Next Article in Journal
WRF-Chem Modeling of Tropospheric Ozone in the Coastal Cities of the Gulf of Finland
Previous Article in Journal
Evaluation of Emission Factors for Particulate Matter and NO2 from Road Transport in Sofia, Bulgaria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

GIS-Based Integrated Multi-Hazard Vulnerability Assessment in Makedonska Kamenica Municipality, North Macedonia

1
Maarif International School—Skopje Campus, Kiro Gligorov 5, 1000 Skopje, North Macedonia
2
Institute of Geography, Faculty of Natural Sciences and Mathematics, Saints Cyril and Methodius University, Arhimedova 3, 1000 Skopje, North Macedonia
3
Department of Physical Geography, Faculty of Geography, University of Belgrade, Studentski trg 3/III, 11000 Belgrade, Serbia
4
Department of Geography, Tourism and Hotel Management, Faculty of Sciences, University of Novi Sad, Trg Dositeja Obradovića 3, 21000 Novi Sad, Serbia
*
Author to whom correspondence should be addressed.
Atmosphere 2024, 15(7), 774; https://doi.org/10.3390/atmos15070774
Submission received: 20 May 2024 / Revised: 14 June 2024 / Accepted: 26 June 2024 / Published: 28 June 2024
(This article belongs to the Section Biosphere/Hydrosphere/Land–Atmosphere Interactions)

Abstract

:
This study presents a comprehensive analysis of natural hazard susceptibility in the Makedonska Kamenica municipality of North Macedonia, encompassing erosion assessment, landslides, flash floods, and forest fire vulnerability. Employing advanced GIS and remote sensing (RS) methodologies, hazard models were meticulously developed and integrated to discern areas facing concurrent vulnerabilities. Findings unveil substantial vulnerabilities prevalent across the area, notably along steep terrain gradients, river valleys, and deforested landscapes. Erosion assessment reveals elevated rates, with a mean erosion coefficient (Z) of 0.61 and an annual erosion production of 182,712.9 m3, equivalent to a specific erosion rate of 961.6 m3/km2/year. Landslide susceptibility analysis identifies 31.8% of the municipality exhibiting a very high probability of landslides, while flash flood susceptibility models depict 3.3% of the area prone to very high flash flood potential. Forest fire susceptibility mapping emphasizes slightly less than one-third of the municipality’s forested area is highly or very highly susceptible to fires. Integration of these hazard models elucidates multi-hazard zones, revealing that 11.0% of the municipality’s territory faces concurrent vulnerabilities from excessive erosion, landslides, flash floods, and forest fires. These zones are predominantly located in upstream areas, valleys of river tributaries, and the estuary region. The identification of multi-hazard zones underscores the critical need for targeted preventive measures and robust land management strategies to mitigate potential disasters and safeguard both human infrastructure and natural ecosystems. Recommendations include the implementation of enhanced monitoring systems, validation methodologies, and community engagement initiatives to bolster hazard preparedness and response capabilities effectively.

1. Introduction

Natural processes and human impacts on the environment cause natural hazards. For example, excessive deforestation may accelerate soil erosion and landslides during heavy rainfall [1]. They profoundly impact health, economy, and society, leading to casualties, infrastructure damage, income loss, and altered consumer behavior [2,3,4,5,6,7,8,9]. While natural hazards may not always present the most significant threat to humanity [10], their impact and consequences can vary greatly, with different levels of vulnerability [11,12,13,14,15,16].
Excessive erosion represents a significant threat, especially to agricultural production. Although our study employs the erosion potential model (EPM) for soil erosion assessment, various other methodologies are available for this purpose. It is also worth noting that most of the global soil erosion assessments were carried out using the Universal Soil Loss Equation (USLE) or its revised versions (e.g., RUSLE) [17,18,19,20,21]. While these methods offer comprehensive approaches to erosion evaluation, we specifically chose the EPM due to its perceived advantages in the context of our study. The EPM’s simplicity and applicability to small municipalities, like ours, align with our research objectives focused on localized erosion risk assessment within specific geographic boundaries. Also, the erosion map of Macedonia was also created using the erosion potential method; it was created in the mid-seventies of the last century, and was published together with an interpreter in 1993. Because of the above, the results of this research are comparable to the results of previous research.
Mapping landslide-prone areas through Landslide Susceptibility Zonation (LSZ) is pivotal for addressing landslide vulnerability [22,23,24,25,26,27,28,29]. While methodologies such as the Analytical Hierarchy Process (AHP) are widely used in landslide susceptibility assessment, the frequency ratio (FR) method offers distinct advantages for our study. FR was chosen for its capability to handle categorical data effectively and its simplicity, which aligns well with our objectives and the available data.
Recent studies [15,30,31,32] emphasize Europe’s vulnerability to hydro-meteorological hazards, particularly flash floods, which threaten environmental stability, population centers, infrastructure, and socio-economic aspects. In SEE, including North Macedonia, flash floods caused by intensive rainfalls are a significant hazard [33,34], and pose a serious threat to both people and the environment. The Flash Flood Potential Index (FFPI) is widely regarded as one of the most effective methods for the assessment of flash flood (torrent) areas worldwide [35]. In addition to the FFPI, several other methodologies have been employed in various studies to analyze flash flood occurrences, such as the Hydrological River Basin Environmental Assessment Model (Hydro-BEAM) [36] and the Flood Intensity Index (Iw) [37]. However, in this study, we opted for FFPI due to its comprehensive evaluation of flash flood potential, considering terrain characteristics, lithological composition, and land cover attributes. GIS is a valuable tool for integrating various factors influencing flash flood hazard susceptibility [38]. While climate is the primary factor of floods, the hydrological response depends on physical geographic characteristics such as relief slope, soil texture, and land use. Deforestation, in particular, can significantly impact slope water propagation [39]. Numerous studies (e.g., [40,41,42,43,44,45]) have further explored the FFPI method for flash flood analysis at the municipal or regional level. Improving land use patterns, restoring forests and arable lands, and enhancing municipal functions are essential to mitigate the threats of flash floods, soil erosion, and landslides. These efforts require comprehensive studies and research to generate relevant information, which can then inform land use decisions and spatial planning in the watershed [46].
Access to forest fire vulnerability zone mapping is crucial to addressing forest fire challenges and devising appropriate solutions effectively. Remote sensing enables qualitative analysis of ecosystems, including forests, across various geographic and spatial scales. Understanding the factors contributing to forest fire-prone environments and fire behavior is crucial for effective forest fire management [47]. The Mediterranean coasts, characterized by a Mediterranean climate and fire-sensitive tree species, face significant forest fire vulnerability. In this study, the objective is to assess fire vulnerability zones in North Macedonia using a GIS-based AHP methodology [48,49,50,51]. While the AHP methodology is employed in this study for forest fire vulnerability zone mapping, other methodologies are also utilized in forest fire research. These include the use of statistical models like logistic regression and machine learning algorithms such as random forest. However, the AHP methodology stands out for its ability to incorporate expert knowledge and prioritize factors based on their relative importance, providing a transparent and systematic approach to vulnerability assessment.
Europe, especially southeastern Europe (SEE), faces significant hydro-meteorological hazards, resulting in environmental, social, and economic damage. Climate change exacerbates these risks, with projections indicating worsening conditions and increased damage [52,53,54,55,56]. As the global population grows and urbanization expands into disaster-prone areas, the vulnerability to natural hazards increases rapidly [57,58]. Given the substantial casualties and damage caused by natural hazards globally, there’s a growing interest in analyzing and assessing their risks [59,60,61,62,63,64,65]. Given the region’s favorable natural conditions and historical human impact, North Macedonia faces significant vulnerability to various natural hazards. Previous studies on natural hazards in the area have been conducted (e.g., [28,29,34,66,67]). However, the importance of employing multi-hazard techniques for comprehensive hazard analysis is emphasized [46,68,69,70,71]. Integrating probabilistic and deterministic stochastic processes can provide further insights.
This study contributes original insights by conducting a comprehensive assessment of natural hazard vulnerability in North Macedonia, particularly focusing on Makedonska Kamenica municipality. Through the development of hazard maps and identification of hot-spot areas, the research aims to facilitate effective mitigation measures at both regional and national scales. Additionally, the study emphasizes the importance of integrating probabilistic and deterministic stochastic processes to gain further insights into hazard analysis. These efforts are informed by rigorous research findings aimed at facilitating informed decision-making and spatial planning processes.

2. Materials and Methods

2.1. Study Area

Makedonska Kamenica municipality (in further text: MKM) is located in the northeastern part of North Macedonia. It is situated at the foot of the Osogovo Mountains on its southern slope. On the northwest are borders with Kriva Palanka municipality, on the west with Kočani municipality, on the south with the municipalities of Vinica and Delčevo, and on the north-east is the country border with the Republic of Bulgaria. It occupies a total area of 190 km2. The region encompasses rocks dating from the Precambrian, Paleozoic, Pliocene, and Quaternary periods. Dominantly, it features Precambrian gneisses and Paleozoic schists. Pliocene clastic lake sediments and Quaternary River sediments are prominently located along the rivers and Kalimanci Lake. A small portion comprises Precambrian amphibolite, Paleozoic compact volcanic rocks, and Paleozoic granitoids [72].
The topography of MKM ranges between 401 m and 2191 m asl. The lowest is the southern part (Kalimanci Lake), and the highest is the mountain area (Osogovo Mountains) in the northern part of the municipality (Figure 1). The municipality’s average slope measures 18.9°. The south aspect is dominant, with the second being eastern, closely connected with the direction of the highest Osogovo Mountain ridges and the Kamenica valley between them. The terrain ruggedness index (TRI) in the specified area ranges between 0 and 173 (m/km2), and the average value of 65.8 m/km2 is relatively high, indicating a significantly rugged and mountainous landscape. This index quantifies the terrain’s roughness or heterogeneity by measuring the variability in elevation within a given area [73].
For the examination of temperature patterns, the ERA-5 dataset was employed due to its superior resolution in temperature measurements. In contrast, for the analysis of precipitation, WorldClim data were selected for their unparalleled resolution in precipitation data, ensuring meticulous accuracy and reliability, especially in regions characterized by pronounced variability. Thus, according to ERA-5 reanalysis data for 1999–2023, the average temperature ranged from 3.8 to 10.6 °C. The average precipitation in the specified period (1980–2020) using the WorldClim 2 dataset [74] ranged from 603 to 816 mm (Figure 2).
Hydrologically, the main watercourse in the study area is the Kamenica River, a right tributary of the Bregalnica River (Figure 3). Its source originates from the Osogovo Mts. at an elevation of 1780 m and through the Kalimanci Lake (reservoir) flows into the Bregalnica River on 510 m asl. Notable left tributaries include the Selška and Vlainska Rivers, while on the right, it is joined by the Gatešnica and Sušica Rivers. Another significant watercourse within the MKM is the Kosevička River, originating to the south of the Šamska Čuka (1514 m); it is also known as Lukovička Reka, ultimately emptying into the Bregalnica River via the Kalimanci Lake. The Bregalnica River (225 km), the second-longest river in North Macedonia, traverses the southern part of the municipality. These rivers are all part of the Bregalnica catchment area. Lake Kalimanci is located at 510 m in the Istibanja Gorge, where the Kamenica River meets Bregalnica.
According to the CORINE Land Cover [75], the predominant land cover types consist of areas characterized by complex cultivation patterns and land primarily dedicated to agriculture. Notably, significant portions are covered by natural vegetation (29.4%), broad-leaved forests (22.4%), and transitional woodland-shrub habitats (19.9%). Mixed forests cover an area of 9.5%, followed by pastures (8.6%) and coniferous forests (8.0%). The forests have undergone significant degradation due to past destruction, likely stemming from various human activities such as deforestation, logging, and agricultural expansion. Water bodies, including parts of Kalimanci Lake, account for 1.4% of the land cover, while artificial surfaces comprise 0.8%.
According to the soil map of Macedonia [76], the municipality’s largest area is covered by Cambisols and a complex of Cambisols and Leptosols, predominantly situated in the northern (mountainous) part. The second largest soil type in the area is Ranker, which also prevails in higher elevations. In the southern part, a dominant presence has a complex comprising Ranker, Regosol, and Leptosol, along with Chromic Luvisols on saprolite. Additionally, Albic Luvisol, Regosol, Ranker, and Leptosol complexes are characterized. The soils are quite eroded and sandy (due to geology). Fluvisols are found adjacent to rivers, particularly at their downstream parts toward Kalimanci Lake.
By socio-economic overview, the municipality comprises nine settlements, eight of which are rural: Todorovci, Lukovica, Kostin Dol, Kosevica, Moštica, Dulica, Sasa, and Cera, with the town of Makedonska Kamenica as the municipal center. According to the last census in 2021 [77], the Municipality of Makedonska Kamenica has a population of 6439, which is significantly less than the census of 2002 (8110 inhabitants), with the highest decrease in the agricultural population. Thus, the municipality has a predominantly rural settlement pattern with an underdeveloped road network, limiting accessibility and economic development. Also, it lies on an important road connecting the eastern and central parts of the republic. Even though the local roads are not great, this main road helps connect the region. All these aspects are important, and the hazard vulnerability assessment focuses on analyzing the causal interaction between natural and anthropogenic systems within the landscape.

