Next Article in Journal
Recognition of Sedentary Behavior by Machine Learning Analysis of Wearable Sensors during Activities of Daily Living for Telemedical Assessment of Cardiovascular Risk
Next Article in Special Issue
Micro-Doppler Feature Extraction of Inverse Synthetic Aperture Imaging Laser Radar Using Singular-Spectrum Analysis
Previous Article in Journal
A Non-Linear Filtering Algorithm Based on Alpha-Divergence Minimization
Previous Article in Special Issue
A Lightweight Convolutional Neural Network Based on Visual Attention for SAR Image Target Classification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Concurrent and Hierarchy Target Learning Architecture for Classification in SAR Application

1
Department of Electronic Engineering, Harbin Institute of Technology, Harbin 150001, China
2
Collaborative Innovation Center of Information Sensing and Understanding, Harbin Institute of Technology, Harbin 150001, China
*
Authors to whom correspondence should be addressed.
Sensors 2018, 18(10), 3218; https://doi.org/10.3390/s18103218
Submission received: 28 June 2018 / Revised: 21 September 2018 / Accepted: 21 September 2018 / Published: 24 September 2018
(This article belongs to the Special Issue Automatic Target Recognition of High Resolution SAR/ISAR Images)

Abstract

:
This article discusses the issue of Automatic Target Recognition (ATR) on Synthetic Aperture Radar (SAR) images. Through learning the hierarchy of features automatically from a massive amount of training data, learning networks such as Convolutional Neural Networks (CNN) has recently achieved state-of-the-art results in many tasks. To extract better features about SAR targets, and to obtain better accuracies, a new framework is proposed: First, three CNN models based on different convolution and pooling kernel sizes are proposed. Second, they are applied simultaneously on the SAR images to generate image features via extracting CNN features from different layers in two scenarios. In the first scenario, the activation vectors obtained from fully connected layers are considered as the final image features; in the second scenario, dense features are extracted from the last convolutional layer and then encoded into global image features through one of the commonly used feature coding approaches, which is Fisher Vectors (FVs). Finally, different combination and fusion approaches between the two sets of experiments are considered to construct the final representation of the SAR images for final classification. Extensive experiments on the Moving and Stationary Target Acquisition and Recognition (MSTAR) dataset are conducted. Experimental results prove the capability of the proposed method, as compared to several state-of-the-art methods.

