Next Article in Journal
Anomaly Detection in Log Files Using Selected Natural Language Processing Methods
Previous Article in Journal
Effect of Ligament Mapping from Different Magnetic Resonance Image Quality on Joint Stability in a Personalized Dynamic Model of the Human Ankle Complex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modelling Soil Temperature by Tree-Based Machine Learning Methods in Different Climatic Regions of China

1
School of Hydraulic and Ecological Engineering, Nanchang Institute of Technology, Nanchang 330099, China
2
State Key Laboratory of Simulation and Regulation of Water Cycle in River Basin, China Institute of Water Resources and Hydropower Research, Beijing 100038, China
3
Key Laboratory of Water Cycle and Related Land Surface Processes, Institute of Geographic Science and Science and Natural Resources Research, Chinese Academy of Sciences, Beijing 100101, China
4
Center for Agricultural Water Research in China, China Agricultural University, Beijing 100083, China
5
Institute of RS and GIS, School of Earth and Space Sciences, Peking University, Beijing 100871, China
6
Jiangxi Key Laboratory of Hydrology-Water Resources and Water Environment, Nanchang Institute of Technology, Nanchang 330099, China
7
School of Civil Engineering and Architecture, Wuhan Polytechnic University, Wuhan 430023, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(10), 5088; https://doi.org/10.3390/app12105088
Submission received: 21 April 2022 / Revised: 12 May 2022 / Accepted: 17 May 2022 / Published: 18 May 2022

Abstract

:
Accurate estimation of soil temperature (Ts) at a national scale under different climatic conditions is important for soil–plant–atmosphere interactions. This study estimated daily Ts at the 0 cm depth for 689 meteorological stations in seven different climate zones of China for the period 1966–2015 with the M5P model tree (M5P), random forests (RF), and the extreme gradient boosting (XGBoost). The results showed that the XGBoost model (averaged coefficient of determination (R2) = 0.964 and root mean square error (RMSE) = 2.066 °C) overall performed better than the RF (averaged R2 = 0.959 and RMSE = 2.130 °C) and M5P (averaged R2 = 0.954 and RMSE = 2.280 °C) models for estimating Ts with higher computational efficiency. With the combination of mean air temperature (Tmean) and global solar radiation (Rs) as inputs, the estimating accuracy of the models was considerably high (averaged R2 = 0.96–0.97 and RMSE = 1.73–1.99 °C). On the basis of Tmean, adding Rs to the model input had a greater degree of influence on model estimating accuracy than adding other climatic factors to the input. Principal component analysis indicated that soil organic matter, soil water content, Tmean, relative humidity (RH), Rs, and wind speed (U2) are the main factors that cause errors in estimating Ts, and the total error interpretation rate was 97.9%. Overall, XGBoost would be a suitable algorithm for estimating Ts in different climate zones of China, and the combination of Tmean and Rs as model inputs would be more practical than other input combinations.

1. Introduction

Soil temperature (Ts), as a consequence of the combined effect of hydrothermal circulation in both the atmosphere and the land surface, is an important factor in atmospheric–ecological–environmental systems [1,2,3]. Ts has significant impacts on the energy balance of global atmospheric–ecological–environmental systems [4,5], thereby directly affecting the growth and development of plants [6]. Under natural conditions, Ts is influenced by other environmental factors such as atmospheric temperature, vegetation type, soil moisture, and topography, which together cause the spatial and temporal variations in Ts [7]. Given its importance and complexity, accurate prediction of Ts is of substantial value in both scientific research and practical applications [8,9]. However, the current approaches for measuring Ts are generally complicated, costly, and time consuming [1], which make the wide range monitoring of Ts inaccessible in many countries, especially for undeveloped countries. The incomplete Ts data caused by this predicament presents an obstacle to the estimation and analysis of Ts conditions and poses a significant challenge for accurate Ts prediction on a large scale [10].
The prevailing approaches for Ts measurements can be divided into three major categories, each of which has its own advantages and disadvantages. The first category is the in situ monitoring of Ts with equipment installed in meteorological stations, which allows the actual Ts to be recorded in real-time [11,12]. However, considering the geographical limitation and the economic cost of constructing meteorological stations and the potential data deficiencies due to equipment problems, this approach is not suitable for studies covering various regions [13]. The second category is the estimation of Ts, which consists of the interpolation method, the Newhorl model estimation, and direct estimation. The interpolation method estimates Ts at any depth based on known Ts at similar depths through certain formulas [14]. The Newhorl model estimation method uses air temperature plus a fixed value to roughly estimate Ts at a certain depth. For instance, Deboer [15] obtained the Ts values at a depth of 50 cm indirectly by adding 2.5 °C to the mean yearly temperature. The direct estimation method simply uses the Ts from known depths to represent Ts at unknown depths [16]. Although approaches in the second category can solve certain problems in Ts measurements to some extent, differences in several factors such as soil depth, elevation, and environment are not considered during estimation, which certainly have impacts on Ts. Therefore, these methods may be suitable for the local scale, but results obtained from them are not universally applicable and cannot be generalized [17]. Therefore, it is necessary to find methods that are applicable to a large-scale range. The third category is to establish a regression equation for Ts calculation. Research has verified similar patterns in the variations between Ts and air temperature [18,19], and therefore relationships between the two temperatures can be used for Ts calculation at various depths through regression equations [20]. Some studies even consider the impacts of geographical factors (e.g., longitude, latitude, and elevation) on Ts and build a multiple linear regression equation to quantify the impact of each geographical factor on Ts [21]. Although approaches in this category overcome the deficiencies of the methods in the other two categories, there are still disadvantages in these approaches as Ts are not necessarily linearly related to geographical factors [22], which might lead to significant biases during calculation [23]. Therefore, it is necessary to explore a more efficient method to perform the calculation.
In recent years, machine learning algorithms have been widely used in studies for estimating hydrological and meteorological indicators, such as air temperature [24], dew point temperature [25], precipitation [26], solar radiation [27], diffuse solar radiation [28], and evapotranspiration [29,30], among which an artificial neural network (ANN) is probably the most common machine learning algorithm used for modeling. For Ts estimation, ANN has been applied in different areas of the world by scholars. The first study on record for estimating Ts with machine learning algorithms was carried out by Yang et al. [31], in which Ts at three depths (i.e., 10, 50, and 150 cm) were estimated with ANN models. Mihalakakou [32] evaluated the potential of ANN in daily and yearly Ts estimation in Athens and Dublin, and the research concluded that the ANN model had an adequate performance in estimating Ts. Bilgili [33] applied an ANN model to estimate monthly Ts at multiple depths in Adana, Turkey, and the result showed that ANN was a suitable model for estimating Ts. Tabari et al. [34] utilized ANN and multiple linear regression model to estimate daily Ts at six soil depths, in which they found that temperature and relative humidity (RH) were the most influential parameters affecting Ts estimation among meteorological factors. Although the ANN model is widely used by everyone for estimating Ts, as scholars continue their research, it was found that the ANN model is not necessarily better than other machine learning algorithms in estimating Ts. Therefore, it becomes more meaningful to keep investigating new models to improve the estimation performance.
Recent advances of newly developed machine learning algorithms have enabled scholars to evaluate and compare the capabilities of various algorithms in estimating Ts, e.g., M5 model tree (M5 Tree), random forests (RF), M5P model tree (M5P), multiple linear regression (MLR) model, support vector machine (SVM) model, and extreme learning machine (ELM) model, etc. Sanikhani et al. [2] compared the capabilities of ELM, ANN, and M5 Tree in modeling monthly Ts at 5, 50, and 100 cm depths, and they found that the ELM model would be a more desirable tool for estimating Ts at a wide range of depths. Mehdizadeh et al. [35] reported that the adaptive neuro-fuzzy inference system (ANFIS) provided superior results in monthly Ts estimation than ANN and gene expression programming (GEP) with data from 31 stations in Iran. Bilgili et al. [36] estimated monthly Ts at different soil depths (i.e., 5, 10, 20, 50, and 100 cm) using nonlinear regression and MLR models, and obtained the most accurate results at the depth of 5 cm. Feng et al. [37] evaluated the capabilities of ELM and RF models in estimating half-hourly Ts for maize fields, and they concluded that the ELM model had better estimated performance. Sihag et al. [38] utilized M5P, MLP, and RF models to estimate daily Ts in arid regions, the results of which showed that MLP outperformed other models with better performance metrics. Kisi et al. [39] also compared the modeling capabilities of MLP, MLR, and radial basis neural networks (RBNN) for monthly Ts estimation at different soil depths, with inconsistent results obtained depending on depth. At depths of 5 and 10 cm, the RBNN model had the best estimating accuracy among all models used, while at other soil depths (50 and 100 cm), MLR performed better. Mehdizadeh et al. [40] applied SVM and multivariate adaptive regression splines (MARS) to estimate month-by-month Ts for 30 stations in Iran, and it was found that MARS outperformed the SVM model.
Overall, according to the literature, machine learning algorithms have been extensively used in the research of Ts estimation, with the scopes of such studies generally covering various time scales, study regions, and/or soil depths. However, to our best knowledge, the performance of machine learning algorithms on Ts estimation has rarely been reported in the different climatic zones of China, especially for studies with large spatial scales (such as the whole country). Accurate estimation of Ts in different climatic zones would be beneficial to studies of crop modeling, hydrological patterns, and soil properties under large-scale conditions, and may also provide theoretical support for better crop production in different regions [2,41,42]. Recently, extreme gradient boosting (XGBoost) is considered as a promising machine learning algorithm, and has been widely applied in many fields such as meteorology [28,43], hydrology [44,45], and agronomy [46,47]. However, to date no study has been conducted on a national wide estimation of Ts with XGBoost. Therefore, in this study we utilize six different combinations of meteorological variables as inputs to develop XGBoost models for estimating daily Ts at the 0 cm depth, with records from th eperiod 1966–2015 from 689 meteorological stations that covers seven different climatic zones in China. To better evaluate the estimating performance of XGBoost, the M5P and RF algorithms that have applied for Ts estimation in previous studies are added for comparison. The primary purpose of this research is to evaluate the performance of tree-based models to estimate daily Ts in seven different climatic zones (large-scale) of China and its applicability. The effects of different climatic and environmental factors on the performance of each model for estimating Ts are then explored.

