Home Spatio-temporal analysis of flood events using GIS and remote sensing-based approach in the Ukrina River Basin, Bosnia and Herzegovina
Article Open Access

Spatio-temporal analysis of flood events using GIS and remote sensing-based approach in the Ukrina River Basin, Bosnia and Herzegovina

  • Luka Sabljić EMAIL logo , Tin Lukić , Davorin Bajić , Slobodan B. Marković , Velibor Spalevic , Vladimir M. Cvetković , Dragica Delić , Dragutin Adžić , Bojana Aleksova , Ivica Milevski and Gordana Petković Srzentić
Published/Copyright: August 4, 2025
Become an author with De Gruyter Brill

Abstract

The subject of this research is the analysis of flood dynamics in the Ukrina River Basin, Bosnia and Herzegovina, using a remote sensing-based approach and geographic information systems during the period 2016–2019. The aim is to identify the spatial and temporal patterns of floods by integrating satellite-derived precipitation data, hydrological data, and Sentinel-1 imagery processed through Google Earth Engine. The methodology included the use of CHIRPS precipitation data and the Standardized Precipitation Index (SPI) for identifying meteorological anomalies, while Sentinel-1 SAR data were used to map flood extent based on radar backscatter change detection. The approach combined temporal analysis with spatial overlays of land use and administrative boundaries to assess affected areas. Flood events were identified in January 2016 (89.98 ha), March 2017 (179.85 ha), March 2018 (58.81 ha), and May 2019 (195.38 ha), coinciding with periods of above-average precipitation (>125%), positive SPI values, and elevated water levels. The spatial analysis of flooded areas, overlaid with land use data, revealed that agricultural land was the most affected category, with 79.21 ha flooded in 2016, 169.15 ha in 2017, 48.89 ha in 2018, and 184.90 ha in 2019. Built-up areas were also significantly impacted, posing risks to infrastructure and economic stability. The cities and municipalities of Derventa, Prnjavor, and Stanari were most frequently affected by floods during the study period. The findings highlight the role of cumulative precipitation and hydrological conditions in triggering flood events and provide insights for flood risk management, including adaptive strategies, early warning, and sustainable land use planning.

1 Introduction

Climate change represents a threat to people worldwide [1]. According to Klein et al. [2] and Meresa et al. [3], changes in the intensity and frequency of climate change lead to the occurrence of extreme events, such as heavy and intense precipitation, which significantly impact water resource management and flood risk occurrence. Asadieh and Krakauer [4] highlight that, globally, the annual maximum daily precipitation increased by an average of 5.73 mm during the period from 1901 to 2010. Furthermore, the global precipitation average showed a pronounced trend toward more intense rainy seasons and drier periods without precipitation between 1979 and 2010 [5]. Additionally, research results [6,7,8] indicate that the frequency of extreme precipitation has significantly increased over the last few decades. Similarly, recent studies [9] suggest that the observed rate of increase in extreme precipitation at the global level aligns with the well-known Clausius–Clapeyron relationship [10]. In Bosnia and Herzegovina (BH), trends in extreme precipitation share similar characteristics with trends in the region, where an increase in the frequency and intensity of precipitation has been recorded [11]. According to the conclusions of the Intergovernmental Panel on Climate Change [12], areas with increasing average precipitation will experience more intense wet events in the future, while areas with decreasing average precipitation will face more pronounced extreme dry periods [13]. Kundzewicz et al. [14] emphasize that prolonged precipitation (>1 day) of high intensity is considered the most common cause of floods in large basins, while floods in smaller basins are often caused by short-term but extremely intense precipitation.

Under the influence of extreme precipitation [15], floods typically occur when a river receives more water than it can handle, causing its channel capacity to be exceeded, and the excess water spills out, flooding surrounding low and vulnerable areas [16,17]. Numerous studies indicate that floods are one of the most devastating and frequent natural hazards [9,18,19,20,21], affecting more people than any other natural hazard [17] and thus leading to increased mortality, as well as significant economic and environmental damage [22,23]. According to the World Meteorological Organization (WMO), floods have caused an average of 6,500 fatalities over the last 20 years [17], as well as economic damage exceeding 30 billion USD annually [24].

On the other hand, in Europe, flood risks and consequences follow trends observed at the global level. Over the past 30 years, Europe has recorded more than 400 major floods (many of which were catastrophic), affecting over 8.7 million people, claiming more than 2,000 lives, and causing material losses exceeding 72 billion euros [25,26]. According to Ionita and Nagavciuc [26], among the most devastating and costly floods in various parts of Europe are the winter floods of 1993 and 1995 in Germany, the Netherlands, and France [27,28,29,30] the 2002 and 2013 floods in the Elbe River Basin [31,32,33]; the 2005, 2008, and 2010 floods in Eastern Europe [34,35,36]; the 2010 floods in Central Europe [37]; the UK floods of 2000, 2007, and 2014 [38,39,40,41]; and the 2014 floods in the Balkan Peninsula [42,43,44,45]. Among recent floods in Europe, the 2021 floods stand out. During that year, extreme precipitation affected Germany, Belgium, Luxembourg, and neighboring countries, resulting in catastrophic floods, particularly pronounced in western Germany, along the Meuse River and its tributaries in Belgium and the Netherlands [46]. In BH, exceptionally high precipitation amounts have caused catastrophic floods multiple times in recent years [11]. Between 2000 and 2021, floods were recorded in 2000, 2001, 2009, 2010, 2014, 2018, 2019, 2020, and 2021 [47]. The most devastating floods occurred in May and August 2014, when 17 and 14% of the average annual precipitation from the 1961–1990 period fell within just a few days [48]. During that year, all historical records in the instrumental period were broken [47]. For example, the maximum water level of the Vrbas River at Delibašino Selo (Banja Luka, BH) on May 16, 2014, was 816 cm, nearly 2 m higher than the previous maximum water level [49]. Before that, similar weather conditions were recorded in June 2010, when 20% of the total annual precipitation occurred within just 2 days [48]. The intensity and frequency of extreme climate events, including floods, are increasing, and according to the RCP8.5 climate scenario, these events are expected to become even more pronounced in the future in BH and the region [47].

For flood risk management, information about the intensity, frequency, and various factors that can influence floods during flood events is essential [17,50,51]. The response to floods requires careful preparation and effective communication and is typically conducted during emergency situations, where a clear understanding of the event’s characteristics is necessary to implement an intervention plan in the affected area [52]. Unfortunately, determining the extent of floods through field data collection (e.g., observing flood perimeters by citizens) is often unreliable, impractical, time-consuming, and financially exhausting [53]. Moreover, data collected using this method usually have limited spatial coverage and low collection frequency [52].

Therefore, data obtained through remote sensing methods are of exceptional importance, as they enable continuous and up-to-date monitoring, as well as access to historical imagery and synoptic representations of large areas [52]. The availability of multi-functional sources (instruments operating in the visible, thermal, and microwave parts of the electromagnetic spectrum), different spatial scales, and resolutions is key for assessing the extent of flooded areas [54]. Additionally, the application of geographic information systems (GIS) is significant, as they, together with remote sensing technologies, enable rapid access to satellite data and efficient management of large data volumes [55,56,57,58] for analyzing natural disasters such as floods. Numerous studies dedicated to flood mapping or identifying flood-prone areas have been conducted using GIS technologies and remote sensing at the global level [59,60,61,62,63], the European level [64,65,66], and in the BH region [67,68,69,70]. Although flood studies based on geographic information systems (GIS) and remote sensing in BH are still relatively rare, there is a growing number of studies indicating increasing interest and potential for applying these technologies [71,72,73,74].

The main research questions this study aims to address are:

  1. What are the key meteorological and hydrological factors that most significantly contribute to the occurrence of floods in the Ukrina River Basin, and how can these factors be effectively identified and monitored using remote sensing and GIS technologies?

  2. How can satellite data, such as Climate Hazards Center InfraRed Precipitation with Station data (CHIRPS), be used to calculate the Standardized Precipitation Index (SPI) and analyze periods with increased flood risk, and what are their limitations in the context of the Ukrina River Basin?

  3. How do floods affect different land categories in the Ukrina River Basin, particularly agricultural areas, and what are the implications for local communities and land management practices?

  4. To what extent can GIS and remote sensing technologies improve flood monitoring and management in areas with limited infrastructure for traditional monitoring?

  5. How can the integration of remote sensing data into flood risk management strategies contribute to more effective prevention, monitoring, and mitigation of flood consequences in the Ukrina River Basin?

The subject of this research is the identification, monitoring, and analysis of floods using advanced remote sensing techniques and GIS. The research aims to identify the factors that lead to flood occurrences, analyze their spatial and temporal characteristics, and examine the socio-geographical consequences, including impacts on the population and land use practices. A key component of the research includes mapping the spatial extent of floods, identifying the most vulnerable areas, and analyzing land use patterns. The study applies advanced technologies, including remote sensing “products” and modern GIS technologies, to provide a comprehensive understanding of the spatial and temporal flood patterns. The results of the research form the basis for developing flood risk management strategies, improving early warning systems, and planning preventive measures to mitigate adverse effects on the environment and local communities.

2 Study area

The study area is the Ukrina River Basin, located in the northern part of BH. The basin covers a total area of 1498.66 km2 [75,76], extending from 44°35′09″N to 45°05′12″N and from 17°23′53″E to 18°07′53″E (Figure 1). The basin belongs to the southern part of the Pannonian Basin and lies at the junction of two macro-regional units: the Pannonian Basin and the mountain ranges [71]. Specifically, it is situated between the Vrbas River Basin to the west and the Bosna River Basin to the east. According to Tošić [77], the basin’s topographical divide is formed by the Dinaric watershed mountains: the Borja and Uzlomac mountains to the south, as well as mountains stretching across the central and northern parts of the basin, including Ljubić, Čavka, Javorovo, Careva Gora, Motacija, and Vučjak at the far east.

Figure 1 
               The study area with locations of MS whose data were used to assess the validity of satellite-based precipitation estimates.
Figure 1

The study area with locations of MS whose data were used to assess the validity of satellite-based precipitation estimates.

The Ukrina River Basin is geologically diverse, with Mesozoic sedimentary, magmatic, and volcano-sedimentary rocks [78], Paleozoic formations linked to the Motajica Massif [79], and extensive Cenozoic deposits. Neogene sediments from the Miocene and Pliocene epochs cover large lowland areas [71], while Quaternary gravel and sand terraces dominate river valleys [80]. Based on the Basic Geological Map at a scale of 1:100,000 [81], the main geological characteristics of the study area are shown in Figure 2a.

Figure 2 
               Characteristics of the study area: (a) geology, (b) elevation, (c) slope, (d) average sum precipitation, (e) average air temperature, and (f) hydrographic network. The markings (a) of the geological classification from the legend have the following meanings: 1 – Albite granites; 2 – Siltstones, clays, flood facies; 3 – Amphibolites; 4 – Amphibolite gneisses; 5 – Amphibolite schists, amphibolite facies; 6 – Amphibolite schists, epidote-amphibolite facies; 7 – Breccias (predominantly diabase), rarely greywackes; 8 – Brecciated limestones; 9 – Brecciated limestones, marly limestones, sandstones; 10 – Dacites; 11 – Deluvial-proluvial sediments: scree, gravels, and clays; 12 – Diabase-chert formation, metamorphosed ophiolitic melange; 13 – Diabase-chert formation: ophiolitic melange; 14 – Diabase-chert formation: sandstones, shales, cherts with small bodies of diabase and spilite; 15 – Diabases; 16 – Diopside amphibolites; 17 – Dolomites and limestones with megalodonts; 18 – Channel facies: alluvium, river islands, and river beaches; 19 – Flysch: massive and thick-bedded sandstones, shales, clayey marls, and conglomerates; 20 – Flysch: sandstones, siltstones, and clay shales with glauconitic mass; 21 – Flysch: fine-grained and medium-grained sandstones with graded bedding in alternation with siltstones and rarely marls; 22 – Gabbros, predominantly olivine gabbro; 23 – Gabbroperidotites; 24 – Clayey and sandy marls, marly-sandy clays, weathered ferruginous sandstones; 25 – Clayey and sandy gravels and scree with loams; 26 – Garnet-diopside amphibolites; 27 – Coarse-clastic formations: sands, coarse-grained sandstones, conglomerates, sandy clays, platy marly limestones, and tuffs; 28 – Keratophyres; 29 – Clinozoisite-amphibolites; 30 – Conglomerates, sandstones, clays, and marls; 31 – Conglomerates, sandstones, subordinately marls and clays, coal (footwall zone); 32 – Quartz-carbonate rocks (listvenites); 33 – Quartz-sericite schists and phyllites; 34 – Marls, calcareous marls, and limestones; 35 – Marlstones, marls, limestones, sandstones, subordinately conglomerates, tuffs, coal; 36 – Marly clays, marly limestones, and sands; 37 – Marly-sandy clays, marls, oolitic limestones, carbonate sandstones, and conglomerates; 38 – Marly micrites, marls, and arenites; 39 – Lithothamnium and Amphistegina limestones, marls, sandy-clayey marls, clays, sands, conglomerates, and tuffs; 40 – Massive, rarely thin-bedded limestones; 41 – Metamorphosed volcano-sedimentary formations; 42 – Lower river terrace; 43 – Ophitic gabbros; 44 – Ophitic normal gabbros; 45 – Olistostrome breccias; 46 – Peridotites; 47 – Sand and gravel; 48 – Sand, clays, and gravels; 49 – Sandstones, fine-grained conglomerates, marls, and shales; 50 – Sandstones, marls, and sands; 51 – Sands, gravels, partially clayey; 52 – Platy to thick-bedded limestones with brachiopods; 53 – Platy marly limestones and organogenic-reef limestones; 54 – Platy marly micrites; 55 – First river terrace; 56 – Variegated sands and clays and chert gravels; 57 – Variegated sands, subordinately gravels; 58 – Cherts, shales, and limestones; 59 – Cherts, greywacke sandstones, shales, and fine-grained conglomerates with lenses of limestone; 60 – Serpentinites; 61 – Scree/Talus; 62 – Fine-grained sands, interbeds of sandy clays, lenses of gravel; 63 – Gravel and sand; 64 – Bedded marly micrites, marls, and subordinately arenites; 65 – Solifluction; 66 – Spilites; 67 – Thin-platy marly and sandy limestones, shaly limestones, shales, and quartz sandstones; 68 – Weathered, coarse conglomerates, gravels, sandy and gravelly loams; 69 – Limestones, marls, conglomerate-breccia.
Figure 2

