Next Article in Journal
Effects of Climate Change and Changes in Land Use and Cover on Water Yield in an Equatorial Andean Basin
Previous Article in Journal
RNN-Based Monthly Inflow Prediction for Dez Dam in Iran Considering the Effect of Wavelet Pre-Processing and Uncertainty Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing Drought Resilience through Groundwater Engineering by Utilizing GIS and Remote Sensing in Southern Lebanon

by
Nasser Farhat
The Lebanese Center for Water and Environment (LCWE), Beirut 1710, Lebanon
Hydrology 2024, 11(9), 156; https://doi.org/10.3390/hydrology11090156
Submission received: 24 August 2024 / Revised: 16 September 2024 / Accepted: 18 September 2024 / Published: 21 September 2024
(This article belongs to the Section Surface Waters and Groundwaters)

Abstract

:
Countries face challenges of excess, scarcity, pollution, and uneven water distribution. This study highlights the benefits of advances in groundwater engineering that improve the understanding of utilizing local geological characteristics due to their crucial role in resisting drought in southern Lebanon. The type of drought in the region was determined using the Standardized Precipitation Index (SPI), Standardized Vegetation Index (NDVI), Vegetation Condition Index (VCI), and Soil Moisture Anomaly Index (SM). The dry aquifer and its characteristics were analyzed using mathematical equations and established hydrogeological principles, including Darcy’s law. Additionally, a morphometric assessment of the Litani River was performed to evaluate its suitability for artificial recharge, where the optimal placement of the water barrier and recharge tunnels was determined using Spearman’s rank correlation coefficient. This analysis involved excluding certain parameters based on the Shapiro–Wilk test for normality. Accordingly, using the Geographic Information System (GIS), we modeled and simulated the potential water table. The results showed the importance and validity of linking groundwater engineering and morphometric characteristics in combating the drought of groundwater layers. The Eocene layer showed a clearer trend for the possibility of being artificially recharged from the Litani River than any other layer. The results showed that the proposed method can enhance artificial recharge, raise the groundwater level to four levels, and transform it into a large, saturated thickness. On the other hand, it was noted that the groundwater levels near the surface will cover most of the area of the studied region and could potentially store more than one billion cubic meters of water, mitigating the effects of climate change for decades.

1. Introduction

Long-term drought is a significant natural obstacle that harms agricultural and socioeconomic activities. It also adversely affects the ecosystem and jeopardizes the sustainability of various natural environments [1,2,3,4,5,6,7,8,9]. The eastern Mediterranean regions are experiencing record climatic anomalies [10,11,12] and more prolonged and severe droughts in the current decade [13,14,15]; these exceptional droughts are about 98 percent drier than the driest periods of the last 500 years [16]. In the past century, more than 10 million people have lost their lives due to major drought events causing several hundred billion dollars (USD) in economic losses worldwide [17], USD 249 billion of which have been in the U.S. alone since 1980 [18], and the numbers are rising.
National Drought Mitigation Center [19] adopts the classification of drought types into three categories: meteorological, agricultural, and hydrological, with indicators including precipitation, NDVI, soil moisture, and groundwater [20]. Groundwater in karst areas is vulnerable to repeated droughts, as it can lose its storage capacity and take longer to replenish [21,22,23]. Scientists have conducted numerous studies to better understand groundwater droughts in the context of the meteorological drivers and the propagation of droughts through hydrological systems [6,24,25,26,27,28]. Numerous scientific investigations have highlighted that three primary factors predominantly have complex nonlinear interactions that influence groundwater configuration and flow rates: geology, climate, and topography [29,30,31,32,33,34,35,36,37].
The geological strata in the study area, situated in the eastern basin of the Mediterranean Sea and the area shared by the continents of Asia and Africa, has been significantly affected by tectonism. This tectonic activity has resulted in the exposure of geological strata to various events, such as lithostatic, tectonic, and thermal pressures, which can lead to the formation of fractures, cracks, and faults [38,39,40]. As a result, the geological strata in this region have the potential to act as aquifers, making them a subject of interest for researchers and experts in the field.
Numerous theoretical, experimental, and simulation studies have been carried out to better understand the intricate mechanisms underlying groundwater flow in fractured and karstic rocks. These studies have significantly contributed to the advancement of our knowledge on the subject matter [41,42,43,44,45,46,47,48,49], and this knowledge benefits the project proposed here.
In 2017, groundwater accounted for 25% of the global freshwater abstractions, making it a significant source of freshwater worldwide [24,50]. Twenty percent of global groundwater recharge is generated in karst areas [51,52], with renewable groundwater being defined as “any groundwater that can be dynamically captured during pumping that leads to a new dynamically stable equilibrium in groundwater levels within human timescales”. This means that we can transform the depleted Eocene aquifer into a sustainable layer if we can secure sufficient recharge to restore the dynamic balance to its previous state.
Drought is a major driver of crop yield volatility and, in particular, causes low yields that can lead to substantial financial losses [53]. Future predictions of precipitation, temperature, and evapotranspiration are uncertain, where the rapid increase in surface temperature correlates with declining biodiversity, including higher extinction rates [54,55]. Quantifying the spatiotemporal interactions between groundwater and the climate is essential [56]. However, the current understanding of the global-scale sensitivity of groundwater systems to climate change and the resulting variation in feedback from groundwater to the climate system is limited [57,58,59,60]. Therefore, we used remote sensing techniques via Google Earth Engine with several indices that showed no effect of climate factors on groundwater systems between 1981 and 2024 in the study area; these techniques allowed us to identify the type of drought experienced by the region, which is hydrological drought; conversely, groundwater may influence the climate through evapotranspiration and has a climate-cooling and wetting effect [61]. We studied this by exploring the potential artificial recharge of the Eocene aquifer.
This study aims to address the issue of drought by exploring the potential of this region to overcome water scarcity through artificial groundwater flow in geological media. Therefore, we resorted to using a Geographic Information System (GIS), and we used many software programs including ArcGIS Pro 3.3, Global Mapper 22, and AutoCAD 2021 due to their importance in modeling and simulating the supposed water table to be raised to a level close to the surface, developing methods to implement artificial groundwater recharge using sustainable and clean surface resources. In addition, GIS and statistical indicators contributed to the evaluation of the optimal design procedures for dam and tunnel construction, as well as their mechanisms of action during groundwater flow and control. The project’s results are presented in detailed maps.

2. Methods

We relied on the inductive approach, considered one of the most closely related to the reality of such studies [62]. This approach discusses many natural topics that have the same target. The methodologies outlined below served as our guiding framework.
The study focuses on southern Lebanon, which is situated between the Mediterranean coast on the west and the mountain peaks on the east. It encompasses numerous hills and valleys that span from sea level up to 800 m in height. The region extends northward to include the highlands of Al-Tuffah, located on the western slopes of Mount Al-Rayhan. Moving westward from there, the region descends toward the coast and stretches to the borders of Palestine in the south. The study area is located between longitude 35°18′ and 35°50′ east and between latitudes 33°5′ and 33°30′ north, covering 1000 km2 (Figure 1); the region is estimated to have a resident population of approximately 200,000.

2.1. Data Sources

We used various satellite data sources, including Landsat 8 and 9, MODIS, Tropical Rainfall Measurement Mission (TRMM), Global Precipitation Measurements (GPMs), Multiple Satellite Constellation (IMERG), and Soil Moisture Active Passive (SMAP) (Table 1), as well as field and practical data from studies of dozens of artesian wells drilled in the region. We also used topographic, geological, and water resource data from the National Remote Sensing Centre and Climatic data from the climate department at Beirut Airport and a private weather station.

2.2. Remote Sensing

We used remote sensing to assess the three drought types and identify areas of vulnerability using the above satellite data sources. We used the Google Earth Engine (GEE) platform, which provides a comprehensive collection of environmental data sources and relevant satellite imagery to perform this task. This platform also allows for the collection of a large number of multi-temporal images of a single location and enables the generation of drought-specific indices (Standardized Precipitation Index (SPI), Vegetation Condition Index (VCI), Normalized Difference Vegetation Index (NDVI), and Soil Moisture Anomaly Index (SM)) by incorporating available data that best serves the project objectives.

2.3. Geographical Information Systems

We utilized engineering software such as ArcGIS Pro v3.3 (overlay, editing, spatial analyst tools), Arc Map v7.1 (interpolation, conversion, analysis tools), Arc Scene v7.1 (3D tool), Global Mapper v22 (classification, combine/compare tools), AutoCAD 2022, and GEO5 to map and analyze the geological strata and aquifers, examining their spatial distribution, inclinations, curvature, and fractures. To assess their physical and chemical properties, we identified suitable points and locations for horizontal tunnel excavation based on the physical characteristics of the Eocene Strata, and we determined and modeled water table levels to address groundwater drought.

2.4. Modeling and Simulation

We utilized modeling and simulation software such as ArcGIS Pro 3.3 (simulation and modeling tools) to map and simplify the representation of the complex reality in the process of recharging groundwater aquifers and predict the path of the water level, whether surface or groundwater, after building the dam through simulations within the standards available in the study area.

2.5. Geomorphic Indices

We conducted a comprehensive morphometric study of river basins and permanent water courses using mathematical equations and algorithms to determine the optimal dam location based on geological considerations.

2.6. Statistical Techniques

Spearman’s correlation coefficient was relied upon to identify morphometric indicators that were strongly related. Morphometric indices that departed from the correlation values between indices with strong correlations were rejected. The selection was dictated by a non-normal probability distribution of one of the parameters (p < 0.05) [63,64] based on the Shapiro–Wilk test for normality [65].

3. Results

3.1. Climate Elements and Water Resources

Precipitation, Temperatures, and Evaporation in the area
Scientists agree that climate change will have a tangible impact on the supply, demand, and quality of water resources in the Mediterranean basin, which will increase pressure on these resources and on ecosystems in general [66]. The latest statistics from the EEA’s [67] State of Lands and Water report forecasts actually predict that by 2025, draft frequencies will increase by 20 to 60%, and by the end of the century, annual rainfall will decrease by between 8 and 10 mm, with large variations across the region and across different seasons, which will affect agricultural seasons. This region falls under the influence of the Mediterranean climate, which is considered a transitional climate between cold temperate regions and semi-tropical desert regions. In winter, it is exposed to moderate climatic influences; that is, it comes under the direct influence of rainy westerly winds, represented by the upper subtropical jet stream, whose meandering allows cold air masses (cold air valleys) to reach the eastern basin of the Mediterranean [68], accompanied by the emergence of air depressions on the surface caused by the warmth of its waters. This results in an unstable situation that leads to rain, which, in turn, forms snow in the mountain highlands.
The westerly current recedes in the summer toward the upper temperate latitudes, and the subtropical high-pressure centers in the upper strata advance toward the eastern basin of the Mediterranean, resulting in a state of stability that leads to a rise in temperatures.
Precipitation of all kinds is the primary support for water resources [69]. The Standardized Groundwater Level Index (SGI) builds on the Standardized Precipitation Index (SPI) to account for differences in the form and characteristics of the groundwater level and precipitation time series [70]. The study area received annual precipitation between 600 and 1300 mm (Figure 2), especially at higher elevations on the mountain slopes. We used the completely regularized spline, which is one of the radial basic function (RBF) methods for evaluation and mapping the annual rainfall data in ArcMap, which is the optimal method compared to others [71]. It is an exact interpolation technique. The resulting smooth surface passes exactly through the input points, and the greater the value of the number of points, the smoother the surface of the output raster. The average annual volume of precipitation is approximately 650 million m3, and more than half of the study area receives an average yearly precipitation exceeding 700 mm. The general average annual rainfall in the study area is 723 mm, which indicates that the region is not considered an arid or semi-arid region. Still, it is classified as a semi-humid region instead, as it depends on rain-fed agriculture. Spring is the reason for its growth and its continuation during the dry season.
Several decades ago, these quantities of precipitation were sufficient to meet the population’s need for water resources due to the simplicity of life and the reliance on the rainwater provided by nature—collecting and storing it for the summer—in addition to the availability of easily accessible surface water, such as that in canals. The governorate’s residents can quickly reach spring streams and small springs. Because the region’s rivers are far from places of settlement, as they flow through steep gorges with rugged paths, residents cannot benefit from these permanent watercourses. In the current era, with its large increases in population, the development of lifestyles and luxury, and technological progress allowing humans to benefit from groundwater—even depleting it to dangerous levels—there has not been adequate investment into rivers, e.g., by building water projects on them. This situation has placed the study area in the dilemma of having insufficient water resources to meet the population’s needs.
Temperatures directly affect water resources in the study area, as they contribute to increased rates of evaporation from the surface and the upper layer of soil [72]; this also negatively affects water infiltration and, thus, groundwater.
The climatic changes occurring in the eastern Mediterranean basin region have been described in several recent relevant studies [14,73,74,75,76].
In the study area, the average minimum recorded temperatures do not fall below 0 °C, except on very few days of the year, and they are limited to between 0 and 16 °C. The average maximum temperatures recorded are limited to between 15 and 35 °C. This does not prevent the temperature on some days from registering a noticeable increase to approximately 40 °C.
This indicates that the climate of the study area is moderate and that the summer months, which are characterized by high temperatures, are considered relatively moderate, as the temperature rarely exceeds 35 °C.
The temperature averages show an upward trend and are statistically significant with an error rate of 5% according to the Mann–Kendal index, and they reach a threshold of 1% at the Beirut station. In the climatic data available to us for the study area, we detected a shift in temperature with an upward trend; a report prepared by an international team within the framework of the Fifth Assessment of the United Nations (67) indicated that the last three decades have witnessed a gradual increase in temperature on the Earth’s surface. This temperature trend is statistically significantly increasing with a 5% error according to Spearman’s arrangement. At the airport station, it has been shown that the annual temperature rates are rising in an unprecedented manner (Table 2, Figure 3).
We calculated the increase in temperature using the equation shown in Figure 3: y = 0.0381x + 23.231, that is, we multiplied the (variable) slope by the number of years (0.0381 × 46 = 1.75), meaning that the temperature at the Beirut station increased. At 1.75 °C, over approximately four decades, this will negatively affect the water resources in the study area due to the increase in evaporation and transpiration with the increase in the temperature because evaporation is directly linked to the water balance equation.
This increase in temperature will change the patterns of entire ecosystems and will even have a direct impact on many other climatic elements, such as the expansion of the dry season and the dominance of drought over more lands in various parts of the world [77]. The rainy season is limited, as we mentioned previously.
There may also be a direct impact in the form of an increase in consumption rates, which, in turn, will reflect negatively on the water resources in the study area.
Temperature is an essential climatic element and plays a vital role in water balance, as well as other climatic elements and geomorphological processes [69]. Average temperatures in the Middle East increased by 1.5 °C according to the IPCC (2018). This increase is the result of global warming [78]. The Middle East will suffer a 17% drop in available freshwater when temperatures rise by 2 °C [79].
The average volume of evaporation in this region (private weather station) is estimated at 43–51% of the volume of precipitation [68]. This percentage of evaporation has a significant and negative impact on the region’s water resources and, particularly, groundwater. As high daily evaporation rates lead to the impoverishment of the soil and the water in the upper rocky Strata, when the soil is dry, water will not flow for a significant distance until the water content in the soil becomes sufficient and the pressure head becomes less harmful. The specific discharge is measured with an equation from Cushman and Tartakovsky [80]:
q = −K (θ) ∂h/∂l
where q is the discharge and θ is the water content of the soil (θ volumetric soil water content (m3/m3).
The maximum infiltration capacity is not reached because the unsaturated hydraulic conductivity is not at its maximum (Equation (1)). When the daily volume of precipitation is low, the rock strata do not benefit from precipitation because of their rapid evaporation.
We thus conclude that the temperatures in the region are a natural obstacle to facilitating the infiltration of precipitation to the underground strata. Conditions tend to become more dangerous with the advancement of time due to climate change because the annual rainfall amounts are expected to decrease, in addition to the decrease in the rainy season in the eastern Mediterranean basin.
Evaporation rates vary depending on various natural conditions; the evaporation rate is considered the engine of the general hydrological cycle and is its first element.
Drought in the Mediterranean basin is expected to increase by 10% if temperatures rise by 1 °C. This would contribute to taking the region toward desertification, exacerbating the current conditions of the water problem, and increasing pressure on water resources [74]. This affects the production of wheat, olives, vines, and all kinds of vegetables, and the population in the study area depends on these crops for their livelihood. The probability of fires occurring will increase, in addition to the effects resulting from human pressure on natural resources [81].
The impact of climate change on the Mediterranean basin may affect the groundwater reserves of freshwater. Climate change has direct and indirect negative effects on the economy. We should not ignore the anomalous weather conditions that may occur due to these changes and fluctuations in climatic elements around the world—especially waves of intense heat or extreme cold in which temperatures are noticeably higher or lower than their annual average in a particular region.

