Next Article in Journal
Spectral Mixture Analysis (SMA) Model for Extracting Urban Fractions from Landsat and Sentinel-2A Images in the Al-Ahsa Oasis, Eastern Region of Saudi Arabia
Previous Article in Journal
Rural Land Management and Revitalization through a Locally Coordinated Integrated Master Plan—A Model from Germany to China
Previous Article in Special Issue
Determining Attribute—Response Relationships of Soils under Different Land Uses: A Case Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Digital Mapping of Soil Organic Carbon Using Machine Learning Algorithms in the Upper Brahmaputra Valley of Northeastern India

by
Amit Kumar
1,2,*,
Pravash Chandra Moharana
3,
Roomesh Kumar Jena
4,
Sandeep Kumar Malyan
5,
Gulshan Kumar Sharma
6,
Ram Kishor Fagodiya
7,
Aftab Ahmad Shabnam
1,
Dharmendra Kumar Jigyasu
1,
Kasthala Mary Vijaya Kumari
1 and
Subramanian Gandhi Doss
2
1
Central Muga Eri Research and Training Institute, Lahdoigarh, Jorhat 785700, Assam, India
2
Central Sericultural Research and Training Institute, Mysuru 570008, Karnataka, India
3
ICAR-National Bureau of Soil Survey and Land Use Planning, Nagpur 440033, Maharashtra, India
4
ICAR-Indian Institute of Water Management, Bhubaneswar 751023, Odisha, India
5
Department of Environmental Studies, Dyal Singh Evening College, University of Delhi, New Delhi 110003, India
6
ICAR-Indian Institute of Soil and Water Conservation, Research Centre, Kota 324002, Rajasthan, India
7
ICAR-Central Soil Salinity Research Institute, Karnal 132001, Haryana, India
*
Author to whom correspondence should be addressed.
Land 2023, 12(10), 1841; https://doi.org/10.3390/land12101841
Submission received: 19 July 2023 / Revised: 30 August 2023 / Accepted: 18 September 2023 / Published: 27 September 2023

Abstract

:
Soil Organic Carbon (SOC) is a crucial indicator of ecosystem health and soil quality. Machine learning (ML) models that predict soil quality based on environmental parameters are becoming more prevalent. However, studies have yet to examine how well each ML technique performs when predicting and mapping SOC, particularly at high spatial resolutions. Model predictors include topographic variables generated from SRTM DEM; vegetation and soil indices derived from Landsat satellite images predict SOC for the Lakhimpur district of the upper Brahmaputra Valley of Assam, India. Four ML models, Random Forest (RF), Cubist, Extreme Gradient Boosting (XGBoost), and Support Vector Machine (SVM), were utilized to predict SOC for the top layer of soil (0–15 cm) at a 30 m resolution. The results showed that the descriptive statistics of the calibration and validation sets were close enough to the total set data and calibration dataset, representing the complete samples. The measured SOC content varied from 0.10 to 1.85%. The RF model’s performance was optimal in the calibration and validation sets (R2c = 0.966, RMSEc = 0.159%, R2v = 0.418, RMSEv = 0.377%). The SVM model, on the other hand, had the next-lowest accuracy, explaining 47% of the variation (R2c = 0.471, RMSEc = 0.293, R2v = 0.081, RMSEv = 0.452), while the Cubist model fared the poorest in both the calibration and validation sets. The most-critical variable in the RF model for predicting SOC was elevation, followed by MAT and MRVBF. The essential variables for the Cubist model were slope, TRI, MAT, and Band4. AP and LS were the most-essential factors in the XGBoost and SVM models. The predicted OC ranged from 0.44 to 1.35%, 0.031 to 1.61%, 0.035 to 1.71%, and 0.47 to 1.36% in the RF, Cubist, XGBoost, and SVM models, respectively. Compared with different ML models, RF was optimal (high accuracy and low uncertainty) for predicting SOC in the investigated region. According to the present modeling results, SOC may be determined simply and accurately. In general, the high-resolution maps might be helpful for decision-makers, stakeholders, and applicants in sericultural management practices towards precision sericulture.

1. Introduction

Soil Organic Carbon (SOC) imparts a significant role in greenhouse gas dynamics by acting as a sink and source based on the prevailing conditions and soil management. Soil is the most-extensive carbon storage system in the terrestrial ecosystem. Thus, the global scientific community requires an understanding of the dynamics of soil organic carbon. As a result, there is a growing demand for SOC information worldwide [1,2]. The best depiction of SOC information is high-resolution soil maps, which are difficult to prepare at the global level simultaneously, and SOC information needs to be continuously upgraded as and when the region-specific information is available. Recently, soil mapping has been achieved through ground-based surveys; the spatial distribution of the soil properties of a landscape is difficult to specify at an appropriate resolution through these traditional mapping methods. Thus, developing a more-accurate and -reliable methodology and process is highly required to forecast the soil properties of specific soil types or locations.
In line with this, Digital Soil Mapping (DSM) is a highly promising methodology combining two advanced technologies, i.e., machine learning and remote sensing, such as Hyper spectral, multispectral, and radar [3,4]. DSM is in high demand globally to map region-specific soil qualities and is influential in achieving various sustainability goals, including land-use management. DSM’s vital machine learning (ML) techniques are artificial neural networks, decision trees, linear models, multivariate adaptive regression splines, regression trees, and support vector machines [5,6]. The statistical relation among the soil, environmental variables, satellite information, topographic characteristics such as digital elevation models, etc., is the merit of these methodologies, which have many advantages over the traditional methods of soil mapping. Due to these advantages, there has been a surge in DSM investigations, especially regarding the spatial variation in soil properties observed in recent decades [7,8]. Hengl et al. [9] used DSM at a resolution of 1 km in Africa to forecast soil parameters such as organic carbon, pH, sand, silt, and clay fractions. Hengl et al. [10] used DSM at a resolution of 250 m to estimate SOC, pH, texture, and bulk density on a global scale. The GlobalSoilMap consortium is one of the best platforms to utilize the historical, as well as the recent information of the different regions and convert it into synchronized global information for further utilization in policy formulation, infrastructural development, biodiversity conservation, disaster risk assessment, etc. [11,12].
Northeastern India is rich in flora and fauna, has specific climates, high endemism, and is one of the most-vulnerable ecoregions of the globe [13]. The Brahmaputra Valley region has a unique status among the different ecosystems of Northeastern India. For example, the world’s most-expensive golden silk, muga, is endemic to this region and has high traditional and ethical values among tribal and non-tribal inhabitants [14]. Muga silkworm rearing is an entirely outdoor activity, i.e., rearing is carried out on the host plant. Som is the primary host plant for muga silk production. Assam contributes > 90% of the muga silk production [14]. Lakhimpur is the highest muga-silk-producing district among all the producing districts in the world. The Brahmaputra Valley, especially the upper part, loses Soil Organic Carbon (SOC) due to enhanced land-use land-cover changes, deforesttion, flooding, leftover fallow land, etc. SOC is highly required to understand the carbon pool dynamic under various agricultural and forest ecosystems. Machine learning can serve as an essential tool to investigate and estimate the SOC at the landscape level by using lower physical inputs than grid-based physical mapping. It also reduces the uncertainty level, which is higher in grid-based physical mapping. Keeping this view, the current investigation was planned using 160 endemic silk (muga and eri)-specific soil organic datasets of the Lakhimpur district of Assam, which bears the first rank in Muga silk production globally. The aims of the investigation were (i) to prepare a digital SOC map with different ML algorithms, (ii) to evaluate the model efficiency, and (iii) to identify the most-influencing environmental covariates in SOC distribution under silk-producing soils.

