Skip to main content

ORIGINAL RESEARCH article

Front. Physiol., 30 August 2022
Sec. Medical Physics and Imaging
This article is part of the Research Topic Artificial Intelligence in Medical Physics and Imaging View all 5 articles

Self-supervised learning for macromolecular structure classification based on cryo-electron tomograms

  • 1Department of Computer Science and Engineering, Indian Institute of Technology, Indore, India
  • 2Department of Electrical and Computer Engineering, University of California, San Diego, San Diego, CA, United States
  • 3Computational Biology Department, Carnegie Mellon University, Pittsburgh, PA, United States
  • 4Irvington High School, Irvington, NY, United States
  • 5Department of Computer Science, University of California, Irvine, Irvine, CA, United States
  • 6Departments of Psychiatry and Cell Biology, University of Pittsburgh, Pittsburgh, PA, United States

Macromolecular structure classification from cryo-electron tomography (cryo-ET) data is important for understanding macro-molecular dynamics. It has a wide range of applications and is essential in enhancing our knowledge of the sub-cellular environment. However, a major limitation has been insufficient labelled cryo-ET data. In this work, we use Contrastive Self-supervised Learning (CSSL) to improve the previous approaches for macromolecular structure classification from cryo-ET data with limited labels. We first pretrain an encoder with unlabelled data using CSSL and then fine-tune the pretrained weights on the downstream classification task. To this end, we design a cryo-ET domain-specific data-augmentation pipeline. The benefit of augmenting cryo-ET datasets is most prominent when the original dataset is limited in size. Overall, extensive experiments performed on real and simulated cryo-ET data in the semi-supervised learning setting demonstrate the effectiveness of our approach in macromolecular labeling and classification.

1 Introduction

Cryo-electron tomography (cryo-ET) is a revolutionary imaging technology with notable applications in the field of cell and structural biology (Gan and Jensen, 2012; Lučić et al., 2013; Zhang, 2013). Our understanding of the structures and accompanying functions of key components of the cellular microenvironment have been significantly expanded by cryo-ET (Grünewald et al., 2003; Cyrklaff et al., 2005; Koning and Koster, 2009). Furthermore, cryo-ET has provided new insights into human disease states including mitochondrial diseases and, most recently, COVID-19 where the structure and function of SARS-CoV-2 was determined in infected host cells (Klein et al., 2020). Another major advantage of cryo-ET is that high-resolution 3D images of subcellular structures (e.g., organelles and macromolecules) are acquired in their near-native states in contrast to earlier approaches that require fixation, sectioning and dehydration steps that may distort or alter cellular architecture (Oikonomou and Jensen, 2017). The 3D images are referred to as tomograms and the small subvolumes of the tomograms that visualize individual macromolecule are termed subtomograms. 3D visualization by cryo-ET enables resolution of the structures of the subcellular components and their spatial interactions in situ within single cells.

To understand macromolecular interactions and dynamics, classifying individual macromolecular structures from the subtomograms is a crucial step (Murata and Wolf, 2018). The classification implies identifying the target macromolecules from subtomograms. However, due to the crowded and heterogeneous cellular environment, each subtomogram closely packs several neighboring potentially unrelated macromolecules alongside the target macromolecule of interest (Best et al., 2007). Thus, the closely packed structures in a single subtomogram makes macromolecular classification challenging (Best et al., 2007). Due to its resemblance to 3D image classification, several deep classification models have been deployed to perform macromolecular classification. VP-Detector (Hao et al., 2022), which uses 3D multiscale convolutional neural network, is one of the recent approaches for cryo-ET classification. However, most of these classification methods are supervised and sample-inefficient. For cryo-ET, availability of labelled data is limited due to the rigorous annotation process. Furthermore, the performance of deep supervised classification models relies on the number of labelled cryo-ET subtomograms (Frazier et al., 2017). One strategy to tackle the scarcity of labelled data is to generate simulated cryo-ET data on which supervised models will be trained - an approach used by several previous studies that simulated cryo-ET subtomogram data (Pei et al., 2016; Liu et al., 2020a,b). There is also a recent study using simulated data for supervised training followed by application to experimental data (Moebel and Kervrann, 2022). Nevertheless, models trained using simulated data often perform poorly when analyzing actual experimental data due to domain shift. In contrast, semi-supervised approaches have the capability to deal with lack of labelled data and avoid the problem of domain shift in simulated data. Thus several approaches (Yu et al., 2020; Du et al., 2021) have been developed that utilise both labelled and unlabelled data for subtomogram classification (Chapelle et al., 2009). However, the accuracy obtained from these approaches is yet to reach near the accuracy from supervised approaches. As a result, improvement of these semi-supervised approaches for subtomogram classification continues to remain a problem.

Recently, self-supervised learning (SSL) (Noroozi and Favaro, 2016; Pathak et al., 2016; Zhang et al., 2016; Komodakis and Gidaris, 2018) has been proven to be an effective unsupervised technique to learn data representations by solving auxiliary tasks on input data, which does not require any human-defined annotations. Contrastive Self-supervised Learning (CSSL) (Hadsell et al., 2006), as a subcategory of SSL, has been widely used to learn better representations of images and has been successful in achieving state-of-the-art results in various domains of image classification (He et al., 2020; Tian et al., 2019; Chen et al., 2020a; Caron et al., 2020; Misra and Maaten, 2020). CSSL learns image representations by optimizing the contrastive loss using positive and negative pairs, where positive pairs refers to pairs of images which are augmentations of the same image, and negative pairs refer to augmentations sourcing from the rest.

