Next Article in Journal
From Planting to Harvesting: The Role of Agricultural Machinery in Crop Cultivation
Previous Article in Journal
Response Prediction and Experimental Validation of Vibration Noise in the Conveyor Trough of a Combine Harvester
Previous Article in Special Issue
Identification of Interannual Variation Frequency of Cropland Cropping Intensity Based on Remote Sensing Spatiotemporal Fusion and Crop Phenological Rhythm: A Case Study of Zhenjiang, Jiangsu
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Early Yield Prediction of Oilseed Rape Using UAV-Based Hyperspectral Imaging Combined with Machine Learning Algorithms

1
Guangxi Key Laboratory of Brain-Inspired Computing and Intelligent Chips, School of Electronic and Information Engineering, Guangxi Normal University, Guilin 541004, China
2
Key Laboratory of Integrated Circuits and Microsystems, Education Department of Guangxi Zhuang Autonomous Region, Guangxi Normal University, Guilin 541004, China
3
School of Biosystems and Food Engineering, University College Dublin (UCD), Belfield, 4 Dublin, Ireland
4
College of Biosystems Engineering and Food Science, Zhejiang University, Hangzhou 310058, China
*
Authors to whom correspondence should be addressed.
Agriculture 2025, 15(10), 1100; https://doi.org/10.3390/agriculture15101100
Submission received: 28 February 2025 / Revised: 22 April 2025 / Accepted: 15 May 2025 / Published: 19 May 2025

Abstract

:
Oilseed rape yield critically reflects varietal superiority. Rapid field-scale estimation enables efficient high-throughput breeding. This study evaluates unmanned aerial vehicle (UAV) hyperspectral imagery’s potential for yield prediction at the pod stage by utilizing wavelength selection and vegetation indices. Meanwhile, optimized feature selection algorithms identified effective wavelengths (EWs) and vegetation indices (VIs) for yield estimation. The optimal yield estimation models based on EWs and VIs were established, respectively, by using multiple linear regression (MLR), partial least squares regression (PLSR), extreme learning machine (ELM), and a least squares support vector machine (LS-SVM). The main results were as follows: (i) The yield prediction of oilseed rape using EWs showed better prediction and robustness compared to the full-spectral model. In particular, the competitive adaptive reweighted sampling–extreme learning machine (CARS-ELM) model (Rpre = 0.8122, RMSEP = 170.4 kg/hm2) achieved the best prediction performance. (ii) The ELM model (Rpre = 0.7674 and RMSEP = 187.6 kg/hm2), using 14 combined VIs, showed excellent performance. These results indicate that the remote sensing image data obtained from the UAV hyperspectral remote sensing system can be used to enable the high-throughput acquisition of oilseed rape yield information in the field. This study provides technical guidance for the crop yield estimation and high-throughput detection of breeding information.

Graphical Abstract

1. Introduction

Oilseed rape (Brassica napus L.) belongs to the cruciferous family (Brassicaceae). It stands as one of the most extensively cultivated oilseed crops globally, with its primary growing regions encompassing Europe, Canada, China, India, Australia, and other countries [1,2]. Until now, oilseed rape is a vital source of forage crops, oil crops, and energy crops. It has great economic benefits and occupies over 40% of China’s total production of oil crops [3,4]. Rapeseed oil, produced from oilseed rape, is a good source of healthy fats and fat-soluble vitamins. So, it is considered a beneficial and affordable edible vegetable oil. In addition, the remaining “cake” after rapeseed oil extraction can also be used as a high-protein animal feed. There are also other potentials in various sectors such as renewable vehicle fuel, environmentally friendly lubricating oil, and raw materials in the chemical industry. Due to its great importance, the research on phenotypic information and the physiological process of winter oilseed rape has attracted growing attention. Most research focuses on guiding oilseed rape producers to be more efficient in planting, breeding, harvesting, and marketing.
Among the numerous studies on crop yield, it is of great significance to enable accurate early prediction for farmers [5]. The traditional method of yield estimation is to establish an approximate relationship between the crop yield and agronomic factors or climate factors after manual field surveys and statistical analysis [6,7]. However, with large-scale commercial oilseed rape breeding and the rapid soaring collaboration of breeding materials in recent years, these conventional methods are cumbersome and time-consuming, and the data’s consistency and timeliness can scarcely be guaranteed. The low-altitude remote sensing platform combined with an unmanned aerial vehicle (UAV) and remote sensing technology, as a versatile, flexible, and low-cost tool, can provide spatial and phenotypic information of field crops because of its high spatial and temporal resolution [8,9]. As a result, UAV-based low-altitude remote sensing is deemed a useful and crucial technology in agriculture with strong suitability and adaptability to complex farmland environments [10].
The past decade has seen an exponential increase in research and the application of UAVs in precision agriculture. The rapid development of UAVs marks the arrival of a new era of remote sensing [5]. UAV-based low-altitude remote sensing has many advantages. It is efficient in acquiring both phenotypic and image information in a real-time manner. In this sense, it is suitable for both qualitative and quantitative research concerning different phenotypic information [11,12]. Therefore, it has been widely applied for the analysis of physiological characteristics, disease diagnosis, insect and pest inspection, and the prediction of chlorophyll content, biomass, drought stress sensitivity, the leaf area index, and yield [5]. Crop structure can be influenced by physiological characteristic differences, leading to the corresponding changes in the crops’ spectral features, such as light absorption, reflection, and transmission. To a certain extent, they can reflect a crop’s phenotypic information, covering nutrient abundant deficiency, growth differences, the degree of plant diseases or pest invasion, etc. This is also the theoretical basis for why spectral imaging sensors enable the representation of the phenotypic information of crops [13,14].
Many researchers have been reported to build predictive models of crop yield by using UAV-based RGB and multispectral cameras [15,16,17], and they have coincidentally confirmed the feasibility of using UAV combined with image sensors for yield prediction. The future development of UAV remote sensing applications should be approached to improve the accuracy and adaptability of the prediction model [18]. The UAV hyperspectral remote sensing system is advantageous as it integrates image and spectral information into one system, and it has gradually grown into the cutting-edge technology in the current remote sensing field due to its high spatial and spectral resolution. However, the research on the estimation of crop yield based on the hyperspectral remote sensing system of micro and small UAVs is rare and inadequate. Therefore, this work aimed to explore the feasibility of a UAV remote sensing platform equipped with a UHD185 hyperspectral imager (Cubert GmbH, ULM, Baden-Wurttemberg, Germany) to predict the yield of oilseed rape at the pod information stage. The specific studies are as follows: (1) Compared the effects of various pretreatment methods on spectral features and selected the optimal spectral pretreatment method. (2) Compared the predictive performance of linear and non-linear models constructed based on eight feature wavelength extraction methods to determine the yield of oilseed rape systematically. (3) Obtained the optimal vegetation indices (VIs) suitable for rape yield estimation and investigated the internal relationship and correlation between rape yield and different VIs. (4) Developed the most desirable model to predict oilseed rape yield and acquired the most relevant spectral bands as well as VIs which are used as inputs of models. This work will provide vital guidance for future work in the high-throughput detection of crop breeding and yield estimation.

2. Materials and Methods

2.1. Experimental Location and Crop Husbandry

This work was carried out at the agricultural research station of Zhejiang University, Hangzhou, China (30°18′26″ N, 120°4′29″ E). The entire experiment process was conducted over the rapeseed growing season. The main procedures of the experiment included sowing seeds in a fertile sandy loam seedbed in early October to ensure healthy growth, cautiously transferring the plants to the experimental field in November, and finally harvesting the oilseed rape in May of the following year. The area and row spacing of each row in the experimental field are 24.4 m × 1.4 m and 0.3 m, respectively. Each row has 10 plots, and the area of each plot is 2.4 m × 1.4 m. Some rows do not actually plant rape due to the setting of protection channels. After removing these plots, there are approximately 420 sample plots in total. The fertilizers used in the experiment were calcium superphosphate (Ca(H2PO4)2·H2O), urea (CO(NH2)2), and potassium chloride (KCl). All plots were supplied with identical levels of phosphorus (P) and potassium (K), specifically 60 kg/hm2 of P and 150 kg/hm2 of K, to ensure adequate nutrition for rapeseed growth. To study the effect of nitrogen on rapeseed growth, different nitrogen concentrations were applied to rapeseed. Based on the normal growth demand of 150 kg/hm2, their concentrations should be correspondingly 0, 75, 150, and 225 kg/hm2, which were marked as N0, N1, N2, and N3, respectively. Once the rapeseed matured, all transplanted plants within the trial field were harvested using hand scythes and transported to the laboratory. To determine crop yield, the total weight of the harvested rapeseed from each plot was divided by the plot’s area, with the result expressed in kg/hm 2. Due to human factors, the rapeseed yield data of some sample plots in the yield measurement and calculation process are missing. Finally, 373 rapeseed sample data with complete yield calculations are obtained.

2.2. Hyperspectral Imaging Data Acquisition

The UHD185(Cubert GmbH, Ulm, Baden-Württemberg, Germany) UAV hyperspectral remote sensing system consisted of a six-rotor UAV (Figure 1a), a UHD185 imaging spectrometer (Figure 1b), a ground control system, and a radiometric calibration system. The weight of the UAV was 4.2 kg, and the endurance time was 15–20 min. UHD185 is a frame-synchronous imaging spectrometer with 125 spectral channels in the 450–950 nm range. The UHD185 imager could simultaneously obtain a 12-bit (4096 DN) hyperspectral image in the dynamic range of 50 × 50 pixels and a grayscale image of 1000 × 1000 pixels. The acquired spectral image data were stored in an inner microcomputer.
The acquisition of hyperspectral images of rape at the pod formation stage was carried out using the UHD185 UAV hyperspectral remote sensing system. This experiment adopted the time-delay shooting method. The entire process was controlled by the POKINE Z software (EXTRA Computer GmbH, Sachsenhausen, Germany). During image collection, the UAV was kept at the 50 m flight altitude, the forward overlap stayed in the range of 40–70%, and the lateral overlap was more than 30%. Meanwhile, the hyperspectral imager worked downward vertically to obtain images with a 23 mm focal length at a fixed 1 ms sampling interval.

