Next Article in Journal
Path Modeling between Urban Factors and Happiness of People with Disabilities: Using Partial Least Squares Multi-Group Analysis
Previous Article in Journal
What Makes Farmers Aware in Adopting Circular Bioeconomy Practices? Evidence from a Greek Rural Region
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Flood Susceptibility Modeling Using an Advanced Deep Learning-Based Iterative Classifier Optimizer

by
Md. Uzzal Mia
1,
Tahmida Naher Chowdhury
2,
Rabin Chakrabortty
3,
Subodh Chandra Pal
3,
Mohammad Khalid Al-Sadoon
4,
Romulus Costache
5,6 and
Abu Reza Md. Towfiqul Islam
1,*
1
Department of Disaster Management, Begum Rokeya University, Rangpur 5400, Bangladesh
2
Department of Civil and Environmental Engineering, Shahjalal University of Science and Technology, Sylhet 3114, Bangladesh
3
Department of Geography, The University of Burdwan, Bardhaman 713104, India
4
Department of Zoology, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
5
Department of Civil Engineering, Transilvania University of Brasov, 5 Turnului Street, 500152 Brasov, Romania
6
Danube Delta National Institute for Research and Development, 165 Babadag Street, 820112 Tulcea, Romania
*
Author to whom correspondence should be addressed.
Land 2023, 12(4), 810; https://doi.org/10.3390/land12040810
Submission received: 31 January 2023 / Revised: 24 March 2023 / Accepted: 28 March 2023 / Published: 3 April 2023
(This article belongs to the Section Land Systems and Global Change)

Abstract

:
We developed a novel iterative classifier optimizer (ICO) with alternating decision tree (ADT), naïve Bayes (NB), artificial neural network (ANN), and deep learning neural network (DLNN) ensemble algorithms to build novel ensemble computational models (ADT-ICO, NB-ICO, ANN-ICO, and DLNN-ICO) for flood susceptibility (FS) mapping in the Padma River basin, Bangladesh. The models consist of environmental, topographical, hydrological, and tectonic circumstances, and the final result was chosen based on the causative attributes using multicollinearity analysis. Statistical techniques were utilized to assess the model’s performance. The results revealed that rainfall, elevation, and distance from the river are the most influencing variables for the occurrence of floods in the basin. The ensemble model of DLNN-ICO has optimal predictive performance (AUC = 0.93, and 0.91, sensitivity = 0.93 and 0.92, specificity = 0.90 and 0.80, F score = 0.91 and 0086 in the training and validation stages, respectively) followed by ADT-ICO, NB-ICO, and ANN-ICO, and might be a viable technique for precisely predicting and visualizing flood events.

1. Introduction

Owing to climatic variability, such as a warmer climate and excessive rainfall, flooding occurs regularly in various locations globally [1,2,3,4]. Flood risks increase as the world’s population, industrialization, and agricultural intensification accelerate [5,6]. For example, river flooding poses a substantial risk to human life and is often a direct result of catastrophic events that cause significant fatalities and economic disruption [7]. Flood events have caused more life and property losses in the 21st century than any other natural disaster, according to the Centre for Research on the Epidemiology of Disasters (CRED) [8]. Therefore, flood hazards, susceptibility, and risk evaluation are crucial to minimizing the forthcoming threat [9,10].
Many studies show floods harm over 200 million people annually [11]. Furthermore, under global warming scenarios, changes in land use patterns, and population growth, it is expected that flood occurrence rates and intensity will be worsened by 2050, potentially causing substantial damage (USD 1 trillion) [12,13,14,15,16,17,18]. As a result, research on flood modeling in affected vulnerable regions is urgently required to adopt mitigation and adaptation practices against disastrous floods [19,20,21]. Floods are divided into four categories depending on their incidence: flash flooding, riverine flooding, coastal flooding, and urbanized flooding [21,22]. Among those types of floods, flash floods are the most catastrophic events, causing massive loss of life and property [23]. With complex landscapes and low-lying areas, the flood-affected areas are stimulated for erosion by intense runoff from streams and rivers. Research scholars have focused on devastating flash floods to reduce the effect of such extreme events [24,25,26,27]. There has also been widespread destruction caused by flash floods in both developing and developed nations [28,29,30,31]. In contrast, developing countries are more affected by flash floods than developed nations due to insufficient infrastructure, vital financial resources, and technological innovation to anticipate flood events or reduce flood repercussions. It is, therefore, essential to develop an accurate model for flood susceptibility in a region.
Flood events have been expected and frequent in Bangladesh during the past two decades, due to irregular rainfall patterns and climatic variabilities [32,33]. The western region of Bangladesh is particularly vulnerable to irregular rainfall or flooding caused by thunderstorm bursting, especially in the rainy season (June–September), causing major socioeconomic and human casualties [34,35,36]. Flood susceptibility (FS) refers to regions’ vulnerability to flooding due to regional geology, geography, and hydrometeorological variables. Using flood susceptibility models (FSMs), it is possible to divide an area into flood-prone zones and develop ways to protect against flooding. The FSM can be performed considering diverse tools, including a geospatial model, data mining techniques, deep and machine learning modeling, hydrologic and hydrodynamic modeling, and so on [37,38]. Among these tools, the uses of deep and machine learning models are frequently described in the existing literature for developing flood susceptibility analysis owing to their popularity and worldwide acceptance [39,40,41,42,43,44,45].
On the other hand, many researchers have used various approaches to construct FSMs such as (1) the hierarchical analytic procedure (expert-based model) [46,47,48]; (2) quantitative and multivariate analysis modeling, such as (FR) frequency ratio [2,39,49,50], information value (IV model) [51,52], certainty factors, logistic regressions (LR), [11,53,54,55], weights-of-evidences (WoE model) [56,57], fuzzy logic (FL) [58,59], neurofuzzy logic (ANFIS) [44,60]; (3) machine learning algorithms (MLAs) [41,61,62,63]; (4) hydrological algorithms such as the Soil–Water Assessment Tools (SWAT) [64] and, among others, the River Analysis System–Hydraulic Engineering Center (HEC-RAS) [65]. The popular machine learning algorithms include alternating decision tree (ADT) [66,67]; naïve Bayes (NB) [54,68]; artificial neural networks (ANN) [29,50,69,70], and deep learning neural network (DLNN) [23,71], which can predict flood inundation areas in susceptible regions. Deep learning models were chosen for the FSMs because they can slowly build high-level features from raw datasets. The deep learning algorithms have high learning and generalization flexibility, detecting complex features hidden in the developed information. Thus, deep learning outperforms other machine learning models in FSM analysis [6,72,73,74]. Despite the advantage of the deep learning model for FSMs, it is yet challenging to generate accurate flood susceptibility mapping [71], this because deep learning models are still inadequate to understand the complicated processes and interrelations that control natural hazards, e.g., floods and landslides.
On the other hand, FSMs face various challenges, such as selecting the optimal modeling approaches, and a diverse method produces various results [75]. Furthermore, modeling or predicting FSMs using any tool has potential drawbacks. Consequently, hybrid ensemble machine learning can overcome these limitations, outperforming conventional and individual methods [76]. Recently, scholars have frequently used hybrid ensemble machine learning methods and information retrieval approaches, such as reptree [40,77] and naïve Bayes [55,78,79] for robust FSMs. Researchers use and create hybrid machine learning algorithms for FSMs [40,79] because of how well they work with hybrid machine learning techniques. In recent times, the developed hybrid ensemble learning tools have substantially enhanced the performance of FSMs. However, there is still no consensus concerning the optimal model for FSMs [56]. In general, iterative classifier optimization (ICO) is used to reduce the number of repetitions for cross-validation purpose in FSMs. The optimization tool is a function of optimized models employed to enhance the prediction accuracy of the model [70,71,72,73,74,75,76,77,78].
Based on the aforementioned discussion, the DLNN and ADT models were seldom applied in FSMs; their hybridization coupled with an optimization system is still rare in the past literature. Additionally, the ICO algorithm is used to optimize the training methods for the four machine learning algorithms.
However, the DLNN model has only been used in environmental studies and catastrophe prevention in a few cases. To the authors’ knowledge, there is no documented cited works for the DLNN and ADT models coupled with ICO for FS mapping. To fill the gap in earlier literature, we applied the proposed model in the Padma River basin in Bangladesh as a case study. The primary goals of the current study are (i) to examine and evaluate the FS map prediction abilities of the hybrid ensemble deep learning ADT-ICO, NB-ICO, ANN-ICO, and DLNN-ICO models, and (ii) to evaluate flood-affecting variables in the Padma River basin, Bangladesh, to generate FS mapping. The Padma River basin, which is very prone to frequent floods, was chosen as a research case within that study to generate FS mapping. The originality of this work is that novel ADT-ICO, NB-ICO, ANN-ICO, and DLNN-ICO hybrid ensemble models were developed to map FS. The findings of the study will encourage regional and municipal authorities and legislators to adopt suitable mitigation measures that will lower flood risks and avert possible harm.

2. Materials and Methods

2.1. Study Area

One of Bangladesh’s most important rivers, the Padma, occupies a total area of more than 2562 km2 (Figure 1). The Ganges River begins at the Gangotri Glacier in the Himalayas and has its most crucial downstream stretch here [80]. In Bangladesh, the Ganges’ lowermost reach (known as the Padma) flows southeasterly for approximately 108 km before joining the Meghna near Chandpur [81]. This research was carried out in the Padma River system. Most of the study is focused on the Kushtia, Pabna, and Rajbari tributaries, and a few others: Manikganj, Faridpur, Rajshahi, and Natore. The research region is centered between latitudes 23°48′ and 25°18′ north and longitude 88°27′ and 89°48′ east. The river has a full bank discharge of around 75,000 m3s−1, while the yearly average discharge of the river is 30,000 m3s−1 (Hydraulics D, DHI (FAP 24) 1996) [82]. More than four hundred and seventy million people in China, India, Nepal, Bhutan, and Bangladesh profit from the Ganges River in many ways. The river profoundly affects the social and ecological settings of these countries [80].
Northwestern Bangladeshi riverine fish species are thought to use the Padma River for critical breeding and feeding habitat. It is also a significant commercial fishing area for many people [83]. People’s way of life depends on how well the Padma River basin can support agriculture and other businesses and provide food and aquaculture space. This freshwater delivery system is essential to the sustainability of the riverine ecology in the southwestern part of Bangladesh, by preserving salinity on the upstream side of the Bay of Bengal (BOB) and avoiding its degradation, which is largely made up of the Sundarbans. Bangladesh regularly faces massive floods of unprecedented magnitude due to this basin’s extreme flow regime fluctuations (water and sediment), which are caused by monsoonal rains. Additionally, this basin frequently experiences river bank erosion and channel shifting, which have caused both human displacement and environmental degradation. Climate change is anticipated to worsen this situation by increasing the quantity of water that flows into the basins of the Ganges, Brahmaputra, and Meghna rivers [80].