2. Materials and Methods

2.1. Study Area

Lakhimpur is located in the northeast corner of Assam State and between 26°48′ and 27°53′ northern latitude and 93°42′ and 94°20′ east longitude (Figure 1). Papump and Siang (Arunachal Pradesh), Dhemji and Majuli (Assam), and Bishwanath (Assam) are situated on north, east, south, and west, respectively. The Brahmaputra and Subansiri Rivers flow east and south of Lakhimpur. Lakhimpur’s total geographical area is about 2277 km2. The rural and urban areas are 2257 and 20 km2, respectively. Lakhimpur’s climate is subtropical and has hot, humid summers and cold winters. The mean annual and soil temperatures are 24.5 °C and 22 °C, respectively. The mean annual precipitation ranges from 2000–4000 mm. The dominant soil of Lakhimpur is alluvial. Lakhimpur contributes to all major types of silk and is the highest in muga silk production among all the districts in India.

2.2. Soil Sampling and Analysis

The topsoil samples (0–15 cm) were collected through a v-notch following systematic randomized sampling from the study region. The initial cleaning was opted to be performed at each sampling point to ensure a debris-free soil sample. The composite samples were prepared in the field through the proportionate method (200 g from each sample) to represent the complete field condition. There were a total of 10 samples that contributed to each composite sample. The minimum quantity of the replicate was restricted to 500 g. One-hundred sixty composite samples were collected from the muga and eri sericulture farms. All the composite replicates were transported to the laboratory. Each replicate was air-dried in the shade and sieved with a 2 mm sieve. The wet oxidation method was used to estimate the SOC in the soil [15] as the OC recovery range was reported to be 60–86% with an average recovery of 76% [15].

2.3. Digital Soil Mapping Technique

The SCORPAN model deduces soil class/properties at a specific place indirectly from environmental factors such as age (a), climate (c), geographical position (n), organisms (o), parent material (p), and relief (r) [16]. The SCORPAN model provides the quantitative (empirical) relationship among the soil characteristics and variables that control spatial variability. The SCORPAN model follows Dokuchaev’s [17] and Jenny’s [18] theories.
S = f ( s , c , o , r , p , a , n )
Soil can be used as the input data for the SCORPAN model in point data format, existing soil maps, or spectral characteristics (remotely sensed). Environmental covariates are raster data processed by a Geographic Information System (GIS) and are digital and geo-graphically explicit. The SCORPAN model makes quantifying the connections between spatially explicit digital environmental variables and the predicted soil properties easier. Uncertainty estimation under the SCORPAN model is more straightforward than other traditional methods.

2.4. Selection of Environmental Covariates

Forty-nine pedogenesis-related environmental covariates (ECs) were collected from multiple sources, i.e., indexes and data products determined from the satellites. These ECs were categorized as climate, organism, parent material, position, relief, remote sensing, and soil. The 30 m spatial resolution and 16 d revisit duration of the Landsat 8 Operational Land Imager have been extensively used in DSM [8]. The Landsat 8 Surface Reflectance Level 1 Tier 1 was used to calculate related indices using land surface reflectance data (https://www.usgs.gov, accessed on 25 May 2023). The spectral indices such as the Normalized Difference Vegetation Index (NDVI) and Landsat data (1–11 bands) were used to predict the SOC and used as variables in the model. A digital elevation model (30 m spatial resolution) was produced by SRTM and used to derive relief covariates. By using SagaGIS Version 6.3.0, the terrain indices such as the Aspect (Asp), Elevation (Elev), Slope (Slp), Relative Slope Position (RSP), Topographic Wetness Index (TWI), Multi-Resolution index of Valley Bottom Flatness (MRVBF), Valley Depth (VD), Convergence Index (CI), Channel Network Base Level (CNBL), Channel Network Distance (CND), LS-factor, Plan Curvature (PlaC), Profile Curvature (PrC), Total Catchment Area (TCA), Topographic Positioning Index (TPI), and Topographic Roughness Index (TRI) were generated. All 19 bioclimatic variables were retrieved from WorldClim products (http://www.worldclim.com/version2, accessed on 25 May 2023) for each sampling point to use in the model. All collected covariates were aggregated/disaggregated through average resampling/bilinear resampling into a 30 × 30 m grid. Quantitative spatial models were developed using these aggregated ECs. All the data used in the prediction modelling were in raster format. The sampling coordinates were converted into UTM WGS84 Zone 46 N before use.
Feature selection was carried out by a multicollinearity test. Based on the SCORPAN conceptual model of soil development, it was possible to generate 49 environmental covariates as predictors from three data sources, i.e., geomorphometric/Digital Elevation Model (DEM), remote sensing, and climatic variables, where the spatial variability of the soil organic carbon was explained. Since the number of covariates was initially 49, there could be a high correlation among them. Therefore, multicollinearity (a common limitation of modeling) might have existed. Consequently, the Variance Inflation Factor (VIF), implemented by using the SPSS software (IBM SPSS Statistics 20.0), was used to evaluate the distributions and relationships between all environmental covariates. In brief, the VIF assesses how much the variance of an estimated regression coefficient increases when the predictors are correlated. This approach tries to remove some irrelevant covariates step by step from the dataset. The covariates were selected based on the VIF values, ranging from 1 to less than 5. The final feature selection step was performed to considerably reduce the number of covariates without significantly decreasing model prediction accuracy. Based on the VIF value, 28 environmental covariates were selected from multiple sources (Table 1).

2.5. Machine Learning Techniques

Four ensemble models (Random Forest (RF), regression tree (Cubist), Extreme Gradient Boosting (XGBoost), and Support Vector Machine (SVM)) were used for the prediction mapping of the SOC in the present investigation. These ensemble models are constructed by integrating multiple learners/algorithms. The ensemble model was constructed by the ensemble learning/method, which integrates multiple algorithms. The respective trend analysis was conducted in the R 4.1.2 software [19], and maps were reproduced in ArcGIS 10.3 software.

