Next Article in Journal
Improved YOLOv7 Algorithm for Small Object Detection in Unmanned Aerial Vehicle Image Scenarios
Next Article in Special Issue
Enhancing Acute Bilirubin Encephalopathy Diagnosis with Multi-Modal MRI: A Deep Learning Approach
Previous Article in Journal
Design of Indoor Visible Light Communication PAM4 System
Previous Article in Special Issue
Advanced Medical Image Segmentation Enhancement: A Particle-Swarm-Optimization-Based Histogram Equalization Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Virtual Staining Method Based on Self-Supervised GAN for Fourier Ptychographic Microscopy Colorful Imaging

Electronics Information Engineering College, Changchun University, Changchun 130022, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Appl. Sci. 2024, 14(4), 1662; https://doi.org/10.3390/app14041662
Submission received: 14 January 2024 / Revised: 6 February 2024 / Accepted: 16 February 2024 / Published: 19 February 2024
(This article belongs to the Special Issue Artificial Intelligence in Biomedical Image Processing)

Abstract

:
Fourier ptychographic microscopy (FPM) is a computational imaging technology that has endless vitality and application potential in digital pathology. Colored pathological image analysis is the foundation of clinical diagnosis, basic research, and most biomedical problems. However, the current colorful FPM reconstruction methods are time-inefficient, resulting in poor image quality due to optical interference and reconstruction errors. This paper combines coloring and FPM to propose a self-supervised generative adversarial network (GAN) for FPM color reconstruction. We design a generator based on the efficient channel residual (ECR) block to adaptively obtain efficient cross-channel interaction information in a lightweight manner, and we introduce content-consistency loss to learn the high-frequency information of the image and improve the image quality of the staining. Furthermore, the effectiveness of our proposed method is demonstrated through objective indicators and visual evaluations.

1. Introduction

The use of light microscopy for minimally invasive high-resolution imaging is fundamental to human understanding of biological systems and processes. In clinical pathology, analyzing high-resolution color pathological images from optical microscopes remains a widely used standard method for diagnosing diseases ranging from cancer to blood-borne infections. To perform quantitative and accelerated pathological image analysis, the whole slide imaging (WSI) [1] system is gradually replacing traditional optical microscopes. Fourier ptychographic microscopy (FPM) [2,3], which has been popularized in recent years, is considered a high-throughput imaging technology that does not require mechanical scanning and has more advantages over conventional whole slide imaging systems in digital pathology. First of all, FPM does not require mechanical scanning to achieve both high resolution (HR) and a large field of view, and it can simply add LED arrays to other microscopes to set up experimental platforms. Secondly, the digital refocusing process is utilized by FPM to correct sample defocusing after acquisition [2,4]. Furthermore, the phase information recovered from the FPM provides more localized scattering details of the sample, which enhances the accuracy and reliability of pathological diagnosis [5,6].
However, the application of FPM in pathology still faces some challenges [6]. On the one hand, the time efficiency of FPM acquisition and reconstruction is relatively poor. FPM usually requires continuous acquisition with red, green, and blue separate illumination, composing hundreds of low-resolution (LR) images for better color reconstruction, leading to slower acquisition and reconstruction speeds. On the other hand, FPM color reconstruction images are usually of poor quality due to coherent artifacts caused by light interference and reconstruction errors. Furthermore, the GlaS dataset in Figure 1 is pathological sections of the differentiation tissue of colon cancer glands stained with hematoxylin and eosin (H&E), and the MoNuSeg dataset is composed of various H&E stained histological images across multiple patients and organs obtained from different hospitals. It can be observed in Figure 1 that the color reconstruction results of the same batch of pathological sections will have significant visual differences due to the influence of factors such as collection personnel, technology, and the external environment. In biological and clinical medicine applications, the color information of the sample after staining with reagents helps the observer quickly locate the regions of interest and interpret the relevant information about tissues and cells. Therefore, we use a virtual staining method for the colorful reconstruction of FPM single-channel grayscale images.
Recently, significant advances have been made in the field of deep learning, paving the way for the application of neural networks in a variety of image-processing tasks, including virtual tissue staining. To better obtain the color information of FPM-stained tissue sections, image translation can be applied to convert the recovered monochromatic FPM image to the style of regular incoherent microscopy. The mathematical mapping between input and output [7] can be found accurately by this deep learning technology based on the deep convolutional neural network (CNN), and the coherence artifacts of FPM recovery can also be reduced [8], thus improving the quality of colored FPM pathology images.
Since its inception, the unsupervised learning image-to-image translation model CycleGAN [9] has shown its powerful transformation ability between two image domains, which makes it a popular method of virtual staining in the field of biomedical imaging. However, the cyclic consistency [9] assumption dictates that the relationship between the two domains must be bijective [10], which is often not ideal, as it may limit the diversity of images. To overcome this limitation, contrastive learning has recently achieved optimal progress [11,12] in the field of self-supervised representation learning. Subsequently, contrastive unpaired translation (CUT) [13] introduced contrastive learning to maximize the learning of mutual information between the two image domains for unpaired image translation.
In this paper, we propose a self-supervised generative adversarial network (GAN) based on contrastive learning for the virtual staining of single-channel images of FPM. While CUT has demonstrated the efficiency of contrastive learning, it may not efficiently capture domain gaps, color, and the high-frequency information of the images in our application. To further take advantage of contrastive learning while avoiding the disadvantages of cyclic consistency, our method trains a GAN with efficient channel residual (ECR) block and content-consistency loss to learn the cross-channel and high-frequency information of images more effectively, as well as to reduce coherent artifacts, and improve the image quality. The network training and testing processes are illustrated in Figure 2. This method provides a better foundation for subsequent research on medical cell recognition, segmentation, classification, and other related fields. The main contributions of this paper are listed below:
  • We adopt the unpaired learning architecture of CUT to treat the virtual staining of the FPM pathology images as a transformation from a single-channel image to a color image. Since CUT cannot accurately identify the light-colored part of the image and will produce unnecessary artifacts, we design the ECR block in the generator to extract cross-channel interaction information and improve the network performance.
  • CUT cannot effectively capture the edges and details of the image. We introduce the content-consistency loss based on multi-scale structural similarity (MS-SSIM) to avoid feature distortion between input and output images and enhance the high-frequency information of images.
We conducted comparative experiments to demonstrate the qualitative and quantitative advantages of our method in single-channel FPM virtual staining tasks compared to other methods. In addition, our method theoretically reduces the collection time of 2/3 of color FPM, standardizes the staining process, and improves the objectivity of disease diagnosis. Moreover, we do not need to obtain paired data, as simply using unpaired input domain X and output domain Y images can achieve good staining results, which effectively solves the problem of difficult matching data acquisition, especially in digital pathology.

