Next Article in Journal
A Preliminary Study of Three-Point Onboard External Calibration for Tracking Radiometric Stability and Accuracy
Next Article in Special Issue
An Under-Ice Hyperspectral and RGB Imaging System to Capture Fine-Scale Biophysical Properties of Sea Ice
Previous Article in Journal
Land Cover Change Detection from High-Resolution Remote Sensing Imagery Using Multitemporal Deep Feature Collaborative Learning and a Semi-supervised Chan–Vese Model
Previous Article in Special Issue
Quantitative Estimation of Soil Salinity Using UAV-Borne Hyperspectral and Satellite Multispectral Images
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tree Species Classification Based on Hybrid Ensembles of a Convolutional Neural Network (CNN) and Random Forest Classifiers

1
Fraunhofer Institute for Factory Operation and Automation IFF, 39106 Magdeburg, Germany
2
Otto von Guericke University, Faculty of Computer Science, 39106 Magdeburg, Germany
3
Forstliches Forschungs- und Kompetenzzentrum, ThüringenForst AöR, 99867 Gotha, Germany
*
Author to whom correspondence should be addressed.
Remote Sens. 2019, 11(23), 2788; https://doi.org/10.3390/rs11232788
Submission received: 29 October 2019 / Revised: 21 November 2019 / Accepted: 22 November 2019 / Published: 26 November 2019

Abstract

:
In this paper, we evaluate different popular voting strategies for fusion of classifier results. A convolutional neural network (CNN) and different variants of random forest (RF) classifiers were trained to discriminate between 15 tree species based on airborne hyperspectral imaging data. The spectral data was preprocessed with a multi-class linear discriminant analysis (MCLDA) as a means to reduce dimensionality and to obtain spatial–spectral features. The best individual classifier was a CNN with a classification accuracy of 0.73 +/− 0.086. The classification performance increased to an accuracy of 0.78 +/− 0.053 by using precision weighted voting for a hybrid ensemble of the CNN and two RF classifiers. This voting strategy clearly outperformed majority voting (0.74), accuracy weighted voting (0.75), and presidential voting (0.75).

Graphical Abstract

1. Introduction

Tree species classification is a challenging and important task for large-area monitoring and managing of forests. For many applications, it is an important step to first retrieve the tree species in order to enable the detection of specific traits such as nutrition state, water content, various stresses, diseases, and other relevant parameters in a species-specific context.
In 2015, Fraunhofer IFF conducted measurement flights over forests in Saxony Anhalt and Thuringia, Germany for two reasons: (1) to develop new methods for detection of biotic stresses related to oak feeding society (green oak-leaf roller, Tortrix viridana), mottled umber (Erannis defoliaria), and winter moth (Operophthera brumata); and (2) to support ground-based forest inventory by airborne assessment of tree species and tree vitality. Both require reliable detection of oak trees as well as other tree species in mixed forests. Hyperspectral imaging was chosen as a means to address these challenging tasks. It is an emerging measurement technique frequently used for optical non-invasive characterization of surfaces and materials. In contrast to the analysis of conventional digital images, multiple image channels that correspond to reflected or transmitted light of a certain small wavelength range, must be taken into account. A typical hyperspectral image consists of hundreds of image bands. A good introduction to hyperspectral image classification was provided by Bioucas-Dias et al. [1]. Challenges mentioned are the limited number of labeled samples and the need to combine spectral and spatial information to obtain good results. Typically, the high-dimensional nature of the data and strong correlations among spectral bands require the extraction of suitable features prior to training of a classifier. The authors also highlight the importance of high spatial image resolution, as most classification techniques assume a single predominant spectral signature per pixel. In contrast, a low image resolution leads to mixed pixels, which requires spectral unmixing as an additional preprocessing step [2].
Wang et al. [3] dedicated a chapter of their book “Remote Sensing of Natural Resources” to the classification of tree species. They listed principal component analysis (PCA), minimum noisef fraction (MNF), canonical discriminant analysis, partial least squares regression (PLS), and wavelet transform as common feature extraction methods. They also provide references where these techniques have been applied. Moreover, an overview on the application of classifiers containing support vector machines (SVM), classification and regression trees (CART), and artificial neural networks (ANN) is given. Fassnacht et al. [4] published a recent survey on tree species classification. The authors considered different types of discriminant analysis (linear, quadratic, canonical, stepwise, regularized, and penalized), SVM, random forest, and maximum likelihood classifiers. With reference to experiences from previous work (e.g., [5]), the authors draw the conclusion that the choice of the classifier is less important than adequate data preprocessing to obtain good results.
Féret et al. [6] investigated the discrimination of 17 tree species in tropical forests. A comparison of different parametric and non-parametric methods showed that SVM with linear or radial basis function kernels outperformed other classifiers given a large number of training samples. For smaller training sets, the authors recommend regularized discriminant analysis. Richter et al. [7] introduced a modified version of discriminant analysis based on PLS. The authors compared their approach to random forests and SVMs on an airborne hyperspectral dataset recorded over a German forest area and obtained better results compared to these classifiers. In this study, random forests performed notably worse than SVMs. To a smaller extent, this is also the result of studies in the Southern Alps conducted by Dalponte et al. [8]. The authors show that the overall accuracy of tree species identification increased by the integration of LiDAR data in addition to hyperspectral measurements.
Xia et al. [9] have demonstrated successful usage of an ensemble classifier on hyperspectral benchmark datasets. In addition, the concept of rotation forests, originally introduced by Rodriguez et al. [10], was successfully applied using CARTs as base classifiers. Compared to MNF, independent component analysis (ICA), and local Fisher discriminant analysis (LFDA) as transformation approaches, the choice of PCA yields the highest accuracy in their studies.
Instead of searching for a favorable transformation of the feature space, certain spectral bands can be selected based on close correspondence to biochemical traits and physical properties of the observed plants. This concept led to a variety of general or trait-specific vegetation indices. Lausch et al. [11] list spectral indices related to greenness and other vitality parameters, mainly using bands in the spectral range of 450–890 nm. A similar overview can be found in a disease detection report published by Sankaran et al. [12]. Mutanga et al. [13] investigated the relationship between water content and spectral variables. Besides known water absorption features located at three different bands, related spectral indices (NDWI [14], WI [15]) and continuum-removed features were included in their correlation analysis.
Airborne hyperspectral imaging generates a large amount of data of which usually only a small fraction is labeled with reference data. Therefore, semi-supervised approaches for classifier training aim to make use of unlabeled data for hyperspectral image classification as well. Here, the term semi-supervised is used for all approaches that includes both labeled and unlabeled data in the training of classifiers to enhance the classification performance. It emphasizes the fact that reference data is still required, but a much larger dataset contributes to the final classifier.
Ayerdi et al. [16] describe a data augmentation method as a means to use the information of unlabeled data. After performing a clustering in the spectral domain, labeled pixels propagate their label to neighboring unlabeled pixels in the spatial domain if they belong to the same cluster.
Wenzhi et al. [17] introduce a semi-supervised feature extraction algorithm named semi-supervised local discriminant analysis (SELD). The combination of an unsupervised local linear feature extraction method with linear discriminant analysis (LDA) results in a projection that separates the different classes while preserving local neighborhoods. In their evaluation, SELD is compared to different feature extraction methods such as PCA and LDA and obtained the best results on several datasets with different classifiers.
Since SVMs are reported to perform well on hyperspectral datasets, an extension to the semi-supervised case is desirable. Vapnik et al. [18] proposed the idea of transductive SVMs, which was applied to hyperspectral data later on (e.g., by Bruzzone et al. [19]).
It is common practice to use both spectral and spatial information for classification. For the problem of tree species classification, a popular approach is to identify individual tree crowns (ITCs), before starting the classification process. Féret et al. [6] used mean shift clustering to segment ITCs and evaluated two different classification methods: object-based classification of the mean spectrum per ITC and majority class assignment based on the decisions per pixel within the ITC. The authors report an increased accuracy for both approaches compared to pixelwise classification without prior tree crown segmentation. Dalponte et al. [20] incorporated the assumption that all pixels belonging to an ITC should have the same label within the training process of a semi-supervised SVM. An improvement over conventional SVMs was observed, especially if the number of training samples is small. The problems related to small numbers of training samples can also be effectively addressed by tensor-based linear and non-linear models as proposed by Makantasis et al. in [21].
Ghamisi et al. [22] proposed a spectral-spatial classifier based on hidden Markov random fields and SVMs. This fully automatic approach performed better than SVMs on widely used datasets. Li et al. [23] integrate spectral and spatial information in a robust Bayesian framework, addressing problems related to noise, the presence of mixed pixels, and small number of training samples. The authors evaluated their algorithm along with several state-of-the-art methods and achieved notably better results in terms of accuracy. Zhang et al. [24] introduced a sparse ensemble learning method, which allows for information sharing between neighboring pixels during the optimization process. While any classifier can potentially be used, the experiments in this study used CARTs. The authors report not only better accuracy, but also a lower runtime of the trained ensemble compared to traditional ensemble methods.
Since the extraction of handcrafted features is time consuming and complex, Makantasis et al. [25] suggest a convolutional neural network (CNN) to automatically construct high-level features. The dimensionality of the input data is reduced by a PCA first and then three convolutional layers hierarchically detect features, which are classified by two fully connected layers in the end. The CNN outperformed different types of SVMs on benchmark datasets.
Ayerdi et al. [16] suggested a regularization step after classification: each pixel adapts its label to the majority class in its neighborhood. If rather homogeneous tree species distributions in the area under consideration can be assumed, this technique leads to more plausible results.
The review of the literature shows a number of promising approaches to classify tree species from hyperspectral imaging data. It was especially shown that recent advances in image classification with CNNs are transferable to hyperspectral image classification. However, for any new application and dataset, several state-of-the-art classifiers and different feature spaces in question must be tested to prove the suitability and superiority of an approach [26]. Given this requirement, the question arises: what benefits the combination of the best available classification techniques and turns them into an ensemble classifier capable of providing for the problem of tree species classification? To answer this question, we focus on popular voting strategies, which are easy to apply to any group of classifiers.

