Next Article in Journal
The Impact of Remanufactured Products’ Similarity on Purchase Intention of New Products
Next Article in Special Issue
A Path Model of the Intention to Adopt Variable Rate Irrigation in Northeast Italy
Previous Article in Journal
Identify Optimal Traffic Condition and Speed Limit for Hard Shoulder Running Strategy
Previous Article in Special Issue
Potential of On-the-Go Gamma-Ray Spectrometry for Estimation and Management of Soil Potassium Site Specifically
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quantitative Evaluation of Soil Quality Using Principal Component Analysis: The Case Study of El-Fayoum Depression Egypt

by
Mohamed K. Abdel-Fattah
1,*,
Elsayed Said Mohamed
2,*,
Enas M. Wagdi
1,
Sahar A. Shahin
3,
Ali A. Aldosari
4,*,
Rosa Lasaponara
5 and
Manal A. Alnaimy
1
1
Soil Science Department, Faculty of Agriculture, Zagazig University, Zagazig 44519, Egypt
2
National Authority for Remote Sensing and Space Sciences, Cairo 11843, Egypt
3
Soils and Water Use Department, Agriculture and Biological Division, National Research Center, Cairo 12622, Egypt
4
Geography Department, King Saud University, Riyadh 11451, Saudi Arabia
5
Italian National Research Council, C.da Santa Loja, Tito Scalo, 85050 Potenza, Italy
*
Authors to whom correspondence should be addressed.
Sustainability 2021, 13(4), 1824; https://doi.org/10.3390/su13041824
Submission received: 11 December 2020 / Revised: 26 January 2021 / Accepted: 29 January 2021 / Published: 8 February 2021
(This article belongs to the Special Issue Smart Farming and Sustainability)

Abstract

:
Soil quality assessment is the first step towards precision farming and agricultural management. In the present study, a multivariate analysis and geographical information system (GIS) were used to assess and map a soil quality index (SQI) in El-Fayoum depression in the Western Desert of Egypt. For this purpose, a total of 36 geo-referenced representative soil samples (0–0.6 m) were collected and analyzed according to standardized protocols. Principal component analysis (PCA) was used to reduce the dataset into new variables, to avoid multi-collinearity, and to determine relative weights (Wi) and soil indicators (Si), which were used to obtain the soil quality index (SQI). The zones of soil quality were determined using principal component scores and cluster analysis of soil properties. A soil quality index map was generated using a geostatistical approach based on ordinary kriging (OK) interpolation. The results show that the soil data can be classified into three clusters: Cluster I represents about 13.89% of soil samples, Cluster II represents about 16.6% of samples, and Cluster III represents the rest of the soil data (69.44% of samples). In addition, the simulation results of cluster analysis using the Monte Carlo method show satisfactory results for all clusters. The SQI results reveal that the study area is classified into three zones: very good, good, and fair soil quality. The areas categorized as very good and good quality occupy about 14.48% and 50.77% of the total surface investigated, and fair soil quality (mainly due to salinity and low soil nutrients) constitutes about 34.75%. As a whole, the results indicate that the joint use of PCA and GIS allows for an accurate and effective assessment of the SQI.

1. Introduction

Precision agriculture is based on the use of a set of techniques and technologies devised to assess the spatial variability of soil and plant properties to facilitate and optimize soil management, which often requires the use of several variables to support decision-making [1,2]. However, in some cases, numerous soil variables are required to assess soil quality. Because some of these variables can be redundant, the ability to identify key parameters/variables can reduce both the time and costs of in situ and laboratory analyses and optimize models and procedures for spatio-temporal soil assessment [3]. In this context, principal component analysis (PCA) is recognized as one of the most widely used methods for reducing the number of variables by identifying those that are most significant in the data. In addition, PCA is useful for different aspects of farming processes, for example, the assessment of vegetative growth and the discrimination of maturation stages needed for optimizing the modeling of crop yield. The spatial variability of the chemical/physical attributes of soil is generally denoted by the Soil Quality Index (SQI) [4]. Since soil quality is linked to soil productivity, a reliable assessment requires an accurate multi-faceted quantification to support sustainable soil management [5]. The SQI is defined as the ability (within the ecosystem) of the soil to supply plants with the nutrients needed to maintain the crop yield throughout growth stages [6,7,8]. The SQI includes physical, chemical, and biological characteristics, which can be used to indicate the fertility status and soil health [9] through a quantitative assessment [10]. The physical indicators include soil depth, bulk density, porosity, aggregate stability, texture, and compaction, whereas the chemical indicators include pH, salinity, organic matter content, phosphorus availability, cation exchange capacity, nutrient cycling, and the level of contaminants in the soil [11]. The physical indicators inform us of the root growth, speed of plant emergence, and water infiltration, while the chemical indicators provide information on organisms and nutrient availability, as well as water for plants and the mobility of contaminants [9]. Rangel-Peraza et al. [12] reported a highly significant correlation between chemical soil properties (electric conductivity (EC), pH, cation exchange capacity (CEC), and organic matter (OM)) and the soil quality index. The major factors that influence soil quality are bulk density, soil CEC, root depth, and soil texture [13]. Hence, understanding the soil quality is important for illustrating the potential steps of proper soil management for sustainable agricultural production [14]. Recently, several methods have been developed for estimating soil quality, including quantitative and qualitative methods [11,15]. The quantitative and qualitative evaluations of soil quality are linked to the diagnostic properties of the soil. Mohamed et al. [16] and Martinez-Salgado, Gutiérrez-Romero, Jannsens, and Ortega-Blu [9] reported that certain features are closely related to the soil quality (including visual indicators), such as erosion, the presence of weeds, color, and the types of coverage. Quantitative assessments of soil quality depend on data obtained from laboratory analyses of physical and chemical properties, while qualitative assessments depend on direct observation of the soil [8]. Moreover, the spatio-temporal dimensions must be considered in the SQI assessment, because soil properties are not permanent characteristics [17,18,19]. The use of Geographic Information System (GIS) technology has facilitated the computation of the spatial variability of different phenomena [20], including investigations on soil properties. Therefore, integrated GIS and geostatistical analyses can be useful for assessing the spatial variation of soil properties and predicting them in unsampled locations. For example, the use of variogram analyses can capture and accurately map the complex spatial relationships between soil data layers [21,22,23]. Kriging is one of the most commonly used interpolation methods [24], and it can suitably support precise farming based on the identification of homogeneous sub-sets of similar yield-limiting factors [25,26]. PCA and cluster analysis are among the most widely used multivariate analysis methods for the recognition, classification, and modeling of data [27], including for soil investigations. PCA is a statistical approach that is useful for reducing the number of features in the dataset through the identification of the most important principal components (PCs), which explain the maximum information content present in the data [28]. PCA has many advantages:
  • Removes correlated features that undermine the statistical significance of an independent variable [29];
  • Improves algorithm performance, which can be significantly degraded if too many features are present in models, and speeds up analyses [30];
  • Reduces overfitting: PCA helps in overcoming the overfitting issue by minimizing the number of variables in the investigated dataset [31];
  • Improves visualization: PCA transforms a high-dimensional dataset into a low-dimensional one while preserving the information content and making data visualization and exploration easier [32,33]
The main aim of the present study is to assess, characterize, and map the SQI using a multivariate analysis based on the joint use of PCA and GIS in El-Fayoum depression in the Western Desert of Egypt.