2. Related Work

2.1. Color-Based FPM Imaging

The combination of virtual staining and FPM can improve imaging efficiency while providing further convenience for the detection and diagnosis of pathological sections. A classical coloring method [14] is used to restore the HR image at three wavelengths with a monochromatic camera and then synthesize the HR full-color image. The Wavelength-Multiplexed FPM (WMFPM) [15] scheme was proposed by Dong et al. with a monochrome camera and multi-wavelength simultaneous illuminations. A color transfer-based FPM virtual staining method called CFPM [16] was reported by Gaod et al. The CFPM scheme sacrifices only about 0.4% imaging accuracy, but the efficiency is improved by three times.
Different deep learning methods represented by CNN have also been widely used in FPM. Zhang et al. [17] proposed color FPM based on a 3D CNN, using the trained CNN to establish the mapping relationship between monochrome HR images and color LR images to generate color HR images. Wang et al. [18] proposed the application of cyclic consistency to generate the adversarial network CycleGAN in FPM imaging, enabling virtual bright field and fluorescence staining of single-channel images. Zhang et al. [19] designed a CNN model for reconstructing FPM HR color pathology images and phase maps.

2.2. Image-to-Image Translation

The application of deep learning in virtual staining is mainly reflected in image-to-image translation. The introduction of this model effectively promotes the development of virtual staining. GANs [20] have been widely used in a series of image applications, achieving significant results in the field of image-to-image translation in particular. Generally speaking, image-to-image translation can be categorized into two methods, i.e., supervised learning and unsupervised learning methods. The widespread application of supervised methods is limited because it requires a large amount of paired data for training.
In recent years, unsupervised methods have become a hot topic of focus for researchers. UNIT [21] proposed a shared latent space assumption that generates high-quality translations by utilizing shared information between two domains. CycleGAN proposed the cyclic consistency assumption, which learns the mapping relationship between two domains through two generator networks and two adversarial discriminator networks to improve the quality and consistency of image conversion. This assumption requires the transformation to be reversible and consistent, resulting in strong limitations. Recently, to alleviate the problems caused by cyclic consistency, some methods [22,23] have adopted different constraints from different aspects in an attempt to break this cycle.

2.3. Contrastive Learning

Contrastive learning facilitates deep learning by training a model to maximize the similarity between similar samples and minimize the similarity between dissimilar samples in the field of self-supervised learning. This method is very suitable for image translation models, especially for image data whose output shape is the same as its input. The introduction of contrastive learning in image translation can preserve the content of input [23] and reduce mode collapse [24,25,26]. CUT introduced noise contrast estimation into the image translation tasks and achieved better performance than that based on cyclic consistency through learning the correspondence between input image patches and generated image patches for the first time.

2.4. Attention Mechanism

Inspired by the fact that humans can naturally and efficiently find significant areas in complex scenes, attention mechanisms have been introduced in computer vision to simulate this process to focus on important features and rationally assign weights. The attention mechanism can overcome the acceptance field limitation of CNN by capturing long-distance dependencies in images. There have also been several unprecedented efforts to integrate GAN with attention mechanisms in recent years. Zhang [27] obtained remote relationships in the images by self-attention. Spatial attention GAN (SPA-GAN) utilized attention maps from the discriminator to guide the generator’s focus on discriminative areas in the image [28]. Hu [29] designed a query-selected attention (QS-Attn) module to intentionally select important anchors for comparative learning. Torbunov [30] equipped the generator of CycleGAN with a Vision Transformer (ViT) to improve non-local pattern learning and network performance.

3. Proposed Method

3.1. Network Architecture

Figure 3 shows the overall architecture of our proposed method, which has been improved based on the CUT model. The core concept is to use only one generator (G) and discriminator (D) pair to map features in one direction, which greatly improves memory efficiency and training speed. We utilize G to convert the single-channel FPM grayscale images (input domain X) into high-resolution FPM color images (output domain Y). The true color image of the target domain Y and the generated color image of the input domain X are distinguished by the trained D.
The generator and discriminator structures of this paper are shown in Figure 4. Generator G includes three downsampling modules, nine ECR blocks, two upsampling modules, and one convolutional layer. The downsampling module extracts low-frequency features of the image through convolution operation and enters it as input into the ECR. The ECR block is formed from a residual block and a channel attention block through the residual connection, which adaptively combines similar features of the image to provide higher-level semantic information. The upsampling module reconstructs the details and resolution of the image through transposed convolution operations. We utilize PatchGAN [31] as the discriminator D, and the receptive field of PatchGAN is set to 70 × 70, which makes PatchGAN faster and still guides the generator to produce realistic results.

