Next Article in Journal
Research Progress on Router Devices for the OAM Optical Communication
Next Article in Special Issue
Synthetic Iris Images: A Comparative Analysis between Cartesian and Polar Representation
Previous Article in Journal
Quantum Efficiency Measurement and Modeling of Silicon Sensors Optimized for Soft X-ray Detection
Previous Article in Special Issue
A Selection of Starting Points for Iterative Position Estimation Algorithms Using Feedforward Neural Networks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Snapshot Multi-Spectral Demosaicing Method for Multi-Spectral Filter Array Images Based on Channel Attention Network

1
Xi’an Institute of Optics and Precision Mechanics, Chinese Academy of Sciences, Xi’an 710119, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
3
Institute of Aerospace Science and Technology, School of Remote Sensing and Information Engineering, Wuhan University, Wuhan 430072, China
*
Authors to whom correspondence should be addressed.
Sensors 2024, 24(3), 943; https://doi.org/10.3390/s24030943
Submission received: 9 January 2024 / Revised: 26 January 2024 / Accepted: 29 January 2024 / Published: 1 February 2024
(This article belongs to the Special Issue Nature-Inspired Algorithms for Sensor Networks and Image Processing)

Abstract

:
Multi-spectral imaging technologies have made great progress in the past few decades. The development of snapshot cameras equipped with a specific multi-spectral filter array (MSFA) allow dynamic scenes to be captured on a miniaturized platform across multiple spectral bands, opening up extensive applications in quantitative and visualized analysis. However, a snapshot camera based on MSFA captures a single band per pixel; thus, the other spectral band components of pixels are all missed. The raw images, which are captured by snapshot multi-spectral imaging systems, require a reconstruction procedure called demosaicing to estimate a fully defined multi-spectral image (MSI). With increasing spectral bands, the challenge of demosaicing becomes more difficult. Furthermore, the existing demosaicing methods will produce adverse artifacts and aliasing because of the adverse effects of spatial interpolation and the inadequacy of the number of layers in the network structure. In this paper, a novel multi-spectral demosaicing method based on a deep convolution neural network (CNN) is proposed for the reconstruction of full-resolution multi-spectral images from raw MSFA-based spectral mosaic images. The CNN is integrated with the channel attention mechanism to protect important channel features. We verify the merits of the proposed method using 5 × 5 raw mosaic images on synthetic as well as real-world data. The experimental results show that the proposed method outperforms the existing demosaicing methods in terms of spatial details and spectral fidelity.

1. Introduction

1.1. Background

Multi-spectral images, which contain rich spectral and spatial information, have been utilized in a wide range of applications in different fields such as food safety inspection [1], medical diagnosis [2], precision agriculture [3], and target tracking [4,5]. A series of multi-channel imaging systems utilizing image sensors have been invented to meet the demands for the capture of multi-spectral images (MSIs). These systems can be classified into the following three categories: (i) single-camera multi-shot systems [6], (ii) multi-camera one-shot systems [7], and (iii) single-camera one-shot systems [8,9,10,11,12,13]. Single-camera multi-shot systems rely on precisely controlled environments and complex equipment, while multi-camera single-shot systems require optical system calibration [14]. Furthermore, single-camera multi-shot systems capture spectral–spatial information at multiple spectral bands by sequentially switching a specific optical filter for each band [15]. Therefore, this technology is unsuitable for capturing dynamic scenes of moving targets because of switching among filters is time-consuming. To overcome these shortcoming, single-camera one-shot systems, which are equipped with special multi-spectral filter arrays (MSFAs), have emerged to acquire spectral–spatial information simultaneously from a single shot [16,17]. The principle is to reduce the acquisition time by adding a multi-spectral color filter array in front of the camera sensor. The structural design has the merit of compacting the sensor size and taking snapshots, which leads to low spatial resolution since each pixel of an MSFA only captures spectral information of a specific band. The recovering procedure, called demosaicing, as shown in Figure 1, is the key step to obtaining a full-resolution and high-quality multi-spectral image cube [18]. As the number of spectral bands and the sparsity of sampling increase, the estimation of missing pixels becomes more difficult, which makes the demosaicing procedure of MSFAs more challenging. For all this, a variety of MSFA patterns and corresponding demosaicing methods [19,20,21] have been proposed to improve demosaiced image quality. Figure 2 shows three different MSFA patterns with 6 bands, 5 bands, and 25 bands.

1.2. Related Works

Brauers and Aach [22] proposed a fast linear interpolation demosaicing method for a six-band MSFA, as shown in Figure 2a. Their demosaicing method initially employs bilinear interpolation for each band based on the color difference between channels, which was developed for the field of demosaicing for color filter arrays (CFAs). The method assumes that the adjacent bands exhibit similarity in the spatial structures, such as textures or edges. The similarity is referred to as inter-band correlation. Nevertheless, the assumption that there is a strong correlation across all channels cannot be satisfied.
Monno et al. [11] proposed a demosaicing method for the five-band MSFA pattern, as shown in Figure 2b. The filters of the MSFA are, respectively, R, Cy, G, Or, and B. The filter G is arranged most densely, allowing it to preserve spatial structures in the image more intricately than the others. Therefore, the proposed demosaicing method utilizes the G band to interpolate other bands based on inter-band correlations.
Wang et al. [23] extended a discrete wavelet transform (DWT)-based CFA demosaicing to the MSFA demosaicing. Their approach independently reconstructs the high-frequency and low-frequency components of the image. The reconstruction of the low-frequency component employs intra-band interpolation, whereas the high-frequency component’s reconstruction follows the “substitution rule” introduced by Driesen and Scheunders [24]. The substitution rule replaces the high-frequency components of each band in an MSI with the high-frequency components corresponding to the middle-spectral channel, assuming it to be the sharpest one. However, this method strongly relies on inter-channel correlations, and the substitution rule is applicable only to Bayer CFA but not fully extendable to MSFAs.
Miao et al. [25] proposed a novel demosaicing method that is binary tree and edge-sensing (BTES) based. In an MSFA, the sampling positions of each channel are mapped onto nodes in a binary tree. They interpolate the unknown pixel iteratively by edge-aware interpolation to improve the quality of the resulting images. However, due to the fact that the estimation of missing pixels in BTES only utilize intra-channel correlation, the reconstructed image failed to retain spatial details.
Mihoubi et al. [26] involved the use of a pseudo-panchromatic image (PPI), which is the average spectral value of all spectral bands, to aid in the demosaicing process. Their approach assumes a strong correlation between the PPI and all channels. The proposed method combines the pseudo-panchromatic image with the original multi-spectral image to improve the accuracy and quality of the demosaicing process, resulting in more accurate multi-spectral images. Nevertheless, the assumption regarding the correlation between PPI and all channels in this method seems to be relatively weak, and it requires complex computations.
In recent years, convolutional neural networks (CNNs) have been employed as a data segmentation method for many low-level image processing problems, such as image deblurring [27], super-resolution [28,29,30], mechanical fault diagnosis [31,32], and denoising [33,34,35]. In recent years, many multi-spectral reconstruction methods have also been proposed, such as MST++ [36], HSCNN+ [37], and AWAN [38]. MST++, which achieves higher-performance metrics with fewer computational and parameter resources, employs a multi-stage spectral-wise transformer to perform spectral reconstruction. HSCNN+ is a CNN-based method for RGB image hyperspectral restoration. This method achieves more accurate solutions by deepening the network structure. AWAN is a deep self-adaptive weighted attention network that can obtain more precise MSI and better reconstruction quality. Similarly, all kinds of deep learning-based methods have been proposed for MSFA deosaicing. Shinda [39] proposed a deep deosaicing network that combines the deep residual network ResNet with three-dimensional (3D) convolution, the final results of which are superior to the pseudo-panchromatic image difference (PPID) method. However, the images generated by this approach may exhibit false color artifacts in areas of high contrast and brightness. The method proposed by Wisotzky [40] involves using raw sparse multi-spectral images (MSIs) cube as the input of the deep convolutional network (DcNet), maintaining the complete spatial information of the original image. However, performing standard convolution on sparse inputs can lead to some artifacts, making the convergence of the network more difficult. Many deep-learning-based methods often validate their effectiveness on publicly available datasets, but they rarely validate their performance on real-world captured images. Furthermore, improving the quality of real-world captured images using models trained on public datasets remains a challenging problem. In this paper, we assume that public datasets can be used to simulate images captured by a real camera based on camera parameters. The aim is to enhance the quality of real captured images by training better models.

