Next Article in Journal
Estimating Daily Global Evapotranspiration Using Penman–Monteith Equation and Remotely Sensed Land Surface Temperature
Next Article in Special Issue
Hyperspectral Image Super-Resolution via Nonlocal Low-Rank Tensor Approximation and Total Variation Regularization
Previous Article in Journal
Satellite-Derived Spatiotemporal Variations in Evapotranspiration over Northeast China during 1982–2010
Previous Article in Special Issue
Spatial Resolution Enhancement of Hyperspectral Images Using Spectral Unmixing and Bayesian Sparse Representation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hyperspectral Image Spatial Super-Resolution via 3D Full Convolutional Neural Network

1
School of Electronics and Information, Northwestern Polytechnical University, Xi’an 710129, China
2
Department of Electrical and Computer Engineering, Mississippi State University, Starkville, MS 39762, USA
*
Author to whom correspondence should be addressed.
Remote Sens. 2017, 9(11), 1139; https://doi.org/10.3390/rs9111139
Submission received: 25 September 2017 / Revised: 22 October 2017 / Accepted: 3 November 2017 / Published: 7 November 2017
(This article belongs to the Special Issue Spatial Enhancement of Hyperspectral Data and Applications)

Abstract

:
Hyperspectral images are well-known for their fine spectral resolution to discriminate different materials. However, their spatial resolution is relatively low due to the trade-off in imaging sensor technologies, resulting in limitations in their applications. Inspired by recent achievements in convolutional neural network (CNN) based super-resolution (SR) for natural images, a novel three-dimensional full CNN (3D-FCNN) is constructed for spatial SR of hyperspectral images in this paper. Specifically, 3D convolution is used to exploit both the spatial context of neighboring pixels and spectral correlation of neighboring bands, such that spectral distortion when directly applying traditional CNN based SR algorithms to hyperspectral images in band-wise manners is alleviated. Furthermore, a sensor-specific mode is designed for the proposed 3D-FCNN such that none of the samples from the target scene are required for training. Fine-tuning by a small number of training samples from the target scene can further improve the performance of such a sensor-specific method. Extensive experimental results on four benchmark datasets from two well-known hyperspectral sensors, namely hyperspectral digital imagery collection experiment (HYDICE) and reflective optics system imaging spectrometer (ROSIS) sensors, demonstrate that our proposed 3D-FCNN outperforms several existing SR methods by ensuring higher quality both in reconstruction and spectral fidelity.

Graphical Abstract

1. Introduction

Hyperspectral remote sensing usually collects reflectance information of objects in hundreds of contiguous bands over a certain electromagnetic spectrum. It collects images with a very high spectral resolution, enabling a fine discrimination of different objects by their spectral signatures. However, due to the limitation of imaging sensor technologies, signal to noise ratio (SNR) and time constraints, there exists a trade-off between the spatial resolution and spectral resolution. Consequently, hyperspectral images (HSIs) are often acquired under a relatively low spatial resolution, degrading their performance in practical applications, including mineralogy, manufacturing and surveillance. Therefore, it is highly desirable to increase the spatial resolution of HSIs via post-processing.
In general, there are several ways to improve spatial resolution of HSIs: (1) image fusion with other high-spatial-resolution sources; (2) sub-pixel based analysis; and (3) single-image super-resolution (SR). The first two approaches have been widely exploited in hyperspectral applications. When an auxiliary image with a higher spatial resolution is available, such as a panchromatic image or multispectral image, image fusion can be applied for spatial-resolution enhancement. Statistics based fusion techniques are firstly proposed, such as maximum a posteriori (MAP) estimation, stochastic mixing model based method, etc. [1,2,3]. Recently, dictionary-based fusion methods dominate hyperspectral and multispectral image fusion, including spectral dictionary [4,5,6] and spatial dictionary based methods, but they cannot effectively utilize both spatial and spectral information equally. The Spatial–Temporal remotely sensed Images and land cover Maps Fusion Model (STIMFM) was proposed to produce land cover maps at both fine spatial and temporal resolutions using a series of coarse spatial resolution images together with a few fine spatial resolution land cover maps that pre- and post-date the series of coarse spatial resolution images [7]. The major limitation in these fusion techniques for HSI spatial-resolution enhancement is that an auxiliary co-registered image with a higher spatial resolution is required, which may be unavailable in practice.
Sub-pixel based analysis aims to exploit the information in the area covered by a pixel for different applications. Spectral mixture analysis (SMA) intends to estimate fractional abundance of pure ground objects within a mixed pixel [8]. Such analysis can be fulfilled by extracting endmembers [9,10] and estimating their fractional abundance [11] separately, or treating these two problems simultaneously as a blind signal decomposition problem, for which non-negative matrix factorization (NMF) [12,13], convex optimization [14], and Neural Network (NN) based techniques [15] are widely used. Sub-pixel level target detection has also been proposed to detect objects of interest within a pixel [16,17]. Soft-classification can be an option to handle the classification problem of low-spatial-resolution HSI [18]. Recently, sub-pixel mapping (SPM) techniques, which predict the location of land cover classes within a coarse pixel (mixed pixel) [19,20], have also been proposed to generate a high-resolution classification map using fractional abundance images. Various methods based on linear optimization technique [21], pixel/sub-pixel spatial attraction model [22], pixel swapping algorithm [23], maximum a posteriori (MAP) model [24,25], Markov random field (MRF) [26,27], artificial neural network (ANN) [28,29,30], simulated annealing [31], total variant model [32], support vector regression [33], and collaborative representation [34] are proposed. In general, sub-pixel based analysis only overcomes the limitation in spatial-resolution for certain applications, e.g., classification and target detection.
Single-image SR, which aims to reconstruct a high-spatial-resolution image only from a low-spatial-resolution image, can break the limitation of the inherent spatial resolution in hyperspectral imaging systems without any other prior or auxiliary information. The basic method of single-image SR is through a traditional linear interpolator, such as bilinear and bicubic interpolation. However, these methods often lead to edge blur, and a jagged and ringing effect. Villa et al. [35] attempted to split pixels into sub-pixels according to a zoom factor and to find the sub-pixel positions. However, sub-pixels are assumed to be pure pixels which may not be a reasonable assumption. In the past decades, the SR of traditional color images has gained great attention and many algorithms have been developed, such as iterative back projection (IBP) based on reconstruction [36,37] and sparse representation based algorithms [38,39]. Recently, deep learning based methods have been applied to the SR of color images and demonstrated to be of great superiority [40,41,42]. Deep convolutional NN (CNN) is designed to directly learn an end-to-end mapping between low- and high-spatial-resolution images [40]. The CNN has also been extended to a very deep version to explore contextual information over large image scenes by cascading small filters many times in a deep network structure [43]. These CNNs for the SR of color images can be directly applied to HSIs for spatial SR in a band-by-band or 3-band-group manner. For example, the msiSRCNN algorithm extended the SRCNN in [42] to spatial SR of multispectral images [44]. However, spectral distortion is often induced in such extensions since spectral correlation in contiguous bands is ignored. Recently, Li et al. applied CNN to the SR of the spectral difference in HSIs to preserve the spectral information [45,46]. The spatial constraint or spatial-error-correction model are also imposed to further correct the spatial error in the SR process. However, the spatial down-sampling function (i.e., spatial filter) is required as complemental information in the training process.
Recently, CNN has also been attempted to the SR of HSIs in the spectral dimension [47], demonstrating the feasibility and superiority of convolution to spectral dimension. Therefore, in order to alleviate spectral distortion by extending existing CNN based SR algorithms to HSIs, effectively utilizing both spatial context and spectral discrimination is of crucial importance. Such integration of spatial context and spectral discrimination has been demonstrated to be of great superiority in many hyperspectral applications, e.g., noise removal [48,49], classification [50,51], and SR [52]. In CNN based hyperspectral applications, Makantasis et al. integrated spatial-spectral information into CNN using a randomized principal component analysis (RPCA) [53]. However, the spatial-spectral features fed to CNN using RPCA cannot be directly extended for SR applications due to information loss. The 3D convolution has been demonstrated to be very effective to explore volumetric data [54,55,56,57] and successfully applied to HSIs to explore both spatial context and spectral discrimination for classification [58]. Therefore, in order to explore both spatial context and spectral discrimination for spatial SR of HSIs, a three-dimensional full CNN (3D-FCNN) framework is proposed. Specifically, the 3D convolution operation is used to explore both spatial context between neighboring pixels and spectral correlation in adjacent band images so that spectral distortion is alleviated. In order to avoid the requirement of a large amount of labelled samples to train such a 3D-FCNN, a sensor-specific manner is designed such that the 3D-FCNN trained on a certain HSI can be directly applied for the SR of other HSIs acquired by the same sensor. Finally, extensive experiments on four HSIs acquired by two well-known hyperspectral sensors, namely HYDICE and ROSIS sensors, are carried out to demonstrate the effectiveness of the proposed algorithms for spatial SR of HSIs.
In summary, the main contributions of this work can be summarized as follows:
(1)
A 3D-FCNN architecture is designed to directly learn an end-to-end mapping between low spatial-resolution and high spatial-resolution HSIs. Specifically, a 3D convolution operation is designed to explore both the spatial context between neighboring pixels and the spectral correlation in adjacent band images so that the spectral distortion is alleviated.
(2)
A sensor-specific manner is designed for the proposed 3D-FCNN to avoid the requirement of a large amount of training samples from the target scene, such that a well-trained 3D-FCNN model from an HSI can be directly applied for spatial SR of other HSIs acquired by the same sensor.
The rest of this paper is organized as follows: our proposed 3D-FCNN architecture for spatial SR of HSIs is proposed in Section 2. The experimental results on HSIs acquired by different sensors are reported in Section 3. Finally, discussions and conclusions are presented in Section 4 and Section 5, respectively.