2.2. Preparation of the Data

Figure 2 displays an overview of the methodological framework followed in the current research and can be presented in the following three steps:
(1)
Flood causal variables were prepared for this study.
(2)
Datasets were demarcated into two components: training and testing, where the training data were then integrated into the deep learning models, and the hyperparameters for each method were tuned properly.
(3)
Deep learning methods were applied to detect how each important conditioning variable is in flood incidence.
Figure 2. Methodology flow chart.
Figure 2. Methodology flow chart.
Land 12 00810 g002

2.2.1. Preparing of Causal Variables

Compiling flood inventory mapping for the area of research is the first step in developing flood vulnerability mapping [23,80,84]. Flooding inventories establish the connection between the flood and the component that generates flooding. To use the datasets described below, we created a flood inventory for the years 1988 through 2021: (i) past flood data from both public and private sources [5,35,85]; (ii) literature review; and (iii) imagery from Landsat 5 and 8, Sentinel-1, and the Moderate Resolution Imaging Spectroradiometer (MODIS) satellites. The flood inventory was prepared to utilize binary classification, with (30 m × 30 m) grid cells having flood assigned 1 and nonflood assigned 0 [35].

2.2.2. Dividing a Database

The resulting flood assessment was sequentially partitioned into two separate sets of data: 70% of the dataset from 350 flood areas was employed to train the algorithms. In contrast, 30% of the dataset from 150 flood-prone areas was utilized to validate the algorithms [86]. We expected to collect sampling points in nonfloodprone areas close to flood zones. Several studies suggested that selecting the same number of nonflood data sets as positive or inundate values [79] depended on topographic mapping, past flooding information, and a field investigation [85]. Moreover, a similar number of nonflood sites inside the research zone (70%) and validation sites (30%) were selected [35].

2.2.3. Flood Conditional Variable Preparation

The 18 relevant attributes were assessed based on previous research [39,85,87,88] and the watershed features of the research location: aspect, elevation (m), slope angle (degree), curvature, plan curvature, profile curvature, flow direction, flow accumulation, land use and land cover (LULC), normalized difference vegetation index (NDVI), distance to the rivers (m), soil types, mean annual rainfall (mm), river density (RD), stream power index (SPI), topographic wetness index (TWI), sediment transport index (STI), and geology. Aspect usually defines the horizontal direction that an elevation’s slope is facing. Figure 3 shows details on the flood-causing agents. Elevation, slope angle, aspect, profile curvature, plan curvature, SPI, TWI, STI, river density, distance to the river, flow accumulation, and flow direction datasets were retrieved from an advanced land-observing satellite (ALOS)-based digital elevation model (DEM) with a resolution of 30 m. Based on past literature, the ALOS-based DEM was the most accurate way to develop the raster layers of these flood conditioning variables [4]. Based on data from [89] and [90], a map illustrating mean annual rainfall from 2000 to 2021 was created in a GIS environment by using kriging spatial interpolation tools (https://chrsdata.eng.uci.edu/) [89]. When the geospatial coverage of the sampling areas was inadequate, kriging interpolation was used for this purpose [91]. A map of LULC and NDVI was acquired from https://earthexplorer.usgs.gov/ [92]. Geological data (scale 1:100,000) was obtained from the United States Geological Survey (USGS) website [93]. The soil-texture-types map (scale of 1:5,000,000) was acquired from the Food and Agriculture Organization (FAO) of the United Nations (UN) [94]. It is worth mentioning that decreasing the uncertainties associated with different spatial resolutions is difficult. A constant grid size of 30 m was adopted to provide meaningful findings, so all causal factors were examined within this approach [95].

2.3. Flood Susceptibility Assessment

Numerous scholars have employed a variety of methodologies to obtain accurate results in FS area assessment. At present, deep learning models have become vital for producing improved results. In this study, four important deep learning models, DLNN, ANN, ADT, and NB, were used to determine the specific flood-prone location with the help of the R package console platform (R-4.2.2).

2.3.1. Alternating Decision Tree (ADT)

Freund and Mason (1999) introduced the alternating decision tree (ADT) as a reliable machine learning classification approach based on boosting [49,66]. It integrates decision trees and boosting algorithms and is frequently used to solve classification and prediction issues [67,96]. ADT structure is more straightforward than decision tree models such as random forest, rotation forest, and classification and regression tree [97]. The decision and prediction nodes are two types of nodes included in ADT algorithms [66]. Prediction nodes represent a single value, while decision nodes define a specific circumstance [98]. ADT was created using a boosting strategy for numerical forecasting, in which a decision node and its two prediction nodes are formed at each iteration of the boosting process. Each prediction network node is given a weight corresponding to how much the node contributed to the overall prediction score. The ultimate prediction probability is obtained by adding all of the contributing weights [67].
If C1 means a precondition, C2 represents a base condition, and a and b represent two real numbers, then a and b are calculated with Equation (1):
a = 0.5 * l n W + ( C 1 C 2 ) W ( C 1 C 2 ) ,       b = 0.5 * l n W + ( C 1 C ¯ 2 ) W ( C 1 C ¯ 2 )
where W is equal to the sum of the values provided by the prediction node, and the optimal C 1 and C 2 are calculated by minimizing Zt ( C 1 , C 2 ), obtained with Equation (2):
Z t ( C 1 ,   C 2 ) = 2 W + ( C 1 C 2 ) * W ( C 1 C 2 ) + W + ( C 1 C ¯ 2 ) * W ( C 1 C ¯ 2 )
The flood-class labels are those that have the best constant prediction and consequently cross-validation [96,99].

2.3.2. Naïve Bayes (NB)

Of the classical machine learning algorithms, one of the few that is based on probability theory is naïve Bayes. Its performance comparable to that of neural networks and learning based on decision trees in some domains [100]. It is a special instance of Bayesian networks in which one node is an attribute node, and the others are feature nodes that are assumed to be independent of one another. NB is a popular classification algorithm based on a basic probability theorem known as Bayes’ theorem, Bayes’ rule, or Bayes’ formula [101].
In the context of evaluating flood susceptibility, the posterior probability of flood occurrence for each cell (x, y) is the conditional probability of variables (Equation (3)):
P ( D = d j | C i ( x , y ) ) = P ( D = D j ) i = 1 v P ( C i ( x , y ) | d j )
where P is the posterior probability; D = {d1 = flooding event, d2 = not flooding event}; j = {1, 2}; v is the number of factors, and Ci is the conditional probability of the ith factor, 1 ≤ i ≤ 18. After determining the posterior probability, the probability of a flood occurring in each cell (x, y) is calculated using Equation (4):
S u s c e p t i b i l i t y = P ( D = d 1 | E i ( x , y ) ) P ( D = d 1 | E i ( x , y ) ) + P ( D = d 2 | E i ( x , y ) )
where susceptibility denotes a cell’s likelihood of flooding, and P (D = d1|Ei(x, y)) and P(D = d2|Ei(x, y)) denote the cell’s posterior probabilities of flooding and nonflooding events, respectively [79].

2.3.3. Artificial Neural Network (ANN)

An artificial neural network (ANN) is a mathematical model of human perception that may be taught to do a specific task using a dataset, particularly to investigate the link between inputs and outputs [87]. Situations in which the unknown relationships between the data can serve as an effective modeling tool [102]. The model identification and classification fields have made substantial use of ANN [37]. ANN generally consists of three interconnected layers: the input, the hidden, and the output layers. The input layer receives data from various sources. Trial and error determine the number of hidden layers and their neurons. The application determines the number of neurons in the output layers, represented by the processed class [87].
The first step in ANN computation is to add a set of numbers, called xi, to the processing node’s input layer. Through connection-specific weights, wt, these signals can be amplified as they travel along connections to each of the nodes in the adjacent layer. The layer nodes next to them serve as summing elements for the incoming signals. The processing units then apply a threshold function to the input signal to convert it into an output signal (Qj), Equation (5):
f ( x ) = 1 1 + e x
In Equation (5), the output, f(x), ranges from 0 to 1. The processing unit’s output is then determined as follows:
Q j = 1 1 + e x i w i
The output signal Qj is transmitted across the weighted connections to the subsequent set of nodes. To ensure the signal reaches the final output layer, the process is repeated. The output signal is then interpreted as the ANN’s response to the given input stimulus [69].

2.3.4. Deep Learning Neural Network (DLNN)

In general, DLNN is classified as an ANN algorithm, but with numerous (deep) hidden layers (depending on the complexity of the features) and a feed-forward network for the back-propagation training technique. Flooding is a process that may present nonlinear relationships with other hydrological variables [103]. Since it yields precise results, it quickly gained popularity in research on susceptibility assessment to natural hazards [104]. Three layers comprise the DLNN: an input layer, several hidden layers, and an output layer. The DLNN algorithm’s basic configuration is to function in such a way that the input layer integrates data from several flood causal factors. These data are saved and evaluated in numerous hidden layers, and the output model effects are ultimately displayed in the last layer, known as the output layer. There are two possible labels for the activation function: a negative (nonflood) and a positive sign (flood). The output layer displays the categorized outcomes obtained from the final hidden layer [37].

2.4. Iterative Classifier Optimizer (ICO)

An iterative classifier optimizer (ICO) determines the best number of iterations for a specific classifier using cross-validation [105]. This approach effectively manages missing, nominal, and binary classes, including nominal, numeric, binary, and empty nominal properties [97]. This algorithm’s initialization consists of two distinct steps: Firstly, following a general model execution, the results are compared with actual measured values. Secondly, the feedback is fed back to the model to continue learning and improving the outcomes [106]. The ICO algorithm builds the model by comparing the measured and observed values and then optimizes it by adding data from the measured and observed values to the model to change the output. This study used ICO algorithms to improve how the four machine learning algorithms mentioned above are trained. For example, in the ANN model, the number of hidden neurons in the hidden layer and the weights across various layers of neurons are optimized using ICO hyperparameters to obtain the lowest RMSE value. The ICO method also assists in optimizing the total number of hidden layers and the number of hidden neurons inside each hidden layer in the DLNN. Hence, the loss and accuracy numbers are modified, and the DNNN architecture is improved. The ICO method was used in the cases of the ADT and NB models to optimize the hyperparameter represented by the maximum depth of the decision tree, allowing the model to achieve the highest level of accuracy. The maximum depth hyperparameter can influence the overall tree complexity.

