Infiltration is crucial in the hydrological cycle, serving as the primary process that increases soil moisture. This study investigates soil infiltration rate (IR) prediction using various techniques, including GMDH, Gaussian Process, SVM, ANN, and MARS. 190 field observations were collected from Alashtar sub-watersheds in Lorestan, Iran. 70% of the observations were used for model preparation, while 30% were used for validation. The input variables for the study are Time, Sand, Clay, Silt, pH, Electrical Conductivity, Moisture Content, Soil Bulk Density, Porosity, Calcium Carbonate, Phosphorus, Organic Carbon, Organic Matter, Nitrogen, and Temperature, while IR is the output variable. Obtained results indicate that the ANN has a higher accuracy with coefficient of correlation values as 0.9366, 0.8624, mean absolute error values as 0.0607, 0.1000, Nash Sutcliffe model efficiency values as 0.8732, 0.7350, scattering index values as 0.3108, 0.5003, and Legates and McCabe's Index values as 0.6585, 0.5654 by using training and testing data sets, respectively. A sensitivity analysis highlighted that time is the parameter that most influences estimating the IR. The study underscores the precision of ANN in predicting soil infiltration rates and the need for AI-based models in hydrological models to improve accuracy and reliability in IR prediction.

  • The infiltration rate (IR) of the soil is estimated using the group method of data handling, Gaussian process, support vector machine, artificial neural networks (ANNs), and multivariate adaptive regression splines.

  • Inter-comparison of different AI-based models revealed that ANN is the most efficient model.

  • Sensitivity analysis suggested that time is the most influential parameter on estimating the IR.

Infiltration is crucial in the hydrological cycle, serving as the primary process that increases soil moisture. It occurs when water enters and moves within the soil's layers. The difference in energy levels propels the movement of water through this porous material. This process's key factors include gravitational pull, capillary action, and water's tendency to adhere to surfaces. The infiltration rate (IR), a vital concept in this context, measures the speed at which water penetrates the soil from the surface, reflecting the soil's ability to take in water over a given time frame. Infiltration, a key element impacting soil's water retention capacity, is crucial for various professionals, including hydrologists, irrigation and agricultural engineers, and soil scientists. It aids in determining multiple aspects, such as soil moisture levels, runoff dynamics, and the movement of sediments and solutes.

IR is a crucial process in the hydrological cycle, regulating water movement and impacting soil moisture, plant growth, and groundwater recharge. It regulates surface runoff, erosion, and soil moisture content, ensuring precipitation is not lost to runoff (Seiler & Gat 2007). It also affects soil properties and water retention, with balanced soils maintaining water movement and retention. It also influences evapotranspiration and climate feedback, with regions with adequate infiltration supporting a balanced cycle (Farmer et al. 2003). Infiltration also has environmental implications, regulating ecosystem health and supporting biodiversity. Proper land management practices, such as no-till farming and cover cropping, are essential for maintaining soil health and moisture retention (Reicosky 2020).

IR is essential for estimating artificial groundwater recharge, designing irrigation and drainage systems, and water balance models (Parhi et al. 2007; Ma & Shao 2008; Bayabil et al. 2019). Accurate measurement of the IR is vital for developing effective hydrological models (Shirmohammadi & Skaggs 1984). However, this quantification faces challenges due to the time and space variability of soil hydraulic properties. Such variability often stems from changes in land use and soil characteristics (Muñoz et al. 2017). Research indicates that land-use patterns significantly impact infiltration, influenced by factors such as soil management strategies, tillage practices, and vegetation types (Brown et al. 2005; García-Ruiz et al. 2008; Wang et al. 2016). Additionally, soil properties such as texture, structure, porosity, hydraulic conductivity, existing moisture conditions, suction head, temperature, humidity, rainfall intensity, and water quality also play a significant role in determining IR (Liu et al. 2011). Various physical soil properties influence the infiltration characteristics. Among these are soil texture, moisture content, and density, which significantly affect infiltration (Angelaki et al. 2013). Soil texture plays a pivotal role in influencing infiltration. The soil's ability to hold water, crucial for water accessibility, is determined by its texture and structure (Al-Azawi 1985). IRs are typically higher in unsaturated soils and decrease over time, stabilizing constantly. Infiltration characteristics show significant variation due to differences in soil texture, type, and various soil conditions. Conducting experimental infiltration measurements is complex and challenging, often described as labor-intensive, tedious, and time-consuming (Vand et al. 2018). The assessment of the infiltration process is complicated by spatial and temporal variations, making it a complex field of study (Pandey & Pandey 2018). Furthermore, because of their significant dependance on soil physical properties, variations in IR can be attributed to the various parameters of infiltration models. The accuracy of IR determination can be enhanced by quantifying the spatial variability of soil properties.

Numerous studies have suggested using conventional infiltration models as alternatives to experimental observations (Mishra et al. 2003; Singh et al. 2018). However, employing any specific model requires a thorough understanding of its boundary conditions and assumptions. Soil water researchers have introduced various models like Kostiakov, Horton, Philip, Holton, Green-Ampt, and Modified Kostiakov for estimating infiltration (Richards 1931; Philips 1957; Mishra et al. 2003; Sihag et al. 2017). Mishra et al. (2003) categorized these models into physical, semi-empirical, and empirical. Most of these models are based on assumptions such as homogeneous water absorption, constant pounding head, and steady IR, which are rarely observed in actual field conditions, potentially leading to inaccurate predictions.

Quantification of IR is a complex phenomenon due to the variability in soil hydraulic properties. These properties vary significantly over time and space, leading to spatial heterogeneity and temporal variability. Soil texture, land use, and microtopography also contribute to these variations. Changes influence temporal variability in soil moisture, precipitation events, temperature, and biological activity. Surface sealing and crusting can also affect IRs. Measurement limitations include small-scale measurements, disturbance of soil structure, and temporal inconsistency. AI-based models face limitations in addressing these interactions. Data requirement, overfitting, and parameter sensitivity are also challenges. Vegetation and land cover also play a role in influencing IRs, but their type and density can vary widely over time and space. Addressing these challenges requires robust data collection methods and advanced modeling techniques, like AI-based models.

The use of AI-based models in civil engineering and water resources engineering problems is increasing day by day. Various researchers used AI-based models to solve their complex problems (Singh et al. 2017; Arora et al. 2019; Sihag et al. 2019; Pandhiani et al. 2020; Singh 2020; Aradhana et al. 2021; Bhoria et al. 2021; Sihag et al. 2021; Sepahvand et al. 2021a, b; Singh et al. 2021a, b; Singh et al. 2022; 2023; Nivesh et al. 2022; Sihag et al. 2022; Arora et al. 2024; Singh & Minocha 2024a, b, c). Some researchers have also employed AI-based models to estimate the infiltration process, focusing on soil properties. These AI-based models have shown high precision in infiltration prediction, as evidenced in studies by Singh et al. (2017), Sihag et al. (2017), which demonstrate that soil physical properties and elapsed time can be effectively used to estimate the infiltration process with greater accuracy. Therefore, in this study, AI-based models are employed to enhance the precision and reliability of the models. These include the group method of data handling (GMDH), support vector machine (SVM), Gaussian process (GP), multivariate adaptive regression splines (MARS), and artificial neural network (ANN). The study aims to develop empirical models for accurately estimating the IR, thereby contributing to enhanced water management and environmental modeling. It includes providing insights and tools that can be used for effective water resource management and addressing ecological challenges in the Alashtar sub-watersheds and similar regions.

Research significance

The study conducted in the Alashtar sub-watersheds in Lorestan, Iran, uniquely contributes to hydrological science and environmental management by addressing critical challenges of climate change and population growth. It is a significant finding that the ANN model excels in predicting IRs in a region with a complex geological and varied climatic profile is of paramount importance. This model's superiority, proven through advanced statistical evaluations, establishes ANN as a vital tool for precise hydrological analysis in similar environments. The practical implications of this research are extensive, benefiting sectors like agriculture, water resource management, and environmental conservation, where accurate infiltration data is crucial. Methodologically, the study sets a new benchmark in hydrological research by employing advanced statistical tools and graphical models for model comparison, enhancing the clarity and reliability of findings and serving as a guide for future research. Additionally, the sensitivity analysis provides insights into the influence of time on IRs and offers valuable information for optimizing resource use in environmental planning. Overall, this research marks a significant stride in advancing hydrological science, offering a more accurate and reliable methodology for managing ecological challenges in regions undergoing similar changes.