2. Materials and Methods

Deep learning networks have achieved great success in the SR of color images. For example, SRCNN [42] aims at learning an end-to-end mapping by taking low spatial-resolution images as input and directly outputs the high spatial-resolution of the input image. However, the 2D convolutional layer utilized in SRCNN mainly takes the spatial information into consideration. When these networks are directly used for SR of HSIs in a band-by-band manner (or three bands treated as a false color image), e.g., msiSRCNN [44], it easily results in spectral distortion because the strong spectral correlation in HSIs is ignored. Therefore, in order to maintain the spectral fidelity of HSIs after spatial SR, both spatial context of adjacent pixels and spectral correlation among neighboring bands should be considered.
In this paper, 3D convolution is used to explore both spatial context and spectral correlation for spatial SR of HSIs. Consequently, a 3D full convolutional network (3D-FCNN) is proposed for single-image spatial SR of HSIs without any auxiliary information. In order to solve the problem of training deep NN in HSIs where it is very difficult to acquire a large amount of training samples, our proposed 3D-FCNN is extended to a sensor-specific manner such that it can be trained with hyperspectral datasets collected by the same sensor as the targeted dataset. As a result, the requirement of a large amount of training samples from the target scene is avoided. As shown in Figure 1, our work can be divided into the following steps: (1) Training: constructing, training and validating 3D-FCNN for SR of HSIs; (2) Testing: applying the trained network to a sensor-specific task. Specifically, the SR of an HSI can be fulfilled by using a 3D-FCNN trained by HSIs acquired with the same sensor without extra training. Moreover, when possible, such sensor-specific 3D-FCNN can be fine-tuned with only a few training data from the target HSIs to further improve the performance of SR.

2.1. Proposed 3D-FCNN for Spatial SR of HSIs

2.1.1. 2D Convolution

In a traditional 2D CNN, 2D convolution is performed to extract features from the previous layer. As shown in Figure 2, a convolution kernel is used to filter a small area of the image, so as to generate the feature value of these small regions. For each pixel of the image, the product of its neighboring pixels and the corresponding elements of the filter matrix is calculated and then added as the feature value of this pixel, which can be expressed as
c x y = f ( i , j w i j a ( x + i ) ( y + j ) + b )
where c x y is the output feature value targeted at position ( x , y ) , a ( x + i ) ( y + j ) is the input unit at position ( x + i , y + j ) with an offset of ( i , j ) to ( x , y ) , w i j is the weight for the input a ( x + i ) ( y + j ) which is located at ( i , j ) in the 2D convolution kernel, b is the bias in the convolution neuron, and f is the activation function. If the kernel has the size F × F and the input image has the size W × W , the output feature map will have a smaller size N × N , in which N = W F + 1 .
In general, a convolution layer consists of a set of learnable filters (or kernels) that have small receptive fields, which are trained to learn specific types of features at the same spatial position in the input. In addition, parameters of kernel windows are forced to be identical to all possible locations of the previous layer, which is called weight sharing. Both the weight sharing technology and small receptive fields strategy can effectively reduce the number of parameters and increase the generalization capability of the network. Weights are replicated over the input image, leading to intrinsic insensitivity to translation in the input. A convolutional layer usually contains multiple feature maps so that multiple features can be detected. The network is trained with the back propagation (BP) gradient-descent procedure.

2.1.2. 3D Convolution

Though CNN has achieved great success in 2D convolution, the 2D convolution is only applied in the 2D space to capture spatial features. In 3D hyperspectral applications, the most straightforward method is to perform 2D CNN processing on each band of HSIs. However, such a 2D convolution on multiple images separately does not explore spectral information encoded in contiguous bands, easily resulting in spectral distortion. To this end, the spectral dimension should also be considered in the convolutional kernel to extract spectral features. Therefore, in this paper, 3D convolution instead of 2D convolution is used to simultaneously conduct convolution in both spatial and spectral dimensions to capture spatial-spectral features. As shown in Figure 3a, 3D convolution is realized by convolving a 3D kernel with the cube formed by stacking multiple contiguous spectral information together. By extending 2D convolution in Equation (1), 3D convolution is calculated as the weighted sum of pixels in a 3D data cube as
c x y z = f ( i , j , k w i j k a ( x + i ) ( y + j ) ( z + k ) + b )
where c x y z is the output feature at position ( x , y , z ) , a ( x + i ) ( y + j ) ( z + k ) represents the input at the position ( x + i , y + j , z + k ) in which ( i , j , k ) denotes its offset to ( x , y , z ) , and w i j k is the weight for input a(x+i)(y+j)(z+k) with an offset of ( i , j , k ) in the 3D convolutional kernel. Similar to 2D convolution, the feature cube has a smaller size.
Similar to the 2D convolution, weight sharing technique—in which the kernel weights are replicated across the entire cube—is also used in the 3D convolution such that one kernel extracts one type of feature all over the image cube. In order to explore different kinds of spatial-spectral local feature patterns, as shown in Figure 3b, multiple 3D convolutions with distinct kernels are applied to the same location in the previous layer.

