Next Article in Journal
A Literature Review of Energy Efficiency and Sustainability in Manufacturing Systems
Previous Article in Journal
Evaluation of the Erosion Characteristics for a Marine Pump Using 3D RANS Simulations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fast and Stable Hyperspectral Multispectral Image Fusion Technique Using Moore–Penrose Inverse Solver

State Key Laboratory of High Performance Computing, College of Computer Science, National University of Defense Technology, Changsha 410073, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(16), 7365; https://doi.org/10.3390/app11167365
Submission received: 4 July 2021 / Revised: 3 August 2021 / Accepted: 9 August 2021 / Published: 10 August 2021

Abstract

:
Fusion low-resolution hyperspectral images (LR-HSI) and high-resolution multispectral images (HR-MSI) are important methods for obtaining high-resolution hyperspectral images (HR-HSI). Some hyperspectral image fusion application areas have strong real-time requirements for image fusion, and a fast fusion method is urgently needed. This paper proposes a fast and stable fusion method (FSF) based on matrix factorization, which can largely reduce the computational workloads of image fusion to achieve fast and efficient image fusion. FSF introduces the Moore–Penrose inverse in the fusion model to simplify the estimation of the coefficient matrix and uses singular value decomposition (SVD) to simplify the estimation of the spectral basis, thus significantly reducing the computational effort of model solving. Meanwhile, FSF introduces two multiplicative iterative processes to optimize the spectral basis and coefficient matrix to achieve stable and high-quality fusion. We have tested the fusion method on remote sensing and ground-based datasets. The experiments show that our proposed method can achieve the performance of several state-of-the-art algorithms while reducing execution time to less than 1% of such algorithms.