2. Materials and Methods

2.1. Datasets

Figure 1 illustrates the acquisition of the datasets. Hyperspectral imaging data has been recorded during two measurement flights over forests in Saxony Anhalt and Thuringia (Germany) to determine tree species and investigate oak tree specific biotic stresses. Used cameras were NEO Hyspex VNIR 1600 and NEO Hyspex SWIR 320m-e. Cameras and inertial measurement unit (IMU) Novatel SPAN CPT were fixed on a stabilized mount (see Figure 1). The recording of the flight path and orientation of the line scanning camera systems by the IMU is crucial for later alignment of the scanlines into an image. Experiments reported here are based on Hyspex VNIR 1600 images recorded during the second flight. The reasons are to minimize potential errors from alignment and interpolation of the low resolution Hyspex SWIR 320m-e camera to match the grid of the VNIR camera, better weather conditions during the second flight (no cloud cover), and the recovery of diseased oak trees after secondary ‘lammas’ shoots [27], which is expected to ease tree species classification.
Orthorectification of the line scanning data was done with parametric geocoding using the software PARGE [28]. Radiometric corrections were performed with the software ATCOR-4 [29].
One of the major challenges while classifying tree species is the availability of training data that represents differences between individual trees of one species, differences between development stages of individual trees as well as the differences between tree species. In order to account for the variability and to minimize the necessary fieldwork, existing databases and high-resolution aerial images were used to determine areas with a single dominant tree species. Figure 2 illustrates the results of the selection process. As shown, potential error sources such as boundary regions with a mix of tree species, individual trees and clearings were excluded to compensate for potential minor misalignment of hyperspectral images as well as discrepancies due to changes between the available historical images and the current state of the forest cover. Given the locations and shapes of the reference sites, the training data can be derived directly from the hyperspectral images. From 23 known tree species at the study sites, a subset of 15 species is represented by sufficiently large reference sites and was therefore selected to develop and evaluate classifiers. In Table 1, these tree species are listed.

