Next Article in Journal
Camera Calibration for Coastal Monitoring Using Available Snapshot Images
Previous Article in Journal
The Moon as a Climate-Quality Radiometric Calibration Reference
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Wheat Lodging Detection from UAS Imagery Using Machine Learning Algorithms

1
Department of Agricultural and Biosystems Engineering, North Dakota State University, Fargo, ND 58102, USA
2
Department of Civil & Environmental Engineering, North Dakota State University, Fargo, ND 58105, USA
3
Department of Plant Sciences, North Dakota State University, Fargo, ND 58108, USA
*
Author to whom correspondence should be addressed.
Remote Sens. 2020, 12(11), 1838; https://0-doi-org.brum.beds.ac.uk/10.3390/rs12111838
Submission received: 27 April 2020 / Revised: 3 June 2020 / Accepted: 4 June 2020 / Published: 5 June 2020
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)

Abstract

:
The current mainstream approach of using manual measurements and visual inspections for crop lodging detection is inefficient, time-consuming, and subjective. An innovative method for wheat lodging detection that can overcome or alleviate these shortcomings would be welcomed. This study proposed a systematic approach for wheat lodging detection in research plots (372 experimental plots), which consisted of using unmanned aerial systems (UAS) for aerial imagery acquisition, manual field evaluation, and machine learning algorithms to detect the occurrence or not of lodging. UAS imagery was collected on three different dates (23 and 30 July 2019, and 8 August 2019) after lodging occurred. Traditional machine learning and deep learning were evaluated and compared in this study in terms of classification accuracy and standard deviation. For traditional machine learning, five types of features (i.e. gray level co-occurrence matrix, local binary pattern, Gabor, intensity, and Hu-moment) were extracted and fed into three traditional machine learning algorithms (i.e., random forest (RF), neural network, and support vector machine) for detecting lodged plots. For the datasets on each imagery collection date, the accuracies of the three algorithms were not significantly different from each other. For any of the three algorithms, accuracies on the first and last date datasets had the lowest and highest values, respectively. Incorporating standard deviation as a measurement of performance robustness, RF was determined as the most satisfactory. Regarding deep learning, three different convolutional neural networks (simple convolutional neural network, VGG-16, and GoogLeNet) were tested. For any of the single date datasets, GoogLeNet consistently had superior performance over the other two methods. Further comparisons between RF and GoogLeNet demonstrated that the detection accuracies of the two methods were not significantly different from each other (p > 0.05); hence, the choice of any of the two would not affect the final detection accuracies. However, considering the fact that the average accuracy of GoogLeNet (93%) was larger than RF (91%), it was recommended to use GoogLeNet for wheat lodging detection. This research demonstrated that UAS RGB imagery, coupled with the GoogLeNet machine learning algorithm, can be a novel, reliable, objective, simple, low-cost, and effective (accuracy > 90%) tool for wheat lodging detection.

1. Introduction

Ranked as one of the top three staple food crops worldwide, wheat is a major source of starch and energy, as well as of essential and beneficial components to health, such as B vitamins, dietary fiber, and phytochemicals [1]. One of the main factors negatively impacting wheat production is crop lodging, defined as the displacement of the above-ground stems from their upright position (stem lodging) or failure of root-soil attachment (root lodging) [2]. Numerous studies have reported that wheat lodging results in yield losses of up to 50% [3,4,5,6] and lodging not only degrades wheat quality, but also delays harvest and increases drying time [7,8]. A variety of reasons can lead to lodging, namely strong winds, heavy rain, hail, and improper crop management, including excessive nitrogen and high crop density [9]. Detecting wheat lodging will be of interest from producers to researchers to assess the extent of damage and to develop new varieties for crop management to reduce the yield and quality loss.
The monitoring and assessing of wheat lodging conditions in a timely fashion would benefit several stakeholders: (i) wheat breeders—it is critical to identify lodging-resistant varieties among thousands of experimental plots [10]; (ii) wheat growers—they need to file a written notice of lodging damage within a short time (two to three days from the lodging discovery) to be eligible for insurance coverage [11]; (iii) insurance loss adjusters—they have to quantify lodging area and severity, on which they base the farmers’ relief compensation [12,13]; (iv) agronomists and plant physiologists—it is ideal to identify lodging areas in a large field so they can study the issue immediately after it occurs [14,15,16].
The currently widely used and accepted approach for wheat lodging assessment relies on manual in situ measurements (using a tape, pole, and/or measuring wheel) and visual inspection, which is time-consuming, labor-intensive, and inaccurate [15,16,17,18,19]. The manual/visual approach is also hindered by some extreme environmental conditions during assessment, such as high temperatures (>40 °C) and flooded fields [11]. Furthermore, the assessment results are subjective, as inspections carried out on the same field by different inspectors can yield different results, which could lead to a compensation dispute between growers and insurance companies [20]. Therefore, there is a need for an unbiased, low-cost, accurate, rapid, repeatable, and objective approach for crop lodging detection.
Remote sensing technology has been identified as a candidate tool for crop lodging detection in recent decades [21]. According to the platforms onto which the sensors are mounted, agricultural remote sensing can be grouped into satellite, aerial, and near-ground remote sensing [22]. Satellite imagery takes advantage of the different spectral reflectance features exhibited by the stalks and leaves to distinguish lodged from non-lodged areas [23]. Despite covering large field areas on the ground, satellite remote sensing has a limited performance for lodging assessment due to its low spatial- and temporal-resolution [12,13]. The currently applied satellite imagery for crop lodging detection presented a spatial resolution > 10 m and a temporal resolution > 3 days, which are larger than the requirements of 1–2 m and 1–2 days [24]. Though some commercial satellites could provide high spatial (sub-meter) and temporal (1 day) resolution images and use this data for large areas, crop lodging assessment only makes sense with ample resources, and is not economically feasible for smaller fields with limited resources that can be readily covered by UAS flights affordably. Additionally, weak spectral differences between lodged and non-lodged crops could occur due to other factors, such as soil nutrient status, crop stress, and water stress [25,26,27]. Aerial remote sensing, based on manned aircrafts or balloons, serves as a supplementary approach to satellites, which can provide higher spatial- and temporal-resolution images. Murakami et al. [28] hung a compact digital camera under a balloon flying 50–150 m above ground level (AGL) to collect images for buckwheat lodging condition detection. However, the operational procedure with this system is complex as the balloon’s motion trajectory in the air is uncontrollable and there can be additional challenges during the image processing phase. Manned aircrafts need a large amount of space to take off and land, and their investment and maintenance are highly expensive, even though their suitable sensors can produce better images and can be applied to larger field areas.
Recently, near-ground remote sensing has been adopted in many agricultural areas due to the rapid development of unmanned aerial systems (UAS), as well as associated data processing software. Compared to satellite and balloon/manned aircraft imagery, UAS imagery can be acquired in a cost-effective manner, with satisfactory maneuverability and increased spatial- and temporal-resolution [29,30]. In addition, the application of UAS is highly flexible and the application procedure is relatively simple compared to the operation of balloons and manned aircrafts. Furthermore, the cost of using UAS for crop monitoring is significantly lower than that of using satellite and manned aircraft images. The use of UAS for crop lodging detection is still in the early experimental stage and related studies are few [15]. Corn lodging detection studies were conducted by incorporating UAS, RGB, and two band spectral images [12,13,19]. However, corn lodging assessment will likely be different from wheat, given their significantly different appearance and crop density. Wheat and canola lodging detection has been conducted by combining RGB and multiple bands spectral images, along with the nearest neighborhood classification algorithm [16,31,32]. In addition, researchers assessed rice lodging conditions using thermal images, along with digital surface modeling, color images, and multi-spectral images [20,21]. In addition to academic efforts, several UAS startups are exploring the potential of using UAS to collect information for crop lodging detection, but their results have not been reported to a wider audience (Aker, MN, US; GK Technology, MN, US). A majority of existing studies took advantage of multi-spectral, thermal, and RGB images for crop lodging detection. However, considering the high cost of multi-spectral and thermal cameras, the research outcomes may be unaffordable to farmers. It is therefore necessary and meaningful to test and validate the method of using low-cost UAS visible light imagery for crop lodging detection.
After collecting high-quality imagery, the development of appropriate classification algorithms is critical for lodging detection. According to feature extraction methods, classification algorithms can be categorized into traditional machine learning and deep learning [33]. For traditional machine learning, proper features that represent lodging plots are first determined using domain knowledge and then extracted manually. The extracted features are then fed into different machine learning algorithms, such as nearest neighbors, linear discriminant analysis, random forest (RF), neural network (NN), and support vector machine (SVM) to distinguish lodging from non-lodging crops [33,34]. Rajapaksa et al. [32] extracted texture features and then applied the SVM for wheat and canola lodging distinction. Liu et al. [21] applied the SVM on color, texture, and thermal features for crop lodging and non-lodging distinction. The modern machine learning approach is deep learning and is associated with artificial intelligence for image processing and data analysis [35]. The convolutional neural network (CNN) of deep learning, combining feature extractor and classifier at the same time, was demonstrated to be superior over many other existing machine learning algorithms [36]. The CNN has been evolving very quickly from the original LeNet to VGGNet/GoogLeNet [37,38,39], and recently it has been applied in the agriculture domain to make assessments, such as plant disease and health condition detection [10,40]. These emerging traditional machine learning and deep learning algorithms (e.g., RF, NN, SVM, VGGNet, and GoogLeNet) have the potential to improve crop lodging detection accuracies. However, based on the authors’ best knowledge, their performances on wheat lodging detection have not been validated and compared.
Given this information, it is therefore proposed to study the application of traditional machine learning and deep learning algorithms to wheat lodging and compare their performances to determine the best methodology. For this study, the UAS visible light imagery of wheat plots, a low-cost approach affordable to users, was acquired on three different dates after the occurrence of lodging. The data were preprocessed and fed to machine learning algorithms, and their classification performances were determined and compared. The specific objectives of this study were to: (1) apply and evaluate the performances of both traditional machine learning methods (i.e., RF, NN, and SVM) and deep learning algorithms (i.e., simple CNN, VGG-16, and GoogLeNet) for wheat lodging assessment using proper performance metrics; (2) determine the desirable algorithms for wheat lodging detection that yield higher accuracy and lower standard deviation.