2.5. Multicollinearity Assessment

The tolerance (TOL) and the variance inflation factor (VIF) were used to evaluate the multicollinearity of all causal variables, since linear collinearity lowers the performance of predictive methods [4,5,6]. In fact, if the TOL or VIF is higher than 5.0 and lower than 0.1, this indicates multicollinearity in the conditioning variable [21].
Equations (7) and (8) are used to compute the VIF [107]:
T O L = 1 R j 2
V I F = 1 T O L
where R j 2 represents the regression value of j on other different variables in a dataset.
If VIF > 10 or TOL < 0.1, the respective predictors are identical and must be eliminated from the prediction models [21].

2.6. Validation Methods for the Models

The predictive performance of the aforementioned models was demonstrated in this research by utilizing five statistical performance metrics: specificity, sensitivity, positive predictive value (PPV), negative predictive value (NPV), and receiver operating characteristics curve (ROC)-AUC analysis [108]. The number of pixels required to identify floodprone and nonfloodprone areas accurately. The validation approaches were calculated using the four statistical metrics true–positive (TP), true–negative (TN), false–negative (FN), and false–positive (FP). The accuracy of the above models depends on the results of different validation methods. When the numbers were higher, the models worked better, and vice versa [109]. Equations (9)–(12) served as the basis for the statistical methods employed in this investigation.
S p e c i f i c i t y = T N F P + T N
S e n s i t i v i t y = T P T P + F N
P P V = T P F P + T P
N P V = T N T N + F N
In this work, ROC-AUC analysis was also performed to evaluate the model, which is a well-established approach. ROC analysis was completed by visualizing sensitivity and specificity on the X and Y axes. It assists in assessing the prediction power of models, which is a well-established approach for this sort of assessment. The AUC values represent the model’s ability to distinguish properly between positive (flooded) and negative (nonflooded) events in the validation sample. Two axes represent the true and false positive rates on the ROC graph. A higher AUC value indicates better goodness of fit for the model. The range 0.5 to 1 shows the models’ mediocre to outstanding performance.

2.7. Graphical Representation

In the current study, a Taylor diagram is also used for model assessment and comparison. The Taylor diagram illustrates a graphical model assessment that simultaneously takes into account RMSE, r, and standard deviation to measure how close the projected models are to the associated observation data.

3. Results

3.1. Multicollinearity Assessment

The multicollinearity test was carried out to identify multicollinearity issues in all causal factors whereas tolerance and VIF were taken into consideration. Table 1 depicts that TWI has the highest tolerance (1.245) and lowest VIF (0.803), while river density has the lowest TOL (0.305) and highest VIF (3.367), implying that all values fall within 0.1 to 5. Thus, all causal variables reveal no multicollinearity issue, and it is free of multicollinearity and is fit for mapping the FS [18].

3.2. Floods Susceptibility (FS) Assessments

Mapping of FS in this basin is evaluated by utilizing the four ensemble learning algorithms (ADT, ANN, NB, and DLNN) with an ICO optimizer. Figure 4 depicts the resulting FS maps from the models, divided into five classes of susceptibility, and lists the areas and percentages of each type. The deep learning neural network–iterative classifier optimizer (DLNN-ICO) highlights highly vulnerable areas, whereas an artificial neural network (ANN-ICO) highlights less susceptibility. In contrast, ANN-ICO found very low susceptibility compared to DLNN-ICO. The areal coverages of very high and very low FS areas within the alternating decision tree–iterative classifier optimizer (ADT-ICO) model were 786.38 and 881.17 km2, respectively. The remainder of this watershed was connected to low, moderate, and high susceptibility zones, with the areal coverage of these zones being 389.25 (12.36%), 388.31 (12.33%), and 704.18 km2 (22.36%), respectively (Figure 4a). Although the whole region shows moderate to high susceptibility, the southwestern section of the watersheds is more susceptible. Naïve Bayes with the iterative classifier optimizer (NB-ICO) estimated 842.12 and 346.11 km2 as very high and high flood-prone zones, accounting for 26.74% and 10.99% of the total flood-prone area (Figure 4b). In the NB-ICO-based model, the very low class covers the highest area (987.93 km2 shown in dark green). By the ANN-ICO (Figure 4c) machine learning technique, the very low flood susceptibility class occupied the largest percentage (36.01) area. About 1134.06 km2 of the general areas have a very low flooding risk. Within this model, the low, moderately high, and very high classes each comprise 12.34%, 8.35%, 22.19%, and 21.31% of the area remaining. Among the four maps of susceptibility, the DLNN-ICO (Figure 4d) combination indicated a maximum area of 1142.25 km2 of very high flood susceptibility and 379.49 km2 of very low susceptibility. According to the DLNN-ICO map (Figure 4d), floods were more likely to be extremely severe in the southwest part of the basin. Besides that, the remaining areas are linked to low, moderate, and high flood-prone zones, with an aerial coverage of 294.77, 13.03, and 292.43 km2, respectively (Figure 5).

3.3. Validation

The FS maps were evaluated using the models through training and validation data sets based on AUC-ROC curve analysis. For the training datasets (Table 2), the AUC values for ADT-ICO, NB-ICO, ANN-ICO, and DLNN-ICO are 0.891, 0.873, 0.835, and 0.932, respectively (Figure 6a). For the validation datasets (Table 2), they are 0.864, 0.845, 0.801, and 0.917, respectively (Figure 6b). The model DLNN-ICO (AUC = 0.932) performs the best on training datasets, followed by ADT-ICO (AUC = 0.891), NB-ICO (AUC = 0.873), and ANN-ICO (AUC = 0.835) (Figure 6a). Similarly, DLNN-ICO has a high prediction accuracy (AUC = 0.917) in the validation datasets.
The training dataset’s sensitivity, specificity, recall, F score, PPV, and NPV value in the ADT-ICO model is 0.88, 0.87, 0.88, 0.87, 0.87, and 0.88, respectively (Figure 7a). In addition, the sensitivity, specificity, recall, F score, PPV, and NPV values of this model’s validation datasets (Figure 7b) are 0.86, 0.87, 0.86, 0.86, 0.87, and 0.85, respectively. The NB-ICO model training and validation datasets of sensitivity, specificity, recall, F score, PPV, and NPV values are (0.87, 0.86, 0.87, 0.86, 0.85, and 0.87) and (0.84, 0.84, 0.84, 0.84, 0.84, and 0.84) (Table 2). However, the validation datasets of the six parameters are 0.87, 0.88, 0.87, 0.87, 0.88, and 0.87, respectively, for DLNN-ICO.
The Taylor diagram helps to compare the four models’ prediction abilities and identifies the most suitable one [99]. In terms of ideal capacity, DLNN-ICO performed best, followed by ADD-ICO, NB-ICO, and ANN-ICO (Figure 8). Convergence of the objective cost function in the ICO-DLNN model is shown in Figure 9. It clearly depicts the optimal nature of the predictive model in FS assessment. It determines and expresses as a single real number the difference between the projected value and expected value. The results of the optimal model (ICO-DLNN) show that costs decrease as the number of iterations increases (Figure 9).

3.4. Importance of the Variables

Table 3 shows that the four variables, i.e., elevation, distance from the river, rainfall, and river density, with high correlation had the most effect on the FSMS. On the other hand, other variable combinations with moderate to low correlations had little effect on the FLMs.

4. Discussion

Our research aimed to suggest hybrid ensemble models (DLNN-ICO, ANN-ICO, NB-ICO, and ADT-ICO) for flash FSMs in the study basin. Many investigations have used spatial ensemble models [35,82,96,110,111,112] for FS mapping. Accurately estimating FS zones is a great challenge for developing realistic and efficient mitigating strategies [39]. Therefore, the FSM has become the global management basis [4,39,61]. As a result, experts are constantly looking for innovative and dependable strategies to produce detailed and accurate findings that can be used to propose flood control measures [4,23,48,61,113]. Various statistical and machine learning tools were used to model FS. However, there is a constraint and a disadvantage to the findings of the analysis of FS using a stand-alone approach [19]. Owing to inadequate datasets, stand-alone models frequently fail to identify the best-fit function in the hypothesis space or the actual probability of the subset [100]. As a result, hybrid or ensemble modeling is used to predict the FS region accurately [3].
This method’s outcomes may vary depending on input environmental variables. The approach is sensitive to this research area. To understand the model’s sensitivity, this approach may be reproduced elsewhere. Input factors may affect results. The OneR method was employed to find that land use, geology, and slope affect flood events in Vietnam [113]. Khosravi et al. [66] identified slope as being the most essential factor in flash-flood mapping in Iran. Distance from the river and rainfall are the two most important factors in flooding in this study’s four models. High-intensity rainfall over a short duration and distance create a large volume of water in the river channels, moving from greater elevations to the basin’s lower flat region. Low height increases flood incidence. Except for river distance, rainfall, elevation, river density, and slope were major flood risk determinants in the study basin. These results agree with earlier works [35,85].
The iterative classifier optimizer (ICO) assisted the four machine learning models to improve their predictive power and accuracy. ADT, NB, ANN, and DLNN have been extensively used for landslide hazard modeling [23,78,108,114,115,116,117,118], flood hazard modeling [3,66,68,85,97,119,120,121], gully erosion [22,112,122,123,124,125,126], and wildfire models [119,127,128,129]. However, little research has used ADT-ICO, NB-ICO, ANN-ICO, and DLNN-ICO to forecast potential calamities such as FS. The results revealed that the ensemble DLNN-ICO modeling outperforms the other three models (AUC for training dataset: 0.932, and AUC for validation datasets: 0.917) [37,112,121]. Thus, DLNN-ICO performed better than the other algorithms because it is a common objective cost approach that has been successfully used to predict natural disasters and other environmental factors [37]. This is supported by the fact that DLNN-ICO requires less computation time and fewer errors. Research described in [130] followed the traditional ANN paradigm in needing more memory, a larger dataset, and a longer time for calculation. Additionally, typical statistical approaches are inappropriate due to lengthy computations, vast datasets, and more input variables [108]. The DLNN-ICO method demonstrated greater accuracy than other models. It is important to note that there is no prior research on DLNN-ICO evaluating FS. It is worth mentioning that the proposed model worked effectively in Bangladesh’s Padma River basin, allowing it to make excellent predictions about the likelihood of FS. Other research scholars have used DLNN models and other methods to forecast landslides and have obtained highly accurate results [131].
These findings confirmed earlier FS studies on a cross-section of the study area. For instance, the authors of [5] studied FS for the northeast Bangladesh using ANN. They found that the Sumanganj areas are more susceptible to flooding. Mia et al. [4] identified sensitive areas in the Brahmaputra River system using an ensemble of CNN algorithms. Our research also identified such areas close to the Padma River. Even though the Padma floodplain is more at risk, no research on western regions was available. Our precise results will help local authorities develop important mitigation methods for those living on the Padma floodplain, which is threatened by floods. In past cited works, there was no discussion on an ensemble of DLNN-ICO in FS mapping. Thus, we could not make a direct comparison of our model with the past cited works. As a consequence, DLNN has higher accuracy than previous ML techniques. Using a deep learning technique has several benefits, one of which is its independence in doing feature engineering [37]. In this method, an algorithm searches the data to find traits that correlate and then combines them to encourage quicker learning without being specifically instructed to do so. In contrast, ICO’s approach has the clear benefit of allowing fewer faults in the final model since iteration allows the model to self-correct whenever there is a mistake. Another crucial factor is that the supervisor has control over the model’s modifications since the error feedback used to train the model is performed externally [99]. This has the drawback that the model does not automatically learn to remove inaccuracies. However, DLNN outperformed other machine learning algorithms, i.e., NB, ANN, ADT, etc., as a single model [4]. Because of its better performance, decision makers and water managers can apply DLNN-ICO to give a precise outcome for local communities in similar environments.
Owing to a lack of hydrodynamic data on floods, this study has some limitations. First, the inconsistent spatial resolution of DEM-derived causative variables may cause issues with the proposed method. Second, we solely studied fluvial flash floods. In the Padma River basin, random nonflood locations were selected, which may have some bias [66]. Future research should use field survey-based nonflood sites to improve input datasets. The study’s results will assist flood hazard managers and researchers in evaluating FSMs and mitigating flood impacts. The proposed robust tool could effectively assess FS in various basins when incorporating geoenvironmental elements and model input variables with hydrodynamic information. The proposed methodology is suggested for evaluating flash-flood risk in a similar basin. Finally, the geographical investigation of the generated factors (such as DEM, land cover, geology, soil, rain, and other indicators) was not always consistent. However, that is the typical scenario and another limitation of this research, which is related to uncertainties that are difficult to precisely quantify.