2.5.1. Random Forest

Random forest consists of multiple decisions (Classification and Regression Trees (CARTs)) based on the binary rule. Each input is considered a tree algorithm to define a relationship between dependent and independent variables. The essential parameters of random forest, i.e., the number of variables (Mtry) and the Number of trees (Ntree), can be selected and adjusted each time during the decision process to obtain the optimum result. The Ntree and Mtry were kept in the range of 1–30 and 100–3000, respectively, as defined parameters (Table 2). The final prediction of RF is the accumulation of the results of all the individual trees (algorithms). Out-Of-Bag (OOB) estimates the error by bootstrapping the original datasets. The Mean-Squared Error (MSEOOB) is determined by combining the OOB predictions from all trees [20]:
MSE OOB   = 1 n i = 1 n ( z i z ^ i O O B ) 2
where Z ^ i O O B  is the average of all OOB predictions. The percentage of explained variance (Varex) is determined after the MSEOOB is normalized because it depends on the unit of the response variable:
V a r e x = 1 M S E O O B V a r z
where Varz denotes the answer variable’s overall variance; thus, RF is handy for predicting the output of ecological variables under a composite environmental system. The “ranger” function in the “caret” package is called in R 4.1.2 for random forest prediction.

2.5.2. Cubist

Cubist is based on the linear regression model [6]. Furthermore, it recently gained importance in DSM investigations in various ecological and geological regions [21,22]. These models (linear regression) have a highly classified relationship with the end value [23]. This linear regression model is used as the leaf nodes of the other regression tree algorithms in the Cubist machine learning techniques. Cubist produces multivariate models, which are composed of various rules. Thus, the prediction model under the Cubist environment is a selection of the rules. In the regression tree (Cubist), the number of model trees (committees) and the number of nearest neighbors (neighbors) were defined as 1–100 and 0–9, respectively (Table 2). This study used the “Cubist” package in the R 4.1.2 software.

2.5.3. Extreme Gradient Boosting

The Extreme Gradient Boosting (XGBoost) algorithm is generally used to enhance the performance of the regression tree and K classification. XGBoost improves the calculation speed and reduces the chance of overestimation by simplifying the objective functions. Automatic simultaneous computation during the training step is another merit of XGBoost. Supplemental training strategies were used to extend “strong” learners from the “weak” learners by “boosting”. Under XGBoost’s training steps, simultaneous computations for the functions are performed automatically [24]. The XGBoost algorithm’s parameters, i.e., the algorithm type, the depth of the tree, the sum of the weights of all observations, the sample number provided to a tree, the variables used in tree construction, and the learning rate used in this investigation, are given in Table 2.

2.5.4. Support Vector Machines

Support Vector Machines (SVMs) are considered as common predictors for any multivariate function up to a specified accuracy. SVMs are used to solve regression problems through linear and nonlinear models. The linear model detects the noise in the datasets, and the nonlinear model converts the input space into a larger dimension [25]. Under the SVM classification and regression process, a set of connected supervised learning algorithms was used. These algorithms are the universal predictors of multivariate functions with high accuracy. In the present investigation, the range of components, input data, kernel type function, parameter for penalty, bandwidth parameters, etc., were used (Table 2).

2.6. Model Evaluation

In this study, four evaluation metrics were calculated to analyze the model performance: the coefficient of determination (R2), Mean Error (ME), Lin’s Concordance Correlation Coefficient (CCC), and Root-Mean-Squared Error (RMSE). The RMSE and R2 were used to measure the accuracy and stability, respectively. A smaller RMSE and higher R2 value depict the accuracy and stability of the model, respectively. The sample set in an 80:20 ratio was used for calibration and validation. The RMSEc and RMSEv were used to express the fitting accuracy of the models’ calibration validation, respectively. These validation metrics are calculated as follows:
R 2 = 1 i = 1 n ( p i o i ) 2 i = 1 n ( p i ^ o i ^ ) 2
ME = 1 n i = 1 n ( p i o i )
RMSE = 1 n i = 1 n ( p i o i ) 2
where p i and o i are the predicted and observed values and p i ^ and o i ^ are the means of the predicted and observed values.
CCC = 2 ρ σ 0 σ p σ 0 2 + σ p 2 + ( µ 0 µ p ) 2
µ 0  and  µ p  are the means of the observed and predicted values, and σ 0 2 and σ p 2 are corresponding variances. ρ is the Pearson correlation coefficient between the observed and predicted values.

2.7. Uncertainty Assessment

The current study used non-parametric bootstrap methods to analyze the spatially detailed quantification of the best model’s SOC prediction uncertainty. It ensured that the sample size and probability distribution of the old and new bootstrap datasets were the same. Fifty bootstrapped datasets were created for the calibration. Through the final evaluation indication and predicted outcome, respectively, it was presumed that the validation dataset’s average accuracy was good. Maps of the 0.05 and 0.95 quantiles were created by the estimate. In other words, there were uncertainty levels for the SOC predictions for each pixel position in the research region. The 90% confidence interval’s upper and lower bounds were mapped. Using the 5th and 95th percentiles of prediction, the mean of the SOC contents in each pixel and the 90% CI were determined. For the top-performing model, three maps of the SOC were created: the mean forecast, lower confidence interval (5%), and higher confidence interval (95%).
The Prediction Interval Coverage Probability (PICP) criteria were used to assess the uncertainties of the prediction [26,27]. The following confidence intervals were utilized in this study: 99, 97.5, 95, 90, 80, 60, 40, 20, 10, and 5%. One should expect the PICP value or proportion to be close to the corresponding confidence level to determine whether the indicated uncertainties have been effectively computed [28,29]. The uncertainty is at its lowest at the PICP, which is approximately 100(1 − a) %, like a 90% confidence interval. The accuracy of the PICP was established by an estimate of 90% for a 90% prediction interval. The standard deviations of the accuracy metrics (R2, RMSE, and ME) made from the 50 bootstrapped datasets were also employed to indicate, to a certain degree, the stability and uncertainty of the predictions.