In this work, we use CSSL to improve the current semi-supervised methods for cryo-ET macro-molecule classification. Specifically, we use SimCLR (Chen et al., 2020a), MoCo (He et al., 2020) and SwAV (Zhu et al., 2020) methods to pretrain weights for the classification. These methods are illustrated in Figure 1. To this end, we design a domain-specific augmentation pipeline for cryo-ET data. The augmentation pipeline consists of 3D affine transformations: translation, rotation and scaling. Given the augmentation pipeline, the CSSL task is to contrast positive pairs against negative pairs, enabling a deep-learning model to learn cryo-ET data representations without the need of labels. The CSSL-pretrained weights are then fine-tuned on the downstream classification task using subsets of the training dataset, so as to mimic semi-supervised learning settings. The overall pipeline is shown in Figure 2. The main contributions of this work are summarised as follows:

• We propose a self-supervised learning framework for classification of macromolecules from subtomograms extracted from cryo-ET images.

• We design a simple yet effective data augmentation strategy for 3D cryo-ET subtomogram images.

• We demonstrate the improvements of self-supervised learning in a semi-supervised learning setting using both labelled and unlabelled cryo-ET data.

• Experiments on simulated and experimentally-derived cryo-ET data show the effectiveness and substantial improvements by our proposed approach.

FIGURE 1
www.frontiersin.org

FIGURE 1. Illustration of methods SimCLR (Chen et al., 2020a), MoCo (He et al., 2020) and SwAV (Zhu et al., 2020), which we use for cryo-ET subtomogram classification.

FIGURE 2
www.frontiersin.org

FIGURE 2. Schematic illustration of the pipeline. The first box represents the CSSL pretraining process. An augmentation pipeline is used to create augmented cryo-ET images which is then fed into a CSSL framework to perform CSSL pretraining. The pretrained encoder is then fine-tuned using labelled cryo-ET data as shown in the second box.

2 Related works

2.1 Pretraining

The most prominent pretraining approach is supervised pretraining (SP) (Pan and Yang, 2009), where the model solves a supervised task, such as predicting class labels, segmenting images etc., to learn the weight updates. Self-supervised learning (Oord et al., 2018; He et al., 2020; Chen et al., 2020a; Misra and Maaten, 2020), has recently gained promising success as an unsupervised pretraining strategy, even outperforming supervised pretraining in certain applications. Self-supervised pretraining (SSP) solves prediction problems, as is the case with SP. However, unlike SP, the labels which are to be predicted by the model are created from input data, rather than being annotated by human beings.

2.2 Data augmentation

Unfortunately, experimental 3D cryo-ET subtomogram image data acquired from cellular imaging, is relatively scarce and hard to collect. Data augmentation is a common method for reducing data bias and helping model generalize better, and can be leveraged to address this issue. Cropping, rotating, occlusion, flipping, shearing, zooming in/out, picture blurring, and adjusting brightness or contrast are all common data-augmentation techniques used in computer vision. In this paper, we propose a brand new data augmentation strategy for 3D cryo-ET subtomogram images, which is especially useful for self-supervised learning.

2.3 Self-supervised learning

Self-supervised learning (SSL) has been widely studied to learn better representations of images. SSL generates a loss from a pretext challenge to learn relevant features without the need for human annotations. It only uses the input data to generate auxiliary tasks, allowing deep neural networks to learn effective latent representations by solving them. Numerous methods have been explored for constructing auxiliary tasks, such as temporal correspondence (Wang et al., 2019b; Liu et al., 2019), cross-modal consistency (Wang et al., 2019a), and so on. Rotation prediction (Komodakis and Gidaris, 2018), picture inpainting (Pathak et al., 2016), automated colorization (Zhang et al., 2016), and instance discrimination (Wu et al., 2018) are only a few examples of auxiliary tasks in computer vision.

2.4 Semi-supervised learning

Semi-supervised learning techniques utilise both labelled and unlabelled data (Chapelle et al., 2009). Unlabelled data often carry important information which can be leveraged via semi-supervised learning. It is particularly useful in domains where getting labelled data is expensive and time-consuming. Recently, SSL is being increasingly used in conjunction with semi-supervised learning techniques (Zhai et al., 2019; Chen et al., 2020b). SSP is first used to learn data representations from unlabelled data, a process termed as the pretraining phase. The weights learned in the pretraining phase are then fine-tuned for the downstream task using labelled data. Therefore, using SSP, one can utilize both labelled and unlabelled data.

In this paper, to evaluate the effectiveness of SSP in a semi-supervised learning setting, the fine-tuning phase only uses a determined portion of the training set. While in the pretraining phase, which does not require labels, we use the whole training set.