3.2. Efficient Channel Residual Block

In this paper, we expect the generator to be able to generate FPM color images similar to the domain Y. The ResNet-based [32] generator is widely used in the field of image-to-image translation. However, CUT failed to accurately identify the unstained part of the pathological section when staining the single-channel FPM input image, resulting in unnecessary artifacts and producing low-quality color images. The weight of each channel in a deep neural network can be recalibrated by using channel attention. This allows for a more adaptive selection process for determining the object of focus in different feature maps, as each channel can represent a different object. However, most of the methods proposed in recent years have focused on designing more complex attention modules. It undoubtedly increases the complexity of the model while delivering better performance. For this purpose, we designed an ECR block to replace the residual block in the CUT.
As shown in Figure 5, we introduce ECANet [33] into the residual block of the generator. Concretely speaking, the ECR block consists of two parts by the shortcut connection. One part is the original residual block, which includes two convolutional layers and one RELU layer, and the other is the efficient channel attention (ECA) block, which includes a global average pooling layer, a one-dimensional convolutional layer with an adaptive convolutional kernel, and a sigmoid layer. These two parts each have a shortcut connection between the input and output.
ResNet can truly deepen the depth of convolutional neural networks without reducing network accuracy, and its powerful representation ability can significantly improve network performance in many computer vision fields. The ECA block improves network performance by avoiding reducing channel dimensions to learn effective channel attention while obtaining cross-channel interaction information in an extremely lightweight manner. We add identity maps to the ECR block, which not only ensures that there is no gradient vanishing problem during backpropagation but also shifts the network to learning residual functions, ensuring a smoother decision function for the entire network and improving generalization performance.

3.3. Loss Functions

3.3.1. PatchNCE Loss

As with the setting of CUT, we enable the network to maximize the mutual information between the input and output corresponding patches by using a noise contrastive estimation framework [34]. The core idea of contrast learning is to relate the two signals of a “query” and its “positive” example, not associated with other examples in the dataset (called “negatives”). Firstly, the probability of choosing “positive” instead of other “negatives” is represented by calculating the cross-entropy loss [35] as follows:
( v , v + , v ) = log [ exp ( v v + / τ ) exp ( v v + / τ ) + n = 1 N exp ( v v n / τ ) ]
where v , v + R K , and v R N × K indicate mapping query, positive, and N negatives to K-dimensional vectors, respectively. v n R K denotes the n-th negative. τ is the parameter used to scale the distance between the query and the other examples, with the default value set to 0.07 in our experiments.
The Genc and Hl are used to share weights to extract features from domain X and domain Y. The L layers are selected first and pass the feature mappings through Hl, as used in SimCLR [12]. A feature stack { Z l } L = { H l ( G enc l ( x ) ) } L is then generated, where the output of the l-th selection layer is represented by G enc l . Similarly, we encode the output image G(x) belonging to domain Y into { Z ^ l } L = { H l ( G enc l ( G ( x ) ) ) } L . Our goal is to match the input and output patches corresponding to a specific position. The PatchNCE Loss [13] for the one-way mapping in this article is represented as follows:
L PatchNCE ( G , H , X ) = E x ~ X l = 1 L s = 1 S l ( Z ^ l S , Z l S , Z l S \ s )
where the number of spatial positions in each layer is represented by S l . Z l S R C l represents the corresponding feature (“positive”), Z l S \ s R ( S l 1 ) × C l represents other features (“negatives”), where the number of channels in each layer is represented by C l .

3.3.2. Content-Consistency Loss

Although CUT has produced competitive results with less computational costs for training, we believe that the model’s performance is limited due to certain design options. CUT may not be able to capture domain gaps efficiently, thus missing edges and details of the image data, as one embedding is shared in two distinct image domains (domain X and domain Y). For the above problems, the content-consistency loss is designed in the loss function for joint training. The content-consistency loss derived from the multiscale structural similarity index (MS-SSIM) enhances the high-frequency information of the image [36]. The definition of content-consistency loss is as below:
L Content ( G , X ) = E x ~ X [ 1 msSSIM ( G ( x ) , x ) ]
where the content change between the generated color image G(x) and the input single-channel image x ∈ X is reduced as much as possible. We aim to avoid color inversion and feature distortion between input and output images through content-consistency loss [37].

3.3.3. Adversarial Loss

We use the adversarial loss [20] to make the output image G(x) of the generator G as visually identical as possible to the color image of the target domain Y. For the one-way mapping with the generator/discriminator (G/D) pair, the GAN loss is represented as follows:
L GAN ( G , D , X , Y ) = E y ~ Y [ log D ( y ) ] + E x ~ X [ log ( 1 D ( G ( x ) ) ) ]
where G tries to generate images G(x) that look similar to images from domain Y, while D aims to distinguish between translated samples G(x) and real samples y.

3.3.4. Identity Loss

We add an identity loss [10] to avoid unnecessary changes from generator G. The main idea of identity loss is to take the image y of domain Y as the input of G, and the output image G(y) should be as consistent with the input image y as possible, without producing other content.
L Identity ( G , Y ) = E y ~ Y [ G ( y ) y 1 ]
The input and output mappings are encouraged to preserve color and brightness composition by such an identity loss.