3. Results

3.1. Descriptive Statistics

Eighty percent of the calibration set and twenty percent of the validation set, each comprising 128 and 32 sample points, were divided from the 160-sample dataset. The calibration set’s coefficient of variation, skewness, and kurtosis values were proximal to the total set data and represented a complete sample, according to the descriptive statistics (Table 3), performed for the calibration and validation sets. The measured SOC content was found in the 0.10–1.85% range. The considerable variation between the minimum and maximum SOC envisaged that the surface soil of the upper Brahmaputra Valley is susceptible to the environment, management practices, and soil disturbance as flooding is a yearly event in the studied region. The higher cumulative variation (47.36%) indicates that the SOC had high spatial variation and a semi-homogenous geographical distribution in the investigated region. This might be due to the topographical variation, crop management practices, and land-use and land-cover changes. A similar argument was devised by Jena et al. [30] and Moharana et al. [4,8] in their respective studies in the northeastern regions. Based on the investigation, contemporary statistical procedures must be prioritized to capture the geographical variability in the SOC as it is a stochastic variable highly influenced by the soil formation and the regional climate.
The frequency distribution of the calibration and validation dataset of the SOC content is depicted in Figure 2. The rug plot on the X-axis represents the locations and SOC contents of the particular sample-collection point. The bell-shaped curve of the SOC contents shows the normal distribution. Further, normalization has yet to be opted for as machine learning models do not require normalization of the data points [31]. The prediction models were fit using the original SOC concentration in the studied location. The probability densities and histograms are depicted by curves and bars, respectively. The uncertainties of the spatial autocorrelation were checked through the Global Moran’s I. It was ensured that there was no spatial autocorrelation in the geographic patterns before fitting the machine learning techniques (prediction models).
The descriptive statistics of the environmental covariates are presented in Table 4. Terrain attributes such as the Asp, CI, CND, LS-factor, MRRTF, MRVBF, RSP, Slp, TCA, and TPI were highly variable (CV > 35%). Landsat bands had a CV < 15%, indicating high data homogeneity. Our findings from the CV analysis suggest that the environmental covariates had a semi-homogeneous geographical distribution, which may be related to variations in the research area’s soil-forming factors. As demonstrated by Falahatkar et al. [32] and Dharumarajan et al. [27], a soil attribute such as SOC is a stochastic variable supported by the continuation of the soil-formation processes and the change in the climate. As a result, it is essential to use contemporary statistical procedures to survey the geographical variability in the SOC.
The pairwise correlation coefficient (r) was computed to determine the relationship between the topographic and remote sensing covariates and the SOC (Figure 3). The correlation matrix showed that the SOC had a positive correlation with the elevation (r = −0.108, p < 0.05), NDVI (r = 0.07, p < 0.01), and CNBL (r = 0.178, p < 0.05). In contrast, a significant negative correlation was observed for the AP (r = −0.091, p < 0.05) and TPI (r = −0.16, p < 0.05). The remaining topographic attributes (Asp, MRVBF, and LS-factor) were not significantly correlated with the SOC in the upper Brahmaputra Valley regions. Among the Landsat 8 remote sensing data, Band1, 2, 6, and 7 were positively correlated with the SOC. In the upper Brahmaputra Valley, soil formed on the top of the slope and, then, moved outward or downward under the influence of gravity and deposited in the valley. Therefore, the elevation and slope always negatively correlated with the SOC. The present study’s correlation coefficient was low, although a similar trend was observed. Anthropogenic activity in the valley may be the reason for the low SOC in that region. This observation is in agreement with the work carried out by previous researchers [3] in similar landscapes.

3.2. Evaluation of Prediction Models

The R2, CCC, RMSE, and ME were computed, and the results are depicted in Table 5. The performance (goodness of fit and errors employed) of the RF, Cubist, SVM, and XGBoost models for the prediction of the SOC in the upper Brahmaputra Valley in the northeast state of India was evaluated. The XGBoost model had the highest accuracy and the lowest error on the calibration set (R2c = 0.998, RMSEc = 0.022%) and had slightly lower accuracy on the test set (R2v = 0.152, RMSEv = 0.424%). In contrast, the RF model performed well on both the calibration and validation sets (R2c = 0.966, RMSEc = 0.159%, R2v = 0.418, RMSEv = 0.377%). However, the SVM model had the next-lowest accuracy, accounting for just around 47% of the variance, and the Cubist model fared the poorest on both the calibration and validation sets (R2c = 0.471, RMSEc = 0.293, R2v = 0.081, RMSEv = 0.452). The RF model provided the lowest RMSE, i.e., 0.159 and 0.377, for the calibration and validation datasets. However, the lowest ME was observed for XGBoost, which was 0 on the calibration dataset and 0.054 on the validation dataset. The ME for the RF model was 0.001 and 0.136 for the calibration and validation datasets. The CCC values of 0.863 and 0.549 on the calibration and validation datasets were found for the RF model, suggesting good agreement between the predicted and observed values. Based on the highest R2 for the validation value of the RF model, it showed the best ability to predict the SOC in the sericultural soil of this northeast region of India.
The RF model’s R2 was not the highest. Still, its RMSE was the lowest when compared to the other models since its predictions were closer to the actual values than those of other models (the predicted–observed fit line was almost 1:1) (Figure 4). Additionally, the slope of the predicted–observed fit curve for the RF and Cubist models dramatically improved, demonstrating a considerable improvement in prediction accuracy, primarily due to the concentration of the scatter distribution. The scatter spots are nearer to the curve of the fitting.

3.3. Importance of Environmental Variables

Figure 5’s comparative relative variable importance plots show that RF chose more variables than the Cubist, SVM, and XGBoost models. The order of the most-crucial factors in the RF model for predicting the SOC was Elev > MAT > Band 3 > Band1 > MRVBF. The most-crucial variables for utilizing a Cubist model to predict the variance of the SOC were Slp, TRI, MAT, and Band4. The AP and LS were the most-important factors in the XGBoost and SVM models. One of the primary benefits of RF models over other ML models is that the former assess the relative relevance of the covariates in the model, in contrast to Cubist, which maintains the model through stepwise selection only with highly correlated predictive variables [33]. Even though there may be relationships between the predictive variables and soil, RF avoids removing those [34]. In a study conducted in the Zahak County of Iran, Pahlavan-Rad and Akbarimoghadam [35] discovered that the CNBL and elevation were the most-crucial covariates for predicting soil qualities. The elevation and slope were the most-significant factors for the SOC prediction for both the RF and Cubist models, followed by the other topography and vegetation factors. In the upper Brahmaputra Valley in the northeast state of India, flooding causes soil sediments to flow and gather in lower-elevation regions. These topography-driven erosional processes transport the SOC from higher elevations, where the SOC concentrations are often lower, to lower elevations, where the SOC concentrations are typically higher [30].

