Next Article in Journal
Ginger Phytotoxicity: Potential Efficacy of Extracts, Metabolites and Derivatives for Weed Control
Previous Article in Journal
The Properties of Microorganisms and Plants in Soils after Amelioration
Previous Article in Special Issue
Nondestructively Determining Soluble Solids Content of Blueberries Using Reflection Hyperspectral Imaging Technique
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Model Optimization and Application of Straw Mulch Quantity Using Remote Sensing

1
College of Information and Technology & Smart Agriculture Research Institute, Jilin Agricultural University, Changchun 130118, China
2
College of Engineering and Technology, Jilin Agricultural University, Changchun 130118, China
3
Changchun Agricultural Machinery Research Institute, Changchun Agricultural Commission, Changchun 130052, China
*
Author to whom correspondence should be addressed.
Agronomy 2024, 14(10), 2352; https://doi.org/10.3390/agronomy14102352
Submission received: 28 August 2024 / Revised: 4 October 2024 / Accepted: 5 October 2024 / Published: 12 October 2024

Abstract

:
Straw mulch quantity is an important indicator in the detection of straw returned to the field in conservation tillage, but there is a lack of large-scale automated measurement methods. In this study, we estimated global straw mulch quantity and completed the detection of straw returned to the field. We used an unmanned aerial vehicle (UAV) carrying a multispectral camera to acquire remote sensing images of straw in the field. First, the spectral index was selected using the Elastic-net (ENET) algorithm. Then, we used the Genetic Algorithm Hybrid Particle Swarm Optimization (GA-HPSO) algorithm, which embeds crossover and mutation operators from the Genetic Algorithm (GA) into the improved Particle Swarm Optimization (PSO) algorithm to solve the problem of machine learning model prediction performance being greatly affected by parameters. Finally, we used the Monte Carlo method to achieve a global estimation of straw mulch quantity and complete the rapid detection of field plots. The results indicate that the inversion model optimized using the GA-HPSO algorithm performed the best, with the coefficient of determination (R2) reaching 0.75 and the root mean square error (RMSE) only being 0.044. At the same time, the Monte Carlo estimation method achieved an average accuracy of 88.69% for the estimation of global straw mulch quantity, which was effective and applicable in the detection of global mulch quantity. This study provides a scientific reference for the detection of straw mulch quantity in conservation tillage and also provides a reliable model inversion estimation method for the estimation of straw mulch quantity in other crops.

1. Introduction

Known for its high organic matter and fertility, black soil is a non-renewable resource that plays a very important role in agricultural development [1]. In order to curb the degradation of black soil and restore and enhance ground strength, conservation tillage has become a world-recognized sustainable production method in agriculture; it is an important part of modern eco-agriculture and plays an important role in environmental protection [2]. Conservation tillage techniques centered on no- and minimum-tillage plus crop straw mulching have become important measures to improve soil degradation on farmland and increase soil organic matter content and moisture retention capacity. Straw mulching has greatly improved the degree of protection of black soil, facilitated the function of carbon metabolism in no-till soil microbial communities, and increased crop yields. Conservation tillage requires that the straw mulch rate be not less than 0.3 [3]. Straw mulch quantity is also an important indicator for determining the level of straw mulching. Straw mulch rate refers to the ratio of the area covered by crop straw or stubble on the surface soil to the total surface area; straw mulch quantity refers to the quantity of crop straw or stubble covered per unit area of surface soil. Our team is committed to the study of straw mulch detection, and we have obtained good results in the detection of straw mulch [4,5,6]. However, it was found that the straw mulch rate can only accomplish the detection of “with” and “without” straw mulch and not the detection of “more” and “less” straw mulch. Liu et al. [7] suggested that different straw mulch quantities have different effects on soil improvement and that a straw return quantity of about 7500 kg/hm2 had the best effect on soil improvement. Therefore, the accurate detection of straw mulch quantity in the field is an important judgment index to measure the level of straw mulching and also reflects the quality and effect of straw return to the field [8]. However, in the detection of straw mulch quantity, there are problems with cumbersome work tasks, diversified influencing factors, and huge measurement errors, and there is no automated measurement method of straw mulch quantity on a large scale yet.
In recent years, remote-sensing technology has been developed with the advantage of being available on temporal and spatial scales, allowing for rapid and accurate large-scale feature detection [9]. Aerial multispectral technology has potential in remote sensing research, which combines the integrated features of UAV monitoring and spectral remote sensing and describes external characterization information more accurately while acquiring rich information [10]. The input–output ratio of UAV multi-spectral cameras is higher, so they have a short data acquisition cycle and are more flexible in operation and suitable for multiple scenarios. Huang et al. [11], Daughtry et al. [12], Daughtry et al. [13], and Burud et al. [14] proposed detecting crops in farmland through remote sensing data and achieved good results. Kavoosi et al. [15] compared the feasibility of satellite and drone images in monitoring straw mulch, and the results showed that drones have certain advantages in straw mulch detection. The mapping relationship established between spectral data and straw dry matter content is one of the simplest methods used to obtain the straw mulch quantity. Yang et al. [16] used a machine learning algorithm combined with a vegetation index approach to estimate the leaf dry matter content. Liu et al. [17] used PCA transform to reconstruct leaf reflectance spectra and retrieve the leaf biochemical composition, which included the leaf mass per area. Machine learning techniques have the advantage of solving nonlinear tasks and have been widely applied to the remote sensing of biochemical composition in recent years. Meanwhile, in the application of remote-sensing technology, machine learning algorithms are considered to be reliable and efficient methods for inverting models; they can greatly improve processing speeds and analysis accuracy [18]. Fu et al. [19], Han et al. [20], Ban et al. [21], and Nasi et al. [22] used machine learning to construct an inversion model for crop performance evaluation under different conditions related to, for example, crop yield estimation and soil salinity. Although machine learning has reached a certain degree of accuracy, the prediction performance is greatly affected by the model parameters [23]. The Meta-Heuristic Algorithm can help models be less likely to fall into a local optimum during the search process and find the global optimal solution with great probability, which is a commonly used optimization method [24]. Zhu et al. [25] and Mo et al. [26] proposed that the hybrid algorithm is superior and can address the shortcomings of a single optimization algorithm well, allowing the model to be further optimized in terms of convergence speed and computational accuracy.
In practical application scenarios, we cannot accurately estimate the global straw mulch quantity using only the inversion model because we can only obtain local samples at each site and are limited by the sampling environment and data quantity. The Monte Carlo method [27], as a flexible and efficient numerical solution, has become a key tool in the field of scientific computing and engineering and plays a crucial role in solving practical problems. The idea of the Monte Carlo method addresses the problem of estimating overall sample distribution by random sampling, which is precisely the problem we wanted to solve. It can approximate the global straw mulch quantity by using local samples [28]. Therefore, to complete the transformation from “with” and “without” to “more” and “less” of the global field straw mulch and return effect, the establishment of an efficient and accurate regional straw mulch quantity estimation model is the most important task.
Aiming at the above problems, this study adopts a UAV carrying a multispectral camera to acquire corn straw mulch remote sensing images of the study areas and simultaneously determines the data of straw mulch quantity at the sampling points. Based on the RF inversion model, we proposed the hybrid GA-HPSO algorithm to improve the performance of the inversion model and optimize the model to obtain the sample data results. The Monte Carlo idea is used to estimate the global straw mulch quantity and complete the detection of straw return to the field, which provides important modern information technology support to effectively solve the hidden dangers and problems of conservation tillage, such as “difficult to verify the area, difficult to monitor the quality, and high risk of subsidy fund issuance”.

