0.30–0.50
Hard rock, erosion resistant0.10–0.30
Coefficient of soil protectionX value
Areas without vegetal cover0.08–1.00
Damaged pasture and cultivated land0.06–0.80
Damaged forest and bushes, pasture0.04–0.06
Coniferous forest with little grove, scarce bushes, bushy prairie0.20–0.40
Thin forest with grove0.05–0.20
Mixed and dense forest0.05–0.20
Coefficient of type and extent of erosionφ value
Whole watershed affected by erosion0.90–1.00
50–80% of the catchment area is affected by surface erosion and landslides0.80–0.90
Erosion in rivers, gullies and alluvial deposits, karstic erosion0.60–0.70
Erosion in waterways on 20–50% of the catchment area0.30–0.50
Little erosion on watershed0.10–0.20
Table 2. EPM erosion categorization and range of the erosion coefficient (Z).
Table 2. EPM erosion categorization and range of the erosion coefficient (Z).
Erosion CategoryErosion IntensityRange of ZAverage of ZRange of W (m3/km2/yr)
IExcessive erosion>1.011.25>3000
IIIntensive erosion0.71–1.000.851200–3000
IIIMedium erosion0.41–0.700.55800–1200
IVWeak erosion0.21–0.400.30400–800
VVery weak erosion0.01–0.200.10100–400
Table 3. Controlling variables used in the correlation analysis (CA) and principal component analysis (PCA).
Table 3. Controlling variables used in the correlation analysis (CA) and principal component analysis (PCA).
Variables–Abbreviation (Units)Calculate
Physical–geographical
Erosion coefficient in 2011
Z2 (–)
Equation (2)X, Z, φTable 1
Mean Altitude
Aav (m)
DEMDEM—Digital Elevation Model
Terrain slope
I (°)
DEMDEM—Digital Elevation Model
Lithology—sediments of Neogen
NSA (%)
NSA = (ANSA/A)∙100ANSA—Neogen sediments area
A—total area of the settlement
Vegetation—Forest cover
FC (%)
FC = (AFC/A)∙100AFC—forest area in 2011
A—total area of the settlement
Agrarian–geographical
Deagrarization index agricultural land
Index AgL (–)
Index AgL = (AgL2/AgL1)∙100AgL2—agricultural land (ha) in 2011
AgL1—agricultural land (ha) in 1961
Deagrarization index arable land
Index ArL (–)
Index ArL = (ArL2/ArL1)∙100ArL2—arable land (ha) in 2011
ArL1—arable land (ha) in 1961
Share of arable land in agricultural land
ArLs (%)
ArLs = (ArL/AgL)∙100ArL—arable land (ha) in 2011
AgL—agricultural land (ha) in 2011
General agrarian population density
GDAgL (rural population/100 ha)
GDAgL = RP/AgL∙100RP—rural population in 2011
AgL—agricultural land (100 ha) in 2011
Specific agrarian population density
SDArL(rural population/100 ha)
SDArL= RP/ArL∙100RP—rural population in 2011
ArL—arable land (100 ha) in 2011
Demographic
Depopulation index
Index RP (–)
Index RP = (RP2/RP1)∙100RP2—rural population in 2011
RP1—rural population in 1961
Density rural population
DRP (rural population/km2)
DRP = RP/ARP—rural population in 2011
A—total area of the settlement
Vitality index
Index V (–)
Index V = (WRP/ORP)∙100WRP—economically active rural population in 2011
ORP—rural population older than 65 yr in 2011
The average age of the rural population
RPav (years)
Age and Sex,
data by settlements 2011
Old rural population
ORP (%)
ORP = (ORP/RP)∙100ORP—rural population older than 65 yr in 2011
RP—rural population in 2011
Household index
Index H (–)
Index H = (H2/H1)∙100H2—total number of households in 2011
H1—total number of households in 1961
Household index size
Index Hs (–)
Index Hs = (Hs2/Hs1)∙100Hs2—average household size in 2011