3.2. Tectonic Influence on Water Resources

Tectonic activity is the predominant source for the formation of geological features [82]. Tectonism determines the inclination of these strata and produces faults, fractures, cracks, and torsions, which are considered one of the quickest and easiest ways to move and direct groundwater [83]. Rainwater that infiltrates through the cracks created by these movements and the joints between the rocks is held there and becomes a factor contributing to the karst process and the formation of groundwater. Tectonic pressure resulting from the formation of the Red Sea between the African and Asian continents in the Miocene era led to the continental break-up between Africa and the Sinai Peninsula on one end and the Arabian Peninsula on the other [84]. A group of longitudinal strike-slip faults appear to the south, starting from Aqaba and passing through the Wadi Araba Depression, the Jordan Valley, the Dead Sea, and Lake Tiberias, all the way to Lebanon, Syria, and the Anatolian Mountains of Turkey, with a length of approximately 1200 km [84,85].
These tectonic forces have led to major faults in the study area with large dimensions in terms of lateral (several kilometers) and vertical (several hundreds of meters) displacement and the appearance of fractured and collapsed zones. In addition, the extension of significant faults has resulted in the fracturing of Cretaceous and Tertiary limestone [83]. This has resulted in several transverse fractures and longitudinal ripples, including the Bint Jbeil syncline and Jabal Amel anticline, which are adjacent to each other. Fractures and ripples aid in forming underground reservoirs that store infiltrated water, where fractures play a positive role in accelerating the passage of this water to the depths.
The southern Lebanese region has experienced relative tectonic stability since the Late Cretaceous–Eocene period, but it was later subjected to a series of minor movements [86].
The deformation that resulted from the pressure exerted on the strata in the study area created many convex and concave ripples (syncline, anticline) that were interrupted by some fractures. Some of this deformation contributed to the formation of important underground reservoirs in the region.
-
Nabatieh–Bint Jbeil Concave: This concave takes the shape of a rectangle from the south, at the Lebanese–Palestinian border, to the north, in the Iqlim al-Tuffah area (Figure 4 and Figure 5). Its surface is covered by continental Neogene limestone rocks (Mcg), followed by Eocene rocks (E2) in the center of its axis, and on the edges, Cretaceous rocks (C4) appear. Moreover, due to the high horizontal pressure on the sides, the rock masses thinned in the middle and formed a concave ripple, or what is referred to as a syncline. The main factor in its formation was the presence of resilient and soft rock strata that tectonic movements could not break (Senonian C6 and Albian C3), which allowed it to form the rocky undulations that contributed to its formation. The axis of this concave inclines from the north and from the south toward the center at the stream of the Litani River [87], while its sides are unequal and incline toward Al-Hula in Palestine (Figure 4).
-
The convex of Jebel Amel: This convex is complementary to the concave of Nabatieh–Bint Jbeil from its western side. It begins to slope toward the sea from its axis at the town of Ansar in the district of Nabatieh and the city of Bint Jbeil (Figure 4 and Figure 5), and it contributes to groundwater flow to the sea, causing springs to discharge several kilometers from the shore, opposite the city of Tyr and the town of Khaizran.
Fractures can develop and evolve due to various factors, such as tectonic activity, heightened fluid pressure, and fluid erosion. Researchers have extensively studied these factors [88,89].
Fractures are well developed in carbonate rocks, which are consolidated rocks [90], such as those found in the study area, and they have secondary porosity that allows water to be filtered and stored in these multi-set voids.
During the search for groundwater, we use fractures as the most important clues, as they create voids and separations in rocks and, thus, increase their porosity and permeability.
Most of the fractures in the study area are of a type involving a thrust fault and a lean-slip fault in which the upper block, above the fault plane, moves up and over the lower block, and there are also secondary faults that appear to be semi-parallel. The general direction of these faults is northeast–southwest. Moreover, there are occasional minor fractures (lineaments) in all directions (Figure 6).
The fractures reach deep underground strata (Figure 6), that is, more than 400 m below the surface of the Earth, but the presence of an impervious and soft rock layer (Senonian) between two hard strata (Eocene and Cenomanian) prevented water from leaking from the upper stratum toward the lower one and contributed to the exacerbation of karst processes within it. This made the Eocene layer an ideal layer for groundwater storage and flow.

3.3. Stratigraphy and Lithology in the Study Area

The lithological characteristics of exposed rocks play a major role in the groundwater recharge rate. They act in two ways: first, they percolate surface water to the stratum underneath, and second, they store water in rock voids, fissures, and bedding planes [91].
The lithology was identified using the available geological maps [92]. The lithological distribution in the form of classes was obtained according to formations attributed to Quaternary, Neogene, Paleogene, Cretaceous, and Jurassic deposits (Figure 4 and Figure 6).
Each rock formation mentioned in the study area has similar depositional and stratigraphic sequences from top to bottom, with few diverse rock beds intervening. The main rock components, including the stratigraphic sequence, the interbedded rock strata, and the thickness of exposed rocks in the total sequence, were determined for each rock formation.
In addition to the Cretaceous and the Paleogene, which are separated by the Senonian insulating layer (C₆), the study area has a small area with other strata (bc, c3, c2b, c1, J, and Mcg), as well as a small area with Quaternary sediments (Q) (Table 3).
The calcareous strata make up 95% of the area (Table 3). These limestone formations have high porosity (30–40%), suggesting high permeability [91,93,94] due to their physical properties. This is firstly due to the presence of small intergranular voids, allowing it to retain quantities of water. Moreover, a connection between these spaces enables the water to pass through them to form vast stores of freshwater inside the depths of the ground, referred to as aquifers. The geological formations located in the study area are considered suitable for storing vast amounts of water and have favorable conditions for the karstification process within their rocks [21].
  • The Cretaceous period:
The formations of this period are considered the most widespread in the study area (Figure 4), with an area of approximately 580 km2. This is a positive indicator regarding groundwater reservoirs [95]. This era is divided into the Lower and Upper Cretaceous. Lower Cretaceous, consisting of the base of Cretaceous (C1) and Albian (C3), includes rocks that are distributed in relatively small areas on the outskirts of the study area.
The Upper Cretaceous, consisting of C4–C6, includes Cenomanian rocks (C4), which are considered the most widespread and important, covering approximately 389 km2 of the general area. They are groups of dense, pale yellow dolomitic rocks at the top with gray and white limestone rocks at the bottom (C4c), and the limestone rocks are interspersed with some marly sections (C4b and C4a). Their thickness reaches more than 600 m, and they are frequently fractured. These units contain quartz nodules and a large number of burials within them such as marine and terrestrial fossils. Senonian rocks (C6) are gray marly rocks, chalky limestone, and impervious marly rocks that often contain siliceous sections. They are considered the ceiling of the Cretaceous period. These are often exposed in isolated low-lying areas. Their thickness reaches 150 m, and they contain some solid black phosphate and petroleum rock conglomerates.
The Eocene (E2) formations are located over a large area of approximately 307 km2 (Figure 6). The rocks of this period are sedimentary and are marly limestone rocks with siliceous sections, dolomitic limestone, soft limestone, and hard coastal limestone. The thickness of this layer in the study area is estimated to be between 300 and 400 m [87].

3.4. Geomorphological Features

Groundwater fluxes are most strongly driven by topographic gradients [96]; the regions where changes influence the movement of groundwater in the elevation of the land surface are highly consistent with water tables that are regulated by topography [97].
The study area contains a large number of plateaus and hills [98] and is divided by the Litani River into two natural parts; the first is the northern part, represented by the Nabatieh region, and the other is the southern part, represented by the areas of Marjayoun and Bint Jbeil (Figure 7).
The northern section is a plateau with an average height between 450 and 500 m above sea level, with the exception of its northern area, represented by the Iqlim al-Tuffah region, which reaches a height of 1400 m at the town of Jbaa. This plateau generally slopes toward the west with an average slope of 5%. The southern section is similar to the northern section in terms of its topography, but its height above sea level varies between 550 and 950 m (Figure 7). The highest part is on the outskirts of the town of Maroun al-Ras. Then, the elevation decreases by 350 m toward the north on the banks of the Litani River. On its east, it is 550 m above sea level, and on the western side, it is 450 m. The general slope of this section is from the south toward the north and west.
The study area is approximately 50 km long and 15 to 17 km wide. It contains a large population, exceeding 200,000, which leads to great pressure on the water resources in this region.

3.5. Karst and Rock Permeability Are Positive Indicators of Groundwater

Geological media are classified into three categories based on their pore structures: porous, cracks, fractured, and karst [99].
The Middle Cretaceous period rocks, especially the Cenomanian (C4) and Eocene (E2) rocks, are characterized by a very high permeability that occupies a large part of the study area. This positive indicator is addressed later, as this very high permeability allows water to penetrate the rock to great depths. When it reaches insulating strata, it forms vast underground reservoirs, especially since the permeable strata in the region are thick.
Senonian Cretaceous roof rocks (C6) are considered impervious rocks. These rocks lie between Cretaceous and Eocene rocks, making the Eocene layer a phreatic subterranean reservoir (free aquifer or saturated thickness).
In Lebanon, carbonate rocks are considered ideal for karstification and the resulting manifestations, such as sinkholes and caves [83]. Many extensive studies have been conducted on karst phenomena in Lebanon [100,101,102,103], where karst geomorphology, such as sinkholes, doline, tunnels, grottoes, and underground rivers, is present.
Karstification is the process of dissolving calcite, CaCO3, the main component of limestone rocks. As water passes through the soil into the ground, it carries monoxide carbon gas with it, which begins the process of dissolving the internal parent rock and leads to the formation of internal gaps, passages, and caves due to the high acidity of the water.
Karstification is a significant and very influential factor in our study in terms of finding solutions, such as when choosing a location for a project (surface or underground).

3.6. Assessing Drought Conditions on the Surface and in the Subsurface Using Remote Sensing

Remote sensing technology is crucial for assessing surface and subsurface drought conditions [98,104]. This technology provides valuable information on various drought types and indicators that are mathematical representations of water deficit (and excess) compared to historical data, such as the Standardized Precipitation Index (SPI), Normalized Difference Vegetation Index (NDVI), Vegetation Condition Index (VCI), Soil Moisture Anomaly Index (SM), and unconfined groundwater levels, which can help experts understand the severity and extent of drought conditions according to NASA and National Drought Mitigation Center criteria [19,20].
By analyzing remote sensing imagery based on the reflectance and/or emission of electromagnetic radiation, passive (Sentinel 2) [105,106,107,108], active (Sentinel 1) [109,110,111,112], thermal (Landsat 8) [113,114,115,116], and optical (most satellites) [116,117,118,119] data can be obtained. Experts can develop effective drought and groundwater management strategies and take timely action to mitigate the impacts of drought on the environment and human activities.
The Standardized Precipitation Index (SPI) (Figure 8) is primarily defined to characterize meteorological drought, whereas historical rainfall data at any location fitted with gamma distribution represent cumulative probability function; therefore, if a rainfall event carries a low probability on the cumulative probability function, it is indicative of a drought event.
The SPI values can be interpreted as the number of standard deviations by which the observed rainfall anomaly deviates from the long-term mean, and its average over different periods (3, 6, 9, and 12 months) indicates the severity and duration of drought [120].
Figure 8 from the Google Earth Engine (GEE) shows the SPI resulting from the combination of the satellite Tropical Rainfall Measurement Mission (TRMM) and Global Precipitation Measurements (GPMs) with multiple satellite constellation (like Sentinel 2A, 2B, etc.), The observed period of the study area for the SPI begins on 1 January 1981 and ends on 25 July 2024; it shows that the region is near level 1, meaning that the study area is not exposed to meteorological drought.
VCI: MODIS data are used to calculate the Vegetation Condition Index (VCI) (Figure 9) for a specific time and location. The VCI is expressed as a percentage and is an indicator of the status of vegetation cover as a function of NDVI minima and maxima encountered for a given ecosystem over many years. It is a better indicator of the water stress condition than the NDVI [19]. The deviation of the vegetation condition is an indicator of the intensity of the impact of drought on vegetation growth.
The Normalized Difference Vegetation Index (NDVI) normalizes green-leaf scattering in near-infrared wavelengths with chlorophyll absorption in red wavelengths. It measures the greenness and density of vegetation captured by a satellite image. By calculating the difference between two bands, visible red and near-infrared, we can determine that healthy vegetation has a very characteristic spectral reflectance curve.
To obtain the VCI in GEE, we define a variable that selects the NDVI band from the MODIS 16-day image collection and map the function across the collection, then we define variables that derive the minimum and maximum NDVI values across the image collection at the pixel level; we then define variables for differencing NDVI images, minima, and maxima encountered. Then, we export the image to Arc-GIS for mapping it.
Figure 9 shows the study area un-exposed to agriculture drought according the NDVI and VCI indices.
The Soil Moisture Anomaly Index (SM) (Figure 10) uses the SMAP satellite, whereas the NASA Land Information System (LIS) provides a framework for the comprehensive characterization of the land surface and hydrological extremes such as drought [121]. The SPoRTLIS provides real-time surface fields and soil moisture percentiles to determine drought. This indicator is considered one of the most accurate indicators of surface and subsurface hydrological drought, as the exploration range reaches 2 m under the soil and at successive and distant periods.
We create this index using GEE by defining a variable for the SMAP 9 km Surface and Root Zone Soil Moisture product.
Figure 10 shows that the region has been extremely exposed to hydrological drought, due to many natural and human reasons, as the recharge of the groundwater layers was affected by the reduction in the rainy season and the abnormal amounts of rainfall in a short period, in addition to the excessive extraction of groundwater in this region.