2.2. Features

As emphasized in the introduction, the selection of an appropriate feature space is crucial for successful classification. By operating in different feature spaces, diversity among the classifiers should increase in many cases. This is a crucial concept in the design of ensemble classifiers [30]. In this study, 15 different feature spaces have been selected to investigate these dependencies between feature spaces and classifiers in terms of improvements in the classification performance. Our choice of feature spaces is explained below.
The measured and preprocessed spectral data itself is a powerful data source for separation of image pixels into classes. Hence, the first feature space is the raw 160-dimensional reflectance data vector per image pixel.
A vector containing a number of spectral indices (calculated based on the reflectance spectra) is used as a second feature space. As presented in the introduction, these indices have been developed to correlate with selected physiological parameters of vegetation, vegetation health, and vegetation nutrition. The following indices have been chosen and are used as features: DI1, GNDVI, MCARI, NDVI, PRI, and WI (see Table 2 for definitions and references).
Classification based on raw spectral data, where individual image bands are simply treated as features as well as using spectral indices as features, represent the most common approaches to the analysis of hyperspectral imaging data. Here, both approaches serve just as a baseline reference for the performance of tree species classification.
The high dimensionality of hyperspectral data imposes a number of problems, which are summarized as the curse of dimensionality in the literature [31]. Therefore, we included approaches for reduction of dimensionality prior to classifier training. Principal component analysis (PCA) is performed to calculate a projection into an orthogonal low dimensional subspace, which covers most of the variation of the original high dimensional spectral dataset. We chose 5 and 14 as the numbers of dimensions in the low dimensional representation. This choice is motivated by the number of most prominent tree species in the observed area. Using reference sites with 15 different tree species, 14 is the number of required dimensions for LDA. On the other hand, the choice of five dimensions is motivated by PCA, where the five largest Eigenvalues explain 99.5% of the variation within the spectral dataset.
Multi-class linear discriminant analysis (MCLDA) aims to generate a low dimensional representation, where a single dimension is a projection allowing a good discrimination between elements of one class and all the others. As mentioned above, 5 and 14 dimensions are used again.
SELD was included as a semi-supervised feature extraction technique. SELD combines the supervised method of LDA with the unsupervised method of local linear embedding (LLE). LLE represents the instances in a graph structure and computes a projection that preserves the local neighborhood of instances in the feature space. The combination of supervised and unsupervised feature extraction enables SELD to find features that maximize the class separation and preserve local neighborhoods in the feature space. Therefore, SELD depends on two parameters: the number of instances to consider for local neighborhoods and the number of extracted features.
Table 3 gives an overview of the above listed feature spaces. In addition to the pixelwise transformation of the spectral data into these representations, so-called spatial-spectral features are derived. Within the predefined square image blocks with a side length of s = 2r + 1, the statistical measures mean, standard deviation and homogeneity are calculated for each feature and used as new features, instead. This approach combines the information from the low dimensional representations of the spectral data with its local spatial distribution. We set r = 5 which corresponds to a side length 4.4 m given the campaigns ground sampling distance of 0.4 m per pixel. This size matches the area of typical tree crowns of boreal trees within the recorded hyperspectral images.

2.3. Classifiers

In addition to the different feature spaces, a number of classifiers were selected to find the best classifier/feature space combination for the task of tree species classification.
Differently parameterized random forests, rotation forests, SVMs, and a CNN provide state-of-the-art classification results as well as a potentially diverse pool of classifiers for integration into an ensemble classifier.
Random forests are a realization of the concept of bootstrap-aggregation (bagging) with decision trees [32]. As single decision trees tend to overfit, the bagging method creates several bootstrapped sample sets from the original data and trains one tree on each sample set. The resulting ensemble of trees can classify a new instance via majority voting. An important parameter here is the number of trees to be trained. This classifier was chosen because it is a standard approach in literature, when dealing with the classification of tree species from hyperspectral data.
Rotation forests are an advanced form of random forests and were proposed by Rodriguez et al. in 2006 [10]. We decided to test this classifier because of the good results for hyperspectral data as previously reported [9]. As in random forests, the trees are trained on bootstrapped sample sets of the original data. The difference is that the bootstrapping is not performed on a subset of the original data, but on the whole data set with only a subset of features. The feature space is divided into k subsets. For each subset, all instances of randomly drawn classes are deleted from the training data and a bootstrapped sample set with a size of 75% of the original data is created. PCA is performed on each sample set and the resulting coefficient matrices are merged into a single one. This is equivalent to k axis rotations. A decision tree is then trained on the whole PCA-transformed data set. This procedure is repeated for each tree in the ensemble. The two parameters this classifier provides are the number of trees in the ensemble and the number of subsets the feature space is divided into.
SVMs are another standard classification and regression approach. We therefore chose to include them in our experiments. An SVM tries to fit a hyperplane that separates two classes while maximizing the margin between the instances and the plane. The plane can then be described by the instances nearest to it. These instances are called support vectors. To classify more than two classes, one can either train an SVM for each pair of classes (1 versus 1) or each class against the rest (1 versus All). The final decision is then found by majority voting (1 versus 1) or by taking the result of the classifier with the highest confidence (1 versus All). In order to classify not linearly separable classes, the data can be transformed into a higher dimensional space where it is linearly separable. To make this method computable in appropriate time one can use the so-called kernel trick, where the kernel is a type of nonlinear transformation. As the number of adjustable parameters for an SVM is very large, we only chose to vary the kernel (linear or polynomial) and whether to use 1 versus 1 or 1 versus All. We tested several SVM implementations ranging from standard MATLAB classes to native C libraries connected to MATLAB via the MEX interface. We settled for the LibLinear library [33] from the National University of Taiwan for 1 versus All and SVMlin [34] by Vikas Sindhwani for 1 versus 1, both specializing in linear SVM solving. Preliminary tests revealed that polynomial SVMs did not converge in appropriate time and we therefore abandoned this kernel from further investigation.
Deep neural networks are a recent hot topic in machine learning. As Makantasis et al. [25] were able to produce good results on hyperspectral data with convolutional deep neural networks, we added this classification method to our experimental setup. Neural networks are inspired by the structure of the human brain to learn classification and regression tasks. A neural network consists of several layers of perceptron units that are interconnected. Via backpropagation, this network can learn from examples. A deep neural network has significantly more layers than a conventional network. Deep convolutional networks use filters that are shifted over the input image to generate new and more high-level features in each layer. It is not possible to test or even enumerate all possible network structures and parameter configurations. We therefore employed the network structure shown in Figure 3, which is based on the one used in the original paper [25]. The network starts with a convolutional layer containing C 1 = 3 F trainable filters (block size 5 × 5 pixels), where F denotes the number of features. The second convolutional layer then contains C 2 = 3 C 1 trainable filters (block size 5 × 5). The output of this layer is then fed to a fully connected network with a hidden layer of size C 3 = 6 F . Additionally, we added the option to use a max-pooling layer in between the convolution layers and an option to apply a specified dropout rate to the fully connected network part. A pooling layer applies a filter mask for each pixel incorporating its surroundings. Common filters are mean or maximum filter. A dropout layer randomly omits a perceptron for the next training instance with a certain probability to avoid early convergence and to improve generalization. The deep learning library MatConvNet [35] provided all functions needed to build our networks.

