Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-23T14:57:11.586Z Has data issue: false hasContentIssue false

Signal enhancement for two-dimensional cryo-EM data processing

Published online by Cambridge University Press:  09 March 2023

Guy Sharon
Affiliation:
School of Electrical Engineering, Tel Aviv University, Tel Aviv, Israel
Yoel Shkolnisky
Affiliation:
School of Mathematical Sciences, Tel Aviv University, Tel Aviv, Israel
Tamir Bendory*
Affiliation:
School of Electrical Engineering, Tel Aviv University, Tel Aviv, Israel
*
*Corresponding author. E-mail: bendory@tauex.tau.ac.il

Abstract

Different tasks in the computational pipeline of single-particle cryo-electron microscopy (cryo-EM) require enhancing the quality of the highly noisy raw images. To this end, we develop an efficient algorithm for signal enhancement of cryo-EM images. The enhanced images can be used for a variety of downstream tasks, such as two-dimensional classification, removing uninformative images, constructing ab initio models, generating templates for particle picking, providing a quick assessment of the data set, dimensionality reduction, and symmetry detection. The algorithm includes built-in quality measures to assess its performance and alleviate the risk of model bias. We demonstrate the effectiveness of the proposed algorithm on several experimental data sets. In particular, we show that the quality of the resulting images is high enough to produce ab initio models of $ \sim 10 $ Å resolution. The algorithm is accompanied by a publicly available, documented, and easy-to-use code.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press

Impact Statement

In the past few years, single-particle cryo-electron microscopy (cryo-EM) has become the state-of-the-art method for resolving the atomic structure and dynamics of biological molecules. We design an efficient algorithm to enhance the quality of the highly noisy cryo-EM experimental images. The enhanced images can be used in a wide variety of tasks in the algorithmic pipeline of cryo-EM, including two-dimensional classification, removal of uninformative images, ab initio modeling, dimensionality reduction, symmetry detection, quick assessment of the data set, and as templates for particle picking. We provide a documented Python code. The algorithm includes built-in quality measures to mitigate the risk of model bias.

1. Introduction

In the past few years, single-particle cryo-electron microscopy (cryo-EM) has become the state-of-the-art method for resolving the atomic structure and dynamics of biological molecules(Reference Bai, McMullan and Scheres 1 Reference Murata and Wolf 5 ). A cryo-EM experiment results in a large set of images, each corresponding to a noisy tomographic projection of the molecule of interest, taken from an unknown viewing direction. In addition, the electron doses transmitted by the microscope must be kept low to prevent damage to the radiation-sensitive biological molecules, inducing signal-to-noise ratio (SNR) levels that might be as low as –20 dB (i.e., the power of the noise is 100 times greater than the signal)(Reference Bendory, Bartesaghi and Singer 6 ). The low SNR level is one of the main challenges in processing cryo-EM data sets. In particular, different tasks in the computational pipeline of cryo-EM require enhancing the quality of the highly noisy raw images. Specifically, the high-quality enhanced images can be used as two-dimensional (2D) class averages, to remove uninformative images (e.g., pure noise images, contamination), to construct ab initio models based on the common-lines property(Reference Greenberg and Shkolnisky 7 , Reference Singer and Shkolnisky 8 ),Footnote 1 as templates for particle picking, to provide a quick assessment of the particles, for dimensionality reduction, and for symmetry detection(Reference Bendory, Bartesaghi and Singer 6 , Reference Singer and Sigworth 11 ).

In this paper, we propose a new signal enhancement algorithm that quickly produces multiple enhanced images that represent different viewing directions of the molecule of interest. The algorithm begins by performing steerable principal component analysis (sPCA) that reduces the dimensionality of the data and allows rotating (steering) the images easily(Reference Zhao, Shkolnisky and Singer 12 ). Next, we randomly choose a subset of the images and find their nearest neighbors based on (approximately) rotationally invariant operations. Hereafter, we refer to each image and its neighbors as a class. Then, we apply two stages for refining the classes. We first remove low-quality classes, and then also remove individual images which are inconsistent with their classes. These stages are based on inspecting the spectra of designed matrices, called synchronization matrices. The spectra of these matrices (that is, the distribution of their eigenvalues) provide a built-in quality measure to assess the consistency of each class. This is essential to support downstream tasks, such as ab initio modeling. Finally, we run an expectation–maximization (EM) algorithm for each class independently. This step aligns and averages the remaining images in each class, producing high SNR enhanced images. The different steps of the algorithm are elaborated in Section 2.