2.5 Subtomogram classification

Identifying macromolecules inside cells essentially implies classifying subtomograms extracted from cryo-ET data. Several supervised and semi-supervised methods have been developed for classifying subtomograms. Popular 3D image classification networks (Simonyan and Zisserman, 2014; He et al., 2016; Che et al., 2018) are used for supervised classification. CB3D, DSRF3D_v2 and RB3D (Che et al., 2018) are the recent examples of deep supervised models that have been used. To deal with limited labelled data, active learning (Du et al., 2021) and few shot learning (Yu et al., 2020) based methods have been used to build classification models for classifying macromolecules from subtomograms.

3 Methods

The basic flow of our method, demonstrated in Figure 1, is as follows: Perform CSSL, using the specifically designed data-augmentation pipeline with RB3D (Che et al., 2018) as an encoder. For the downstream classification task, we use the CSSL weights to initialize the RB3D architecture and perform supervised classification using the labelled subset of cryo-ET subtomogram images. The steps in our workflow are described in detail in the sequel.

3.1 Contrastive self supervised learning techniques

We chose three representative self-supervised learning approaches for our studies: SimCLR (Chen et al., 2020a), MoCo (He et al., 2020), and SwAV (Zhu et al., 2020). All of them are based on contrastive learning (Hadsell et al., 2006). The core principle behind contrastive self-supervised learning is to construct augmented instances from original data samples, design a prediction task that asks if two augmented instances are augmented from a single data sample or not, and train the model by solving this auxiliary task. SimCLR (Chen et al., 2020a) is a simple framework for contrastive learning with bigger batch sizes and considerable data augmentation that yields competitive performance as supervised learning. MoCo (Wu et al., 2018) uses a queue, which holds a dynamic collection of augmented data instances (called keys), to accomplish contrastive learning. For the sake of efficiency, a momentum encoder is used to encode the keys. With a query augmentation, a contrastive loss is defined on the query and keys based on whether they come from the same source. SwAV performs contrastive SSP without requiring computation of pairwise comparisons. In SwAV, clustering is performed on the augmentations of data examples. For cluster assignments for different augmentations from the same image, SwAV encourages them to be consistent. Specifically, the code of one augmentation is predicted based on the representation of another augmentation. Because it does not ask for a big memory bank, this technique is considered to be more efficient interms of memory. We introduce detailed descriptions of contrastive learning for self-supervision and a momentum encoder that is equipped with a queue-structured dictionary in the following sections.

3.1.1 Contrastive learning for self-supervision

Based upon an original subtomogram image from the dataset, CSSL (Hadsell et al., 2006) creates two augmented versions of this image denoted by xq and xk, where xq is considered as query and xk as key. The query encoder fq (⋅; θq) and the key encoder fk (⋅; θk), with weights θq and θk respectively, are adopted to gain latent representations q = fq (xq; θq) and k = fk (xk; θk) for xq and xk.

A positive pair consists of a query and a key from the same image, while a negative pair contains a query and a key from different images. The auxiliary task is designed to tell if the given pair is positive or not.

CSSL employs a queue to hold a collection of keys ki from different images, and the contrastive loss is computed by:

LCL=logexpqjkj/τexpqjkj/τ+iexpqjki/τ,(1)

with (qj, kj) being a pair obtained from an image instance and τ being a temperature parameter (He et al., 2020). During the training process, the encoders are updated by optimizing this loss.

3.1.2 Momentum encoder with queue-structured dictionary

To maintain and perform sampling over key vectors, existing approaches use a variety of strategies (Hadsell et al., 2006; Hjelm et al., 2018; Oord et al., 2018; Chen et al., 2020a). Resorting to the same network fk = fq on xk and xq at the same time, the Siamese-like approach is proposed and has been proven to be effective (Chen X. et al., 2020). However, learning discriminative features from comparing fk and fq requires a very big mini-batch size (Chen et al., 2020a). This Siamese-like approach is simple to use, but it is of high computation complexity and is quite resource intensive. As an alternative, a memory bank can be used to store the representations of historical keys in a negative key dictionary Dk = {ki} (Wu et al., 2018). Instead of utilising fk, a mini-batch of keys is sampled from the memory bank at each iteration. The memory bank is updated with the current mini-batch of queries. With an expanded buffer pool, this approach eliminates big batch sizes by default. However, the key sampling step leads to inconsistency when training the encoder. Momentum Contrastive (MoCo) (He et al., 2020) incorporates both types of learning strategies. The memory bank is replaced with a queue-structured key dictionary with a preset length. The oldest key mini-batch will act as the negative key and will be substituted by fresh queries due to the queue’s first-in-first-out (FIFO) principle. This method can avoid negative sampling from being irregular.

An additional important feature of this approach is that parameters of query encoder and key encoder are fixed and do not receive gradient updates. Instead, a running average of the key encoder fk is used to update the query encoder (Tarvainen and Valpola, 2017; He et al., 2020), referred as momentum encoder. Thereby, θk and θq are updated as follows:

θqθqαLθqθkmθk+1mθq,(2)

where the momentum coefficient is denoted by m, and the query encoder’s learning rate is represented by α. As can be seen, θq is updated via the back propagating, while θk from the key encoder always keeps a running average of previous states.

3.2 Encoder

We have used RB3D (Che et al., 2018) as the encoder in MoCo. The architecture of RB3D is illustrated in Figure 3. RB3D is a 3D residual block based neural network, which was designed specifically for classifying 3D cryo-ET images.

