Home Inversion method of organic matter content of different types of soils in black soil area based on hyperspectral indices
Article Open Access

Inversion method of organic matter content of different types of soils in black soil area based on hyperspectral indices

  • Nan Lin , Yanlong Liu , Qiang Liu EMAIL logo , Ranzhe Jiang and Xunhu Ma
Published/Copyright: December 24, 2024
Become an author with De Gruyter Brill

Abstract

Soil organic matter content (SOMC) is a key factor in improving the soil fertility of arable land. Determining how to quickly and accurately grasp SOMC on a regional scale has become an important task for farmland quality monitoring. Hyperspectral imaging remote sensing technology can enable large-scale SOMC estimation, owing to its large-scale and fine spectral resolution. Enhancing the accuracy and reliability of SOM estimation models based on hyperspectral satellite remote sensing has emerged as a prominent topic of study. In this study, feature spectral indices such as difference indices (DI), ratio indices, and normalized indices were extracted using the correlation coefficient method and used as variables to construct a regression model for SOM, with a split-sample regression method employed to account for the complexity of soil types and map the corresponding spatial distribution of SOM. The results showed that the SOM estimation model, built using these feature spectral indices from hyperspectral satellite imagery, achieved high predictive accuracy, with R² values approaching 0.80 for most soil types. This demonstrates that the model effectively captures variations in SOM content across diverse soil backgrounds, highlighting its robustness and adaptability. The DI499/576 combinations, in particular, contributed significantly to prediction accuracy, demonstrating their importance as key spectral parameters for SOM estimation. Furthermore, among the three sets of feature model variables derived from the split-sample regression strategy, the enhanced vegetation indices and Soil-Adjusted Total Vegetation Index exhibited distinct contributions to different soil sample groups. This variation reveals the specific responsiveness of these indices to soil properties, which further enhances model performance in varied soil contexts. This study provides innovative methods for large-scale SOMC estimation, particularly by utilizing hyperspectral indices to enhance model accuracy across various soil types, demonstrating substantial practical significance.

1 Introduction

Soil is an indispensable environmental element, and soil quality and health play important roles in maintaining the security of agro-ecosystems. Changes in soil fertility can affect crop growth and sustainable land use [1]. Soil organic matter (SOM) refers to a variety of organic components, such as plant and animal residues, microorganisms, root secretions, and some highly stabilized humus deposited in the soil, and is an important component of soil [2]. Over the past decades, the black soil in Northeast China has experienced intensive mechanized farming and unbalanced inputs of exogenous fertilizers and nutrients, which have significantly reduced the carbon sequestration potential of cropland soils [3]. Cultivated farmland faces a series of soil degradation processes, such as the reduction in organic matter, severe carbon and nitrogen imbalances, and destabilization of microbial communities [4]. For this reason, many researchers have monitored the organic matter content in farmland to provide standardized and scientific references for agricultural production to maintain the stability of the soil ecosystem [5]. Rapid and accurate estimation of SOM in agricultural areas is of great significance for ensuring food security and promoting the digital and sustainable development of agriculture.

While traditional chemical analysis methods face time-consuming and laborious sampling processes [6], hyperspectral imaging provides a scalable, efficient and non-destructive alternative for large-scale soil monitoring, effectively solving the problems of cumbersome sampling processes and long laboratory analysis cycles. It offers the advantages of high efficiency, real-time analysis, and non-polluting characteristics and is widely used in the rapid estimation of the physical and chemical properties of soil [7,8]. However, this advantage is accompanied by problems such as high autocorrelation and low information density in the hyperspectral image bands. To some extent, these issues contribute to the instability of the estimation model [9]. Shi et al. [10] demonstrated that redundant spectral variables can reduce model accuracy, as evidenced by their comparison of full-band and feature-band models. To address the issue of redundancy in hyperspectral data and information, feature selection algorithms such as the correlation coefficient thresholding method and successive projection algorithm are utilized to select appropriate spectral feature bands that can effectively extract feature information and accelerate the training speed [11,12]. Yu et al. [13] extracted significant bands by analyzing the correlation coefficients to establish a quantitative inversion model for SOM and achieved high estimation accuracy. Ding et al. [14] screened spectral features using three feature selection algorithms, Ant Colony Optimization-interval Partial Least Squares, Recursive Feature Elimination-Support Vector Machine (SVM), and Random Forest (RF), to determine the characteristic wavelength range of the SOM. However, the spectral reflectance of soil in the visible and near-infrared wavelength ranges is usually low, the absorption characteristics are not significant, and they are easily affected by environmental factors, such as moisture, temperature, and atmosphere [15]. Therefore, the SOM inversion model constructed based on the characteristic wavelength bands suffers from problems of instability and difficulty in improving the accuracy.

The spectral composition of radiative fluxes at the Earth’s surface reflects the physical properties of soil, moisture, vegetation, etc., in physical and climatic environments, exhibiting consistent spectral response patterns [16]. This analysis aims to establish an empirical or semi-empirical, non-contact, qualitative, or quantitative measurement of features. Compared with single-band information, spectral indices, with their clear physical meaning and concise expressions, have proven to be effective in realizing remote sensing evaluation and monitoring of surface cover information on a large spatial scale, which provides a reference for the estimation of soil properties on a large spatial scale using remote sensing technology [17,18]. When using satellite image inversion to predict soil components, the soil reflectance is significantly affected by soil conditions (e.g., soil moisture and roughness) and soil surface conditions (e.g., crop and crop residue cover), thus reducing the prediction accuracy of soil properties [19]. It has been shown that using spectral indices to estimate soil constitution helps mitigate the interference of environmental factors such as moisture and atmosphere [20]. The spectral index contains spectral information that is more sensitive than single-band spectra, leading to more accurate results in the statistical analysis [21,22]. Guo et al. [23] successfully predicted farmland SOMD using vegetation indices derived from Sentinel-2 and Landsat-8 images. He et al. [24] generated 17 phenological parameters and accurately predicted surface SOM content using a RF algorithm in combination with 15 Sentinel-2-based spectral indices. These include a vegetation index, brightness correlation index, and humidity index, which demonstrate the high predictability of SOM based on spectral indices.

In addition, the spatial heterogeneity of the soil affects estimation accuracy to a certain extent. It is important to separately construct regression models according to soil types to improve the accuracy of the hyperspectral SOM estimation model [25]. Soil spectral reflectance is an integrated response to the spectral behavior of the inherent physical and chemical properties of the soil [26]. Differences in the physical and chemical properties of various soil types lead to distinct spectral characteristics [27]. Liu et al. [28] found that the differences in the spectral characteristics between black soil and meadow soil were mainly expressed in five spectral absorption valleys. The first two absorption valleys were mainly caused by the organic matter, iron content, and mechanical composition of the soil, whereas the last three absorption valleys were caused by the soil moisture. This result suggests that different types of soil exhibit significant variations in their reflectance spectra. These differences complicate the model-building process, thereby reducing the prediction accuracy and reliability. Xiao et al. [29] identified spectral bands that are sensitive to nitrogen content across various soil types, significantly enhancing the accuracy of nitrogen estimation. Liu et al. [30] utilized visible near-infrared spectroscopy to differentiate between soil types in the samples. They used a hierarchical calibration strategy to increase the estimation accuracy of the soil organic carbon content. Therefore, considering the input variables of different soil types when constructing an SOM estimation model and establishing a split-sample regression model for Soil organic matter content (SOMC) can enhance the model’s accuracy and robustness across various soil types.

The aim of this study is to propose a method for constructing a SOM estimation model using hyperspectral indices, while considering different soil types to perform sub-sample regression to improve the estimation accuracy. Using ZY1-02D hyperspectral imagery as the data source, Pearson correlation coefficients were applied to extract characteristic spectral indices. The differences in spectral indices and model accuracy between the full sample and different soil types were analyzed, highlighting the impact of split-sample regression on model precision. Based on these results, a spatial distribution map of SOM was generated. The main contributions of this study are as follows: (1) The application of spectral index in monitoring SOMC was investigated, and the characteristic band combinations were extracted; (2) to study the differences in eigen-spectral indices under split-sample regression as well as the contribution and impact on the estimation model; and (3) to plot the distribution of SOMC in the study area and validate the accuracy of the split-sample regression model constructed using hyperspectral indices.

2 Materials and methods

2.1 Study area

The study area is shown in Figure 1. This area, situated in the black-soil region in and around Sihe Town, Yushu City, Jilin Province, experiences a temperate continental monsoon climate. The winter climate is cold, long, and snowy, whereas the summer rainfall is abundant. The average annual temperature is 5.3°C and the average annual precipitation is approximately 580 mm. The study area is situated in the hinterland of the Songliao Plain. The terrain within the area is relatively flat, with slightly undulating undulations and an average elevation of approximately 186 m. The region is rich in natural resources, including ample surface water resources, fertile soil, and a wide range of soils, mainly black soils, meadow soils, and albic soils, which are suitable for its growth, making it an important black soil planting area in China. Crops in the area are harvested once a year, with corn and rice being the main crops. Corn planting area accounts for >70% of the total cultivated area in the study area, which is a unique condition for the development of agricultural production.