2. Materials and Methods

The various processes followed in the study of wheat lodging detection, such as field data acquisition, machine learning model applications, and model performance evaluations, are summarized in Figure 1. Datasets created from the UAS images and field observations were processed through two groups of machine learning methods, namely, traditional machine learning, and deep learning (three selected models each). Model performances were finally evaluated and compared, with the most desirable one(s) recommended. All of the processes involved in the study are subsequently presented in detail in the appropriate sections.

2.1. Test Fields

The experimental field used in this study was located near Thompson, ND (UTM WGS 84 14 N), US, which consisted of 372 plots, as shown in Figure 2. The plots belonged to different research trials and their dimensions were as follows: 1.5 m × 3.6 m (51 rows × 4 columns = 204 plots, ID1), 1.5 m × 5.4 m (10 rows × 12 columns = 120 plots, ID2), and 1.5 m × 14.6 m (12 rows × 4 columns = 48 plots, ID3). The field was planted on 15 May 2019, at a rate of ~100 kg/ha, with a row spacing of 0.19 m. Immediately after planting, eight ground control points (GCP) were installed in the field, as shown in Figure 2.

2.2. Data Collection

Followed by seed germination (one week after sowing), crop growth conditions were monitored on weekly basis. Only after mid-July of 2019, were the first symptoms of wheat lodging noticed because of heavy rain and strong winds. The data collection was then started using a DJI Phantom 4D RTK UAS (DJI-Innovations, Inc., ShenZhen, China). The UAS is outfitted with a 20 megapixel (5472 × 3648 pixels) color camera, mounted on a three-axis stabilization gimbal. This study did not use a multispectral camera for data collection, due to its high cost, and our focus of developing an affordable technology for farmers. DJI’s ground station app (DJI GS RTK, V2.1.1) was used to set up the flight mission map and parameters. Instead of 3D Photogrammetry, 2D Photogrammetry was applied for setting up the mission because the experimental plots were flat. The UAS was flown 25 m AGL (image resolution of ~0.7 cm/pixel), and the speed was set to 2.5 m/s. The following settings were applied: shooting mode was set to “Timed Shooting”; photo ratio was 3:2; white balance mode was “sunny”; gimbal angle was -90° (nadir position); both side and forward overlap were set at 80%; the margin setting was kept at “Auto” mode. Three flights were carried out on 23 July 2019, 30 July 2019, and 8 August 2019. The georeferenced images were stored in a SD card during flight, and they were transferred to a desktop computer back in the office for processing. After the UAS imagery collection, a group of inspectors visited the field and manually classified each plot as lodging or non-lodging based on visual observations, and the results were recorded, as shown in Figure 3. If the wheat stems were off their original vertical position in a permanent manner, they were judged as lodged crops.

2.3. Image Preprocessing

After each UAS flight, images were stitched together using PIX4DMapper (Pix4D V4.3.33, S.A., Prilly, Switzerland) to generate an orthomosaic map. Eight GCPs established in the field, shown in Figure 2, were used as references to overlap the orthomosaic maps of the three dates, using the first date’s imagery as a reference for the other two dates. The plot images datasets (372 plot images) for each of the dates were created by manually cropping the individual plot image, and the recorded results (lodged or not-lodged plot) for each plot were organized to be associated with the individual image. All three data collections occurred in the morning time between 9:00–11:00 am in sunny weather conditions, and each flight mission lasted for about 12~15 min. The consistent illumination conditions (measured by MQ-200, Apogee Instruments, Logan, UT, USA) during the flight mission avoided the calibration among the datasets of different dates.

2.4. Traditional Machine Learning Algorithms for Lodging Detection

2.4.1. Feature Extraction for RF, NN, and SVM

Image classification using RF, NN, and SVM requires individual images to be represented by a number of discriminative features [33]. The domain knowledge suggested that textural features (the second-order statistics of an image domain) should be proper indicators for lodging and non-lodging plot classifications [20,21].
The right photo in Figure 3 shows two lodging and two non-lodging plots. A sample of lodging and non-lodging plot images illustrates easily observable differences in the color and textural characteristics of these two categories, as is visible in Figure 3. A variety of image characteristics could be used in machine learning algorithms for the accurate and efficient distinction of lodging or non-lodging. Color features may represent differences, such as leaves being darker green, and stems being lighter green [41]. Texture features could also be desirable indicators for lodging and non-lodging plot distinctions [21]. Compared to the lodging plots, shown in Figure 3 as having non-uniform and heterogeneous patters, the non-lodging plots have a more uniform and homogeneous format. Considering the fact that color characteristics are highly related to a variety of factors, such as growth stages, crop variety, and nutrients of soil, textural features tend to be more consistent. Therefore, five types of image texture features, namely, Haralick, local binary pattern (LBP), Gabor, intensity, and Hu-moment were extracted and used in the algorithms. Haralick features (measurement of the variation in the intensity of an image) were calculated from a Gray Level Co-occurrence Matrix (GLCM) for the mean, sum, variance, and standard deviation of different textural measures [41]. A total of 88 Haralick textural features (22 features in four directions) were extracted [42,43]. For LBP, 59 features (labeling the pixels of an image by thresholding the neighborhood of each pixel) were extracted by combining one non-uniform LBP and 58 uniform LBP features. The LBP uniformity was determined on the occurrence of two bitwise transitions when circularly sampled by a 3 × 3 filter [44]. For the Gabor features (analyzing if there are any specific frequency contents in the image in specific directions), each image was filtered with a bank of Gabor filters generated with four dilations and four rotations as a spatial mask of a 4 × 4 pixel square [45]. In total, 160 Gabor features were extracted from one plot image. Additionally, six basic intensity features of the individual image were extracted, including mean, standard deviation, kurtosis, skewness, average gradient, and Laplacian mean [46]. Furthermore, seven invariant descriptors for an image were extracted as Hu-moment features [47]. After all of the features were extracted, they were concatenated in tandem for every plot image to form a matrix of features (also referred to as master dataset), and by doing so, one image was represented by 320 features.