Characteristics of the study area: (a) geology, (b) elevation, (c) slope, (d) average sum precipitation, (e) average air temperature, and (f) hydrographic network. The markings (a) of the geological classification from the legend have the following meanings: 1 – Albite granites; 2 – Siltstones, clays, flood facies; 3 – Amphibolites; 4 – Amphibolite gneisses; 5 – Amphibolite schists, amphibolite facies; 6 – Amphibolite schists, epidote-amphibolite facies; 7 – Breccias (predominantly diabase), rarely greywackes; 8 – Brecciated limestones; 9 – Brecciated limestones, marly limestones, sandstones; 10 – Dacites; 11 – Deluvial-proluvial sediments: scree, gravels, and clays; 12 – Diabase-chert formation, metamorphosed ophiolitic melange; 13 – Diabase-chert formation: ophiolitic melange; 14 – Diabase-chert formation: sandstones, shales, cherts with small bodies of diabase and spilite; 15 – Diabases; 16 – Diopside amphibolites; 17 – Dolomites and limestones with megalodonts; 18 – Channel facies: alluvium, river islands, and river beaches; 19 – Flysch: massive and thick-bedded sandstones, shales, clayey marls, and conglomerates; 20 – Flysch: sandstones, siltstones, and clay shales with glauconitic mass; 21 – Flysch: fine-grained and medium-grained sandstones with graded bedding in alternation with siltstones and rarely marls; 22 – Gabbros, predominantly olivine gabbro; 23 – Gabbroperidotites; 24 – Clayey and sandy marls, marly-sandy clays, weathered ferruginous sandstones; 25 – Clayey and sandy gravels and scree with loams; 26 – Garnet-diopside amphibolites; 27 – Coarse-clastic formations: sands, coarse-grained sandstones, conglomerates, sandy clays, platy marly limestones, and tuffs; 28 – Keratophyres; 29 – Clinozoisite-amphibolites; 30 – Conglomerates, sandstones, clays, and marls; 31 – Conglomerates, sandstones, subordinately marls and clays, coal (footwall zone); 32 – Quartz-carbonate rocks (listvenites); 33 – Quartz-sericite schists and phyllites; 34 – Marls, calcareous marls, and limestones; 35 – Marlstones, marls, limestones, sandstones, subordinately conglomerates, tuffs, coal; 36 – Marly clays, marly limestones, and sands; 37 – Marly-sandy clays, marls, oolitic limestones, carbonate sandstones, and conglomerates; 38 – Marly micrites, marls, and arenites; 39 – Lithothamnium and Amphistegina limestones, marls, sandy-clayey marls, clays, sands, conglomerates, and tuffs; 40 – Massive, rarely thin-bedded limestones; 41 – Metamorphosed volcano-sedimentary formations; 42 – Lower river terrace; 43 – Ophitic gabbros; 44 – Ophitic normal gabbros; 45 – Olistostrome breccias; 46 – Peridotites; 47 – Sand and gravel; 48 – Sand, clays, and gravels; 49 – Sandstones, fine-grained conglomerates, marls, and shales; 50 – Sandstones, marls, and sands; 51 – Sands, gravels, partially clayey; 52 – Platy to thick-bedded limestones with brachiopods; 53 – Platy marly limestones and organogenic-reef limestones; 54 – Platy marly micrites; 55 – First river terrace; 56 – Variegated sands and clays and chert gravels; 57 – Variegated sands, subordinately gravels; 58 – Cherts, shales, and limestones; 59 – Cherts, greywacke sandstones, shales, and fine-grained conglomerates with lenses of limestone; 60 – Serpentinites; 61 – Scree/Talus; 62 – Fine-grained sands, interbeds of sandy clays, lenses of gravel; 63 – Gravel and sand; 64 – Bedded marly micrites, marls, and subordinately arenites; 65 – Solifluction; 66 – Spilites; 67 – Thin-platy marly and sandy limestones, shaly limestones, shales, and quartz sandstones; 68 – Weathered, coarse conglomerates, gravels, sandy and gravelly loams; 69 – Limestones, marls, conglomerate-breccia.

Elevation within the basin ranges from 88 to 950 m (Figure 2b). The northern and central parts are dominated by low-lying terrain along the main course of the Ukrina River and its major tributaries, forming typical river valleys. In contrast, the southern and southwestern parts are characterized by higher terrains and steep slopes. The slope map (Figure 2c) shows that flat and gently sloping terrain is located along the riverbanks of the Ukrina. According to the Köppen–Geiger classification [82], the Ukrina River Basin is partly classified as Cfb (southwestern part), characterized by mild winters and moderately warm summers, while the northeastern part falls under the Cfc type, marked by mild winters and short, cool summers. The climatic characteristics of the study area are illustrated through the spatial distribution of the average annual sum of precipitation (Figure 2d) and air temperature (Figure 2e) for the period 2000–2020. The highest precipitation values are recorded in the southern parts of the basin, where the average annual total reaches nearly 996 mm [83], which can be attributed to the orographic influence of elevated areas. In contrast, the northern part of the basin receives lower amounts of precipitation, averaging around 785 mm [83], consistent with the predominantly lowland character of this region. Temperatures range from 12.58°C in the southern, higher-altitude areas to 13.53°C in the northern part of the basin [84]. The spatial distribution of air temperature largely reflects the morphological structure of the terrain, with cooler values associated with higher elevations and warmer areas located in the lower valleys along the Ukrina River.

The Ukrina River Basin represents a hydrologically complex area characterized by a dense and well-developed river network (Figure 2f). The main watercourse, the Ukrina River, originates in the southern part of the basin from the confluence of the Velika and Mala Ukrina rivers and flows from south to north until it joins the Sava River as a right tributary. The total length of the Ukrina River is 134.9 km, while the length of the main watershed ridge within the basin reaches 248.9 km [77]. The horizontal asymmetry of the basin affects the spatial distribution of tributaries and runoff dynamics [77], which is particularly relevant in the formation of rapid and intense flood waves.

According to hydrological indicators, the average annual discharge of the Ukrina at its confluence is 19.8 m3/s, while the maximum discharge during a 100-year flood event (Q1/100) reaches 635.4 m3/s [71,85].

The river network (Figure 2f) is dense, especially in the southern and central parts of the basin, where tributaries such as the Vijaka, Lupljanica, and Ilova contribute to the hydrological system [71]. In addition to natural hydrographic elements (Figure 2f), hydrotechnical structures (dams and embankments) have been mapped within the basin.

The Ukrina River Basin encompasses a total of 405 settlements, which had 242,655 inhabitants and 91,394 households [86]. It is important to emphasize that the population within the basin is smaller than the stated number, as the basin does not cover the entire area of the listed cities and municipalities. Also, according to the Corine Land Cover (CLC) geospatial database (https://land.copernicus.eu/) from 2018, the dominant land use in the Ukrina River Basin consists of agricultural land, which covers 818.10 km2 (54.59%), followed by forested areas 649.47 km2 (43.34%), built-up areas 22.80 km2 (1.52%), and water bodies 8.27 km2 (0.55%). In terms of employment structure, the largest share of employed persons (approximately 40%) is in the economy and manufacturing industries, which rely on agriculture.

3 Methodology and data

The spatial and temporal investigation of floods in the study area is based on the application of various geoinformation technologies, including GIS and remote sensing. The application of these technologies involves the processing and analysis of remote sensing “products” in the form of satellite imagery. The purpose of processing and analyzing satellite data is to identify the occurrence of extreme precipitation, monitor the effects of extreme precipitation on water level rise, and spatially map floods that result from these causal factors. Additionally, this approach aims to assess the population exposed to flood risk and analyze the patterns of land use affected by flooding. The processing of satellite data is conducted using the Google Earth Engine platform and QGIS 3.40 “Bratislava” software (https://qgis.org/). A comprehensive methodological approach to the spatial and temporal investigation of floods is presented in Figure 3, and the same will be explained in detail in the following sections.

Figure 3 
               Algorithm for processing satellite and meteorological data (step 1 – validation of Climate Hazards Group InfraRed Precipitation with Station Data; step 2 – calculation of SPI; step 3 – identification of high water levels; step 4 – mapping of flooded areas and assessment of adverse impacts).
Figure 3

Algorithm for processing satellite and meteorological data (step 1 – validation of Climate Hazards Group InfraRed Precipitation with Station Data; step 2 – calculation of SPI; step 3 – identification of high water levels; step 4 – mapping of flooded areas and assessment of adverse impacts).

3.1 Assessment of the validity of satellite precipitation data

The primary input data for identifying extreme precipitation in the study area are satellite-based precipitation estimates known as CHIRPS. CHIRPS is a relatively new precipitation database, developed by the Climate Hazards Group at the University of California, Santa Barbara [83]. It is based on a combination of rain gauge data, satellite imagery, and (re)analysis products [87]. The data are available from 1981 to near-real-time [88], with a relatively high spatial resolution of 0.05° [89]. Before use, they have been validated and applied in scientific research across various parts of the world, including Cyprus [90,91,92], Argentina [93], Nepal [94], Italy [95], Mozambique [96], Vietnam [97], Pakistan [88], China [98,99], and BH [74,100,101].

The validation of CHIRPS satellite data in the Ukrina River Basin is based on the methodological framework of the study by Sabljić et al. [74], which involves comparing CHIRPS data with meteorological station (MS) data on monthly precipitation for the period 1981–2023. Data from the MS were obtained from the Republic Hydrometeorological Service of the Republic of Srpska. The analysis included data from MS located within the basin boundaries, as well as those in close proximity to the basin (Figure 1, Table 1). The inclusion of stations outside the basin boundaries is justified by the limited number of stations within the basin relative to its area, the short operational period of certain stations, and the relatively “low” spatial resolution of satellite data compared to the extent of the study area.

Table 1

MS whose data were used for the validation assessment

Name of MS Location Time period Elevation (m)
Derventa 44°58′N, 17°54′E 1981–2012 108
Stanari 44°45′N, 17°47′E 2006–2018 169
Banja Vrućica 44°35′N, 17°53′E 2009–2023 206
Prnjavor 44°52′N, 17°39′E 2019–2023 186
Čelinac 44°43′N, 17°19′E 2017–2023 199
Srbac 45°05′N, 17°31′E 2000–2023 107
Modriča 44°57′N, 18°18′E 1997–2002 111

The validation of CHIRPS data at the monthly level is based on the following formula:

(1) V month , i = 1 | P sat , i P met , i | P met , i × 100 ,

where V month,i represents the percentage agreement for month (i), expressed as a percentage; P sat,i is the precipitation value based on satellite data for month (i); P met,i represents the precipitation value based on meteorological data for month (i); and |Psat,i P met,i | is the absolute difference between satellite and meteorological data for month (i).

The overall percentage agreement or validation of CHIRPS data for the period 1981–2023 was determined using the following formula:

(2) V o verall = 1 i = 1 12 | P sat , i P met , i | i = 1 12 P met , i × 100 ,

where V overall is the overall percentage agreement; ∑₌₁12 |Pₛₐₜ,Pₘₑₜ,| is the total absolute difference between satellite and meteorological data for all months, and i = 1 12 P met , i is the total precipitation value based on meteorological data for all months.

In addition to the presented methodological framework, the validity of the CHIRPS data was further confirmed by calculating the Pearson correlation coefficient (R p), following the methodology proposed by Dehaghani et al. [102]. The values of this coefficient range between −1 and 1, and it serves as an indicator of the agreement between precipitation measured by MS (P o i ) and satellite-based CHIRPS precipitation estimates (P s i ). The closer the coefficient is to 1, the stronger the agreement between the satellite data and ground-based measurements. The Pearson correlation coefficient was calculated using the following formula [102]:

(3) Rp = i = 1 N ( P s i P ˆ s ) ( P o i P ˆ o ) i = 1 N ( P s i P ˆ s ) 2 i = 1 N ( P o i P ˆ o ) 2

where P ˆ o is the mean value of observed precipitation, P ˆ s is the mean value of satellite-derived precipitation, and N is the number of precipitation data.

3.2 Meteorological and hydrological characteristics of the flood

Extreme precipitation refers to types of precipitation that significantly deviate from normal values for a specific area or time period. To identify the occurrence of extreme precipitation in the Ukrina River Basin, the average monthly precipitation for a 42-year period (1981–2023) was calculated and compared to the average monthly precipitation during the reference years (2016, 2017, 2018, and 2019). In this study, months in the reference years were considered extreme if their average monthly precipitation exceeded 125% of the corresponding 42-year monthly average. Accordingly, in line with the recommendations of the WMO, a meteorological element encompassing a full climatological cycle was observed [103]. Previously validated CHIRPS satellite precipitation estimates were used as input data for the calculation. The months during the reference years in which precipitation exceeded the 42-year average (>125%) were identified as periods of extreme precipitation that could potentially cause flooding in the study area.