Hs1—average household size in 1961
Table 4. Erosion category 1971 and 2011 in the Velika Morava River Basin.
Table 4. Erosion category 1971 and 2011 in the Velika Morava River Basin.
Erosion CategoryErosion IntensityF (km2) 1971F (%) 1971F (km2) 2011F (%) 2011
IExcessive erosion310.34.64.00.1
IIIntensive erosion2111.831.41094.016.2
IIIMedium erosion2093.231.12580.938.3
IVWeak erosion1060.115.71021.115.2
VVery weak erosion1159.217.22034.530.2
Table 5. Correlation matrix of the principal component analysis [Pearson (n)] for high–high cluster. Values in bold are different from 0 with a significance level α = 0.05.
Table 5. Correlation matrix of the principal component analysis [Pearson (n)] for high–high cluster. Values in bold are different from 0 with a significance level α = 0.05.
Z2AavINSAFCIndex AgLIndex ArLArLsGDAgLSDArLIndex RPDRPIndex VRPavORPIndex HIndex Hs
Z21
Aav−0.3791
I−0.4450.8931
NSA0.254−0.856−0.8371
FC−0.4050.0100.1740.0631
Index AgL0.0910.028−0.0260.034−0.2721
Index ArL0.473−0.236−0.2880.235−0.1820.7111
ArLs0.583−0.692−0.7050.6430.0420.0540.5961
GDAgL−0.0910.0200.065−0.125−0.0400.002−0.0080.0181
SDArL−0.2900.1770.232−0.253−0.0510.052−0.176−0.2900.9321
Index RP−0.0520.0510.087−0.164−0.1090.1640.069−0.0740.4830.5211
DRP−0.186−0.219−0.1320.1090.1200.0200.0010.0980.5080.5040.7771
Index V−0.014−0.214−0.1510.109−0.1540.0440.0140.0490.4150.4040.7060.6591
RPav0.1620.1250.056−0.0950.095−0.0870.0350.054−0.314−0.343−0.547−0.521−0.8481
ORP−0.0070.2390.184−0.1420.119−0.0270.018−0.050−0.358−0.351−0.606−0.597−0.9510.8591
Index H−0.0500.0590.081−0.181−0.0670.1300.060−0.0530.4620.4840.9800.7690.713−0.534−0.6211
Index Hs0.049−0.0340.0130.064−0.2420.1620.074−0.0870.0460.1010.055−0.029−0.045−0.0520.091−0.1361
Table 6. Percentage of variance and cumulative variance after varimax rotation (high–high).
Table 6. Percentage of variance and cumulative variance after varimax rotation (high–high).
Factor 1Factor 2Factor 3Factor 4
Variability (%)26.2321.2811.3911.59
Cumulative (%)26.2347.5158.9070.50
Table 7. Results after varimax rotation (high–high). Rotation method: varimax with Kaiser normalization; bold values indicate correlated variables included in the PCs > 0.40.
Table 7. Results after varimax rotation (high–high). Rotation method: varimax with Kaiser normalization; bold values indicate correlated variables included in the PCs > 0.40.
VariableFactor 1Factor 2Factor 3Factor 4Communality
Z2−0.0900.4870.498−0.1420.887
Aav−0.121−0.9280.0450.0210.889
I−0.069−0.905−0.0610.0700.842
NSA0.0300.876−0.064−0.0990.858
FC−0.104−0.008−0.4170.0540.447
Index AgL0.084−0.0640.836−0.0080.941
Index ArL−0.0170.3280.801−0.0260.800
ArLs−0.0430.8190.2520.0030.810
GDAgL0.307−0.0130.0260.8700.872
SDArL0.327−0.233−0.0360.8851.000
Index RP0.790−0.1420.1890.3390.866
DRP0.7010.130−0.0710.4120.775
Index V0.9510.117−0.0090.0930.951
RPav−0.831−0.0250.052−0.0190.776
ORP−0.909−0.1470.056−0.0330.889
Index H0.818−0.1440.1580.3040.970
Index Hs−0.048−0.0110.1640.0700.111
Table 8. Correlation matrix of the principal component analysis [Pearson (n)] for the low–low cluster. Values in bold are different from 0 with a significance level α = 0.05.
Table 8. Correlation matrix of the principal component analysis [Pearson (n)] for the low–low cluster. Values in bold are different from 0 with a significance level α = 0.05.