5. Conclusions

Accurate and reliable FSMs are beneficial for reducing the effects of floods. Using tolerance and VIF approaches, the result was evaluated for sensitivity, specificity, NPV, and PPV. Identifying the floodprone region of the Padma River basin, Bangladesh, we presented the coupled ensemble algorithms (ADT-ICO, NB-ICO, ANN-ICO, and DLNN-ICO). In this respect, 18 environmental and topographical flood conditioning variables were determined, and the models were run employing a realistic and empirical technique. The ICO-based component significance analyses were used to select and rank the 18 flood conditioning variables for geospatial simulation. The DLNN-ICO model’s findings showed that river density (50), elevation (65), rainfall (70), and distance from the river (80) were the key variables affecting flood occurrence in the study basin. The DLNN-ICO model outscored the other three methods tested in this research. This model’s high and very high flash-flood vulnerable areas cover 922 km2 (29.29%) and 1142 km2 (36.27%) of the total area, respectively. The remaining area is covered by moderate, low, and very low floodprone zones, which were mostly located in lower coarse river basins around the northwest and northern regions. The findings of this study demonstrate that the flash-flood-sensitive zone may be identified in a similar climate region using the DLNN-ICO modeling technique. The suggested models are robust overall for analyzing FS. ICO methods and multiverse optimization could improve the single DNNN model for FSMs. However, these methods can also solve different hydrological problems, which could be considered in future research. The findings of this study can help policymakers at the regional and national levels implement a viable approach to reducing flood risk, thereby limiting financial damage and loss of life in the studied region, particularly for people who live near the river system. To improve the suggested model, more research should be performed on how ensemble models that use deep learning and hydrodynamics can take uncertainty into account.

Author Contributions

Conceptualization, M.U.M. and A.R.M.T.I.; methodology, S.C.P.; software, R.C. (Rabin Chakraborty); validation, T.N.C.; formal analysis, M.U.M.; investigation, A.R.M.T.I.; resources, M.K.A.-S.; data curation, M.U.M.; writing—original draft preparation, R.C. (Romulus Costache) and A.R.M.T.I.; writing—review and editing, R.C. (Rabin Chakraborty), S.C.P., and R.C. (Romulus Costache); visualization, A.R.M.T.I.; supervision, M.K.A.-S.; project administration, M.K.A.-S.; funding acquisition, M.K.A.-S. All authors have read and agreed to the published version of the manuscript.

Funding

Supported by Researchers Supporting Project Number (RSP2023R410), King Saud University, Riyadh, Saudi Arabia.

Data Availability Statement

Data are available based on reasonable request on corresponding author.