1. Introduction

Hyperspectral images with rich spectral information play an important role in agriculture [1], medicine [2], geological surveying [3], remote sensing [4], computer vision [5,6], and other fields. Generally, the resolution of hyperspectral images directly affects the application effect of hyperspectral images. However, the spatial resolution of hyperspectral images directly obtained with hyperspectral cameras is relatively low. On the contrary, the multispectral image directly obtained by the multispectral camera has a higher spatial resolution. Obtaining high-resolution hyperspectral images by improving hardware processes is usually costly, so fusing high-resolution multispectral images with low-resolution hyperspectral images is one of the most economical ways to obtain high-resolution hyperspectral images.
In the practical application of the fusion method, two images of the same area obtained by the hyperspectral camera and the multispectral camera can be pre-processed by image alignment [7] and enhancement. Then, the feature points in the two images can be extracted and matched separately, and then the two images can be geometrically aligned. After alignment, the two images can be used as input images for the fusion algorithm. The application flow chart is shown in Figure 1a. However, in this paper, in order to compare the fusion ability of various algorithms more effectively, we use the original image in the dataset as the target HR-HSI in the experimental part, simulate the LR-HSI obtained by the hyperspectral camera by spatially blurred downsampling of the target HR-HSI, and simulate the HR-MSI obtained by the multispectral camera by spectral downsampling of the target HR-HSI. The LR-HSI and HR-MSI involved in the fusion are already aligned images, so that we can easily check the fusion performance of the fusion algorithm by comparing the similarity between the fused images and the target image (the better the similarity, the better the fusion effect). The application flow chart is shown in Figure 1b.
The earliest methods of spatial-spectral image fusion can be traced back to pan-sharpening techniques, which fuse low spatial resolution RGB images with high resolution panchromatic images to obtain high-resolution RGB images. Representative pan-sharpening techniques include component substitution [8], methods based on principal component analysis [9], and methods based on compressed sensing [10]. Since panchromatic images contain only one piece of spectral information, pan-sharpening techniques usually lead to spectral distortion. To enhance the fusion quality, HR-MSI is usually used instead of panchromatic images to fuse with LR-HSI. Typical fusion methods include Bayesian-based fusion methods, spectral unmixing-based fusion methods, matrix decomposition-based fusion methods, and tensor decomposition-based fusion methods.
Eismannet et al. first used the Bayesian method based on maximum posterior probability (MAP) estimation to realize fusion. This method uses a stochastic mixed model (SMM) to estimate potential spectral scene features and formulates a cost function to optimize the corresponding hyperspectral image data estimated by LR-HSI and HR-MSI. The optimization of MAP-SMM is carried out in the principal component subspace. The idea of using two input images to merge spectral information in a subspace is the primary source of inspiration for many later fusion methods [11,12,13]. Compared with pan-sharpening technology, this fusion method has made a significant breakthrough in improving the spatial resolution of hyperspectral images [14,15].
With the development of fusion technology, fusion methods based on spectral unmixing [16,17,18] have been proposed, based on the fact that hyperspectral images often have a large number of so-called “mixed pixels”, i.e., overlapping spectral responses of several different surface materials, due to the much broader pixel coverage. The spectral unmixing approach tries to separate these responses, i.e., to recover the underlying pure material spectra (called endmembers) and their relative proportions (called abundances). Since LR-HSI and HR-MSI capture the same scene, the underlying pure matter members are identical. Therefore, it can extract a spectral dictionary from one image to interpret the other, thus achieving high spectral image super-resolution. Based on linear spectral unmixing, the literature [19] has proposed a coupled spectral unmixing method, which solves the super-resolution problem by alternating the spectral unmixing of HR-MSI and LR-HSI simultaneously while updating spectral and abundance. However, the fusion quality of this method is closely related to the estimation of the number of underlying endmembers. The requirement for estimating the underlying endmembers is high, and it is not easy to achieve an accurate estimation. Therefore, in the literature [20], a sparse constraint is added to the fusion. As a result, a hyperspectral, multispectral image fusion algorithm based on sparse representation is proposed (NSSR). Through updating the number of the underlying endmembers by an effective non-negative dictionary learning algorithm, NSSR effectively improves the fusion quality. However, this method demands more parameters to be determined, and the fusion quality fluctuates wildly.
In recent years, matrix factorization-based fusion methods have started to emerge. The matrix factorization-based fusion methods assume HR-HSI as the product of the spectral basis matrix and a coefficient matrix. They estimate the spectral basis from LR-HSI and coefficient matrix from HR-MSI, respectively, to achieve hyperspectral super-resolution. The authors of [21] propose the coupled non-negative matrix factorization (CNMF) technology, which extracts the spectral basis and coefficient matrix by performing non-negative matrix factorization (NMF) on LR-HSI and HR-MSI, respectively. However, since the NMF is usually not unique, the results produced in [21] are not always good. In [13], a standard linear inverse problem model for LR-HSI and HR-MSI fusion was developed to solve the optimized spectral basis and coefficient matrices by introducing full variational regularization terms in the pairwise matrix equations. However, this approach usually requires splitting of variables requiring relatively more computational time as well. Further studies began to add a priori information to the fusion to improve the quality of the fusion. For example, in [22], by adding spatial similarity to the fusion, and in [23], by using the nonlocal coefficient representation method and the self-similarity constraint to impose a priori constraints on the fusion, the quality of image fusion can be improved to some extent. However, the fusion process becomes more complicated due to the added constraints, which leads to a further increase in the fusion time and makes the efficiency for fusion decrease. Thus, for some application scenarios that require real-time response, this fusion method is not suitable.
Unlike matrix factorization-based methods, tensor factorization-based methods can preserve the three-dimensional characteristics of image data. This has led to the continuous application of tensor factorization-based methods to image processing work, such as multi-frame data denoising [24], completion [25], compression perception [26], and classification [27]. Tensor factorization-based methods for solving hyperspectral image fusion have also been proposed. Based on the Tucker decomposition, [28] views HR-HSI as the product of a kernel tensor and three-factor matrices and estimates the kernel tensor and factor matrices from HR-MSI and LR-HSI, respectively. The autocorrelation of hyperspectral images is cited in [29], and the method of nonlocal tensor factorization is proposed to effectively use the nonlocal similarity information to realize hyperspectral image fusion. However, this method based on Tucker decomposition requires more intermediate variables to be estimated, which requires more computational effort and usually results in a slower fusion rate. A fusion method based on canonical polyadic decomposition (CPD) is proposed in [30], which decomposes the target HR-HSI into the sum of vector products of three dimensions. It effectively reduces the computational effort of tensor decomposition-based methods. However, this approach assumes that the ranks of the three dimensions are the same (which is not valid in some cases), so this usually leads to undesirable fusion results. In [31], the advantages of Tucker decomposition and CP decomposition are combined, and a fusion method based on tensor ring factorization is proposed. This method views the target HR-HSI as a matrix product of three dimensions and achieves hyperspectral image super-resolution by estimating the spectral dimension matrix from LR-HSI and the two matrices of spatial dimension from HR-MSI. This fusion method usually requires the estimation of three matrices, which also has a significant computation and can lead to poor timeliness of fusion.
Although the recently proposed tensor factorization-based method can effectively preserve the three-dimensional characteristics of hyperspectral images, the tensor factorization-based method also has to be operated by converting to a two-dimensional matrix, and the fusion quality is not substantially improved. Compared with the tensor factorization-based method, the matrix factorization-based method has better timeliness. This paper proposes a fast and stable fusion method based on matrix factorization to ensure the fusion quality and shorten the fusion time. This method first solves the fuzzy estimation of the spectral basis through a simple singular value decomposition method. Then, through the Moore–Penrose inversion operation, it calculates the fuzzy estimate of the spectral coefficient matrix and then obtains the optimized spectral base and coefficient matrix estimate through a multiplication iterative optimization process. Finally, the exact HR-HSI estimate is obtained by doing another multiplicative iterative optimization operation on the fuzzy version of the HR-HSI obtained in the previous step.
Overall, this paper presents a fast and stable hyperspectral multispectral image fusion technique. The significant contributions of this work are as follows.
  • It is the first time that the Moore–Penrose inverse has been used to determine the solution of the spectral coefficient matrix, which simplifies the solution process of the fusion model, dramatically reduces the computational workload, and shortens the fusion time.
  • A multiplicative iterative optimization process is used to optimize the spectral basis and coefficient matrices, enabling further enhancement of the fused results and improving the quality of the fusion.
  • Previous fusion methods stop optimizing the fusion results after obtaining the spectral basis and coefficient matrices. In contrast, the method proposed in this paper continues to optimize the rough HR-HSI estimates after fusion to further improve the fusion quality after obtaining the spectral basis and coefficient matrices.
The rest of this paper is organized as follows. Section 2 details the proposed method. Section 3 gives the experimental settings. The results and analysis are presented in Section 4, and the conclusion is given in Section 5.

2. Proposed Method

Our proposed method is a matrix factorization-based fusion method, which requires first transforming the 3D hyperspectral image data into 2D matrix data and then performing the corresponding calculations. Usually, we denote the target HR-HSI as Z R W × H × L , with W × H pixels and L bands; the LR-HSI involved in the fusion as H R w × h × L , with w × h pixels and L bands; and the HR-MSI as M R W × H × l , with W × H pixels and l bands, and their mode-3 matricization matrices are Z R L × W H , H R L × w h , and M R l × W H , respectively. We denote a three-dimensional tensor as X R I 1 × I 2 × I 3 , its ( i , j , k ) elements as X i j k or x i j k , and its Frobenius norm as X F = i j k x i j k 2 .