2.2. Methodology for Soil Erosion Assessment

The research procedure for the GIS-Based Integrated Multi-Hazard Vulnerability Assessment in MKM, North Macedonia involves a comprehensive approach utilizing various models to assess the most pronounced hydro-meteorological and climatological hazards within the case study. The procedure encompassed gathering the relevant geospatial data, utilizing the erosion potential model (EPM) to assess the susceptibility of the area to erosion, quantifying the landslide susceptibility index (LSI) to identify areas prone to landslides, implementing the Flash Flood Potential Index (FFPI) to evaluate the likelihood of flash floods in the area, assessing the risk of forest fires using a dedicated model that considers variables such as vegetation type, temperature, humidity, and historical fire occurrence data, and integrating the outputs of the individual hazard models into a comprehensive multi-hazard assessment. This model combines the results of the EPM, LSI, FFPI, and forest fires model to identify areas vulnerable to multiple hazards simultaneously in a spatial-temporal context. In the final step, validation of the models and assessment results using ground-based data and expert knowledge was performed to acquire the information necessary for refinements of the models and methodologies based on feedback and additional data if necessary.
The flow chart in Figure 4 presents all procedures and approaches utilized for this research.
For the assessment of soil erosion intensity and sediment yield, the erosion potential model (EPM), also known as the Gavrilović method [78], is widely utilized in the Balkans and worldwide, including Serbia, Croatia, Slovenia, Germany, Italy, Argentina, Belgium, North Macedonia, Greece, etc. (e.g., [29,79,80,81,82,83,84]). It reliably assesses soil erosion rate, mean annual soil loss, sediment yield, erosion control works, and torrent regulation on a regional scale. While convenient for areas with limited erosion data, the EPM does not delve extensively into the underlying physics of erosion processes. The method relies on the following equation:
Wy = T ∙ H ∙ π ∙ √Z³ ∙ F
where W represents the average annual sediment yield in m3; T is the temperature coefficient, calculated as T = (0.1 ∙ t + 0.1) ^ 0.5, with t being the annual mean air temperature; H indicates the annual precipitation in mm; Z is the erosion coefficient, ranging from 0.1 to 1.5 and beyond; and f is the studied area in km2. Among these factors, the coefficient Z holds the utmost importance, as it combines rock erodibility/erosion resistance (Y), land cover index (Xa), index of visible erosion processes (φ), and mean slope of the catchment (J) in the ratio:
Z = Y ∙ Xa ∙ (φ + √J)
Contrary to the traditional method, the GIS approach for implementing the erosion potential model (EPM) heavily relies on deriving most parameters from digital elevation models and satellite images [85,86,87,88]. For the Y coefficient, a previously digitized geological (100 k) [72] and soil map (50 k) [76] was rasterized to 15 m to correspond to the DEM resolution used, and erodibility values were added according to Gavrilović’s approach [78]. These values generally range from 0.1 for very resistant rocks to 2.0 for non-resistant rocks and soils. However, accurately quantifying coefficient Y poses challenges because of the different soil and rock compactness of the field. Thus, a fitting procedure in the form of rooting Y = sqrt (Y1) is used [82]. The land cover index, Xa, is derived from the CORINE Land Cover model [75], with values ranging from 0.1 (indicating dense forests) to 1.0 (indicating bare rocks), and they are based on suggestions from the original model. The values for coefficient φ (visible erosion processes) are determined from the spectral band 4 of suitable cloud-free Landsat 8 scene in the form φ = a/255, where a is the greyscale pixel value. The results are between 0 and 1, where the lower values indicate areas without visible erosion processes, while higher values show sites with severe erosion processes. Recently, for the coefficient φ, some researchers have used the so-called BSI (Bare Soil Index), based on Landsat 8 or Sentinel-2 imagery [45]. However, further research will show the eventual advantage of this approach. The average terrain slope (J) was calculated from a 15 m digital elevation model (DEM), expressed in decimal percentage. It is recognized that higher slopes correspond to reduced stability, increased erosion potential, and increased susceptibility to torrential floods, as noted by [89]. Using this, the GIS-calibrated coefficient Z is determined by the equation:
Z = √(Y) × φ × ((Xa + φ) × log(a +1) + √(a/57.3))
Climate parameters (T and H) were derived from ERA-5 average monthly data for precipitation and air temperature, as well as MODIS land surface temperature (LST) via Google Earth Engine. These data, spanning from January 1999 to December 2023, produce raster models of average air temperatures and precipitation for the MKM, forming the basis for T and H coefficients in the erosion potential model (EPM). Their accuracy is verified against measurements from nearby meteorological stations (Kočani, Berovo, and Delčevo), with a high accuracy level of 97.5%. Multiplied, climate potential and erosion coefficient Z determine the mean annual erosion loss (Wy) for the catchment. The GIS-based EPM approach has been tested in various regions in North Macedonia [29,34,82,90], and other countries as well [79,91,92,93,94,95,96]. In assessing the accuracy of the EPM, a common approach involves comparing its results with the measurement of sediment deposition in reservoirs [97,98,99]. This comparison consistently reveals strong correlations between the model’s outputs and measured and observed data. Field investigations were conducted to check and correct the obtained coefficient Z.

2.3. Methodology for Landslide Susceptibility Assessment

Determining landslide-prone areas on a regional scale is a multifaceted endeavor influenced by many natural and anthropogenic factors. Previous research on areas of similar size has provided valuable insights into landslide susceptibility assessment (LSA) [28,100]. Moreover, the work of Milevski and Dragićević [101] underscores evaluating six primary triggering factors: lithology, slope angle, land cover, terrain curvature, distance from rivers, and distance from roads. These factors are the foundation for various LSA methods [102]. They are typically categorized into internal and external factors, with internal factors affecting the mass’s shear strength and external factors surpassing its strength. Topography, lithology, and land use are considered essential factors for GIS analysis [23], although other factors like tectonic features and rainfall distribution can also be influential [103]. In this research, following the study of Milevski et al. [28], we specifically consider six causative factors: lithology, slope, plan curvature, land use, distance from streams, and distance from roads (elaborated further in the following Table 1).
After the preceding steps, the next crucial phase was selecting a method for assessing landslide susceptibility (LS). Statistical analysis and frequency ratio methods are effective in extensive study areas lacking comprehensive landslide inventories. This involves correlating landslides’ spatial distribution with conditioning parameters to determine the LS approach [104]. This study evaluates landslide susceptibility occurrence with conditioning factors (e.g., slope, lithology, land cover) through the landslide susceptibility index (LSI). This involves calculating the index for each factor category and estimating susceptibility based on individual characteristics. Weighting factors compare calculated landslide density with the overall density in the study area. This comparison helps assess the relative significance of each variable concerning landslide susceptibility [105]. It can be formulated as:
Wij = 1000(fij − f) = 1000(Aij × /Aij × A × /A)
This model indicates areas with varying levels of susceptibility to landslides, considering multiple factors. Weight values for the LSA model factors are provided as follows:
  • Wij: the importance of a certain class (i) of parameter (j).
  • fij: the landslide density within the class (i) of the parameter (j).
  • f: the overall landslide density across the entire model.
  • Aij*: the landslide susceptibility in a specific class (i) of the parameter (j).
  • Aij: the surface area of a particular class (i) of the parameter (j).
  • A*: the total area of landslides across the entire model.
  • A: the total area of the entire model.
As per [28], a methodology akin to the previous one is adopted for reclassifying LSI values into distinct susceptibility zones and conducting map validation. Each factor is assigned a weighted value, with 30 for slope, 15 for lithology, 10 for land cover, 8 for curvature, and 2 for stream and road distance, totaling 67. These weights signify the relative importance of each factor in triggering landslides within the catchment, with slope being the most influential factor, followed by lithology, land cover, planar curvature, and distance to streams and roads. The cumulative sum of all six parameters determines landslide susceptibility, computed by aggregating values for each grid cell across all layers. The resulting model is classified into five distinct susceptibility classes using natural breaks classification, ranging from very low to very high susceptibility.

2.4. Methodology for Flash Flood Potential Index (FFPI)

The Flash Flood Potential Index (FFPI) is a widely utilized statistical method for identifying areas vulnerable to flash flooding [34,35,43,45,106,107,108,109]. It integrates various factors such as slope, vegetation cover, land use, and soil type (in this case, represented by lithological type due to data constraints) to provide a quantitative susceptibility assessment. Flash floods often occur in municipality areas with reduced soil infiltration rates, typically due to bare terrain, steep slopes, deforestation, urbanization, or industrialization. The combination of intense precipitation, snowmelt, terrain characteristics, soil properties, and vegetation density (including forests, shrubs, and grasses) influences susceptibility to flash flooding (e.g., [110,111,112,113,114,115]). Less vegetated areas generally have a higher vulnerability.
To enhance the accuracy of this study, GIS and RS analysis were conducted to identify areas prone to flash floods within the MKM. These analyses employed the Flash Flood Potential Index (FFPI), a statistical method used to assess flash flood susceptibility. Due to limited soil data from the available soil map of Macedonia (50 k) [76], lithological type was considered instead of soil type. Thus, the model incorporated weighting factors, including slope, vegetation cover, lithology, and land use. The FFPI method is formulated as follows [35]:
FFPI = (M + S + L + V)/4
The Flash Flood Potential Index (FFPI) incorporates several factors, specifically terrain slope (M), lithology (S), land use (L), and vegetation index (V).
  • Terrain slope (M) was derived from a 15 m digital elevation model (DEM). Slope percentages were calculated, and the formula M = 10n/30 was applied, where n represents the average slope. If n is 30% or higher, M is set to 10. The municipality’s average slope is 32.1%, resulting in M = 10.
  • For lithology (S) analysis, a digital lithological map based on the geological map of Macedonia (100 k) [72] was utilized (basic geological map). Rocks were classified on a scale from 1 to 9 based on erosion susceptibility.
  • Land use (L) data were obtained from the CORINE Land Cover CLC2018 database [75]. Classes were assigned values from 1 to 10 based on their impact on flash flood occurrence.
The Bare Soil Index (BSI) served as the calculation’s vegetation density index (V). Sentinel-2 satellite images from Google Earth Engine script, spanning the period from 1999 to 2023, were utilized for BSI derivation, offering a robust method for assessing erosion rates, often associated with flash floods. Remote sensing, particularly in erosion and flood assessment, provides significant advantages. The BSI was calculated using the formula:
BSI = ((Red + SWIR) − (NIR + Blue))/((Red + SWIR) + (NIR + Blue))
where Red represents the red spectral channel, SWIR is the short-wave infrared spectral channel, NIR is the near-infrared spectral channel, and Blue is the blue spectral channel. BSI values range from −1.9 to 0.7, with an average of −0.9. To prevent negative values in the vegetation index (V) formula, a constant value of 1 was added. Thus, V = 7.68 × ln(BSI + 1) + 8.

