Next Article in Journal
Effects of Selenium in Different Valences on the Community Structure and Microbial Functions of Biofilms
Previous Article in Journal
On the Relationship of Arctic Oscillation with Atmospheric Rivers and Snowpack in the Western United States Using Long-Term Multi-Platform Dataset
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Application of Analytical Hierarchy Process and Geophysical Method for Groundwater Potential Mapping in the Tata Basin, Morocco

by
Fatima Zahra Echogdali
1,
Said Boutaleb
1,
Amine Bendarma
2,
Mohamed Elmehdi Saidi
3,
Mohamed Aadraoui
4,
Mohamed Abioui
1,*,
Mohammed Ouchchen
1,
Kamal Abdelrahman
5,
Mohammed S. Fnais
5 and
Kochappi Sathyan Sajinkumar
6,7
1
Department of Earth Sciences, Faculty of Sciences, Ibn Zohr University, Agadir 80000, Morocco
2
Laboratory for Sustainable Innovation and Applied Research, Universiapolis—International University of Agadir, Agadir 80000, Morocco
3
Department of Earth Sciences, Faculty of Sciences and Techniques, Cadi Ayyad University, Marrakesh 40000, Morocco
4
Department of Earth Sciences, Faculty of Science and Technology, Sultan Moulay Slimane University, Beni-Mellal 23000, Morocco
5
Department of Geology & Geophysics, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
6
Department of Geology, University of Kerala, Thiruvananthapuram 695 581, Kerala, India
7
Department of Geological & Mining Engineering & Sciences, Michigan Technological University, Houghton, MI 49931, USA
*
Author to whom correspondence should be addressed.
Water 2022, 14(15), 2393; https://doi.org/10.3390/w14152393
Submission received: 11 June 2022 / Revised: 24 July 2022 / Accepted: 1 August 2022 / Published: 2 August 2022
(This article belongs to the Section Hydrogeology)

Abstract

:
Ensuring water availability for agriculture and drinking water supply in semi-arid mountainous regions requires control of factors influencing groundwater availability. In most cases, the population draws its water needs from the alluvial aquifers close to villages that are already limited and influenced by current climatic change. In addition, the establishment of deep wells in the hard rock aquifers depletes the aquifer. Hence, understanding the factors influencing water availability is an urgent requirement. The use of geographic information system (GIS), and remote sensing (RS), together with decision-making methods like analytical hierarchy process (AHP) will be of good aid in this regard. In the Tata basin, located in SE Morocco, ten factors were used to explain the groundwater potentiality map (GWPM). Five categories of potential zones were determined: very low (8.67%), low (17.74%), moderate (46.77%), high (19.95%), and very high (6.87%). The efficiency of the AHP model is validated using the ROC curve (receiver operating characteristics) which revealed a good correlation between the high potential groundwater zones and the spatial distribution of high flow wells. Geophysical prospecting, using electrical resistivity profiles, has made it possible to propose new well sites. It corresponds to conductive resistivity zones that coincide with the intersection of hydrogeological lineaments.

1. Introduction

Global climate change, an omnipresent reality, is manifested in the Morocco Kingdom as rising temperature [1], which has increased the risk of water scarcity in the coming decades. Furthermore, the scenario is aggravated by severely polluted water from agriculture, the main economic activity in this region. Out of the 13 billion m3 of surface water that can be mobilized annually, 8.8 billion m3 are lost through evaporation, discharge into the Atlantic Sea or existing leaks in drinking-water networks. Therefore, to meet the needs stated, the country’s natural water resources must be developed through a coherent, holistic plan, considering the need for conservation, prevention, and protection, both in water quantity and water quality, particularly in areas characterized by strong climatic contrast and increased scarcity of water resources. The development of these regions seems inevitably related to the shortage of water resources.
One of the main basins in this region is the Tata basin (sub-basin of the Drâa Wadi) since its economy depends mainly on agriculture, an activity whose production depends primarily on the availability of water resources [2]. Indeed, this activity is mainly developed at oases where traditional palm groves are irrigated from alluvial groundwater captured by springs, wells, and surface water. The total gross water need for this agricultural activity amounts to 82 million m3 [2]. The current status of water use shows that the annual volume of water used in the Tata basin is 69.3 million m3 [2], distributed between 8.3 million m3/year (12%) from surface water and 61 million m3/year (88%) from groundwater.
Comparing the theoretical water needs (82 million m3/year) with the volume of water that is currently available (69.3 million m3/year) shows a significant deficit for agricultural water demands. The coverage rate is 85%; the 15% deficit in the balance sheet is a matter of concern, as the irrigated areas are constantly increasing. This deficit will probably increase in the future, which will undoubtedly hinder the region’s economic development, an area already ruined by climate change [2].
This paper aims to update the knowledge on the existing water resources in this basin and to establish a groundwater potentiality map by diversifying the water supplies from aquifers other than the alluvial ones whose hydraulic equipment is subject to flood phenomena [3]. An elaboration of the groundwater potential map (GWPM) is useful to predict this resource’s development, conservation, and management strategies [4]. GWPM is generally provided by hydrogeological studies integrating, geophysical, geological, and hydrogeological surveys. It remains expensive, complex, and requires experts [5,6]. In contrast, the integration of geographic information system (GIS) and remote sensing (RS) methods allows for the efficient assessment of water resources, given their low cost and high capacity to analyze a dataset in a short time [7,8]. The delineation of GWPM mainly relies on integrating multiple factors such as topography, geomorphology, geological structures, lithology, and drainage networks [9,10]. These factors have rarely been studied together due to the unavailability of data and integration tools [11]. However, GIS and RS tools have emerged as effective tools for spatial data execution and decision-making [12].
The identification of factors affecting groundwater availability differs among researchers. Khan et al. [10] suggested rainfall, geology, lineaments, slope, drainage, and geomorphology map GWPM in the Central Eastern Desert in Egypt. Zghibi et al. [13] add land use and soil in the Korba aquifer, Tunisia. Naghibi et al. [14] consider elevation, plan curvature, profile curvature, distance from the river, distance from faults, stream power index (SPI), and topographic wetness index (TWI) for delineating potential areas in the Moghan basin, Iran. For the Tata basin, following the availability of data received through satellite image processing, a digital elevation model (DEM), and also through a geological map, ten factors were used: permeability, slope, topographic wetness index (TWI), plan curvature, profile curvature, stream transport index (STI), stream power index (SPI), lineament density, node density, and drainage density.
To integrate and analyze the set of factors, several methods were used: frequency ratio (FR) [14,15], evidential belief function (EBF) [16], Shannon’s entropy (SE) models [14], boosted regression tree (BRT) [17], logistic regression (LR) [18], statistical index (SI) [19] and analytical hierarchy process (AHP) [20,21]. Murmu et al. [22], Abijith et al. [23] and Arefin [24] showed that the latter method coupled with GIS is the most effective model for editing GWPM. The AHP model is a multi-criteria decision-making method based on GIS by assigning weights for different factors based on expert opinion [25,26]. This method has been chosen to be applied in the Tata basin to optimize the various choices for future wells. The performance of AHP was evaluated using the receiver operating characteristics curve (ROC) and confirmed using a geophysical study.