Additional confirmation of the meteorological conditions for flood occurrence was conducted by calculating the SPI. The mentioned index, developed by McKee et al. [104], represents the z-score deviation from the mean expressed in units of standard deviation. In this study, the SPI was calculated following the previously implemented methodological framework of Sabljić et al. [101]. The calculation was performed based on validated CHIRPS precipitation estimates at the location of each pixel for the composite period of each year during the reference period. The formula for calculating the SPI is as follows:

(4) SP I ijk = ( P ijk P ij ̅ ) σ ij ,

where SPI ijk is the z-value for the pixel (i) during timeframe (j) for year (k), P ijk is the precipitation value for pixel (i) during timeframe (j) for year (k), P ij is the mean for pixel (i) during timeframe (j) over n years, and σ ij is the standard deviation of pixel (i) during week (j) over n years.

The obtained SPI results were classified according to McKee et al. [104] and Moccia et al. [105] into multiple classes representing the severity of wet or dry events (Table 2). Each class has an established probability of occurrence, which is, by definition, the same for both wet and dry events.

Table 2

Classification of events based on SPI values and their probability of occurrence (according to McKee et al. [104] and Moccia et al. [105])

SPI values SPI classes
≥2.0 Extremely wet
1.5 to 2.00 Severely wet
1.0 to 1.5 Moderately wet
0 to 1.0 Mildly wet
−1.0 to 0 Mildly drought
−1.5 to −1.0 Moderate drought
−2.0 to −1.5 Severe drought
≤2.00 Extreme drought

Meteorological extremes can significantly influence the hydrological characteristics of the study area. Unfortunately, these characteristics cannot be monitored at the level of the Ukrina River Basin because there are no active hydrological stations (HS) within the study area. For this reason, to assess the impact of meteorological extremes manifested through extreme precipitation and exceptionally wet events, data from the HS Srbac (Vrbas River Basin), located near the Ukrina River Basin, were used. The Vrbas River Basin and the Ukrina River Basin belong to the same larger Sava River Basin [106], indicating their hydrological and spatial connection. This connection further justifies the use of data from HS Srbac in this study, as it is expected that meteorological and hydrological conditions within the broader Sava River Basin exhibit similar patterns. Although the station is not directly located within the Ukrina River Basin, its proximity and connection to the surrounding environment provide relevant data that can serve as an indicator of the hydrological state of the study area. The inclusion of data from HS Srbac partially compensates for the absence of HS within the Ukrina River Basin and provides a foundation for assessing the impact of extreme precipitation and exceptionally wet conditions on water resources. The obtained data form a basis for evaluating potential floods in the Ukrina River Basin, considering broader hydrological processes within the larger Sava River Basin.

To assess the impact of extreme precipitation on water levels, the average mean water level of the Sava River (HS Srbac) for the period 1997–2023 was calculated. The average mean water level for the 26-year period was compared with the average mean water level during the reference years (2016, 2017, 2018, and 2019). Months in which the mean water level exceeded 100% of the long-term monthly average were considered hydrologically significant. Given the station’s indirect spatial relevance to the study area, a more conservative threshold was adopted to ensure that even moderate hydrological responses to meteorological extremes were not overlooked. Identifying higher water levels at HS Srbac, although not directly within the study area, allows for the establishment of a chronology of events and serves as an indicator of the cause-and-effect relationships between extreme precipitation, high SPI values, and flood occurrence. The periods (months) where these three components coincide – extreme precipitation, high SPI, and elevated water levels – represent key timeframes for further analysis and flood identification. During these periods, using remote sensing methods, it is possible to map flooded areas and determine the spatial extent of floods within the study area, provided that the limitations of the temporal resolution of satellite data are taken into account.

3.3 Flood mapping

The methodology for mapping flooded areas is based on processing remote sensing “products” in the form of radar satellite images known as Sentinel-1 [107]. This satellite mission is part of the European Space Agency’s (ESA) radar mission under the Copernicus Sentinel-1 program. It provides continuous imaging in the C-band, regardless of weather conditions or time of day. The mission consists of two satellites, Sentinel-1A and Sentinel-1B, which operate in the same orbital plane [108]. The first satellite, Sentinel-1A, was launched on April 3, 2014, while Sentinel-1B was launched on April 25, 2016 [109]. Mastrorosa et al. [109] emphasize that the Sentinel-1 mission operates in four unique imaging modes with varying resolutions (up to 5 m) and coverage (up to 400 km), and that it enables dual polarization, very short revisit intervals, and fast product delivery. The main characteristics of the Sentinel-1 satellite mission are provided in Table 3 [110].

Table 3

Characteristics of the Sentinel-1 satellite mission (according to Fletcher [110])

Interferometric wide swath mode (IW) Wave mode Strip map mode Extra wide swath mode (EW)
Parameters Polarization Dual (HH + VV, VV + VH) Single (HH, VV) Dual (HH + HV, VV + VH) Dual (HH + HV, VV + VH)
Azimuth resolution 20 m 5 m 5 m 40 m
Ground range resolution 5 m 5 m 5 m 20 m
Azimuth and range looks Single Single Single Single
Products Level 2 Ocean Ocean Ocean Ocean
Level 1 Single look complex Single look complex Single look complex
Level 0 Raw data Raw data Raw data
Characteristic Lifetime 7 years (consumables for 12 years)
Launch date 1A – 3 April 2014 | 1B – 25 April 2016 timeframe
Launcher/location Soyuz, Kourou (both launches)
Orbit Near-polar, Sun-synchronous, about 690 km, 12-day repeat cycle
Orbital period 98.6 min

As part of the research aimed at spatial flood identification, Sentinel-1 images were filtered using the GEE platform based on cloud technology [111]. The images were selected with the following characteristics: vertical polarization (VV), interferometric wide swath (IW) imaging mode, and orbital paths in either descending or ascending mode. The study area was defined by the boundaries of the Ukrina River Basin, obtained from the HydroSHEDS geospatial database (https://www.hydrosheds.org/) [75,76]. Data on hydrotechnical structures (dams and embankments) were obtained from the OpenStreetMap geospatial database (https://www.openstreetmap.org/). The satellite images were divided into two temporal groups: pre-flood period and post-flood period (for each reference year of the study). For each time interval, a mosaic was created by merging available images and reducing them to the study area, which enabled a clear depiction of surface conditions before and after the floods. The data processing included a step to re-move “speckle” noise, characteristic of radar images, using the Refined Lee filter. This algorithm uses local statistics, such as mean and variance, within defined “windows”, reducing noise while preserving spatial details. For the filter application, the input satellite data were first converted from dB to natural units, and after processing, they were converted back to dB format. The identification of flooded areas in the study was based on analyzing changes in backscatter signal values (dB). Flooded areas were defined as pixels that had pre-flood values above −18 dB (non-water surfaces) but showed values below −18 dB (water surfaces) after the flood. The threshold of −18 dB was tested on multiple smaller samples within the basin area before applying it across the entire basin to ensure validation. A similar methodology was applied by Arora et al. [112] in the Ghaggar River Basin in India, where Sentinel-1 SAR data and threshold-based classification were used to detect flood-affected areas. Their approach also included testing of backscatter thresholds and integration with land use and elevation data to improve flood impact assessment, especially in agricultural and built-up zones, which supports the methodological choices in this study. Permanent water surfaces in this study were identified as pixels with values below −18 dB in both time intervals. Based on the presented analysis, binary masks were created, and the flooded areas were spatially represented. To enhance the spatial interpretation and visualization of the flood extent, the European Digital Elevation Model (EU-DEM) with a 25 m resolution was used as a topographic base layer for the final flood maps.

3.4 Assessment of flood damage impact

To assess the damage impact of floods, the results of the mapped floods were overlaid with the administrative boundaries of cities, municipalities, and their settlements, obtained from the Republic Administration for Geodetic and Property-Legal Affairs of the Republic of Srpska (https://www.rgurs.org/). This analysis enabled the identification of administrative units and an estimation of the population at risk of flooding during the period from 2016 to 2019. Additionally, the impact of floods on different land use types in the study area was assessed. The land use data were obtained from the CLC geospatial database for the year 2018, provided by the Copernicus Land Monitoring Service under the European Environment Agency. The CLC database offers pan-European standardized land cover information with a spatial resolution of 100 m, allowing for consistent and comparable analyses across regions. The 2018 dataset was selected as it corresponds to the mid-point of the study period and provides the most up-to-date, validated classification available at the time of analysis. For the purposes of this study, detailed land cover classes were reclassified into three major categories – built-up areas, agricultural land, and forested areas (Table 4) – to simplify the overlay with mapped flood extents and to assess flood exposure in a spatially clear and policy-relevant manner. The processing and analysis of data were performed using QGIS 3.40 “Bratislava” software (https://qgis.org/). By intersecting flood data with land use categories, the quantitative effects of floods were determined for each class. Spatial–temporal insights into the dynamics of flood events highlighted the most vulnerable categories and their spatial distribution.

Table 4

Generalization of land use types based on the CLC 2018 geospatial database

Level 1 Level 2 Level 3 Generalized
1 Artificial surface 11 Urban fabric 111 Continuous urban fabric Built-up area
112 Discontinuous urban fabric
12 Industrial, commercial, and transport units 121 Industrial or commercial units
122 Road and rail networks and associated land
123 Port areas
124 Airports
13 Mine, dump, and transport units 131 Mineral extraction site
132 Dump sites
133 Construction site
14 Artificial, non-agricultural vegetated areas 141 Green urban areas
142 Sport and leisure facilities
2 Agricultural areas 21 Arable land 211 Non-irrigated arable land Agriculture areas
212 Permanently irrigated land
213 Rice fields
22 Permanent crops 221 Vineyards
222 Fruit trees and berry plantations
223 Olive groves
23 Pastures 231 Pastures
24 Heterogeneous agricultural areas 241 Annual crops associated with permanent crops
242 Complex cultivation patterns
243 Land principally occupied by agriculture, with significant areas of natural vegetation
244 Agro-forestry areas
3 Forest and semi natural areas 31 Forests 311 Broad-leaved forest Forest areas
312 Coniferous forest
313 Mixed forest
32 Scrub and/or herbaceous vegetation associations 321 Natural grasslands
322 Moors and heathland
323 Sclerophyllous vegetation
324 Transitional woodland-shrub
33 Open spaces with little or no vegetation 331 Beaches, dunes, sands
332 Bare rocks
333 Sparsely vegetated areas
334 Burnt areas
335 Glaciers and perpetual snow

4 Results

4.1 Validation of satellite precipitation data

The validation of CHIRPS satellite precipitation estimates was conducted according to the previously described methodology for the time period 1981–2023. A comparison of the CHIRPS data with MS data indicates a high degree of validity of the CHIRPS data (Figure 4a). May and December are characterized by a >85% agreement between CHIRPS data and MS data. April and June exhibit >90% agreement, while 6 out of 12 months (February, March, July, September, October, and November) show >95% agreement. Lower agreement percentages were recorded in August (82.42%) and January (79.69%). The overall percentage agreement of CHIRPS satellite precipitation estimates with MS data for the period 1981–2023 is 94.18%, which confirms that these data are valid for use in the Ukrina River Basin. To further support this conclusion, a Pearson correlation analysis was performed, revealing a very strong and statistically significant positive correlation between the monthly CHIRPS estimates and MS observations (R p = 0.875). This result, presented in Figure 4b, provides additional confirmation of the reliability of CHIRPS satellite precipitation data for the study area.

Figure 4 
                  Validation of CHIRPS data (1981–2023): (a) Monthly precipitation comparison (MS vs CHIRPS) and (b) Pearson correlation analysis (R
                     p = 0.875).
Figure 4

Validation of CHIRPS data (1981–2023): (a) Monthly precipitation comparison (MS vs CHIRPS) and (b) Pearson correlation analysis (R p = 0.875).

4.2 Identification of meteorological and hydrological preconditions for flood occurrence

To identify extreme precipitation in the Ukrina River Basin, a comparison was made between the CHIRPS average monthly precipitation for the period 1981–2023 and the CHIRPS average monthly precipitation during the reference years (2016, 2017, 2018, and 2019). Months in the reference years with precipitation exceeding 125% of the 42-year monthly average were classified as extreme. This threshold represents a 25% increase over the long-term average and serves as an indicator of significant precipitation anomalies.

In 2016 (Figure 5a), extreme precipitation was recorded in February (190.60%) and July (161.20%). January (127.60%) also exceeded the 125% threshold. March (123.10%) was slightly below but close to the threshold. In 2017 (Figure 5b), extreme values were observed in February (131.40%), April (133.60%), and December (144.10%). March (124.20%) and September (123.60%) were near the threshold. During 2018 (Figure 5c), March (189.60%), May (134.30%), and July (156.40%) exceeded the 125% threshold. February (127.60%) and June (124.40%) also approached or slightly exceeded this value. December recorded 123.00%, close to the extreme threshold. In 2019 (Figure 5d), only May (141.00%) surpassed the 125% threshold. June (117.10%) and November (114.00%) were elevated but below the defined limit.

Figure 5 
                  Percentage-based comparison of CHIRPS monthly precipitation during (a) 2016, (b) 2017, (c) 2018, and (d) 2019 with the 1981–2023 average.
Figure 5

Percentage-based comparison of CHIRPS monthly precipitation during (a) 2016, (b) 2017, (c) 2018, and (d) 2019 with the 1981–2023 average.

For a more precise analysis of extreme precipitation, the SPI index provides significant insights into moisture intensity, complementing the results of CHIRPS average precipitation data and enabling a better understanding of water conditions on a monthly level. While CHIRPS satellite precipitation estimates identify months with above-average precipitation, the SPI index allows for the assessment of moisture levels and the potential risk of hydrological extremes, such as floods.