2.1.3. The Architecture of the Proposed 3D-FCNN

In this paper, as shown in Figure 4, a 3D-FCNN is constructed for SR of HSIs by using 3D convolution to fully explore spatial-spectral features in the 3D hyperspectral data cube. Considering a single-source low-resolution HSI X , it is firstly up-scaled to X with the same size of desired output Y using bicubic interpolation. Since only spatial information is taken into account in bicubic interpolation, 3D convolution is used to further improve the initial SR result X by alleviating spectral distortion. Therefore, the initial SR result X is entered into our 3D-FCNN to generate a high-spatial-resolution HSI F ( X ) approaching the desired output Y as accurately as possible, which can fully take advantages of both spatial context and spectral correlation.
The structure of our proposed 3D-FCNN is shown in Figure 4. It contains five layers including one input layer, and four convolutional layers where the output of the last convolutional layer is the output of the whole network. Generally, the number of parameters to be optimized is proportional to that of neurons in a CNN. In SR problems, the scale of the input, namely the initial super-resolution result X , heavily influences the scale of the network. In this paper, a sub-image, rather than the entire hyperspectral image, is fed to the 3D-FCNN. Specifically, as shown in Table 1, the input is restricted as a 33 × 33 × c × 1 -pixel sub-image cube, where 33 × 33 is spatial dimensions, c is the special dimension depending on the sensor properties, and the color channel is set as 1 for HSIs. Therefore, all the filers in the successive convolution layers are designed to learn spectral information from c contiguous spectral bands. It should be noted that larger sub-images can also be used as input to design the 3D-DCNN and its construction is similar to that in this paper. We restrict the network to a relatively small scale to be easily and quickly trained.
As shown in Figure 4, four 3D convolutional layers are sequentially connected to the input layer to improve the performance of initial SR result by exploring both spatial context and spectral correlation. For the first three convolutional layers, the ‘ReLU’ function is adopted as activation function for nonlinear mapping since it improves models fitting without extra computational cost and over-fitting risk. Suppose the input of neurons in activation layers is represented as I i , i = 1 , 2 , , their output is calculated in an element-wise manner as follows
O i = max { I i , 0 } , i = 1 , 2 ,
For the fourth convolutional layer, its output without the effect of activation function is used as the output of the whole network. The size of kernels in convolution layers is of crucial importance in CNN since it greatly affects the performance of the network. The size of different convolutional layers and ‘ReLU’ layers in our proposed 3D-FCNN are listed in Table 1.
According to Figure 4 and Table 1, in the proposed 3D-FCNN, 64 different 3D kernels with size 9 × 9 × 7 ( 9 × 9 in spatial dimension and 7 in the spectral dimension) are conducted on the input data to generate 64 feature maps of the size 25 × 25 × ( c 7 + 1 ) for Layer ‘Conv1’. Layer ‘Conv2’ consists of 32 features maps of the size 25 × 25 × ( c 6 ) , which are obtained by applying 32 different 3D kernels size of 1 × 1 × 1 on the input of Conv1. The third convolution layer ‘Conv3’ is obtained by applying nine different 3D kernels of size 1 × 1 × 1 , resulting in nine feature maps with a size of 25 × 25 × ( c 6 ) . Finally, the output layer ‘Conv4’ applies a 3D kernel of the size 5 × 5 × 3 to produce the output image F ( X ) with size 21 × 21 × ( c 8 ) . In order to prevent the border effects that occur during training, all convolutions are not padding. Thus, as mentioned earlier, the actual output of the network is smaller than the input. Specifically, according to Table 1, the proposed 3D-FCNN uses an initial SR result of 33 × 33 × c to generate an image of 21 × 21 × ( c 8 ) .
Training an end-to-end network requires a cost function to optimize the parameters in the network. In practice, we use the mean squared error (MSE) between outputs and ground-truth as cost function. Considering the actual output of the network is a smaller image, MSE loss function is evaluated only by the difference between the central pixels of ground-truth high-resolution image Y and the network output F ( X ) , which can be calculated as
M S E = 1 m n = 0 m ( 1 h w c i = 0 h j = 0 w k = 0 c ( Y i j k F ( X ) i j k ) )
where m is the number of training samples, h and w represent the length and width of the output of the network F ( X ) , respectively, and c is the band of the training samples. Generally, high Peak Signal to Noise Ratio (PSNR) can also be guaranteed by using MSE, which is widely used to evaluate image quality. Let n represent the number of bits per pixel. The PSNR can be calculated as
P S N R = 10 × log 10 ( ( 2 n 1 ) 2 M S E )
The loss function defined by Equation (4) is minimized using adaptive moment estimation (ADAM) with the standard BP algorithm, because it requires less memory, can calculate different adaptive learning rates for different parameters, and is suitable for large datasets and high dimensional space. The main advantage of ADAM is that the learning rate of each iteration has a definite range after the correction of bias, which makes the parameters more stable. In particular, the weight matrices are updated as
m t = μ × m t 1 + ( 1 μ ) × g t n t = v × n t 1 + ( 1 v ) × g t 2 m ^ t = m t 1 μ t n ^ t = n t 1 v t Δ θ t = m ^ t n ^ t + ε × η
where t represents time-step, g t represents the gradients at time-step t , m t is the biased first moment estimate, n t is the biased second moment estimate, m ^ t is the bias-corrected first moment estimate, n ^ t is the bias-corrected second moment estimate, η is the learning rate, θ t is the parameter vector, and μ , v [ 0 , 1 ) are exponential decay rates for the moment estimates.

2.2. Sensor-Specific Implementation

Our proposed 3D-FCNN can be extended to the spatial SR problem of hyperspectral sensors, such as ROSIS and HYDICE. Once the 3D-FCNN is well trained, its parameters can be used for a sensor-specific mission, which means that our trained network can be directly applied to different datasets obtained from the same sensor.
As shown in Figure 1, when the proposed 3D-FCNN is used for the sensor-specific mission, there are two modes as follows:
(1)
Unsupervised sensor-specific mode: the parameters involved in the network do not vary when being directly transferred to the target data obtained from the same sensor as the data trained on the proposed 3D-FCNN. As a result, any training data of the target dataset is not required. Thus, this network can be viewed as an unsupervised sensor-specific spatial SR system, which can be directly used for reconstructing high spatial-resolution HSI from the target low spatial-resolution HSI without changing the parameters of networks. In this mode, a large amount of training data and training time is avoided when it is used for the SR of the target scene. It should also be noted that such unsupervised sensor-specific mode is especially effective for the target scene which is acquired with similar imaging condition with the dataset on which the 3D-FCNN is trained.
(2)
Fine-tuning based sensor-specific mode: the parameters in the network trained under unsupervised sensor-specific mode can be further fine-tuned using the samples from the target dataset obtained by the same sensor. Through such fine-tuning, the performance of unsupervised sensor-specific spatial SR can be further improved. Such fine-tuning based sensor-specific mode is especially effective when the target scene is acquired with different imaging condition with the dataset on which the 3D-FCNN is trained. Only a small number of training samples is required, so the training based on sensor-specific initialization is very fast, compared with previous supervised training with random initialization.

3. Results

In this section, extensive experiments are conducted to verify the performance of the proposed 3D-FCNN for spatial SR of HSIs.

3.1. Datasets

Datasets from two well-known hyperspectral sensors, namely ROSIS and HYDICE, are used for evaluation. For the ROSIS sensor, two scenes that are acquired during a flight campaign over Pavia, northern Italy, i.e., Pavia Centre and Pavia University are selected. The Pavia Centre scene contains 1096 × 1096 pixels, while the Pavia University scene contains 610 × 340 pixels. In the Pavia Centre scene, samples that contain no information are discarded in this experiment and only 1096 × 715 valid pixels are used. The numbers of spectral bands are 102 for Pavia Centre and similarly 103 for Pavia University. For the HYDICE sensor, two datasets are adopted, namely the Washington DC Mall datasets and the Urban dataset. The Washington DC Mall datasets image contains 1280 × 307 pixels over 191 bands of the original 224 atmospherically corrected bands which are adopted by removing the channels associated with H2O and OH absorption bands. The Urban HYDICE datasets is of size 307 × 307 and contains 210 atmospherically corrected bands.
These four datasets are used as ground-truth of high spatial-resolution HSIs to train and evaluate the performance of the proposed 3D-FCNN based framework for spatial SR. In order to simulate low spatial-resolution HSIs, these two images are firstly down-sampled using Gauss kernels. The size of sub-images used for training is set as 33 × 33 × c × 1 , which is obtained by overlapping from the original dataset. The proposed 3D-FCNN based framework is designed, trained, and tested based on the Keras framework using the TensorFlow backend. The BP strategy is adopted to train the network. For the ADAM based training, the base learning rate is 0.00005.

3.2. Performance Evaluation

Many assessment methods have been used to evaluate the quality of a reconstructed hyperspectral image for SR and restoration [59]. In this experiment, the performance of SR for HSIs is evaluated by comparing the reconstructed high spatial-resolution HSI with the ground-truth data from two aspects: the spatial reconstruction quality of each band image at image levels and the spectral reconstruction quality of each spectrum at pixel-levels.
To evaluate the spatial reconstruction quality, the mean peak signal-to-noise ratio (MPSNR) and the mean structural similarity (MSSIM) index are adopted. The MPSNR is defined as
M P S N R = 1 c i = 0 c 10 × log 10 ( M A X i 2 M S E i )
where MAXi is the maximal pixel value in the i-th band image, and M S E i is the M S E of the i-th band image. The MSSIM between reconstructed image F ( X ) and its ground truth Y is defined as
M S S I M = 1 c i = 0 c ( 2 μ F ( X ) i μ Y i + c 1 ) ( 2 σ F ( X ) i Y i + c 2 ) ( μ 2 F ( X ) i + μ 2 Y i + c 1 ) ( σ F ( X ) i 2 + σ Y i 2 + c 2 )
where F ( X ) i and Y i represents the i-th band image of F ( X ) and Y , respectively, μ F ( X ) i and μ Y i are the mean of F ( X ) i and Y i , respectively, σ F ( X ) i 2 and σ Y i 2 are the variance of F ( X ) i and Y i , respectively, σ F ( X ) i Y i is the covariance of F ( X ) i and Y i , and c 1 and c 2 are constants that are set as 0.0001 and 0.0009, respectively.
In order to evaluate the spectral reconstruction quality, the spectral angle mapper (SAM) between the reconstructed spectra and their corresponding ground-truth spectra is used. The SAM between two spectra z and z is defined as
S A M = arccos ( z , z z 2 z 2 )
where z , z represents the dot product of z and z , and 2 represents 2-norm of vectors.
Generally, a higher MPSNR and MSSIM value means a better visual quality, and a lower SAM value implies less spectral distortion and a higher quality of spectral reconstruction.