2.1. Fusion Model

In the fusion-based approach, LR-HSI and HR-MSI are usually assumed to be the spatial blur downsampling and spectral downsampling of HR-HSI, respectively. Furthermore, the spectral response function can be estimated from the hyperspectral camera and multispectral camera parameters, and the estimation of spatial blur downsampling can be approximated by several spatial blur estimation functions. Thus, the following equation can be obtained:
H = Z BS
where B R W H × W H and S R W H × w h represent the convolution blur and the downsampling matrix, respectively. The HR-MSI can be seen as the spectrally downsampled version of HR-HSI, i.e.,
M = R Z
where R R l × L denotes the spectral response matrix. As mentioned above, both the spectral response function and spatial fuzzy downsampling can be obtained. Therefore, the matrices B , S , and R in Equations (1) and (2) are all determined, so solving HR-HSI means solving the unknown matrix Z in the above matrix equation. In matrix factorization-based methods, it is usually assumed that the spectral basis exists in a lower subspace [13], i.e.,
Z = D C
where D R L × L s and C R L s × W H are the spectral bases and spectral basis coefficient matrix, respectively. Therefore, the following equations can be obtained:
H = DC BS
M = R DC

2.2. Moore–Penrose Inverse

From Equations (4) and (5), only the matrices D and C are unknown, and combining Equation (3), Z can be solved by solving the matrices D and C . Since the spectral information in HR-HSI is mainly from LR-HSI, the idea that LR-HSI and HR-HSI contain the same spectral basis was proposed and verified in [22]. Therefore, we can perform SVD on LR-HSI to obtain the preliminary spectral basis.
H = U Σ V T
Since the first few singular values of a matrix can contain more than 90% of the matrix information, we save the first q singular values in Σ to obtain the following spectral basis, i.e.,
D = U ( : , 1 : q )
In Equation (5), only the matrix C is unknown, and the matrix C can be obtained by inverting the matrix RD .
C = ( RD ) M
Since the matrix RD is not necessarily a square matrix, the inverse of the matrix RD in Equation (8) is not unique, which will cause errors in the solution results. However, the Moore–Penrose inverse (“+” inverse) is unique among all inverse matrices of the matrix, so the fuzzy estimate of the matrix C can be obtained by introducing the Moore–Penrose inverse as follows.
C = ( RD ) + M
According to the definition of the matrix Moore–Penrose inverse, we can solve the Moore–Penrose inverse of the matrix RD through the Algorithm 1.
Algorithm 1 Moore–Penrose inverse
Require: RD
U 1 , Σ 1 , V 1 = SVD RD
m , n = size Σ 1
K = min m , n
for k = 1:K do
  if Σ 1 ( k , k ) ! = 0 then
    Σ 1 ( k , k ) = 1 / Σ 1 ( k , k )
  end if
end for
Σ 2 = zeros n , m
for k = 1:K do
   Σ 2 ( k , k ) = Σ 1 ( k , k )
end for
( RD ) + = V 1 Σ 2 U 1 T
return ( RD ) +

2.3. Optimized Fusion Solution

From above, we have obtained the fuzzy estimates of spectral basis D and coefficients C . However, the fused images generated from the fuzzy estimates D , C and the fusion quality are unsatisfactory. Therefore, based on Equation (4), we can optimize the spectral basis D by solving the optimization procedure of the following equation.
min D H D C B S F 2
In Equation (10), we assume that X = CBS ; then, Equation (10) can be resolved as
min D H D X F 2
We extend the multiplicative update rule in [21] to obtain an approximate solution for solving Equation (11), i.e.,
D = D . × H X T . / D X X T
After obtaining the optimized spectral base D , we can obtain the HR-HSI fuzzy estimation matrix Z 1 as
Z 1 = D C
Obtaining the HR-HSI estimation result Z 1 , fusion quality still has room for further improvement. Combining Equation (2), we can obtain
M = R Z 1
Combining the known matrix M and R to further optimize the fusion result Z 1 ,
min Z 1 M R Z 1 F 2
Similarly, we use the extended multiplication iteration rule to solve Equation (15), and we can obtain
Z 1 = Z 1 . × R T M . / R T R Z 1
After iterative optimization, we can obtain a more accurate fused image Z = Z 1 . Finally, our proposed method can be described by Algorithm 2, where q represents the parameter to be adjusted.
Algorithm 2 FSF Algorithm
Require: H M R B S q
 Get D by Equations (6) and (7)
 Get C by Algorithm 2
for k = 1:K do
  Update D by Equation (12)
end for
Z 1 = D C
for k = 1:K do
  Update Z 1 by Equation (16)
end for
Z = Z 1
return Z

3. Experiments

To validate the superior performance of our proposed method, we compare it with six state-of-the-art methods on two commonly used datasets, Pavia University and CAVE. Six commonly used quality metrics are used to measure the effectiveness of the fusion.