3.7. Hydrogeological Concept and Groundwater in the Study Area

Seepage is the transfer of precipitation or surface water through the pores and small cracks of rocks into the ground.
Seepage is the displacement of precipitation or surface water through rock pores and cracks into the ground. This water typically originates from precipitation and minor seepage from rivers and lakes, though riverbeds and lake bottoms often have clay layers that limit this process. Without such layers, water seeps through and is halted by an impermeable layer. The hydraulic properties of rocks affect how water seeps; porous rocks have uniform properties, while karst and fractured limestone are more variable due to localized karst formations [122].
In the study area, 90% of rocks are prone to karst processes, affecting groundwater estimations. Surface water-bearing strata in the area include the Eocene Strata, which is an unconfined aquifer or saturated thickness and is influenced by atmospheric pressure and gravity. The Senonian (C6) layer blocks seepage, leading to impermeable strata. An aquifer can be described as a physiographic unit containing a single large aquifer or several interconnected aquifers [123]. There are two water-bearing strata in the study area: the Eocene layer (E2) and the Cenomanian layer (C4) (Figure 4 and Figure 6).

3.7.1. Eocene Aquifers

Physical Characteristics

The Eocene aquifer is a free aquifer with a saturated thickness, extending from the Triassic period and largely exposed in the study area, except for a small portion covered by recent Quaternary strata. It drains non-artesian, flowing into karst springs or where it meets insulating strata, particularly the Senonian, which separates it from the Cretaceous strata. Composed of highly cracked limestone, dolomite, siliceous, and marl sections, it stretches from the northern Iqlim al-Tuffah area to the southern Aitaroun, near the Palestinian border (Figure 4 and Figure 6). Its thickness ranges from 300 to 400 m [87] and is subject to significant karst processes due to precipitation and other factors, allowing rainwater to penetrate through cracks and fissures.

The Direction of Groundwater Flow

Groundwater moves through rock pores, cracks, and fissures according to Darcy’s Law, influenced by permeability and hydraulic gradient [123]. This movement can be tracked using non-toxic tracers [124]. Water velocity in limestone ranges from 0.6 to 6 m per day. The Eocene aquifer’s flow extends into the Nabatieh–Bint Jbeil ripple from both east and west, with some springs appearing along the Litani River.

Natural Recharge Eocene Aquifers

Water balance equation:
GWr = [P − (ET + R)] × S
GWr = the average volume of precipitation recharge;
P = the average annual volume of precipitation over the area of the underground reservoir;
ET = the average evapotranspiration, direct evaporation from bodies of water and precipitation, and evaporation through plants;
R = the average surface runoff volume over the exposed area of the underground storage;
S = the exposed area of the rock Strata.
Thus,
The surface of the Eocene aquifer receives abundant amounts of precipitation, estimated at approximately 745 mm.
Surface runoff from precipitation: The percentage of surface runoff from precipitation in the Eocene layer averages approximately 7.5% of the annual precipitation rate, so the surface runoff value becomes approximately 56 mm. Therefore, the equation becomes the following:
GWr/Eocene according to Equation (2) = [745 − (341 + 56)] × 3−10 × 307 ×106 = 106,000,000 m3.
Darcy’s law [125] of laminar flow is as follows:
Re = ρvd/µ
where Re is Reynold’s number, ρ is the water density, v is the average pore velocity, d is the average pore diameter, and µ is the dynamic viscosity of water at a given temperature.
Hydrogeologists generally accept the assumption of laminar flow throughout granular porous media with Re < 100 for the sake of simplifying the mathematics of flow Equation (3) [126].
Transmissivity is the rate at which water is transmitted through a unit width of an aquifer under a unit hydraulic gradient. It is expressed as the product of the average hydraulic conductivity and thickness of the saturated portion of an aquifer (Table 4).
Darcy’s Law (Equation (4)) describes the flow of groundwater through a porous medium.
Q = K A d h d l
Q is the volumetric flow rate. K is the hydraulic conductivity of the aquifer. A is the cross-sectional area perpendicular to the direction of flow. dh/dl is the hydraulic gradient. To simplify Darcy’s Law, we can use the Dupuit–Forchheimer assumption:
Q = K A d H d l
where dH/dl is the slope of the water table. This is often applied to steady-state horizontal flow in free aquifers.

Geometry of the Eocene Aquifer

The Eocene aquifer extends from north to south for a length of approximately 50 km and is geometrically divided into two parts: the northern and southern parts (Figure 11).
The engineering study of this aquifer using geographic information systems and engineering software (Figure 12) showed that the northern section has a general lean of approximately 0.3% from the north toward the center at the Litani River.
The maximum elevation above sea level is approximately 620 m in the area of Iqlim al-Tuffah, and the lowest elevation is approximately 132 m at the Litani River (Figure 13).
The thickness of this layer ranges between 400 m in the northern region and 370 m at the river’s edge, decreasing to approximately 30 m in the river’s downstream area. This is the smallest thickness of this layer in the study area. The southern part of the concave layer of the Eocene layer leans from the south toward the center at the Litani River, with a lean of approximately 0.3% (Figure 11).
The highest elevation of this section above sea level in the southern region is 935 m; it then descends toward the north at the edge of the Litani River by 350 m.
The thickness of this part is approximately 540 m in the south and 400 m in the middle. That is, this part is approximately 140 m thicker than the northern part of the Eocene Strata. The width of this layer that is exposed on the surface ranges from 8 km in the south to 12 km in the center, reaching 3 km in the extreme north.

3.8. Support for the Eocene Aquifer with Clean Water Resources from the Litani River

3.8.1. Morphometric Characteristics of the Litani River Basin in the Study Area

Drainage basin engineering is used to determine watershed surface flow and the drainage mechanism in their basins through indicators based on mathematical relationships, through which one can form a quantitative value for qualitative aspects to help in a more realistic understanding of the aspects of a river basin.
These indicators are numerical indications that provide for the concept and type of flow in waterways during their permanent flow; as a result, one can determine whether a river is suitable for constructing dams or has a great capacity for sedimentation in its stream.
The Litani River is the focus point of these studies, as its annual volume discharge is up to 350 million cubic meters [94].
All values associated with the study area used in the analysis of morphometric indicators were extracted (Table 5) through the use of remote sensing (RS) and geographic information systems (GISs) (Figure 14 and Figure 15).
  • Form factor (Ff)
The shape factor of a basin (Ff) (Equation (6)) is a parameter that makes it possible to determine its geometry and is related to the flows of the drainage network.
Ff = A/Lc2 = 0.25
A: Basin Area; Lc: Basin length
2.
Circularity ratio (Rc)
This parameter is the ratio between the area of a basin and the area of a circle that has a circumference equal to the length of the basin’s perimeter (Equation (7)); therefore, when values are not close to unity, this indicates that the shape of the basin does not resemble a circle.
Rc = A/P2/4π = 0.32
P: Basin Perimeter
3.
Elongation ratio (Re)
This parameter takes a maximum value of 1 for perfectly round basins (Equation (8)). This is the ratio between the diameter of a circle with the same area as that of the basin (Dc) and the maximum basin length (Lc). Furthermore, these values can be grouped into the following forms: a circular form when Re > 0.9, an oval form when 0.9 > Re > 0.8, and a less elongated form when Re < 0.7; this is the form in the study area.
Re = Dc/Lc = 0.57
4.
Compactness coefficient (Kc)
The compactness coefficient describes the relationship between a basin’s perimeter and a circle with the same area (Equation (9)); the more circular the basin is, the closer the result is to unity; however, this is not the case in the study area.
(Kc) = P/√ πA = 3.53
5.
Basin relief (H)
The basin relief (H) is defined as the difference between the maximum and minimum height (Equation (10)). It plays an important role in the superficial flow and development of the drainage network, permeability, and susceptibility to soil erosion.
H = HmaxHmin = 1151 m
6.
Hypsometric Curve of the Basin
The hypsometric curve is the area in km2 between two levels. It provides a graphic representation of the basin relief, where the ordinate represents the elevation above sea level in meters and its abscissa.
7.
Rainwater Harvesting Potential Index (RWHPI)
The planning of water harvesting projects is a multi-criterion problem because it depends on several factors. Rainwater harvesting (RWH) is one of the most common practices for mitigating water scarcity.
8.
Drainage density (Dd)
(Dd) = L/A = 0.08
L: Main channel length of the stream
Drainage density (D): The ratio that defines the drainage density represents the number of rivers in the catchment needed to drain a basin [127] (Equation (11)), and it describes the drainage distribution and spacing in a watershed or the total length of channels per unit area [128].
9.
Constant of channel maintenance (Cm)
(Cm) = 1/Dd = 12.5
The constant of channel maintenance (Cm) is the ratio between the area of a drainage basin and the total length of all channels (Equation (12)); it is equal to the drainage density [129].
10.
Stream frequency (F)
(F) = N/Ad = 1.26
N: Total stream number of the Basin
The stream frequency (F) is the number of stream segments per unit drainage area (usually per square kilometer) (Equation (13)). Dense networks have high stream frequencies.
11.
Stream length ratio (Rl)
(Rl) = Lu¯/Lu¯−1 = 3.34
The stream length ratio (R1) (Equation (14)) is the ratio between the mean length of streams of a particular order and that of the next-lowest order.
12.
Bifurcation ratio (Br)
(Br) = Nu/Nu+1 = 0.998
This indicates an understanding of the runoff behavior and the shape of the basin (Equation (15)), and it is measured for flooding-prone areas. The probability of flooding is high when a short period of concentration is indicated.
13.
Sinuosity (S)
(S) = L/La = 1.28
Sinuosity is a consequence of channel migration processes in meandering rivers (Equation (16)), where flow through a channel sets up instabilities that drive bank erosion and accretion, which change the planform curvature [130].
14.
Mean stream length (Lm)
(Lm) = Lu/Nu = 0.76
Average stream length (Lm) (Equation (17)): This is defined as the ratio of the length of a stream to the number of streams [131].
15.
Relief gradient (Rg)
(Rg) = Mean Elevation − Mini elevation/Maxi Elevation − Mini elevation = 0.5
The selection was dictated by a non-normal probability distribution of one of the parameters according to Equations (18) and (19) (p < 0.05) [63,64], based on the Shapiro–Wilk test for normality [65].
W = ( i = 1 n a i x i ) ^ 2 i = 1 n ( x i x ¯ ) ^ 2 = 0.160 < 0.850   so   p < 0.05
The results for the morphometric parameters of the drainage network allowed us to understand the basin’s hydrological behavior. This would make the Litani River’s course suitable for establishing a dam with which one could reserve a huge amount of water, which we will estimate.

3.8.2. Artificial Recharge

Subsurface fluid flow is a complex and interdisciplinary topic involving various scientific and engineering fields, such as hydrogeology, rock mechanics, geotechnical engineering, and earth resource engineering [132,133,134].
This area of study requires a comprehensive understanding of the dynamics of fluid flow through rocks and the associated geological properties.
The artificial recharge of groundwater with available natural resources involves an increase in the volume of infiltration of rainfall or surface water into geological formations carrying groundwater [135]; this can occur via the injection of water through wells or the mitigation of runoff [136], provided that this process is adapted to the topographical and geological conditions and soil type, in addition to the quality of the water with which the recharging will occur, so as not to lead to the destruction of the underground storage and its exclusion from service [137].
Establishing obstacles or dams to inject aquifers using horizontal wells (or trenches perpendicular to the riverbed) from surface watercourses for transport over long distances allows for subsequent leakage [138], especially in the fractured and cracked Eocene rocks, which are characterized by multi-scale heterogeneity due to their geometrical features, which exhibit variations in surface roughness, fracture apertures, fracture density, and connectivity [118,139]. The hydraulic and geomechanical properties of fractured rocks are strongly dependent on these spatially varying geometrical parameters [140,141,142].
This method aims to increase the area of contact between water and soil and extend the time taken for seepage into the aquifer strata [143]. Accurately estimating the pore size and geomorphological features of the rock strata is crucial for ensuring the successful artificial recharge of underground aquifers; for example, karst processes, which are dependent on the injection of water, carry carbon dioxide (CO2), and this prevents several adverse effects in the process of artificial recharging underground aquifers [144]. Moreover, groundwater flow mainly induces the deformation of fractured and cracked media [145]. This increases the capacity of the aquifer according to an extension of Biot’s poroelasticity theory [146] with the equation of continuum mechanics for coupled hydromechanical processes [145,147,148].
Based on what we studied through the use of geographical information systems and remote sensing of a number of morphometric parameters of the river courses, the characteristics of the Eocene aquifer, and its geometry, as well as the use of several algorithms and mathematical equations associated with them, we confirmed the possibility of establishing a dam on the Litani River to support and recharge the Eocene aquifer.

3.8.3. Dam Properties

Dams are built on river courses based on the functions that are assigned to them.
In addition to natural characteristics, the intended role also has an impact on the specifications for building dams. The study area suffers from water scarcity and a lack of sustainable electricity supply, in addition to slow economic growth and a lack of job opportunities in light of the high unemployment rate and the migration and displacement of a large number of residents of this region [68].
Locating the dam
The first goal of building this dam is to support the Eocene aquifer during the dry season. Accordingly, the location of the dam must be on the rocks of this layer that are visible on the ground due to the geometric characteristics of this Strata, which we discussed previously, especially in terms of its presence within a tectonic concave. Then, we must take advantage of the largest area of water in contact with this Strata, which necessitates the presence of this dam at the furthest point on either end of the boundaries of this Strata.
The topographical factor also has an impact on the determination of the location for building the dam, as it has become clear through data obtained from geographic information systems that the optimal place is where the two edges of the valley converge with each other.
Geologically, the rocks on which the dam is based must be hard and thick so that they are not exposed to natural accidents.
This approach provides opportunities to explore the geological, geomorphological, and morphodynamic controls on the position of the dam.
The distinctive hydromorphological attributes of the Litani River make it a spatial unit that is of interest for dam construction (Figure 16).