2.5. Methodology for Forest Fires

Forest fires, besides weather conditions, are influenced by various factors [116]. Human negligence and the rise in global temperature led to extensive fire outbreaks annually worldwide, including in North Macedonia. While 97% of forest fires result from human activities, natural conditions also significantly affect a region’s fire susceptibility [53]. In MKM, forest fires, especially in summer, regularly occur, adversely affecting the environment. Despite the municipality’s forest cover (39.9% of the area), the forested area diminishes yearly due to human-induced factors like forest fires and illegal logging. Conducting a vulnerability AHP analysis for forest fire occurrence could aid in implementing preventive measures to safeguard forests. The Analytic Hierarchy Process (AHP) is used for multicriteria analysis and synthesis mapping [117,118]. It prioritizes criteria by quantifying them hierarchically [119,120,121], requiring an understanding of the research domain and physical laws [122,123]. AHP’s key feature is its use of subjective judgment to assign weights, informed by previous research [124]. This enhances objectivity by reflecting expert opinions on the importance hierarchy. AHP is esteemed for expert scenario analysis and decision making, systematically evaluating objectives, criteria, sub-criteria, and alternatives. In forest fire research, criteria vary in importance, necessitating different weight coefficients. AHP uses a 9-point scale for pairwise comparisons, from 1 (equal importance) to 9 (extreme importance) [117,118]. This study simplifies the scale to 1–5 to reduce subjectivity, with 1 as the most significant and 5 as the least. SAGA GIS 9.3.0 software was used for data processing, incorporating various factors from previous research. Thus, for forest fire hazard analysis, the formula developed in one study [125] was used:
RC = 7VT + 5(S + A) + 3(DR + DS)
In this equation, RC is the numerical index of forest fire vulnerability zones where VT indicates vegetation type with 5 classes, S the slope factor with 5 classes, A the aspect variable with 4 classes, and DR and DS indicates distance factor from road and settlement [47]. Terrain slope and aspect were determined using a 15 m resolution digital elevation model (DEM). Slope values, expressed in degrees, were categorized, with steeper terrains receiving higher values due to reduced accessibility. The vegetation type index was derived from the CLC2018 geospatial database [75], assigning values from 1 to 5 based on the predominant vegetation in each area. Water surfaces and large settlements were deemed the least susceptible, as fire vulnerability in such areas is minimal [45]. To calculate the distance index from roads, settlements, and buildings, satellite images and topographic maps were digitized, creating buffer zones of suitable width. Finally, based on this analysis carried out, a fire vulnerability zone map was produced.

2.6. Methodology for Multi-Hazards

A hierarchical framework was established to model overall susceptibility to natural hazards in MKM, considering occurrence probability, frequency, and consequences of disasters. The AHP method quantifies relative priority using an appropriate value scale. With the global natural risk rising, innovative approaches are emerging to assess and zone areas susceptible to landslides, flash floods, excessive erosion, and forest fires. According to the UNDRR [126], the multi-hazard concept involves identifying major hazards and considering their potential simultaneous, gradual, or cumulative effects, including interrelated or cascading impacts.
Multi-hazard techniques are vital for hazard event analysis [29,34,46,68,69,70,71,127,128], with probability and deterministic stochastic processes offering an alternative approach. Thus, it is necessary to develop models for assessing potential erosion, landslides, flash floods, and forest fires-prone areas in the MKM, North Macedonia. This study aims to (1) assess multiple hydro-meteorological hazards using a modified erosion assessment model and systematic approaches, (2) create natural susceptibility and vulnerability maps, and (3) pinpoint high- vulnerability erosion areas for implementing prevention measures.
To identify multi-hazard areas, QGIS (v. 3.34.2) and SAGA GIS (v. 9.3.0) software were utilized with appropriate GIS tools. Values showing high vulnerability and susceptibility to natural hazards were considered, followed by reclassification and area calculations using the Grid Calculus tool. Different models were developed to enhance the analytical process.

3. Results

3.1. Erosion Hazard Modeling

A map was prepared for the MKM (Figure 5) using SAGA GIS (v. 9.3.0) to calculate the total amount of eroded material based on the previously obtained coefficient of erosion Z (erosion vulnerability).
The obtained results of the erosion coefficient (Z) reveal a substantial prevalence of areas with medium, high, and very high vulnerability (values exceeding 0.4) within the MKM (Table 2). These areas cover 82.7 km2, equivalent to 43.6% of the municipality’s territory. They are prone to significant soil erosion and sediment transport, even during moderate rainfall. The land undergoes intense rain splash and erosion processes, leading to the production, transportation, and accumulation of deposited material. This phenomenon is especially pronounced during intense rainfalls with rates surpassing 0.5 mm/min or extended periods of heavy rainfall episodes. The accumulated eroded material challenges cultivated areas, roads, and other infrastructure. The mean value of the erosion coefficient Z within the MKM stands at 0.61. Several factors contribute to these elevated erosion coefficients, including the dominance of erodible lithologies, such as Precambrian mica shists, gneisses, Pliocene, and Quaternary clastic sediments, as well as the absence of vegetation and steep terrain slopes.
The model calculation suggests a total erosion production (W) of 182,712.9 m3/year and a specific erosion rate (Wspec) of 961.6 m3/km2/year. The municipality’s most erosive sites are the steep valley sides of the Kamenica River and tributaries, which signify a considerable erosion rate. These areas, especially lacking vegetation or sparse grassy cover, are highly susceptible to erosion processes exacerbated by rainfall (Figure 6).
A substantial 9.4% of the entire area is under very high erosion vulnerability, exceeding 2000 m3/km2/year, equivalent to a soil loss of 2 mm per year (Table 3). These high-vulnerability areas experience excessive erosion, resulting in various relief landforms, loss of fertile land, and significant sediment yield accumulating in the riverbeds of Kamenica River and its tributaries as well as in Kalimanci Lake (Figure 7). In contrast, in the northwestern, higher parts of the municipality, erosion potential remains within natural values below 500 m3/km2/year, owing to the presence of well-forested areas. To tackle degradation and mitigate the further loss of natural resources, particularly soil and water, targeted preventive erosion control works must be implemented in areas where erosion intensity exceeds 1000 m3/km2/year. These measures are vital for preserving ecosystem integrity and safeguarding the valuable resources within the region.
Also, the procedure calculates the proportion of sediments reaching Kalimanci Lake (reservoir) as not all sediments exit the Kamenica River catchment. Hence, with the second part of the EPM approach, the sediment delivery ratio (Ru) is estimated using the following equation:
Ru = √(OD)/0.25 × (L + 10)
where O represents the length of the watershed border in km and D signifies the difference between the average altitude and the altitude of the catchment outlet in km [78]. The sediment yield (G) is then calculated as:
G = W × Ru
According to the calculations performed, the Ru value for the study area is 0.7 and the average annual sediment yield (G) from the MKM to the Kalimanci Lake totals 127,970 m3/y (Table 4). This value aligns with one study’s echo sonar measurements of sediment deposition in Kalimanci Lake [129]. According to this study, the measured average annual sediment deposition in Kalimanci Lake is 214,325 m3/year, from which about 1/2 originate from the Kamenica and Lukovica Rivers, both in the scope of the municipality. That shows the high accuracy of our EPM approach, which needs to be further checked with additional measurements on the mentioned rivers.

3.2. Landslides Susceptibility Modeling

Based on the probability map of landslide occurrence in the MKM, generated using the frequency ratio (FR) method and the data provided in Table 5, it is apparent that class 5 prevails with the highest representation at 31.8%. This class signifies a very high probability of landslides occurring, notably observed along the steep valley sides of the municipality and the valley areas of Kamenica and Bregalnica rivers, its left and right tributaries.
Class 5, representing the largest proportion at 31.8%, predominantly occurs in areas characterized by mica shists, gneiss, clay and sand, shales, and shale granites. This distribution signifies a pronounced likelihood of landslides, particularly prevalent along the left and right banks of the Kamenica River downstream of Makedonska Kamenica, within the Lukovica River valley, and adjacent to the confluence with Kalimanci Lake. The second-largest dominant class is 3 (31.2%), which indicates average susceptibility to landslides, particularly notable on the steep valley sides of MKM, especially in the valley parts of the right Kamenica River tributaries. This area predominantly consists of mica shists and gneisses. Class 4, comprising 18.7% of the total area, signifies a high intensity of landslide occurrence, predominantly along the valley and left tributaries of the Kamenica River and in the higher mountainous part, with shists being the predominant lithological composition. Class 2, constituting 11.5% of the total area, denotes a low probability of landslide occurrences, primarily confined to a small region within the mountainous parts of the left and right tributaries of the Kamenica River. This class encompasses areas where quartz latites and amphibolites are prevalent. The first susceptibility class (class 1), which constitutes 6.7% of the total area, is defined by areas exhibiting a very low intensity of landslide occurrence. These regions are predominantly situated in the northeastern part of the municipality, particularly in the higher (source) regions of the Kamenica River’s left tributaries, where quartz latites are notably present.
Validation of the landslide susceptibility analysis (LSA) involved comparing recorded landslide locations from field surveys (totaling 20) with the LSA zonation. The comparison results are detailed in Table 6.
Among the 20 landslides observed, 10 (or 50%) are situated within the highly susceptible LSI zone. Considering the very high (class 5) and high susceptible zones (class 4), this combination encompasses 85.0% of the recorded landslides, indicating a robust accuracy level for the model employed. No landslides were recorded in the field survey within class 1, which denotes a low probability of landslides. This validation reinforces our results’ accuracy and affirms the LSI model’s reliability.
The landslide hazard map was cross-validated with the national study of North Macedonia [28], revealing an overlap with areas highly susceptible to landslides (Figure 8). Factors contributing to landslides include the terrain slope, soils, lithology, and land use. The predominant clay composition of the soil poses a significant risk during heavy rainfall. Alterations in surrounding forest land use can exacerbate this vulnerability. Implementing preventive measures to safeguard the residents of local villages and towns is crucial.
Biosoft easyROC online platform (biosoft.erciyes), was employed to utilize the Receiver Operating Characteristic (ROC) curve as a multi-hazard validation technique, which is widely utilized in geospatial modeling. This curve allows for the visualization of the false positive fraction in relation to the true positive across all values utilized in generating the modeling results. The ROC curve graphically represents specificity on the x-axis and sensitivity on the y-axis (Figure 9). Furthermore, for validation purposes, an automatically calculated Area Under the Curve (AUC) was generated on the ROC curve. AUC values serve as indicators of the success and accuracy of a given model concerning the reference data, ranging from excellent (AUC = 0.9–1) to failed models (AUC = 0.5–0.6), with corresponding categories of good, fair, and poor models falling within specified ranges [29]. To ensure proper validation, it is recommended to have 2–3 times more false-positive landslides than true ones in the validation dataset. In this study, 60 false-positive landslides were randomly selected and thoroughly inspected, resulting in an AUC value of 0.82, showing a high level of accuracy for the employed model.

3.3. Flash Floods Hazard Modeling

The slope map, derived from the 15 m digital elevation model (DEM), plays a crucial role in hydrological processes, influencing runoff timing and infiltration rates. Infiltration rates typically decrease as the slope angle increases [130]. Across the entire area, the average relief slope is 18.9 degrees. Slope values are then converted to percentages and classified accordingly. Subsequently, the model assigns an FFPI value ranging from 1 to 10. Any slope exceeding 30 degrees is categorized with an FFPI value of 10.
In this study, the lithology index was derived from the lithological map, as documented in [72], based on five primary lithological units: clastic sediments, compact volcanic rocks, granitoids, amphibolite, gneisses, and mica schists. The susceptibility of these lithological units to torrential floods was analyzed and classified accordingly (see Table 7). River clastic sediments received the highest coefficient (9), indicating their heightened susceptibility to flash floods. In contrast, compact volcanic rocks and amphibolite received the lowest coefficient (1), suggesting their minimal susceptibility to this natural hazard. This index holds significant importance as the composition of these units influences infiltration rates and runoff dynamics during intense rainfall events. Compact volcanic rocks, amphibolite, and granitoids are less likely to contribute to flash floods due to their erosion-resistant nature. Conversely, clastic sediments and mica schists are more prone to erosion and transport during flash floods, increasing the risk of flooding.
Table 8 illustrates the various land use types present in the study area. The land types most vulnerable to the Flash Flood Potential Index (FFPI) are those characterized by intricate cultivation patterns and predominantly utilized for agriculture, albeit with sizable portions of natural vegetation. Conversely, areas covered by coniferous and mixed forests exhibit the lowest susceptibility to FFPI (water bodies as well). Utilizing the land use map, we generated and classified the land use index into the FFPI (Table 8).
The methodology outlined above has provided comprehensive insights into the potential for flash floods and erosion intensity across the study area. By analyzing Sentinel-2 satellite imagery and utilizing the Bare Soil Index (BSI), we have identified areas more vulnerable to flash floods with enhanced precision (in the GEE platform). Furthermore, the correlation between vegetation density and erosion rate has furnished valuable insights for formulating effective land management strategies and implementing measures to mitigate the impact of flash flood events. Integrating remote sensing techniques and BSI computation has significantly advanced our understanding of erosion dynamics and the susceptibility of the study area to flash floods. The calculated coefficient V ranges from 6.5 to 10.4, with an average of 9.6. A vegetation index is generated within a range of values from 1 to 9.
Utilizing the GIS and RS database, we analyzed the FFPI index to assess flash flood vulnerability across the MKM (Figure 10). Our calculations (Table 9) reveal that most of the municipality area, constituting 41.7%, falls within the class characterized by a high probability of flash floods. The second by representation is the moderate class with 70.3%, and areas with very high susceptibility to floods collectively occupy 3.3%. The very high susceptibility class primarily encompasses steep slopes along the tributaries flowing into the Kamenica River, particularly downstream towards the estuary of the Kalimanci Lake, where the Kamenica River converges with the Bregalnica River. The Flash Flood Potential Index (FFPI) yields an average value of 5.7, ranging from 2.4 to 8.1.
For the validation of the model, comparisons with the location of actual flash flood events recorded in the media and the reports of the Crisis Management Center of North Macedonia are made. Thus, in the last twenty years, frequent flash floods have been recorded in the catchments of Mostička Reka, Lukovička Reka, Sušica, and the middle part of the Kamenica River. By field prospections, huge amounts (approximately 1.5–2.0 million m3) of fresh deposits are visible in the floodplains of these rivers, indicating recent flash floods, which overlap with the model zoning. Given the limitations of basic data sources (systematic historical observation), the availability of flash flood data needed for the ROC analysis in this study is limited.