2. Study Area

The Tata basin, with an area of 2567 km2, is situated in the southeast of Morocco. It is located between the coordinates of latitudes 29°34′00″ N to 30°14′00″ N and longitudes 8°32′00″ W to 7°46′00″ W. The basin has a mountainous topography where the altitudes vary from 474 to 2518 m (Figure 1). The continental and semi-arid climate prevailing in this area is characterized by low rainfall (<150 mm/year) [3,27].
During a 70-year rainfall monitoring period (1931–2001), it appears that there were alternate humid periods, with a duration of 8 years, followed by more extended dry periods. Monthly (Figure 2a), two periods can be distinguished: a rainy period from September to March (with a maximum of 18mm in October) followed by a dry period from April to August. It is worth noting that during summer, violent rainstorms occurred, as confirmed by average precipitation of around 7 mm (Figure 2b). The average monthly temperatures registered at the Tata station for the 1985–2020 time-period show an alternating pattern of two contrasting periods with a cool winter (11.49 °C) and a warm and dry summer (33.09 °C) (Figure 2b). The average relative moisture values show significant monthly variations, exceeding 40% in summer. They are over 40% in November, December, January, and February, when low temperatures (<16 °C) are registered (Figure 2b). The average monthly flows of Wadi Tata are highly variable due to the high variability of rainfall in the basin. The average annual discharge is 3.2 m3/s.
Geologically, the basin is occupied by Precambrian formations composed of granite, gabbro, and schist at Ighrem, Tagragra, and Agouliz inliers (Figure 3). The Paleozoic cover occupies a large part of the basin, including volcano-sedimentary deposits progressing to carbonate deposits of “Lower limestones” topped with pelites of purplish-red color “Lie-de-vin” followed by carbonate deposits of “Upper limestones” and schists–calcareous series [28,29,30,31,32,33,34]. Sandstones and quartzites of the Ordovician age occupy the southern part of the basin, followed by Devonian shale and limestone [35]. Alluvial deposits fill the valleys and form alluvial aquifers. Structurally, three major fault directions exist in the basin: NNE–SSW, NE–SW, and E–W [29] (Figure 3). NE–SW sills and doleritic dykes related to the opening of the Atlantic Ocean in the Triassic/Lias period, traverse all these terrains [36].

3. Materials and Methods

The search for favorable sites for implementing wells requires the integration of hydrogeological, geological, and geomorphological characteristics within the framework of the AHP model, which is summarized in the flowchart (Figure 4). The methodological approach consists of the following three steps: (1) identification and elaboration of decision factors, (2) classification and standardization of these factors, and (3) weight assignment of the factor and their combination according to the multi-criteria approach.

3.1. Analytical Hierarchy Process (AHP) Model

The analytical hierarchy process (AHP) is a multi-criteria decision-making approach that evaluates quantitative as well as qualitative criteria and alternatives on the same scale of preferences [37]. It integrates both empirical data and the subjective opinion of specialists to reach a robust decision-making process [38].

3.2. Determination of Decision Factors

To make a more accurate prediction of GWPM, the selection of a model and the quality of the conditioning factors are essential [39]. Based on the research conducted by [14,17,40,41], ten influential factors were used: permeability, slope, topographic wetness index (TWI), plan curvature, profile curvature, stream transport index (STI), stream power index (SPI), lineament density, node density, and drainage density. The combination of these leads to the development of the GWPM, which can be valid only after verification of the consistency ratio (CR).

3.3. Development of Decision Factor Maps

3.3.1. Drainage Density (DD)

DD is the ratio of the sum of the lengths of the wadi sections to the basin’s surface. This criterion reduces infiltration at the expense of runoff [42] and, consequently, reduces the probability of having significant groundwater reserves [43,44]. The DD map is classified into five categories (Figure 5a). The high DD areas are located in the Tagmout syncline and downstream of Tata city.

3.3.2. Lineaments Density (LD)

The digital processing of satellite images to identify and extract lineaments is fundamental in geological and structural mapping. Some examples of the processing carried out in Morocco [45,46,47,48] have proved that applying filtering processes (spatial and directional) facilitate the identification of lineaments. For this purpose, Landsat 8 OLI satellite images, acquired in 2021, are used to extract lineaments. In the Tata basin, Sobel-type filters oriented at 0°, 45°, 90°, and 135° are used to maximize the extraction of lineaments in all directions [49,50].
The LD map represents the segments derived from the superposition of information contained in the four filtered images and the ones corresponding to the faults extracted from geological maps. Three major prominent groups of preferred directions: NNE–SSW, NE–SW, and ENE–WSW, are distinguished. The NE–SW trending lineaments are predominant and can facilitate the accumulation of a large volume of water (Figure 5b).
The LD map includes five classes: very low, low, moderate, high, and very high (Figure 5b). It highlights a difference in this map’s arrangement and concentration of lineaments. Indeed, the core of the Tagmout syncline shows a low density of lineaments occupied by younger sediments (Quaternary). On the other hand, the northwest of the Tagmout and Tata areas, and the northeast of Tata city, show a high density of lineaments, which corresponds to a favorable accumulation of groundwater zones.

3.3.3. Slope (SL)

Geomorphology is the main factor that influences the recharge of aquifers. Thus, the slope greatly influences the runoff to infiltration ratio [44,51,52]. The SL map has been divided into five classes: very low, low, moderate, high, and very high. The slopes greater than 16% are considered areas of very low infiltration [53,54] and cover more than half of the total surface of the basin (Figure 5c). In contrast, less than 16% of slopes are poorly represented and located in the central part and downstream of the basin.

3.3.4. Node Density (ND)

Chowdhury et al. [41] consider that high and moderate density areas of lineaments and their nodes correspond to high fracture permeability regions. These structures provide secondary porosity that plays an important role in water accumulation (Figure 5d). ND varies from 0.35 to 2.05 km/km2. Areas with high ND are located in the northeastern and northwestern parts of Tata city and southeastern Tagmout village. This offers interesting possibilities for groundwater development.

3.3.5. Topographic Wetness Index (TWI)

The TWI is a classic index used to quantify the effect of topography on hydrological processes (including waterlogging). It highlights relatively flat and naturally wet areas due to their position in the basin [55] (Figure 5e). The TWI is calculated using the following equation [56]:
TWI   =   ln ( As tan β )
where As is flow accumulation and β is slope gradient.

3.3.6. Permeability (PER)

Low permeability, consisting of Precambrian bedrock formations, is located in the northern and southeastern parts of the basin (Figure 5f). In contrast, high permeability areas occur along the alluvial deposit surrounding Tata Wadi, locally in the Tagmout syncline, and the downstream part of Jbel Bani. More than 50% of the surface of the basin is characterized by moderate permeability. Regions of high permeability located outside the alluvial zones are characterized by highly fractured bedrock with large openings favorable for storing a significant quantity of water.

3.3.7. Plan Curvature (PLC)

The plan curvature is perpendicular to the direction of the maximum slope. Therefore, it affects the convergence and divergence of the wadi section flows. A positive value indicates that the surface is convex on the side of the treated cell. A negative value indicates that the surface is concave laterally to this cell. A zero value indicates that the surface is linear [57] (Figure 6g).

3.3.8. Profile Curvature (PRC)

The profile curvature is parallel to the direction of the maximum slope. Therefore, it affects the acceleration of water flows at the basin. A negative value indicates that the surface is convex upward at this cell. A positive profile suggests that the surface is concave upward on this cell. A zero value indicates that the surface is linear [58] (Figure 6h).

3.3.9. Stream Power Index (SPI)

The stream power index describes the potential erosion of the flow at a particular point in the basin. As this basin and the slope gradient increase, the amount of water input from upstream areas and the velocity of water flow also increases, thus increasing the power index and the potential for erosion (Figure 6i). The SPI is calculated using the following equation [56]:
SPI = As · tanβ
where As is flow accumulation and β is slope gradient.

3.3.10. Sediment Transport Index (STI)

The sediment transport index combines slope and slope length. This parameter is used to measure the sediment transport capacity of overland flow in the universal soil loss equation [59] (Figure 6j). STI is calculated using the following equation [60]:
STI   = ( As 22.13 ) 0.6 ( sin β 0.0896 ) 1.3
where As is flow accumulation and β is slope gradient.