2. Materials and Methods

2.1. Study Area

According to previous studies, there are seven different climatic zones in China based on geographical and meteorological data [48,49], which are the arid desert of northwest China (NWC), the semi-arid steppe of Inner Mongolia (IM), the (semi-) humid cold-temperate northeast China (NEC), the semi-humid warm-temperate north China (NC), the humid subtropical central China (CC), the humid tropical south China (SC), and the Qinghai-Tibetan Plateau (QTP) [50] (Figure 1). Among these climatic zones, the QTP has the highest average elevation (up to 4000 m), with considerably high UV intensity and relatively low vegetation coverage. The region is mainly composed of mountains, plateaus, river valleys, and basins. The geology is mainly schist, millstone, sandstone, shale, and volcanic rocks, The QTP region has a variety of ecosystem types and mainly grows barley. The NWC generally has longer daylight hours than other climatic zones. The landscape is dominated by mountains, basins, and deserts. The area is sparsely vegetated, and the desert covers a large area. The crops and fruits are of high yield with excellent quality in places where water is relatively abundant. The IM region receives less rainfalls than other regions. The altitude is generally between 1000 and 1200 m. The terrain is high in the south and low in the north, with 80% of the plateau covered by grassland, which is mainly clumped with grasses. The thickness of the grass layer is between 10 and 60 cm. There are seven main lake areas in the region, including the Hetao and Hulunbuir Lakes. The NEC region has long winters and short summers, with a significant amount of snow and a wet climate. The region is dominated by plains and mountains. There are many marshes and thick mounds of earth. The proximity to the Bohai Sea leads to a long rainy season in the NEC region. The region has large areas of coniferous and mixed coniferous forests and rich black soils. Forestry accounts for 14.7% of the country’s land. The main food crop is wheat. The NC is a relatively small region of China with less than 1000 m3 of water resources per capita. Summers are hot and rainy. Winters are cold and dry. Rainfall varies between 400 and 800 mm. The land type is dominated by the great plains and plateaus, with the terrain decreasing from west to east. The soil type is mainly brown loam. The CC and SC regions generally receive heavy rainfalls in spring, where flooding tend to occur during that time period. The CC region has red loamy soils of low fertility. The hills, plains, basins, rivers and lakes dominate. The main crops grown are rice and oilseed rape. The topography of the SC region is mainly hilly and plain. The average annual temperature is 18–24 °C. Moreover, the area has a large variety of plants, mainly tropical scrub, grassy slopes, and secondary forests. The soil types are mainly brick red loam and russet red loam.

2.2. Case Study and Data

In this study, daily meteorological data from 1966 to 2015 from 689 meteorological stations located in the seven climatic zones of China were collected. Meteorological data consisted of mean air temperature (Tmean), maximum air temperature (Tmax), minimum air temperature (Tmin), relative humidity (RH), global solar radiation (Rs), and wind speed (U2). The selected data were quality controlled. All weather-related variables were measured at a 2 m height. However, it should be noted that Rs was not recorded in all the 689 stations. To fix this problem, the extraterrestrial solar radiation (Ra) was used for interpolation through modelling, if the Rs data were absent. Ts at a depth of 0 cm were measured through probes on the soil surface. Six input combinations were utilized for training the M5P, RF, and XGBoost models. Table 1 showed the types of input combinations. Raw meteorological data used in the study were provided and quality inspected by the National Meteorological Information Center (NMIC) of the China Meteorological Administration (CMA). A further examination was applied to datasets before computing. A number of data were deleted from the original data because they were missing or the ratio between measured Ts and theoretical Ts was greater than one. Tmean was converted from Tmax and Tmin. The M5P, RF, and XGBoost programs were written in R software (version 3.2.3; The R Project for Statistical Computing). All the simulations were performed in a computer with a single Intel Core i7-6700 at 3.4–4.0 GHz and 16 GB of random-access memory (RAM).

2.3. Methodology

2.3.1. M5P Model Tree (M5P)

The M5P model tree (M5P) is a regression tree algorithm that develops conventional decision trees with the addition of linear regression functions to the nodes [51]. This technique has been successful in predicting continuous values, which can be achieved by employing the conversion of the classification problem into a functional optimization problem [52]. The M5 Tree denotes a segmented linear function, testing the value of a specific property at each internal node and predicting the class value at each leaf node. To predict the class value of a new sample, the tree is interpreted starting from the root node. At each internal node, the left or the right branch is selected based on the value of a particular attribute of the sample. The advantage of model trees over regression trees is that regression trees’ computational load increases rapidly as the dimensionality increases, while model trees are significantly less than conventional models. Therefore, model trees are more efficient when handling with high-dimensional data as leaf nodes use linear functions rather than constants, leading to more accurate predictions. All enumerated attributes in the M5P algorithm are converted into binary variables before the tree is constructed. This algorithm can efficiently handle missing values. There are three main steps in the M5P tree algorithm, which are tree construction, tree pruning, and tree smoothing [53]. The basic tree is constructed with the splitting criterion, which takes the standard deviation of the class values reaching a node as the error of that node. It computes the expected error reduction resulting from testing each property at that node. The property that maximizes the expected error reduction is then selected.

2.3.2. Random Forests (RF)

The random forests (RF) algorithm, proposed by Breiman [54], was developed using classification and regression trees (CART) and the concept of “bagging”. As a machine learning algorithm that can effectively solve high-dimensional regression problems, RF has been used extensively in research of regression and estimation, which uses subsets of data through bootstrap to process random binary trees. By repeatedly selecting random T (T < N) sample sets, a new training sample set is generated from the N original training samples. In the whole process of selecting samples, the same part of the samples may be collected repeatedly. A random subset of the training dataset needs to be randomly extracted from the original dataset for the development and training of the model (see the flowchart in Figure 2). Datasets that are not used in the model training are often called out-of-bag (OOB) data. These OOB datasets will not be used for model fitting, but in turn will be used for testing the estimation ability of the model [25].
The RF algorithm is a feature selection based on the Gini coefficient. The criterion for selecting the Gini coefficient is that each child node needs to achieve the highest purity. The smaller the Gini coefficient, the higher the stability of the model and the higher the purity. CART is a binary tree, which means that each non-leaf node can only produce two branches. If multiple (taller than two) discrete variables are generated on a non-leaf node, the variable may be reused multiple times. Each feature selected from the RF tree is randomly produced from all the features, which reduces the risk of overfitting. In contrast to other decision trees, each RF tree is part of the selected feature [25]. Among the selected features in this part, the best feature is picked to partition the left and right subtrees of the decision tree, thereby providing more randomness and further improving the model’s inductive power [25]. In short, the final estimation of the RF algorithm is the average of all factors. Breiman’s research has more details about the RF algorithm [54].

2.3.3. Extreme Gradient Boosting (XGBoost)

Extreme gradient boosting (XGBoost) is a new paradigm for Gradient Boosting Machines (GBMs), which improves the processing of databases through the optimization of decision tree algorithms [55]. The XGBoost algorithm solves the overfitting problem through regularization and built-in cross-validation, which improve the computational accuracy and allow for optimum speed. Furthermore, functions in the XGBoost algorithm are operated and computed automatically, and therefore XGBoost is extensively applied in applications such as classification [56] and estimation [57]. The XGBoost algorithm is derived from the concept of “boosting”, which combines the forecasts of all “weak” learners with special training to foster “strong” learners. The expressions are as follows:
f i t = k = 1 t f k x i = f i t 1 + f t x i
where f k x i and f t x i are the predicted values for the k-th and t-th iterations of the XGBoost model, respectively; f i t and f i t 1 are the predicted values for the t and t−1 iterations of the i-th sample; x i is the input variable; k = [1, 2, …, t], i = [1, 2, …, n].
To prevent overfitting problems without compromising the speed of the algorithm, the XGBoost algorithm can be derived as follows:
O b j t = i = 1 n l f i t ¯ , f i t + i = 1 n Ω f i
where O b j t is the objective function; l is the loss function; f i t ¯ is the true value of the t-th iteration of the i-th sample; and Ω f i is the canonical term of the objective function, which is given by:
Ω f = β T + 1 2 λ ω 2
where β and λ are regularization parameters and T is the number of leaf nodes.

2.3.4. Model Evaluation

In this study, three statistical indicators, including coefficient of determination (R2), root mean square error (RMSE), and mean absolute error (MAE), were chosen to analyze and compare the accuracy and stability of different models for estimating Ts [50,58]. The corresponding formulas are:
R 2 = i = 1 n ( Y i , m Y ¯ i , m ) ( Y i , e Y ¯ i , e ) 2 i = 1 n ( Y i , m Y ¯ i , m ) 2 i = 1 n ( Y i , e Y ¯ i , e ) 2
RMSE = 1 n i = 1 n ( Y i , m Y i , e ) 2
MAE = 1 n i = 1 n Y i , m Y i , e
where Y i , m , Y i , e , Y ¯ i , m , Y ¯ i , e , and n are the measured daily Ts, the estimated daily Ts, the measured average value of daily Ts, the estimated average value of daily Ts, and the number of observations, respectively. Higher R2 values (closer to 1) and lower RMSE and MAE values indicate better estimation performance of the model.