3.4. Forest Fires Hazard Modeling

Utilizing geographic information systems and remote sensing data, we conducted analysis and data processing to generate a forest fire susceptibility map for the municipality of Makedonska Kamenica. The susceptibility was categorized into four classes: low, medium, high, and very high, as illustrated in Figure 11.
Considering the extensive forested area covering 75.9 km2 and the considerable terrain slope, the municipality is susceptible to forest fires (Table 10). Areas with steep slopes and southern exposure are particularly prone to such hazards. These locations typically consist of coniferous, broad-leaved, or mixed forests and are relatively proximate to roads and settlements.
The analysis of forest fire susceptibility indicates that slightly less than one-third of the municipality’s area is very highly susceptible to forest fires (31.5%). Additionally, 24.2% of the territory is highly susceptible to forest fires. Moderate susceptibility encompasses 18.6% of the municipality, and low and very low susceptibility together cover 25.7% of the municipality’s area.
Similarly to flash floods, forest fire data in this study is limited. Therefore, further research should focus on expanding the data sources for these parameters in order to improve the modeling accuracy, resulting in more precise ROC/AUC values.

3.5. Modeling the Overall Susceptibility to Natural Hazards (Multi-Hazard Modeling)

In line with the study’s objectives, a multi-hazard model was constructed by integrating erosion, flash floods, landslide susceptibility, and forest fire maps (Figure 12). Thus, regions exhibiting high erosion potential were overlaid with those highly prone to landslides, floods, and forest fires as well, within the territory of MKM. This process identified areas facing concurrent vulnerability from fourth hazards, termed multi-hazard zones.
The analysis disclosed that segments of MKM are susceptible to multiple hazards, specifically flash floods, forest fires, landslides, and excessive erosion. Designated as the “total vulnerability” areas within the study, these zones encompass around 11.0% of the municipality’s total area (refer to Table 11).
Located predominantly in the upstream areas of MKM, the valleys of the right tributaries of the Kamenica River, and the estuary in the Bregalnica River, these multi-hazard zones exhibit distinctive terrain characteristics. These include deforested areas, exposed soil, steep slopes, and susceptible rock formations. It is worth noting that infrastructure within these high-vulnerability zones faces potential danger. To enhance environmental management effectiveness, it is crucial to pinpoint high and very high susceptibility areas to natural disasters across all settlements. Implementing protective measures in these identified areas will mitigate the likelihood of these analyzed disasters occurring.

4. Discussion

Compared to the entire country area of North Macedonia, the MKM exhibits higher erosivity, with a coefficient Z of 0.6 compared to 0.35. Additionally, the average specific erosion rate (Wspec) for the MKM is estimated to be 961 m3/km2/year, whereas for the entire North Macedonia is reported as 681 m3/km2/year according to [82]. These results indicate that the average annual erosion in the study area is nearly twice as high as the average specific erosion rate for the entire territory of North Macedonia. Also, a comparison of the mean value of the coefficient Z with Pehčevo municipality (in the eastern part of N. Macedonia) is made, which is 0.44 [67]. Additionally, a comparative analysis was conducted with a study area of similar size, namely the municipality of Štrpce in southern Serbia. Research [45] indicated a mean Z erosion coefficient of 0.34 for Štrpce. This finding is consistent with observations from other studies (e.g., [29,34,131,132,133]) emphasizing the interconnectedness of soil erosion intensity with various factors such as natural conditions, demographic characteristics, settlement patterns, and land use dynamics.
The municipality’s total sediment yield (G) is 127,970 m3/year. Compared to [129], the annually measured sediment deposition within Kalimanci Lake totals 214,325 m3, with equitable distribution, approximately half originating from the Kamenica and Lukovica Rivers in Makedonska Kamenica River (Figure 13). The specific sediment yield (Gspec) for the area is calculated to be 673.5 m3/km2/year. Despite proposed erosion control works, this area remains critically prone to erosion, necessitating focused attention on municipality and land management strategies geared towards the conservation and protection of the land.
This study analyzed six factors influencing landslide occurrence: slope, lithology, land cover, plan curvature, distance from streams, and distance from roads. Results show that about one-third of the catchment area has a very high probability of landslides, especially in the lower part of MKM. Overall, 31.8% of the municipality is in the very high susceptibility class. In North Macedonia, regions prone to landslides are mainly hilly areas with Neogene lacustrine sediments and slopes steeper than 10° (Figure 14). The MKM shows even higher susceptibility than the national average. Urgent action is needed to mitigate landslide vulnerability in this area [28]. Exploring alternative model validation approaches, as suggested by [134,135,136], is advisable to enhance the reliability of susceptibility models. A comprehensive assessment of total susceptibility to natural hazards, advocated by [29,137], offers valuable insights for protecting vulnerable areas and cultural heritage. Validating the landslide susceptibility model is straightforward, as it only requires a basic inventory. Comparing the 20 recorded landslides with the model’s zones, we found an 85.0% overlap with the high and very high susceptibility areas. Validation was performed using the Area Under the Curve (AUC) from Receiver Operating Characteristic (ROC) curve analyses, resulting in a value of 0.82. Further research, including precise LIDAR measurements and advanced technologies such as deep learning, is needed to assess the region’s erosion and landslide vulnerability changes.
In this study, we noticed a flash flood susceptibility. Unlike studies that predominantly examine natural flash floods within municipalities [138,139,140], our study area experiences urban flash floods influenced by infrastructure such as drainage systems. The absence of certain key factors in our analysis may have impacted our findings, including the significant role of precipitation in triggering flash floods during intense rainfall. Precipitation is a key factor in flash flood validation due to its intensity and spatial spread. Integrating WorldClim 2 data [74] enhances the vulnerability assessment. Intense rainfall often triggers flash floods in the municipality, peaking in May and November. While lacking a local pluviometry station, GIS and remote sensing models helped analyze rainfall data. Areas at high vulnerability typically receive 600–815 mm of precipitation. Research confirms these findings, suggesting lower rainfall areas experience reduced vegetation and slower hazard recovery. Geological composition and deforestation also heighten erosion vulnerability. The study [34] about high-vulnerability areas in the catchment also validates this result.
The FFPI ranges from 1 to 10, with MKM’s FFPI index of 5.7, indicating a medium vulnerability for flash floods. Additional factors like SPI, TWI, TPI, and Soil Index may improve results [42]. Establishing a monitoring system is crucial due to changing natural and human factors [141]. Advanced technologies aid in data collection for predictive modeling through GIS, aiding hazard prevention [139]. Various approaches can enhance effectiveness in assessing flash flood vulnerability, and selecting the most suitable method for each situation to optimize solutions. Acquiring and analyzing databases is made simple by spatial analysis software, aiding professionals in spatial planning [142]. The model proposed in this study offers practical insights into land management and assists local authorities in mitigating flash flood vulnerability [39]. Its versatility allows implementation on a national scale, which is crucial given changing land use and increasing extreme weather events. Collaborative efforts between local, provincial, and national bodies are essential for deploying protective measures effectively. Integrating GIS and remote sensing provides a powerful tool for assessing flash flood potential, supported by the FFPI method’s accurate vulnerability reflection. Future research should explore machine learning methods to enhance susceptibility modeling, reducing subjectivity and increasing relevance for specific regional areas [44]. For more detailed validation, higher resolution historical data are needed to perform a ROC curve analysis to quantitatively assess the flash flood model’s predictive performance.
A GIS-based forest fire susceptibility map for MKM, located in a high-vulnerability forest fire area in the northeastern part of North Macedonia, evaluated factors like forest structure, topography, and proximity to vulnerable points. The map categorized forest areas into five vulnerability levels: very low, low, moderate, high, and very high. Nearly half of the forests in the study area were classified as very high or high vulnerability. This assessment helps decision-makers plan actions before, during, and after fires. Effective firefighting strategies can be implemented by considering vulnerability levels, including reviewing fire action plans, organizing fire response teams, and strategically positioning fire-watch towers [48]. The model also aids in fire prevention efforts by informing the deployment of firefighting teams, assessing road network efficiency, and identifying suitable locations for water resources. Additionally, they can guide the creation of buffer zones between forests and neighboring residential and agricultural areas in high-vulnerability zones. Satellite data and GIS integration offer effective tools for identifying and categorizing forest areas based on factors like topography, vegetation type, average temperature (ERA-5 data [143]), and proximity to roads and settlements. This integration facilitates the identification of high-vulnerability areas and aids in forestry management planning post-fire [47,49,50,51,144,145]. According to the study [146], vandalism generally exhibited high values of forest fires in North Macedonia, especially in the eastern part of the country (MKM). More attention is needed to understand the social causes of forest fires and effectively target prevention efforts. Prioritizing communication to specific population segments is crucial. Emphasis should shift from intervention to prevention, with a balanced budget allocation and community engagement to reduce negligent fires.
As is the case with flash floods, the situation concerning forest fires is similarly characterized by insufficient systematic historical observations. Authors recognize that factors beyond geophysical ones (e.g., anthropogenic factors) must be considered when analyzing forest fires. This complexity adds a layer of delicacy to estimating model accuracy associated with ROC/AUC values. Furthermore, the use of MODIS remote sensing products makes it difficult to attribute the given climatological hazard within the existing methodological framework due to the limitations of their high resolution.
While the study provides invaluable insights into the assessment of natural hazard vulnerability, it is imperative to acknowledge certain limitations in the methodologies employed. Firstly, despite exhaustive validation efforts encompassing various calculations, measurements, and field research, inherent uncertainties and limitations persist within the datasets utilized, potentially undermining the accuracy of the findings. Secondly, the absence of local pluviometry stations may have compromised the precision of the flash flood susceptibility assessment, given its heavy reliance on precipitation data. Thirdly, a notable constraint in assessing landslide susceptibility lies in the inadequacy of the inventory database, which may have impinged upon the completeness and accuracy of the model. Additionally, it is crucial to recognize the limitations in assessing forest fire susceptibility, including the reliance on remote sensing data and the inherent uncertainties in fire behavior modeling. Moreover, while the FFPI and RC method offers a comprehensive evaluation of flash flood and forest fire potential, its applicability is subject to variation based on regional characteristics, underscoring the necessity for further validation and refinement tailored to specific geographical areas.

5. Conclusions

The study provides comprehensive insights into sediment transport, erosion rates, landslide susceptibility, flash flood vulnerability, and forest fire susceptibility in the MKM region. The findings emphasize the critical need for proactive measures to mitigate the impact of natural hazards on the local environment and community. Collaborative efforts involving stakeholders at various levels are essential for implementing effective conservation and management strategies. Further research and field studies are recommended to enhance understanding and address the complex interplay of factors influencing vulnerability to natural hazards in the region.
Integrating the findings with field research is crucial for preserving and enhancing geodiversity in this municipality area. According to one study [46], these areas must be managed carefully, so it is necessary to conduct advanced research to identify the kind of measurement or management action to be implemented in the future. Collaborative efforts involving public nature protection agencies, provincial and national environmental institutions, and local environmental groups are essential for implementing conservation measures and monitoring environmental elements to reduce the vulnerability of natural hazards in MKM. Standardizing and implementing multi-hazard methodologies would improve monitoring and identifying natural disasters in North Macedonia on local and regional scales. This underscores the importance of developing southeastern Europe vulnerability assessments and management programs.