2.4.2. Classification

Image classification was conducted separately on three dates datasets. Each date dataset was partitioned through a random selection process (without replacement) to obtain a training and test dataset. While the training dataset consisted of 70% (260 images) of instances from the master dataset, the test dataset consisted of 30% (112 images) of the instances from the master dataset.
The RF is an ensemble learning algorithm that predicts the class labels of an unlabeled instance by aggregating the results from multiple decision trees [48]. Unlike a single decision tree-based classification model, in which a complete training dataset is chosen to determine the decision rules, a subset of the training dataset via boot strapping (i.e., via random sampling with replacement) is chosen to build multiple decision trees in the case of RF, as shown in Figure 4. Details of the decision tree algorithms are not provided in this paper, but can be found in other publications [34,49]. Decision trees built in the RF use only a subset of features (drawn randomly) to obtain the splitting of a node instead of all the features [50]. The number of features in the subset is approximately chosen to be m , where m = 320, which is the number of features for this study. Each tree of the RF then provides the outcome (i.e., class label) of the unlabeled instance [49]. Attributed to the fact that each tree of the RF may result in different outcomes, the majority vote of the total outcomes is carried out to determine the final class label [49], as shown in Figure 4.
The prediction accuracies of RF depend on the number of decision trees used to train the model. To assess the performance of the RF, typically the out-of-bag error (OOBE) is evaluated with respect to the number of decision trees. Performance in this study refers to the ability of the model to predict correct class labels. The OOBE refers to the percentage of the misclassified class labels on the data instances that are left out after bootstrapping samples. The number of decision trees at which the OOBE drops significantly and stabilizes is generally chosen to build a RF model [51,52]. From the multiple preliminary tests carried out in this study, it was observed that OOBE decreased rapidly and then stabilized when the tree number exceeded 80, as shown in Figure 5. Note that no prior feature selection was performed in this study.
A NN is a computational system that is made of several highly interconnected process elements (neurons), which receive, process, and transmit information to other neurons [53]. In other words, NN is a mathematical model that maps a given input to an output. In the current study, with 320 features extracted, NN aims to predict if the features belong to “lodging” or “non-lodging”. A multi-layer perception feed-forward neural network was used in this study, which is configured with two hidden layers—10 and 5 neurons each for the first and second hidden layers, respectively, as shown in Figure 6. The first (input) layer consists of 320 features, and the last (output) layer represents the class labels (i.e., lodging and non-lodging). The neurons in the hidden layers represent a computational unit that performs a transformation of the linear sum of inputs that is received from preceding neurons. Transformation was carried out using a non-linear activation function, as shown in Figure 6. Training NN involves computing the weights (w in Figure 6) and bias (b in Figure 6) of the model to minimize the classification error in an iterative manner. In this study, a MATLAB® built-in program “net” was used to train the multi-layer feed-forward neural network model where a number of 1000 iterations are used [53]. ‘Sigmoid’ was used as an activation function in the hidden neurons, and ‘softmax’ function was applied to determine the class label in the output layer, as shown in Figure 6.
The SVM is a supervised machine learning algorithm that has gained a lot of attention in recent years and is used to perform the task of classification [54]. The objective of the SVM is to find a hyperplane in a F-dimensional space (F is the number of extracted features) that can distinctly classify the data points. Given that there are many possible hyperplanes that can be chosen, it uses the one that has the maximum margin—maximum distance between different class points, as denoted by ‘Margin’ in Figure 7. Support vectors are data points that are closer to the hyperplane, which determine the position and orientation of the hyperplane. To facilitate the distinction of data that are non-linearly separable, kernel functions that can transform the original data into higher dimensions are generally adopted, shown in Figure 7C. There are various kernel functions, such as “linear”, “polynomial”, “Gaussian or Radial Basis Function” (RBF), and “Sigmoid”, among which RBF was applied in this study, because of its robustness and proven performance [54,55]. As there were only two classes (lodging and non-lodging) in this study, a binary SVM was selected for image classification [56]. To perform classification, a built-in MATLAB® function “fitcsvm” was executed where the auto kernel scale was chosen as the input arguments. Applying an “auto” kernel scale indicates that the algorithm automatically selected an appropriate scale factor using a heuristic procedure.
All plot images were re-sized to 80 × 250 pixels before feature extraction. The above procedures for image resizing, feature extraction, and model training/testing with RF, NN, and SVM were performed in MATLAB® R2019a (The Mathworks, Inc., Natick, Mass., USA).

2.5. Deep Learning

In the deep learning CNN algorithm, wherein the whole image is fed as inputs and the various aspects/objects in the image are assigned importance (e.g., learnable weights and biases) for establishing a distinction between different objects [37]. Compared to the traditional machine learning methods, which require the manual feature extraction of individual images, the CNN deep learning approach requires minimal image pre-processing [57]. In CNN, an image is passed through a sequence of convolutional layers or kernel filters to extract the features (not directly accessible to users). Kernel filters are composed of weights that are determined through an iteration process. Multiple convolutional layers may be required, with different layers extracting different levels of features—low-level features include edges, colors, and gradient, while high-level features have a wholesome understanding of images.
Followed by convolution, pooling operation was carried out in each layer with the purpose of reducing the spatial size of the convolved features. Simultaneously, this procedure reduced the computational power requirement for further data processing. There are two commonly used pooling approaches—max pooling and average pooling. Max pooling is generally superior as it suppresses noise, which was applied in this study. The third layer is a fully connected layer—a classification layer using the Softmax classification technique to provide the predicted label of the image.
Similar to a majority of other machine learning algorithms, CNN needs a large set of training images to avoid overfitting. Considering the relatively small data samples in this study (372 samples), data augmentation (increasing the data sample) was performed. There are two popular ways for data augmentation. The first approach is to physically increase the dataset—the current samples number is 372 and it could be increased to 1116 samples (three times). Then, the physically augmented dataset is used for training and testing. The other method is to conduct data augmentation before each epoch, which means slightly different versions of images are fed into the algorithm for training. Since avoiding physically increasing the dataset (saving disk space), the second data augmentation approach was implemented by applying a variety of geometric transformation to the original images [58]. These geometric transformations include reflection, translation, rotation, horizontal/vertical scaling, zooming, and flipping, among which the first four were applied. Figure 8 shows a sample of the images generated by these transformations for image data augmentation.

2.5.1. Simple Convolutional Neural Network for Classification

A simple convolutional neural network (SCNN) consisting of three convolutional layers, two pooling layers, and one fully connected layer was generated and trained. For the first convolutional-pooling layer, eight convolution filters (resulting in 24 feature maps) of 3 × 3 pixels with a stride step of one pixel were applied, followed by a 2 × 2 max pooling layer, as shown in Figure 9. A second convolutional-pooling layer consisted of 16 convolution filters of 3 × 3 pixels with one stride step and a 2 × 2 max pooling layer, shown in Figure 9 as a hidden layer. Then, it was followed by another convolution layer, consisting of 32 convolution filters and a fully connected layer. A Softmax layer activation function normalized the output of the fully connected layer, and the classification layers (final layers) used parameters from Softmax activation to make a classification. A rectified linear unit was applied as the activation function in all the hidden layers.
In this study, the network was trained using a stochastic gradient descent algorithm with a momentum (an initial learning rate) of 0.01. Four epochs were applied, and the data were shuffled and geometrically transformed before being fed into every epoch (data augmentation).