1.3. Our Contribution

Multi-spectral imaging systems, which are based on MSFA, often capture objects of interest in complex environments, resulting in poor image quality. Therefore, the proposed demosaicing method must have strong robustness. Moreover, the above demosaicing methods have certain limitations in the quality of image restoration.
In this paper, based on the spectral and spatial correlations existing in mosaic images captured by the snapshot MSFA imaging sensors, we propose a deep neural network that combines channel attention mechanisms with CNNs, which is able to extract features automatically and protect important channel features, for multi-spectral images based on 25-band MSFA, as shown in Figure 2c. The proposed demosaicing method makes the following main contributions:
  • We analyze the acquisition process of multi-spectral images, laying the theoretical groundwork for synthesizing the mosaic images and the radiance label images using the spectral sensitive functions (SSFs) of the IMEC camera we purchased and the available illuminants.
  • We present a simple and feasible end-to-end deep convolution neural network that introduces the channel attention mechanism, which is able to adaptively adjust channel feature response and protect important channel features.
  • The methodology we propose exhibits superior demosaicing performance on both simulated datasets and real-world scenarios compared to other existing methods, offering significant potential for the application of IMEC’s camera in both commercial and industrial sectors.

2. Observation Model

In this section, we perform the description of the observation model for the snapshot imaging systems based on MSFA. A single-sensor multi-spectral camera fitted with MSFA of K band provides a raw image I M S F A with M × N pixels. As depicted in Figure 3, at each pixel of a raw mosaic image I M S F A , only one out of the K bands is available while the levels of the remaining bands are missing. Furthermore, we consider that a fully defined MSI with K bands { G λ } 1 K can be modulated through a series of sampling matrices { X λ } 1 K . Finally, the raw mosaic image can be formulated as
I M S F A = Σ K λ = 1 X λ · G λ
where · represents the pixel-wise product.
Our expectation is to use the CNN network to learn a mapping model T that can estimate a high-precision MSI. The entire process can be formulated as
θ = a r g m i n l M S I ( T ( I M S F A ; θ ) , G )
where l M S I ( · ) refers to the loss function, and θ represent the parameters of the networks.

3. Reference Image Simulation

We refer to the multi-spectral image I = { I j } j = 1 K , which is composed of K fully defined channels corresponding to K bands, as the reference image. The reference image is often utilized as a reference for evaluating the demosaicing quality, although it cannot be provided by multi-spectral cameras fitted with a single-sensor. To obtain the reference image, we employ the multi-spectral image formation model described in Section 3.1 to simulate the acquisition process. Subsequently, we will utilize the acquired reference image to generate spectral mosaic images, which serve as inputs to our network.

3.1. Multi-Spectral Image Formation

Assuming ideal optics and uniform spectral sensitivity of the sensors, the image for the j-th channel can be expressed as
I j = Q ( Ω E ( λ ) · R ( λ ) · T j ( λ ) d λ )
where the term E ( λ ) refers to the relative spectral power distribution of the light source, which uniformly illuminates all surface elements within the scene. Ω stands for the working spectral range. R ( λ ) represents the reflectance function of all surface elements. The camera captures the radiance spectrum represented by E ( λ ) · R ( λ ) . This spectrum is then filtered based on the transmittance T j ( λ ) of the band j, which is centered at the wavelength λ j . The pixel values in the image of the last j-th channel are determined by quantifying the received energy using the function Q.

3.2. Simulation of Radiance Data

In order to simulate the radiance data, we need (i) illumination data and (ii) reflectance. (i) We specifically consider three standard light sources (F12, A, and D65) to generate radiance data. Their relative spectral power distributions E ( λ ) are defined for all [420 nm, 1000 nm]. (ii) The T T -59 database consists of high-quality hyperspectral images, which are saved in the form of spectral reflectance. The reflectance is defined on 59 bands, ranging from 420 nm to 1000 nm at 10 nm intervals and centered at {420 nm, 430 nm, … 1000 nm}. Assuming linear continuity of reflectance, we can use linear interpolation of the reflectance data from the T T -59 database to get R ( λ ) for all integer wavelengths λ within the range of [420 nm, 1000 nm].