Author Contributions

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

Funding

This research received no external funding.

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. The data are not publicly available due to privacy.

Acknowledgments

Tin Lukić gratefully acknowledges the support of the Ministry of Science, Technological Development and Innovation of the Republic of Serbia (Grants No. 451-03-66/2024-03/200125 & 451-03-65/2024-03/200125) and the Provincial Secretariat for Higher Education and Scientific Research of Vojvodina (Serbia), No. 000871816 2024 09418 003 000 000 001 04 002 (GLOMERO), under Program 0201 and Program Activity 1012. The authors are grateful to the anonymous reviewers whose comments and suggestions greatly improved the manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Agrawal, N. Defining Natural Hazards–Large Scale Hazards. In Natural Disasters and Risk Management in Canada; Springer: Berlin/Heidelberg, Germany, 2018; Volume 49, pp. 1–40. [Google Scholar] [CrossRef]
  2. Salvati, P.; Petrucci, O.; Rossi, M.; Bianchi, C.; Pasqua, A.A.; Guzzetti, F. Gender, Age and Circumstances Analysis of Flood and Landslide Fatalities in Italy. Sci. Total Environ. 2018, 610–611, 867–879. [Google Scholar] [CrossRef] [PubMed]
  3. Alam, E.; Ray-Bennett, N.S. Disaster Risk Governance for District-Level Landslide Risk Management in Bangladesh. Int. J. Disaster Risk Reduct. 2021, 59, 102220. [Google Scholar] [CrossRef]
  4. Li, X.; Huang, X.; Zhang, Y. Spatio-Temporal Analysis of Groundwater Chemistry, Quality and Potential Human Health Risks in the Pinggu Basin of North China Plain: Evidence from High-Resolution Monitoring Dataset of 2015–2017. Sci. Total Environ. 2021, 800, 149568. [Google Scholar] [CrossRef] [PubMed]
  5. Shinohara, Y.; Kume, T. Changes in the Factors Contributing to the Reduction of Landslide Fatalities between 1945 and 2019 in Japan. Sci. Total Environ. 2022, 827, 154392. [Google Scholar] [CrossRef] [PubMed]
  6. Yang, H.-Q.; Zhang, L.; Gao, L.; Phoon, K.-K.; Wei, X. On the Importance of Landslide Management: Insights from a 32-year Database of Landslide Consequences and Rainfall in Hong Kong. Eng. Geol. 2022, 299, 106578. [Google Scholar] [CrossRef]
  7. Yao, D.; Xu, Y.; Zhang, P. How a Disaster Affects Household Saving: Evidence from China’s 2008 Wenchuan Earthquake. J. Asian Econ. 2019, 64, 101133. [Google Scholar] [CrossRef]
  8. Vereb, V. Geoheritage and Resilience. Selected Studies of Volcanic Geo-Heritage Areas from Different Geographical Environments and Different Levels of Protection. Ph.D. Thesis, Doctoral School of Earth Sciences, Eötvös Loránd University, Budapest, Hungary, 2020. [Google Scholar]
  9. Valjarević, A.; Filipović, D.; Valjarević, D.; Milanović, M.; Milošević, S.; Živić, N.; Lukić, T. GIS and remote sensing techniques for the estimation of dew volume in the Republic of Serbia. Meteorol. Appl. 2020, 27, 1930. [Google Scholar] [CrossRef]
  10. Blaikie, P.; Cannon, T.; Davis, I.; Wisner, B. At Risk: Natural Hazards, People’s Vulnerability and Disasters; Routledge: Abingdon, UK, 2014. [Google Scholar] [CrossRef]
  11. Lukić, T.; Leščešen, I.; Sakulski, D.; Basarin, B.; Jordaan, A. Rainfall Erosivity as an Indicator of Sliding Occurrence along the Southern Slopes of the Bačka Loess Plateau: A Case Study of the Kula Settlement, Vojvodina (North Serbia). Carpathian J. Earth Environ. Sci. 2016, 11, 303–318. [Google Scholar]
  12. Lukić, T.; Marić, P.; Hrnjak, I.; Gavrilov, M.B.; Mladjan, D.; Zorn, M.; Komac, B.; Milošević, Z.; Marković, S.B.; Sakulski, D.; et al. Forest Fire Analysis and Classification Based on a Serbian Case Study. Acta Geogr. Slov. 2017, 57, 51–63. [Google Scholar] [CrossRef]
  13. Lukić, T.; Bjelajac, D.; Fitzsimmons, K.E.; Marković, S.B.; Basarin, B.; Mlađan, D.; Micić, T.; Schaetzl, R.J.; Gavrilov, M.B.; Milanović, M.; et al. Factors Triggering Landslide Occurrence on the Zemun Loess Plateau, Belgrade Area, Serbia. Environ. Earth Sci. 2018, 77, 519. [Google Scholar] [CrossRef]
  14. Lukić, T.; Dunjić, J.; Đerčan, B.; Penjišević, I.; Milosavljević, S.; Bubalo-Živković, M.; Solarević, M. Local Resilience to Natural Hazards in Serbia. Case Study: The West Morava River Valley. Sustainability 2018, 10, 2866. [Google Scholar] [CrossRef]
  15. Lukić, T.; Lukić, A.; Basarin, B.; Ponjiger, T.M.; Blagojević, D.; Mesaroš, M.; Milanović, M.; Gavrilov, M.; Pavić, D.; Zorn, M.; et al. Rainfall erosivity and extreme precipitation in the Pannonian basin. Open Geosci. 2019, 11, 664–681. [Google Scholar] [CrossRef]
  16. Durlević, U.; Momčilović, A.; Čurić, V.; Dragojević, M. GIS Application in the Analysis of Erosion Intensity in the Vlasina River Basin. Bull. Serbian Geogr. Soc. 2019, 99, 17–36. [Google Scholar] [CrossRef]
  17. Sud, A.; Sajan, B.; Kanga, S.; Singh, S.K.; Singh, S.; Durin, B.; Kumar, P.; Meraj, G.; Sahariah, D.; Debnath, J.; et al. Integrating RUSLE Model with Cloud-Based Geospatial Analysis: A Google Earth Engine Approach for Soil Erosion Assessment in the Satluj Watershed. Water 2024, 16, 1073. [Google Scholar] [CrossRef]
  18. Hanane, M.; Naïma, E.A.; Charaf Eddine, B. Quantification of water erosion using the RUSLE method in the Mazer watershed (Berrechid region, Morocco). E3S Web Conf. 2024, 489, 04013. [Google Scholar] [CrossRef]
  19. Jodhani, K.H.; Patel, D.; Madhavan, N.; Singh, S.K. Soil Erosion Assessment by RUSLE, Google Earth Engine, and Geospatial Techniques over Rel River Watershed, Gujarat, India. Water Conserv. Sci. Eng. 2023, 8, 49. [Google Scholar] [CrossRef]
  20. Batjes, N.H. Global assessment of land vulnerability to water erosion on a 1/2° by 1/2° grid. Land Degrad. Dev. 1996, 7, 353–365. [Google Scholar] [CrossRef]
  21. Borrelli, P.; Robinson, D.A.; Fleischer, L.R.; Lugato, E.; Ballabio, C.; Alewell, C.; Meusburger, K.; Modugno, S.; Schütt, B.; Ferro, V.; et al. An assessment of the global impact of 21st century land use change on soil erosion. Nat. Commun. 2017, 8, 2013. [Google Scholar] [CrossRef] [PubMed]
  22. Varnes, D.J. Landslide Hazard Zonation: A Review of Principles and Practice; United Nations Educational, Scientific and Cultural Organization: Paris, France, 1984. [Google Scholar]
  23. Crozier, M. Landslides: Causes, Consequences and Environment; Croom Helm: London, UK, 1986. [Google Scholar]
  24. Soeters, R.; van Westen, C.J. Slope instability recognition, analysis, and zonation. In Landslides: Investigation and Mitigation; Turner, A.K., Schuster, R.L., Eds.; Transportation Research Board: Washington, DC, USA, 1996; pp. 129–177. [Google Scholar]
  25. Guzzetti, F.; Carrara, A.; Cardinali, M.; Reichenbach, P. Landslide Hazard Evaluation: A Review of Current Techniques and Their Application in a Multi-Scale Study, Central Italy. Geomorphology 1999, 31, 181–216. [Google Scholar] [CrossRef]
  26. Crozier, M.J.; Glade, T. Landslide Hazard and Risk: Issues, Concepts and Approach. In Landslide Hazard and Risk; Glade, T., Anderson, M., Crozier, M.J., Eds.; Wiley: Chichester, UK, 2012; pp. 1–40. [Google Scholar] [CrossRef]
  27. Guzzetti, F.; Reichenbach, P.; Ardizzone, F.; Cardinali, M.; Galli, M. Estimating the Quality of Landslide Susceptibility Models. Geomorphology 2006, 81, 166–184. [Google Scholar] [CrossRef]
  28. Milevski, I.; Dragićević, S.; Zorn, M. Statistical and Expert-Based Landslide Susceptibility Modeling on a National Scale Applied to North Macedonia. Open Geosci. 2019, 11, 750–764. [Google Scholar] [CrossRef]
  29. Aleksova, B.; Lukić, T.; Milevski, I.; Spalević, V.; Marković, S.B. Modelling Water Erosion and Mass Movements (Wet) by Using GIS-Based Multi-Hazard Susceptibility Assessment Approaches: A Case Study—Kratovska Reka Catchment (North Macedonia). Atmosphere 2023, 14, 1139. [Google Scholar] [CrossRef]
  30. Blöschl, G.; Hall, J.; Viglione, A.; Perdigão, R.A.P.; Parajka, J.; Merz, B.; Lun, D.; Arheimer, B.; Aronica, G.T.; Bilibashi, A.; et al. Changing Climate Both Increases and Decreases European River Floods. Nature 2019, 573, 108–111. [Google Scholar] [CrossRef]
  31. Turconi, L.; Faccini, F.; Marchese, A.; Paliaga, G.; Casazza, M.; Vojinovic, Z.; Luino, F. Implementation of Nature-Based Solutions for Hydro-Meteorological Risk Reduction in Small Mediterranean Catchments: The Case of Portofino Natural Regional Park, Italy. Sustainability 2020, 12, 1240. [Google Scholar] [CrossRef]
  32. Morar, C.; Lukić, T.; Basarin, B.; Valjarević, A.; Vujičić, M.; Niemets, L.; Telebienieva, I.; Boros, L.; Nagy, G. Shaping Sustainable Urban Environments by Addressing the Hydro-Meteorological Factors in Landslide Occurrence: Ciuperca Hill (Oradea, Romania). Int. J. Environ. Res. Public Health 2021, 18, 5022. [Google Scholar] [CrossRef] [PubMed]
  33. Milevski, I. Natural Hazards in the Republic of Macedonia with Special Emphasis on Flood and Earthquake in Skopje. Geogr. Rev. 2017, 50, 5–22. [Google Scholar]
  34. Aleksova, B.; Lukić, T.; Milevski, I.; Puhar, D.; Marković, S.B. Preliminary Assessment of Geohazards’ Impacts on Geodiversity in the Kratovska Reka Catchment (North Macedonia). Geosciences 2024, 14, 62. [Google Scholar] [CrossRef]
  35. Smith, G.E. Flash Flood Potential: Determining the Hydrologic Response of FFMP Basins to Heavy Rain by Analyzing Their Physiographic Characteristics. 2003. Available online: http://www.cbrfc.noaa.gov/papers/ffp_wpap.pdf (accessed on 5 April 2024).
  36. Sumi, T.; Saber, M.; Kantoush, S.A. Japan-Egypt hydro network: Science and technology collaborative research for flash flood management. J. Disaster Res. 2013, 8, 28–36. [Google Scholar] [CrossRef]
  37. Dottori, F.; Martina, M.L.V.; Figueiredo, R. A methodology for flood susceptibility and vulnerability analysis in complex flood scenarios. J. Flood Risk Manag. 2016, 11, S632–S645. [Google Scholar] [CrossRef]
  38. Cao, C.; Xu, P.; Wang, Y.; Chen, J.; Zheng, L.; Niu, C. Flash Flood Hazard Susceptibility Mapping Using Frequency Ratio and Statistical Index Methods in Coalmine Subsidence Areas. Sustainability 2016, 8, 948. [Google Scholar] [CrossRef]
  39. Popa, M.C.; Simion, A.; Peptenatu, D.; Dima, C.; Draghici, C.; Florescu, M.S.; Dobrea, R.C.; Diaconu, D.C. Spatial Assessment of Flash-Flood Vulnerability in the Moldova River Catchment (N Romania) Using the FFPI. J. Flood Risk Manag. 2020, 13, e12624. [Google Scholar] [CrossRef]