Our algorithm is inspired by and builds on the general scheme of(Reference Zhao and Singer 13 ). In particular, the authors of(Reference Zhao and Singer 13 ) suggest finding the rotationally invariant nearest neighbors of each image based on the bispectrum: a third-order rotationally invariant feature(Reference Bendory, Boumal, Ma, Zhao and Singer 14 ). Then, a high-quality image is produced by aligning all neighbors and averaging. While this method works quite well in many cases, the bispectrum inflates the dimensionality of the problem, boosts the noise level (which is already high in typical data sets), and does not offer a systematic way to assess the performance of the algorithm. Section 2 elaborates on the differences between(Reference Zhao and Singer 13 ) and our proposed algorithm.

Our work also shares similarities with 2D classification algorithms, which cluster the particle images and average them to produce high SNR images, dubbed class averages. 2D classification is a standard routine in all contemporary cryo-EM computational pipelines and is mostly used to remove uninformative images that are associated with low-quality class averages and to provide a quick assessment of the particles; our algorithm can be used for those tasks as well. A popular solution to the 2D classification task, implemented in the software RELION, is based on maximizing the posterior distribution of the classes, while marginalizing over the rotations and translations, using an EM algorithm(Reference Scheres 15 ); we describe this methodology in more detail in Section 2.5. A large number of class averages leads, however, to high computational complexity and to low-quality results because only a few images are assigned to each class. In addition, EM tends to suffer from the “rich get richer” phenomenon (also dubbed the attraction problem): most experimental images would correlate well with, and thus be assigned to, the class averages that enjoy higher SNR. As a result, EM tends to output only a few informative classes(Reference Sorzano, Bilbao-Castro and Shkolnisky 16 ). We circumvent this phenomenon since we apply the EM to each class separately.

Another related research thread considers denoising at the image or at the micrograph level. One example of the former is denoising based on Wiener filtering(Reference Bhamre, Zhang and Singer 17 ). A popular micrograph denoising technique is TOPAZ(Reference Bepler, Kelley, Noble and Berger 18 ), which is based on deep learning methods; see also(Reference Li, Zhang, Wan, Yang, Li, Li, Han, Zhu and Zhang 19 , Reference Palovcak, Asarnow, Campbell, Yu and Cheng 20 ). However, these techniques do not harness the similarity between particle images to suppress the noise, and thus their denoising quality is limited. For example, we use the enhanced images (the outputs of our algorithm) to directly construct molecular structures at a low to intermediate resolution. As far as we know, this has not been done using the said methods.

The paper is organized as follows. Section 2 outlines our method. In Section 3, we present results on four experimental data sets. We attain high-quality images that can be used to construct ab initio models with resolutions between $ \sim 10 $ Å to $ \sim 20 $ Å. Section 4 concludes the paper and delineates future work to improve the algorithm.

2. Method

This section describes the main steps of the proposed algorithm. A documented Python code is available at https://github.com/TamirBendory/CryoEMSignalEnhancement.

2.1. Preprocessing

The algorithm begins with a few standard preprocessing steps: we apply phase-flipping to approximately correct the effect of the CTF, down-sample the images to a size of $ 89\times 89 $ pixels, and whiten the noise in the images. The down-sampling has a minor effect on the nearest neighbors search, and is thus used to accelerate the running time of the algorithm. Then, we further reduce the dimension of the images using sPCA, which learns a steerable, data-driven basis for the data set(Reference Zhao, Shkolnisky and Singer 12 ). Under this basis, the $ i $ th image can be approximated by a finite expansion