In the study at the Alashtar sub-watersheds, Lorestan, Iran, the methodology included collecting 190 field observations to develop empirical models to estimate the soil IR. It involved integrating soil characteristics such as sand, clay, and pH. cutting-edge AI-based models, such as GP, SVM, GMDH, MARS, and ANN, have been utilized to estimate parameters accurately. The models' effectiveness was assessed using metrics such as correlation coefficient (CC) and mean absolute error, offering insights for hydrological science and environmental management in climate change and population growth.

Gaussian process

GP models represent a class of stochastic models where groups of random variables, organized by time or space, each have a non-linear model. Unique to this approach is that any finite linear combination of these variables follows the same distribution. Rooted in the Gaussian distribution, this concept is named after Carl Friedrich Gauss, who is synonymous with the normal distribution. GPs can be seen as a form of multivariate normal distribution but with an infinite-dimensional space. GPs are employed as a probabilistic, non-parametric supervised learning methodology in AI-based models. Their primary role is to generalize complex and non-linear function mappings found within datasets, a feature that has captured the interest of researchers in diverse fields. GPs excel in managing non-linear data, a capability attributed to their use of kernel functions. One of the most notable benefits of Gaussian processes, as outlined by Omran et al. (2016), is their ability to yield consistent responses to varying input data. Gaussian functions use kernel machines to analyze and interpret models in practical applications. This process illustrates a hands-on approach to understanding the functionality of kernel machines. A critical step in the GP regressor fitting process involves optimizing the kernel's hyperparameters. This optimization is accomplished by increasing the log-marginal-likelihood, which varies depending on the optimizer (Williams & Rasmussen 2006). This process typically begins with a designated training set in probabilistic regression, setting the stage for the model's learning and adaptation.
(1)
The input dataset X, represented as XRDxn, is known as the design matrix, while yRn stands for the vector of targeted outputs. A fundamental premise of Gaussian process regression is that the output y is derived as follows (Hoang et al. 2016).
(2)
In this context, εN (0, σ²n) ∈ R symbolizes a consistent noise distribution across all samples x. GP regression is a Bayesian approach that assumes a GP before accomplished functions, implying that function values exhibit a predictable pattern.
(3)
where f = [y1, y2, …, yn] > represent a vector of latent function values, where fi = f(yi), and K denotes a covariance matrix with elements determined by the covariance function, Kij = k (yi, yj). In GP modeling, these latent function values (fi) are treated as random variables associated with their respective inputs.

Support vector machine

SVM models are supervised learning algorithms used for classification and regression. This method was established in the statistical learning framework (Cortes & Vapnik 1995). SVM employs a technique known as the kernel trick to project input data into higher-dimensional feature spaces, facilitating indirect manipulation of inputs. This approach is known for achieving enhanced generalization capabilities by effectively separating data using a hyperplane. This hyperplane was chosen to maximize its distance from the nearest training data points, thereby minimizing generalization error. A key advantage of SVMs, as highlighted by Park et al. (2019), lies in their distinctive approach to optimization. These models are also recognized for their effective handling of high-dimensional spaces and their foundation in the principles of computational learning theory. In practical applications, an SVM learning algorithm builds a model that sorts training examples into one of two distinct categories, serving as a non-probabilistic binary linear classifier. Methods like Platt scaling have been introduced to adapt SVMs for probabilistic classification scenarios. The operational mechanism of an SVM involves projecting training examples into a dimensional space to maximize the distance between the two predefined categories. Subsequently, new examples are added to this space, and a type is assigned based on their position relative to the established division. In terms of execution, the SVM analysis process involves utilizing training and testing datasets that correlate with designated input and output variables. Two main strategies exist for executing SVM analysis. The first involves employing an optimal margin classifier, which functions as a linear classifier to establish a decision boundary. The second strategy uses a kernel function that facilitates calculating the dot products between two vectors in a feature space of n-dimensions. When data are processed through a non-linear kernel function in this expanded dimensional space, it achieves a linear separation without modifying the original input space. This approach, detailed by Goh & Goh (2007), allows for efficient classification in complex data scenarios. When provided with a training dataset, a linear SVM will identify n points characterized by the following formula.
(4)
In this scenario, y1 takes the value of either 1 or −1, contingent on the class to which the point xi belongs. Each xi represents a vector in p-dimensional real space. The objective is to determine the ‘maximum-margin hyperplane,’ which separates the set of points xi for which yi = 1 from those where yi = −1. This separation is defined to maximize the distance between the hyperplane and the closest point xi from either set. The representation of any hyperplane can be articulated through the set of points x that fulfill the equation (WTXb = 0).
(5)
where W acts as the average vector to the hyperplane; interestingly, it does not require normalization. This concept is somewhat like Hesse's standard form but with a notable distinction: W is not necessarily a unit vector.

Group method of data handling

In the mid-1960s, Ivakhnenko, a notable Russian mathematician and cyberneticist, introduced a groundbreaking method for modeling complex systems without the necessity of understanding their internal mechanisms. This method, recognized as the GMDH, described by Ziari et al. (2016), focuses on creating self-organizing models, specifically high-order polynomials based on input variables. This model has been widely applied in various domains, including prediction, classification, and control composition. The GMDH algorithm stands out due to its inductive strategy for modeling multi-parametric datasets mathematically. Its key characteristic is the complete automation of structural and parametric optimization of models. This aspect makes GMDH particularly valuable in data mining, knowledge discovery, modeling of complex systems, optimization, and pattern recognition.

A key feature of GMDH algorithms is their systematic evaluation of progressively intricate polynomial models in an inductive manner. The selection of the best model is based on an external criterion. Moreover, using the Volterra functional series, these algorithms can approximate the relationship between input and output variables. As noted by Anastasakis & Mort (2001), the discrete equivalent of this series is the Kolmogorov–Gabor polynomial. This attribute enhances the GMDH's adaptability for modeling intricate relationships across diverse scientific and engineering disciplines.

A GMDH model, featuring multiple inputs and a single output, represents a select set of elements from the foundational function.
(6)
In a GMDH model, ki and ai represent coefficients. At the same time, m signifies the count of base function components, and ki and zi are elementary functions with distinct sets of input variables. To derive the most effective solution, GMDH methods assess various component subsets from the base function called partial models. The coefficients of these models are calculated using the least-squares method. In the GMDH framework, the quantity of components in a partial model continuously increases until the minimal value of an external criterion suggests an optimally complex model structure, a process termed model self-organization. The initial foundational function in GMDH was the incrementally complex Kolmogorov–Gabor polynomial (2).
(7)

Within GMDH methodologies, simple partial models are standard, typically encompassing functions up to the second degree. Such inductive approaches are frequently called polynomial neural networks (Madala 2019). The development of GMDH revealed a notable correlation between the challenges of modeling noisy data and the concept of signal transmission through a noisy channel. This understanding paved the way for the establishment of a theory of noise-immune modeling.

A fundamental concept of this theory is that the intricacy of an optimal predictive model should be in direct proportion to the amount of uncertainty inherent in the data. Essentially, this means that more significant uncertainty in the data, often caused by noise, necessitates a more streamlined optimization technique characterized by fewer estimated parameters. Consequently, GMDH theory evolved as an inductive approach that automatically adjusts the model's complexity to suit the noise variance in fuzzy data. This adaptive feature of GMDH, aligning model complexity with data uncertainty, led to its recognition as one of the pioneering information technologies for extracting knowledge from experimental data. This conceptualization and its practical application were notably discussed by Ivakhnenko & Stepashko (1985), the importance of GMDH's significance in data analysis and modeling.

Multivariate adaptive regression splines

MARS, a non-parametric regression method developed by Friedman (1991), expands upon conventional linear models by automatically integrating non-linear impacts and interactions among variables. This method is particularly valued for its ability to articulate complex non-linear relationships between predictor variables and the response variable. A notable feature of the MARS model is its utilization of both forward and backward stepwise procedures. As de Andrés et al. (2011b) outlined, the forward stepwise approach in MARS is akin to selecting an appropriate set of input variables. This step incrementally builds the model by adding variables and their interactions that significantly improve performance.