Z2AavINSAFCIndex AgLIndex ArLArLsGDAgLSDArLIndex RPDRPIndex VRPavORPIndex HIndex Hs
Z21
Aav−0.7251
I−0.7820.8211
NSA0.498−0.428−0.5521
FC−0.7890.7350.890−0.3611
Index AgL0.409−0.383−0.4740.179−0.5611
Index ArL0.254−0.250−0.2400.276−0.2220.6851
ArLs0.602−0.736−0.6280.396−0.5150.4710.6521
GDAgL−0.1340.1700.0540.1210.157−0.249−0.213−0.2571
SDArL−0.1450.1860.0650.1120.167−0.253−0.222−0.2740.9981
Index RP−0.037−0.143−0.0570.003−0.026−0.048−0.084−0.055−0.024−0.0401
DRP0.353−0.458−0.4630.414−0.4450.2710.1570.300−0.005−0.0470.2631
Index V0.095−0.368−0.3550.244−0.2670.3430.2890.3380.0470.0210.3480.5361
RPav−0.1570.4090.389−0.3190.281−0.387−0.411−0.468−0.039−0.016−0.364−0.432−0.8561
ORP−0.1740.4360.415−0.3510.294−0.372−0.348−0.417−0.070−0.054−0.303−0.391−0.8390.9361
Index H−0.020−0.168−0.079−0.010−0.061−0.002−0.057−0.022−0.044−0.0620.9870.3110.381−0.379−0.3131
Index Hs0.124−0.255−0.3110.367−0.1810.1730.2380.2180.1840.1580.4550.4350.634−0.741−0.6960.3931
Table 9. Percentage of variance and cumulative variance after varimax rotation (low–low).
Table 9. Percentage of variance and cumulative variance after varimax rotation (low–low).
Factor 1Factor 2Factor 3Factor 4
Variability (%)25.2322.2413.2612.74
Cumulative (%)25.2347.4760.7273.47
Table 10. Results after varimax rotation (low–low). Rotation method: varimax with Kaiser normalization. Bold values indicate correlated variables included in the PCs > 0.40.
Table 10. Results after varimax rotation (low–low). Rotation method: varimax with Kaiser normalization. Bold values indicate correlated variables included in the PCs > 0.40.
VariableFactor 1Factor 2Factor 3Factor 4Communality
Z20.888−0.007−0.084−0.0470.798
Aav−0.818−0.2360.134−0.1060.755
I−0.935−0.208−0.008−0.0370.918
NSA0.5080.2710.157−0.0480.359
FC−0.869−0.1060.115−0.0280.780
Index AgL0.4180.395−0.313−0.1900.465
Index ArL0.2350.490−0.313−0.2940.480
ArLs0.6140.396−0.293−0.1930.657
GDAgL−0.0630.0740.969−0.0400.950
SDArL−0.0760.0530.964−0.0550.941
Index RP−0.0110.226−0.0280.9430.942
DRP0.4320.3620.0210.2450.378
Index V0.1720.8160.0210.2300.748
RPav−0.189−0.9460.001−0.1780.962
ORP−0.221−0.881−0.050−0.1420.847
Index H0.0150.236−0.0590.9150.897
Index Hs0.1590.6800.1820.3040.613
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

Srejić, T.; Manojlović, S.; Sibinović, M.; Bajat, B.; Novković, I.; Milošević, M.V.; Carević, I.; Todosijević, M.; Sedlak, M.G. Agricultural Land Use Changes as a Driving Force of Soil Erosion in the Velika Morava River Basin, Serbia. Agriculture 2023, 13, 778. https://doi.org/10.3390/agriculture13040778

AMA Style

Srejić T, Manojlović S, Sibinović M, Bajat B, Novković I, Milošević MV, Carević I, Todosijević M, Sedlak MG. Agricultural Land Use Changes as a Driving Force of Soil Erosion in the Velika Morava River Basin, Serbia. Agriculture. 2023; 13(4):778. https://doi.org/10.3390/agriculture13040778

Chicago/Turabian Style

Srejić, Tanja, Sanja Manojlović, Mikica Sibinović, Branislav Bajat, Ivan Novković, Marko V. Milošević, Ivana Carević, Mirjana Todosijević, and Marko G. Sedlak. 2023. "Agricultural Land Use Changes as a Driving Force of Soil Erosion in the Velika Morava River Basin, Serbia" Agriculture 13, no. 4: 778. https://doi.org/10.3390/agriculture13040778

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