3.8.4. Geomorphic Indices for Dam Construction

Wc is the channel width: 11 m. Wv is the valley width: 50 m. Lv is the length along the axis of the valley: 528 m. Lc is the channel length: 1010 m. A is the amplitude of sinuosity: 252 m. λ is the wavelength of sinuosity: 314 m. Rc is the radius of curvature of a meandering river: 54 m. D is the maximum depth of the valley: 190 m.
SI is the wavelength of the sinuosity of the meandering reaches; SI > 1.25 corresponds to highly sinuous or meandering rivers (Equation (20)), whilst values of SI < 1.1 represent straight rivers.
SI = Lc/Lv = 1.9
Variations in the SI index are mainly controlled by changes in water discharge and sediment load in response to tectonic, climatic, and anthropogenic factors [129,149].
The Rc/Wc ratio is an indicator of meander maturity [150] (Equation (21)). When this ratio is between 2 and 3, the lateral erosion rate is high.
Rc/Wc = 4.9
This means that the lateral erosion rate is low, and this is the optimal condition for dam construction.
Next, we calculate the river incision factor (If) (Equation (22)), which quantifies the incision in a valley, by comparing the difference between the valley width (Wv) and channel width (Wc) with the valley depth (D) at a specific location.
If = [(Wv − Wc)/D] = 0.2
Measuring the If index allows for the quantification of the narrowing or widening of a river valley along its course, and a low If value (<10) typically indicates the “V” shape of mountainous valleys.
Since its value is very low (0.2), this indicates that the two edges of the river are very close together; this helps in the construction of a dam with the smallest amount of building materials due to the close distance between the two edges in this location.
Through the intersection of these indicators and relevant natural, ecological, and social data, the specific location for building a dam on the Litani River was determined: “Arid el Ajrad” between the villages of Alman and Zawtar El Gharbiyeh (33.3105°; 35.4500°).

3.8.5. Characteristics of the Dam

To determine the dam’s dimensions, we characterized the dam in relation to the topography of the area and the purpose of its construction. The height of the dam body above the base is the most important factor, as it can contribute to the understanding and study of the quantities of water retained behind the dam, as well as the quantities of water that can feed the Eocene aquifer.
Based on the previous data, morphometric equations, and determination of the optimal location for this dam, we determined the width of the dam at the crest between the two edges of the valley to be approximately 559 m at an altitude of 290 m above sea level. That is, the height of the dam body above the base was 160 m. This was the maximum height of the dam body, and, accordingly, the areas that will benefit from the height of its water table are determined. However, it is possible to benefit from the lower heights of the dam, as the mere presence of the dam in this area will lead to the flooding of the river basin and, thus, support the Eocene aquifer. However, the greater the height, the more the recharging of this layer extends north and south toward areas that are farther away and in need of water sources.
The maximum area of the dam body or the wall facing the river (Equation (23)) was calculated using geographic information systems to determine the dimensions of the trapezoid-shaped dam:
A = (a + b) × h/2 = 47,000 m2
A is the area, a is the smallest base, b is the largest base, and h is the height.
Worldwide, there are many types of dams, as the type is determined by some natural factors, such as topography, geology, and soil, while other factors are human and economic, such as the goal of constructing the dam. There are also political and military considerations. The aggregate or rock type is the most suitable for the study area because longitudinal tunnels need to be excavated to facilitate the passage of water toward the aquifer. From these excavation materials, we can backfill the dam over a long distance.
For this dam to be safe, the area behind the dam must be backfilled for a distance of more than 100 m, most of which can be secured from tunnels that must be excavated into the rock mass of the Eocene Strata. Then, it is expected that the total volume of the dam in its maximum state will reach about 4 million cubic meters.

3.8.6. Artificial Recharge by Tunnel

The construction of this dam is accompanied by the construction of two main tunnels. The first tunnel extends from the dam lake toward the south for a distance of up to 16 km; spatial analysis of data from the geographic information system showed that when the water level in the dam rises to a level of 290 m above sea level, artificial springs can be created for a distance of 6 km in Wadi Al-Hujair and Wadi Al-Saluki. The height of these two valleys above sea level is approximately 290 m.
Then, several cross-tunnels branch out from the tunnel, with each tunnel not exceeding 2 km in length. They can be used in two cases: either direct artificial springs or a lower water tank very close to the surface. From this idea, we support the reason for the extension of this first main tunnel to a distance of 16 km, as the water in this tunnel will come very close to the surface. Then, the residents of these areas (Shakra, Mays al-Jabal, Aitaroun, Bint Jbeil, etc.) will be able to install wells to depths of 100, 150, and 250 m (maximum) to reach the permanent water source provided by the dam on the course of the Litani River. Note that the region has wells at a depth of 700 m and often suffers from the loss of water sources or the inability to obtain sufficient energy to draw water from these overwhelming depths.
The Eocene aquifer will remain saturated with water that seeped into it from precipitation in the winter because the drainage process was carried out through small springs in the course of the Litani River, and since the dam contains a water level that is approximately 200 m higher than the base of the Eocene Strata, water will be unable to seep out due to counter pressure; this will facilitate the planned artificial refeeding.
This applies to the second main tunnel that extends north toward the Nabatieh area, where the water source will come close to the surface.
The capacity of the Eocene aquifer will be enhanced after filling the proposed dam. On the one hand, it will retain the water seeping into it from 45% of the annual precipitation; this is the percentage of the Eocene rocks’ ability to pass precipitation into the aquifer storage. On the other hand, additional quantities will leak into this layer from the reservoir behind the dam, which holds water, reaching a height of approximately 160 m.
Areas shown in blue indicate where the groundwater level corresponds to the surface of the Earth and where there may be artificial springs or small dams from which to benefit in the summer. Areas in yellow indicate where the water table is within 150 m from the surface. Areas in olive indicate where the water table is within 250 m from the surface (Figure 17), while dark green indicates where the water table is lies between 250 and 540 m from the surface. This is the maximum depth of the Eocene Strata.
The first horizontal tunnel proposed for excavation into the Eocene layer in the study area is shown in Figure 18. This tunnel is supposed to be at an altitude of 200 m above sea level, that is, 70 m above the base of the dam, which would lead to the emergence of several springs in some of the neighboring valleys; these would be considered subsidiary streams of the main river. The streams will flow in these valleys throughout the year because the river will feed them after the dam is built. More importantly, we can build small dams on these streams to reserve larger quantities of water for use later in the summer and to supply the surrounding towns with water for domestic, agricultural, and industrial use.
The groundwater level will rise above the tunnel level. That is, it will rise to 290 m above sea level due to the hydraulic head in the reservoir, where it will be equal to the upper level in the dam. This will allow water to flow rapidly into the rocky voids whose level is less than 290 m, forming several seasonal springs that are likely to cease flowing when the water level in the dam drops below this height. The Saluki and Al-Hojair valleys are among the most significant valleys that will benefit from this process because of their long paths; in these valleys, we can also build small dams with which we can reserve large quantities of water to recharge the Eocene aquifer in the summer.

3.9. Using GISs to Determine Areas Benefiting from Groundwater According to Their Proximity to the Surface

To determine the areas benefiting from groundwater based on their proximity to the surface, after the simulation of the building of the dam, we analyzed the geospatial data using GISs to identify the locations where the water table was closest to the surface. We achieved this by using various GIS tools, such as elevation models, hydrological models, and ArcScene 10.7 software, to create a 3D map that showed the areas closest to the recharged groundwater. By analyzing this map, we identified the areas that were likely to benefit from groundwater according to the depth and accordingly planned four depth levels: 0, 150, 250, and deeper than 250 m (Figure 19).
The quantities of water that can be retained within the Eocene aquifer after artificially recharging using the water of the dam located on the course of the Litani River are determined by the ability of the rock to retain water, which, in turn, is linked to the percentage of voids present in these rocks (porosity), which was estimated to be about 5% (Table 4).

3.9.1. The 0 m Depth of the Water Table

This area extends about 6 km beyond the reservoir; it is located below the towns of Qantara, Toulin, and Burj Qalawiyah. It is situated between 130 and 290 m above sea level, covering an area of approximately 15.5 km2. This area is the lower part of Wadi Al-Hujayr, and it is connected at its end with the course of the Litani River (Figure 20).
To ensure sufficient water storage until the end of the summer, it is recommended to establish a small dam in the middle of this valley near the town of Froun. This dam will be able to store more than one million cubic meters of water. The reason for this is the presence of the impervious Sinonian rock layer at the bottom of this valley, which will help to prevent the leakage of water into underlying permeable strata, such as the Cenomanian Strata.

3.9.2. The 150 m Depth of the Water Table

This is regarded as a key region of the water table due to the dam’s construction along the Litani River. The groundwater level will be found at depths between 0 and 150 m for considerable distances to both the north and south of the dam (Figure 21). The altitude in this area ranges from 290 to 440 m above sea level. In the north, water within the Eocene aquifer will reach the towns of Nabatieh, such as Kfarsir, Aba, Ansar, and Harouf, which are among the towns that suffer most from severe shortages of water resources during the summer.
More importantly, the southern regions will benefit from the surface waters being close to 150 m. Because these towns are currently required to drill to depths of more than 750 m to obtain water, this in itself is a major challenge to economic investment in the region.
This level extends away from the course of the Litani River by about 14 km south, reaching the middle of Wadi Saluki, which is located between the towns of Mays al-Jabal, Shaqra, Hula, Majdal Salam, and Bani Hayyan. The residents of these towns will be able to easily benefit from the sustainable groundwater in the Eocene layer due to its proximity to the surface; thus, it does not require excessive investment in drilling operations.
The area covered by this level of groundwater in the Eocene layer is approximately 89 km2, and the height of the water there, as we mentioned previously, is 150 m.

3.9.3. The 250 m Depth of the Water Table

Naturally, the greater the height above sea level, the greater the depth from the surface to the water table. This can be applied to our third Strata, in which the groundwater level lies at a depth of between 150 and 250 m (Figure 22). This area includes the towns located next to the city of Nabatieh, such as Kfar Rumman, Habbouch, Al-Duwair, Deir Al-Zahrani, Romain, and Houmin.
Many of these towns experience a significant shortage of water resources during the summer. Thus, the shallower groundwater levels, at depths of 150 to 250 m, represent a notable advantage for easier water extraction. In the southern region, water continues to flow towards the far south, reaching about 18 km from the Litani River. It extends to the borders of the towns of Blida and Aitaroun in the far south, as well as the towns of Al-Suwanah, Khirbet Selm, and Tulin in the west.
The quantities that this level of the Eocene aquifer can store depend on the area of this Strata, which is estimated to be 97 km2 based on data from geographic information systems.

3.9.4. The Layer of the Water Table at a Depth of up to 250 m

This layer is concentrated in the far south due to the region’s elevation above sea level, which is more than 900 m. This means that it is normal for the depth of the water table to increase. However, of course, the situation is much better than it was before, as drilling previously reached 800 m to obtain small amounts of water at a high financial cost. However, if this project is implemented, these areas will be able to obtain sustainable groundwater with a maximum depth of only 400 m (Figure 23).
This area is approximately 121 km2, and it includes the city of Bint Jbeil and its neighboring villages and towns.

4. Discussion

4.1. Water Projects and Proposed Recommendations to Confront Drought

The Cenomanian aquifer in the region has ample groundwater, but it is located at a depth of over 800 m (Figure 13). The high cost of extraction, coupled with the water often being mineralized and of low natural quality, makes it less favorable to access. On the other hand, the Eocene aquifer near the surface is experiencing a groundwater drought. Therefore, extracting water from the Cenomanian aquifer requires considerable financial investment and the use of high-capacity submersible pumps powered by energy sources that require maintenance throughout the year.
To achieve maximum efficiency in the artificial recharge process, careful hydrological, hydrogeological, and engineering investigations are necessary to select suitable sites for recharge structures [151,152,153,154,155]. By augmenting the natural water supply through increased infiltration of rainwater and surface water into saturated geological formations, we can overcome some of the challenges and dilemmas associated with water resource management [139] by using several methods and techniques, such as injecting water through wells, mitigating runoff, or creating watersheds over hollows, sinkholes, shafts, and some natural geomorphological features [136,156,157,158,159,160,161]; moreover, the quantification of subsurface discharges in highly heterogeneous areas such as alluvial surficial aquifers is ruled by several uncertainty sources [162]
During summer, the water table of the Eocene aquifer drops below 4 m, and its rock thickness can reach up to 540 m (as reported by the Artesian Well Technical Report in Chakra). The aquifer’s southern part, situated in the Bint Jbeil District, leans from the south to the north, while its northern part, situated in the Nabatieh District, leans from the north to the south, with the Litani River separating the two parts. This means that groundwater generally does not flow beyond the study area. Consequently, the thickness of this aquifer increases in the south and is limited at the center (as shown in Figure 11). Based on our analysis, we estimate that the general inclination of these strata is approximately 0.3% from the south and north toward the center at the Litani River.

4.2. Potential Environmental and Ecological Issues

Sustainable water management is a significant challenge in the basin. Thus, the Litani River Basin represents a representative hydrological system for setting water-related SDGs (especially Goal 6).
The Litani River contributes to meeting the social, industrial, energy, and ecosystem functions of more than a million people living in its water basin. It is essential to the national economy.
The lower basin of the Litani River is still untouched by what the upper basin in the Bekaa region is suffering from, where the most significant part of the basin is located where the quality and quantity of water are deteriorating, affecting food safety, threatening sustainable agricultural practices and affecting the functions of ecosystems. The Lower Litani benefits from multiple springs with good water quality [83].
This proposal contributes to meeting the expected outputs from the invested resources and efforts and implementing Sustainable Development Goal 6, specifically Goals 6.1, 6.3, 6.4, and 6.5. These objectives apply to integrated river basin management, considering critical components of river basins, including water, ecosystems, socio-economic development, capacity, and data [163]. Therefore, future missions and programs dealing with environmental management of the Litani River do not need much work to bridge the gap between research and policy and improve water governance, legislation, and institutional capacities.

