Next Article in Journal
Electrochemical Performance of Graphene Oxide/Black Arsenic Phosphorus/Carbon Nanotubes as Anode Material for LIBs
Next Article in Special Issue
Machine Learning Techniques for Evaluating Concrete Strength with Waste Marble Powder
Previous Article in Journal
Computational AI Models for Investigating the Radiation Shielding Potential of High-Density Concrete
Previous Article in Special Issue
A Pragmatic Transfer Learning Approach for Oxygen Vacancy Formation Energies in Oxidic Ceramics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Strength Predictive Modelling of Soils Treated with Calcium-Based Additives Blended with Eco-Friendly Pozzolans—A Machine Learning Approach

1
Faculty of Environment and Technology, Department of Engineering, Design and Mathematics, Civil Engineering Cluster, University of the West of England, Bristol BS16 1QY, UK
2
Faculty of Environment and Technology, Centre for Architecture and Built Environment Research, University of the West of England, Bristol BS16 1QY, UK
*
Author to whom correspondence should be addressed.
Materials 2022, 15(13), 4575; https://doi.org/10.3390/ma15134575
Submission received: 1 June 2022 / Revised: 21 June 2022 / Accepted: 23 June 2022 / Published: 29 June 2022
(This article belongs to the Special Issue Functional Materials, Machine Learning, and Optimization)

Abstract

:
The unconfined compressive strength (UCS) of a stabilised soil is a major mechanical parameter in understanding and developing geomechanical models, and it can be estimated directly by either lab testing of retrieved core samples or remoulded samples. However, due to the effort, high cost and time associated with these methods, there is a need to develop a new technique for predicting UCS values in real time. An artificial intelligence paradigm of machine learning (ML) using the gradient boosting (GB) technique is applied in this study to model the unconfined compressive strength of soils stabilised by cementitious additive-enriched agro-based pozzolans. Both ML regression and multinomial classification of the UCS of the stabilised mix are investigated. Rigorous sensitivity-driven diagnostic testing is also performed to validate and provide an understanding of the intricacies of the decisions made by the algorithm. Results indicate that the well-tuned and optimised GB algorithm has a very high capacity to distinguish between positive and negative UCS categories (‘firm’, ‘very stiff’ and ‘hard’). An overall accuracy of 0.920, weighted recall rates and precision scores of 0.920 and 0.938, respectively, were produced by the GB model. Multiclass prediction in this regard shows that only 12.5% of misclassified instances was achieved. When applied to a regression problem, a coefficient of determination of approximately 0.900 and a mean error of about 0.335 were obtained, thus lending further credence to the high performance of the GB algorithm used. Finally, among the eight input features utilised as independent variables, the additives seemed to exhibit the strongest influence on the ML predictive modelling.

1. Introduction

Ground improvement using the technique of shallow or deep soil mixing has received much interest and acceptance in recent years largely due to its extensive applications in construction projects. In the UK, EU and US where the uptake and implementation of the technology has increased exponentially over the past three decades, environmental policies and laws, taxes, landfill directives and the ever-increasing cost of excavating and moving poor soils has made this method of ground stabilisation even more imminent [1,2,3,4].
Cementitious materials, such as cement and lime, have been used traditionally over the past 50 decades as hydraulic binders to stabilise poor soils. However, the attendant negative environmental impacts associated with the production of these energy intensive binders are a present global concern. Hence, based on current developments in knowledge and research, attention is gradually shifting from an over-reliance on solely cement and lime to the utilisation of waste materials, industrial and agricultural by-products, organics, etc., in soil stabilisation [5,6].
Agro-based environmentally friendly pozzolanic materials, such as rice husk ash, palm oil fuel ash, bagasse ash, coconut shell ash, coconut husk ash, corn cob ash, almond shell ash, etc., have gained considerable attention in soil stabilisation given the ever-growing costs of their disposal [7,8,9,10,11,12,13,14,15,16,17,18]. The major chemical composition of these plant-based pozzolans are alumino-silicates [19,20,21]. Moreover, in order to achieve the desired effect on the mechanical properties of soils, most applications in soil stabilisation have tended towards the partial substitution of calcium-based agents by agro-based pozzolans. The soil-binder mix in this regard can speed up the rate of development of the calcium alumino-silicate hydrated gel (CAH or CASH) as well as the sodium alumino-silicate hydrated gel (NASH) [22]. These binding gels will develop inside the soil voids, and aid in the formation of a more compact soil-binder mix and thus enable a further improvement in the strength of the stabilised soil. Approximately 50–80% reduction in the quantity of calcium-oxide-based additives as a result of the addition of agro-based pozzolanic materials has been reported [23]. Important literature surveys reflecting the composite mix of calcium-based and agro-based agents used in soil stabilisation were carried out recently [15,24,25]. Table 1 is an indication of some of the binders used in the recent past and the target strength properties considered in the improvement.
In general, the determination of the strength properties of soils stabilised by using a composite binder mixture is often a crucial first step towards establishing the correct design mix guideline for field application [26,27,28]. For soils stabilised by using multiple binder combinations, the challenges of establishing a parameter, such as the UCS, may require laborious laboratory experimentation and time-consuming trial batching of binder type, quantities, optimal combinations, choice of curing duration, and the determination of other influencing factors.
Moreover, conventional techniques of predicting or modelling the UCS of stabilised soils do rely essentially on relationships that are developed empirically from statistical concepts, employing mostly linear, and occasionally nonlinear regression methods [29,30,31]. The equations generated analytically from these methods do tend to determine several unknown coefficients that may affect relationships between the dependent and independent features or variables. Hence, the resulting models, although effective in certain situations, are inherently riddled with shortcomings due mainly to the complexities of the stabilised soil mix.
In recent times, artificial intelligence paradigms relying on several machine learning (ML) techniques have begun to gain traction as alternatives for the determination of the UCS of soils [32,33,34]. That notwithstanding, the adoption of ML methods for the predictive modelling of improved ground properties has only been reported in a few studies [32,35,36,37,38,39,40,41,42]. Moreover, an application that involves predictive modelling of compressive strength of soils stabilised by eco-friendly pozzolans enriched by cementitious additives has not been reported.
In this study, the gradient boosting (GB) machine learning technique is utilised for the predictive modelling of compressive strength of soils stabilised by cementitious additive-enriched eco-friendly pozzolans. This research shall take into account both regression and multinomial classification of the compressive strength of the stabilised soils. Rigorous sensitivity-driven diagnostic testing to validate the algorithm used and the corresponding statistical outcomes are also undertaken. Finally, it is recommended that an implementation of the concepts derived from this study be applied during the preliminary stages of soil stabilisation for civil construction and related ground improvement applications.
Table 1. Studies on soil stabilisation using agro-based and cementitious additive blends.
Table 1. Studies on soil stabilisation using agro-based and cementitious additive blends.
Pozzolanic AshCementitious AdditiveTarget StrengthReference
Rice husk ashLimeCBR[43]
Bagasse ashLimeUCS, CBR[44]
Rice husk ashCement, LimeUCS[45]
Rice husk ashCement, LimeUCS, Shear[23]
Bagasse ashLimeCBR[46]
Rice husk ashCementUCS, CBR[47]
Palm oil fuel ashLimeUCS[21]
Rice husk ashLimeCBR, shear[48]
Bagasse ashCalcium carbideUCS, shear[49]
Rice husk ashCementUCS, tensile strength, flexural strength [50]
Rice husk ashCement CBR[51]
Sawdust ashLimeCBR[17]
Coconut shell ash, coconut husk ashLimeUCS, CBR[52]
Palm oil fuel ashHigh calcium pulverized fuel ashUCS, CBR[53]
Rice husk ashCementUCS, tensile strength[54]
Palm oil fuel ash, rice husk ashCalcium carbideUCS, shear[55]
Rice husk ashLimeUCS, shear, tensile strength, CBR[56]
Rice husk ashLimeUCS, CBR[7]
Rice husk ashLimeUCS, CBR[57]
Bagasse ashCement UCS, CBR[58]
Rice husk ashLimeUCS, shear, CBR[59]
Palm oil fuel ashCementShear [60]
Rice husk ashCement Shear, CBR[11]
Rice husk ashLime UCS[61]
Rice husk ashLimeCBR[62]
Bagasse ashLimeUCS, CBR[63]
Rice husk ashLime, calcium chlorideUCS, CBR[64]
Corn cob ashCalcium carbideUCS[65]
Bagasse ashCement UCS[66]
Almond shell ashLimeUCS[67]
Plant ashCement, calcium chlorideUCS[14]
Rice husk ashCement, limeUCS[35]