3.3. Multi-Spectral Image Simulation

In order to simulate the reference channels provided by the IMEC camera we purchased according to Equation (1), we also need transmittance T I M E C i ( λ ) of the the IMEC camera. Transmittance T i ( λ ) differs based on the specific camera. The IMEC camera samples 25 bands with known transmittance T I M E C i ( λ ) . By summing discretely with d ( λ ) = 1 , Equation (1) transforms to
I j = Q ( Σ 1000 λ = 420 E ( λ ) · R ( λ ) · T I M E C i ( λ ) )

4. Proposed Demosaicing Method

4.1. Network Framework

Multi-spectral demosaicing can be seen as an interpolation method that utilizes known pixel information from the raw mosaic image to estimate the missing pixel information. The bilinear interpolation, which is fast and easy to achieve, can effectively recover low-frequency information in the image. Furthermore, the CNN has the ability to reconstruct the high-frequency information of the image by learning the mapping relationship between the original mosaic image and the multi-spectral image. Therefore, we utilize the strengths of both to achieve the reconstruction process.
First of all, we regard the multi-spectral demosaicing process as an end-to-end learning task. The input to our network is the two-dimensional raw mosaic image generated by Section 3.3. Based on the spatial and spectral correlation existing in multi-spectral filter array mosaic images, we propose an end-to-end deep convolution neural network that introduces the channel attention mechanism to reconstruct an fully defined and high-quality MSIs. The proposed method first utilizes K sampling matrices to separate the band of the initial mosaic input image, as shown in Figure 4, generating K sparse single-spectral images. Sparse single-spectral images exhibit values exclusively at particular spectral band pixels, with information absent at the remaining pixels. The proposed method utilizes a 9 × 9 convolution filter, as shown in Equation (6), to perform a rough and quick reconstruction. The 9 × 9 convolution is performed independently on sparse spectral images for each spectral band. The low-frequency information of the images can be effectively recovered, but the details and textures of the images suffer from more severe loss. We utilize parallel convolution and a deep residual network (ResNet) to reduce artifacts in the initial demosaiced image. The efficient channel attention (ECA) module [41] is introduced into both the parallel convolution and ResNet to adaptively adjust channel feature response and protect important channel features. The ECA module eliminates the fully connected layer typically present in traditional attention mechanism modules. Instead, it directly employs a 1D convolution on the features following global average pooling for learning. Because experimental evidence has shown that convolution has excellent cross-channel information-gathering capabilities. The ECA module achieves the fusion of global contextual information by squeezing the feature maps. This step is accomplished by performing global average pooling, transforming the feature maps from a size of (N, C, H, W) to (N, C, 1, 1). Then, ECA calculates the size of the adaptive convolutional kernel using Equation (5) and maps the weights between 0 and 1 using the sigmoid activation function. In Equation (5), C represents the number of input channels, we set γ and b to 2 and 1. Finally, ECA performs element-wise multiplication between the reshaped weight values and the original feature maps to obtain feature maps with different weights. The diagram of the ECA module is shown in Figure 5. The detailed structure of the proposed method network is illustrated in Figure 6.
k = l o g 2 ( C ) γ + b γ
F = 1 / 25 1 2 3 4 5 4 3 2 1 2 4 6 8 10 8 6 4 2 3 6 9 12 15 12 9 6 3 4 8 12 16 20 16 12 8 4 5 10 15 20 25 20 15 10 5 4 8 12 16 20 16 12 8 4 3 6 9 12 15 12 9 6 3 2 4 6 8 10 8 6 4 2 1 2 3 4 5 4 3 2 1
The proposed network consists of three steps: initial demosaicing, convolutional network, and residual network. The initial spectral mosaic image is the input of the network. The process of the initial demosaicing is used to transform the input from 480 × 480 × 1 to 480 × 480 × K . The value of K is set to 25, and it is determined by the number of spectral bands of the specific snapshot camera. The relationship between the initial demosaicing image and the reference image is utilized to train the whole network model, which is able to restore the high-frequency information of the multi-spectral image. The convolutional network uses three convolution kernels of different sizes to implement parallel convolution operations on the initial demosaiced image, respectively, which is able to obtain the shallow features of the multi-spectral image. The three paths of the convolutional network adopt convolution kernels of 1 × 1 , 3 × 3 , and 5 × 5 , respectively, and the number of kernels is 128. Each path has the same input, which is processed in the same way: convolution operation, rectified linear unit (ReLU), convolution operation, and ECA. The convolution operation can automatically extract important features from multi-spectral images without any manual supervision. The main function of ReLUs is to improve the expression ability of the CNN model by making the network sparse and reducing the possibility of gradient disappearing. The ECA module with channel attention mechanism is added at the end of each convolutional pathway, allowing for the adaptive adjustment of channel feature responses and preserving important information. The results of the three paths are summed as the input to the residual network. The residual network [42] can effectively extract deep features from multi-spectral images while preserving important information. The unique architecture can effectively solve the problem of gradient vanishing and explosion. Each residual block has the same structure with skip connections, which consists of two convolutional layers with the filter size of 3 × 3 , ReLU activation functions, and ECA. The number of extracted feature maps is 128. There are 12 residual blocks in the residual network. The output of the residual network performs one convolutional operation to obtain the final fully defined multi-spectral image. A 5 × 5 convolution kernel is used for convolution, and the number of images obtained through convolution should be consistent with the number of spectral bands. The convolution layer has a stride with padding to ensure that the size of the input and output arrays remains the same. In the proposed network structure, all convolution kernel coefficients are determined by learning. The ReLU function in the proposed network is f = m a x ( x , 0 ) , which has the ability to improve the speed of network training [43].

4.2. Loss Function

In order to enable the trained network model to more accurately reconstruct the fully defined MSI, we design a weighted combination loss function to minimize the signal reconstrucion errors of the demosaiced image. We define the real image as I and define the final reconstructed images as I . The overall loss function can be formulated as
L d e m o s a i c = L m s e + α · L w a v e l e t
Typically, the MSE (mean squared error) is commonly used as a loss function in the field of image processing because it encourages estimated pixel values to be closer to the ground truth. The MSE loss function can be formulated as
L m s e = 1 N · Σ N p I p k I p k 2 2
where k and p represent the channel index and the pixel index, and N is the number of pixels in the image.
Morever, we introduced an additional edge loss to enhance the shaperness and texture richness of the final estimated images. We transform the final estimated images and the reference images into wavelet domain, respectively, and calculate their MSE after transformation in the high-frequency sub-bands. The edge loss can be calculated as
L w a v e l e t = 1 N w · Σ N w q w q I w q I 2 2
where w q I and w q I represent the qth wavelet coefficients of the final estimated images and the reference images, respectively. N w is the number of high-frequency wavelet coefficients of the image decomposed by stationary wavelet transform. We use Haar filters in wavelet transformation and set the transformation level to 2.