Acknowledgments

The authors would like to extend their sincere appreciation to the Researchers Supporting Project Number (RSP2023R410), King Saud University, Riyadh, Saudi Arabia. We greatly acknowledge Begum Rokeya University, Rangpur, for all of its support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hirabayashi, Y.; Mahendran, R.; Koirala, S.; Konoshima, L.; Yamazaki, D.; Watanabe, S.; Kim, H.; Kanae, S. Global flood risk under climate change. Nat. Clim. Change 2013, 3, 816–821. [Google Scholar] [CrossRef]
  2. Khosravi, K.; Melesse, A.M.; Shahabi, H.; Shirzadi, A.; Chapi, K.; Hong, H. Flood susceptibility mapping at Ningdu catchment, China using bivariate and data mining techniques. In Extreme Hydrology and Climate Variability; Elsevier: Amsterdam, The Netherlands, 2019; pp. 419–434. [Google Scholar]
  3. Ruidas, D.; Chakrabortty, R.; Islam, A.R.M.T.; Saha, A.; Pal, S.C. A novel hybrid of meta-optimization approach for flash flood-susceptibility assessment in a monsoon-dominated watershed, Eastern India. Environ. Earth Sci. 2022, 81, 145. [Google Scholar] [CrossRef]
  4. Mia, M.U.; Rahman, M.; Elbeltagi, A.; Abdullah-Al-Mahbub, M.; Sharma, G.; Islam, H.T.; Pal, S.C.; Costache, R.; Islam, A.R.M.T.; Islam, M.M.; et al. Sustainable flood risk assessment using deep learning-based algorithms with a blockchain technology. Geocarto Int. 2022, 1–31. [Google Scholar] [CrossRef]
  5. Rahman, M.; Chen, N.; Islam, M.M.; Dewan, A.; Pourghasemi, H.R.; Washakh, R.M.A.; Nepal, N.; Tian, S.; Faiz, H.; Alam, M. Location-allocation modeling for emergency evacuation planning with GIS and remote sensing: A case study of Northeast Bangladesh. Geosci. Front. 2020, 12, 101095. [Google Scholar] [CrossRef]
  6. Zhang, Q.; Gu, X.; Singh, V.P.; Shi, P.; Sun, P. More frequent flooding? Changes in flood frequency in the Pearl River basin, China, since 1951 and over the past 1000 years. Hydrol. Earth Syst. Sci. 2018, 22, 2637–2653. [Google Scholar] [CrossRef] [Green Version]
  7. Dottori, F.; Salamon, P.; Bianchi, A.; Alfieri, L.; Hirpa, F.A.; Feyen, L. Development and evaluation of a framework for global flood hazard mapping. Adv. Water Resour. 2016, 94, 87–102. [Google Scholar] [CrossRef]
  8. Hoque, M.A.A.; Tasfia, S.; Ahmed, N.; Pradhan, B. Assessing spatial flood vulnerability at KalaparaUpazila in Bangladesh using an analytic hierarchy process. Sensors 2019, 19, 1302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Fernandez, P.; Mourato, S.; Moreira, M. Social vulnerability assessment of flood risk using GIS-based multicriteria decision analysis. A case study of Vila Nova de Gaia (Portugal). Geomat. Nat. Hazards Risk 2016, 7, 1367–1389. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, Y.; Li, Z.; Tang, Z.; Zeng, G. A GIS-based spatial multi-criteria approach for flood risk assessment in the Dongting Lake Region, Hunan, Central China. Water Resour. Manag. 2011, 25, 3465–3484. [Google Scholar] [CrossRef]
  11. Tien Bui, D.; Khosravi, K.; Shahabi, H.; Daggupati, P.; Adamowski, J.F.; Melesse, A.M.; Thai Pham, B.; Pourghasemi, H.R.; Mahmoudi, M.; Bahrami, S.; et al. Flood spatial modeling in northern Iran using remote sensing and gis: A comparison between evidential belief functions and its ensemble with a multivariate logistic regression model. Remote Sens. 2019, 11, 1589. [Google Scholar] [CrossRef] [Green Version]
  12. Bubeck, P.; Thieken, A.H. What helps people recover from floods? Insights from a survey among flood-affected residents in Germany. Reg. Environ. Change 2018, 18, 287–296. [Google Scholar] [CrossRef]
  13. Ali, R.; Kuriqi, A.; Abubaker, S.; Kisi, O. Long-term trends and seasonality detection of the observed flow in Yangtze River using Mann-Kendall and Sen’s innovative trend method. Water 2019, 11, 1855. [Google Scholar] [CrossRef] [Green Version]
  14. Alexander, K.; Hettiarachchi, S.; Ou, Y.; Sharma, A. Can integrated green spaces and storage facilities absorb the increased risk of flooding due to climate change in developed urban environments. J. Hydrol. 2019, 579, 124201. [Google Scholar] [CrossRef]
  15. Huang, K.; Li, X.; Liu, X.; Seto, K.C. Projecting global urban land expansion and heat island intensification through 2050. Environ. Res. Lett. 2019, 14, 114037. [Google Scholar] [CrossRef] [Green Version]
  16. Jahandideh-Tehrani, M.; Zhang, H.; Helfer, F.; Yu, Y. Review of climate change impacts on predicted river streamflow in tropical rivers. Environ. Monit. Assess. 2019, 191, 752. [Google Scholar] [CrossRef]
  17. Xu, L.; Wang, X.; Liu, J.; He, Y.; Tang, J.; Nguyen, M.; Cui, S. Identifying the trade-offs between climate change mitigation and adaptation in urban land use planning: An empirical study in a coastal city. Environ. Int. 2019, 133, 105162. [Google Scholar] [CrossRef]
  18. Sarker, M.N.I.; Alam, G.M.M.; Firdaus, R.B.R.; Biswas, J.C.; Islam, A.R.M.T.; Raihan, M.L.; Hattori, T.; Alam, K.; Joshi, N.P.; Shaw, R. Assessment of flood vulnerability of riverine island community using a composite flood vulnerability index. Int. J. Disaster Risk Reduct. 2022, 82, 103306. [Google Scholar] [CrossRef]
  19. Cao, Q.; Mehran, A.; Ralph, F.M.; Lettenmaier, D.P. The role of hydrological initial conditions on Atmospheric River floods in the Russian River basin. J. Hydrometeorol. 2019, 20, 1667–1686. [Google Scholar] [CrossRef]
  20. Su, J.; Lü, H.; Zhu, Y.; Cui, Y.; Wang, X. Evaluating the hydrological utility of latest IMERG products over the Upper Huaihe River Basin, China. Atmos. Res. 2019, 225, 17–29. [Google Scholar] [CrossRef]
  21. Costache, R.; Hong, H.; Pham, Q.B. Comparative assessment of the flash-flood potential within small mountain catchments using bivariate statistics and their novel hybrid integration with machine learning models. Sci. Total Environ. 2020, 711, 134514. [Google Scholar] [CrossRef]
  22. Tien Bui, D.; Shirzadi, A.; Shahabi, H.; Chapi, K.; Omidavr, E.; Pham, B.T.; Talebpour Asl, D.; Khaledian, H.; Pradhan, B.; Panahi, M.; et al. A novel ensemble artificial intelligence approach for gully erosion mapping in a semi-arid watershed (Iran). Sensors 2019, 19, 2444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Bui, D.T.; Hoang, N.D.; Martínez-Álvarez, F.; Ngo, P.T.T.; Hoa, P.V.; Pham, T.D.; Samui, P.; Costache, R. A novel deep learning neural network approach for predicting flash flood susceptibility: A case study at a high frequency tropical storm area. Sci. Total Environ. 2020, 701, 134413. [Google Scholar]
  24. Siegel, F.R. Structures That Protect Coastal Populations, Assets, and GDPs: Sea Dikes, Breakwaters, Seawalls. In Adaptations of Coastal Cities to Global Warming, Sea Level Rise, Climate Change and Endemic Hazards; Springer Briefs in Environmental Science; Springer: Cham, Switzerland, 2020; pp. 11–25. [Google Scholar] [CrossRef]
  25. Brinkmann, R. Wicked Problems and Disasters. Environmental Sustainability in a Time of Change. Palgrave Studies in Environmental Sustainability; Palgrave Macmillan: Cham, Switzerland, 2020; pp. 55–82. [Google Scholar] [CrossRef]
  26. Luino, F.; De Graff, J.; Roccati, A.; Biddoccu, M.; Cirio, C.G.; Faccini, F.; Turconi, L. Eighty years of data collected for the determination of rainfall threshold triggering shallow landslides and mud-debris flows in the Alps. Water 2020, 12, 133. [Google Scholar] [CrossRef] [Green Version]
  27. Kuriqi, A.; Ardiçlioglu, M.; Muceku, Y. Investigation of seepage effect on river dike’s stability under steady state and transient conditions. Pollack Period. 2016, 11, 87–104. [Google Scholar] [CrossRef] [Green Version]
  28. Kuriqi, A.; Ardiçlioǧlu, M. Investigation of hydraulic regime at middle part of the Loire River in context of floods and low flow events. Pollack Period. 2018, 13, 145–156. [Google Scholar] [CrossRef]
  29. Bui, Q.T.; Nguyen, Q.H.; Nguyen, X.L.; Pham, V.D.; Nguyen, H.D.; Pham, V.M. Verification of novel integrations of swarm intelligence algorithms into deep learning neural network for flood susceptibility mapping. J. Hydrol. 2020, 581, 124379. [Google Scholar] [CrossRef]
  30. Munir, B.A.; Ahmad, S.R.; Hafeez, S. Integrated hazard modeling for simulating torrential stream response to flash flood events. ISPRS Int. J. Geo. Inf. 2020, 9, 1. [Google Scholar] [CrossRef] [Green Version]
  31. Okaka, W.T. Climate Change-Induced Flood Disaster Policy Communication Issues for Local Community Adaptation Resilience Management in Uganda: Climate Information Services for Effective National Flood Risk Assessment Decision Communication. In Decision Support Methods for Assessing Flood Risk and Vulnerability; IGI Global: Pennsylvania, PA, USA, 2020; pp. 230–249. [Google Scholar]
  32. Sinha, D.K. Natural Disaster Reduction: South East Asian Realities, Risk Perception and Global Strategies; Anthem Press: London, UK, 2007. [Google Scholar]
  33. Mallick, J.; Salam, R.; Amin, R.; Islam, A.R.M.T.; Islam, A.; Siddik, M.N.A.; Alam, G.M. Assessing factors affecting drought, earthquake, and flood risk perception: Empirical evidence from Bangladesh. Nat. Hazards 2022, 112, 1633–1656. [Google Scholar] [CrossRef]
  34. Gazi, M.Y.; Islam, M.A.; Hossain, S. Flood-hazard mapping in a regional scale way forward to the future hazard atlas in Bangladesh. Malays. J. Geosci. 2019, 3, 1–11. [Google Scholar] [CrossRef]
  35. Rahman, M.; Chen, N.; Elbeltagi, A.; Islam, M.M.; Alam, M.; Pourghasemi, H.R.; Tao, W.; Zhang, J.; Shufeng, T.; Faiz, H.; et al. Application of stacking hybrid machine learning algorithms in delineating multi-type flooding in Bangladesh. J. Environ. Manag. 2021, 295, 113086. [Google Scholar] [CrossRef]
  36. Sarkar, S.K.; Ansar, S.B.; Ekram, K.M.M.; Khan, M.H.; Talukdar, S.; Naikoo, M.W.; Islam, A.R.M.T.; Rahman, A.; Mosavi, A. Developing Robust Flood Susceptibility Model with Small Numbers of Parameters in Highly Fertile Regions of Northwest Bangladesh for Sustainable Flood and Agriculture Management. Sustainability 2022, 14, 3982. [Google Scholar] [CrossRef]
  37. Ahmed, N.; Hoque, M.A.A.; Arabameri, A.; Pal, S.C.; Chakrabortty, R.; Jui, J. Flood susceptibility mapping in Brahmaputra floodplain of Bangladesh using deep boost, deep learning neural network, and artificial neural network. Geocarto Int. 2021, 37, 8770–8791. [Google Scholar] [CrossRef]