3.4. Spatial Prediction of SOC

Significant discrepancies between the predictions are shown in Figure 6, which depicts the spatial predictions of the SOC made by the various models. Although the four models’ cross-validation accuracy metrics were comparable, RF’s predictions more closely matched the spatial pattern of the OC we anticipated in the research area. The SOC in the research area appeared to be underestimated by the SVM prediction. All of the maps for the prediction models displayed abrupt and gradual shifts throughout the research region. In the RF, Cubist, XGBoost, and SVM models, the predicted OC ranged from 0.44 to 1.35%, 0.031 to 1.61%, 0.035 to 1.71%, and 0.47 to 1.36%, respectively. It is impossible to say which model is the most-accurate without independently confirming these predictions; however, we selected the RF model as the “best” because the cross-validation accuracy metrics were similar, and the spatial predictions visually matched our perception of the terrain. The spatial patterns of the SOC in all models were reasonable, with large values in the study area’s western region, which is dominated by forest and covered in dense vegetation, and minor values in its eastern regions, which have soils subject to significant erosion and crop cultivation. Low-elevation cultivable fields appear more uncertain in the upper Brahmaputra Valley than in high-elevation cultivable lands. This can be explained by the variation in the management techniques used in the more-intensively farmed regions [30]. The covariates utilized in these predictive models may have needed to adequately account for the management aspects contributing to the spatial heterogeneity in the SOC content levels in these downstream-farmed regions. The high-elevation areas are less actively farmed, which may have less impact due to management and because the terrain features have a stronger hold over them. SOC concentrations are the outcome of the balance between carbon imports and outputs in soils; several variables, including local characteristics such as vegetation and topography and environmental conditions, affect this equilibrium [36]. According to Pahlavan-Rad et al. [37], topography was the most-significant covariate influencing the distribution of the SOC in our study.

3.5. Uncertainty Prediction

To quantify the spatial distribution of prediction uncertainty, bootstrapping was used. Figure 7 displays the level of uncertainty in the best model’s (RF) predictions. The uncertainty analysis partially confirmed the trend of the ML algorithms’ ability to predict the SOC. With a 90% confidence interval, uncertainty was found in the lower and upper predicted bounds. The effectiveness of the predicted uncertainties during testing was evaluated using the PICP technique. The uncertainty analysis exhibited the same trends as the ML algorithms’ capacity to predict the SOC.
The 90% confidence interval with lower and higher prediction bounds showed the degree of uncertainty. The PICP methodology is a method for assessing how well the anticipated uncertainties perform throughout testing. The percentage of observations within the associated prediction interval is the PICP. Simply assessing the coverage of the prediction intervals around different degrees of observed confidence allows for this. When the coverage probability and confidence level are closely monitored along the 1:1 line, as shown in Figure 8’s graphs, it is obvious what to expect. According to Jena et al. [3], the probabilities above the 1:1 line show a slight overprediction of the uncertainty range. This may be because of the uncertainty approach. First, the sample density is where the forecast uncertainty comes from. Our study indicated more uncertainty in the region with a high elevation, few sampling locations, and generally dispersed fields. This resulted from the need for further knowledge regarding the connections between the soil and environmental covariates.
While this is happening, capturing the spatial variability of the environmental variables with more accuracy is possible, significantly reducing the deviation of the sample sites, particularly in the region of dramatic change for the environment and soil characteristics [38]. Second, the quality and amount of the environmental variables contribute to the uncertainty. To predict the SOC, we gathered as many environmental variables as we could. The sample density should be increased; useful variables should be captured; spatial modelling techniques should be optimized to reduce uncertainty. The reliability and accuracy of the data products may be increased by minimizing sampling and modelling uncertainty using high-quality sample data and environmental variables, particularly in areas where soils and landscape changes occur quickly.

4. Conclusions

In this work, four tree-based machine-learning algorithms were examined. It calculated the SOC’s spatial distribution patterns in northeastern India’s upper Brahmaputra Valley. This study used optimum models, environmental covariates, and information to determine the spatial distribution of the SOC. A succinct summary of the conclusions is as follows:
  • RF had the best accuracy and the lowest uncertainty for predicting the regional SOC compared to XGBoost, SVM, and Cubist.
  • Compared to XGBoost, SVM, and Cubist, the RF showed higher R2 and RMSE values for predicting SOC based on the validation data.
  • The order of the most-crucial factors in the RF model for predicting the SOC was Elev > MAT > Band 3 > Band1 > MRVBF. The most-crucial variables for utilizing a Cubist model to predict the variance of the SOC were Slp, TRI, MAT, and Band4. The AP and LS were the most-essential factors in the XGBoost and SVM models.
  • The predicted SOC ranged from 0.44 to 1.35%, 0.031 to 1.61%, 0.035 to 1.71%, and 0.47 to 1.36% with the RF, Cubist, XGBoost, and SVM models.
Based on the combination of DSM methodologies with currently available, high-resolution soil-formation environmental data, the results updated the accuracy of the regional variation of the SOC prediction at both the national scale and a detailed level. The SOC maps highlight probable reasons for the predicted uncertainty, which may be used to assess changes in soil quality following extensive cropping and direct further high-resolution DSM research. In general, fine-resolution soil maps are valuable to many soil and environmental professionals and land managers in northeast India. As a result, we advocate employing the comparative technique used in this study area to map the SOC in other parts of India, given that the agroecological zones differ significantly throughout the northeast state of India.

Author Contributions

Software, P.C.M. and R.K.J.; validation, R.K.J. and P.C.M.; formal analysis, D.K.J. and A.A.S.; investigation, A.K. and D.K.J.; resources, P.C.M., K.M.V.K. and S.G.D.; data curation, D.K.J.; writing—A.K., R.K.F. and S.K.M. original draft preparation, A.K., P.C.M., G.K.S. and R.K.J.; reviewing and editing—G.K.S., R.K.F., S.K.M. and A.A.S.; visualization, S.G.D. and K.M.V.K.; supervision, S.G.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Most of the data are available in the tables and figures given in the manuscript.