5. Datasets and Training

We use the T T -59 database [44] to evaluate the effectiveness of the proposed demosaicing method. Multi-spectral images in the database are saved in the form of spectral reflectance, which is more convenient for us in synthesizing the radiance data. The T T -59 database has 16 scenes, which are divided into 11 scenarios for training the models and 5 scenarios for evaluating the performance of the model. We utilize the SSFs of the IMEC camera we purchased to synthesize the raw images with a 5 × 5 mosaic, illustrated in Figure 2c, where the spectral bands of the IMEC camera were centered at λ i = 699, 713, 726, 739, 752, 766, 780, 794, 807, 821, 834, 847, 859, 872, 884, 896, 908, 919, 930, 941, 952, 962, 972, 982, and 992 nm.
In the practical training of the proposed multi-spectral demosaicing approach, in order to artificially augment the number of training samples, we use two CIE standard light sources (D65 and F12) and rotate 90 , 180 , 270 during the training of the model. During the training of the network, we randomly extract 32 mosaic images with a size of 480 × 480 as a batch input. The optimizer of the network is Adam, and the learning rate is initialized to 0.0001 for all of the layers. We implemented the network model using the TensorFlow neural network framework. The deep learning model was trained for 500 epochs. In this paper, we used Python 3.6 to train the models, and all the experiments were conducted on a laboratory server with the configuration of an NVIDIA GeForce RTX 3090 12 GB GPU.

6. Experimental Results with Simulated Data and Real-Word Data

To validate the effectiveness of the model, we compare the proposed demosaicing method with three existing traditional demosaicing methods including weighted bilinear interpolation (WB), binary tree-based edge-sensing (BTES), pseudo-panchromatic image difference (PPID), and two deep-learning-based methods, including ResNet-3D [39] and deep convolutional network (DcNet) [40]. Furthermore, three metrics were utilized to represent the quality of the reconstructed multi-spectral image [45,46], which are the peak signal-to-noise ratio (PSNR), the structural similarity (SSIM), and the spectral angle mapper (SAM), respectively. The PSNR is to calculate the difference in pixel values between the reconstructed image and the reference image. The larger the value of PSNR between the two images, the better the quality of the reconstructed image. The SSIM is an image quality evaluation index that is utilized to measure the image degradation between the reconstructed image and the reference image. The value range of SSIM is [ 0 , 1 ] . The larger the SSIM value, the smaller the distortion degree of the reconstructed image. The SAM, which is widely used in pixel classification in multi-spectral imaging, is utilized to measures the spectral fidelity between the reconstructed image and the reference image by calculating the spectral angle. The smaller the value of SAM, the higher the spectral similarity between the reconstructed multi-spectral image and the reference image. We select two scenes Cloth3 and Spray in the database to visually compare the reconstructed images, which were generated by the above demosaicing methods. We crop and zoom in an area of 140 × 140 pixels from the reconstructed images for display. The visual comparison of reconstructed results by the above demosaicing methods is shown in Figure 7 and Figure 8.
As shown in Figure 7 and Figure 8, the edge of the image reconstructed by WB and BTES is highly blurred, and high-frequency information is seriously lost. The preservation of details in the PPID is significantly superior to WB and BTES, but the PPID still suffers from poor artifact prevention. The results from the ResNet-3D and DcNet are significantly superior to conventional methods, but still exhibit some artifacts. The image reconstructed of the proposed method provides better edge information with almost no artifacts and noise.
The PSNR and SSIM quantitative comparison results are shown in Table 1. The best results are highlighted in bold. Furthermore, The PSNR and SSIM values are calculated from the average of all test scenes in the database. Compared to other demosaicing methods in the spatial domain, the proposed demosaicing method has a significant improvement in both the PSNR and the SSIM. Table 1 indicates that the proposed demosaiking method outperforms the existing methods. The SAM comparison results are given in Table 2. In order to verify the effect of the proposed method, we randomly select four points from the reconstructed image to calculate SAM. The average SAM is reduced to 0.4168. Table 2 indicates that the proposed demosaiking method has an obvious superiority over other existing methods in spectral domain.
Due to the fact that the visual quality of an image cannot be strictly reflected by the PSNR/SSIM metrics, we compare the subjective quality of the demosaicing results on three test scenes from the TT-59 database using the error maps, which are the absolute errors between the demosaicing results and the ground truth. As shown in Figure 9, our method has superior demosaicing results in MSI images, which indicates that our method outperforms other demosaicing methods in terms of spatial accuracy. To further compare the superiority of the proposed method compared to other demosaicing methods in terms of spectral performance, we compared the absolute errors of all methods along the spectrum using the three scenes in Figure 9. As shown in Figure 10, the proposed method is much closer to the ground truth, which demonstrates higher spectral fidelity than other methods.
To evaluate the generalization ability of the proposed model, we select scene peppers in the CAVE database to visually compare the reconstructed images. We crop an area of 140 × 140 pixels from the reconstructed images for display. The reconstructed results by the above demosaicing methods are shown in Figure 11. As shown in Figure 11, The image reconstructed of the proposed method provides better edge information with almost no artifacts and noise. To further validate the effectiveness of our method, we conducted quantitative comparisons using three metrics: PSNR, SSIM, and SAM. The results are shown in Table 3 and Table 4. The quantitative comparison results indicate that the proposed method outperforms the other methods.
The proposed demosaicing method is applied in real-world images acquired by a real 5 × 5 snapshot camera. This camera has a spatial resolution of 409 × 217 (per band), and the frame rate can reach up to 170 data cubes per second (full sensor frame). As shown in Figure 12 and Figure 13, the proposed method is better than the traditional methods of WB and PPID. In terms of the details of the overall image, our method is also superior to ResNet-3D and DcNet.

7. Discussion

In this paper, we propose a snapshot multi-spectral demosaicing method for MSFA images based on the channel attention network. Extensive experiments indicate that the proposed demosaicing method has an obvious superiority over other existing methods in image quality and spectral domain. We also demonstrate that our method can guarantee real-time demosaicing performance in real-world shooting scenarios in ablation studies.
However, due to the limited number of training samples and the influence of the network architecture, the results of the proposed demosaicing method on real-world data still have some artifacts. How to more effectively utilize public datasets to simulate images captured based on camera parameters is also a factor. This is also a direction on which we will focus our future research efforts.