3.4. Classification and Standardization of Factors

The factor classification is an equally delicate phase and must be carefully carried out. The choice of classes considers the variance of the data, and for better interpretation, their number has been reduced to five [61,62,63,64,65,66]. Standardization is necessary for good multi-criteria analysis as the factors are measured on different scales with other units. A standard range, from 1 to 10, is adopted for this purpose [63,67,68]. A score of 10 is assigned to the “very low” or “very high” classes depending on whether they contribute to the excellent performance of the considered indicator. Contrarily, an opposite score is given to the other extreme classes. Following the same logic, intermediate values are assigned to intermediate classes according to a linear distribution. Table 1 shows the classification of the factors for creating the potentiality map.

3.5. Weighting of the Deciding Factors

The weighting of the deciding factors was performed using the pairwise comparison method in the analytical hierarchy process (AHP) model developed by [69,70,71] and used by [22,72]. Each factor influencing groundwater was given a score from 1 to 4 based on its relative importance to other factors [73] and where 1 and 4 indicate equal and extreme influences of a factor relative to other factors [73]. This method is more appropriate than the direct weight assignment method because its consistency can be checked by calculating the consistency ratio [74]. Moreover, the direct weight assignment method depends only on the decision-maker’s preference [70]. In contrast to the weighting technique based on the arbitrary choice of weights, AHP is based on mathematical calculations that generate standardized weights whose sum is equal to 1. The pairwise comparison matrix and the weights for the selected factors are presented in Table 2 and Table 3.
After defining the matrix, the consistency of the choice made is checked by the consistency ratio (CR), which is calculated by the ratio between the consistency index (CI) and the random index (RI) proposed by [69]. The reasoning is consistent if its consistency ratio is less than or equal to 0.1. Finally, the values of the random index (RI) are given in Table 4. These values are given as a function of the number of parameters compared.
CR = CI RI
where CI is the consistency index and RI is a random index.
CI = λ max n n 1
λmax is the maximum eigenvalue of the comparison matrix and n is the number of factors [69].
The consistency ratio (CR) of the pairwise comparison matrix is 0.07 with a CI calculated for: λ max = 10.92, n = 10, and RI = 1.49. The CR ratio less than 0.1 indicates that the weighting coefficients obtained can be used for water resources potentiality mapping.

3.6. Determination of the GWPM

The weighted method was used in this study to explain the GWPM [44,61,62,72]. It consists of summing the standardized and weighted values of each criterion involved in the development of a given indicator [21,75], according to the following equation:
GWPM = i = 1 n w i · x i
where wi is the weight of the criterion and xi is the standardized value of criterion i.
GWPM = (drainage density × 1.85) + (lineament density × 1.35) + (node density × 1.86) + (slope × 0.96) + (permeability × 1.27) + (profile curvature × 0.49) + (plan curvature × 0.49) + (TWI × 0.55) + (STI × 0.64) + (SPI × 0.53)
The elaboration of the GWPM consists of transferring in space the various values resulting from the summation of each criterion’s standardized and weighted values. All combinations required for the thematic map were performed in Raster mode using the Map algebra tool of the Spatial Analyst module of ArcGIS 10.4.

3.7. Validation of the GWPM

The GWPM is validated by applying the receiver operating characteristic (ROC) [40,76], which will be followed by geophysical prospection. Moreover, the wells not used in the model were retained and then superimposed on the GWPM for its validation. The ROC curve is a graphical representation of the true positive rate on the x–axis and the false positive rate on the y–axis [77,78]. The area under the ROC curve (AUC) demonstrates the accuracy of a prediction system by describing the ability of the system to predict the correct occurrence or non-occurrence of predefined “events” [79,80]. The values of the ROC curve, ranging from 0.5 to 1, are used to determine the accuracy of the applied model. They are classified as: excellent (0.9–1), very good (0.8–0.9), good (0.7–0.8), moderate (0.6–0.7), and poor (0.5–0.6) [81].

4. Results and Discussion

4.1. GWPMusing AHP Model

According to the AHP model, the combination of factors has been integrated into the GIS environment, generating the GWPM (Figure 7). The spatial distribution of the groundwater potential map (GWPM) in the Tata basin is mainly controlled by node density, drainage density, and lineament density, with respective weights of 1.86, 1.85, and 1.35. They confirm the results obtained by [61,82] applied in the Arghren basin in Morocco and the basin in Kurdistan Province in Iran. The weights assigned to the permeability and slope factors are less important than ND, DD, and LD. However, areas with high permeability (alluvial deposits) can comprise areas with high potential, provided the slope is low (less than 16%). On the other hand, plane curvature and profile curvature weakly contribute to water infiltration.
Five classes have been determined on the GWPM in terms of occupied surface: very low (8.67%), low (17.74%), moderate (46.77%), high (19.95%), and very high (6.87%). The areas of high to very high potentiality are mainly located in the downstream part of the basin and on the Tagmout plain where alluvial formations with concave surfaces and a slight slope are dominant. Likewise, the northwestern and central parts of the basin show high potential due to a high lineament density and node density that also coincide with moderately permeable rock outcrops. Lineaments provide favorable conditions for groundwater accumulation and, consequently, groundwater reserves. These results are in agreement with several studies [22,61,82].
The low and very low potential areas (26.41%) are limited to sites with a low density of lineaments and nodes, steep slopes, and low-permeability formations. These areas are located in the southwestern and eastern part of the basin, in the northeast of Tagmout village, and on mountain ridges.
The areas with high water potential near Tata city led us to optimize the choice of future well locations by avoiding areas of high flooding for sustainable security of well equipment. Geophysical prospecting using electrical profiles was applied to confirm and validate these delimited potential zones.

4.2. Validation of GWPM

Validation of results is an essential step in predictive modeling. The validation dataset must be independent of the training dataset, and the ability of a model to predict the potential zones from the validation dataset is the ultimate determinant of the quality and accuracy of models [17]. For this reason, the validation of GWPM was performed using data from existing wells in the basin.
ROC curve analysis was performed to validate the AHP model, comparing the existing 64 wells with GWPM [18,67]. The AUC value of the success rate curve is around 0.804, indicating a prediction accuracy of 80.4% (Figure 8). These results corroborate the results of [83,84,85], who have revealed the effectiveness of this model in identifying GWPM.
The generated GWPM has some limitations due to the difficulty of having detailed geospatial data in the study area. The lack of climatological stations, which can provide a real spatial variation in precipitation, limits the integration of this factor in this study. Thus, it seems necessary to equip this basin with a set of homogeneously distributed stations. Bhattacharya et al. [44], Kumar et al. [86], and Al-Abadi et al. [87] faced the same situation as the Tata basin and were able to generate the GWPM without integrating precipitation. On the other hand, the errors related to image classification are mainly due to their very-low resolution [88]. Mohammadzadeh et al. [89] demonstrated that high-resolution satellite imagery enables more efficient data extraction and detection. The digital elevation model (DEM) and Landsat 8OLI satellite image used have a medium resolution of 30 × 30 m. The high resolution of input data can improve generated factor maps.
It is essential to also draw attention to the assignment of weights to factors, as the peer comparison is based on expert opinion. Inadequate judgment on a factor can influence the assigned weight [90,91]. Finally, the validation of the GWPM, based only on wells located near Tata wadi, is insufficient. The absence of wells in the other parts of the basin constitutes an obstacle to GWPM validation, even though the ROC curve gives 80.4% satisfaction.