(2.1) $$ {I}_i\left(\xi, \theta \right)\approx \sum \limits_{k=-{k}_{max}}^{k_{max}}\sum \limits_{q=1}^{q_k}{a}_{k,q}^i{\psi}_c^{k,q}\left(\xi, \theta \right),\hskip1em i=1,\dots, N, $$

where $ N $ is the number of images in the data set, $ \left(\xi, \theta \right) $ are polar coordinates, $ {\psi}_c^{k,q} $ are the sPCA eigenfunctions, $ {a}_{k,q}^i $ are the corresponding coefficients, $ c=1/2 $ is the bandlimit of $ I $ , and $ {k}_{max} $ and $ {q}_k $ are determined as described in(Reference Zhao, Shkolnisky and Singer 12 ). Remarkably, under this representation, an in-plane rotation translates into a phase shift in the expansion coefficients

(2.2) $$ {I}_i(\xi, \theta -\alpha )\approx \sum \limits_{k=-{k}_{max}}^{k_{max}}\sum \limits_{q=1}^{q_k}{a}_{k,q}^i{e}^{-\iota k\alpha }{\psi}_c^{k,q}(\xi, \theta ), $$

where $ \iota =\sqrt{-1} $ is the imaginary unit, and, for real-valued images, a reflection translates into conjugation

(2.3) $$ {I}_i\left(\xi, \pi -\theta \right)\approx \sum \limits_{k=-{k}_{max}}^{k_{max}}\sum \limits_{q=1}^{q_k}\overline{a_{k,q}^i}{\psi}_c^{k,q}\left(\xi, \theta \right). $$

The sPCA dramatically reduces the dimensionality of the images. We use 500 sPCA coefficients to represent the images. Henceforth, with a slight abuse of notation, we refer to the vector of sPCA coefficients of the $ i $ th image $ {\mathbf{a}}_i:= {\left\{{a}_{k,q}^i\right\}}_{k,q} $ as the image.

We mention that the final stage of the algorithm, the EM step, uses the raw images, which are not affected by these preprocessing steps.

2.2. Nearest neighbor search

Next, we randomly choose $ {N}_c $ images $ {I}_{r_1},\dots, {I}_{r_{N_c}} $ from the data set and find the $ K $ nearest neighbors of each image. The underlying assumption is that the nearest neighbors arise from similar viewing directions. We refer to an image and its $ K $ neighbors as a class.

The nearest neighbors search is based on a correlation measure which is approximately invariant under in-plane rotations and reflection. Let $ \Theta $ be a predefined set of $ {N}_{\theta } $ angles; we typically use $ \Theta =\frac{i}{36}\pi, i=0,\dots, 71 $ so that $ {N}_{\theta }=72. $ We define the approximately invariant correlation, between two images $ {\mathbf{a}}_i $ and $ {\mathbf{a}}_j $ , by

(2.4) $$ \underset{\theta \in \Theta}{\max}\left[\max \right(\mathrm{corr}\left({e}^{i\boldsymbol{k}\theta}\cdot {\boldsymbol{a}}_{\boldsymbol{i}},{\boldsymbol{a}}_{\boldsymbol{j}}\right),\mathrm{corr}\left({e}^{i\boldsymbol{k}\theta}\cdot {\boldsymbol{a}}_{\boldsymbol{i}},\overline{{\boldsymbol{a}}_{\boldsymbol{j}}}\right)\Big], $$

where

(2.5) $$ \mathrm{corr}\left(\boldsymbol{u},\boldsymbol{v}\right)=\frac{{\left(\boldsymbol{u}-\overline{\boldsymbol{u}}\right)}^{\ast}\left(\boldsymbol{v}-\overline{\boldsymbol{v}}\right)}{\sigma_{\boldsymbol{u}}{\sigma}_{\boldsymbol{v}}}, $$

and where $ \boldsymbol{k} $ is a radial frequency vector, $ \cdot $ denotes element-wise product, and $ {\sigma}_{\boldsymbol{a}} $ is the standard deviation of a vector $ \boldsymbol{a} $ . The nearest neighbors of the $ i $ -th image are chosen as the $ K $ images with the highest correlation (2.4).