1. Introduction

The classification of Synthetic Aperture Radar (SAR) targets into different classes is one of the most challenging algorithmic aspects of radar structures. Unlike optical remote sensing, which cannot achieve its role in bad weather and at night, SAR can operate in all-weather conditions day-and-night and make very high resolution images, and it has played an important role in military and civil applications, such as target classification, reconnaissance, and surveillance. However, the comprehension of SAR images requires specialists, because unlike natural images, SAR images reflect the backscattering electromagnetic wave intensity of targets and speckle. Moreover, for humans, searching for targets of interest in the massive SAR images will take a lot of time. Furthermore, SAR images are covered with speckle noise, which is an important reason behind the reduction of the images quality. Besides this, they are very sensitive to the variation of target pose, and vary suddenly and quickly with small change in aspect angles. Therefore, SAR Automatic Target Recognition (ATR) is a demanding mission and has become a serious research topic for many applications. In the last decade, several methods were proposed in order to analyze and classify SAR images. A basic architecture of SAR ATR was defined as three phases: Detection, discrimination, and classification [1]. The primary two stages, also known as prescreening and low-level classification (LLC), are ordinarily referred to as the focus-of-attention module [2]. Detection is to excerpt candidate targets from SAR images via a Constant False Alarm Rate (CFAR) detector [3]. The output includes targets of interests such as armored vehicles, missile launchers, and tanks as well as false alarm clutter such as buildings, bridges, trees, and cars. In order to dispose of false alarms, several features are chosen to train a discriminator to distinguish between the two classes (target and clutter), which is the discrimination stage [4]. The third stage, also called high-level classification (HLC) [5], aims to classify the targets into different categories, which is the focus of this research.
Of the numerous methods, each approach that has been proposed to deal with the classification stage is included in one of the three main taxonomies, namely template-based [6], model-based [7] and pattern-based [8]. In recent years, pattern-based algorithms have made significant breakthroughs in the image classification process. A variety of architectures have been proposed, and they are based on first designing a set of different features to represent the targets by extracting them directly from the raw images, and then using these feature vectors to train a classifier and hence the classification problem is solved. To obtain a suitable classification performance, the features together with the classifiers should be carefully constructed. On one hand, numerous algorithms are employed for feature extraction, for instance, scale invariant feature transform (SIFT) [9], histogram of oriented gradients (HOG) [10], non-negative Matrix Factorization (NMF) [11,12], Principal Component Analysis (PCA) [13,14], deep Convolutional Neural Networks (CNNs) [15,16]. On the other hand, several robust trainable classifiers have been developed such as mean square error (MSE) classifier [17], Bayes classifier [18] and support vector machine (SVM) [19].
Deep CNN is currently one of the most promising technologies for classification. It has obtained the state-of-the-art results for object detection and automated land cover classification in high-resolution remote-sensing imagery. Deep CNN has been used to classify a wide variety of remotely sensed imagery, including SAR images [20], hyperspectral [21], and high resolution electro-optical imagery [22]. Different from methods based on hand-crafted feature extraction, the deep CNN based method automatically learns the feature from large-scale datasets, and achieves extraordinary performance in object recognition. The goal of CNN is to catch exceptional representations, usually at various levels, using the unknown structure in the input data, where lower-level features make it possible to define and learn higher-level features and make them more abstract, with their individual features. Then, these discovered features are more invariant to most of the variations commonly present in the raw training distribution, while collectively conserving the essential information in the input. Taking the tremendous progress deep learning has made in object recognition into consideration, deep CNNs are expected to also solve the SAR target recognition issue. Several researchers replicate architectures of deep CNN, which were successfully applied on object classification on SAR images. However, they did not obtain the same results because training a deep CNN requires a very large dataset such as ImageNet [23,24,25,26], which is clearly not the case of SAR images. Nonetheless, the limited labeled SAR target data have not stopped researchers from achieving better classification rates. To address this problem, numerous approaches have been employed to achieve better results on SAR images like transfer learning and training CNN from scratch.
One technique consists of applying transfer learning from pre-trained deep CNNs that can be modified to produce accurate classifications for specific applications. The main idea of CNN transfer learning is to use the power of networks that are already trained on a large-scale data set of general images in different application areas. These types of networks can be regarded as general feature extractors or classifiers. This technique can be successfully applied in cases where we do not have a large data set for training a CNN from scratch. There are three famous scenarios for transfer learning of a CNN: A pre-trained net is used as a fixed feature extractor, a pre-trained net is used for fine-tuning and a pre-trained net is used for initialization in training from scratch [27,28,29]. In the first approach, a neural network is trained with an independent general data set of images and the output of the network can be interpreted as a feature vector and used for further classification. In addition, features from the previous to last hidden layer can be used for classification. The fine-tuning method is based on the idea of training the last layers of the network to specialize them for a particular data set. The main benefits of this method are reduced training time and the possibility of effective training with a small data set. However, this technique does not perform well on SAR images since they refer to the backscattering characteristics of the ground features, representing a list of scattering centers, and each pixel intensity of the image depends on a range of factors, such as shapes, orientations and types of the scatterers in the area where the target is located. Other techniques were proposed to overcome this problem, one of which consist of training a CNN from scratch using data augmentation to increase the size of data [20]. Another method uses transfer learning from pre-trained deep CNN. However, instead of using optical images data set in the training, it uses a large number of unlabeled SAR scene images [27].
Another interesting point that is worth noting is the use of feature coding approaches in the process of images classification. The Bag-of-Words (BoW) [30] approach is one of the famous models in this field. In order to have good classification accuracy, three steps were performed: First extract the features, then generate a codebook, and after that a histogram is generated to represent each image. Many modern approaches are based on the BoW model; for example Reference [31] proposed a supervised incremental coding method based on the BoW model and proved that this method yielded much better features for SAR image classification. One recent method uses Fisher Vectors (FVs), which are in essence an image representation obtained by pooling local image features, and they are used as a global image descriptor in image classification. Compared to the previous coding approaches, its advantages are: First, its ability to store second-order information about the features. Secondly, FVs utilize Gaussian Mixture Models (GMMs) to generate the feature vocabulary. Therefore, it generates a probabilistic visual vocabulary instead of using a hard codebook, which allows it to be more flexible. This is an important feature, which helps in increasing the accuracy performance.
Another important technique that can be utilized to improve classification accuracy is fusion. Sensor fusion is a common technique in signal processing to combine data from various sensors. Feature fusion is another current method, ranging from simple concatenation to very advanced methods like fuzzy integrals. Finally, information fusion merges independent results from signal processing techniques that otherwise can be used alone as the final signal processing result. For example, Reference [32] has obtained good results by using some of these techniques on Land Cover High-Resolution Imagery.
The CNN architectures proposed in the literature have proven that the activations from high-level layers of CNNs can generate powerful feature representations with outstanding performance. However, we noticed that the best classification accuracy for individual SAR classes varied among the different CNN architectures. In addition, features extracted from lower-level layers, particularly the convolutional layers, lack sufficient study. Only a few works has been conducted in this area; for example, Reference [33] used cross-layer CNN features extracted from multiple layers of CNN for generic classification tasks. Reference [34] on the other hand has used features extracted from the last convolutional layer of deep CNN using transfer learning for the scene classification of high-Resolution remote sensing images. They both demonstrate the capacity of lower-level layers of CNN architectures to achieve better performance.
The two reasons mentioned above encourage us to propose a new framework. The framework is drawn from the most contemporary techniques in image processing and deep learning. It is a combination of three different CNNs. The three CNNs have the same architecture; the main difference is the sizes of the convolution and pooling kernel in each one of them: Coarse Grain CNN (CG-CNN), Middle Grain CNN (MG-CNN) and Fine Grain CNN (FG-CNN). We train each CNN from scratch using the chosen sizes of convolutional and pooling kernel. We show that through a combination of recent techniques, we can obtain significant performance improvement on the Moving and Stationary Target Acquisition and Recognition (MSTAR) dataset classification task. We investigate how to obtain better accuracy classification compared to the state-of-the-art results that use the same dataset without data augmentation by forming better representations from the SAR images using CNNs activations. By deleting the last few layers of a CNN, we handle the remainder of the CNN as a fixed feature extractor. Considering that these CNNs are broad multi-layer architectures, we consider two ways of extracting CNN features with reference to different layers:
  • We simply calculate the CNN activations for the entire SAR images and consider the Fully-Connected (FC) layer activation vectors as the global feature representations for all images.
  • We first compute dense CNN activations from the last convolutional layer of the input image, and then we convert them into a global representation using the Fisher encoding. Then, the global features of the image are fed to a simple classifier for the classification task.
Extensive experiments prove that powerful features SAR images can be generated. The resulting features and classification outputs from each of these CNNs are then either combined or fused using a variety of methods into a final refined classification. Evaluation of the proposed approach is conducted with the MSTAR benchmark data set. Experimental results validate the superiority and effectiveness of the proposed approach.
The paper is formulated as follows. In Section 2, we briefly review some related works corresponding to some state-of-the-art SAR images classification methods. Section 3 describes the main methods and tools used in this work, followed by the fusion method used as well as the fuse classifiers for SAR images classification. Section 4 presents the proposed framework used to extract, process, and classify the features. In Section 5, experiments are carried out with the MSTAR database, and the performance of the proposed approach is described. Finally, Section 6 lists the conclusions along with the discussion of the results.

2. Related Work

In this section, a quick overview of the previous studies related to SAR target recognition is given. Cui Z [35] reviewed the effectiveness of some traditional methods such as PCA and NMF, which have an accuracy of 86.07% and 89.47%, respectively. These traditional methods have a relatively low accuracy rate compared with recent approaches. Agrawal and Mangalraj [36] proposed a feature extraction algorithm based on SIFT and they achieved an accuracy of 90.99% in a 3-class classification task.
Over the last decade, many researchers have focused on the classification of SAR ATR using CNN, which normally refers to supervised classification. Using randomly sampled SAR targets patches, Reference [30] extracts SAR targets feature representation by a single convolutional layer and achieves the accuracy of 84.7% in 10-class classification tasks. Reference [16] learned discriminative feature sets directly from training data instead of requiring pre-specification by a human designer. In order to achieve that the author used an architecture of three convolutional layers, followed by an FC layer of Softmax as a classifier, and recorded an accuracy of 92.3%. Since then, various learning algorithms were explored such as AdaDelta [35], which can update the different learning rates of hyper-parameters and outperformed other methods like stochastic gradient descent (SGD) and AdaGrad. Reference [27] proposed a framework based on transfer learning where an assembled CNN architecture consisting of a classification pathway and a reconstruction pathway were designed. The author used a large number of unlabeled SAR scene images to train the reconstruction pathway using stacked convolutional auto-encoders instead of training the CNN with the limited dataset. The results showed an accuracy of 99.05%. Reference [20] proposed an architecture of five-layer all-convolutional layers (A-ConvNets) using the data augmentation technique to enlarge the dataset, since the limited training data was insufficient to train the deep CNN, and used two techniques to avoid over-fitting which are removing the FC layer and a drop-out in a convolution layer. The experimental results illustrate that A-ConvNets achieved an average accuracy of 99% on the classification of ten-class targets.