  38. Bonafilia, D.; Tellman, B.; Anderson, T.; Issenberg, E. Sen1Floods11: A georeferenced dataset to train and test deep learning flood algorithms for sentinel-1. In Proceedings of the IEEE/CVF Conference on Computer Vision and Pattern Recognition Workshops, Seattle, WA, USA, 14–19 June 2020; pp. 210–211. [Google Scholar]
  39. Sahana, M.; Rehman, S.; Sajjad, H.; Hong, H. Exploring effectiveness of frequency ratio and support vector machine models in storm surge flood susceptibility assessment: A study of Sundarban Biosphere Reserve, India. Catena 2020, 189, 104450. [Google Scholar] [CrossRef]
  40. Chen, W.; Hong, H.; Li, S.; Shahabi, H.; Wang, Y.; Wang, X.; Ahmad, B.B. Flood susceptibility modelling using novel hybrid approach of reduced-error pruning trees with bagging and random subspace ensembles. J. Hydrol. 2019, 575, 864–873. [Google Scholar] [CrossRef]
  41. Bui, D.T.; Ngo, P.T.T.; Pham, T.D.; Jaafari, A.; Minh, N.Q.; Hoa, P.V.; Samui, P. A novel hybrid approach based on a swarm intelligence optimized extreme learning machine for flash flood susceptibility mapping. Catena 2019, 179, 184–196. [Google Scholar] [CrossRef]
  42. Mahmood, S.; Rahman, A.U. Flash flood susceptibility modeling using geo-morphometric and hydrological approaches in Panjkora Basin, Eastern Hindu Kush, Pakistan. Environ. Earth Sci. 2019, 78, 1–16. [Google Scholar] [CrossRef]
  43. Hong, H.; Panahi, M.; Shirzadi, A.; Ma, T.; Liu, J.; Zhu, A.X.; Chen, W.; Kougias, I.; Kazakis, N. Flood susceptibility assessment in Hengfeng area coupling adaptive neuro-fuzzy inference system with genetic algorithm and differential evolution. Sci. Total Environ. 2018, 621, 1124–1141. [Google Scholar] [CrossRef]
  44. Termeh, S.V.R.; Kornejady, A.; Pourghasemi, H.R.; Keesstra, S. Flood susceptibility mapping using novel ensembles of adaptive neuro fuzzy inference system and metaheuristic algorithms. Sci. Total Environ. 2018, 615, 438–451. [Google Scholar] [CrossRef]
  45. Siahkamari, S.; Haghizadeh, A.; Zeinivand, H.; Tahmasebipour, N.; Rahmati, O. Spa- tial prediction of flood-susceptible areas using frequency ratio and maximum en- tropy models. Geocarto Int. 2018, 33, 927–941. [Google Scholar] [CrossRef]
  46. Souissi, D.; Zouhri, L.; Hammami, S.; Msaddek, M.H.; Zghibi, A.; Dlala, M. GIS-based MCDM–AHP modeling for flood susceptibility mapping of arid areas, southeastern Tunisia. Geocarto Int. 2019, 35, 991–1017. [Google Scholar] [CrossRef]
  47. Dano, U.L.; Balogun, A.L.; Matori, A.N.; Wan Yusouf, K.; Abubakar, I.R.; Said Mohamed, M.A.; Aina, Y.A.; Pradhan, B. Flood susceptibility mapping using GIS-based analytic network process: A case study of Perlis, Malaysia. Water 2019, 11, 615. [Google Scholar] [CrossRef] [Green Version]
  48. Nachappa, T.G.; Piralilou, S.T.; Gholamnia, K.; Ghorbanzadeh, O.; Rahmati, O.; Blaschke, T. Flood susceptibility mapping with machine learning, multi-criteria decision analysis and ensemble using Dempster Shafer Theory. J. Hydrol. 2020, 590, 125275. [Google Scholar] [CrossRef]
  49. Chen, J.; Li, Q.; Wang, H.; Deng, M. A machine learning ensemble approach based on random forest and radial basis function neural network for risk evaluation of regional flood disaster: A case study of the Yangtze River Delta, China. Int. J. Environ. Res. Public Health 2020, 17, 49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Moghaddam, D.D.; Pourghasemi, H.R.; Rahmati, O. Assessment of the contribution of geo-environmental factors to flood inundation in a semi-arid region of SW Iran: Comparison of different advanced modeling approaches. In Natural Hazards GIS-Based Spatial Modeling Using Data Mining Techniques; Springer: Cham, Switzerland, 2019; pp. 59–78. [Google Scholar]
  51. Xu, C. Assessment of earthquake-triggered landslide susceptibility based on expert knowledge and information value methods: A case study of the 20 April 2013 Lushan, China Mw6. 6 earthquakes. Disaster Adv. 2013, 6, 119–130. [Google Scholar]
  52. Chen, W.; Li, W.; Hou, E.; Zhao, Z.; Deng, N.; Bai, H.; Wang, D. Landslide susceptibility mapping based on GIS and information value model for the Chencang District of Baoji, China. Arab. J. Geosci. 2014, 7, 4499–4511. [Google Scholar] [CrossRef]
  53. ShafapourTehrany, M.; Kumar, L.; NeamahJebur, M.; Shabani, F. Evaluating the application of the statistical index method in flood susceptibility mapping and its comparison with frequency ratio and logistic regression methods. Geomat. Nat. Haz. Risk. 2019, 10, 79–101. [Google Scholar] [CrossRef] [Green Version]
  54. Pham, B.T.; Phong, T.V.; Nguyen, H.D.; Qi, C.; Al-Ansari, N.; Amini, A.; Ho, L.S.; Tuyen, T.T.; Yen, H.P.H.; Ly, H.B.; et al. A comparative study of kernel logistic regression, radial basis function classifier, multinomial naïve bayes, and logistic model tree for flash flood susceptibility mapping. Water 2020, 12, 239. [Google Scholar] [CrossRef] [Green Version]
  55. Ali, S.A.; Parvin, F.; Pham, Q.B.; Vojtek, M.; Vojteková, J.; Costache, R.; Linh, N.T.T.; Nguyen, H.Q.; Ahmad, A.; Ghorbani, M.A. GIS-based comparative assessment of flood susceptibility mapping using hybrid multi-criteria decision-making approach, naïve Bayes tree, bivariate statistics and logistic regression: A case of Topľa basin, Slovakia. Ecol. Indic. 2020, 117, 106620. [Google Scholar] [CrossRef]
  56. Chen, W.; Panahi, M.; Tsangaratos, P.; Shahabi, H.; Ilia, I.; Panahi, S.; Li, S.; Jaafari, A.; Ahmad, B.B. Applying population-based evolutionary algorithms and a neuro-fuzzy system for modeling landslide susceptibility. Catena 2019, 172, 212–231. [Google Scholar] [CrossRef]
  57. Paul, G.C.; Saha, S.; Hembram, T.K. Application of the GIS-based probabilistic models for mapping the flood susceptibility in Bansloi sub-basin of Ganga-Bhagirathi River and their comparison. Remote Sens. Earth Syst. Sci. 2019, 2, 120–146. [Google Scholar] [CrossRef]
  58. Wang, Y.; Hong, H.; Chen, W.; Li, S.; Pamučar, D.; Gigović, L.; Drobnjak, S.; Tien Bui, D.; Duan, H. A hybrid GIS multi-criteria decision-making method for flood susceptibility mapping at Shangyou, China. Remote Sens. 2018, 11, 62. [Google Scholar] [CrossRef] [Green Version]
  59. Sahana, M.; Patel, P.P. A comparison of frequency ratio and fuzzy logic models for flood susceptibility assessment of the lower Kosi River Basin in India. Environ. Earth Sci. 2019, 78, 289. [Google Scholar] [CrossRef]
  60. Hong, H.; Tsangaratos, P.; Ilia, I.; Liu, J.; Zhu, A.X.; Chen, W. Application of fuzzy weight of evidence and data mining techniques in construction of flood susceptibility map of Poyang County, China. Sci. Total Environ. 2018, 625, 575–588. [Google Scholar] [CrossRef] [PubMed]
  61. Wang, Y.; Fang, Z.; Hong, H.; Peng, L. Flood susceptibility mapping using convolutional neural network frameworks. J. Hydrol. 2020, 582, 124482. [Google Scholar] [CrossRef]
  62. Shahabi, H.; Shirzadi, A.; Ghaderi, K.; Omidvar, E.; Al-Ansari, N.; Clague, J.J.; Geertsema, M.; Khosravi, K.; Amini, A.; Bahrami, S.; et al. Flood detection and susceptibility mapping using sentinel-1 remote sensing data and a machine learning approach: Hybrid intelligence of bagging ensemble based on k-nearest neighbor classifier. Remote Sens. 2020, 12, 266. [Google Scholar] [CrossRef] [Green Version]
  63. Dodangeh, E.; Choubin, B.; Eigdir, A.N.; Nabipour, N.; Panahi, M.; Shamshirband, S.; Mosavi, A. Integrated machine learning methods with resampling algorithms for flood susceptibility prediction. Sci. Total 2020, 705, 135983. [Google Scholar] [CrossRef]
  64. Oeurng, C.; Sauvage, S.; Sánchez-Pérez, J.M. Assessment of hydrology, sediment and particulate organic carbon yield in a large agricultural catchment using the SWAT model. J. Hydrol. 2011, 401, 145–153. [Google Scholar] [CrossRef] [Green Version]
  65. Getahun, Y.S.; Gebre, S.L. Flood hazard assessment and mapping of flood inundation area of the Awash River Basin in Ethiopia using GIS and HEC-GeoRAS/HEC-RAS model. J. Civ. Environ. Eng. 2015, 5, 1. [Google Scholar]
  66. Khosravi, K.; Pham, B.T.; Chapi, K.; Shirzadi, A.; Shahabi, H.; Revhaug, I.; Prakash, I.; Bui, D.T. A comparative assessment of decision trees algorithms for flash flood susceptibility modeling at Haraz watershed, northern Iran. Sci. Total Environ. 2018, 627, 744–755. [Google Scholar] [CrossRef]
  67. Janizadeh, S.; Avand, M.; Jaafari, A.; Phong, T.V.; Bayat, M.; Ahmadisharaf, E.; Prakash, I.; Pham, B.T.; Lee, S. Prediction success of machine learning methods for flash flood susceptibility mapping in the Tafresh watershed, Iran. Sustainability 2019, 11, 5426. [Google Scholar] [CrossRef] [Green Version]
  68. Chen, W.; Li, Y.; Xue, W.; Shahabi, H.; Li, S.; Hong, H.; Wang, X.; Bian, H.; Zhang, S.; Pradhan, B.; et al. Modeling flood susceptibility using data-driven approaches of naïve bayes tree, alternating decision tree, and random forest methods. Sci. Total Environ. 2020, 701, 134979. [Google Scholar] [CrossRef]
  69. Falah, F.; Rahmati, O.; Rostami, M.; Ahmadisharaf, E.; Daliakopoulos, I.N.; Pourghasemi, H.R. Artificial neural networks for flood susceptibility mapping in data-scarce urban areas. In Spatial Modeling in GIS and R for Earth and Environmental Sciences; Elsevier: Amsterdam, The Netherlands, 2019; pp. 323–336. [Google Scholar]
  70. Pham, B.T.; Jaafari, A.; Avand, M.; Al-Ansari, N.; Dinh Du, T.; Yen, H.P.H.; Phong, T.V.; Nguyen, D.H.; Le, H.V.; Mafi-Gholami, D.; et al. Performance evaluation of machine learning methods for forest fire modeling and prediction. Symmetry 2020, 12, 1022. [Google Scholar] [CrossRef]
  71. Chakrabortty, R.; Pal, S.C.; Janizadeh, S.; Santosh, M.; Roy, P.; Chowdhuri, I.; Saha, A. Impact of climate change on future flood susceptibility: An evaluation based on deep learning algorithms and GCM model. Water Resour. Manag. 2021, 35, 4251–4274. [Google Scholar] [CrossRef]
  72. Garcia-Garcia, A.; Orts-Escolano, S.; Oprea, S.; Villena-Martinez, V.; Martinez-Gonzalez, P.; GarciaRodriguez, J. A survey on deep learning techniques for image and video semantic segmentation. Appl. Soft Comput. 2018, 70, 41. [Google Scholar] [CrossRef]