2. Methodology

2.1. Database Development, Pre-Processing, and Exploratory Analysis

A dataset of 392 soils stabilised using cementitious additives’-enriched agro-based pozzolans in various proportions and combinations, compacted and cured for 7, 14 and 28 days were compiled from a very intensive literature search [7,17,21,22,35,44,47,49,50,53,54,55,56,63,67]. As stated previously, most agro-based pozzolanic materials are composed mainly of alumino-silicates. The range of some of the main chemical compositions of the pozzolans and those of the cementitious materials (containing mostly of calcium-oxide compounds) utilised to stabilise the soils are given in Table 2. A very broad range of agro-based pozzolans (rice husk ash, palm oil fuel ash, bagasse ash, coconut shell ash, coconut husk ash, corn cob ash, and almond shell ash) were used to stabilise the soils. On the other hand, the cementitious materials consist of cement, lime, cement kiln dust, high calcium fly ash and calcium carbide. As could be observed in Table 2, X-ray fluorescence (XRF) measurements conducted on the binding agents used in this study indicates the maximum proportion of alumino-silicates in the agro-based pozzolans as being about 93%, while that of cementitious additive is about 25%. On the other hand, the highest amount of calcium oxide compound realized in the pozzolanic ashes is about 14% compared to 95% in the cementitious binders. It seems from Table 2 that the proportions of both agents that can be used to stabilize the soil could be an important trade-off between their innate chemical compositions, the target compressive strength to be achieved and the impact of their usage on the environment. It is also pertinent to bear in mind that the clay soils to be stabilized are themselves mostly siliceous as indicated by their chemical compositions in Table 2; hence, they should be taken into account in deriving a suitable design mix. Standard preparation methods including those involving slight modifications of traditional or standardized measurement procedures carried out to reflect special laboratory testing conditions were followed to achieve the aims of stabilisation. Since the nature of the dataset of UCS are diverse in this regard, it was necessary to normalise these data in order to enhance the significance of the overall modelling and the reliability of the results of findings. A two-step inverse-normal data transformation approach was applied on the dataset of UCS regarded in this study as the target variable [68]. As could be observed from Figure 1 and Table 3, normally distributed data and relatively lower values of kurtosis (−0.16) and skewness (1.63 × 10−6) suggest that the dataset can be reliable for ML modelling.
A total of 8 independent variables are used as input features in the ML modelling namely, values of agro-based pozzolans, cementitious additives, soil class, liquid limit, plasticity index, plastic limit, curing duration and strength class. Exploratory dataset analysis carried out on these variables yields the statistical metrics and distributions shown in Table 3 and Figure 2, respectively.
Given that the independent variable dataset will be used in its raw form for the ML modelling, it is very refreshing to note how reasonably low their kurtosis and skewness scores are as observed in Table 3. It is very necessary to use raw independent data of the variables in the predictive modelling to preserve as well as ensure an accurate representation of the natural reality of random distributions. As Table 3 shows, the range and proportion (calculated by weight of dry soil) of the agro-based pozzolans (lowest of 0.1% and highest of 25%) and the cementitious materials (lowest of 0.5% and highest of 11.25%) used, demonstrate a very diverse mix of binder quantities that are used to stabilise the soils. Except for the soil plasticity properties, Figure 2a,b shows a fairly uniform frequency distribution for the binders used. The distribution of the soil class is observed as being skewed to favour mostly soils of lower plasticity. However, there is almost a satisfying balance between the soil classes if both high plasticity classes (CH and MH) are considered together compared to the lower plasticity class (CL). The frequency distribution of curing duration (Figure 2g) seems fair, except for the UCS strength classes, which, according to Figure 2h, is imbalanced and skewed towards the hardened stabilised soils. Thus, the frequency distribution for the stabilised soil’s compressive strength class (or consistency) in this regard indicates that most of the stabilised soils in the dataset were greater than approximately 400 kPa, as Table 4 shows. The UCS class shall serve as a target (or dependent) variable in the multinomial classification ML prediction, while, as already stated above, the actual numerical values of the UCS shall be used as dependent variables in the ML regression modelling considered subsequently in this research.

2.2. Gradient Boosting Machine (GBM)

Boosting is generally an ensemble machine learning technique that involves an aggregation of based learners to enable better predictions of mostly classification and regression problems. Gradient boosting (GB) machine works by optimising a differentiable loss function (an example is the ‘squared error’ for regression and ‘logarithmic’ for classification) as well as an additive modelling that involves taking a weighted sum of several suitable base learners in order to minimise the loss function. In its simplest from, the GBM as an additive model can be represented mathematically as
F m ( X ) = F m 1 ( X ) + η · f m ( X )
where F = ensemble model, f = base(weak) learner, η = rate of learning or shrinkage and X is the input vector.
Fm (X) is the result of each iteration obtained by minimising a loss function and therefore can be considered as a directional vector (rm−1), which points to the steepest decent. Hence, the GB machine can then be expressed alternatively as
F m ( X ) = F m 1 ( X ) + η · r m 1
If the function that is being approximated is given as
F ( x ) = f 1 ( x ) + f 2 ( x ) + f 3 ( x ) + + f n
If there are n number of samples in a dataset (D), with each sample having m set of features in a vector x and a real target or dependent value of y all expressed as
D = x i , y i ( | D | = n , x i   m , y i     )
Then, an ensemble of trees considering additive modelling can be given as below:
y i = ϕ ( x i ) = m = 1 M f m ( x i ) , f m    
where M = number of base leaners and = regression tree space.
Additionally, if the differentiable loss function is given as
( ϕ ) = i l ( y ι ^ ,   y i )
then the first step in initialising the model with a constant value by minimizing the loss function becomes
F 0 ( x ) = arg min ρ i = 1 n L ( y i , b 0 )
where b0 = minimisation of loss function at 0th iteration.
For m = 1 to M, the following is computed for all the n samples for i = 1, …, n
r i m = [ L ( y i , F m 1 ( x i ) ) F m 1 ( x i ) ]
Next, we fit a regression (or classification tree) to rim, allowing each tree to be denoted by Rjm for j = 1, …, Jm, where Jm is the number of leaves in the trees created in the mth iteration.
For j = 1, …, Jm the following is then computed:
ρ m = arg min ρ i = 1 n L ( y i , F m 1 ( x ) + ρ m j = 1 J b j m
where bjm = least square coefficient or the basis function, ρm = leaf weight or scaling factor
The equation then simplifies to
ρ j m = arg min ρ i = 1 n L ( y i , F m 1 ( x ) + ρ m · b m ) ,
With the update given as
F m ( x ) = F m 1 ( x ) + η · ρ m · j = 1 J m b j m
It is important to note that the parameters or hyperparameters required (among other factors) will have to be carefully selected in order to obtain an optimised or desirable results of the ML prediction. The following section shall describe the methods adopted to optimise the GB machine to ensure higher performance on multinomial classification and regression.

2.3. Model Optimisation

2.3.1. Hyperparameter Tuning

In order to ensure the best performance of the GB model, a series of stepwise-randomized searching were implemented to select the best performing hyperparameters using python’s sklearn searching class type called ‘‘RandomizedSearchCV”. Table 5 shows the hyperparameters eventually chosen to optimise the algorithm on both the training and testing datasets.

2.3.2. Cross-Validation

The k-fold cross-validation technique was applied to enhance learning and validation on 80% of the dataset. Cross validation also ensures that undue overfitting of the algorithm on the training set was prevented. After several trials, 10-fold cross validation was regarded as the most effective in the ML modelling. It is important to note that a further 20% of the data was set aside for model testing.