3. Methods and Materials

3.1. Deep Convolutional Neural Networks (CNNs)

Deep CNNs are neural networks whose topology encodes a spatial relationship between nodes in successive layers that is analogous to the convolution operation. A deep CNN is an algorithm, which learns a nonlinear classification function from the labeled input data. It is characterized in this way: First, an input layer, where the algorithm receives the SAR images of the MSTAR dataset and transfer them to the next layers for further processing. Next, a number of hidden layers follows the input layer. The role of hidden layers is to perform feature extraction and aggregation from the raw data. In a typical CNN, the hidden layers are alternating sequences of convolutional, batch normalization, ReLu and pooling layers. Finally, the architecture has an output layer that acts as a classification mechanism. Each layer has neurons that act upon the input features and pass the result to the next layer. An effective network learns weights in each layer, which produce accurate classification outputs for their input images. When training a deep neural network, the main challenges are in setting up the correct topology, choosing appropriate parametric initializations, defining an objective function to quantify misclassification error and an appropriate algorithm for training. All these factors affect the performance of the network. A multitude of extensions to the basic architecture and training mechanism have been proposed, the most recent one which have better results is composed of multiple cascaded stages, as is shown in Figure 1.
The convolutional layer’s outputs are feature maps; each one is computed by a dot product between the local regions (receptive field). Training CNN is complicated since the distribution of each layer’s inputs changes during training, as the parameters of the previous layers change (internal covariate shift). This slows down the training by requiring lower learning rates and careful parameters initialization, and makes it notoriously hard to train models with saturating nonlinearities. Therefore, a Batch Normalization (BN) layer is added before the nonlinearity where elements of feature maps, at different locations, are normalized. The BN layer’s role is to reduce the internal covariant shift and the gradients dependence on their initial values or on the scale of the parameters. It also reduces the need for dropout and regularizes the model [37]. Principally, an elementwise non-linear activation function is practiced to these feature maps like ReLu or sigmoid. By computing a local region maximum, the max-pooling layers complete a down sampling operation on the feature maps spatial dimensions. The FC layer has full connections to the totality of activations in the previous layer, and their activations are computed with a matrix multiplication followed by a bias offset. The last layer is a Softmax layer that computes the scores for each defined class. CNNs transform the input from original values to the final class scores through the structure in a feedforward style. The CNNs parameters (i.e., the weights and bias) are trained with classic stochastic gradient descent with momentum based on the backpropagation algorithm [38,39].

3.2. Feature Coding

Feature coding is a key component in image classification. It has been extensively studied over the past decade, where numerous coding algorithms have been proposed. The bag-of-features (BoF) [40], developed from the BoW model in document analysis [30], is one of the most popular and effective image classification frameworks in the recent literature. It has achieved impressive performance in several databases and competitions. Another method, which can be viewed as a generalization of the BoW, is the Fisher coding-based method, which has gained popularity in the last few years. Due to its special characteristics compared with the other image representation methods, it is well suited for SAR images, as FVs method is based on a visual vocabulary, with the assignment of patches to visual words and can extract a larger image signature for a given number of visual words than other representations such as the BoW. Moreover, the FVs record also the mean and variance of the points per dimension in each cell, which means having more information for same visual words. Therefore, compared with the other methods the FVs lead to high-dimensional feature vectors. In addition, FVs can be computed from much smaller vocabularies, and therefore at a lower computational cost. FVs are designed to estimate the probability density distribution (PDD) of features and focus on the global description of all features, achieving translation and rotation-invariance, which is crucial to classifying SAR images.

Fisher Vectors

FVs coding-based methods [41,42] estimate the distribution of features with the GMM, consisting of the weights, the means, and the covariance matrix of multiple Gaussian distributions, each of which reflects one pattern of features. FVs coding was originally derived from the Fisher kernel, which aims to combine the benefit of generative and discriminative approaches for pattern recognition by describing a signal by a gradient vector from its probability density function [43]. The gradient vector indicates the direction in which parameters should be adjusted to best fit the data.
Generally, FVs coding process can be divided into front-end stage and back-end stage. In our case, the front-end stage consists of extracting features from the last convolutional layer. The back-end stage employ GMM as the generative model for these local features. The FVs are then derived from the gradients with respect to the GMM parameters. Most existing methods employ a diagonal GMM as a generative model to represent the distribution of local features, for instance SIFT [9] and HoG [10]. Several versions of Fisher coding have been proposed after the original one [41]. As far as we can tell, Improved Fisher Kernel(IFK) [42] achieves the best performance.
Let X = { x n , n = 1 N } R D × N be the features extracted from an image in a D-dimensional descriptor space, B = { b m , m = 1 M } R D × M be a codebook with M codewords (obtained typically by clustering over features), and V = { v n , n = 1 N } be the corresponding representation of these N features. Each x is represented by the codebook B in feature coding. This process generates responses on M codewords, consisting of a coding vector v with M elements. In most coding algorithms, only a part of codewords is chosen to represent a feature, and therefore the coding vector v is usually sparse, i.e., the responses are zeros on the codewords which are not chosen. The GMMs define the probability density distribution of features in IFK. Its parameters θ n = { α n , μ n , σ n } are the weight α n , the mean vector μ n , and the covariance matrix σ n of the nth Gaussian distribution can be generally estimated by the Expectation Maximization (EM) algorithm [44]. Supposing features are independent from each other, each image can be expressed by the log likelihood of all extracted features:
L ( X / θ ) = n = 1 N log p ( x n / θ )  
where p ( x n / θ ) is the GMM-based probability density function. The normalized gradient vector is represented as:
Γ = F θ 1 / 2 θ L ( X / θ )  
θ L ( X / θ ) = [ δ L δ α , δ L δ μ , δ L δ σ ]
The derivative to α has very little contribution to the performance according to Reference [42]. Thus, it is removed in IFK.
θ L ( X / θ ) [ δ L δ μ , δ L δ σ ]  
F θ = E X p , X q [ θ L ( X p / θ ) θ L ( X q / θ ) ]  
where F θ is the Fisher information matrix, X p and X q denote two sets of features extracted from two arbitrary images. The Fisher information has an approximated close solution according to References [42,45], with which the coding vector of a feature, i.e., the FVs, can be represented as:
v ( k ) = [ Γ μ , k ,   Γ σ , k ] , k = 1 , , M  
Γ α k X = 1 w k n = 1 N γ n ( k ) ( x n μ k σ k )
Γ σ k X = 1 w k n = 1 N γ n ( k ) 1 2 ( ( x n μ k ) 2 σ k 2 1 )  
γ n , k = w k p k ( x n / θ ) / j = 1 M w j p j ( x n / θ )  
The FVs are obtained via quantizing the set of local feature descriptors with a small codebook and aggregating first and second order residual statistics for features quantized to each centroid. In other words, the fisher encoding captures the average first and second order differences between the image descriptors and the centers of a GMM, which can be thought of as a soft visual vocabulary. The residual statistics of centroids are concatenated together to attain the high-dimensional global descriptor representation. The performance has a proportional relation with the dimensionality of the global descriptor as shown in Reference [46].
FVs can be aggregated on descriptors in the image [45] which is popular for image classification, or around interest points [47] which are invariant to scale and orientation. The latter is usually used in image retrieval as the Difference of Gaussian interest points provide invariance to scale and rotation, which make FVs in image retrieval more invariant to certain geometrical transformations of the images and robust to training data [48].