Figure 1 
                  Location map of the study area and photographs of three selected sampling sites (a) Jilin Province, (b) part of Yushu, (c) sampling area, and (d) sampling sites within the quadrangle (five-point method).
Figure 1

Location map of the study area and photographs of three selected sampling sites (a) Jilin Province, (b) part of Yushu, (c) sampling area, and (d) sampling sites within the quadrangle (five-point method).

2.2 Sample collection and analysis

The distribution of agricultural croplands in the study area was investigated. Sampling points and routes were predetermined at regular intervals of 500 m in each direction, considering the available map data on soil and land use types. The samples were collected in mid-April 2022, and all collection sites were agricultural fields. Figure 1 shows the locations of the 518 sample plots. Each sample plot was 30 m × 30 m in size. Five soil samples were collected from each plot, containing three soil types: black soil, albic soil, and meadow soil. During the data collection period, the crops were harvested; there were no crops or weeds on the surface, and only a small amount of straw remained in some of the fields, which facilitated the collection of soil samples for analysis. In this study, a five-point sampling method was used to collect 500 g of soil at a surface depth of 5 cm [31]. The distance between the sampling point and other features (roads, buildings, etc.) was ensured to be >100 m to ensure that the image pixel at the sampling point was a “pure soil pixel,” meaning the pixel predominantly contained soil with minimal interference from other surface features, such as vegetation, roads, or water bodies [32]. Simultaneously, a handheld GPS locator was used to record the geographic coordinates of sampling points. After the soil samples were collected, they were placed in an oven for 48 h at 60°C to remove fine root impurities and ground [33]. The soil samples were subjected to indoor spectral measurements and organic matter content determination after being sieved through a 0.15 mm sieve [34]. In this experiment, the spectral reflectance of the soil samples was measured using an ASD FieldSpec 4 spectrometer with a wavelength range of 350–2,500 nm, spectral resolution of 3 nm, and wavelength accuracy of less than 0.5 nm [35]. To avoid measurement errors caused by the light source, all samples were measured in a dark room, and a standard whiteboard was used for calibration to reduce the noise and maintain a consistent distance of 3 cm between the probe and the surface of the sample. The measurement was repeated ten times for each soil sample, with an interval of 1 s between each measurement. The final results of the ten measurements were averaged to obtain the measured spectral data on the ground. The SOMC of the soil samples was determined using an external heating method with potassium dichromate-concentrated sulfuric acid. This method is accurate and suitable for the analysis of a large number of samples because of the non-interfering effects of carbonates in the soil [36].

2.3 Data preprocessing

The ZY1-02D hyperspectral image data used in this study were provided by the China Center for Resources Satellite Data and Application and contained hyperspectral data in 166 spectral bands with a spectral coverage of 400–2,500 nm. For this study, the satellite image acquired on 26 April 2022 during a cloud-free observation window was selected to ensure synchronization with soil sample collection. During the period from sample collection to imaging, there was no agricultural production activity in the study area, and the change in SOMC under natural conditions was extremely weak, which did not affect the sample collection. The streaking phenomenon is evident in the ZY1-02D hyperspectral short-wave infrared (SWIR) data; therefore, the global debanding method is used to repair the streaks. For the effect of water vapor in the atmosphere, the satellite hyperspectral images are removed from the water vapor absorption bands at 1,350–1,450 nm and 1,800–1,900 nm wavelength range of the band that is seriously affected by water vapor, as well as the edge spectral channels containing a large amount of noise data before 400 nm and after 2,470 nm wavelengths [37]. A total of 21 spectral bands were cumulatively eliminated and the processed hyperspectral image contained 145 bands. The processed hyperspectral image contained 145 bands. Radiometric calibration was performed according to the satellite calibration file and the FLAASH model was used to correct the atmosphere of the radiometrically calibrated ZY1-02D hyperspectral data. To accurately identify the study area, the SVM classification method was used to supervise the classification of non-agricultural land parcels, and cloud shadows, roads, and residential areas in the images were eliminated (Figure 2).

Figure 2 
                  SVM method for supervised classification extracted bare soil image element results.
Figure 2

SVM method for supervised classification extracted bare soil image element results.

2.4 Selection and construction of spectral index

Difference indices (DI), normalized indices (NDI), and ratio indices (RI) are commonly used data processing and analysis methods that are widely applied to extract relevant surface information [38]. The DI characterizes the spectral difference by calculating the difference in reflectance between two spectral bands; the NDI is used to normalize the difference between two bands; and the RI characterizes the relative difference between bands by calculating the ratio of the reflectance of two bands. These indices help highlight the spectral differences of specific features and can effectively eliminate the differences caused by light, sensors, atmosphere, and other factors, making the changes in feature characteristics more obvious, thus enhancing the interpretation and analysis of remote sensing data [39]. The calculation method of DI can be used to highlight the features of different features, RI can strengthen the reflectance differences of features in certain bands, and NDI can reduce scale differences. The formulae are as follows:

(1) DI = R i R j ,

(2) NDI = ( R i R j ) / ( R i + R j ) ,

(3) RI = R i / R j .

Most currently published spectral indices are derived from the physical characteristic laws of spectral reflectance, including the vibrations and combinations of reactive groups such as C–H groups, C–N groups, C═O groups, and –OH groups, and electromagnetic activity resulting from electron jumps occurring in the VN-SWIR spectral interval [40]. These spectral indices consider the interactions between electromagnetic radiation, atmosphere, vegetation cover, and soil background [41]. These have been gradually improved and developed through continuous improvements in mathematical simulation statistics. For example, Soil Adjusted Total Vegetation Index (SATVI) detects the reflectance of vigorous and senescent vegetation through the ratio of red and two SWIR radiations. In combination with the soil correction factor, this can reflect the nature of the soil in densely vegetated areas [42]. Brightness Index (BI) explains the relationship between soil brightness and remote sensing data through the mathematical operation of reflectance in the two visible bands and is more sensitive to the texture of the soil [43]. However, the relationship between spectral indices and organic matter content in bare soil areas has not yet been studied in detail. To investigate the potential of this remote sensing index in predicting the organic matter content of bare soil areas, this study was designed to develop two types of spectral indices: the vegetation index and the luminance correlation index. These indices were based on six types of spectral indices, using the waveband of ZY1-02D. The spectral index calculation formula and the corresponding waveband channels are presented in Table 1.

Table 1

Spectral index calculation formula

Classes Abbreviation Remote sensing predictors Formula based on ZY1-02D Reference
Vegetation index TVI Transformed Vegetation Index TVI = R 700-1,100 - R 600-700 R 700-1,100 - R 600-700 + 0 .5 1/2 × 100 Liu et al. [44]
SATVI Soil Adjusted Total Vegetation Index SATVI = R 1,550 1,750 R 630 690 R 1,550 1,750 + R 630 690 + 1 × ( 1 + 1 ) R 2,080 2,350 2 Villarreal et al. [45]
EVI Enhanced Vegetation Index EVI = 2 .5 × R 700 1,100 R 600 700 R 700 1,100 + 6 × R 600 700 7.52 × R 450 520 + 1 Liu et al. [44]
Bright-related index BI Brightness Index BI = ( R 645 683 × R 645 683 ) + ( R 537 582 × R 537 582 ) 2 Escadafal [46]
BI 2 The second Brightness Index BI2 = ( R 645 683 × R 645 683 ) + ( R 537 582 × R 537 582 ) + ( R 743 928 × R 743 928 ) 2 Escadafal [46]
RI Redness Index RI = R 645 683 × R 645 683 R 448 537 × R 537 582 × R 537 582 × R 537 582 Mahmood et al. [47]

The Pearson correlation coefficient (PCC) is a statistical measure that quantifies the strength of the linear relationship between two variables, with values ranging from −1 to 1 [48]. For two-band indices, calculating the correlation coefficients between the spectral indices of the sample set and SOMC is the most commonly used and effective method for extracting the characteristic spectral indices. In this study, two-band indices were calculated based on ZY1-02D hyperspectral remote sensing images. The spectral indices DI, NDI, and RI of any two bands R i and R j were calculated by traversing the full-band soil spectral data from 400 to 2,500 nm. The combination of TVI and BI was restricted within the band channel range. PCCs between various indices and SOM were calculated, and a correlation coefficient matrix was constructed to analyze the characteristic spectral indices of SOM. Most existing studies on SOM estimation using spectral indices use two-band spectral indices for environmental modeling and attribute quantification. However, only a limited number of studies have explored the potential of three-band spectral indices to estimate soil attributes. For the three-band indices, the spatial scatter plots of the correlation coefficients between the three-band index and SOM were generated by calculating the combinations of spectral indices within the three spectral band channels [α1n], [β1n], and [γ1n], which correspond to any three bands Rα, Rβ, and Rγ. The methodological workflow of this study is shown in Figure 3.