3. Results and Discussion

3.1. Comparison of the Accuracy of Various Machine Learning Models

The overall and average performance of each data-driven model in the seven different climatic zones is shown in Table 2 and Table 3, respectively. As listed in Table 2, averaged across the six input combinations and the seven climatic zones, the XGBoost model (on average R2 = 0.964, RMSE = 2.066 °C, MAE = 1.597 °C) overall performed better in Ts estimation than the RF (on average R2 = 0.959, RMSE = 2.130 °C, MAE = 1.647 °C) and M5P models (on average R2 = 0.954, RMSE = 2.280 °C, MAE = 1.742 °C). Among the six input combinations, models with the complete combination of meteorological variables as inputs (i.e., the input combination 1) had the greatest estimating accuracy in each of the seven climatic zones (Table 3). Under this input combination, averaged across the meteorological stations in each climatic zone, the ranges of the mean statistical indicator values for M5P1, RF1, and XGBoost1 at different climatic zones were 1.722–2.875 °C, 1.354–2.257 °C, and 1.342–2.208 °C for RMSE, 0.928–0.975, 0.956–0.984, and 0.959–0.985 for R2, and 1.245–2.106 °C, 1.028–1.702 °C, and 0.992–1.651 °C for MAE, respectively. The second best estimating accuracy among input combinations was observed in models with the combination of Tmean and Rs (i.e., the input combination 4) as inputs, which were M5P4, RF4, and XGBoost4 (with corresponding ranges of 1.658–2.830 °C, 1.506–2.556 °C, and 1.446–2.435 °C for RMSE, 0.929–0.975, 0.941–0.980, and 0.9494–0.983 for R2, and 1.214–2.093 °C, 1.119–1.915 °C, and 1.084–1.840 °C for MAE, respectively). Although input combination 4 yielded slightly worse model accuracy than the complete combination, it required less input variables and a much smaller dataset. It was noted that models with Tmean as inputs all achieved relatively good performance in estimation, including models with the complete input combination (M5P1, RF1, and XGBoost1) and the combinations of Tmean and RH (M5P3, RF3, and XGBoost3), Tmean and Rs (M5P4, RF4, and XGBoost4) and Tmean and U2 (M5P6, RF6, and XGBoost6). These results confirm the argument that temperature is the most fundamental meteorological factor for data-driven models in estimating Ts. Among the other three meteorological factors, Rs had the highest degree of influence on estimating accuracy, followed by RH, and then U2. Therefore, considering the model estimating accuracy and the size of input data jointly, models using the input combination of Tmean and Rs would have a higher potential in Ts estimation across different climatic zones in China than models with other input combinations. For the three algorithms in this study, the overall model performance was generally ranked as XGBoost > RF > M5P with any of the input combinations in any of the climatic zones, with the exception that M5P2 was always better than RF2 in all climatic zones and even better than XGBoost2 in some climatic zones (e.g., NWC, IM, and QTP).
The box plots of the RMSE values of the three data-driven models (i.e., M5P, RF, and XGBoost models) for estimating Ts with the six input combinations at the 689 stations of China are shown in Figure 3. The RMSE values of the XGBoost model under the six input combinations were 1.001–2.823 °C, 1.272–3.480 °C, 1.121–3.259 °C, 1.050–2.981 °C, 1.173–3.540 °C, and 1.231–3.422 °C, respectively, which overall were lower than the corresponding values of the M5P model (1.241–3.618 °C, 1.056–3.447 °C, 1.133–3.638 °C, 1.174–3.478 °C, 1.250–4.059 °C, and 1.107–3.758 °C, respectively) or the RF model (1.051–2.873 °C, 1.082–3.611 °C, 1.111–3.431 °C, 1.090–3.123 °C, 1.170–3.761 °C, and 1.153–3.619 °C, respectively). Averaged across all stations, the mean RMSE values for the XGBoost model under the six input combinations decreased by 23.1%, −1.4%, 5.5%, 13.2%, 11.2%, and 3.4% when compared with the M5P model, and by 3.0%, 1.4%, 1.9%, 4.3%, 4.5%, and 1.7% when compared with the RF model. The median RMSE values of the three models showed a similar pattern to the mean RMSE values. These results together indicated that, among the three models in this study, the XGBoost model had the highest overall performance, followed by the RF model and then the M5P model.

3.2. Comparison of the Spatial Distribution of Errors in Estimating Soil Temperature

Spatial distributions of the performance in estimating Ts for the data-driven models with all climatic variables as inputs (M5P1, RF1, and XGBoost1 models) are shown in Figure 4. Generally, the models all performed well in estimating Ts at the majority of stations, with the values of RMSE less than 2.0 °C and the values of MAE less than 1.50 °C at 76.8% of all stations studied. Overall, there was some spatial variation in model performance between climatic zones. The accuracy of the model was relatively better in the CC and SC regions than in other regions, while the model performance was relatively poor in the NEC region, especially in the northeast provinces of China. Among the three models, M5P1 had the worst performance in estimation, with RMSE values greater than 2.01 °C and MAE values greater than 1.50 °C in up to 47.7% of the stations in all regions, except for the CC region where only a small fraction of stations had similar poor performance. On the contrary, the XGBoost1 model had the most satisfactory in estimating Ts, with RMSE values less than 1.51 °C and MAE values less than 1.25 °C in most stations of all regions, except for the NEC region.
The performance of the M5P2, RF2, and XGBoost2 models for estimating Ts are illustrated in Figure S1. The three models were trained using Tmean as the only input available. The three models generally had a positive performance in estimating Ts at the majority of stations, with RMSE values greater than 2.01 °C and MAE values greater than 1.50 °C at 48.8% of the stations. Although there was still relatively good estimating accuracy in some stations of the CC and SC regions, the overall performance of the three models was poor in other regions of this study. In particular, the IM and NEC regions had up to 71.9% of the stations with model RMSE values greater than 2.51 °C and MAE values greater than 2.00 °C. Overall, the estimation accuracy of the XGBoost2 model was slightly greater than both other models. These results suggest that there is space for improvement in the estimation accuracy of the three models when using Tmean as the only input. Our findings also confirm the conclusion of Talaee’s study that the appropriate addition of other meteorological factors to Tmean could effectively improve model performance in estimating Ts [59].
The spatial distribution of the performance for the data-driven models (M5P3, RF3, and XGBoost3 models) on Ts estimation with Tmean and RH as inputs is illustrated in Figure S2. The models generally had positive performance in estimating Ts at the majority of stations, with RMSE values less than 2.0 °C and MAE values less than 1.50 °C at 56.2% of the stations. Among the climatic zones, the three models performed relatively better in the CC region than other regions, while the poorest model performance was observed in the NEC region, where the RMSE and MAE values in 78.5% of the stations were greater than 2.51 °C and 2.00 °C, respectively. Models showed very similar estimating performance between the NWC and the QTP regions. Compared with the M5P1, RF1, and XGBoost1 models, the accuracy of models with only Tmean and RH as inputs was slightly worse. Therefore, reducing the number of meteorological factors as inputs would lower the estimation performance of each model, which is consistent with findings obtained by Feng et al. [60]. This result suggests that using a complete set of relevant meteorological factors as inputs in studies of Ts estimation could effectively improve the estimating performance of models, which confirms the findings from previous studies [61].
The spatial distribution of estimating performance for M5P4, RF4, and XGBoost4 models that were trained with Tmean and Rs as inputs are illustrated in Figure S3. The models all showed positive performance in modeling Ts for a large proportion of stations, with RMSE values less than 2.0 °C and MAE value less than 1.50 °C at up to 71.6% of the stations. Among the climatic zones, the estimating accuracy of the three models was relatively excellent in the CC and SC regions, with RMSE values less than 1.50 °C and MAE values less than 1.25 °C at 45.5% of the stations, while in contact the model performance in the NEC region was relatively poor, with up to 64.9% of the stations showing values of RMSE and MAE greater than 2.51 °C and 1.75 °C, respectively. Among the three models, the estimating performance of XGBoost4 was slighter better than RF4, but both of them obviously outperformed M5P4. Overall, patterns of the spatial variability in model performance for this input combination (i.e., Tmean and Rs) were very similar to patterns shown in models with the complete combination of meteorological variables as inputs, indicating that the removal of RH and U2 from the inputs did not noticeably decrease the estimation accuracy of models in this study. Moreover, the input combination of Tmean and Rs required far less meteorological data than the complete input combination. Therefore, it can be concluded that Tmean and Rs are key input variables for estimating Ts in different climate zones, consistent with findings obtained by Talaee [59] and Bilgili [21].
For the data-driven models with Tmax and Tmin as inputs (i.e., M5P5, RF5, and XGBoost5), the spatial distribution of their estimating performance is demonstrated in Figure S4. Overall, the three models performed poorly in estimating Ts at most stations, with 55.8% of the stations showing the values of RMSE and MAE above 2.01 °C and 1.50 °C, respectively. Models with Tmax and Tmin as input variables performed worse than the M5P1, RF1, and XGBoost1 models, indicating that using temperature as the only input climatic variable would not be sufficient enough to obtain satisfactory model performance.
The spatial distributions of the estimating performance for the M5P6, RF6, and XGBoost6 models that were trained using Tmean and U2 as inputs are shown in Figure S5. The three models showed poor performance in modeling Ts at most stations, with RMSE values greater than 2.01 °C and MAE values greater than 1.50 °C at 58.7% of the stations. Although the estimating accuracy of the three models was relatively good in the CC region, their estimating performance was poor in other regions. For instance, in the IM and NEC regions, up to 78.4% of the stations showed values of RMSE and MAE greater than 2.51 °C and 2.00 °C, respectively.
A combined comparison of Figures S2–S5 indicates that the models showed an overall trend of decreasing accuracy from south to north. This would lead us to think that it is related to latitude, whereas latitude could affect Ts due to the angle of incidence by solar radiation [62]. Regarding the effect of latitude on Ts, scholars have reached similar conclusions in studies from other countries. In a study conducted by Fitton and Brooks [63] in the USA, it was concluded that Ts decreased with increasing latitude. The model with inputs Tmean and Rs would be slightly more accurate in estimation than under several other combinations of inputs, which was very close to the performance of each model at full input of the parameter variables. Based on Tmean, the factor Rs had a greater influence on the accuracy of each model than several other meteorological factors, followed by RH and U2. Because this research estimated Ts at 0 cm depth, the air temperature was most closely related to Ts [64]. However, the greater degree of influence of Rs might be due to the source of soil heat. Ts is one of the components of soil heat. However, soil heat is mainly formed by solar radiation, heat emitted from the Earth’s interior, and heat generated by the decomposition of microorganisms in the soil [65]. Of these, solar radiation dominates. Therefore, Rs had a greater degree of influence in estimating Ts. Nahvi et al. [66] and Huang et al. [67] also concluded that Tmean was significantly relevant in estimating Ts. It suggested that the input combination of Tmean and Rs was the most generalizable for studies estimating Ts in the different climatic zones of China.