2. Materials and Methods

2.1. Overview of the Study Area

To facilitate a comprehensive discussion on straw mulch quantity estimation, the study area was divided into experimental and validation segments. The experimental areas were primarily designed to assess the feasibility of the optimal inversion model and estimation methods, while the validation areas aimed to demonstrate the practical applicability of these methodologies. The study site was located in Changchun City, Jilin Province, China, situated in the temperate semi-humid continental monsoon climate zone, characterized by an average annual temperature of 5.5 °C, average annual precipitation of 558 mm, and an average elevation of approximately 220 m. Jilin Province has actively promoted conservation tillage practices involving straw mulch, with widespread adoption of full straw mulch applications. Details of the study area selection are presented in Table 1.
The experimental fields across the eight study areas employed crop rotation and wide–narrow row planting patterns, predominantly utilizing the Udi 598 maize hybrid (Hongxiang Seed Industry Company Limited, Jilin, China). Spring corn is typically sown in early May and harvested in early October. The terrain within the study area is relatively flat, with rows fully covered by returned straw post-crushing, as illustrated in Figure 1. The return-row plowing method allows for substantial or complete straw return to the field, thereby reducing soil erosion from wind and water while enhancing soil fertility [29]. This setting provides an optimal environment for conducting maize straw mulch detection research.

2.2. Data Acquisition and Analysis

2.2.1. Determination of Straw Mulch Quantity

To provide a better understanding of the study areas, the weather information under sunny skies, including wind direction, wind speed, and temperature, was collected and shown in Table 2. The experimental data were collected after harvesting in the autumn but before planting in the spring of the second year concerning the timeliness and diversity of the data. Meanwhile, the locations and number of sampling points were determined based on the size of the area. The corn straw mulch quantity parameters were measured at the sampling points.
We defined each sampling unit as approximately 2500 m2, randomly selecting six 0.5 m × 0.5 m squares within each unit as sampling points. After measuring straw mulch quantity within these points, we assessed thickness at five designated corners and the center for error verification, utilizing an electronic scale (Hengzi KC-833, accuracy: 0.001 kg). A total of 408 sampling points were collected during the experimental phase and 258 during the validation phase, with the sampling process illustrated in Figure 2. For example, in Study Area 1, the area was divided into nine sampling units, each containing six randomly distributed sampling points. The UAV captured multispectral images above these points to measure straw mulch quantity.

2.2.2. UAV Remote Sensing Data

The DJI Matric 300 RTK drone (DJI Technology Company Limited, Shenzhen, China) was equipped with an AQ600 PRO multispectral camera (Changguang Yuchen Information Technology and Equipment Company Limited, Qingdao, China). The AQ600 PRO sensor collects data across discrete spectral bands: blue (450 ± 35 nm), green (555 ± 25 nm), red (660 ± 20 nm), red edge (720 ± 10 nm), and near-infrared (840 ± 35 nm), with a pixel resolution of 2048 × 1536. Real-time radiometric correction was supported, with radiometric calibration images captured prior to each flight. Aerial image acquisition was conducted during the optimal timeframe of 10:00 to 15:00.
To ensure comprehensive data collection from the study area plots, we utilized a high-altitude UAV flight pattern. The UAV followed a pre-planned route at a speed of 3 m/s, maintaining an 80% overlap for both bypass and heading. After conducting several trial flights, an optimal flight altitude of 40 m was determined. Additionally, to guarantee the accuracy of spectral images, we employed a fixed-point manual control mode, hovering at a determined altitude of 12 m for image capture above each sampling point.
The multispectral images acquired via UAV were processed and analyzed using Yusense Map V 2.2.3 software, which is specifically designed for the multispectral camera. Initially, we employed the band alignment function within Yusense Map V 2.2.3 software. Subsequently, geometric and radiometric corrections were applied to eliminate potential noise and interference. The corrected UAV multispectral images were then imported into ENVI 5.3 software, where the corresponding spectral image portions were cropped according to the predefined sampling points. Finally, we calculated the average reflectance spectra for the straw mulch samples within the region of interest, yielding the spectral reflectance data.

2.3. Spectral Reflectance Measurement

The objective of the spectral reflectance measurements was to simulate and analyze variations in spectral reflectance associated with different straw mulch quantities within the same area, thereby assessing the impact of these variations on remote sensing estimates of straw mulch quantity and validating the feasibility of this study.
To facilitate this, we collected straw samples from the experimental area across four datasets in 2022 and 2023, placing approximately 0.7 kg of straw into black bags. The experimental environment was maintained under natural light conditions. A square area measuring 0.5 m × 0.5 m was designated for sampling, with a layer of black soil (0.01 m thick) spread evenly. The multispectral camera (AQ600 PRO) was fixed at a height of 1 m above the black soil. We collected multispectral data corresponding to varying straw mulch quantities, starting with 0.1 kg of straw evenly distributed across the sampling area and increasing in 0.1 kg increments up to 0.5 kg. This acquisition process is illustrated in Figure 3, and we captured radiometric calibration target images prior to each set of measurements.

2.4. Spectral Index Construction

Spectral indices serve as metrics for analyzing and characterizing spectral data, enabling the study of the composition, structure, and properties of substances. By enhancing the separability of targets, spectral indices improve the accuracy of feature detection. Given their sensitivity to feature characteristics, we selected 12 widely used spectral indices for our experiments, with their calculation formulas detailed in Table 3.

2.5. Optimization of the Inverse Model for Straw Mulch Quantity

Previously, our team compared various inversion models in the context of straw mulch quantity estimation, concluding that the Random Forest (RF) algorithm yielded the most effective inversion model [42]. Therefore, we employed the RF algorithm in this study [43]. This algorithm integrates ensemble learning with decision trees, utilizing resampling methods to enhance the prediction accuracy and convergence speed. However, empirical evidence suggests that hyperparameters—specifically, the number of decision trees (n_estimators), maximum tree depth (max_depth), and maximum number of leaf nodes (max_leaf_nodes)—significantly influence RF performance. Consequently, it is essential to balance accuracy and computational efficiency in parameter selection.
To address this, we introduced a hybrid optimization approach combining Genetic Algorithm (GA) [44] and Particle Swarm Optimization (PSO) [45], resulting in the GA-HPSO optimization algorithm. This method enables the RF model to autonomously identify optimal solutions, mitigating the subjectivity associated with manual parameter tuning and alleviating issues related to parameter sensitivity in model prediction performance.
The GA is advantageous due to its population diversity and global search capabilities; however, it lacks individual memory and exhibits slower convergence. Conversely, the PSO benefits from memory and rapid convergence but is susceptible to premature convergence and diminished diversity. By merging these two algorithms, we aim to leverage their complementary strengths. Parameter values such as inertia weights and learning factors in the PSO significantly affect convergence performance. Conventional approaches to these parameters can hinder convergence speed and overall efficiency. Additionally, initialization plays a critical role in accuracy and speed; uneven solution space distribution may lead to local optima.
Traditional PSO algorithms typically employ random initialization to determine initial population distributions. This method can result in uneven initial particle distribution within the solution space, adversely affecting early convergence and increasing the likelihood of local optima. To overcome this, we adopted chaotic initialization to replace the conventional random method. Chaotic initialization incorporates randomness, traversal, and regularity, allowing the search space to be traversed systematically without repetition, thereby enhancing both solution accuracy and convergence speed. In this study, we employed the Logistic Chaos Model for initialization [46], as represented in Equation (1):
X n + 1 = μ × X n × 1 X n ,   μ ϵ 0 , 4
where μ is the chaotic model mapping parameter. When μ ϵ 3.57,4 , the system exhibits chaotic behavior; at μ = 4 , it reaches a fully chaotic state, which we employed in this study.
Next, the inertia weight parameter in the PSO algorithm significantly influences convergence performance. A larger inertia weight favors global exploration, while a smaller weight accelerates local convergence. To balance these capabilities, we introduced a dynamic weight factor (ω) to enhance both convergence speed and accuracy. The formulation of ω is shown in Equation (2):
ω = ω m i n ω m a x ω m i n × f f m i n f a v g f m i n , f f a v g ω m a x , f > f a v g
where f represents the real-time objective function value of the particle. f a v g and f m i n denote the average and minimum objective value of all current particles, respectively. A larger weight at the search’s initial phase promotes global exploration, while a smaller weight in later phases fosters local optimization.
Furthermore, the learning factors c 1 and c 2 in the PSO algorithm impact overall performance and are typically fixed. In this study, to prevent rapid convergence around local optima during the initial stages, we allowed c 1 to be larger and c 2 to be smaller. Conversely, in the later stages, we adjusted c 1 to be smaller and c 2 to be larger to facilitate quick and accurate convergence to the global optimum. Thus, c 1 and c 2 are represented in Equations (3) and (4) as follows:
c 1 = c 1 i + t × c 1 f c 1 i / T m a x
c 2 = c 2 i + t × c 2 f c 2 i / T m a x
where t is the current iteration number. T m a x is the maximum iteration number of the particle swarm. c 1 i , c 2 i are the c 1 , c 2 initial values and c 1 f , c 2 f are the c 1 , c 2 final values. When c 1 i c 2 f and c 2 i c 1 f , it is an asymmetric learning factor method.
Lastly, we incorporated crossover and mutation operators from GA into the PSO framework. These operations enhance population diversity and broaden the algorithm’s search range, thereby aiding in the attainment of a global optimum. The computational flow of the GA-HPSO algorithm is illustrated in Figure 4.
The research environment utilized MATLAB 2018a, running on a Windows 10 operating system with an Intel (R) Core (TM) i7-8550U [email protected] GHz 2.00 GHz processor and an NVIDIA GeForce 940MX graphics card.