2.3. Data Processing

The overall procedure for predicting oilseed rape yield from UAV hyperspectral data, as depicted in Figure 2, involved several sequential steps beginning with data acquisition and preparation. Initially, raw hyperspectral image cubes and associated panchromatic images captured by the UAV system underwent image fusion to integrate spatial and spectral details. Crucially, radiometric calibration was then performed to convert sensor digital numbers into physically meaningful spectral reflectance values, followed by image stitching using Agisoft Photoscan (version 1.2.6, AgiSoft LLC, St. Petersburg, Russia) to assemble the individual calibrated captures into a complete hyperspectral orthophoto mosaic of the experimental field. Following the creation of the orthophoto, the extracted spectral reflectance data required preprocessing to minimize noise and environmental interference. A comparative analysis of various methods, including smoothing (MAS and SG), scatter correction (MSC and SNV), derivatives (1-Der and 2-Der), and wavelet transform (WT), was conducted using PLSR models. This evaluation identified WT as the optimal technique for enhancing data quality before further analysis.
Subsequently, two primary pathways were explored to extract the most relevant spectral features for yield prediction and reduce data dimensionality. The first pathway focused on selecting a subset of effective wavelengths (EWs) directly from the preprocessed spectra using algorithms like SPA, GAPLS, UVE, CARS, and others. The second pathway involved calculating various established and derived vegetation indices (VIs) and selecting the best ones through correlation analysis and ANOVA with the measured yield data. Finally, using either the selected EWs or the chosen combination of VIs as input features, the dataset was split into calibration and prediction sets. Four machine learning models (MLR, PLSR, ELM, and LS-SVM) were trained on the calibration data, and their predictive performance was rigorously evaluated using the independent prediction set to identify the most accurate and robust model for estimating oilseed rape yield. This comprehensive workflow ensured systematic data refinement and comparative model assessment [19,20].

2.3.1. Hyperspectral Remote Sensing Image Processing

Processing the hyperspectral imagery acquired by the UAV system was a critical multi-step procedure essential for generating accurate spectral data for yield analysis, as illustrated in the workflow diagram (Figure 3). Given that individual remote sensing images captured by the UAV typically cover only a small area, image stitching is a key technology to create a complete spectral map of the entire experimental field from numerous overlapping frames. The process began with the raw data from the UHD185 sensor, which simultaneously captures lower-resolution (50 × 50 pixel) hyperspectral cubes and higher-resolution (1000 × 1000 pixel) panchromatic images. Initial processing, which was performed using the Cubert-Pilot software (Cubert GmbH, Ulm, Baden-Württemberg, Germany), involved fusing each hyperspectral cube with its corresponding panchromatic image. This step, often referred to as pansharpening, significantly enhanced the spatial resolution of the hyperspectral data to match that of the panchromatic image while also incorporating radiometric calibration based on diffuse calibration data to convert raw digital numbers into reflectance values.
Following the fusion and calibration of individual image pairs, the Agisoft Photoscan software was employed to mosaic these multiple overlapping high-resolution hyperspectral images into a single, continuous representation of the test field. Leveraging the rich point cloud information derived primarily from the panchromatic component of the fused images, along with ground control points for accurate georeferencing, Agisoft Photoscan performed photo alignment and applied dense multi-view stereo matching algorithms. This resulted in the generation of a comprehensive, 125-band hyperspectral digital orthophoto covering the 450–950 nm spectral range with a 4 nm sampling interval. A quality check using quick stitching in Photoscan was advisable after each flight to ensure data integrity and arrange re-flights if necessary due to poor stitching or missing imagery. Finally, to obtain the spectral information relevant to each experimental plot, the ENVI (version 4.6, ITT, Visual Information Solutions, Boulder, CO, USA) software was used.

2.3.2. Radiation Calibration

As the basis of remote sensing quantification, the radiometric calibration of UAV remote sensing images is necessary, i.e., converting the digital number (DN) values of images obtained by spectral image sensors into spectral reflectance with actual physical meaning. Therefore, radiometric calibration using black and white plates was needed before data collection using the UHD185 UAV hyperspectral remote sensing system, which enabled the calibration of the spectral reflectance for the target area to obtain more accurate spectral reflectance data. The radiation calibration was carried out using the following equation:
Ref target = D N t a r g e t D N d a r k D N p a n e l D N d a r k × Ref panel
where
DNdark represents the systematic error of the hyperspectral imager;
DNtarget represents the DN values of the target;
DNpanel represents the standard reference plate in the original hyperspectral image;
Reftarget represents the reflectance of the target;
Refpanel represents the standard reference plate.

2.3.3. Spectral Preprocessing

To reduce the impact of external disturbances and improve the reliability and accuracy of the model, it is essential to apply spectral preprocessing. There were several preprocessing methods used: (1) Savitzky–Golay smoothing (SG) [21] and moving average smoothing (MAS) were performed to reduce the influence of noise. (2) A standard normalized variate (SNV) was used to correct the scattered noise interference [22]. (3) Multiplicative scatter correction (MSC) was applied to alleviate the interference caused by the scattering effect [23]. (4) A de-trending algorithm was employed in conjunction with SNV [22]. (5) A differential analysis algorithm including the first derivative (1-Der) and second derivative (2-Der) was implemented to reduce the influence of background interference or baseline drift [24]. (6) Wavelet transform (WT) was used for spectral data compression and smoothing processes in spectral analysis [25].

2.3.4. Wavelength Variable Selection

Data redundancy or collinearity is likely to result in an inferior model. Therefore, the selection of the optimal wavelengths is desired to trim the obtained spectral data, and it can save stored space and shorten data processing time. It is beneficial for reducing computational complexity and enhancing predictive ability, and the result can be used to develop online or portable multispectral devices [26].
In this study, multiple spectral variable selection algorithms were used. As a feed-forward variable selection method for multivariate calibration, successive projection algorithms (SPAs) can significantly decrease the number of variables, thereby increasing the speed and efficiency of modelling [26,27,28]. By calculating straightforward projections in a vector space, the SPA chose the wavelength variable, which has the lowest correlation and redundancy, as well as the largest projection vector. A weighted regression coefficient (BW) method is used to select the optimal wavelength recognized with a larger absolute value of BW, which is regarded as having a greater impact on the prediction result [29,30]. Genetic algorithm partial least squares (GAPLS), as one of the most effective variable selection algorithms, has been widely used in effective variable selection [31]. The genetic algorithm (GA) excels in global exploration and remains unencumbered by the premise of a restricted solution domain. For some discrete, multi-polar, and high-dimensional problems, the GA can still find solutions to these problems that are globally optimal, even in the presence of noise. In 1996, Centner et al. proposed a variable selection method named the uninformative variable elimination (UVE) algorithm. The UVE can eliminate wavelength variables which contribute little to the establishment of the PLS regression model and retain wavelengths which contain useful information. Thus, we can decrease the number of variables, reduce computation time and model complexity, and improve accuracy by using the SPA algorithm [32,33]. Derivative transformation can highlight useful information in the spectrum, weaken background interference, and improve spectral resolution. Different peaks and troughs tend to contain useful information. In this sense, this work selected feature wavelengths located at the major peaks and troughs of the second derivative (2-Der) spectra. Competitive adaptive reweighted sampling (CARS) is a feature variable selection method based on the regression coefficients of the least square model. After the PLS regression model is established using Monte Carlo sampling, adaptive weighted sampling can get rid of the corresponding wavelength ranges whose regression coefficient weight in the PLS model is relatively less than others but retain the rest of the wavelength ranges and take them as a new subset. Then, based on the new subset, the selection will be executed again. After several repeated cycle calculations, when they have the smallest root mean square error for the PLS model, the rest wavelength ranges will be selected as a combination [34]. The random frog (RF) algorithm is a variable selection method based on an iterative and reversible jumping Markov chain Monte Carlo method [35]. RF chooses the best combination of feature wavelength by repetitively screening variable subsets.

2.3.5. Vegetation Indices

The 53 common VIs and 6 other VIs are listed in Table S1 of the Supplementary Materials. These common VIs are divided into three categories: the characteristic index, the soil line vegetation index, and the atmospheric adaptability index. The latter two are mainly used to remove soil and atmospheric effects. The characteristic index is correlated with the epoxidation state of the lutein cycle, chlorophyll a + b concentration, the blue/green/red ratio index, carotenoid concentration, and the canopy structure [36]. The lutein pigment index mainly refers to the photochemical reflectance index (Pri570 and Pri515). When the wavelength is 515 nm, the calculated photochemical reflectance index will hardly suffer from the inference of the structure. Chlorophyll a + b concentration is mainly correlated with the Zarco-Tejada–Miller index (ZM), Vogelmann red edge index 1 (VOG1), Vogelmann red edge index 2 (VOG2), the chlorophyll absorption ratio index (CARI), the transformed chlorophyll absorption ratio index (TCARI), and the TCARI/optimized soil-adjusted vegetation index (OSAVI). The blue/green/red ratio index mainly includes the greenness index (R550/R670), the blue/green pigment index, the blue/red pigment index, and the Lichtenthaler index. Vegetation indices were associated with carotenoid concentration, which includes R520/R500, R515/R570, and R515/R670. The canopy structure index, including the normalized difference vegetation index (NDVI), the simple ratio (SR), the modified simple ratio (MSR), OSAVI, the triangular vegetation index (TVI), modified triangular vegetation index 1 (MTVI1), and other VIs, can be used to evaluate the changes in canopy structure caused by various biotic and abiotic stress.