3.3. Information Fusion

In this section, we focus on information fusion of the CNN classifier outputs to obtain improved results for the classification of the MSTAR data sets. The techniques summarized in the following are used in this paper to fuse the different CNN architectures outputs to obtain the final classification result. In our subsequent classifier fusion discussion, we use the following notation:
Suppose that we have a set of K classifiers, each of which classifies targets into one of distinct M classes. The output vector of classifier ck, given a target X, is represented by a column vector.
y k = { y k , m ; m = 1 , 2 , , M } , w h e r e : k = 1 , 2 , , K  
where m is the component of the output vector yk,m represents the estimated posterior probability that target X belongs to the class m, estimated by classifier ck. yk,m satisfies the following two requirements:
0 y k , m 1
m = 1 M y k , m = 1
The classification decision of classifier θk is:
θ k = arg max 1 m M { y k , m }
C is the classification matrix of size K × M , corresponding to K classifiers for M classes; p k is per-class classification accuracy from the CNN, for k [ 1 , , K ] calculated for each network as follows:
p k = [ ( correctly   predicted   targets   in   class   m   /   total   testing   of   class   m )   ×   100 % ; m = 1 , 2 , , M ]
The fusion takes place using the K classification vectors with c k arranged into a K × M classification matrix C .

3.3.1. Arrogance

Each network has a different classification output from the other networks. This type of fusion select the max per-element (for each target) classification output across all classifiers. In simple terms, the network that has the greatest confidence in its classification is accepted as the final output classification.
V ( C ) = ( max k = 1 , , K C k 1 , , max k = 1 , , K C k M )
Then, the predicted class is computed using arg max i = 1 , , M ( V ( C ) ) .

3.3.2. Simple Voting

This takes the highest confidence class for each c k , and then selects the class by the most occurring class i . In other words, we compute a vector V of a selected class per classifier by
V ( C ) = [ arg max m = 1 , , M ( c 1 ) , , arg max ( m = 1 , , M c K ) ]
Then, the classification is simply m o d e ( V ( C ) ) . A null classification is generated, if the majority of CNNs do not agree on a class. This is counted as a missed classification during evaluation.

3.3.3. Accuracy Weighted Sum

This takes the classification matrix C and computes a weighted confidence across all classifiers for each class using the classification accuracy p j as the weights in the sum. The vector p j is computed as the per class, average cross-validation performance for each network, j [ 1 , N ] . Let C j denote the jth row of the classification matrix, then
V ( C ) = [ p 1 . C 1 T , , p M . C M T ]  
The predicted class, i , is then arg max i ( V ( C ) ) . Intuitively, these weights are calculated from each network by its past performance measured from the cross-validation testing.

4. Methodology

For SAR image classification, several CNN architectures were proposed in the literature and have proven that the activations from high-level layers of CNNs can generate powerful feature representations with outstanding performance. However, two important points were noticed. First, the best classification accuracy for individual SAR classes is varied between the different CNN architecture and second, features extracted from lower-level layers, particularly the convolutional layers, lack sufficient study. These two reasons have encouraged us to propose different scenarios for utilizing CNN features for SAR images classification for the sake of investigating the effectiveness of features from the last convolutional layer and FC layer as well as the effect of combination and fusion different features and classifiers. Figure 2 and Figure 3 illustrate the framework of the proposed method.
The proposed method of SAR image classification is based on the following steps:
  • We use a set of different convolutional neural networks learning at different kernel sizes of convolution and pooling to produce high level invariant features;
  • The high-level features obtained from the FC layer are classified for each CNN model separately.
  • The features obtained from the last convolutional layer are classified after being encoded using the FVs coding.
  • A final feature vector from the FC layer is obtained by different combination of results of steps 2 and 3.
  • A final classification decision using different fusion methods based on results of steps 2 and 3 is obtained.
In this work, we have elected to apply the network topology presented in Section 3. Three different CNNs models at different kernel sizes of convolution and pooling were designed and trained on the MSTAR training set.
The choice of different parameters of each CNN model was mostly based on both previous types of research architecture [16,20,27,28,29] and experiments where we change the different parameters of CNN to obtain the architecture that achieves the best performances. We remark that several types of architecture can generate good SAR image representation. However, the best classification accuracy for individual classes is varied between the different forms of CNN architecture, which encourage us to exploit three kinds of CNN models with different convolution and pooling kernel size in each layer. We manage to have the same size of the FC layer in each architecture for reasons of simplicity in the two last steps of the proposed method, which are feature combination and the fusion. We chose the following appellation for the three architecture based on their respective convolution and pooling kernel sizes: CL-CNN, coarse grain with larger size, MM-CNN, middle grain with medium size, FS-CNN, fine grain with small size. The architecture of each network is given in Table 1. The convolution layer receives inputs from a local region of the input volume located in the small neighborhood of the previous layer, which is called the local receptive field. A typical convolutional layer has several feature maps. Weight vectors between different feature maps are different but all the units within one feature map share the same set of weights. Due to the use of local receptive fields and weight sharing, the number of free parameters to be learned is significantly reduced. BN layer helps in obtaining higher overall accuracy and faster learning. ReLu layer improves the networks by speeding up the training since it keeps the computation of the gradient very simple. Subsampling layer usually implemented as max-pooling layer further reduces feature dimension with translational invariance. FC layer is similar to classical neural networks computing a dot product between their input vector and their weight vector. The Softmax nonlinearity is utilized as the final output layer to deal with the multiclass classification issue.
Accuracy is calculated from each FC layer of the three CNN models using the SoftMax classifier, as well as from the dense features obtained from the last convolutional layers of each model, using the SVM classifier after being encoded by FVC. Figure 2 resumes the first three aforementioned steps of the proposed method.
Figure 3 resumes the last two above-mentioned steps of the proposed method. The combination module in Figure 3a could be either the concatenation method or the addition. In Figure 3b, the fusion module is one of the three aforementioned fusion methods in Section 3.3. The average cross-validation weights are needed only in the case where the fusion module is the Accuracy Weighted Sum.
The results of each individual CNN model motivate us to use different techniques of combination and fusion. Since the CNN which produced the best classification accuracy for individual classes is varied between the different CNNs, this suggests that differences in the CNN architecture types may allow one network to consistently perform better on a subset of the SAR classes even though it underperforms, on average, across all the classes compared with another network. Consequently, feature combination and fusion of the CNNs information outputs as shown in Figure 3 should improve the overall robustness and accuracy of the result compared to a single CNN.