3.4. Overall Objective

Our ultimate goal is to make the patches of the network input and output images learn the same correspondence, thereby generating more realistic color images. The total loss is estimated by:
L ( G , D , H ) = λ G A N L GAN ( G , D , X , Y ) + λ N C E L PatchNCE ( G , H , X ) + λ C C L L Content ( G , X ) + λ I d t L Identity ( G , Y )
where λ G A N , λ N C E , λ C C L and λ Idt are the weight coefficients of adversarial loss, patchNCE loss, content-consistency loss and identity loss, respectively. The weight coefficients of losses are set to λ G A N = 1 , λ N C E = 10 , λ C C L = 1 , and λ I d t = 1 by default.

4. Experiments

4.1. Datasets

We need to construct a large-scale dataset containing a training set and a test set to train the neural network. However, directly collecting high-resolution color images from FPM has a series of disadvantages: (1) the amount of raw data collected is too large to build datasets of sufficient size; (2) the ground truth of the dataset is obtained through traditional reconstruction methods, so the effect of network reconstruction is limited by traditional methods; (3) the dataset is bound to a specific system, so it is difficult to flexibly apply to other systems.
To avoid the above problems, we directly used the publicly available dataset GlaS@MICCAI’ 2015 [38] as the ground truth (the image resolution was 775 × 522) and generated the corresponding single-channel high-resolution input data through a Fourier ptychographic microscopy imaging model simulation. Our simulation dataset includes 1363 single-channel FPM images and 1362 unpaired color FPM images. Additionally, 1208 single-channel grayscale images and 1279 color images were used for network training, while the rest were used for testing.

4.2. Training Settings

Our proposed model was optimized and improved based on CUT. To better compare with the baseline, we trained our model primarily using the settings specified in CUT. More specifically, the initial learning rate was set to 0.0002, and due to GPU memory limitations, the batch size was set to 1. Because the Adam [39] optimizer has the advantages of an adaptive adjusted learning rate, integrated gradient and momentum, robustness, and parallel computation-friendliness, it was selected in this paper to adaptively update the model parameters. The size of all training images was resized to 256 × 256 pixels and 200 epochs were trained. Cropping the image to 256 × 256 was done because the size is large enough to include the main details in the image while maintaining computational efficiency, making it easy to input them into the neural network for training and processing. The learning rate began to decay linearly when the network was trained to half of the total epochs. The model presented in this paper was deployed and trained on the PyTorch deep learning framework and a single NVIDIA GPU (RTX 3080, 10 GB).

4.3. Comparison Results with Unsupervised Deep Learning Methods

Here, we provide a comprehensive qualitative and quantitative evaluation of our approach by comparing it to recent state-of-the-art unsupervised methods, such as CUT [14], CycleGAN [10], UNIT [21], GCGAN [22], and DCLGAN [23]. These methods are retrained in our datasets and evaluated using ground truth values.

4.3.1. Visual Comparisons

Figure 6 illustrates the results of the qualitative comparison of the different contrasting methods. Our method successfully stains the unpaired single-channel FPM input image into a three-channel color FPM target domain image through image translation. The method proposed in this article can preserve the structure and content of the input image to the maximum extent while performing good virtual coloring on the input image, achieving the highest color image quality.
Specifically, CUT cannot accurately identify the unstained part of the image. CycleGAN and UNIT can recognize the white part of pathological sections well, but lose the texture details inside. DCLGAN introduces artifacts and has low image quality. In addition, all baselines have a certain difference in color compared to the ground truth value. Our proposed method has the highest visual similarity, which proves the effectiveness of the network.

4.3.2. Quantitative Examinations

We mainly choose frechet inception distance (FID) [40], LPIPS [41], structural similarity (SSIM), and peak signal-to-noise ratio (PSNR) as the evaluation indicators in this paper to measure the quality of the generated color images. It is worth noting that FID and LPIPS show high correspondence with human visual perception. Lower FID and LPIPS mean generated images are more realistic. On the contrary, larger values of SSIM and PSNR indicate better image quality.
Table 1 shows the evaluation metrics for the different comparison experiments. It can be observed that our method significantly outperforms the other baselines in the FID, LPIPS, SSIM, and PSNR metrics. The specific indicators strongly demonstrate the effectiveness of our proposed method.

4.4. Comparative Experiments under Noisy Conditions

In the process of actual data acquisition by FPM, it is easy to be affected by the equipment, light source, dust, and other environments producing a certain degree of noise, and thus affecting the color image quality of the FPM acquisition and reconstruction. We added Gaussian noise with mean 0 and standard deviation 6 × 10−4 as the main interference condition for the single-channel input data of the network to simulate the potential noise effects during the actual acquisition process, thus further verifying the effectiveness of our proposed algorithm.
A qualitative comparison and quantitative analysis under the same noise condition are shown in Figure 7 and Table 2, respectively. With the same noise added, our method has an optimal visual performance compared to contrast experiments. Our method also performs the best on the four indicators in Table 2.
Different degrees of noise may be generated during the actual acquisition process, and to further verify the robustness and generalization ability of our proposed method, we conducted three sets of experiments. Gaussian noise with a mean of 0 and standard deviation of 6 × 10−4, 6 × 10−3, and 6 × 10−2 was added to the input images of the network. Figure 8 shows the visual analysis results of our method under different noise conditions, and Table 3 shows the evaluation index of stained images under three kinds of noise interference. As can be seen from the specific metrics in Table 3, the PSNR metric performs well as the Gaussian noise increases, and the FID, LPIPS, and SSIM metrics produce some degree of deterioration, but the degree of change is small and essentially flat. The visual analysis and quantitative results demonstrate the effectiveness of our proposed method against noise.