2.3.6. Machine Learning Algorithms

Artificial intelligence techniques play indispensable roles in the hyperspectral remote sensing of UAVs at low latitudes, especially for image segmentation, ground object classification, and phenotypic index modelling and prediction [37,38]. There are multiple machine learning algorithms used in this work, namely multiple linear regression (MLR), partial least squares regression (PLSR), extreme learning machine (ELM), and a least squares support vector machine (LS-SVM).
MLR aims to establish a linear relationship between the spectral data and the measured indicators (e.g., physiological and biochemical indicators) [26,39]. Partial least squares (PLS) project the original variable to the orthogonal dimension to produce the principal component or principal factor which is also known as the latent variable (LV). LV carries the peak information content and optimal covariance linking spectral data and physical and chemical index readings, enabling it to account for the vast majority of the principal variable information, so PLS is frequently employed for feature extraction [26,29]. In this study, PLSR derived from PLS was used to solve quantitative regression problems. ELM suits well in the generalization of feed-forward neural networks, as it can select input weights and hidden layer deviations randomly during model development and overcome overfitting and local minimum problems effectively in the process of training [40,41]. A support vector machine (SVM) is a suitable modelling method when dealing with a small number of samples [42,43]. The LS-SVM can be used to obtain support vectors using a set of linear equations and can solve multivariate analysis problems, which are both linear and non-linear [26].

2.3.7. Model Evaluation

The correlation coefficient (R), with values closer to 1 indicating stronger predictive power, reflects the linear relationship between predicted and observed values. The root mean square error (RMSE) quantifies the standard deviation of the prediction errors, calculated as the difference between the reference and predicted values. Therefore, a smaller RMSE closer to 0 reveals a smaller prediction error, and the model performs better. In this work, we adopted R and RMSE as evaluation indicators, which were used to evaluate the performance of the model. The R and RMSE were calculated using the following equations:
R = i = 1 n ( y i ^ y ^ ¯ ) ( y i y ¯ ) i = 1 n ( y i ^ y ^ ¯ ) 2 ( y i y ¯ ) 2
R M S E = 1 n 1 · i = 1 n ( y i y i ^ ) 2
where
y i represents the ith sample’s reference value;
y i ^ represents the i th sample’s predicted value;
y ¯ represents the averaged reference value;
y ^ ¯ represents the averaged predicted value;
n is the number of samples.

3. Results and Analysis

3.1. Investigating Yield Metrics and Spectral Attributes in Oilseed Rape Treatment

3.1.1. Descriptive Statistics of Measured Yield

Table 1 presents a statistical overview of the yield data collected for field-grown oilseed rape across varying nitrogen levels. For the 373 samples, yields varied from 2217.3 to 3586.3 kg/hm2. In this experiment, we randomly split all samples into a calibration set and a prediction set in a 2:1 ratio. The calibration set (248 samples) was used to establish a stable and reliable prediction model and evaluate the model objectively. The prediction set (125 samples) served primarily to evaluate the effectiveness of the developed model. As shown in Table 1, the calibration set had a mean yield of 2942.68 kg/hm2 and a median yield of 2977.70 kg/hm2, while the prediction set had a mean of 2947.24 kg/hm2 and a median of 3023.80 kg/hm2. The closeness of the mean and median values within each set, along with the negative skewness values (−0.37 for calibration, −0.53 for prediction), indicates a slight negative skew in the yield distributions for both groups. This suggests a tendency for yields to cluster towards the higher end of the observed range, with fewer instances of very low yields pulling the mean slightly below the median. Furthermore, the kurtosis values (−0.41 and −0.20, respectively) are negative, indicating platykurtic distributions, which are somewhat flatter than a normal distribution and possess fewer extreme outlier yields. The yield of different oilseed rape plots showed noticeable variation, as evidenced by the standard deviations (274.23 and 292.97 kg/hm 2 for the calibration and prediction sets, respectively). Moreover, the sample groups, comprising the calibration and prediction groups, exhibited a considerable span of yield figures (see Table 1). The coefficients of variation (CV) were 9.31% for the calibration set and 9.94% for the prediction set, signifying moderate and comparable relative variability within both sample groups. These characteristics, including the range and distribution of yields, were conducive to establishing a robust quantitative prediction model and providing a reliable basis for its validation.

3.1.2. The Effect of Different Pretreatment Methods on the Yield Prediction of Oilseed Rape

Due to the complex environment of farmland, there are many influencing factors during the flight of the UAV, making the collected image and spectral data prone to interference information, such as noise, background, and stray light. This interference information is highly likely to cause a reduction in the accuracy and stability of the quantitative prediction model. In this study, eight spectral analysis methods (MAS, SG, SNV, MSC, standard normalized variable detrending (SNV-Detrending), 1-Der, 2-Der, and WT) were used to preprocess hyperspectral data. Therein, the wavelet basis functions and decomposition scale for WT preprocessing were db7 and 2, respectively. For different pre-processed spectral and raw spectral data, yield estimation models for oilseed rape were developed based on the PLSR algorithm. At the same time, we used the leave-one-out method of cross-validation (CV) to ascertain the optimal number of LVs needed to evaluate the effectiveness of different methods and select the optimal preprocessing method for subsequent studies.
The results of PLSR analysis models under different spectral pretreatments are shown in Table 2. Compared with the raw spectral curves, the five preprocessing methods, MAS, SG, MSC, SNV, and WT, maintained the trend of the raw spectral curves based on the smoothed spectral noise. The MSC, SNV, and SNV-Detrending reinforced the differences between the spectral curves from different perspectives. It can be seen from Table 2 that there was a certain effect of different spectral preprocessing methods for oilseed rape yield prediction. But Rcal and Rpre were not enhanced or reduced significantly; i.e., the model performance obtained by different preprocessing methods was not significantly different. Among them, the WT model achieved optimal prediction performance for the calibration set and the prediction set, with higher Rcal and Rpre and lower RMSEC and RMSEP. Therefore, the UAV hyperspectral data were chosen to be analyzed using wavelet transform processed spectral data.

3.1.3. Analysis of Spectral Features

The spectral profile of the canopy of oilseed rape is shown in Figure 4. While the precise positions and sizes of the peaks and troughs exhibited specificity, the broad tendency of the spectral response patterns aligned with that of other vegetations, including tobacco [27], maize [44], and tomato [28]. Prominent peaks were observed at wavelengths of 560, 760, and 890 nm, while the primary troughs were found in proximity to 480 and 680 nm, respectively. The peaks and troughs were caused by chlorophyll absorption in the plant leaves around 560 nm and 680 nm. The peak change was related to water absorption around 760 nm and 890 nm, and the spectral reflectance was dramatically increased by the red-edge effect from 680 to 760 nm [45].
As shown in Figure 4b, the average spectral patterns of oilseed rape canopies exhibited considerable uniformity across varying nitrogen levels. Nevertheless, within the 750–1000 nm wavelength region, the spectral profiles showed visually distinct patterns, and the reflectivity of profiles associated with higher nitrogen fertilization was elevated compared to that from oilseed rape treated with lower nitrogen fertilization. The difference might be affected by oilseed rape pods (intercellular space and cell arrangement) and the canopy structure (such as oilseed rape leaf inclination, the leaf area index, and biomass) [46]. In the range of 900–950 nm, the curves were not perfect and the reflectance decreased faster, which may be related to the imaging method and the construction of the instrument itself. Therefore, we chose to maintain all the spectral bands for subsequent analysis to select the optimal effective wavelengths (EWs) and VIs.

3.2. Feature Wavelength Selection and Comparison

Hyperspectral data are highly inter-wavelength correlated and exist with information redundancy, which can cause the instability of convergence and complex computation and be time-consuming in multivariate prediction models. Using EWs can simplify models, reduce computational complexity, improve the predictive power of models, and facilitate more stable model output [45]. Therefore, selecting appropriate EWs for oilseed rape yield prediction was necessary. In this study, eight methods (SPA, GAPLS, UVE, a combination of UVE and SPA (UVE-SPA), BW, 2-Der, CARS, and RF) were used to reduce the dimensionality of the full-spectrum data to select the optimal EWs. Figure 5 and Figure 6 show the process of the selection of EWs using SPA and GAPLS.
Different feature wavelength selection methods selected the EWs according to different principles, and the selection results are presented in Table 3. Figure 7a shows the stability of each wavelength variable based on the UVE analysis. The wavelength variable and the random variable were located on the left and right sides, respectively. The two horizontal lines were the lower and upper limits of the maximum absolute value of the stable value in the random variable matrix. For the UVE algorithm, variables whose stability was inside the cutoff line were considered uninformative and eliminated. Finally, we picked 52 wavelength variables from 125 raw variables, which means that 58.4% of the wavelength variables were eliminated. Although the modelling variables were reduced by UVE analysis, the number of sensitive wavelengths was still large. To solve this problem, a combination of UVE and SPA was used for the selection of EWs. The EWs selected using SPA were ranked based on importance. The EWs selected using GA and RF were ranked according to the frequency or probability of variable selection, with the wavelengths ranked ahead being more important and selected frequently. In contrast, the EWs selected using 2-Der are selected based on the peaks and valleys of the second-order derivatives. Similarly, the EWs were selected using BW based on the peaks and troughs of the weighted regression coefficients of the PLSR algorithm. For the CARS method, Monte Carlo and adaptive weighted multiple sampling were used to select the EWs according to the wavelength corresponding to the subset with the lowest root mean square error in the PLSR algorithm.
As can be seen from Table 3, the number of wavelengths was reduced by more than 90% after the UVE-SPA or SPA treatment. It could effectively simplify the model and decrease the computational complexity. In contrast, the GA, CARS, and RF selected a larger number of EWs and contained more information, which would make the model more stable in the training phase. As a whole, the EWs selected were mainly related to photosynthetic capacity (495–680 nm), the red inflexion point (680–780 nm), plant water status (900–980 nm), etc. The EWs chosen around 450 nm and 550 nm may be associated with the absorption of anthocyanins [27]. Moreover, chlorophyll a and chlorophyll b had feature peaks at 675 nm and 620 nm, respectively, which can explain to some extent the EWs in this vicinity. At 700–750 nm, the wavelengths selected were associated with the red-edge index, etc. Labelling the shape of the red to near-infrared transition can indicate signs of canopy stress and senescence.