8. Ablation Studies

To demonstrate the effect of the proposed ECA module, we use the plain network without any attention as our base model, and then we study the networks with different attention in Table 5:
(1)
The base model without any attention, labeled as NT-NA;
(2)
The base model with ECA added to the convolutional network, labeled as NT-CN;
(3)
The base model with ECA added to the residual network, labeled as NT-RN;
(4)
The base model with ECA, which is our full model, is labeled as NT-FM.
All of these networks are trained under the same default settings for fair comparison. We test and average all the test images in the T T -59 database. Table 5 indicates that the proposed method are all better in terms of spatial accuracy and spectral fidelity than the other attention modules.
To further validate the effectiveness of the ECA module, we selected scenes from the dataset for visual quality comparisons. As shown in Figure 14, the proposed method is better than the other attention modules.
The running time and computational cost of the demosaicing method are of great importance when implemented on a realistic multi-spectral imaging system. Therefore, we compared the running time and GFLOPs of several deep-learning-based methods in Table 4. All experiments are implemented using the same machine (Intel Xeon Platinum 8260 CPU 2.40GHz, and NVIDIA GeForce RTX 3090 12 G GPU) and implemented on all test images of the T T -59 database. As shown in Table 6, the proposed method can guarantee real-time demosaicing performance.

9. Conclusions

In this paper, we present an end-to-end network based on the channel attention network to demosaic spectral mosaic images acquired by an MSFA-based imager. To achieve high-accuracy demosaicing of MSFA images, the proposed method takes advantage of the channel attention mechanism to protect important channel features, which is able to avoid artifacts and aliasing during the process of demosaicing. Our experimental results demonstrated that the proposed method generates more accurate demosaicing results in terms of spatial accuracy and spectral fidelity compared to other existing methods. It also serves as a universal multi-spectral demosaicing approach that can be adapted to different MSFA patterns with different spectral resolutions. Finally, We validated the effectiveness of the proposed method on real-world mosaic images with a pattern size of 5 × 5 , which offers an alternative pathway for the real-time application of multi-spectral imaging in complex environments. Because the results of the proposed demosaicing method on real-world data still have some artifacts, we need to improve the network structure (e.g., change from bilinear interpolation to PPID in the initial demosaicing) in the future. In summary, the proposed method could be potentially utilized to build high-quality MSFA-based image acquisition systems working well in object detection, medical diagnosis, and food safety inspection applications.

Author Contributions

Conceptualization, B.H. and G.Z.; methodology, X.Z. (Xuejun Zhang); software, Y.D.; validation, X.Z. (Xuejun Zhang) and Y.D.; formal analysis, X.Z. (Xuejun Zhang); investigation, X.Z. (Xuejun Zhang); resources, X.Z. (Xuejun Zhang); data curation, Y.D.; writing—original draft preparation, X.Z. (Xuejun Zhang); writing—review and editing, X.Z. (Xuejun Zhang); visualization, X.Z. (Xuemin Zhang); supervision, G.Z.; project administration, X.Z. (Xuejun Zhang); funding acquisition, G.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Natural Science Foundation of China (Grant No.: 42176182), the National Science Basic Research Foundation of Shaanxi Province (Grant No.: 2023-YBGY-390), and the CAS “Light of West China” Program.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