2. Materials and Methods

2.1. Study Area

The study area is located in El-Fayoum Governorate, Western Desert of Egypt. It is bounded by latitudes 29°15′–29°35′ N and longitudes 30°32′30″–30°52′30.59″ E. The study area is characterized by an elevation that reaches 23 m above sea level. The area is connected to the Nile River by the Hawara canal through the Bahr Yousef canal (Figure 1). The physiographic units of El-Fayoum depression include three main landscapes, i.e., lacustrine plain, fluvio-lacustrine plain, and alluvial plain [34]. The main landforms in the area are recent and old lake terraces, depressions, plains, and basins [35] with varying vegetation cover; therefore, the sensitivity to desertification differs widely in the study area [36]. The climate of the study area is characterized by a hot and dry summer with limited winter rainfall and bright sunshine throughout the year. The area has low annual rainfall of around 7.2 mm/year, and the mean minimum and maximum annual temperatures are 14.5 and 31.0 °C, respectively. The lowest evaporation rate (1.9 mm/day) is recorded in January, while the highest value (7.3 mm/day) is recorded in June [34].

2.2. Sampling and Soil Analysis

The soil samples were collected using GPS and a soil cylinder auger (Figure 1) at depths of 0–60 cm in 36 different locations. One mixed sample in each location was collected that represents the soil of root zone. The selected sites represent spatial changes in the study area, which is characterized by wide variation of physiographic features, such as lacustrine plain, fluvio-lacustrine plain, and alluvial plain [34]. The area is characterized by slope levels ranging between −15 and 45 m above sea level, and the change in slope has directly affected vegetation density and land suitability [37]. The soil classifications in the study area include Vertic Torrifluvent, Typic Haplocalcids, Typic Torrifluvents, Typic Haplogypsids, Typic Haplosalids, Typic Torripsamments, and Typic Haplargids [34].
The samples were air-dried, ground, and passed through a 2 mm sieve to prepare them for physical and chemical analyses according to standardized protocols described in [38,39,40]. The soil reaction (pH) of a 1:2.5 soil-to-water suspension was measured using a glass electrode [39]. The soil electrical conductivity was assessed in saturated soil paste extract (ECe) [39]. The Walkley and Black method was used to determine the soil organic matter [38,40]. Available nitrogen was determined by distillation using the micro-Kjeldahl method [40]. Available phosphorus was determined calorimetrically using the ascorbic acid method [40]. Available potassium was extracted with 1N NH4OAc at pH 7 and was measured using a flame-photometer device [40]. The exchangeable sodium percentage (ESP) was computed based on the mathematical equation described by van Reeuwijk [38]. The sodium acetate method was used to measure CEC [38]. Soil particle analyses were performed according to an international pipette method and based on the percentage of sand, silt, and clay; the soil texture was determined using the international texture triangle [38].

2.3. Statistical Analysis

Descriptive statistics of the studied soil characteristics include the minimum, maximum, arithmetic mean, and standard deviation, which were computed using SPSS version 25. The Shapiro–Wilk test was used to assess the normal distribution of the data. The Pearson correlation coefficient (r) was used to examine the linear relationships between the variables. XLSTAT software 2016 and SPSS version 25 were used to conduct the principal component analysis (PCA). PCA was used to reduce the dataset into new variables, which are called principal components (PCs), as well as to avoid multicollinearity between the original variables. These PCs explain most of the variation present in the original variables.

2.4. Soil Quality Index (SQI) Calculation and Mapping

The SQI was calculated using Equation (1) according to Cude [41]:
SQI = i = 1 N W i × S i
where Wi is the relative weight of each indicator and has values ranging between 0 and 1, and Si is the value of each soil indicator.
Wi expresses the component score coefficient (CSC) that is obtained from the PCA results. Because the soil indicators have different scales and units, the Si values are standardized using Equation (2) [42]:
z = x   x ¯ σ
where z, x, x ¯ and σ refer to the standardized value, the value of a soil indicator, the average of a soil indicator, and the standard deviation of a soil indicator, respectively.
Therefore, the SQI equation based on principal components (PCs) becomes the following (Equation (3)):
SQI PC = i = 1 N C S C × z
Thus, the comprehensive SQI (CSQI) is computed using Equation (4):
CSQI = i = 1 N V a r i a b i l i t y   o f   e a c h   P C     ×     S Q I P C
The CSQI, which is calculated using z scores, is transformed into a standard normal distribution (which has a mean of zero and a standard deviation of one) using Equation (5) [42]:
f ( x ) = 1 2 π e ( z ) 2 2
where e and z refer to the natural logarithm, equal to approximately 2.718, and the CSQI, which is computed using z scores, respectively.
Aprisal, Bambang, and Harianti [13] reported that the soil quality could be classified into the following conditions: very good (0.8–1), good (0.6–0.79), fair (0.35–0.59), bad (0.20–0.34), and very bad (0–0.19).

2.5. Cluster Analysis

From the PC scores of soil samples, a cluster analysis was performed using k-means to categorize the observations into groups [43,44,45]. This analysis was applied to classify the soils into specific zones according to their properties. A one-way ANOVA test and Duncan multiple range (DMR) test were performed for comparisons between the different soil zones that were generated.
The cluster analysis results were also simulated using the Monte Carlo approach, one of the most popular and widely used methods for simulation and probabilistic analyses based on the generation of a large number of random samples. This step was adopted to confirm the clusters obtained from the previous analyses [46,47].

2.6. Geostatistical Analyses

The geostatistical approach was adopted to predict the values of variables in unsampled locations using the ordinary kriging (OK) method. Semivariograms of the soil parameters were generated using the average squared differences among all pairs (Equation (6)) [48]:
γ ( h ) =   1 2 N ( h ) i = 1 N ( h ) [ Z ( x i ) Z ( x i + h ) ] 2
where γ(h) is the semivariance of the distance interval h, N(h) is the number of pairs of the lag interval, Z(xi) is the measured sample value at point i, and Z(xi + h) is the measured sample value at position (i + h).
The best semivariogram models were selected based on strong spatial dependence (SDC), mean error (ME), root-mean-square error (RMSE), mean standardized error (MSE), root-mean-square standardized error (RMSSE), and average standard error (ASE). If the values of ME, MSE, and ASE are close to zero and the RMSE is close to one, this indicates that the quality and suitability of the predicting model are high [49]. In addition, ratios of nugget to sill (SDC) of <0.25, 0.25–0.75, and >0.75 indicate strong, moderate, and weak spatial dependence, respectively [50].
A spatial distribution map of the soil quality index was generated using ordinary kriging interpolation in ArcGIS software version 10.2, where the kriging method was applied to predict the values of variables in unsampled locations and to interpolate the spatial soil properties using Equation (7) [51]:
Z * ( x o ) = i = 1 N λ i Z ( X i )
where Z*(xo) is an estimated variable at location xo, Z*(Xi) is the value of an inspected variable at location Xi, λi is the statistical weight that is attributed to Z*(Xi) for a sample located near xo, and N is the number of observations in the neighborhood of the inspected point. The flowchart of the procedures used to determine the soil quality index in this study is shown in Figure 2.

3. Results and Discussion

3.1. Soil Characteristics of the Study Area

The soil characteristics of the study area are listed in Table 1. In particular, the pH values range from 7.09 to 8.65, with an average value of 7.86 ± 0.47, which indicates that the conditions of the study area are mildly/strongly alkaline [40]. The results indicate that the study area is characterized by moderate to high salinity soils, with ECe values varying from 0.87 to 20.33 dSm−1 with an average value of 5.30 ± 5.05 dSm−1 [52]. The CEC of the study area varies within a wide range, between 3.45 and 40.23 cmolckg−1 soil, with an average of 20.62 ± 8.79 cmolckg−1 soil.
The ESP values range from 1.86 to 17.13, with an average of 9.75 ± 3.67, which indicates that the area is not exposed to sodicity hazards [53]. The OM contents range from low to high in the study area, in agreement with [40], with an average of 0.69 ± 0.46. The available N ranges between 1.33 mg kg1 (2.98 kg N ha−1) and 61.6 mg kg−1(138 kg N ha−1) with an average of 19.91 ± 17.42 mg kg−1 (44.6 ± 39 kg N ha−1), indicating that the nitrogen content in the area is low [40]. The available P content ranges from low (2.33 mg kg−1; 12.0 kg P ha−1) to high (19.84 mg kg−1; 101 kg P ha−1), with an average of 9.50 ± 4.51 mg kg−1 (48.7 kg P ha−1), and available K ranges from low (32.76 mg kg−1; 88.1 kg K ha−1) to high (734 mg kg−1; 1972 kg K ha−1), with an average of 183.5 ± 193 mg kg−1 (493 ± 519 kg K ha−1), which is classified as high according to [40]. The soil texture, which refers to the proportions of silt, clay, and sand, varies from 8.19 to 44.76%, 24.98 to 62.09%, and 12.98 to 55.95%, respectively.