5. Experiment Results and Analysis

In our experiments, we implement different CNN models for feature extraction and classification and try different combinations of CNNs to improve the classification accuracy. Moreover, further treatment was applied using the FVs to improve the quality of the obtained features, followed by several combination and fusion methods.

5.1. Experimental Setup

In this section, we investigate the representative power of CNNs features and evaluate the proposed models on real SAR image classification problem. The detailed experimental setup and numerous experiments with reasonable analysis are presented. We evaluate the effectiveness of CNN features on the following publicly available SAR dataset.
The MSTAR benchmark data set is generally used to compare and test the performance of SAR-ATR algorithms. It was collected by the Sandia National Laboratory SAR platform [49]. Hundreds of thousands SAR images enclosing ground targets were collected, containing different target types, aspect angles, depression angles, serial number, and articulation. However, only a limited subset is publicly accessible on the website [50]. The publicly data set available consists of ten different classes of ground targets: (air defense unit: ZSU-234; armored personnel carrier: BMP-2, BRDM-2, BTR-60, and BTR-70; tank: T-62, T-72; rocket launcher: 2S1; truck: ZIL-131; bulldozer: D7). They were collected using an X-band SAR sensor, in a one feet resolution spotlight mode, with full aspect coverage. Each image has a size of around 128 by 128. The images are already centered with 0 °C to 360 °C degrees of orientation. Figure 4 depicts examples of SAR images of ten types of targets at similar aspect angles and their corresponding optical images. The data set is composed of 5172 image decomposed into 2747 training images and 2425 testing images. The distribution of training and testing data for the different targets is illustrated in Table 2.
Before introducing the experimental setup, it is worth noting that only the original number of images is used i.e., No data augmentation algorithm is applied. Each target with the same serial number in the training and testing sets differs in the azimuth and depression angle. Images for training are captured at a 17 ° depression angle, and images for testing are acquired at a 15 °C depression angle. We resize all images to 128 × 128 size where the target is in the center of the image. Also after this step, we have subtracted the mean from the images and divided it by the standard deviation of the images, in other words, we have normalized the data.

5.2. Results of the First Set of Experiments

For the first set of experiments, all the CNN models introduced in Section 4 are evaluated and compared. The training subset is introduced as a cell array of image data, where the data in each cell have the same dimension ( 128 × 128 ), and then it is fed to the CNN catalog. Finally, the SAR images are classified into ten classes. The classification performances of the three different CNN models are shown in Table 3. The resulting high classification accuracies reveal the powerful ability of the chosen CNNs models.
We evaluate the time consumption (measured in terms of seconds) of training and testing all the three CNNs models all image in the MSTAR dataset, shown in Figure 5. As expected, CL-CNN has the biggest computational time consumption followed by MM-CNN and finally the FS-CNN. Although the three models have the same structure, they have very different computational time, due to the number of operations in each network. In our case, the size of the first pooling layer in each network has much more effects on the computation time than the other layers’ parameters. The bigger the size, the smaller the time consumption.

5.3. Results of the Second Set of Experiments

For the second set of experiments, we evaluate the global CNN features of the three CNN models. Dense features are extracted from the last convolutional layer and then encoded into global image features through one of the commonly used remote sensing feature coding approaches which is FVs coding. The dense features are L2-normalized prior to applying feature coding. We empirically set the number of Gaussian components in the GMM to 100.
We randomly select samples of each class for training the SVM classifier. The classification accuracy is measured by A = N c / N t , where N c denotes the number of correctly classified samples in the testing samples and N t denotes the total number of testing samples. We evaluate the final classification performance with the average accuracy A over 50 runs. The public LIBLINEAR library [51] is used for SVM training and testing with the linear kernel. The open source library VLFeat is also used [52] for implementing the FVs coding. The overall performance of the IFK when using the three CNN models convolutional features from the last convolutional layer are shown in Table 4. The accuracies of IFK on MSTAR dataset on the three models exceed all the results of using FC features.

5.4. Combining Features from the FC Layer

In the first set of experiments, we generated the global image representations solely via the CNN models. We can take it a step further by combining the features computed by these three frameworks. We use two techniques of combination: (1) addition and (2) concatenation, to combine among the extracted features. For addition, the size of transformed features is the same. In the concatenation combination, the size of the features is multiplied by 3. Finally, SVM is used for classification. Here, we present some tentative combinations using the three different CNNs, and the results are shown in Table 5. We observe that on the MSTAR dataset, the combined features improve performance with approximately 1–2% gains compared to the separate cases.

5.5. Combining Features from the Last Convolutional Layer

In the second set of experiments, we generated the global image representations via the FVs after extracting the dense features from the last convolutional layer for each CNN model. We combine the features computed by these three frameworks by using the same techniques (addition and concatenation). Finally, SVM is used for classification. Some tentative combinations using the three different CNNs are shown in Table 6.
It can be easily observed the combined features generated by the last convolutional layers improve performance with approximately 1–2% gains compared to the separate situations. So in both cases, the combination of features improves the accuracy performances even though the same structure was used in the three CNN models, which proves that differences in the CNN architecture allow one network to consistently perform better on a subset of the SAR images. In another words the gain would not be possible if the three CNN models classify the SAR images in the same way. Consequently, feature combination of the CNNs information improve the overall robustness and accuracy of the result compared to a single CNN.

5.6. Decision Fusion of the First and the Second Sets of Experiments Classifiers

After classification of the three CNN models separately by their individual classifiers for the two sets of experiments, decision fusion algorithms mentioned in Section 3.3 are used sequentially and the results of different combinations for fusion are summarized in Table 7 and Table 8. It is clear from the two tables that fusing the results of the second set of experiments (with FVC) outperforms those of the first set. The simple voting method achieved 98.39% classification accuracy on results of the second set of experiments when combining CL-CNN and FS-CNN. The arrogance method achieved the best accuracy, with a 98.60% classification. The accuracy weighted sum method achieved its best rate when fusing the three models with a 98.35% classification accuracy. Each method used for fusion outperformed the single best CNN baseline result.