Acknowledgments

The authors are highly thankful to the respective Directors for supporting the present investigation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schillaci, C.; Acutis, M.; Vesely, F.; Saia, S. A simple pipeline for the assessment of legacy soil datasets: An example and test with soil organic carbon from a highly variable area. Catena 2019, 1, 110–122. [Google Scholar] [CrossRef]
  2. Yigini, Y.; Panagos, P. Assessment of soil organic carbon stocks under future climate and land cover changes in Europe. Sci. Total Environ. 2016, 557–558, 838–850. [Google Scholar] [CrossRef] [PubMed]
  3. Jena, R.K.; Moharana, P.C.; Dharumarajan, S.; Sharma, G.K.; Ray, P.; Deb Roy, P.; Ghosh, D.; Das, B.; Alsuhaibani, A.M.; Gaber, A.; et al. Spatial Prediction of Soil Particle-Size Fractions Using Digital Soil Mapping in the North Eastern Region of India. Land 2023, 12, 1295. [Google Scholar] [CrossRef]
  4. Moharana, P.C.; Meena, R.L.; Nogiya, M.; Jena, R.K.; Sharma, G.K.; Sahoo, S.; Jha, P.K.; Aditi, K.; Vara Prasad, P.V. Impacts of Land Use on Pools and Indices of Soil Organic Carbon and Nitrogen in the Ghaggar Flood Plains of Arid India. Land 2022, 11, 1180. [Google Scholar] [CrossRef]
  5. Taghizadeh-Mehrjardi, R.; Nabiollahi, K.; Kerry, R. Digital mapping of soil organic carbon at multiple depths using different data mining techniques in Baneh region, Iran. Geoderma 2016, 266, 98–110. [Google Scholar] [CrossRef]
  6. Taghizadeh-Mehrjardi, R.; Minasny, B.; Sarmadian, F.; Malone, B.P. Digital mapping of soil salinity in Ardakan region, central Iran. Geoderma 2014, 213, 15–28. [Google Scholar] [CrossRef]
  7. Liu, F.; Zhang, G.-L.; Song, X.; Li, D.; Zhao, Y.; Yang, J.; Wu, H.; Yang, F. High-resolution and three-dimensional mapping of soil texture of China. Geoderma 2020, 361, 114061. [Google Scholar] [CrossRef]
  8. Moharana, P.C.; Dharumarajan, S.; Kumar, N.; Jena, R.K.; Pradhan, U.K.; Meena, R.M.; Sahoo, S.; Kumar, S.; Meena, R.L.; Tailor, B.; et al. Modelling and Prediction of Soil Organic Carbon using Digital Soil Mapping in the Thar Desert Region of India. J. Indian Soc. Soil. Sci. 2022, 70, 86–96. [Google Scholar] [CrossRef]
  9. Hengl, T.; de Jesus, J.M.; MacMillan, R.A.; Batjes, N.H.; Heuvelink, G.B.M.; Ribeiro, E.; Samuel-Rosa, A.; Kempen, B.; Leenaars, J.G.B.; Walsh, M.G.; et al. SoilGrids1km—Global soil information based on automated mapping. PLoS ONE 2014, 9, e105992. [Google Scholar] [CrossRef]
  10. Hengl, T.; Heuvelink, G.B.; Kempen, B.; Leenaars, J.G.B.; Walsh, M.G.; Shepherd, K.D.; Sila, A.; MacMillan, R.A.; de Jesus, J.M.; Tamene, L.; et al. Mapping soil properties of Africa at 250 m resolution: Random forests significantly improve current predictions. PLoS ONE 2015, 10, e0125814. [Google Scholar] [CrossRef]
  11. Arrouays, D.; Grundy, M.G.; Hartemink, A.E.; Hempel, J.W.; Heuvelink, G.B.M.; Hong, S.Y.; Lagacherie, P.; Lelyk, G.; McBratney, A.B.; McKenzie, N.J.; et al. GlobalSoilMap: Toward a fine-resolution global grid of soil properties. Adv. Agron. 2014, 125, 93–134. [Google Scholar]
  12. Dharumarajan, S.; Hegde, R.; Janani, N.; Singh, S.K. The need for digital soil mapping in India. Geoderma Reg. 2019, 16, e00204. [Google Scholar] [CrossRef]
  13. Mishra, G.; Giri, K.; Jangir, A.; Francaviglia, R. Projected trends of soil organic carbon stocks in Meghalaya state of Northeast Himalayas, India. Implications for a policy perspective. Sci. Total Environ. 2020, 698, 134266. [Google Scholar] [CrossRef] [PubMed]
  14. Jigyasu, D.K.; Kumar, A.; Shabnam, A.A.; Sharma, G.K.; Jena, R.K.; Das, B.; Naik, V.S.; Ahmed, S.A.; Kumari, K.M.V. Spatial Distribution of the Fertility Parameters in Sericulture Soil: A Case Study of Dimapur District, Nagaland. Land 2023, 12, 956. [Google Scholar] [CrossRef]
  15. Walkley, A.; Black, I.A. An examination of the Degtjareff method for determining soil organic matter, and a proposed modification of the chromic acid titration method. Soil. Sci. 1934, 37, 29–38. [Google Scholar] [CrossRef]
  16. McBratney, A.B.; Santos, M.M.; Minasny, B. On digital soil mapping. Geoderma 2003, 117, 3–52. [Google Scholar] [CrossRef]
  17. Dokuchaev, V.V. Russian Chernozems (Russkii Chernozems); Kaner, N., Ed.; US Department of Commerce: Springfield, VA, USA, 1976.
  18. Jenny, H. Factors of Soil Formation; McGraw Hill: New York, NY, USA, 1941. [Google Scholar]
  19. R Core Team. R: A Language and Environment for Statistical Computing; R Foundation for Statistical Computing: Vienna, Austria, 2019; Available online: https://www.R-project.org/ (accessed on 5 February 2023).
  20. Liaw, A.; Wiener, M. Classification and regression by randomForest. R News 2002, 2, 18–22. [Google Scholar]
  21. Emadi, M.; Taghizadeh-Mehrjardi, R.; Cherati, A.; Danesh, M.; Mosavi, A.; Scholten, T. Predicting and mapping of soil organic carbon using machine learning algorithms in Northern Iran. Remote Sens. 2020, 12, 2234. [Google Scholar] [CrossRef]
  22. Mahmoudzadeh, H.; Matinfar, H.R.; Taghizadeh-Mehrjardi, R.; Kerry, R. Spatial prediction of soil organic carbon using machine learning techniques in western Iran. Geoderma Reg. 2020, 21, e00260. [Google Scholar] [CrossRef]
  23. Mikkonen, H.G.; van de Graaff, R.; Clarke, B.O.; Dasika, R.; Wallis, C.J.; Reichman, S.M. Geochemical indices and regression tree models for estimation of ambient background concentrations of copper, chromium, nickel and zinc in soil. Chemosphere 2018, 210, 193–203. [Google Scholar] [CrossRef]
  24. Fan, J.; Wang, X.; Wu, L.; Zhou, H.; Zhang, F.; Yu, X.; Lu, X.; Xiang, Y. Comparison of Support Vector Machine and Extreme Gradient Boosting for predicting daily global solar radiation using temperature and precipitation in humid subtropical climates: A case study in China. Energy Convers. Manag. 2018, 164, 102–111. [Google Scholar] [CrossRef]
  25. Zhang, H.T.; Gao, M.X. The Application of Support Vector Machine (SVM) Regression Method in Tunnel Fires. Procedia Eng. 2018, 211, 1004–1011. [Google Scholar] [CrossRef]