FIGURE 3
www.frontiersin.org

FIGURE 3. RB3D model (Che et al., 2018). ‘3×3×3 Conv, 64’ represents a 3D convolutional layer with kernel dimensions 3×3×3 and 64 filters. Other convolutional layers follow similar definitions. All the convolutional layers have a stride of 1. ‘2×2×2 MaxPool’ represents a max-pooling operation over the input signal with kernel size 3×3×3 and stride of 2. ‘Concatenation’ denotes the concatenation of the filters of the same dimensions. ‘FC-1024’ represents a fully connected layer with 1024 neurons. The ‘L’ in ‘FC-L’ corresponds to the output dimension. ‘RelU’ and ‘Softmax’ are activation functions.

3.3 Data-augmentation pipeline

Original data-augmentation pipelines used in CSSL methods such as SimCLR, MoCo and SwAV were designed primarily for traditional 2D RGB image-datasets such as ImageNet (Deng et al., 2009). The augmentation pipeline used random changes in brightness, contrast, saturation and hue of RGB images, along with random horizontal flips and random resized cropping. This augmentation pipeline is very specific to ImageNet like datasets, and needs to be modified to be applied to a different domain (Chaitanya et al., 2020).

Due to the expensive annotation process, experimentally acquired, biological cryo-ET dataset sizes are quite small. Further, the dimensions of subtomograms are also usually small (323 and 283 in the two datasets we use in this paper). In such a case, using strong augmentations can make the pretraining process difficult. We experimented with various permutations and combinations of strong augmentations such as Gaussian blur, Gamma correction, elastic transformations, bias-field etc (Pérez-García et al., 2021). However, due to small-sized datasets coupled with small dimensions of subtomograms, such a augmentation pipeline proved to be too complex for the model to learn useful features during the pretraining phase.

We also considered other image-level augmentations such as the tomography artefacts, e.g. missing wedge effect and electron optical factors [using Contrast Transfer Function (CTF) and Modulation Transfer Function (MTF)], but they are mostly specific to 3D tomography reconstruction from 2D tilt series of cryo-ET images. However, since subtomogram classification is a far downstream task from reconstructing 3D tomograms, it may not be possible to include artefacts that are encountered in a far upstream step in our augmentation pipeline for subtomogram classification. Nevertheless, simulated subtomograms are extracted from the simulated tomograms and, while generating simulated tomograms, we have incorporated the aforementioned tomography artefacts. Consequently, the tomography artefacts are inherent in the simulated subtomogram dataset too. Since the contrastive-learning methods give promising results for simulated data in the presence of tomography artefacts, we consider that contrastive learning methods are robust towards the presence of such artefacts.

Based on the above arguments, we propose a simple yet effective and fine-tuned data-augmentation pipeline as follows:

1) A random resized crop of the image is taken with a probability of 50%. The scale range of the cropped image before resizing is between 0.5 and 1.

2) A random affine transformation is applied with a probability of 50%. This affine transformation includes rotation, translation, and scaling. Image rotation is done by a random angle in the range -45 to 45° along the z axis. Horizontal translation of the image is done by a random fraction 0.1 of horizontal dimension of the image. Similarly, the vertical translation is done by a random fraction 0.1 of the vertical dimension of the image. The image may be scaled up or down by a random scale-factor 0.1.

The intuition behind the above augmentation-pipeline is that to judge if a pair of augmented images originate from a common subtomogram image or not, the model would have to learn global 3D spatial features. These features would then be helpful in downstream classification tasks and may prevent overfitting upon transfer to smaller datasets (Newell and Deng, 2020).

4 Experiments

4.1 Datasets

4.1.1 Simulated data

Several different methods exist for simulating cryo-ET data (Pei et al., 2016; Liu et al., 2020a,b). Here, we use the framework designed by Liu et al. (Liu et al., 2020b). They proposed an efficient gradient descent based method to generate 3D cryo-ET subtomogram images of a target macromolecule with a crowded environment having several random neighbouring macromolecules. The macromolecules are randomly rotated and translated. Further, the simulation procedure includes tomographic artefacts such as the missing wedge effect and electron optical factors to mimic experimentally-acquired cryo-ET images. For illustration, the 3D visualization formed using Chimera (Pettersen et al., 2004) and the 2D slices of a simulated 2h12 macromolecule are shown in Figures 4, 5.

FIGURE 4
www.frontiersin.org

FIGURE 4. 3D isosurface visualization of simulated 2h12 macromolecule, along with randomly simulated macromolecules depicting a crowded subcellular environment.

FIGURE 5
www.frontiersin.org

FIGURE 5. 2D subtomogram slice visualization of simulated 2h12 macromolecule, along with randomly simulated macromolecules depicting a crowded subcellular environment.

For our experiments, we use three simulated datasets with signal to noise ratio (SNR) as , 0.05 and 0.03. Each dataset has 500 images per class for 10 classes and each subtomogram is of size 323 (32 × 32 × 32). For our experiments, the three simulated datasets are split in ratio 60:20:20 for training, validation and testing respectively.