3.1. Date Set

We use two datasets, among which Pavia University [32] belongs to the remote sensing dataset. It is one of two scenes acquired by the ROSIS sensor during a flight campaign over Pavia, northern Italy. It contains 103 spectral bands (0.43–0.86 μ m), and each band includes 610 × 610 pixels, but some of the samples in the images contain no information and have to be discarded before the analysis. The geometric resolution is 1.3 m. After removing the absorbing bands and some undesirable pixel points, we finally keep 93 bands and 256 × 256 pixel points as the reference HR-HSI image. We generate the 4-band HR-MSI (4 × 256 × 256) using the IKONOS class reflection spectral response filter [12] as a spectral response function. We used a 7 × 7 Gaussian blur (standard deviation 2) for the reference image and then sampled down every 32 pixels in both spatial dimensions to simulate LR-HSI (93 × 8 × 8 ).
The CAVE dataset [33] is a ground-based hyperspectral dataset containing 32 high-quality indoor HSIs, each containing 512 × 512 pixels in 31 spectral bands, of which the first two bands are blurred. Its spectral range is 400 nm 700 nm, and the wavelength interval is 10 nm. We remove the first two bands and finally keep 29 bands and 512 × 512 pixel points as the reference HR-HSI image (29 × 512 × 512). Also, to simulate LR-HSI, we use 7 × 7 Gaussian blur (standard deviation 2) and then sample down every 32 pixels in both spatial dimensions (29 × 16 × 16). To simulate HR-MSI, we use Gf-1-16m multispectral camera response to generate the three-band HR-MSI (RGB image) (3 × 512 × 512).

3.2. Compared Methods

We use six state-of-the-art algorithms for comparison, including three methods based on Tucker decomposition: low tensor-train rank (LTTR)-based method [34], coupled sparse tensor factorization (CSTF) [28], nonlocal sparse tensor factorization (NLSTF) [29], and a method based on CP decomposition Super-resolution TEnsor-REcOnstruction (STEREO) [30] (we use the non-blind version). In addition, we use the matrix decomposition-based method Hysure [13] and the spectral unmixing-based method non-negative structured sparse representation (NSSR) [20]. To demonstrate the reliability of the experimental results, all source codes were obtained through open access, and the codes for the various experiments were run in the same environment. This paper’s experiments were implemented in MATLAB R2014a on localhost with 2.40 GHz Intel i7-7700 CPU and 8.0 GB DDR3.

3.3. Quantitative Metrics

To comprehensively evaluate the superior performance of our proposed method, we used six quantitative metrics in our experiments. These include the peak signal-to-noise ratio (PSNR) [22] based on an error-sensitive quality metric; an index to evaluate the mean square error between the fused image and the reference image (ERGAS) [35]; the spectral angle mapper (SAM) [12] to evaluate the spectral angle shift of the fused image; and the universal quality index (UIQI) [36] and the structural similarity (SSIM) [37], the last of which is an important evaluation metric for evaluating the time required for fusion.
The first five metrics evaluate the similarity between the fused image and the target image from several aspects, and the better the result of the metrics, the better the effect of the fusion. A higher similarity between the fused image and the target image indicates that the LR-HSI obtained after spatial blurring downsampling achieves better resolution enhancement after fusion.

3.4. Parameters Discussion

For our proposed method FSF, only one parameter is to be adjusted, the dimensionality of spectral bases q. When extracting spectral bases by SVD on LR-HSI, the dimensionality of spectral bases will affect the fusion results. Usually, the spectral bases exist in a small subspace, and we vary the dimension of the spectral bases from 1 to 10 to find the most suitable number of spectral bases. Figure 2a shows the PSNR values of the fused images when the number of spectral bases is varied from 1 to 10 in the Pavia University dataset, and it can be seen that the best fusion results are obtained when the dimension of spectral bases is 4. In Figure 2b, corresponding to the results on the CAVE dataset, it is clear that the best results are obtained when the spectral basis dimension is 3. The other methods used for comparison, CSTF https://github.com/renweidian/CSTF, Hysure https://github.com/alfaiate/HySure, LTTR https://github.com/renweidian/LTTR, NLSTF https://github.com/renweidian/NLSTF, STEREO https://sites.google.com/site/harikanats/, and NSSR http://see.xidian.edu.cn/faculty/wsdong, adjusted the parameters that affect their fusion results to achieve the best fusion results. The best parameters after tuning corresponding to these methods are shown in Table 1 and Table 2.

3.5. Experimental Setting

To fully demonstrate the timeliness and stability of our proposed method, we conduct comparative experiments in the following order. First, we conduct comparative experiments on the CAVE and Pavia University datasets to test the fusion effect and the excellent timeliness of our proposed method. Then, we test the fusion effects of the various fusion methods at downsampling factors of 4, 8, 16, 32, and 64 to verify the stability of our proposed method. Note that the downsampling factor is the number of pixel points per pixel point interval taken when simulating LR-HSI. For example, when the downsampling factor is 4, the size of the LR-HSI simulated from Pavia University data is 93 × 64 × 64, and when the downsampling factor is 16, the size of the corresponding LR-HSI is 93 × 16 × 16. The larger the downsampling factor is, the stronger the super-resolution capability of the fusion is. Here, we take the maximum downsampling factor to 64, i.e., to achieve a super-resolution of 64 times the spatial dimension, and the size of the LR-HSI involved in the fusion is only 93 × 4 × 4 at this time.

4. Results and Analysis

We first conducted experiments on the Pavia University and CAVE datasets at a downsampling factor of 32. Then, to describe the experimental results objectively and fairly, we show the subjective results of the fused images and the objective results of the six fusion metrics, respectively. Figure 3 shows false-color images from LR-HSI, fusion images, and ground truth images. It can be seen that the fused image is clearer than LR-HSI.