During 2016 (Figure 6a), above-average precipitation in January, March, and July corresponds with positive SPI values. Mildly wet conditions were recorded in January (0.59) and March (0.34), while severely wet conditions were observed in July (1.48). In 2017 (Figure 6b), positive SPI values “follow” above-average precipitation over several months. Mildly wet conditions were recorded in February (0.70), March (0.53), May (0.09), September (0.38), October (0.06), and December (0.90). On the other hand, moderately wet conditions were observed in April (1.01). During 2018 (Figure 6c), significant positive SPI values were identified in months with above-average precipitation: mildly wet conditions were recorded in January (0.21), February (0.61), June (0.66), and December (0.49), while moderately wet conditions were observed in May (1.15), and extremely wet conditions in March (2.06). In 2019 (Figure 6d), mildly wet conditions were recorded in April (0.57), June (0.48), July (0.35), August (0.03), November (0.32), and December (0.09), while moderately wet conditions were observed in May (1.32).

Figure 6 
                  Monthly SPI values for (a) 2016, (b) 2017, (c) 2018, and (d) 2019.
Figure 6

Monthly SPI values for (a) 2016, (b) 2017, (c) 2018, and (d) 2019.

Extreme precipitation, as well as high SPI values identified using CHIRPS satellite data, can significantly influence water levels both in the Ukrina River Basin and the surrounding area. Although there is no HS within the Ukrina River Basin, data from the HS Srbac, located near the basin, were used for the purposes of this research. To identify relevant hydrological responses, particular attention was given to months in which recorded water levels exceeded the long-term monthly average (>100%). This approach enables monitoring the reflection of meteorological characteristics on hydrological conditions in the region, providing an approximate picture of the relationship between precipitation and water levels in the study area.

In 2016 (Figure 7a), months with water levels above the long-term average included January (102.78%), February (135.24%), March (166.13%), May (127.71%), June (104.84%), July (118.01%), August (102.62%), and November (133.12%). The most pronounced anomalies occurred in February and March, both significantly exceeding the threshold and indicating a high flood risk. January, May, and July also recorded noticeable increases in water levels. In 2017 (Figure 7b), elevated water levels were observed in February (113.91%), March (104.06%), September (136.98%), and December (143.36%). December recorded the highest value for the year, suggesting cumulative hydrological impacts, while February and March aligned with elevated precipitation and SPI values, reinforcing their hydrological relevance. During 2018 (Figure 7c), significantly elevated water levels were registered in January (154.16%), February (121.18%), March (157.09%), April (154.79%), and July (122.47%). March stands out as the peak month, followed closely by January and April, pointing to a clustering of extreme hydrological responses in the first half of the year. In 2019 (Figure 7d), water levels exceeded the average in February (107.35%), May (154.93%), June (162.55%), November (123.54%), and December (117.71%). May and June recorded the highest values, making them critical months for flood monitoring, while late-year elevations in November and December indicate sustained hydrological activity beyond the main precipitation season.

Figure 7 
                  Percentage-based comparison of average mean water levels during (a) 2016, (b) 2017, (c) 2018, and (d) 2019 with the 1997–2023 average.
Figure 7

Percentage-based comparison of average mean water levels during (a) 2016, (b) 2017, (c) 2018, and (d) 2019 with the 1997–2023 average.

4.3 Flood mapping and assessment of flood damage impact

Flood mapping in the study area relies on identifying months characterized by above-average precipitation, positive SPI index values, and elevated water levels. Considering the temporal resolution of Sentinel-1 satellite imagery, the aim of the research was to map at least one flood event in each observed year.

For the purpose of enhancing the spatial interpretation of flood extent, a digital elevation model (DEM) was used as a base layer in Figure 8. The elevation data allow a clearer visualization of flood distribution in relation to terrain morphology. As visible from the Figure 8, flooded areas predominantly occur in low-lying agricultural zones, typically located in alluvial plains along the Ukrina River and its tributaries.

Figure 8 
                  Mapping of flooded areas in the Ukrina River basin during (a) 2016, (b) 2017, (c) 2018, and (d) 2019.
Figure 8

Mapping of flooded areas in the Ukrina River basin during (a) 2016, (b) 2017, (c) 2018, and (d) 2019.

In 2016, favorable conditions for flood mapping were recorded in January, February, and March. Floods in January 2016 were successfully mapped, covering an area of 89.98 ha (Figure 8a). In 2017, favorable conditions were identified in February, March, September, and December. Floods from March 2017 were successfully mapped, covering 179.85 ha (Figure 8b), and their development is associated with favorable wet conditions observed in February, indicating continuity in the processes that led to the floods. In contrast, although wet conditions were recorded in September of the same year, floods were not considered because the preceding months, including June, July, and August, were marked by below-average precipitation, significantly negative SPI values, and falling water levels. It was concluded that the soil during this period was unsaturated, making the September precipitation insufficient to cause flooding. Similarly, flood conditions in December 2017 were deemed inadequate for the same reasons. In 2018, favorable conditions for flood mapping were recorded in January, February, March, and July. Wet conditions from the first three months allowed for the successful mapping of floods in March 2018 (Figure 8c), when the flooded area amounted to 58.81 ha. However, floods from July 2018 could not be mapped due to the temporal resolution limitations of Sentinel-1 imagery. In 2019, favorable conditions for floods were identified in May, June, November, and December. A successfully mapped flooded area of 195.38 ha was recorded in May 2019 (Figure 8d). On the other hand, the conditions in November and December were not analyzed because the preceding months (September and October) were characterized by below-average precipitation and negative SPI values, which indicated unfavorable conditions for the occurrence of floods.

The variation in the timing of the mapped flood events across different months and years reflects the inherent irregularity of hydrometeorological conditions in the Ukrina River Basin. Each flood was associated with a unique combination of meteorological and hydrological drivers (above-average precipitation, positive SPI values, and elevated water levels), and thus, the temporal distribution of events does not follow a strict seasonal pattern. This irregularity highlights the importance of a multi-indicator approach and emphasizes the need for continuous monitoring throughout the year.

The socio-geographical analysis of floods during the period 2016–2019 highlights the vulnerability of the Derventa, Prnjavor, and, in 2017, Stanari municipalities. In the Derventa municipality, floods regularly affected settlements such as Miškovci, Rapčani, Lužani, Begluci, Polje, and Pojezina. In the Prnjavor municipality, the most frequently impacted settlements included Donji Vijačani, Gornji and Donji Štrpci, and Drenova. In the Stanari municipality, the floods of 2017 were recorded in the settlements of Dragalovci, Osredak, Ljeb, Stanari, Cvrtkovci, and Brestovo.

According to the 2013 census [86], the number of people affected by floods in the settlements of Derventa ranged between 15,000 and 20,000, while in Prnjavor, it varied from 3,000 to 10,000, depending on the year. In the Stanari municipality, the 2017 floods impacted settlements with approximately 4,500 residents. The largest spatial extent of floods was recorded in 2019, while the 2017 floods affected the greatest number of settlements across all three municipalities.

The spatial extents of the mapped floods were also analyzed in the context of land use types to assess their impact on different surface categories (Figure 9). The results indicate that agricultural areas were the most affected, while the extent of flooded built-up and forest areas varied depending on the year. During the 2016 floods, 4.90 ha of built-up areas, 79.21 ha of agricultural areas, and 5.87 ha of forest areas were flooded. The 2017 floods brought a significant increase in the extent of flooded agricultural areas (169.15 ha), while the extent of flooded built-up (3.76 ha) and forest areas (6.94 ha) remained relatively low. In 2018, there was a notable decline in the total flood extent, with 2.66 ha of built-up areas, 48.89 ha of agricultural areas, and 7.26 ha of forest areas affected. However, 2019 was characterized by a renewed increase in the flood extent, with 2.84 ha of built-up areas, 184.90 ha of agricultural areas, and 6.64 ha of forest areas flooded.

Figure 9 
                  Flooded land use types.
Figure 9

Flooded land use types.

The analysis of the spatial extent of mapped floods indicates that agricultural areas are the most vulnerable land category in the basin. These areas are primarily located in close proximity to rivers, which is directly related to the soil fertility of alluvial plains. While such a spatial arrangement is favorable for agricultural production, it simultaneously makes these areas particularly susceptible to flood risk. Field photographs (Figure 10a–h) provide visual confirmation of the locations of agricultural areas in river valleys, as well as their connection to the hydrological conditions of the area. In addition to agricultural land, a significant share of built-up areas is exposed to flood risk. Built-up areas in the basin, which include residential, infrastructural, and industrial structures, pose a serious threat to the safety of the local population in the event of flooding. Furthermore, since the basin's population is heavily engaged in agriculture, the loss of crops and damage to residential and industrial structures have a direct impact on the economic and social stability of the community in the study area.

Figure 10 
                  (a–h) Natural characteristics and agricultural areas of the Ukrina River Basin (photos by: L. Sabljić and D. Delić).
Figure 10

(a–h) Natural characteristics and agricultural areas of the Ukrina River Basin (photos by: L. Sabljić and D. Delić).

The presented results highlight key aspects of the vulnerability of various land categories and settlements to floods, providing a valuable foundation for decision-making in flood risk management. Detailed mapping and analysis of flood-prone areas enable the responsible authorities to more accurately plan preventive measures, such as the development of spatial plans, protection strategies, and emergency interventions. Integrating these results into local and regional policy documents can improve flood management by directing resources to the most vulnerable areas and developing adaptive measures, including the restoration of natural floodplains, infrastructure protection, and public education. The practical application of this data represents an important step toward sustainable flood risk management and reducing their harmful consequences.

5 Discussion

Until a few decades ago, data from rain gauges represented the primary source of precipitation data. According to Michaelides et al. [113] and Brocca et al. [114], these data are characterized by limited spatial coverage and an unrealistic representation of precipitation distribution in study areas. On the other hand, data obtained from meteorological radars have limitations in data quality due to signal distortion [115]. Satellite-based precipitation data cover large areas with high spatial and temporal resolution [74]. However, this type of data also has certain limitations and shortcomings, which can be mitigated through post-processing of satellite data [116]. Accordingly, as highlighted by Loew et al. [117] and Kumar et al. [118], it is necessary to perform a validation assessment of these data using MS data before applying them in research. The CHIRPS satellite precipitation data used in this study have previously been validated in specific areas of BH. According to Sabljić et al. [100], these data were validated with 89.68% accuracy in the municipality of Stanari (BH) for the period 2000–2020. Similarly, the validity of CHIRPS data was confirmed for the Sana River Basin (BH), with accuracy reaching 92.63% for the period 1992–2022 [74] and 94.95% for the period 1981–2023 [101]. The results of CHIRPS data validation for the Ukrina River Basin are consistent with these findings, as a percentage match of 94.18% with MS data was determined for the period 1981–2023.

In recent decades, the impact of extreme precipitation on flood occurrences has been observed not only in BH but throughout Southeastern Europe, including the neighboring Republic of Serbia during May 2014 [119,120]. Precipitation during this period exceeded 200 mm within 72 h, surpassing average May values and causing some of the most catastrophic floods in Serbia's modern history [121,122]. In certain areas, the daily precipitation reached between 190 and 219 l/m2 [123]. On the other hand, Popov et al. [11] highlighted that during the same period in BH, 146.6 mm of precipitation was recorded within a few days, representing 17% of the annual average precipitation for the standard climatological period (1961–1990). More recently, Leščešen et al. [124] identified a positive trend in winter precipitation (December–January–February) in the upper Vardar River Basin in North Macedonia, where one of the most severe flood events was recorded in January and February 2016. In line with these findings, Sabljić et al. [101] identified above-average precipitation in BH during the winter months of 2016. Specifically, extreme precipitation and significantly positive SPI values were recorded in the Sana River Basin, with 0.39 in January and 2.69 in February [101]. Similar to these findings, this study identified extreme precipitation and distinctly positive SPI values during January and February 2016, which ultimately resulted in the occurrence of floods in the Ukrina River Basin. Moreover, during this period, a shift of hydro-meteorological hazards was evident, as both floods and droughts occurred in the same year across BH. This is further supported by the findings of Milanović Pešić et al. [125], who analyzed hydroclimatic trends in the lower courses of the Una, Sana, and Vrbanja rivers (basins geographically adjacent to the Ukrina River Basin). Their study revealed that, despite relatively stable precipitation trends, discharge levels exhibited a declining tendency due to increased air temperatures and enhanced evapotranspiration. These results align with the notion that thermal factors play a crucial role in modulating hydrological responses in small basins of northern BH, particularly under extreme weather conditions.

The results of Sabljić et al. [74] regarding floods caused by extreme precipitation in February 2017 in the Sana River Basin align with the pattern of extreme precipitation and its consequences in the form of floods in the Ukrina River Basin in March 2017. The 1-month difference highlights similar hydrological conditions in both cases, where intensive precipitation over a short period can significantly impact small river basins, causing floods with comparable consequences. In March 2018, extreme precipitation led to a significant rise in the Sana River water level (428 cm), resulting in the river overflowing its banks and flooding 518 ha [73]. These findings align with the occurrences of extreme precipitation and elevated SPI values identified in the Ukrina River Basin, where floods were also successfully spatially identified during March 2018, pointing to the same pattern of hydrological extremes in the region. In May 2019, the water level of the Sana River exceeded the emergency flood defense threshold, reaching a maximum value of 474 cm with a discharge of 902 m3/s [73], which caused floods successfully spatially identified using satellite data [72]. Similarly, during the same period, flooded areas in the Ukrina River Basin were also successfully spatially identified, demonstrating the efficiency of applying satellite data in analyzing hydrological extremes in the region. Additional insight into regional flood dynamics is provided by Gnjato et al. [126], who analyzed trends in the frequency and timing of maximum river discharges across BH. Their results revealed a noticeable increase in flood-prone events over the last two decades, particularly in the northwestern part of the country – including the Una, Sana, Vrbas, and Bosna river basins. These basins are geographically and climatologically comparable to the Ukrina River Basin, which lies between the Vrbas and Bosna. The study highlighted a shift toward more frequent and later spring peak discharges (May–June), attributed to intensified rainfall episodes. These findings reinforce the conclusion that the flood dynamics observed in the Ukrina basin are part of a broader regional hydrological trend affecting small and medium-sized catchments in northern BH.