3.3. Analysis of Factors Affecting the Model’s Estimation of Soil Temperature

To compare the level of linearity between the estimated and the measured values of Ts for each model, we randomly selected one meteorological station per region (including stations 51,477, 50,618, 50,353, 52,986, 56,188, 59,287, and 52,984, respectively) out of the 689 stations in the seven different climatic zones across the country as examples for description. Outputs from models with input combinations 1, 5, and 6 were plotted for comparison (Figure 5, Figures S6 and S7). It is clear from Figure 5, Figures S6 and S7 that the three types of models (i.e., the M5P, RF, and XGBoost models) all performed very well in estimating Ts. For any of the models constructed with the selected input combinations in the randomly selected stations, there was a significant positive correlation between the estimated and the measured Ts values, with R2 values at least higher than 0.92.
For models with the complete combination of climatic variables (i.e., combination 1) as inputs, XGBoost outperformed both M5P and RF at each of the randomly selected stations (Figure 5). Across the stations, the average R2 values of XGBoost models were 0.981, higher than RF (averaged R2 = 0.978) and M5P (averaged R2 = 0.966) models. For each type of model, the estimating accuracy differed among stations, suggesting that variations in geographical and environmental factors among different climatic zones could affect the model performance in Ts estimation to some extent, as suggested by Kassaye et al. [68]. Taking XGBoost as an example, except for station 59,287 (R2 = 0.966), the model accuracy was considerably high at all stations (all R2 values no less than 0.980; averaged R2 = 0.983), with the highest accuracy observed at station 51,477 (R2 = 0.989). The significant difference in model estimating accuracy between station 51,477 and station 59,287 could be attributed to the differences in climatic factors (e.g., temperature) between climatic zones. Station 51,477 is located in the NWC region, with sufficient light, low soil water contents, and low variation of Ts. In contrast, station 59,287 is within the SC region, where the soil moisture is sufficient and the variation in Ts is high. First of all, we selected stations that are all weather stations. The terrain of the station measurement points is short grass and flat, so the slope and some human factors are not considered to influence Ts. However, soil characteristics differ between climatic zones. For example, the soil in the CC area is mainly red loam. The main characteristic of red loam soil is strong clay and little organic matter. Therefore, the soil heat transfer is faster, which means that Ts does not change much. Thus, the model estimated Ts with high accuracy. However, the soil in the NEC area is mainly black soil. It is characterized by fertile land and high organic matter content. However, as we know, high organic matter content means higher heat requirements for microbial decomposition. Therefore, the variation in Ts varied widely, leading to a decrease in the accuracy of model estimation. Zhang et al. [69] concluded that the variation of Ts in different climatic zones was influenced by a combination of air temperature and precipitation. More precipitation occurs in humid areas and less in arid regions. These findings are consistent with the conclusion made by Knight et al. [70], where they suggested that Ts could be directly affected by air temperature and indirectly impacted by soil moisture. In addition, other factors such as soil depth, soil characteristics, irrigation levels, and precipitation could also affect the estimation of Ts to some extent, as described in the literature [9,37,71,72,73]. For example, the variability of Ts at different soil depths generally vary, with greater variability between measured and estimated values observed in deeper soils. The literature indicates that the thermal conductivity of soil decreases with increasing soil porosity but increases with increasing soil water content [74]. Therefore, both enhanced rainfall and reduced porosity could improve the thermal conductivity of the soil, which could make Ts less variable [75]. In addition, the increase in the organic matter content of the soil also reduces the thermal conductivity [76]. The degree of irrigation, similar to precipitation, can indirectly affect the soil water content. The increase in soil water content increases the heat capacity in the soil, which results in a smaller range of variation in Ts [74].
With combination 5 (i.e., including Tmax and Tmin) as inputs (Figure S6), the accuracy of each model at each station was lower than the accuracy of its corresponding model with the complete input combination, indirectly suggesting that adding more relevant meteorological factors to the inputs would improve the model performance in estimating Ts. This result is consistent with the conclusion made by Sanikhani et al. [2] and Kim and Singh [6]. Among the selected stations, the best-performing models were observed at station 56,188, where the R2 values of the M5P, RF, and XGBoost models were 0.969, 0.971, and 0.976, respectively (Figure S6). For station 56,188 (within the CC region) and station 59,287 (within the SC region), the range of variations in Ts was smaller than other stations within other climatic zones, resulting in the scatters of the regression at the two stations being denser than at other stations. This outcome might be due to the fact that these two stations are located at regions where the climatic difference between them is similar, but they are distinct from other stations in terms of climate. Similar patterns were found by other studies, in which the estimation of reference crop evapotranspiration (ET0) in different climatic zones was conducted [50].
For combination 6 (i.e., including Tmean and U2 as inputs), XGBoost models outperformed both RF and M5P models in terms of estimating accuracy, similar to other input combinations (Figure S7). Across the selected stations, the R2 values of XGBoost models averaged 0.966, slightly higher than the RF and M5P models, which had the same estimating accuracy on average (both averaged R2 = 0.960). Under this input combination, the accuracies of the M5P and RF models were very close at each of the stations. At some of the stations (e.g., stations 56,188 and 59,287), the M5P model even slightly outperformed the RF model. Among these stations, the best model performance in estimating Ts was observed in station 51,744, where the XGBoost model had the highest accuracy (R2 = 0.976). Compared to models with input combination 5, there were improvements to some extent in the estimating accuracy of models with input combination 6 at all selected stations, except for station 59,287. The estimating accuracies of the M5P, RF, and XGBoost models at station 59,287 decreased by 2.0%, 2.5%, and 2.1%, respectively, when comparing the input combinations. This result might be due to the high variability of wind speeds in the region where station 59,287 was located, which could impair the model accuracy in estimation. A previous study by Kim and Singh [6] had similar findings, which confirmed that U2 could significantly affect the performance of estimated Ts.
To quantify the effects of soil water content, soil organic matter content, and other meteorological factors on Ts estimation by models in different climatic zones, the relationships between these factors and the RMSE values of XGBoost1 were analyzed (Figure 6). The spatial distribution of soil organic matter content and soil water content for the selected stations are plotted in Figure 7. Data from different climatic zones were fitted separately. For each of the soil texture factors or the meteorological factors, its correlation with the RMSE of Ts differed between climatic zones, suggesting that the model accuracy in estimating Ts was related to the soil texture and the climate of the region where the meteorological stations were located. Different soil textures have different permeability and nutrient contents, which partially leads to differences of Ts variability and consequently affects the model accuracy for Ts estimation. This pattern is consistent with results described in the previous section.
As shown in Figure 6a, the most significant correlation between soil organic matter content and model estimating accuracy was found in the NC region (R2 = 0.255), followed in order by the IM, the NWC, and the CC regions. In other regions, including NEC, SC, and QTP, no significant correlation was found. In regions where the correlation was significant, the model accuracy for estimating Ts decreased with the increasing soil organic matter content. This is due to the fact that the specific heat capacity of soil organic matter is less than the soil specific heat capacity. Under certain conditions, the higher the organic matter content is, the more heat absorption should be required for decomposition, leading to more variable and unstable Ts. Wang et al. [77] also concluded that soil organic matter could significantly affect the variation of Ts. According to the results, the NEC, SC, and QTP regions had relatively high soil organic matter contents, indicating that too much soil organic matter may lead to other types of reactions within the soil and make the model estimation of Ts less stable [78,79].
The correlation between soil water content and model estimating accuracy was most significant in the IM region (R2 = 0.175; Figure 6b), followed by the CC region, and then the QTP and the NC regions. Soil water content was not significantly correlated with the accuracy of estimating Ts in the NWC, NEC, and SC regions. In most regions where the correlation was significant, the model estimating accuracy increased with the increasing soil water content. This might be attributed to the fact that the specific heat capacity of water is about four times higher than that of soil. The temperature change of the soil water is much smaller than the soil itself when equally heated or cooled, suggesting that soil water can indirectly regulate Ts and reduce the range of Ts variation. Yuan [80] also concluded that soil water content had an important influence on soil thermal properties and could effectively regulate Ts. Therefore, under certain conditions, the higher the soil water content is, the smaller is the expected range of Ts variation, leading to higher model accuracy in Ts estimation. This finding is consistent with conclusions from previous studies [81,82]. It should be noted that a contrasting pattern was observed in the IM region, where the model estimating accuracy decreased with the increasing soil water content. This phenomenon might be attributed to the relatively low soil water content in this region. For most stations in the IM region, their soil water contents were lower than 10%, which was distinct from the other six regions in the study. Therefore, soil water content might not play an important role in determining the pattern of model estimating accuracy in the IM region, in which other factors such as soil organic matter might be more influential.
RH was highly significantly correlated with the accuracy of model-estimated Ts in the CC region, which might be due to the fact that the region is in a humid area with high perennial air humidity (Figure 6c). When the relative humidity is high, there is considerable moisture in the air. Given the fact that the specific heat capacity of water is larger than air, the heat transferred by the air temperature and solar radiation to soils is partially absorbed by the moisture in the air, thus reaching the amount of heat to the surface of the soil. Sawan [83] also suggested that RH had a strong influence on air temperature variation and indirectly affected the heat absorbed by the soil.
Solar radiation is one of the main sources of soil heat. In our study, Rs showed significant correlation with model accuracy for estimating Ts in the NWC and CC regions, which might be caused by the longer insolation time in the NWC region and the higher insolation intensity in the low-latitude CC region (Figure 6d). Solar radiation can be directed to the surface, allowing the surface to absorb heat and thus increase the Ts. High latitudes have a high tilt of solar irradiation, so that less solar radiation energy is absorbed on the soil per unit area. This results in lower Ts. In contrast, the Ts is higher at low latitudes because the solar radiation absorbed per unit area of soil is higher due to direct sunlight to the ground [62]. Olchev et al. [84] also confirmed the important role of solar radiation in the conversion of Ts.
Wind speed affects the flow of air. As shown Figure 6e, U2 was highly significantly correlated with the accuracy of model-estimated Ts in the NEC region, which might be due to the fact that the NEC is in a cold-temperate region with frequent cold air and high wind speed all year round. When the U2 is low, heat exchange between soil and air is reduced, thus reducing the variation in Ts. Conversely, the heat loss from air is generally fast when U2 is high, resulting in significant variation in Ts. Kong et al. [85] studied the effect of seasonal variation in U2 on Ts and suggested that U2 affects the heat uptake by soils to some extent, thus causing changes in Ts.
We also explored the extent of the contribution of these factors to the estimating error of the model and the relationship between them through principal component analysis (PCA) and correlation coefficients (Figure 8). PCA was used to explain the degree of influence of the factors on the model error. As shown in Figure 8a, the total model error explained by each factor was 97.9%. Among them, PC1 accounted for 96.2% and PC2 accounted for 1.7%. Soil water content, Tmean, Rs, and U2 were all located in the same quadrant. These factors, illustrated in Figure 8a, are shown to be the main factors contributing to the error in estimating Ts. Zhang [86], when he studied the effect of snow cover on Ts, also proposed that air temperature alone does not fully explain the variation in Ts. The Rs, soil water content, and RH can also combine to influence the variation of Ts, whereas soil organic matter content and U2 were additionally proposed to explain the Ts variation in this study. For each of the factors, there was a significant relationship between it and the estimated RMSE values of Ts (Figure 8b). Among these factors, U2 had the lowest correlation (R = 0.14), whereas RH had the highest correlation (R = −0.48) and soil water content had the second highest correlation (R = 0.38), indicating that the heat absorption capacity of water had a more obvious effect on Ts than other factors such as Tmean, Rs, and soil organic matter content. To summarize, soil organic matter, soil water content, Tmean, RH, Rs, and U2 are the main factors that cause errors in estimating Ts, all of which would have, to some extent, a critical impact on the model estimation of Ts.