2.6. Global Estimation Methods

To relate sample point data to global straw mulch quantity, this study employed the Monte Carlo method. This method’s fundamental principle involves obtaining desired results through random sampling and statistical averaging. The steps typically include defining the problem and required data; calculating necessary functions and probability distributions; designing a random sampling approach (e.g., utilizing pseudo-random numbers); and deriving statistical quantities from the sampled data, such as sums, averages, and variances. Conclusions are drawn based on these results.
Utilizing the Monte Carlo method, the straw mulch quantities obtained from sampling points via the optimal inversion model are treated as known samples. From these samples, an empirical distribution is established, leading to the formulation of a Gaussian mixture model. The probability density function of this model, constituted by the superposition of k Gaussian distributions, is expressed in Equation (5):
p x | θ = k = 1 k π k p x , μ k , k
where θ is the full Gaussian model parameter, k = 1 k π k is 1, π k is the weight, x is the sample, μ is the model expectation, and is the model variance.
The parameters for each class in the hybrid model are derived using an Expectation-Maximization (EM) algorithm with an iterative optimization strategy, comprising an Expectation step (E step) and a maximization step (M step). The E step is represented in Equation (6):
Q θ , θ i = E Z l o g X , Z | θ X , θ i = l o g p X , Z | θ P Z | X , θ i
where Z denotes the hidden variable associated with the unknown straw mulch quantity data. θ = μ k , k , π k is the parameter estimate of the initial Gaussian model. μ k i , k i , π k i is the parameter estimate of the Gaussian model with i iterations. The joint distribution is P Z | X , θ . Given the observed data X and parameter estimates θ i , the conditional probability distribution of the hidden variable Z is sought P Z | X , θ i .
The M step is shown in Equation (7):
θ i + 1 = a r g m a x Q θ , θ i
This involves maximizing Q θ , θ i to determine the parameter estimates for the subsequent iteration.
Employing the Monte Carlo principle, the global estimation method entails n random draws from a known distribution function, yielding n known sample data. Given the unknown probability distribution formula and normalization constant of the Gaussian mixture model, direct sampling from the distribution function proves infeasible. Thus, we employed the choice algorithm for random sampling, employing the accept–reject principle. In this process, a random point is accepted if it lies within the designated region; otherwise, it is rejected.
The ultimate goal of this research is to assess straw mulch quantity in the field and evaluate the effectiveness of straw return in conservation tillage. To achieve this, we need to establish the theoretical straw resources to facilitate subsequent calculations of straw volume return rates. The grass-to-grain ratio method [47], as outlined in NY/T 1701–2009 [48], was utilized to obtain the theoretical straw resources. The grass-to-grain ratio μ is defined in Equation (8):
μ = m s 1 A s / 1 15 % m G 1 A G / 1 12.5 %
where m s is the mass of straw, m G is the mass of grain, A s is the moisture content of straw, and A G is the moisture content of grain. The standard moisture content of air-dried straw is 15%, while the national standard moisture content for grain crops is 12.5%. The theoretical corn straw resource P is articulated in Equation (9):
P = μ i G i
where G i is the maize yield. The straw collectible resource P c is shown in Equation (10):
P c = δ μ i G i
where δ is the straw collection coefficient, as shown in Equation (11):
δ = 1 L i , j L i J i + 1 L i , r L i 1 J i 1 Z i / 100
where L i is the average plant height of maize, L i , j is the stubble height of mechanically harvested crops, L i , r is the stubble height of manually harvested crops, J i is the mechanised harvesting rate of the crop, and Z i is the rate of loss of the crop stalks during harvesting and transportation. The theoretical straw resources and collectible straw resources were calculated based on relevant data obtained from the study area, with results presented in Table 4.

3. Results and Analysis

3.1. Spectral Reflectance Analysis

In this study, spectral reflectance data corresponding to varying straw mulch quantities were processed, with the results presented in Figure 5. A positive correlation was observed between straw mulch quantity and spectral reflectance across four data groups. Notably, the spectral reflectance differences between bands diminished as the mulch quantity increased. Additionally, it was evident that spectral reflectance in spring was lower than in fall, with a slower decrease at 450 nm compared to other wavelengths. This analysis suggests that utilizing spectral indices to differentiate spectral characteristics, followed by inverse modeling to simulate reflectance under various mulch amounts, can effectively estimate the corn straw mulch quantity.

3.2. Correlation Analysis

In remote sensing image processing, efficient spectral index screening algorithms significantly reduce data dimensionality, streamline processing, and enhance computational efficiency while minimizing storage needs. The Elastic-net (ENET) algorithm [49], a linear regression model that integrates Lasso’s L1 and Ridge’s L2 regularization terms, serves to balance model sparsity and non-sparsity [50].
To evaluate the relationship strength between spectral variables and straw mulch quantity, this study employed the ENET algorithm to screen 12 spectral index variables. The findings are summarized in Table 5, where regression coefficients indicate the valid spectral information conveyed by each index and its explanatory power in the model [51]. Coefficients nearing zero were eliminated, and the remaining spectral variables were selected as input for the model.