6. Conclusions

In this paper, we have investigated the SAR images classification task on the MSTAR dataset using the activations of different CNN models. In order to obtain strong global image representations for classification, various sets of experiments are proposed by directly extracting features from the FC layers and encoding dense features from convolutional layers. We evaluate our methods on the public MSTAR dataset and achieve remarkable overall classification accuracies of 98.60% for the ten-classification task. As expected, our method obtained remarkable results competitive with the state-of-the-art methods. Even though we do not apply any data augmentation strategies in our method in contrast with the methods that also explore CNNs, our method is only 0.4% worse than the best method that elaborately uses transfer learning from SAR images scenes from the CNN, and better than methods which use pre-trained CNN models. Furthermore, combining and fusing scenarios improves performance by about 2–3%. These impressive results demonstrate that not only the features of FC layers but also dense convolutional features of CNN are discriminative and powerful image representations. In future studies, we plan to investigate more sophisticated strategies for encoding the dense CNN features to improve the invariance of representations.

Author Contributions

Methodology, M.T.; Supervision, Q.Y.; Validation, Q.Y.; Writing—original draft, M.T.

Funding

This research was funded by the National Natural Science Foundation of China under grants 61171182 and 61032011 and the fundamental Research Funds for the Central Universities under grants HIT.MKSTISP.2016 13 and HIT.MKSTISP.2016 26.