2.5.2. VGG-16

The VGG-16 is a pre-trained CNN architecture whose weights were determined through the training conducted on approximately one million images from the ImageNet Dataset (http://image-net.org/index). The VGG architecture, shown in Figure 10, consists of 13 convolutional layers (extracting image features with a 3 × 3 size filter), five max pooling layers (reducing the spatial size of images), and three fully connected layers (classifying images into labels). The model can classify images into 1000 object categories (e.g., keyboard, mouse, and pencil). Compared to the SCNN, which requires the training of the whole network with randomly initialized weights, pre-trained VGG-16 saves model training time as the weights were already determined [38].
The procedures of using VGG-16 for lodging detection are described in Figure 11, starting with loading pre-training VGG-16. The late layers (e.g., “loss3-classifier” and “output” of loaded VGG-16), which function to combine network extracted features into class probabilities, were replaced with new layers adapting to the datasets for this study. Before re-training the network, the weights of earlier layers in the network were frozen by setting the learning rates to 0. In addition, to significantly reduce the time required for network training, freezing the weights of initial layers can prevent their overfitting effect on the new dataset. The training of the network starts from resizing all images into a standard size of 224 × 224 × 3, which was realized by the function of imageAugmenter. MaxEpochs and InitialLearningRate, as the two key options for network training, were setup as 6 and 3 × 10−4 in this study, respectively. Then the testing dataset was fed into the newly trained VGG-16 network and the predicted results were used to calculate the accuracy of the updated VGG-16 model.

2.5.3. GoogLeNet

Compared to the VGG-16 net, in which convolution layers are stacked linearly for better performance [39], GoogLeNet applies the inception module for feature extraction [59]. The inception module is a block of parallel convolutional layers with three differently sized filters (i.e., 1 × 1, 3 × 3, and 5 × 5) and a 3 × 3 max pooling layer, and the results are concatenated, as shown in Figure 12 [60]. Since large (5 × 5) and small (1 × 1) size filters extract general and local features, respectively, the inception module extracts features in a more inclusive manner. This study takes advantage of a pre-trained 22-layer deep GoogLeNet for detecting wheat lodging. Other than loading GoogLeNet instead of VGG-16, the procedure of applying GoogLeNet is exactly the same as applying VGG-16, described earlier and shown in Figure 11.
In this study, we tested three CNNs for their accuracy performance on detecting lodging plots, and for all these three methods, the datasets were randomly portioned into training and testing sets according to a ratio of 7:3 before they were fed into algorithms. The above procedures for constructing, loading, modifying, and running three different CNNs were performed in MATLAB® R2019a (The Mathworks, Inc., Natick, Mass.).

2.6. Accuracy Evaluation and Model Comparison

Image classification results for the test dataset are most commonly assessed using the following three performance metrics: precision (PRE), recall (REC), and overall accuracy (OAC) [34]. Besides these, another performance metric called F-measurement (F1 score) is also used, which combines precision and recall [61].
PRE (%) = #TP/(#TP + #FP) × 100
REC (%) = #TP/(#TP + #FN) × 100
OAC (%) = (#TP + #TN)/(#TP + #TN + #FP + #FN) × 100
F1 (%) = 2 × PRE × REC/(PRE + REC) × 100
where # stands for “number of”, TP is true positive (lodging plots classified as lodging), TN is true negative (non-lodging plots classified as non-lodging), FP is false positive (non-lodging plots classified as lodging), and FN is false negative (lodging plots classified as non-lodging). The F1 was calculated based on PRE and REC. The four metrics, Equations (1)–(4), were calculated as an average from 10 replications.
Model accuracies were compared among machine learning models as well as among the three dates within each group, and finally between selected models from the groups. Tukey’s test was performed at 0.05 significance level using SAS (Version 9.4, SAS Institute Inc., Cary, NC, USA) for the comparison.

3. Results and Discussion

3.1. Traditional Machine Learning for Lodging Detection

The model performance metrics, PRE, REC, OAC, and F1 results for the three classifiers (RF, NN, and SVM) on the three individual date datasets are given in Figure 13. Detection accuracies, as measured by PRE, REC, OAC, and F1, varied with different classifiers and performance metrics. The REC ranged from 73% to 87%, 67% to 92%, and 70% to 92% for RF, NN, and SVM, respectively, and overall, it ranked the lowest accuracy among the four parameters for all three classifiers. The PRE ranges were 87%–88%, 77%–85%, 87%–92%, and F1 were 79%–88%, 71%–88%, 77%–91%, for RF, NN, and SVM, respectively. Additionally, the OAC values ranges were 85%–88%, 85%–91%, 88%–93%, for RF, NN, and SVM, respectively. Compared to PRE, REC and F1, OAC performed more desirably. In addition to ranking the highest accuracy among the three parameters, the OAC resulted in the smallest averaged fluctuation among the three classifiers, as shown in Table 1, indicating its robust performance. Therefore, only OAC was chosen in the following discussion for model performance comparisons, and hereafter, accuracy denotes OAC.
For all three classifiers, it was observed that the overall trend of accuracy increased with time, from 89% to 91%, 85% to 91%, and 88% to 93%, for RF, NN, and SVM, respectively, as shown in Figure 14. For all three classifiers, their accuracies ranked highest on the last date dataset (August 8th 2019), and the first and the last dates were significantly different (p < 0.05), as shown in Figure 14. This can be explained that the lodging was a dynamic process, which required time to complete. In addition, the average standard deviation for RF, NN, and SVM on the three dates datasets were calculated as 0.005, 0.046, and 0.027, respectively, with RF producing the least deviations.
Further comparisons of three classifiers’ performances on the individual date datasets are shown in Figure 15. For any of the three dates datasets, the classifiers’ accuracies were not significantly different from each other (p > 0.05), shown in Figure 15, indicating the selection of the classifiers does not affect the accuracy. However, RF was determined as the most satisfactory approach for its lowest standard deviation, shown in Figure 14. Generally, to achieve higher accuracy, it is desirable to avoid the use of the images collected immediately after lodging occurred.

3.2. Deep Learning for Lodging Detection

For the deep learning algorithms of SCNN and GoogLeNet, the statistical accuracy comparisons of different date datasets showed that the last date (August 8th 2019) was significantly higher than the second date (July 30th 2019), but not significantly higher than the first date (July 23rd 2019), as shown in Figure 16. This was probably because the second date was a transitional lodging stage, and the automatically extracted features did not perform well during this transitional stage. The VGG-16 deep learning algorithm’s accuracy was not significantly different for the individual date dataset, indicating its robust performance on all three dates. In addition, GoogLeNet resulted in the lowest average value on standard deviation (0.027), followed by SCNN (0.032) and VGG-16 (0.044). Similar to the results of the previous section, it is generally recommended to apply deep learning algorithms on the last date dataset for lodging detection for the purpose of higher accuracy.
Further comparisons of the results of three deep learning algorithms at all three dates showed a similar accuracy patter, shown in Figure 17. The GoogLeNet ranked the highest based on accuracy and was significantly different from the other two (SCNN and VGG-16), which in turn were not significantly different from each other. The GoogLeNet accuracies for July 23rd 2019, July 30th 2019, and August 8th 2019 were 91%, 89%, and 93%, respectively. This result indicates that it is preferable to use GoogLeNet on the August 8th 2019 dataset for higher detection accuracy.
Crop lodging detection accuracies through the application of machine learning can also be found in the literature. Yang et al. [20] achieved a 96% rice lodging detection by using visible light and spectral images. Kumpumäki et al. [62] reported a detection accuracy of 73% for rye by using Sentinel-2 images. Chauhan et al. [16] and Rajapaksa et al. [32] achieved 90% and 92% accuracies, respectively, on wheat lodging detection based on UAS multispectral data. Rajapaksa et al. [32] further reported canola lodging detection accuracies of 90% and 87% using five channel spectral imagery (red, blue, green, near infrared, and red edge). In our study, the 93% accuracy of GoogLeNet on wheat lodging detection can be considered as an improvement compared to the reported results and a satisfactory performance. In addition, we only used visible light images, which significantly decreased the technology cost.