3.3. The Yield Prediction of Oilseed Rape Based on the Effective Wavelength

Using EWs and the full spectra as input variables, respectively, we constructed regression analysis models for the yield of oilseed rape based on the MLR, PLSR, ELM, and LS-SVM algorithms. The prediction results and the settings of model parameters are presented in Table 4. In the regression analysis model built using the full spectra, the PLSR, LS-SVM, and ELM models achieved good prediction performance, with Rcal and Rpre greater than 0.77. In particular, the ELM model performed the best, with Rcal and Rpre greater than 0.8. In the regression analysis model established based on the EWs selected using SPA, GAPLS, UVE-SPA, BW, CARS, and RF, the PLSR, MLR, LS-SVM, and ELM models all obtained well-prediction results, with Rcal and Rpre greater than 0.77, while the regression analysis models developed using UVE and 2-Der had the next-best prediction performance. Among them, the ELM prediction models established using UVE-SPA, CARS, and RF performed better, with Rcal and Rpre greater than 0.8. In general, the optimal prediction model was CARS-ELM, with Rcal and Rpre values of 0.8022 and 0.8122 and RMSEC and RMSEP values of 163.4 kg/hm2 and 170.4 kg/hm2, respectively, as shown in Figure 8.
In summary, the models using EWs with key information showed better robustness and performance than the models using full spectra. Among them, four feature band selection methods (GAPLS, CARS, RF, and UVE-SPA) showed better performances, and the ELM regression analysis model had better prediction. Employing diverse machine learning methodologies grounded in distinct principles, regression analyses were conducted on the datasets by examining them through varied lenses. From these analyses, the selection of a suitable regression modelling approach would prove advantageous for realizing oilseed rape yield prediction.

3.4. Optimal Narrow-Band Vegetation Index Selection and Analysis

In this study, we calculated different VIs and investigated the intrinsic link and correlation connecting oilseed rape yield and diverse VIs to identify the optimal VIs for estimating oilseed rape yield. For any two spectral bands (reflectance spectra and first-order derivative spectra), colour mapping of the absolute values of correlation coefficients between NDSI(Rλ1, Rλ2), RSI(Rλ1, Rλ2), DSI(Rλ1, Rλ2), SAVI(Rλ1, Rλ2), NDSI(Dλ1, Dλ2), RSI(Dλ1, Dλ2), and oilseed rape yield was plotted simultaneously to determine the optimal band combination for oilseed rape yield prediction. The calculation results for all combinations are shown in Figure 9. In the two-dimensional plot, the intersection points of each pair of bands corresponded to the absolute value of the correlation coefficient R between the VI and yield. The vegetation index was calculated using reflectance under this pair of bands, with redder-coloured points representing larger R and bluer representing smaller R. The most effective band combination and wavelength range for predicting field-based oilseed rape yield can be derived from the correlation coefficient plot.
For the vegetation index reflectance spectra NDSI(Rλ1, Rλ2) and RSI(Rλ1, Rλ2), the optimal correlation band combinations with the yield of oilseed rape were 550 nm and 582 nm, exhibiting absolute correlation coefficient values of 0.5493 and 0.5507, respectively. For the NDSI(Dλ1, Dλ2) of first-order derivatives, the best correlation band combinations were 522 nm and 650 nm, with a maximum absolute correlation coefficient value of 0.4943. As for the RSI(Dλ1, Dλ2) of first-order derivatives, Rmax = 0.5042, and the band combinations were 522 nm and 654 nm. The optimal band combinations obtained from DSI(Rλ1, Rλ2) and SAVI(Rλ1, Rλ2) were (670 nm, 710 nm) and (622 nm, 666 nm), and the highest absolute values of correlation coefficients were 0.5960 and 0.5356, respectively. Finally, we obtained six optimal VIs based on these isopotential maps, which are NDSI(550,582), RSI(550,582), NDSI(D522,D650), RSI(D522,D654), DSI(670,710), and SAVI(622,666).
Due to the limitation of the band range, except for the 6 narrow-band VIs mentioned above, 53 VIs were used for correlation analysis and one-way analysis of variance (ANOVA) with the yield of oilseed rape. These outcomes are presented in Tables S2 and S3 in the Supplementary Materials. While the correlation analysis ranking did not perfectly align with the ANOVA-based order, the overall outcomes showed resemblance, and vegetation indices with higher F values tended to be associated with stronger correlation coefficients. Drawing from the aforementioned analytical findings and with model parsimony in subsequent phases in mind, we chose nine VIs (|R| > 0.5) as inputs to the machine learning model, namely CRI1, CRI2, TCARI, CARI, TCARI/OSAVI, RVSI, TVI, MTVI1, and PRI512.

3.5. The Yield Prediction of Oilseed Rape Based on the Vegetation Index

Based on the correlation coefficient isopotential plots of narrow-band VIs and the results of the correlation analysis and ANOVA for other VIs, we selected 14 VIs with absolute values of correlation coefficients greater than 0.5 to establish the yield estimation model, namely NDSI(550, 582), RSI(550, 582), NDSI(D522, D650), RSI(D522, D654), DSI(670, 710), SAVI(622, 666), CRI1, CRI2, TCARI, CARI, TCARI/OSAVI, RVSI, TVI, MTVI1, and PRI512.
Using the combination of VIs (CRI1 with a linear fitting equation of y = −463.4x + 5569 (|R| = 0.7405, RMSE = 188.6 kg/hm2)) as the input, we established regression-based prediction models developed using a range of machine learning algorithms (PLSR, MLR, LS-SVM, and ELM), and the model parameters were configured in line with the specifications outlined earlier (see Table 4). The predicted results for the oilseed rape yield model derived from the optimal combination of VIs are presented in Table 5.
During the establishment of the regression model based on combining VIs, the PLSR, MLR, LS-SVM, and ELM models showed better prediction, with Rcal and Rpre greater than 0.74, and the PLSR achieved excellent predictive performance with Rpre up to 0.7736 and an RMSEP value of 185.4 kg/hm2. However, when combining the results of the calibration set and the prediction set, the ELM performed the best, with Rcal and Rpre values of 0.7964 and 0.7674 and RMSEC and RMSEP values of 165.5 kg/hm2 and 187.6 kg/hm2, respectively, as shown in Figure 10. Overall, the predictive models constructed based on the combined VIs predicted better than single VIs. The results show that fusing multiple VIs was more beneficial to constructing highly accurate and general yield prediction models, which can effectively promote the application of UAV remote sensing in the yield estimation of oilseed rape. Moreover, the performance of non-linear models (ELM and LS-SVM) was mostly better than that of linear models (MLR and PLSR) in the machine learning algorithms used.

4. Discussion

The yield estimation model for oilseed rape was developed based on raw spectra and different preprocessing spectral data. We evaluated the effects of different spectral pre-processing methods systematically (MAS, SG, SNV, MSC, SNV-Detrending, 1-Der,2-Der, and WT) on the spectral characteristics. Considering the performance of the model on the calibration and test sets, we concluded that WT was the optimal preprocessing method. The experimental results indicated that, compared with the raw spectral data, there was no significant difference in the prediction performance between the prediction models constructed based on different preprocessing methods (Table 2), but most of the noise and interference information in the raw spectral data could be effectively removed by using spectral preprocessing (Figure 4a,b). This phenomenon may be caused by the fact that the hyperspectral data still have highly tabulated redundancy or covariance after spectral preprocessing. The spectral data still contained a large amount of invalid information which was not relevant to the prediction of the yield of oilseed rape. Therefore, it is necessary to preprocess the spectral data and carry out feature analysis and extraction further.
To construct a yield estimation model for oilseed rape, we evaluated the effectiveness of eight wavelength selection methods (SPA, GAPLS, UVE, UVE-SPA, BW, 2-Der, CARS, and RF) for the extraction of EWs. In this context, effectiveness refers to the ability of the method to identify a subset of wavelengths that allows for the construction of simpler predictive models using fewer input variables. Based on the results of R and RMSE, the predictive performance is comparable to the models constructed using the full spectrum, thus successfully reducing the data dimensionality while preserving or enhancing the predictive power. As shown in Table 3, the characteristic wavelength variables were effectively reduced after wavelength selection. Compared to using full spectra, the regression analysis models using the EWs could achieve excellent predictions even with a significant reduction in the model input variables (Table 4). In addition, the prediction performance of the non-linear models was better than that of linear models. The results showed that it is feasible to select EWs containing key information by using wavelength selection algorithms and establish a quantitative analysis model based on EWs to achieve the yield prediction of oilseed rape.
To develop yield estimation models for oilseed rape based on VIs, we explored the intrinsic link and correlation between the oilseed rape yield and different VIs. Compared to the model prediction results based on single VIs and combined VIs (Table 5), we found that the prediction models based on combined VIs outperformed single VIs in terms of accuracy and stability. Here, accuracy is mainly measured by the R and RMSE between the model predictions and the measured rapeseed yield. As shown in Table 5, the ELM model using the combination of VIs obtained higher R values and lower RMSE values. Moreover, for the prediction models built by individual VIs, the prediction accuracy of the models was low, and there were significant differences in the fitted equations between the different sample sets. This may be caused by the fact that the model had fewer input feature variables, which affects the stability of the model. The research results indicated that fusing multiple VIs was more beneficial to constructing yield prediction models with high accuracy and generality.
The early prediction of crop yield relies heavily on obtaining the best spectral, spatial, and temporal signatures of yield by using appropriate platforms, sensors, and analytical methods. UAV low-altitude remote sensing technology with the advantages of low cost, high flexibility, high resolution, and easy operation provides a new strategy for the early prediction of crop yields at the field scale. We successfully achieved the early yield prediction of oilseed rape using a UAV remote sensing platform equipped with the UHD185 hyperspectral imaging system. Compared with existing international research, Yu et al. [17] collected large-scale soybean breeding trial multispectral images using drones and employed random forests to measure crop geometric features to improve soybean yield estimation, achieving a result of r = 0.82. Rajveer et al. [18] used machine learning regression models to predict corn and soybean yields in Ohio based on weather parameters, with R2 values of 0.73 for corn and 0.64 for soybeans. Hassan et al. [15] explained significant variations in biomass and grain yield through UAV-NDVI by utilizing vegetation indices related to yield and physiological traits. In conjunction with this study, it is suggested that future research utilizing canopy spectral information from UAV hyperspectral remote sensing, along with optimized EWs and VIs, can establish a quantitative analysis model for canola yield, providing technical support for yield prediction at the field scale.
However, we have encountered several limitations. Firstly, this study is mainly based on the hyperspectral remote sensing images of rapeseed at the angular fruiting stage. In the future, we will try to use hyperspectral remote sensing images of other growth periods of oilseed rape to explore the feasibility of early yield prediction of oilseed rape. Second, this study mainly used the spectral features of hyperspectral images to predict oilseed rape yield, but the image information of hyperspectral images was not fully utilized. In future research, we will continue to expand the number and type of samples, combine cutting-edge data mining methods such as deep learning to resolve the spectral image features of UAV low-altitude remote sensing, construct more robust crop yield prediction models, and provide technical support for the high-throughput acquisition of crop yield estimation and breeding information.