2.4. ML Performance Evaluation Metrics

For an assessment of the performance of the ML model on the multinomial classification problem, accuracy, precision, recall and F1 score were used. Additionally, in order to depict the capacity of the model to predict the probability of the compressive strength of the stabilised soils belonging to different categories across a specified decision threshold, the receiver operating characteristic curve (ROC) and corresponding area under curve (AUC) were used. ROC is a plot of true positive rate (TPR) or sensitivity versus false positive rate (FPR) (or one less specificity) under some threshold values hence, separating “noise” from “signals”. AUC is a measure of the actual ability of a model to distinguish between class labels. For the regression problem, coefficient of determination (R2) and mean absolute percentage error (MAPE) metrics shall be adopted.

3. Results and Discussion

3.1. Behaviour of the Stabilised Soils

The UCS of soils stabilised using cementitious additives-enriched pozzolanic materials and compacted at optimal conditions (dry density (DD) and moisture content (MC)) as shown in Figure 3. The stabilised materials were thoroughly synthesized and slightly modified from their original sources with the horizontal axis showing a combined total of the binders (agro-based pozzolans and cementitious additives) used (by weight of the dry soil). Generally, a rising trend is observed as represented by the shaded bands signifying increments up to 28 days of curing. The increase in strength corresponds to an increase in the proportion of the binders with the curing duration. However, a few slight deviations from this trend are also observed, where there seems to be a threshold signifying the highest rise in UCS corresponding to the optimal binder mixes beyond which there is a minimal decrease in UCS as the total binder proportion increases (Figure 3a,c,g,k,l,p). 28-day strength development of twice and slightly more than doubled those observed at 7 days of curing are noticed in Figure 3b,d,h. As indicated in Figure 3b, 100% strength increase at 28 days, up from 7 days of curing, is also realised when the binders with combined proportion of 25% are used to stabilise the soil. On the other hand, the strength increases of about 23% (from 7 to 28 days of curing), considered the least of all, are indicated in Figure 3k,o.
In general, the total amount of agro-based pozzolans combined with calcium-based additives required to stabilise the soil to produce the highest strength is approximately 15% by weight of the dry soil. Moreover, it has been reported that the proportion of agro-based pozzolanic ashes should be more than calcium oxide-rich additives in the stabilised soil in order to achieve the desired aim of reducing carbon footprinting [51,54,69]. It should also be noted that although the behaviour of the stabilised soil in terms of strength increase is mostly due to the proportion of binders used and curing duration, the methods of preparation of the stabilised soil, curing conditions, method of compaction, soil type, laboratory instrumentation, etc., can also contribute to this trend. It is suggested that these secondary factors be further investigated and modelled to include environmental factors (temperature, freezing and thawing, wetting and drying, etc.), which could also have potential effect in terms of the durability of a stabilised soil.

3.2. ML Classification of UCS

Three compressive strength categories or classes (‘firm’, ‘very stiff’ and ‘hard’) are obvious from the dataset as earlier indicated by their statistical distribution in Figure 2h. Hence, an application of ML multiclass classification was used to gauge the performance of the GB ML algorithm by investigating its capacity to properly learn the stabilised soil strength patterns and therefore provide predictions accordingly. In most cases, it is very essential that for multinomial classification problems, a threshold point be set for the classifier’s boundary across the different categories. Hence, a sensitivity analysis was performed by using the ROC (receiver operating characteristics curve) and resulting area under curve (AUC) as relevant tools for the assessment of the capacity of a ML algorithm to identify or decern the true positives among the class labels.
As shown in Figure 4, the GB algorithm possesses a very high capacity to identify the positive and negative UCS consistency classes. Table 6 also shows very high values of the multinomial classification metrics resulting from the ML prediction. With a ‘non weighted’ accuracy of 0.920, it can then be inferred that the ability of the GB ML algorithm in the multiclass prediction to ensure that both type 1 and type 2 prediction errors have been kept at the barest minimum has been achieved. Table 6 also shows that the assignment of weighted average to take care of under-represented class instances does clearly improve the individual multiclass performance with ‘weighted’ accuracy of 0.966 compared to the overall accuracy metric of 0.920. Further, it is important to note that the averaging that considers each class instances (i.e., micro-averaging) gives higher scores when compared to the averaging, whereby all equal class instances (i.e., macro-averaging) are taken into account with respect to the most frequently occurring class labels. Additionally, by comparing other metrics, such as the recall rates and precision scores, it is observed in Table 6 that using micro-averaging in this regard does provide relatively higher sensitivity and thus better performance.
The number of misclassified instances indicated in the normalised confusion matrix could have been due to an imbalance of the strength categories or classes. However, it is very important to note that in order to ensure an overfitting of the data by the algorithm was prevented, a 10-fold cross validation was used. The distribution of the strength categories as earlier indicated in Figure 5 indicated that most of the stabilised soils in the dataset were classed as being ‘hard’ (i.e., >400 kPa). Hence the prediction has unsurprisingly shown slight biases towards the stabilised soils’ ‘hard’, ‘very stiff’ and ‘firm’ categories in that order. Based on the nature of dataset used in this study, this behaviour of the ML model is also partly a reaffirmation of the effectiveness of the eco-friendly pozzolanas combined with the cementitious binders in stabilising the soft soils. A total of 12.5% misclassified instances was achieved.
Another means of validating a classifier’s performance is by using the cumulative gains (CG) and its corresponding lift curve (LC). By relying on the CG or LC, further justification can be provided to show the classifier’s effectiveness in the ML prediction especially when compared to another model exhibiting an act of random guessing between multinomial categories. LC is extrapolated from the CG and represents the cumulative gains between a model that also includes the baseline lifting portion (or horizontal percentile axis) of the curve. The lift is viewed mathematically as the ratio of 1 s on a certain sampled data point, to that of 1 s on the entire dataset as given in the equation below. This then can be seen as the predictions that a random model would be making.
lift = prediction   rate average   rate = ratio   of   1   s   in   a   particular   samples   dataset ratio   of   1   s   in   the   whole   dataset
Figure 6 shows the CG and LC of the multiclass classification. A high lift indicates a high performing algorithm. The lift indicated by micro-averaging indicates the highest rising from the baseline percentile axis up to the value of about 5 as seen in Figure 6. Although LC could also be used to compare two or more models, the performance of GB in this regard can be viewed as very reasonable. Hence, if 5% of the sampled data is considered, it therefore means that there are 5 times more positive class labels for each of the classes than an average. Put differently, for 5% of the dataset sampled, about 50% of the entire dataset would contain all the positive class labels.
It is pertinent to note that the performance of the ML model as indicated by the LC resulting from micro-averaging is generally higher than those of macro- and weighted lift scores. This behaviour confirms further, the multiclass ML prediction outcomes given by the ROC and CG curves.

3.3. ML Regression of UCS

GB demonstrates a very high capacity to learn the complexities of the soil stabilised using the different additives and their various combinations as already seen from the multiclass classification above. As can be observed from Figure 7, both the coefficient of determination for the train and test sets are equal at 0.900. Further evidence of this high performance is indicated by the very low percentage error of both the training and testing datasets being 0.213 and 0.335, respectively. Tabarsa et al. [35], in their studies utilising both artificial neural networks and support vector machine algorithms, obtained a coefficient of determination of about 0.990, albeit on a smaller number of dataset (137 datapoints). Unlike the present study, they did not utilise the technique of cross validation, and hence, overfitting of the dataset could not be ascertained. Additionally, Tabarsa et al. [35] only considered a single agro-based ash (rice husk ash) and two cementitious binders (cement and lime) in the stabilised soil mix, whereas the present study utilised a multitude of the agro-based and calcium-oxide based agents in the soil stabilisation. While this could mean that the stabilised soils’ model in this study is more non-linear, it is also important to add that the eight input variables utilised in the model may have also impacted the non-linearity of the model compared to six such input features used in the research conducted by Tabarsa et al. [35].
It is important to emphasise that cross-validation was performed on the training dataset used in this study in order to prevent undue overfitting of the model. Further sensitivity analyses to justify the capability of the GB algorithm used are discussed in the sections following.