2.4. Fusion Algorithms

Equations (1) and (2) describe the applied framework for fusion of different base classifier results. Given a feature vector   x i , its label l = y ˜ ( x i ) is calculated as follows.
Using Equation (1), a score s l is calculated for each class label from the results of the N base classifiers. For each label in question, the score is a sum of weights ω for the corresponding base classifiers. The indicator function I equals one, if its argument is true and zero, if not. Hence, only the weights of base classifiers with output label L are included.
s l ( x i ) = j = 1 N ω j I ( y j ( x i ) = l )
y ˜ ( x i ) = argmax l s l ( x i )
Equation (2) simply denotes the selection of the label with the highest score. Different voting strategies are implemented by variation of the weights   ω .
Majority voting is defined by setting i , j :   ω i = ω j . If a majority of the N base classifiers assigns the same label, the corresponding score s L   is maximized. In our study, majority voting serves as a baseline for the performance of classifier fusion as it is a commonly used method.
Presidential voting is defined by setting ω i = N 1.5 and j i :   ω j = 1 , where the index i denotes a prior chosen classifier, e.g., the one with the best overall accuracy. Due to this constraint, the label assigned by the chosen classifier determines the ensemble result in most cases. Only if all other classifiers agree on the same divergent result, a different label is assigned.
Accuracy weighted voting uses a classification accuracy estimate to weight the different classifiers. According to the definitions in [35], the average accuracy is calculated from the elements of the confusion matrix C of each classifier. A subset of the reference data is used to determine C and to estimate the weights ω .
Hence, for the j-th classifier the weight ω j is defined with Equation (3) as the average accuracy [36] which measures the average per-class effectiveness of a classifier.
ω j = i = 1 L t p i + t n i t p i + t n i + f p i + f n i L
The L × L confusion matrix C is transformed into L 2 × 2 confusion matrices to obtain the required true positive t p i , true negative t n i , false positive f p i , and false negative f n i values.
For precision weighted voting, a similar approach to calculate the weights is used. The precision measure is calculated with Equation (4) as follows
ω j = i = 1 L t p i i = 1 L ( t p i + f p i )
This fusion framework is a means to investigate how different popular approaches to weight a classifier influence the quality of the joint decision-making of an ensemble classifier.

3. Results

Table 4, Table 5, Table 6 and Table 7 summarize the results of SVM, random forest, rotation forest, and CNN classifiers. Each table contains the mean accuracy values and standard deviations obtained with multiple runs of holdout testing. In the tables, for each feature space the best results per table row are emphasized as bold text. In addition, the feature space with the best overall classification accuracy is emphasized and the corresponding accuracy has been underlined. In each run, a complete reference site with a single dominant tree species was only used for testing the classifier, which was trained with data from the remaining reference sites. See Table 1 for the total number of reference sites and areas. If only a single reference site is available it is divided into non-overlapping sites for training and validation of equal size. Each row corresponds to one of the selected feature spaces. We discriminate between the results of pixelwise and spatial classification, where statistical measures (mean, standard deviation, homogeneity) of the original features are used as features instead.
We report results for different ensemble sizes of random forest (Table 5) and rotation forest (Table 6) classifiers separately to assess the impact of this parameter and the trade-off for using a compact ensemble of 20 decision trees instead of a large ensemble of 100 decision trees.
Transformation of the reflectance data with MCLDA into a low dimensional representation yields the best results for all tested classifiers in separating between the 15 tree species. Moreover, inclusion of the spectral-spatial features significantly improves classification accuracy compared to pixelwise classification of spectral features in most cases. The best individual classifier was the CNN (see Table 7) with an overall accuracy of 0.732 ± 0.086, followed by a rotation forest of 100 decision trees with an overall accuracy of 0.705 ± 0.044. The CNN intrinsically combines spectral and spatial information. Hence, the CNN operating on raw or preprocessed hyperspectral image data achieved a similar performance to fandom rorests and SVMs, which make use of handcrafted spectral and spatial-spectral features. However, even the CNN the initial transformation with MCLDA was crucial to achieve the gain in overall accuracy.
The results of combining a small number of base classifiers into a hybrid ensemble are summarized in Figure 4. The diagram shows the accuracies and standard deviations of individual base classifiers together with results of different voting strategies. An accuracy gain of 5.1% was achieved by precision weighted voting compared to the best individual classifier. It significantly outperformed the three other tested voting methods.
To better understand these improvements, Figure 5 shows net diagrams of class-wise performance measures precision and recall. The subplots A–C show the performances of the included base classifiers, while subplot D represents the performance of the fusion framework with precision weighted voting using Equation (4) to calculate the weights.
The expected general improvement of class-wise precision values is shown by the more convex shape and the much larger area within the precision curve in subplot D. The recall value of the hybrid ensemble is still dominated by the best base classifier, the CNN (compare subplots B and D, dotted curves). However, a class0wise direct comparison reveals a few minor differences. For some tree species the recall values increase (e.g., poplar, robinia, oak), while for a few others (e.g., larch, Douglas dir) they decrease slightly.

4. Discussion

The need for a reliable tree species classification motivated our investigation of several state-of-the-art classifiers as well as their combination into an ensemble classifier. This demand has led to the development of a processing pipeline, where the positions of suitable training sites with one dominant tree species are obtained from existing databases of the forest authorities. Hence, it was possible to create a large training dataset covering 15 tree species without additional groundwork. The alternative, the assessment of tree species of individual trees in mixed forests by experts to create a reference database, either from ground or from aerial stereo imagery is expensive, time-consuming, and error-prone.
Although, we developed a promising framework for classifier fusion and did careful validation, the method was not validated against ground truth data from real mixed forests. This is due to the fact that a large-scale validation would require the same efforts to obtain the reference data as mentioned above for the classifier training. However, the method was successfully applied to classify the complete forest in our hyperspectral dataset. This tree species map can then be used to select a number of individual trees and to validate the assigned class label in the field.
The performance of classification models based on this kind of training data was extensively tested with a combination of cross-validation and hold-out-testing. Compared to standard cross-validation not only a subset of randomly chosen pixels or patches were excluded from training for testing, but complete areas. This allowed us to study the performances on independent data. However, the spectral data belongs to a single measurement flight and has undergone the same preprocessing (e.g., radiometric and atmospheric corrections). Hence, the trained classifiers are adapted to the development stages of the trees and the conditions on the day of flight. To our best knowledge, the proposed method to acquire training data directly from the hyperspectral images at predefined locations with dominant tree species is the best way to learn classification models for future flights. Otherwise, many test flights and expenses are required to acquire training data that cover all possible appearances of leaves and needles beforehand. Locations with a single dominant tree species can be determined by an expert using existing databases and aerial images. For any region, this could be done once stored in a database and then be used for future measurement flights.
The per-class assessment of the classification performance of the fusion approach shows differences depending on the tree species. The net diagram in Figure 5 reveals that the precision of the class assignment was improved to large extends, but for some tree species the fraction of successfully detected trees remains low. However, compared to the state-of-the-art CNN a significant loss is only observed for Douglas firs. On the other hand, the rate of true Douglas firs among the reported ones increased significantly. There are different reasons, which possibly explain this behavior. First, the dataset contains a number of tree species with only subtle differences like oaks, sessile oaks, red oaks, and pedunculate oaks. While this information is of interest for the forest authorities and motivates the use of hyperspectral imaging to detect subtle differences, it might be better to take a hierarchical approach, which first detects all oaks and then tries to discriminate between oak species. Second, some of the tree species are more common than others. We account for this by balancing the dataset to have an equal number of samples for all tree species. However, the number of reference sites for rare species is also low and the natural variation between trees is better covered for frequent tree species. Here, our strict validation with holding back complete reference sites for testing may penalize the rare species.
With Scotch pines, Norway spruces, oaks, and beeches being the most frequent tree species in Saxony-Anhalt as well as Thuringia the results show, that our approach for analysis of airborne hyperspectral images already provides a useful tool to support forest inventory and to detect oak trees for subsequent analysis of vitality parameters. Moreover, the proposed fusion framework allows to easily add any other classifier.