Figure 3 
                  Flowchart of research methods.
Figure 3

Flowchart of research methods.

2.5 Model establishment and accuracy evaluation

In this study, the XGBoost algorithm was used to build the SOM prediction model. XGBoost is an efficient and scalable gradient boosting algorithm [49]. It is based on the concept of boosting trees, which integrates multiple weak learners (typically decision trees) to create a strong learner, thereby enhancing the model’s predictive performance. Each new tree is built by iteratively reducing prediction errors, allowing the model to better correct the mistakes of the previous models. To prevent overfitting, XGBoost incorporates regularization terms (L1 and L2 regularization) during the construction of each tree, effectively controlling the model’s complexity. After the model creates the booster tree, it is relatively straightforward to obtain the importance score for each attribute [50]. In general, the importance score measures the value of a feature during the construction of a boosting decision tree in the model. The more an attribute is used in the construction of the decision tree in the model, the higher its relative importance. The importance was calculated in a single decision tree based on the amount of improvement in the performance metrics by each attribute split point. Nodes are responsible for evaluating and recording the frequency with which an attribute carries more weight to enhance the split-point performance metric. The importance of an attribute increases as it is selected more frequently by decision trees. The evaluation metrics selected for this study were the coefficient of determination (R 2), normalized root mean square error (NRMSE), and Residual Predictive Deviation (RPD). The expressions for the above evaluation metrics are as follows:

(4) R 2 = 1 i = 1 n ( y i y ˆ i ) 2 i = 1 n ( y ¯ i y i ) 2 ,

(5) RMSE = 1 n i = 1 n ( y i y ˆ i ) 2 ,

(6) NRMSE = RMSE y Max y Min ,

(7) RPD = 1 n 1 i = 1 n ( y i y ˆ i ) 2 RMSE ,

where n denotes the number of samples in the dataset, y i is the true value, y ˆ i is the predicted value, and y ¯ i is the average of the true values.

3 Results

3.1 Soil reflectance spectral characteristics

To analyze the spectral reflectance characteristics of different soils in the study area, the ground spectra of 518 soil samples collected were statistically analyzed by grouping them according to soil type. Subsequently, the measured indoor spectra and satellite spectra were compared separately. Figure 4(a) shows the average spectral reflectance of three soils with the same organic matter content. Overall, the spectral reflectance of meadow soil is higher, while the reflectances of black soil and albic soil are similar in the visible region; however, after 1,000 nm, the reflectance of black soil exceeds that of albic soil. In the near-infrared region, there are significant differences in the spectral reflectance of the three soils. The average spectral reflectances of the black soil, albic soil, and meadow soil at different SOMCs are represented by b, c, and d in Figure 4. In the wavelength range of 400–2,500 nm, the overall soil reflectance exhibited a decreasing trend with increasing SOMC. It is noteworthy that the differences in reflectance among the three groups become more pronounced in the NIR region, especially beyond 1,000 nm, where the reflectance of meadow soils is more affected by SOMC. A comprehensive comparison of the three sets of spectral data revealed that although the ZY1-02D hyperspectral data were affected by conditions such as soil roughness and soil moisture, the change rule with the indoor measured spectra was generally aligned with that of the pixel spectral reflectance in a specific wavelength range, with the variation in SOMC being more significant.

Figure 4 
                  Comparison of average spectral curves of three soils under the same SOMC and different SOMC conditions.
Figure 4

Comparison of average spectral curves of three soils under the same SOMC and different SOMC conditions.

To elucidate the relationship between the satellite spectra and indoor measured spectra, the correlation coefficients of the two spectra of the three soil-type samples with SOM were calculated. As shown in Figure 5, the spectral correlation coefficient curves of the three soil types indicate that SOMC primarily affects the VNIR spectral range, with the overall correlation ranked from high to low as black, meadow, and albic soils. However, in the range of 460–1,000 nm, the correlation of albic soil began to surpass that of meadow soil. The correlation coefficients of the three soil types in the visible region exceeded 0.4, indicating a strong correlation. This demonstrated that the spectrally corrected image element spectra retained the majority of soil spectral features and could be effectively used for SOM estimation.

Figure 5 
                  Correlation between reflectance of two spectra (measured spectrum and ZY1-02D spectrum) and SOMC.
Figure 5

Correlation between reflectance of two spectra (measured spectrum and ZY1-02D spectrum) and SOMC.

3.2 Spectral indices

To accurately extract the wavelength combinations with the highest correlation between SOM and dual-band indices, this study was designed to compute five spectral indices (DI, NDI, RI, TVI, and BI) for the combinations of two bands in the range of 450–2,500 nm, and to analyze the correlation between them and SOM. Figure 6(a)–(e) shows the five dual-band index and SOM correlation heat maps. The results indicated that the two strongest correlations of the band were primarily found in the visible and near-infrared wavelengths and were predominant at 1,000 nm or below. The high correlation with SOM in DI is the combination of the 400–580 nm and 580–860 nm bands, NDI is the combination of the 490–540 nm and 550–590 nm bands, and RI is the 490–540 nm and 550–580 nm. Although the correlation between TVI and SOM was generally low, the index still provided information that could reflect the nature of SOM in the spectral combination of 700–800 nm and 600–700 nm. To make full use of bands rich in characteristic information in the images, this study selected spectral indices of the band combinations with correlation coefficients >0.5 in the satellite spectral combinations for analysis. In total, 56 DIs, 40 NDIs, 36 RIs, 33 TVIs, and 15 BIs were screened.

Figure 6 
                  Heat map of correlation between DI, NDI, RI, and SOM. (a) DI, (b) NDI, (c) RI, (d) TVI, and (e) BI.
Figure 6

Heat map of correlation between DI, NDI, RI, and SOM. (a) DI, (b) NDI, (c) RI, (d) TVI, and (e) BI.

A scatter plot of the correlation between the four types of three-band indices and the organic matter content is shown in Figure 7. The spectral band combinations with higher correlations between SATVI and SOM were mainly concentrated in the bands of 630–690 nm, 1,560–1,700 nm, and 2,080–2,300 nm, and the band regions with higher correlation of EVI were 470–510 nm, 600–640 nm, and 840–1,020 nm. Owing to the broad spectral bands of SATVI and EVI, there were up to 2,000 spectral band combinations. To minimize information redundancy, only the band combinations with correlation coefficients >0.5 were selected, resulting in 24 and 27 band combinations for SATVI and EVI, respectively. The spectral combinations of The Second Brightness Index (BI2) and Redness Index (RI) were generally moderately correlated with organic matter content, with a few combinations having weak correlations, and a total of 24 and 18 band combinations with correlations >0.5 were selected.

Figure 7 
                  Degree of correlation between four three-band indices and SOMC.
Figure 7

Degree of correlation between four three-band indices and SOMC.

3.3 Model construction for split-sample strategy

Considering the information redundancy of hyperspectral indices, XGBoost ranked the importance of the 273 spectral index features in the modeling. Feature variables were iteratively selected in descending order of importance, with optimal variable selection determined using K-fold cross-validation to obtain the model with the lowest NRMSE. Figure 8 shows the model NRMSE for all types of soil samples when cross-validation was used to select the optimal features. The model performance gradually improved with an increase in the spectral index. However, after the index increased beyond the optimal spectral index, the model performance tended to become weakened. The best inversion model accuracy was achieved by setting the retained feature variables to the first 27 in the sample groups of all soil types and meadow soils, with NRMSEs of 10.07 and 8.77%, respectively. The highest model accuracy was achieved by selecting the first 23 and 25 spectral indices for the black soil group and albic soil group, resulting in NRMSEs of 8.17 and 9.42%, respectively. Table S1 lists the optimal set of variables for the four soil types.

Figure 8 
                  Iterative analysis of the number of features and XGBoost model performance for different sample groups.
Figure 8

Iterative analysis of the number of features and XGBoost model performance for different sample groups.

3.4 Spatial estimation and mapping of SOMC

In this study, the optimal spectral indices for each group were used as independent variables in the model. SOMC was used as the dependent variable, the number of samples was randomly divided into training and validation groups at a 3:1 ratio, and local regression estimation was performed on different soil sample types using the XGBoost model. Table 2 presents the regression accuracies of the three SOM estimation models. The estimation model for black soil used 23 spectral indices as dependent variables, and the model’s coefficient of determination was 0.845. The estimation model for albic soil used 25 spectral indices as dependent variables, and the coefficient of determination of the model was 0.744. The estimation model for meadow soil used 27 spectral indices as the independent variables, and the model’s coefficient of determination was 0.788. This indicates that the estimation models under the split-sample strategy exhibited better performance.