3.4. ML Model Diagnostics

3.4.1. GB Model Residuals

Residual plots of ML predictions can provide an insight into the behaviour of a model while also offering the means of validating such a model. By utilising the residual plots, a ML model’s predicted or observed errors can be assessed and evaluated in order to judge its consistency with that of its corresponding stochastic errors (i.e., its measure of unpredictability or randomness). Figure 8 indicates the level of independence of the residuals for both the training and the testing sets when applying GB ML model. As could be observed, ML predictions on the dataset have demonstrated very high independence of the residuals or stochastic prediction errors, given that the training and testing dataset are relatively very close and concentrated around the zero-line including a very random distribution of the data points. Almost similar trends of very high symmetry about the origin are exhibited by both training and testing sets of data thus further confirming their matching R2 metric score as indicated above followed by a minute difference in the mean absolute percentage error metric. That notwithstanding, as could be observed from the residual plot of Figure 8, a few data points seem to be far away from the zero-line, which also confirms that the residuals may not be entirely independent.

3.4.2. Distribution of Residuals

The normal frequency distribution plot of the residuals is another means for which a ML algorithm’s authenticity and effectiveness can be assessed. Figure 9 shows the cumulative distributions on histogram plots of the GB regression on training and testing sets. Much like the residual plots given above, there is a higher symmetry with very balanced distributions of the GB model predictions about the centre, although the residuals of the training sets are slightly higher, given they contain most of the data points compared to the testing set. Furthermore, a very high performing model will tend to possess residuals with peaks at the centre or the origin in some cases but with only marginal stochastic errors at the seams, whereas an inaccurate model with very low predictive performance will tend to have much higher errors at the seams and fewer peaks at the centre of the distribution. Based on this concept, it could be observed that although GB prediction on the training set does possess a higher residual peak given the greater quantity of the dataset used, both it and the predictions on the testing set tend to have very low residuals at the extremes of the distribution, hence indicating further the high level of accuracy provided by the model in the ML prediction.
Overall, when the algorithm is applied on the training dataset set, the total of residual data points is about 0.04% of each of the sum of the predicted and actual data (Figure 9b). Meanwhile, ML learning regression using the testing set would produce a total of approximately 1.60% of the sum of the predicted and actual data (Figure 9d). Although the performance of the ML algorithm on both the training and testing set are equal when considering the regression coefficient of 0.900 for both sets of data, further diagnostics has revealed that the errors (although very marginal) inherent in the predictive model are slightly higher on the testing dataset than on the training set. This is also a confirmation of their different mean percentage error scores as observed previously.

3.5. ML Model Interpretation

3.5.1. ML Feature Importance

Indicators of the importance of input features or independent variables in ML prediction can provide an insight into the dataset utilised in the stabilised soil modelling, while also enabling in some instances an improvement of the effectiveness of the adopted algorithm in the preliminary stages of the model selection. The relative usefulness and significance of the input features used in the modelling or forecast of target variables are shown in Figure 10 for both the multiclass classification and ML regression.
For the classification problem, the feature importance of the input variables aggregated over the three target categories (strength classes) shows the cementitious additives as having the greatest influence on the prediction of the strength class. Although the importance of the soil class is greater than that of the pozzolanic materials, it could be concluded generally that the binders used in the soil stabilisation do possess the highest effect on the prediction of the strength class within the bounds of dataset used in this study. When considering the performance of the features in the prediction of each of the different class categories, it is observed from Figure 10 that the binders again do possess the highest significance in the ML classification. In the same vein, the binders are also observed as consistently having an important bearing on the class categories as noted by the similar levels of feature importance values for each of pozzolans and calcium-oxide-rich additives across the three class categories. It is interesting to observe that although the soil class has the second highest importance after calcium-oxide-rich additives, it does not seem to consistently contribute to the performance or multiclass prediction of the strength categories. Again, Figure 10 does also indicate how the soil plasticity features are lower in the ranking of importance in the ML classification.
Almost a similar trend in terms of feature importance in ML classification is demonstrated by the input features when considering ML regression on the actual compressive strength values of the stabilised soils. It is clearly seen from Figure 11 that here, the binders do possess much stronger influence on the prediction. Interestingly, the soil class seems to have the least importance in the prediction of the target variable. It is also very refreshing to note how the soil plasticity properties used as input features in the ML forecast have a stronger bearing on the regression forecast when compared to the ML classification. It is important to stress that although curing duration is an important consideration in determining the strength gain in soil stabilisation, its contribution as an input feature is judged, but the GB model is only fair both in the classification and regression problems. This could be seen as positive at least within the context of the dataset used in this study. This does suggest that the GB ML algorithm can show high performance by determining either what strength class the stabilised soil belongs or what compressive strength it possesses, irrespective of the knowledge of the age of curing.

3.5.2. Partial Dependence and Interaction Plots

A partial dependence plot (PDP) and an individual conditional expectation (ICE) plot could be utilised to analyse and visualise the interaction that exit between the target response variable and a set of explanatory or input variables of interest. PDP indicates the dependence (taken as an average) between the target response variable and a set of independent variables of interest, marginalizing over the average values of the other input variables (or the ‘complementary’ features). Meanwhile, an ICE allows a visualisation of the dependence of a prediction on a variable for each sample when considered separately with one line per sample.
Figure 12 indicates one-way PDP plots with ICE lines between the explanatory features and the response variable. As could be observed, the strongest non-linear relationship seems to exist mostly between the binders (cementitious additive and agro-based pozzolans) and the target or response variable of UCS, although there are some exceptions, where the response variable remains almost constant in some ranges of the average amount of cementitious additive and agro-based pozzolans. This indicates a dependence of the predicted variable on the binders. However, there appears to be almost no relationship between the duration and the strength class with the target variable as indicated by the ICE lines being constant throughout in most ranges of the median of the explanatory features.
A much better visualisation of the interaction between each of the two most influential and less influential features in the prediction is shown in Figure 13. The two-way interaction PDP plots indicate the dependence of the average of the target variable (UCS) on the joint values of the most important and the less influential variables. When considered together, it can be clearly seen from Figure 13a–d that there is a strong interaction between the two most important features: for an average agro-based pozzolan of between 7 and 15%, the UCS is nearly independent of the cementitious additive used (indicated by the nearly horizontal contours of Figure 13c), whereas for median values between 0 and 7% (indicated by vertical contours), there is a strong dependence on the cementitious additives. Although there are pockets of exceptions to these relationships as observed from Figure 13c, it could be concluded generally that the quantity of agro-based binder required in combination with a calcium-based cementitious agent for high stabilised soil performance in terms of unconfined compressive strength should be limited to about 15% but not less than 7% calculated by weight of the dry soil. With regards to the soil plasticity properties, for an average plasticity index (PI) of less than 40%, the UCS is nearly independent of the plastic limit (PL) (Figure 14a–d). Finally, there seems to be almost no dependence of UCS on the soil class for all average values of the curing duration (Figure 15a–d).

3.6. Comparison to Other Tree-Based ML Ensembles

A confirmation of the performance of the GB ML algorithm used in this study when compared to other commonly used tree-based ensemble ML techniques is given in Table 7. GB does clearly outperform its counterparts and base estimators as indicated by the statistical metrics. Granted, these algorithms are characterised by slightly different parameters; however, these were optimised and used for the training and testing. The order of performance and ranking of the algorithms on the stabilised soil dataset is Gradient Boost > Adaboost > Extremely Randomised Trees > Random Forest > Decision Tree.

4. Study Significance

The importance of machine learning to civil and environmental engineering design and construction especially in ground improvement works including but not limited to road subgrades, building foundations, embankments and cut slopes, bridge abutments, exclusion barriers, liquefaction mitigation, backfills, contaminated ground remediation, etc. cannot be over- emphasized. The concept of artificial intelligence as applied in this study can save time, cost and money during the planning and design stages of ground improvement. Examples of some of the preliminary exercises that can be circumvented are laborious laboratory experimentation and time-consuming trial batching of binder type, quantities, optimum combinations, choice of curing duration, and the determination of other influencing factors. In order to practically implement the model studied herein, all its resources, including background scripting, would have to be deployed and persisted on an organisation’s server to be used to train and test on new data of stabilised soil’s unconfined compressive strength.
However, it should be borne in mind that only a few of the plethora of factors that can directly or indirectly influence soil stabilisation were considered. Hence, it is suggested that the principles applied in this research be extended to include modelling of not just soil strength behaviours but also other important serviceability design parameters that involve settlement and soil swelling or expansion. Moreover, predictive modelling using artificial intelligence is recommended for stabilised soils subjected to different curing and environmental durability conditions in future studies. It is also suggested that various other environmentally friendly materials be utilised to stabilise the soils and predictions made using machine learning models.