3.3. Experimental Results on Hyperspectral Datasets Acquired by ROSIS Sensor

Two datasets acquired by ROSIS sensor, namely Pavia Centre dataset and Pavia University dataset, were used to evaluate the performance of the proposed 3D-FCNN for SR. These two datasets are complemented to 103 bands by setting the missing bands as 0.

3.3.1. Experimental Results of Spatial SR with the Proposed 3D-FCNN Method

The performance of the proposed 3D-FCNN for spatial SR is evaluated over the two selected datasets. For each dataset, a 150 × 150 sub-region is selected to validate the performance of our proposed 3D-FCNN, while the remaining pixels are used for training. In order to simulate low spatial-resolution HSIs, these two images are firstly down-sampled by a factor of 2. In addition, different levels of additive Gaussian white noise, measured by signal-to-noise ratio (SNR), are added to the low spatial-resolution images to verify the robustness of the proposed 3D-FCNN to noises.
Five strategies are chosen as the baseline for comparison: Bicubic SR, Bilinear SR, SR by nearest neighbor, SRCNN [40,42] in a band-by-band manner (namely misSRCNN in [44]) and three-band-wise manner (denoted as 3B-SRCNN). The qualitative comparison results of all the considered algorithms are listed in Table 2. Their corresponding visual results are shown in Figure 5 and Figure 6, respectively. It is also observed that: (1) the proposed 3D-FCNN obviously outperforms all the other algorithms all over these two datasets by ROSIS sensor with the highest MPSNR and MSSIM values and lowest SAM for all the cases; (2) The proposed 3D-FCNN provides better results no matter how noisy the dataset is. The average improvements in MPSNR are 2.572 dB and 2.878 dB, and in MSSIM are 0.033 and 0.032 compared with Bicubic when S N R = 30 dB and S N R = 60 dB, respectively, indicating that the proposed 3D-FCNN is robust to noises in low spatial-resolution HSI; (3) The proposed 3D-FCNN based algorithm provides much better spectral fidelity (i.e., lower SAM) compared with SRCNN (msiSRCNN and 3B-SRCNN). Figure 7 further lists the spectra of several typical ground materials after spatial SR. It can be observed that the spectra reconstructed by the proposed 3D-FCNN is closer to their ground-truth than that by the bicubic spatial interpolation, demonstrating that the proposed 3D-FCNN can better maintain spectral characteristics during spatial SR. This is because both spatial context in neighboring areas and spectral correlation in full-band images are considered in the proposed method by 3D convolution; (4) The proposed 3D-FCNN based algorithm also achieves best spatial reconstruction for all the cases, which has the highest MPSNR and MSSIM values. Compared with the Bicubic algorithm, the average improvements are 2.849 dB in terms of MPSNR and 0.035 for MSSIM.

3.3.2. Experimental Results of SR for Different Up-Sampling Factors

In this subsection, the SR results of different up-sampling factors are discussed. The low spatial-resolution HSIs are simulated by down-sampling the high spatial-resolution HSI with the factor 2, 3 and 4, respectively. Then different SR algorithms are used for spatial SR. Similar to previous experiments, only a 150 × 150 sub-region from the Pavia Centre datasets is used for validation of the performance of our proposed 3D-FCNN, while all the remaining areas are used for training. Table 3 lists the quantitative results by different SR algorithms for different up-sampling factors over the Pavia Centre area. It is observed that, the performance of our proposed 3D-FCNN outperforms all the other algorithms with different up-sampling factors. When the up-sampling factor increases, the superiority of our proposed 3D-FCNN becomes less obvious. This is because the problem of SR with a higher up-sampling factor is more difficult. However, even for 4 times spatial SR, the proposed 3D-FCNN achieves an improvement of 1.74 in MPSNR, 0.045 in MSSIM, and 0.287 in SAM, respectively, compared with the Bicubic method.

3.3.3. Experimental Results of Sensor-Specific SR by the Proposed 3D-FCNN

In this experiment, the 3D-FCNN is used for spatial SR of HSIs in a sensor-specific manner, which means that it is trained on one dataset and directly applied for spatial SR of other datasets acquired by the same sensor without any tuning of parameters. Specifically, the 3D-FCNN trained on the Pavia Centre dataset is used for spatial SR of the Pavia University dataset, and the model trained on the Pavia University dataset is used for spatial SR of the Pavia Centre dataset. Such sensor-specific manner can be viewed as an unsupervised version of our proposed 3D-FCNN for spatial SR of HSIs since training samples from the target scene are not required. In addition, the performance of fine-tuning for sensor-specific SR, in which the parameters of sensor-specific 3D-FCNN is fine-tuned with a limited number of available training samples from target datasets, is also tested.
Table 4 lists the quantitative results from our proposed 3D-FCNN for spatial SR of the two datasets from the ROSIS sensor. The results of previous supervised 3D-FCNN are also listed as benchmark results. It is observed that, the performance of spatial SR by the proposed 3D-FCNN does not degrade even if it is not trained by the targeted datasets, demonstrating the effectiveness of the proposed sensor-specific strategy for spatial SR of hyperspectral sensors. The visual results of our proposed 3D-FCNN for spatial SR of the Pavia Centre dataset and the Pavia University dataset are shown in Figure 8 and Figure 9, respectively, including both supervised mode that is trained by samples from the target scene and sensor-specific mode that is trained by other datasets from the same sensor. These results also confirmed that our proposed sensor-specific mode is very effective for spatial SR of HSIs. However, it avoids the requirement of a huge amount of training samples from the target scene and the time-consuming training.
It is also observed from Table 4 that the fine-tuning process of training sensor-specific 3D-FCNN with a few samples from the target dataset cannot improve the performance of sensor-specific very much. This may be because these two datasets are acquired during one flight and the imaging condition, e.g., acquisition time, weather, atmosphere, texture of images, etc., does not vary much. Therefore, we further evaluated the performance of our proposed 3D-FCNN on two datasets acquired at different imaging conditions by the HYDICE sensor.

3.4. Experiments on Hyperspectral Datasets Acquired by the HYDICE Sensor

Two datasets acquired by the HYDICE sensor, namely the Washington DC Mall and Urban datasets, are used to evaluate the performance of the proposed 3D-FCNN. As with the ROSISI sensor, the two datasets are complemented to 210 bands by setting the missing bands as 0.

3.4.1. Experimental Results of SR with the Proposed 3D-FCNN Method

Since the Urban dataset is too small to train the proposed 3D-FCNN, only the Washington DC Mall dataset is selected to evaluate the performance of the proposed 3D-FCNN for SR in supervised mode. Unlike the ROSIS sensor, we selected a larger region with 600 × 307 pixels to test the performance of our proposed 3D-FCNN, while the remaining pixels are used for training. The image is firstly down-sampled by a factor of 2 and additive Gaussian white noises are added to the simulated low-dimensional HSI such that the robustness of the 3D-FCNN is tested for images with different SNRs. Five strategies are chosen as the baseline for comparison: Bicubic SR, Bilinear SR, SR by nearest neighbor, SRCNN [40,42] in a band-by-band manner (namely misSRCNN in [44]) and three-band-wise manner (denoted as 3B-SRCNN).
The qualitative comparison results of all the considered algorithms are listed in Table 5. These results also confirm similar conclusions from previous experiments: (1) the performance of our 3D-FCNN obviously outperforms traditional methods, i.e., Bicubic SR, Bilinear SR and SR by nearest neighbor; (2) the spectral distortion caused by applying SRCNN of natural images directly to HSIs in band-wise manner or 3-band-group manner can be greatly alleviated by the 3D convolution in the proposed 3D-FCNN algorithm; (3) the proposed 3D-FCNN is robust to noise. Even in very noisy cases when S N R = 30 dB, our proposed 3D-FCNN improves the performance of 3B-FCNN (the second best results) by about 2.5 dB in MPSNR, 0.02 in MSSIM, and 0.2 in SAM. The visual results of all these algorithms are also listed in Figure 10, which demonstrate the superiority of the proposed 3D-FCNN for spatial SR of HSIs.