  26. Lagacherie, P.; Arrouays, D.; Bourennane, H.; Gomez, C.; Martin, M.; Saby, N.P.A. How far can the uncertainty on a Digital Soil Map be known?: A numerical experiment using pseudo values of clay content obtained from Vis-SWIR hyperspectral imagery. Geoderma 2019, 337, 1320–1328. [Google Scholar] [CrossRef]
  27. Dharumarajan, S.; Kalaiselvi, B.; Suputhra, A.; Lalitha, M.; Vasundhara, R.; Anil Kumar, K.S.; Nair, K.M.; Hegde, R.; Singh, S.K.; Lagacherie, P. Digital soil mapping of soil organic carbon stocks in Western Ghats, South India. Geoderma Reg. 2021, 25, e00387. [Google Scholar] [CrossRef]
  28. Solomatine, D.P.; Shrestha, D.L. A novel method to estimate model uncertainty using machine learning techniques. Water Resour. Res. 2009, 45, W00B11. [Google Scholar] [CrossRef]
  29. Malone, B.P.; McBratney, A.B.; Minasny, B. Empirical estimates of uncertainty for mapping continuous depth functions of soil attributes. Geoderma 2011, 160, 614–626. [Google Scholar] [CrossRef]
  30. Jena, R.K.; Bandyopadhyay, S.; Pradhan, U.K.; Moharana, P.C.; Kumar, N.; Sharma, G.K.; Roy, P.D.; Ghosh, D.; Ray, P.; Padua, S.; et al. Geospatial Modelling for Delineation of Crop Management Zones Using Local Terrain Attributes and Soil Properties. Remote Sens. 2022, 14, 2101. [Google Scholar] [CrossRef]
  31. Lamichhane, S.; Adhikari, K.; Kumar, L. Use of multi-seasonal satellite images to predict SOC from cultivated lands in a Montane ecosystem. Remote Sens. 2021, 13, 4772. [Google Scholar] [CrossRef]
  32. Falahatkar, S.; Hosseini, S.M.; Ayoubi, S.; Salmanmahiny, A. Predicting soil organic carbon density using auxiliary environmental variables in northern Iran. Arch. Agron. Soil. Sci. 2016, 62, 375–393. [Google Scholar] [CrossRef]
  33. da Silva Chagas, S.; de Carvalho Junior, W.; Bhering, S.B.; Calderano Filho, B. Spatial prediction of soil surface texture in a semiarid region using random forest and multiple linear regressions. Catena 2016, 139, 232–240. [Google Scholar] [CrossRef]
  34. Akpa, S.I.; Odeh, I.O.; Bishop, T.F.; Hartemink, A.E. Digital mapping of soil particle-size fractions for Nigeria. Soil. Sci. Soc. Am. J. 2014, 78, 1953–1966. [Google Scholar] [CrossRef]
  35. Pahlavan-Rad, M.R.; Akbarimoghaddam, A. Spatial variability of soil texture fractions and pH in a flood plain (case study from eastern Iran). Catena 2018, 160, 275–281. [Google Scholar] [CrossRef]
  36. Sahoo, U.K.; Singh, S.L.; Gogoi, A.; Kenye, A.; Sahoo, S.S. Active and passive soil organic carbon pools as affected by different land use types in Mizoram, Northeast India. PLoS ONE 2019, 14, e0219969. [Google Scholar] [CrossRef] [PubMed]
  37. Pahlavan-Rad, M.R.; Dahmardeh, K.; Hadizadeh, M.; Keykha, G.; Mohammadnia, N.; Gangali, M.; Keikha, M.; Davatgar, N.; Brungard, C. Prediction of soil water infiltration using multiple linear regression and random forest in a dry flood plain, eastern Iran. Catena 2020, 194, 104715. [Google Scholar] [CrossRef]
  38. Liang, Z.; Chen, S.; Yang, Y.; Zhao, R.; Shi, Z.; Rossel, R.A.V. National digital soil map of organic matter in topsoil and its associated uncertainty in 1980’s China. Geoderma 2019, 335, 47–56. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions, and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions, or products referred to in the content.
Figure 1. Spatial distribution of soil sampling points in the study area of the Upper Brahmaputra Valley in northeast India.
Figure 1. Spatial distribution of soil sampling points in the study area of the Upper Brahmaputra Valley in northeast India.
Land 12 01841 g001
Figure 2. Soil Organic Carbon (SOC) content distribution in the calibration and validation dataset is represented by the histogram, density plot, and rug plot. Vertical dashed lines represent the mean values.
Figure 2. Soil Organic Carbon (SOC) content distribution in the calibration and validation dataset is represented by the histogram, density plot, and rug plot. Vertical dashed lines represent the mean values.
Land 12 01841 g002
Figure 3. Correlation analysis between covariates and SOC in the study area of Upper Brahmaputra Valley in northeast India. Abbreviations for the variables are listed in Table 1. ** Correlation is significant at the 0.01 level. * Correlation is significant at the 0.05 level.
Figure 3. Correlation analysis between covariates and SOC in the study area of Upper Brahmaputra Valley in northeast India. Abbreviations for the variables are listed in Table 1. ** Correlation is significant at the 0.01 level. * Correlation is significant at the 0.05 level.
Land 12 01841 g003
Figure 4. Predicted vs. observed parameters of soil organic carbon by various models in the study area of Upper Brahmaputra Valley in the northeast state of India.
Figure 4. Predicted vs. observed parameters of soil organic carbon by various models in the study area of Upper Brahmaputra Valley in the northeast state of India.
Land 12 01841 g004
Figure 5. Importance variables in predicting soil organic carbon by various models in the study area of Upper Brahmaputra Valley in northeast India.
Figure 5. Importance variables in predicting soil organic carbon by various models in the study area of Upper Brahmaputra Valley in northeast India.
Land 12 01841 g005
Figure 6. Distribution of soil organic carbon predicted by various model in the study area of Upper Brahmaputra Valley in northeast India.
Figure 6. Distribution of soil organic carbon predicted by various model in the study area of Upper Brahmaputra Valley in northeast India.
Land 12 01841 g006
Figure 7. Distribution of lower and upper limits of 90% prediction interval of soil organic carbon predicted by the best model (RF) in the study area of Upper Brahmaputra Valley in northeast India.
Figure 7. Distribution of lower and upper limits of 90% prediction interval of soil organic carbon predicted by the best model (RF) in the study area of Upper Brahmaputra Valley in northeast India.
Land 12 01841 g007
Figure 8. Based on the validation of the RF model for soil organic carbon between the Prediction Interval Coverage Probability (PICP) and confidence level (CI).