5. Conclusions

An artificial intelligent approach using machine learning for predictive modelling of soil stabilised by eco-friendly pozzolans rich in cementitious additives was studied. Sensitivity-driven diagnostic testing was also performed to validate the algorithm used along with the corresponding statistical outcomes. Highlights and the summary of this study are as follows:
  • The combined total quantity of agro-based pozzolans and calcium oxide-rich required to stabilise a soil to produce the highest strength is approximately 15% (by weight of the dry soil) but with the pozzolan having the greater amount in the binder mix.
  • GB algorithm shows a very high capacity to distinguish between positive and negative unconfined compressive strength categories (‘firm’, ‘very stiff’ and ‘hard’) with an overall accuracy of 0.920, weighted scores of 0.938 and 0.920 for precision and recall rates respectively and an overall lift of 5 in a multinomial classification.
  • Although ML multiclass prediction was slightly biases towards the stabilised soils’ ‘hard’, ‘very stiff’ and ‘firm’ categories in that order, only 12.5% of misclassified instances was achieved by the algorithm.
  • A coefficient of determination of approximately 0.900 and a mean absolute percentage error of 0.335 lends further evidence to the high performance of the GB algorithm when applied to a regression problem.
  • Rigorous diagnostic tests performed on the GB algorithm revealed only marginal inherent stochastic errors in the prediction with the errors only being slightly higher on the testing set when compared to the training set.
  • Overall, among the 8 input features used as independent variables, the binders (both agro-based pozzolans and the cementitious additives) tend to exhibit the strongest influence on the ML predictive modelling in the multinomial classification and regression modelling.