Table 2

Estimation accuracy of SOM in three groups

R² NRMSE (%) RPD
Black soils 0.845 8.17 2.37
Albic soils 0.744 9.42 2.05
Meadow soils 0.788 8.77 2.21

The study area was divided into three soil subzones based on the soil type. The content was estimated and the distribution of SOMC was mapped using the XGBoost model. This was based on the characteristic spectral indices and extracted bare soil image. Figure 9 illustrates the spatial distribution of SOMC within the study area, which ranged from 0.4 to 2.89%, with an average SOMC of 1.24%. The overall SOMC level was predominantly 1.5–2%, accounting for 43.39% of the total study area, with the highest 2–3% range accounting for 15.73% of the study area. The spatial distribution of SOMC was generally high in the east-central region, while areas with lower values were primarily dispersed in the southern region. The high-value areas are roughly banded due to the cultivation of paddy fields in the eastern region, where long-term application of nitrogen fertilizers and irrigation results in a high humus content.

Figure 9 
                  Spatial distribution of predicted SOMC under split-sample regression conditions.
Figure 9

Spatial distribution of predicted SOMC under split-sample regression conditions.

4 Discussion

4.1 Accuracy analysis of SOM estimation model based on hyperspectral index under split-sample strategy

To further analyze the superiority of the split-sample strategy, the R² and NRMSE values of the SOM estimation models were used to assess predictive accuracy across soil types, with black soil achieving the highest R² of 0.845 and the lowest NRMSE of 8.17%. Figure 10 shows the error scatter plots of the split-sample and full-sample regression models categorized according to the three soil types. The SOMC estimation model accuracies were in the following descending order: black soil, meadow soil, albic soil, and full-sample groups. The estimation model of the full-sample regression had 27 spectral indices as independent variables and the coefficient of determination of the model was 0.560. Among the regression models for the three soils, the R² of the SOMC estimation model was 0.845, 0.744, and 0.786 for the black, albic, and meadow soils, respectively. It was observed that the spectral index model of the full-sample group was limited by the complexity of the soil type and limited accuracy improvement. These results indicate that the models using the split-sample strategy demonstrated superior predictive ability and a better fit to the data. Among them, black soil had the best accuracy for the SOMC estimation model, which may occur for two reasons. First, it is caused by different physical properties of the soil structure. The structure of black soil is looser and mainly arranged in agglomerates or blocks, which helps to maintain aeration and drainage of the soil [51]. In contrast, the soil structures of the albic soil and meadow soil were mainly dominated by blocks or columns, leading to relatively poor soil drainage [52]. Particularly in albic soil, the albic layer has a high hardness and water capacity, inevitably causing water stagnation in the upper layer. Considering the varying intrinsic soil storage water content, this difference may result in variations in the accuracy of estimation among different datasets. Second, it may be caused by different chemical components of different soil types. Compared with albic soils, black soils and meadow soils have rich humus layers and higher SOMC accumulation, which increases the variability of the spectral signals. Lu et al. [53] found that the accuracy of the spectral prediction model for SOM was not satisfactory when considering traditional soil types. This is attributed to the presence of sand ginger black and yellow-brown soils, both of which contain high levels of clay minerals, particularly montmorillonite. The SOMC values in these two soil types were relatively similar, leading to spectral curves that exhibited similar characteristics. As a result, the regression accuracy for SOMC was low. In contrast, the clay minerals in the black and meadow soils in this study exhibited significant differences, which can enhance the acquisition of valuable information, thereby improving the accuracy of the model.

Figure 10 
                  Comparison of accuracy between split-sample regression and full-sample regression based on three soil type classifications.
Figure 10

Comparison of accuracy between split-sample regression and full-sample regression based on three soil type classifications.

Therefore, by analyzing the spectral differences of different soils and then constructing a spectral index applicable to the assessment of different soil properties using methods such as correlation analysis and feature selection, it helps in soil quality evaluation. However, although the hyperspectral index developed in this study enhanced the differences in the independent variables and the accuracy of the estimation model, it still excluded a significant amount of information from the bands, which may have resulted in the omission of information and limited the accuracy improvement of the estimated model. To maximize the use of valuable information within the bands, the derivative method can be applied to hyperspectral data to highlight the distinctions among spectral information [8]. This approach can improve the estimation accuracy of the surface parameters while effectively capturing the fine details of the spectral curves.

4.2 Analysis of the contribution of eigenspectral index variables in regression models

Combined with the analysis of the spectral reflectance of various soil types, it was found that different soils exhibit different response characteristics in satellite spectral bands, as well as in different hyperspectral indices. To explore the degree of response of various soil types to these spectral indices, the importance scores of the spectral index features in the estimation models for the different soil types were assessed, as shown in Figure 11. This was performed by calculating and ranking each attribute in the dataset, as shown in Figure 11. This result was obtained from the XGBoost model by weighting and summing the results of each attribute in all boosted trees and then averaging them. This process highlights the differences in the spectral indices of each group within the sub-sample while also reflecting the varying degrees of contribution of different spectral indices in the model. In both the full-sample and split-sample regressions, the DI index, with R499 and R576 as the dependent variables, exhibited a significant influence on the prediction of the four model groups. The proportions of DI and NDI in the spectral parameters of the black soil estimation model reached 39%. Moreover, the proportions of DI and NDI were higher in the albic and meadow soil estimation models. In contrast, ratio indices (RI), although not a large proportion of the model and does not have the ability to significantly improve its predictive ability, may play a role in stabilizing the model when participating in the model construction process together with other spectral indices.

Figure 11 
                  Characteristic scores of spectral indices involved in modeling under different soil types.
Figure 11

Characteristic scores of spectral indices involved in modeling under different soil types.

When distinguishing between the three soil types for analyzing the F-score, an increasing trend in the importance of vegetation indices in the model was observed. In particular, the EVI, which consists of the R456, R679, and R937 bands, exhibited a higher degree of characteristic response to SOM across the three soil backgrounds. The average NRMSE of the model decreased by 0.2% at the second iteration in the EVI for the black soil group. Additionally, the average NRMSE decreased by 0.3 and 0.1% at the third and twenty-first iterations in EVI and SATVI for the albic soil group, respectively. These results indicate that the vegetation indices, such as EVI and SATVI, exhibit varying degrees of responsiveness to organic matter across different soil types, aiding in SOMC estimation in areas with complex soil types. This is consistent with He et al. [24], who found that the near-infrared band and SATVI are crucial for SOC prediction. This finding supports the notion that the vegetation index responds differently to organic matter in various soil types during the process of soil type differentiation, highlighting its potential to enhance the accuracy of estimating SOMC in bare soil areas.

4.3 Uncertainty analysis of modeling using hyperspectral indices

As demonstrated by the accuracy of the SOM estimation model, spectral indices serve as reliable inputs for the construction of estimation models. Specifically, spectral indices such as DI and NDI played a pivotal role in enhancing both model stability and prediction accuracy. They effectively captured critical variations in soil reflectance at key wavelengths, particularly at 499, 516, 559, and 576 nm, which are highly sensitive to organic matter content. This ability to comprehensively capture spectral responses related to SOM significantly contributes to the overall effectiveness of the estimation models.

In the optimal variable set for the black soil group, the combination of bands R576 and R585 was used most frequently, each accounting for 14%. For the albic soil and meadow soil groups, R576 had the highest utilization rate. The correlation coefficients of these bands in previous correlation analyses exceeded 0.48. Therefore, it can be seen that the band combinations of spectral indices in various prediction models were primarily concentrated in the 500–600 nm range, which is consistent with the broad absorption bands identified by Angelopoulou et al. [54] in their spectral study of soil properties. Different types of soil contain different forms of humus with diverse chemical compositions. Weak overtones and combinations of these vibrations are present in NIR-SWIR (700–2,500 nm), owing to the stretching and bending of the N–H, O–H, and C–H bonds in the soil components. From the results of the iterative analysis, it is clear that vegetation indices such as EVI and SATVI have high feature scores in the regression model for distinguishing soil types. They help provide more comprehensive information about soil properties.

However, there are still some uncertainties in the estimation of organic matter content using spectral indices. Factors such as small quantities of straw, moisture, and atmospheric conditions can influence the accuracy of organic matter content estimation. Many of these variables change over time, which significantly exacerbates the uncertainty problem. These include temperature, precipitation, soil moisture content, and the physical canopy structure. Therefore, finding and synergizing important environmental variables with the specific conditions of the study area, and developing a bare soil spectral index that can correct for atmospheric and vegetation effects to optimize the estimation model is one of the important directions to improve the accuracy of SOM inversion tools in the future.