Figure 8. Based on the validation of the RF model for soil organic carbon between the Prediction Interval Coverage Probability (PICP) and confidence level (CI).
Land 12 01841 g008
Table 1. Different environmental covariates used in the models.
Table 1. Different environmental covariates used in the models.
GroupsPredictorAbbreviationResolutionDescription
Terrain indicesElevation (m)Elev30 mVertical distance above sea level
SlopeSlp30 mInclination of the land surface from the horizontal
Relative Slope PositionRSP Relative slope position
Topographic Wetness IndexTWI30 mRatio of local catchment area to slope
Multi-Resolution index of Valley Bottom FlatnessMRVBF30 mMeasure of flatness and lowness
Valley DepthVD30 mRelative position of the valley
Channel Network Base LevelCNBL30 mCalculates the distance to a channel network base level
Channel Network DistanceCND30 mCalculates the distance to a channel network
Spectral indicesNormalized Difference Vegetation IndexNDVI30 mAmount of vegetation
Landsat data (11 bands)Band1–1130 mLandsat OLI spectral band
ClimateAnnual PrecipitationAP1 kmBioclimatic variables (BIO1)
Mean Annual Temperature MAT1 kmBioclimatic variables (BIO12)
Table 2. Hyperparameters of machine learning algorithms used in this study.
Table 2. Hyperparameters of machine learning algorithms used in this study.
ML AlgorithmsRandom Forest (RF)Cubist (Regression Tree)Extreme Gradient Boosting (XGBoost)Support Vector Machine (SVM)
HyperparametersMtryNtreeCommitteesNeighborsBooster Max_DepthMin_Child_WeightColsample_BytreeSubsampleetaKernel TypeCσ
Defined Parameters1–30100–30001–1000–9gbtree3–100–50.5–10.5–10.01–0.5RBF0.01–1000.01–100
Definitionthe number of input variablesthe number of treesthe number of model treesthe number of nearest neighborsthe type of modelthe depth of the treethe minimum sum of weights of all observationsthe number of variables supplied to a treethe number of samples supplied to a treelearning ratethe kernel functionthe penalty parameterthe bandwidth parameter
Table 3. Descriptive statistics of soil organic carbon in the study area of Upper Brahmaputra Valley in northeast India.
Table 3. Descriptive statistics of soil organic carbon in the study area of Upper Brahmaputra Valley in northeast India.
DatasetnMinMaxMeanSEMedianSDCVSkewnessKurtosis
Total1600.101.850.810.030.750.3847.360.28−0.66
Calibration1280.101.540.810.030.810.3745.780.04−0.88
Validation320.291.850.780.070.630.4254.371.050.27
SD, Standard Deviation; SE, Standard Error; CV, Coefficient of Variation.
Table 4. Descriptive statistics of environmental covariates used in the models.
Table 4. Descriptive statistics of environmental covariates used in the models.
CovariatesMinMaxMeanMedianSDCVSkewnessKurtosis
Asp0.006.283.183.141.8658.32−0.02−1.09
CI−47.1851.215.393.8414.96277.650.331.46
CNBL70.75109.9782.4379.618.5910.421.180.67
CND0.0022.517.035.985.3776.440.750.02
Elev59.67116.4088.5386.1311.1712.620.34−0.46
LS-Factor0.029.902.672.561.6260.831.313.54
MRRTF0.004.420.650.330.83128.102.064.23
MRVBF0.005.800.730.301.10150.042.516.26
NDVI0.080.460.370.370.0514.19−1.154.44
RSP0.000.430.110.100.0983.051.171.26
Slp0.000.260.100.100.0550.570.310.11
TCA9594,892,34078,6382317522,500664871
TPI−11.1910.490.851.262.98352.22−0.391.59
TRI1.088.403.233.081.2237.731.072.27
VD18.95100.1559.4559.2017.6429.680.09−0.63
Band110,693.9014,115.7011,112.3111,038.65355.063.204.0131.11
Band29613.7413,974.5010,121.2710,015.80448.334.434.2233.29
Band38952.7214,430.609615.189455.46554.815.774.3834.80
Band47539.6015,172.708488.408311.01796.629.384.0230.28
Band514,978.7021,984.6018,392.1518,427.751178.436.410.100.37
Band611,266.5020,536.8013,582.2113,334.001183.538.711.797.28
Band77657.7318,707.309506.829174.761252.4313.172.9017.33
Band88331.0114,683.009091.328789.09739.248.133.1919.53
Band95024.325068.925048.365048.277.030.140.000.93
Band1027,930.4030,232.7028,404.282343.90268.270.942.5613.60
Band1124,840.8026,183.8025,195.6625,178.80178.820.711.666.24
AP2554.713253.063019.873126.40222.677.37−0.59−1.18
MAT23.6524.1523.9823.950.100.43−0.24−0.41
SD, Standard Deviation; CV, Coefficient of Variation.
Table 5. Performance of various models for predicting soil carbon in the study area of Upper Brahmaputra Valley in northeast India.
Table 5. Performance of various models for predicting soil carbon in the study area of Upper Brahmaputra Valley in northeast India.
ModelCalibration Validation
R2cCCCcRMSEcMEcR2vCCCvRMSEvMEv
RF0.9660.8630.1590.0010.4180.5490.3770.136
Cubist0.3960.5710.2910.0390.2300.3140.4850.062
SVM0.4710.4530.2930.0150.0810.1750.4520.049
XGBoost0.9980.9900.0220.0000.1520.1900.4240.054
R2: coefficient of determination, ME: Mean Error, RMSE: Root-Mean-Squared Error, CCC: Lin’s Concordance Correlation Coefficient.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kumar, A.; Moharana, P.C.; Jena, R.K.; Malyan, S.K.; Sharma, G.K.; Fagodiya, R.K.; Shabnam, A.A.; Jigyasu, D.K.; Kumari, K.M.V.; Doss, S.G. Digital Mapping of Soil Organic Carbon Using Machine Learning Algorithms in the Upper Brahmaputra Valley of Northeastern India. Land 2023, 12, 1841. https://doi.org/10.3390/land12101841

AMA Style

Kumar A, Moharana PC, Jena RK, Malyan SK, Sharma GK, Fagodiya RK, Shabnam AA, Jigyasu DK, Kumari KMV, Doss SG. Digital Mapping of Soil Organic Carbon Using Machine Learning Algorithms in the Upper Brahmaputra Valley of Northeastern India. Land. 2023; 12(10):1841. https://doi.org/10.3390/land12101841

Chicago/Turabian Style

Kumar, Amit, Pravash Chandra Moharana, Roomesh Kumar Jena, Sandeep Kumar Malyan, Gulshan Kumar Sharma, Ram Kishor Fagodiya, Aftab Ahmad Shabnam, Dharmendra Kumar Jigyasu, Kasthala Mary Vijaya Kumari, and Subramanian Gandhi Doss. 2023. "Digital Mapping of Soil Organic Carbon Using Machine Learning Algorithms in the Upper Brahmaputra Valley of Northeastern India" Land 12, no. 10: 1841. https://doi.org/10.3390/land12101841

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