5. Conclusions

In this study, we focus on the early prediction of crop yield. Using the canopy spectrum of oilseed rape at the pod information stage acquired by the UAV hyperspectral remote sensing system, we conducted a series of studies about the yield prediction of oilseed rape at the field scale. Using the preferred EWs and VIs, a quantitative analytical model of rapeseed yield was established. We evaluated the effects of various spectral preprocessing algorithms: EW selection algorithms and optimal VIs on the prediction performance of the yield of oilseed rape. The following conclusions were obtained: (1) Compared with using the full spectral band, the estimation model constructed with EWs containing key information had good prediction effects and robustness. (2) The integration of the multi-vegetation index was more conducive to the construction of a yield prediction model with high accuracy and strong versatility. In subsequent studies, we will focus on the early prediction of the yield of oilseed rape based on time-series hyperspectral images to construct more accurate and versatile crop yield prediction models to provide technical support for the high-throughput acquisition of the yield estimation of oilseed rape and breeding information.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/agriculture15101100/s1, Table S1. The common vegetation indices; Table S2. The correlation coefficient between the canopy vegetation index and the yield of oilseed rape; Table S3. The results of ANOVA using vegetation indices. References [47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83] are cited in the supplementary materials.

Author Contributions

Conceptualization, H.Z.; methodology, H.Z.; funding acquisition, H.Z.; Project administration, H.Z.; supervision, H.Z., J.-L.X. and Y.H.; validation, H.Z. and C.L.; formal analysis, C.L.; writing—original draft, H.Z.; writing—review and editing, H.Z., C.L., Z.D., J.-L.X. and Y.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Guangxi (No. 2024 GXNSFBA010381), the National Natural Science Foundation of China (No. 62361006), the grant (No. NCOC-24-02) from Key Laboratory of Nonlinear Circuit and Optical Communications (Guangxi Normal University), Education Department of Guangxi Zhuang Autonomous Region, and the Guangxi Key Laboratory of Brain-inspired Computing and Intelligent Chips under grant (No. BCIC-24-Z1).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors upon request.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
UAVunmanned aerial vehicle
EWseffective wavelengths
VIsvegetation indices
MLRmultiple linear regression
PLSRpartial least squares regression
ELMextreme learning machine
LS-SVMleast squares support vector machine
CARS-ELMcompetitive adaptive reweighted sampling-extreme learning machine
DNdigital number
SGSavitzky–Golay smoothing
MASmoving average smoothing
SNVstandard normalized variate
MSCmultiplicative scatter correction
1-Derfirst derivative
2-Dersecond derivative
WTwavelet transform
SPAsuccessive projection algorithms
BWweighted regression coefficient
GAPLSgenetic algorithm partial least squares
GAgenetic algorithm
UVEuninformative variable elimination
CARScompetitive adaptive reweighted sampling
RFrandom frog
PLSpartial least squares
LVlatent variable
SVMsupport vector machine
Rcorrelation coefficient
RMSEroot mean square error
SNV-Detrendingstandard normalized variable detrending
CVcross-validation
ANOVAone-way analysis of variance