4.5. Ablation Experiment

To analyze the validity of our approach in depth, we conducted several experiments to study each of our contributions individually. The first ablation experiment removed content-consistency loss and identity loss while using the patchNCE loss in CUT as a learnable and domain-specific identity loss to prevent unnecessary changes by the generator. Another ablation experiment replaced the proposed ECR block with the residue block from CUT in the generator.
The experiments are summarized as described in Figure 9 and Table 4. The removal of content-consistency loss and identity loss results in the loss of high-frequency information in the output image, and the visual effect of coloring deteriorates. The ability of our method to accurately recognize the blank portion of a pathology section is diminished with the removal of the ECR block, resulting in poor image quality. Our proposed method shows superior performance in all metrics and visual comparisons.

4.6. Comparison Results with Classical Virtual Staining Methods

We compared our model with four other classical and recent virtual coloring methods to further validate the effectiveness and time efficiency of the network in grayscale image coloring. Reinhard [42] and Macenko [43] are classic methods for image color transfer, while CFPM [16] and CFFPM [44] are recently proposed methods for FPM virtual staining based on color transfer. The above method was applied to our dataset for training to achieve virtual coloring of FPM single-channel grayscale images. The experimental results show that the proposed method has advantages in terms of color and time.
We randomly selected three images for display in Figure 10, and through visual analysis, it can be seen that the method proposed in this article has achieved the best results in terms of color and content. In our data, the Reinhard and Macenko methods performed poorly, with a wide difference in color styles and an inability to accurately stain cells at the appropriate locations. CFPM and CFFPM had relatively good results in distinguishing and coloring areas with different shades of color, but the overall color was relatively light and did not meet the expected goals. Moreover, CFFPM introduced significant artifacts. In conclusion, the deep learning method proposed in this paper had the best experimental results.
Specific evaluation indicators are shown in Table 5. We choose the number of iterations and the inference time as indicators to measure the time efficiency. Our proposed approach has the least time cost and the best performance. CFFPM requires iteration and takes the longest time. The Reinhard method is also fast in time, but ineffective.

5. Conclusions

We propose an effective self-supervised deep learning method based on contrastive learning for virtual staining of single-channel FPM pathological sections. The proposed method designs an ECR-based generator and a content-consistency loss for joint training to obtain efficient cross-channel interaction information and improve the quality of stained images in an extremely lightweight manner. The experimental results show that this method maximizes the high-frequency information of the input image while effectively staining the FPM images with the best visual similarity. This virtual staining method can avoid the differences in staining results due to the stochasticity of personnel and technology, which can standardize staining, and it has various potential applications in digital pathology in the future.

Author Contributions

Conceptualization, Y.W., N.G. and J.L.; methodology, Y.W. and N.G.; software, N.G.; validation, N.G.; investigation, N.G.; writing—original draft preparation, N.G.; writing—review and editing, Y.W., N.G. and X.W.; supervision, Y.W., J.L. and X.W.; project administration, Y.W. and N.G. contributed equally to this paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science and Technology Development Plan Projects of Jilin Province [NO: YDZJ202301ZYTS180].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the raw data used for training and evaluating the performance of the models are publicly available datasets that can be obtained from [36].