3.4.2. Experimental Results of Sensor Specific SR for the HYDICE Sensor

In this experiment, the 3D-FCNN trained on the Washington DC Mall dataset is directly applied for spatial SR of the Urban dataset in a sensor-specific manner. The performance of fine-tuning is also tested. The spatial SR results of the Urban dataset using sensor-specific 3D-FCNN are listed in Table 6. It is observed that the performance of sensor-specific 3D-FCNN can be effectively improved by the fine-tuning process, especially for the quality of spectral reconstruction. When more training samples from the target scene are used for fine-tuning, slightly better results can be achieved. This is because these two datasets are acquired under different conditions, i.e., different time, weather, atmosphere, etc. As shown in these results, the performance of sensor-specific spatial SR can be further improved when even a small amount of training samples from the target scene is used for fine-tuning. These conclusions are also confirmed by the visual results listed in Figure 11.

4. Discussion

The 3D-FCNN is proposed to learn an end-to-end full-band mapping between low and high spatial-resolution HSIs. It can effectively reconstruct high spatial-resolution HSI using a single low spatial-resolution HSI without any auxiliary information. According to previous experiments, its sensor-specific version can restore a high spatial-resolution HSI without the requirement of training on it. Generally, fine-tuning by using training samples from the target scene can improve the performance of SR. However, if the target scene is acquired with similar conditions to the scene that the 3D-FCNN is trained over, there is no need to fine-tune the network for SR since the performance of SR cannot be further improved.
Table 7 lists the computation time of three CNN based spatial SR algorithms for the Pavia Centre dataset by the ROSIS sensor. All these algorithms are implemented on a computer with one GPU card (NVIDI GTX1070 with 16GB memory). It is observed that, in order to learn the spectral correlation in adjacent band images, our proposed 3D-FCNN takes more time for training. However, once our network is well trained, it takes less than one second to reconstruct an image, indicating that it is computationally effective when working under an unsupervised sensor-specific manner.
In order to further evaluate the performance of our proposed 3D-FCNN with different parameters for SR, various experiments are conducted over the Pavia University dataset by varying different parameters involved in the proposed 3D-FCNN, including the number of convolutional layers, the size of input, activation function, filter size, and the size of receptive field. The corresponding results are listed in Table 8, Table 9, Table 10, Table 11 and Table 12.
As shown in Table 8, the 3D-FCNN with four convolutional layers slightly outperforms that with other numbers of convolutional layers. Actually, the 3D-FCNN with different convolutional layers does not vary much. Here, four convolutional layers are adopted since more convolutional layers will bring about many more parameters to be learned and the computational time for training and testing will increase rapidly.
In Table 9, the 3D-FCNN with the input of 33x33xc slightly outperforms that with other inputs. Moreover, larger inputs also result in more parameters to be trained.
The 3D-FCNN with the activation function of ‘Relu’ slightly outperforms the other two well-known activation functions, namely ‘Tanh’ and ‘PRelu’, as tabulated in Table 10.
In terms of the filter size, as shown in Table 11, the 3D-FCNN with ‘64-32-9-1’ slightly outperforms that with other filters, where the numbers in these filters represent the last dimension of filters listed in ‘Table 1’ (in the manuscript) for the four convolutional layers, respectively.
As for the size of receptive field in Table 12, the 3D-FCNN with a receptive field of 16 produces the best results. However, the number of parameters are much more than that with a smaller receptive field. Therefore, 12 is adopted to achieve the balance between the performance of SR and computational complexity.

5. Conclusions and Future Work

In this paper, a novel 3D-FCNN model is proposed for spatial SR of HSIs by learning an end-to-end full-band mapping between low and high spatial-resolution HSIs. Compared with traditional CNN based SR algorithms in color images, 3D convolution is used to reconstruct a high spatial-resolution HSI by exploring both spatial context in neighboring areas and spectral correlation in neighboring bands, such that the spectral distortion can be alleviated. The proposed 3D-FCNN is also extended to a sensor-specific version such that well-trained 3D-FCNN from a dataset can be directly used for spatial SR of other HSIs from the same sensor. The training samples from the target scene, if available, can also be used to further improve the performance of sensor-specific spatial SR. Experiments on four simulated datasets by two well-known hyperspectral sensors have demonstrated the proposed 3D-FCNN based spatial SR algorithm obviously outperforms existing methods and it is also very effective to reconstruct sensor-specific spatial SR which is more practically useful.
Recent work in SR of HSIs proposed to use CNN for SR of spectral difference of HSIs [44,45]. The proposed 3D-FCNN can also be extended for this kind of SR. In addition, using 3D-FCNN for SR of HSIs in both spatial and spectral domains is also of interest for future research.

Acknowledgments

This work is partially supported by the Fundamental Research Funds for the Central Universities (3102016ZY019, 3102016ZB012), National Natural Science Foundation of China (61671383), and National Defense Basic Scientific Research Project (JCKY2016203C067).