5. Conclusions

In this paper, we applied a general and easy-to-use fusion framework based on voting to the problem of tree species classification from hyperspectral aerial images. The proposed hybrid multiple classifier system enhances the results of a state-of-the-art CNN with two random forest classifiers of different size and operating in different feature spaces. It was shown that this approach yields a significant gain in overall classification accuracy. This improvement results from a gain in precision of the class assignments by weighted fusion of the CNN and random forest results by an estimate of their individual precisions. The comparison to other popular voting techniques showed the superiority of the approach. The results provide evidence that even the best available classifiers for image data analysis can be further improved by incorporating their decisions into a multiple classifier system. MCLDA performs best among the different dimensionality reduction methods for hyperspectral imaging data. Even the CNN performance is enhanced by using MCLDA transformed images instead of the hyperspectral images as input data.

Author Contributions

conceptualization, S.C., U.K., and U.S.; methodology, H.S. and U.K.; software, C.S.v.R., M.S., T.K.,T.P.M.,V.H., and U.K.; validation, U.K., C.S.v.R., M.S., T.K., T.P.M., and V.H.; formal analysis, U.K., C.S.v.R., M.S., T.K., T.P.M., and V.H.; investigation, U.K., D.K., C.S.v.R., M.S., T.K., T.P.M, and V.H.; resources, H.S. and U.K.; data curation, H.S. and U.K.; writing—original draft preparation, U.K.; writing—review and editing, U.K., C.S.v.R., U.S., H.S., and S.C.; visualization, U.K. and H.S.; supervision, U.S. and U.K.; project administration, I.E.; funding acquisition, I.E., U.K., and U.S.

Funding

This research was partially funded by Ministry of Environment, Agriculture and Food of Saxony-Anhalt under grants A02/2014 and A02/2016.