Acknowledgments

This work has been supported by the National Natural Science Foundation of Chinaand the fundamental Research Funds for the Central Universities.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Dudgeon, D.E.; Lacoss, R.T. An overview ofautomatic target recognition. Lincoln Lab. J. 1993, 6, 3–9. [Google Scholar]
  2. El-Darymli, K.; Gill, E.W.; McGuire, P.; Power, D.; Moloney, C. Automatic target recognition in synthetic aperture radar imagery: A state-of-the-art review. IEEE Access 2016, 4, 6014–6058. [Google Scholar] [CrossRef]
  3. Cui, Y.; Zhou, G.; Yang, J.; Yamaguchi, Y. On the iterative censoring for target detection in SAR images. IEEE Geosci. Remote Sens. Lett. 2011, 8, 641–645. [Google Scholar] [CrossRef]
  4. Park, J.-I.; Kim, K.-T. Modified polar mapping classifier for SAR automatic target recognition. IEEE Trans. Aerosp. Electron. Syst. 2014, 50, 1092–1107. [Google Scholar] [CrossRef]
  5. Patil, M.R.; Dubey, M.S. A review on automatic target recognition and detection image preprocessing approaches for SAR images. Int. Res. J. Eng. Technol. 2017, 4, 579–585. [Google Scholar]
  6. Kaplan, L.M. Analysis of multiplicative speckle models for template-based SAR ATR. IEEE Trans. Aerosp. Electron. Syst. 2001, 37, 1424–1432. [Google Scholar] [CrossRef]
  7. O’Sullivan, J.A.; DeVore, M.D.; Kedia, V.; Miller, M.I. SAR ATR performance using a conditionally Gaussian model. IEEE Trans. Aerosp. Electron. Syst. 2001, 37, 91–108. [Google Scholar] [CrossRef]
  8. Qiu, C.; Ren, H.; Zou, H.; Zhou, S. Performance comparison of target classification in SAR images based on PCA and 2D-PCA features. In Proceedings of the 2009 2nd Asian-Pacific Conference on Synthetic Aperture Radar, Xi’an, China, 26–30 October 2009; pp. 868–871. [Google Scholar]
  9. Gray, G.J.; Aouf, N.; Richardson, M.A.; Butters, B.; Walmsley, R.; Nicholls, E. Feature-based target recognition in infrared images for future unmanned aerial vehicles. J. Battlef. Technol. 2011, 14, 27. [Google Scholar]
  10. Cao, Z.; Zhang, X.; Wang, W. Forward-looking infrared target recognition based on histograms of oriented gradients. In Proceedings of the MIPPR 2011: Automatic Target Recognition and Image Analysis, Guilin, China, 4–6 November 2011. [Google Scholar]
  11. Huan, R.; Yun, P.; Keji, M. SAR image target recognition based on NMF feature extraction and Bayesian decision fusion. In Proceedings of the 2010 Second IITA International Conference on Geoscience and Remote Sensing, Qingdao, China, 28–31 August 2010; pp. 496–499. [Google Scholar]
  12. Riasati, V.; Srinivas, U.; Monga, V. Sar automatic target recognition via non-negative matrix approximations. In Proceedings of the Automatic Target Recognition XXII, Baltimore, MD, USA, 23–27 April 2012. [Google Scholar]
  13. Chamundeeswari, V.V.; Singh, D.; Singh, K. An analysis of texture measures in PCA-based unsupervised classification of SAR images. IEEE Geosci. Remote Sens. Lett. 2009, 6, 214–218. [Google Scholar] [CrossRef]
  14. Hu, L.; Xing, X. Sar target feature extraction and recognition based multilinear principal component analysis. In Proceedings of the International Symposium on Optoelectronic Technology and Application 2014: Image Processing and Pattern Recognition, Beijing, China, 13–15 May 2014. [Google Scholar]
  15. Bengio, Y. Deep learning of representations for unsupervised and transfer learning. In Proceedings of the ICML Workshop on Unsupervised and Transfer Learning, Edinburgh, UK, 26 June–1 July 2012; pp. 17–36. [Google Scholar]
  16. Morgan, D.A. Deep convolutional neural networks for ATR from SAR imagery. In Proceedings of the Algorithms for Synthetic Aperture Radar Imagery XXII, Baltimore, MD, USA, 20–24 April 2015. [Google Scholar]
  17. Novak, L.M.; Owirka, G.J.; Brower, W.S. Performance of 10-and 20-target MSE classifiers. IEEE Trans. Aerosp. Electron. Syst. 2000, 36, 1279–1289. [Google Scholar]
  18. Mehra, R.K.; Huff, M.; Ravichandran, R.B.; Williams, A.C. Nonparametric error estimation techniques applied to MSTAR data sets. In Proceedings of the Algorithms for Synthetic Aperture Radar Imagery V, Orlando, FL, USA, 14–17 April 1998; pp. 614–625. [Google Scholar]
  19. Zhao, Q.; Principe, J.C. Support vector machines for SAR automatic target recognition. IEEE Trans. Aerosp. Electron. Syst. 2001, 37, 643–654. [Google Scholar] [CrossRef]
  20. Chen, S.; Wang, H.; Xu, F.; Jin, Y.-Q. Target classification using the deep convolutional networks for SAR images. IEEE Trans. Geosci. Remote Sens. 2016, 54, 4806–4817. [Google Scholar] [CrossRef]
  21. 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]
  22. Scott, G.J.; England, M.R.; Starms, W.A.; Marcum, R.A.; Davis, C.H. Training deep convolutional neural networks for land–cover classification of high-resolution imagery. IEEE Geosci. Remote Sens. Lett. 2017, 14, 549–553. [Google Scholar] [CrossRef]
  23. Krizhevsky, A.; Sutskever, I.; Hinton, G.E. Imagenet classification with deep convolutional neural networks. In Advances in Neural Information Processing Systems; NIPS Conference: Lake Tahoe, NV, USA, 2012; pp. 1097–1105. [Google Scholar]
  24. He, K.; Zhang, X.; Ren, S.; Sun, J. Deep residual learning for image recognition. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition, Las Vegas, NV, USA, 26 June–1 July 2016; pp. 770–778. [Google Scholar]
  25. Simonyan, K.; Zisserman, A. Very deep convolutional networks for large-scale image recognition. arXiv, 2014; arXiv:1409.1556. [Google Scholar]
  26. 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. [Google Scholar]
  27. Huang, Z.; Pan, Z.; Lei, B. Transfer learning with deep convolutional neural network for SAR target classification with limited labeled data. Remote Sens. 2017, 9, 907. [Google Scholar] [CrossRef]
  28. Castelluccio, M.; Poggi, G.; Sansone, C.; Verdoliva, L. Land use classification in remote sensing images by convolutional neural networks. arXiv arXiv:1508.00092, 2015.
  29. Nogueira, K.; Penatti, O.A.; dos Santos, J.A. Towards better exploiting convolutional neural networks for remote sensing scene classification. Pattern Recognit. 2017, 61, 539–556. [Google Scholar] [CrossRef] [Green Version]
  30. Joachims, T. Text categorization with support vector machines: Learning with many relevant features. In Proceedings of the European Conference on Machine Learning, Chemnitz, Germany, 21–24 April 1998; Springer: Berlin, Germany, 1998; pp. 137–142. [Google Scholar]
  31. Cui, S.; Datcu, M. Supervised incremental feature coding for SAR image classification. In Proceedings of the ESA-EUSC-JRC 2014-9th Conference on Image Information Mining Conference: The Sentinels Era, Bucharest, Romania, 5–7 March 2014; pp. 85–88. [Google Scholar]
  32. Scott, G.J.; Marcum, R.A.; Davis, C.H.; Nivin, T.W. Fusion of deep convolutional neural networks for land cover classification of high-resolution imagery. IEEE Geosci. Remote Sens. Lett. 2017, 14, 1638–1642. [Google Scholar] [CrossRef]
  33. Peng, K.C.; Chen, T. Cross-layer features in convolutional neural networks for generic classification tasks. In Proceedings of the IEEE International Conference on Image Processing, Quebec City, QC, Canada, 27–30 September 2015; pp. 3057–3061. [Google Scholar]
  34. Hu, F.; Xia, G.S.; Hu, J.; Zhang, L. Transferring deep convolutional neural networks for the scene classification of high-resolution remote sensing imagery. Remote Sens. 2015, 7, 14680–14707. [Google Scholar] [CrossRef]
  35. Cui, Z.; Cao, Z.; Yang, J.; Feng, J.; Ren, H. Target recognition in synthetic aperture radar images via non-negative matrix factorisation. IET Radar Sonar Navig. 2015, 9, 1376–1385. [Google Scholar] [CrossRef]
  36. Agrawal, A.; Mangalraj, P.; Bisherwal, M.A. Target detection in SAR images using sift. In Proceedings of the 2015 IEEE International Symposium on Signal Processing and Information Technology (ISSPIT), Abu Dhabi, UAE, 7–10 December 2015; pp. 90–94. [Google Scholar]
  37. Ioffe, S.; Szegedy, C. Batch normalization: Accelerating deep network training by reducing internal covariate shift. arXiv, 2015; arXiv:1502.03167. [Google Scholar]
  38. Rumelhart, D.E.; Hinton, G.E.; Williams, R.J. Learning representations by back-propagating errors. Nature 1986, 323, 533. [Google Scholar] [CrossRef]