4.3. Application of Geophysics: Analysis and Interpretation of Results for Area A (NE of Tata City)

Geophysical prospecting was conducted in two areas close to Tata city to showcase the efficacy of AHP-derived GWPM. In these areas, geophysical validation using the geo-electrical profiles methods was conducted. An initial area (A) was detailed to understand the geophysical behavior of the most permeable geological formations. It was chosen as a calibration area since it contains wells with different yields. Subsequently, once the geophysical parameters were defined, the same prospecting was applied to the NE area of Tata city (area B) to determine new potential sites to propose to the water agency. For the first calibration area (A), nine similar NW–SE electrical profiles were executed, separated by 500 m. For each resistivity measurement, the distance between the two transmitting electrodes was kept at 300 m. The measurements were made with a 50-m step.
The apparent resistivity values correspond to the geological formations located between 50 and 60 m depth. A resistivity variation map has been produced, showing a significant variation in this parameter that ranges from 85 to 4438 Ohm.m (Figure 9a). High resistivity is located at the NE and SW extremities of the prospected area. They probably correspond to the Lower Cambrian limestone. A low resistivity anomaly separates them with a NW–SE direction, where all the wells are embedded. These are probably basement formations (schists) covered by alluvial terraces which cover the passage zones of the hydrogeological lineaments (faults) detected through the processing of satellite images. Further, as we move away from the NW–SE lineament (Figure 9a), the well yields become low and the resistivity increases (Figure 9b). The probability of high yield increases with proximity to the hydrogeological lineaments, which are represented by low resistivity.
A high potential zone for water resources located NE of Tata city (area B) (Figure 10a) was also subjected to geophysical prospecting by electrical profiles to optimize the choice of future wells. The apparent resistivity map shows values ranging from 55 to 2000 Ohm.m (Figure 10b).
A decrease in resistivity values compared to the first prospected area (A) was observed due to clay and silty formations interspersed between the limestone banks. The conductive anomalies that may be hydrogeologically interesting are:
  • Anomaly 1, which corresponds to the basement formations at a depth affected by faults and hydrogeological lineaments with an alluvial cover above and
  • Anomaly 2, which corresponds to NW–SE faults affecting the limestone formations without alluvial cover.
Three points have been proposed for the establishment of future wells in the junction zones of hydrogeological lineaments with low resistivity values. In addition, they are located in areas with a low risk of flooding and can be tested by the local water agency.

5. Conclusions

Elaboration of GWPM in the Tata basin, located in southeastern Morocco, was carried out using GIS mapping tools, remote sensing, an AHP model, and geophysical prospection. According to the results of the applied model, factors influencing water availability are mainly geomorphological since they correspond to node density, drainage density, and lineament density factors. Furthermore, the validation by the ROC curve showed the model’s efficiency despite using low-resolution satellite data.
The geo-electrical prospection has been used to locate future well sites outside flood risk areas. Moreover, the proposed study model provides maximum optimization of time and financial means since it enables, by using accessible data, the precise location of well sites in a short time.

Author Contributions

Conceptualization, F.Z.E., S.B., A.B., M.E.S., M.A. (Mohamed Aadraoui), M.A. (Mohamed Abioui) and M.O.; methodology, F.Z.E., M.A. (Mohamed Abioui) and K.S.S.; software, F.Z.E. and M.O.; validation, M.A. (Mohamed Abioui) and K.S.S.; formal analysis, F.Z.E., S.B., A.B., M.E.S., M.A. (Mohamed Aadraoui) and M.O.; investigation, F.Z.E.; resources, F.Z.E.; data curation, F.Z.E.; writing—original draft preparation, F.Z.E., S.B., A.B., M.E.S., M.A. (Mohamed Aadraoui) and M.O.; writing—review and editing, M.A. (Mohamed Abioui), K.A., M.S.F. and K.S.S.; visualization, F.Z.E.; supervision, S.B. and M.A. (Mohamed Abioui); project administration, M.A. (Mohamed Abioui); funding acquisition, K.A. and M.S.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Researchers Supporting Project number (RSP–2021/249), King Saud University, Riyadh, Saudi Arabia. The funding source was not involved in any way in the conceptualization, writing, and submission of this work.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank all reviewers and editor for their valuable comments on this paper.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

AHPAnalytical hierarchy process
PERPermeability
SLSlope
NDNode density
TWITopographic wetness index
STIStream transport index
SPIStream power index
PLCPlan curvature
PRCProfile curvature
DDDrainage density
GISGeographic Information System
RSTeledetection
GWPMGroundwater potential map
ROCReceiver operating characteristics
AUCArea under the curve
DEMDigital elevation model
FRFrequency ratio
EBFEvidential belief function
SEShannon’s entropy
BRTBoosted regression tree
LRLogistic regression
SIStatistical index
CRConsistency ratio
RIRandom index
CIConsistency index