Acknowledgments

The authors would like to thank Landesforstbetrieb Sachsen-Anhalt, Thüringen Forst, and Landeszentrum Wald Sachsen-Anhalt for supporting the project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bioucas-Dias, J.M.; Plaza, A.; Camps-Valls, G.; Scheunders, P.; Nasrabadi, N.M.; Chanussot, J. Hyperspectral Remote Sensing Data Analysis and Future Challenges. IEEE Geosci. Remote Sens. Mag. 2013, 1, 6–36. [Google Scholar] [CrossRef]
  2. Bioucas-Dias, J.M.; Plaza, A.; Dobigeon, N.; Parente, M.; Du, Q.; Gader, P.; Chanussot, J. Hyperspectral Unmixing Overview: Geometrical, Statistical, and Sparse Regression-Based Approaches. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2012, 5, 354–379. [Google Scholar] [CrossRef]
  3. Wang, G.; Weng, Q. Remote Sensing of Natural Resources; Taylor & Francis: Boca Raton, FA, USA, 2013. [Google Scholar]
  4. Fassnacht, F.E.; Latifi, H.; Sterenczak, K.; Modzelewska, A.; Lefsky, M.; Waser, L.T.; Straub, C.; Ghosh, A. Review of studies on tree species classification from remotely sensed data. Remote Sens. Environ. 2016, 186, 64–87. [Google Scholar] [CrossRef]
  5. Fassnacht, F.E.; Neumann, C.; Förster, M.; Buddenbaum, H.; Ghosh, A.; Clasen, A.; Joshi, P.K.; Koch, B. Comparison of Feature Reduction Algorithms for Classifying Tree Species with Hyperspectral Data on Three Central European Test Sites. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2014, 7, 2547–2561. [Google Scholar] [CrossRef]
  6. Feret, J.B.; Asner, P. Tree Species Discrimination in Tropical Forests Using Airborne Imaging Spectroscopy. IEEE Trans. Geosci. Remote Sens. 2013, 51, 73–84. [Google Scholar] [CrossRef]
  7. Richter, R.; Reu, B.; Wirth, C.; Doktor, D. The use of airborne hyperspectral data for tree species classification in a species-rich Central European forest area. Int. J. Appl. Earth Obs. Geoinf. 2016, 52, 464–474. [Google Scholar] [CrossRef]
  8. Dalponte, M.; Bruzzone, L.; Gianelle, D. Tree species classification in the Southern Alps based on the fusion of very high geometrical resolution multispectral/hyperspectral images and LiDAR data. Remote Sens. Environ. 2012, 123, 258–270. [Google Scholar] [CrossRef]
  9. Xia, J.; Du, P.; He, X.; Chanussot, J. Hyperspectral remote sensing image classification based on rotation forest. IEEE Geosci. Remote Sens. Lett. 2014, 11, 239–243. [Google Scholar] [CrossRef]
  10. Rodríguez, J.J.; Kuncheva, L.I.; Alonso, C.J. Rotation Forest: A new classifier ensemble method. IEEE Trans. Pattern Anal. Mach. Intell. 2006, 28, 1619–1630. [Google Scholar] [CrossRef]
  11. Lausch, A.; Heurich, M.; Gordalla, D.; Dobner, H.J.; Gwillym-Margianto, S.; Salbach, C. Forecasting potential bark beetle outbreaks based on spruce forest vitality using hyperspectral remote sensing techniques at different scales. For. Ecol. Manag. 2013, 308, 76–89. [Google Scholar] [CrossRef]
  12. Sankaran, S.; Mishra, A.; Ehsani, R.; Davis, C. A review of advanced techniques for detecting plant diseases. Comput. Electron. Agric. 2010, 72, 1–13. [Google Scholar] [CrossRef]
  13. Mutanga, O.; Ismail, R. Variation in foliar water content and hyperspectral reflectance of Pinus patula trees infested by Sirex noctilio. South. For. 2010, 72, 1–7. [Google Scholar] [CrossRef]
  14. Hardinsky, M.A.; Klemas, V.; Smart, M. The influence of soil salinity, growth form, and leaf moisture on the spectral radiance of Spartina alterniflora canopies. Photogramm. Eng. Remote Sens. 1983, 49, 77–83. [Google Scholar]
  15. Penuelas, J.; Pinol, J.; Ogaya, R.; Filella, I. Estimation of plant water concentration by the reflectance water index (R900/R970). Int. J. Remote Sens. 1997, 18, 2869–2875. [Google Scholar] [CrossRef]
  16. Ayerdi, B.; Marqués, I.; Grana, M. Spatially regularized semi-supervised Ensembles of Extreme Learning Machines for hyperspectral image segmentation. Neurocomputing 2015, 149, 373–386. [Google Scholar] [CrossRef]
  17. Wenzhi, L.; Pizurica, A.; Scheunders, P.; Philips, W.; Youguo, P. Semi-supervised Local Discriminant Analysis for Feature Extraction in Hyperspectral Images. IEEE Trans. Geosci. Remote Sens. 2013, 51, 184–198. [Google Scholar] [CrossRef]
  18. Vapnik, V.N. Statistical Learning Theory; John Wiley&Sons: New York, NY, USA, 2013. [Google Scholar]
  19. Bruzzone, L.; Chi, M.; Marconcini, M. A novel transductive SVM for semi-supervised classification of remote-sensing images. IEEE Trans. Geosci. Remote Sens. 2006, 44, 3363–3372. [Google Scholar] [CrossRef]
  20. Dalponte, M.; Ene, L.T.; Marconcini, M.; Gobakken, T.; Naesset, E. Semi-supervised SVM for individual tree crown species classification. ISPRS J. Photogramm. Remote Sens. 2015, 110, 77–87. [Google Scholar] [CrossRef]
  21. Makantasis, K.; Doulamis, A.D.; Doulamis, N.D.; Nikitakis, A. Tensor-Based Classification Models for Hyperspectral Data Analysis. IEEE Trans. Geosci. Remote Sens. 2018, 56, 6884–6898. [Google Scholar] [CrossRef]
  22. Ghamisi, P.; Benediktsson, J.A.; Ulfarsson, M.O. Spectral-Spatial Classification of Hyperspectral Images Based on Hidden Markov Random Fields. IEEE Trans. Geosci. Remote Sens. 2014, 52, 2565–2574. [Google Scholar] [CrossRef]
  23. Li, J.; Bioucas-Dias, J.M.; Plaza, A. Spectral-spatial hyperspectral image segmentation using subspace multinomial logistic regression and Markov random fields. IEEE Trans. Geosci. Remote Sens. 2012, 50, 809–823. [Google Scholar] [CrossRef]
  24. Zhang, E.; Zhang, X.; Jiao, L.; Liu, H.; Wang, S.; Hou, B. Spectral-spatial hyperspectral image ensemble classification via joint sparse representation. Pattern Recognit. 2016, 59, 42–54. [Google Scholar] [CrossRef]
  25. Makantasis, K.; Karantzalos, K.; Doulamis, A.; Doulamis, N. Deep Supervised Learning for Hyperspectral Data Classification through Convolutional Neural Networks. In Proceedings of the 2015 IEEE International Geoscience and Remote Sensing Symposium (IGARSS), Milan, Italy, 26–31 July 2015; pp. 4959–4962. [Google Scholar]
  26. Wolpert, D.H.; Macready, W.G. Coevolutionary free lunches. IEEE Trans. Evol. Comput. 2015, 9, 721–735. [Google Scholar] [CrossRef]
  27. Crawley, M.J.; Akhteruzzaman, M. Individual Variation in the Phenology of Oak Trees and Its Consequences for Herbivorous Insects. Funct. Ecol. 1988, 2, 409–415. [Google Scholar] [CrossRef]
  28. Schläpfer, D.; Schaepman, M.E.; Itten, K.I. PARGE: Parametric geocoding based on GCP-calibrated auxiliary data. Proc. SPIE 1998, 3438, 334–344. [Google Scholar]
  29. Schläpfer, D.; Richter, R.; Hueni, A. Recent developments in operational atmospheric and radiometric correction of hyperspectral imagery. In Proceedings of the Earsel SIG-IS Workshop, Tel-Aviv, Israel, 16–19 March 2009. [Google Scholar]
  30. Didaci, L.; Fumera, G.; Roli, F. Diversity in Classifier Ensembles: Fertile Concept or Dead End? Mult. Cl. Syst. Lect. Notes Comput. Sci. 2013, 7872, 37–48. [Google Scholar]
  31. Trunk, G.V. A Problem of Dimensionality: A Simple Example. IEEE Trans. Pattern Anal. Mach. Intell. 1979, 1, 306–307. [Google Scholar] [CrossRef]
  32. Breiman, L. Random Forests. Mach. Learn. 2001, 45, 5–32. [Google Scholar] [CrossRef]
  33. Fan, R.E.; Chang, K.W.; Hsieh, C.J.; Wang, X.R.; Lin, C.J. LIBLINEAR: A library for large linear classification. J. Mach. Learn. Res. 2008, 9, 1871–1874. [Google Scholar]
  34. Sindhwani, V.; Keerthi, S. Newton Methods for Fast Solution of Semi-Supervised Linear SVMs; Book Chapter in Large Scale Kernel Machines; MIT Press: Cambridge, MA, USA, 2006. [Google Scholar]
  35. Vedaldi, A.; Lenc, K.; Gupta, A. MatConvNet Convolutional Networks for Matlab. Available online: http://www.vlfeat.org/matconvnet/matconvnet-manual.pdf (accessed on 29 October 2019).
  36. Sokolova, M.; Lapalme, G. A systematic analysis of performance measures for classification tasks. Inf. Process. Manag. 2009, 45, 427–437. [Google Scholar] [CrossRef]
Figure 1. Left: Calendar depicts the dates and scopes of the measurement flights. Middle: Measurement setup for hyperspectral imaging data: (A) Hyspex VNIR 1600, (B) Hyspex SWIR 320m-e, and (C) Novatel SPAN CPT inertial measurement unit (IMU) mounted on stabilized platform (not depicted). Images have been recorded at an altitude of ~1000 m above ground level with FOV of 17° (VNIR) and 14° (SWIR). Right: Trajectory of the flight on 3 August.
Figure 1. Left: Calendar depicts the dates and scopes of the measurement flights. Middle: Measurement setup for hyperspectral imaging data: (A) Hyspex VNIR 1600, (B) Hyspex SWIR 320m-e, and (C) Novatel SPAN CPT inertial measurement unit (IMU) mounted on stabilized platform (not depicted). Images have been recorded at an altitude of ~1000 m above ground level with FOV of 17° (VNIR) and 14° (SWIR). Right: Trajectory of the flight on 3 August.
Remotesensing 11 02788 g001
Figure 2. Identification of reference sites with a dominant tree species using existing information systems (forest inventory) and validation with existing aerial images. Left image shows a spruce reference site (1) surrounded by a mix of tree species (2). The right image shows another spruce reference site (1) with a group of birch and maple (2), large (3) as well as small clearings (4), and other isolated trees (5). The blue-bordered regions are the finally derived reference sites.
Figure 2. Identification of reference sites with a dominant tree species using existing information systems (forest inventory) and validation with existing aerial images. Left image shows a spruce reference site (1) surrounded by a mix of tree species (2). The right image shows another spruce reference site (1) with a group of birch and maple (2), large (3) as well as small clearings (4), and other isolated trees (5). The blue-bordered regions are the finally derived reference sites.