4.1.2 Experimentally acquired biological data

The real dataset has been constructed from the Noble single particle dataset (Noble et al., 2018). For each tomogram in the Noble single particle dataset, potential structural regions have been extracted using the Difference-of-Gaussians (DoG) method (Pei et al., 2016). The top 1000 sub-volumes were selected according to cross-correlation scores (Zeng et al., 2018) and then 400 subtomograms were selected manually for each class (Liu et al., 2019). The final constructed dataset has 400 samples for seven classes and each subtomogram is of size 283 (28 × 28 × 28). For illustration, 3D visualization formed using Chimera (Pettersen et al., 2004) and 2D slices of an extracted T20 S proteasome macromolecule is shown in Figures 6, 7. For our experiments, the dataset is split with the ratio of 3:1:1 for training, validation and testing respectively.

FIGURE 6
www.frontiersin.org

FIGURE 6. 3D isosurface visualization of T20 S proteasome (EMPIAR 10143) macromolecule, extracted from Noble single particle dataset.

FIGURE 7
www.frontiersin.org

FIGURE 7. 2D subtomogram slice visualization of T20 S proteasome (EMPIAR 10143) macromolecule, extracted from Noble single particle dataset.

4.2 Experimental settings

Simulated data: For the MoCo pretraining phase, the MoCo queue size is set to 128. The momentum variable for updating the key encoder is kept as 0.999 and the temperature parameter τ is set as 0.2. Adam optimiser (Kingma and Ba, 2014) is used, with learning rate 1e−4, weight-decay 1e−4 and batch-size 16. The training is done for 200 epochs. For SimCLR and SwAV pretraining phase, the settings are directly inherited from (Chen et al., 2020a) and (Zhu et al., 2020). For the fine-tuning phase, we use the SGD optimiser with cosine decay schedule (Loshchilov and Hutter, 2016). The learning rate is 5e−4, weight-decay 1e−4 and batch-size 16. The fine-tuning is done for 50 epochs and the model with best validation accuracy is chosen. For normal supervised learning with random initialization, the same hyper-parameters as that of MoCo fine-tuning phase are used.

Experimental biological data: For the MoCo pretraining phase, the hyper-parameters are the same as those for simulated data except for MoCo queue size, which is set as 64. In the fine-tuning phase, for 100 and 75% labelled experiments, the hyper-parameters are same as that for simulated data. For 25 and 50% labelled experiment, the learning rate is 1e−4.

4.3 Experimental results

For both simulated and real datasets, we randomly select 25, 50, 75 and 100% of the training set size, and then fine-tune the classification models on these subsets. All the experiments are run 5 times and the average accuracy and the standard deviation are reported. Our results for simulated data have been shown in Table 1 and the results for experimental data in Table 2. We found that subtomogram classification accuracy for our experimentally acquired dataset is comparatively higher than the simulated dataset. This is because of the higher complexity of the simulated dataset due to higher resolution and more closely packed macromolecules. As a result of dataset complexity and small training set, the highest accuracy achieved for simulated dataset is around 69%. At the same time, MoCo outperforms the other two CSSL baselines in most experiments. This may be because MoCo extends the idea of contrastive learning by leveraging an extra dictionary along with a momentum encoder, and is more robust and adaptable to be applied to cryo-ET data. We use two-tailed student’s t-test to reject null-hypothesis (Cox, 1982). The p-value of the MoCo results obtained is 0.046. Considering significance level, α = 0.05, we reject the null hypothesis.

TABLE 1
www.frontiersin.org

TABLE 1. Comparison of subtomogram classification accuracy (%) with standard deviation on experimental biological data. Classifier with CSSL pretrained initial weights performs much better than classifier with random initial weights.

TABLE 2
www.frontiersin.org

TABLE 2. Comparison of subtomogram classification accuracy (%) with standard deviation on real data. Classifiers with CSSL pretrained initial weights always perform better than classifiers with random initial weights proving the efficacy of CSSL pretraining.

We further show the Grad-CAM visualizations (Selvaraju et al., 2017) of a sample subtomogram image for CSSL-pretrained (MoCo) and randomly initialized models in Figure 8, which roughly highlights the region important for making the classification decision. We have used M3d-CAM (Gotkowski et al., 2020) to make these visualizations. It can be observed that the CSSL-pretrained model along with giving higher accuracy also considers wider regions of the 3D environment. These data indicate that CSSL pretraining has a regularization effect on the model (Newell and Deng, 2020). The improvements of CSSL methods over Random Init in all experiments show that the classifier can leverage knowledge gained from CSSL and effectively exploit the representations obtained via pretraining.

FIGURE 8
www.frontiersin.org

FIGURE 8. Grad-CAM Visualizations. CSSL (MoCo) pretrained model shows wider regions of 3D space, indicating regularization effect of CSSL pretraining (Newell and Deng, 2020).

5 Conclusion

In this paper, we addressed the problem of utilizing unlabelled data for macromolecular structure classification from cryo-ET subtomograms. We developed a pipeline that uses the unlabelled subtomogram data for pretraining weights of a classifier using CSSL methods: SimCLR, MoCo and SwAV, yielding a regularization effect over the classification model. To this end, we designed a unique data-augmentation pipeline for cryo-ET subtomogram data. Our pipeline was able to generate cryo-ET subtomogram images, and those generated images worked well as a source of augmentation for self-supervised learning. We fine-tune the CSSL pretrained weights using labelled subtomograms for the downstream classification task. Taken together, we present a novel workflow that provides significant improvement over traditional classification methods on both simulated and real data.