The public dataset used in this paper can be downloaded at: http://www.ok.sc.e.titech.ac.jp/res/MSI/MSIdata59.html (accessed on 1 October 2023).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Pu, H.; Lin, L.; Sun, D.W. Principles of hyperspectral microscope imaging techniques and their applications in food quality and safety detection: A review. Compr. Rev. Food Sci. Food Saf. 2019, 18, 853–866. [Google Scholar] [CrossRef]
  2. Hashimoto, E. Tissue classifification of liver pathological tissue specimens image using spectral features. Proc. SPIE 2017, 10140, 101400Z. [Google Scholar] [CrossRef]
  3. Hadoux, X.; Gorretta, N.; Roger, J.M.; Bendoula, R.; Rabatel, G. Comparison of the efficacy of spectral pre-treatments for wheat and weed discrimination in outdoor conditions. Comput. Electron. Agric. 2014, 108, 242–249. [Google Scholar] [CrossRef]
  4. Uzkent, B.; Rangnekar, A.; Hoffman, M.J. Tracking in aerial hyperspectral videos using deep kernelized correlation filters. IEEE Trans. Geosci. Remote Sens. 2019, 57, 449–461. [Google Scholar] [CrossRef]
  5. Xiong, F.; Zhou, J.; Qian, Y. Material based object tracking in hyperspectral videos. IEEE Trans. Image Process. 2020, 29, 3719–3733. [Google Scholar] [CrossRef] [PubMed]
  6. Chi, C.; Yoo, H.; Ben-Ezra, M. Multi-spectral imaging by optimized wide band illumination. Int. J. Comput. Vis. 2010, 86, 140–151. [Google Scholar] [CrossRef]
  7. Ohsawa, K.; Ajito, T.; Komiya, Y.; Fukuda, H.; Haneishi, H.; Yamaguchi, M.; Ohyama, N. Six-band HDTV camera system for spectrum-based color reproduction. J. Imaging Sci. Technol. 2004, 48, 85–92. [Google Scholar] [CrossRef]
  8. Lapray, P.J.; Wang, X.; Thomas, J.B.; Gouton, P. Multispectral filter arrays: Recent advances and practical implementation. Sensors 2014, 14, 21626–21659. [Google Scholar] [CrossRef]
  9. Geelen, B.; Tack, N.; Lambrechts, A. A compact snapshot multispectral imager with a monolithically integrated per-pixel filter mosaic. Proc. SPIE 2014, 8974, 80–87. [Google Scholar] [CrossRef]
  10. Martínez, M.A.; Valero, E.M.; Hernández-Andrés, J.; Romero, J.; Langfelder, G. Combining transverse field detectors and color filter arrays to improve multispectral imaging systems. Appl. Opt. 2014, 53, C14–C24. [Google Scholar] [CrossRef]
  11. Monno, Y.; Kikuchi, S.; Tanaka, M.; Okutomi, M. A practical one-shot multispectral imaging system using a single image sensor. IEEE Trans. Image Process. 2015, 24, 3048–3059. [Google Scholar] [CrossRef]
  12. Thomas, J.B.; Lapray, P.J.; Gouton, P.; Clerc, C. Spectral characterization of a prototype SFA camera for joint visible and NIR acquisition. Sensors 2016, 16, 993. [Google Scholar] [CrossRef] [PubMed]
  13. Murakami, Y.; Yamaguchi, M.; Ohyama, N. Spectral Hybrid-resolution multispectral imaging using color filter array. Opt. Express 2012, 20, 7173–7183. [Google Scholar] [CrossRef]
  14. Jaiswal, S.P.; Fang, L.; Jakhetiya, V.; Pang, J.; Mueller, K.; Au, O.C. Adaptive multispectral demosaicing based on frequency domain analysis of spectral correlation. IEEE Trans. Image Process. 2017, 26, 953–968. [Google Scholar] [CrossRef]
  15. Fukuda, H.; Uchiyama, T.; Haneishi, H.; Yamaguchi, M.; Ohyama, N. Development of a 16-band multispectral image archiving system. In Proceedings of the SPIE Electronic Imaging: Color Imaging X, San Jose, CA, USA, 17–20 January 2005; Volume 5667, pp. 136–145. [Google Scholar] [CrossRef]
  16. Monakhova, K.; Yanny, K.; Aggarwal, N.; Waller, L. Spectral DiffuserCam: Lensless snapshot hyperspectral imaging with a spectral filter array. Optica 2020, 7, 1298–1307. [Google Scholar] [CrossRef]
  17. Gunturk, B.K.; Glotzbach, J.; Altunbasak, Y.; Schafer, R.W.; Mersereau, R.M. Demosaicing: Color filter array interpolation. IEEE Signal Process. Mag. 2005, 22, 44–54. [Google Scholar] [CrossRef]
  18. Monno, Y.; Tanaka, M.; Okutomi, M. Multispectral demosaicing using guided filter. Digit. Photogr. VIII 2012, 8299, 204–210. [Google Scholar] [CrossRef]
  19. Zhao, B.; Zheng, J.; Dong, Y.; Shen, N.; Yang, J.; Cao, Y.; Cao, Y. PPI Edge Infused Spatial–Spectral Adaptive Residual Network for Multispectral Filter Array Image Demosaicing. IEEE Trans. Geosci. Remote Sens. 2023, 61, 1–14. [Google Scholar] [CrossRef]
  20. Rathi, V.; Goyal, P. Generic multispectral demosaicking using spectral correlation between spectral bands and pseudo-panchromatic image. Signal Process. Image Commun. 2023, 110, 116893. [Google Scholar] [CrossRef]
  21. Gupta, M.; Rathi, V.; Goyal, P. Adaptive and Progressive Multispectral Image Demosaicking. IEEE Trans. Comput. Imaging 2022, 8, 69–80. [Google Scholar] [CrossRef]
  22. Brauers, J.; Aach, T. A Color Filter Array based Multispectral Camera. In Proceedings of the Workshop Farbbildverarbeitung, Ilmenau, Germany, 5–6 October 2006. [Google Scholar]
  23. Wang, X.; Thomas, J.B.; Hardeberg, J.Y.; Gouton, P. Discrete wavelet transform based multispectral filter array demosaicing. In Proceedings of the 2013 Colour and Visual Computing Symposium (CVCS), Gjovik, Norway, 5–6 September 2013; pp. 1–6. [Google Scholar] [CrossRef]
  24. Driesen, J.; Scheunders, P. Wavelet-based color filter array demosaicing. In Proceedings of the International Conference on Image Processing (ICIP), Singapore, 24–27 October 2004; Volume 5, pp. 3311–3314. [Google Scholar] [CrossRef]
  25. Miao, L.; Qi, H.; Ramanath, R.; Snyder, W.E. Binary tree-based generic demosaicing algorithm for multispectral filter arrays. IEEE Trans. Image Process. 2006, 15, 3550–3558. [Google Scholar] [CrossRef]
  26. Mihoubi, S.; Losson, O.; Mathon, B.; Macaire, L. Multispectral demosaicing using pseudo-panchromatic image. IEEE Trans. Comput. Imaging 2017, 3, 982–995. [Google Scholar] [CrossRef]
  27. Xu, L.; Ren, J.S.; Liu, C.; Jia, J. Deep convolutional neural network for image deconvolution. In Proceedings of the International Conference on Neural Information Processing Systems, Montreal, QC, Canada, 8–13 December 2014; pp. 1790–1798. [Google Scholar]
  28. Dong, C.; Loy, C.C.; He, K.; Tang, X. Learning a deep convolutional network for image super-resolution. In Proceedings of the European Conference on Computer Vision, Zurich, Switzerland, 6–12 September 2014; pp. 184–199. [Google Scholar] [CrossRef]
  29. Kim, J.; Lee, J.K.; Lee, K.M. Deeply-recursive convolutional network for image super-resolution. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition, Las Vegas, NV, USA, 27–30 June 2016; pp. 1637–1645. [Google Scholar] [CrossRef]