3.2. Pearson Correlation Matrix, Bartlett’s, and Kaiser Meyer Olkin (KMO) Tests

The correlations between soil indicators are listed in Table 2. The soil pH has a statistically significant negative relationship (p < 0.05) with all other soil indicators except for silt content (which exhibits a significant positive relationship). Soil EC has significant positive correlations (p < 0.05) with N (r = 0.59), P (r = 0.43), ESP (r = 0.55), and clay (0.35), while its correlations with K (r = 0.30), CEC (r = 0.23), and organic matter (r = 0.26) are positive but not significant. The results show that ECe has a significant negative relationship with pH and silt (p < 0.05). The soil organic matter has a significant positive correlation (p < 0.05) with CEC, available N, available P, available K, and clay, while it has a non-significant positive correlation (r = 0.24) with ESP and a non-significant negative correlation (r = −0.06) with silt. CEC is significantly positively correlated (p < 0.05) with available N, P, K, ESP, and clay, while it has a non-significant positive correlation with silt. Available N is significantly positively correlated (p < 0.05) with available P, available K, ESP, and clay, and it has a significant positive correlation with silt. Available P is significantly positively correlated (p < 0.05) with available K, ESP, and clay and negatively correlated with silt. Available K has a significant positive correlation (p < 0.05) with clay, while it has a non-significant positive correlation with ESP and non-significant negative correlation with silt.
Soil pH affects other soil variables and controls the soil physical, chemical, and biological properties [54,55]; thus, pH demonstrates significant correlations with other properties. The negative correlation between EC and pH is largely dependent on the leaching process of major cations (Ca, Mg, Na, and K), as the reduction of these cations increases the pH and decreases EC and ESP. This process is also accompanied by increased mineralization and dissociation processes of organic matter, which explains the negative relation [56]. The increased decomposition of OM at low pH values leads to increases in H+ ion content, soil CEC, and the availability of macronutrients (N, P, and K) [54,57]. Additionally, higher soil pH leads to increases in the mineralizable fractions of N and C ratios, where the bonds between clays and organic constituents are broken [58]. Clay content is associated with an increase in CEC and basic alkali cation adsorption, which is negatively correlated with pH [57]. There are positive correlations between soil EC, macronutrients (N, P and K), base cations, and clay content; these results agree with those in [59]. Increased OM has a positive correlation with clay content, which leads to increases in CEC and N, P, and K contents; in addition, OM improves soil physical and chemical properties [60,61]. Negative correlations were identified between clay and silt. The reverse effects exerted by clay and silt on other soil properties mainly depend on the ratio in which they contribute to soil particle size distribution because the surface area and CEC of clay are greater than those of silt [57].
Table 3 shows the results of Bartlett’s test of sphericity and the KMO test of sampling adequacy. The significance level of Bartlett’s test of sphericity was <0.0001, and the observed chi-square value was 334.63, which is larger than the critical chi-square value of 61.66; therefore, the variables are not completely uncorrelated, and PCA is appropriate for the dataset [62]. The results show that the KMO value is greater than 0.6, which indicates that the sample size is suitable for assessing the factor structure, in agreement with Barrett and Morgan [63]. According to results of these tests, the variables are not completely uncorrelated; the variables included in the model can explain the phenomenon, and a Principal Component Analysis is suitable [64,65,66].

3.3. Soil Quality Index Using Principal Component Analysis

3.3.1. Principal Component Analysis

The results of PCA are summarized in Table 4. The first three Principal Components (PCs) have eigenvalues greater than 1; therefore, these PCs were used according to the method described by Kaiser [67], while the other PCs were excluded (Table 4 and Figure 3). The results show that the first three PCs explain 83.63% of the total variance. According to the factor loadings, the first PC, which explains 56.45% of the total variance, has higher positive correlations with EC, OM, CEC, available NPK, ESP, and clay, while the second PC, which explains 16.76% of the total variance, is strongly correlated with silt. The third PC explains 10.41% of the total variance and is correlated with ESP. The PCA biplot in Figure 4 shows both the PC scores of samples and the loadings of variables.
The soil quality index was generated using the results of PCA using Equation (3) as follows:
SQI-PC1 = −0.14 × z_pH + 0.11 × z_EC+0.13 × z_OM + 0.13 × z_CEC + 0.17 × z_N+0.16 × z_P + 0.15 × z_K + 0.11 × z_ESP−0.07 × z_SILT + 0.13 × z_CLAY
SQI-PC2 = 0.27 × z_pH − 0.34 × z_EC + 0.25 × z_OM + 0.33 × z_CEC + 0.05 × z_N + 0.05 × z_P + 0.12 × z_K − 0.19 × z_ESP − 0.42 × z_SILT + 0.09 × z_CLAY
SQI-PC3 = −0.01 × z_pH + 0.14 × z_EC−0.26 × z_OM + 0.23 × z_CEC − 0.16 × z_N − 0.28 × z_P − 0.23 × z_K + 0.58 × z_ESP + 0.38 × z_SILT + 0.43 × z_CLAY
Using Equation (4), the CSQI was computed as follows:
CSQI = 0.564 5 × PC1 + 0.1676 × PC2 + 0.1041 × PC3
The CSQI, which was computed using z scores, was transformed into a standard normal distribution using Equation (5). The results of CSQI are presented in Table 5 and Figure 5. The results reveal highly significant correlations between the different soil indicators and SQI.

3.3.2. Cluster Analysis (k-Means Clustering)

Clustering is an effective statistical approach to data analysis that can be used to classify a large number of variables into specific groups. Each group represents a specific class of soil quality. According to the PC scores of samples, the data were divided into three clusters (Table 6). Cluster I occupies about 13.89% of the total data, Cluster II occupies about 16.67%, and Cluster III occupies the rest of the data, which represents about 69.44%. The results of ANOVA show that a statistically significant difference exists between different clusters, mainly in the SQI.

3.3.3. Simulation of Cluster Analysis

The cluster analysis results were confirmed using Monte Carlo simulations based on 200 random values of the SQI for the three clusters (first, second, and third). Figure 6 shows the normal probability distribution, where the p value of the Anderson–Darling normality test is >0.05. The SQI simulation results are acceptable, with standard deviations of 0.03, 0.07, and 0.10 and mean values of 0.88, 0.67, and 0.37 for the first, second, and third cluster, respectively. The coefficient of variance (CV) was used to assure the quality of the cluster analysis [47]; the resulting CVs are 3.18%, 9.55%, and 26.25% from the average values of the first, second, and third cluster, respectively. Additionally, the mean values are close to the median values of 0.88, 0.68, and 0.38 for the first, second, and third cluster, respectively. Therefore, the mean values of the obtained SQIs are representative of the most probable SQI values of this study area.