  40. Ceru, J. Flash Flood Potential Index for Pennsylvania. In Proceedings of the 2012 ESRI Federal GIS Conference, Washington, DC, USA, 22–24 February 2012; Available online: https://proceedings.esri.com/library/userconf/fed12/index.html (accessed on 5 April 2024).
  41. Karagiorgos, K.; Thaler, T.; Hübl, J.; Maris, F.; Fuchs, S. Multi-Vulnerability Analysis for Flash Flood Risk Management. Nat. Hazards 2016, 82, 63–87. [Google Scholar] [CrossRef]
  42. Yassin, N.A.; Nor Adnan, N.A.; Sadek, E.M. Analysis of Flash Flood Potential Index (FFPI) and scenarios assessment in Shah Alam using GIS approach. Plan. Malays. 2023, 21. [Google Scholar] [CrossRef]
  43. Petrović, A.; Kostadinov, S.; Dragićević, S. The Inventory and Characterization of Torrential Flood Phenomenon in Serbia. Pol. J. Environ. Stud. 2014, 23, 823–830. [Google Scholar]
  44. Šiljeg, S.; Milošević, R.; Mamut, M. Pluvial Flood Susceptibility in the Local Community of the City of Gospić (Croatia). Sustainability 2024, 16, 1701. [Google Scholar] [CrossRef]
  45. Durlević, U.; Novković, I.; Lukić, T.; Valjarević, A.; Samardžić, I.; Krstić, F.; Batocanin, 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]
  46. Muwahhid, A.; Alwi, M.; Sari, S.P.; Tilova, U.D.N.; Pratama, I.D. Mapping of Multiple Hazards in the Cilongkrang Sub-Watershed, Majalengka, Indonesia. IOP Conf. Ser. Earth Environ. Sci. 2024, 1313, 012028. [Google Scholar] [CrossRef]
  47. Erten, E.; Kurgun, V.; Musaoglu, N. Forest Fire Risk Zone Mapping from Satellite Imagery and GIS: A Case Study. In Proceedings of the 20th Congress of the International Society for Photogrammetry and Remote Sensing, Istanbul, Turkey, 12–23 July 2004. [Google Scholar]
  48. Demir, A.; Akay, A.E. Forest Fire Risk Mapping Using GIS-Based Analytical Hierarchy Process Approach. Eur. J. For. Eng. 2024, 10, 15–28. [Google Scholar] [CrossRef]
  49. Mendhe, S.S. Forest Fire Risk Zone Mapping from Satellite Imagery and GIS: A Case Study of Madhya Pradesh. For. Fire Risk Zone Mapp. 2023, preprint. [Google Scholar] [CrossRef]
  50. Das, R.R.; Nandi, D. Forest Cover Loss and Forest Fire Monitoring Using GIS And GEE over Similipal Tiger Reserve, Odisha, India. Open Geosci. 2024, preprint. [Google Scholar] [CrossRef]
  51. Baykal, T.M. GIS-Based Spatiotemporal Analysis of Forest Fires in Turkey from 2010 to 2020. Trans. GIS 2023, 27, 1289–1317. [Google Scholar] [CrossRef]
  52. Rüttinger, L.; van Ackern, P.; Gordon, N.; Adrian, F. Regional Assessment for South-Eastern Europe: Security Implications of Climate Change; OSCE: Adelphi, MD, USA; Berlin, German, 2021; pp. 1–73. [Google Scholar]
  53. Kurnik, B.; van der Linden, P.; Mysiak, J.; Swart, R.J.; Füssel, H.M.; Christiansen, T.; Cavicchia, L.; Gualdi, S.; Mercogliano, P.; Rianna, G.; et al. Climate Change Adaptation and Disaster Risk Reduction in Europe: Enhancing Coherence of the Knowledge Base, Policies and Practices; EEA-European Environment Agency: Copenhagen, Denmark, 2017; pp. 1–176. [Google Scholar]
  54. Spinoni, J.; Vogt, J.V.; Naumann, G.; Barbosa, P.; Dosio, A. Will drought events become more frequent and severe in Europe? Int. J. Climatol. 2018, 38, 1718–1736. [Google Scholar] [CrossRef]
  55. Insana, A.; Beroya-Eitner, M.A.; Barla, M.; Zachert, H.; Žlender, B.; van Marle, M.; Kalsnes, B.; Bračko, T.; Pereira, C.; Prodan, I.; et al. Climate Change Adaptation of Geo-Structures in Europe: Emerging Issues and Future Steps. Geosciences 2021, 11, 488. [Google Scholar] [CrossRef]
  56. Cvetković, V.M.; Grbić, L. Public Perception of Climate Change and Its Impact on Natural Disasters. J. Geogr. Inst. Jovan Cvijic SASA 2021, 71, 43–58. [Google Scholar] [CrossRef]
  57. Güneralp, B.; Güneralp, I.; Liu, Y. Changing Global Patterns of Urban Exposure to Flood and Drought Hazards. Glob. Environ. Chang. 2015, 31, 217–225. [Google Scholar] [CrossRef]
  58. Pörtner, H.-O.; Roberts, D.C.; Tignor, M.; Poloczanska, E.S.; Mintenbeck, K.; Alegría, A.; Craig, M.; Langsdorf, S.; Löschke, S.; Möller, V.; et al. (Eds.) Climate Change 2022: Impacts, Adaptation and Vulnerability. In Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2022; p. 3056. [Google Scholar]
  59. Dilley, M.; Chen, R.S.; Deichmann, M.; Lerner-Lam, A.; Arnold, M.; Agwe, J.; Buys, P.; Kjekstad, O.; Lyon, B.; Yetman, G. Natural Disaster Hotspots: A Global Risk Analysis; World Bank Publications: Washington, DC, USA, 2005. [Google Scholar]
  60. Gaume, E.; Borga, M. Post-Flood Field Investigations in Upland Catchments after Major Flash Floods: Proposal of a Methodology and Illustrations. J. Flood Risk Manag. 2008, 1, 175–189. [Google Scholar] [CrossRef]
  61. Smith, K.; Petley, D.N. Environmental Hazards: Assessing Risk and Reducing Disaster, 5th ed.; Routledge: London, UK; New York, NY, USA, 2009; p. 383. [Google Scholar]
  62. Marchi, L.; Borga, M.; Preciso, E.; Gaume, E. Characterization of Selected Extreme Flash Floods in Europe and Implications for Flood Risk Management. J. Hydrol. 2010, 394, 118–133. [Google Scholar] [CrossRef]
  63. Gencer, E.A. Natural Disasters, Urban Vulnerability, and Risk Management: A Theoretical Overview. In The Interplay between Urban Development, Vulnerability, and Risk Management, Mediterranean Studies; Gencer, E.A., Ed.; Springer: Berlin/Heidelberg, Germany, 2013; Volume 7, pp. 7–43. [Google Scholar]
  64. Arnell, N.; Gosling, S. The Impacts of Climate Change on River Flow Regimes at the Global Scale. J. Hydrol. 2013, 486, 351–364. [Google Scholar] [CrossRef]
  65. Hirabayashi, Y.; Mahendran, R.; Koirala, S.; Konoshima, L.; Yamazaki, D.; Watanabe, S.; Kim, H.; Kanae, S. Global Flood Risk under Climate Change. Nat. Clim. Chang. 2013, 3, 816–821. [Google Scholar] [CrossRef]
  66. Milevski, I.; Dragićević, S.; Radevski, I. GIS and Remote Sensing Based Natural Hazard Modeling of Kriva Reka Catchment, Republic of Macedonia. Z. Geomorphol. 2017, 58 (Suppl. S3), 213–228. [Google Scholar] [CrossRef]
  67. Milevski, I.; Dragićević, S.; Tošić, R. GIS and Remote Sensing Assessment of Erosion Risk Areas in Pehchevo Municipality. In Proceedings of the Hilly-Mountain Areas–Problems and Perspectives Conference, Ohrid, Republic of Macedonia, 12–15 September 2013. [Google Scholar]
  68. Bathrellos, G.D.; Skilodimou, H.D.; Chousianitis, K.; Youssef, A.M.; Pradhan, B. Suitability Estimation for Urban Development Using Multi-Hazard Assessment Map. Sci. Total Environ. 2017, 575, 119–134. [Google Scholar] [CrossRef] [PubMed]
  69. Skilodimou, H.D.; Bathrellos, G.D.; Chousianitis, K.; Youssef, A.M.; Pradhan, B. Multi-hazard assessment modeling via multicriteria analysis and GIS: A case study. Environ. Earth Sci. 2019, 78, 47. [Google Scholar] [CrossRef]
  70. Sevieri, G.; Galasso, C.; D’Ayala, D.; De Jesus, R.; Oreta, A.; Grio, M.E.D.A.; Ibabao, R. A multi-hazard risk prioritization framework for cultural heritage assets. Nat. Hazards Earth Syst. Sci. 2020, 20, 1391–1414. [Google Scholar] [CrossRef]
  71. Lombardo, L.; Tanyas, H.; Nicu, C.I. Spatial Modeling of Multi-Hazard Threat to Cultural Heritage Sites. Eng. Geol. 2020, 277, 105776. [Google Scholar] [CrossRef]
  72. Geological Survey of SR Macedonia. Interpreters of the Basic Geological Maps 100k for SR Macedonia; Professional Fund of the Federal Geological Survey of SFRY: Belgrade, Yugoslavia, 1963–1985. (In Macedonian) [Google Scholar]
  73. Różycka, M.; Migoń, P.; Michniewicz, A. Topographic Wetness Index and Terrain Ruggedness Index in geomorphic characterisation of landslide terrains, on examples from the Sudetes, SW Poland. Z. Geomorphol. 2017, 61, 61–80. [Google Scholar] [CrossRef]
  74. Fick, S.E.; Hijmans, R.J. Worldclim 2: New 1-km Spatial Resolution Climate Surfaces for Global Land Areas. Int. J. Climatol. 2017, 37, 4302–4315. [Google Scholar] [CrossRef]
  75. Corine Land Cover 2018. Copernicus Land Monitoring Service. In Technical Report; Office for Official Publications of the European Communities: Luxembourg, 2016. [Google Scholar]
  76. FAO. Soil Map of the Republic of Macedonia, 1:50,000; FAO: Rome, Italy; Agricultural Institute-UKIM: Skopje, North Macedonia, 2015. (In Macedonian) [Google Scholar]
  77. State Statistics Office of the Republic of North Macedonia (Census 2002; 2021). Available online: www.stat.gov.mk/ (accessed on 2 May 2024).
  78. Gavrilović, S. Engineering of Torrents and Erosion. J. Constr. 1972, 292. (In Serbian) [Google Scholar]
  79. 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]
  80. De Vente, J.; Poesen, J. Predicting Soil Erosion and Sediment Yield at the Basin Scale: Scale Issue and Semi-Quantitative Models. Earth-Sci. Rev. 2005, 71, 95–125. [Google Scholar] [CrossRef]
  81. 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]
  82. Milevski, I. An Approach of GIS Based Assessment of Soil Erosion Rate on Country Level in the Case of Macedonia. Geobalcanica 2015, 1, 97–104. [Google Scholar] [CrossRef]
  83. Dragićević, N.; Karleuša, B.; Ožanić, N. A Review of the Gavrilović Method (Erosion Potential Method) Application. Građevinar J. Croat. Assoc. Civ. Eng. 2016, 68, 715–725. [Google Scholar]
  84. Bezak, N.; Borrelli, P.; Mikoš, M.; Auflič, M.J.; Panagos, P. Towards Multi-Model Soil Erosion Modelling: An Evaluation of the Erosion Potential Method (EPM) for Global Soil Erosion Assessments. Catena 2024, 234, 107596. [Google Scholar] [CrossRef]
  85. Tosić, 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]
  86. Globevnić, L.; Holjević, D.; Petkovšek, G.; Rubinić, J. Applicability of the Gavrilović Method in Erosion Calculation Using Spatial Data Manipulation Techniques. In Erosion Prediction in Ungauged Basins: Integrating Methods and Techniques; IAHS Publication: Wallingford, UK, 2003; p. 279. [Google Scholar]
  87. Milevski, I. Factors, Forms, Assessment and Human Impact on Excess Erosion and Deposition in Upper Bregalnica Watershed (Republic of Macedonia). Z. Geomorphol. Suppl. Issues 2011, 55, 77–94. [Google Scholar] [CrossRef]
  88. Petras, J.; Holjević, D.; Kunstek, D. Implementation of GIS-Technology in Gavrilovič’s Method for Estimation Soil Erosion Production and Sediment Transport. In Proceedings of the International Conference Erosion and Torrent Control as a Factor in Sustainable River Basin Management, Belgrade, Serbia, 25–28 September 2007. [Google Scholar]