3.3. Evaluation of Inversion Models

To demonstrate the superiority of the proposed GA-HPSO algorithm over GA and PSO algorithms, we evaluated three standard test functions from the CEC: Sphere, Griewank, and Rastrigin. The Sphere function is articulated in Equation (12):
f X = i N X i 2
where the range of values of the independent variable X i : −100 < X i < 100. The Griewank function is shown in Equation (13):
f X = 1 4000 i N X i 2 i N cos X i i + 1
where the range of values of the independent variable X i : −600 < X i < 600. The Rastrigin function is shown in Equation (14):
f X = i N X i 2 10 cos 2 π X i + 10
where the range of values of the independent variable variable X i : −5.12 < X i < 5.12.
The results depicted in Figure 6 show that PSO exhibited the least effective convergence, while GA-HPSO demonstrated superior convergence performance, efficiently identifying optimal individual fitness with fewer iterations while maintaining stability. This indicates that the GA-HPSO algorithm significantly enhances global optimization performance and convergence speed compared to standalone GA and PSO algorithms.
To comprehensively assess the straw mulch quantity inversion model, spectral variables filtered by the ENET algorithm were utilized as input variables for models based on RF, GA-RF, PSO-RF, and GA-HPSO-RF algorithms. Model accuracy was evaluated using R2 and RMSE metrics. Higher R2 values (closer to 1) and lower RMSE values (closer to 0) indicate better model accuracy. The results displayed in Table 6 reveal that the inversion model constructed using the ENET algorithm, with spectral indices as input variables, achieved substantial accuracy and stability. The closeness of R2c and R2v values indicates that the models are not “overfitted” while maintaining high accuracy. Notably, the GA-HPSO-RF algorithm outperformed the others, achieving the highest R2c and R2v values exceeding 0.70, with both RMSEc and RMSEv below 0.061. Comparatively, GA-RF and PSO-RF models showed improved accuracy over the original model, with R2v values exceeding 0.66. Thus, the GA-HPSO-RF algorithm demonstrated the best accuracy among the four models.
For a more intuitive comparison of model inversion effects across different periods and regions, the scatter plot of the validation set is illustrated in Figure 7. The GA-HPSO-RF algorithm yielded the best inversion accuracy, with autumn data from both years displaying similar results significantly higher than those obtained in spring. This finding aligns with correlation analysis outcomes and reflects real-world conditions, confirming that the GA-HPSO-RF inversion model based on autumn data is the most effective.

3.4. Assessment of Estimation Models

To assess the empirical distribution of the samples, we analyzed the acquired data presented in Figure 8. The histogram illustrates the initial distribution, and the QQ plot suggests that the samples conform closely to a normal distribution. Therefore, we assumed the global straw distribution to be a mixture of multiple Gaussian distributions, leading to the conclusion that its probability model can be represented as a Gaussian mixture model. Under this assumption, we employed the Expectation-Maximization algorithm to determine each parameter within the Gaussian mixture model. The total data samples were fitted into a mixture distribution to establish the global straw distribution probability model, enabling the global straw mulch quantity to be derived through random sampling from the Gaussian mixture model.
During the experimental phase, eight data sets were collected to determine the straw mulch quantity. To evaluate the effectiveness and feasibility of the proposed Monte Carlo method, we utilized both accuracy assessment mean methods and Monte Carlo estimation. The core concept of the Monte Carlo approach relies on random sampling; as more sample values are obtained, the accuracy increases. We recommended a distribution range of y = 0.6, which encompasses the entire Gaussian mixture model, to facilitate random sampling. The number of iterations was determined based on the total area of the study site and the sampling area, ultimately yielding total sample values that align with the probability distribution model as the global straw mulch quantity solution. Estimated results derived from the GA-HPSO-RF inversion model are summarized in Table 7.
As shown in the table, the Monte Carlo method demonstrated superior estimation accuracy, fluctuating between 87.39% and 89.85%. In contrast, the mean method yielded lower accuracy, ranging from 74.60% to 81.58%. The accuracy of the estimated results based on autumn data was notably higher, between 88.00% and 91.61%, while estimates derived from spring data were lower, with accuracy rates between 87.39% and 87.92%. This discrepancy can be attributed to the limitations of the Gaussian mixture model in handling large data values, resulting in higher estimation accuracy during the autumn season. The comparative experiments validate the feasibility of employing Monte Carlo methods for calculating global straw mulch quantities.

3.5. Global Estimation Validation

To further assess the applicability of the proposed global straw mulch quantity estimation method, we collected validation area data, resulting in 258 samples obtained during the autumn of 2023. The validation process is illustrated in Figure 9.
We utilized six spectral indices—SRER, NDVI, SR, NDRE, GNDVI, and DVI—selected by the ENET algorithm as model input variables. The inversion results for the validation area based on the GA-HPSO-RF algorithm, shown in Figure 10, aligned closely with the measured results, achieving an average R2 of 0.74 and RMSE of 0.0483. These results were consistent with those from the experimental phase, confirming the effectiveness of the proposed optimization algorithm in estimating straw mulch quantities.
Using the inversion model results from the GA-HPSO-RF algorithm, the Monte Carlo estimation method was applied to assess global straw mulch quantity in the validation phase, as summarized in Table 8. The estimation accuracy across six validation areas reached a commendable level, ranging from 88.25% to 90.27%. These results were stable and comparable to those obtained in the experimental areas, further demonstrating the method’s robustness and applicability.
Detecting the straw return rate is essential, as some straw is buried during the plowing process, making only surface straw detectable in image acquisition. Data from the experimental area over two years indicated that approximately 0.05 kg/m2 of straw is buried in the soil. The straw return quantity and return rate, calculated from the results in Table 9, revealed return rates ranging from 82.32% to 85.34%. This method proves feasible for estimating both straw mulch and return quantities, providing valuable insights for global straw return estimations.

4. Discussion

4.1. Methodology for Estimation of Global Straw Mulch Quantity

In conservation tillage practices, the intelligent detection of straw mulch quantity remains undeveloped, relying on traditional field measurements that are time-consuming and labor-intensive [52]. By integrating multispectral data with model inversion and employing the Monte Carlo method, we reduced the need for extensive sampling, thus saving time and labor while accurately reflecting regional straw mulch quantities. This approach offers a scientific reference for estimating straw mulch in conservation tillage.
Reflectance data analysis revealed changes in the spectral reflectance of corn straw at varying mulch quantities. Although the straws all showed similar spectral profiles, their spectral profiles were within a range and regular. Light interaction with crops is complex; incident light is partly reflected while the remainder penetrates, leading to absorption and diffuse reflection within the plant structure. Light absorption is primarily influenced by the chemical composition, while diffuse reflection is affected by the physical properties of the straw. Variations in straw thickness and structure result in differing light attenuation [53]. Consequently, establishing a statistical relationship between straw mulch quantity and its optical properties facilitates accurate estimations.
Seasonal changes also impact reflectance spectra, likely related to straw decomposition, akin to findings by Daughtry et al. [13]. The observed higher accuracy of global estimations in fall compared to spring stems from the probabilistic model’s reliance on initial sample quality. While fewer samples might yield smaller errors, the accuracy of global straw estimations improves with higher-quality inversion model results. Over time, changes in the chemical and physical properties of straw during decomposition affect spectral features, weakening remote sensing assessments of straw mulch [54]. Therefore, the fall data yielded more accurate inversion model assessments. Additionally, the uneven global distribution of straw due to environmental factors, such as prolonged exposure to bare conditions and erosion, further complicates global straw mulch quantity estimations.

4.2. Advantages of Algorithms

The RF algorithm demonstrated improved performance in estimating straw mulch quantities; however, its predictive accuracy was significantly influenced by model parameters. The GA and PSO algorithms, as established optimization techniques, can address some of the RF algorithm’s limitations, yet each has its specific applicability and constraints [55,56]. In this study, we employed embedded hybridization to integrate the strengths of both algorithms, enhancing the initial population’s solution accuracy and convergence speed. This complementary approach mitigated the risk of local optima during the search process, resulting in a straw mulch quantity inversion model that effectively extracted features and enhanced detection accuracy.

4.3. Deficiency and Prospect

The methodology yielded promising results in global straw mulch quantity estimation, indicating broad applicability across most of the Corn Belt. However, it also has limitations and may not directly apply to regions with different crops, cropping patterns, or weather conditions. The field data were collected over two years (2022 and 2023), encompassing varied climatic conditions, yet did not account for all potential weather variations and soil moisture levels. Further research is essential for a comprehensive understanding of these dynamics. Additionally, integrating multi-sensor fusion techniques could enhance the accuracy of conservation tillage detection and support sustainable agricultural practices.