Conversely, MARS employs the backward stepwise method to refine the model further. As described by Sharda et al. (2006), this involves eliminating extra variables from the set initially chosen in the forward step, thereby enhancing the model's predictive accuracy. MARS accomplishes its objective by transforming a variable X into a new variable Y using two basis functions (BFs). These functions, or the variable's value that determines a turning point, are utilized over the entire span of input values. This process allows for a finer and more precise depiction of the data.
(8)
(9)
The threshold value is denoted by the letter ‘a’. To maintain uniformity in the BFs, MARS ensures that two neighboring splines intersect at the knot. This model finds diverse applications across various fields, such as managing financial systems, analyzing time-series data, and forecasting pesticide levels. However, it is surprising that it has not yet been used to predict river quality (Kisi 2015). The spline function comprises truncated functions positioned to the left in Equation (10) and to the right in Equation (11), differentiated by a knot location, as illustrated in the following.
(10)
(11)

In the MARS algorithm framework, v denotes the knot's location, while bn (u, v) and bn + (u, v) refer to specific spline functions. The algorithm operates in three distinct phases: first, a forward stepwise approach is employed to select spline BFs; second, a backward stepwise process is used to remove BFs until an optimal set is identified iteratively; and third, a smoothing technique is applied to enhance the consistency of the final MARS model approximation. A generalized cross-validation (GVC) method is utilized to prioritize BFs for elimination based on their minimal contribution, as detailed by Dutta et al. (2018). While this model effectively fits estimation data, its ability to predict new instances is limited. To refine its predictive capabilities, surplus BFs are methodically discarded using a backward stepwise approach. The inclusion of BF in the model is guided by GVC. This GVC value is calculated by amplifying the mean squared residual error with a penalty that escalates with the model's complexity (De Andrés et al. 2011a).

Artificial neural networks

ANNs are sophisticated computational systems inspired by biological neural networks. Their ability to use a database of inputs to approximate unknown functions is a crucial characteristic. One significant aspect of ANNs is their capacity to address complex and non-linear problems through basic mathematical operations (Thakur et al. 2021; Upadhya et al. 2022). The operating principle of ANNs is comparable to the functioning of biological neurons in the brain, where the network of cells can learn and recognize patterns, such as faces and sounds, based on past experiences. In ANN terminology, the process of determining model parameters is known as training. An ANN is structured into three layers: the input, hidden, and output layers. Each layer can contain multiple units, fully interconnected with the subsequent layer. Each connection in the network possesses an adjustable weight (Silverman & Dracup 2000). The total number of input parameters equals the nodes in the initial layer, as per Min Equation (12). As the inputs transition to the subsequent layer, they are multiplied by the weights of the connections. Each node (I) in a layer receives inputs from every node j in the preceding layer, with each incoming signal (qi) being assigned a weight (wki). The effective signal (Pj) to node j is the cumulative sum of all weighted incoming signals. Initially, these weights (wji) are set randomly and are fine-tuned during the early stages of training:
(12)
In the later stages of neural networks, the input signal (Pj) is processed through a transfer function (as shown in Equation (12)) to produce the output signal (yi) from node k. Transfer functions can be non-sigmoidal, such as logistic and hyperbolic tangent functions, or hard limit functions that are restricted to the bounds of 0 or 1, such as linear, polynomial, rational functions (ratios of polynomials), and Fourier series (sums of cosine functions). The latter is beneficial for extrapolating beyond the training data. In ANN architecture, sigmoidal transfer functions are primarily used in hidden layers, while linear transfer functions (where yi equals Pk) are typically used in the output layer. According to Zealand et al. (1999), these are among the most used transfer functions in neural network models:
(13)
Backpropagation is a technique used to calculate the gradient in non-linear multilayer networks. It encompasses a series of learning rules that clarify the process of propagating errors backward. This method involves computing the error function concerning the network's weights, refining the model's accuracy. As detailed by Yang et al. (2019), the sigmoid transfer function is employed to compress input values varying from 0 to infinity. The fundamental architecture of the ANN model is depicted in Figure 1.
Figure 1

ANN model structure.

Figure 1

ANN model structure.

Close modal
A double-ring infiltrometer was used to measure the soil's IR. It consists of two rings, i.e., the outer and inner rings. These rings were driven into the soil to create a controlled environment for water infiltration. A constant water supply ensured that both rings were filled to a certain level. The process involves initial saturation to simulate natural field conditions, filling the rings simultaneously, measuring drops in water, and continuously adding water until the IR stabilizes. Measurements were taken at regular intervals until a steady IR was achieved. Field observations were conducted at multiple locations within the Alashtar sub-watersheds in Lorestan, Iran. One hundred ninety observations were measured and used to predict the soil IR. The data were randomly divided into two segments: 70% (133 observations) for training and 30% (57 observations) for testing the predictive models. The study utilized various ML techniques for prediction, including GP, SVM, GMDH, MARS, and ANN. The dataset encompassed 15 input variables, including time, sand, clay, silt, PH, electrical conductivity (EC), moisture content, soil bulk density, porosity, calcium carbonate (CaCO3), phosphorus (P), organic carbon (OC), organic matter (OM), nitrogen (N), and temperature. These variables were critical in determining the IR, the primary output variable of the study. A variety of statistical tools were utilized for precise prediction of the IR, including the CC, mean absolute error (MAE), root mean squared error (RMSE), Nash–Sutcliffe efficiency (NSE), the scattering index (SI), and the Legates and McCabe's index (LMI). A flow chart outlining the study is depicted in Figure 2. Table 1 of the research document elaborates on the specific statistical attributes of each input variable.
Table 1

Statistical properties of input parameters

ParametersUnitsTraining dataset
Testing dataset
MeanStandard deviationMinMaxMeanStandard DeviationMinMax
Time Min 26.40 19.26 2.5 60 25.26 20.38 2.5 70 
Sand 48.03 12.81 26 69 46.27 12.11 26 68.15 
Clay 13.43 7.92 6.7 35.55 13.66 8.52 6.7 35.55 
Silt 38.48 8.87 24 54.33 40.00 7.70 25.15 54.33 
PH – 7.97 0.11 7.74 8.14 7.98 0.09 7.74 8.14 
EC – 305.40 74.92 176.5 501 311 75.60 176.5 501 
Moisture content 2.18 1.02 1.21 5.6 2.08 0.97 1.21 5.6 
Soil bulk density g/cm3 1.48 0.21 1.1 1.87 1.51 0.18 1.1 1.87 
Porosity 53.56 3.65 48 62 53.78 4.11 48 62 
CaCO3 25.10 10.24 48.25 23.47 11.28 48.25 
mg/kg 1.10 0.62 0.55 3.06 1.07 0.54 0.55 3.06 
OC 1.09 0.65 0.11 2.76 0.98 0.58 0.12 2.10 
OM 1.88 1.12 0.20 4.77 1.69 1.00 0.20 3.63 
0.11 0.06 0.01 0.25 0.10 0.06 0.01 0.24 
Temperature °C 31.48 0.50 31 32 31.43 0.50 31 32 
IR cm/min 0.28 0.24 0.01 1.44 0.30 0.29 0.01 1.28 
ParametersUnitsTraining dataset
Testing dataset
MeanStandard deviationMinMaxMeanStandard DeviationMinMax
Time Min 26.40 19.26 2.5 60 25.26 20.38 2.5 70 
Sand 48.03 12.81 26 69 46.27 12.11 26 68.15 
Clay 13.43 7.92 6.7 35.55 13.66 8.52 6.7 35.55 
Silt 38.48 8.87 24 54.33 40.00 7.70 25.15 54.33 
PH – 7.97 0.11 7.74 8.14 7.98 0.09 7.74 8.14 
EC – 305.40 74.92 176.5 501 311 75.60 176.5 501 
Moisture content 2.18 1.02 1.21 5.6 2.08 0.97 1.21 5.6 
Soil bulk density g/cm3 1.48 0.21 1.1 1.87 1.51 0.18 1.1 1.87 
Porosity 53.56 3.65 48 62 53.78 4.11 48 62 
CaCO3 25.10 10.24 48.25 23.47 11.28 48.25 
mg/kg 1.10 0.62 0.55 3.06 1.07 0.54 0.55 3.06 
OC 1.09 0.65 0.11 2.76 0.98 0.58 0.12 2.10 
OM 1.88 1.12 0.20 4.77 1.69 1.00 0.20 3.63 
0.11 0.06 0.01 0.25 0.10 0.06 0.01 0.24 
Temperature °C 31.48 0.50 31 32 31.43 0.50 31 32 
IR cm/min 0.28 0.24 0.01 1.44 0.30 0.29 0.01 1.28 
Figure 2

Flow chart of the study.