Author Contributions

Shaohui Mei, Xin Yuan, and Jingyu Ji provided the overall conception of this research, and designed the methodology and experiments. Xin Yuan and Shaohui Mei wrote the manuscript. Yifan Zhang, Shuai Wan and Qian Du provided improvements on the methodology and experiments, reviewed and edited the manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hardie, R.C.; Eismann, M.T.; Wilson, G.L. Map estimation for hyperspectral image resolution enhancement using an auxiliary sensor. IEEE Trans. Image Process. 2004, 13, 1174–1184. [Google Scholar] [CrossRef] [PubMed]
  2. Eismann, M.T.; Hardie, R.C. Application of the stochastic mixing model to hyperspectral resolution enhancement using auxiliary sensor. IEEE Trans. Geosci. Remote Sens. 2004, 42, 1924–1933. [Google Scholar] [CrossRef]
  3. Eismann, M.T.; Hardie, R.C. Hyperspectral resolution enhancement using high-resolution multispectral imagery with arbitrary response functions. IEEE Trans. Geosci. Remote Sens. 2005, 43, 455–465. [Google Scholar] [CrossRef]
  4. Yokoya, N.; Yairi, T.; Iwasaki, A. Coupled nonnegative matrix factorization unmixing for hyperspectral and multispectral data fusion. IEEE Trans. Geosci. Remote Sens. 2012, 50, 528–537. [Google Scholar] [CrossRef]
  5. Bendoumi, M.A.; He, M.; Mei, S. Hyperspectral image resolution enhancement using high-resolution multispectral image based on spectral unmixing. IEEE Trans. Geosci. Remote Sens. 2014, 52, 6574–6583. [Google Scholar] [CrossRef]
  6. Zhang, Y. Spatial resolution enhancement of hyperspectral image based on the combination of spectral mixing model and observation model. Proc. SPIE 2014, 9244, 201–204. [Google Scholar]
  7. Li, X.; Ling, F.; Foody, G.M.; Ge, Y.; Zhang, Y.; Du, Y. Generating a series of fine spatial and temporal resolution land cover maps by fusing coarse spatial resolution remotely sensed images and fine spatial resolution land cover maps. Remote Sens. Environ. 2017, 196, 293–311. [Google Scholar] [CrossRef]
  8. Keshava, N.; Mustard, J.F. Spectral unmixing. IEEE Signal Process. Mag. 2002, 19, 44–57. [Google Scholar] [CrossRef]
  9. Plaza, A.; Martinez, P.; Perez, R.; Plaza, J. Spatial/spectral endmember extraction by multidimensional morphological operations. IEEE Trans. Geosci. Remote Sens. 2012, 40, 2025–2041. [Google Scholar] [CrossRef]
  10. Mei, S.; He, M.; Wang, Z.; Feng, D. Spatial Purity Based Endmember Extraction for Spectral Mixture Analysis. IEEE Trans. Geosci. Remote Sens. 2010, 48, 3434–3445. [Google Scholar] [CrossRef]
  11. Mei, S.; He, M.; Wang, Z.; Feng, D. Mixture Analysis by Multichannel Hopfield Neural Network. IEEE Trans. Geosci. Remote Sens. Lett. 2010, 7, 455–459. [Google Scholar] [CrossRef]
  12. Huck, A.; Guillaume, M.; Blanc-Talon, J. Minimum Dispersion Constrained Nonnegative Matrix Factorization to Unmix Hyperspectral Data. IEEE Trans. Geosci. Remote Sens. 2010, 48, 2590–2602. [Google Scholar] [CrossRef]
  13. Qian, Y.; Jia, S.; Zhou, J.; Robles-Kelly, A. Hyperspectral Unmixing via L1/2 Sparsity-Constrained Nonnegative Matrix Factorization. IEEE Trans. Geosci. Remote Sens. 2011, 49, 4282–4297. [Google Scholar] [CrossRef]
  14. Hendrix, E.M.T.; Garcia, I.; Plaza, J.; Martin, G.; Plaza, A. A New Minimum-Volume Enclosing Algorithm for Endmember Identification and Abundance Estimation in Hyperspectral Data. IEEE Trans. Geosci. Remote Sens. 2012, 50, 2744–2757. [Google Scholar] [CrossRef]
  15. Mei, S.; He, M.; Wang, Z.; Feng, D. Unsupervised Spectral Mixture Analysis of Highly Mixed Data with Hopfield Neural Network. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2014, 7, 1922–1935. [Google Scholar] [CrossRef]
  16. Manolakis, D.; Siracusa, C.; Shaw, G. Hyperspectral subpixel target detection using the linear mixing model. IEEE Trans. Geosci. Remote Sens. 2001, 39, 1392–1409. [Google Scholar] [CrossRef]
  17. Zhang, L.; Zhang, L.; Tao, D.; Huang, X.; Du, B. Hyperspectral Remote Sensing Image Subpixel Target Detection Based on Supervised Metric Learning. IEEE Trans. Geosci. Remote Sens. 2014, 52, 4955–4965. [Google Scholar] [CrossRef]
  18. Gu, Y.; Liu, Y.; Zhang, Y. A Soft Classification Algorithm based on Spectral-spatial Kernels in Hyperspectral Images. In Proceedings of the IEEE International Conference on Innovative Computing, Information and Control, Kumamoto, Japan, 5–7 September 2007; p. 548. [Google Scholar]
  19. Atkinson, P.M. Mapping subpixel boundaries from remotely sensed images. In Innovations in GIS; Talor and Francis: London, UK, 1997; Volume 4, pp. 166–180. [Google Scholar]
  20. Foody, G.M. Sharpening fuzzy classification output to refine the representation of sub-pixel land cover distribution. Int. J. Remote Sens. 2006, 19, 2593–2599. [Google Scholar] [CrossRef]
  21. Verhoeye, J.; De Wulf, R. Land cover mapping at sub-pixel scales using linear optimization techniques. Remote Sens. Environ. 2002, 79, 96–104. [Google Scholar] [CrossRef]
  22. Mertens, K.C.; De Baets, B.; Verbeke, L.; De Wulf, R. A sub-pixel mapping algorithm based on sub-pixel/pixel spatial attraction models. Int. J. Remote Sens. 2006, 27, 3293–3310. [Google Scholar] [CrossRef]
  23. Atkinson, P.M. Sub-pixel target mapping from soft-classified, remotely sensed imagery. Photogramm. Eng. Remote Sens. 2005, 71, 839–846. [Google Scholar] [CrossRef]
  24. Pickup, L.C.; Capel, D.P.; Roberts, S.J.; Zisserman, A. Bayesian image super-resolution, continued. Adv. Neural Inf. Process. Syst. 2006, 19, 1089–1096. [Google Scholar]
  25. Xu, X.; Zhong, Y.; Zhang, L.; Zhang, H. Sub-pixel mapping based on a MAP model with multiple shifted hyperspectral imagery. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2013, 6, 580–593. [Google Scholar] [CrossRef]
  26. Kasetkasem, T.; Arora, M.K.; Varshney, P.K. Super-resolution land cover mapping using a Markov random field based approach. Remote Sens. Environ. 2005, 96, 302–314. [Google Scholar] [CrossRef]
  27. Wang, L.; Wang, Q. Subpixel mapping using Markov random field with multiple spectral constraints from subpixel shifted remote sensing images. IEEE Geosci. Remote Sens. Lett. 2013, 10, 598–602. [Google Scholar] [CrossRef]
  28. Ling, F.; Du, Y.; Xiao, F.; Xue, H.; Wu, S. Super-resolution land-cover mapping using multiple sub-pixel shifted remotely sensed images. Int. J. Remote Sens. 2010, 31, 5023–5040. [Google Scholar] [CrossRef]
  29. Mertens, K.C.; Verbeke, L.P.C.; Westra, T.; De Wulf, R.R. Subpixel mapping and sub-pixel sharpening using neural network predicted wavelet coefficients. Remote Sens. Environ. 2004, 91, 225–236. [Google Scholar] [CrossRef]
  30. Tatem, A.J.; Lewis, H.G.; Atkinson, P.M.; Nixon, M.S. Super-resolution land cover pattern prediction using a Hopfield neural network. Remote Sens. Environ. 2002, 79, 1–14. [Google Scholar] [CrossRef]
  31. Villa, A.; Chanussot, J.; Benediktsson, J.A.; Jutten, C. Spectral unmixing for the classification of hyperspectral images at a finer spatial resolution. IEEE J. Sel. Top. Signal Process. 2011, 5, 521–533. [Google Scholar] [CrossRef]
  32. Feng, R.; Zhong, Y.; Xu, X.; Zhang, L. Adaptive sparse subpixel mapping with a total variation model for remote sensing imagery. IEEE Trans. Geosci. Remote Sens. 2016, 54, 2855–2872. [Google Scholar] [CrossRef]
  33. Zhang, Y.; Du, Y.; Ling, F.; Fang, S.; Li, X. Example-based super-resolution land cover mapping using support vector regression. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2014, 7, 1271–1283. [Google Scholar] [CrossRef]
  34. Zhang, Y.; Xue, X.; Wang, T.; He, M. A Hybrid Subpixel Mapping Framework for Hyperspectral Images Using Collaborative Representation. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2017. [Google Scholar] [CrossRef]
  35. Villa, A.; Chanussot, J.; Benediktsson, J.A.; Ulfarsson, M.; Jutten, C. Super-resolution: An efficient method to improve spatial resolution of hyperspectral image. In Proceedings of the 2010 IEEE International Geoscience and Remote Sensing Symposium, Honolulu, HI, USA, 25–30 July 2010; Volume 38. [Google Scholar]
  36. Dong, W.S.; Zhang, L.; Shi, G.M.; Wu, X. Nonlocal Back-Projection for Adaptive Image Enlargement. In Proceedings of the 2009 16th IEEE International Conference on Image Processing (ICIP), Cairo, Egypt, 7–10 November 2009; pp. 349–352. [Google Scholar]