3.3. Comparison of RF and GoogLeNet

RF and GoogLeNet were identified as the most desirable algorithms for detecting lodging plots from the traditional machine learning and deep learning approach, respectively. The accuracy comparisons of RF and GoogLeNet on the wheat lodging detection of the individual date dataset are shown in Table 2. For any of the three dates, the detection accuracies of the two classifiers were not significantly different from each other (p > 0.05). Additionally, both methods had the highest accuracies on the last date datasets (August 8th 2019). It can be concluded that the choice of the methods (either RF or GoogLeNet) would not affect the accuracies. However, considering that GoogLeNet resulted in a higher average accuracy value over RF (93% > 91%), it is recommended to use GoogLeNet for wheat lodging detection.

3.4. Future Research Direction

This research focused on distinguishing lodging plots from non-lodging plots in a binary sense, and the future research direction should specifically look into determining quantitative information on lodging severity expressed on a scale ranging from non-lodging to complete lodging. Additionally, the versatility of this method should be tested and validated on other crops susceptible to lodging, such as canola, corn, and soybean. Since the current method is entirely off-line, future efforts should be directed to near real- or real-time lodging detection for the benefits of end-users. Embedded systems attached to UAS, coupled with RF or GoogLeNet algorithms, should be another promising area of research [63,64]. In this study, UAS was flown 25 m AGL, with an image resolution of 0.7 cm. With a larger AGL, the UAS would have a higher area coverage per flight but would only obtain low resolution images. More studies are needed to check the performance of these machine learning algorithms with a higher AGL (e.g., 35 m and 40 m). Considering the current model prediction accuracy is >90% by only using textual features, incorporating visible color and spectral information has the potential to further improve the model performance and this can be addressed through additional research.

4. Conclusions

In this study, UAS imagery was collected over wheat plots (372 individual plots) on three different dates, along with ground truth data (lodging or non-lodging) for each plot. After the stitching process, individual plot images were manually cropped from orthomosaic imagery to create datasets for each date. For the traditional machine learning approach, 320 extracted features were fed into three algorithms (random forest, neural network, and support vector machine). Though the three algorithms did not perform significantly different detection accuracies on any of the three dates datasets (p > 0.05), RF was determined as the most satisfactory and robust method, due to its lowest standard deviation value. For the deep learning method, which had the advantage of avoiding the manual feature extraction and using the images directly, the detection accuracies of GoogLeNet was consistently higher than the other two algorithms, and it had the highest detection accuracy (93%) on the last date (August 8th 2019) dataset, where the crop lodging was complete. Detection accuracy comparisons between RF and GoogLeNet demonstrated that there were no significant differences (p > 0.05) in the two methods on any of the three dates datasets. It can be recommended that the users could choose either of the two methods (RF or GoogLeNet) based on their preference or the availability of resources. Future research should investigate the quantification of lodging severity expressed in numerical scale, extending to other crops susceptible to lodging, and the use of real-time detection and embedded systems. It should be noted that the UAS used in this study can fly a mission with maximum of ~25 min, and this significantly limits its application in large-area fields. In addition, the flight is significantly dependent on weather conditions, and the UAS cannot carry out scout work in rainy and windy weather conditions. This study had demonstrated that UAS imagery, coupled with machine learning algorithms, has the potential to be used as a novel, objective, and a promising ready-to-use tool for wheat lodging detection because of its simplicity and efficiency (accuracy > 90%). This developed technology could benefit wheat breeders and growers, insurance loss adjusters, as well as agronomists and plant physiologists.

Author Contributions

Conceptualization, Z.Z., P.F. and C.I.; methodology, Z.Z., C.I. and D.L.N.; validation, Z.Z., P.F., C.I., D.L.N., and R.K.; writing—original draft preparation, Z.Z. N.K., and C.I.; writing—review and editing, Z.Z., P.F., C.I., J.K.R.; supervision, Z.Z., and P.F.; project administration, Z.Z., and P.F.; funding acquisition, Z.Z., and P.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research is financially supported by the United States Department of Agriculture (USDA).