Lovrić et al. [71] emphasized in their research that 54 and 41% of the Ukrina River Basin area, based on the Index-Based Method and Flash Flood Potential Index, is classified into high and very high flood susceptibility categories. The topographical characteristics of the basin significantly contribute to this high susceptibility. As noted in the description of the study area, the higher terrains in the southern and southwestern parts are characterized by steep slopes. Such morphology, while reducing the probability of flooding in those specific zones, contributes to increased erosion potential and accelerated water flow toward lower-lying areas, thereby intensifying hydrological pressure in the valley regions. Furthermore, the flat and gently sloping terrain along the riverbanks of the Ukrina represents zones of increased flood risk. These areas, due to their limited natural drainage capacity and low slope gradients, are characterized by slower water runoff and more frequent occurrences of water stagnation following extreme precipitation events. This aligns with findings in this study, which show that flooded areas are predominantly located in lowland agricultural zones. This high susceptibility, combined with the recorded extreme precipitation and frequent flood occurrences, confirms the importance of continuous analysis of hydro-meteorological extremes in the Ukrina River Basin. By applying advanced satellite technologies, it is possible not only to effectively identify floods but also to better understand their causes and mitigate their consequences on local communities and natural systems. Additionally, a significant segment of flood analysis involves assessing their damage impact. According to the results of flood mapping by Ivanišević et al. [73] in the lower Sana River Basin during 2018 and 2019, >95% of the flooded areas were agricultural land (> 400 ha). These results are expected since areas near watercourses, due to their natural characteristics, are favorable for agriculture [127]. Similarly, the results of this study indicate that agricultural areas are the most affected by floods, highlighting the need for priority planning and flood protection infrastructure development in zones where floods regularly recur. Furthermore, adequate infrastructural measures, such as levees, are needed to protect built-up areas, which, according to this study, are also affected by floods in the study area. While existing hydrotechnical structures such as dams and embankments play a role in water level regulation, particularly in the lower course of the river where flood exposure is greatest, this study recognizes that their current capacity and spatial distribution are often insufficient for effective protection under extreme hydrometeorological conditions. Moreover, areas identified as chronically vulnerable require integration into broader risk management strategies, including the adoption of sustainable land management methods and improving early warning systems. These approaches can not only reduce economic and environmental damages but also enhance the resilience of local communities to hydro-meteorological hazards such as floods.

6 Conclusion

This research demonstrated the effectiveness of integrating Sentinel-1 SAR data, CHIRPS precipitation estimates, SPI index values, and hydrological information for identifying and analyzing flood events in the Ukrina River Basin during 2016–2019. The remote sensing-based approach enabled the detection of flooded areas in January 2016, March 2017 and 2018, and May 2019. Among the different land use types, agricultural areas were the most affected, which is consistent with their spatial concentration in lowland zones along river corridors.

The results highlight the potential of remote sensing and GIS tools to improve flood monitoring in areas where ground-based hydrometeorological infrastructure is limited. By combining flood extent data with administrative boundaries and land use layers, we identified both vulnerable land categories and population exposure. This spatial information can assist decision-makers in prioritizing flood protection measures, infrastructure investment, and adaptation planning.

Nevertheless, the study has certain limitations. The spatial resolution of Sentinel-1 imagery (10 m) may not detect localized flood events in small or narrow floodplains. The temporal resolution of satellite imagery (typically 6–12 days) may result in missed events or delayed mapping, especially during short-duration floods. Additionally, the absence of active HSs within the Ukrina River Basin limits the ability to validate flood peaks and water discharge directly, relying instead on proxy data from nearby stations. The land use dataset (CLC 2018) may also not fully capture land cover changes during the entire study period. Moreover, the adopted methodology, based on coinciding meteorological and hydrological indicators, may overlook certain flood events triggered by rapid runoff over dry antecedent soils. Such events, while lacking pronounced cumulative wetness, can still pose significant flood hazards due to reduced infiltration capacity. Future research could address this by incorporating satellite-derived soil moisture products (e.g., SMAP or ASCAT), which would enable a more comprehensive assessment of antecedent conditions and improve the accuracy of flood detection under diverse hydrometeorological scenarios. Also, future research should focus on integrating higher-resolution satellite products (e.g., Sentinel-2 or commercial imagery), as well as topographic data such as DEMs, to refine floodplain delineation and hydrological modeling. Expanding the analysis toward multi-hazard assessment (including droughts and wildfires) would contribute to the development of an integrated risk management system in the Ukrina River Basin. The presented results form a valuable foundation for designing adaptive land use policies and improving local flood preparedness through spatially informed early warning systems.

Acknowledgments

The authors are grateful to the anonymous reviewers whose comments and suggestions improved the manuscript. The authors would like to thank the Ministry of Agriculture, Forestry, and Water Management of the Republic of Srpska, as well as the Republic Hydrometeorological Institute of the Republic of Srpska, for providing meteorological data (No. 12/1.03-79-1/24). Also, L.S. acknowledges the support of the Ministry of Scientific and Technological Development and Higher Education of the Republic of Srpska through the project “Analysis of Flood Frequency in the Sava River Basin in the Republic of Srpska” (No. 1259111). Furthermore, T.L. and S.B.M. acknowledge the support by the Science Fund of the Republic of Serbia, #17807, The Loess Plateau Margins: Towards Innovative Sustainable Conservation – LAMINATION.

  1. Funding information: This research received no external funding.

  2. Author contributions: Conceptualization, L.S., T.L., and D.B.; methodology, L.S.; software, L.S., D.B., and D.A.; validation, L.S., D.B., T.L., and V.S.; formal analysis, L.S. and D.D.; investigation, L.S. and D.D.; resources, T.L.; data curation, L.S., D.B., and D.A.; writing – original draft preparation, L.S. and T.L.; writing – review and editing, L.S., T.L., D.B., S.B.M., V.S., V.M.C., D.D., D.A., B.A., I.M., and G.P.S.; visualization, L.S.; supervision, T.L., D.B., and D.A.; project administration, L.S., B.A., and I.M.; funding acquisition, T.L.

  3. Conflict of interest: The authors declare no conflicts of interest.

  4. Data availability statement: The datasets used during the current study are available from the corresponding author on reasonable request.

References

[1] IPCC. Summary for policymakers. In: Masson-Delmotte V, Zhai P, Pirani A, Connors SL, Péan C, Berger S, et al. Climate Change 2021: The Physical science basis. Contribution of working group I to the sixth assessment report of the intergovernmental panel on climate change. Cambridge, UK: Cambridge University Press; 2021. p. 3–32. 10.1017/9781009157896.001.Search in Google Scholar

[2] Klein RJT, Midgley GF, Preston BL, Alam M, Berkhout FGH, Dow K, et al. Adaptation opportunities, constraints, and limits. In: Field CB, Barros VR, Dokken DJ, Mach KJ, Mastrandrea MD, Bilir TE, et al. editors. Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK: Cambridge University Press; 2014. p. 899–943. https://www.ipcc.ch/site/assets/uploads/2018/02/WGIIAR5-Chap16_FINAL.pdf. Accessed October 10, 2024.Search in Google Scholar

[3] Meresa H, Tischbein B, Mekonnen T. Climate change impact on extreme precipitation and peak flood magnitude and frequency: Observations from CMIP6 and hydrological models, Nat Hazards, 2022;111:2649–79. 10.1007/s11069-021-05152-3.Search in Google Scholar

[4] Asadieh B, Krakauer NY. Global trends in extreme precipitation: Climate models versus observations. Hydrol Earth Syst Sci. 2015; 19:877–91. 10.5194/hess-19-877-2015.Search in Google Scholar

[5] Chou C, Chiang JCH, Lan CW, Chung CH, Liao YC, Lee CJ. Increase in the range between wet and dry season precipitation. Nat Geosci. 2013;6:263–7. 10.1038/ngeo1744.Search in Google Scholar

[6] Donat MG, Alexander LV, Yang H, Durre I, Vose R, Dunn RJH, et al. Updated analyses of temperature and precipitation extreme indices since the beginning of the twentieth century: The HadEX2 dataset. J Geophys Res Atmos. 2013;118(5):2098–118. 10.1002/jgrd.50150.Search in Google Scholar

[7] Westra S, Alexander LV, Zwiers FW. Global increasing trends in annual maximum daily precipitation. J Clim. 2013;26(11):3904–18. 10.1175/JCLI-D-12-00502.1.Search in Google Scholar

[8] Papalexiou SM, Montanari A. Global and regional increase of precipitation extremes under global warming. Water Resour Res. 2019;55(6):4901–14. 10.1029/2018WR024067.Search in Google Scholar

[9] Seneviratne SI, Zhang X, Adnan M, Badi W, Dereczynski C, Di Luca A, et al. Weather and climate extreme events in a changing climate. In: Masson-Delmotte V, Zhai P, Pirani A, Connors SL, Péan C, Berger S, et al., editors. Climate change 2021: The physical science basis. contribution of working group I to the sixth assessment report of the intergovernmental panel on climate change. Cambridge, UK: Cambridge University Press; 2021. p. 1513–766. 10.1017/9781009157896.013.Search in Google Scholar

[10] Pall P, Allen MR, Stone DA. Testing the Clausius-Clapeyron constraint on changes in extreme precipitation under CO2 warming. Clim Dyn. 2007;28:351–63. 10.1007/s00382-006-0180-2.Search in Google Scholar

[11] Popov T, Gnjato S, Trbić G, Ivanišević M. Trends in extreme daily precipitation indices in Bosnia and Herzegovina. Coll Pap Fac Geogr Univ Belgrade. 2017;65(1):5–24. 10.5937/zrgfub1765005P.Search in Google Scholar

[12] Solomon S, Qin D, Manning M, Chen Z, Marquis M, Averyt KB, et al. Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK: Cambridge University Press; 2007.Search in Google Scholar

[13] Guhathakurta P, Sreejith OP, Menon PA. Impact of climate change on extreme rainfall events and flood risk in India. J Earth Syst Sci. 2011;120:359–73. 10.1007/s12040-011-0082-5.Search in Google Scholar

[14] Kundzewicz ZW, Kanae S, Seneviratne SI, Handmer J, Nicholls N, Peduzzi P, et al. Flood risk and climate change: Global and regional perspectives. Hydrol Sci J. 2014;59(1):1–28. 10.1080/02626667.2013.857411.Search in Google Scholar

[15] Khouz A, Trindade J, Santos PP, Oliveira SC, El Bchari F, Bougadir B, et al. Flood susceptibility assessment through statistical models and HEC-RAS analysis for sustainable management in Essaouira Province, Morocco. Geosciences. 2023;13(12):382. 10.3390/geosciences13120382.Search in Google Scholar

[16] Rincón D, Khan UT, Armenakis C. Flood risk mapping using GIS and multi-criteria analysis: A Greater Toronto Area case study. Geosciences. 2018;8(8):275. 10.3390/geosciences8080275.Search in Google Scholar

[17] Ghosh A, Chatterjee U, Pal SC, Towfiqul Islam ARM, Alam E, Islam MK. Flood hazard mapping using GIS-based statistical model in vulnerable riparian regions of sub-tropical environment. Geocarto Int. 2023;38(1):2285355. 10.1080/10106049.2023.2285355.Search in Google Scholar

[18] Coumou D, Rahmstorf S. A decade of weather extremes. Nat Clim Change. 2012;2:491–6. 10.1038/nclimate1452.Search in Google Scholar

[19] Merz B, Blöschl G, Vorogushyn S, Dottori F, Aerts JCJH, Bates P, et al. Causes, impacts, and patterns of disastrous river floods. Nat Rev Earth Environ. 2021;2:592–609. 10.1038/s43017-021-00195-3.Search in Google Scholar

[20] Manoj A, Pérez Ciria T, Chioasadiehgna G, Salzmann N, Agarwal A. Characterising the coincidence of soil moisture–precipitation extremes as a possible precursor to European floods. J Hydrol. 2023;620(A):129445. 10.1016/j.jhydrol.2023.129445.Search in Google Scholar

[21] Aktar A, Shohani K, Hasan N, Hasan K. Flood vulnerability assessment by Flood Vulnerability Index (FVI) method: A study on Sirajganj Sadar Upazila. Int J Disaster Risk Manag. 2021;3(1):1–14. 10.18485/ijdrm.2021.3.1.1.Search in Google Scholar

[22] Lukić T, Lukić A, Basarin B, Ponjiger T, Blagojević D, Mesaroš M, et al. Rainfall erosivity and extreme precipitation in the Pannonian Basin. Open Geosci. 2019;11(1):664–81. 10.1515/geo-2019-0053.Search in Google Scholar

[23] Odero NA, Mahiri I. The complacency of flood victims, socio-economic factors, and effects and vulnerabilities of floods in Lower Kano Plains, Kisumu County, Kenya. Int J Disaster Risk Manag. 2022;4(2):59–77. 10.18485/ijdrm.2022.4.2.4.Search in Google Scholar

[24] Roxy MK, Ghosh S, Pathak A, Athulya R, Mujumdar M, Murtugudde R, et al. A threefold rise in widespread extreme rain events over Central India. Nat Commun. 2017;8(708):9. 10.1038/s41467-017-00744-.Search in Google Scholar