Author Contributions

Conceptualisation, E.U.E.; methodology, E.U.E.; software, E.U.E.; validation, E.U.E. and S.J.A.; formal analysis, E.U.E. and S.J.A.; investigation, E.U.E.; resources, C.A.B.; data curation, E.U.E.; writing—original draft preparation, E.U.E. and S.J.A.; writing—review and editing, S.J.A.; visualisation, S.J.A.; project administration, C.A.B. 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.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Puppala, A.J. Advances in ground modification with chemical additives: From theory to practice. Transp. Geotech. 2016, 9, 123–138. [Google Scholar] [CrossRef]
  2. Corrêa-Silva, M.; Miranda, T.; Rouainia, M.; Araújo, N.; Glendinning, S.; Cristelo, N. Geomechanical behaviour of a soft soil stabilised with alkali-activated blast-furnace slags. J. Clean. Prod. 2020, 267, 122017. [Google Scholar] [CrossRef]
  3. Kranina, E.I. China on the Way to Achieving Carbon Neutrality. Financ. J. 2021, 13, 51–61. [Google Scholar] [CrossRef]
  4. Moiseev, N.; Mikhaylov, A.; Varyash, I.; Saqib, A. Investigating the relation of GDP per capita and corruption index. Entrep. Sustain. Issues 2020, 8, 780–794. [Google Scholar] [CrossRef]
  5. Tiwari, N.; Satyam, N.; Puppala, A.J. Strength and durability assessment of expansive soil stabilized with recycled ash and natural fibers. Transp. Geotech. 2021, 29, 100556. [Google Scholar] [CrossRef]
  6. Renjith, R.; Robert, D.; Setunge, S.; Costa, S.; Mohajerani, A. Optimization of fly ash based soil stabilization using secondary admixtures for sustainable road construction. J. Clean. Prod. 2021, 294, 126264. [Google Scholar] [CrossRef]
  7. Lakshmi, S.M.; Geetha, S.; Selvakumar, M.; Susanna, K.D. Strength enhancement of Clayey Sand subgrade using lime and rice husk ash. Mater. Today Proc. 2021, 46, 7430–7435. [Google Scholar] [CrossRef]
  8. Oluremi, J.R.; Elsaigh, W.A.; Ikotun, B.D.; Osuolale, O.M.; Adedokun, S.I.; Oyelakin, S.E.; Ayodele, O.P. Strength enhancement in high silica wood ash stabilized lateritic soil using sodium tetraoxosulphate VI (Na2SO4) as activator. Int. J. Pavement Res. Technol. 2020, 14, 410–420. [Google Scholar] [CrossRef]
  9. Oyebisi, S.; Ede, A.; Olutoge, F.; Olukanni, D. Assessment of activity moduli and acidic resistance of slag-based geopolymer concrete incorporating pozzolan. Case Stud. Constr. Mater. 2020, 13, e00394. [Google Scholar] [CrossRef]
  10. De Azevedo, A.R.G.; Amin, M.; Hadzima-Nyarko, M.; Agwa, I.S.; Zeyad, A.M.; Tayeh, B.A.; Adesina, A. Possibilities for the application of agro-industrial wastes in cementitious materials: A brief review of the Brazilian perspective. Clean. Mater. 2021, 3, 100040. [Google Scholar] [CrossRef]
  11. Rahgozar, M.A.; Saberian, M.; Li, J. Soil stabilization with non-conventional eco-friendly agricultural waste materials: An experimental study. Transp. Geotech. 2018, 14, 52–60. [Google Scholar] [CrossRef]
  12. Singh, N.K.; Kalita, A. Use of Bagasse Ash and Coconut Fiber in Stabilization of Black Cotton Soil; Springer: Singapore, 2021; Volume 124 LNCE. [Google Scholar] [CrossRef]
  13. Senbagam, T.; Tharani, K.; Karunakaran, G.; Manikandan, G. Improve the physical properties of black cotton soil by using Prosopis julifloa ash. Mater. Today Proc. 2021; in press. [Google Scholar] [CrossRef]
  14. Yang, Q.; Du, C. Experimental Study on the Effect of Plant Ash on Soft Clay Stabilized with Cement-Based Composites. Geotech. Geol. Eng. 2020, 39, 105–117. [Google Scholar] [CrossRef]
  15. Vijayan, D.S.; Parthiban, D. Effect of Solid waste based stabilizing material for strengthening of Expansive soil—A review. Environ. Technol. Innov. 2020, 20, 101108. [Google Scholar] [CrossRef]
  16. Oyediran, I.A.; Ayeni, O.O. Comparative effect of microbial induced calcite precipitate, cement and rice husk ash on the geotechnical properties of soils. SN Appl. Sci. 2020, 2, 1157. [Google Scholar] [CrossRef]
  17. Ikeagwuani, C.; Obeta, I.; Agunwamba, J. Stabilization of black cotton soil subgrade using sawdust ash and lime. Soils Found. 2019, 59, 162–175. [Google Scholar] [CrossRef]
  18. Aziz, A.; Benzaouak, A.; Bellil, A.; Alomayri, T.; el Hassani, I.-E.E.N.; Achab, M.; El Azhari, H.; Et-Tayea, Y.; Shaikh, F.U.A. Effect of acidic volcanic perlite rock on physio-mechanical properties and microstructure of natural pozzolan based geopolymers. Case Stud. Constr. Mater. 2021, 15, e00712. [Google Scholar] [CrossRef]
  19. Jongpradist, P.; Homtragoon, W.; Sukkarak, R.; Kongkitkul, W.; Jamsawang, P. Efficiency of Rice Husk Ash as Cementitious Material in High-Strength Cement-Admixed Clay. Adv. Civ. Eng. 2018, 2018, 8346319. [Google Scholar] [CrossRef]
  20. Juluru, S.; Joy, M.; Obeth, A.; Mohammed, S. Enhancement of Expansive Soil with Addition of Wood Husk Ash and Silica Fume. Int. J. Innov. Technol. Explor. Eng. 2019, 9, 320–323. [Google Scholar] [CrossRef]
  21. Khalid, N.; Mukri, M.; Kamarudin, F. Strength of Soft Soil Stabilized Using Lime-POFA Mixtures. In InCIEC 2013: Proceedings of the International Civil and Infrastructure Engineering Conference 2013; Springer: Singapore, 2014. [Google Scholar] [CrossRef]
  22. Abdeldjouad, L.; Asadi, A.; Ball, R.; Nahazanan, H.; Huat, B.B. Application of alkali-activated palm oil fuel ash reinforced with glass fibers in soil stabilization. Soils Found. 2019, 59, 1552–1561. [Google Scholar] [CrossRef]
  23. Bagheri, Y.; Ahmad, F.; Ismail, M.A.M. Strength and mechanical behavior of soil–cement–lime–rice husk ash (soil–CLR) mixture. Mater. Struct. 2013, 47, 55–66. [Google Scholar] [CrossRef]
  24. Anupam, A.K.; Kumar, P.; Ransinchung R.N., G.D. Use of Various Agricultural and Industrial Waste Materials in Road Construction. Procedia Soc. Behav. Sci. 2013, 104, 264–273. [Google Scholar] [CrossRef] [Green Version]
  25. James, J.; Pandian, P.K. Soil Stabilization as an Avenue for Reuse of Solid Wastes: A Review. Acta Technol. Napoc. Civ. Eng. Archit. 2015, 58, 50–76. [Google Scholar]
  26. Kitazume, M.; Terashi, M. The Deep Mixing Method, 1st ed.; CRC Press: Boca Raton, FL, USA, 2013. [Google Scholar]
  27. EuroSoilStab. Design Guide Soft Soil Stabilisation; Ct97-0351; IHS BRE Press: Bracknell, UK, 1997; p. 95. [Google Scholar]
  28. Eyo, E.U.; Ng’Ambi, S.; Abbey, S.J. Performance of clay stabilized by cementitious materials and inclusion of zeolite/alkaline metals-based additive. Transp. Geotech. 2020, 23, 100330. [Google Scholar] [CrossRef]
  29. Abbey, S.J.; Ngambi, S.; Ganjian, E. Development of Strength Models for Prediction of Unconfined Compressive Strength of Cement/Byproduct Material Improved Soils. Geotech. Test. J. 2017, 40, 928–935. [Google Scholar] [CrossRef]
  30. Motamedi, S.; Song, K.-I.; Hashim, R. Prediction of unconfined compressive strength of pulverized fuel ash–cement–sand mixture. Mater. Struct. 2013, 48, 1061–1073. [Google Scholar] [CrossRef]
  31. Gunaydin, O.; Gokoglu, A.; Fener, M. Prediction of artificial soil’s unconfined compression strength test using statistical analyses and artificial neural networks. Adv. Eng. Softw. 2010, 41, 1115–1123. [Google Scholar] [CrossRef]
  32. Eyo, E.; Abbey, S. Machine learning regression and classification algorithms utilised for strength prediction of OPC/by-product materials improved soils. Constr. Build. Mater. 2021, 284, 122817. [Google Scholar] [CrossRef]
  33. Gajurel, A.; Mukherjee, P.S.; Chittoori, B. Estimating Optimal Additive Content for Soil Stabilization Using Machine Learning Methods. In Geo-Congress 2019: Geotechnical Materials, Modeling, and Testing 2019; American Society of Engineers (ASCE): Reston, VA, USA, 2019. [Google Scholar] [CrossRef]
  34. Soleimani, S.; Rajaei, S.; Jiao, P.; Sabz, A.; Soheilinia, S. New prediction models for unconfined compressive strength of geopolymer stabilized soil using multi-gen genetic programming. Measurement 2018, 113, 99–107. [Google Scholar] [CrossRef]
  35. Tabarsa, A.; Latifi, N.; Osouli, A.; Bagheri, Y. Unconfined compressive strength prediction of soils stabilized using artificial neural networks and support vector machines. Front. Struct. Civ. Eng. 2021, 15, 520–536. [Google Scholar] [CrossRef]
  36. Suman, S.; Mahamaya, M.; Das, S.K. Prediction of Maximum Dry Density and Unconfined Compressive Strength of Cement Stabilised Soil Using Artificial Intelligence Techniques. Int. J. Geosynth. Ground Eng. 2016, 2, 11. [Google Scholar] [CrossRef] [Green Version]