Acknowledgments

We would like to thank Hao Wang and Xinbo Wang for their support and contributions to the work of this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Treanor, D. Virtual Slides: An Introduction. Diagn. Histopathol. 2009, 15, 99–103. [Google Scholar] [CrossRef]
  2. Zheng, G.; Horstmeyer, R.; Yang, C. Wide-field, high-resolution Fourier ptychographic microscopy. Nat. Photon 2013, 7, 739–745. [Google Scholar] [CrossRef] [PubMed]
  3. Zheng, G.; Shen, C.; Jiang, S.; Song, P.; Yang, C. Concept, implementations and applications of Fourier ptychography. Nat. Rev. Phys. 2021, 3, 207–223. [Google Scholar] [CrossRef]
  4. Chung, J.; Lu, H.; Ou, X.; Zhou, H.; Yang, C. Wide-field Fourier ptychographic microscopy using laser illumination source. Biomed. Opt. Express 2016, 7, 4787–4802. [Google Scholar] [CrossRef] [PubMed]
  5. Ou, X.; Horstmeyer, R.; Yang, C.; Zheng, G. Quantitative phase imaging via Fourier ptychographic microscopy. Opt. Lett. 2013, 38, 4845–4848. [Google Scholar] [CrossRef] [PubMed]
  6. Horstmeyer, R.; Ou, X.; Zheng, G.; Willems, P.; Yang, C. Digital pathology with Fourier ptychography. Comput. Med. Imaging Graph. 2015, 42, 38–43. [Google Scholar] [CrossRef] [PubMed]
  7. LeCun, Y.; Bengio, Y.; Hinton, G. Deep learning. Nature 2015, 521, 436–444. [Google Scholar] [CrossRef] [PubMed]
  8. Guo, C.; Jiang, S.; Yang, L.; Song, P.; Wang, T.; Shao, X.; Zhang, Z.; Murphy, M.; Zheng, G. Deep learning-enabled whole slide imaging (DeepWSI): Oil-immersion quality using dry objectives, longer depth of field, higher system throughput, and better functionality. Opt. Express 2021, 29, 39669–39684. [Google Scholar] [CrossRef]
  9. Zhu, J.; Park, T.; Isola, P.; Efros, A.A. Unpaired image-to-image translation using cycle-consistent adversarial networks. In Proceedings of the 2017 IEEE International Conference on Computer Vision (ICCV), Venice, Italy, 22–29 October 2017. [Google Scholar]
  10. Li, C.; Liu, H.; Chen, C.; Pu, Y.; Chen, L.; Henao, R.; Carin, L. Alice: Towards understanding adversarial learning for joint distribution matching. In Proceedings of the Advances in Neural Information Processing Systems (NIPS), Long Beach, CA, USA, 4–9 December 2017. [Google Scholar]
  11. He, K.; Fan, H.; Wu, Y.; Xie, S.; Girshick, R. Momentum contrast for unsupervised visual representation learning. In Proceedings of the 2020 IEEE/CVF Conference on Computer Vision and Pattern Recognition (CVPR), Seattle, WA, USA, 13–19 June 2020. [Google Scholar]
  12. Chen, T.; Kornblith, S.; Norouzi, M.; Hinton, G. A simple framework for contrastive learning of visual representations. In Proceedings of the International Conference on Machine Learning (ICML), Vienna, Austria, 14–16 July 2020. [Google Scholar]
  13. Park, T.; Efros, A.A.; Zhang, R.; Zhu, J.Y. Contrastive learning for unpaired image-to-image translation. In Proceedings of the European Conference on Computer Vision (ECCV), Glasgow, UK, 23–28 August 2020. [Google Scholar]
  14. Zhou, Y.; Wu, J.; Bian, Z.; Suo, J.; Zheng, G.; Dai, Q. Fourier ptychographic microscopy using wavelength multiplexing. J. Biomed. Opt. 2017, 22, 066006. [Google Scholar] [CrossRef]
  15. Dong, S.; Shiradkar, R.; Nanda, P.; Zheng, G. Spectral multiplexing and coherent-state decomposition in Fourier ptychographic imaging. Biomed. Opt. Express 2014, 5, 1757–1767. [Google Scholar] [CrossRef]
  16. Gao, Y.; Chen, J.; Wang, A.; Pan, A.; Ma, C.; Yao, B. High-throughput fast full-color digital pathology based on Fourier ptychographic microscopy via color transfer. Sci. China Phys. Mech. Astron. 2021, 64, 114211. [Google Scholar] [CrossRef]
  17. Zhang, M.; Liang, Y. Color Fourier stacked microscopy based on three-dimensional convolutional neural networks. J. Opt. 2020, 40, 2011001. [Google Scholar]
  18. Wang, R.; Song, P.; Jiang, S.; Yan, C.; Zhu, J.; Guo, C.; Bian, Z.; Wang, T.; Zheng, G. Virtual brightfield and fluorescence staining for Fourier ptychography via unsupervised deep learning. Opt. Lett. 2020, 45, 5405–5408. [Google Scholar] [CrossRef] [PubMed]
  19. Zhang, J.; Li, J.; Sun, H.; Jiang, S.; Zhang, Y.; Chen, Y.; Zhang, J.; Xu, T. Edge-enabled anti-noise telepathology imaging reconstruction technology in harsh environments. IEEE Netw. 2022, 36, 92–99. [Google Scholar] [CrossRef]
  20. Goodfellow, I.; Pouget-Abadie, J.; Mirza, M.; Xu, B.; Warde-Farley, D.; Ozair, S.; Courville, A.; Bengio, Y. Generative adversarial nets. In Proceedings of the Advances in Neural Information Processing Systems (NIPS), Montreal, QC, Canada, 8–13 December 2014. [Google Scholar]
  21. Liu, M.Y.; Breuel, T.; Kautz, J. Unsupervised image-to-image translation networks. In Proceedings of the Advances in Neural Information Processing Systems (NIPS), Long Beach, CA, USA, 4–9 December 2017. [Google Scholar]
  22. Fu, H.; Gong, M.; Wang, C.; Batmanghelich, K.; Zhang, K.; Tao, D. Geometry-consistent generative adversarial networks for one-sided unsupervised domain mapping. In Proceedings of the IEEE/CVF Conference on Computer Vision and Pattern Recognition (CVPR), Long Beach, CA, USA, 16–20 June 2019. [Google Scholar]
  23. Han, J.; Shoeiby, M.; Petersson, M.; Armin, M.A. Dual contrastive learning for unsupervised image-to-image translation. In Proceedings of the 2021 IEEE/CVF Conference on Computer Vision and Pattern Recognition Workshops (CVPRW), Nashville, TN, USA, 19–25 June 2021. [Google Scholar]