  89. Novković, D.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. (In Serbian with English Summary). [Google Scholar]
  90. Blinkov, I.; Kostadinov, S. Applicability of Various Erosion Risk Assessment Methods for Engineering Purposes. In Proceedings of the BALWOIS Conference, Ohrid, Macedonia, 25–29 May 2010; Available online: https://dcs.ufla.br/images/imagens_dcs/pdf/Prof%20Marx/Aula%205/Outro%20art%20interesse/Blinkov%20&%20Kostadinov%202010_USLE_PESERA_KINEROS_WEP_WEPP_EPM.pdf (accessed on 5 April 2024).
  91. Lazarević, R. The New Method for Erosion Coefficient Determination-Z. Eros.-Prof. Inf. Bull. 1985, 13, 54–61. [Google Scholar]
  92. Spalević, V.; Barović, G.; Vujacić, D.; Curović, M.; Behzadfar, M.; Djurović, N.; Dudić, B.; Billi, P. The Impact of Land Use Changes on Soil Erosion in the River Basin of MiockiPotok, Montenegro. Water 2020, 12, 2973. [Google Scholar] [CrossRef]
  93. 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]
  94. Zorn, M.; Komac, B. The importance of measuring erosion processes on the example of Slovenia. Hrvat. Geogr. Glas. 2011, 73, 19–34. [Google Scholar] [CrossRef]
  95. Tavares, A.S.; Spalević, V.; Avanzi, J.C.; Alves, D. Modeling of water erosion by the erosion potential method in a pilot subbasin in southern Minas Gerais. Semin. Ciências Agrárias Londrina 2019, 40, 555–572. [Google Scholar] [CrossRef]
  96. Ali, S.; Al-Umary, F.A.; Salar, S.G.; Al-Ansari, N.; Knutsson, S. GIS-Based Soil Erosion Estimation Using EPM Method, Garmiyan Area, Kurdistan Region, Iraq. J. Civ. Eng. Archit. 2016, 10, 291–308. [Google Scholar] [CrossRef]
  97. Tošić, R.; Dragićević, S.; Belanović, S.; Brčeski, I.; Lovrić, N. Considerations on reservoir sedimentation and heavy metals content within the Drenova Reservoir (B&H). Carpathian J. Earth Environ. Sci. 2013, 8, 175–184. [Google Scholar]
  98. Kazimierski, L.D.; Irigoyen, M.; Re, M.; Menendez, A.N.; Spalletti, P.; Brea, J.D. Impact of Climate Change on Sediment Yield from the Upper Plata Basin. Int. J. River Basin Manag. 2013, 11, 411–421. [Google Scholar] [CrossRef]
  99. Behzadfar, M.; Tazioli, A.; Vukleic-Shutoska, M.; Simunic, I.; Spalevic, V. Calculation of Sediment Yield in the S1-1 Watershed, Shirindareh Watershed, Iran. Agric. For. 2014, 60, 207–216. [Google Scholar]
  100. Ivanova, E.; Milevski, I. Landslide Susceptibility Mapping of the Territory of Municipalities Pehčevo and Simitli by Means of GIS Modeling. In Proceedings of the Conference Space, Ecology, Safety SES, Sofia, Bulgaria, 24–26 October 2013. [Google Scholar]
  101. Milevski, I.; Dragićević, S. Landslides Susceptibility Zonation of the Territory of North Macedonia Using the Analytical Hierarchy Process Approach. Sect. Nat. Math. Biotech. Sci. MASA 2019, 40, 115–126. [Google Scholar] [CrossRef]
  102. Dragićević, S.; Carević, I.; Kostadinov, S.; Novković, I.; Abolmasov, B.; Milojković, B.; Simić, D. Landslide Susceptibility Zonation in the Kolubara River Basin (Western Serbia)—Analysis of Input Data. Carpathian J. Earth Environ. Sci. 2012, 7, 37–47. [Google Scholar]
  103. Donati, L.; Turrini, M.C. An Objective Method to Rank the Importance of the Factors Predisposing to Landslides with the GIS Methodology: Application to an Area of the Apennines (Valnerina; Perugia, Italy). Eng. Geol. 2002, 63, 277–289. [Google Scholar] [CrossRef]
  104. Magliulo, P.; Di Lisio, A.; Russo, F.; Zelando, A. Geomorphology and Landslide Susceptibility Assessment Using GIS and Bivariate Statistics: A Case Study in Southern Italy. Nat. Hazards 2008, 47, 411–435. [Google Scholar]
  105. Süzen, M.; Doyuran, V. Data-driven bivariate landslide susceptibility assessment using geographical information systems: A method and application to Asarsuyu catchment, Turkey. Eng. Geol. 2004, 71, 303–321. [Google Scholar] [CrossRef]
  106. Minea, G. Assessment of the Flash Flood Potential of Bâsca River Catchment (Romania) Based on Physiographic Factors. Cent. Eur. J. Geosci. 2013, 5, 344–353. [Google Scholar] [CrossRef]
  107. Costache, R.; Pravalie, R.; Mitof, I.; Popescu, C. Flood Vulnerability Assessment in the Low Sector of Saratel Catchment. Case Study: Joseni Village. Carpathian J. Earth Environ. Sci. 2015, 10, 161–169. [Google Scholar]
  108. Tincu, R.; Lazar, G.; Lazar, I. Modified flash flood potential index in order to estimate areas with predisposition to water accumulation. Open Geosci. 2018, 10, 593–606. [Google Scholar] [CrossRef]
  109. Petrović, A.; Dragićević, S.; Radić, B.; Milanović Pešić, A. Historical Torrential Flood Events in the Kolubara River Basin. Nat. Hazards 2015, 79, 537–547. [Google Scholar] [CrossRef]
  110. Ristić, R.; Radić, B.; Vasiljević, N. Characteristics of maximal discharges on torrential watersheds in Serbia. Bull. Serbian Geogr. Soc. 2009, 89, 161–189. [Google Scholar]
  111. Pravalie, R.; Costache, R. The analysis of the susceptibility of the flash-floods’ genesis in the area of the hydrographical basin of Basca Chiojdului river. Forum Geogr. 2014, 13, 39–49. [Google Scholar] [CrossRef]
  112. Minea, G.; Iliescu, M.; Dedu, F. Temporal Rainfall Properties at Events Scale in the Curvature Subcarpathians (Romania). Forum Geogr. 2016, 15, 115–123. [Google Scholar] [CrossRef]
  113. 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]
  114. Saleh, A.; Yuzir, A.; Abustan, I. Flash Flood Susceptibility Modelling: A Review. IOP Conf. Ser. Mater. Sci. Eng. 2020, 712, 012005. [Google Scholar] [CrossRef]
  115. Band, S.S.; Janizadeh, S.; Chandra Pal, S.; Saha, A.; Chakrabortty, R.; Melesse, A.M.; Mosavi, A. Flash Flood Susceptibility Modeling Using New Approaches of Hybrid and Ensemble Tree-Based Machine Learning Algorithms. Remote Sens. 2020, 12, 3568. [Google Scholar] [CrossRef]
  116. Tošić, I.; Mladjan, D.; Gavrilov, M.B.; Živanović, S.; Radaković, M.G.; Putniković, S.; Petrović, P.; Krstić Mistridželović, I.; Marković, S.B. Potential influence of meteorological variables on forest fire risk in Serbia during the period 2000–2017. De Gruyter 2019, 11, 414–425. [Google Scholar] [CrossRef]
  117. Saaty, T.L. Analytic Hierarchy Process; McGrawHill: New York, NY, USA, 1980. [Google Scholar]
  118. Saaty, T.L. How to make a decision: The analytic hierarchy process. Eur. J. Oper. Res. 1990, 48, 9–26. [Google Scholar] [CrossRef]
  119. Das, S. Comparison among influencing factor, frequency ratio, and analytical hierarchy process techniques for groundwater potential zonation in Vaitarna basin, Maharashtra, India. Groundw. Sustain. Dev. 2019, 8, 617–629. [Google Scholar] [CrossRef]
  120. Lentswe, G.B.; Molwalefhe, L. Delineation of potential groundwater recharge zones using analytic hierarchy process-guided GIS in the semi-arid Motloutse watershed, eastern Botswana. J. Hydrol. Reg. Stud. 2020, 28, 100674. [Google Scholar] [CrossRef]
  121. Ahmed, A.; Ranasinghe-Arachchilage, C.; Alrajhi, A.; Hewa, G. Comparison of Multicriteria Decision-Making Techniques for Groundwater Recharge Potential Zonation: Case Study of the Willochra Basin, South Australia. Water 2021, 13, 525. [Google Scholar] [CrossRef]
  122. Ghosh, B. Spatial mapping of groundwater potential using data-driven evidential belief function, knowledge-based analytic hierarchy process and an ensemble approach. Environ. Earth Sci. 2021, 80, 625. [Google Scholar] [CrossRef]
  123. Tabarestani, E.S.; Afzalimehr, H. Artificial neural network and multi-criteria decision-making models for flood simulation in GIS: Mazandaran Province, Iran. Stoch. Environ. Res. Risk Assess. 2021, 35, 2439–2457. [Google Scholar] [CrossRef]
  124. Liang, J.; Yang, J. Application of the AHP method on the optimization with undesirable priorities. Eng. Comput. 2021, 38, 2137–2153. [Google Scholar] [CrossRef]
  125. Bentekhici, N.; Bellal, S.A.; Zegrar, A. Contribution of remote sensing and GIS to mapping the fire risk of Mediterranean forest case of the forest massif of Tlemcen (North-West Algeria). Nat. Hazards 2020, 104, 811–831. [Google Scholar] [CrossRef]
  126. United Nations Office for Disaster Risk Reduction. Available online: https://www.undrr.org/ (accessed on 19 April 2024).
  127. Dragićević, S.; Milevski, I.; Novković, I.; Milojković, B. Natural Conditions as a Limiting Factor in the Development of the Border Area of Serbia and Macedonia. SGD Glas. 2010, 90, 29–38. (In Serbian) [Google Scholar]
  128. Tošić, R.; Dragićević, S.; Lovrić, N.; Milevski, I. Multi-hazard assessment using GIS in urban areas: Case study–Banja Luka municipality, Bosnia and Herzegovina. Bull. Serbian Geogr. Soc. 2013, 93, 41–50. [Google Scholar] [CrossRef]
  129. Minčev, I. Development of a Methodology for Determining Protective Zones Around Water Accumulation from the Aspect of Erosion and Sediment Transport. Ph.D. Thesis, Department of Land and Water, Faculty of Forestry, University “St. Cyril and Methodius”, Skopje, North Macedonia, 2014. [Google Scholar]
  130. Fox, D.M.; Bryan, R.B. The Relationship of Soil Loss by Interrill Erosion to Slope Gradient. Catena 1999, 38, 211–222. [Google Scholar] [CrossRef]
  131. Ristanović, B.; Cimbaljević, M.; Miljković, Đ.; Ostojić, M.; Fekete, R. GIS Application for Determining Geographical Factors on Intensity of Erosion in Serbian River Basins. Case Study: River Basin Likodra. Atmosphere 2019, 10, 526. [Google Scholar] [CrossRef]
  132. Gocić, M.; Dragićević, S.; Živanović, S.; Ivanović, R.; Martić Bursać, N.; Stričević, L.; Radivojević, A.; Živković, J. Assessment of soil erosion intensity in the Kutinska river basin in the period 1971–2016. Fresenius Environ. Bull. 2021, 30, 10890–10898. [Google Scholar]
  133. 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. [Google Scholar] [CrossRef]
  134. De Graff, J.V.; Romesburg, H.C.; Ahmad, R.; McCalpin, J.P. Producing Landslide-Susceptibility Maps for Regional Planning in Data-Scarce Regions. Nat. Hazards 2012, 64, 729–749. [Google Scholar] [CrossRef]
  135. Wang, L.J.; Guo, M.; Sawada, K.; Lin, J.; Zhang, J. A comparative study of landslide susceptibility maps using logistic regression, frequency ratio, decision tree, weights of evidence and artificial neural network. Geosci. J. 2016, 20, 117–136. [Google Scholar] [CrossRef]
  136. Saro, L.; Woo, J.; Kwan-Young, O.; Moung-Jin, L. The spatial prediction of landslide susceptibility applying artificial neural network and logistic regression models: A case study of Inje, Korea. Open Geosci. 2016, 8, 117–132. [Google Scholar] [CrossRef]