3.4. Comparison of Estimation Performance in Different Climate Zones

To compare the estimating performance for each of the 18 models (3 machine learning algorithms × 6 input combinations) at each of the stations within each of the climatic zones in this study, a combined metric based on the RMSE and MAE values of each model was calculated. The value of the combined metric ranged from 0 to 1, in which the closer the value to 0, the better the model performance was, while on the contrary the closer the value to 1, the worse the model performance was. Values of the combined metric at different stations for each climate zone are illustrated with heatmaps (see Figure 9 and Figures S8–S13). Overall, under the same input combination, the estimation performance of the XGBoost model was generally better than that of the RF and M5P models at most stations (Figure 9 and Figures S8–S13). Among all the 18 models with different input combinations, the XGBoost1 and RF1 models generally performed best at most stations in the seven climate zones, of which XGBoost1 had better performance than RF1. Among the incomplete input combinations, models with combination 4 (i.e., including Tmean and Rs) generally performed better than models with other combinations, of which the models could be ranked as XGBoost4 > RF4 > M5P4 in terms of estimation performance. It is noted that the performance of the XGBoost4 model was relatively close to that of the XGBoost1 model, while the former required fewer input datasets than the latter. Therefore, the XGBoost4 model would be more generalizable than the XGBoost1 model for estimating Ts across different climatic zones. Overall, the model performance differed between different climatic zones. For example, the M5P5 and RF5 models had relatively poor estimating performance in the NWC, IM, NEC, NC, and QTP regions, while the estimation performance of the M5P6, RF6, and XGBoost6 models was poor in the SC and CC regions. These results confirmed that the model estimating performance would be impacted by variations among different climatic zones.

3.5. Comparison of Stability of Various Machine Learning Models

To further compare the performance of different machine learning models, the absolute errors (AE) for each model was also adopted as a statistical metric. The frequency distribution of AE between measured and estimated daily Ts derived from the three machine learning models under various input combinations during the testing stage was plotted (Figure 10). As shown in the histograms, the AE of the three models mainly distributed at the ranks of <0.3 °C, while the proportions of AE greater than 0.5 °C for the models only ranged from 1.0% to 3.9%, which together suggested that the three types of models all performed well in Ts estimation. The proportions of AE less than 0.3 °C differed among model types and input combinations. For the input combination of Tmean and Rs (i.e., combination 4), the proportions of M5P and XGBoost models with AE less than 0.3 °C ranged from 87.4% to 90.9%, higher than the proportions of their corresponding counterparts with the input combination of Tmax and Tmin (ranging from 81.6% to 85.1%). Kisi et al. [39] also concluded that the input combination of Tmean and Rs provided a more effective estimation of Ts in the Turkish region. For the complete input combination, the proportions of M5P and XGBoost models with AE less than 0.3 °C ranged from 87.0% to 93.1%. Moreover, the XGBoost model had the highest proportion of AE less than 0.1 (up to 55.2%) with this input combination among all models, indicating that the XGBoost model had stronger overall estimating performance than both other models. In addition, the XGBoost model outperformed the RF and M5P models in data processing speed, with less computing time required, consistent with results obtained by former research [29,87,88].
Taking all the previous results together, it can be concluded that, among the different climate zones of China in this study, the best-performed Ts estimations were observed in the CC region, followed by the SC region. In contrast, models constructed in the study generally showed poor performance in the NEC and IM regions. Models with the combination of Tmean and Rs as inputs could archive relatively high performance in estimation, approaching the performance of models in the fully parametric input pattern. Moreover, the former required fewer datasets and was more generalizable. In terms of different types of models, the XGBoost model would outperform the other two models. Moreover, soil organic matter content, soil water content, Tmean, RH, Rs, and U2 were significantly related to the accuracy of the model estimation of Ts and were the main factors that caused errors in the model estimation of Ts.

4. Conclusions

Soil temperature (Ts) has an important influence on the energy conversion between the atmosphere and the soil. In this study, three data-driven algorithms (i.e., M5P, RF, and XGBoost) were selected to model Ts at a 0 cm depth using different combinations of meteorological factors (including Tmax, Tmin, Tmean, RH, Rs, and U2) as inputs. The three types of tree-based models were evaluated and compared at 689 meteorological stations in seven different climate zones of China. The results showed that the XGBoost model performed better and was more computationally efficient for estimating Ts compared to the M5P and RF models in different climate zones. With the inputs of Tmean and Rs, the models exhibited similar performance to that of models with the complete combination of meteorological factors as inputs, both of which had good estimation performance, but the former required fewer meteorological factors and datasets. On the basis of Tmean, adding Rs to the model input had more significant influence and importance on model estimating accuracy than adding other climatic factors such as RH and U2 to the input. There was diversity in the estimating performance of the models across climate zones for the same combination of inputs. Comparing different climatic zones, the lower variability of temperature and radiation in the CC and SC regions resulted in a better model estimating accuracy, while the larger variability of radiation in the IM and NEC regions led to poorer model performance for estimating Ts. Soil organic matter content, soil water content, Tmean, RH, Rs, and U2 were significantly related to model accuracy for estimating Ts, suggesting that these factors all contributed to model errors during the estimation of Ts. Overall, for estimating Ts at the 0 cm depth, the XGBoost model is highly recommended, and if the study area covers different climatic zones of China, the XGBoost model with the combination of Tmean and Rs as inputs would obtain satisfactory results for Ts estimation across different regions. However, it is to be noted that there is variability in Ts at different soil depths, and other meteorological factors (e.g., precipitation) would also impact Ts. Therefore, for further studies of Ts estimation, the use of different soil depths as well as additional meteorological and environmental factors should be considered.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/app12105088/s1. Figure S1. Performance of data-driven models (M5P2: a, b; RF2: c, d; XGBoost2: e, f) for the estimation of Ts with Tmean as inputs. Figure S2. Performance of data-driven models (M5P3: a, b; RF3: c, d; XGBoost3: e, f) for the estimation of Ts with Tmean and RH as inputs. Figure S3. Performance of data-driven models (M5P4: a, b; RF4: c, d; XGBoost4: e, f) for the estimation of Ts with Tmean and Rs as inputs. Figure S4. Performance of data-driven models (M5P5: a, b; RF5: c, d; XGBoost5: e, f) for the estimation of Ts with Tmax and Tmin as inputs. Figure S5. Performance of data-driven models (M5P6: a, b; RF6: c, d; XGBoost6: e, f) for the estimation of Ts with Tmean and U2 as inputs. Figure S6. Scatter plot of estimated Ts versus measured Ts from the three machine learning models with Tmax and Tmin as inputs. Figure S7. Scatter plot of estimated Ts versus measured Ts from the three machine learning models with Tmean and U2 as inputs. Figure S8. Combined indicator values for each model at all stations in the IM region. Figure S9. Combined indicator values for each model at all stations in the NEC region. Figure S10. Combined indicator values for each model at all stations in the NC region. Figure S11. Combined indicator values for each model at all stations in the CC region. Figure S12. Combined indicator values for each model at all stations in the SC region. Figure S13. Combined indicator values for each model at all stations in the QTP region.

Author Contributions

Conceptualization, J.D. and L.W.; methodology, J.D. and L.W.; software, L.W.; validation, G.H., J.D. and L.W.; formal analysis, J.D. and L.W.; investigation, F.L., S.L., and Y.C.; data curation, J.D. and L.W.; writing—original draft preparation, J.D.; writing—review and editing, L.W. and G.H.; visualization, Y.W., M.L., J.W. and S.W.; funding acquisition, L.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 51709143, and Jiangxi Natural Science Foundation of China (No. 20181BBG78078 and 20212BDH80016).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The meteorological station data for this study were obtained from the China Meteorological Data Sharing Network (http://www.cma.cn, accessed on 1 December 2020).