5 Conclusion

In this study, we investigated the methodology and feasibility of using hyperspectral indices to estimate SOMC at large scales, while considering different soil types for sub-sample regression, which effectively improved the accuracy and stability of the XGBoost estimation model. The results show that the spectral bands of the characteristic spectral indices were primarily concentrated in the 499–600 nm range. Among these, the combinations of DI, NDI, and ratio indices (RI), consisting of bands around 580 nm with other bands, contained abundant information related to SOM, with DI499/576 having the most significant impact on model construction. This finding confirms the critical role these spectral indices play in improving model accuracy. Additionally, in the split-sample regression model, SATVI and EVI demonstrated higher explanatory power for SOM across the three soil backgrounds, with particularly strong responsiveness in black soils and albic soils. These results indicate that vegetation indices are more effective in estimating SOM in regions with complex soil types, as they can better capture organic matter-related information for different soil types. However, this study also has some limitations. Since the model was primarily trained and tested on specific soil types, its applicability to regions with different soil characteristics may be limited. Furthermore, this method may require a substantial amount of field data for model training, which could hinder its application in data-scarce environments. Future studies should consider other typical soil types to validate the model’s performance across diverse geographic regions. Additionally, developing more cost-effective data collection methods and exploring machine learning algorithms that require fewer training samples will be the key to broadening the practical use of SOM estimation models.

Acknowledgements

The authors would like to express their sincere gratitude to the following individuals and organizations for their valuable contributions and support throughout the course of this research. First and foremost, we extend our thanks to Northeast Institute of Geography and Agroecology, Chinese Academy of Sciences for providing the necessary resources and laboratory facilities that were instrumental in conducting our experiments. Special thanks are due to Xingming Zheng for their expert technical assistance and guidance throughout the research process. We are also grateful to Qian Yang for their assistance with data collection and analysis. Finally, we are deeply grateful to all the participants of the study for their willingness to contribute to our research.

  1. Funding information: This research was supported by the Natural Science Foundation of Jilin Province (Grant No. 20230101373JC) and the Science and Technology Development Project of Jilin Province (Grant No. 20210203016SF), enabling advancements in hyperspectral soil analysis techniques.

  2. Author contributions: Nan Lin performed the data analysis, collated the data, and drafted the manuscript. Liu Yanlong conducted data acquisition and data analysis, visualized experimental results, and edited manuscripts for review. Liu Qiang verified the method, participated in the fund acquisition of the project, and participated in the management of the project. Ranzhe Jiang developed the methodology and managed the project. Mazur conducted the survey and developed the software used in the study. Each author played a pivotal role in the research, ensuring the integrity and innovation of the manuscript.

  3. Conflict of interest: The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

[1] Francaviglia R, Almagro M, Vicente-Vicente JL. Conservation agriculture and soil organic carbon: Principles, processes, practices and policy options. Soil Syst. 2023;7(1):17. 10.3390/soilsystems7010017.Search in Google Scholar

[2] Ondrasek G, Bakić Begić H, Zovko M, Filipović L, Meriño-Gergichevich C, Savić R, et al. Biogeochemistry of soil organic matter in agroecosystems & environmental implications. Sci Total Environ. 2019;658:1559–73. 10.1016/j.scitotenv.2018.12.243.Search in Google Scholar PubMed

[3] Fan M, Lal R, Zhang H, Margenot AJ, Wu J, Wu P, et al. Variability and determinants of soil organic matter under different land uses and soil types in eastern China. Soil Tillage Res. 2020;198:104544. 10.1016/j.still.2019.104544.Search in Google Scholar

[4] Yu P, Li Y, Liu S, Ding Z, Zhang A, Tang X. The quantity and stability of soil organic carbon following vegetation degradation in a salt-affected region of Northeastern China. CATENA. 2022;211:105984. 10.1016/j.catena.2021.105984.Search in Google Scholar

[5] Carter MR. Soil quality for sustainable land management: organic matter and aggregation interactions that maintain soil functions. Agron J. 2002;94(1):38–47. 10.2134/agronj2002.3800.Search in Google Scholar

[6] Bowman R, Guenzi W, Savory DJ. Spectroscopic method for estimation of soil organic carbon. Soil Sci Soc Am J. 1991;55(2):563–6. 10.2136/sssaj1991.03615995005500020048x.Search in Google Scholar

[7] Fu Y, Yang G, Pu R, Li Z, Li H, Xu X, et al. An overview of crop nitrogen status assessment using hyperspectral remote sensing: Current status and perspectives. Eur J Agron. 2021;124:126241. 10.1016/j.eja.2021.126241.Search in Google Scholar

[8] Shen L, Gao M, Yan J, Li ZL, Leng P, Yang Q, et al. Hyperspectral estimation of soil organic matter content using different spectral preprocessing techniques and PLSR method. Remote Sens. 2020;12(7):1206. 10.3390/rs12071206.Search in Google Scholar

[9] Bajwa SG, Bajcsy P, Groves P, Tian LF. Hyperspectral image data mining for band selection in agricultural applications. Trans ASAE. 2004;47(3):895–907. 10.13031/2013.16087.Search in Google Scholar

[10] Shi T, Chen Y, Liu H, Wang J, Wu G. Soil organic carbon content estimation with laboratory-based visible near-infrared reflectance spectroscopy: Feature selection. Appl Spectrosc. 2014;68(8):831–7. 10.1366/13-07294.Search in Google Scholar PubMed

[11] Xing Z, Chen J, Zhao X, Li Y, Li X, Zhang Z, et al. Quantitative estimation of wastewater quality parameters by hyperspectral band screening using GC, VIP and SPA. PeerJ. 2019;7:e8255. 10.7717/peerj.8255.Search in Google Scholar PubMed PubMed Central

[12] Zhang W, Li X, Zhao L. A fast hyperspectral feature selection method based on band correlation analysis. IEEE Geosci Remote Sens Lett. 2018;15(11):1750–4. 10.1109/lgrs.2018.2853805.Search in Google Scholar

[13] Yu L, Hong Y, Geng L, Zhou Y, Zhu Q, Cao J, et al. Hyperspectral estimation of soil organic matter content based on partial least squares regression. Trans Chin Soc Agric Eng. 2015;31(14):103–9. 10.11975/j.issn.1002-6819.2015.14.015.Search in Google Scholar

[14] Ding J, Yang A, Wang J, Sagan V, Yu D. Machine-learning-based quantitative estimation of soil organic carbon content by VIS/NIR spectroscopy. PeerJ. 2018;6:e5714. 10.7717/peerj.5714.Search in Google Scholar PubMed PubMed Central

[15] Stenberg B, Viscarra Rossel RA, Mouazen AM, Wetterlind J. Visible and near infrared spectroscopy in soil science. Adv Agron. 2010;107:163–215. 10.1016/S0065-2113(10)07005-7.Search in Google Scholar

[16] Ahmad F. Spectral vegetation indices performance evaluated for Cholistan Desert. J Geogr Reg Plan. 2012;5(6):165–72. 10.5897/JGRP11.098.Search in Google Scholar

[17] Chen L, Lai J, Tan K, Wang X, Chen Y, Ding J. Development of a soil heavy metal estimation method based on a spectral index: Combining fractional-order derivative pretreatment and the absorption mechanism. Sci Total Environ. 2022;813:151882. 10.1016/j.scitotenv.2021.151882.Search in Google Scholar PubMed

[18] Yang J, Weisberg PJ, Bristow NA. Landsat remote sensing approaches for monitoring long-term tree cover dynamics in semi-arid woodlands: Comparison of vegetation indices and spectral mixture analysis. Remote Sens Environ. 2012;119:62–71. 10.1016/j.rse.2011.12.004.Search in Google Scholar

[19] Vaudour E, Gomez C, Loiseau T, Baghdadi N, Loubet B, Arrouays D, et al. The impact of acquisition date on the prediction performance of topsoil organic carbon from Sentinel-2 for croplands. Remote Sens. 2019;11(18):2143. 10.3390/rs11182143.Search in Google Scholar

[20] Dou X, Wang X, Liu H, Zhang X, Meng L, Pan Y, et al. Prediction of soil organic matter using multi-temporal satellite images in the Songnen Plain, China. Geoderma. 2019;356:113896. 10.1016/j.geoderma.2019.113896.Search in Google Scholar

[21] Pisek J, Chen JM. Comparison and validation of MODIS and VEGETATION global LAI products over four BigFoot sites in North America. Remote Sens Environ. 2007;109(1):81–94. 10.1016/j.rse.2006.12.004.Search in Google Scholar

[22] Sims DA, Gamon JA. Relationships between leaf pigment content and spectral reflectance across a wide range of species, leaf structures and developmental stages. Remote Sens Environ. 2002;81(2–3):337–54. 10.1016/s0034-4257(02)00010-x.Search in Google Scholar

[23] Guo L, Sun X, Fu P, Shi T, Dang L, Chen Y, et al. Mapping soil organic carbon stock by hyperspectral and time-series multispectral remote sensing images in low-relief agricultural areas. Geoderma. 2021;398:115118. 10.1016/j.geoderma.2021.115118.Search in Google Scholar