Figure 2

Flow chart of the study.

Close modal

Study area

The study is on the Alashtar sub-watersheds, part of the Lorestan province, in southwestern Iran. This area is geographically positioned between 33°45′ 17″ N and 33°51′ 23″ N latitude and 48°10′ 28″ E and 48°23′ 29″ E longitude. Covering approximately 800 km², the sub-watersheds' boundaries are demarcated based on the locations of various sampling sites. The elevation within this study region ranges significantly from 1,481 to 3,613 m above sea level. Characterized by a cold, semiarid climate, the area receives an average annual rainfall of about 570 mm. The surface geology of the Alashtar watersheds is quite diverse, predominantly covered by formations from different geological periods. A significant portion, approximately 35.29%, comprises quaternary period formations. It is followed by 28.95% of the area being covered by Lower Cretaceous and Upper Jurassic period formations and 7.37% by Eocene period formations. The study meticulously outlines the soil texture and GIS coordinate system for all locations, as shown in Figure 3.
Figure 3

Map of the study area showing a geological map of sampling sites.

Figure 3

Map of the study area showing a geological map of sampling sites.

Close modal

Model assessment

The study focused on refining the precision and dependability of a predictive model using statistical metrics. The performance assessment of the model was conducted using vital metrics, including the CC, MAE, RMSE, Nash–Sutcliffe efficiency (NSE), SI, and LMI (Saqr et al. 2022; Abd-Elmaboud et al. 2024). These metrics provided critical insights into the model's accuracy, reliability, and overall suitability.
(14)
(15)
(16)
(17)
(18)
(19)
where li and mi are individual sample points. and are the mean values of the sample series. n represents the number of observations.

Evaluation of the GP model

The GP, in regression analysis, employs a combination of a radial basis function kernel (RBF kernel) and a Pearson kernel function (PUK), along with customizable parameters such as ‘L,’ gamma (Ɣ), omega (O), and sigma (S). Numerous trials were performed to determine their ideal values, which align with the highest CC and lowest error rates. An evaluation of the model's performance, as detailed in Table 2, contrasted the outcomes of GP–PUK and GP–RBF during both the training and testing phases. This comparison revealed that the GP–RBF model surpassed the GP–PUK model in predicting the IR. The superiority of GP–RBF was evident in its higher CC value of 0.9631 compared to 0.8273 for GP–PUK, lower MAE at 0.0357 alongside 0.1109, reduced RMSE value of 0.1174 versus 0.1789, higher Nash–Sutcliffe efficiency at 0.9274 compared to 0.6240, smaller SI at 0.2352 against 0.5959, and better the LMI value of 0.7991 as opposed to 0.5180 in both training and testing stages. Figure 4(a) and 4(b) illustrates a scatter plot comparing the observed and predicted values using the GP–PUK and GP–RBF models. The clustering of most data points around the line of perfect agreement in this figure indicated a closer alignment between observed and predicted values, signifying a higher level of consistency in the results.
Table 2

Performance of GP–PUK, GP–RBF, SVM–PUK, and SVM–RBF model

ModelsCCMAERMSENSESILMI
 Training stage 
GP–PUK 0.9781 0.0189 0.0517 0.9567 0.1817 0.8938 
GP–RBF 0.9631 0.0357 0.0669 0.9274 0.2352 0.7991 
SVM–PUK 0.8846 0.0454 0.1174 0.7763 0.4128 0.7444 
SVM–RBF 0.8027 0.0597 0.1495 0.6368 0.5260 0.6638 
 Testing stage 
GP–PUK 0.7620 0.1531 0.2130 0.4671 0.7094 0.3346 
GP–RBF 0.8273 0.1109 0.1789 0.6240 0.5959 0.5180 
SVM–PUK 0.7604 0.1257 0.1986 0.5366 0.6615 0.4537 
SVM–RBF 0.7851 0.1149 0.1838 0.6029 0.6124 0.5007 
ModelsCCMAERMSENSESILMI
 Training stage 
GP–PUK 0.9781 0.0189 0.0517 0.9567 0.1817 0.8938 
GP–RBF 0.9631 0.0357 0.0669 0.9274 0.2352 0.7991 
SVM–PUK 0.8846 0.0454 0.1174 0.7763 0.4128 0.7444 
SVM–RBF 0.8027 0.0597 0.1495 0.6368 0.5260 0.6638 
 Testing stage 
GP–PUK 0.7620 0.1531 0.2130 0.4671 0.7094 0.3346 
GP–RBF 0.8273 0.1109 0.1789 0.6240 0.5959 0.5180 
SVM–PUK 0.7604 0.1257 0.1986 0.5366 0.6615 0.4537 
SVM–RBF 0.7851 0.1149 0.1838 0.6029 0.6124 0.5007 
Figure 4

Scatter graph between observed and predicted values using (a), (b) GP–PUK, GP–RBF (c), (d) SVM–PUK, SVM–RBF (e), (f) GMDH (g), (h) MARS and (i), (j) the ANN model for training, and testing stages.

Figure 4

Scatter graph between observed and predicted values using (a), (b) GP–PUK, GP–RBF (c), (d) SVM–PUK, SVM–RBF (e), (f) GMDH (g), (h) MARS and (i), (j) the ANN model for training, and testing stages.

Close modal

Evaluation of SVM

SVM, in its regression application, integrates an RBF kernel and a PUK alongside adjustable parameters such as ‘L,’ gamma (Ɣ), precision to a certain number of decimal places, and O and S. A comparative analysis of the models' performances, encapsulated in Table 2, examined SVM–PUK and SVM–RBF across training and testing phases. This comparison from Table 2 highlighted that the SVM–RBF model exhibited a more accurate prediction of the IR than the SVM–PUK model. These metrics emphasized that the SVM–RBF model outperformed the SVM–PUK model in several ways. The SVM–RBF model has a higher CC value of 0.8027 compared to 0.7851 for the SVM–PUK model, a lower MAE value of 0.0597 compared to 0.1149, a reduced RMSE of 0.1174 compared to 0.1838, and a higher NSE of 0.6368 compared to 0.6029. Additionally, it has a smaller SI value of 0.5260 compared to 0.6124 and a better LMI value of 0.6638 compared to 0.5007 for both the training and testing stages. Figure 4(c) and 4(d) presents a scatter plot that maps the observed versus predicted values using the SVM–PUK and SVM–RBF models. The concentration of data points around the line of perfect agreement in this figure suggested greater accuracy and reliability in the predictions, indicating a solid alignment between observed and predicted values.

Evaluation of GMDH

The GMDH employed inductive methods for the computer-assisted mathematical modeling of datasets with multiple parameters. This approach enables the complete automation of models' structural and parametric optimization. Table 3 in the study presented the coefficients and constants used in the function transfer of the GMDH-based model. Additionally, the table included performance evaluation metrics for the model's IR predictions across the training and testing phases. The results in Table 5 demonstrated the efficacy of the GMDH model in predicting the IR. The model showed CC values of 0.8027 and 0.7702, MAE values of 0.2170 and 0.1357, RMSE values of 0.1459 and 0.1861, NSE values of 0.6543 and 0.5930, SI values of 0.5131 and 0.6200, and LMI scores of 0.3884 and 0.4100 for the training and testing stages, respectively. Figure 4(e) and 4(f) illustrated this relationship through a scatter plot, comparing observed values with those predicted by the GMDH-based models for the IR. The scatter plot's alignment with the line of perfect agreement indicated a strong correlation between the observed and predicted values, suggesting a high level of accuracy in the model's predictions for the IR.
Table 3