  24. Jeong, J.; Shin, J. Training GANs with stronger augmentations via contrastive discriminator. In Proceedings of the International Conference on Learning Representations (ICLR), Vienna, Austria, 4–8 May 2021. [Google Scholar]
  25. Kang, M.; Park, J. ContraGAN: Contrastive learning for conditional image generation. In Proceedings of the Advances in Neural Information Processing Systems (NIPS), Vancouver, BC, Canada, 6–12 December 2020. [Google Scholar]
  26. Liu, R.; Ge, Y.; Choi, C.L.; Wang, X.; Li, H. DivCo: Diverse conditional image synthesis via contrastive generative adversarial network. In Proceedings of the 2021 IEEE/CVF Conference on Computer Vision and Pattern Recognition (CVPR), Nashville, TN, USA, 20–25 June 2021. [Google Scholar]
  27. Zhang, H.; Goodfollow, I.; Metaxes, D.; Odena, A. Self-attention generative adversarial networks. In Proceedings of the International Conference on Machine Learning (ICML), Long Beach, CA, USA, 10–15 June 2019. [Google Scholar]
  28. Emami, H.; Aliabadi, M.M.; Dong, M.; Chinnam, R.B. SPA-GAN: Spatial attention gan for image-to-image translation. IEEE Trans. Multimed. 2020, 23, 391–401. [Google Scholar] [CrossRef]
  29. Hu, X.; Zhou, X.; Huang, Q.; Shi, Z.; Sun, L.; Li, Q. QS-Attn: Query-selected attention for contrastive learning in I2I translation. In Proceedings of the 2022 IEEE/CVF Conference on Computer Vision and Pattern Recognition (CVPR), New Orleans, LA, USA, 18–24 June 2022. [Google Scholar]
  30. Torbunov, D.; Huang, Y.; Yu, H.; Huang, G.; Yoo, S.; Lin, M.; Viren, B.; Ren, Y. UVCGAN: UNet Vision Transformer cycle-consistent GAN for unpaired image-to-image translation. In Proceedings of the 2023 IEEE/CVF Winter Conference on Applications of Computer Vision (WACV), Waikoloa, HI, USA, 2–7 January 2023. [Google Scholar]
  31. Isola, P.; Zhu, J.Y.; Zhou, T.; Efros, A.A. Image-to-image translation with conditional adversarial networks. In Proceedings of the 2017 IEEE Conference on Computer Vision and Pattern Recognition (CVPR), Hawaii, HI, USA, 21–26 July 2017. [Google Scholar]
  32. 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 (CVPR), Las Vegas, NV, USA, 26 June–1 July 2016. [Google Scholar]
  33. 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 IEEE/CVF Conference on Computer Vision and Pattern Recognition (CVPR), Seattle, WA, USA, 13–19 June 2020. [Google Scholar]
  34. Oord, A.v.d.; Li, Y.; Vinyals, O. Representation learning with contrastive predictive coding. arXiv 2018, arXiv:1807.03748. [Google Scholar]
  35. Michael, G.; Aapo, H. Noise-contrastive estimation: A new estimation principle for unnormalized statistical models. In Proceedings of the Thirteenth International Conference on Artificial Intelligence and Statistics, Sardinia, Italy, 13–15 May 2010. [Google Scholar]
  36. Zhao, H.; Gallo, O.; Frosio, I.; Kautz, J. Loss functions for image restoration with neural networks. IEEE Trans. Comput. Imaging 2016, 3, 47–57. [Google Scholar] [CrossRef]
  37. Ouyang, W.; Aristov, A.; Lelek, M.; Hao, X.; Zimmer, C. Deep learning massively accelerates super-resolution localization microscopy. Nat. Biotechnol. 2018, 36, 460–468. [Google Scholar] [CrossRef]
  38. Sirinukunwattana, K.; Pluim, J.P.; Chen, H.; Qi, X.; Heng, P.A.; Guo, Y.B.; Wang, L.Y.; Matuszewski, B.; Bruni, E.; Sanchez, U.; et al. Gland segmentation in colon histology images: The glas challenge contest. Med. Image Anal. 2017, 35, 489–502. [Google Scholar] [CrossRef]
  39. Kingma, D.P.; Ba, J. Adam: A method for stochastic optimization. arXiv 2019, arXiv:1412.6980. [Google Scholar]
  40. Heusel, M.; Ramsauer, H.; Unterthiner, T.; Nessler, B.; Hochreiter, S. Gans trained by a two time-scale update rule converge to a local nash equilibrium. In Proceedings of the Advances in Neural Information Processing Systems (NeurIPS), Long Beach, CA, USA, 8–14 December 2017. [Google Scholar]
  41. Zhang, R.; Isola, P.; Efros, A.A.; Shechtman, E.; Wang, O. The unreasonable effectiveness of deep features as a perceptual metric. In Proceedings of the 2018 IEEE/CVF Conference on Computer Vision and Pattern Recognition (CVPR), Salt Lake City, UT, USA, 18–22 June 2018. [Google Scholar]
  42. Reinhard, E.; Adhikhmin, M.; Gooch, B.; Shirley, P. Color transfer between images. IEEE Comput. Graph. Appl. 2001, 21, 34–41. [Google Scholar] [CrossRef]
  43. Macenko, M.; Niethammer, M.; Marron, J.S.; Borland, D.; Woosley, J.T.; Guan, X.; Schmitt, C.; Thomas, N.E. A method for normalizing histology slides for quantitative analysis. In Proceedings of the 2009 IEEE International Symposium on Biomedical Imaging: From Nano to Macro, Boston, MA, USA, 28 June–1 July 2009. [Google Scholar]
  44. Chen, J.; Wang, A.; Pan, A.; Zheng, G.; Ma, C.; Yao, B. High-throughput fast full-color Fourier ptychographic microscopy via color transfer and spatial filtering. Photonics Res. 2022, 10, 2410–2421. [Google Scholar] [CrossRef]
Figure 1. Different H&E color styles on the two datasets. (a) GlaS dataset, (b) MoNuSeg dataset.
Figure 1. Different H&E color styles on the two datasets. (a) GlaS dataset, (b) MoNuSeg dataset.
Applsci 14 01662 g001
Figure 2. Training and testing process framework for the proposed method. (a) Train one generator–discriminator pair using two unpaired datasets (X and Y). (b) The generator converts a FPM single-channel grayscale image into a three-channel virtual stained image with improved image quality.