Acknowledgments

Authors would like to thank Kalin Morgen and Jensen Kenton for their help for data collection.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shewry, P.R.; Hey, S.J. The contribution of wheat to human diet and health. Food Energy Secur. 2015, 4, 178–202. [Google Scholar] [CrossRef] [PubMed]
  2. Wu, W.; Ma, B.L. A new method for assessing plant lodging and the impact of management options on lodging in canola crop production. Sci. Rep. 2016, 6, 31890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Weibel, R.O.; Pendleton, J.W. Effect of Artificial Lodging on Winter Wheat Grain Yield and Quality1. Agron. J. 1907, 56, 487–488. [Google Scholar] [CrossRef] [Green Version]
  4. Berry, P.; Sterling, M.; Spink, J.; Baker, C.; Sylvester-Bradley, R.; Mooney, S.J.; Tams, A.; Ennos, A. Understanding and Reducing Lodging in Cereals. Adv. Agron. 2004, 84, 217–271. [Google Scholar]
  5. Berry, P.M.; Spink, J. Predicting yield losses caused by lodging in wheat. Field Crop. Res. 2012, 137, 19–26. [Google Scholar] [CrossRef]
  6. Berry, P.M.; Sterling, M.; Baker, C.; Spink, J.; Sparkes, D.L. A calibrated model of wheat lodging compared with field measurements. Agric. For. Meteorol. 2003, 119, 167–180. [Google Scholar] [CrossRef]
  7. Pinthus, M.J. Lodging in Wheat, Barley, and Oats: The Phenomenon, its Causes, and Preventive Measures. Adv. Agron. 1974, 25, 209–263. [Google Scholar]
  8. Islam, M.S.; Peng, S.; Visperas, R.M.; Ereful, N.; Bhuiya, M.S.U.; Julfiquar, A.W. Lodging-related morphological traits of hybrid rice in a tropical irrigated ecosystem. Field Crop. Res. 2007, 101, 240–248. [Google Scholar] [CrossRef]
  9. Duy, P.Q.; Hirano, M.; Sagawa, S.; Kuroda, E. Analysis of the Dry Matter Production Process Related to Yield and Yield Components of Rice Plants Grown under the Practice of Nitrogen-Free Basal Dressing Accompanied with Sparse Planting Density. Plant Prod. Sci. 2004, 7, 155–164. [Google Scholar] [CrossRef] [Green Version]
  10. Mardanisamani, S.; Maleki, F.; Kassani, S.H.; Rajapaksa, S.; Duddu, H.; Wang, M.; Shirtliffe, S.; Ryu, S.; Josuttes, A.; Zhang, T.; et al. Crop Lodging Prediction from UAV-Acquired Images of Wheat and Canola Using a DCNN Augmented with Handcrafted Texture Features. In Proceedings of the 2019 IEEE/CVF Conference on Computer Vision and Pattern Recognition Workshops (CVPRW), Long Beach, CA, USA, 16–17 June 2019; pp. 2657–2664. [Google Scholar]
  11. Horvatic, B. Using Drone Mapping for Crop Insurance. 2019. Available online: https://www.precisionag.com/in-field-technologies/drones-uavs/using-drone-mapping-for-crop-insurance/ (accessed on 15 April 2020).
  12. Li, X.; Wang, K.; Ma, Z.; Wang, H. Early detection of wheat disease based on thermal infrared imaging. Trans. Chin. Soc. Agric. Eng. 2014, 30, 183–189. [Google Scholar]
  13. Li, Z.; Chen, Z.; Wang, L.; Liu, J.; Zhou, Q. Area extraction of maize lodging based on remote sensing by small unmanned aerial vehicle. Trans. Chin. Soc. Agric. Eng. 2014, 30, 207–213. [Google Scholar]
  14. Sposaro, M.M.; Berry, P.M.; Sterling, M.; Hall, A.J.; Chimenti, C.A. Modelling root and stem lodging in sunflower. Field Crop. Res. 2010, 119, 125–134. [Google Scholar] [CrossRef]
  15. Chauhan, S.; Darvishzadeh, R.; Boschetti, M.; Pepe, M.P.L.; Nelson, A. Remote sensing-based crop lodging assessment: Current status and perspectives. ISPRS J. Photogramm. Remote Sens. 2019, 151, 124–140. [Google Scholar] [CrossRef] [Green Version]
  16. Chauhan, S.; Darvishzadeh, R.; Lu, Y.; Stroppiana, D.; Boschetti, M.; Pepe, M.; Nelson, A. Wheat lodging assessment using multispectral UAV data. ISPRS—Int. Arch. Photogramm. Remote Sens. Spat. Inf. Sci. 2019, 31, 235–240. [Google Scholar] [CrossRef] [Green Version]
  17. Bock, C.H.; Poole, G.H.; Parker, P.E.; Gottwald, T.R. Plant Disease Severity Estimated Visually, by Digital Photography and Image Analysis, and by Hyperspectral Imaging. Crit. Rev. Plant Sci. 2010, 29, 59–107. [Google Scholar] [CrossRef]
  18. Robertson, D.J.; Julias, M.; Gardunia, B.W.; Barten, T.; Cook, D.D. Corn Stalk Lodging: A Forensic Engineering Approach Provides Insights into Failure Patterns and Mechanisms. Crop. Sci. 2015, 55, 2833–2841. [Google Scholar] [CrossRef] [Green Version]
  19. Chu, T.; Starek, M.J.; Brewer, M.J.; Murray, S.C.; Pruter, L.S. Assessing Lodging Severity over an Experimental Maize (Zea mays L.) Field Using UAS Images. Remote Sens. 2017, 9, 923. [Google Scholar] [CrossRef] [Green Version]
  20. Yang, M.-D.; Huang, K.-S.; Kuo, Y.-H.; Tsai, H.P.; Lin, L.-M. Spatial and Spectral Hybrid Image Classification for Rice Lodging Assessment through UAV Imagery. Remote Sens. 2017, 9, 583. [Google Scholar] [CrossRef] [Green Version]
  21. Liu, T.; Li, R.; XiaoChun, Z.; Jiang, M.; Jin, X.; Zhou, P.; Liu, S.; Sun, C.; Guo, W. Estimates of rice lodging using indices derived from UAV visible and thermal infrared images. Agric. For. Meteorol. 2018, 252, 144–154. [Google Scholar] [CrossRef]
  22. Du, M.; Noboru, N. Monitoring of Wheat Growth Status and Mapping of Wheat Yield’s within-Field Spatial Variations Using Color Images Acquired from UAV-camera System. Remote Sens. 2017, 9, 289. [Google Scholar] [CrossRef] [Green Version]
  23. Liu, L.Y.; Wang, J.H.; Song, X.Y.; Li, C.J.; Huang, W.J.; Zhao, C.J. The canopy spectral features and remote sensing of wheat lodging. J. Remote Sens. 2005, 9, 323. [Google Scholar]
  24. Chauhan, S.; Darvishzadeh, R.; Lu, Y.; Boschetti, M.; Nelson, A. Understanding wheat lodging using multi-temporal Sentinel-1 and Sentinel-2 data. Remote Sens. Environ. 2020, 243, 111804. [Google Scholar] [CrossRef]
  25. Carter, G.A.; McCain, D.C. Relationship of leaf spectral reflectance to chloroplast water content determined using NMR microscopy. Remote Sens. Environ. 1993, 46, 305–310. [Google Scholar] [CrossRef]
  26. Miphokasap, P.; Honda, K.; Vaiphasa, C.; Souris, M.; Nagai, M. Estimating Canopy Nitrogen Concentration in Sugarcane Using Field Imaging Spectroscopy. Remote Sens. 2012, 4, 1651–1670. [Google Scholar] [CrossRef] [Green Version]
  27. Yang, H.; Chen, E.; Li, Z.; Zhao, C.; Yang, G.; Pignatti, S.; Casa, R.; Zhao, L. Wheat lodging monitoring using polarimetric index from RADARSAT-2 data. Int. J. Appl. Earth Obs. Geoinf. 2015, 34, 157–166. [Google Scholar] [CrossRef]
  28. Murakami, T.; Yui, M.; Amaha, K. Canopy height measurement by photogrammetric analysis of aerial images: Application to buckwheat (Fagopyrum esculentum Moench) lodging evaluation. Comput. Electron. Agric. 2012, 89, 70–75. [Google Scholar] [CrossRef]
  29. Bendig, J.; Yu, K.; Aasen, H.; Bolten, A.; Bennertz, S.; Broscheit, J.; Gnyp, M.L.; Bareth, G. Combining UAV-based plant height from crop surface models, visible, and near infrared vegetation indices for biomass monitoring in barley. Int. J. Appl. Earth Obs. Geoinf. 2015, 39, 79–87. [Google Scholar] [CrossRef]
  30. Du, M.M.; Noboru, N. Multi-temporal monitoring of wheat growth through correlation analysis of satellite images, Unmanned aerial vehicle images with ground variable. In Proceedings of the 5th IFAC Conference on Sensing, Control and Automation Technologies for Agriculture, Seattle, WA, USA, 14–17 August 2016; pp. 14–17. [Google Scholar]
  31. Zhang, C.; Walters, D.; Kovacs, J. Applications of Low Altitude Remote Sensing in Agriculture upon Farmers’ Requests– A Case Study in Northeastern Ontario, Canada. PLoS ONE 2014, 9, e112894. [Google Scholar] [CrossRef] [PubMed]
  32. Rajapaksa, S.; Eramian, M.; Duddu, H.; Wang, M.; Shirtliffe, S.; Ryu, S.; Josuttes, A.; Zhang, T.; Vail, S.; Pozniak, C.; et al. Classification of crop lodging with gray level co-occurrence matrix. In Proceedings of the 2018 IEEE Winter Conference on Applications of Computer Vision (WACV), Lake Tahoe, NV, USA, 12–15 March 2018; pp. 251–258. [Google Scholar]
  33. Lu, Y.; Lu, R. Detection of surface and subsurface defects of apples using structured-illumination reflectance imaging with machine learning algorithms. Trans. ASABE 2018, 61, 1831–1842. [Google Scholar] [CrossRef]
  34. Naik, D.L.; Kiran, R. Identification and characterization of fracture in metals using machine learning based texture recognition algorithms. Eng. Fract. Mech. 2019, 219, 106618. [Google Scholar] [CrossRef]