[25] World Health Organization (WHO). Flooding: managing health risks in the WHO European Region. Copenhagen, Denmark: World Health Organization Regional Office for Europe; 2017. ISBN: 978-9-28905-279-5.Search in Google Scholar

[26] Ionita M, Nagavciuc V. Extreme floods in the eastern part of Europe: Large-scale drivers and associated impacts. Water. 2021;13(8):1122. 10.3390/w13081122.Search in Google Scholar

[27] Chbab EH. How extreme were the 1995 flood waves on the rivers Rhine and Meuse? Phys Chem Earth. 1995;20(5–6):455–8. 10.1016/S0079-1946(96)00005-5.Search in Google Scholar

[28] Fink A, Ulbrich U, Engel H. Aspects of the January 1995 flood in Germany. Weather. 1996;51(2):34–9. 10.1002/j.1477-8696.1996.tb06182.x.Search in Google Scholar

[29] Engel H, Busch N, Wilke K, Krahe P, Mendel HG, Giebel H, et al. Das Hochwasser 1993/94 im Rheingebiet. Koblenz, Germany: Bundesanstalt für Gewässerkunde; 1994. BfG-Nr.Search in Google Scholar

[30] Ionita M, Nagavciuc V, Guan B. Rivers in the sky, flooding on the ground: The role of atmospheric rivers in inland flooding in Central Europe. Hydrol Earth Syst Sci. 2020;24(11):5125–47. 10.5194/hess-24-5125-2020.Search in Google Scholar

[31] Ulbrich U, Brücher T, Fink AH, Leckebusch GC, Krüger A, Pinto JG. The Central European floods of August 2002: Part 1—Rainfall periods and flood development. Weather. 2003;58(10):371–7. 10.1256/wea.61.03A.Search in Google Scholar

[32] Ulbrich U, Brücher T, Fink AH, Leckebusch GC, Krüger A, Pinto JG. The Central European floods of August 2002: Part 2—Synoptic causes and considerations with respect to climatic change. Weather. 2003;58(11):434–42. 10.1256/wea.61.03B.Search in Google Scholar

[33] Ionita M, Dima M, Lohmann G, Scholz P, Rimbu N. Predicting the June 2013 European flooding based on precipitation, soil moisture, and sea level pressure. J Hydrometeorol. 2015;16(2):598–614. 10.1175/JHM-D-14-0156.1.Search in Google Scholar

[34] Ionita M. Interannual summer streamflow variability over Romania and its connection to large-scale atmospheric circulation. Int J Climatol. 2015;35(14):4186–96. 10.1002/joc.4278.Search in Google Scholar

[35] Barredo JI. Major flood disasters in Europe: 1950–2005. Nat Hazards. 2007;42:125–48. 10.1007/s11069-006-9065-2.Search in Google Scholar

[36] Barredo JI. Normalised flood losses in Europe: 1970–2006. Nat Hazards Earth Syst Sci. 2009;9(1):97–104. 10.5194/nhess-9-97-2009.Search in Google Scholar

[37] Bissolli P, Friedrich K, Rapp J, Ziese M. Flooding in Eastern Central Europe in May 2010 – Reasons, evolution and climatological assessment. Weather. 2011;66(6):147–53. 10.1002/wea.759.Search in Google Scholar

[38] Kelman I. The autumn 2000 floods in England and flood management. Weather. 2001;56(10):346–60. 10.1002/j.1477-8696.2001.tb06507.x.Search in Google Scholar

[39] Posthumus H, Morris J, Hess TM, Neville D, Phillips E, Baylis A. Impacts of the summer 2007 floods on agriculture in England. J Flood Risk Manag. 2009;2(3):182–9. 10.1111/j.1753-318X.2009.01031.x.Search in Google Scholar

[40] Muchan K, Lewis M, Hannaford J, Parry S. The winter storms of 2013/2014 in the UK: Hydrological responses and impacts. Weather. 2015;70(2):55–61. 10.1002/wea.2469.Search in Google Scholar

[41] Stevens AJ, Clarke D, Nicholls RJ. Trends in reported flooding in the UK: 1884–2013. Hydrol Sci J. 2016;61(1):50–63. 10.1080/02626667.2014.950581.Search in Google Scholar

[42] Stadtherr L, Coumou D, Petoukhov V, Petri S, Rahmstorf S. Record Balkan floods of 2014 linked to planetary wave resonance. Sci Adv. 2016;2(4):e1501428. 10.1126/sciadv.1501428.Search in Google Scholar PubMed PubMed Central

[43] Vidić D, Delić D. Analysis of the climate change in the Doboj municipality and adaptation options. In: Leal Filho W, Trbić G, Filipović D, editors. Climate change adaptation in Eastern Europe. Cham, Switzerland: Springer; 2019. p. 43–59. 10.1007/978-3-030-03383-5_4.Search in Google Scholar

[44] Lapčević Z, Mandić-Rajčević S, Lepić M, Jovanović M. Evaluating a primary healthcare centre's preparedness for disasters using the Hospital Safety Index: Lessons learned from the 2014 floods in Obrenovac, Serbia. Int J Disaster Risk Reduct. 2019;34:436–42. 10.1016/j.ijdrr.2018.12.014.Search in Google Scholar

[45] Andjelkovic A, Djekovic V, Janic M, Spalevic V, Djukanovic G, Nikolic V. Flood on the River Balica at Jagodina, Serbia in 2014. J Environ Prot Ecol. 2020;21(1):308–16.Search in Google Scholar

[46] Tradowsky JS, Philip SY, Kreienkamp F, Kew SF, Lorenz P, Arrighi J, et al. Attribution of the heavy rainfall events leading to severe flooding in Western Europe during July 2021. Clim Change. 2023;176:90. 10.1007/s10584-023-03502-7.Search in Google Scholar

[47] Trbic G, Popov T, Djurdjevic V, Milunovic I, Dejanovic T, Gnjato S, et al. Climate change in Bosnia and Herzegovina according to climate scenario RCP8.5 and possible impact on fruit production. Atmosphere. 2022;13(1):1. 10.3390/atmos13010001.Search in Google Scholar

[48] Popov T, Gnjato S, Trbić G. Analysis of extreme precipitation over the Peripannonian region of Bosnia and Herzegovina. IDŐJÁRÁS. 2018;122(4):433–52. 10.28974/idojaras.2018.4.5.Search in Google Scholar

[49] Oprašić S, Avdić S, Selmanagić Bajrović A, Muharemović A, Prašović S, Trbić G, et al. Third National Communication Bosnia and Herzegovina under UNFCCC. Sarajevo, Bosnia and Herzegovina: UNDP; 2016. https://www.undp.org/sites/g/files/zskgke326/files/migration/ba/TNC-Report-ENG.pdf.Search in Google Scholar

[50] Das S. Flood susceptibility mapping of the Western Ghat coastal belt using multi-source geospatial data and analytical hierarchy process (AHP). Remote Sens Appl Soc Environ. 2020;20:100379. 10.1016/j.rsase.2020.100379.Search in Google Scholar

[51] Mishra K, Sinha R. Flood risk assessment in the Kosi megafan using multi-criteria decision analysis: A hydro-geomorphic approach. Geomorphology. 2020;350:106861. 10.1016/j.geomorph.2019.106861.Search in Google Scholar

[52] Samela C, Coluzzi R, Imbrenda V, Manfreda S, Lanfredi M. Satellite flood detection integrating hydrogeomorphic and spectral indices. GIScience Remote Sens. 2022;59(1):1997–2018. 10.1080/15481603.2022.2143670.Search in Google Scholar

[53] Anusha N, Bharathi B. Flood detection and flood mapping using multi-temporal synthetic aperture radar and optical data. Egypt J Remote Sens Space Sci. 2020;23(2):207–19. 10.1016/j.ejrs.2019.01.001.Search in Google Scholar

[54] Goffi A, Stroppiana D, Brivio PA, Bordogna G, Boschetti M. Towards an automated approach to map flooded areas from Sentinel-2 MSI data and soft integration of water spectral features. Int J Appl Earth Obs Geoinf. 2020;84:101951. 10.1016/j.jag.2019.101951.Search in Google Scholar

[55] Chen H, Ito Y, Sawamukai M, Tokunaga T. Flood hazard assessment in the Kujukuri Plain of Chiba Prefecture, Japan, based on GIS and multicriteria decision analysis. Nat Hazards. 2015;78(1):105–20. 10.1007/s11069-015-1699-5.Search in Google Scholar

[56] Tehrany MS, Pradhan B, Mansor S, Ahmad N. Flood susceptibility assessment using GIS-based support vector machine model with different kernel types. Catena. 2015;125:91–101. 10.1016/j.catena.2014.10.017.Search in Google Scholar

[57] Melkamu T, Bagyaraj M, Adimaw M, Ngusie A, Karuppannan S. Detecting and mapping flood inundation areas in Fogera-Dera floodplain, Ethiopia during an extreme wet season using Sentinel-1 data. Phys Chem Earth A/B/C. 2022;127:103189. 10.1016/j.pce.2022.103189.Search in Google Scholar

[58] Diriba D, Takele T, Karuppannan S, Husein M. Flood hazard analysis and risk assessment using remote sensing, GIS, and AHP techniques: A case study of the Gidabo watershed, Main Ethiopian Rift, Ethiopia. Geomat Nat Hazards Risk. 2024;15(1):236813. 10.1080/19475705.2024.2361813.Search in Google Scholar

[59] Akram A, Tahir A, Alam A. GIS-based flood extent assessment using MODIS satellite remote sensing and spatial analysis. Front Earth Sci. 2024;12:1309629. 10.3389/feart.2024.1309629.Search in Google Scholar

[60] Tanim AH, McRae CB, Tavakol-Davani H, Goharian E. Flood detection in urban areas using satellite imagery and machine learning. Water. 2022;14(7):1140. 10.3390/w14071140.Search in Google Scholar

[61] Wedajo GK, Lemma TD, Fufa T, Gamba P. Integrating satellite images and machine learning for flood prediction and susceptibility mapping for the case of Amibara, Awash Basin, Ethiopia. Remote Sens. 2024;16(12):2163. 10.3390/rs16122163.Search in Google Scholar

[62] Kumar D, Bhattacharyya RK, Shyam S, Ratnayke UR. Indicator based assessment of integrated flood vulnerability index for Brunei Darussalam. Int J Disaster Risk Manag. 2020;2(2):47–70. 10.18485/ijdrm.2020.2.2.4.Search in Google Scholar

[63] Rebouh N, Tout F, Dinar H, Benzid Y, Zouak Z. Integrating multi-source geospatial data and AHP for flood susceptibility mapping in Ain Smara, Constantine, Algeria. Int J Disaster Risk Manag. 2024;6(2):245–64. 10.18485/ijdrm.2024.6.2.16.Search in Google Scholar

[64] Notti D, Giordan D, Caló F, Pepe A, Zucca F, Galve JP. Potential and limitations of open satellite data for flood mapping. Remote Sens. 2018;10(11):1673. 10.3390/rs10111673.Search in Google Scholar

[65] Lahsaini M, Albano F, Albano R, Mazzariello A, Lacava T. A synthetic aperture radar-based robust satellite technique (RST) for timely mapping of floods. Remote Sens. 2024;16(12):2193. 10.3390/rs16122193.Search in Google Scholar

[66] Satriano V, Ciancia E, Pergola N, Tramutoli V. A first extension of the robust satellite technique RST-FLOOD to Sentinel-2 data for the mapping of flooded areas: The case of the Emilia Romagna (Italy) 2023 event. Remote Sens. 2024;16(18):3450. 10.3390/rs16183450.Search in Google Scholar

[67] Breznik J, Oštir K, Ivačič M, Rak G. Water dynamics analysis in karst flood areas using Sentinel-1 time series. Remote Sens. 2023;15(15):3861. 10.3390/rs15153861.Search in Google Scholar

[68] Gašparović M, Klobučar D. Mapping floods in lowland forest using Sentinel-1 and Sentinel-2 data and an object-based approach. Forests. 2021;12(5):553. 10.3390/f12050553.Search in Google Scholar

[69] Lazarević K, Todosijević M, Vulević T, Polovina S, Momirović N, Caković M. Determination of flash flood hazard areas in the Likodra watershed. Water. 2023;15(15):2698. 10.3390/w15152698.Search in Google Scholar

[70] 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. 10.3390/atmos15070774.Search in Google Scholar

[71] Lovrić N, Tošić R, Dragićević S, Novković I. Assessment of torrential flood susceptibility: Case study—Ukrina River Basin (B&H). Bull Serbian Geogr Soc. 2019;99(2):1–16. 10.2298/GSGD1902001L.Search in Google Scholar

[72] Sabljić L, Bajić D. Mapping of flooded areas using remote sensing on the example of Sana River. Herald. 2021;25:109–20. 10.7251/HER2125109S.Search in Google Scholar

[73] Ivanišević M, Savić S, Pavić D, Gnjato S, Popov T. Spatio-temporal patterns of flooded areas in the lower part of the Sana River Basin. Bull Serbian Geogr Soc. 2022;102(2):67–82. 10.2298/GSGD2202067I.Search in Google Scholar

[74] Sabljić L, Pavić D, Savić S, Bajić D. Extreme precipitations and their influence on the river flood hazards: A case study of the Sana River Basin in Bosnia and Herzegovina. Geogr Pannonica. 2023;27(3):184–98. 10.5937/gp27-45600.Search in Google Scholar

[75] Lehner B, Verdin K, Jarvis A. New global hydrography derived from spaceborne elevation data. EOS Trans Am Geophys Union. 2011;89(10):93–4. 10.1029/2008EO100001.Search in Google Scholar