Data availability statement

The method to generate the datasets used this paper has been described in Section 4.1.

Author contributions

MX envisioned the research. TG conducted the experiments. XH designed the experiments. MU and XZ helped in data simulation and preparation. AZ, JZ, and ZF contributed in manuscript writing and preliminary studies.

Funding

This work was supported in part by the U.S. NIH grants R01GM134020, R01DK124219, R21DA052419, and P41GM103712, NSF grants DBI-1949629 and IIS-2007595, DoD grant 192,466, and Mark Foundation for Cancer Research 19-044-ASP. We thank the computational resources support from AMD COVID-19 HPC Fund. JZ, was supported in part by the U.S. NIH grant K01MH123896. XZ was supported in part by a fellowship from CMU CMLH.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Best C., Nickell S., Baumeister W. (2007). Localization of protein complexes by pattern recognition. Methods Cell Biol. 79, 615–638. doi:10.1016/S0091-679X(06)79025-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Caron M., Misra I., Mairal J., Goyal P., Bojanowski P., Joulin A. (2020). Unsupervised learning of visual features by contrasting cluster assignments. Adv. Neural Inf. Process. Syst. 33, 9912–9924.

Google Scholar

Chaitanya K., Erdil E., Karani N., Konukoglu E. (2020). Contrastive learning of global and local features for medical image segmentation with limited annotations. Adv. Neural. Inf. Process Syst. 33, 12546–12558.

Google Scholar

Chapelle O., Scholkopf B., Zien A. (Editors) (2009). Semi-supervised learning. Transactions on Neural Networks. (IEEE Transactions on Neural Networks), 20 (3), 542–542.

CrossRef Full Text | Google Scholar

Che C., Lin R., Zeng X., Elmaaroufi K., Galeotti J., Xu M. (2018). Improved deep learning-based macromolecules structure classification from electron cryo-tomograms. Mach. Vis. Appl. 29, 1227–1236. doi:10.1007/s00138-018-0949-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen T., Kornblith S., Norouzi M., Hinton G. (2020a). “A simple framework for contrastive learning of visual representations,” in International conference on machine learning, PMLR, 1597–1607.

Google Scholar

Chen T., Kornblith S., Swersky K., Norouzi M., Hinton G. (2020b). Big self-supervised models are strong semi-supervised learners Adv. Neural. Inf. Process Syst. 33, 22243–22255.

Google Scholar

Chen X., Fan H., Girshick R., He K. (2020). Improved baselines with momentum contrastive learning. arXiv preprint arXiv:2003.04297

Google Scholar

Cox D. R. (1982). Statistical significance tests. Br. J. Clin. Pharmacol. 14, 325–331. doi:10.1111/j.1365-2125.1982.tb01987.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Cyrklaff M., Risco C., Fernández J. J., Jiménez M. V., Estéban M., Baumeister W., et al. (2005). Cryo-electron tomography of vaccinia virus. Proc. Natl. Acad. Sci. U. S. A. 102, 2772–2777. doi:10.1073/pnas.0409825102

PubMed Abstract | CrossRef Full Text | Google Scholar

Deng J., Dong W., Socher R., Li L.-J., Li K., Fei-Fei L. (2009). “Imagenet: A large-scale hierarchical image database,” in Cvpr.

CrossRef Full Text | Google Scholar

Du X., Wang H., Zhu Z., Zeng X., Chang Y.-W., Zhang J., et al. (2021). Active learning to classify macromolecular structures in situ for less supervision in cryo-electron tomography. Bioinformatics 37, 2340–2346. doi:10.1093/bioinformatics/btab123

CrossRef Full Text | Google Scholar

Frazier Z., Xu M., Alber F. (2017). Tomominer and tomominercloud: A software platform for large-scale subtomogram structural analysis. Structure 25, 951–961.e2. doi:10.1016/j.str.2017.04.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Gan L., Jensen G. J. (2012). Electron tomography of cells. Q. Rev. Biophys. 45, 27–56. doi:10.1017/S0033583511000102

PubMed Abstract | CrossRef Full Text | Google Scholar

Gotkowski K., Gonzalez C., Bucher A., Mukhopadhyay A. (2020). M3d-cam: A pytorch library to generate 3d data attention maps for medical deep learning. arXiv preprint arXiv:2007.00453

Google Scholar

Grünewald K., Desai P., Winkler D. C., Heymann J. B., Belnap D. M., Baumeister W., et al. (2003). Three-dimensional structure of herpes simplex virus from cryo-electron tomography. Science 302, 1396–1398. doi:10.1126/science.1090284

PubMed Abstract | CrossRef Full Text | Google Scholar

Hadsell R., Chopra S., LeCun Y. (2006). “Dimensionality reduction by learning an invariant mapping,” in 2006 IEEE Computer Society Conference on Computer Vision and Pattern Recognition (CVPR’06) (IEEE), 1735–1742.