References

  1. Chen, X.; Lütken, H.; Liang, K.; Liu, F.; Favero, B.T. Superior osmotic stress tolerance in oilseed rape transformed with wild-type Rhizobium rhizogenes. Plant Cell Rep. 2024, 43, 1–12. [Google Scholar] [CrossRef] [PubMed]
  2. Kortesniemi, M.; Vuorinen, A.L.; Sinkkonen, J.; Yang, B.; Rajala, A.; Kallio, H. NMR metabolomics of ripened and developing oilseed rape (Brassica napus) and turnip rape (Brassica rapa). Food Chem. 2015, 172, 63–70. [Google Scholar] [CrossRef] [PubMed]
  3. Msanne, J.; Kim, H.; Cahoon, E.B. Biotechnology tools and applications for development of oilseed crops with healthy vegetable oils. Biochimie 2020, 178, 4–14. [Google Scholar] [CrossRef]
  4. Yong, K.J.; Wu, T.Y. Second-generation bioenergy from oilseed crop residues: Recent technologies, techno-economic assessments and policies. Energy Convers. Manag. 2022, 267, 115869. [Google Scholar] [CrossRef]
  5. Maes, W.H.; Steppe, K. Perspectives for Remote Sensing with Unmanned Aerial Vehicles in Precision Agriculture. Trends Plant Sci. 2019, 24, 152–164. [Google Scholar] [CrossRef]
  6. Overgaard, S.I.; Isaksson, T.; Kvaal, K.; Korsaeth, A. Comparisons of two hand-held, multispectral field radiometers and a hyperspectral airborne imager in terms of predicting spring wheat grain yield and quality by means of powered partial least squares regression. J. Near Infrared Spectrosc. 2010, 18, 247–261. [Google Scholar] [CrossRef]
  7. Swain, K.C.; Thomson, S.J.; Jayasuriya, H.P. Adoption of an unmanned helicopter for low-altitude remote sensing to estimate yield and total biomass of a rice crop. Trans. ASABE 2010, 53, 21–27. [Google Scholar] [CrossRef]
  8. Sunoj, S.; Yeh, B.; Marcaida, M., III; Longchamps, L.; van Aardt, J.; Ketterings, Q.M. Maize grain and silage yield prediction of commercial fields using high-resolution UAS imagery. Biosyst. Eng. 2023, 235, 137–149. [Google Scholar] [CrossRef]
  9. Verger, A.; Vigneau, N.; Cheron, C.; Gilliot, J.M.; Comar, A.; Baret, F. Green area index from an unmanned aerial system over wheat and rapeseed crops. Remote Sens. Environ. 2014, 152, 654–664. [Google Scholar] [CrossRef]
  10. Zhang, C.H.; Kovacs, J.M. The application of small unmanned aerial systems for precision agriculture: A review. Precis. Agric. 2012, 13, 693–712. [Google Scholar] [CrossRef]
  11. Ninomiya, S. High-throughput field crop phenotyping: Current status and challenges. Breed. Sci. 2022, 72, 3–18. [Google Scholar] [CrossRef] [PubMed]
  12. Araus, J.L.; Cairns, J.E. Field high-throughput phenotyping: The new crop breeding frontier. Trends Plant Sci. 2014, 19, 52–61. [Google Scholar] [CrossRef]
  13. López-García, P.; Ortega, J.F.; Pérez-Álvarez, E.P.; Moreno, M.A.; Ramírez, J.M.; Intrigliolo, D.S.; Ballesteros, R. Yield estimations in a vineyard based on high-resolution spatial imagery acquired by a UAV. Biosyst. Eng. 2022, 224, 227–245. [Google Scholar] [CrossRef]
  14. Adao, T.; Hruska, J.; Padua, L.; Bessa, J.; Peres, E.; Morais, R.; Sousa, J.J. Hyperspectral Imaging: A Review on UAV-Based Sensors, Data Processing and Applications for Agriculture and Forestry. Remote Sens. 2017, 9, 110. [Google Scholar] [CrossRef]
  15. Hassan, M.A.; Yang, M.; Rasheed, A.; Yang, G.; Reynolds, M.; Xia, X.; He, Z. A rapid monitoring of NDVI across the wheat growth cycle for grain yield prediction using a multi-spectral UAV platform. Plant Sci. 2019, 282, 95–103. [Google Scholar] [CrossRef]
  16. Kyratzis, A.C.; Skarlatos, D.P.; Menexes, G.C.; Vamvakousis, V.F.; Katsiotis, A. Assessment of Vegetation Indices Derived by UAV Imagery for Durum Wheat Phenotyping under a Water Limited and Heat Stressed Mediterranean Environment. Front. Plant Sci. 2017, 8, 1114. [Google Scholar] [CrossRef]
  17. Yu, N.; Li, L.J.; Schmitz, N.; Tiaz, L.F.; Greenberg, J.A.; Diers, B.W. Development of methods to improve soybean yield estimation and predict plant maturity with an unmanned aerial vehicle based platform. Remote Sens. Environ. 2016, 187, 91–101. [Google Scholar] [CrossRef]
  18. Dhillon, R.; Takoo, G.; Sharma, V.; Nagle, M. Utilizing Machine Learning Framework to Evaluate the Effect of Climate Change on Maize and Soybean Yield. Comput. Electron. Agric. 2024, 221, 108982. [Google Scholar] [CrossRef]
  19. Tmušić, G.; Manfreda, S.; Aasen, H.; James, M.R.; Gonçalves, G.; Ben-Dor, E.; Brook, A.; Polinova, M.; Arranz, J.J.; Mészáros, J.; et al. Current Practices in UAS-based Environmental Monitoring. Remote Sens. 2020, 12, 1001. [Google Scholar] [CrossRef]
  20. Bazzo, C.O.G.; Kamali, B.; Hütt, C.; Bareth, G.; Gaiser, T. A review of estimation methods for aboveground biomass in grasslands using UAV. Remote Sens. 2023, 15, 639. [Google Scholar] [CrossRef]
  21. Savitzky, A.; Golay, M.J. Smoothing and differentiation of data by simplified least squares procedures. Anal. Chem. 1964, 36, 1627–1639. [Google Scholar] [CrossRef]
  22. Barnes, R.; Dhanoa, M.S.; Lister, S.J. Standard normal variate transformation and de-trending of near-infrared diffuse reflectance spectra. Appl. Spectrosc. 1989, 43, 772–777. [Google Scholar] [CrossRef]
  23. Helland, I.S.; Næs, T.; Isaksson, T. Related versions of the multiplicative scatter correction method for preprocessing spectroscopic data. Chemom. Intell. Lab. Syst. 1995, 29, 233–241. [Google Scholar] [CrossRef]
  24. Norris, K.; Williams, P. Optimization of mathematical treatments of raw near-infrared signal in the. Cereal Chem. 1984, 61, 158–165. [Google Scholar]
  25. Subasi, A.; Ahmed, A.; Alickovic, E.; Hassan, A.R. Effect of photic stimulation for migraine detection using random forest and discrete wavelet transform. Biomed. Signal Process. Control 2019, 49, 231–239. [Google Scholar] [CrossRef]
  26. Zhu, H.; Chu, B.; Fan, Y.; Tao, X.; Yin, W.; He, Y. Hyperspectral imaging for predicting the internal quality of kiwifruits based on variable selection algorithms and chemometric models. Sci. Rep. 2017, 7, 7845. [Google Scholar] [CrossRef]
  27. Zhu, H.Y.; Chu, B.Q.; Zhang, C.; Liu, F.; Jiang, L.J.; He, Y. Hyperspectral Imaging for Presymptomatic Detection of Tobacco Disease with Successive Projections Algorithm and Machine-learning Classifiers. Sci. Rep. 2017, 7, 4125. [Google Scholar] [CrossRef]
  28. Xie, C.; Shao, Y.; Li, X.; He, Y. Detection of early blight and late blight diseases on tomato leaves using hyperspectral imaging. Sci. Rep. 2015, 5, 16564. [Google Scholar] [CrossRef]
  29. ElMasry, G.; Sun, D.W.; Allen, P. Near-infrared hyperspectral imaging for predicting colour, pH and tenderness of fresh beef. J. Food Eng. 2012, 110, 127–140. [Google Scholar] [CrossRef]
  30. Kamruzzaman, M.; ElMasry, G.; Sun, D.W.; Allen, P. Prediction of some quality attributes of lamb meat using near-infrared hyperspectral imaging and multivariate analysis. Anal. Chim. Acta 2012, 714, 57–67. [Google Scholar] [CrossRef]
  31. Leardi, R. Application of genetic algorithm–PLS for feature selection in spectral data sets. J. Chemom. 2000, 14, 643–655. [Google Scholar] [CrossRef]
  32. Han, Q.J.; Wu, H.L.; Cai, C.B.; Xu, L.; Yu, R.Q. An ensemble of Monte Carlo uninformative variable elimination for wavelength selection. Anal. Chim. Acta 2008, 612, 121–125. [Google Scholar] [CrossRef] [PubMed]
  33. Ye, S.; Wang, D.; Min, S. Successive projections algorithm combined with uninformative variable elimination for spectral variable selection. Chemom. Intell. Lab. Syst. 2008, 91, 194–199. [Google Scholar] [CrossRef]
  34. Li, H.; Liang, Y.; Xu, Q.; Cao, D. Key wavelengths screening using competitive adaptive reweighted sampling method for multivariate calibration. Anal. Chim. Acta 2009, 648, 77–84. [Google Scholar] [CrossRef]
  35. Li, H.D.; Xu, Q.S.; Liang, Y.Z. Random frog: An efficient reversible jump Markov Chain Monte Carlo-like approach for variable selection with applications to gene selection and disease classification. Anal. Chim. Acta 2012, 740, 20–26. [Google Scholar] [CrossRef]
  36. Zarco-Tejada, P.J.; Gonzalez-Dugo, V.; Berni, J.A.J. Fluorescence, temperature and narrow-band indices acquired from a UAV platform for water stress detection using a micro-hyperspectral imager and a thermal camera. Remote Sens. Environ. 2012, 117, 322–337. [Google Scholar] [CrossRef]
  37. Ma, L.; Li, M.C.; Ma, X.X.; Cheng, L.; Du, P.J.; Liu, Y.X. A review of supervised object-based land-cover image classification. Isprs J. Photogramm. Remote Sens. 2017, 130, 277–293. [Google Scholar] [CrossRef]
  38. Singh, A.; Ganapathysubramanian, B.; Singh, A.K.; Sarkar, S. Machine Learning for High-Throughput Stress Phenotyping in Plants. Trends Plant Sci. 2016, 21, 110–124. [Google Scholar] [CrossRef]
  39. Wu, D.; Sun, D.W. Advanced applications of hyperspectral imaging technology for food quality and safety analysis and assessment: A review—Part I: Fundamentals. Innov. Food Sci. Emerg. Technol. 2013, 19, 1–14. [Google Scholar] [CrossRef]
  40. Ding, S.F.; Zhao, H.; Zhang, Y.N.; Xu, X.Z.; Nie, R. Extreme learning machine: Algorithm, theory and applications. Artif. Intell. Rev. 2015, 44, 103–115. [Google Scholar] [CrossRef]
  41. Wang, J.; Lu, S.; Wang, S.H.; Zhang, Y.D. A review on extreme learning machine. Multimed. Tools Appl. 2022, 81, 41611–41660. [Google Scholar] [CrossRef]
  42. Raccuglia, P.; Elbert, K.C.; Adler, P.D.; Falk, C.; Wenny, M.B.; Mollo, A.; Norquist, A.J. Machine-learning-assisted materials discovery using failed experiments. Nature 2016, 533, 73–76. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, F.; O’Donnell, L.J. Support vector regression. In Machine Learning; Academic Press: Cambridge, MA, USA, 2020; pp. 123–140. [Google Scholar] [CrossRef]
  44. Weber, V.S.; Araus, J.L.; Cairns, J.E.; Sanchez, C.; Melchinger, A.E.; Orsini, E. Prediction of grain yield using reflectance spectra of canopy and leaves in maize plants grown under different water regimes. Field Crops Res. 2012, 128, 82–90. [Google Scholar] [CrossRef]
  45. Liu, D.; Sun, D.W.; Zeng, X.A. Recent Advances in Wavelength Selection Techniques for Hyperspectral Image Processing in the Food Industry. Food Bioprocess Technol. 2014, 7, 307–323. [Google Scholar] [CrossRef]
  46. Hansen, P.M.; Schjoerring, J.K. Reflectance measurement of canopy biomass and nitrogen status in wheat crops using normalized difference vegetation indices and partial least squares regression. Remote Sens. Environ. 2003, 86, 542–553. [Google Scholar] [CrossRef]
  47. Blackburn, G.A. Spectral indices for estimating photosynthetic pigment concentrations: A test using senescent tree leaves. Int. J. Remote Sens. 1998, 19, 657–675. [Google Scholar] [CrossRef]
  48. Broge, N.H.; Leblanc, E. Comparing prediction power and stability of broadband and hyperspectral vegetation indices for estimation of green leaf area index and canopy chlorophyll density. Remote Sens. Environ. 2001, 76, 156–172. [Google Scholar] [CrossRef]
  49. Carter, G.A. Ratios of leaf reflectances in narrow wavebands as indicators of plant stress. Remote Sens. 1994, 15, 697–703. [Google Scholar] [CrossRef]
  50. Chen, J.M. Evaluation of vegetation indices and a modified simple ratio for boreal applications. Can. J. Remote Sens. 1996, 22, 229–242. [Google Scholar] [CrossRef]
  51. Dash, J.; Curran, P.J. The MERIS terrestrial chlorophyll index. Int. J. Remote Sens. 2004, 25, 5403–5413. [Google Scholar] [CrossRef]
  52. Datt, B. Remote sensing of water content in Eucalyptus leaves. Aust. J. Bot. 1999, 47, 909–923. [Google Scholar] [CrossRef]
  53. Daughtry, C.S.; Walthall, C.L.; Kim, M.S.; De Colstoun, E.B.; McMurtrey Iii, J.E. Estimating corn leaf chlorophyll concentration from leaf and canopy reflectance. Remote Sens. Environ. 2000, 74, 229–239. [Google Scholar] [CrossRef]
  54. Diker, K.; Bausch, W.C. Potential use of nitrogen reflectance index to estimate plant parameters and yield of maize. Biosyst. Eng. 2003, 85, 437–447. [Google Scholar] [CrossRef]
  55. Gamon, J.A.; Penuelas, J.; Field, C.B. A narrow-waveband spectral index that tracks diurnal changes in photosynthetic efficiency. Remote Sens. Environ. 1992, 41, 35–44. [Google Scholar] [CrossRef]
  56. 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]
  57. Gitelson, A.A.; Zur, Y.; Chivkunova, O.B.; Merzlyak, M.N. Assessing carotenoid content in plant leaves with reflectance spectroscopy. Photochem. Photobiol. 2002, 75, 272–281. [Google Scholar] [CrossRef]
  58. Haboudane, D.; Miller, J.R.; Tremblay, N.; Zarco-Tejada, P.J.; Dextraze, L. Integrated narrow-band vegetation indices for prediction of crop chlorophyll content for application to precision agriculture. Remote Sens. Environ. 2002, 81, 416–426. [Google Scholar] [CrossRef]
  59. Haboudane, D.; Miller, J.R.; Pattey, E.; Zarco-Tejada, P.J.; Strachan, I.B. Hyperspectral vegetation indices and novel algorithms for predicting green LAI of crop canopies: Modeling and validation in the context of precision agriculture. Remote Sens. Environ. 2004, 90, 337–352. [Google Scholar] [CrossRef]
  60. Hernández-Clemente, R.; Navarro-Cerrillo, R.M.; Suárez, L.; Morales, F.; Zarco-Tejada, P.J. Assessing structural effects on PRI for stress detection in conifer forests. Remote Sens. Environ. 2011, 115, 2360–2375. [Google Scholar] [CrossRef]
  61. Huete, A.R. A soil-adjusted vegetation index (SAVI). Remote Sens. Environ. 1988, 25, 295–309. [Google Scholar] [CrossRef]
  62. Huete, A.; Didan, K.; Miura, T.; Rodriguez, E.P.; Gao, X.; Ferreira, L.G. Overview of the radiometric and biophysical performance of the MODIS vegetation indices. Remote Sens. Environ. 2002, 83, 195–213. [Google Scholar] [CrossRef]
  63. Inoue, Y.; Sakaiya, E.; Zhu, Y.; Takahashi, W. Diagnostic mapping of canopy nitrogen content in rice based on hyperspectral measurements. Remote Sens. Environ. 2012, 126, 210–221. [Google Scholar] [CrossRef]