4.3. Future Research

Groundwater Modeling and Simulation

A model of a ground-water system can be described as a simplified representation of a real-world system that closely approximates its relevant excitation–response relationships. Models are typically simplified compared to reality, which is inherently complex, proposing a set of assumptions that capture the essence of the system and the features of its behavior that are pertinent to the specific problem being studied. Because of this simplification, there is no single model that fits every groundwater system.
The maps presented in Figure 24 describe the modeling of groundwater flow in the free Eocene aquifer (saturated thickness), which was completed using geographic information systems in the Darcy Flow and Darcy Velocity tools, where we obtained raster data to calculate head height from data from artesian wells drilled in the area (Figure 24).
The assumptions made in the model are those through which we can predict the movement of water within the Eocene layer and its directions, in addition to simulating the water table and its levels after the implementation of the dam project.
Although less complex than others, this modeling shows that the groundwater level will rise with the rise in the water level behind the dam to be completed, as these are considered preliminary simulations of what could become. The maps in Figure 24, each of which indicates a distinctive feature and hypothesis that contributes to understanding the movement of water and what the results of the project may lead to from a natural and operational perspective, support improvement of the vision for decision makers. These models confirm, from a scientific perspective and based on hydrogeological data and characteristics, that if this project is completed, the groundwater level will rise significantly, through which we can secure the community’s requirements of clean and sustainable water resources.
Groundwater flow assessment in rock masses is a key issue in the resolution of many engineering, geotechnical, and hydrogeological problems to reconstruct the flow path induced by tunneling and forecast the water inflow location, discharge rate, and water table drawdown [164].
The results of the modeling and simulation confirmed that the groundwater level will rise by the height of the dam on the Litani River, meaning that if we build a dam 100 m high from the base, the groundwater level will rise by 100 m from the bottom of the Eocene layer and over several kilometers to the south and north.
The modeling hypotheses also show that the stakeholders’ benefit from groundwater will be available at low economic costs due to the rise in the water table to areas close to the surface, which will secure large quantities of sustainable water resources for most of the towns located above the Eocene layer, which extends over a length of fifty kilometers.

5. Conclusions

This study created a practical scientific foundation for managing the issue of drought in the southern Lebanon region by utilizing groundwater engineering to address this transboundary problem.
It has been shown that there is a positive connection between groundwater drought and drought in general and that treating the former will have consequences for the latter.
According to the natural study, remote sensing, geographic information systems, and advanced mathematical algorithms were crucial in determining the optimal approach to recharging and supporting the Eocene aquifer for drought resistance and sustainable groundwater management. The research findings indicate that the successful implementation of geological engineering techniques to combat groundwater drought relies heavily on the distribution and structure of geological units. It was demonstrated that morphometric studies and the remote sensing of surface water resources effectively contributed to the identification of the best locations for constructing structures that artificially recharge groundwater. As noted, a dam that is between 30 and 160 m high should be built on the Litani River. This dam will help retain the abundant river water, which gravity can then direct into the Eocene aquifer through horizontal tunnels.
Through this process, we anticipate that the water table in the region will rise toward the surface, leading to the emergence of new springs in the valleys’ depths, which can then be contained through the construction of small dams. We categorized the water table into four distinct levels:
  • Level zero: This pertains to areas where the water table reaches the surface, forming sustainable streams, rivers, and swamps.
  • The 150 m level: These areas encompass locations where the water table is situated at a depth of 0 to 150 m below the ground surface.
  • The 250 m level: This level corresponds to areas where the water table is 150 to 250 m below the ground surface.
  • Depths exceeding 250 m: This level encompasses the remaining areas situated above the Eocene Strata, with the water table depth ranging from 250 to 540 m, extending to the base of the Eocene Strata.
This project aims to deliver over one billion cubic meters of clean and sustainable groundwater, meeting the domestic, agricultural, and industrial water demands in the study area.

Funding

IHE Delft, Netherlands 2023/080/111457/SMK.