Results of the GMDH transfer function for parameter adjustment

LayersNeuronsa0a1a2a3a4a5
−7.62902 −0.00048 0.272431 0.000206 −0.00223 −0.00032 
 0.908387 −0.0218 −0.03955 0.000247 0.000575 0.000196 
 0.011921 −0.01107 0.007782 0.000239 6.09E-05 −0.00017 
 7.447267 −0.28382 −0.89401 0.002707 −0.03294 0.020808 
 7.413539 −0.28259 −1.53771 0.002696 −0.09802 0.035807 
0.193914 −0.91559 −0.60218 2.677236 2.198473 0.473449 
 0.192743 −0.91687 −0.59041 2.679242 2.177974 0.473984 
 0.09301 −0.20286 −0.20651 1.073433 0.893554 1.298476 
 0.09191 −0.20317 −0.19571 1.074827 0.87471 1.296485 
 0.145266 −0.17225 −0.0065 −4.77314 −5.36011 12.38166 
0.048078 −0.38963 0.92695 2.258161 0.001442 −1.64676 
LayersNeuronsa0a1a2a3a4a5
−7.62902 −0.00048 0.272431 0.000206 −0.00223 −0.00032 
 0.908387 −0.0218 −0.03955 0.000247 0.000575 0.000196 
 0.011921 −0.01107 0.007782 0.000239 6.09E-05 −0.00017 
 7.447267 −0.28382 −0.89401 0.002707 −0.03294 0.020808 
 7.413539 −0.28259 −1.53771 0.002696 −0.09802 0.035807 
0.193914 −0.91559 −0.60218 2.677236 2.198473 0.473449 
 0.192743 −0.91687 −0.59041 2.679242 2.177974 0.473984 
 0.09301 −0.20286 −0.20651 1.073433 0.893554 1.298476 
 0.09191 −0.20317 −0.19571 1.074827 0.87471 1.296485 
 0.145266 −0.17225 −0.0065 −4.77314 −5.36011 12.38166 
0.048078 −0.38963 0.92695 2.258161 0.001442 −1.64676 

Evaluation of MARS

The model in question was formulated using the MARS approach to create an adaptive regression splines model. This process involved both forward and backward phases and integrating a set of user-defined parameters. These parameters included the count of BFs in the final model (14 and 6), the number of partitioning folds (10), the total best number of parameters (33.5 and 13.5), and the maximum degree of interactions within the final model (2). The model achieved a mean squared error of 0.0205 and a GCV value of 0.0253. For calculating the IR using the MARS model on the testing dataset, the formulae are defined in Equation (20). Additionally, Table 4 in the study outlines the primary BFs employed by the MARS methodology in predicting the IR, highlighting the model's intricate structure and predictive capabilities.
(20)
Table 4

The basic function of the MARS model is to predict the IR

Basic function using MARS modelIR
BF-1 max (0, Time - 5) 
BF-2 max (0, 5 - Time) 
BF-3 max (0, 58 - Porosity) 
BF-4 max (0, 0.093 - N) 
BF-5 = BF-2* max (0, 55 - Sand) 
Basic function using MARS modelIR
BF-1 max (0, Time - 5) 
BF-2 max (0, 5 - Time) 
BF-3 max (0, 58 - Porosity) 
BF-4 max (0, 0.093 - N) 
BF-5 = BF-2* max (0, 55 - Sand) 
Table 5

Performance evaluation parameter of all applied models

ModelsCCMAERMSENSESILMI
 Training stage 
GP–RBF 0.9631 0.0357 0.0669 0.9274 0.2352 0.7991 
SVM–RBF 0.8027 0.0597 0.1495 0.6368 0.5260 0.6638 
MARS 0.8171 0.1061 0.1430 0.6677 0.5031 0.4027 
GMDH 0.8096 0.2170 0.1459 0.6543 0.5131 0.3884 
ANN 0.9366 0.0607 0.0884 0.8732 0.3108 0.6585 
 Testing stage 
GP–RBF 0.8273 0.1109 0.1789 0.6240 0.5959 0.5180 
SVM–RBF 0.7851 0.1149 0.1838 0.6029 0.6124 0.5007 
MARS 0.7445 0.1440 0.1978 0.5404 0.6588 0.3738 
GMDH 0.7702 0.1357 0.1861 0.5930 0.6200 0.4100 
ANN 0.8624 0.1000 0.1502 0.7350 0.5003 0.5654 
ModelsCCMAERMSENSESILMI
 Training stage 
GP–RBF 0.9631 0.0357 0.0669 0.9274 0.2352 0.7991 
SVM–RBF 0.8027 0.0597 0.1495 0.6368 0.5260 0.6638 
MARS 0.8171 0.1061 0.1430 0.6677 0.5031 0.4027 
GMDH 0.8096 0.2170 0.1459 0.6543 0.5131 0.3884 
ANN 0.9366 0.0607 0.0884 0.8732 0.3108 0.6585 
 Testing stage 
GP–RBF 0.8273 0.1109 0.1789 0.6240 0.5959 0.5180 
SVM–RBF 0.7851 0.1149 0.1838 0.6029 0.6124 0.5007 
MARS 0.7445 0.1440 0.1978 0.5404 0.6588 0.3738 
GMDH 0.7702 0.1357 0.1861 0.5930 0.6200 0.4100 
ANN 0.8624 0.1000 0.1502 0.7350 0.5003 0.5654 

According to the data presented in Table 5, the MARS model demonstrated reliable performance in predicting IR. The model's efficacy is evidenced by its statistical metrics: it achieved CC values of 0.8171 and 0.7445, MAE values of 0.1061 and 0.1440, RMSE values of 0.1430 and 0.1978, NSE values of 0.6677 and 0.5404, the SI values of 0.5031 and 0.6588, and the LMI values of 0.4027 and 0.3738 for the training and testing phases, respectively. The scatter plot in Figure 4(g), (h) visually represents the correlation between the observed and the MARS model-predicted values for the IR. The alignment of the data points with the line of perfect agreement in this graph suggested a strong concordance between the observed and predicted values, indicating the model's potential for accurate and reliable predictions in estimating the IR.

Evaluation of ANN

A multilayer perceptron framework involving an iterative process created an ANN-based model. Numerous trials were conducted with various combinations of user-defined parameters to achieve the best configuration, characterized by the highest CC value and the lowest error margins for training and testing datasets in the context of model prediction assessment. The combinations of user-defined parameters, i.e., learning rate, momentum, hidden layers, neurons per hidden layers, and iterations, achieving the best combination were 0.2, 0.1, 1, 20, and 1,000, respectively. The performance evaluation metrics for the best trials are detailed in Table 5. The ANN-based model demonstrated superior performance in predicting the IR compared to all other applied models. This superiority is reflected in its statistical metrics: for the training and testing stages, respectively, the model registered a CC values of 0.9366 and 0.8624, MAE values of 0.0607 and 0.1000, RMSE values of 0.0884 and 0.1502, Nash–Sutcliffe efficiency values of 0.8732 and 0.7350, the SI values of 0.3108 and 0.5003, and the LMI values of 0.6585 and 0.5654. These results indicated a higher accuracy of the ANN model in predicting the IR. Figure 4(i), (j) presents a scatter plot comparing observed values with those predicted by the ANN-based models. The close alignment of these points with the line of perfect agreement further confirmed the model's enhanced ability to accurately match observed and predicted outcomes, particularly in the context of IR predictions.

This study aimed to determine the most effective AI-based model for analyzing 15 input variables – time, sand, clay, silt, PH, EC, moisture content, soil bulk density, porosity, CaCO3, P, OC, OM, N, and temperature to predict the IR accurately. The considered ML techniques included the GP, SVM, MARS, GMDH, and ANN. The comparative analysis of these models, as illustrated in Table 5, utilized seven different goodness-of-fit parameters: CC, MAE, RMSE, NSE, SI, and LMI. The findings revealed that the ANN-based model excelled over others in predicting the IR. It was evidenced by its superior metrics: a CC values of 0.9366 and 0.8624, MAE values of 0.0607 and 0.1000, RMSE values of 0.0884 and 0.1502, NSE values of 0.8732 and 0.7350, SI values of 0.3108 and 0.5003, and LMI values of 0.6585 and 0.5654 for training and testing stages, respectively. However, the results from Tables 2 and 5 also indicated that the GP–RBF and SVM–RBF models were notably reliable in predicting the IR. Figure 5(a) in the study showcased the performance of observed versus predicted values across all tested models, highlighting the proximity of the ANN model's predictions to the line of perfect agreement, thus underscoring its accuracy. Additionally, Figure 5(b) presents a pulse graph demonstrating that all models, to varying degrees, effectively predict the IR during the testing stages when using the complete dataset.
Figure 5

(a) and (b) Observed and predicted values with all applied models using a testing dataset.

Figure 5

(a) and (b) Observed and predicted values with all applied models using a testing dataset.