4.1. Timeliness of The FSF

Figure 4 shows the fusion results of the compared methods on the Pavia University dataset. When there are more black pixels in the error image, the fusion result is better. It is evident from the figure that NLSTF has the worst fusion result, and the error image contains a large number of white spots. The STEREO fusion result is also unsatisfactory, and the error image contains many white pixel points. The main reason for the unsatisfactory fusion results of NLSTF and STEREO is that the downsampling factor influences the fusion effect of NLSTF and STEREO, and the larger the downsampling factor is, the worse the fusion effect is. The fusion quality of NSSR is usually closely related to the selection of endmembers, and the fusion quality of pixel points at the edges of endmembers is generally poor. Therefore, we can see that the error images of the fusion results obtained by the NSSR method also contain many noticeable white pixels. Among all the methods, only our proposed method, FSF, gives the best fusion results with the least number of white pixel points in the error image.
Table 3 corresponds to the quantitative metrics of the fusion results of the competing approaches on the Pavia University dataset. Our proposed method achieves the best results in all metrics, and only the CSTF and Hysure methods achieve fusion results similar to ours when the time required for fusion is not taken into account. When considering the running time, CSTF and Hysure consume 289.643 s and 60.61 s, respectively, while our proposed method FSF takes only 0.26 s, consuming less than 0.1% of the time for CSTF and less than 1% that for Hysure. Aside from our FSF, the fastest algorithm is STEREO; this is because STEREO reduces the tensor calculation in three dimensions to a one-dimensional vector to calculate, which significantly improves the speed of fusion, but even STEREO takes 8.088 s.
Figure 5 shows the fusion results of the compared methods on the CAVE dataset. Similarly, the NLSTF and STEREO methods obtained the worst fusion results due to the poor fusion of these two methods at 32× super-resolution. Our FSF method still achieves better fusion results.
Table 4 corresponds to the quantitative metrics of the fusion results of the competing approaches on the CAVE dataset. In terms of the quantitative evaluation metrics of fusion, our method achieves similar or even slightly better fusion results than Hysure and CSTF methods. However, in terms of the time required for fusion, Hysure and CSTF required 239.961 s and 386.388 s, respectively, while our proposed FSF method required only 0.561 s. Thus, the time required was less than 0.2% of that for CSTF and less than 0.3% of that for Hysure. In addition to FSF, the shortest fusion time, STEREO, also required 25.213 s. Thus, FSF needed only 1/20 of STEREO, while the fusion effect was much better than STEREO.

4.2. Stability of The FSF

Figure 6 shows the PSNR value of the fusion image when the downsampling factors are 4, 8, 16, 32, and 64 for the comparison method. Our proposed method FSF can maintain a high fusion effect under different downsampling factors, while other methods will decrease the fusion quality as the downsampling factor increases. The stability of our proposed method FSF is based on the fact that in matrix singular value decomposition, only the first few singular values need to be retained to recover the information of the original matrix. Furthermore, CSTF and Hysure are also stable for the downsampling factors since both methods also use SVD to extract part of the fusion information in the fusion process. On the other hand, the NLSTF method only achieves better fusion results when the downsampling factor is 4, while the fusion results are the worst in all other cases since NLSTF uses a nonlocal clustering operation before fusion, limiting the super-resolution capability.

5. Conclusions

In this paper, we propose a new fast and stable fusion algorithm based on matrix factorization. Unlike the traditional matrix factorization approach, our proposed factorization method takes the lead in using the Moore–Penrose inverse to simplify the estimation process of the spectral basis coefficient matrix, which significantly reduces the computation time of the fusion model. In addition, to further stabilize the quality of the fusion, we use two multiplicative iterative processes to optimize the quality of the fused image. In addition, we are the first to propose a different optimization method for the first generation of spectral basis and coefficient matrices to make the fusion results more accurate. Finally, our method is tested on two commonly used public datasets, verifying that our proposed method has better timeliness and stability than other state-of-the-art algorithms. In future research, more a priori information can be introduced into the fusion model proposed in this paper to improve the quality of fusion further.

Author Contributions

Conceptualization, Y.P. and T.Z.; methodology, J.L. (Jian Long); software, J.L. (Jian Long); validation, J.L. (Jian Long), J.L. (Jun Li) and L.Z.; data curation, J.L.; writing—original draft preparation, J.L. (Jian Long); writing—review and editing, Y.P., T.Z., J.L. (Jun Li) and L.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially supported by the National Key Research and Development Program of China (Nos. 2017YFB1301104 and 2017YFB1001900) and the National Natural Science Foundation of China (Nos. 91648204 and 61803375).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Sample Availability

The code in the article is available from the authors.