References

  1. Pouffary, S.; De Laboulaye, G.; Antonini, A.; Quefelec, S.; Dittrick, L. Les Défis Du Changement Climatique En Méditerranée. Transformer Les Contraintes En Opportunités D’agir; ENERGIES 2050: Biot, France, 2016. [Google Scholar]
  2. DRPE. Etude D’actualisation Du Plan Directeur D’aménagement Intégré Des Ressources En Eau (PDAIRE) Du Bassin Hydraulique Du Draa; Secrétariat d’Etat auprès du Ministère de l’Energie, des Mines, de l’Eau et de l’Environnement, chargé de l’Eau et de l’Environnement: Rabat, Morocco, 2008.
  3. Echogdali, F.Z.; Boutaleb, S.; Jauregui, J.; Elmouden, A. Cartography of Flooding Hazard in Semi–Arid Climate: The Case of Tata Valley (South–East of Morocco). J. Geogr. Nat. Disast. 2018, 8, 214. [Google Scholar] [CrossRef]
  4. Panahi, M.; Sadhasivam, N.; Pourghasemi, H.R.; Rezaie, F.; Lee, S. Spatial prediction of groundwater potential mapping based on convolutional neural network (CNN) and support vector regression (SVR). J. Hydrol. 2020, 588, 125033. [Google Scholar] [CrossRef]
  5. Sander, P.; Minor, T.B.; Chesley, M.M. Ground-Water Exploration Based on Lineament Analysis and Reproducibility Tests. Groundwater 1997, 35, 888–894. [Google Scholar] [CrossRef]
  6. Israil, M.; Al-Hadithi, M.; Singhal, D.C. Application of a resistivity survey and geographical information system (GIS) analysis for hydrogeological zoning of a piedmont area, Himalayan foothill region, India. Appl. Hydrogeol. 2005, 14, 753–759. [Google Scholar] [CrossRef]
  7. Wieland, M.; Pittore, M. A Spatio-Temporal Building Exposure Database and Information Life-Cycle Management Solution. ISPRS Int. J. Geo-Inf. 2017, 6, 114. [Google Scholar] [CrossRef]
  8. Souissi, D.; Msaddek, M.H.; Zouhri, L.; Chenini, I.; El May, M.; Dlala, M. Mapping groundwater recharge potential zones in arid region using GIS and Landsat approaches, southeast Tunisia. Hydrol. Sci. J. 2018, 63, 251–268. [Google Scholar] [CrossRef]
  9. Ghosh, D.; Mandal, M.; Banerjee, M.; Karmakar, M. Impact of hydro-geological environment on availability of groundwater using analytical hierarchy process (AHP) and geospatial techniques: A study from the upper Kangsabati river basin. Groundw. Sustain. Dev. 2020, 11, 100419. [Google Scholar] [CrossRef]
  10. Khan, M.Y.A.; ElKashouty, M.; Tian, F. Mapping Groundwater Potential Zones Using Analytical Hierarchical Process and Multicriteria Evaluation in the Central Eastern Desert, Egypt. Water 2022, 14, 1041. [Google Scholar] [CrossRef]
  11. Hussein, A.-A.; Govindu, V.; Nigusse, A.G.M. Evaluation of groundwater potential using geospatial techniques. Appl. Water Sci. 2016, 7, 2447–2461. [Google Scholar] [CrossRef] [Green Version]
  12. Stafford, K.W.; Rosales-Lagarde, L.; Boston, P.J. Castile evaporite karst potential map of the Gypsum Plain, Eddy County, New Mexico and Culberson County, Texas: A GIS methodological comparison. J. Cave Karst Stud. 2008, 70, 35–46. [Google Scholar]
  13. Zghibi, A.; Mirchi, A.; Msaddek, M.H.; Merzougui, A.; Zouhri, L.; Taupin, J.-D.; Chekirbane, A.; Chenini, I.; Tarhouni, J. Using Analytical Hierarchy Process and Multi-Influencing Factors to Map Groundwater Recharge Zones in a Semi-arid Mediterranean Coastal Aquifer. Water 2020, 12, 2525. [Google Scholar] [CrossRef]
  14. Naghibi, S.A.; Pourghasemi, H.R.; Pourtaghi, Z.S.; Rezaei, A. Groundwater qanat potential mapping using frequency ratio and Shannon’s entropy models in the Moghan watershed, Iran. Earth Sci. Inform. 2014, 8, 171–186. [Google Scholar] [CrossRef]
  15. Regmi, A.D.; Devkota, K.C.; Yoshida, K.; Pradhan, B.; Pourghasemi, H.R.; Kumamoto, T.; Akgun, A. Application of frequency ratio, statistical index, and weights–of–evidence models and their comparison in landslide susceptibility mapping in Central Nepal Himalaya. Arab. J. Geosci. 2014, 7, 725–742. [Google Scholar] [CrossRef]
  16. Mogaji, K.A.; Lim, H.S.; Abdullah, K. Regional prediction of groundwater potential mapping in a multifaceted geology terrain using GIS-based Dempster–Shafer model. Arab. J. Geosci. 2014, 8, 3235–3258. [Google Scholar] [CrossRef]
  17. Naghibi, S.A.; Pourghasemi, H.R.; Dixon, B. GIS–based groundwater potential mapping using boosted regression tree, classification and regression tree, and random forest machine learning models in Iran. Environ. Monit. Assess. 2016, 188, 44. [Google Scholar] [CrossRef] [PubMed]
  18. Pourtaghi, Z.S.; Pourghasemi, H.R. GIS–based groundwater spring potential assessment and mapping in the Birjand Township, southern Khorasan Province. Iran. Hydrogeol. J. 2014, 22, 643–662. [Google Scholar] [CrossRef]
  19. Falah, F.; Nejad, S.G.; Rahmati, O.; Daneshfar, M.; Zeinivand, H. Applicability of generalized additive model in groundwater potential modelling and comparison its performance by bivariate statistical methods. Geocarto Int. 2016, 32, 1069–1089. [Google Scholar] [CrossRef]
  20. Shao, Z.; Huq, E.; Cai, B.; Altan, O.; Li, Y. Integrated remote sensing and GIS approach using Fuzzy-AHP to delineate and identify groundwater potential zones in semi-arid Shanxi Province, China. Environ. Model. Softw. 2020, 134, 104868. [Google Scholar] [CrossRef]
  21. Kumar, T.; Gautam, A.K.; Kumar, T. Appraising the accuracy of GIS–based multi–criteria decision making technique for delineation of groundwater potential zones. Water Resour. Manag. 2014, 28, 4449–4466. [Google Scholar] [CrossRef]
  22. Murmu, P.; Kumar, M.; Lal, D.; Sonker, I.; Singh, S.K. Delineation of groundwater potential zones using geospatial techniques and analytical hierarchy process in Dumka district, Jharkhand, India. Groundw. Sustain. Dev. 2019, 9, 100239. [Google Scholar] [CrossRef]
  23. Abijith, D.; Saravanan, S.; Singh, L.; Jennifer, J.J.; Saranya, T.; Parthasarathy, K. GIS-based multi-criteria analysis for identification of potential groundwater recharge zones—A case study from Ponnaniyaru watershed, Tamil Nadu, India. HydroResearch 2020, 3, 1–14. [Google Scholar] [CrossRef]
  24. Arefin, R. Groundwater potential zone identification at Plio-Pleistocene elevated tract, Bangladesh: AHP-GIS and remote sensing approach. Groundw. Sustain. Dev. 2020, 10, 100340. [Google Scholar] [CrossRef]
  25. Al-Ruzouq, R.; Shanableh, A.; Merabtene, T.; Siddique, M.; Khalil, M.A.; Idris, A.; Almulla, E. Potential groundwater zone mapping based on geo-hydrological considerations and multi-criteria spatial analysis: North UAE. CATENA 2018, 173, 511–524. [Google Scholar] [CrossRef]
  26. Jothibasu, A.; Anbazhagan, S. Modeling groundwater probability index in Ponnaiyar River basin of South India using analytic hierarchy process. Model. Earth Syst. Environ. 2016, 2, 109. [Google Scholar] [CrossRef] [Green Version]
  27. Echogdali, F.Z.; Boutaleb, S.; Taia, S.; Ouchchen, M.; Id–Belqas, M.; Kpan, R.B.; Abioui, M.; Aswathi, J.; Sajinkumar, K.S. Assessment of soil erosion risk in a semi–arid climate watershed using SWAT model: Case of Tata basin, South–East of Morocco. Appl. Water Sci. 2022, 12, 137. [Google Scholar] [CrossRef]
  28. Choubert, G. Histoire géologique du précambrien de l’Anti–Atlas. Notes Mem. Serv. Geol. Maroc 1963, 162, 352. [Google Scholar]
  29. Faik, F.; Belfoul, M.A.; Bouabdelli, M.; Hassenforder, B. Les structures de la couverture Néoprotérozoïque terminal et Paléozoïque de la région de Tata, Anti–Atlas centre–occidental, Maroc: Déformation polyphasée, ou interactions socle/couverture pendant l’orogenèse hercynienne? J. Afr. Earth Sci. 2001, 32, 765–776. [Google Scholar] [CrossRef]