  30. Wang, X.; Chen, J.; Wei, Q.; Richard, C. Hyperspectral image super resolution via deep prior regularization with parameter estimation. IEEE Trans. Circuits Syst. Video Technol. 2022, 32, 1708–1723. [Google Scholar] [CrossRef]
  31. Xu, Y.; Feng, K.; Yan, X.; Sheng, X.; Sun, B.; Liu, Z.; Yan, R. Cross-modal Fusion Convolutional Neural Networks with Online Soft Label Training Strategy for Mechanical Fault Diagnosis. IEEE Trans. Ind. Inform. 2023, 20, 73–84. [Google Scholar] [CrossRef]
  32. Xu, Y.; Feng, K.; Yan, X.; Yan, R.; Ni, Q.; Sun, B.; Lei, Z.; Zhang, Y.; Liu, Z. CFCNN: A novel convolutional fusion framework for collaborative fault identification of rotating machinery. Inf. Fusion 2023, 95, 1–16. [Google Scholar] [CrossRef]
  33. Xie, J.; Xu, L.; Chen, E. Image denoising and inpainting with deep neural networks. In Proceedings of the International Conference on Neural Information Processing Systems, Lake Tahoe, NV, USA, 3–6 December 2012; pp. 341–349. [Google Scholar]
  34. Zeng, X.; Bian, W.; Liu, W.; Shen, J.; Tao, D. Dictionary pair learning on Grassmann manifolds for image denoising. IEEE Trans. Image Process. 2015, 24, 4556–4569. [Google Scholar] [CrossRef] [PubMed]
  35. Pan, Z.; Li, B.; Cheng, H.; Bao, Y. Deep residual network for MSFA raw image denoising. In Proceedings of the IEEE International Conference on Acoustics, Speech and Signal Processing, Barcelona, Spain, 4–8 May 2020; pp. 2413–2417. [Google Scholar] [CrossRef]
  36. Cai, Y.; Lin, J.; Lin, Z.; Wang, H.; Zhang, Y.; Pfister, H.; Timofte, R.; Van Gool, L. Mst++: Multi-stage spectral-wise transformer for efficient spectral reconstruction. In Proceedings of the IEEE/CVF Conference on Computer Vision and Pattern Recognition, New Orleans, LA, USA, 19–20 June 2022; pp. 745–755. [Google Scholar] [CrossRef]
  37. Shi, Z.; Chen, C.; Xiong, Z.; Liu, D.; Wu, F. Hscnn+: Advanced cnn-based hyperspectral recovery from rgb images. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition Workshops, Salt Lake City, UT, USA, 18–22 June 2018; pp. 939–947. [Google Scholar] [CrossRef]
  38. Li, J.; Wu, C.; Song, R.; Li, Y.; Liu, F. Adaptive weighted attention network with camera spectral sensitivity prior for spectral reconstruction from RGB images. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition Workshops, Seattle, WA, USA, 14–19 June 2020; pp. 462–463. [Google Scholar] [CrossRef]
  39. Shinoda, K.; Yoshiba, S.; Hasegawa, M. Deep demosaicing for multispectral filter arrays. arXiv 2018, arXiv:1808.08021. [Google Scholar]
  40. Wisotzky, E.L.; Daudkane, C.; Hilsmann, A.; Eisert, P. Hyperspectral Demosaicing of Snapshot Camera Images Using Deep Learning. In Proceedings of the DAGM German Conference on Pattern Recognition, Konstanz, Germany, 27–30 September 2022; Springer International Publishing: Cham, Switzerland, 2022; pp. 198–212. [Google Scholar] [CrossRef]
  41. Wang, Q.; Wu, B.; Zhu, P.; Li, P.; Zuo, W.; Hu, Q. ECA-Net: Efficient Channel Attention for Deep Convolutional Neural Networks. In Proceedings of the 2020 IEEE/CVF Conference on Computer Vision and Pattern Recognition (CVPR), Seattle, WA, USA, 13–19 June 2020; pp. 11531–11539. [Google Scholar] [CrossRef]
  42. He, K.; Zhang, X.; Ren, S.; Sun, J. Deep residual learning for image recognition. In Proceedings of the 2016 IEEE Conference on Computer Vision and Pattern Recognition, Las Vegas, NV, USA, 27–30 June 2016; pp. 770–778. [Google Scholar] [CrossRef]
  43. Krizhevsky, A.; Sutskever, I.; Hinton, G.E. ImageNet Classifification with Deep Convolutional Neural Networks. In Proceedings of the Advances in Neural Information Processing Systems 25 (NIPS 2012), Lake Tahoe, NV, USA, 3–6 December 2012; pp. 1097–1105. [Google Scholar] [CrossRef]
  44. Monno, Y.; Teranaka, H.; Yoshizaki, K.; Tanaka, M.; Okutomi, M. Single-sensor RGB-NIR imaging: High-quality system design and prototype implementation. IEEE Sens. J. 2019, 19, 497–507. [Google Scholar] [CrossRef]
  45. Setiadi, D.R.I.M. PSNR vs. SSIM: Imperceptibility quality assessment for image steganography. Multimed. Tools Appl. 2021, 80, 8423–8444. [Google Scholar] [CrossRef]
  46. 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]
Figure 1. The procedure of multi-spectral demosaicing.
Figure 1. The procedure of multi-spectral demosaicing.
Sensors 24 00943 g001
Figure 2. Several different MSFA patterns: (a) Brauers and Aach’s MSFA; (b) Monno et al.’s MSFA; (c) our MSFA.
Figure 2. Several different MSFA patterns: (a) Brauers and Aach’s MSFA; (b) Monno et al.’s MSFA; (c) our MSFA.
Sensors 24 00943 g002
Figure 3. The model of a snapshot mosaic imaging system fitted with an MSFA.
Figure 3. The model of a snapshot mosaic imaging system fitted with an MSFA.
Sensors 24 00943 g003
Figure 4. Band separation.
Figure 4. Band separation.
Sensors 24 00943 g004
Figure 5. A diagram of the efficient channel attention (ECA) module.
Figure 5. A diagram of the efficient channel attention (ECA) module.
Sensors 24 00943 g005
Figure 6. A schematic diagram of the proposed demosaicing method.
Figure 6. A schematic diagram of the proposed demosaicing method.
Sensors 24 00943 g006
Figure 7. The visual quality comparison results of different demosaicing methods for cloth3 scenes. (a,b) is the reference image and the ground truth. (c) is the result of the WB. (d) is the result of the BTES. (e) is the result of the PPID. (f) is the result of the ResNet-3D. (g) is the result of the DcNet. (h) is the result of ours.
Figure 7. The visual quality comparison results of different demosaicing methods for cloth3 scenes. (a,b) is the reference image and the ground truth. (c) is the result of the WB. (d) is the result of the BTES. (e) is the result of the PPID. (f) is the result of the ResNet-3D. (g) is the result of the DcNet. (h) is the result of ours.
Sensors 24 00943 g007
Figure 8. The visual quality comparison results of different demosaicing methods. (Images are converted to sRGB for display). (a,b) is the reference image and the Ground Truth. (c) is the result of the WB. (d) is the result of the BTES. (e) is the result of the PPID. (f) is the result of the ResNet-3D. (g) is the result of the DcNet. (h) is the result of ours.
Figure 8. The visual quality comparison results of different demosaicing methods. (Images are converted to sRGB for display). (a,b) is the reference image and the Ground Truth. (c) is the result of the WB. (d) is the result of the BTES. (e) is the result of the PPID. (f) is the result of the ResNet-3D. (g) is the result of the DcNet. (h) is the result of ours.