Google Scholar

Hao Y., Wan X., Yan R., Liu Z., Li J., Zhang S., et al. (2022). Vp-detector: A 3d multi-scale dense convolutional neural network for macromolecule localization and classification in cryo-electron tomograms. Comput. Methods Programs Biomed. 221, 106871. doi:10.1016/j.cmpb.2022.106871

PubMed Abstract | CrossRef Full Text | Google Scholar

He K., Fan H., Wu Y., Xie S., Girshick R. (2020). “Momentum contrast for unsupervised visual representation learning,” in Proceedings of the IEEE/CVF conference on computer vision and pattern recognition, 9729–9738.

CrossRef Full Text | Google Scholar

He K., Zhang X., Ren S., Sun J. (2016). “Deep residual learning for image recognition,” in Proceedings of the IEEE conference on computer vision and pattern recognition, 770–778.

CrossRef Full Text | Google Scholar

Hjelm R. D., Fedorov A., Lavoie-Marchildon S., Grewal K., Bachman P., Trischler A., et al. (2018). Learning deep representations by mutual information estimation and maximization. arXiv preprint arXiv:1808.06670

Google Scholar

Kingma D., Ba J. (2014). “Adam: A method for stochastic optimization,” in International Conference on Learning Representations.

Google Scholar

Klein S., Cortese M., Winter S. L., Wachsmuth-Melm M., Neufeldt C. J., Cerikan B., et al. (2020). Sars-cov-2 structure and replication characterized by in situ cryo-electron tomography. Nat. Commun. 11, 5885–5910. doi:10.1038/s41467-020-19619-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Komodakis N., Gidaris S. (2018). “Unsupervised representation learning by predicting image rotations,” in International Conference on Learning Representations (ICLR).

Google Scholar

Koning R. I., Koster A. J. (2009). Cryo-electron tomography in biology and medicine. Ann. Anatomy-Anatomischer Anzeiger 191, 427–445. doi:10.1016/j.aanat.2009.04.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu S., Ban X., Zeng X., Zhao F., Gao Y., Wu W., et al. (2020a). A unified framework for packing deformable and non-deformable subcellular structures in crowded cryo-electron tomogram simulation. BMC Bioinforma. 21, 1–24. doi:10.1186/s12859-020-03660-w

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu S., Ma Y., Ban X., Zeng X., Nallapareddy V., Chaudhari A., et al. (2020b). “Efficient cryo-electron tomogram simulation of macromolecular crowding with application to sars-cov-2,” in 2020 IEEE International Conference on Bioinformatics and Biomedicine (BIBM), 80. doi:10.1109/BIBM49941.2020.9313185

CrossRef Full Text | Google Scholar

Liu S., Du X., Xi R., Xu F., Zeng X., Zhou B., et al. (2019). Semi-supervised macromolecule structural classification in cellular electron cryo-tomograms using 3d autoencoding classifier. BMVC 30. doi:10.5244/C.33.67

CrossRef Full Text | Google Scholar

Loshchilov I., Hutter F. (2016). Sgdr: Stochastic gradient descent with warm restarts. arXiv preprint arXiv:1608.03983

Google Scholar

Lučić V., Rigort A., Baumeister W. (2013). Cryo-electron tomography: The challenge of doing structural biology in situ. J. Cell Biol. 202, 407–419. doi:10.1083/jcb.201304193

PubMed Abstract | CrossRef Full Text | Google Scholar

Misra I., Maaten L. v. d. (2020). “Self-supervised learning of pretext-invariant representations,” in Proceedings of the IEEE/CVF Conference on Computer Vision and Pattern Recognition, 6707–6717.

CrossRef Full Text | Google Scholar

Moebel E., Kervrann C. (2022). Towards unsupervised classification of macromolecular complexes in cryo electron tomography: Challenges and opportunities. Comput. Methods Programs Biomed. 225, 107017. doi:10.1016/j.cmpb.2022.107017

PubMed Abstract | CrossRef Full Text | Google Scholar

Murata K., Wolf M. (2018). Cryo-electron microscopy for structural analysis of dynamic biological macromolecules. Biochim. Biophys. Acta. Gen. Subj. 1862, 324–334. doi:10.1016/j.bbagen.2017.07.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Newell A., Deng J. (2020). “How useful is self-supervised pretraining for visual tasks?,” in Proceedings of the IEEE/CVF Conference on Computer Vision and Pattern Recognition, 7345–7354.

Google Scholar

Noble A. J., Dandey V. P., Wei H., Brasch J., Chase J., Acharya P., et al. (2018). Routine single particle cryoem sample and grid characterization by tomography. Elife 7, e34257. doi:10.7554/eLife.34257

PubMed Abstract | CrossRef Full Text | Google Scholar

Noroozi M., Favaro P. (2016). “Unsupervised learning of visual representations by solving jigsaw puzzles,” in European Conference on Computer Vision (Springer), 69–84.

CrossRef Full Text | Google Scholar

Oikonomou C. M., Jensen G. J. (2017). Cellular electron cryotomography: Toward structural biology in situ. Annu. Rev. Biochem. 86, 873–896. doi:10.1146/annurev-biochem-061516-044741

PubMed Abstract | CrossRef Full Text | Google Scholar

Oord A. v. d., Li Y., Vinyals O. (2018). Representation learning with contrastive predictive coding. arXiv preprint arXiv:1807.03748

Google Scholar

Pan S. J., Yang Q. (2009). A survey on transfer learning. IEEE Trans. Knowl. Data Eng. 22, 1345–1359. doi:10.1109/tkde.2009.191

CrossRef Full Text | Google Scholar

Pathak D., Krahenbuhl P., Donahue J., Darrell T., Efros A. A. (2016). “Context encoders: Feature learning by inpainting,” in Proceedings of the IEEE conference on computer vision and pattern recognition, 2536–2544.