5. Conclusions

This study combined remote sensing imagery with ground-measured straw mulch quantity data to establish an inversion model specific to the study area by selecting relevant straw spectral indices. We proposed an optimization algorithm to enhance model performance and employed Monte Carlo methods to estimate global straw mulch quantities. The following conclusions were drawn from the experimental analysis:
(1)
Reflectance changes of corn straw in relation to varying mulch quantities were observed, revealing corresponding shifts in spectral composition. The reflectance spectra across different mulch scenarios were simulated using an inverse model, allowing for the estimation of corn straw mulch quantity based on optimal spectral indices.
(2)
The GA-HPSO-RF inversion model achieved an R2 of 0.75 and RMSEv of 0.044, demonstrating strong nonlinear fitting capability, the highest among the four models, and fulfilling local estimation requirements. The GA-HPSO algorithm addressed the RF model’s sensitivity to parameters, thereby enhancing inversion performance and facilitating more effective corn straw mulch quantity estimations.
(3)
The Monte Carlo method attained an average accuracy of 89.61% for global straw mulch quantity estimation, with relatively stable resultant errors, validating its applicability for large-scale mulch quantity detection.
In the context of global corn straw mulch quantity estimation, the precise application of the inversion model and estimation methods provides a scientific foundation for estimating the volume of corn straw returned to the field. This approach also achieves the quantitative goals of straw return detection, offering scientific and efficient automated means for promoting conservation tillage.

Author Contributions

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

Funding

This research was funded by the Jilin Science and Technology Development Program Project, product number 20220402023GH; the National Natural Science Foundation of China, product number 42001256; and the Jilin Science and Technology Development Program Project, product number 20230202039NC.