  73. Costache, R.; Tin, T.T.; Arabameri, A.; Crăciun, A.; Ajin, R.S.; Costache, I.; Islam, A.R.M.T.; Abba, S.I.; Sahana, M.; Avand, M.; et al. Flash-flood hazard using deep learning based on H2O R package and fuzzy-multicriteria decision-making analysis. J. Hydrol. 2022, 609, 127747. [Google Scholar] [CrossRef]
  74. Torres, J.; Galicia, A.; Troncoso, A.; Martínez-Álvarez, F. A scalable approach based on deep learning for big data time series forecasting. Integr. Comput. Aided Eng. 2018, 25, 335–348. [Google Scholar] [CrossRef]
  75. Shafizadeh-Moghadam, H.; Valavi, R.; Shahabi, H.; Chapi, K.; Shirzadi, A. Novel forecasting approaches using combination of machine learning and statistical models for flood susceptibility mapping. J. Environ. Manag. 2018, 217, 1–11. [Google Scholar] [CrossRef] [Green Version]
  76. Pham, B.T.; Bui, D.T.; Prakash, I.; Dholakia, M.B. Hybrid integration of Multilayer Perceptron Neural Networks and machine learning ensembles for landslide susceptibility assessment at Himalayan area (India) using GIS. Catena 2017, 149, 52–63. [Google Scholar] [CrossRef]
  77. Ghasemain, B.; Asl, D.T.; Pham, B.T.; Avand, M.; Nguyen, H.D.; Janizadeh, S.J.V.J.O.E.S. Shallow landslide susceptibility mapping: A comparison between classification and regression tree and reduced error pruning tree algorithms. Vietnam J. Earth Sci. 2020, 42, 208–227. [Google Scholar] [CrossRef]
  78. Pham, B.T.; Phong, T.V.; Nguyen-Thoi, T.; Parial, K.; Singh, K.; Ly, H.B.; Nguyen, K.T.; Ho, L.S.; Le, H.V.; Prakash, I. Ensemble modeling of landslide susceptibility using random subspace learner and different decision tree classifiers. Geocarto Int. 2021, 37, 735–757. [Google Scholar] [CrossRef]
  79. Tang, X.; Li, J.; Liu, M.; Liu, W.; Hong, H. Flood susceptibility assessment based on a novel random Naïve Bayes method: A comparison between different factor discretization methods. Catena 2020, 190, 104536. [Google Scholar] [CrossRef]
  80. Islam, A.R.M.T.; Talukdar, S.; Akhter, S.; Eibek, K.U.; Rahman, M.; Pal, S.; Naikoo, M.W.; Rahman, A.; Mosavi, A. Assessing the Impact of the Farakka Barrage on Hydrological Alteration in the Padma River with Future Insight. Sustainability 2022, 14, 5233. [Google Scholar] [CrossRef]
  81. Dewan, A.; Corner, R.; Saleem, A.; Rahman, M.M.; Haider, M.R.; Rahman, M.M.; Sarker, M.H. Assessing channel changes of the Ganges-Padma River system in Bangladesh using Landsat and hydrological data. Geomorphology 2017, 276, 257–279. [Google Scholar] [CrossRef]
  82. Islam, A.R.M.T.; Talukdar, S.; Mahato, S.; Kundu, S.; Eibek, K.U.; Pham, Q.B.; Kuriqi, A.; Linh, N.T.T. Machine learning algorithm-based risk assessment of riparian wetlands in Padma River Basin of Northwest Bangladesh. Environ. Sci. Pollut. Res. 2021, 28, 34450–34471. [Google Scholar] [CrossRef]
  83. Hossain, M.Y. Morphometric relationships of length-weight and length-length of four Cyprinid small indigenous fish species from the Padma River (NW Bangladesh). Turk. J. Fish. Aquat. Sci. 2010, 10, 131–134. [Google Scholar] [CrossRef]
  84. Sarkar, D.; Mondal, P. Flood vulnerability mapping using frequency ratio (FR) model: A case study on Kulik river basin, Indo-Bangladesh Barind region. Appl. Water Sci. 2020, 10, 17. [Google Scholar] [CrossRef] [Green Version]
  85. Islam, A.R.M.T.; Talukdar, S.; Mahato, S.; Kundu, S.; Eibek, K.U.; Pham, Q.B.; Kuriqi, A.; Linh, N.T.T. Flood susceptibility modelling using advanced ensemble machine learning models. Geosci. Front. 2021, 12, 12101075. [Google Scholar]
  86. Wubalem, A.; Tesfaw, G.; Dawit, Z.; Getahun, B.; Mekuria, T.; Jothimani, M. Comparison of statistical and analytical hierarchy process methods on flood susceptibility mapping: In a case study of Tana sub-basin in northwestern Ethiopia. Nat. Hazards Earth Syst. Sci. Discuss 2020, 332, 1–43. [Google Scholar]
  87. Rahman, M.; Ningsheng, C.; Islam, M.M.; Dewan, A.; Iqbal, J.; Washakh, R.M.A.; Shufeng, T. Flood susceptibility assessment in Bangladesh using machine learning and multi-criteria decision analysis. Earth Syst. Environ. 2019, 3, 585–601. [Google Scholar] [CrossRef]
  88. Kim, H.I.; Han, K.Y. Urban flood prediction using deep neural network with data augmentation. Water 2019, 12, 899. [Google Scholar] [CrossRef] [Green Version]
  89. Available online: https://chrsdata.eng.uci.edu/ (accessed on 13 October 2022).
  90. Nguyen, V.T.; Tran, T.H.; Ha, N.A.; Ngo, V.L.; Nadhir, A.A.; Tran, V.P.; Duy Nguyen, H.; MA, M.; Amini, A.; Prakash, I.; et al. GIS based novel hybrid computational intelligence models for mapping landslide susceptibility: A case study at da lat city, Vietnam. Sustainability 2019, 11, 7118. [Google Scholar] [CrossRef] [Green Version]
  91. Kourgialas, N.N.; Karatzas, G.P. Flood management and a GIS modelling method to assess flood-hazard areas—A case study. Hydrol. Sci. J. J. Des Sci. Hydrol. 2011, 56, 212–225. [Google Scholar] [CrossRef]
  92. Available online: https://earthexplorer.usgs.gov/ (accessed on 15 October 2022).
  93. Available online: https://www.usgs.gov/programs/energy (accessed on 20 November 2022).
  94. Available online: https://www.fao.org (accessed on 12 December 2022).
  95. Yi, Y.; Zhang, Z.; Zhang, W.; Jia, H.; Zhang, J. Landslide susceptibility mapping using multiscale sampling strategy and convolutional neural network: A case study in Jiuzhaigou region. Catena 2020, 195, 104851. [Google Scholar] [CrossRef]
  96. Pham, B.T.; Luu, C.; Van Phong, T.; Trinh, P.T.; Shirzadi, A.; Renoud, S.; Asadi, S.; Van Le, H.; von Meding, J.; Clague, J.J. Can deep learning algorithms outperform benchmark machine learning algorithms in flood susceptibility modeling. J. Hydrol. 2021, 592, 125615. [Google Scholar] [CrossRef]
  97. Costache, R.; Arabameri, A.; Costache, I.; Crăciun, A.; Islam, A.R.M.T.; Abba, S.I.; Sahana, M.; Pham, B.T. Flood susceptibility evaluation through deep learning optimizer ensembles and GIS techniques. J. Environ. Manag. 2022, 316, 115316. [Google Scholar] [CrossRef]
  98. Nhu, V.H.; Zandi, D.; Shahabi, H.; Chapi, K.; Shirzadi, A.; Al-Ansari, N.; Singh, S.K.; Dou, J.; Nguyen, H. Comparison of support vector machine, Bayesian logistic regression, and alternating decision tree algorithms for shallow landslide susceptibility mapping along a mountainous road in the west of Iran. Appl. Sci. 2020, 10, 5047. [Google Scholar] [CrossRef]
  99. Costache, R.; Arabameri, A.; Blaschke, T.; Pham, Q.B.; Pham, B.T.; Pandey, M.; Arora, A.; Linh, N.T.T.; Costache, I. Flash-flood potential mapping using deep learning, alternating decision trees and data provided by remote sensing sensors. Sensors 2021, 21, 280. [Google Scholar] [CrossRef] [PubMed]
  100. Wang, H.; Wang, H.; Wu, Z.; Zhou, Y. Using multi-factor analysis to predict urban flood depth based on Naive Bayes. Water 2021, 13, 432. [Google Scholar] [CrossRef]
  101. Sellami, E.M.; Maanan, M.; Rhinane, H. Performance of machine learning algorithms for mapping and forecasting of flash flood susceptibility in Tetouan, Morocco. Int. Arch. Photogramm. Remote Sens. Spat. Inf. Sci. 2022, 46, 305–313. [Google Scholar] [CrossRef]
  102. Kia, M.B.; Pirasteh, S.; Pradhan, B.; Mahmud, A.R.; Sulaiman, W.N.A.; Moradi, A. An artificial neural network model for flood simulation using GIS: Johor River Basin, Malaysia. Environ. Earth Sci. 2012, 67, 251–264. [Google Scholar] [CrossRef]
  103. Kalantar, B.; Ueda, N.; Saeidi, V.; Janizadeh, S.; Shabani, F.; Ahmadi, K.; Shabani, F. Deep neural network utilizing remote sensing datasets for flood hazard susceptibility mapping in Brisbane, Australia. Remote Sens. 2021, 13, 2638. [Google Scholar] [CrossRef]
  104. Costache, R.; Arabameri, A.; Moayedi, H.; Pham, Q.B.; Santosh, M.; Nguyen, H.; Pandey, M.; Pham, B.T. Flash-flood potential index estimation using fuzzy logic combined with deep learning neural network, naïve Bayes, XGBoost and classification and regression tree. Geocarto Int. 2021, 37, 6780–6807. [Google Scholar] [CrossRef]
  105. Khosravi, K.; Safari, M.J.S.; Cooper, J.R. Clear-water scour depth prediction in long channel contractions: Application of new hybrid machine learning algorithms. Ocean. Eng. 2021, 238, 109721. [Google Scholar] [CrossRef]
  106. Khosravi, K.; Barzegar, R.; Golkarian, A.; Busico, G.; Cuoco, E.; Mastrocicco, M.; Colombani, N.; Tedesco, D.; Ntona, M.M.; Kazakis, N. Predictive modeling of selected trace elements in groundwater using hybrid algorithms of iterative classifier optimizer. J. Contam. Hydrol. 2021, 242, 103849. [Google Scholar] [CrossRef] [PubMed]
  107. Mitra, R.; Saha, P.; Das, J. Assessment of the performance of GIS-based analytical hierarchical process (AHP) approach for flood modelling in Uttar Dinajpur district of West Bengal, India. Geomat. Nat. Hazards Risk 2022, 13, 2183–2226. [Google Scholar] [CrossRef]
  108. Saha, S.; Saha, A.; Hembram, T.K.; Kundu, B.; Sarkar, R. Novel Ensemble of Deep Learning Neural Network and Support Vector Machine for Landslide Susceptibility Mapping in Tehri Region, Garhwal Himalaya. Geocarto Int. 2022, 37, 17018–17043. [Google Scholar] [CrossRef]
  109. Khosravi, K.; Shahabi, H.; Pham, B.T.; Adamowski, J.; Shirzadi, A.; Pradhan, B.; Dou, J.; Ly, H.B.; Gróf, G.; Ho, H.L.; et al. A comparative assessment of flood susceptibility modeling using multi-criteria decision-making analysis and machine learning methods. J. Hydrol. 2019, 573, 311–323. [Google Scholar] [CrossRef]
  110. Arora, A.; Arabameri, A.; Pandey, M.; Siddiqui, M.A.; Shukla, U.K.; Bui, D.T.; Mishra, V.N.; Bhardwaj, A. Optimization of state-of-the-art fuzzy metaheuristic ANFIS-based machine learning models for flood susceptibility prediction mapping in the Middle Ganga Plain, India. Sci. Total Environ. 2021, 750, 141565. [Google Scholar] [CrossRef] [PubMed]
  111. Al-Aizari, A.R.; Al-Masnay, Y.A.; Aydda, A.; Zhang, J.; Ullah, K.; Islam, A.R.M.T.; Habib, T.; Kaku, D.U.; Nizeyimana, J.C.; Al-Shaibah, B.; et al. Assessment Analysis of Flood Susceptibility in Tropical Desert Area: A Case Study of Yemen. Remote Sens. 2022, 14, 4050. [Google Scholar] [CrossRef]
  112. Band, S.S.; Janizadeh, S.; Chandra Pal, S.; Saha, A.; Chakrabortty, R.; Shokri, M.; Mosavi, A. Novel ensemble approach of deep learning neural network (DLNN) model and particle swarm optimization (PSO) algorithm for prediction of gully erosion susceptibility. Sensors 2020, 20, 5609. [Google Scholar] [CrossRef] [PubMed]