  37. Chen, H.C.; Wang, W.J. Two Stage Interpolation Algorithm Based on Fuzzy Logics and Edges Features for Image Zooming. EURASIP J. Adv. Signal Process. 2009, 1, 121–128. [Google Scholar] [CrossRef]
  38. Yang, J.; Wright, J.; Huang, T.S.; Ma, Y. Image super-resolution via sparse representation. IEEE Trans. Image Process. 2010, 19, 2861–2873. [Google Scholar] [CrossRef] [PubMed]
  39. Gao, X.; Zhang, K.; Tao, D.; Li, X. Image super-resolution with sparse neighbor embedding. IEEE Trans. Image Process. 2012, 21, 3194–3205. [Google Scholar] [PubMed]
  40. Dong, C.; Loy, C.C.; He, K.; Tang, X. Learning a Deep Convolutional Network for Image Super-Resolution. In European Conference on Computer Vision; Springer: Zurich, Switzerland, 2014; pp. 184–199. [Google Scholar]
  41. Shi, W.; Caballero, J.; Huszar, F.; Totz, J.; Aitken, A.P.; Bishop, R.; Rueckert, D.; Wang, Z. Real-Time Single Image and Video Super-Resolution Using an Efficient Sub-Pixel Convolutional Neural Network. In Proceedings of the 2016 IEEE Conference on Computer Vision and Pattern Recognition, Las Vegas, NV, USA, 27–30 June 2016; pp. 1874–1883. [Google Scholar]
  42. Dong, C.; Loy, C.C.; He, K.; Tang, X. Image Super-Resolution Using Deep Convolutional Networks. IEEE Trans. Pattern Anal. Mach. Intell. 2016, 38, 295–307. [Google Scholar] [CrossRef] [PubMed]
  43. Kim, J.; Lee, J.K.; Lee, K.M. Accurate Image Super-Resolution Using Very Deep Convolutional Networks. In Proceedings of the 2016 IEEE Conference on Computer Vision and Pattern Recognition (CVPR), Las Vegas, NV, USA, 27–30 June 2016. [Google Scholar]
  44. Liebel, L.; Körner, M. Single-Image Super Resolution for Multispectral Remote Sensing Data Using Convolutional Neural Networks. ISPRS-Int. Arch. Photogramm. Remote Sens. Spat. Inf. Sci. 2016, XLI-B3, 883–890. [Google Scholar] [CrossRef]
  45. Li, Y.; Hu, J.; Zhao, X.; Xie, W.; Li, J. Hyperspectral image super-resolution using deep convolutional neural network. Neurocomputing 2017, 266, 29–41. [Google Scholar] [CrossRef]
  46. Hu, J.; Li, Y.; Xie, W. Hyperspectral Image Super-Resolution by Spectral Difference Learning and Spatial Error Correction. IEEE Geosci. Remote Sens. Lett. 2017, 14, 1825–1829. [Google Scholar] [CrossRef]
  47. Galliani, S.; Lanaras, C.; Marmanis, D.; Baltsavias, E.; Schindler, K. Learned Spectral Super-Resolution. arXiv, 2017; arXiv:1703.09470. [Google Scholar]
  48. Liu, S.; Jiao, L.; Yang, S. Hierarchical sparse learning with spectral-spatial information for hyperspectral imagery denoising. Sensors 2016, 16, 1718. [Google Scholar] [CrossRef] [PubMed]
  49. Li, J.; Yuan, Q.; Shen, H.; Zhang, L. Noise Removal from Hyperspectral Image with Joint Spectral-Spatial Distributed Sparse Representation. IEEE Trans. Geosci. Remote Sens. 2016, 54, 5425–5439. [Google Scholar] [CrossRef]
  50. Fauvel, M.; Tarabalka, Y.; Benediktsson, J.A.; Chanussot, J.; Tilton, J.C. Advances in Spectral-Spatial Classification of Hyperspectral Images. Proc. IEEE 2013, 101, 652–675. [Google Scholar] [CrossRef]
  51. Wang, Q.; Meng, Z.; Li, X. Locality Adaptive Discriminant Analysis for Spectral-Spatial Classification of Hyperspectral Images. IEEE Geosci. Remote Sens. Lett. 2017, 11, 2077–2081. [Google Scholar] [CrossRef]
  52. Li, J.; Yuan, Q.; Shen, H.; Meng, X.; Zhang, L. Hyperspectral Image Super-Resolution by Spectral Mixture Analysis and Spatial-Spectral Group Sparsity. IEEE Geosci. Remote Sens. Lett. 2016, 13, 1250–1254. [Google Scholar] [CrossRef]
  53. 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]
  54. Ji, S.; Xu, W.; Yang, M.; Yu, K. 3D Convolutional Neural Networks for Human Action Recognition. IEEE Trans. Pattern Anal. Mach. Intell. 2013, 35, 221–231. [Google Scholar] [CrossRef] [PubMed]
  55. Wu, Z.; Song, S.; Khosla, A.; Yu, F.; Zhang, L.; Tang, X.; Xiao, J. 3D ShapeNets: A deep representation for volumetric shapes. In Proceedings of the 2015 IEEE Conference on Computer Vision and Pattern Recognition (CVPR), Boston, MA, USA, 7–12 June 2015; pp. 1912–1920. [Google Scholar]
  56. Qi, C.R.; Su, H.; Niebner, M.; Dai, A.; Yan, M.; Guibas, L.J. Volumetric and Multi-view CNNs for Object Classification on 3D Data. In Proceedings of the 2016 IEEE Conference on Computer Vision and Pattern Recognition (CVPR), Las Vegas, NV, USA, 27–30 June 2016; pp. 5648–5656. [Google Scholar]
  57. Milletari, F.; Navab, N.; Ahmadi, S.A. V-Net: Fully Convolutional Neural Networks for Volumetric Medical Image Segmentation. In Proceedings of the 2016 Fourth International Conference on 3D Vision (3DV), Stanford, CA, USA, 25–28 October 2016; pp. 565–571. [Google Scholar]
  58. Li, Y.; Zhang, H.; Shen, Q. Spectral-spatial classification of hyperspectral imagery with 3D convolutional neural network. Remote Sens. 2017, 9, 67. [Google Scholar] [CrossRef]
  59. Yang, J.; Zhao, Y.; Yi, C.; Chan, J.-W. No-Reference Hyperspectral Image Quality Assessment via Quality-Sensitive Features Learning. Remote Sens. 2017, 9, 305. [Google Scholar] [CrossRef]
Figure 1. Illustration of the proposed 3D-FCNN for SR of HSIs.
Figure 1. Illustration of the proposed 3D-FCNN for SR of HSIs.
Remotesensing 09 01139 g001
Figure 2. Illustration of 2D convolution to extract spatial features.
Figure 2. Illustration of 2D convolution to extract spatial features.
Remotesensing 09 01139 g002
Figure 3. Illustration of 3D convolution: (a) illustration of a 3D kernel to extract spatial-spectral features; (b) illustration of multiple 3D kernels to extract different kinds of spatial-spectral local feature patterns.
Figure 3. Illustration of 3D convolution: (a) illustration of a 3D kernel to extract spatial-spectral features; (b) illustration of multiple 3D kernels to extract different kinds of spatial-spectral local feature patterns.
Remotesensing 09 01139 g003
Figure 4. Framework of the proposed 3D-FCNN for SR of HSIs.
Figure 4. Framework of the proposed 3D-FCNN for SR of HSIs.
Remotesensing 09 01139 g004
Figure 5. The visual results of spatial SR on the Pavia Centre dataset, in which the area in the red rectangle is enlarged in the upper left corner of the image.
Figure 5. The visual results of spatial SR on the Pavia Centre dataset, in which the area in the red rectangle is enlarged in the upper left corner of the image.
Remotesensing 09 01139 g005
Figure 6. The visual results of spatial SR on the Pavia University dataset, in which the area in the red rectangle is enlarged in the upper left corner of the image.
Figure 6. The visual results of spatial SR on the Pavia University dataset, in which the area in the red rectangle is enlarged in the upper left corner of the image.
Remotesensing 09 01139 g006
Figure 7. Example spectra of several typical materials in this image scene: (a) trees; (b) meadows; and (c) asphalt.
Figure 7. Example spectra of several typical materials in this image scene: (a) trees; (b) meadows; and (c) asphalt.
Remotesensing 09 01139 g007
Figure 8. Visual results of spatial SR for the Pavia Centre dataset by the proposed 3D-FCNN: (a) supervised mode by training the network using samples from the Pavia Centre dataset; (b) sensor-specific mode by training the network using sample from the ROSIS sensor, i.e., the Pavia University dataset. The area in the red rectangle is enlarged in the upper left corner of the image.
Figure 8. Visual results of spatial SR for the Pavia Centre dataset by the proposed 3D-FCNN: (a) supervised mode by training the network using samples from the Pavia Centre dataset; (b) sensor-specific mode by training the network using sample from the ROSIS sensor, i.e., the Pavia University dataset. The area in the red rectangle is enlarged in the upper left corner of the image.
Remotesensing 09 01139 g008
Figure 9. Visual results of spatial SR for the Pavia University dataset by the proposed 3D-FCNN: (a) supervised mode by training the network using samples from the Pavia University dataset; (b) sensor-specific mode by training the network using sample from the ROSIS sensor, i.e., the Pavia Centre dataset. The area in the red rectangle is enlarged in the lower right corner of the image.