Data Availability Statement

The datasets in this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wang, Q.; Cao, X.; Wang, C.; Li, H.; He, J.; Lu, C. Research Progress of No/Minimum Tillage Corn Seeding Technology and Machine in Northeast Black Land of China. Trans. Chin. Soc. Agric. Mach. 2021, 52, 1–15. [Google Scholar]
  2. Li, Z.; Wang, K.; Pan, X.; Liu, Y.; Li, Y.; Shi, R. Estimation of wheat residue cover using simulated Landsat-8 OLI datas. Trans. Chin. Soc. Agric. Eng. 2016, 32, 145–152. [Google Scholar]
  3. Liu, Z.; Liu, Z.; Wang, W.; Huang, J.; Wang, J.; Zheng, M. Estimation of maize residue cover on the basis of SAR and optical remote sensing image. Natl. Remote Sens. Bull. 2021, 25, 1308–1323. [Google Scholar] [CrossRef]
  4. Liu, Y.; Sun, J.; Zhang, S.; Yu, H.; Wang, Y. Detection of straw coverage based on multi-threshold and multi-target UAV image segmentation optimization algorithm. Trans. Chin. Soc. Agric. Eng. 2020, 36, 134–143. [Google Scholar]
  5. Liu, Y.; Zhou, X.; Wang, Y.; Yu, H.; Geng, C.; He, M. Straw coverage detection of conservation tillage farmland based on improved U-Net model. Opt. Precis. Eng. 2022, 30, 1101–1112. [Google Scholar] [CrossRef]
  6. Wang, Y.; Gao, X.; Sun, Y.; Liu, Y.; Wang, L.; Liu, M. Sh-DeepLabv3+: An Improved Semantic Segmentation Lightweight Network for Corn Straw Cover Form Plot Classification. Agriculture 2024, 14, 628. [Google Scholar] [CrossRef]
  7. Liu, Y.; Zhu, X.; Bao, X.; Zheng, T.; He, H.; Liang, C.; Xie, H. Impact of Corn Stover Mulch Quantity on Microbial Carbon Metabolic Activities and Diversity of Black Soil. Chin. J. Soil Sci. 2023, 54, 407–415. [Google Scholar]
  8. Tian, S.; Zhang, Y.; Bian, W.; Dong, L.; Jiafa, L.; Guo, H. Effects of subsoiling and straw return on soil labile organic carbon fractions in continuous rotary tillage cropland. Trans. Chin. Soc. Agric. Eng. 2020, 36, 185–192. [Google Scholar]
  9. Zheng, B.; Campbell, J.B.; Serbin, G.; Galbraith, J.M. Remote sensing of crop residue and tillage practices: Present capabilities and future prospects. Soil Tillage Res. 2014, 138, 26–34. [Google Scholar] [CrossRef]
  10. Zhang, Z.; Zhu, L. A Review on Unmanned Aerial Vehicle Remote Sensing: Platforms, Sensors, Data Processing Methods, and Applications. Drones 2023, 7, 398. [Google Scholar] [CrossRef]
  11. Huang, J.; Liu, Z.; Wan, W.; Liu, Z.; Wang, J.; Wang, S. Remote sensing retrieval of maize residue cover on soil heterogeneous background. Chin. J. Appl. Ecol. 2020, 31, 474–482. [Google Scholar]
  12. Daughtry, C.S.T. Discriminating crop residues from soil by shortwave infrared reflectance. Agron. J. 2001, 93, 125–131. [Google Scholar] [CrossRef]
  13. Daughtry, C.S.T.; Guy, S.; Reeves, J.B., III; Doraiswamy, P.C.; Hunt, E.R., Jr. Spectral reflectance of wheat residue during decomposition and remotely sensed estimates of residue cover. Remote Sens. 2010, 2, 416–431. [Google Scholar] [CrossRef]
  14. Burud, I.; Lange, G.; Lillemo, M.; Bleken, E.; Grimstad, L.; From, P.J. Exploring Robots and UAVs as Phenotyping Tools in Plant Breeding. IFAC World Congr. 2018, 50, 11479–11484. [Google Scholar] [CrossRef]
  15. Kavoosi, Z.; Raoufat, M.H.; Dehghani, M.; Abdolabbas, J.; Kazemeini, S.A.; Nazemossadat, M.J. Feasibility of satellite and drone images for monitoring soil residue cover. J. Saudi Soc. Agric. Sci. 2020, 19, 56–64. [Google Scholar] [CrossRef]
  16. Yang, B.; Lin, H.; He, Y. Data-Driven Methods for the Estimation of Leaf Water and Dry Matter Content: Performances, Potential and Limitations. Sensors 2020, 20, 5394. [Google Scholar] [CrossRef]
  17. Liu, L.; Song, B.; Zhang, S.; Liu, X. A Novel Principal Component Analysis Method for the Reconstruction of Leaf Reflectance Spectra and Retrieval of Leaf Biochemical Contents. Remote Sens. 2017, 9, 1113. [Google Scholar] [CrossRef]
  18. Fei, S.; Hassan, M.A.; Xiao, Y.; Rasheed, A.; Xia, X.; Ma, Y.; Fu, L.; Chen, Z.; He, Z. Application of multi-layer neural network and hyperspectral reflectance in genome-wide association study for grain yield in bread wheat. Field Crops Res. 2022, 289, 108730. [Google Scholar] [CrossRef]
  19. Fu, Y.; Yang, G.; Wang, J.; Song, X.; Feng, H. Winter wheat biomass estimation based on spectral indices, band depth analysis and partial least squares regression using hyperspectral measurements. Comput. Electron. Agric. 2014, 100, 51–59. [Google Scholar] [CrossRef]
  20. Han, W.; Cui, J.; Cui, X.; Ma, W.; Li, G. Estimation of Farmland Soil Salinity Content Based on Feature Optimization and Machine Learning Algorithms. Trans. Chin. Soc. Agric. Mach. 2023, 54, 328–337. [Google Scholar]
  21. Ban, S.; Liu, W.; Tian, M.; Wang, Q.; Yuan, T.; Chang, Q.; Li, L. Rice Leaf Chlorophyll Content Estimation Using UAV-Based Spectral Images in Different Regions. Agronomy 2022, 12, 2832. [Google Scholar] [CrossRef]
  22. Näsi, R.; Vilianen, N.; Kaivosoja, J.; Alhonoja, K.; Hakala, T.; Markelin, L.; Honkavaara, E. Estimating biomass and nitrogen amount of barley and grass using UAV and aircraft based spectral and photogrammetric 3D features. Remote Sens. 2018, 10, 1082. [Google Scholar] [CrossRef]
  23. Feng, Y.; Cui, N.; Hao, W.; Gao, L.; Cong, D. Estimation of soil temperature from meteorological data using different machine learning models. Geoderma 2019, 338, 67–77. [Google Scholar] [CrossRef]
  24. Jia, Y.; Su, Y.; Zhang, R.; Zhang, Z.; Lu, Y.; Shi, D.; Xu, C.; Huang, D. Optimization of an extreme learning machine model with the sparrow search algorithm to estimate spring maize evapotranspiration with film mulching in the semiarid regions of China. Comput. Electron. Agric. 2022, 201, 107298. [Google Scholar] [CrossRef]
  25. Zhu, B.; Feng, Y.; Gong, D.; Jiang, S.; Zhao, L.; Cui, N. Hybrid particle swarm optimization with extreme learning machine for daily reference evapotranspiration prediction from limited climatic data. Comput. Electron. Agric. 2020, 173, 105430. [Google Scholar] [CrossRef]
  26. Mo, L.; Chen, H.; Chen, W.; Feng, Q.; Xu, L. Study on evolution methods for the optimization of machine learning models based on FT-NIR spectroscopy. Infrared Phys. Technol. 2020, 108, 103366. [Google Scholar] [CrossRef]
  27. Cao, J.; Zhang, J.; Xiao, Q.; Wang, F.; Han, X.; Huang, Y. Estimation and Uncertainty Analysis of Aboveground Carbon Storage of Pinus densata based on Random Forests and Monte Carlo. For. Res. 2023, 36, 131–139. [Google Scholar]
  28. Yang, T.; Wang, X.; Hou, F.; Wang, X.; Li, D.; Guo, M.; Lin, R.; Zhang, S. Monte Carlo method and its application in oil and gas reserve assessment. Prog. Geophys. 2021, 36, 2109–2118. [Google Scholar]
  29. Xie, H.; Du, H.; Wang, Y.; Li, S. Protective Cultivation Technology Model of Strip Tillage with Full Coverage of Corn Straw. Agric. Eng. 2020, 10, 24–26. [Google Scholar]
  30. Schnell, J.A. Monitoring the Vernal Advancement and Retrogradation (Greenwave Effect) of Natural Vegetation. Nasa/gsfct Type Final Report. U.S. Patent E75-10354, 1 November 1974. [Google Scholar]
  31. Tucker, C.J. Red and Photographic Infrared Linear Combinations for Monitoring Vegetation. Remote Sens. Environ. 1979, 8, 127–150. [Google Scholar] [CrossRef]
  32. Glenn, E.P.; Huete, A.R.; Nagler, P.L.; Nelson, S.G. Relationship Between Remotely-sensed Vegetation Indices, Canopy Attributes and Plant Physiological Processes: What Vegetation Indices Can and Cannot Tell Us About the Landscape. Sensors 2008, 8, 2136–2160. [Google Scholar] [CrossRef] [PubMed]
  33. Jiang, Y.; Hu, M.; Li, M.; Zhang, X. Research on Remote Sensing Estimation Method of Forest Aboveground Biomass at County Scale. J. Southwest For. Univ. 2015, 35, 53–59. [Google Scholar]