Remotesensing 11 02788 g002
Figure 3. Structure of convolutional neural network (CNN) for classification of tree species. The variable F denotes the dimension of the input feature space.
Figure 3. Structure of convolutional neural network (CNN) for classification of tree species. The variable F denotes the dimension of the input feature space.
Remotesensing 11 02788 g003
Figure 4. Classification accuracies for best base classifiers and ensembles of classifiers.
Figure 4. Classification accuracies for best base classifiers and ensembles of classifiers.
Remotesensing 11 02788 g004
Figure 5. Performance measures of base classifiers (AC) and hybrid ensemble based upon precision weighted fusion (D) per class. Net diagrams show precision and recall values of the following base classifiers: Random forest classifier for spectral indices (A), CNN classifier for MCLDA-transformed images (B), and random forest classifier for MCLDA-transformed images (C).
Figure 5. Performance measures of base classifiers (AC) and hybrid ensemble based upon precision weighted fusion (D) per class. Net diagrams show precision and recall values of the following base classifiers: Random forest classifier for spectral indices (A), CNN classifier for MCLDA-transformed images (B), and random forest classifier for MCLDA-transformed images (C).
Remotesensing 11 02788 g005
Table 1. List of 23 tree species from existing databases (forest inventory). For 15 tree species (excluded species marked with *) references sites for the classifier training have been provided forest authorities. The number of available reference sites and the total area are listed for each species.
Table 1. List of 23 tree species from existing databases (forest inventory). For 15 tree species (excluded species marked with *) references sites for the classifier training have been provided forest authorities. The number of available reference sites and the total area are listed for each species.
Tree Species Pedunculate oak926,046 m2Littleleaf lime12295 m2
European aspen * Norway spruce1020,340 m2Black pine *
Sycamore maple * Scotch pine1337,637 m2Sessile oak11203 m2
Birch * European white birch11117 m2Weymouth pine1969 m2
Douglas fir11102 m2Japanese larch * Serbian spruce11630 m2
Oak1032,808 m2European beech415,720 m2Poplar217,649 m2
European larch48128 m2Red oak35732 m2Robinia12439 m2
Ash * Red alder * Sitka spruce*
Table 2. Selected spectral indices. The values of the indices are combined into a six-dimensional feature vector and used for classification of tree species. The abbreviation R800 denotes reflectance value in band with central wavelength of 800 nm.
Table 2. Selected spectral indices. The values of the indices are combined into a six-dimensional feature vector and used for classification of tree species. The abbreviation R800 denotes reflectance value in band with central wavelength of 800 nm.
Spectral IndicesFormulasReferences
DI1R800 − R550Lausch et al. (2013)
GNDVI(R780 − R550)/(R780 + R550)Gitelson et al. (1996)
MCARI[(R700 − R670) − 0.2*(R700 − R550)]*R700/R670Haboudane et al. (2004)
NDVI(R800 − R670)/(R800 + R670)Rouse et al. (1974)
PRI(R531 − R570)/(R531 + R570)Gamon et al. (1992)
WIR900/R970Penuelas et al. (1997)
Table 3. Overview of the 15 selected feature spaces. First row: pixel-based feature spaces, Second row: spatial-spectral feature spaces. Due to the high number of channels of the hyperspectral images, the calculation of spatial features was performed for images of reduced dimensionality. PCA: principal component analysis; MCLDA: multi-class linear discriminant analysis; SELD: supervised local discriminant analysis.
Table 3. Overview of the 15 selected feature spaces. First row: pixel-based feature spaces, Second row: spatial-spectral feature spaces. Due to the high number of channels of the hyperspectral images, the calculation of spatial features was performed for images of reduced dimensionality. PCA: principal component analysis; MCLDA: multi-class linear discriminant analysis; SELD: supervised local discriminant analysis.
Reflectance
Spectra
IndicesPCA
5-dim
PCA
14-dim
MCLDA 5-dimMCLDA 14-dimSELD 5-dimSELD
14-dim
Pixel········
Spatial ·······
Table 4. Overall accuracies and standard deviations for Multi-Class-SVM classifier with linear kernel using a 1-vs.-all strategy depending on the choice of feature space (1st column), pixelwise (2nd column) or with inclusion of spatial information from surrounding pixels (3rd column). The best performance within each row and the best method are emphasized with bold font to highlight the better performance of the spatial approaches.
Table 4. Overall accuracies and standard deviations for Multi-Class-SVM classifier with linear kernel using a 1-vs.-all strategy depending on the choice of feature space (1st column), pixelwise (2nd column) or with inclusion of spatial information from surrounding pixels (3rd column). The best performance within each row and the best method are emphasized with bold font to highlight the better performance of the spatial approaches.
Feature SpacePixelwiseSpatial
Reflectances (160 dim)0.227 ± 0.075-
Indices (7 dim)0.097 ± 0.0610.098 ± 0.04
PCA (5 dim)0.231 ± 0.0620.195 ± 0.082
PCA (14 dim)0.26 ± 0.1140.273 ± 0.113
MCLDA (5 dim)0.428 ± 0.0490.616 ± 0.064
MCLDA (14 dim)0.459 ± 0.0330.651 ± 0.08
SELD (5 dim)0.398 ± 0.10.421 ± 0.071
SELD (14 dim)0.415 ± 0.0970.571 ± 0.053
Table 5. Overall accuracies and standard deviations for the random forest classifier with different choices of feature space (1st column) and varying complexity of models (# trees) for pixelwise classification and with consideration of spatial information from surrounding pixels. The best performance within each row and the best method are emphasized with bold font to highlight the better performance of the spatial approaches.
Table 5. Overall accuracies and standard deviations for the random forest classifier with different choices of feature space (1st column) and varying complexity of models (# trees) for pixelwise classification and with consideration of spatial information from surrounding pixels. The best performance within each row and the best method are emphasized with bold font to highlight the better performance of the spatial approaches.
PixelwiseSpatialPixelwiseSpatial
# trees2020100100
Reflectances0.435 ± 0.055-0.447 ± 0.057-
Indices0.352 ± 0.0320.622 ± 0.0390.361 ± 0.0340.633 ± 0.040
PCA (5 dim)0.357 ± 0.0290.582 ± 0.0560.366 ± 0.0310.594 ± 0.085
PCA (14 dim)0.441 ± 0.050.553 ± 0.0880.452 ± 0.050.572 ± 0.053
MCLDA (5 dim)0.487 ± 0.040.642 ± 0.0550.495 ± 0.0420.653 ± 0.055
MCLDA (14 dim)0.522 ± 0.0380.663 ± 0.0520.533 ± 0.0390.684 ± 0.049
SELD (5 dim)0.367 ± 0.0270.621 ± 0.030.377 ± 0.0290.63 ± 0.03
SELD (14 dim)0.486 ± 0.0410.622 ± 0.060.499 ± 0.0420.639 ± 0.059
Table 6. Overall accuracies and standard deviations for the Rotation Forest classifier with different choices of feature space (first column) and varying complexity of models (number of trees) for pixelwise classification and with consideration of spatial information from surrounding pixels (neighborhood). The best performance within each row and the best method are emphasized with bold font to highlight the better performance of the spatial approaches.
Table 6. Overall accuracies and standard deviations for the Rotation Forest classifier with different choices of feature space (first column) and varying complexity of models (number of trees) for pixelwise classification and with consideration of spatial information from surrounding pixels (neighborhood). The best performance within each row and the best method are emphasized with bold font to highlight the better performance of the spatial approaches.
PixelwiseSpatialPixelwiseSpatial
# trees2020100100
Reflectances0.501 ± 0.043-0.517 ± 0.043-
Indices0.33 ± 0.0250.609 ± 0.0390.331 ± 0.0280.619 ± 0.038
PCA (5 dim)0.342 ± 0.0310.561 ± 0.0530.347 ± 0.0320.575 ± 0.055
PCA (14 dim)0.442 ± 0.050.525 ± 0.0780.45 ± 0.0520.548 ± 0.081
MCLDA (5 dim)0.467 ± 0.0420.658 ± 0.0480.472 ± 0.0420.665 ± 0.047
MCLDA (14 dim)0.522 ± 0.0380.683 ± 0.0440.539 ± 0.040.705 ± 0.044
SELD (5 dim)0.345 ± 0.0310.615 ± 0.0330.351 ± 0.030.624 ± 0.033
SELD (14 dim)0.487 ± 0.0410.615 ± 0.0590.498 ± 0.0420.634 ± 0.058
Table 7. Overall accuracies and standard deviations for the CNN classifier. The gaps in the table reflect the stepwise approach to identify the best CNN configuration. First, we identified a best performing image transformation, second we evaluated different promising modifications of the chosen CNN. Bold setting and underlining highlight the parameter combination with best overall accuracy.
Table 7. Overall accuracies and standard deviations for the CNN classifier. The gaps in the table reflect the stepwise approach to identify the best CNN configuration. First, we identified a best performing image transformation, second we evaluated different promising modifications of the chosen CNN. Bold setting and underlining highlight the parameter combination with best overall accuracy.
Pixelwise
Dropout = 0
no Pooling
Pixelwise
Dropout = 0
with Pooling
Pixelwise
Dropout = 0.5
no Pooling
Pixelwise
Dropout = 0.5
with Pooling
Spatial
Dropout = 0
no Pooling
Indices0.485 ± 0.054----
PCA (5 dim)0.443 ± 0.072----
PCA (14 dim)0.439 ± 0.086----
MCLDA (5 dim)0.589 ± 0.074----
MCLDA (14 dim)0.665 ± 0.0720.656 ± 0.0770.675 ± 0.080.664 ± 0.0870.732 ± 0.086
SELD (5 dim)0.47 ± 0.058----
SELD (14 dim)0.412 ± 0.058----