Figure 9. Visual results of spatial SR for the Pavia University dataset by the proposed 3D-FCNN: (a) supervised mode by training the network using samples from the Pavia University dataset; (b) sensor-specific mode by training the network using sample from the ROSIS sensor, i.e., the Pavia Centre dataset. The area in the red rectangle is enlarged in the lower right corner of the image.
Remotesensing 09 01139 g009
Figure 10. The spatial SR results of the Washington DC Mall dataset for low-dimensional images with different SNRs. The results of msiSRCNN, 3B-SRCNN and the proposed 3D-FCNN are enlarged in the upper left corner.
Figure 10. The spatial SR results of the Washington DC Mall dataset for low-dimensional images with different SNRs. The results of msiSRCNN, 3B-SRCNN and the proposed 3D-FCNN are enlarged in the upper left corner.
Remotesensing 09 01139 g010
Figure 11. Visual results of spatial SR over the Urban dataset. (a) Ground-Truth (High spatial-resolution HSI); (b) Sensor-specific SR; (c) Fine-tuning with 50 samples; (d) Fine-tuning with 100 samples; (e) Fine-tuning with 150 samples.
Figure 11. Visual results of spatial SR over the Urban dataset. (a) Ground-Truth (High spatial-resolution HSI); (b) Sensor-specific SR; (c) Fine-tuning with 50 samples; (d) Fine-tuning with 100 samples; (e) Fine-tuning with 150 samples.
Remotesensing 09 01139 g011
Table 1. Details of the proposed 3D-CNN for SR of HSIs.
Table 1. Details of the proposed 3D-CNN for SR of HSIs.
LayerInput SizeKernel SizeOutput Size
Conv1 33 × 33 × c × 1 9 × 9 × 7 × 64 25 × 25 × ( c 6 ) × 64
Relu1 25 × 25 × ( c 6 ) × 64 - 25 × 25 × ( c 6 ) × 64
Conv2 25 × 25 × ( c 6 ) × 64 1 × 1 × 1 × 32 25 × 25 × ( c 6 ) × 32
Relu2 25 × 25 × ( c 6 ) × 32 - 25 × 25 × ( c 6 ) × 32
Conv3 25 × 25 × ( c 6 ) × 32 1 × 1 × 1 × 9 25 × 25 × ( c 6 ) × 9
Relu3 25 × 25 × ( c 6 ) × 9 25 × 25 × ( c 6 ) × 9
Conv4 25 × 25 × ( c 6 ) × 9 5 × 5 × 3 × 1 21 × 21 × ( c 8 ) × 1
Table 2. Quantitative comparison results of different SR methods on HSIs by ROSIS sensor.
Table 2. Quantitative comparison results of different SR methods on HSIs by ROSIS sensor.
SNR (dB)BicubicBilinearNearestmsiSRCNN3B-SRCNN3D-FCNN
Pavia Centre MPSNR31.129.90929.98332.47932.68633.916
MSSIM0.9370.9150.9210.9570.960.969
SAM4.5925.0094.7864.6174.6614.14
60MPSNR31.09729.90729.97932.23832.74633.97
MSSIM0.9370.9150.9210.9550.960.969
SAM4.6065.0194.8054.7654.6934.163
30MPSNR31.08329.929.96632.24832.76233.86
MSSIM0.9370.9140.9210.9550.960.968
SAM4.6065.0464.864.7934.6874.218
Pavia University MPSNR29.49928.59928.75229.25330.08332.381
MSSIM0.9250.9060.9140.9230.9360.964
SAM3.9464.384.2434.1974.0713.454
60MPSNR29.49528.59728.74829.09630.1331.701
MSSIM0.9250.9060.9140.920.9390.958
SAM3.9654.3914.2674.3234.0413.807
30MPSNR29.48428.59128.73629.17329.89831.852
MSSIM0.9250.9060.9130.9210.9350.958
SAM4.0184.4254.3334.3354.1423.778
Table 3. Quantitative comparison of spatial SR for different up-sampling factors over the Pavia Centre dataset.
Table 3. Quantitative comparison of spatial SR for different up-sampling factors over the Pavia Centre dataset.
BicubicBilinearNearestmsiSRCNN3B-SRCNN3D-FCNN
Pavia Centre n = 2 MPSNR31.129.90929.98332.47932.68633.916
MSSIM0.9370.9150.9210.9570.960.969
SAM4.5925.0094.7864.6174.6614.14
n = 3 MPSNR28.51927.90727.5729.15929.26129.853
MSSIM0.8820.8590.8570.8890.9050.921
SAM5.5715.855.8645.495.3765.300
n = 4 MPSNR26.8326.32726.2126.2127.25627.59
MSSIM0.8080.7830.7920.7920.8510.862
SAM6.3626.6946.6166.6166.4046.336
n is the factor of up-sampling.
Table 4. Quantitative comparative results of the proposed 3D-FCNN for SR of datasets from the ROSIS sensor in different modes.
Table 4. Quantitative comparative results of the proposed 3D-FCNN for SR of datasets from the ROSIS sensor in different modes.
SupervisedSensor-SpecificFine-Tune
50100150
Pavia CentreMPSNR34.90834.89934.93734.97834.998
MSSIM0.9790.9790.9770.9790.979
SAM4.3964.4474.4294.4234.424
SupervisedSensor-SpecificFine-Tune
50100150
Pavia UniversityMPSNR34.51934.50634.4934.50234.559
MSSIM0.9770.9770.9760.9760.976
SAM3.163.1723.1683.1463.122
50, 100, 150 is the number of fine-tuning samples.
Table 5. Quantitative results of different methods for spatial SR of the Washington DC Mall dataset.
Table 5. Quantitative results of different methods for spatial SR of the Washington DC Mall dataset.
SNR(dB)BicubicBilinearNearestmsiSRCNN3B-SRCNN3D-FCNN
Washington DC Mall MPSNR47.58846.7646.8945.86649.6253.556
MSSIM0.9810.9790.980.9560.9760.989
SAM0.4550.490.4620.7180.560.357
60MPSNR42.50341.87641.46542.92943.90846.028
MSSIM0.9780.9750.9730.9660.9710.987
SAM0.5440.5640.6020.6470.6140.412
30MPSNR39.00438.71337.92440.9441.74344.288
MSSIM0.9630.9650.9540.960.9680.984
SAM0.7490.7110.8440.7350.6680.481
Table 6. Quantitative results of the proposed 3D-FCNN for sensor-specific spatial SR.
Table 6. Quantitative results of the proposed 3D-FCNN for sensor-specific spatial SR.
Sensor-SpecificFine-Tune
50100150
UrbanMPSNR27.56328.92229.03929.091
MSSIM0.890.930.9320.933
SAM14.0411.6711.52511.457
50, 100, 150 is the number of fine-tuning samples.
Table 7. The computation time of different SR methods on the Pavia Centre dataset.
Table 7. The computation time of different SR methods on the Pavia Centre dataset.
msiSRCNN3B-SRCNN3D-FCNN
Training7 h9 h20 h
Test0.492 s0.536 s0.426 s
Table 8. Experimental results of the proposed 3D-FCNN with different numbers of convolutional layers over the Pavia University datasets while other parameters are fixed as listed in Table 1.
Table 8. Experimental results of the proposed 3D-FCNN with different numbers of convolutional layers over the Pavia University datasets while other parameters are fixed as listed in Table 1.
Number of 3D Convonlution LayersMPSNRSAM
333.644.27
433.924.14
533.784.19
633.704.29
Table 9. Experimental results of the proposed 3D-FCNN with different sizes of input over the Pavia University datasets while other parameters are fixed as listed in Table 1.
Table 9. Experimental results of the proposed 3D-FCNN with different sizes of input over the Pavia University datasets while other parameters are fixed as listed in Table 1.
Size of InputMPSNRSAM
33 × 33 × c33.924.14
44 × 44 × c33.844.21
55 × 55 × c33.794.25
Table 10. Experimental results of the proposed 3D-FCNN with different activation functions over the Pavia University datasets while other parameters are fixed as listed in Table 1.
Table 10. Experimental results of the proposed 3D-FCNN with different activation functions over the Pavia University datasets while other parameters are fixed as listed in Table 1.
Activation FunctionMPSNRSAM
Relu33.924.14
Tanh33.374.19
PRelu33.764.29
Table 11. Experimental results of the proposed 3D-FCNN with different filter sizes over the Pavia University datasets while other parameters are fixed as listed in Table 1.
Table 11. Experimental results of the proposed 3D-FCNN with different filter sizes over the Pavia University datasets while other parameters are fixed as listed in Table 1.
Filter SizeMPSNRSAM
64-32-9-133.924.14
64-32-16-133.854.17
32-16-8-133.674.24
Table 12. Experimental results of the proposed 3D-FCNN with different sizes of receptive field over the Pavia University datasets while other parameters are fixed as listed in Table 1.
Table 12. Experimental results of the proposed 3D-FCNN with different sizes of receptive field over the Pavia University datasets while other parameters are fixed as listed in Table 1.
Receptive FieldMPSNRSAMParameters
1233.924.1439403
1433.944.1355789
1634.034.1188557

Share and Cite

MDPI and ACS Style

Mei, S.; Yuan, X.; Ji, J.; Zhang, Y.; Wan, S.; Du, Q. Hyperspectral Image Spatial Super-Resolution via 3D Full Convolutional Neural Network. Remote Sens. 2017, 9, 1139. https://doi.org/10.3390/rs9111139

AMA Style

Mei S, Yuan X, Ji J, Zhang Y, Wan S, Du Q. Hyperspectral Image Spatial Super-Resolution via 3D Full Convolutional Neural Network. Remote Sensing. 2017; 9(11):1139. https://doi.org/10.3390/rs9111139

Chicago/Turabian Style

Mei, Shaohui, Xin Yuan, Jingyu Ji, Yifan Zhang, Shuai Wan, and Qian Du. 2017. "Hyperspectral Image Spatial Super-Resolution via 3D Full Convolutional Neural Network" Remote Sensing 9, no. 11: 1139. https://doi.org/10.3390/rs9111139

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