Acknowledgments

Thanks to the National Meteorological Information Center of China Meteorological Administration for offering the meteorological data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hu, G.; Zhao, L.; Wu, X.; Li, R.; Wu, T.; Xie, C.; Qiao, Y.; Shi, J.; Cheng, G. An analytical model for estimating soil temperature profiles on the Qinghai-Tibet Plateau of China. J. Arid Land 2016, 8, 232–240. [Google Scholar] [CrossRef]
  2. Sanikhani, H.; Deo, R.C.; Yaseen, Z.M.; Eray, O.; Kisi, O. Non-tuned data intelligent model for soil temperature estimation: A new approach. Geoderma 2018, 330, 52–64. [Google Scholar] [CrossRef]
  3. Lizcano-Toledo, R.; Reyes-Martín, M.P.; Celi, L.; Fernández-Ondoño, E. Phosphorus dynamics in the Soil–Plant–Environment relationship in cropping systems: A review. Appl. Sci. 2021, 11, 11133. [Google Scholar] [CrossRef]
  4. Liu, Z.; Li, D.; Zhang, J.; Saleem, M.; Zhang, Y.; Ma, R.; He, Y.; Yang, J.; Xiang, H.; Wei, H. Effect of simulated acid rain on soil CO2, CH4 and N2O emissions and microbial communities in an agricultural soil. Geoderma 2020, 366, 114222. [Google Scholar] [CrossRef]
  5. Barros, N. Thermodynamics of soil microbial metabolism: Applications and functions. Appl. Sci. 2021, 11, 4962. [Google Scholar] [CrossRef]
  6. Kim, S.; Singh, V.P. Modeling daily soil temperature using data-driven models and spatial distribution. Theor. Appl. Climatol. 2014, 118, 465–479. [Google Scholar] [CrossRef]
  7. Karnieli, A.; Agam, N.; Pinker, R.T.; Anderson, M.; Imhoff, M.L.; Gutman, G.G.; Panov, N.; Goldberg, A. Use of NDVI and land surface temperature for drought assessment: Merits and limitations. J. Clim. 2010, 23, 618–633. [Google Scholar] [CrossRef]
  8. Jackson, T.; Mansfield, K.; Saafi, M.; Colman, T.; Romine, P. Measuring soil temperature and moisture using wireless MEMS sensors. Measurement 2008, 41, 381–390. [Google Scholar] [CrossRef]
  9. Samadianfard, S.; Asadi, E.; Jarhan, S.; Kazemi, H.; Kheshtgar, S.; Kisi, O.; Sajjadi, S.; Manaf, A.A. Wavelet neural networks and gene expression programming models to predict short-term soil temperature at different depths. Soil Tillage Res. 2018, 175, 37–50. [Google Scholar] [CrossRef]
  10. Porada, P.; Ekici, A.; Beer, C. Effects of bryophyte and lichen cover on permafrost soil temperature at large scale. Cryosphere 2016, 10, 2291–2315. [Google Scholar] [CrossRef] [Green Version]
  11. Hu, Q.; Feng, S. A daily soil temperature dataset and soil temperature climatology of the contiguous United States. J. Appl. Meteorol. 2003, 42, 1139–1156. [Google Scholar] [CrossRef]
  12. Ozturk, M.; Salman, O.; Koc, M. Artificial neural network model for estimating the soil temperature. Can. J. Soil Sci. 2011, 91, 551–562. [Google Scholar] [CrossRef]
  13. Yue, J.; Chen, X.; Zhang, W. Research on an evaluation method for the Job-Housing Spaces of megacities using different scales based on multisource data integration: A case study from shenzhen. IOP Conf. Ser. Earth Environ. Sci. IOP Publ. 2019, 264, 012016. [Google Scholar] [CrossRef]
  14. Changnon, S.A. A rare long record of deep soil temperatures defines temporal temperature changes and an urban heat island. Clim. Chang. 1999, 42, 531–538. [Google Scholar] [CrossRef]
  15. DeBoer, T.A. Relationships between the Newhall Simulation Model and Dryland Corn Yield in the Major Land Resource Areas of Nebraska. Master’s Thesis, University of Nebraska at Omaha, Omaha, NE, USA, 2007. [Google Scholar]
  16. Watson, C.L. Seasonal soil temperature regimes in south-eastern Australia. Soil Res. 1980, 18, 325–331. [Google Scholar] [CrossRef]
  17. Wu, W.; Tang, X.; Ma, X.; Liu, H. A comparison of spatial interpolation methods for soil temperature over a complex topographical region. Theor. Appl. Climatol. 2016, 125, 657–667. [Google Scholar] [CrossRef]
  18. Yang, Y.; Wu, Z.; He, H.; Du, H.; Wang, L.; Guo, X.; Zhao, W. Differences of the changes in soil temperature of cold and mid-temperate zones, Northeast China. Theor. Appl. Climatol. 2018, 134, 633–643. [Google Scholar] [CrossRef]
  19. Cheon, J.; Ham, B.; Lee, J.; Park, Y.; Lee, K. Soil temperatures in four metropolitan cities of Korea from 1960 to 2010: Implications for climate change and urban heat. Environ. Earth Sci. 2014, 71, 5215–5230. [Google Scholar] [CrossRef]
  20. Abdul-Wahab, S.A.; Bakheit, C.S.; Al-Alawi, S.M. Principal component and multiple regression analysis in modelling of ground-level ozone and factors affecting its concentrations. Environ. Modell. Softw. 2005, 20, 1263–1271. [Google Scholar] [CrossRef]
  21. Bilgili, M. Prediction of soil temperature using regression and artificial neural network models. Meteorol. Atmos. Phys. 2010, 110, 59–70. [Google Scholar] [CrossRef]
  22. An, K.; Wang, W.; Wang, Z.; Zhao, Y.; Yang, Z.; Chen, L.; Zhang, Z.; Duan, L. Estimation of ground heat flux from soil temperature over a bare soil. Theor. Appl. Climatol. 2017, 129, 913–922. [Google Scholar] [CrossRef]
  23. 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. Energ. Convers. Manag. 2018, 164, 102–111. [Google Scholar] [CrossRef]
  24. Zhou, B.; Erell, E.; Hough, I.; Rosenblatt, J.; Just, A.C.; Novack, V.; Kloog, I. Estimating near-surface air temperature across Israel using a machine learning based hybrid approach. Int. J. Climatol. 2020, 40, 6106–6121. [Google Scholar] [CrossRef] [Green Version]
  25. Dong, J.; Wu, L.; Liu, X.; Li, Z.; Gao, Y.; Zhang, Y.; Yang, Q. Estimation of daily dew point temperature by using bat algorithm optimization based extreme learning machine. Appl. Therm. Eng. 2020, 165, 114569. [Google Scholar] [CrossRef]
  26. Adnan, R.M.; Petroselli, A.; Heddam, S.; Santos, C.A.G.; Kisi, O. Short term rainfall-runoff modelling using several machine learning methods and a conceptual event-based model. Stoch. Env. Res. Risk A 2021, 35, 597–616. [Google Scholar] [CrossRef]
  27. Fan, J.; Wu, L.; Zhang, F.; Cai, H.; Zeng, W.; Wang, X.; Zou, H. Empirical and machine learning models for predicting daily global solar radiation from sunshine duration: A review and case study in China. Renew. Sust. Energ. Rev. 2019, 100, 186–212. [Google Scholar] [CrossRef]
  28. Dong, J.; Wu, L.; Liu, X.; Fan, C.; Leng, M.; Yang, Q. Simulation of daily diffuse solar radiation based on three machine learning models. Comput. Modeling Eng. Sci. 2020, 123, 49–73. [Google Scholar] [CrossRef]
  29. Wu, L.; Peng, Y.; Fan, J.; Wang, Y.; Huang, G. A novel kernel extreme learning machine model coupled with K-means clustering and firefly algorithm for estimating monthly reference evapotranspiration in parallel computation. Agr. Water Manag. 2021, 245, 106624. [Google Scholar] [CrossRef]
  30. Yan, S.; Wu, L.; Fan, J.; Zhang, F.; Zou, Y.; Wu, Y. A novel hybrid WOA-XGB model for estimating daily reference evapotranspiration using local and external meteorological data: Applications in arid and humid regions of China. Agr. Water Manag. 2021, 244, 106594. [Google Scholar] [CrossRef]
  31. Yang, C.; Prasher, S.O.; Mehuys, G.R. An artificial neural network to estimate soil temperature. Can. J. Soil Sci. 1997, 77, 421–429. [Google Scholar] [CrossRef]
  32. Mihalakakou, G. On estimating soil surface temperature profiles. Energ. Build. 2002, 34, 251–259. [Google Scholar] [CrossRef]
  33. Bilgili, M. The use of artificial neural networks for forecasting the monthly mean soil temperatures in Adana, Turkey. Turk. J. Agric. For. 2011, 35, 83–93. [Google Scholar] [CrossRef]
  34. Tabari, H.; Sabziparvar, A.; Ahmadi, M. Comparison of artificial neural network and multivariate linear regression methods for estimation of daily soil temperature in an arid region. Meteorol. Atmos. Phys. 2011, 110, 135–142. [Google Scholar] [CrossRef]
  35. Mehdizadeh, S.; Behmanesh, J.; Khalili, K. Evaluating the performance of artificial intelligence methods for estimation of monthly mean soil temperature without using meteorological data. Environ. Earth Sci. 2017, 76, 325. [Google Scholar] [CrossRef]
  36. Bilgili, M.; Sahin, B.; Sangun, L. Estimating soil temperature using neighboring station data via multi-nonlinear regression and artificial neural network models. Environ. Monit. Assess. 2012, 185, 347–358. [Google Scholar] [CrossRef]
  37. Feng, Y.; Cui, N.; Hao, W.; Gao, L.; Gong, D. Estimation of soil temperature from meteorological data using different machine learning models. Geoderma 2019, 338, 67–77. [Google Scholar] [CrossRef]
  38. Sihag, P.; Esmaeilbeiki, F.; Singh, B.; Pandhiani, S.M. Model-based soil temperature estimation using climatic parameters: The case of Azerbaijan Province, Iran. Geol. Ecol. Landsc. 2020, 4, 203–215. [Google Scholar] [CrossRef] [Green Version]
  39. Kisi, O.; Tombul, M.; Kermani, M.Z. Modeling soil temperatures at different depths by using three different neural computing techniques. Theor. Appl. Climatol. 2015, 121, 377–387. [Google Scholar] [CrossRef]
  40. Mehdizadeh, S.; Behmanesh, J.; Khalili, K. Comprehensive modeling of monthly mean soil temperature using multivariate adaptive regression splines and support vector machine. Theor. Appl. Climatol. 2018, 133, 911–924. [Google Scholar] [CrossRef]