Data Availability Statement

The current study required sets of geospatial data from different sources. These sources were mentioned in the study. However, the sources did not provide all the geospatial datasets needed; therefore, the authors of this study worked to generate a large amount of geospatial data, which were retrieved and processed from various satellite images, as was mentioned. In addition, the manipulated geospatial data were also obtained by the authors, who used GIS for this purpose. Therefore, the data availability of this work is owned by the authors and these data are ready to be provided to whomever needs to apply them in scientific research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, L.; Peng, Q.; Li, Z.; Cai, H. Evolution of drought characteristics and propagation from meteorological to agricultural drought under the influences of climate change and human activities. Environ. Sci. Pollut. Res. 2024, 31, 26713–26736. [Google Scholar] [CrossRef] [PubMed]
  2. Cheng, X.; Xu, Y.; Chen, J.; Liu, Q. The Impact of Climatic Conditions, Human Activities, and Catchment Characteristics on the Propagation from Meteorological to Agricultural and Hydrological Droughts in China. J. Geophys. Res. 2023, 128, 24. [Google Scholar] [CrossRef]
  3. Saha, A.; Pal, S.; Chowdhuri, I.; Roy, P.; Chakrabortty, R.; Shit, M. Vulnerability assessment of drought in India: Insights from meteorological, hydrological, agricultural and socio-economic perspectives. Glob. Planet. Chang. 2022, 123, 68–88. [Google Scholar] [CrossRef]
  4. Van Lanen, H.A.J.; Wanders, N.; Tallaksen, L.M.; Van Loon, A.F. Hydrological drought across the world: Impact of climate and physical catchment structure. Hydrol. Earth Syst. Sci. 2013, 17, 1715–1732. [Google Scholar] [CrossRef]
  5. Mishra, A.K.; Singh, V.P. A review of drought concepts. J. Hydrol. 2010, 391, 202–216. [Google Scholar] [CrossRef]
  6. Tallaksen, L.M.; Hisdal, H.; van Lanen, H.A.J. Space-time modelling of catchment scale drought characteristics. J. Hydrol. 2009, 375, 363–372. [Google Scholar] [CrossRef]
  7. Tallaksen, L.M.; Van Lanen, H.A.J. (Eds.) Hydrological Drought. Processes and Estimation Methods for Streamflow and Groundwater. In Developments in Water Science; Elsevier Science B.V.: Amsterdam, The Netherlands, 2004; Volume 48, p. 579. [Google Scholar]
  8. Fink, A.H.; Brucher, T.; Kruger, A.; Leckebush, G.C.; Pinto, J.G.; Ulbrich, U. The 2003 European summer heatwaves and drought—Synoptic diagnosis and impacts. Weather 2006, 59, 209–216. [Google Scholar] [CrossRef]
  9. Lloyd-Hughes, B.; Saunders, M.A. A drought climatology for Europe. Int. J. Climatol. 2002, 22, 1571–1592. [Google Scholar] [CrossRef]
  10. Zittis, G.; Almazroui, M.; Alpert, P.; Ciais, P.; Cramer, W.; Dahdal, Y.; Fnais, M.; Francis, D.; Hadjinicolaou, P.; Howari, F.; et al. Climate Change and Weather Extremes in the Eastern Mediterranean and Middle East. Rev. Geophys. 2022, 60, e2021RG000762. [Google Scholar] [CrossRef]
  11. Cramer, W.; Guiot, J.; Fader, M.; Garrabou, J.; Gattuso, J.P.; Iglesias, A.; Lange, M.A.; Lionello, P.; Llasat, M.C.; Paz, S.; et al. Climate change and interconnected risks to sustainable development in the Mediterranean. Nature Clim. Chang. 2018, 8, 972–980. [Google Scholar] [CrossRef]
  12. Lelieveld, J.; Hadjinicolaou, P.; Kostopoulou, E.; Chenoweth, J.; El Maayar, M.; Giannakopoulos, C.; Hannides, C.; Lange, M.A.; Tanarhte, M.; Tyrlis, E.; et al. Climate change and impacts in the Eastern Mediterranean and the Middle East. Clim. Chang. 2012, 114, 667–687. [Google Scholar] [CrossRef] [PubMed]
  13. Hochman, A.; Marra, F.; Messori, G.; Raveh-Rubin, S.; Yosef, I.; Zittis, G.; Pinto, J.G. Extreme weather and related societal impacts over the Eastern Mediterranean: A systematic review. Earth Syst. Dyn. 2022, 13, 749–777. [Google Scholar] [CrossRef]
  14. Spinoni, J.; Barbosa, P.; Füssel, H.-M.; McCormick, N.; Vogt, J.V.; Dosio, A. Global population-weighted degree-day projections for a combination of climate and socio-economic scenarios. Int. J. Climatol. 2021, 41, 5447–5464. [Google Scholar] [CrossRef]
  15. Zittis, G.; Hadjinicolaou, P.; Klangidou, M.; Proestos, Y.; Lelieveld, J. A multi-model, multi-scenario, and multi-domain analysis of regional climate projections for the Mediterranean. Reg. Environ. Chang. 2019, 19, 2621–2635. [Google Scholar] [CrossRef]
  16. Cook, I.; Anchukaitis, K.J.; Touchan, R.; Meko, M.; Cook, B.I. Spatiotemporal drought variability in the Mediterranean over the last 900 years. J. Geophys. Res. Atmos. 2016, 121, 1184–1199. [Google Scholar] [CrossRef] [PubMed]
  17. Guha-Sapir, D.; Below, R.; Hoyois, P. EM-DAT: International Disaster Database. Universitécatholique de Louvain—CRED, 2015, Brussels, Belgium. Available online: http://www.emdat.be/ (accessed on 25 December 2023).
  18. NOAA-NCEI. U.S. Billion-Dollar Weather and Climate Disasters. 2021. Available online: https://www.ncdc.noaa.gov/billions/ (accessed on 14 June 2023). [CrossRef]
  19. National Drought Mitigation Center. University of Nebraska-Lincoln. 2024. Available online: https://drought.unl.edu/ (accessed on 2 May 2024).
  20. ARSET—Drought Monitoring, Prediction, and Projection Using NASA Earth System Data. NASA Applied Remote Sensing Training Program. 2024. Available online: http://appliedsciences.nasa.gov/get-involved/training/english/arset-drought-monitoring-prediction-and-projection-using-nasa-earth (accessed on 24 July 2024).
  21. Farhat, N.; Sleiman, W. Potential natural obstacles to groundwater engineering. Civil Eng. Appl. J. 2023, 4, 46–61. [Google Scholar] [CrossRef]
  22. Stevanović, Z. Karst waters in potable water supply: A global scale overview. Environ. Earth Sci. 2019, 78, 1–12. [Google Scholar] [CrossRef]
  23. Peters, E.; Bier, G.; van Lanen, H.A.J.; Torfs, P.J.J.F. Propagation and spatial distribution of drought in a groundwater catchment. J. Hydrol. 2006, 321, 257–275. [Google Scholar] [CrossRef]
  24. Müller, J.D.; Gruber, N.; Carter, B.; Feely, R.; Ishii, M.; Lange, N.; Lauvset, S.K.; Murata, A.; Olsen, A.; Pérez, F.F.; et al. Decadal Trends in the Oceanic Storage of Anthropogenic Carbon from 1994 to 2014. AGU Adv. 2023, 4, e2023AV000875. [Google Scholar] [CrossRef]
  25. Leblanc, M.J.; Tregoning, P.; Ramillien, G.; Tweed, S.O.; Fakes, A. Basin-scale, integrated observations of the early 21st century multiyear drought in southeast Australia. Water Resour. Res. 2009, 45, W04408. [Google Scholar] [CrossRef]
  26. Peters, E.; van Lanen, H.A.J.; Torfs, P.J.J.F.; Bier, G. Drought in groundwater—Drought distribution and performance indicators. J. Hydrol. 2005, 306, 302–317. [Google Scholar] [CrossRef]
  27. Peters, E.; Torfs, P.J.J.F.; van Lanen, H.A.J.; Bier, G. Propagation of drought through groundwater—A new approach using linear reservoir theory. Hydrol. Process. 2003, 17, 3023–3040. [Google Scholar] [CrossRef]
  28. Eltahir, E.A.B.; Yeh, P.J.-F. On the asymmetric response of aquifer water level to floods and droughts in Illinois. Water Resour. Res. 1999, 35, 1199–1217. [Google Scholar] [CrossRef]
  29. Aranguren-Díaz, Y.; Galán-Freyle, N.J.; Guerra, A.; Manares-Romero, A.; Pacheco-Londoño, L.C.; Romero-Coronado, A.; Vidal-Figueroa, N.; Machado-Sierra, E. Aquifers and Groundwater: Challenges and Opportunities in Water Resource Management in Colombia. Water 2024, 16, 685. [Google Scholar] [CrossRef]
  30. Eduvie, M.; Musilim, I. Groundwater recharge, rainwater harvesting and regulations for sustainable water resources development in Nigeria. In Managed Groundwater Recharge and Rainwater Harvesting; Springer: Berlin/Heidelberg, Germany, 2024; pp. 1–15. [Google Scholar] [CrossRef]
  31. Fan, Y.; Miguez-Macho, G.; Weaver, C.P.; Walko, R.L.; Robock, A. Incorporating water table dynamics in climate modeling: 1. Water table observations and equilibrium water table simulations. J. Geophys. Res. 2007, 112, D10125. [Google Scholar] [CrossRef]
  32. Haitjema, H.M.; Mitchell-Bruker, S. Are water tables a subdued replica of the topography? Ground Water 2005, 43, 781–786. [Google Scholar] [CrossRef]
  33. Devito, K.; Creed, I.; Gan, T.; Mendoza, C.; Petrone, R.; Silins, U.; Smerdon, B. A framework for broad-scale classification of hydrologic response units on the Boreal Plain: Is topography the last thing to consider? Hydrol. Process. 2005, 19, 1705–1714. [Google Scholar] [CrossRef]
  34. Wolock, D.M.; Winter, T.C.; McMahon, G. Delineation and evaluation of hydrologic-landscape regions in the United States using geographic information system tools and multivariate statistical analyses. Environ. Manag. 2004, 34, S71–S88. [Google Scholar] [CrossRef]
  35. Winter, T.C. Relation of streams, lakes, and wetlands to groundwater flow systems. Hydrogeol. J. 1999, 7, 28–45. [Google Scholar] [CrossRef]
  36. Salvucci, G.D.; Entekhabi, D. Hillslope and climatic controls on hydrologic fluxes. Water Resour. Res. 1995, 31, 1725–1739. [Google Scholar] [CrossRef]
  37. Toth, J. A conceptual model of the groundwater regime and the hydrogeological environment. J. Hydrol. 1970, 10, 164–176. [Google Scholar] [CrossRef]
  38. Chen, F.; Crow, W.T.; Bindlish, R.; Colliander, A.; Burgin, M.S.; Asanuma, J.; Aida, K. Global-scale evaluation of SMAP, SMOS and ASCAT soil moisture products using triple collocation. Remote Sens. Environ. 2018, 214, 1–13. [Google Scholar] [CrossRef]
  39. Frash, L.P.; Carey, J.W.; Welch, N.J. Scalable en echelon shear-fracture aperture-roughness mechanism: Theory, validation, and implications. J. Geophys. Res. Solid Earth 2019, 124, 957–977. [Google Scholar] [CrossRef]
  40. McBeck, J.; Kandula, N.; Aiken, J.M.; Cordonnier, B.; Renard, F. Isolating the factors that govern fracture development in rocks throughout dynamic in situ X-ray tomography experiments. Geophys. Res. Lett. 2019, 46, 11127–11135. [Google Scholar] [CrossRef]
  41. Shoemaker, W.B.; Kuniansky, E.L.; Birk, S.; Bauer, S.; Swain, E.D. Documentation of a Conduit Flow Process (CFP) for MODFLOW-2005 (Vol. 6); US Department of the Interior, US Geological Survey: Reston, VA, USA, 2008. Available online: http://www.usgs.gov/pubprod (accessed on 12 June 2024).
  42. Bonnet, E.; Bour, O.; Odling, N.E.; Davy, P.; Main, I.; Cowie, P.; Berkowitz, B. Scaling of fracture systems in geological media. Rev. Geophys. 2001, 39, 347–383. [Google Scholar] [CrossRef]
  43. Javadi, M.; Sharifzadeh, M.; Shahriar, K.; Mitani, Y. Critical Reynolds number for nonlinear flow through rough-walled fractures: The role of shear processes. Water Resour. Res. 2014, 50, 1789–1804. [Google Scholar] [CrossRef]
  44. Rutqvist, J.; Tsang, C.F. Multiphysics processes in partially saturated fractured rock: Experiments and models from Yucca Mountain. Rev. Geophys. 2012, 50, 65–81. [Google Scholar] [CrossRef]
  45. Hyman, J.D.; Aldrich, G.; Viswanathan, H.; Makedonska, N.; Karra, S. Fracture size and transmissivity correlations: Implications for transport simulations in sparse three-dimensional discrete fracture networks following a truncated power law distribution of fracture size. Water Resour. Res. 2016, 52, 6472–6489. [Google Scholar] [CrossRef]
  46. Antonellini, M.; Mollema, P.N.; Del Sole, L. Application of analytical diffusion models to outcrop observations: Implications for mass transport by fluid flow through fractures. Water Resour. Res. 2017, 53, 5545–5566. [Google Scholar] [CrossRef]
  47. Maldaner, C.H.; Munn, J.D.; Coleman, T.I.; Molson, J.W.; Parker, B.L. Groundwater flow quantification in fractured rock boreholes using active distributed temperature sensing under natural gradient conditions. Water Resour. Res. 2019, 55, 3285–3306. [Google Scholar] [CrossRef]
  48. Ford, D.; Williams, P.D. Karst Hydrogeology and Geomorphology; John Wiley & Sons: Hoboken, NJ, USA, 2007. [Google Scholar] [CrossRef]
  49. Stevanović, Z. Karst Aquifers-Characterization and Engineering; Springer: Berlin/Heidelberg, Germany, 2015; Volume 44. [Google Scholar] [CrossRef]
  50. IGRAC. Groundwater overview. In Making the Invisible Visible; International Groundwater Resources Assessment Centre: Delft, The Netherlands, 2018; pp. 1–60. [Google Scholar]
  51. Wan, W.; Döll, P.; Schmied, H.M. Global-scale groundwater recharge modeling is improved by tuning against ground-based estimates for karst and non-karst areas. Water Resour. Res. 2024, 60, e2023WR036182. [Google Scholar] [CrossRef]
  52. Gleeson, T.; Cuthbert, M.O.; Ferguson, G.; Perrone, D. Global groundwater sustainability, resources, and systems in the Anthropocene. Annu. Rev. Earth Planet. Sci. 2020, 48, 431–463. [Google Scholar] [CrossRef]
  53. Bucheli, J.; Dalhaus, T.; Finger, R. The optimal drought index for designing weather index insurance. Eur. Rev. Agric. Econ. 2021, 48, 573–597. [Google Scholar] [CrossRef]
  54. Nath, S.; Shyanti, R.K.; Nath, Y. Influence of anthropocene climate change on biodiversity loss in different ecosystems. In Global Climate Change; Elsevier: Amsterdam, The Netherlands, 2021; pp. 63–78. [Google Scholar] [CrossRef]
  55. Peace, N. Impact of climate change on insect, pest, disease, and animal biodiversity. Int. J. Environ. Sci. Nat. Res. 2020, 23, 151–153. [Google Scholar] [CrossRef]
  56. Cuthbert, M.O.; Gleeson, T.; Moosdorf, N.; Befus, K.M.; Schneider, A.; Hartmann, J.; Lehner, B. Global patterns and dynamics of climate–groundwater interactions. Nature Clim. Chang. 2019, 9, 137–141. [Google Scholar] [CrossRef]
  57. Döll, P. Vulnerability to the impact of climate change on renewable groundwater resources: A global-scale assessment. Environ. Res. Lett. 2009, 4, 035006. [Google Scholar] [CrossRef]
  58. Green, T.R.; Taniguchi, M.; Kooi, H.; Gurdak, J.J.; Allen, D.M.; Hiscock, K.M.; Aureli, A. Beneath the surface of global change: Impacts of climate change on groundwater. J. Hydrol. 2011, 405, 532–560. [Google Scholar] [CrossRef]
  59. Maxwell, R.M.; Condon, L.E. Connections between groundwater flow and transpiration partitioning. Science 2016, 353, 377–380. [Google Scholar] [CrossRef]
  60. Maxwell, R.M.; Kollet, S.J. Interdependence of groundwater dynamics and land-energy feedbacks under climate change. Nat. Geosci. 2008, 1, 665–669. [Google Scholar] [CrossRef]
  61. Colin, J.; Decharme, B.; Cattiaux, J.; Saint-Martin, D. Groundwater Feedbacks on Climate Change in the CNRM Global Climate Model. J. Climate 2023, 36, 7599–7617. [Google Scholar] [CrossRef]
  62. Berry, B. Inductive Approach to the Regionalization of Economic Development. Res. Pap. Univ. Chicago Dep. Geogr. 1960, 6, 83–84. [Google Scholar] [CrossRef]
  63. Pearson, K. Mathematical contributions to the theory of evolution. III. Regression, heredity, and panmixia. Phil. Trans. R. Soc. A 1896, 187, 253–318. [Google Scholar]
  64. Schober, P.; Boer, C.; Schwarte, L.A. Correlation Coefficients: Appropriate Use and Interpretation. Anesth. Analg. 2018, 126, 1763–1768. [Google Scholar] [CrossRef] [PubMed]
  65. Royston, P. Remark AS R94, a remark on Algorithm AS 181, the W-test for normality. J. R. Stat. Soc. Ser. C Appl. Stat. 1995, 44, 547–551. [Google Scholar] [CrossRef]
  66. Alibou, J. Impact des Changements Climatiques sur les Ressources en eau et les Zones Humides du Maroc; Centre d’Etudes et de Recherches sur les Systèmes Hydrauliques et Environnementaux, EHTP: Maroc, Rabat, 2002. [Google Scholar]
  67. European Environment Agency (EEA). Climate Change, Impacts and Vulnerability in Europe (2012)—An Indicator-Based Report. Report No 12/2012. Available online: http://www.eea.europa.eu/publications/climate-impacts-and-vulnerability-2012 (accessed on 14 June 2023).
  68. Farhat, N. Effect of relative humidity on evaporation rates in Nabatieh Region. Leb. Sci. J. 2018, 19, 59–66. [Google Scholar] [CrossRef]
  69. Panizza, M. Geomorfologia; Pitagora Editrice: Bologna, Italia, 2007. [Google Scholar]
  70. Bloomfield, J.P.; Marchant, B.P. Analysis of groundwater drought building on the standardised precipitation index approach. Hydrol. Earth Syst. Sci. 2013, 17, 4769–4787. [Google Scholar] [CrossRef]