The nearest neighbors search requires computing $ N $ correlations for each of the $ {N}_c $ selected images, resulting in a total of $ {N}_cN $ correlations. For each image, the correlations can be computed using a couple of matrix multiplications using established linear algebra libraries. The computational complexity of this stage is governed by the multiplication of matrices of size $ {N}_{\theta}\times {N}_{\mathrm{co}} $ and $ {N}_{\mathrm{co}}\times N $ , where $ {N}_{\mathrm{co}} $ is the number of sPCA coefficients. For the experiments in Section 3, this stage took less than a minute.

Two comments are in order. First, the approximately invariant correlation can be, in principle, replaced with invariant polynomials called the bispectra, giving rise to analytical rotationally invariant features(Reference Zhao and Singer 13 , Reference Bendory, Boumal, Ma, Zhao and Singer 14 ) or approximately rotationally and translationally invariant features(Reference Bendory, Hadi and Sharon 21 ) However, the dimension of the bispectrum far exceeds the dimension of the image, and thus we preferred to use the more direct expression of (2.4). Second, we choose the $ {N}_c $ images at random in order to cover different viewing directions. In a future work, we hope to replace this random strategy with a deterministic technique that finds a set of images covering all viewing directions.

We next describe a method to rank and remove low-quality classes resulting from our random sampling strategy. This method provides a built-in measure of the quality of the classes, and thus of the enhanced images.

2.3. Sorting the classes

Until now, we have randomly chosen a set of images $ {I}_i,i=1,\dots .{N}_c $ , and found $ K $ nearest neighbors per class. However, since the images $ {I}_i $ were chosen randomly, it is plausible that some of them will be uninformative in the sense that they do not have close neighbors. To discard uninformative images and their classes, we aim to rank the classes according to their quality.

We define a good class as a class where all of its members were taken from a similar viewing direction, up to an in-plane rotation and, possibly, a reflection. For each pair of images in the class, we compute the most likely relative in-plane rotation and reflection; this is a by-product of computing the correlations in (2.4) so no additional computations are required. We denote the estimated relative rotation angle between the $ i $ th and $ j $ th members of the $ k $ th class by $ {\theta}_{i,j}^{(k)} $ . If no reflection is involved, the relative rotation can be represented by a $ 2\times 2 $ rotation matrix

(2.6) $$ {\boldsymbol{R}}_{\boldsymbol{i},\boldsymbol{j}}^{\left(\boldsymbol{k}\right)}=\left[\begin{array}{cc}\cos {\theta}_{i,j}^{(k)}& -\sin {\theta}_{i,j}^{(k)}\\ {}\sin {\theta}_{i,j}^{(k)}& \cos {\theta}_{i,j}^{(k)}\end{array}\right]. $$

If the pair of images are also reflected, then

(2.7) $$ {\boldsymbol{R}}_{\boldsymbol{i},\boldsymbol{j}}^{\left(\boldsymbol{k}\right)}=\left[\begin{array}{cc}\cos {\theta}_{i,j}^{(k)}& -\sin {\theta}_{i,j}^{(k)}\\ {}-\sin {\theta}_{i,j}^{(k)}& -\cos {\theta}_{i,j}^{(k)}\end{array}\right]. $$

We then construct a Hermitian block matrix $ {\boldsymbol{R}}^{\left(\boldsymbol{k}\right)}\in {\mathrm{\mathbb{R}}}^{2K\times 2K} $