  41. Qi, J.; Li, S.; Li, Q.; Xing, Z.; Bourque, C.P.; Meng, F. A new soil-temperature module for SWAT application in regions with seasonal snow cover. J. Hydrol. 2016, 538, 863–877. [Google Scholar] [CrossRef]
  42. Li, Z.; Yang, J.Y.; Drury, C.F.; Yang, X.M.; Reynolds, W.D.; Li, X.; Hu, C. Evaluation of the DNDC model for simulating soil temperature, moisture and respiration from monoculture and rotational corn, soybean and winter wheat in Canada. Ecol. Model. 2017, 360, 230–243. [Google Scholar] [CrossRef]
  43. Lu, X.; Fan, J.; Wu, L.; Dong, J. Forecasting multi-step ahead monthly reference evapotranspiration using hybrid extreme gradient boosting with grey wolf optimization algorithm. Comput. Model. Eng. Sci. 2020, 125, 699–723. [Google Scholar] [CrossRef]
  44. Jin, Q.; Fan, X.; Liu, J.; Xue, Z.; Jian, H. Estimating tropical cyclone intensity in the South China Sea using the XGBoost Model and FengYun Satellite images. Atmosphere 2020, 11, 423. [Google Scholar] [CrossRef] [Green Version]
  45. Dong, J.; Zeng, W.; Lei, G.; Wu, L.; Chen, H.; Wu, J.; Huang, J.; Gaiser, T.; Srivastava, A.K. Simulation of dew point temperature in different time scales based on grasshopper algorithm optimized extreme gradient boosting. J. Hydrol. 2022, 127452. [Google Scholar] [CrossRef]
  46. Goydaragh, M.G.; Taghizadeh-Mehrjardi, R.; Jafarzadeh, A.A.; Triantafilis, J.; Lado, M. Using environmental variables and Fourier Transform Infrared Spectroscopy to predict soil organic carbon. Catena 2021, 202, 105280. [Google Scholar] [CrossRef]
  47. Fan, J.; Zheng, J.; Wu, L.; Zhang, F. Estimation of daily maize transpiration using support vector machines, extreme gradient boosting, artificial and deep neural networks models. Agr. Water Manag. 2021, 245, 106547. [Google Scholar] [CrossRef]
  48. Zhao, S. A new scheme for comprehensive physicl regionalization in China. Acta Geographica Sinica. 1983, 38, 1–10. (In Chinese) [Google Scholar]
  49. Yao, N.; Li, Y.; Lei, T.; Peng, L. Drought evolution, severity and trends in mainland China over 1961–2013. Sci. Total Environ. 2018, 616, 73–89. [Google Scholar] [CrossRef]
  50. Dong, J.; Liu, X.; Huang, G.; Fan, J.; Wu, L.; Wu, J. Comparison of four bio-inspired algorithms to optimize KNEA for predicting monthly reference evapotranspiration in different climate zones of China. Comput. Electron. Agr. 2021, 186, 106211. [Google Scholar] [CrossRef]
  51. Quinlan, J.R. Learning with continuous classes. In 5th Australian Joint Conference on Artificial Intelligence; World Scientific: Singapore, 1992; pp. 343–348. [Google Scholar] [CrossRef]
  52. Wang, Y.; Witten, I.H. Induction of Model Trees for Predicting Continuous Classes; University of Waikato: Hamilton, New Zealand, 1996; Available online: https://hdl.handle.net/10289/1183 (accessed on 20 March 2022).
  53. Yi, H.; Lee, B.; Park, S.; Kwak, K.; An, K. Prediction of short-term algal bloom using the M5P model-tree and extreme learning machine. Environ. Eng. Res. 2019, 24, 404–411. [Google Scholar] [CrossRef]
  54. Breiman, L. Random forests. Mach. Learn. 2001, 45, 5–32. [Google Scholar] [CrossRef] [Green Version]
  55. Chen, T.; Guestrin, C. Xgboost: A scalable tree boosting system. In Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, San Francisco, CA, USA, 13–17 August 2016; pp. 785–794. [Google Scholar] [CrossRef] [Green Version]
  56. Dong, H.; Xu, X.; Wang, L.; Pu, F. Gaofen-3 PolSAR image classification via XGBoost and polarimetric spatial information. Sensors 2018, 18, 611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Tziachris, P.; Aschonitis, V.; Chatzistathis, T.; Papadopoulou, M. Assessment of spatial hybrid methods for predicting soil organic matter using DEM derivatives and soil parameters. Catena 2019, 174, 206–216. [Google Scholar] [CrossRef]
  58. Malone, B.P.; Styc, Q.; Minasny, B.; McBratney, A.B. Digital soil mapping of soil carbon at the farm scale: A spatial downscaling approach in consideration of measured and uncertain data. Geoderma 2017, 290, 91–99. [Google Scholar] [CrossRef]
  59. Talaee, P.H. Daily soil temperature modeling using neuro-fuzzy approach. Theor. Appl. Climatol. 2014, 118, 481–489. [Google Scholar] [CrossRef]
  60. Feng, Y.; Jia, Y.; Zhang, Q.; Gong, D.; Cui, N. National-scale assessment of pan evaporation models across different climatic zones of China. J. Hydrol. 2018, 564, 314–328. [Google Scholar] [CrossRef]
  61. Wang, L.; Kisi, O.; Zounemat-Kermani, M.; Li, H. Pan evaporation modeling using six different heuristic computing methods in different climates of China. J. Hydrol. 2017, 544, 407–427. [Google Scholar] [CrossRef]
  62. Hara, M. New estimation trial for the soil temperature of Morioka from the meteorological data of Morioka. In New Trends on System Sciences and Engineering; IOS Press: Amsterdam, The Netherlands, 2015; pp. 595–608. [Google Scholar]
  63. Fitton, E.M.; Brooks, C.F. Soil temperatures in the United States. Mon. Weather Rev. 1931, 59, 6–16. [Google Scholar] [CrossRef]
  64. Bai, Y.; Scott, T.A.; Min, Q. Climate change implications of soil temperature in the Mojave Desert, USA. Front. Earth Sci.-PRC. 2014, 8, 302–308. [Google Scholar] [CrossRef]
  65. Davies, A.; Thomas, H. Rates of leaf and tiller production in young spaced perennial ryegrass plants in relation to soil temperature and solar radiation. Ann. Bot. 1983, 51, 591–597. [Google Scholar] [CrossRef]
  66. Nahvi, B.; Habibi, J.; Mohammadi, K.; Shamshirband, S.; Al Razgan, O.S. Using self-adaptive evolutionary algorithm to improve the performance of an extreme learning machine for estimating soil temperature. Comput. Electron. Agr. 2016, 124, 150–160. [Google Scholar] [CrossRef]
  67. Huang, R.; Huang, J.; Zhang, C.; Wen, Z.; Chen, Y.; Zhu, D.; Qingling, W.U.; Mansaray, L.R. Soil temperature estimation at different depths, using remotely-sensed data. J. Integr. Agr. 2020, 19, 277–290. [Google Scholar] [CrossRef]
  68. Kassaye, K.T.; Boulange, J.; Saito, H.; Watanabe, H. Soil water content and soil temperature modeling in a vadose zone of Andosol under temperate monsoon climate. Geoderma 2021, 384, 114797. [Google Scholar] [CrossRef]
  69. Zhang, H.; Shi, X.; Yu, D.; Wang, H.; Zhao, Y.; Sun, W.; Huang, B. Spatial prediction of soil temperature in China. Acta Pedol. Sin. 2009, 46, 3718–3723. (In Chinese) [Google Scholar] [CrossRef]
  70. Knight, J.H.; Minasny, B.; McBratney, A.B.; Koen, T.B.; Murphy, B.W. Soil temperature increase in eastern Australia for the past 50 years. Geoderma 2018, 313, 241–249. [Google Scholar] [CrossRef]
  71. Hu, G.; Zhao, L.; Li, R.; Wu, X.; Wu, T.; Xie, C.; Zhu, X.; Su, Y. Variations in soil temperature from 1980 to 2015 in permafrost regions on the Qinghai-Tibetan Plateau based on observed and reanalysis products. Geoderma 2019, 337, 893–905. [Google Scholar] [CrossRef]
  72. Kunkel, V.; Wells, T.; Hancock, G.R. Soil temperature dynamics at the catchment scale. Geoderma 2016, 273, 32–44. [Google Scholar] [CrossRef]
  73. Zhang, Y.; Hou, W.; Chi, M.; Sun, Y.; An, J.; Yu, N.; Zou, H. Simulating the effects of soil temperature and soil moisture on CO2 and CH4 emissions in rice straw-enriched paddy soil. Catena 2020, 194, 104677. [Google Scholar] [CrossRef]
  74. Zhang, L.L.; Zhao, L.; Li, R.; Gao, L.M.; Xiao, Y.; Qiao, Y.P.; Shi, J.Z. Investigating the influence of soil moisture on albedo and soil ther modynamic parameters during the warm season in Tanggula Range, Tibetan Plateau. J. Glaciol. Geocryol. 2016, 38, 351–358. (In Chinese) [Google Scholar]
  75. Nikolaev, I.V.; Leong, W.H.; Rosen, M.A. Experimental investigation of soil thermal conductivity over a wide temperature range. Int. J. Thermophys. 2013, 34, 1110–1129. [Google Scholar] [CrossRef]
  76. Hurrass, J.; Schaumann, G.E. Influence of the sample history and the moisture status on the thermal behavior of soil organic matter. Geochim. Cosmochim. Ac. 2007, 71, 691–702. [Google Scholar] [CrossRef]
  77. Wang, Y.; Lu, Y.; Horton, R.; Ren, T. Specific heat capacity of soil solids: Influences of clay content, organic matter, and tightly bound water. Soil Sci. Soc. Am. J. 2019, 83, 1062–1066. [Google Scholar] [CrossRef]
  78. Curtin, D.; Beare, M.H.; Hernandez-Ramirez, G. Temperature and moisture effects on microbial biomass and soil organic matter mineralization. Soil Sci. Soc. Am. J. 2012, 76, 2055–2067. [Google Scholar] [CrossRef]
  79. Grunwald, D.; Kaiser, M.; Junker, S.; Marhan, S.; Piepho, H.; Poll, C.; Bamminger, C.; Ludwig, B. Influence of elevated soil temperature and biochar application on organic matter associated with aggregate-size and density fractions in an arable soil. Agric. Ecosyst. Environ. 2017, 241, 79–87. [Google Scholar] [CrossRef]
  80. Yuan, Q. Prediction for the effect of temperature and water content on the soil specific heat by BP neural network. Trans. Chin. Soc. Agric. Mach. 2008, 5, 108–111. (In Chinese) [Google Scholar] [CrossRef] [Green Version]
  81. Epron, D.; Farque, L.; Lucot, É.; Badot, P. Soil CO2 efflux in a beech forest: Dependence on soil temperature and soil water content. Ann. Forest Sci. 1999, 56, 221–226. [Google Scholar] [CrossRef] [Green Version]
  82. Gaumont-Guay, D.; Black, T.A.; Griffis, T.J.; Barr, A.G.; Jassal, R.S.; Nesic, Z. Interpreting the dependence of soil respiration on soil temperature and water content in a boreal aspen stand. Agr. For. Meteorol. 2006, 140, 220–235. [Google Scholar] [CrossRef]