  71. Apaydin, H.; Sonmez, F.K.; Yildirim, Y.E. Spatial interpolation techniques for climate data in the GAP region in Turkey. Clim. Res. 2004, 28, 31–40. [Google Scholar] [CrossRef]
  72. Farhat, N.; Fadel, B.; Damirji, B. A model based on GIS and climate criteria to locate optimal solar power plant sites. J. Agric. For. Meteorol. Stud. 2023, 2, 1–13. Available online: https://manuscriptscientific.com/article/5/journal-of-agriculture-and-forest-meteorology-studies/180/1 (accessed on 11 June 2023). [CrossRef]
  73. Holper, P. Cloud Seeding, Trying to Artificially Generate Additional Rainfall from Clouds. CSIRO, Marine and Atmospheric Research. 2001. Available online: https://www.cmar.csiro.au/e-print/open/holper_2001c.htm (accessed on 5 June 2024).
  74. Kelley, C.; Mohtadi, S.; Cane, M.A.; Seager, R.; Kushnir, Y. Climate change in the Fertile Crescent and implications of the recent Syrian drought. Proc. Natl. Acad. Sci. USA 2015, 112, 3241–3246. [Google Scholar] [CrossRef]
  75. Sgobbi, A.; Carraro, C. Climate Change Impacts and Adaptation Strategies In Italy. An Economic Assessment. Fondazione Eni Enrico Mattei. 2008. Available online: https://ideas.repec.org/p/fem/femwpa/2008.6.html (accessed on 3 April 2024).
  76. Friedrich, T.; Timmermann, A.; Tigchelaar, M.; Timm, O.E.; Ganopolski, A. Nonlinear climate sensitivity and its implications for future greenhouse warming. Sci. Adv. 2016, 2, e1501923. [Google Scholar] [CrossRef]
  77. McEnvoy, D.; Huntington, J.L.; Hobbins, M.T.; Wood, A.; Morton, C.; Anderson, M.; Hain, C.; Verdin, J. The Evaporative Demand Drought Index. Part II: CONUS-Wide Assessment against Common Drought Indicators. J. Hydrometeorol. 2016, 17, 1763–1779. [Google Scholar] [CrossRef]
  78. Flageollet, J.C. Atlas du réchauffement climatique 1971–2010. Physio-Géo 2012, 6. [Google Scholar] [CrossRef]
  79. Schleussner, C.-F.; Lissner, T.; Fischer, E.; Wohland, J.; Perrette, M.; Golly, A.; Rogelj, J.; Childers, K.; Schewe, J.; Frieler, K.; et al. Differential climate impacts for policy-relevant limits to global warming: The case of 1.5 °C and 2 °C. Earth Syst. Dynam. 2016, 7, 327–351. [Google Scholar] [CrossRef]
  80. Cushman, J.H.; Tartakovsky, D.M. (Eds.) The Handbook of Groundwater Engineering, 3rd ed.; CRC Press: Boca Raton, FL, USA, 2016. [Google Scholar] [CrossRef]
  81. Bergaoui, K.; Mitchell, D.; Otto, F.; Allen, M.; Zaaboul, R.; McDonnell, R. The Contribution of Human-Induced Climate Change to the Drought of 2014 in the Southern Levant Region. Bull. Am. Meteorol. Soc. 2015, 96, S66–S70. [Google Scholar] [CrossRef]
  82. Rodríguez, A. Water Resource Distribution and Tectonics in the Tulum Valley, Western Central Argentina. Int. J. Hydrol. 2018, 2, 55–59. [Google Scholar] [CrossRef]
  83. Shaban, A.; Keyrouz, L. The geological controls of geothermal groundwater sources in Lebanon. Int. J. Energy Environ. 2013, 4, 787–796. Available online: https://www.ijee.ieefoundation.org/vol4/issue5/IJEE_04_v4n5.pdf (accessed on 5 January 2024).
  84. Freund, R.; Garfunkel, Z.; Zak, I.; Goldberg, M.; Weisbrod, T.; Derin, B. The shear along the Dead Sea rift. Philos. Trans. R. Soc. London Ser. A 1970, 267, 107–130. Available online: https://www.jstor.org/stable/73612 (accessed on 2 March 2023).
  85. Quennell, A.M. Tectonics of the Dead Sea rift. In Proceedings of the 20th International Geological Congress, Mexico City, Mexico, 4–11 September 1956; pp. 385–403. [Google Scholar]
  86. Sanlaville, P. Etude Géomorphologique de la Région Littorale du Liban; Université Libanaise: Beirut, Lebanon, 1977. [Google Scholar]
  87. PNUD. Programme des Nations Unies pour le Développement. In Rapport Final Intitulé: “Liban Etude des Eaux Souterraines”; NATIONS UNIES: New York, NY, USA, 1970. [Google Scholar]
  88. Brown, S.R. Simple mathematical model of a rough fracture. J. Geophys. Res. Solid Earth 1995, 100, 5941–5952. [Google Scholar] [CrossRef]
  89. Yang, Z.; Xue, S.; Zheng, X.; Chen, Y.F. Partitioning dynamics of gravity-driven unsaturated flow through simple T-shaped fracture intersections. Water Resour. Res. 2019, 55, 7130–7142. [Google Scholar] [CrossRef]
  90. Guo, J.; Zhan, L.; Gou, B.; Zhang, R.; Liu, C.; Li, X.; Ren, J. Formation of fractures in carbonate rocks by pad acid fracturing with different states of carbon dioxide. Petroleum Explor. Dev. 2021, 48, 744–751. [Google Scholar] [CrossRef]
  91. Shaban, A.; Khawlie, M.; Abdallah, C. Use of remote sensing and GIS to determine recharge potential zones: The case of Occidental Lebanon. Hydrogeol. J. 2006, 14, 433–443. [Google Scholar] [CrossRef]
  92. Dubertret, L. Carte géologique de la Syrie et de Liban au 1/50000me; Minister des Travaux Publics: Beyrouth, Lebanon, 1953. [Google Scholar]
  93. Shaban, A.; Farhat, N.; Fadel, B.; Sheib, A.; Bitar, A.; Darwish, D. LithoSFR Model for Mapping Groundwater Potential Zones Using Remote Sensing and GIS. Water 2024, 16, 1951. [Google Scholar] [CrossRef]
  94. Shaban, A.; Bou Kheir, R.; Froidefond, J.; Khawlie, M.; Girard, M. Characterisation of morphometric factors of hydrographic networks corresponding to the infiltration capacities of rocks in western Lebanon. Z. Für Geomorphol. 2004, 48, 79–94. [Google Scholar] [CrossRef]
  95. Nader, F. The Petroleum prospectivity of Lebanon: An overview. J. Petrol. Geol. 2011, 34, 135–156. [Google Scholar] [CrossRef]
  96. Condon, L.E.; Maxwell, R.M. Evaluating the relationship between topography and groundwater using outputs from a continental-scale integrated hydrology model. Water Resour. Res. 2015, 51, 6602–6621. [Google Scholar] [CrossRef]
  97. Gleeson, T.; Marklund, L.; Smith, L.; Manning, A.H. Classifying the water table at regional to continental scales. Geophys. Res. Lett. 2011, 38, L05401. [Google Scholar] [CrossRef]
  98. Farhat, N.; Makke, H.; Fadel, B.; Ayoub, M.; Koubeissi, A. Remote sensing and GIS to assess land degradation in Wadi El Hjeir, Lebanon. Plants Environ. 2022, 4, 10–15. [Google Scholar] [CrossRef]
  99. Zhou, C.; Chen, Y.; Hu, R.; Yang, Z. Groundwater flow through fractured rocks and seepage control in geotechnical engineering: Theories and practices. J. Rock Mech. Geotech. Eng. 2022, 15, 1–36. [Google Scholar] [CrossRef]
  100. Guerre, A. Étude hydrogéologique préliminaire des karsts libanais. Hannon 1969, 4, 64–92. [Google Scholar]
  101. Hakim, B. Recherches Hydrologiques et Hydrochimiques sur Quelques Karsts Méditerranéens: Liban, Syrie et Maroc; Publications de l’Université Libanaise; Section des Études Géographiques Lebanon: Beirut, Lebanon, 1985; Volume II, 701p. [Google Scholar]
  102. Edgell, H.S. Karst and hydrogeology of Lebanon. Carbonates Evaporites 1997, 12, 220–235. [Google Scholar] [CrossRef]
  103. Azar, D. Contribution à L'étude de la Source Karstique d'Afka; Université Saint-Joseph: Liban, Lebanon, 2000. [Google Scholar]
  104. Babaeian, E.; Sadeghi, M.; Jones, S.B.; Montzka, C.; Vereecken, H.; Tuller, M. Ground, Proximal, and Satellite Remote Sensing of Soil Moisture. Rev. Geophys. 2019, 57, 530–616. [Google Scholar] [CrossRef]
  105. Chen, Y.F.; Ling, X.M.; Liu, M.M.; Hu, R.; Yang, Z. Statistical distribution of hydraulic conductivity of rocks in deep-incised valleys, southwest China. J. Hydrol. 2018, 566, 216–226. [Google Scholar] [CrossRef]
  106. Pierdicca, N.; Fascetti, F.; Pulvirenti, L.; Crapolicchio, R. Error characterization of soil moisture satellite products: Retrieving error cross-correlation through extended quadruple collocation. IEEE J. Sel. Topics Appl. Earth Obs. Rem. Sens. 2017, 10, 4522–4530. [Google Scholar] [CrossRef]
  107. Gruber, A.; Su, C.; Zwieback, S.; Crow, W.; Dorigo, W.; Wagner, W. Recent advances in (soil moisture) triple collocation analysis. Int. J. Appl. Earth Obs. Geoinf. 2016, 45, 200–211. [Google Scholar] [CrossRef]
  108. Kerr, Y.H.; Al-Yaari, A.; Rodríguez-Fernández, N.; Parrens, M.; Molero, B.; Leroux, D.; Bircher, S.; Mahmoodi, A.; Mialon, A.; Richaume, P.; et al. Overview of SMOS performance in terms of global soil moisture monitoring after six years in operation. Remote Sens. Environ. 2016, 180, 40–63. [Google Scholar] [CrossRef]
  109. Bartalis, Z.; Wagner, W.; Naeimi, V.; Hasenauer, S.; Scipal, K.; Bonekamp, H. Initial soil moisture retrievals from the METOP-A Advanced Scatterometer (ASCAT). Geophys. Res. Lett. 2018, 34, L20401. [Google Scholar] [CrossRef]
  110. Bartalis, Z.; Scipal, K.; Wagner, W. Azimuthal anisotropy of scatterometer measurements over land. IEEE Trans. Geosci. Remote Sens. 2006, 44, 2083–2092. [Google Scholar] [CrossRef]
  111. Wagner, W.; Dorigo, W.; de Jeu, R.; Fernandez, D.; Benveniste, J.; Haas, E.; Ertl, M. Fusion of active and passive microwave observations to create an essential climate variable data record on soil moisture. ISPRS Ann. Photogramm. Remote Sens. Spatial Inf. Sci. 2012, 7, 315–321. [Google Scholar] [CrossRef]
  112. Ceballos, A.; Scipal, K.; Wagner, W.; Martínez-Fernández, J. Validation of ERS scatterometer-derived soil moisture data in the central part of the Duero Basin, Spain. Hydrol. Process. 2005, 19, 1549–1566. [Google Scholar] [CrossRef]
  113. Lei, S.J.; Bian, Z.F.; Daniels, J.L.; Liu, D.L. Improved spatial resolution in soil moisture retrieval at arid mining areas using apparent thermal inertia. Trans. Nonferrous Met. Soc. China 2014, 24, 1866–1873. [Google Scholar] [CrossRef]
  114. Rahimzadeh-Bajgiran, P.; Berg, A.A.; Champagne, C.; Omasa, K. Estimation of soil moisture using optical/thermal infrared remote sensing in the Canadian Prairies. ISPRS J. Photogramm. Remote Sens. 2013, 83, 94–103. [Google Scholar] [CrossRef]
  115. Champagne, C.; McNairn, H.; Berg, A. Monitoring agricultural soil moisture extremes in Canada using passive microwave remote sensing. Remote Sens. Environ. 2011, 115, 2434–2444. [Google Scholar] [CrossRef]
  116. Verstraeten, W.W.; Veroustraete, F.; Van der Sande, C.J.; Grootaers, I.; Feyen, J. Soil moisture retrieval using thermal inertia, determined with visible and thermal spaceborne data, validated for European forests. Remote Sens. Environ. 2006, 101, 299–314. [Google Scholar] [CrossRef]
  117. Sadeghi, M.; Jones, B.S.; Philpot, W.D. A linear physically-based model for remote sensing of soil moisture using shortwave infrared bands. Remote Sens. Environ. 2015, 164, 66–76. [Google Scholar] [CrossRef]
  118. Yang, Z.; Neuweiler, I.; Méheust, Y.; Fagerlund, F.; Niemi, A. Fluid trapping during capillary displacement in fractures. Adv. Water Resour. 2016, 95, 264–275. [Google Scholar] [CrossRef]
  119. Al-Yaari, A.; Wigneron, J.; Ducharne, A.; Kerr, Y.; De Rosnay, P.; De Jeu, R.; Govind, A.; Al Bitar, A.; Albergel, C.; Muñoz-Sabater, J.; et al. Global-scale evaluation of two satellite-based passive microwave soil moisture datasets (SMOS and AMSR-E) with respect to Land Data Assimilation System estimates. Remote Sens. Environ. 2014, 149, 181–195. [Google Scholar] [CrossRef]
  120. Gutman, G.; Ignatov, A. The derivation of the green vegetation fraction from NOAA/AVHRR data for use in numerical weather prediction models. Int. J. Remote Sens. 1998, 19, 1533–1543. [Google Scholar] [CrossRef]
  121. ARSET—Application of NASA SPoRT-Land Information System (SPoRT-LIS) Soil Moisture Data for Drought. NASA Applied Remote Sensing Training Program. 2023. Available online: http://appliedsciences.nasa.gov/get-involved/training/english/arset-application-nasa-sport-land-information-system-sport-lis-soil (accessed on 25 May 2024).
  122. Khawlie, M. The Impacts of Climate Change on Water Resources of Lebanon-Eastern Mediterranean; Fondazione Eni Enrico Mattei FEEM: Milan, Italy, 2001. [Google Scholar] [CrossRef]
  123. Todd, D.K. Discussion of infiltration galleries. Proc. Amer. Soc. Civil Engrs. 1955, 81, 7–9. [Google Scholar]
  124. Dole, B. Use of fluoresceine in the study of underground water. U.S. Geol. Surv. Water-Supply Pap. 1906, 160, 73–86. [Google Scholar]
  125. Darcy, H. Les Fontaines Publiques de la ville de Dijon; V. Dahmont: Paris, France, 1856. [Google Scholar]
  126. Leap, D.I.; Belmonte, P.M. Influence of pore pressure on apparent dispersivity of a fissured dolomitic aquifer. Groundwater 1991, 30, 87–95. [Google Scholar] [CrossRef]
  127. Gallagher, A.S. Aquatic Habitat Assessment; American Fisheries Society: Bethesda, MD, USA, 1999; pp. 25–34. ISBN 1-888569-18-2. [Google Scholar]
  128. Glennon, A.; Groves, C. An examination of perennial stream drainage patterns within the Mammoth Cave watershed, Kentucky. J. Cave Karst Stud. 2002, 64, 82–91. [Google Scholar]
  129. Schumm, S.A. The Fluvial System; Wiley: New York, NY, USA, 1977; 338p. [Google Scholar] [CrossRef]
  130. Furbish, D.J. Spatial autoregressive structure in meander evolution. Geol. Soc. Am. Bull. 1991, 103, 1576–1589. [Google Scholar] [CrossRef]
  131. Horton, R.E. Erosional development of streams and their drainage basins; hydro-physical approach to quantitative morphology. Bull. Geol. Soc. Am. 1945, 56, 275–370. [Google Scholar] [CrossRef]
  132. Berkowitz, B. Characterizing flow and transport in fractured geological media: A review. Adv. Water Resour. 2002, 25, 861–884. [Google Scholar] [CrossRef]
  133. Meakin, P.; Tartakovsky, A.M. Modeling and simulation of pore-scale multiphase fluid flow and reactive transport in fractured and porous media. Rev. Geophys. 2009, 47. [Google Scholar] [CrossRef]
  134. Lei, Q.; Latham, J.P.; Tsang, C.F. The use of discrete fracture networks for modelling coupled geomechanical and hydrological behaviour of fractured rocks. Comput. Geotech. 2017, 85, 151–176. [Google Scholar] [CrossRef]
  135. Brashears, M.L. Recharging ground-water reservoirs with wells and basins. Min. Eng. 1953, 5, 1029–1032. Available online: https://www.jstor.org/stable/i40056498 (accessed on 19 September 2023).
  136. Jansa, V. Artificial Replenishment of Underground Water; International Water Supply Association: Paris, France, 1952; 105p. [Google Scholar]
  137. Butler, R.G.; Orlob, G.T.; McGauhey, P.H. Underground Movement of Bacterial and Chemical Pollutants. J. Am. Water Works Assoc. 1954, 46, 97–111. Available online: http://www.jstor.org/stable/41253595 (accessed on 19 September 2023). [CrossRef]
  138. Ferris, J.G. Water spreading and recharge wells. In Proc. Indiana Water Conserv. Conf.; Indiana Dept. Conserv.: Indianapolis, IN, USA, 1950; pp. 52–59. [Google Scholar]
  139. Neuman, S.P. Trends, prospects and challenges in quantifying flow and transport through fractured rocks. Hydrogeol. J. 2005, 13, 124–147. [Google Scholar] [CrossRef]
  140. Hu, R.; Zhou, C.X.; Wu, D.S.; Yang, Z.; Chen, Y.F. Roughness control on multiphase flow in rock fractures. Geophys. Res. Lett. 2019, 46, 12002–12011. [Google Scholar] [CrossRef]
  141. Olsson, R.; Barton, N. An improved model for hydromechanical coupling during shearing of rock joints. Int. J. Rock Mech. Min. Sci. 2001, 38, 317–329. [Google Scholar] [CrossRef]
  142. Da Silva, J.A.; Kang, P.K.; Yang, Z.; Cueto-Felgueroso, L.; Juanes, R. Impact of confining stress on capillary pressure behavior during drainage through rough fractures. Geophys. Res. Lett. 2019, 46, 7424–7436. [Google Scholar] [CrossRef]
  143. Sonderegger, A. Hydraulic phenomena and the effect of spreading of flood water southern california. Trans. Am. Soc. Civ. Eng. 1918, 82, 802–825. [Google Scholar] [CrossRef]
  144. Ruiz, C.; Recreo, F.; Prado, P.; Campos, R.; Pelayo, M.; de la Losa, A.; Hurtado, A.; Lomba, L.; Pérez del Villar, L.; Martínez, R.; et al. Almacenamiento geológico de CO2. Criterios de selección de emplazamientos. Inf. Técnicos CIEMAT 2007, 1106, 100. [Google Scholar]
  145. Chen, Y.F.; Ye, Y.; Hu, R.; Yang, Z.; Zhou, C.B. Modeling unsaturated flow in fractured rocks with scaling relationships between hydraulic parameters. J. Rock Mech. Geotech. Eng. 2022, 14, 1697–1709. [Google Scholar] [CrossRef]
  146. Biot, M.A. General theory of three-dimensional consolidation. J. Appl. Phys. 1941, 12, 155–164. [Google Scholar] [CrossRef]