  35. Marsland, S. Machine Learning: An Algorithmic Perspective; CRC Press: Boca Raton, FL, USA, 2015. [Google Scholar]
  36. Krizhevsky, A.; Sutskever, I.; Hinton, G.E. Imagenet classification with deep convolutional neural networks. In Proceedings of the Advances in Neural Information Processing Systems, Lake Tahoe, NV, USA, 3–6 December 2012; pp. 1097–1105. [Google Scholar]
  37. LeCun, Y.; Bottou, L.; Bengio, Y.; Haffner, P. Gradient-based learning applied to document recognition. Proc. IEEE 1998, 86, 2278–2324. [Google Scholar] [CrossRef] [Green Version]
  38. Carvalho, T.; De Rezende, E.R.; Alves, M.T.; Balieiro, F.K.; Sovat, R.B. Exposing computer generated images by eye’s region classification via transfer learning of VGG19 CNN. In Proceedings of the 2017 16th IEEE International Conference on Machine Learning and Applications (ICMLA), Cancun, Mexico, 18–21 December 2017; pp. 866–870. [Google Scholar]
  39. Bharath, R. A Simple Guide to the Versions of the Inception Network. 2018. Available online: https://towardsdatascience.com/a-simple-guide-to-the-versions-of-the-inception-network-7fc52b863202 (accessed on 15 April 2020).
  40. Mohanty, S.P.; Hughes, D.P.; Salathé, M. Using Deep Learning for Image-Based Plant Disease Detection. Front. Plant Sci. 2016, 7, 1419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Jia, L.; Buerkert, E.S.A.; Chen, X.; Roemheld, V.; Zhang, F. Low-altitude aerial photography for optimum N fertilization of winter wheat on the North China Plain. Field Crop. Res. 2004, 89, 389–395. [Google Scholar] [CrossRef]
  42. Haralick, R.M.; Shanmugam, K.; Dinstein, I.H. Textural features for image classification. IEEE Trans. Syst. Man Cybern. 1973, 6, 610–621. [Google Scholar] [CrossRef] [Green Version]
  43. Soh, L.-K.; Tsatsoulis, C. Texture analysis of SAR sea ice imagery using gray level co-occurrence matrices. IEEE Trans. Geosci. Remote Sens. 1999, 37, 780–795. [Google Scholar] [CrossRef] [Green Version]
  44. Ojala, T.; Pietikäinen, M.; Mäenpää, T. Multiresolution gray-scale and rotation invariant texture classification with local binary patterns. IEEE Trans. Pattern Anal. Mach. Intell. 2002, 24, 971–987. [Google Scholar] [CrossRef]
  45. Kumar, A.; Pang, G.K.H. Defect detection in textured materials using Gabor filters. IEEE Trans. Ind. Appl. 2002, 38, 425–440. [Google Scholar] [CrossRef] [Green Version]
  46. Nixon, M.S.; Aguado, A.S. Feature Extraction and Image Processing for Computer Vision, 3rd ed.; Elsevier: Singapore, 2012. [Google Scholar]
  47. Hu, M.-K. Visual pattern recognition by moment invariants. IEEE Trans. Inf. Theory 1962, 8, 179–187. [Google Scholar] [CrossRef] [Green Version]
  48. Zhou, Z.H. Ensemble Methods: Foundations and Algorithms; CRC Press: Boca Raton, FL, USA, 2012. [Google Scholar]
  49. Pal, M. Random forest classifier for remote sensing classification. Int. J. Remote Sens. 2005, 26, 217–222. [Google Scholar] [CrossRef]
  50. Liaw, A.; Wiener, M. Classification and regression by randomForest. R News 2002, 2, 18–22. [Google Scholar]
  51. Breiman, L. Random Forests—Random Features; Technical Report 567; Statistics Department, University of California: Berkeley, CA, USA, 1999; Available online: https://www.stat.berkeley.edu/~breiman/random-forests.pdf (accessed on 10 April 2020).
  52. Fukushima, K. Neocognitron: A hierarchical neural network capable of visual pattern recognition. Neural Netw. 1988, 1, 119–130. [Google Scholar] [CrossRef]
  53. Hajian, A.; Styles, P. Application of Soft Computing and Intelligent Methods in Geophysics; Springer International Publishing: New York, NY, USA, 2018; pp. 25–26. [Google Scholar]
  54. Han, S.; Qubo, C.; Meng, H. Parameter selection in SVM with RBF kernel function. In Proceedings of the World Automation Congress 2012, Puerto Vallarta, Mexico, 24–28 June 2012; IEEE: New York, NY, USA, 2012; pp. 1–4. [Google Scholar]
  55. Amari, S.-I.; Wu, S. Improving support vector machine classifiers by modifying kernel functions. Neural Netw. 1999, 12, 783–789. [Google Scholar] [CrossRef]
  56. Shao, Y.-H.; Chen, W.-J.; Deng, N.-Y. Nonparallel hyperplane support vector machine for binary classification problems. Inf. Sci. 2014, 263, 22–35. [Google Scholar] [CrossRef]
  57. Sumit, S. A Comprehensive Guide to Convolutional Neural Networks—The ELI5 Way. Towards Data Science. 2018. Available online: https://towardsdatascience.com/a-comprehensive-guide-to-convolutional-neural-networks-the-eli5-way-3bd2b1164a53 (accessed on 1 April 2020).
  58. Simard, P.Y.; Steinkraus, D.; Platt, J.C. Best practices for convolutional neural networks applied to visual document analysis. In Proceedings of the 7th International Conference Document Analysis and Recognition (ICDAR 2003), Edinburgh, Scotland, UK, 3–6 August 2003; IEEE Computer Society: Piscataway, NJ, USA; pp. 958–962. [Google Scholar]
  59. Szegedy, C.; Liu, W.; Jia, Y.; Sermanet, P.; Reed, S.; Anguelov, D.; Erhan, D.; Vanhoucke, V.; Rabinovich, A. Going deeper with convolutions. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition, Boston, MA, USA, 7–12 June 2015; pp. 1–9. [Google Scholar]
  60. Brownlee, J. How to Develop VGG, Inception and ResNet Modules from Scratch in Keras. 2019. Available online: https://machinelearningmastery.com/how-to-implement-major-architecture-innovations-for-convolutional-neural-networks/ (accessed on 5 April 2020).
  61. Akosa, J. Predictive accuracy: A misleading performance measure for highly imbalanced data. In Proceedings of the SAS Global Forum, Orlando, FL, USA, 2–5 April 2017; pp. 2–5. [Google Scholar]
  62. Kumpumäki, T.; Linna, P.; Lipping, T. Crop Lodging Analysis from Uas Orthophoto Mosaic, Sentinel-2 Image and Crop Yield Monitor Data. In Proceedings of the IGARSS 2018–2018 IEEE International Geoscience and Remote Sensing Symposium, Valencia, Spain, 22–27 July 2018; pp. 7723–7726. [Google Scholar]
  63. Zhu, J.; Cen, H.; He, L. Development and performance evaluation of a multi-rotor unmanned aircraft system for agricultural monitoring. Smart Agric. 2019, 1, 43. [Google Scholar]
  64. Cen, H.; Wan, L.; Zhu, J.; Li, Y.; Li, X.; Zhu, Y.; Haiyong, W.; Wu, W.; Yin, W.; Xu, C.; et al. Dynamic monitoring of biomass of rice under different nitrogen treatments using a lightweight UAV with dual image-frame snapshot cameras. Plant Methods 2019, 15, 32. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overall process flowchart of lodging detection using traditional machine learning and deep learning with visual field observation results.
Figure 1. Overall process flowchart of lodging detection using traditional machine learning and deep learning with visual field observation results.
Remotesensing 12 01838 g001
Figure 2. Location and details of the experimental wheat field layout established near Thompson County, ND (UTM WGS 84 14 N), USA with pre-installed control points. ID1, ID2, and ID3 are identification numbers of research field blocks of 204, 120, and 48 plots of different dimensions.
Figure 2. Location and details of the experimental wheat field layout established near Thompson County, ND (UTM WGS 84 14 N), USA with pre-installed control points. ID1, ID2, and ID3 are identification numbers of research field blocks of 204, 120, and 48 plots of different dimensions.
Remotesensing 12 01838 g002
Figure 3. Field wheat lodging data collection: left, aerial imagery collection; middle, manual/visual evaluation of plot lodging or non-lodging; right, UAS image samples of lodging and non-lodging plots.
Figure 3. Field wheat lodging data collection: left, aerial imagery collection; middle, manual/visual evaluation of plot lodging or non-lodging; right, UAS image samples of lodging and non-lodging plots.
Remotesensing 12 01838 g003
Figure 4. Random forest classifier classification based on the majority vote.
Figure 4. Random forest classifier classification based on the majority vote.
Remotesensing 12 01838 g004
Figure 5. Number of grown trees vs. out-of-bag classification (OOBC) error results generated from four random forest model replicate runs.
Figure 5. Number of grown trees vs. out-of-bag classification (OOBC) error results generated from four random forest model replicate runs.
Remotesensing 12 01838 g005