  37. Ghorbani, A.; Hasanzadehshooiili, H. Prediction of UCS and CBR of microsilica-lime stabilized sulfate silty sand using ANN and EPR models; application to the deep soil mixing. Soils Found. 2018, 58, 34–49. [Google Scholar] [CrossRef]
  38. Wang, J.; Xing, Y.; Cheng, L.; Qin, F.; Ma, T. The Prediction of Mechanical Properties of Cement Soil Based on PSO-SVM. In Proceedings of the International Conference on Computational Intelligence and Software Engineering (CiSE), Wuhan, China, 10–12 December 2010; pp. 1–4. [Google Scholar] [CrossRef]
  39. Mozumder, R.A.; Laskar, A.I.; Hussain, M. Empirical approach for strength prediction of geopolymer stabilized clayey soil using support vector machines. Constr. Build. Mater. 2017, 132, 412–424. [Google Scholar] [CrossRef]
  40. Chou, J.-S.; Yang, K.-H.; Lin, J.-Y. Peak Shear Strength of Discrete Fiber-Reinforced Soils Computed by Machine Learning and Metaensemble Methods. J. Comput. Civ. Eng. 2016, 30, 04016036. [Google Scholar] [CrossRef] [Green Version]
  41. Hanandeh, S.; Ardah, A.; Abu-Farsakh, M. Using artificial neural network and genetics algorithm to estimate the resilient modulus for stabilized subgrade and propose new empirical formula. Transp. Geotech. 2020, 24, 100358. [Google Scholar] [CrossRef]
  42. Tinoco, J.; Alberto, A.; da Venda, P.; Correia, A.G.; Lemos, L. A Data-driven Approach for qu Prediction of Laboratory Soil-cement Mixtures. Procedia Eng. 2016, 143, 566–573. [Google Scholar] [CrossRef] [Green Version]
  43. Roy, T.K. Influence of lime on Alluvial Soil Strengthened with Pond Ash and Rice Husk Ash for Construction of Subgrade of Road. In Proceedings of the GeoShanghai International Conference 2010, Shanghai, China, 3–5 June 2010; pp. 34–41. [Google Scholar]
  44. Hasan, H.; Dang, L.; Khabbaz, H.; Fatahi, B.; Terzaghi, S.; Hasan, H.; Dang, L.; Khabbaz, H.; Fatahi, B.; Terzaghi, S. Remediation of Expansive Soils Using Agricultural Waste Bagasse Ash. Procedia Eng. 2016, 143, 1368–1375. [Google Scholar] [CrossRef] [Green Version]
  45. Ali, F.H.; Adnan, A.; Choy, C.K. Geotechnical properties of a chemically stabilized soil from Malaysia with rice husk ash as an additive. Geotech. Geol. Eng. 1992, 10, 117–134. [Google Scholar] [CrossRef]
  46. Dang, L.C.; Hasan, H.; Fatahi, B.; Jones, R.; Khabbaz, H. Enhancing the engineering properties of expansive soil using bagasse ash and hydrated lime. Int. J. Geomate 2016, 11, 2447–2454. [Google Scholar] [CrossRef]
  47. Basha, E.A.; Hashim, R.; Mahmud, H.B.; Muntohar, A.S. Stabilization of residual soil with rice husk ash and cement. Constr. Build. Mater. 2005, 19, 448–453. [Google Scholar] [CrossRef] [Green Version]
  48. Choobbasti, A.J.; Ghodrat, H.; Vahdatirad, M.J.; Firouzianbandpey, S.; Barari, A.; Torabi, M.; Bagherian, A. Influence of using rice husk ash in soil stabilization method with lime. Front. Earth Sci. China 2010, 4, 471–480. [Google Scholar] [CrossRef]
  49. Hatmoko, J.T.; Suryadharma, H. Shear Behavior of Calcium Carbide Residue—Bagasse Ash Stabilized Expansive Soil. Procedia Eng. 2017, 171, 476–483. [Google Scholar] [CrossRef]
  50. Ghorbani, A.; Salimzadehshooiili, M.; Medzvieckas, J.; Kliukas, R. Strength Characteristics of Cement-Rice Husk Ash Stabilised Sand-Clay Mixture Reinforced with Polypropylene Fibers. Balt. J. Road Bridg. Eng. 2018, 13, 447–474. [Google Scholar] [CrossRef] [Green Version]
  51. Gupta, D.; Kumar, A. Performance evaluation of cement-stabilized pond ash-rice husk ash-clay mixture as a highway construction material. J. Rock Mech. Geotech. Eng. 2017, 9, 159–169. [Google Scholar] [CrossRef]
  52. Ikeagwuani, C.C. Comparative Assessment of the Stabilization of Lime-Stabilized Lateritic Soil as Subbase Material Using Coconut Shell Ash and Coconut Husk Ash. Geotech. Geol. Eng. 2019, 37, 3065–3076. [Google Scholar] [CrossRef]
  53. Jafer, H.; Atherton, W.; Sadique, M.; Ruddock, F.; Loffill, E. Stabilisation of soft soil using binary blending of high calcium fly ash and palm oil fuel ash. Appl. Clay Sci. 2018, 152, 323–332. [Google Scholar] [CrossRef]
  54. Kumar, A.; Gupta, D. Behavior of cement-stabilized fiber-reinforced pond ash, rice husk ash–soil mixtures. Geotext. Geomembranes 2016, 44, 466–474. [Google Scholar] [CrossRef]
  55. Liu, Y.; Chang, C.-W.; Namdar, A.; She, Y.; Lin, C.-H.; Yuan, X.; Yang, Q. Stabilization of expansive soil using cementing material from rice husk ash and calcium carbide residue. Constr. Build. Mater. 2019, 221, 1–11. [Google Scholar] [CrossRef]
  56. Muntohar, A.S.; Widianti, A.; Hartono, E.; Diana, W. Engineering Properties of Silty Soil Stabilized with Lime and Rice Husk Ash and Reinforced with Waste Plastic Fiber. J. Mater. Civ. Eng. 2013, 25, 1260–1270. [Google Scholar] [CrossRef]
  57. Nasiri, M.; Lotfalian, M.; Modarres, A.; Wu, W. Optimum utilization of rice husk ash for stabilization of sub-base materials in construction and repair projects of forest roads. Croat. J. For. Eng. 2016, 37, 333–343. [Google Scholar]
  58. Okonkwo, U.; Agunwamba, J. Classical optimization of bagasse ash content in cement-stabilized lateritic soil. Niger. J. Technol. 2016, 35, 481. [Google Scholar] [CrossRef] [Green Version]
  59. Patel, P. Mahiyar HK. An experimental study of black cotton soil, stabilized with rice husk ash, fly ash and lime. Int. J. Eng. Res. Technol. (IJERT) 2014, 4, 660–665. [Google Scholar]
  60. Pourakbar, S.; Asadi, A.; Huat, B.B.; Fasihnikoutalab, M.H. Stabilization of clayey soil using ultrafine palm oil fuel ash (POFA) and cement. Transp. Geotech. 2015, 3, 24–35. [Google Scholar] [CrossRef]
  61. Ramesh, H.N.; Manjunatha, B.V. Justification of strength properties of microstructural changes in the black cotton soil stabilized with rice husk ash and carbide lime in the presence of sodium salts. SN Appl. Sci. 2020, 2, 457. [Google Scholar] [CrossRef] [Green Version]
  62. Rao, D.K.; Rao, G.R.; Pranav, P.R.T. A laboratory study on the effect of rice husk ash & lime on the properties of marine clay. Int. J. Eng. Innov. Technol. 2012, 2, 345–353. [Google Scholar]
  63. Sabat, A.K. Utilization of bagasse ash and lime sludge for construction of flexible pavements in expansive soil areas. Electron. J. Geotech. Eng. 2012, 17, 1037–1046. [Google Scholar]
  64. Sharma, R.S.; Phanikumar, B.R.; Rao, B.V. Engineering Behavior of a Remolded Expansive Clay Blended with Lime, Calcium Chloride, and Rice-Husk Ash. J. Mater. Civ. Eng. 2008, 20, 509–515. [Google Scholar] [CrossRef]
  65. Singh, S. Experimental investigation of corn cob ash on silty clay stabilized with calcium carbide. Mater. Today Proc. 2020, 37, 3658–3660. [Google Scholar] [CrossRef]
  66. Sujatha, E.R.; Umarajeswari, S.; Vidhya, B.; Deepa, D. A Study on the Strength of Bagasse Ash Blended Clay. In Lecture Notes in Civil Engineering; Springer: Berlin/Heidelberg, Germany, 2021; pp. 251–258. [Google Scholar] [CrossRef]
  67. Vydehi, K.V.; Moghal, A.A.B.; Mohammed, S.A.S.; Lateef, M.A.; Ali, M.; Basheer, A.; Handoo, A.A.; Nabi, S.Z.; Wafa, B.U. Efficacy of Almond Shell Ash Inclusion on the Geotechnical Behavior of Lime Blended Kaolinitic Soil. In Proceedings of the Indian Geotechnical Conference 2019; Springer: Singapore, 2021; Volume 136. [Google Scholar] [CrossRef]
  68. Gary, F. Templeton, Department of Management and Information Systems, Mississippi State University A Two-Step Approach for Transforming Continuous Variables to Normal: Implications and Recommendations for IS Research. Commun. Assoc. Inf. Syst. 2011, 28, 41–58. [Google Scholar] [CrossRef] [Green Version]
  69. Hossain, K.M.A. Stabilized Soils Incorporating Combinations of Rice Husk Ash and Cement Kiln Dust. J. Mater. Civ. Eng. 2011, 23, 1320–1327. [Google Scholar] [CrossRef]
Figure 1. Normal distribution of values of UCS.
Figure 1. Normal distribution of values of UCS.
Materials 15 04575 g001
Figure 2. Distribution of input features for (a) Agro-based pozzolans; (b) Cementitious additives; (c) Plasticity index (PI); (d) Liquid limit (LL); (e) Plastic limit; (f) Soil class; (g) Curing duration; (h) Strength class.
Figure 2. Distribution of input features for (a) Agro-based pozzolans; (b) Cementitious additives; (c) Plasticity index (PI); (d) Liquid limit (LL); (e) Plastic limit; (f) Soil class; (g) Curing duration; (h) Strength class.
Materials 15 04575 g002