  30. Algouti, A.; Algouti, A.; Chbani, B.; Zaim, M. Sédimentation et volcanisme synsédimentaire de la série de base de l’Adoudounien infra–cambrien à travers deux exemples de l’Anti–Atlas du Maroc. J. Afr. Earth Sci. 2001, 32, 541–556. [Google Scholar] [CrossRef]
  31. Benssaou, M.; Hamoumi, N. The western Anti-Atlas of Morocco: Sedimentological and palæogeographical formation studies in the Early Cambrian. J. Afr. Earth Sci. 2001, 32, 351–372. [Google Scholar] [CrossRef]
  32. Benssaou, M.; Hamoumi, N. Le graben de l’Anti–Atlas occidental (Maroc): Contrôle tectonique de la paléogéographie et des séquences au Cambrien inférieur. Comptes Rendus Geosci. 2003, 335, 297–305. [Google Scholar] [CrossRef]
  33. Ouchchen, M.; Boutaleb, S.; El Azzab, D.; Abioui, M.; Mickus, K.L.; Miftah, A.; Echogdali, F.Z.; Dadi, B. Structural interpretation of the Igherm region (Western Anti Atlas, Morocco) from an aeromagnetic analysis: Implications for copper exploration. J. Afr. Earth Sci. 2021, 176, 104140. [Google Scholar] [CrossRef]
  34. Echogdali, F.Z.; Boutaleb, S.; Abia, E.H.; Ouchchen, M.; Dadi, B.; Id-Belqas, M.; Abioui, M.; Pham, L.T.; Abu-Alam, T.; Mickus, K.L. Mineral prospectivity mapping: A potential technique for sustainable mineral exploration and mining activities—a case study using the copper deposits of the Tagmout basin, Morocco. Geocarto Int. 2021, 1–22. [Google Scholar] [CrossRef]
  35. Wendt, J. Disintegration of the continental margin of northwestern Gondwana: Late Devonian of the eastern Anti-Atlas (Morocco). Geology 1985, 13, 815. [Google Scholar] [CrossRef]
  36. Sahabi, M.; Aslanian, D.; Olivet, J.-L. Un nouveau point de départ pour l’histoire de l’Atlantique central. C. R. Geosci. 2004, 336, 1041–1052. [Google Scholar] [CrossRef] [Green Version]
  37. Ishizaka, A.; Labib, A. Review of the main developments in the analytic hierarchy process. Expert Syst. Appl. 2011, 38, 14336–14345. [Google Scholar] [CrossRef] [Green Version]
  38. Şener, Ş.; Sener, E.; Karagüzel, R. Solid waste disposal site selection with GIS and AHP methodology: A case study in Senirkent–Uluborlu (Isparta) Basin, Turkey. Environ. Monit Assess. 2011, 173, 533–554. [Google Scholar] [CrossRef]
  39. Pradhan, B. A comparative study on the predictive ability of the decision tree, support vector machine and neuro-fuzzy models in landslide susceptibility mapping using GIS. Comput. Geosci. 2013, 51, 350–365. [Google Scholar] [CrossRef]
  40. Chen, W.; Li, H.; Hou, E.; Wang, S.; Wang, G.; Panahi, M.; Li, T.; Peng, T.; Guo, C.; Niu, C.; et al. GIS–based groundwater potential analysis using novel ensemble weights–of–evidence with logistic regression and functional tree models. Sci. Total Environ. 2018, 634, 853–867. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Chowdhury, A.; Jha, M.K.; Chowdary, V.M.; Mal, B.C. Integrated remote sensing and GIS-based approach for assessing groundwater potential in West Medinipur district, West Bengal, India. Int. J. Remote Sens. 2009, 30, 231–250. [Google Scholar] [CrossRef]
  42. Laurent, F. Outils De Modélisation Spatiale Pour La Gestion Intégrée Des Ressources En Eau: Application Aux Schémas D’Aménagement Et De Gestion Des Eaux. Ph.D. Thesis, École Nationale Supérieure des Mines de Saint–Étienne/École Nationale Supérieure des Mines de Paris, Lyon, Paris, 1996. [Google Scholar]
  43. Khan, M.Y.A.; ElKashouty, M.; Subyani, A.M.; Tian, F.; Gusti, W. GIS and RS intelligence in delineating the groundwater potential zones in Arid Regions: A case study of southern Aseer, southwestern Saudi Arabia. Appl. Water Sci. 2021, 12, 3. [Google Scholar] [CrossRef]
  44. Bhattacharya, S.; Das, S.; Das, S.; Kalashetty, M.; Warghat, S.R. An integrated approach for mapping groundwater potential applying geospatial and MIF techniques in the semiarid region. Environ. Dev. Sustain. 2020, 23, 495–510. [Google Scholar] [CrossRef]
  45. Emran, A.; Chorowicz, J.; Cervelle, B.; Lyberis, N.; Tamain, G.; Alem, E.M. Cartographie géologique et analyse de la fracturation du sud de l’Anti–Atlas central (Maroc) à partir d’une image Landsat MSS. Photo Interprétation 1988, 27, 1–10. [Google Scholar]
  46. Boutaleb, S.; Boualoul, M.; Oudra, M.; Bouchaou, L.; Dindane, K. Apports du traitement d’image et de la géophysique à l’étude des ressources en eau en milieu fissuré: Cas de l’Anti–Atlas marocain. Afr. Geosci. Rev. 2008, 15, 129–141. [Google Scholar]
  47. Boutaleb, S.; Hammichi, F.; Tabyaoui, H.; Bouchaou, L.; Dindane, K. Détermination des écoulements préférentiels en zone karstique (Tafrata, Maroc), Apport des données satellitaires SAR ERS–1 et Landsat ETM+ et de la prospection géophysique. J. Water Sci. 2009, 22, 407–419. [Google Scholar]
  48. Chotin, P.; Ait Brahim, L.; Deffontaines, B.; Rudant, J.P.; Chaouni, A. Apports des données ERS1–SAR sur la reconnaissance du réseau de failles dans la péninsule de Tanger, Maroc. Photo Interprétation 1995, 33, 146–152. [Google Scholar]
  49. Bouchaou, L.; Boutaleb, S.; Boualoul, M.; Oudra, M. Application of remote-sensing and surface geophysics for groundwater prospecting in a hard rock terrain, Morocco. In Applied Groundwater Studies in Africa; CRC Press: London, UK, 2008; pp. 215–230. [Google Scholar] [CrossRef]
  50. Hssaisoune, M.; Boutaleb, S.; Bouchaou, L.; Benssaou, M.; Tagma, T. Use of remote sensing and electrical resistivity tomography to determine Tidsi spring recharge and underground drainage. Eur. Water 2017, 57, 429–434. [Google Scholar]
  51. Ozdemir, A. Using a binary logistic regression method and GIS for evaluating and mapping the groundwater spring potential in the Sultan Mountains (Aksehir, Turkey). J. Hydrol. 2011, 405, 123–136. [Google Scholar] [CrossRef]
  52. Selvam, S.; Magesh, N.S.; Sivasubramanian, P.; Soundranayagam, J.P.; Manimaran, G.; Seshunarayana, T. Deciphering of groundwater potential zones in Tuticorin, Tamil Nadu, using remote sensing and GIS techniques. J. Geol. Soc. India 2014, 84, 597–608. [Google Scholar] [CrossRef]
  53. Sharma, K.D.; Singh, H.P.; Pareek, O.P. Rainwater infiltration into a bare loamy sand. Hydrol. Sci. J. 1983, 28, 417–424. [Google Scholar] [CrossRef] [Green Version]
  54. Fox, D.; Bryan, R.; Price, A. The influence of slope angle on final infiltration rate for interrill conditions. Geoderma 1997, 80, 181–194. [Google Scholar] [CrossRef]
  55. Lee, S.; Kim, Y.-S.; Oh, H.-J. Application of a weights-of-evidence method and GIS to regional groundwater productivity potential mapping. J. Environ. Manag. 2012, 96, 91–105. [Google Scholar] [CrossRef] [PubMed]
  56. Moore, I.D.; Grayson, R.B.; Ladson, A.R. Digital terrain modelling: A review of hydrological, geomorphological, and biological applications. Hydrol. Process. 1991, 5, 3–30. [Google Scholar] [CrossRef]
  57. Kimerling, A.J.; Buckley, A.R.; Muehrcke, P.C.; Muehrcke, J.O. Map Use: Reading, Analysis, Interpretation, 8th ed.; ESRI Press Academic: Redlands, CA, USA, 2016. [Google Scholar]
  58. Oh, H.J.; Lee, S. Cross-validation of logistic regression model for landslide susceptibility mapping at Geneoung areas, Korea. Disaster Adv. 2010, 3, 44–55. [Google Scholar]