3.4. Mapping Soil Properties and Soil Quality Index

3.4.1. Mapping Soil Properties

The ordinary kriging interpolation method was used to estimate and map the unknown values of soil properties. The model’s accuracy was confirmed for each soil property based on ME, RMSE, MSE, and RMSSE, as shown in Table 7. The results show that the exponential model is the most suitable for predicting the unknown values of most of soil properties (CEC, ESP, Av.N, Av.P, clay, and silt), followed by the K-Bessel model for pH and OM. Tetraspherical is the most suitable for ECe and Av.K. Finally, the spherical model is suitable for sand content. In addition, the results indicate that RMSSE is close to one and the MSE is close to zero for the selected soil properties; therefore, the selected models fit the data and are suitable for predicting the unsampled soil properties [68,69].
The results show that the spatial dependence (SD) is strong for all soil properties except for ESP and OM, for which SD is moderate and weak, respectively. A strong dependence may be attributable to natural factors, such as soil texture and terrain factors, while moderate and weak dependence may be due to other factors, such as inappropriate agricultural practices and agricultural management [70,71]. Figure 7 shows the spatial distribution maps of soil properties; pH varies from 7.06 to 9.28, and soil ECe ranges from low to high soil salinity (0.88–21 dS/m). This difference in soil salinity is a result of the activity of land degradation processes in the Fayoum depression, where inadequate drainage conditions reduce salinity, which is also a common feature in the soils of the North Delta [72,73]. The results show that the study area has low soil OM contents, ranging between 0.4% and 1%. The results indicate that the area is poor in nutrient content, except for some spots in the north of the area that contain reasonable values of soil nutrients; the maximum values are 66, 18, and 860 for Av. N, P, and K, respectively (Figure 7).
Spatial distribution maps of soil properties affecting the SQI in the study area shown in Figure 7.

3.4.2. Mapping the Soil Quality Index

OK interpolation was used to interpolate the spatial variability of soil quality in the study area based on the results of CSQI, which was calculated using Equation (4). The results are shown in Table 5.
The results of the SQI range from 0.88 to 0.37. The SQI is classified into three quality zones according to Aprisal, Bambang, and Harianti [13], as shown in Figure 8 and Table 8. The soil is affected by its composition as well as the surrounding environmental and climatic conditions [74,75,76]; the first zone is characterized by a very good quality index that represents about 14.48% (70.52 × 106 ha) of the total area. The soils of this zone are characterized by adequate values of all soil characteristics. The second zone is characterized by good soil quality: this class covers about 50.77% of the area (247.19 × 106 ha). The third zone is fair (low quality) and covers about 34.7% (169.17 × 106 ha). The soil pH is mild in the first zone and strong in the second and third zones. The status of available N and available P in the second and third zones is low and medium, respectively. Available K is classified as high in the second zone and medium in the third zone.
The organic matter, clay, EC, available N, available P, available K, and CEC are the most effective factors contributing to the SQI in the Fayoum depression [12,13]. The low values of these parameters lead to negative effects on the SQI [9]. The physical indicators (depth, bulk density, porosity, aggregate stability, texture, and compaction) affect the organization of the particles and pores, explaining their impacts on root growth, speed of plant emergence, and water infiltration [9].

4. Conclusions

The precise evaluation of soil quality is a very important issue for precise farming (in particular) and for the proper management of sustainable agricultural practices (in general). This evaluation facilitates the identification of the most suitable crops and the potential agricultural uses of the area. Soil quality is affected by agricultural practices and climatic conditions, which, in turn, affect the physical, chemical, and fertility properties of the soil. In this study, the nutrients and physical and chemical properties of the soil were used to assess the SQI in the El-Fayum depression, in the Western Desert of Egypt. For the purpose of these investigations, PCA analysis was jointly used with GIS to capture, quantify, and map the soil quality index of the study area. The results showed that the PCs of PCA explained 83.6% of the total variance of soil data. In addition, the soil data were classified into three clusters: Cluster I represented about 13.89% of soil samples, Cluster II represented about 16.67%, and Cluster III represented the rest of the soil data, i.e., 69.44% of samples. The use of GIS to map soil properties immediately highlighted the changes and spatial variation in SQI from one place to another. The exponential model was the most suitable for predicting the unknown values of the majority of the soil properties (CEC, ESP, Av.N, Av.P, clay, and silt), followed by the K-Bessel model for pH and OM. The study area was classified into three zones based on the variation in CSQI values. These zones differed in both the number and type of the limiting factors that reduced the soil quality. In particular, zone 1 was characterized by significant improvement in the soil nutrients and chemical properties, whereas zones 2 and 3 were affected by a decrease in the soil’s nutrient contents, in addition to an increase in soil salinity in zone 3. The areas categorized as very good and good quality occupied about 14.48% and 50.77%, respectively, of the total surface investigated, and fair soil quality (mainly due to salinity and low soil nutrients) constituted about 34.75%. As a whole, the results reflect that the joint use of PCA and GIS allows for an accurate and effective assessment of the SQI.