Figure 6. Pattern recognition neural network used in this study for wheat lodging detection, consisting of two hidden layers (10 and 5 neurons) with input (320 features), and output (lodging and non-lodging). w and b stand for weight matrix and bias, respectively.
Figure 6. Pattern recognition neural network used in this study for wheat lodging detection, consisting of two hidden layers (10 and 5 neurons) with input (320 features), and output (lodging and non-lodging). w and b stand for weight matrix and bias, respectively.
Remotesensing 12 01838 g006
Figure 7. Geometric illustration of support vector machine principle applied to lodging and non-lodging binary classification: (a) Linear SVM; (b1) and (b2) Non-linear SVM; (c1) and (c2) Kernel method for non-linear classification.
Figure 7. Geometric illustration of support vector machine principle applied to lodging and non-lodging binary classification: (a) Linear SVM; (b1) and (b2) Non-linear SVM; (c1) and (c2) Kernel method for non-linear classification.
Remotesensing 12 01838 g007
Figure 8. Original UAS image and three augmented images generated by geometric transformations.
Figure 8. Original UAS image and three augmented images generated by geometric transformations.
Remotesensing 12 01838 g008
Figure 9. Architecture of a simple convolutional neural network. The input for each sample is a 3-D tensor (height, width, and channel), with channels corresponding to R, G, and B channels. The two output neurons represent two categories (lodging and non-lodging).
Figure 9. Architecture of a simple convolutional neural network. The input for each sample is a 3-D tensor (height, width, and channel), with channels corresponding to R, G, and B channels. The two output neurons represent two categories (lodging and non-lodging).
Remotesensing 12 01838 g009
Figure 10. Architecture of VGG-16 net showing 13 convolutional layers (Conv #), 5 pooling layers (Pool #), and 3 fully connected (fc#) layers.
Figure 10. Architecture of VGG-16 net showing 13 convolutional layers (Conv #), 5 pooling layers (Pool #), and 3 fully connected (fc#) layers.
Remotesensing 12 01838 g010
Figure 11. Procedure of updating VGG-16 for lodging detection. * Early layers learn low-level features (e.g., edges and colors); # late layers learn task-specific features; *# new layers learn features related to lodging; ** training dataset are 260 randomly selected plots (70%) of the entire dataset; ## options set for training (e.g., data augmenter and epochs); #* testing dataset is the other part (112 plot images) of the entire dataset.
Figure 11. Procedure of updating VGG-16 for lodging detection. * Early layers learn low-level features (e.g., edges and colors); # late layers learn task-specific features; *# new layers learn features related to lodging; ** training dataset are 260 randomly selected plots (70%) of the entire dataset; ## options set for training (e.g., data augmenter and epochs); #* testing dataset is the other part (112 plot images) of the entire dataset.
Remotesensing 12 01838 g011
Figure 12. Inception module of GoogLeNet with three different size filters and one pooling layer.
Figure 12. Inception module of GoogLeNet with three different size filters and one pooling layer.
Remotesensing 12 01838 g012
Figure 13. Classification accuracies by random forest, neural network, and support vector machine for lodging detection on three different dates datasets, where PER, REC, OAC, and F1 denote precision, recall, overall accuracy, and F1 measurement, respectively.
Figure 13. Classification accuracies by random forest, neural network, and support vector machine for lodging detection on three different dates datasets, where PER, REC, OAC, and F1 denote precision, recall, overall accuracy, and F1 measurement, respectively.
Remotesensing 12 01838 g013
Figure 14. Accuracy comparisons of different dates datasets for different classifiers. Whiskers on bars represent two standard deviations calculated from 10 replicates. Bars with different letters are significantly different by Tukey’s test at 0.05 significance level.
Figure 14. Accuracy comparisons of different dates datasets for different classifiers. Whiskers on bars represent two standard deviations calculated from 10 replicates. Bars with different letters are significantly different by Tukey’s test at 0.05 significance level.
Remotesensing 12 01838 g014
Figure 15. Accuracy comparisons of three classifiers (random forest, neural network, and support vector machine) on individual date datasets. Whiskers on bars represent two standard deviations calculated from 10 replicates. Bars with different letters are significantly different by Tukey’s test at 0.05 significance level.
Figure 15. Accuracy comparisons of three classifiers (random forest, neural network, and support vector machine) on individual date datasets. Whiskers on bars represent two standard deviations calculated from 10 replicates. Bars with different letters are significantly different by Tukey’s test at 0.05 significance level.
Remotesensing 12 01838 g015
Figure 16. Accuracy comparisons of different dates datasets for different deep learning algorithms (simple convolutional neural network, VGG-16, and GoogLeNet). Whiskers on bars represent two standard deviations calculated from 10 replicates. Bars with different letters are significantly different by Tukey’s test at 0.05 significance level.
Figure 16. Accuracy comparisons of different dates datasets for different deep learning algorithms (simple convolutional neural network, VGG-16, and GoogLeNet). Whiskers on bars represent two standard deviations calculated from 10 replicates. Bars with different letters are significantly different by Tukey’s test at 0.05 significance level.
Remotesensing 12 01838 g016
Figure 17. Accuracy comparisons of three deep learning algorithms (simple convolutional neural network, VGG-16, and GoogLeNet) on individual date datasets. Whiskers on bars represent two standard deviations calculated from 10 replicates. Bars with different letters are significantly different by Tukey’s test at 0.05 significance level.
Figure 17. Accuracy comparisons of three deep learning algorithms (simple convolutional neural network, VGG-16, and GoogLeNet) on individual date datasets. Whiskers on bars represent two standard deviations calculated from 10 replicates. Bars with different letters are significantly different by Tukey’s test at 0.05 significance level.
Remotesensing 12 01838 g017
Table 1. Accuracy fluctuation of four parameters for three classifiers.
Table 1. Accuracy fluctuation of four parameters for three classifiers.
Performance MetricsRandom ForestNeural NetworkSupport Vector MachineAverage
Precision1.40%7.83%4.84%4.69%
Recall14.94%25.00%22.68%20.87%
Overall Accuracy3.23%6.01%4.72%4.66%
F1 Score8.67%16.62%13.13%12.81%
Table 2. Accuracy comparisons of Random Forest (RF) and GoogLeNet on wheat lodging plot detection for three individual date datasets. Different letters in the same date are significantly different by Tukey’s test at 0.05 significance level.
Table 2. Accuracy comparisons of Random Forest (RF) and GoogLeNet on wheat lodging plot detection for three individual date datasets. Different letters in the same date are significantly different by Tukey’s test at 0.05 significance level.
July 23rd 2019July 30th 2019Aug 8th 2019
ClassifierRFGoogLeNetRFGoogLeNetRFGoogLeNet
Average Accuracy89% + 1% a91% + 2% a89% + 1% a89% + 2% a91% + 1% a93% + 4% a

Share and Cite

MDPI and ACS Style

Zhang, Z.; Flores, P.; Igathinathane, C.; L. Naik, D.; Kiran, R.; Ransom, J.K. Wheat Lodging Detection from UAS Imagery Using Machine Learning Algorithms. Remote Sens. 2020, 12, 1838. https://0-doi-org.brum.beds.ac.uk/10.3390/rs12111838

AMA Style

Zhang Z, Flores P, Igathinathane C, L. Naik D, Kiran R, Ransom JK. Wheat Lodging Detection from UAS Imagery Using Machine Learning Algorithms. Remote Sensing. 2020; 12(11):1838. https://0-doi-org.brum.beds.ac.uk/10.3390/rs12111838

Chicago/Turabian Style

Zhang, Zhao, Paulo Flores, C. Igathinathane, Dayakar L. Naik, Ravi Kiran, and Joel K. Ransom. 2020. "Wheat Lodging Detection from UAS Imagery Using Machine Learning Algorithms" Remote Sensing 12, no. 11: 1838. https://0-doi-org.brum.beds.ac.uk/10.3390/rs12111838

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