  34. Jordan, C.F. Derivation of leaf-area index from quality of light on the forest floor. Ecology 1969, 50, 663–666. [Google Scholar] [CrossRef]
  35. Gitelson, A.A.; Kaufman, Y.J.; Merzlyak, M.N. Use of a green channel in remote sensing of global vegetation from EOS-MODIS. Remote Sens. Environ. 1996, 58, 289–298. [Google Scholar] [CrossRef]
  36. Zhao, W.; Ma, F.; Ma, H.; Zhou, C. Soil salinity inversion model based on unmanned aerial vehicle multispectral images. J. Agric. Eng. 2022, 38, 93–101. [Google Scholar]
  37. Bao, Y.; Tian, Q.; Wang, L. A Study on Crop Straw Spectrum Diagnosis Model and Extraction Using HJ-1 CCD Remote Sensing Data. Remote Sens. Inf. 2011, 15–19+122. [Google Scholar]
  38. Tian, Y.; Yang, J.; Yao, X.; Cao, W.; Zhu, Y. Prediction of total nitrogen content in the leaf layer of rice population using leaf hyperspectral index. J. Crop Sci. 2010, 36, 1529–1537. [Google Scholar]
  39. Tomoaki, M.; Hiroki, Y.; Kayo, F.; Hirokazu, Y. Inter-Comparison of ASTER and MODIS Surface Reflectance and Vegetation Index Products for Synergistic Applications to Natural Resource Monitoring. Sensors 2008, 8, 2480–2499. [Google Scholar] [CrossRef] [PubMed]
  40. Barnes, E.M.; Clarke, T.R.; Richards, S.E. Coincident detection of crop water stress, nitrogen status and canopy density using ground-based multispectral data. In Proceedings of the 5th International Conference on Precision Agriculture, Bloomington, MN, USA, 16–19 July 2000; Volume 7, pp. 16–19. [Google Scholar]
  41. Tian, Y. Research on Monitoring Nitrogen Nutritional Parameters of Rice Based on Hyperspectral Remote Sensing; Nanjing Agricultural University: Nanjing, China, 2008. [Google Scholar]
  42. Zhang, X.; Shen, H.; Huang, T.; Wu, Y.; Guo, B.; Liu, Z.; Luo, H.; Tang, J.; Zhou, H.; Wang, L.; et al. Improved random forest algorithms for increasing the accuracy of forest aboveground biomass estimation using Sentinel-2 imagery. Ecol. Indic. 2024, 159, 111752. [Google Scholar] [CrossRef]
  43. Liu, Y.; Sun, Y.; Gao, X.; Wang, L.; Wang, Y.; Liu, M.; Cui, S. Establishment and optimization of aerial multispectral field straw mulch quantity inversion mode. Opt. Precis. Eng. 2024, 32, 1773–1787. [Google Scholar] [CrossRef]
  44. Claudia, C.; Vincenzo, C.; Mario, P.; Francesco, Z. Machine Learning and Genetic Algorithms: A case study on image reconstruction. Knowl.-Based Syst. 2024, 284, 111194. [Google Scholar]
  45. Mbula, N.N.; Shu, J.; Grant, C.M.; Allou, K.F.K.; Paulin, K.H.; Konan, Y.H.R. Brittleness index prediction using modified random forest based on particle swarm optimization of Upper Ordovician Wufeng to Lower Silurian Longmaxi shale gas reservoir in the Weiyuan Shale Gas Field, Sichuan Basin, China. Geoenergy Sci. Eng. 2024, 233, 212518. [Google Scholar]
  46. Zhao, F.; Cheng, Y.; Pan, H.; Cheng, Y.; Zhang, X.; Wu, B.; Hu, Y. Solving redundant inverse kinematics of CMOR based on chaos-driven particle swarm optimization algorithm. Fusion Eng. Des. 2023, 192, 113712. [Google Scholar] [CrossRef]
  47. Huo, L.; Zhao, L.; Yao, Z.; Jia, J.; Zhao, Y.; Fu, G.; Cong, H. Difference of the ratio of maize stovers to grain and spatiotemporal variation characteristics of maize stovers in China. Trans. Chin. Soc. Agric. Eng. 2020, 36, 227–234. [Google Scholar]
  48. Tian, Y. Interpretation of Technical code of crop straw surveying and evaluating. Agric. Eng. Technol. (New Energy Industry) 2009, 6, 7–13. [Google Scholar]
  49. Zhang, Z.; Lai, Z.; Xu, Y.; Shao, L.; Wu, J.; Xie, G.S. Discriminative elastic-net regularized linear regression. IEEE Trans. Image Process. 2017, 26, 1466–1481. [Google Scholar] [CrossRef]
  50. Liu, W.; Li, Q. An efficient elastic net with regression coefficients method for variable selection of spectrum data. PLoS ONE 2017, 12, e0171122. [Google Scholar] [CrossRef]
  51. Jiang, D.; Zhuang, D.; Fu, J.; Huang, Y.; Wen, K. Bioenergy potential from crop residues in china: Availability and distribution. Renew. Sustain. Energy Rev. 2012, 16, 1377–1382. [Google Scholar] [CrossRef]
  52. Morellos, A.; Pantazi, X.E.; Moshou, D.; Alexandridis, T.; Whetton, R.; Tziotzios, G.; Wiebensohn, J.; Bill, R.; Mouazen, A.M. Machine learning based prediction of soil total nitrogen, organic carbon and moisture content by using vis-nir spectroscopy. Biosyst. Eng. 2016, 152, 104–116. [Google Scholar] [CrossRef]
  53. Huang, M.; Wan, X.; Zhu, Q.; Zhang, M. Thickness Measurement of Green Soybean Using Hyperspectral Imaging Technology. J. Food Sci. Biotechnol. 2012, 31, 1142–1147. [Google Scholar]
  54. Yue, J.; Tian, Q.; Dong, X.; Xu, K.; Zhou, C. Using Hyperspectral Crop Residue Angle Index to Estimate Maize and Winter-Wheat Residue Cover: A Laboratory Study. Remote Sens. 2019, 11, 807. [Google Scholar] [CrossRef]
  55. Ye, F.; Zhou, B. Mangrove Classification from Unmanned Aerial Vehicle Hyperspectral Images Using Object-Oriented Methods Based on Feature Combination and Optimization. Sensors 2024, 24, 4108. [Google Scholar] [CrossRef] [PubMed]
  56. Pan, X.; Yuan, J.; Yang, Z.; Tansey, K.; Xie, W.; Song, H.; Wu, Y.; Yang, Y. Spatio-Temporal Variation of Cyanobacteria Blooms in Taihu Lake Using Multiple Remote Sensing Indices and Machine Learning. Remote Sens. 2024, 16, 889. [Google Scholar] [CrossRef]
Figure 1. The diagram of the locations of different study areas.
Figure 1. The diagram of the locations of different study areas.
Agronomy 14 02352 g001
Figure 2. Sampling and measurement process.
Figure 2. Sampling and measurement process.
Agronomy 14 02352 g002
Figure 3. Spectral reflectance acquisition.
Figure 3. Spectral reflectance acquisition.
Agronomy 14 02352 g003
Figure 4. GA-HPSO algorithm process.
Figure 4. GA-HPSO algorithm process.
Agronomy 14 02352 g004
Figure 5. Comparison of different evolutionary curves. (a) Study area 1, Autumn 22; (b) Study area 1, Spring 23; (c) Study area 2, Autumn 22; (d) Study area 2, Spring 23.
Figure 5. Comparison of different evolutionary curves. (a) Study area 1, Autumn 22; (b) Study area 1, Spring 23; (c) Study area 2, Autumn 22; (d) Study area 2, Spring 23.
Agronomy 14 02352 g005aAgronomy 14 02352 g005b
Figure 6. Comparison of different evolutionary curves. (a) Sphere Evolution; (b) Griewank Evolution; (c) Rastrigin Evolution.
Figure 6. Comparison of different evolutionary curves. (a) Sphere Evolution; (b) Griewank Evolution; (c) Rastrigin Evolution.
Agronomy 14 02352 g006
Figure 7. Scatterplot of measured versus and predicted values. (a) Study area 1, Spring 22; (b) Study area 1, Autumn 22; (c) Study area 1, Spring 23; (d) Study area 1, Autumn 23; (e) Study area 2, Spring 22; (f) Study area 2, Autumn 22; (g) Study area 2, Spring 23; (h) Study area 2, Autumn 23.
Figure 7. Scatterplot of measured versus and predicted values. (a) Study area 1, Spring 22; (b) Study area 1, Autumn 22; (c) Study area 1, Spring 23; (d) Study area 1, Autumn 23; (e) Study area 2, Spring 22; (f) Study area 2, Autumn 22; (g) Study area 2, Spring 23; (h) Study area 2, Autumn 23.
Agronomy 14 02352 g007aAgronomy 14 02352 g007b
Figure 8. Experience distribution validation chart.
Figure 8. Experience distribution validation chart.
Agronomy 14 02352 g008
Figure 9. Validation experiment process.
Figure 9. Validation experiment process.
Agronomy 14 02352 g009
Figure 10. Comparison results between predicted and actual values in the validation area. (a) Study area 3; (b) Study area 4; (c) Study area 5; (d) Study area 6; (e) Study area 7; (f) Study area 8.