  39. Zeiler, M.D. Adadelta: An adaptive learning rate method. arXiv, 2012; arXiv:1212.5701. [Google Scholar]
  40. Csurka, G.; Dance, C.; Fan, L.; Willamowski, J.; Bray, C. Visual categorization with bags of keypoints. In Proceedings of the Workshop on Statistical Learning in Computer Vision, Prague, Czech Republic, 11–14 May 2004; pp. 1–2. [Google Scholar]
  41. Perronnin, F.; Dance, C. Fisher kernels on visual vocabularies for image categorization. In Proceedings of the 2007 IEEE Conference on Computer Vision and Pattern Recognition (2007), Minneapolis, MN, USA, 17–22 June 2007; pp. 1–8. [Google Scholar]
  42. Perronnin, F.; Sánchez, J.; Mensink, T. Improving the fisher kernel for large-scale image classification. In Proceedings of the European Conference on Computer Vision, Heraklion, Greece, 5–11 September 2010; Springer: Berlin, Germany, 2010; pp. 143–156. [Google Scholar]
  43. Jaakkola, T.; Haussler, D. Exploiting generative models in discriminative classifiers. In Advances in Neural Information Processing Systems; NIPS Conference: Denver, CO, USA, 1999; pp. 487–493. [Google Scholar]
  44. McLachlan, G.; Peel, D. Finite Mixture Models; John Wiley & Sons: Hoboken, NJ, USA, 2004. [Google Scholar]
  45. Sánchez, J.; Perronnin, F.; Mensink, T.; Verbeek, J. Image classification with the fisher vector: Theory and practice. Int. J. Comput. Vis. 2013, 105, 222–245. [Google Scholar] [CrossRef]
  46. Perronnin, F.; Liu, Y.; Sánchez, J.; Poirier, H. Large-scale image retrieval with compressed fisher vectors. In Proceedings of the 2010 IEEE Computer Society Conference on Computer Vision and Pattern Recognition, San Francisco, CA, USA, 13–18 June 2010; pp. 3384–3391. [Google Scholar]
  47. Lowe, D.G. Distinctive image features from scale-invariant keypoints. Int. J. Comput. Vis. 2004, 60, 91–110. [Google Scholar] [CrossRef]
  48. Chandrasekhar, V.; Lin, J.; Morère, O.; Goh, H.; Veillard, A. A practical guide to CNNs and fisher vectors for image instance retrieval. Signal Process. 2016, 128, 426–439. [Google Scholar] [CrossRef]
  49. Keydel, E.R.; Lee, S.W.; Moore, J.T. Mstar extended operating conditions: A tutorial. In Proceedings of the Algorithms for Synthetic Aperture Radar Imagery III, Orlando, FL, USA, 8–12 April 1996; pp. 228–243. [Google Scholar]
  50. The Air Force Moving and Stationary Target Recognition Database. Available online: https://www.sdms.afrl.af.mil/datasets/mstar/ (accessed on 7 February 2018).
  51. 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]
  52. Vedaldi, A.; Fulkerson, B. Vlfeat: An open and portable library of computer vision algorithms. In Proceedings of the 18th ACM international conference on Multimedia, Firenze, Italy, 25–29 October 2010; pp. 1469–1472. [Google Scholar]
Figure 1. Typical architecture of a CNN.
Figure 1. Typical architecture of a CNN.
Sensors 18 03218 g001
Figure 2. The framework of the proposed method.
Figure 2. The framework of the proposed method.
Sensors 18 03218 g002
Figure 3. Final classification vector based on the three CNN architectures (a). Feature combination of the three CNN architectures (b) Fusion of the three CNN architectures.
Figure 3. Final classification vector based on the three CNN architectures (a). Feature combination of the three CNN architectures (b) Fusion of the three CNN architectures.
Sensors 18 03218 g003aSensors 18 03218 g003b
Figure 4. Examples of the ten class SAR targets.
Figure 4. Examples of the ten class SAR targets.
Sensors 18 03218 g004
Figure 5. The time consumption of computing CNN activations with different CNN architectures for all SAR images in the MSTAR dataset.
Figure 5. The time consumption of computing CNN activations with different CNN architectures for all SAR images in the MSTAR dataset.
Sensors 18 03218 g005
Table 1. The architecture of the three CNN networks.
Table 1. The architecture of the three CNN networks.
CL-CNNMM-CNNFS-CNN
Image SizeFeature MapsKernel SizeImage SizeFeature MapsKernel SizeImage SizeFeature MapsKernel Size
Input layer128 × 1281-128 × 1281-128 × 1281-
Convolutional104 × 104825 × 25111 × 111818 × 18120 × 12089 × 9
BN104 × 1048-111 × 1118-120 × 1208-
ReLu104 × 1048-111 × 1118-120 × 1208-
pooling52 × 5282 × 237 × 3783 × 330 × 3084 × 4
Convolutional40 × 401613 × 1332 × 32166 × 626 × 26165 × 5
BN40 × 4016-32 × 3216-26 × 2616-
ReLu40 × 4016-32 × 3216-26 × 2616-
pooling20 × 20162 × 216 × 16162 × 213 × 13162 × 2
Convolutional12 × 12329 × 912 × 12325 × 512 × 12322 × 2
BN12 × 1232-12 × 1232-12 × 1232-
ReLu12 × 1232-12 × 1232-12 × 1232-
pooling6 × 6322 × 26 × 6322 × 26 × 6322 × 2
FC11152-11152-11152-
Output110-110-110-
Table 2. Type and sample number of training and testing set.
Table 2. Type and sample number of training and testing set.
TargetZIL131ZSU_234T72T62D72S1BMP2BRDM2BTR_60BTR_70
Training set299299232299299299233298256233
Testing set274274196273274274195274195196
Table 3. Type and sample number of training and testing set.
Table 3. Type and sample number of training and testing set.
ModelClassification Accuracy (%)
CL-CNN95.95
MM-CNN 96.12
FS-CNN 96.17
Table 4. Overall Classification Accuracy of FVs Coding Method Using Different CNN architecture.
Table 4. Overall Classification Accuracy of FVs Coding Method Using Different CNN architecture.
ModelClassification Accuracy (%)
CL-CNN97.29
MM-CNN 97.18
FS-CNN 96.77
Table 5. Performance of Combining Features Generated by the FC layer.
Table 5. Performance of Combining Features Generated by the FC layer.
CombinationType of CombinationAccuracy Performance (%)
CL-CNNMM-CNN Addition98.45
CL-CNN FS-CNNAddition98.31
MM-CNNFS-CNNAddition98.56
CL-CNNMM-CNN Concatenation98.37
CL-CNN FS-CNNConcatenation98.42
MM-CNNFS-CNNConcatenation98.53
CL-CNNMM-CNNFS-CNNAddition98.27
CL-CNNMM-CNNFS-CNNConcatenation98.41
Table 6. Performance of Combining Features Generated by the last convolutional layers.
Table 6. Performance of Combining Features Generated by the last convolutional layers.
CombinationType of CombinationAccuracy Performance (%)
CL-CNNMM-CNN Addition98.45
CL-CNN FS-CNNAddition98.31
MM-CNNFS-CNNAddition98.56
CL-CNNMM-CNN Concatenation98.34
CL-CNN FS-CNNConcatenation98.42
MM-CNNFS-CNNConcatenation98.53
CL-CNNMM-CNNFS-CNNAddition98.42
CL-CNNMM-CNNFS-CNNConcatenation98.53
Table 7. Performance of fusing classifiers outputs from the first set of experiments.
Table 7. Performance of fusing classifiers outputs from the first set of experiments.
Fusion CombinationType of FusionAccuracy Performance (%)
CL-CNNMM-CNN Arrogance97.48
CL-CNNMM-CNN Simple voting97.53
CL-CNNMM-CNN Weighted Sum97.40
CL-CNN FS-CNNArrogance98.19
CL-CNN FS-CNNSimple voting97.57
CL-CNN FS-CNNWeighted Sum98.19
MM-CNNFS-CNNArrogance97.94
MM-CNNFS-CNNSimple voting98.02
MM-CNNFS-CNNWeighted Sum97.90
CL-CNNMM-CNNFS-CNNArrogance98.19
CL-CNNMM-CNNFS-CNNSimple voting97.98
CL-CNNMM-CNNFS-CNNWeighted Sum98.14
Table 8. Performance of fusing classifiers outputs from the second set of experiments.
Table 8. Performance of fusing classifiers outputs from the second set of experiments.
Fusion CombinationType of fusionAccuracy Performance (%)
CL-CNNMM-CNN Arrogance98.60
CL-CNNMM-CNN Simple voting98.14
CL-CNNMM-CNN Weighted Sum98.23
CL-CNN FS-CNNArrogance98.60
CL-CNN FS-CNNSimple voting98.39
CL-CNN FS-CNNWeighted Sum98.31
MM-CNNFS-CNNArrogance98.60
MM-CNNFS-CNNSimple voting98.27
MM-CNNFS-CNNWeighted Sum98.39
CL-CNNMM-CNNFS-CNNArrogance98.60
CL-CNNMM-CNNFS-CNNSimple voting98.27
CL-CNNMM-CNNFS-CNNWeighted Sum98.35

Share and Cite

MDPI and ACS Style

Touafria, M.; Yang, Q. A Concurrent and Hierarchy Target Learning Architecture for Classification in SAR Application. Sensors 2018, 18, 3218. https://doi.org/10.3390/s18103218

AMA Style

Touafria M, Yang Q. A Concurrent and Hierarchy Target Learning Architecture for Classification in SAR Application. Sensors. 2018; 18(10):3218. https://doi.org/10.3390/s18103218

Chicago/Turabian Style

Touafria, Mohamed, and Qiang Yang. 2018. "A Concurrent and Hierarchy Target Learning Architecture for Classification in SAR Application" Sensors 18, no. 10: 3218. https://doi.org/10.3390/s18103218

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