  64. Jiang, Z.; Huete, A.R.; Didan, K.; Miura, T. Development of a two-band enhanced vegetation index without a blue band. Remote Sens. Environ. 2008, 112, 3833–3845. [Google Scholar] [CrossRef]
  65. Jordan, C.F. Derivation of leaf-area index from quality of light on the forest floor. Ecology 1969, 50, 663–666. [Google Scholar] [CrossRef]
  66. Kaufman, Y.J.; Tanre, D. Strategy for direct and indirect methods for correcting the aerosol effect on remote sensing: From AVHRR to EOS-MODIS. Remote Sens. Environ. 1996, 55, 65–79. [Google Scholar] [CrossRef]
  67. Kim, M.S.; Daughtry, C.; Chappelle, E.; McMurtrey, J.; Walthall, C. The use of high spectral resolution bands for estimating absorbed photosynthetically active radiation (A par). In CNES, Proceedings of the 6th International Symposium on Physical Measurements and Signatures in Remote Sensing, Val D’Isere, France, 1 January 1994; NASA: Washington, DC, USA, 1994; No. GSFC-E-DAA-TN72921. [Google Scholar]
  68. Maccioni, A.; Agati, G.; Mazzinghi, P. New vegetation indices for remote measurement of chlorophylls based on leaf directional reflectance spectra. J. Photochem. Photobiol. B Biol. 2001, 61, 52–61. [Google Scholar] [CrossRef]
  69. Mahlein, A.K.; Rumpf, T.; Welke, P.; Dehne, H.W.; Plümer, L.; Steiner, U.; Oerke, E.C. Development of spectral indices for detecting and identifying plant diseases. Remote Sens. Environ. 2013, 128, 21–30. [Google Scholar] [CrossRef]
  70. Merton, R. Monitoring community hysteresis using spectral shift analysis and the red-edge vegetation stress index. In Proceedings of the Seventh Annual JPL Airborne Earth Science Workshop, Pasadena, CA, USA, 12–16 January 1998; pp. 12–16. [Google Scholar]
  71. Pearson, R.L.; Miller, L.D. Remote mapping of standing crop biomass for estimation of the productivity of the shortgrass prairie, Pawnee National Grasslands, Colorado. In Proceedings of the Eighth International Symposium on Remote Sensing of Environment, Ann Arbor, MI, USA, 2–6 October 1972. [Google Scholar]
  72. Peñuelas, J.; Gamon, J.; Fredeen, A.; Merino, J.; Field, C. Reflectance indices associated with physiological changes in nitrogen-and water-limited sunflower leaves. Remote Sens. Environ. 1994, 48, 135–146. [Google Scholar] [CrossRef]
  73. Peñuelas, J.; Filella, I.; Lloret, P.; Mun Oz, F.; Vilajeliu, M. Reflectance assessment of mite effects on apple trees. Int. J. Remote Sens. 1995, 16, 2727–2733. [Google Scholar] [CrossRef]
  74. Qi, J.; Chehbouni, A.; Huete, A.R.; Kerr, Y.H.; Sorooshian, S. A modified soil adjusted vegetation index. Remote Sens. Environ. 1994, 48, 119–126. [Google Scholar] [CrossRef]
  75. Rouse, J.W.; Haas, R.H.; Schell, J.A.; Deering, D.W. Monitoring vegetation systems in the Great Plains with ERTS. NASA Spec. Publ 1974, 351, 309. [Google Scholar]
  76. Rondeaux, G.; Steven, M.; Baret, F. Optimization of soil-adjusted vegetation indices. Remote Sens. Environ. 1996, 55, 95–107. [Google Scholar] [CrossRef]
  77. Roujean, J.-L.; Breon, F.-M. Estimating PAR absorbed by vegetation from bidirectional reflectance measurements. Remote Sens. Environ. 1995, 51, 375–384. [Google Scholar] [CrossRef]
  78. Sims, D.A.; Gamon, J.A. Relationships between leaf pigment content and spectral reflectance across a wide range of species, leaf structures and developmental stages. Remote Sens. Environ. 2002, 81, 337–354. [Google Scholar] [CrossRef]
  79. Vogelmann, J.E.; Rock, B.N.; Moss, D.M. Red edge spectral measurements from sugar maple leaves. Remote Sens. 1993, 14, 1563–1575. [Google Scholar] [CrossRef]
  80. Wu, C.; Niu, Z.; Tang, Q.; Huang, W. Estimating chlorophyll content from hyperspectral vegetation indices: Modeling and validation. Agric. For. Meteorol. 2008, 148, 1230–1241. [Google Scholar] [CrossRef]
  81. Zarco-Tejada, P.J.; Miller, J.R.; Noland, T.L.; Mohammed, G.H.; Sampson, P.H. Scaling-up and model inversion methods with narrowband optical indices for chlorophyll content estimation in closed forest canopies with hyperspectral data. IEEE Trans. Geosci. Remote Sens. 2001, 39, 1491–1507. [Google Scholar] [CrossRef]
  82. Zarco-Tejada, P.J.; Berjón, A.; Lopez-Lozano, R.; Miller, J.R.; Martín, P.; Cachorro, V.; De Frutos, A. Assessing vineyard condition with hyperspectral indices: Leaf and canopy reflectance simulation in a row-structured discontinuous canopy. Remote Sens. Environ. 2005, 99, 271–287. [Google Scholar] [CrossRef]
  83. Zarco-Tejada, P.J.; Berni, J.A.; Suárez, L.; Sepulcre-Cantó, G.; Morales, F.; Miller, J.R. Imaging chlorophyll fluorescence with an airborne narrow-band multispectral camera for vegetation stress detection. Remote Sens. Environ. 2009, 113, 1262–1275. [Google Scholar] [CrossRef]
Figure 1. The UHD185 UAV hyperspectral remote sensing system and an overview map of the study area. (a) The six-rotor UAV and (b) the UHD185 hyperspectral imager. (c) The overview map.
Figure 1. The UHD185 UAV hyperspectral remote sensing system and an overview map of the study area. (a) The six-rotor UAV and (b) the UHD185 hyperspectral imager. (c) The overview map.
Agriculture 15 01100 g001
Figure 2. The flow chart of yield prediction of oilseed rape based on the UAV hyperspectral remote sensing system. MAS: moving average smoothing; SG: Savitzky–Golay smoothing; MSC: multiplicative scatter correction; SNV: standard normalized variate; 1-Der: first derivative; 2-Der: second derivative; WT: wavelet transform; SPAs: successive projection algorithms; GAPLS: genetic algorithm partial least squares; UVE: uninformative variable elimination; BW: weighted regression coefficient; CARS: competitive adaptive reweighted sampling; RF: random frog; MLR: multiple linear regression; PLSR: partial least squares regression; ELM: extreme learning machine; LS-SVM: least squares support vector machine.
Figure 2. The flow chart of yield prediction of oilseed rape based on the UAV hyperspectral remote sensing system. MAS: moving average smoothing; SG: Savitzky–Golay smoothing; MSC: multiplicative scatter correction; SNV: standard normalized variate; 1-Der: first derivative; 2-Der: second derivative; WT: wavelet transform; SPAs: successive projection algorithms; GAPLS: genetic algorithm partial least squares; UVE: uninformative variable elimination; BW: weighted regression coefficient; CARS: competitive adaptive reweighted sampling; RF: random frog; MLR: multiple linear regression; PLSR: partial least squares regression; ELM: extreme learning machine; LS-SVM: least squares support vector machine.
Agriculture 15 01100 g002
Figure 3. The image processing flow of UHD185 UAV hyperspectral remote sensing.
Figure 3. The image processing flow of UHD185 UAV hyperspectral remote sensing.
Agriculture 15 01100 g003
Figure 4. (a) The spectra of oilseed rape canopy after wavelet transform; (b) the average reflectance spectra and standard deviation (SD) of ‘0 kg/hm2’, ‘75 kg/hm2’, ‘150 kg/hm2’, and ‘225 kg/hm2’. The colored lines in Figure 4a represent the canopy spectral curves of individual plots in the experiment.
Figure 4. (a) The spectra of oilseed rape canopy after wavelet transform; (b) the average reflectance spectra and standard deviation (SD) of ‘0 kg/hm2’, ‘75 kg/hm2’, ‘150 kg/hm2’, and ‘225 kg/hm2’. The colored lines in Figure 4a represent the canopy spectral curves of individual plots in the experiment.
Agriculture 15 01100 g004
Figure 5. Selection of EWs using SPA. (a) Final number of selected variables according to the minimum RMSEV in the validation set of MLR calibration; (b) EWs shown in box markers.
Figure 5. Selection of EWs using SPA. (a) Final number of selected variables according to the minimum RMSEV in the validation set of MLR calibration; (b) EWs shown in box markers.
Agriculture 15 01100 g005
Figure 6. Selection of EWs using GAPLS. (a) Smoothed frequency of selections; (b) final number of selected variables according to the minimum RMSE cross-validation.
Figure 6. Selection of EWs using GAPLS. (a) Smoothed frequency of selections; (b) final number of selected variables according to the minimum RMSE cross-validation.
Agriculture 15 01100 g006
Figure 7. Selection of EWs. (a) The stability of each wavelength variable based on UVE analysis, (b) the combination of UVE and SPA, (c) the selection of EWs using BW, (d) the selection of EWs using 2−De, (e) the selection of EWs using CARS, (f) the selection of EWs using 2−Der.
Figure 7. Selection of EWs. (a) The stability of each wavelength variable based on UVE analysis, (b) the combination of UVE and SPA, (c) the selection of EWs using BW, (d) the selection of EWs using 2−De, (e) the selection of EWs using CARS, (f) the selection of EWs using 2−Der.
Agriculture 15 01100 g007
Figure 8. Performance of the best prediction model using the EWs for yield prediction.
Figure 8. Performance of the best prediction model using the EWs for yield prediction.
Agriculture 15 01100 g008
Figure 9. Correlation coefficient contour plots for yield and NDSI(Rλ1,Rλ2), NDSI(Dλ1,Dλ2), RSI(Rλ1,Rλ2), RSI(Dλ1,Dλ2), DSI(Rλ1,Rλ2), and SAVI(Rλ1,Rλ2), generated using all combinations of two wavebands at λ1 and λ2 nm.
Figure 9. Correlation coefficient contour plots for yield and NDSI(Rλ1,Rλ2), NDSI(Dλ1,Dλ2), RSI(Rλ1,Rλ2), RSI(Dλ1,Dλ2), DSI(Rλ1,Rλ2), and SAVI(Rλ1,Rλ2), generated using all combinations of two wavebands at λ1 and λ2 nm.
Agriculture 15 01100 g009
Figure 10. Performance of the best prediction model using the combination of VIs for yield prediction.
Figure 10. Performance of the best prediction model using the combination of VIs for yield prediction.
Agriculture 15 01100 g010
Table 1. Statistics of the measured yield values of oilseed rape under different nitrogen gradients using standard methods.
Table 1. Statistics of the measured yield values of oilseed rape under different nitrogen gradients using standard methods.
Sample SetsNumberMinimum (kg/hm2)Maximum (kg/hm2)Mean (kg/hm2)Median (kg/hm2)Standard Deviation (SD)Coefficient of Variation (%)KurtosisSkewness
Calibration2482217.33586.32942.682977.70274.239.31−0.41−0.37
Prediction1252217.33526.82947.243023.80292.979.94−0.20−0.53
Table 2. The comparison of different pretreatment methods based on PLSR for the yield prediction of oilseed rape.
Table 2. The comparison of different pretreatment methods based on PLSR for the yield prediction of oilseed rape.
Preprocessing MethodModelsLVsCalibrationPrediction
RcalRMSECRpreRMSEP
-PLSR40.7771172.30.7770183.8
MASPLSR40.7750173.00.7733207.4
SGPLSR40.7768172.40.7771183.8
MSCPLSR50.7178190.50.6365225.5
SNVPLSR90.7182190.40.6384225.1
SNV-DetrendingPLSR90.7141191.60.6122231.5
1-DerPLSR50.7144191.50.6722216.1
2-DerPLSR20.6430209.60.5026257.1
WTPLSR40.7767172.40.7775183.6
Table 3. The selected EWs for yield prediction using different wavelength selection algorithms.
Table 3. The selected EWs for yield prediction using different wavelength selection algorithms.
Algorithms(No.)Selected EWs (nm)
SPA7858, 474, 746, 638, 510, 554, and 910
GAPLS32918, 514, 922, 934, 518, 510, 914, 938, 542, 930, 738, 734, 926, 538, 858, 546, 730, 742, 862, 782, 474, 778, 854, 470, 786, 942, 698, 478, 534, 702, 506, and 746
UVE52486–522, 582–754, 914–922, and 938
UVE-SPA7730, 494, 486, 650, 914, 754, and 938
BW12450, 466, 498, 542, 622, 638, 646, 734, 862, 898, 922, and 946
2-Der13506, 526, 538, 550, 574, 606, 618, 634, 686, 734, 770, 894, and 934
CARS8478, 498, 506, 542, 598, 914, 918, and 922
RF29510, 514, 474, 478, 506, 542, 518, 502, 642, 538, 498, 522, 494, 570, 638, 566, 918, 546, 666, 670, 646, 470, 738, 742, 590, 574, 634, 934, and 782
Table 4. The full spectra and the selected EWs for yield prediction using different machine learning algorithms.
Table 4. The full spectra and the selected EWs for yield prediction using different machine learning algorithms.
Wavelength Selection AlgorithmsEWsModelsLVs/(γ, σ2)/
Hidden Neurons
CalibrationPrediction
RcalRMSECRpreRMSEP
-125PLSR40.7767172.40.7775183.6
-125MLR-0.8988140.10.6588247.2
-125LS-SVM(76.4, 5.6 × 103)0.8012163.90.7703186.1
-125ELM480.8221155.80.8019175.2
SPA7PLSR40.7760172.60.7787183.2
SPA7MLR-0.7805171.10.7919179.0
SPA7LS-SVM(127.0, 392.1)0.7894168.00.7783183.2
SPA7ELM200.7963165.50.8056173.1
GAPLS32PLSR40.7803171.10.7824181.9
GAPLS32MLR-0.7999164.30.7846181.0
GAPLS32LS-SVM(6.4 × 103, 3.6 × 104)0.7893168.10.7888179.4
GAPLS32ELM390.8100160.50.8060172.9
UVE52PLSR40.7793171.50.7803182.6
UVE52MLR-0.8459147.30.7037214.0
UVE52LS-SVM(347.2, 2.7 × 103)0.8006164.00.7464194.4
UVE52ELM470.8281153.40.7888180.2
UVE-SPA7PLSR40.7756172.80.7773183.8
UVE-SPA7MLR-0.7816170.70.7858180.6
UVE-SPA7LS-SVM(9.4 × 103, 5.1 × 103)0.7895168.00.7881179.7
UVE-SPA7ELM290.8069161.70.8050173.4
BW12PLSR40.7798171.70.7750184.5
BW12MLR-0.7853169.80.7707186.1
BW12LS-SVM(2.2 × 103, 5.3 × 103)0.7914167.30.7798182.7
BW12ELM310.8101160.50.7980176.0
2-Der13PLSR40.7705174.50.7694186.5
2-Der13MLR-0.7810170.90.7594189.9
2-Der13LS-SVM(44.5, 279.1)0.7916167.30.7453194.6
2-Der13ELM80.7734173.50.7879179.9
CARS8PLSR40.7750173.00.7853180.8
CARS8MLR-0.7844169.80.7979176.0
CARS8LS-SVM(381.5, 333.1)0.7952166.00.7737184.9
CARS8ELM210.8022163.40.8122170.3
RF29PLSR40.7792171.50.7906179.0
RF29MLR-0.8136160.50.7743186.2
RF29LS-SVM(2.1 × 103, 2.8 × 103)0.8078161.40.7746184.6
RF29ELM390.8224155.70.8018175.0
Note: number of LVs for PLSR, (γ, σ2) for LS-SVM, number of hidden neurons for ELM.
Table 5. Vegetation index combinations for yield prediction across various machine learning algorithms.
Table 5. Vegetation index combinations for yield prediction across various machine learning algorithms.
(No.)ModelsLVs/(γ, σ2)/
Hidden Neurons
CalibrationPrediction
RcalRMSECRpreRMSEP
14PLSR20.7605177.70.7736185.4
14MLR-0.7813170.80.7670187.4
14LS-SVM(0.80, 72.2)0.7952166.30.7493193.5
14ELM270.7964165.50.7674187.6
Note: parameters and abbreviations are shown in Table 4.
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