  147. Hu, R.; Chen, Y.F.; Liu, H.H.; Zhou, C.B. A coupled two-phase fluid flow and elastoplastic deformation model for unsaturated soils: Theory, implementation, and application. Int. J. Numer. Anal. Methods Geomech. 2016, 40, 1023–1058. [Google Scholar] [CrossRef]
  148. Zhou, C.; Chen, Y.; Jiang, Q.; Lu, W. A generalized multi-field coupling approach and its application to stability and deformation control of a high slope. J. Rock Mech. Geotech. Eng. 2011, 3, 193–206. [Google Scholar] [CrossRef]
  149. Schumm, S.A.; Dumont, J.F.; Holbrook, J.M. Active Tectonics and Alluvial Rivers; Cambridge University Press: New York, NY, USA, 2002; pp. 210–243. Available online: https://ui.adsabs.harvard.edu/link_gateway/2002JQS....17..800S/doi:10.1002/jqs.698 (accessed on 13 March 2024).
  150. Hickin, E.J.; Nanson, G.C. The character of channel migration on the Beatton river, northeast British Columbia, Canada. Geol. Soc. Am. Bull. 1975, 86, 487–494. [Google Scholar] [CrossRef]
  151. Adham, M.I.; Jahan, C.S.; Mazumder, Q.H. Study on groundwater recharge potentiality of Barind Tract, Rajshahi District, Bangladesh using GIS and Remote Sensing technique. J. Geol. Soc. India 2010, 75, 432–438. [Google Scholar] [CrossRef]
  152. Bouwer, H. Artificial recharge of groundwater: Hydrogeology and engineering. Hydrogeol. J. 2002, 10, 121–142. [Google Scholar] [CrossRef]
  153. Chowdhury, A.; Jha, M.K.; Chowdary, V.M. Delineation of groundwater recharge zones and identification of artificial recharge sites in West Medinipur district, West Bengal, using RS, GIS and MCDM techniques. Environ. Earth Sci. 2010, 59, 1209–1222. [Google Scholar] [CrossRef]
  154. Christy, R.M.; Lakshmanan, E. Percolation pond as a method of managed aquifer recharge in a coastal saline aquifer: A case study on the criteria for site selection and its impacts. J. Earth Syst. Sci. 2017, 126, 66. [Google Scholar] [CrossRef]
  155. Russo, T.A.; Fisher, A.T.; Lockwood, B.S. Assessment of Managed Aquifer Recharge Site Suitability Using a GIS and Modeling. Groundwater 2015, 53, 389–400. [Google Scholar] [CrossRef]
  156. Dar, T.; Rai, N.; Bhat, A. Delineation of potential groundwater recharge zones using analytical hierarchy process (AHP). Geol. Ecol. Landsc. 2021, 5, 292–307. [Google Scholar] [CrossRef]
  157. Gesim, N.A.; Okazaki, T. Identification of groundwater artificial recharge sites in Herat city, Afghanistan, using Fuzzy logic. Int. J. Eng. Technol. Res. 2018, 8, 40–45. [Google Scholar]
  158. Heil, J.; Häring, V.; Marschner, B.; Stumpe, B. Advantages of fuzzy k-means over k-means clustering in the classification of diffuse reflectance soil spectra: A case study with West African soils. Geoderma 2019, 337, 11–21. [Google Scholar] [CrossRef]
  159. Javadi, S.; Saatsaz, M.; Hashemy Shahdany, S.M.; Neshat, A.; Ghordoyee Milan, S.; Akbari, S. A new hybrid framework of site selection for groundwater recharge. Geosci. Front. 2021, 12, 101144. [Google Scholar] [CrossRef]
  160. Laverty, F.B. Ground-water recharge. J. Amer. Water Works Assoc. 1952, 44, 677–681. [Google Scholar] [CrossRef]
  161. Prabhu, M.V.; Venkateswaran, S. Delineation of Artificial Recharge Zones Using Geospatial Techniques in Sarabanga Sub Basin Cauvery River, Tamil Nadu. Aquat. Procedia 2014, 4, 1265–1274. [Google Scholar] [CrossRef]
  162. Schiavo, M. The role of different sources of uncertainty on the stochastic quantification of subsurface discharges in heterogeneous aquifers. J. Hydrol. 2023, 617, 128930. [Google Scholar] [CrossRef]
  163. Ge, Y.; Li, X.; Cai, X.; Deng, X.; Wu, F.; Li, Z.; Luan, W. Converting UN Sustainable Development Goals (SDGs) to Decision-Making Objectives and Implementation Options at the River Basin Scale. Sustainability 2018, 10, 1056. [Google Scholar] [CrossRef]
  164. Gattinoni, P.; Scesi, L. Short review of some methods for groundwater flow assessment in fractured rock masses. Acque Sotter. Ital. J. Groundw. 2018, 7, 7–17. [Google Scholar] [CrossRef]
Figure 1. The study area in Lebanon.
Figure 1. The study area in Lebanon.
Hydrology 11 00156 g001
Figure 2. Annual rainfall in the study area.
Figure 2. Annual rainfall in the study area.
Hydrology 11 00156 g002
Figure 3. Trend in temperature increase.
Figure 3. Trend in temperature increase.
Hydrology 11 00156 g003
Figure 4. The lithological map of the study area.
Figure 4. The lithological map of the study area.
Hydrology 11 00156 g004
Figure 5. A geological section from east to west showing the fractures and the inclination of the geological strata in addition to the Bint Jbeil syncline. E2: Eocene; C6: Senonian; C4: Cenomanian.
Figure 5. A geological section from east to west showing the fractures and the inclination of the geological strata in addition to the Bint Jbeil syncline. E2: Eocene; C6: Senonian; C4: Cenomanian.
Hydrology 11 00156 g005
Figure 6. Stratum with the lineaments and faults in the study area.
Figure 6. Stratum with the lineaments and faults in the study area.
Hydrology 11 00156 g006
Figure 7. The topography of the region.
Figure 7. The topography of the region.
Hydrology 11 00156 g007
Figure 8. The Standardized Precipitation Index (SPI) from 1981 to 2024.
Figure 8. The Standardized Precipitation Index (SPI) from 1981 to 2024.
Hydrology 11 00156 g008
Figure 9. Vegetation Condition Index (VCI).
Figure 9. Vegetation Condition Index (VCI).
Hydrology 11 00156 g009
Figure 10. Soil Moisture Anomaly Index (SM).
Figure 10. Soil Moisture Anomaly Index (SM).
Hydrology 11 00156 g010
Figure 11. DEM with dimension of the Eocene aquifer.
Figure 11. DEM with dimension of the Eocene aquifer.
Hydrology 11 00156 g011
Figure 12. Geological section of the syncline of the Eocene aquifer. C6 is the Senonian isolate layer; C4 is the Cenomanian aquifer; E2 is the Eocene saturated thickness.
Figure 12. Geological section of the syncline of the Eocene aquifer. C6 is the Senonian isolate layer; C4 is the Cenomanian aquifer; E2 is the Eocene saturated thickness.
Hydrology 11 00156 g012
Figure 13. Aquifers in the study area.
Figure 13. Aquifers in the study area.
Hydrology 11 00156 g013
Figure 14. Hydrological map of the study area.
Figure 14. Hydrological map of the study area.
Hydrology 11 00156 g014
Figure 15. Litani watershed in the study area.
Figure 15. Litani watershed in the study area.
Hydrology 11 00156 g015
Figure 16. The Litani River in its valley and the optimal site of the dam.
Figure 16. The Litani River in its valley and the optimal site of the dam.
Hydrology 11 00156 g016aHydrology 11 00156 g016b
Figure 17. The water table in the Eocene aquifer showing its relationship with the proposed tunnels.
Figure 17. The water table in the Eocene aquifer showing its relationship with the proposed tunnels.
Hydrology 11 00156 g017
Figure 18. Topographic section from the dam to the southern part of the study area showing the tunnels and the water table.
Figure 18. Topographic section from the dam to the southern part of the study area showing the tunnels and the water table.
Hydrology 11 00156 g018
Figure 19. Regional depth of the water table following filling of the Litani River reservoir.
Figure 19. Regional depth of the water table following filling of the Litani River reservoir.
Hydrology 11 00156 g019
Figure 20. This figure shows the potential groundwater level in a very small area within deep valleys where the water table is theoretically supposed to approach the ground surface.
Figure 20. This figure shows the potential groundwater level in a very small area within deep valleys where the water table is theoretically supposed to approach the ground surface.
Hydrology 11 00156 g020
Figure 21. Areas where the water table will be at depths ranging from 1 to 150 m.
Figure 21. Areas where the water table will be at depths ranging from 1 to 150 m.
Hydrology 11 00156 g021
Figure 22. Areas where the water table will be at depths ranging from 150 to 250 m.
Figure 22. Areas where the water table will be at depths ranging from 150 to 250 m.
Hydrology 11 00156 g022
Figure 23. Areas where the water table will be at depths greater than 250 m.
Figure 23. Areas where the water table will be at depths greater than 250 m.
Hydrology 11 00156 g023
Figure 24. Calculation of the groundwater volume balance residual and seepage velocity vector (direction and magnitude) for steady flow in an aquifer by GIS. (A) Head elevation: the head elevation raster comes from various sources. It has been interpolated from borehole data using the surface interpolation tool Kriging. This head is consistent with the transmissivity raster and reflects its flow through its field. (B) Porosity is defined as the volume of void space that contributes to fluid flow divided by the entire volume. The effective porosity field, a physical property of the aquifer, is estimated from geological data. It was expressed as a value of around 35 to 41 percent of the volume of the porous medium contributing to fluid flow. (C) Transmissivity, measured in area square per day units, was interpreted from geological information. Transmissivity is the rate at which water is transmitted through a unit width of an aquifer under a unit hydraulic gradient. It is expressed as the product of the average hydraulic conductivity and thickness of the saturated portion of an aquifer. (D) The saturated thickness, measured in length units, was interpreted from geological information. The saturated thickness of an unconfined Eocene aquifer is the distance between the water table and the lower confining layer. (E) The magnitude is in units of length over time and is an optional output raster where each cell value represents the magnitude of the seepage velocity vector (average linear velocity) at the center of the cell. (F) The velocity vector’s direction is expressed in compass coordinates (degrees clockwise from north). Each cell value corresponds to the direction of the seepage velocity vector (average linear velocity) at the cell’s center. This is calculated as the average value of the seepage velocity through the four faces of the cell. (G) Darcy flow is the volume balance residual raster. Each cell value represents the groundwater volume balance residual for steady flow in an aquifer, as determined by Darcy’s Law.
Figure 24. Calculation of the groundwater volume balance residual and seepage velocity vector (direction and magnitude) for steady flow in an aquifer by GIS. (A) Head elevation: the head elevation raster comes from various sources. It has been interpolated from borehole data using the surface interpolation tool Kriging. This head is consistent with the transmissivity raster and reflects its flow through its field. (B) Porosity is defined as the volume of void space that contributes to fluid flow divided by the entire volume. The effective porosity field, a physical property of the aquifer, is estimated from geological data. It was expressed as a value of around 35 to 41 percent of the volume of the porous medium contributing to fluid flow. (C) Transmissivity, measured in area square per day units, was interpreted from geological information. Transmissivity is the rate at which water is transmitted through a unit width of an aquifer under a unit hydraulic gradient. It is expressed as the product of the average hydraulic conductivity and thickness of the saturated portion of an aquifer. (D) The saturated thickness, measured in length units, was interpreted from geological information. The saturated thickness of an unconfined Eocene aquifer is the distance between the water table and the lower confining layer. (E) The magnitude is in units of length over time and is an optional output raster where each cell value represents the magnitude of the seepage velocity vector (average linear velocity) at the center of the cell. (F) The velocity vector’s direction is expressed in compass coordinates (degrees clockwise from north). Each cell value corresponds to the direction of the seepage velocity vector (average linear velocity) at the cell’s center. This is calculated as the average value of the seepage velocity through the four faces of the cell. (G) Darcy flow is the volume balance residual raster. Each cell value represents the groundwater volume balance residual for steady flow in an aquifer, as determined by Darcy’s Law.
Hydrology 11 00156 g024aHydrology 11 00156 g024b
Table 1. Satellite and sensor data types and their resolutions [19].
Table 1. Satellite and sensor data types and their resolutions [19].
ParameterSatelliteSensorsSpatial/Temporal Resolutions and Coverage
PrecipitationCombined TRMM and GPMs With Multiple Satellite Constellation IMERGMicrowave radiometer (TMI, GMI) and RADAR (PR, DPR),
microwave imagers and sounders calibrated with GPM sensor data
0.1° × 0.1°
30 min, daily, monthly,
06/2000 to present
Soil MoistureSMAPL Band
Microwave Radiometer
9 km × 9 km and 36 km × 36 km
every 2–3 days, 3/2015 to present
NDVIBased on MODIS Terra Vegetation Indices 16-Day Global 250 m, long-term meanMODerate-resolution imaging spectroradiometer (MODIS)2350 km swath 12/1999–present;
generated from the MODIS/061/MOD13Q1
Landsat 8Operational land imager (OLI and OLI2)30 m, 185 km swath, every 16 days, 02/2013–present
Landsat 9Thermal infrared sensor (TIRS and TIRS2)30 m, 185 km swath, every 16 days, 09/2021–present
Sentinel 2A and 2BMulti-spectral imager (MSI)290 km swath; 10 m, 20 m, 60 m; every 5 days, 6/23/2015 and 3/7/2017–present
Table 2. Annual average temperature in Beirut, Lebanon.
Table 2. Annual average temperature in Beirut, Lebanon.
year19751976197719781979198019811982198319841985198619871988198919901991199219931994199519961997
degree22.520.125.122.322.622.522.420.723.223.223.423.823.223.623.223.124.122.623.224.424.423.323.5
year19981999200020012002200320042005200620072008200920102011201220132014201520162017201820192020
degree23.924.924.124.624.424.424.325.024.124.024.824.825.625.625.324.124.124.324.723.625.924.525.6
Table 3. Lithological characteristics of strata in the study area.
Table 3. Lithological characteristics of strata in the study area.
TimeAgeCodeName and Nature of RocksArea/km2Permeability % from Annual Precipitation Amount
SecondaryJurassicJ7Portlandian oolitic limestone1315–35
CretaceousbcNeocomian–Barremian basalt6.70
C1Neocomian–Barremian sandstone31.410–20
C2bAbtian dolomite16.730–35
C3Albian marly limestone and marl151–3
C4Cenomanian dolomitic limestone388.634–41
C6Senonian marl1192–3
TertiaryPaleogeneE2Upper Eocene marly and chalky limestone307.520–35
NeogeneMcgMiocene conglomeritic limestone16.49–20
Quaternary QQuaternary deposits11.65–10
Total926
Table 4. Characteristics of the Eocene Strata.
Table 4. Characteristics of the Eocene Strata.
Rock
Formation
Average Thickness (b)AreaHydraulic Properties
Porosity (Ø)Permeability (P) % from Annual PrecipitationMoisture Content (ɱ)Hydraulic Conductivity (k)Transmissivity T = k×b
(m)km2%%%m/sm2/s
Eocene30033035–41387–910−43 × 10−3
Table 5. Morphometric parameters obtained from ArcGIS and global mapper software.
Table 5. Morphometric parameters obtained from ArcGIS and global mapper software.
Basin area400 km2Total number of streams (permanent and intermittent)506Average stream length 0.77 km
Basin perimeter125.2 kmShorter stream0.02 kmH max1181 m
Main channel length32 kmLonger stream3.2 kmH min30 m
Stream axial length25 kmLength of all streams387 kmBasin length40 km
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

Farhat, N. Enhancing Drought Resilience through Groundwater Engineering by Utilizing GIS and Remote Sensing in Southern Lebanon. Hydrology 2024, 11, 156. https://doi.org/10.3390/hydrology11090156

AMA Style

Farhat N. Enhancing Drought Resilience through Groundwater Engineering by Utilizing GIS and Remote Sensing in Southern Lebanon. Hydrology. 2024; 11(9):156. https://doi.org/10.3390/hydrology11090156

Chicago/Turabian Style

Farhat, Nasser. 2024. "Enhancing Drought Resilience through Groundwater Engineering by Utilizing GIS and Remote Sensing in Southern Lebanon" Hydrology 11, no. 9: 156. https://doi.org/10.3390/hydrology11090156

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