Figure 10. Comparison results between predicted and actual values in the validation area. (a) Study area 3; (b) Study area 4; (c) Study area 5; (d) Study area 6; (e) Study area 7; (f) Study area 8.
Agronomy 14 02352 g010
Table 1. Description of study area selection.
Table 1. Description of study area selection.
Study AreaSiteLatitude and LongitudeArea [m2]
Experimental areaYingjun Town, Erdao (study area 1)125°29′56″ E 43°52′06″ N22,500
Bahao Town, Yushu (study area 2)126°13′59″ E 45°05′44″ N20,000
Verification areaHuancheng Town, Yushu (study area 3)126°28′14″ E 44°51′24″ N25,000
Biangang Town, Dehui (study area 4)125°39′02″ E 44°35′04″ N15,000
Tiantai Town, Dehui (study area 5)125°35′57″ E 44°34′04″ N22,500
Kai’an Town, Nong’an (study area 6)125°11′34″ E 44°07′27″ N15,000
Nong’an Town, Nong’an (study area 7)125°14′44″ E 44°36′20″ N10,000
Chaoyangpo Town, Gongzhuling (study area 8)124°47′45″ E 43°36′54″ N20,000
Table 2. The weather information used for collecting data on different study areas.
Table 2. The weather information used for collecting data on different study areas.
Study AreaSiteCollection TimeCollecting WeatherSample Size
Experimental areaStudy area 123 April 2022Sunny, West wind, 3.4–5.4 m/s, 12–3 °C54
1 November 2022Sunny, West wind, 3.4–5.4 m/s, 12–4 °C54
23 April 2023Sunny, Northwest wind, 3.4–5.4 m/s, 11–3 °C54
26 October 2023Sunny, Northwest wind, 3.4–5.4 m/s 10–4 °C54
Study area 225 April 2022Sunny, West wind, 5.5–7.9 m/s, 13–4 °C48
2 November 2022Sunny, West wind, 3.4–5.4 m/s, 6–4 °C48
11 May 2023Sunny, Southwest wind, 3.4–5.4 m/s, 14–1 °C48
27 October 2023Sunny, West wind, 0.3–1.5 m/s, 12–3 °C48
Verification areaStudy area 328 October 2023Sunny, Southeast wind, 5.5–7.9 m/s, 12–2 °C60
Study area 42 November 2023Sunny, Southeast wind, 5.5–7.9 m/s, 16–4 °C36
Study area 53 November 2023Sunny, Northeast wind, 3.4–5.4 m/s, 2–7 °C54
Study area 624 October 2023Sunny, Northeast wind, 3.4–5.4 m/s, 19–1 °C36
Study area 725 October 2023Sunny, Southwest wind, 5.5–7.9 m/s, 19–1 °C24
Study area 829 October 2023Sunny, Southwest wind, 3.4–5.4 m/s, 18–7 °C48
Table 3. Spectral index and its formula.
Table 3. Spectral index and its formula.
Spectral IndexAbbreviationFormulaReference
Normalized difference vegetation indexNDVINDVI = (NIR − R)/(NIR + R)[30]
Difference vegetation indexDVIDVI = NIR − R[31]
Enhanced vegetation indexEVIEVI = 2.5(NIR − R)/(NIR + 6R − 7.5B + 1)[32]
Brightness indexBIBI = (R + NIR)0.5[33]
Simple ratio indexSRSR = NIR/R[34]
Green-light normalized difference vegetation indexGNDVIGNDVI = (NIR − G)/(NIR + G)[35]
Green-light ratio vegetation indexGRVIGRVI = NIR/G[36]
Straw multiplier indexSMISMI = NIR × (NIR − G) × (R − G)[37]
Red edge ratio indexRERIRERI = RE/NIR[38]
Near-infrared normalized indexNRNINRNI = NIR/(NIR + R + G)[39]
Normalized difference red edge indexNDRENDRE = (NIR − RE)/(NIR + RE)[40]
Red-light normalized difference indexRNDIRNDI = (RE − R)/(RE + R)[41]
B, G, R, RE, and NIR are the spectral reflectance at 450, 555, 660, 720, and 840 nm wavelengths, respectively.
Table 4. Research data related to straw in the study area.
Table 4. Research data related to straw in the study area.
Study AreasPeriodsμ [%]P [t] δ [%]Pc [t]
Study area 1Spring 220.9021.5520.9320.043
Autumn 220.9121.9300.9220.176
Spring 230.9121.9300.9220.176
Autumn 230.8922.0010.9220.241
Study area 2Spring 220.9116.5420.9215.219
Autumn 220.9217.0270.9315.835
Spring 230.9217.0270.9315.835
Autumn 230.9016.4760.9114.993
Study area 3Autumn 230.9021.8070.9119.844
Study area 40.8912.0910.9311.245
Study area 50.8918.3350.9216.868
Study area 60.8912.1660.9211.193
Study area 70.898.2910.907.462
Study area 80.9017.0700.9315.875
Table 5. Spectral variable screening based on ENET.
Table 5. Spectral variable screening based on ENET.
Study AreasPeriodsNDVIDVIEVIBISRGNDVIGRVISMIRERINRNINDRERNDI
Study area 1Spring 220.390000.470.43000.5600.540
Autumn 220.560.53000.570.50000.6100.550
Spring 230.380000.450.42000.5600.500
Autumn 230.580.49000.550.51000.6000.540
Study area 2Spring 220.380000.460.42000.5700.540
Autumn 220.550.53000.570.50000.6000.550
Spring 230.410000.460.40000.5800.540
Autumn 230.550.50000.560.53000.6100.550
Table 6. Model inversion results based on different data.
Table 6. Model inversion results based on different data.
PeriodsModeling AlgorithmStudy Area 1Study Area 2
Modeling SetValidation SetModeling SetValidation Set
R2cRMSEcR2vRMSEvR2cRMSEcR2vRMSEv
Spring 22RF0.650.06650.620.06790.650.06500.630.0651
GA-RF0.710.05830.690.06140.740.04620.680.0542
PSO-RF0.710.06010.670.06300.730.04670.670.0571
GA-HPSO0.780.04150.710.05960.760.04510.700.0515
Autumn 22RF0.690.05960.670.06310.700.05280.670.0564
GA-RF0.740.04490.730.04620.770.04240.720.0447
PSO-RF0.750.04510.710.04660.790.04150.710.0454
GA-HPSO0.770.04080.750.04530.790.04270.750.0441
Spring 23RF0.690.06260.630.06670.690.06340.640.0640
GA-RF0.710.05780.690.05970.730.05130.680.0630
PSO-RF0.710.06010.670.06240.720.05470.660.0676
GA-HPSO0.770.04740.720.05740.760.04740.700.0544
Autumn 23RF0.710.05870.680.06040.720.05430.680.0637
GA-RF0.750.04810.710.05230.750.04580.730.0482
PSO-RF0.740.04820.720.05010.750.04710.710.0511
GA-HPSO0.780.04320.740.04850.780.04350.750.0453
Table 7. Estimation results based on the different data.
Table 7. Estimation results based on the different data.
Study AreasPeriodsPc [t]Mean Value MethodMonte Carlo Method
Estimated Value [t]Accuracy [%]Estimated Value [t]Accuracy [%]
Study area 1Spring 2220.04315.98374.6017.79887.39
Autumn 2220.17616.18475.3318.01488.00
Spring 2320.17616.32876.4317.96887.71
Autumn 2320.24116.79679.4918.33789.62
Study area 2Spring 2215.21912.37677.0313.57987.92
Autumn 2215.83512.74775.7714.37689.85
Spring 2315.83513.10779.1914.08587.58
Autumn 2314.99312.66181.5813.49488.89
Table 8. Estimation results of global straw mulch quantity.
Table 8. Estimation results of global straw mulch quantity.
Study AreasΜ [%]Pc [t]Monte Carlo Method
Estimated Value [t]Accuracy [%]
Study area 30.9019.84417.85788.87
Study area 40.8911.24510.05988.21
Study area 50.8916.86815.23489.27
Study area 60.8911.19310.01688.25
Study area 70.897.4628.37589.10
Study area 80.9015.87514.46790.27
Table 9. Estimation results of the straw return rate in the validation area.
Table 9. Estimation results of the straw return rate in the validation area.
Study AreasP [t]Straw Return Quantity to The Field [t]Straw Return Rate [%]
Study area 321.80717.98282.46
Study area 412.09110.13483.81
Study area 518.33515.34783.70
Study area 612.16610.09182.94
Study area 78.2916.82582.32
Study area 817.0714.56785.34
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, Y.; Sun, Y.; Wang, Y.; Wang, J.; Gao, X.; Wang, L.; Liu, M. Model Optimization and Application of Straw Mulch Quantity Using Remote Sensing. Agronomy 2024, 14, 2352. https://doi.org/10.3390/agronomy14102352

AMA Style

Liu Y, Sun Y, Wang Y, Wang J, Gao X, Wang L, Liu M. Model Optimization and Application of Straw Mulch Quantity Using Remote Sensing. Agronomy. 2024; 14(10):2352. https://doi.org/10.3390/agronomy14102352

Chicago/Turabian Style

Liu, Yuanyuan, Yu Sun, Yueyong Wang, Jun Wang, Xuebing Gao, Libin Wang, and Mengqi Liu. 2024. "Model Optimization and Application of Straw Mulch Quantity Using Remote Sensing" Agronomy 14, no. 10: 2352. https://doi.org/10.3390/agronomy14102352

APA Style

Liu, Y., Sun, Y., Wang, Y., Wang, J., Gao, X., Wang, L., & Liu, M. (2024). Model Optimization and Application of Straw Mulch Quantity Using Remote Sensing. Agronomy, 14(10), 2352. https://doi.org/10.3390/agronomy14102352

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