  113. Luu, C.; Pham, B.T.; Van Phong, T.; Costache, R.; Nguyen, H.D.; Amiri, M.; Bui, Q.D.; Nguyen, L.T.; Van Le, H.; Prakash, I.; et al. GIS-Based Ensemble Computational models for Flood Susceptibility Prediction in the Quang Binh Province, Vietnam. J. Hydrol. 2021, 599, 126500. [Google Scholar] [CrossRef]
  114. Tehrany, M.S.; 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. [Google Scholar] [CrossRef]
  115. Riazi, M.; Khosravi, K.; Shahedi, K.; Ahmad, S.; Jun, C.; Bateni, S.M.; Kazaki, N. Enhancing flood susceptibility modeling using multi temporal Sar images, CHIRPS data and hybrid machine learning algorithms. Sci. Total Environ. 2023, 871, 161066. [Google Scholar] [CrossRef] [PubMed]
  116. Hong, H.; Pradhan, B.; Xu, C.; Bui, D.T. Spatial prediction of landslide hazard at the Yihuang area (China) using two-class kernel logistic regression, alternating decision tree and support vector machines. Catena 2015, 133, 266–281. [Google Scholar] [CrossRef]
  117. Sevgen, E.; Kocaman, S.; Nefeslioglu, H.A.; Gokceoglu, C. A novel performance assessment approach using photogrammetric techniques for landslide susceptibility mapping with logistic regression, ANN and random forest. Sensors 2019, 19, 3940. [Google Scholar] [CrossRef] [Green Version]
  118. Lee, S.; Lee, M.J.; Jung, H.S.; Lee, S. Landslide susceptibility mapping using Naïve Bayes and Bayesian network models in Umyeonsan, Korea. Geocarto Int. 2020, 35, 1665–1679. [Google Scholar] [CrossRef]
  119. Pham, B.T.; Avand, M.; Janizadeh, S.; Phong, T.V.; Al-Ansari, N.; Ho, L.S.; Das, S.; Le, H.V.; Amini, A.; Bozchaloei, S.K.; et al. GIS based hybrid computational approaches for flash flood susceptibility assessment. Water 2020, 12, 683. [Google Scholar] [CrossRef] [Green Version]
  120. Zhao, G.; Pang, B.; Xu, Z.; Yue, J.; Tu, T. Mapping flood susceptibility in mountainous areas on a national scale in China. Sci. Total Environ. 2018, 615, 1133–1142. [Google Scholar] [CrossRef]
  121. Chakrabortty, R.; Chandra Pal, S.; Rezaie, F.; Arabameri, A.; Lee, S.; Roy, P.; Saha, A.; Chowdhuri, I.; Moayedi, H. Flash-flood hazard susceptibility mapping in Kangsabati River Basin, India. Geocarto Int. 2021, 37, 6713–6735. [Google Scholar] [CrossRef]
  122. Pal, S.C.; Chakrabortty, R.; Arabameri, A.; Santosh, M.; Saha, A.; Chowdhuri, I.; Roy, P.; Shit, M. Chemical weathering and gully erosion causing land degradation in a complex river basin of Eastern India: An integrated field, analytical and artificial intelligence approach. Nat. Hazards 2022, 110, 847–879. [Google Scholar] [CrossRef]
  123. Arabameri, A.; AsadiNalivan, O.; Saha, S.; Roy, J.; Pradhan, B.; Tiefenbacher, J.P.; Thi Ngo, P.T. Novel ensemble approaches of machine learning techniques in modeling the gully erosion susceptibility. Remote Sens. 2020, 12, 1890. [Google Scholar] [CrossRef]
  124. Saha, A.; Pal, S.C.; Chowdhuri, I.; Islam, A.R.M.T.; Chakrabortty, R.; Roy, P. Application of neural network model-based framework approach to identify gully erosion potential hotspot zones in sub-tropical environment. Geocarto Int. 2022, 37, 14758–14784. [Google Scholar] [CrossRef]
  125. Arabameri, A.; Chen, W.; Loche, M.; Zhao, X.; Li, Y.; Lombardo, L.; Cerda, A.; Pradhan, B.; Bui, D.T. Comparison of machine learning models for gully erosion susceptibility mapping. Geosci. Front. 2020, 11, 1609–1620. [Google Scholar] [CrossRef]
  126. Chowdhuri, I.; Pal, S.C.; Saha, A.; Chakrabortty, R.; Roy, P. Evaluation of different DEMs for gully erosion susceptibility mapping using in-situ field measurement and validation. Ecol. Inform. 2021, 65, 101425. [Google Scholar] [CrossRef]
  127. Chen, W.; Zhou, Y.; Zhou, E.; Xiang, Z.; Zhou, W.; Lu, J. Wildfire risk assessment of transmission-line corridors based on naïve bayes network and remote sensing data. Sensors 2021, 21, 634. [Google Scholar] [CrossRef]
  128. Tella, A.; Balogun, A.L.; Adebisi, N.; Abdullah, S. Spatial assessment of PM10 hotspots using random forest, K-nearest neighbour and Naïve Bayes. Atmos. Pollut. Res. 2021, 12, 101202. [Google Scholar] [CrossRef]
  129. Nguyen, H.D. Hybrid models based on deep learning neural network and optimization algorithms for the spatial prediction of tropical forest fire susceptibility in Nghe An province, Vietnam. Geocarto Int. 2022, 37, 11281–11305. [Google Scholar] [CrossRef]
  130. Costache, R.; Tin, T.T.; Arabameri, A.; Crăciun, A.; Costache, I.; Islam, A.R.M.T.; Sahana, M.; Pham, B.T. Stacking state-of-the-art ensemble for flash-flood potential assessment. Geocarto Int. 2022, 37, 13812–13838. [Google Scholar] [CrossRef]
  131. Bui, D.T.; Tsangaratos, P.; Nguyen, V.T.; Van Liem, N.; Trinh, P.T. Comparing the prediction performance of a Deep Learning Neural Network model with conventional machine learning models in landslide susceptibility assessment. Catena 2020, 188, 104426. [Google Scholar] [CrossRef]
Figure 1. Map showing the location of the study region.
Figure 1. Map showing the location of the study region.
Land 12 00810 g001
Figure 3. Flood susceptibility causal variables: (a) aspect, (b) elevation (m), (c) slope angle (degree), (d) curvature, (e) plan curvature, (f) profile curvature, (g) flow direction, (h) flow accumulation, (i) land use and land cover (LULC), (j) normalized difference vegetation index (NDVI), (k) distance to the rivers (m), (l) soil, (m) mean annual rainfall (mm), (n) river density (RD), (o) stream power index (SPI), (p) topographic wetness index (TWI), (q) sediment transport index (STI), (r) geology.
Figure 3. Flood susceptibility causal variables: (a) aspect, (b) elevation (m), (c) slope angle (degree), (d) curvature, (e) plan curvature, (f) profile curvature, (g) flow direction, (h) flow accumulation, (i) land use and land cover (LULC), (j) normalized difference vegetation index (NDVI), (k) distance to the rivers (m), (l) soil, (m) mean annual rainfall (mm), (n) river density (RD), (o) stream power index (SPI), (p) topographic wetness index (TWI), (q) sediment transport index (STI), (r) geology.
Land 12 00810 g003aLand 12 00810 g003bLand 12 00810 g003c
Figure 4. Flood susceptibility assessment using ADT-ICO (a), NB-ICO (b), ANN-ICO (c), and DLNN-ICO (d).
Figure 4. Flood susceptibility assessment using ADT-ICO (a), NB-ICO (b), ANN-ICO (c), and DLNN-ICO (d).
Land 12 00810 g004
Figure 5. Areal percentage of different flood susceptibility classes.
Figure 5. Areal percentage of different flood susceptibility classes.
Land 12 00810 g005
Figure 6. ROC curve using training (a) and validation (b) datasets.
Figure 6. ROC curve using training (a) and validation (b) datasets.
Land 12 00810 g006
Figure 7. Validation of the model when considering the training (a) and validation (b) datasets.
Figure 7. Validation of the model when considering the training (a) and validation (b) datasets.
Land 12 00810 g007
Figure 8. Taylor diagram.
Figure 8. Taylor diagram.
Land 12 00810 g008
Figure 9. Convergence of the objective cost function in the ICO-DLNN model.
Figure 9. Convergence of the objective cost function in the ICO-DLNN model.
Land 12 00810 g009
Table 1. Multicollinearity assessment.
Table 1. Multicollinearity assessment.
Sl. No.VariablesCollinearity
TOLVIF
1Aspect0.4532.208
2Elevation0.3362.976
3Slope0.4992.004
4Curvature0.6381.567
5Plan curvature0.9131.095
6Profile curvature0.5611.783
7Flow direction0.3912.558
8Flow accumulation0.5591.789
9LULC0.3692.710
10NDVI0.7791.284
11Distance from River0.6521.534
12Soil0.4492.227
13Rainfall0.3053.279
14River density0.2973.367
15SPI0.3692.710
16TWI1.2450.803
17STI0.9511.052
18Geology0.9931.007
Table 2. Validation of the models.
Table 2. Validation of the models.
ModelsComplete DatasetPredictionTotalCorrect (%)Wrong (%)SensitivitySpecificityRecallF scorePPVNPVAUC
Nonflood (0)Flood (1)
TrainingDLNN-ICO027440314195.71418.5710.9090.8730.9090.8870.8670.9130.932
126260286185.71420.000
Total300300600190.71419.286
ADT-ICO026440304188.57125.7140.8780.8680.8780.8720.8670.8800.891
136260296185.71420.000
Total300300600187.14322.857
NB-ICO026044304185.71428.5710.8650.8550.8650.8590.8530.8670.873
140256296182.85722.000
Total300300600184.28625.286
ANN-ICO025852310184.28630.0000.8550.8320.8550.8410.8270.8600.835
142248290177.14326.000
Total300300600180.71428.000
ValidationDLNN-ICO01301814892.85714.2860.8680.8780.8680.8740.8800.8670.917
12013215294.2869.000
Total15015030093.57111.643
ADT-ICO01282014891.42915.7140.8550.8650.8550.8610.8670.8530.864
12213015292.85710.000
Total15015030092.14312.857
NB-ICO01262415090.00017.1430.8400.8400.8400.8400.8400.8400.845
12412615090.00012.000
Total15015030090.00014.571
ANN-ICO01232414787.85719.2860.8240.8370.8240.8320.8400.8200.801
12712615390.00012.000
Total15015030088.92915.643
Table 3. Importance of the variables.
Table 3. Importance of the variables.
ParametersAD-ICONB-ICOANN-ICODLNN-ICO
Aspect1491215
Elevation50405565
Slope35353040
Curvature20152520
Plan curvature20203015
Profile curvature10201520
Flow direction15252010
Flow accumulation30352015
LULC20203525
NDVI20302520
Distance from River70757080
Soil30352530
Rainfall70606570
River density40455050
SPI30303530
TWI15202010
STI20252030
Geology20202525
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mia, M.U.; Chowdhury, T.N.; Chakrabortty, R.; Pal, S.C.; Al-Sadoon, M.K.; Costache, R.; Islam, A.R.M.T. Flood Susceptibility Modeling Using an Advanced Deep Learning-Based Iterative Classifier Optimizer. Land 2023, 12, 810. https://doi.org/10.3390/land12040810

AMA Style

Mia MU, Chowdhury TN, Chakrabortty R, Pal SC, Al-Sadoon MK, Costache R, Islam ARMT. Flood Susceptibility Modeling Using an Advanced Deep Learning-Based Iterative Classifier Optimizer. Land. 2023; 12(4):810. https://doi.org/10.3390/land12040810

Chicago/Turabian Style

Mia, Md. Uzzal, Tahmida Naher Chowdhury, Rabin Chakrabortty, Subodh Chandra Pal, Mohammad Khalid Al-Sadoon, Romulus Costache, and Abu Reza Md. Towfiqul Islam. 2023. "Flood Susceptibility Modeling Using an Advanced Deep Learning-Based Iterative Classifier Optimizer" Land 12, no. 4: 810. https://doi.org/10.3390/land12040810

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

Article Metrics

Back to TopTop