  59. Wischmeier, W.H.; Smith, D.D. Predicting Rainfall Erosion Losses: A Guide to Conservation Planning; United States Department of Agriculture: Washington, DC, USA, 1978.
  60. Moore, I.D.; Burch, G.J. Sediment Transport Capacity of Sheet and Rill Flow: Application of Unit Stream Power Theory. Water Resour. Res. 1986, 22, 1350–1360. [Google Scholar] [CrossRef]
  61. Rahmati, O.; Nazari Samani, A.; Mahdavi, M.; Pourghasemi, H.R.; Zeinivand, H. Groundwater potential mapping at Kurdistan region of Iran using analytic hierarchy process and GIS. Arab. J. Geosci. 2015, 8, 7059–7071. [Google Scholar] [CrossRef]
  62. Saranya, T.; Saravanan, S. Groundwater potential zone mapping using analytical hierarchy process (AHP) and GIS for Kancheepuram District, Tamilnadu, India. Model. Earth Syst. Environ. 2020, 6, 1105–1122. [Google Scholar] [CrossRef]
  63. Maity, D.K.; Mandal, S. Identification of groundwater potential zones of the Kumari river basin, India: An RS & GIS based semi-quantitative approach. Environ. Dev. Sustain. 2017, 21, 1013–1034. [Google Scholar] [CrossRef]
  64. Ikirri, M.; Faik, F.; Boutaleb, S.; Echogdali, F.Z.; Abioui, M.; Al-Ansari, N. Application of HEC-RAS/WMS and FHI models for extreme hydrological events under climate change in the Ifni River arid watershed from Morocco. In Climate and Land Use Impacts on Natural and Artificial Systems; Elsevier: Amsterdam, The Netherlands, 2021; pp. 251–270. [Google Scholar] [CrossRef]
  65. Echogdali, F.Z.; Boutaleb, S.; Elmouden, A.; Ouchchen, M. Assessing Flood Hazard at River Basin Scale: Comparison between HECRAS-WMS and Flood Hazard Index (FHI) Methods Applied to El Maleh Basin, Morocco. J. Water Resour. Prot. 2018, 10, 957–977. [Google Scholar] [CrossRef] [Green Version]
  66. Echogdali, F.Z.; Kpan, R.B.; Ouchchen, M.; Id-Belqas, M.; Dadi, B.; Ikirri, M.; Abioui, M.; Boutaleb, S. Spatial Prediction of Flood Frequency Analysis in a Semi-Arid Zone: A Case Study from the Seyad Basin (Guelmim Region, Morocco). In Geospatial Technology for Landscape and Environmental Management; Springer: Berlin, Germany, 2022; pp. 49–71. [Google Scholar] [CrossRef]
  67. Razandi, Y.; Pourghasemi, H.R.; Neisani, N.S.; Rahmati, O. Application of analytical hierarchy process, frequency ratio, and certainty factor models for groundwater potential mapping using GIS. Earth Sci. Inform. 2015, 8, 867–883. [Google Scholar] [CrossRef]
  68. Dar, T.; Rai, N.; Bhat, A. Delineation of potential groundwater recharge zones using analytical hierarchy process (AHP). Geol. Ecol. Landsc. 2020, 5, 292–307. [Google Scholar] [CrossRef] [Green Version]
  69. Saaty, T.L. A scaling method for priorities in hierarchical structures. J. Math. Psychol. 1977, 15, 234–281. [Google Scholar] [CrossRef]
  70. Saaty, T.L. How to make a decision: The analytic hierarchy process. Eur. J. Oper. Res. 1990, 48, 9–26. [Google Scholar] [CrossRef]
  71. Saaty, T.L. Decision making with the analytic hierarchy process. Int. J. Serv. Sci. 2008, 1, 83–98. [Google Scholar] [CrossRef] [Green Version]
  72. Sutradhar, S.; Mondal, P.; Das, N. Delineation of groundwater potential zones using MIF and AHP models: A micro–level study on Suri Sadar Sub–Division, Birbhum District, West Bengal, India. Groundw. Sustain. Dev. 2021, 12, 100547. [Google Scholar] [CrossRef]
  73. Saaty, T.L. The Analytic Hierarchy Process: Planning, Priority Setting, Resources Allocation; McGraw: New York, NY, USA, 1980. [Google Scholar] [CrossRef]
  74. Kumar, V.; Jain, K. Site suitability evaluation for urban development using remote sensing, GIS and analytic hierarchy process (AHP). In Proceedings of the International Conference on Computer Vision and Image Processing; Springer: Singapore, 2017; pp. 377–388. [Google Scholar] [CrossRef]
  75. Vellaikannu, A.; Palaniraj, U.; Karthikeyan, S.; Senapathi, V.; Viswanathan, P.M.; Sekar, S. Identification of groundwater potential zones using geospatial approach in Sivagangai district, South India. Arab. J. Geosci. 2021, 14, 8. [Google Scholar] [CrossRef]
  76. Rahmati, O.; Pourghasemi, H.R.; Melesse, A.M. Application of GIS-based data driven random forest and maximum entropy models for groundwater potential mapping: A case study at Mehran Region, Iran. CATENA 2016, 137, 360–372. [Google Scholar] [CrossRef]
  77. Pourghasemi, H.R.; Moradi, H.R.; Aghda, S.M.F. Landslide susceptibility mapping by binary logistic regression, analytical hierarchy process, and statistical index models and assessment of their performances. Nat. Hazards 2013, 69, 749–779. [Google Scholar] [CrossRef]
  78. Chen, W.; Peng, J.; Hong, H.; Shahabi, H.; Pradhan, B.; Liu, J.; Zhu, A.X.; Pei, X.; Duan, Z. Landslide susceptibility modelling using GIS–based machine learning techniques for Chongren County, Jiangxi Province, China. Sci. Total Environ. 2018, 626, 1121–1135. [Google Scholar] [CrossRef]
  79. Bui, D.T.; Pradhan, B.; Lofman, O.; Revhaug, I.; Dick, O.B. Landslide susceptibility mapping at Hoa Binh province (Vietnam) using an adaptive neuro-fuzzy inference system and GIS. Comput. Geosci. 2012, 45, 199–211. [Google Scholar] [CrossRef]
  80. Jaafari, A.; Najafi, A.; Pourghasemi, H.R.; Rezaeian, J.; Sattarian, A. GIS-based frequency ratio and index of entropy models for landslide susceptibility assessment in the Caspian forest, northern Iran. Int. J. Environ. Sci. Technol. 2014, 11, 909–926. [Google Scholar] [CrossRef] [Green Version]
  81. Yesilnacar, E.K. The Application of Computational Intelligence to Landslide Susceptibility Mapping in Turkey. Ph.D. Thesis, University of Melbourne, Melbourne, Australia, 2005. [Google Scholar]
  82. Benjmel, K.; Amraoui, F.; Boutaleb, S.; Ouchchen, M.; Tahiri, A.; Touab, A. Mapping of groundwater potential zones in crystalline terrain using remote sensing, GIS techniques, and multicriteria data analysis (Case of the Ighrem Region, Western Anti-Atlas, Morocco). Water 2020, 12, 471. [Google Scholar] [CrossRef] [Green Version]
  83. Pal, S.C.; Ghosh, C.; Chowdhuri, I. Assessment of groundwater potentiality using geospatial techniques in Purba Bardhaman district, West Bengal. Appl. Water Sci. 2020, 10, 221. [Google Scholar] [CrossRef]
  84. Das, B.; Pal, S.C. Assessment of groundwater vulnerability to over-exploitation using MCDA, AHP, fuzzy logic and novel ensemble models: A case study of Goghat-I and II blocks of West Bengal, India. Environ. Earth Sci. 2020, 79, 104. [Google Scholar] [CrossRef]
  85. Pathmanandakumar, V.; Thasarathan, N.; Ranagalage, M. An Approach to Delineate Potential Groundwater Zones in Kilinochchi District, Sri Lanka, Using GIS Techniques. ISPRS Int. J. Geo-Inf. 2021, 10, 730. [Google Scholar] [CrossRef]
  86. Kumar, V.A.; Mondal, N.C.; Ahmed, S. Identification of Groundwater Potential Zones Using RS, GIS and AHP Techniques: A Case Study in a Part of Deccan Volcanic Province (DVP), Maharashtra, India. J. Indian Soc. Remote Sens. 2020, 48, 497–511. [Google Scholar] [CrossRef]
  87. Al-Abadi, A.M. Modeling of groundwater productivity in northeastern Wasit Governorate, Iraq using frequency ratio and Shannon’s entropy models. Appl. Water Sci. 2015, 7, 699–716. [Google Scholar] [CrossRef] [Green Version]
  88. Akula, A.; Singh, A.; Ghosh, R.; Kumar, S.; Sardana, H.K. Target recognition in infrared imagery using convolutional neural network. In Proceedings of the International Conference on Computer Vision and Image Processing; Springer: Singapore, 2017; pp. 25–34. [Google Scholar] [CrossRef]