[24] He X, Yang L, Li A, Zhang L, Shen F, Cai Y, et al. Soil organic carbon prediction using phenological parameters and remote sensing variables generated from Sentinel-2 images. CATENA. 2021;205:105442. 10.1016/j.catena.2021.105442.Search in Google Scholar

[25] Wang K, Qi Y, Guo W, Zhang J, Chang Q. Retrieval and mapping of soil organic carbon using sentinel-2A spectral images from bare cropland in autumn. Remote Sens. 2021;13(6):1072. 10.3390/rs13061072.Search in Google Scholar

[26] Silvero NEQ, Demattê JAM, Minasny B, Rosin NA, Nascimento JG, Rodríguez Albarracín HS, et al. Sensing technologies for characterizing and monitoring soil functions: A review. Adv Agron. 2023;177:125–68. 10.1016/bs.agron.2022.08.002.Search in Google Scholar

[27] Naimi S, Ayoubi S, Di Raimo LADL, Dematte JAM. Quantification of some intrinsic soil properties using proximal sensing in arid lands: Application of Vis-NIR, MIR, and pXRF spectroscopy. Geoderma Reg. 2022;28:e00484. 10.1016/j.geodrs.2022.e00484.Search in Google Scholar

[28] Liu H, Zhang Y, Zhang B. Novel hyperspectral reflectance models for estimating black-soil organic matter in Northeast China. Environ Monit Assess. 2009;154:147–54. 10.1007/s10661-008-0385-4.Search in Google Scholar PubMed

[29] Xiao S, He Y, Dong T, Nie P. Spectral analysis and sensitive waveband determination based on nitrogen detection of different soil types using near infrared sensors. Sensors (Basel, Switzerland). 2018;18(2):523. 10.3390/s18020523.Search in Google Scholar PubMed PubMed Central

[30] Liu Y, Shi Z, Zhang G, Chen Y, Li S, Hong Y, et al. Application of spectrally derived soil type as ancillary data to improve the estimation of soil organic carbon by using the Chinese soil vis-NIR spectral library. Remote Sens. 2018;10(11):1747. 10.3390/rs10111747.Search in Google Scholar

[31] Xie S, Ding F, Chen S, Wang X, Li Y, Ma K. Prediction of soil organic matter content based on characteristic band selection method. Spectrochim Acta Part A, Mol Biomol Spectrosc. 2022;273:120949. 10.1016/j.saa.2022.120949.Search in Google Scholar PubMed

[32] Pan Y, Zhang X, Liu H, Wu D, Dou X, Xu M, et al. Remote sensing inversion of soil organic matter by using the subregion method at the field scale. Precis Agric. 2022;23(5):1813–35. 10.1007/s11119-022-09914-2.Search in Google Scholar

[33] Brown DJ, Bricklemyer RS, Miller PRJG. Validation requirements for diffuse reflectance soil characterization models with a case study of VNIR soil C prediction in Montana. Geoderma. 2005;129(3-4):251–67. 10.1016/j.geoderma.2005.01.001.Search in Google Scholar

[34] Tamburini E, Vincenzi F, Costa S, Mantovi P, Pedrini P, Castaldelli G. Effects of moisture and particle size on quantitative determination of total organic carbon (TOC) in soils using near-infrared spectroscopy. Sensors (Basel, Switzerland). 2017;17(10):2366. 10.3390/s17102366.Search in Google Scholar PubMed PubMed Central

[35] Ba Y, Liu J, Han J, Zhang X. Application of Vis-NIR spectroscopy for determination the content of organic matter in saline-alkali soils. Spectrochim Acta Part A, Mol Biomol Spectrosc. 2020;229:117863. 10.1016/j.saa.2019.117863.Search in Google Scholar PubMed

[36] Guo Y, Qi W, Yao Y, Qin F, Gao Y, Zhao W. Relations between soil organic carbon and other soil physicochemical indices under different land use patterns. Surf Rev Lett. 2017;24(3):1750039. 10.1142/s0218625x17500391.Search in Google Scholar

[37] Calin MA, Calin AC, Nicolae DN. Application of airborne and spaceborne hyperspectral imaging techniques for atmospheric research: Past, present, and future. Appl Spectrosc Rev. 2021;56(4):289–323. 10.1080/05704928.2020.1774381.Search in Google Scholar

[38] Wang X, Zhang F, Ding J, Kung HT, Latif A, Johnson VC. Estimation of soil salt content (SSC) in the Ebinur Lake Wetland National Nature Reserve (ELWNNR), Northwest China, based on a Bootstrap-BP neural network model and optimal spectral indices. Sci Total Environ. 2018;615:918–30. 10.1016/j.scitotenv.2017.10.025.Search in Google Scholar PubMed

[39] Matsushita B, Yang W, Chen J, Onda Y, Qiu G. Sensitivity of the enhanced vegetation index (EVI) and normalized difference vegetation index (NDVI) to topographic effects: A case study in high-density cypress forest. Sensors (Basel, Switzerland). 2007;7(11):2636–51. 10.3390/s7112636.Search in Google Scholar PubMed PubMed Central

[40] Mohamed ES, Saleh AM, Belal AB, Gad A. Application of near-infrared reflectance for quantitative assessment of soil properties. Egypt J Remote Sens Space Sci. 2018;21(1):1–14. 10.1016/j.ejrs.2017.02.001.Search in Google Scholar

[41] Xue J, Su B. Significant remote sensing vegetation indices: A review of developments and applications. J Sens. 2017;2017:1–17. 10.1155/2017/1353691.Search in Google Scholar

[42] Goirán S, Aranibar J, Gomez M. Heterogeneous spatial distribution of traditional livestock settlements and their effects on vegetation cover in arid groundwater coupled ecosystems in the Monte Desert (Argentina). J Arid Environ. 2012;87:188–97. 10.1016/j.jaridenv.2012.07.011.Search in Google Scholar

[43] Gobrecht A, Bendoula R, Roger JM, Bellon-Maurel V. A new optical method coupling light polarization and Vis–NIR spectroscopy to improve the measurement of soil carbon content. Soil Tillage Res. 2016;155:461–70. 10.1016/j.still.2015.06.003.Search in Google Scholar

[44] Liu Q, He L, Guo L, Wang M, Deng D, Lv P, et al. Digital mapping of soil organic carbon density using newly developed bare soil spectral indices and deep neural network. SSRN Electron J. 2022;219:106603. 10.2139/ssrn.4160571.Search in Google Scholar

[45] Villarreal ML, Norman LM, Buckley S, Wallace CSA, Coe MA. Multi-index time series monitoring of drought and fire effects on desert grasslands. Remote Sens Environ. 2016;183:186–97. 10.1016/j.rse.2016.05.026.Search in Google Scholar

[46] Escadafal R. Remote sensing of arid soil surface color with Landsat thematic mapper. Adv Space Res. 1989;9(1):159–63. 10.1016/0273-1177(89)90481-x.Search in Google Scholar

[47] Mahmood K, Batool SA, Chaudhry MN. Studying bio-thermal effects at and around MSW dumps using Satellite Remote Sensing and GIS. Waste Manag (New York, NY). 2016;55:118–28. 10.1016/j.wasman.2016.04.020.Search in Google Scholar PubMed

[48] Yang Y, Shang K, Xiao C, Wang C, Tang H. Spectral index for mapping topsoil organic matter content based on ZY1-02D satellite hyperspectral data in Jiangsu Province, China. ISPRS Int J Geo-Inf. 2022;11(2):111. 10.3390/ijgi11020111.Search in Google Scholar

[49] Chen T, Guestrin CJA. XGboost: A scalable tree boosting system. Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining. New York, NY; 2016. p. 785–94. 10.1145/2939672.2939785.Search in Google Scholar

[50] Shi X, Wong YD, Li MZ, Palanisamy C, Chai C. A feature learning approach based on XGBoost for driving assessment and risk prediction. Accid Anal Prev. 2019;129:170–9. 10.1016/j.aap.2019.05.005.Search in Google Scholar PubMed

[51] Yu G, Fang H, Gao L, Zhang W. Soil organic carbon budget and fertility variation of black soils in Northeast China. Ecol Res. 2006;21(6):855–67. 10.1007/s11284-006-0033-9.Search in Google Scholar

[52] Chizhikova N, Khitrov N. Diversity of clay minerals in soils of solonetzic complexes in the southeast of Western Siberia. Eurasian Soil Sci. 2016;49:1419–31. 10.1134/S106422931612005X.Search in Google Scholar

[53] Lu L, Zhang P, Lu HL, Liu BY, Zhao MS. Hyperspectral characteristics of soils in huaibei plain and estimation of SOM content. 2019;51(2):378–80. 10.13758/j.cnki.tr.2019.02.023.Search in Google Scholar