References

  1. Samiappan, S. Spectral Band Selection for Ensemble Classification of Hyperspectral Images with Applications to Agriculture and Food Safety; Mississippi State University: Starkwell, MS, USA, 2014. [Google Scholar]
  2. Liu, Z.; Yan, J.Q.; Zhang, D.; Li, Q.L. Automated tongue segmentation in hyperspectral images for medicine. Appl. Opt. 2007, 46, 8328–8334. [Google Scholar] [CrossRef] [Green Version]
  3. Pechanec, V.; Mráz, A.; Rozkošný, L.; Vyvlečka, P. Usage of Airborne Hyperspectral Imaging Data for Identifying Spatial Variability of Soil Nitrogen Content. ISPRS Int. J. Geoinf. 2021, 10, 355. [Google Scholar] [CrossRef]
  4. Meer, F.D.V.D.; Werff, H.M.A.V.D.; Ruitenbeek, F.J.A.V.; Hecker, C.A.; Bakker, W.H.; Noomen, M.F.; Meijde, M.V.D.; Carranza, E.J.M.; Smeth, J.B.D.; Woldai, T. Multi- and hyperspectral geologic remote sensing: A review. Int. J. Appl. Earth Obs. Geoinf. 2012, 14, 112–128. [Google Scholar] [CrossRef]
  5. Cui, Y.; Zhang, B.; Yang, W.; Yi, X.; Tang, Y. Deep CNN-based Visual Target Tracking System Relying on Monocular Image Sensing. In Proceedings of the 2018 International Joint Conference on Neural Networks (IJCNN), Rio de Janeiro, Brazil, 8–13 July 2018. [Google Scholar]
  6. Cui, Y.; Zhang, B.; Yang, W.; Wang, Z.; Li, Y.; Yi, X.; Tang, Y. End-to-End Visual Target Tracking in Multi-robot Systems Based on Deep Convolutional Neural Network. In Proceedings of the IEEE International Conference on Computer Vision Workshops, Venice, Italy, 22–29 October 2017; pp. 1113–1121. [Google Scholar]
  7. Irani, M.; Anandan, P. Robust Multi-Sensor Image Alignment. In Proceedings of the Sixth International Conference on Computer Vision, Bombay, India, 7 January 1998. [Google Scholar]
  8. Carper, W.J.; Lillesand, T.M.; Kiefer, P.W. The use of intensity-hue-saturation transformations for merging SPOT panchromatic and multispectral image data. Photogramm. Eng. Remote Sens. 1990, 56, 459–467. [Google Scholar]
  9. Shah, V.P.; Younan, N.H.; King, R.L. An Efficient Pan-Sharpening Method via a Combined Adaptive PCA Approach and Contourlets. IEEE Trans. Geosci. Remote Sens. 2008, 46, 1323–1335. [Google Scholar] [CrossRef]
  10. Cheng, J.; Zhang, H.; Shen, H.; Zhang, L. A Practical Compressed Sensing-Based Pan-Sharpening Method. IEEE Geosci. Remote Sens. Lett. 2012, 9, 629–633. [Google Scholar] [CrossRef]
  11. Zhang, Y.; Backer, S.D.; Scheunders, P. Noise-Resistant Wavelet-Based Bayesian Fusion of Multispectral and Hyperspectral Images. IEEE Trans. Geosci. Remote Sens. 2009, 47, 3834–3843. [Google Scholar] [CrossRef]
  12. Wei, Q.; Bioucas-Dias, J.; Dobigeon, N.; Tourneret, J.Y. Hyperspectral and Multispectral Image Fusion based on a Sparse Representation. IEEE Trans. Geosci. Remote Sens. 2015, 53, 3658–3668. [Google Scholar] [CrossRef] [Green Version]
  13. Simoes, M.; Bioucas-Dias, J.; Almeida, L.B.; Chanussot, J. A Convex Formulation for Hyperspectral Image Superresolution via Subspace-Based Regularization. IEEE Trans. Geosci. Remote Sens. 2015, 53, 3373–3388. [Google Scholar] [CrossRef] [Green Version]
  14. Hardie, R.C..; Russell, C.; Eismann, M.T.; Michael, T.; Wilson, G.L.; Gregory, L. MAP Estimation for Hyperspectral Image Resolution Enhancement Using an Auxiliary Sensor. IEEE Trans. Image Process. 2004, 13, 1174–1184. [Google Scholar] [CrossRef] [PubMed]
  15. Eismann, M.T. Resolution Enhancement of Hyperspectral Imagery Using Maximum a Posteriori Estimation with a Stochastic Mixing Model. Ph.D. Thesis, University of Dayton, Dayton, OH, USA, 2004. [Google Scholar]
  16. Iordache, M.D.; Bioucas-Dias, J.M.; Plaza, A. Sparse Unmixing of Hyperspectral Data. IEEE Trans. Geosci. Remote Sens. 2011, 49, 2014–2039. [Google Scholar] [CrossRef] [Green Version]
  17. Zhukov, B.; Oertel, D.; Lanzl, F.; Reinhackel, G. Unmixing-based multisensor multiresolution image fusion. IEEE Trans. Geosci. Remote Sens. 1999, 37, 1212–1226. [Google Scholar] [CrossRef]
  18. Minghelli-Roman, A.; Polidori, L.; Mathieu-Blanc, S.; Loubersac, L.; Auneau, F.C. Spatial resolution improvement by merging MERIS-ETM images for coastal water monitoring. IEEE Geosci. Remote Sens. Lett. 2006, 3, 227–231. [Google Scholar] [CrossRef]
  19. Grohnfeldt, C.; Bamler, R. Jointly Sparse Fusion of Hyperspectral and Multispectral Imagery. In Proceedings of the 2013 IEEE International Geoscience and Remote Sensing Symposium—IGARSS, Melbourne, VIC, Australia, 21–26 July 2013. [Google Scholar]
  20. Dong, W.; Fu, F.; Shi, G.; Cao, X.; Wu, J.; Li, G.; Li, X. Hyperspectral Image Super-Resolution via Non-Negative Structured Sparse Representation. IEEE Trans. Image Process. 2016, 25, 2337–2352. [Google Scholar] [CrossRef]