Author Contributions

Conceptualization: M.K.A.-F., E.S.M., E.M.W., S.A.S., A.A.A., M.A.A., and R.L.; introduction: M.K.A.-F., M.A.A., and E.S.M.; methodology: M.K.A.-F., E.M.W., and S.A.S.; results and discussion: M.K.A.-F., E.M.W., S.A.S., M.A.A., and E.S.M.; review and conclusions: E.S.M., E.M.W., S.A.S., A.A.A., M.A.A., and R.L.; funding acquisition: A.A.A. and M.K.A.-F. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the Faculty of Agriculture, Zagazig University, for the laboratory analysis and support. The authors would like to extend thanks to the National Authority for Remote Sensing and Space Sciences for its support. The authors would like to extend their sincere appreciation to the Deanship of Scientific Research at King Saud University, through Research Group Project no. RGP-VPP-275, for support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roy, T. Precision farming: A step towards sustainable, climate-smart agriculture. In Global Climate Change: Resilient and Smart Agriculture; Springer: Singapore, 2020; pp. 199–220. [Google Scholar]
  2. Mohamed, E.S.; Baroudy, A.; El-beshbeshy, T.; Emam, M.; Belal, A.; Elfadaly, A.; Aldosari, A.A.; Ali, A.; Lasaponara, R. Vis-NIR spectroscopy and satellite Landsat-8 OLI data to map soil nutrients in arid conditions: A case study of the northwest coast of Egypt. Remote Sens. 2020, 12, 3716. [Google Scholar] [CrossRef]
  3. Said, M.E.S.; Ali, A.; Borin, M.; Abd-Elmabod, S.K.; Aldosari, A.A.; Khalil, M.; Abdel-Fattah, M.K. On the use of multivariate analysis and land evaluation for potential agricultural development of the northwestern coast of Egypt. Agronomy 2020, 10, 1318. [Google Scholar] [CrossRef]
  4. Baroudy, A.A.E.; Ali, A.; Mohamed, E.S.; Moghanm, F.S.; Shokr, M.S.; Savin, I.; Poddubsky, A.; Ding, Z.; Kheir, A.; Aldosari, A.A. Modeling land suitability for rice crop using remote sensing and soil quality indicators: The case study of the Nile Delta. Sustainability 2020, 12, 9653. [Google Scholar] [CrossRef]
  5. Lal, R.; Stewart, B.A. Principles of Sustainable Soil Management in Agroecosystems; CRC Press; Taylor and Francis: Boca Raton, FL, USA, 2013; p. 568. [Google Scholar]
  6. Doran, J.; Coleman, D.; Bezdicek, D.; Stewart, B. A Framework for Evaluating Physical and Chemical Indicators of Soil Quality; Soil Science Society of America: Madison, WI, USA, 1994; Volume 35, pp. 53–72. [Google Scholar]
  7. Karlen, D.L.; Stott, D.E. A framework for evaluating physical and chemical indicators of soil quality. Defin. Soil Qual. Sustain. Environ. 1994, 35, 53–72. [Google Scholar]
  8. Mukherjee, A.; Lal, R. Comparison of soil quality index using three methods. PLoS ONE 2014, 9, e105981. [Google Scholar] [CrossRef] [Green Version]
  9. Martinez-Salgado, M.; Gutiérrez-Romero, V.; Jannsens, M.; Ortega-Blu, R. Biological soil quality indicators: A review. Curr. Res. Technol. Educ. Top. Appl. Microbiol. Microb. Biotechnol. 2010, 1, 319–328. [Google Scholar]
  10. Bouma, J. Land quality indicators of sustainable land management across scales. Agric. Ecosyst. Environ. 2002, 88, 129–136. [Google Scholar] [CrossRef]
  11. AbdelRahman, M.A.; Shalaby, A.; Mohamed, E. Comparison of two soil quality indices using two methods based on geographic information system. Egypt. J. Remote Sens. Space Sci. 2019, 22, 127–136. [Google Scholar] [CrossRef]
  12. Rangel-Peraza, J.G.; Padilla-Gasca, E.; López-Corrales, R.; Medina, J.R.; Bustos-Terrones, Y.; Amabilis-Sosa, L.E.; Rodríguez-Mata, A.E.; Osuna-Enciso, T. Robust soil quality index for tropical soils influenced by agricultural activities. J. Agric. Chem. Environ. 2017, 6, 199. [Google Scholar] [CrossRef] [Green Version]
  13. Aprisal, I.; Bambang, J.; Harianti, M. Soil quality index analysis under horticultural farming in Sumani upper watershed. Int. J. Geomate 2019, 16, 191–196. [Google Scholar] [CrossRef]
  14. Moore, F.; Sheykhi, V.; Salari, M.; Bagheri, A. Soil quality assessment using GIS-based chemometric approach and pollution indices: Nakhlak mining district, Central Iran. Environ. Monit. Assess. 2016, 188, 214. [Google Scholar] [CrossRef] [PubMed]
  15. Tesfahunegn, G.B. Soil quality assessment strategies for evaluating soil degradation in Northern Ethiopia. Appl. Environ. Soil Sci. 2014, 2014, 1–14. [Google Scholar] [CrossRef] [Green Version]
  16. Mohamed, E.S.; Belal, A.; Shalaby, A. Impacts of soil sealing on potential agriculture in Egypt using remote sensing and GIS techniques. Eurasian Soil Sci. 2015, 48, 1159–1169. [Google Scholar] [CrossRef]
  17. Smith, J.L.; Doran, J.W.; Jones, A.J. Methods for Assessing Soil Quality; SSSA Spec. Pub. 49; Soil Science Society of America: Madison, WI, USA, 1996. [Google Scholar]
  18. Karlen, D.; Mausbach, M.; Doran, J.; Cline, R.; Harris, R.; Schuman, G. Soil quality: Concept, rationale, and research needs. Soil Sci. Soc. Am. J. 1997, 61, 4–10. [Google Scholar] [CrossRef] [Green Version]
  19. Estrada-Herrera, I.R.; Hidalgo-Moreno, C.; Guzmán-Plazola, R.; Almaraz Suárez, J.J.; Navarro-Garza, H.; Etchevers-Barra, J.D. Indicadores de calidad de suelo para evaluar su fertilidad. Agrociencia 2017, 51, 813–831. [Google Scholar]
  20. Burrough, P.A.; McDonnell, R.; McDonnell, R.A.; Lloyd, C.D. Principles of Geographical Information Systems; Oxford University Press: Oxford, UK, 2015. [Google Scholar]
  21. Ali, A.; Soliman, K.; Nasr-Alla, A.; Kamh, R. Evaluation of soil fertility status at West Delta Area using GIS technique. Zagazig J. Agric. Res. 2008, 35. Available online: https://scholar.google.com/scholar?cluster=5574441466090192910&hl=en&oi=scholarr (accessed on 3 February 2021).
  22. Hammam, A.; Mohamed, E. Mapping soil salinity in the East Nile Delta using several methodological approaches of salinity assessment. Egypt. J. Remote Sens. Space Sci. 2018, 23, 125–131. [Google Scholar] [CrossRef]
  23. Hendawy, E.; Belal, A.; Mohamed, E.; Elfadaly, A.; Murgante, B.; Aldosari, A.A.; Lasaponara, R. The prediction and assessment of the impacts of soil sealing on agricultural land in the North Nile Delta (Egypt) using satellite data and GIS modeling. Sustainability 2019, 11, 4662. [Google Scholar] [CrossRef] [Green Version]
  24. Mohamed, E.S.; Saleh, A.M.; Belal, A.B.; Gad, A. Application of near-infrared reflectance for quantitative assessment of soil properties. Egypt. J. Remote Sens. Space Sci. 2018, 21, 1–14. [Google Scholar] [CrossRef]
  25. Doerge, T. Defining management zones for precision farming. Crop Insights 1999, 8, 1–5. [Google Scholar]
  26. Khosla, R.; Shaver, T. Zoning in on nitrogen needs. Colo. State Univ. Agron. Newsl. 2001, 21, 24–26. [Google Scholar]
  27. Csomós, E.; Héberger, K.; Simon-Sarkadi, L. Principal component analysis of biogenic amines and polyphenols in Hungarian wines. J. Agric. Food Chem. 2002, 50, 3768–3774. [Google Scholar] [CrossRef]
  28. Jolliffe, I.T.; Cadima, J. Principal component analysis: A review and recent developments. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2016, 374, 20150202. [Google Scholar] [CrossRef] [PubMed]