[54] Angelopoulou T, Tziolas N, Balafoutis A, Zalidis G, Bochtis D. Remote sensing techniques for soil organic carbon estimation: A review. Remote Sens. 2019;11(6):676. 10.3390/rs11060676.Search in Google Scholar

Received: 2024-09-01
Revised: 2024-11-03
Accepted: 2024-11-06
Published Online: 2024-12-24

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Regular Articles
  2. Theoretical magnetotelluric response of stratiform earth consisting of alternative homogeneous and transitional layers
  3. The research of common drought indexes for the application to the drought monitoring in the region of Jin Sha river
  4. Evolutionary game analysis of government, businesses, and consumers in high-standard farmland low-carbon construction
  5. On the use of low-frequency passive seismic as a direct hydrocarbon indicator: A case study at Banyubang oil field, Indonesia
  6. Water transportation planning in connection with extreme weather conditions; case study – Port of Novi Sad, Serbia
  7. Zircon U–Pb ages of the Paleozoic volcaniclastic strata in the Junggar Basin, NW China
  8. Monitoring of mangrove forests vegetation based on optical versus microwave data: A case study western coast of Saudi Arabia
  9. Microfacies analysis of marine shale: A case study of the shales of the Wufeng–Longmaxi formation in the western Chongqing, Sichuan Basin, China
  10. Multisource remote sensing image fusion processing in plateau seismic region feature information extraction and application analysis – An example of the Menyuan Ms6.9 earthquake on January 8, 2022
  11. Identification of magnetic mineralogy and paleo-flow direction of the Miocene-quaternary volcanic products in the north of Lake Van, Eastern Turkey
  12. Impact of fully rotating steel casing bored pile on adjacent tunnels
  13. Adolescents’ consumption intentions toward leisure tourism in high-risk leisure environments in riverine areas
  14. Petrogenesis of Jurassic granitic rocks in South China Block: Implications for events related to subduction of Paleo-Pacific plate
  15. Differences in urban daytime and night block vitality based on mobile phone signaling data: A case study of Kunming’s urban district
  16. Random forest and artificial neural network-based tsunami forests classification using data fusion of Sentinel-2 and Airbus Vision-1 satellites: A case study of Garhi Chandan, Pakistan
  17. Integrated geophysical approach for detection and size-geometry characterization of a multiscale karst system in carbonate units, semiarid Brazil
  18. Spatial and temporal changes in ecosystem services value and analysis of driving factors in the Yangtze River Delta Region
  19. Deep fault sliding rates for Ka-Ping block of Xinjiang based on repeating earthquakes
  20. Improved deep learning segmentation of outdoor point clouds with different sampling strategies and using intensities
  21. Platform margin belt structure and sedimentation characteristics of Changxing Formation reefs on both sides of the Kaijiang-Liangping trough, eastern Sichuan Basin, China
  22. Enhancing attapulgite and cement-modified loess for effective landfill lining: A study on seepage prevention and Cu/Pb ion adsorption
  23. Flood risk assessment, a case study in an arid environment of Southeast Morocco
  24. Lower limits of physical properties and classification evaluation criteria of the tight reservoir in the Ahe Formation in the Dibei Area of the Kuqa depression
  25. Evaluation of Viaducts’ contribution to road network accessibility in the Yunnan–Guizhou area based on the node deletion method
  26. Permian tectonic switch of the southern Central Asian Orogenic Belt: Constraints from magmatism in the southern Alxa region, NW China
  27. Element geochemical differences in lower Cambrian black shales with hydrothermal sedimentation in the Yangtze block, South China
  28. Three-dimensional finite-memory quasi-Newton inversion of the magnetotelluric based on unstructured grids
  29. Obliquity-paced summer monsoon from the Shilou red clay section on the eastern Chinese Loess Plateau
  30. Classification and logging identification of reservoir space near the upper Ordovician pinch-out line in Tahe Oilfield
  31. Ultra-deep channel sand body target recognition method based on improved deep learning under UAV cluster
  32. New formula to determine flyrock distance on sedimentary rocks with low strength
  33. Assessing the ecological security of tourism in Northeast China
  34. Effective reservoir identification and sweet spot prediction in Chang 8 Member tight oil reservoirs in Huanjiang area, Ordos Basin
  35. Detecting heterogeneity of spatial accessibility to sports facilities for adolescents at fine scale: A case study in Changsha, China
  36. Effects of freeze–thaw cycles on soil nutrients by soft rock and sand remodeling
  37. Vibration prediction with a method based on the absorption property of blast-induced seismic waves: A case study
  38. A new look at the geodynamic development of the Ediacaran–early Cambrian forearc basalts of the Tannuola-Khamsara Island Arc (Central Asia, Russia): Conclusions from geological, geochemical, and Nd-isotope data
  39. Spatio-temporal analysis of the driving factors of urban land use expansion in China: A study of the Yangtze River Delta region
  40. Selection of Euler deconvolution solutions using the enhanced horizontal gradient and stable vertical differentiation
  41. Phase change of the Ordovician hydrocarbon in the Tarim Basin: A case study from the Halahatang–Shunbei area
  42. Using interpretative structure model and analytical network process for optimum site selection of airport locations in Delta Egypt
  43. Geochemistry of magnetite from Fe-skarn deposits along the central Loei Fold Belt, Thailand
  44. Functional typology of settlements in the Srem region, Serbia
  45. Hunger Games Search for the elucidation of gravity anomalies with application to geothermal energy investigations and volcanic activity studies
  46. Addressing incomplete tile phenomena in image tiling: Introducing the grid six-intersection model
  47. Evaluation and control model for resilience of water resource building system based on fuzzy comprehensive evaluation method and its application
  48. MIF and AHP methods for delineation of groundwater potential zones using remote sensing and GIS techniques in Tirunelveli, Tenkasi District, India
  49. New database for the estimation of dynamic coefficient of friction of snow
  50. Measuring urban growth dynamics: A study in Hue city, Vietnam
  51. Comparative models of support-vector machine, multilayer perceptron, and decision tree ‎predication approaches for landslide ‎susceptibility analysis
  52. Experimental study on the influence of clay content on the shear strength of silty soil and mechanism analysis
  53. Geosite assessment as a contribution to the sustainable development of Babušnica, Serbia
  54. Using fuzzy analytical hierarchy process for road transportation services management based on remote sensing and GIS technology
  55. Accumulation mechanism of multi-type unconventional oil and gas reservoirs in Northern China: Taking Hari Sag of the Yin’e Basin as an example
  56. TOC prediction of source rocks based on the convolutional neural network and logging curves – A case study of Pinghu Formation in Xihu Sag
  57. A method for fast detection of wind farms from remote sensing images using deep learning and geospatial analysis
  58. Spatial distribution and driving factors of karst rocky desertification in Southwest China based on GIS and geodetector
  59. Physicochemical and mineralogical composition studies of clays from Share and Tshonga areas, Northern Bida Basin, Nigeria: Implications for Geophagia
  60. Geochemical sedimentary records of eutrophication and environmental change in Chaohu Lake, East China
  61. Research progress of freeze–thaw rock using bibliometric analysis
  62. Mixed irrigation affects the composition and diversity of the soil bacterial community
  63. Examining the swelling potential of cohesive soils with high plasticity according to their index properties using GIS
  64. Geological genesis and identification of high-porosity and low-permeability sandstones in the Cretaceous Bashkirchik Formation, northern Tarim Basin
  65. Usability of PPGIS tools exemplified by geodiscussion – a tool for public participation in shaping public space
  66. Efficient development technology of Upper Paleozoic Lower Shihezi tight sandstone gas reservoir in northeastern Ordos Basin
  67. Assessment of soil resources of agricultural landscapes in Turkestan region of the Republic of Kazakhstan based on agrochemical indexes
  68. Evaluating the impact of DEM interpolation algorithms on relief index for soil resource management
  69. Petrogenetic relationship between plutonic and subvolcanic rocks in the Jurassic Shuikoushan complex, South China
  70. A novel workflow for shale lithology identification – A case study in the Gulong Depression, Songliao Basin, China
  71. Characteristics and main controlling factors of dolomite reservoirs in Fei-3 Member of Feixianguan Formation of Lower Triassic, Puguang area
  72. Impact of high-speed railway network on county-level accessibility and economic linkage in Jiangxi Province, China: A spatio-temporal data analysis
  73. Estimation model of wild fractional vegetation cover based on RGB vegetation index and its application
  74. Lithofacies, petrography, and geochemistry of the Lamphun oceanic plate stratigraphy: As a record of the subduction history of Paleo-Tethys in Chiang Mai-Chiang Rai Suture Zone of Thailand
  75. Structural features and tectonic activity of the Weihe Fault, central China
  76. Application of the wavelet transform and Hilbert–Huang transform in stratigraphic sequence division of Jurassic Shaximiao Formation in Southwest Sichuan Basin