  21. Yokoya, N.; Member, S.; 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]
  22. Dian, R.; Li, S. Hyperspectral Image Super-Resolution via Subspace-Based Low Tensor Multi-Rank Regularization. IEEE Trans. Image Process. 2019, 28, 5135–5146. [Google Scholar] [CrossRef] [PubMed]
  23. Long, J.; Peng, Y.; Li, J.; Zhang, L.; Xu, Y. Hyperspectral Image Super-resolution via Subspace-based Fast Low Tensor Multi-Rank Regularization. Infrared Phys. Technol. 2021, 116, 103631. [Google Scholar] [CrossRef]
  24. Dong, W.; Li, G.; Shi, G.; Xin, L.; Yi, M. Low-Rank Tensor Approximation with Laplacian Scale Mixture Modeling for Multiframe Image Denoising. In Proceedings of the 2015 IEEE International Conference on Computer Vision (ICCV), Santiago, Chile, 7–13 December 2015. [Google Scholar]
  25. Ji, L.; Musialski, P.; Wonka, P.; Ye, J. Tensor completion for estimating missing values in visual data. IEEE Trans. Pattern Anal. Mach. Intell. 2013, 35, 208–220. [Google Scholar]
  26. Zhao, Q.; Zhang, L.; Cichocki, A. Bayesian CP Factorization of Incomplete Tensors with Automatic Rank Determination. IEEE Trans. Pattern Anal. Mach. Intell. 2015, 37, 1751–1763. [Google Scholar] [CrossRef] [Green Version]
  27. Xian, G.; Xin, H.; Zhang, L.; Zhang, L.; Plaza, A.; Benediktsson, J.A. Support Tensor Machines for Classification of Hyperspectral Remote Sensing Imagery. IEEE Trans. Geosci. Remote Sens. 2016, 54, 3248–3264. [Google Scholar]
  28. Li, S.; Dian, R.; Fang, L.; Bioucas-Dias, J.M. Fusing Hyperspectral and Multispectral Images via Coupled Sparse Tensor Factorization. IEEE Trans. Image Process. 2018, 27, 4118–4130. [Google Scholar] [CrossRef]
  29. Dian, R.; Fang, L.; Li, S. Hyperspectral Image Super-Resolution via Non-local Sparse Tensor Factorization. In Proceedings of the 2017 IEEE Conference on Computer Vision and Pattern Recognition (CVPR), Honolulu, HI, USA, 21–26 July 2017; pp. 3862–3871. [Google Scholar] [CrossRef]
  30. Kanatsoulis, C.I.; Fu, X.; Sidiropoulos, N.D.; Ma, W. Hyperspectral Super-Resolution: A Coupled Tensor Factorization Approach. IEEE Trans. Signal Process. 2018, 66, 6503–6517. [Google Scholar] [CrossRef] [Green Version]
  31. He, W.; Chen, Y.; Yokoya, N.; Li, C.; Zhao, Q. Hyperspectral Super-Resolution via Coupled Tensor Ring Factorization. arXiv 2020, arXiv:2001.01547. [Google Scholar]
  32. Dell’Acqua, F.; Gamba, P.; Ferrari, A.; Palmason, J.A.; Benediktsson, J.A.; Arnason, K. Exploiting spectral and spatial information in hyperspectral urban data with high resolution. IEEE Geosci. Remote Sens. Lett. 2004, 1, 322–326. [Google Scholar] [CrossRef]
  33. Yasuma, F.; Mitsunaga, T.; Iso, D.; Nayar, S.K. Generalized Assorted Pixel Camera: Postcapture Control of Resolution, Dynamic Range, and Spectrum. IEEE Trans. Image Process. 2010, 19, 2241–2253. [Google Scholar] [CrossRef] [Green Version]
  34. Dian, R.; Li, S.; Fang, L. Learning a Low Tensor-Train Rank Representation for Hyperspectral Image Super-Resolution. IEEE Trans. Neural Netw. Learn. Syst. 2019, 30, 2672–2683. [Google Scholar] [CrossRef]
  35. Wald, L. Quality of high resolution synthesised images: Is there a simple criterion? In Third Conference Fusion of Earth Data: Merging Point Measurements, Raster Maps and Remotely Sensed Images; SEE/URISCA: Nice, France, 2000; pp. 99–103. [Google Scholar]
  36. Wang, Z.; Bovik, A.C. A universal image quality index. IEEE Signal Process. Lett. 2002, 9, 81–84. [Google Scholar] [CrossRef]
  37. Yokoya, N.; Grohnfeldt, C.; Chanussot, J. Hyperspectral and multispectral data fusion: A comparative review of the recent literature. IEEE Geosci. Remote Sens. Mag. 2017, 5, 29–56. [Google Scholar] [CrossRef]
Figure 1. Flowchart for achieving super-resolution using a fusion approach.
Figure 1. Flowchart for achieving super-resolution using a fusion approach.
Applsci 11 07365 g001
Figure 2. FSF fusion quality on two datasets with different number of subspaces.
Figure 2. FSF fusion quality on two datasets with different number of subspaces.
Applsci 11 07365 g002
Figure 3. First row: LR-HSI, fusion images, and ground truth images’ false-color images (formed by bands 30th, 60th, and 90th) in Pavia university. Second row: LR-HSI, fusion images, and ground truth images’ false-color images (formed by bands 10th, 19th, and 28th) of oil_painting in the CAVE database.
Figure 3. First row: LR-HSI, fusion images, and ground truth images’ false-color images (formed by bands 30th, 60th, and 90th) in Pavia university. Second row: LR-HSI, fusion images, and ground truth images’ false-color images (formed by bands 10th, 19th, and 28th) of oil_painting in the CAVE database.
Applsci 11 07365 g003