  29. Perez, L.V. Principal Component Analysis to Address Multicollinearity. Walla Walla, WA 99362, USA. Walla Walla Whitman Coll. 2017. Available online: https://www.whitman.edu/Documents/Academics/Mathematics/2017/Perez.pdf (accessed on 3 February 2021).
  30. Dandpat, S.K.; Meher, S. Performance improvement for face recognition using PCA and two-dimensional PCA. In Proceedings of the 2013 International Conference on Computer Communication and Informatics, Tamil Nadu, India, 9–11 January 2013; pp. 1–5. [Google Scholar]
  31. Ilin, A.; Raiko, T. Practical approaches to principal component analysis in the presence of missing values. J. Mach. Learn. Res. 2010, 11, 1957–2000. [Google Scholar]
  32. Müller, W.; Nocke, T.; Schumann, H. Enhancing the visualization process with principal component analysis to support the exploration of trends. In Proceedings of the 2006 Asia-Pacific Symposium on Information Visualisation, Tokyo, Japan, 1–3 February 2006; Volume 60, pp. 121–130. [Google Scholar]
  33. Ivosev, G.; Burton, L.; Bonner, R. Dimensionality reduction and visualization in principal component analysis. Anal. Chem. 2008, 80, 4933–4944. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Ali, R.; Kawy, W.A. Land degradation risk assessment of El Fayoum depression, Egypt. Arab. J. Geosci. 2013, 6, 2767–2776. [Google Scholar] [CrossRef]
  35. Abo-El-Ennan, S. Pedogenesis of the Fayoum Area’; Agriculture, Azhar Univ.: Cairoc, Egypt, 1985; p. 156. [Google Scholar]
  36. Mohamed, E.S. Spatial assessment of desertification in north Sinai using modified MEDLAUS model. Arab. J. Geosci. 2013, 6, 4647–4659. [Google Scholar] [CrossRef]
  37. Abd-Elmabod, S.K.; Bakr, N.; Muñoz-Rojas, M.; Pereira, P.; Zhang, Z.; Cerdà, A.; Jordán, A.; Mansour, H.; De la Rosa, D.; Jones, L. Assessment of soil suitability for improvement of soil factors and agricultural management. Sustainability 2019, 11, 1588. [Google Scholar] [CrossRef] [Green Version]
  38. Van Reeuwijk, L.P. Procedures for Soil Analysis. Tech. Pap.-Int. Soil Ref. Inf. Cent. 1986, 9, 106p. Available online: https://pascal-francis.inist.fr/vibad/index.php?action=getRecordDetail&idt=8009739 (accessed on 3 February 2021).
  39. Richards, L.A. Diagnosis and Improvement of Saline and Alkali Soils; U.S. Dept. of Agriculture: Washington, DC, USA, 1954; Volume 78, p. 154.
  40. Baruah, T.; Barthakur, H. A Textbook of Soil Analysis; Vikas Publishing House PVT Ltd.: New Delhi, India, 1997. [Google Scholar]
  41. Cude, C.G. Oregon water quality index a tool for evaluating water quality management effectiveness 1. J. Am. Water Resour. Assoc. (JAWRA) 2001, 37, 125–137. [Google Scholar] [CrossRef]
  42. Walck, C. Handbook on Statistical Distributions for Experimentalists; University of Stockholm: Stockholm, Sweden, 1996; p. 190. [Google Scholar]
  43. Everitt, B.; Landau, S.; Leese, M. Cluster Analysis, 4th ed.; Arnold: London, UK, 2001. [Google Scholar]
  44. Rencher, A.C.; Christensen, W. Methods of Multivariate Analysis; A John Wiley & Sons.: New York, NY, USA, 2002; p. 727. [Google Scholar]
  45. Manly, B.F.; Alberto, J.A.N. Multivariate Statistical Methods: A Primer; CRC Press Group; Taylor and Francis: Boca Raton, FL, USA; London, UK; New York, NY, USA, 2016; p. 252. [Google Scholar]
  46. Wang, X.; Yang, Z.; Jivkov, A.P. Monte Carlo simulations of mesoscale fracture of concrete with random aggregates and pores: A size effect study. Constr. Build. Mater. 2015, 80, 262–272. [Google Scholar] [CrossRef]
  47. Pham, B.T.; Nguyen-Thoi, T.; Ly, H.-B.; Nguyen, M.D.; Al-Ansari, N.; Tran, V.-Q.; Le, T.-T. Extreme learning machine based prediction of soil shear strength: A sensitivity analysis using Monte Carlo simulations and feature backward elimination. Sustainability 2020, 12, 2339. [Google Scholar] [CrossRef] [Green Version]
  48. Chiverton, A.; Hannaford, J.; Holman, I.; Corstanje, R.; Prudhomme, C.; Bloomfield, J.; Hess, T.M. Which catchment characteristics control the temporal dependence structure of daily river flows? Hydrol. Process. 2015, 29, 1353–1369. [Google Scholar] [CrossRef] [Green Version]
  49. Gundogdu, K.S.; Guney, I. Spatial analyses of groundwater levels using universal kriging. J. Earth Syst. Sci. 2007, 116, 49–55. [Google Scholar] [CrossRef] [Green Version]
  50. Cambardella, C.A.; Moorman, T.B.; Novak, J.; Parkin, T.; Karlen, D.; Turco, R.; Konopka, A. Field-scale variability of soil properties in central Iowa soils. Soil Sci. Soc. Am. J. 1994, 58, 1501–1511. [Google Scholar] [CrossRef]
  51. Cafarelli, B.; Castrignanò, A.; De Benedetto, D.; Palumbo, A.D.; Buttafuoco, G. A linear mixed effect (LME) model for soil water content estimation based on geophysical sensing: A comparison of an LME model and kriging with external drift. Environ. Earth Sci. 2015, 73, 1951–1960. [Google Scholar] [CrossRef]
  52. FAO; United Nations Environment Programme; United Nations Educational, Scientific, and Cultural Organization. A Provisional Methodology for Soil Degradation Assessment; FAO: Rome, Italy, 1980. [Google Scholar]
  53. Abrol, I.; Yadav, J.S.P.; Massoud, F. Salt-Affected Soils and Their Management; Food & Agriculture Organization: Rome, Italy, 1988; p. 131. [Google Scholar]
  54. Neina, D. The role of soil pH in plant nutrition and soil remediation. Appl. Environ. Soil Sci. 2019, 2009, 5794869. Available online: https://www.hindawi.com/journals/aess/2019/5794869/ (accessed on 3 February 2021). [CrossRef]
  55. Brady, N.C.; Weil, R.R.; Weil, R.R. The Nature and Properties of Soils; Prentice Hall: Upper Saddle River, NJ, USA, 2008; Volume 13. [Google Scholar]
  56. White, R.E. Principles and Practice of Soil Science: The Soil as a Natural Resource; John Wiley & Sons: Malden, MA, USA, 2013. [Google Scholar]
  57. Zhenghu, D.; Honglang, X. Effects of soil properties on ammonia volatilization. Soil Sci. Plant Nutr. 2000, 46, 845–852. [Google Scholar] [CrossRef]
  58. Curtin, D.; Campbell, C.; Jalil, A. Effects of acidity on mineralization: pH-dependence of organic matter mineralization in weakly acidic soils. Soil Biol. Biochem. 1998, 30, 57–64. [Google Scholar] [CrossRef]
  59. Clay, D.; Chang, J.; Malo, D.; Carlson, C.; Reese, C.; Clay, S.; Ellsbury, M.; Berg, B. Factors influencing spatial variability of soil apparent electrical conductivity. Commun. Soil Sci. Plant Anal. 2001, 32, 2993–3008. [Google Scholar] [CrossRef] [Green Version]
  60. Alam, M.; Mishra, A.; Singh, K.; Singh, S.K.; David, A. Response of sulphur and FYM on soil physico-chemical properties and growth, yield and quality of mustard (Brassica nigra L.). J. Agric. Phys. 2014, 14, 156–160. [Google Scholar]
  61. Fabrizio, A.; Tambone, F.; Genevini, P. Effect of compost application rate on carbon degradation and retention in soils. Waste Manag. 2009, 29, 174–179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Hutcheson, G.D.; Sofroniou, N. The Multivariate Social Scientist: Introductory Statistics Using Generalized Linear Models; Sage: Thousand Oaks, CA, USA, 1999; p. 275. [Google Scholar]