CrossRef Full Text | Google Scholar

Pei L., Xu M., Frazier Z., Alber F. (2016). Simulating cryo electron tomograms of crowded cell cytoplasm for assessment of automated particle picking. BMC Bioinforma. 17, 405–413. doi:10.1186/s12859-016-1283-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Pérez-García F., Sparks R., Ourselin S. (2021). Torchio: A python library for efficient loading, preprocessing, augmentation and patch-based sampling of medical images in deep learning. Comput. Methods Programs Biomed. 208, 106236doi. doi:10.1016/j.cmpb.2021.106236

CrossRef Full Text | Google Scholar

Pettersen E. F., Goddard T. D., Huang C. C., Couch G. S., Greenblatt D. M., Meng E. C., et al. (2004). Ucsf chimera—A visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612. doi:10.1002/jcc.20084

PubMed Abstract | CrossRef Full Text | Google Scholar

Selvaraju R. R., Cogswell M., Das A., Vedantam R., Parikh D., Batra D. (2017). “Grad-cam: Visual explanations from deep networks via gradient-based localization,” in Proceedings of the IEEE international conference on computer vision, 618–626.

CrossRef Full Text | Google Scholar

Simonyan K., Zisserman A. (2014). Very deep convolutional networks for large-scale image recognition. arXiv preprint arXiv:1409.1556

Google Scholar

Tarvainen A., Valpola H. (2017). “Mean teachers are better role models: Weight-averaged consistency targets improve semi-supervised deep learning results,” in NIPS.

Google Scholar

Tian Y., Krishnan D., Isola P. (2019). “Contrastive multiview coding,” in European conference on computer vision. (Cham: Springer), 776–794.

Google Scholar

Wang X., Huang Q., Celikyilmaz A., Gao J., Shen D., Wang Y.-F., et al. (2019a). “Reinforced cross-modal matching and self-supervised imitation learning for vision-language navigation,” in Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition, 6629–6638.

CrossRef Full Text | Google Scholar

Wang X., Jabri A., Efros A. A. (2019b). “Learning correspondence from the cycle-consistency of time,” in Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition, 2566–2576.

CrossRef Full Text | Google Scholar

Wu Z., Xiong Y., Yu S. X., Lin D. (2018). “Unsupervised feature learning via non-parametric instance discrimination,” in Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition, 3733–3742.

CrossRef Full Text | Google Scholar

Yu L., Li R., Zeng X., Wang H., Jin J., Yang G., et al. (2020). Few shot domain adaptation for in situ macromolecule structural classification in cryo-electron tomograms. Bioinformatics 37, 185–191. doi:10.1093/bioinformatics/btaa671

PubMed Abstract | CrossRef Full Text | Google Scholar

Zeng X., Leung M. R., Zeev-Ben-Mordehai T., Xu M. (2018). A convolutional autoencoder approach for mining features in cellular electron cryo-tomograms and weakly supervised coarse segmentation. J. Struct. Biol. 202, 150–160. doi:10.1016/j.jsb.2017.12.015

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhai X., Oliver A., Kolesnikov A., Beyer L. (2019). “S4l: Self-supervised semi-supervised learning,” in Proceedings of the IEEE/CVF International Conference on Computer Vision, 1476–1485.

CrossRef Full Text | Google Scholar

Zhang P. (2013). Correlative cryo-electron tomography and optical microscopy of cells. Curr. Opin. Struct. Biol. 23, 763–770. doi:10.1016/j.sbi.2013.07.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang R., Isola P., Efros A. A. (2016). “Colorful image colorization,” in European conference on computer vision (Springer), 649–666.

CrossRef Full Text | Google Scholar

Zhu Z., Wang Y., Zhou X., Yang L., Meng G., Zhang Z. (2020). Swav: A web-based visualization browser for sliding window analysis. Sci. Rep. 10, 149–154. doi:10.1038/s41598-019-57038-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: self-supervised learning, macromolecular structure classification, electron cryo tomograms, data augmentation, contrastive learning

Citation: Gupta T, He X, Uddin MR, Zeng X, Zhou A, Zhang J, Freyberg Z and Xu M (2022) Self-supervised learning for macromolecular structure classification based on cryo-electron tomograms. Front. Physiol. 13:957484. doi: 10.3389/fphys.2022.957484

Received: 31 May 2022; Accepted: 02 August 2022;
Published: 30 August 2022.

Edited by:

Renmin Han, Shandong University, China

Reviewed by:

Arnab Bhattacherjee, Jawaharlal Nehru University, India
JJ Fernandez, Spanish National Research Council (CSIC), Spain

Copyright © 2022 Gupta, He, Uddin, Zeng, Zhou, Zhang, Freyberg and Xu. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Min Xu, mxu1@cs.cmu.edu

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.