[76] Lehner B, Grill G. Global river hydrography and network routing: Baseline data and new approaches to study the world’s large river systems. Hydrol Process. 2013;27(15):2171–86. 10.1002/hyp.9740.Search in Google Scholar

[77] Tošić R. Analysis of droughty and katey cyclic period in Ukrina River. Her Geogr Soc Repub Srpska. 2003;8:61–74. (In Serbian).Search in Google Scholar

[78] Tošić R. Geomorphological research in the Velika Ukrina. Geogr Pannonica. 2002;6:4–7.10.5937/GeoPan0206004TSearch in Google Scholar

[79] Tošić R, Dragićević S, Zlatić M, Todosijević M, Kostadinov S. The impact of socio-demographic changes on land use and soil erosion (case study: Ukrina river catchment). Geogr Razgledi. 2012;46:69–78.Search in Google Scholar

[80] Sofilj J, Marinković R, Pamić J, Đorđević D. Tumač za list Derventa L 33–120. Sarajevo: RO Geoinženjering; 1985.Search in Google Scholar

[81] “Geoinženjering” Institute of Geology. Basic geological map of Bosnia and Herzegovina, scale 1:100,000. Sheets L33-120 (1964–1984), L33-132 (1963–1974), L34-109 (1962–1976). Sarajevo (Bosnia and Herzegovina): “Geoinženjering” Institute of Geology.Search in Google Scholar

[82] Kottek M, Grieser J, Beck C, Rudolf B, Rubel F. World map of the Köppen-Geiger climate classification updated, Meteorol Z. 2006;15(3):259–63. 10.1127/0941-2948/2006/0130.Search in Google Scholar

[83] Funk CC, Peterson PJ, Landsfeld MF, Pedreros DH, Verdin JP, Rowland JD, et al. A quasi-global precipitation time series for drought monitoring. US Geol Surv Data Ser. 2014;832:4. 10.3133/ds832.Search in Google Scholar

[84] Copernicus Climate Change Service (C3S). ERA5: Fifth generation of ECMWF atmospheric reanalyses of the global climate [Internet]. Copernicus Climate Change Service Climate Data Store (CDS); [cited 2024 Apr 25]. https://cds.climate.copernicus.eu/cdsapp#!/home.Search in Google Scholar

[85] Tošić R. Erosion in the Ukrina River Basin. Banja Luka: Geographical Society of the Republic of Srpska; 2006. Special Editions Book 12. (In Serbian).Search in Google Scholar

[86] Agency of Statistics of Bosnia and Herzegovina. Census of Population, Households and Dwellings in Bosnia and Herzegovina. Sarajevo: Agency of Statistics of Bosnia and Herzegovina; 2019; 2013.Search in Google Scholar

[87] He Q, Yang J, Chen H, Liu J, Ji Q, Wang Y, et al. Evaluation of extreme precipitation based on three long-term gridded products over the Qinghai-Tibet Plateau. Remote Sens. 2021;13(15):30. 10. 10.3390/rs13153010.Search in Google Scholar

[88] Shahid M, Rahman KU, Haider S, Gabriel HF, Khan AJ, Pham QB, et al. Assessing the potential and hydrological usefulness of the CHIRPS precipitation dataset over a complex topography in Pakistan. Hydrol Sci J. 2021;66(11):1664–84. 10.1080/02626667.2021.1957476.Search in Google Scholar

[89] Funk C, Peterson P, Landsfeld M, Pedreros D, Verdin J, Shukla S, et al. The Climate Hazards Infrared Precipitation with Stations – A new environmental record for monitoring extremes. Sci Data. 2015;2:150066. 10.1038/sdata.2015.66.Search in Google Scholar PubMed PubMed Central

[90] Paredes-Trejo FJ, Barbosa H, Kumar TV. Validating CHIRPS-based satellite precipitation estimates in Northeast Brazil. J Arid Environ. 2017;139:26–40. 10.1016/j.jaridenv.2016.12.009.Search in Google Scholar

[91] Katsanos D, Retalis A, Tymvios F, Michaelides S. Analysis of precipitation extremes based on satellite (CHIRPS) and in situ data set over Cyprus. Nat Hazards. 2016;83:53–63. 10.1007/s11069-016-2335-8.Search in Google Scholar

[92] Katsanos D, Retalis A, Michaelides S. Validation of a high-resolution precipitation database (CHIRPS) over Cyprus for a 30-year period. Atmos Res. 2016;169(B):459–64. 10.1016/j.atmosres.2015.05.015.Search in Google Scholar

[93] Rivera JA, Marianetti G, Hinrichs S. Validation of CHIRPS precipitation dataset along the Central Andes of Argentina. Atmos Res. 2018;213:437–49. 10.1016/j.atmosres.2018.06.023.Search in Google Scholar

[94] Mainali J, Pricope NG. High-resolution spatial assessment of population vulnerability to climate change in Nepal. Appl Geogr. 2017;82:66–82. 10.1016/j.apgeog.2017.03.008.Search in Google Scholar

[95] Tuo Y, Duan Z, Disse M, Chiogna G. Evaluation of precipitation input for SWAT modeling in Alpine catchment: A case study in the Adige River Basin (Italy). Sci Total Environ. 2016;573:66–82. 10.1016/j.scitotenv.2016.08.034.Search in Google Scholar PubMed

[96] Toté C, Patricio D, Boogaard H, van der Wijngaart R, Tarnavsky E, Funk C. Evaluation of satellite rainfall estimates for drought and flood monitoring in Mozambique. Remote Sens. 2015;7(2):1758–76. 10.3390/rs70201758.Search in Google Scholar

[97] Guo H, Bao A, Liu T, Ndayisaba F, He D, Kurban A, et al. Meteorological drought analysis in the Lower Mekong Basin using satellite-based long-term CHIRPS product. Sustainability. 2017;9(6):901. 10.3390/su9060901.Search in Google Scholar

[98] Bai L, Shi C, Li L, Yang Y, Wu J. Accuracy of CHIRPS satellite-rainfall products over mainland China. Remote Sens. 2018;10(3):362. 10.3390/rs10030362.Search in Google Scholar

[99] Gao F, Zhang Y, Ren X, Yao Y, Hao Z, Cai W. Evaluation of CHIRPS and its application for drought monitoring over the Haihe River Basin, China. Nat Hazards. 2018;92:155–72. 10.1007/s11069-018-3196-0.Search in Google Scholar

[100] Sabljić L, Lukić T, Bajić D, Marković S, Delić D. Application of remote sensing in monitoring land degradation: A case study of Stanari Municipality (Bosnia and Herzegovina). Open Geosci. 2024;16(1):20220671. 10.1515/geo-2022-0671.Search in Google Scholar

[101] Sabljić L, Lukić T, Marković SB, Bajić D. Potential of remote sensing techniques for integrated spatio-temporal monitoring and analysis of drought in the Sana River Basin, Bosnia and Herzegovina. IDŐJÁRÁS. 2024;128(4):399–423. 10.28974/idojaras.2024.4.1.Search in Google Scholar

[102] Dehaghani AM, Gohario A, Zareian MJ, Haghighi AT. A comprehensive evaluation of the satellite precipitation products across Iran. J Hydrol Reg Stud. 2023;46:101360. 10.1016/j.ejrh.2023.101360.Search in Google Scholar

[103] Lukić T, Micić Ponjiger T, Basarin B, Sakulski D, Gavrilov M, Marković S, et al. Application of Angot Precipitation Index in the assessment of rainfall erosivity: Vojvodina region case study (North Serbia). Acta Geogr Slov. 2021;61(2):123–53. 10.3986/AGS.8754.Search in Google Scholar

[104] McKee TB, Doesken NJ, Kleist J. The relationship of drought frequency and duration to time scales. In: Proceedings of the 8th Conference on Applied Climatology. Vol. 17, Anaheim, CA, USA: 1993. p. 179–83.Search in Google Scholar

[105] Moccia B, Mineo C, Ridolfi E, Russo F, Napolitano F. SPI-based drought classification in Italy: Influence of different probability distribution functions. Water. 2022;14(22):3668. 10.3390/w14223668.Search in Google Scholar

[106] Leščešen I, Gnjato S, Galinović I, Basarin B. Hydrological drought assessment of the Sava River Basin in South-Eastern Europe. J Water Clim Change. 2024;15(8):3902–18. 10.2166/wcc.2024.157.Search in Google Scholar

[107] European Space Agency (ESA). Sentinel-1 [Internet]. https://sentinels.copernicus.eu/web/sentinel/missions/sentinel-1. Accessed November 24, 2024.Search in Google Scholar

[108] Kakoullis D, Fotiou K, Ibarrola Subiza N, Brcic R, Eineder M, Danezis C. An advanced quality assessment and monitoring of ESA Sentinel-1 SAR products via the CyCLOPS infrastructure in the southeastern Mediterranean region. Remote Sens. 2024;16(10):1696. 10.3390/rs16101696.Search in Google Scholar

[109] Mastrorosa S, Crespi M, Congedo L, Munafò M. Land consumption classification using Sentinel-1 data: A systematic review. Land. 2023;12(4):932. 10.3390/land12040932.Search in Google Scholar

[110] Fletcher K; European Space Agency. Sentinel-1: ESA’s radar observatory mission for GMES operational services (ESA SP-1322/1, March 2012). Noordwijk, The Netherlands: ESA Communications; 2012. ISBN: 978-92-9221-418-0.Search in Google Scholar

[111] Gorelick N, Hancher M, Dixon M, Ilyushchenko S, Thau D, Moore R. Google Earth Engine: Planetary-scale geospatial analysis for everyone. Remote Sens Environ. 2017;202:18–27. 10.1016/j.rse.2017.06.031.Search in Google Scholar

[112] Arora M, Sahoo S, Bhatt CM, Litoria PK, Pateriya B. Rapid flood inundation mapping and impact assessment using Sentinel-1 SAR data over Ghaggar River basin of Punjab, India. J Earth Syst Sci. 2023;132(183):7. 10.1007/s12040-023-02199-.Search in Google Scholar

[113] Michaelides S, Levizzani V, Anagnostou E, Bauer P, Kasparis T, Lane JE. Precipitation: Measurement, remote sensing, climatology, and modeling. Atmos Res. 2009;94(4):512–33. 10.1016/j.atmosres.2009.08.017.Search in Google Scholar

[114] Brocca L, Filippucci P, Hahn S, Ciabatta L, Massari C, Camici S, et al. SM2RAIN–ASCAT (2007–2018): Global daily satellite rainfall data from ASCAT soil moisture observations. Earth Syst Sci Data. 2019;11(4):1583–601. 10.5194/essd-11-1583-2019.Search in Google Scholar

[115] Raghavan S. Radar Meteorology. Dordrecht: Springer; 2013. 10.1007/978-94-017-0201-0.Search in Google Scholar

[116] Maggioni V, Sapiano MRP, Adler RF. Estimating uncertainties in high-resolution satellite precipitation products: Systematic or random error? J Hydrometeorol. 2016;17(4):1119–29. 10.1175/JHM-D-15-0094.1.Search in Google Scholar

[117] Loew A, Bell W, Brocca L, Bulgin CE, Burdanowitz J, Calbet X, et al. Validation practices for satellite-based Earth observation data across communities. Rev Geophys. 2017;55(3):779–817. 10.1002/2017RG000562.Search in Google Scholar

[118] Kumar TVL, Barbosa HA, Thakur MK, Paredes-Trejo F. Validation of satellite (TMPA and IMERG) rainfall products with the IMD gridded data sets over monsoon core region of India. In: Rustamov RB, editor. Satellite Information Classification and Interpretation. London, UK: IntechOpen; 2019. 10.5772/intechopen.84999.Search in Google Scholar

[119] Anđelković G, Jovanović S, Manojlović S, Samardžić I, Živković L, Šabić D, et al. Extreme precipitation events in Serbia: Defining the threshold criteria for emergency preparedness. Atmosphere. 2018;9(5):188. 10.3390/atmos9050188.Search in Google Scholar

[120] Nikolić Popadić S. Flood prevention in Serbia and legal challenges in obtaining the land for flood risk management. Environ Sci Policy. 2021;116:213–9. 10.1016/j.envsci.2020.11.007.Search in Google Scholar

[121] Tošić I, Unkašević M, Putniković S. Extreme daily precipitation: The case of Serbia in 2014. Theor Appl Climatol. 2017;128:785–94. 10.1007/s00704-016-1749-2.Search in Google Scholar

[122] Babić Mladenović M, Kolarov V. May 2014 floods in Serbia: Source of flood. Water Res Manag. 2015;5(1):3–8.Search in Google Scholar

[123] Ristić R, Polovina S, Malušević I, Milčanović V. Serbia and torrential floods: Three years after disaster in May 2014. Vodoprivreda. 2017;49:245–52.Search in Google Scholar

[124] Leščešen I, Basarin B, Podraščanin Z, Mesaroš M. Changes in annual and seasonal extreme precipitation over Southeastern Europe. Envirn Sci Proc. 2023;26(1):48. 10.3390/environsciproc2023026048.Search in Google Scholar

[125] Milanović Pešić A, Jakovljević D, Rajčević V, Gnjato S. Assessment of hydroclimatic trends in Southeast Europe – examples from two adjacent countries (Bosnia & Herzegovina and Serbia). Időjárás. 2025;129(1):69–87. 10.28974/idojaras.2025.1.5.Search in Google Scholar

[126] Gnjato S, Leščešen I, Basarin B, Popov T. What is happening with frequency and occurrence of the maximum river discharges in Bosnia and Herzegovina? Acta Geogr Slov. 2024;64(1):129–39. 10.3986/AGS.13461.Search in Google Scholar