Close modal
Table 6 displays descriptive statistics of observed and predicted values in terms of quartile, indicating that the range of expected values for the outperforming model, ANN, has a minimum and maximum value of (−0.0150 to 1.0150) and an interquartile range (IQR) value of (0.3150). Figure 6 shows a box plot diagram demonstrating all applied models' suitability in predicting the IR. In contrast, Figure 7 shows the relative error for all applied models, confirming that ANN models are reliable in predicting the IR with a minimum number of errors using the testing dataset. Table 7 provides the descriptive statistics of observed and projected values in terms of relative error for all applied models, suggesting that the ANN model has a minimum error of (−0.4310), a maximum error of (0.4540), and a mean value of (0.4540), (0.0228). Figure 8 shows the violin graph, which shows the density of each variable and data peaks, and the relative estimate error distribution, demonstrating that the best-predicted model, ANN, has the slightest error in predicting the inflation rate.
Table 6

Descriptive statistics for all models used, based on observed and predicted values

StatisticObservedGP–RBFSVM–RBFGMDHMARSANN
Minimum 0.0120 −0.0710 0.0020 0.0945 −0.0237 −0.0150 
Maximum 1.2800 1.3130 0.9290 1.0477 1.0889 1.0150 
First quartile 0.1000 0.0830 0.1020 0.1237 0.1156 0.0920 
Median 0.1600 0.1770 0.1690 0.2023 0.1908 0.1400 
Third quartile 0.4000 0.4150 0.3420 0.3832 0.3759 0.4070 
IQR 0.3000 0.3320 0.2400 0.2594 0.2603 0.3150 
StatisticObservedGP–RBFSVM–RBFGMDHMARSANN
Minimum 0.0120 −0.0710 0.0020 0.0945 −0.0237 −0.0150 
Maximum 1.2800 1.3130 0.9290 1.0477 1.0889 1.0150 
First quartile 0.1000 0.0830 0.1020 0.1237 0.1156 0.0920 
Median 0.1600 0.1770 0.1690 0.2023 0.1908 0.1400 
Third quartile 0.4000 0.4150 0.3420 0.3832 0.3759 0.4070 
IQR 0.3000 0.3320 0.2400 0.2594 0.2603 0.3150 
Table 7

Descriptive statistics of observed and predicted values of all applied models in terms of error

StatisticGP–RBFSVM–RBFGMDHMARSANN
Minimum −0.5730 −0.5700 −0.4163 −0.5041 −0.4310 
Maximum 0.4390 0.4840 0.4987 0.5063 0.4540 
First quartile −0.0400 −0.0200 −0.0947 −0.0906 −0.0470 
Median 0.0080 0.0010 −0.0197 0.0148 0.0150 
Third quartile 0.0480 0.0850 0.0989 0.0968 0.0730 
Mean −0.0026 0.0329 0.0024 0.0192 0.0228 
StatisticGP–RBFSVM–RBFGMDHMARSANN
Minimum −0.5730 −0.5700 −0.4163 −0.5041 −0.4310 
Maximum 0.4390 0.4840 0.4987 0.5063 0.4540 
First quartile −0.0400 −0.0200 −0.0947 −0.0906 −0.0470 
Median 0.0080 0.0010 −0.0197 0.0148 0.0150 
Third quartile 0.0480 0.0850 0.0989 0.0968 0.0730 
Mean −0.0026 0.0329 0.0024 0.0192 0.0228 
Figure 6

Box plot for all applied models using the testing dataset.

Figure 6

Box plot for all applied models using the testing dataset.

Close modal
Figure 7

The relative error for all applied models using the testing dataset.

Figure 7

The relative error for all applied models using the testing dataset.

Close modal
Figure 8

Violin graph with all applied models using the testing dataset.

Figure 8

Violin graph with all applied models using the testing dataset.

Close modal

ANN has surpassed the SVM, GP, GMDH, and MARS in predicting soil IR. It is due to their ability to handle intricate, non-linear interactions, flexible structure, and proficiency in capturing temporal fluctuations. Also, the ANN was optimized using the trial-and-error method, which significantly improved the model accuracy. Despite the efficacy of SVM and other AI-based models, their performance in this work was constrained by structural limitations. It reduced adaptability to the intricate and dynamic characteristics of the IR data. The ANN model shows supremacy over all other models. Thus, this model can be used for predicting the IR of soil and could be a better replacement for the experimentation as it requires less time and effort. Overall, the numerous advantages of ANN, together with meticulous parameter optimization and tactics to address its limits, position it as a dependable alternative to experimental approaches for predicting soil IRs. This method provides an efficient solution in terms of time and effort, while more refining of computational algorithms may improve its scalability and application.

Taylor diagram

Figure 9 in the study uses the Taylor diagram, an effective graphical tool, to compare the performance of various models during the testing phase. This diagram facilitated the assessment of the models based on two critical statistical measures: the standard deviation and the CC. Regarding predicting the IR, the model indicated by the yellow dot on the Taylor diagram, which is nearest to the reference hollow circle, stands out for its superior performance. This model, identified as the ANN model, demonstrated the highest CC compared to its counterparts. Models based on the GP, SVM, MARS, and GMDH followed ANN in effectiveness.
Figure 9

Taylor diagram using the testing stage.

Figure 9

Taylor diagram using the testing stage.

Close modal

Sensitivity analysis

A sensitivity analysis was conducted to find the most influential input variables for predicting the IR. This analysis, using an optimized model such as the ANN-based one, examined the impact of various combinations of input parameters. A total of 16 scenarios were created. The first scenario included all input variables, while the remaining 15 were created by removing one variable (diagonally) from each scenario. In Table 8, this process is depicted, where models based on different combinations of inputs were developed. The effectiveness of each scenario variant was evaluated using metrics like CC, MAE, and RMSE. As presented in Table 8, the findings revealed that among all the input parameters, time (T) emerged as the most critical factor in accurately predicting the IR, as there were large variations in the evaluation metrics when T was removed in the modeling process. In other scenarios, there was low sensitivity toward IR, as in these scenarios, the CC varied from 0.8619 to 0.8365, whereas the ideal value was 0.8624. This was evidenced by a notable decrease in the CC value and an increase in the error rates (MAE and RMSE) when time was excluded from the model inputs, underscoring its sensitivity and importance in the prediction process.

Table 8

Sensitivity analysis with the ANN-based model

Input parametersOutput IRANN-based model
TSCSipHECMCSdPoCaCO3POCOMNTempCCMAERMSE
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8624 0.1000 0.1502 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.6113 0.2328 0.2662 
 ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8500 0.0995 0.1568 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8610 0.0959 0.1510 
✓ ✓ ✓ Si ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8548 0.0989 0.1538 
✓ ✓ ✓ ✓ pH ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8365 0.1034 0.1633 
✓ ✓ ✓ ✓ ✓ EC ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8490 0.1064 0.1579 
✓ ✓ ✓ ✓ ✓ ✓ MC ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8470 0.1002 0.1571 
✓ ✓ ✓ ✓ ✓ ✓ ✓ Sd ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8411 0.0986 0.1591 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ PO ✓ ✓ ✓ ✓ ✓ ✓  0.8436 0.1084 0.1591 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ CaCO3 ✓ ✓ ✓ ✓ ✓  0.8488 0.1011 0.1567 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8584 0.1066 0.1549 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ OC ✓ ✓ ✓  0.8539 0.1005 0.1558 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ OM ✓ ✓  0.8537 0.1004 0.1559 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8563 0.0985 0.154 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ Temp  0.8460 0.1026 0.159 
Input parametersOutput IRANN-based model
TSCSipHECMCSdPoCaCO3POCOMNTempCCMAERMSE
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8624 0.1000 0.1502 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.6113 0.2328 0.2662 
 ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8500 0.0995 0.1568 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8610 0.0959 0.1510 
✓ ✓ ✓ Si ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8548 0.0989 0.1538 
✓ ✓ ✓ ✓ pH ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8365 0.1034 0.1633 
✓ ✓ ✓ ✓ ✓ EC ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8490 0.1064 0.1579 
✓ ✓ ✓ ✓ ✓ ✓ MC ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8470 0.1002 0.1571 
✓ ✓ ✓ ✓ ✓ ✓ ✓ Sd ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8411 0.0986 0.1591 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ PO ✓ ✓ ✓ ✓ ✓ ✓  0.8436 0.1084 0.1591 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ CaCO3 ✓ ✓ ✓ ✓ ✓  0.8488 0.1011 0.1567 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8584 0.1066 0.1549 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ OC ✓ ✓ ✓  0.8539 0.1005 0.1558 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ OM ✓ ✓  0.8537 0.1004 0.1559 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓  0.8563 0.0985 0.154 
✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ Temp  0.8460 0.1026 0.159 

This investigation identifies the optimal model for predicting IR using several computing approaches, such as GP, SVM, GMDH, MARS, and ANN-based models. In this study, 15 input variables, including time, sand, clay, silt, pH, EC, moisture content, soil bulk density, porosity, (CaCO3, P, OC, OM, N, and temperature, were evaluated using six key performance metrics to assess the efficacy of various predictive models. These metrics comprised the CC, MAE, RMSE, Nash–Sutcliffe efficiency (NSE), the SI, and the LMI. The evaluation results highlighted the superior performance of the ANN-based model in predicting the IR. This model demonstrated outstanding results in both the training and testing phases, with CC values of 0.9366 and 0.8624, MAE values of 0.0607 and 0.1000, RMSE values of 0.0884 and 0.1502, NSE values of 0.8732 and 0.7350, the SI values of 0.3108 and 0.5003, and the LMI values of 0.6585 and 0.5654, indicative of lower error rates.