  137. Liu, Q.; Yang, Z.; Shi, H.; Wang, Z. Ecological Risk Assessment of Geo-Hazards in Natural World Heritage Sites: An Empirical Analysis of Bogda, Tianshan. Open Geosci. 2019, 11, 327–340. [Google Scholar] [CrossRef]
  138. Petrović, A. Flash Floods in Serbia; Serbian Geographical Society: Belgrade, Serbia, 2021; ISBN 978-86-82751-62-5. (In Serbian) [Google Scholar]
  139. Jovanović, D. Application of the Flash Flood Potential Index in Torrential Floods Risk Assessment (FFPI): A Case Study of Svilajnac Municipality. Proc.–Fac. Geogr. Univ. Belgrade 2023, 71, 65–78. [Google Scholar] [CrossRef]
  140. Kourgialas, N.N.; Karatzas, G.P. Flood Management and a GIS Modelling Method to Assess Flood-Hazard Areas—A Case Study. Hydrol. Sci. J. 2011, 56, 212–225. [Google Scholar] [CrossRef]
  141. Vujačić, D.; Milevski, I.; Mijanović, D.; Vujović, F.; Lukić, T. Initial results of comparative assessment of soil erosion intensity using the WIntErO model: A case study of Polimlje and Shirindareh drainage basins. Carpath. J. Earth Environ. Sci. 2023, 18, 385–404. [Google Scholar] [CrossRef]
  142. Kocsis, I.; Stefan, B.; Ioan-Aurel, I.; Dohotar, V.; Rusu, R.; Rosca, S. Flash Flood Vulnerability Mapping Based on FFPI Using GIS Spatial Analysis Case Study: Valea Rea Catchment Area, Romania. Sensors 2022, 22, 3573. [Google Scholar] [CrossRef] [PubMed]
  143. European Centre for Medium-Range Weather Forecasts (ECMWF). ERA5-Land Dataset (2013–2023); European Centre for Medium-Range Weather Forecasts (ECMWF): Reading, UK, 2023. [Google Scholar]
  144. Novković, I.; Marković, G.B.; Lukić, D.; Dragićević, S.; Milošević, M.; Djurdjić, S.; Samardzić, I.; Ležaić, T.; Tadić, M. GIS-Based Forest Fire Susceptibility Zonation with IoT Sensor Network Support, Case Study–Nature Park Golija, Serbia. Sensors 2021, 21, 6520. [Google Scholar] [CrossRef] [PubMed]
  145. Jaiswal, R.K.; Mukherjee, S.; Raju, D.K.; Saxena, R. Forest Fire Risk Zone Mapping from Satellite Imagery and GIS. Int. J. Appl. Earth Obs. Geoinf. 2002, 4, 1–10. [Google Scholar] [CrossRef]
  146. Tedim, F.; Leone, V.; Lovreglio, R.; Xanthopoulos, G.; Chas-Amil, M.-L.; Ganteaume, A.; Efe, R.; Royé, D.; Fuerst-Bjeliš, B.; Nikolov, N.; et al. Forest Fire Causes and Motivations in the Southern and South-Eastern Europe through Experts’ Perception and Applications to Current Policies. Forests 2022, 13, 562. [Google Scholar] [CrossRef]
Figure 1. Overview of the study area: MKM (left); the case study location in Europe (upper-right); and in North Macedonia (bottom-right).
Figure 1. Overview of the study area: MKM (left); the case study location in Europe (upper-right); and in North Macedonia (bottom-right).
Atmosphere 15 00774 g001
Figure 2. Surface air temperature map (A), and precipitation map (B) of the study area.
Figure 2. Surface air temperature map (A), and precipitation map (B) of the study area.
Atmosphere 15 00774 g002
Figure 3. The Kamenica River valley in the village of Sasa (left) and the Kamenica River (right) located within the MKM.
Figure 3. The Kamenica River valley in the village of Sasa (left) and the Kamenica River (right) located within the MKM.
Atmosphere 15 00774 g003
Figure 4. Flow chart with all the procedures and methods used in this research.
Figure 4. Flow chart with all the procedures and methods used in this research.
Atmosphere 15 00774 g004
Figure 5. Erosion vulnerability map of MKM based on Z coefficient values.
Figure 5. Erosion vulnerability map of MKM based on Z coefficient values.
Atmosphere 15 00774 g005
Figure 6. Soil erosion map of the MKM.
Figure 6. Soil erosion map of the MKM.
Atmosphere 15 00774 g006
Figure 7. (A) Erosion landforms in the Lukovica River valley; (B) excessive erosion on the steep slopes of the Lukovica River valley; (C) deposition in the Kamenica River valley; (D) excessive deposition in the valley bottom of the Kamenica River.
Figure 7. (A) Erosion landforms in the Lukovica River valley; (B) excessive erosion on the steep slopes of the Lukovica River valley; (C) deposition in the Kamenica River valley; (D) excessive deposition in the valley bottom of the Kamenica River.
Atmosphere 15 00774 g007
Figure 8. A landslide susceptibility map (LSI) for the MKM, categorized using the natural breaks classification (m3/km2 year).
Figure 8. A landslide susceptibility map (LSI) for the MKM, categorized using the natural breaks classification (m3/km2 year).
Atmosphere 15 00774 g008
Figure 9. ROC curve and AUC graph of the used LSA model.
Figure 9. ROC curve and AUC graph of the used LSA model.
Atmosphere 15 00774 g009
Figure 10. Susceptibility map to flash floods (FFPI) in the MKM.
Figure 10. Susceptibility map to flash floods (FFPI) in the MKM.
Atmosphere 15 00774 g010
Figure 11. Terrain susceptibility map for forest fires.
Figure 11. Terrain susceptibility map for forest fires.
Atmosphere 15 00774 g011
Figure 12. Multi-hazard map for the MKM.
Figure 12. Multi-hazard map for the MKM.
Atmosphere 15 00774 g012
Figure 13. Flood plains of excessive sediment deposition in the Kamenica River valley.
Figure 13. Flood plains of excessive sediment deposition in the Kamenica River valley.
Atmosphere 15 00774 g013
Figure 14. (A) Huge colluvial fan as a result of very high erosion and deposition; (B) active landslide in Sasa village.
Figure 14. (A) Huge colluvial fan as a result of very high erosion and deposition; (B) active landslide in Sasa village.
Atmosphere 15 00774 g014
Table 1. Data sources for landslide susceptibility analysis (LSA).
Table 1. Data sources for landslide susceptibility analysis (LSA).
Data TypeSourceDetailsRelevance
LithologyDigitalized Geological Map (100 k)Focused on seven distinct lithological units within the municipality area.Provides lithological composition data for landslide analysis.
Slope and Plan Curvature15 m Digital Elevation Model (DEM)Derived from a 15 m DEM.Key for understanding terrain morphology and its influence on landslides.
Land UseCORINE Land Cover 2018 Classification HierarchyExtracted land cover types within the study area.Aids in assessing landslide susceptibility based on land cover types.
Distance from StreamsTopographic River NetworksDistances were calculated with buffer zones (20 m increments) and converted to raster format.Proximity to streams influences landslide susceptibility.
Distance from RoadsOpen Street Map Road NetworksDistances were calculated with buffer zones (20 m increments) and converted to raster format.Proximity to roads influences landslide susceptibility.
Table 2. Erosion assessment modeling in classes.
Table 2. Erosion assessment modeling in classes.
Class NameClassArea in km2Area in %
Very low0–0.428.415.0
Low0.4–0.878.441.4
Moderate0.8–1.251.126.9
High1.2–1.622.411.8
Very high>1.69.34.9
Total189.6100.0
Table 3. Areas within the MKM are categorized by the extent of erosion.
Table 3. Areas within the MKM are categorized by the extent of erosion.
Class in m3/km2/yArea in km2Area in %
0–50060.732.0
500–100059.031.1
1000–150033.817.8
1500–200018.49.7
>200017.89.4
Total189.6100.0
Table 4. Average annual deposition of eroded sediments from the area of MKM to Kalimanci Lake.
Table 4. Average annual deposition of eroded sediments from the area of MKM to Kalimanci Lake.
W = Annual Sediment
Production from EPM
Ru = Retention CoefficientG = Annual Sediment Yield from EPMMeasured Average Annual Sediment Yield
(Kalimanci Lake)
182,713 m3/year0.7127,970 m3/year214,325 m3/year
Table 5. Landslide susceptibility areas in the MKM, categorized through the frequency ratio (FR) method.
Table 5. Landslide susceptibility areas in the MKM, categorized through the frequency ratio (FR) method.
ClassClass NameArea in km2Area in %
1Very low12.76.7
2Low21.911.5
3Moderate59.331.2
4High35.618.7
5Very high60.531.8
Total190.0100.0
Table 6. Landslide distribution relative to LSI classes.
Table 6. Landslide distribution relative to LSI classes.
ClassClass NameNu%
1Very low/0.0
2Low15.0
3Moderate210.0
4High735.0
5Very high1050.0
Total20100.0
Table 7. Coefficient of lithological units (S).
Table 7. Coefficient of lithological units (S).
RocksCoefficient
Amphibolite1
Granitoids2
Gneiss6
Compact volcanic rocks1
Mica shists7
Clastic sediments9
Clastic sediments9
Table 8. Types of land use (L).
Table 8. Types of land use (L).
Classes of LandArea (%)Value
Complex cultivation patterns: Land principally occupied by agriculture, with significant areas of natural vegetation29.47
Broad-leaved forest22.43
Transitional woodland-shrub19.95
Mixed forest; natural grasslands9.53
Pastures8.66
Coniferous forest8.02
Water bodies1.41
Discontinuous urban fabric0.64
Dump sites0.25
Table 9. Terrains susceptible to flash floods.
Table 9. Terrains susceptible to flash floods.
Class NameIn km2In %
Very low16.98.9
Low17.29.1
Moderate70.337.0
High79.341.7
Very high6.23.3
Total190.0100.0
Table 10. Forest fire susceptibility areas.
Table 10. Forest fire susceptibility areas.
Class NameRangeIn km2In %
Very low0–117.29.0
Low1–231.816.7
Moderate2–335.318.6
High3–446.024.2
Very high4–559.831.5
Total 190.0100.0
Table 11. High-vulnerability natural hazard areas in the MKM.
Table 11. High-vulnerability natural hazard areas in the MKM.
Natural HazardIn km2In %
Flash floods6.23.3
Landslides27.914.7
Forest fires59.831.5
Excessive erosion29.415.5
Multi-hazards20.911.0
Total vulnerability102.454.0
Total190.0100.0
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

Aleksova, B.; Milevski, I.; Dragićević, S.; Lukić, T. GIS-Based Integrated Multi-Hazard Vulnerability Assessment in Makedonska Kamenica Municipality, North Macedonia. Atmosphere 2024, 15, 774. https://doi.org/10.3390/atmos15070774

AMA Style

Aleksova B, Milevski I, Dragićević S, Lukić T. GIS-Based Integrated Multi-Hazard Vulnerability Assessment in Makedonska Kamenica Municipality, North Macedonia. Atmosphere. 2024; 15(7):774. https://doi.org/10.3390/atmos15070774

Chicago/Turabian Style

Aleksova, Bojana, Ivica Milevski, Slavoljub Dragićević, and Tin Lukić. 2024. "GIS-Based Integrated Multi-Hazard Vulnerability Assessment in Makedonska Kamenica Municipality, North Macedonia" Atmosphere 15, no. 7: 774. https://doi.org/10.3390/atmos15070774

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