  83. Sawan, Z.M. Climatic variables: Evaporation, sunshine, relative humidity, soil and air temperature and its adverse effects on cotton production. Inf. Process Agric. 2018, 5, 134–148. [Google Scholar] [CrossRef]
  84. Olchev, A.; Radler, K.; Sogachev, A.; Panferov, O.; Gravenhorst, G. Application of a three-dimensional model for assessing effects of small clear-cuttings on radiation and soil temperature. Ecol. Model. 2009, 220, 3046–3056. [Google Scholar] [CrossRef]
  85. Kong, K.; Nandintsetseg, B.; Shinoda, M.; Ishizuka, M.; Kurosaki, Y.; Bat-Oyun, T.; Gantsetseg, B. Seasonal variations in threshold wind speed for saltation depending on soil temperature and vegetation: A case study in the Gobi Desert. Aeolian Res. 2021, 52, 100716. [Google Scholar] [CrossRef]
  86. Zhang, T. Influence of the seasonal snow cover on the ground thermal regime: An overview. Rev. Geophys. 2005, 43, 1–23. [Google Scholar] [CrossRef]
  87. Huang, G.; Wu, L.; Ma, X.; Zhang, W.; Fan, J.; Yu, X.; Zeng, W.; Zhou, H. Evaluation of CatBoost method for prediction of reference evapotranspiration in humid regions. J. Hydrol. 2019, 574, 1029–1041. [Google Scholar] [CrossRef]
  88. Nanda, A.; Sen, S.; Sharma, A.N.; Sudheer, K.P. Soil temperature dynamics at hillslope scale—Field observation and machine learning-based approach. Water 2020, 12, 713. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The geographical distribution of meteorological stations used in this study.
Figure 1. The geographical distribution of meteorological stations used in this study.
Applsci 12 05088 g001
Figure 2. The flow chart of the RF algorithm.
Figure 2. The flow chart of the RF algorithm.
Applsci 12 05088 g002
Figure 3. Box plots of the RMSEs of the three models for estimating Ts under six input combinations (af).
Figure 3. Box plots of the RMSEs of the three models for estimating Ts under six input combinations (af).
Applsci 12 05088 g003
Figure 4. Performance of data-driven models (M5P1: (a,b); RF1: (c,d); XGBoost1: (e,f)) for the estimation of Ts with Tmean, RH, Rs, and U2 as inputs.
Figure 4. Performance of data-driven models (M5P1: (a,b); RF1: (c,d); XGBoost1: (e,f)) for the estimation of Ts with Tmean, RH, Rs, and U2 as inputs.
Applsci 12 05088 g004
Figure 5. Scatter plot of estimated Ts versus measured Ts from the three machine learning models with Tmean, RH, Rs, and U2 as inputs.
Figure 5. Scatter plot of estimated Ts versus measured Ts from the three machine learning models with Tmean, RH, Rs, and U2 as inputs.
Applsci 12 05088 g005
Figure 6. Fitting relationships between several factors ((a) soil organic matter; (b) soil water content; (c) RH; (d) Rs; (e) U2) and model performance in estimating Ts (* indicates significant correlation. ** indicates highly significant correlation.).
Figure 6. Fitting relationships between several factors ((a) soil organic matter; (b) soil water content; (c) RH; (d) Rs; (e) U2) and model performance in estimating Ts (* indicates significant correlation. ** indicates highly significant correlation.).
Applsci 12 05088 g006
Figure 7. Spatial distribution of soil organic matter content (a) and soil water content (b).
Figure 7. Spatial distribution of soil organic matter content (a) and soil water content (b).
Applsci 12 05088 g007
Figure 8. Principal component analysis (a) and correlation analysis (b) between several factors and model estimation Ts errors.
Figure 8. Principal component analysis (a) and correlation analysis (b) between several factors and model estimation Ts errors.
Applsci 12 05088 g008
Figure 9. Combined indicator values for each model at all stations in the NEC region. (Model 1–18 in the Y axis represents M5P1, RF1, XGBoost1, M5P2, RF2, XGBoost2, M5P3, RF3, XGBoost3, M5P4, RF4, XGBoost4, M5P5, RF5, XGBoost5, M5P6, RF6, and XGBoost6, respectively, which also applies for Figures S8–S13).
Figure 9. Combined indicator values for each model at all stations in the NEC region. (Model 1–18 in the Y axis represents M5P1, RF1, XGBoost1, M5P2, RF2, XGBoost2, M5P3, RF3, XGBoost3, M5P4, RF4, XGBoost4, M5P5, RF5, XGBoost5, M5P6, RF6, and XGBoost6, respectively, which also applies for Figures S8–S13).
Applsci 12 05088 g009
Figure 10. Frequency distribution of absolute errors (AE) between measured and estimated daily Ts obtained by three machine learning models under various input combinations (af) during testing.
Figure 10. Frequency distribution of absolute errors (AE) between measured and estimated daily Ts obtained by three machine learning models under various input combinations (af) during testing.
Applsci 12 05088 g010
Table 1. The input combinations of meteorological variables for various machine learning models.
Table 1. The input combinations of meteorological variables for various machine learning models.
ModelsInput Combinations
M5PRFXGBoost
M5P1RF1XGBoost1Tmean, RH, Rs, U2
M5P2RF2XGBoost2Tmean
M5P3RF3XGBoost3Tmean, RH
M5P4RF4XGBoost4Tmean, Rs
M5P5RF5XGBoost5Tmax, Tmin
M5P6RF6XGBoost6Tmean, U2
Table 2. Summary of the performance of data-driven models for soil temperature in different climate zones of China.
Table 2. Summary of the performance of data-driven models for soil temperature in different climate zones of China.
Different ZoneM5PRFXGBoost
RMSER2MAERMSER2MAERMSER2MAE
NWC2.358 0.971 1.800 2.213 0.974 1.709 2.175 0.976 1.694
IM2.679 0.967 2.071 2.486 0.971 1.944 2.417 0.974 1.907
NEC2.912 0.957 2.211 2.700 0.963 2.075 2.584 0.967 1.988
NC2.197 0.966 1.688 2.042 0.970 1.588 1.984 0.973 1.539
CC1.748 0.961 1.302 1.643 0.965 1.239 1.608 0.970 1.202
SC1.735 0.925 1.330 1.645 0.931 1.274 1.604 0.941 1.226
QTP2.330 0.930 1.794 2.181 0.938 1.697 2.090 0.945 1.625
Mean2.280 0.954 1.742 2.130 0.959 1.647 2.066 0.964 1.597
Note: RMSE and MAE are in °C.
Table 3. Performance of data-driven models for the estimation of soil temperature in the different climatic zones of China.
Table 3. Performance of data-driven models for the estimation of soil temperature in the different climatic zones of China.
ModelNWCIMNECNCCCSCQTP
RMSER2MAERMSER2MAERMSER2MAERMSER2MAERMSER2MAERMSER2MAERMSER2MAE
M5P12.1550.9751.6012.5150.9711.8782.8750.9582.1062.0770.9691.5521.7220.9621.2451.7280.9281.2832.2390.9341.688
RF11.7380.9841.3191.9960.9811.5332.2570.9741.7021.6420.9811.2571.3740.9751.0281.3540.9561.0351.7600.9591.352
XGBoost11.6560.9851.2581.9040.9851.4532.2080.9761.6511.5800.9831.2001.3420.9790.9921.3670.9591.0361.6710.9631.274
M5P22.2120.9741.7232.5060.9721.9942.7370.9622.1462.1030.9691.6731.7090.9631.3211.7130.9251.3652.2190.9361.748
RF22.3010.9721.7952.6070.9692.0702.8510.9592.2282.1620.9671.7191.7450.9621.3491.7360.9231.3842.2710.9331.788
XGBoost22.4740.9701.9372.7100.9682.1542.7190.9642.1142.1660.9691.6841.6730.9681.2671.6260.9401.2502.3040.9351.806
M5P32.3640.9711.8102.6560.9682.0542.8770.9582.1912.1830.9661.6681.6980.9631.2541.6950.9271.2902.2610.9351.735
RF32.2730.9731.7562.5150.9711.9622.7290.9622.0992.0730.9691.6011.6550.9651.2331.6940.9271.2962.1870.9381.697
XGBoost32.2120.9751.7352.4550.9741.9522.6510.9662.0572.0390.9721.5971.6390.9691.2231.6640.9371.2742.1110.9451.647
M5P42.1530.9751.6232.4380.9731.8472.8300.9602.0932.0060.9721.5201.6580.9651.2141.6780.9291.2622.2340.9351.707
RF41.9510.9801.4892.2030.9781.6842.5560.9671.9151.8110.9771.3891.5060.9711.1191.5320.9411.1682.0350.9461.571
XGBoost41.8700.9831.4402.1180.9801.6402.4350.9711.8401.7590.9781.3561.4480.9741.0841.4460.9491.1051.8960.9541.465
M5P52.8490.9582.1853.1760.9552.4743.1540.9502.4142.4900.9571.9061.8580.9571.3751.7740.9211.3512.6320.9132.025
RF52.6580.9632.0572.9260.9612.3042.9210.9572.2582.3030.9631.7801.7460.9621.3021.7160.9251.3122.4580.9231.913
XGBoost52.5200.9681.9732.7370.9672.1762.7300.9632.1242.1660.9681.6811.6970.9671.2771.6450.9381.2592.3020.9341.797
M5P62.4160.9691.8552.7810.9652.1822.9990.9552.3162.3220.9621.8121.8440.9571.4011.8210.9161.4302.3970.9271.860
RF62.3600.9711.8342.6710.9672.1142.8840.9582.2492.2610.9641.7801.8340.9571.4051.8380.9141.4482.3770.9271.864
XGBoost62.3170.9741.8222.5800.9722.0672.7580.9632.1452.1960.9691.7151.8490.9631.3691.8750.9271.4322.2550.9381.758
Note: RMSE and MAE are in °C. The bolded numbers are the best values for model estimation performance.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dong, J.; Huang, G.; Wu, L.; Liu, F.; Li, S.; Cui, Y.; Wang, Y.; Leng, M.; Wu, J.; Wu, S. Modelling Soil Temperature by Tree-Based Machine Learning Methods in Different Climatic Regions of China. Appl. Sci. 2022, 12, 5088. https://doi.org/10.3390/app12105088

AMA Style

Dong J, Huang G, Wu L, Liu F, Li S, Cui Y, Wang Y, Leng M, Wu J, Wu S. Modelling Soil Temperature by Tree-Based Machine Learning Methods in Different Climatic Regions of China. Applied Sciences. 2022; 12(10):5088. https://doi.org/10.3390/app12105088

Chicago/Turabian Style

Dong, Jianhua, Guomin Huang, Lifeng Wu, Fa Liu, Sien Li, Yaokui Cui, Yicheng Wang, Menghui Leng, Jie Wu, and Shaofei Wu. 2022. "Modelling Soil Temperature by Tree-Based Machine Learning Methods in Different Climatic Regions of China" Applied Sciences 12, no. 10: 5088. https://doi.org/10.3390/app12105088

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