The scatter plot analysis further reinforced the ANN model's reduced error margins and optimal prediction fit. Visual representations such as box plots and Violin graphs corroborated the model's efficacy in predicting the IR with minimal errors. In the Taylor diagram, the ANN model was depicted as excelling over other models, confirming its suitability for accurately predicting the IR. Additionally, the sensitivity analysis pointed out the significant impact of the time (T) variable on the IR, underscoring its importance in the model's predictive accuracy.

No funding was reported for this research.

This article contains no studies with human participants or animals performed by any of the authors.

Data cannot be made publicly available; readers should contact the corresponding author for details.

The authors declare there is no conflict.

Abd-Elmaboud
M. E.
,
Saqr
A. M.
,
El-Rawy
M.
,
Al-Arifi
N.
&
Ezzeldin
R.
(
2024
)
Evaluation of groundwater potential using ANN-based mountain gazelle optimization: A framework to achieve SDGs in East El Oweinat, Egypt
,
Journal of Hydrology: Regional Studies
,
52
,
101703
.
Al-Azawi
S. A.
(
1985
)
Experimental evaluation of infiltration models
,
Journal of Hydrology (New Zealand)
,
24
(
2
),
77
88
.
Anastasakis
L.
&
Mort
N.
(
2001
)
The Development of Self-Organization Techniques in Modeling: A Review of the Group Method of Data Handling (GMDH). Research report-University of Sheffied, UK, Department Of Automatic Control And Systems Engineering
.
Angelaki
A.
,
Sakellariou-Makrantonaki
M.
&
Tzimopoulos
C.
(
2013
)
Theoretical and experimental research of cumulative infiltration
,
Transport in Porous Media
,
100
(
2
),
247
257
.
Aradhana
A.
,
Singh
B.
&
Sihag
P.
(
2021
)
Predictive models for estimation of labyrinth weir aeration efficiency
,
Journal of Achievements in Materials and Manufacturing Engineering
,
105
(
1
),
18
32
.
Arora
G.
,
Kumar
D.
&
Singh
B.
(
2024
). '
Tree based regression models for predicting the compressive strength of concrete at high temperature
',
IOP Conference Series: Earth and Environmental Science
, Vol.
1327
(
1
), p.
012015
.
IOP Publishing
.
Bayabil
H. K.
,
Dile
Y. T.
,
Tebebu
T. Y.
,
Engda
T. A.
&
Steenhuis
T. S.
(
2019
)
Evaluating infiltration models and pedotransfer functions: Implications for hydrologic modeling
,
Geoderma
,
338
,
159
169
.
Bhoria
S.
,
Sihag
P.
,
Singh
B.
,
Ebtehaj
I.
&
Bonakdari
H.
(
2021
)
Evaluating Parshall flume aeration with experimental observations and advance soft computing techniques
,
Neural Computing and Applications
,
33
,
17257
17271
.
Brown
A. E.
,
Zhang
L.
,
McMahon
T. A.
,
Western
A. W.
&
Vertessy
R. A.
(
2005
)
A review of paired catchment studies for determining changes in water yield resulting from alterations in vegetation
,
Journal of Hydrology
,
310
(
1–4
),
28
61
.
Cortes
C.
&
Vapnik
V.
(
1995
)
Support-vector networks
,
Machine Learning
,
20
(
3
),
273
297
.
Performance concrete with Gaussian process regression model. Advances in Civil Engineering, 2016
.
De Andrés
J.
,
Lorca
P.
,
de Cos Juez
F. J.
&
Sánchez-Lasheras
F.
(
2011
)
Bankruptcy forecasting: A hybrid approach using fuzzy c-means clustering and multivariate adaptive regression splines (MARS)
,
Expert Systems with Applications
,
38
(
3
),
1866
1875
.
de Andrés
J.
,
Sánchez-Lasheras
F.
,
Lorca
P.
&
De Cos Juez
F. J.
(
2011
)
A hybrid device of self-organizing maps (SOM) and multivariate adaptive regression splines (MARS) for the forecasting of firms'bankruptcy
,
Accounting & Management Information Systems/Contabilitate Si Informatica de Gestiune
,
10
(
3
),
351
374
.
Dutta
S.
,
Samui
P.
&
Kim
D.
(
2018
)
Comparison of machine learning techniques to predict compressive strength of concrete
,
Computers and Concrete
,
21
(
4
),
463
470
.
Friedman
J. H.
(
1991
)
Multivariate adaptive regression splines
,
The Annals of Statistics
,
19
(
1
),
1
67
.
García-Ruiz
J. M.
,
Regüés
D.
,
Alvera
B.
,
Lana-Renault
N.
,
Serrano-Muela
P.
,
Nadal-Romero
E.
,
Navas
A.
,
Latron
J.
,
Martí-Bono
C.
&
Arnáez
J.
(
2008
)
Flood generation and sediment transport in experimental catchments affected by land use changes in the central Pyrenees
,
Journal of Hydrology
,
356
(
1–2
),
245
260
.
Hoang
N. D.
,
Pham
A. D.
,
Nguyen
Q. L.
&
Pham
Q. N.
(
2016
)
Estimating compressive strength of high- .
Ivakhnenko
A. G.
&
Stepashko
V. S.
(
1985
)
Noise Immunity of Modeling
.
Kiev
:
Naukova Dumka
, p.
216
.
Liu
H.
,
Lei
T. W.
,
Zhao
J.
,
Yuan
C. P.
,
Fan
Y. T.
&
Qu
L. Q.
(
2011
)
Effects of rainfall intensity and antecedent soil water content on soil infiltrability under rainfall conditions using the run-off-on-out method
,
Journal of Hydrology
,
396
(
1–2
),
24
32
.
Ma
D.
&
Shao
M.
(
2008
)
Simulating infiltration into stony soils with a dual-porosity model
,
European Journal of Soil Science
,
59
(
5
),
950
959
.
Madala
H. R.
(
2019
)
Inductive Learning Algorithms for Complex Systems Modeling
.
CRC press
,
Boca Raton
.
Mishra
S. K.
,
Tyagi
J. V.
&
Singh
V. P.
(
2003
)
Comparison of infiltration models
,
Hydrological Processes
,
17
(
13
),
2629
2652
.
Muñoz
D. R.
,
Villas
D. B.
,
Arnedo
M. T. E.
,
Negrell
M. A. G.
,
Monreal
N. L. R.
,
León
J.
,
Romero
M. E. N.
,
Pardini
G.
&
Muela
M. P. S.
(
2017
)
Analysing the effect of land use and vegetation cover on soil infiltration in three contrasting environments in northeast Spain
,
Cuadernos de Investigación Geográfica/Geographical Research Letters
,
43
,
141
169
.
Nivesh
S.
,
Negi
D.
,
Kashyap
P. S.
,
Aggarwal
S.
,
Singh
B.
,
Saran
B.
,
Sawant
P. N.
&
Sihag
P.
(
2022
)
Prediction of river discharge of Kesinga sub catchment of Mahanadi basin using machine learning approaches
,
Arabian Journal of Geosciences
,
15
(
16
),
1369
.
Omran
B. A.
,
Chen
Q.
&
Jin
R.
(
2016
)
Comparison of data mining techniques for predicting compressive strength of environmentally friendly concrete
,
Journal of Computing in Civil Engineering
,
30
(
6
),
04016029
.
Pandey
P. K.
&
Pandey
V.
(
2018
)
Estimation of infiltration rate from readily available soil properties (RASPs) in fallow cultivated land
,
Sustainable Water Resources Management
,
5
(
2
),