Sensors 24 00943 g008
Figure 9. Visual quality comparison of representative scenes of the TT-59 database at 699 nm. The error maps for WB/BTES/PPID/ResNet-3D/DcNet/our demosaicing results.
Figure 9. Visual quality comparison of representative scenes of the TT-59 database at 699 nm. The error maps for WB/BTES/PPID/ResNet-3D/DcNet/our demosaicing results.
Sensors 24 00943 g009
Figure 10. The absolute error between the demosaicing results and the ground truth of the scenes in Figure 9 along the spectrum for all methods.
Figure 10. The absolute error between the demosaicing results and the ground truth of the scenes in Figure 9 along the spectrum for all methods.
Sensors 24 00943 g010
Figure 11. The visual quality comparison results of different demosaicing methods for peppers scenes. (a,b) is the reference image and the Ground Truth. (c) is the result of the WB. (d) is the result of the BTES. (e) is the result of the PPID. (f) is the result of the ResNet-3D. (g) is the result of the DcNet. (h) is the result of ours.
Figure 11. The visual quality comparison results of different demosaicing methods for peppers scenes. (a,b) is the reference image and the Ground Truth. (c) is the result of the WB. (d) is the result of the BTES. (e) is the result of the PPID. (f) is the result of the ResNet-3D. (g) is the result of the DcNet. (h) is the result of ours.
Sensors 24 00943 g011
Figure 12. The visual results of a building using different demosaicing methods at 699 nm. (a) is the mosaic image. (b) is the result of the WB. (c) is the result of the PPID. (d) is the result of the ResNet-3D. (e) is the result of the DcNet. (f) is the result of ours.
Figure 12. The visual results of a building using different demosaicing methods at 699 nm. (a) is the mosaic image. (b) is the result of the WB. (c) is the result of the PPID. (d) is the result of the ResNet-3D. (e) is the result of the DcNet. (f) is the result of ours.
Sensors 24 00943 g012
Figure 13. The visual results of a signboard using different demosaicing methods at 699 nm. (a) is the mosaic image. (b) is the result of the WB. (c) is the result of the PPID. (d) is the result of the ResNet-3D. (e) is the result of the DcNet. (f) is the result of ours.
Figure 13. The visual results of a signboard using different demosaicing methods at 699 nm. (a) is the mosaic image. (b) is the result of the WB. (c) is the result of the PPID. (d) is the result of the ResNet-3D. (e) is the result of the DcNet. (f) is the result of ours.
Sensors 24 00943 g013
Figure 14. The visual results of different cases.
Figure 14. The visual results of different cases.
Sensors 24 00943 g014
Table 1. Demosaicing results (PSNR/SSIM) for three typical scenarios and the average results for all test scenarios on the T T -59 dataset. (Bold indicates the best result).
Table 1. Demosaicing results (PSNR/SSIM) for three typical scenarios and the average results for all test scenarios on the T T -59 dataset. (Bold indicates the best result).
MethodsSprayCloth3Doll2Average of All Test
PSNR ↑ SSIM ↑ PSNR ↑ SSIM ↑ PSNR ↑ SSIM ↑ PSNR ↑ SSIM ↑
WB26.600.900625.420.899826.880.906826.420.9016
BTES26.720.907625.780.901427.560.910627.640.9048
PPID35.460.981033.780.962436.780.976435.780.9784
ResNet-3D41.870.999239.410.997540.260.997041.530.9978
DcNet42.760.999341.120.998041.770.997241.980.9982
OURS45.870.999942.500.998843.550.998444.880.9990
Table 2. Demosaicing results (SAM) at different pixel points for doll2 scenes. (Bold indicates the best result).
Table 2. Demosaicing results (SAM) at different pixel points for doll2 scenes. (Bold indicates the best result).
ImageMethodsPoint 1Point 2Point 3Point 4Average
Sensors 24 00943 i001WB1.29731.17681.16181.61281.2482
BTES1.12361.17521.15271.50231.2378
PPID0.89520.86810.84860.87490.8254
ResNet-3D0.80110.83400.79810.82920.8022
DcNet0.64200.66900.74470.67910.6814
OURS0.42660.43310.47200.41540.4168
Table 3. Demosaicing results (PSNR/SSIM) for three typical scenarios and the average results for all test scenarios on the cave dataset. (Bold indicates the best result).
Table 3. Demosaicing results (PSNR/SSIM) for three typical scenarios and the average results for all test scenarios on the cave dataset. (Bold indicates the best result).
MethodsSpongesPaintsFeathersAverage of All Test
PSNR ↑ SSIM ↑ PSNR ↑ SSIM ↑ PSNR ↑ SSIM ↑ PSNR ↑ SSIM ↑
WB25.580.911226.180.901826.460.913225.840.9106
BTES25.880.917826.720.912426.880.918626.950.9162
PPID36.620.978834.220.972837.460.965236.120.9720
ResNet-3D40.260.998639.580.996439.860.997240.840.9980
DcNet41.520.999040.460.998641.040.998241.080.9986
OURS44.120.999741.860.998643.200.998843.240.9989
Table 4. Demosaicing results (SAM) at different pixel points for peppers scenes. (Bold indicates the best result).
Table 4. Demosaicing results (SAM) at different pixel points for peppers scenes. (Bold indicates the best result).
ImageMethodsPoint 1Point 2Point 3Point 4Average
Sensors 24 00943 i002WB1.38201.25261.27641.52081.3206
BTES1.29601.25421.23281.48561.2726
PPID0.90420.85280.86640.88120.8652
ResNet-3D0.83240.82260.81680.84560.8288
DcNet0.78120.74160.78520.70180.7182
OURS0.51260.48640.49120.43280.4418
Table 5. The ablation study of the network with different attention. (Bold indicates the best result).
Table 5. The ablation study of the network with different attention. (Bold indicates the best result).
CasePSNRSAM
NT-NA41.560.5436
NT-CN42.880.5324
NT-RN42.860.5312
NT-FM44.880.5132
Table 6. The demosaicing performance comparison of different methods.
Table 6. The demosaicing performance comparison of different methods.
MethodsRunning Times (ms)GFLOPs
ResNet-3D2.87932.4
DcNet2.1650.6
Ours2.4668.2
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, X.; Dai, Y.; Zhang, G.; Zhang, X.; Hu, B. A Snapshot Multi-Spectral Demosaicing Method for Multi-Spectral Filter Array Images Based on Channel Attention Network. Sensors 2024, 24, 943. https://doi.org/10.3390/s24030943

AMA Style

Zhang X, Dai Y, Zhang G, Zhang X, Hu B. A Snapshot Multi-Spectral Demosaicing Method for Multi-Spectral Filter Array Images Based on Channel Attention Network. Sensors. 2024; 24(3):943. https://doi.org/10.3390/s24030943

Chicago/Turabian Style

Zhang, Xuejun, Yidan Dai, Geng Zhang, Xuemin Zhang, and Bingliang Hu. 2024. "A Snapshot Multi-Spectral Demosaicing Method for Multi-Spectral Filter Array Images Based on Channel Attention Network" Sensors 24, no. 3: 943. https://doi.org/10.3390/s24030943

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