  77. Structural detachment influences the shale gas preservation in the Wufeng-Longmaxi Formation, Northern Guizhou Province
  78. Distribution law of Chang 7 Member tight oil in the western Ordos Basin based on geological, logging and numerical simulation techniques
  79. Evaluation of alteration in the geothermal province west of Cappadocia, Türkiye: Mineralogical, petrographical, geochemical, and remote sensing data
  80. Numerical modeling of site response at large strains with simplified nonlinear models: Application to Lotung seismic array
  81. Quantitative characterization of granite failure intensity under dynamic disturbance from energy standpoint
  82. Characteristics of debris flow dynamics and prediction of the hazardous area in Bangou Village, Yanqing District, Beijing, China
  83. Rockfall mapping and susceptibility evaluation based on UAV high-resolution imagery and support vector machine method
  84. Statistical comparison analysis of different real-time kinematic methods for the development of photogrammetric products: CORS-RTK, CORS-RTK + PPK, RTK-DRTK2, and RTK + DRTK2 + GCP
  85. Hydrogeological mapping of fracture networks using earth observation data to improve rainfall–runoff modeling in arid mountains, Saudi Arabia
  86. Petrography and geochemistry of pegmatite and leucogranite of Ntega-Marangara area, Burundi, in relation to rare metal mineralisation
  87. Prediction of formation fracture pressure based on reinforcement learning and XGBoost
  88. Hazard zonation for potential earthquake-induced landslide in the eastern East Kunlun fault zone
  89. Monitoring water infiltration in multiple layers of sandstone coal mining model with cracks using ERT
  90. Study of the patterns of ice lake variation and the factors influencing these changes in the western Nyingchi area
  91. Productive conservation at the landslide prone area under the threat of rapid land cover changes
  92. Sedimentary processes and patterns in deposits corresponding to freshwater lake-facies of hyperpycnal flow – An experimental study based on flume depositional simulations
  93. Study on time-dependent injectability evaluation of mudstone considering the self-healing effect
  94. Detection of objects with diverse geometric shapes in GPR images using deep-learning methods
  95. Behavior of trace metals in sedimentary cores from marine and lacustrine environments in Algeria
  96. Spatiotemporal variation pattern and spatial coupling relationship between NDVI and LST in Mu Us Sandy Land
  97. Formation mechanism and oil-bearing properties of gravity flow sand body of Chang 63 sub-member of Yanchang Formation in Huaqing area, Ordos Basin
  98. Diagenesis of marine-continental transitional shale from the Upper Permian Longtan Formation in southern Sichuan Basin, China
  99. Vertical high-velocity structures and seismic activity in western Shandong Rise, China: Case study inspired by double-difference seismic tomography
  100. Spatial coupling relationship between metamorphic core complex and gold deposits: Constraints from geophysical electromagnetics
  101. Disparities in the geospatial allocation of public facilities from the perspective of living circles
  102. Research on spatial correlation structure of war heritage based on field theory. A case study of Jinzhai County, China
  103. Formation mechanisms of Qiaoba-Zhongdu Danxia landforms in southwestern Sichuan Province, China
  104. Magnetic data interpretation: Implication for structure and hydrocarbon potentiality at Delta Wadi Diit, Southeastern Egypt
  105. Deeply buried clastic rock diagenesis evolution mechanism of Dongdaohaizi sag in the center of Junggar fault basin, Northwest China
  106. Application of LS-RAPID to simulate the motion of two contrasting landslides triggered by earthquakes
  107. The new insight of tectonic setting in Sunda–Banda transition zone using tomography seismic. Case study: 7.1 M deep earthquake 29 August 2023
  108. The critical role of c and φ in ensuring stability: A study on rockfill dams
  109. Evidence of late quaternary activity of the Weining-Shuicheng Fault in Guizhou, China
  110. Extreme hydroclimatic events and response of vegetation in the eastern QTP since 10 ka
  111. Spatial–temporal effect of sea–land gradient on landscape pattern and ecological risk in the coastal zone: A case study of Dalian City
  112. Study on the influence mechanism of land use on carbon storage under multiple scenarios: A case study of Wenzhou
  113. A new method for identifying reservoir fluid properties based on well logging data: A case study from PL block of Bohai Bay Basin, North China
  114. Comparison between thermal models across the Middle Magdalena Valley, Eastern Cordillera, and Eastern Llanos basins in Colombia
  115. Mineralogical and elemental analysis of Kazakh coals from three mines: Preliminary insights from mode of occurrence to environmental impacts
  116. Chlorite-induced porosity evolution in multi-source tight sandstone reservoirs: A case study of the Shaximiao Formation in western Sichuan Basin
  117. Predicting stability factors for rotational failures in earth slopes and embankments using artificial intelligence techniques
  118. Origin of Late Cretaceous A-type granitoids in South China: Response to the rollback and retreat of the Paleo-Pacific plate
  119. Modification of dolomitization on reservoir spaces in reef–shoal complex: A case study of Permian Changxing Formation, Sichuan Basin, SW China
  120. Geological characteristics of the Daduhe gold belt, western Sichuan, China: Implications for exploration
  121. Rock physics model for deep coal-bed methane reservoir based on equivalent medium theory: A case study of Carboniferous-Permian in Eastern Ordos Basin
  122. Enhancing the total-field magnetic anomaly using the normalized source strength
  123. Shear wave velocity profiling of Riyadh City, Saudi Arabia, utilizing the multi-channel analysis of surface waves method
  124. Effect of coal facies on pore structure heterogeneity of coal measures: Quantitative characterization and comparative study
  125. Inversion method of organic matter content of different types of soils in black soil area based on hyperspectral indices
  126. Detection of seepage zones in artificial levees: A case study at the Körös River, Hungary
  127. Tight sandstone fluid detection technology based on multi-wave seismic data
  128. Characteristics and control techniques of soft rock tunnel lining cracks in high geo-stress environments: Case study of Wushaoling tunnel group
  129. Influence of pore structure characteristics on the Permian Shan-1 reservoir in Longdong, Southwest Ordos Basin, China
  130. Study on sedimentary model of Shanxi Formation – Lower Shihezi Formation in Da 17 well area of Daniudi gas field, Ordos Basin
  131. Multi-scenario territorial spatial simulation and dynamic changes: A case study of Jilin Province in China from 1985 to 2030
  132. Review Articles
  133. Major ascidian species with negative impacts on bivalve aquaculture: Current knowledge and future research aims
  134. Prediction and assessment of meteorological drought in southwest China using long short-term memory model
  135. Communication
  136. Essential questions in earth and geosciences according to large language models
  137. Erratum
  138. Erratum to “Random forest and artificial neural network-based tsunami forests classification using data fusion of Sentinel-2 and Airbus Vision-1 satellites: A case study of Garhi Chandan, Pakistan”
  139. Special Issue: Natural Resources and Environmental Risks: Towards a Sustainable Future - Part I
  140. Spatial-temporal and trend analysis of traffic accidents in AP Vojvodina (North Serbia)
  141. Exploring environmental awareness, knowledge, and safety: A comparative study among students in Montenegro and North Macedonia
  142. Determinants influencing tourists’ willingness to visit Türkiye – Impact of earthquake hazards on Serbian visitors’ preferences
  143. Application of remote sensing in monitoring land degradation: A case study of Stanari municipality (Bosnia and Herzegovina)
  144. Optimizing agricultural land use: A GIS-based assessment of suitability in the Sana River Basin, Bosnia and Herzegovina
  145. Assessing risk-prone areas in the Kratovska Reka catchment (North Macedonia) by integrating advanced geospatial analytics and flash flood potential index
  146. Analysis of the intensity of erosive processes and state of vegetation cover in the zone of influence of the Kolubara Mining Basin
  147. GIS-based spatial modeling of landslide susceptibility using BWM-LSI: A case study – city of Smederevo (Serbia)
  148. Geospatial modeling of wildfire susceptibility on a national scale in Montenegro: A comparative evaluation of F-AHP and FR methodologies
  149. Geosite assessment as the first step for the development of canyoning activities in North Montenegro
  150. Urban geoheritage and degradation risk assessment of the Sokograd fortress (Sokobanja, Eastern Serbia)
  151. Multi-hazard modeling of erosion and landslide susceptibility at the national scale in the example of North Macedonia
  152. Understanding seismic hazard resilience in Montenegro: A qualitative analysis of community preparedness and response capabilities
  153. Forest soil CO2 emission in Quercus robur level II monitoring site
  154. Characterization of glomalin proteins in soil: A potential indicator of erosion intensity
  155. Power of Terroir: Case study of Grašac at the Fruška Gora wine region (North Serbia)
  156. Special Issue: Geospatial and Environmental Dynamics - Part I
  157. Qualitative insights into cultural heritage protection in Serbia: Addressing legal and institutional gaps for disaster risk resilience
Downloaded on 10.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/geo-2022-0739/html
Scroll to top button