921
934
.
Pandhiani
S. M.
,
Sihag
P.
,
Shabri
A. B.
,
Singh
B.
&
Pham
Q. B.
(
2020
)
Time-series prediction of streamflows of Malaysian rivers using data-driven techniques
,
Journal of Irrigation and Drainage Engineering
,
146
(
7
),
04020013
.
Parhi
P. K.
,
Mishra
S. K.
&
Singh
R.
(
2007
)
A modification to Kostiakov and modified Kostiakov infiltration models
,
Water Resources Management
,
21
(
11
),
1973
1989
.
Reicosky
D.
(
2020
)
Conservation agriculture systems: Soil health and landscape management. In Advances in Conservation Agriculture (pp. 87–154). Burleigh Dodds Science Publishing
.
Saqr
A. M.
,
Nasr
M.
,
Fujii
M.
,
Yoshimura
C.
&
Ibrahim
M. G.
(
2022
). '
Optimal solution for increasing groundwater pumping by integrating MODFLOW-USG and particle swarm optimization algorithm: A case study of Wadi El-Natrun, Egypt
',
International Conference on Environment Science and Engineering
, pp.
59
73
.
Singapore
:
Springer Nature Singapore
.
Seiler
K. P.
&
Gat
J. R.
(
2007
)
Groundwater Recharge From run-off, Infiltration and Percolation
, Vol.
55
.
Springer Science & Business Media
.
Verlag, Dordrecht, Netherlands
.
Sepahvand
A.
,
Singh
B.
,
Ghobadi
M.
&
Sihag
P.
(
2021a
)
Estimation of infiltration rate using data-driven models
,
Arabian Journal of Geosciences
,
14
,
1
11
.
Sepahvand
A.
,
Singh
B.
,
Sihag
P.
,
Nazari Samani
A.
,
Ahmadi
H.
&
Fiz Nia
S.
(
2021b
)
Assessment of the various soft computing techniques to predict sodium absorption ratio (SAR)
,
ISH Journal of Hydraulic Engineering
,
27
(
sup1
),
124
135
.
Sharda
V. N.
,
Patel
R. M.
,
Prasher
S. O.
,
Ojasvi
P. R.
&
Prakash
C.
(
2006
)
Modeling runoff from middle Himalayan watersheds employing artificial intelligence techniques
,
Agricultural Water Management
,
83
(
3
),
233
242
.
Shirmohammadi
A.
&
Skaggs
R. W.
(
1984
)
Effect of soil surface conditions on infiltration for shallow water table soils
,
Transactions of the ASAE
,
27
(
6
),
1780
1787
.
Sihag
P.
,
Tiwari
N. K.
&
Ranjan
S.
(
2017
)
Prediction and inter-comparison of infiltration models
,
Water Science
,
31
(
1
),
34
43
.
Sihag
P.
,
Esmaeilbeiki
F.
,
Singh
B.
,
Ebtehaj
I.
&
Bonakdari
H.
(
2019
)
Modeling unsaturated hydraulic conductivity by hybrid soft computing techniques
,
Soft Computing
,
23
,
12897
12910
.
Sihag
P.
,
Kumar
M.
&
Singh
B.
(
2021
)
Assessment of infiltration models developed using soft computing techniques
,
Geology, Ecology, and Landscapes
,
5
(
4
),
241
251
.
Sihag
P.
,
Singh
B.
,
Said
M. A. B. M.
&
Azamathulla
H. M.
(
2022
)
Prediction of Manning's coefficient of roughness for high-gradient streams using M5P
,
Water Supply
,
22
(
3
),
2707
2720
.
Silverman
D.
&
Dracup
J. A.
(
2000
)
Artificial neural networks and long-range precipitation prediction in California
,
Journal of Applied Meteorology
,
39
(
1
),
57
66
.
Singh
B.
&
Minocha
V. K.
(
2024a
). '
Comparative study of machine learning techniques for prediction of scour depth around Spur Dikes
’,
World Environmental and Water Resources Congress 2024
. pp.
635
651
.
Singh
B.
&
Minocha
V. K.
(
2024b
). '
Clear water scour depth prediction using gradient boosting machine and deep learning
',
IOP Conference Series: Earth and Environmental Science
, Vol.
1327
(
1
), p.
012030
.
IOP Publishing
.
Singh
B.
&
Minocha
V. K.
(
2024c
)
Approximation in scour depth around spur dikes using novel hybrid ensemble data-driven model
,
Water Science & Technology
,
89
(
4
),
962
975
.
Singh
B.
,
Sihag
P.
&
Singh
K.
(
2017
)
Modelling of impact of water quality on infiltration rate of soil by random forest regression
,
Modeling Earth Systems and Environment
,
3
,
999
1004
.
Singh
B.
,
Sihag
P.
&
Singh
K.
(
2018
)
Comparison of infiltration models in NIT Kurukshetra campus
,
Applied Water Science
,
8
(
2
),
63
.
Singh
A.
,
Singh
B.
&
Sihag
P.
(
2021a
)
Experimental investigation and modeling of aeration efficiency at labyrinth weirs
,
Journal of Soft Computing in Civil Engineering
,
5
(
3
),
15
31
.
Singh
B.
,
Sihag
P.
,
Parsaie
A.
&
Angelaki
A.
(
2021b
)
Comparative analysis of artificial intelligence techniques for the prediction of infiltration process
,
Geology, Ecology, and Landscapes
,
5
(
2
),
109
118
.
Singh
B.
,
Ebtehaj
I.
,
Sihag
P.
&
Bonakdari
H.
(
2022
)
An expert system for predicting the infiltration characteristics
,
Water Supply
,
22
(
3
),
2847
2862
.
Singh
K.
,
Singh
B.
,
Sihag
P.
,
Kumar
V.
&
Sharma
K. V.
(
2023
)
Development and application of modeling techniques to estimate the unsaturated hydraulic conductivity
,
Modeling Earth Systems and Environment
,
9
(
4
),
4557
4571
.
Thakur
M. S.
,
Pandhiani
S. M.
,
Kashyap
V.
,
Upadhya
A.
&
Sihag
P.
(
2021
)
Predicting bond strength of FRP bars in concrete using soft computing techniques
,
Arabian Journal for Science and Engineering
,
46
(
5
),
4951
4969
.
Upadhya
A.
,
Thakur
M. S.
,
Mashat
A.
,
Gupta
G.
&
Adbo
M. S.
(
2022
)
Prediction of binder content in glass fiber reinforced asphalt Mix using machine learning techniques
,
IEEE Access.
10, 33866–33881.
Vand
A. S.
,
Sihag
P.
,
Singh
B.
&
Zand
M.
(
2018
)
Comparative evaluation of infiltration models
,
KSCE Journal of Civil Engineering
,
22
(
10
),
4173
4184
.
Wang
Y.
,
Fan
J.
,
Cao
L.
&
Liang
Y.
(
2016
)
Infiltration and runoff generation under various cropping patterns in the red soil region of China
,
Land Degradation & Development
,
27
(
1
),
83
91
.
Williams
C. K.
&
Rasmussen
C. E.
(
2006
)
Gaussian Processes for Machine Learning 2(3), 4
.
Cambridge, MA
:
MIT Press
.
Yang
G.
,
Yu
W.
,
Li
Q. J.
,
Wang
K.
,
Peng
Y.
&
Zhang
A.
(
2019
)
Random forest-based pavement surface friction prediction using high-resolution 3D image data
,
Journal of Testing and Evaluation.
,
49
(
2
),
1141
1152
.
Zealand
C. M.
,
Burn
D. H.
&
Simonovic
S. P.
(
1999
)
Short-term streamflow forecasting using artificial neural networks
,
Journal of Hydrology
,
214
(
1–4
),
32
48
.
Ziari
H.
,
Sobhani
J.
,
Ayoubinejad
J.
&
Hartmann
T.
(
2016
)
Prediction of IRI in short and long terms for flexible pavements: ANN and GMDH methods
,
International Journal of Pavement Engineering
,
17
(
9
),
776
788
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).