  63. Barrett, K.; Morgan, G. SPSS for Intermediate Statistics; Use and Interpretation; Lawrence Erlbaum Associates Publishers: Mahwah, NJ, USA; London, UK, 2005. [Google Scholar]
  64. Huck, S.W.; Cormier, W.H.; Bounds, W.G. Reading Statistics and Research; Pearson: Boston, MA, USA, 2012; Volume 566. [Google Scholar]
  65. Tabachnick, B.; Fidell, L. Multivariate Regression. Using Multivariate Statistics, 5th ed.; Pearson Education: Boston, MA, USA, 2007; pp. 117–159. [Google Scholar]
  66. Pallant, J. SPSS Survival Manual; McGraw-Hill Education: London, UK, 2013. [Google Scholar]
  67. Kaiser, H.F. The application of electronic computers to factor analysis. Educ. Psychol. Meas. 1960, 20, 141–151. [Google Scholar] [CrossRef]
  68. Mohamed, E.; Abdellatif, M.; Abd-Elmabod, S.K.; Khalil, M. Estimation of surface runoff using NRCS curve number in some areas in northwest coast, Egypt. In Proceedings of the E3S Web of Conferences, Barcelona, Spain, 10–12 February 2020; EDP Sciences. Volume 167, p. 02002. [Google Scholar]
  69. Abd-Elmabod, S.K.; Mansour, H.; Hussein, A.; Mohamed, E.S.; Zhang, Z.; Anaya-Romero, M.; Jordán, A. Influence of irrigation water quantity on the land capability classification. Plant Arch. 2019, 2, 2253–2261. [Google Scholar]
  70. Yasrebi, J.; Saffari, M.; Fathi, H.; Karimian, N.; Moazallahi, M.; Gazni, R. Evaluation and comparison of ordinary kriging and inverse distance weighting methods for prediction of spatial variability of some soil chemical parameters. Res. J. Biol. Sci. 2009, 4, 93–102. [Google Scholar]
  71. Aldabaa, A.; Yousif, I.A.H. Geostatistical approach for land suitability assessment of some desert soils. Egypt. J. Soil Sci. 2020, 60, 195–205. [Google Scholar] [CrossRef]
  72. Abdel-Fattah, M.K.; Abd-Elmabod, S.K.; Aldosari, A.A.; Elrys, A.S.; Mohamed, E.S. Multivariate analysis for assessing irrigation water quality: A case study of the Bahr Mouise Canal, Eastern Nile Delta. Water 2020, 12, 2537. [Google Scholar] [CrossRef]
  73. Hassan, A.; Belal, A.; Hassan, M.; Farag, F.; Mohamed, E. Potential of thermal remote sensing techniques in monitoring waterlogged area based on surface soil moisture retrieval. J. Afr. Earth Sci. 2019, 155, 64–74. [Google Scholar] [CrossRef]
  74. Saleh, A.; Belal, A.; Mohamed, E. Land resources assessment of El-Galaba basin, South Egypt for the potentiality of agriculture expansion using remote sensing and GIS techniques. Egypt. J. Remote Sens. Space Sci. 2015, 18, S19–S30. [Google Scholar] [CrossRef] [Green Version]
  75. Mansour, H.A.; Abd-Elmabod, S.K.; AbdelGawad, S. The impact of sub-surface drip irrigation and different water deficit treatments on the spatial distribution of soil moisture and salinity. Plant. Arch 2019, 2, 384–392. [Google Scholar]
  76. Abd-Elmabod, S.K.; Jordán, A.; Fleskens, L.; Phillips, J.D.; Muñoz-Rojas, M.; van der Ploeg, M.; Anaya-Romero, M.; El-Ashry, S.; de la Rosa, D. Modeling agricultural suitability along soil transects under current conditions and improved scenario of soil factors. In Soil Mapping and Process Modeling for Sustainable Land Use Management; Elsevier: Amsterdam, The Netherlands, 2017; pp. 193–219. [Google Scholar]
Figure 1. Study area and locations of soil samples.
Figure 1. Study area and locations of soil samples.
Sustainability 13 01824 g001
Figure 2. The flowchart of the procedures for the soil quality evaluation.
Figure 2. The flowchart of the procedures for the soil quality evaluation.
Sustainability 13 01824 g002
Figure 3. Scree plot for the different components considered for the principal component analysis with eigenvalues greater Table 4. PCA biplot (biplot shows both PC scores of samples and loadings of variables).
Figure 3. Scree plot for the different components considered for the principal component analysis with eigenvalues greater Table 4. PCA biplot (biplot shows both PC scores of samples and loadings of variables).
Sustainability 13 01824 g003
Figure 4. PCA biplot (biplot shows both PC scores of samples and loadings of variables).
Figure 4. PCA biplot (biplot shows both PC scores of samples and loadings of variables).
Sustainability 13 01824 g004
Figure 5. Correlation coefficients among the different soil indicators with SQI.
Figure 5. Correlation coefficients among the different soil indicators with SQI.
Sustainability 13 01824 g005
Figure 6. Soil Quality Index simulation for the first (a), second (b), and third cluster (c).
Figure 6. Soil Quality Index simulation for the first (a), second (b), and third cluster (c).
Sustainability 13 01824 g006
Figure 7. The spatial distribution maps of the soil parameters in the study area.
Figure 7. The spatial distribution maps of the soil parameters in the study area.
Sustainability 13 01824 g007
Figure 8. The spatial distribution patterns of the SQI.
Figure 8. The spatial distribution patterns of the SQI.
Sustainability 13 01824 g008
Table 1. Soil characteristics of the study area.
Table 1. Soil characteristics of the study area.
PropertyNMinimumMaximumMeanStd. DeviationShapiro–Wilk
pH367.098.657.860.470.06
EC, dS/m360.8720.335.35.05<0.0001
CEC, cmolc/kg soil363.4540.2320.628.970.7
ESP361.8617.139.753.670.23
OM, %360.071.770.690.460.04
N, mg kg−1361.3361.5519.9117.420
P, mg kg−1362.3319.849.54.510.08
K, mg kg−13632.76733.77183.52193.09<0.0001
Silt,368.1944.7626.588.980.615
Clay, %3624.9862.0942.5410.0550.193
Sand, %3612.9855.9430.8812.310.078
Table 2. Correlation coefficients among soil properties.
Table 2. Correlation coefficients among soil properties.
VariablespHECOMCECNPKESPSilt
EC, dS/m−0.72
OM, %−0.360.26
CEC, cmolc/l−0.330.230.66
N, mg kg−1−0.710.590.770.70
P, mg kg−1−0.640.430.780.630.84
K, mg kg−1−0.610.300.690.670.890.82
ESP−0.580.550.240.340.390.400.32
Silt, %0.52−0.47−0.060.15−0.34−0.41−0.33−0.24
Clay, %−0.480.350.410.760.650.520.640.60−0.16
Note: Values in bold are different from 0 with a significance level alpha = 0.05.
Table 3. Kaiser Meyer Olkin measure of sampling adequacy and Bartlett’s sphericity test.
Table 3. Kaiser Meyer Olkin measure of sampling adequacy and Bartlett’s sphericity test.
Kaiser Meyer Olkin Measure of Sampling AdequacyKMO0.70
Bartlett’s sphericity testChi-square (Observed value)334.63
Chi-square (Critical value)61.66
DF45.00
p-value<0.0001
Alpha0.05
Table 4. Summarization of Principal Component Analysis.
Table 4. Summarization of Principal Component Analysis.
PC1PC2PC3
Eigenvalue5.651.681.04
Variability (%)56.4516.7610.41
Cumulative %56.4573.2283.63
pHFactor loadings−0.790.45−0.01
EC, dS/m0.63−0.570.15
OM, %0.740.42−0.27
CEC, cmolc/l0.740.550.24
N, mg kg−10.950.09−0.17
P, mg kg−10.890.09−0.29
K, mg kg−10.870.21−0.24
ESP0.60−0.320.60
Silt, %−0.400.700.40
Clay, %0.760.160.45
pHComponent Score Coefficient Matrix (CSC)−0.140.27−0.01