Share and Cite

MDPI and ACS Style

Knauer, U.; von Rekowski, C.S.; Stecklina, M.; Krokotsch, T.; Pham Minh, T.; Hauffe, V.; Kilias, D.; Ehrhardt, I.; Sagischewski, H.; Chmara, S.; et al. Tree Species Classification Based on Hybrid Ensembles of a Convolutional Neural Network (CNN) and Random Forest Classifiers. Remote Sens. 2019, 11, 2788. https://doi.org/10.3390/rs11232788

AMA Style

Knauer U, von Rekowski CS, Stecklina M, Krokotsch T, Pham Minh T, Hauffe V, Kilias D, Ehrhardt I, Sagischewski H, Chmara S, et al. Tree Species Classification Based on Hybrid Ensembles of a Convolutional Neural Network (CNN) and Random Forest Classifiers. Remote Sensing. 2019; 11(23):2788. https://doi.org/10.3390/rs11232788

Chicago/Turabian Style

Knauer, Uwe, Cornelius Styp von Rekowski, Marianne Stecklina, Tilman Krokotsch, Tuan Pham Minh, Viola Hauffe, David Kilias, Ina Ehrhardt, Herbert Sagischewski, Sergej Chmara, and et al. 2019. "Tree Species Classification Based on Hybrid Ensembles of a Convolutional Neural Network (CNN) and Random Forest Classifiers" Remote Sensing 11, no. 23: 2788. https://doi.org/10.3390/rs11232788

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