Figure 2. Training and testing process framework for the proposed method. (a) Train one generator–discriminator pair using two unpaired datasets (X and Y). (b) The generator converts a FPM single-channel grayscale image into a three-channel virtual stained image with improved image quality.
Applsci 14 01662 g002
Figure 3. The overall architecture of the network. The generator G is broken down into two components: an encoder Genc and a decoder Gdec. We use Genc and the MLP (Hl) as the embedding of domain X and domain Y. We expect the PatchNCE loss to make the red patch in the fake image generated by generator G more consistent with the yellow patch in the real input image while minimizing the similarity with other blue patches.
Figure 3. The overall architecture of the network. The generator G is broken down into two components: an encoder Genc and a decoder Gdec. We use Genc and the MLP (Hl) as the embedding of domain X and domain Y. We expect the PatchNCE loss to make the red patch in the fake image generated by generator G more consistent with the yellow patch in the real input image while minimizing the similarity with other blue patches.
Applsci 14 01662 g003
Figure 4. The specific structure of generator and discriminator networks. The generator adopts an encoder–decoder architecture with nine ECR blocks. The discriminator utilizes the classic PatchGAN network, which has five convolutional layers.
Figure 4. The specific structure of generator and discriminator networks. The generator adopts an encoder–decoder architecture with nine ECR blocks. The discriminator utilizes the classic PatchGAN network, which has five convolutional layers.
Applsci 14 01662 g004
Figure 5. The structure of the ECR block.
Figure 5. The structure of the ECR block.
Applsci 14 01662 g005
Figure 6. Qualitative comparison between the best five methods.
Figure 6. Qualitative comparison between the best five methods.
Applsci 14 01662 g006
Figure 7. Qualitative comparison under the same noise condition.
Figure 7. Qualitative comparison under the same noise condition.
Applsci 14 01662 g007
Figure 8. Qualitative comparison of our method under the different noise conditions.
Figure 8. Qualitative comparison of our method under the different noise conditions.
Applsci 14 01662 g008
Figure 9. The qualitative comparison results between the two ablation experiments and the proposed method. (a) Input, (b) Ours w/o Loss, (c) Ours w/o ECR, (d) Ours, (e) Ground Truth.
Figure 9. The qualitative comparison results between the two ablation experiments and the proposed method. (a) Input, (b) Ours w/o Loss, (c) Ours w/o ECR, (d) Ours, (e) Ground Truth.
Applsci 14 01662 g009
Figure 10. Qualitative comparison with classical virtual staining methods.
Figure 10. Qualitative comparison with classical virtual staining methods.
Applsci 14 01662 g010
Table 1. Quantitative results of different approaches.
Table 1. Quantitative results of different approaches.
MethodFID↓LPIPS↓SSIM↑PSNR↑
CUT [14]58.70770.18320.907820.1739
CycleGAN [10]63.11890.16880.909519.4760
UNIT [21]81.03820.17090.888419.1821
GCGAN [22]76.10180.16490.893418.3975
DCLGAN [23]95.70420.21310.872217.5598
Ours49.36860.13340.924120.4459
Table 2. Quantitative results under the same noise condition.
Table 2. Quantitative results under the same noise condition.
MethodFID↓LPIPS↓SSIM↑PSNR↑
CUT [14]67.67800.18320.906620.0794
CycleGAN [10]72.82580.16490.909319.4901
UNIT [21]90.36500.16990.888419.3234
GCGAN [22]83.33630.16920.887518.2488
DCLGAN [23]104.51460.21090.871617.6097
Ours56.47090.13270.922020.3231
Table 3. Quantitative results of our method under the different noise conditions.
Table 3. Quantitative results of our method under the different noise conditions.
NoiseFID↓LPIPS↓SSIM↑PSNR↑
6 × 10−456.47090.13270.922020.3231
6 × 10−356.48540.13430.921420.3286
6 × 10−256.52980.13780.920320.4288
Table 4. Quantitative results for ablations.
Table 4. Quantitative results for ablations.
MethodFID↓LPIPS↓SSIM↑PSNR↑
Ours w/o Loss67.80050.18660.908620.3208
Ours w/o ECR57.37470.13720.918019.8297
Ours49.36860.13340.924120.4459
Table 5. Comparison of time efficiency between different staining methods.
Table 5. Comparison of time efficiency between different staining methods.
Virtual Staining MethodIterations↓Inference Time↓
Reinhard [40]01.354 s
Macenko [41]017.130 s
CFPM [16]050.426 s
CFFPM [42]5799.343 s
Ours00.313 s
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

Wang, Y.; Guan, N.; Li, J.; Wang, X. A Virtual Staining Method Based on Self-Supervised GAN for Fourier Ptychographic Microscopy Colorful Imaging. Appl. Sci. 2024, 14, 1662. https://doi.org/10.3390/app14041662

AMA Style

Wang Y, Guan N, Li J, Wang X. A Virtual Staining Method Based on Self-Supervised GAN for Fourier Ptychographic Microscopy Colorful Imaging. Applied Sciences. 2024; 14(4):1662. https://doi.org/10.3390/app14041662

Chicago/Turabian Style

Wang, Yan, Nan Guan, Jie Li, and Xiaoli Wang. 2024. "A Virtual Staining Method Based on Self-Supervised GAN for Fourier Ptychographic Microscopy Colorful Imaging" Applied Sciences 14, no. 4: 1662. https://doi.org/10.3390/app14041662

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