Zhu, H.; Lin, C.; Dong, Z.; Xu, J.-L.; He, Y. Early Yield Prediction of Oilseed Rape Using UAV-Based Hyperspectral Imaging Combined with Machine Learning Algorithms. Agriculture 2025, 15, 1100. https://doi.org/10.3390/agriculture15101100

AMA Style

Zhu H, Lin C, Dong Z, Xu J-L, He Y. Early Yield Prediction of Oilseed Rape Using UAV-Based Hyperspectral Imaging Combined with Machine Learning Algorithms. Agriculture. 2025; 15(10):1100. https://doi.org/10.3390/agriculture15101100

Chicago/Turabian Style

Zhu, Hongyan, Chengzhi Lin, Zhihao Dong, Jun-Li Xu, and Yong He. 2025. "Early Yield Prediction of Oilseed Rape Using UAV-Based Hyperspectral Imaging Combined with Machine Learning Algorithms" Agriculture 15, no. 10: 1100. https://doi.org/10.3390/agriculture15101100

APA Style

Zhu, H., Lin, C., Dong, Z., Xu, J.-L., & He, Y. (2025). Early Yield Prediction of Oilseed Rape Using UAV-Based Hyperspectral Imaging Combined with Machine Learning Algorithms. Agriculture, 15(10), 1100. https://doi.org/10.3390/agriculture15101100

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