EC, dS/m0.11−0.340.14
OM, %0.130.25−0.26
CEC, cmolc/l0.130.330.23
N, mg kg−10.170.05−0.16
P, mg kg−10.160.05−0.28
K, mg kg−10.150.12−0.23
ESP0.11−0.190.58
Silt, %−0.070.420.38
Clay, %0.130.090.43
Table 5. Comprehensive soil quality index (CSQI) calculation based on studied soil indicators using PCA.
Table 5. Comprehensive soil quality index (CSQI) calculation based on studied soil indicators using PCA.
Sample NoStandardized z ScoreSQI-PC1SQI-PC2SQI-PC3CSQI 1CSQI 2
pHECOMCECNPKESPSiltClay
11.551.082.320.582.392.292.162.010.640.701.640.73−0.151.030.85
21.640.641.180.931.891.861.790.250.921.331.140.91−0.190.780.78
31.600.600.740.030.700.381.060.380.731.030.400.310.360.320.62
41.300.492.061.812.211.802.850.150.811.231.591.58−0.661.090.86
51.170.400.901.282.051.532.790.011.180.841.200.92−0.480.780.78
61.250.290.870.490.380.280.090.340.500.740.250.450.460.270.61
71.100.561.762.191.522.011.630.860.031.941.591.640.281.200.89
80.630.770.621.930.910.230.290.530.111.810.850.811.240.740.77
90.030.620.241.280.530.010.170.470.251.520.610.271.210.520.70
101.272.970.140.340.690.480.331.211.160.070.52−1.131.300.240.59
111.082.770.400.250.750.030.421.290.900.340.56−0.971.340.290.61
121.132.421.130.210.140.170.581.480.240.030.57−0.460.840.330.63
130.310.190.250.980.630.740.450.580.961.180.580.120.980.450.67
140.200.440.961.380.730.030.560.532.050.660.54−0.191.240.400.66
150.420.371.341.870.950.830.720.051.580.950.820.530.610.620.73
160.420.390.380.760.610.450.420.310.401.750.600.360.890.490.69
170.610.240.510.530.540.240.560.370.360.370.350.310.240.280.61
180.310.351.301.730.931.200.780.050.861.190.950.800.310.710.76
191.250.470.920.120.040.310.311.960.801.360.45−0.081.700.420.66
201.510.500.570.320.020.710.532.150.311.460.560.181.650.520.70
211.700.621.241.910.981.250.712.140.750.861.010.741.410.840.80
220.270.190.960.020.290.230.360.511.370.750.31−0.280.720.200.58
230.570.320.320.370.100.460.430.430.590.370.250.030.430.190.58
240.500.410.910.990.691.270.660.731.030.450.710.190.440.480.68
250.610.800.170.470.480.250.470.212.020.030.16−0.710.820.060.52
260.540.760.350.130.611.120.510.371.550.190.39−0.540.400.170.57
270.500.861.150.991.071.560.760.961.480.460.91−0.080.570.560.71
281.270.880.140.860.540.390.351.681.210.390.41−0.341.590.340.63
291.100.700.110.490.730.900.560.451.830.540.34−0.410.860.220.59
300.180.821.040.570.981.590.700.580.960.720.89−0.020.220.520.70
310.440.700.170.150.660.520.220.530.410.370.36−0.190.400.220.59
320.800.721.130.210.910.030.300.460.640.780.450.120.470.320.63
331.510.871.320.011.030.560.610.540.230.460.520.44−0.100.350.64
340.820.630.810.560.120.250.241.300.081.750.600.321.390.530.70
350.800.660.270.300.310.440.171.480.371.060.46−0.121.320.380.65
360.720.721.020.420.831.020.590.240.450.360.600.30−0.170.370.64
1 Calculated according to standardized z scores; 2 the CSQI, which was computed using standardized z scores, was transformed into a standard normal distribution (which has a mean of zero and a standard deviation of one) using Equation (5).
Table 6. Results of clustering analysis (k-means clustering).
Table 6. Results of clustering analysis (k-means clustering).
ObservationClassDistance to CentroidObservationClassDistance to Centroid
Sample 1120,405Sample 19342,857
Sample 2188,513Sample 20314,292
Sample 32150,358Sample 21342,775
Sample 41117,091Sample 22334,889
Sample 51106,341Sample 23321,216
Sample 6237,283Sample 24329,534
Sample 71119,473Sample 25321,470
Sample 8215,665Sample 26314,196
Sample 9287,178Sample 27350,381
Sample 10347,456Sample 28335,095
Sample 11335,431Sample 29319,728
Sample 12321,836Sample 30337,255
Sample 13319,036Sample 31358,097
Sample 14323,596Sample 32342,278
Sample 15344,994Sample 33320,554
Sample 16322,714Sample 34212,668
Sample 17310,775Sample 35226,036
Sample 18354,312Sample 36319,239
Table 7. Geostatistical analyses and semivariogram parameters of soil properties.
Table 7. Geostatistical analyses and semivariogram parameters of soil properties.
Soil
Attribute
ModelNuggetPartial SillSillNugget/SillMajor RangeSDCMERMSEMSERMSSEASE
pHK-Bessel0.000.250.250.006834.73Strong−0.0250.26−0.071.970.16
ECeTetraspherical0.0028.4328.430.004097.26Strong0.1493.750.030.814.92
CECExponential0.00109.91109.910.0013,286.94Strong0.2057.480.021.136.59
ESPExponential6.097.8413.930.446767.76Moderate0.2613.310.070.943.53
OMK-Bessel0.210.020.240.908069.50Weak0.0040.490.010.990.50
Av. NExponential0.00436.44436.440.008997.19Strong0.41813.390.020.8615.45
Av. PExponential2.1621.1123.270.094117.22Strong0.1234.700.021.024.66
Av. KTetraspherical0.0064,817.4064,817.400.0010,056.06Strong0.814181.60.011.24146.1
ClayExponential0.00113.55113.550.007915.15Strong0.2847.320.030.908.28
SiltExponential0.0098.4198.410.007638.38Strong−0.1036.91−0.020.897.81
SandSpherical0.00171.92171.920.007698.37Strong−0.3237.85−0.020.998.09
Table 8. Statistical analysis of the zonation of soil parameters and quality index in the study area.
Table 8. Statistical analysis of the zonation of soil parameters and quality index in the study area.
IndicatorsFirst ZoneSecond ZoneThird ZonePr > FSig.
ValueRatingValueRatingValueRating
pH7.22 bNormal7.71aHigh8.02 aHigh0.001Yes
EC, dS/m4.46 cMedium6.14 bMedium8.50 aHigh0.244Yes
OM, %1.46 aHigh0.68 bMedium0.54 bLow0.000Yes
CEC, cmolc/kg32.8 aHigh27.49 aHigh16.54 bMedium0.000Yes
N, mg kg−154.96 aLow25.99 bLow11.45 cLow0.000Yes
P, mg kg−118.08 aHigh10.03 bMedium7.67 bMedium0.000Yes
K, mg kg−1616.84 aHigh237.47 bHigh83.9 cMedium0.000Yes
ESP11.93 abLow12.5 aLow8.65 bLow0.020Yes
Silt, %20.28 a 24.62 a 28.31 a 0.161No
Clay, %54.67 a 55.9 a 36.93 b 0.000Yes
SQI0.88 aVery good0.67 bGood0.37 cFair0.000Yes
Area, ha (%)7052 (14.48%) 24,719 (50.77%) 16,919 (34.75%)
Note: The letters are symbols of Duncan test that the means followed by the same letter in each row are not significantly different from one another at a 5% probability level (Duncan Multiple Range Test).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Abdel-Fattah, M.K.; Mohamed, E.S.; Wagdi, E.M.; Shahin, S.A.; Aldosari, A.A.; Lasaponara, R.; Alnaimy, M.A. Quantitative Evaluation of Soil Quality Using Principal Component Analysis: The Case Study of El-Fayoum Depression Egypt. Sustainability 2021, 13, 1824. https://doi.org/10.3390/su13041824

AMA Style

Abdel-Fattah MK, Mohamed ES, Wagdi EM, Shahin SA, Aldosari AA, Lasaponara R, Alnaimy MA. Quantitative Evaluation of Soil Quality Using Principal Component Analysis: The Case Study of El-Fayoum Depression Egypt. Sustainability. 2021; 13(4):1824. https://doi.org/10.3390/su13041824

Chicago/Turabian Style

Abdel-Fattah, Mohamed K., Elsayed Said Mohamed, Enas M. Wagdi, Sahar A. Shahin, Ali A. Aldosari, Rosa Lasaponara, and Manal A. Alnaimy. 2021. "Quantitative Evaluation of Soil Quality Using Principal Component Analysis: The Case Study of El-Fayoum Depression Egypt" Sustainability 13, no. 4: 1824. https://doi.org/10.3390/su13041824

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