[127] Sabljić L, Lukić T, Bajić D, Marković R, Spalević V, Delić D, et al. Optimizing agricultural land use: A GIS-based assessment of suitability in the Sana River Basin, Bosnia and Herzegovina. Open Geosci. 2024;16(1):20220683. 10.1515/geo-2022-0683.Search in Google Scholar

Received: 2025-01-28
Revised: 2025-06-24
Accepted: 2025-07-08
Published Online: 2025-08-04

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Seismic response and damage model analysis of rocky slopes with weak interlayers
  3. Multi-scenario simulation and eco-environmental effect analysis of “Production–Living–Ecological space” based on PLUS model: A case study of Anyang City
  4. Remote sensing estimation of chlorophyll content in rape leaves in Weibei dryland region of China
  5. GIS-based frequency ratio and Shannon entropy modeling for landslide susceptibility mapping: A case study in Kundah Taluk, Nilgiris District, India
  6. Natural gas origin and accumulation of the Changxing–Feixianguan Formation in the Puguang area, China
  7. Spatial variations of shear-wave velocity anomaly derived from Love wave ambient noise seismic tomography along Lembang Fault (West Java, Indonesia)
  8. Evaluation of cumulative rainfall and rainfall event–duration threshold based on triggering and non-triggering rainfalls: Northern Thailand case
  9. Pixel and region-oriented classification of Sentinel-2 imagery to assess LULC dynamics and their climate impact in Nowshera, Pakistan
  10. The use of radar-optical remote sensing data and geographic information system–analytical hierarchy process–multicriteria decision analysis techniques for revealing groundwater recharge prospective zones in arid-semi arid lands
  11. Effect of pore throats on the reservoir quality of tight sandstone: A case study of the Yanchang Formation in the Zhidan area, Ordos Basin
  12. Hydroelectric simulation of the phreatic water response of mining cracked soil based on microbial solidification
  13. Spatial-temporal evolution of habitat quality in tropical monsoon climate region based on “pattern–process–quality” – a case study of Cambodia
  14. Early Permian to Middle Triassic Formation petroleum potentials of Sydney Basin, Australia: A geochemical analysis
  15. Micro-mechanism analysis of Zhongchuan loess liquefaction disaster induced by Jishishan M6.2 earthquake in 2023
  16. Prediction method of S-wave velocities in tight sandstone reservoirs – a case study of CO2 geological storage area in Ordos Basin
  17. Ecological restoration in valley area of semiarid region damaged by shallow buried coal seam mining
  18. Hydrocarbon-generating characteristics of Xujiahe coal-bearing source rocks in the continuous sedimentary environment of the Southwest Sichuan
  19. Hazard analysis of future surface displacements on active faults based on the recurrence interval of strong earthquakes
  20. Structural characterization of the Zalm district, West Saudi Arabia, using aeromagnetic data: An approach for gold mineral exploration
  21. Research on the variation in the Shields curve of silt initiation
  22. Reuse of agricultural drainage water and wastewater for crop irrigation in southeastern Algeria
  23. Assessing the effectiveness of utilizing low-cost inertial measurement unit sensors for producing as-built plans
  24. Analysis of the formation process of a natural fertilizer in the loess area
  25. Machine learning methods for landslide mapping studies: A comparative study of SVM and RF algorithms in the Oued Aoulai watershed (Morocco)
  26. Chemical dissolution and the source of salt efflorescence in weathering of sandstone cultural relics
  27. Molecular simulation of methane adsorption capacity in transitional shale – a case study of Longtan Formation shale in Southern Sichuan Basin, SW China
  28. Evolution characteristics of extreme maximum temperature events in Central China and adaptation strategies under different future warming scenarios
  29. Estimating Bowen ratio in local environment based on satellite imagery
  30. 3D fusion modeling of multi-scale geological structures based on subdivision-NURBS surfaces and stratigraphic sequence formalization
  31. Comparative analysis of machine learning algorithms in Google Earth Engine for urban land use dynamics in rapidly urbanizing South Asian cities
  32. Study on the mechanism of plant root influence on soil properties in expansive soil areas
  33. Simulation of seismic hazard parameters and earthquakes source mechanisms along the Red Sea rift, western Saudi Arabia
  34. Tectonics vs sedimentation in foredeep basins: A tale from the Oligo-Miocene Monte Falterona Formation (Northern Apennines, Italy)
  35. Investigation of landslide areas in Tokat-Almus road between Bakımlı-Almus by the PS-InSAR method (Türkiye)
  36. Predicting coastal variations in non-storm conditions with machine learning
  37. Cross-dimensional adaptivity research on a 3D earth observation data cube model
  38. Geochronology and geochemistry of late Paleozoic volcanic rocks in eastern Inner Mongolia and their geological significance
  39. Spatial and temporal evolution of land use and habitat quality in arid regions – a case of Northwest China
  40. Ground-penetrating radar imaging of subsurface karst features controlling water leakage across Wadi Namar dam, south Riyadh, Saudi Arabia
  41. Rayleigh wave dispersion inversion via modified sine cosine algorithm: Application to Hangzhou, China passive surface wave data
  42. Fractal insights into permeability control by pore structure in tight sandstone reservoirs, Heshui area, Ordos Basin
  43. Debris flow hazard characteristic and mitigation in Yusitong Gully, Hengduan Mountainous Region
  44. Research on community characteristics of vegetation restoration in hilly power engineering based on multi temporal remote sensing technology
  45. Identification of radial drainage networks based on topographic and geometric features
  46. Trace elements and melt inclusion in zircon within the Qunji porphyry Cu deposit: Application to the metallogenic potential of the reduced magma-hydrothermal system
  47. Pore, fracture characteristics and diagenetic evolution of medium-maturity marine shales from the Silurian Longmaxi Formation, NE Sichuan Basin, China
  48. Study of the earthquakes source parameters, site response, and path attenuation using P and S-waves spectral inversion, Aswan region, south Egypt
  49. Source of contamination and assessment of potential health risks of potentially toxic metal(loid)s in agricultural soil from Al Lith, Saudi Arabia
  50. Regional spatiotemporal evolution and influencing factors of rural construction areas in the Nanxi River Basin via GIS
  51. An efficient network for object detection in scale-imbalanced remote sensing images
  52. Effect of microscopic pore–throat structure heterogeneity on waterflooding seepage characteristics of tight sandstone reservoirs
  53. Environmental health risk assessment of Zn, Cd, Pb, Fe, and Co in coastal sediments of the southeastern Gulf of Aqaba
  54. A modified Hoek–Brown model considering softening effects and its applications
  55. Evaluation of engineering properties of soil for sustainable urban development
  56. The spatio-temporal characteristics and influencing factors of sustainable development in China’s provincial areas
  57. Application of a mixed additive and multiplicative random error model to generate DTM products from LiDAR data
  58. Gold vein mineralogy and oxygen isotopes of Wadi Abu Khusheiba, Jordan
  59. Prediction of surface deformation time series in closed mines based on LSTM and optimization algorithms
  60. 2D–3D Geological features collaborative identification of surrounding rock structural planes in hydraulic adit based on OC-AINet
  61. Spatiotemporal patterns and drivers of Chl-a in Chinese lakes between 1986 and 2023
  62. Land use classification through fusion of remote sensing images and multi-source data
  63. Nexus between renewable energy, technological innovation, and carbon dioxide emissions in Saudi Arabia
  64. Analysis of the spillover effects of green organic transformation on sustainable development in ethnic regions’ agriculture and animal husbandry
  65. Factors impacting spatial distribution of black and odorous water bodies in Hebei
  66. Large-scale shaking table tests on the liquefaction and deformation responses of an ultra-deep overburden
  67. Impacts of climate change and sea-level rise on the coastal geological environment of Quang Nam province, Vietnam
  68. Reservoir characterization and exploration potential of shale reservoir near denudation area: A case study of Ordovician–Silurian marine shale, China
  69. Seismic prediction of Permian volcanic rock reservoirs in Southwest Sichuan Basin
  70. Application of CBERS-04 IRS data to land surface temperature inversion: A case study based on Minqin arid area
  71. Geological characteristics and prospecting direction of Sanjiaoding gold mine in Saishiteng area
  72. Research on the deformation prediction model of surrounding rock based on SSA-VMD-GRU
  73. Geochronology, geochemical characteristics, and tectonic significance of the granites, Menghewula, Southern Great Xing’an range
  74. Hazard classification of active faults in Yunnan base on probabilistic seismic hazard assessment
  75. Characteristics analysis of hydrate reservoirs with different geological structures developed by vertical well depressurization
  76. Estimating the travel distance of channelized rock avalanches using genetic programming method
  77. Landscape preferences of hikers in Three Parallel Rivers Region and its adjacent regions by content analysis of user-generated photography
  78. New age constraints of the LGM onset in the Bohemian Forest – Central Europe
  79. Characteristics of geological evolution based on the multifractal singularity theory: A case study of Heyu granite and Mesozoic tectonics
  80. Soil water content and longitudinal microbiota distribution in disturbed areas of tower foundations of power transmission and transformation projects
  81. Oil accumulation process of the Kongdian reservoir in the deep subsag zone of the Cangdong Sag, Bohai Bay Basin, China
  82. Investigation of velocity profile in rock–ice avalanche by particle image velocimetry measurement
  83. Optimizing 3D seismic survey geometries using ray tracing and illumination modeling: A case study from Penobscot field
  84. Sedimentology of the Phra That and Pha Daeng Formations: A preliminary evaluation of geological CO2 storage potential in the Lampang Basin, Thailand
  85. Improved classification algorithm for hyperspectral remote sensing images based on the hybrid spectral network model
  86. Map analysis of soil erodibility rates and gully erosion sites in Anambra State, South Eastern Nigeria
  87. Identification and driving mechanism of land use conflict in China’s South-North transition zone: A case study of Huaihe River Basin
  88. Evaluation of the impact of land-use change on earthquake risk distribution in different periods: An empirical analysis from Sichuan Province
  89. A test site case study on the long-term behavior of geotextile tubes
  90. An experimental investigation into carbon dioxide flooding and rock dissolution in low-permeability reservoirs of the South China Sea
  91. Detection and semi-quantitative analysis of naphthenic acids in coal and gangue from mining areas in China
  92. Comparative effects of olivine and sand on KOH-treated clayey soil
  93. YOLO-MC: An algorithm for early forest fire recognition based on drone image
  94. Earthquake building damage classification based on full suite of Sentinel-1 features
  95. Potential landslide detection and influencing factors analysis in the upper Yellow River based on SBAS-InSAR technology
  96. Assessing green area changes in Najran City, Saudi Arabia (2013–2022) using hybrid deep learning techniques
  97. An advanced approach integrating methods to estimate hydraulic conductivity of different soil types supported by a machine learning model
  98. Hybrid methods for land use and land cover classification using remote sensing and combined spectral feature extraction: A case study of Najran City, KSA
  99. Streamlining digital elevation model construction from historical aerial photographs: The impact of reference elevation data on spatial accuracy
  100. Analysis of urban expansion patterns in the Yangtze River Delta based on the fusion impervious surfaces dataset
  101. A metaverse-based visual analysis approach for 3D reservoir models
  102. Late Quaternary record of 100 ka depositional cycles on the Larache shelf (NW Morocco)
  103. Review Articles
  104. Humic substances influence on the distribution of dissolved iron in seawater: A review of electrochemical methods and other techniques
  105. Applications of physics-informed neural networks in geosciences: From basic seismology to comprehensive environmental studies
  106. Ore-controlling structures of granite-related uranium deposits in South China: A review
  107. Shallow geological structure features in Balikpapan Bay East Kalimantan Province – Indonesia
  108. A review on the tectonic affinity of microcontinents and evolution of the Proto-Tethys Ocean in Northeastern Tibet
  109. Special Issue: Natural Resources and Environmental Risks: Towards a Sustainable Future - Part II
  110. Depopulation in the Visok micro-region: Toward demographic and economic revitalization
  111. Special Issue: Geospatial and Environmental Dynamics - Part II
  112. Advancing urban sustainability: Applying GIS technologies to assess SDG indicators – a case study of Podgorica (Montenegro)
  113. Spatiotemporal and trend analysis of common cancers in men in Central Serbia (1999–2021)
  114. Minerals for the green agenda, implications, stalemates, and alternatives
  115. Spatiotemporal water quality analysis of Vrana Lake, Croatia
  116. Functional transformation of settlements in coal exploitation zones: A case study of the municipality of Stanari in Republic of Srpska (Bosnia and Herzegovina)
  117. Hypertension in AP Vojvodina (Northern Serbia): A spatio-temporal analysis of patients at the Institute for Cardiovascular Diseases of Vojvodina
  118. Regional patterns in cause-specific mortality in Montenegro, 1991–2019
  119. Spatio-temporal analysis of flood events using GIS and remote sensing-based approach in the Ukrina River Basin, Bosnia and Herzegovina
  120. Flash flood susceptibility mapping using LiDAR-Derived DEM and machine learning algorithms: Ljuboviđa case study, Serbia
  121. Geocultural heritage as a basis for geotourism development: Banjska Monastery, Zvečan (Serbia)
  122. Assessment of groundwater potential zones using GIS and AHP techniques – A case study of the zone of influence of Kolubara Mining Basin
  123. Impact of the agri-geographical transformation of rural settlements on the geospatial dynamics of soil erosion intensity in municipalities of Central Serbia
  124. Where faith meets geomorphology: The cultural and religious significance of geodiversity explored through geospatial technologies
  125. Applications of local climate zone classification in European cities: A review of in situ and mobile monitoring methods in urban climate studies
  126. Complex multivariate water quality impact assessment on Krivaja River
Downloaded on 3.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/geo-2025-0856/html?lang=en&srsltid=AfmBOoq9NffqaNsUM0KuzRBtOJ9Q09htws11k432CCL5wcN_WksBfguA
Scroll to top button