Figure 4. First row: the competing approaches to reconstructed images of Pavia University at 30th band. Second row: the corresponding error images of the 30th band. Third row: the competing approaches to reconstructed images’ false-color images formed by bands 30th, 60th, and 90th. Fourth row: the corresponding error images of the false-color images.
Figure 4. First row: the competing approaches to reconstructed images of Pavia University at 30th band. Second row: the corresponding error images of the 30th band. Third row: the competing approaches to reconstructed images’ false-color images formed by bands 30th, 60th, and 90th. Fourth row: the corresponding error images of the false-color images.
Applsci 11 07365 g004
Figure 5. First row: the competing approaches to reconstructed images of oil_painting in the CAVE database at 10th band. Second row: the corresponding error images of the 10th band. Third row: the competing approaches to reconstructed images’ false-color images formed by bands 10th, 19th, and 28th. Fourth row: the corresponding error images of the false-color images.
Figure 5. First row: the competing approaches to reconstructed images of oil_painting in the CAVE database at 10th band. Second row: the corresponding error images of the 10th band. Third row: the competing approaches to reconstructed images’ false-color images formed by bands 10th, 19th, and 28th. Fourth row: the corresponding error images of the false-color images.
Applsci 11 07365 g005
Figure 6. Comparison of the fusion results of the methods with different downsampling factors.
Figure 6. Comparison of the fusion results of the methods with different downsampling factors.
Applsci 11 07365 g006
Table 1. Parameter settings on the Pavia University dataset.
Table 1. Parameter settings on the Pavia University dataset.
Pavia University
MethodParameters
CSTFW = 400; H = 400; S = 8; λ = 1 × 10 6 ;
Hysure λ _ m = 1 × 10 5 ; λ _ p h i = 1 × 10 1 ;
LTTRK = 3900; η = 0.002 ;
NLSTFW = 10; H = 10; S = 14; λ = 1 × 10 6 ; K = 151; λ 1 = 1 × 10 5 ; λ 2 = 1 × 10 5 ; λ 3 = 1 × 10 5 ;
STEREOs_iter = 1; kernel_length = 31; t_rank = 40; maxit = 100; λ = 1 × 10 4 ;
NSSRpar.ro = 1.3; par.Iter = 30; par.K = 40; par.lambda = 1 × 10 5 ;
Table 2. Parameter settings on the CAVE dataset.
Table 2. Parameter settings on the CAVE dataset.
CAVE
MethodParameters
CSTFW = 400; H = 400; S = 10; λ = 1 × 10 5 ;
Hysure λ _ m = 1 × 10 3 ; λ _ p h i = 1 × 10 7 ;
LTTRK = 900; η = 0.002 ;
NLSTFW = 10; H = 10; S = 14; λ = 1 × 10 6 ; K = 151; λ 1 = 1 × 10 5 ; λ 2 = 1 × 10 5 ; λ 3 = 1 × 10 5 ;
STEREOs_iter = 2; kernel_length = 21; t_rank = 40; maxit = 5; λ = 1 × 10 3 ;
NSSRpar.ro = 1.3; par.Iter = 30; par.K = 40; par.lambda = 1 × 10 5 ;
Table 3. The quantitative results of the experiment on the Pavia University dataset. The best results are shown in bold among all methods.
Table 3. The quantitative results of the experiment on the Pavia University dataset. The best results are shown in bold among all methods.
Pavia University
MethodPSNRERGASSAMUIQISSIMTime
Best Value + 00110
CSTF43.0720.1442.1230.9930.989289.643
Hysure42.3130.1582.2550.9920.98860.61
LTTR32.9320.6237.2630.9180.937485.533
NLSTF29.5510.6727.3880.9240.93685.319
STEREO27.0510.8129.1080.7670.6898.088
NSSR38.270.2292.4660.9850.98160.676
FSF(Ours)43.0770.1412.0580.9930.9890.26
Table 4. The quantitative results of the experiment on the CAVE dataset. The best results are shown in bold among all methods.
Table 4. The quantitative results of the experiment on the CAVE dataset. The best results are shown in bold among all methods.
CAVE
MethodPSNRERGASSAMUIQISSIMTime
Best Value + 00110
CSTF40.6910.4637.9530.9140.957386.388
Hysure40.2630.42115.230.9340.957239.961
LTTR34.4821.13812.2680.810.914768.284
NLSTF24.631.52114.4750.7170.813246.843
STEREO31.2750.69517.1710.7380.83525.213
NSSR38.3790.55214.3370.9220.952125.605
FSF(Ours)41.0670.38813.9260.9370.9590.561
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Long, J.; Peng, Y.; Zhou, T.; Zhao, L.; Li, J. Fast and Stable Hyperspectral Multispectral Image Fusion Technique Using Moore–Penrose Inverse Solver. Appl. Sci. 2021, 11, 7365. https://doi.org/10.3390/app11167365

AMA Style

Long J, Peng Y, Zhou T, Zhao L, Li J. Fast and Stable Hyperspectral Multispectral Image Fusion Technique Using Moore–Penrose Inverse Solver. Applied Sciences. 2021; 11(16):7365. https://doi.org/10.3390/app11167365

Chicago/Turabian Style

Long, Jian, Yuanxi Peng, Tong Zhou, Liyuan Zhao, and Jun Li. 2021. "Fast and Stable Hyperspectral Multispectral Image Fusion Technique Using Moore–Penrose Inverse Solver" Applied Sciences 11, no. 16: 7365. https://doi.org/10.3390/app11167365

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