Figure 3. Variation of UCS of stabilised soils with duration and binder proportion. (a) 39% increase in UCS; (b) 100% increase in UCS; (c) 74% increase in UCS; (d) 109% increase in UCS; (e) 48% increase in UCS; (f) 44% increase in UCS; (g) 43% increase in UCS; (h) 115% increase in UCS; (i) 83% increase in UCS; (j) 60% increase in UCS; (k) 23% increase in UCS; (l) 29% increase in UCS; (m) 52% increase in UCS; (n) 36% increase in UCS; (o) 23% increase in UCS; (p) 77% increase in UCS.
Figure 3. Variation of UCS of stabilised soils with duration and binder proportion. (a) 39% increase in UCS; (b) 100% increase in UCS; (c) 74% increase in UCS; (d) 109% increase in UCS; (e) 48% increase in UCS; (f) 44% increase in UCS; (g) 43% increase in UCS; (h) 115% increase in UCS; (i) 83% increase in UCS; (j) 60% increase in UCS; (k) 23% increase in UCS; (l) 29% increase in UCS; (m) 52% increase in UCS; (n) 36% increase in UCS; (o) 23% increase in UCS; (p) 77% increase in UCS.
Materials 15 04575 g003
Figure 4. Classification metrics. (a) ROC curve, (b) precision–recall curve.
Figure 4. Classification metrics. (a) ROC curve, (b) precision–recall curve.
Materials 15 04575 g004
Figure 5. Normalised confusion matrix of multinomial classification.
Figure 5. Normalised confusion matrix of multinomial classification.
Materials 15 04575 g005
Figure 6. Multinomial classification graphs. (a) Cumulative gains curve, (b) lift curve.
Figure 6. Multinomial classification graphs. (a) Cumulative gains curve, (b) lift curve.
Materials 15 04575 g006
Figure 7. Regression curves. (a) Training set, (b) testing set.
Figure 7. Regression curves. (a) Training set, (b) testing set.
Materials 15 04575 g007
Figure 8. Residual curves. (a) Training set, (b) testing set.
Figure 8. Residual curves. (a) Training set, (b) testing set.
Materials 15 04575 g008
Figure 9. Distribution of residuals. (a) Normal distribution for training set, (b) cumulative curve of training set, (c) normal distribution for testing set, (d) cumulative curve of testing set.
Figure 9. Distribution of residuals. (a) Normal distribution for training set, (b) cumulative curve of training set, (c) normal distribution for testing set, (d) cumulative curve of testing set.
Materials 15 04575 g009
Figure 10. Multinomial classification feature importance. (a) Aggregate importance, (b) individual importance.
Figure 10. Multinomial classification feature importance. (a) Aggregate importance, (b) individual importance.
Materials 15 04575 g010
Figure 11. Regression aggregate feature importance.
Figure 11. Regression aggregate feature importance.
Materials 15 04575 g011
Figure 12. Partial dependence and individual conditional exception plots.
Figure 12. Partial dependence and individual conditional exception plots.
Materials 15 04575 g012aMaterials 15 04575 g012b
Figure 13. Partial dependence plots: (a) one-way dependence for agro-based pozzolan, (b) one-way dependence for cementitious additive, (c) two-way dependence for additives’ features, (d) 3-dimensional two-way dependence for additives’ feature.
Figure 13. Partial dependence plots: (a) one-way dependence for agro-based pozzolan, (b) one-way dependence for cementitious additive, (c) two-way dependence for additives’ features, (d) 3-dimensional two-way dependence for additives’ feature.
Materials 15 04575 g013
Figure 14. Partial dependence plots: (a) one-way dependence for PI, (b) one-way dependence for LL, (c) two-way dependence for soil plasticity features, (d) 3-dimensional two-way dependence for soil plasticity features (Python, version 3.8).
Figure 14. Partial dependence plots: (a) one-way dependence for PI, (b) one-way dependence for LL, (c) two-way dependence for soil plasticity features, (d) 3-dimensional two-way dependence for soil plasticity features (Python, version 3.8).
Materials 15 04575 g014
Figure 15. Partial dependence plots: (a) one-way dependence for duration of curing, (b) one-way dependence for soil class, (c) two-way dependence for curing duration and soil class, (d) 3-dimensional two-way dependence for curing duration and soil class (Python, version 3.8).
Figure 15. Partial dependence plots: (a) one-way dependence for duration of curing, (b) one-way dependence for soil class, (c) two-way dependence for curing duration and soil class, (d) 3-dimensional two-way dependence for curing duration and soil class (Python, version 3.8).
Materials 15 04575 g015
Table 2. Range of chemical compositions of the materials.
Table 2. Range of chemical compositions of the materials.
OxideAgro-Based PozzolansCementitious AdditivesSoils
(%)(%)(%)
Silica (SiO2)9.00–93.150.90–25.1241.45–71.16
Alumina (Al2O3)0.21–19.390.13–12.203.41–19.00
Iron oxide (Fe2O3)0.21–6.100.15–4.463.59–21.34
Calcium oxide (CaO)0.30–13.878.47–95.300.15–5.76
Potash (K2O)1.38–27.240.10–5.100.15–3.45
Magnesia (MgO)0.20–3.700.25–3.000.25–2.86
Loss on ignition (LOI)1.76–15.211.24–4.701.40–5.61
Sodium oxide (Na2O)0.14–1.500.17–1.720.21–5.20
Sodium trioxide (SO3)0.10–2.530.20–2.660.34–0.67
Table 3. Statistical measures of input features and target variable.
Table 3. Statistical measures of input features and target variable.
StatisticAgro-Based PozzolanCementitious AdditivePILLPLDurationUCS
(%)(%)(%)(%)(%)(Days)kPa
Mean8.225.0327.9253.6425.7216.30893.12
Standard Deviation5.932.5411.7215.457.789.40623.42
Kurtosis−0.73−0.61−1.02−1.05−1.50−1.71−0.16
Skewness0.640.420.780.620.120.311.63 × 106
Minimum0.100.5014.0035.0014.007.0060.00
Maximum25.0011.2550.0086.0037.0028.003306.51
Table 4. Range of values of strength category for UCS.
Table 4. Range of values of strength category for UCS.
Strength CategoryUCS (kPa)
Very soft<25
Soft25–50
Firm50–200
Stiff200–400
Hard>400
Table 5. Optimised parameters of GB algorithm.
Table 5. Optimised parameters of GB algorithm.
CategoryParameter
Tree-basedmin_samples_split = 6; max_depth = 5; min_samples_leaf = 11; max_leaf_nodes = 100; max_features = sqrt; min_weight_fraction_leaf = 0.005
BoostingSubsmaple = 0.7999905; n_estimators = 501; Learning rate = 0.0805;
General functionalityloss = huber; random state = 20; criterion = friedman_mse;
min_impurity_decrease = 8.999999999; init = None; alpha = 0.9; n_iter_no_change = None
Table 6. Multinomial classification metrics.
Table 6. Multinomial classification metrics.
Multiclass MetricType Value
AccuracyOverall0.920
Weighted 0.966
AUCMacro0.951
Micro0.980
Weighted0.958
f1 scoreMacro0.818
Micro0.920
Weighted0.915
PrecisionMacro0.882
Micro0.920
Weighted0.938
RecallMacro0.833
Micro0.920
Weighted0.920
Table 7. ML performance of tree ensembles on the stabilised soil dataset.
Table 7. ML performance of tree ensembles on the stabilised soil dataset.
ModelOptimised ParametersR2MAPE
TrainingTestingTrainingTesting
Decision Treesplitter = ‘random’, random_state = 51, max_depth = 20, min_samples_split = 5, criterion = ‘squared_error’, min_samples_leaf = 2, max_features = ‘auto’, min_impurity_decrease = 0.9, max_leaf_nodes = 500.8200.8130.3320.915
Random Forestn_estimators = 500, random_state = 5, max_depth = 9, min_samples_split = 5, criterion = ‘squared_error’, min_samples_leaf = 2, max_features = ‘log2’, oob_score = False, bootstrap = True, max_leaf_nodes = 1000.8610.8270.3280.845
Extremely Randomised Treesn_estimators = 100, random_state = 90, max_depth = 9, min_samples_split = 6, criterion = ‘squared_error’, min_samples_leaf = 2, max_features = ‘auto’, oob_score = False, bootstrap = False, max_leaf_nodes = 700.8670.8490.2110.513
Adaboostbase_estimator = Decision tree (max_depth = 7), n_estimators = 20, random_state = 45, loss = ‘exponential’, learning_rate = 0.92050.8670.8560.2000.478
Gradient Boosting (this study)Table 50.9000.9000.2130.335
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Eyo, E.U.; Abbey, S.J.; Booth, C.A. Strength Predictive Modelling of Soils Treated with Calcium-Based Additives Blended with Eco-Friendly Pozzolans—A Machine Learning Approach. Materials 2022, 15, 4575. https://doi.org/10.3390/ma15134575

AMA Style

Eyo EU, Abbey SJ, Booth CA. Strength Predictive Modelling of Soils Treated with Calcium-Based Additives Blended with Eco-Friendly Pozzolans—A Machine Learning Approach. Materials. 2022; 15(13):4575. https://doi.org/10.3390/ma15134575

Chicago/Turabian Style

Eyo, Eyo U., Samuel J. Abbey, and Colin A. Booth. 2022. "Strength Predictive Modelling of Soils Treated with Calcium-Based Additives Blended with Eco-Friendly Pozzolans—A Machine Learning Approach" Materials 15, no. 13: 4575. https://doi.org/10.3390/ma15134575

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