  89. Mohammadzadeh, A.; Zoej, M.J.V.; Tavakoli, A. Automatic main road extraction from high resolution satellite imageries by means of particle swarm optimization applied to a fuzzy-based mean calculation approach. J. Indian Soc. Remote Sens. 2009, 37, 173–184. [Google Scholar] [CrossRef]
  90. Benjmel, K.; Amraoui, F.; Aydda, A.; Tahiri, A.; Yousif, M.; Pradhan, B.; Abdelrahman, K.; Fnais, M.S.; Abioui, M. A multidisciplinary approach for groundwater potential mapping in a fractured semi-arid terrain (Kerdous Inlier, Western Anti-Atlas, Morocco). Water 2022, 14, 1553. [Google Scholar] [CrossRef]
  91. Aswathi, J.; Sajinkumar, K.; Rajaneesh, A.; Oommen, T.; Bouali, E.; Binojkumar, R.; Rani, V.; Thomas, J.; Thrivikramji, K.; Ajin, R.; et al. Furthering the precision of RUSLE soil erosion with PSInSAR data: An innovative model. Geocarto Int. 2022, 1–22. [Google Scholar] [CrossRef]
Figure 1. Location of the Tata basin with digital elevation model (DEM).
Figure 1. Location of the Tata basin with digital elevation model (DEM).
Water 14 02393 g001
Figure 2. (a) Average annual rainfall registered at the Tata station (1931–2001) and (b) average monthly rainfall, temperature, and humidity at the Tata station (1985–2020).
Figure 2. (a) Average annual rainfall registered at the Tata station (1931–2001) and (b) average monthly rainfall, temperature, and humidity at the Tata station (1985–2020).
Water 14 02393 g002
Figure 3. Geological map of the Tata basin.
Figure 3. Geological map of the Tata basin.
Water 14 02393 g003
Figure 4. Flowchart of the methodology used in the Tata basin for the implementation of future wells.
Figure 4. Flowchart of the methodology used in the Tata basin for the implementation of future wells.
Water 14 02393 g004
Figure 5. Variation maps of the factor influencing the groundwater potential map (part 1): (a) drainage density (DD); (b) lineament density (LD); (c) slope (SL); (d) node density (ND); (e) topographic wetness index (TWI); and (f) permeability (PER).
Figure 5. Variation maps of the factor influencing the groundwater potential map (part 1): (a) drainage density (DD); (b) lineament density (LD); (c) slope (SL); (d) node density (ND); (e) topographic wetness index (TWI); and (f) permeability (PER).
Water 14 02393 g005
Figure 6. Variation maps of the factor influencing the groundwater potential map (part 2): (g) plan curvature (PLC); (h) profile curvature (PRC); (i) stream power index (SPI); and (j) sediment transport index (STI).
Figure 6. Variation maps of the factor influencing the groundwater potential map (part 2): (g) plan curvature (PLC); (h) profile curvature (PRC); (i) stream power index (SPI); and (j) sediment transport index (STI).
Water 14 02393 g006
Figure 7. Groundwater potential map of the Tata basin.
Figure 7. Groundwater potential map of the Tata basin.
Water 14 02393 g007
Figure 8. ROC curves of GWPM using AHP.
Figure 8. ROC curves of GWPM using AHP.
Water 14 02393 g008
Figure 9. (a) Variation map of apparent resistivities in Ohm.m in area A deduced from the electrical traces for a length AB = 300 m and (b) correlation between apparent resistivity values in Ohm.m and flow rates in m3/s of the wells in area A.
Figure 9. (a) Variation map of apparent resistivities in Ohm.m in area A deduced from the electrical traces for a length AB = 300 m and (b) correlation between apparent resistivity values in Ohm.m and flow rates in m3/s of the wells in area A.
Water 14 02393 g009
Figure 10. (a) Groundwater potential map in area B and (b) variation map of the apparent resistivities in Ohm.m in area B deduced from the electrical traces for a length AB = 300 m.
Figure 10. (a) Groundwater potential map in area B and (b) variation map of the apparent resistivities in Ohm.m in area B deduced from the electrical traces for a length AB = 300 m.
Water 14 02393 g010
Table 1. Classification and standardization of groundwater potentiality factors.
Table 1. Classification and standardization of groundwater potentiality factors.
Factor (units)ClassRatingWeightFactor (units)ClassRatingWeight
Drainage density (m/km2)1.56–1.8511.85TWI5.68–8.7810.55
1.25–1.5638.78–10.423
0.84–1.25510.42–12.695
0.59–0.84812.69–16.288
0.33–0.591016.28–26.4710
Node density (m/km2)0–0.2111.86STI1568–370210.64
0.21–0.493726–15683
0.49–0.825275–7265
0.82–1.15858–2758
1.15–2.04100–5810
Lineament density (m/km2)0–0.3211.35SPI509,811–1,108,33110.53
0.32–0.643295,555–509,8113
0.64–0.9755998–295,5555
0.97–1.378160–59988
1.37–2.19100–16010
PermeabilityLow permeability11.27Plan curvature0.42–17.8710.49
Medium permeability3−0.55–−0.423
High permeability8−13.46–−0.555
Slope (%)16<10.96Profile curvature–16.98–−0.6910.49
12–163–0.69–0.443
8–125−0.44–−15.215
4–88
0–410
Table 2. Comparison matrix by pair of factors.
Table 2. Comparison matrix by pair of factors.
FactorDDNDLDPERSLTWISTISPIPLCPRC
DD1213233233
ND1/2141224444
LD11/411/2323333
PER1/3121232222
SL1/21/21/31/2142222
TWI1/31/21/21/31/411/21/222
STI1/31/41/31/21/221122
SPI1/21/41/31/21/22111/21/2
PLC1/31/41/31/21/21/21/2211
PRC1/31/41/31/21/21/21/2211
Note: DD, drainage density; ND, node density; LD, lineament density; PER, permeability; SL, slope; TWI, topographic wetness index; STI, sediment transport index; SPI, stream power index; PLC, plan curvature; and PRC, profile curvature.
Table 3. The weighting of the factors.
Table 3. The weighting of the factors.
Factor DDNDLDPERSLTWISTISPIPLCPRCWeight
DD0.190.320.100.360.160.150.170.100.150.151.85
ND0.100.160.390.120.160.100.230.210.200.201.86
LD0.190.040.100.060.240.100.170.150.150.151.35
PER0.060.160.200.120.160.150.110.100.100.101.27
SL0.100.080.030.060.080.200.110.100.100.100.96
TWI0.060.080.050.040.020.050.030.030.100.100.55
STI0.060.040.030.060.040.100.060.050.100.100.64
SPI0.100.040.030.060.040.100.060.050.020.020.53
PLC0.060.040.030.060.040.030.030.100.050.050.49
PRC0.060.040.030.060.040.030.030.100.050.050.49
Table 4. Random indices function of the number of elements compared [69].
Table 4. Random indices function of the number of elements compared [69].
n12345678910
RI000.580.91.121.241.321.411.451.49
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Echogdali, F.Z.; Boutaleb, S.; Bendarma, A.; Saidi, M.E.; Aadraoui, M.; Abioui, M.; Ouchchen, M.; Abdelrahman, K.; Fnais, M.S.; Sajinkumar, K.S. Application of Analytical Hierarchy Process and Geophysical Method for Groundwater Potential Mapping in the Tata Basin, Morocco. Water 2022, 14, 2393. https://doi.org/10.3390/w14152393

AMA Style

Echogdali FZ, Boutaleb S, Bendarma A, Saidi ME, Aadraoui M, Abioui M, Ouchchen M, Abdelrahman K, Fnais MS, Sajinkumar KS. Application of Analytical Hierarchy Process and Geophysical Method for Groundwater Potential Mapping in the Tata Basin, Morocco. Water. 2022; 14(15):2393. https://doi.org/10.3390/w14152393

Chicago/Turabian Style

Echogdali, Fatima Zahra, Said Boutaleb, Amine Bendarma, Mohamed Elmehdi Saidi, Mohamed Aadraoui, Mohamed Abioui, Mohammed Ouchchen, Kamal Abdelrahman, Mohammed S. Fnais, and Kochappi Sathyan Sajinkumar. 2022. "Application of Analytical Hierarchy Process and Geophysical Method for Groundwater Potential Mapping in the Tata Basin, Morocco" Water 14, no. 15: 2393. https://doi.org/10.3390/w14152393

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