(2.8) $$ {\boldsymbol{R}}^{\left(\boldsymbol{k}\right)}=\left[\begin{array}{cccc}{\boldsymbol{R}}_{\mathbf{1},\mathbf{1}}^{\left(\boldsymbol{k}\right)}& {\boldsymbol{R}}_{\mathbf{1},\mathbf{2}}^{\left(\boldsymbol{k}\right)}& \cdots & {\boldsymbol{R}}_{\mathbf{1},\boldsymbol{K}}^{\left(\boldsymbol{k}\right)}\\ {}{\boldsymbol{R}}_{\mathbf{2},\mathbf{1}}^{\left(\boldsymbol{k}\right)}& {\boldsymbol{R}}_{\mathbf{2},\mathbf{2}}^{\left(\boldsymbol{k}\right)}& \cdots & {\boldsymbol{R}}_{\mathbf{2},\boldsymbol{K}}^{\left(\boldsymbol{k}\right)}\\ {}\vdots & \vdots & \ddots & \vdots \\ {}{\boldsymbol{R}}_{\boldsymbol{K},\mathbf{1}}^{\left(\boldsymbol{k}\right)}& {\boldsymbol{R}}_{\boldsymbol{K},\mathbf{2}}^{\left(\boldsymbol{k}\right)}& \cdots & {\boldsymbol{R}}_{\boldsymbol{K},\boldsymbol{K}}^{\left(\boldsymbol{k}\right)}\end{array}\right]. $$

The matrix $ {\boldsymbol{R}}^{\left(\boldsymbol{k}\right)} $ is a synchronization matrix over the dihedral group(Reference Bendory, Edidin, Leeb and Sharon 22 ). If indeed all $ K $ class members are the same image up to an in-plane rotation and, possibly, a reflection (namely, an element of the dihedral group), then $ {\boldsymbol{R}}^{\left(\boldsymbol{k}\right)} $ is of rank two. In other words, only the first two largest eigenvalues $ {\lambda}_1^{(k)},{\lambda}_2^{(k)} $ of $ {\boldsymbol{R}}^{\left(\boldsymbol{k}\right)} $ are nonzero. In practice, since the images were not taken precisely from the same viewing direction, and because of the noise, the matrix is not of rank two. Therefore, as a measure of the quality of this class, it is only natural to compute how “close” is $ {\boldsymbol{R}}^{\left(\boldsymbol{k}\right)} $ to a rank two matrix, namely,

(2.9) $$ {G}^{(k)}=\frac{\lambda_1^{(k)}+{\lambda}_2^{(k)}}{2K}, $$

where we note that $ {\sum}_i{\lambda}_i^{(k)}=\mathrm{Tr}\left({\boldsymbol{R}}^{\left(\boldsymbol{k}\right)}\right)=2K $ . We refer to $ {G}^{(k)}\in \left[0,1\right] $ as the grade of the $ k $ th class. We repeat this procedure for each class, and remove the classes with the lowest grades. In practice, we found that removing half of the classes yields good results. Since we need to extract the two leading eigenvalues of $ {N}_c $ synchronization matrices, the typical computational complexity of this stage is $ O\left({K}^2{N}_c\right) $ .

The empirical distributions of the eigenvalues, or the grades $ {G}^{(k)} $ , provide a measure to assess the performance of the algorithm. This is important since the output of the signal enhancement algorithm may be used in downstream procedures, for example, to construct ab initio models. Thus, producing poor output may bias the entire computational pipeline with unpredictable consequences(Reference Henderson 23 ).

2.4. Sorting images within classes

After removing classes of low quality, we wish to improve each of the remaining classes by removing inconsistent images. We follow the same strategy as before, and look for, within each class, a subset of images that are consistent with each other, namely, that form an approximately rank-two synchronization matrix (2.8).

Let $ {\hat{\mathbf{R}}}^{(k)}={\mathbf{V}}^{(k)}{\left({\mathbf{V}}^{(k)}\right)}^{\ast } $ be the best rank-two approximation of $ {\mathbf{R}}^{(k)} $ , where the columns of $ {\mathbf{V}}^{(k)}\in {\mathrm{\mathbb{R}}}^{2K\times 2} $ are the eigenvectors of $ {\mathbf{R}}^{(k)} $ associated with the two leading eigenvalues. Let $ {\hat{\mathbf{R}}}^{(k)}\left[i,j\right] $ and $ {\mathbf{R}}^{(k)}\left[i,j\right] $ be the $ \left(i,j\right) $ th entries of $ {\hat{\mathbf{R}}}^{(k)} $ and $ {\mathbf{R}}^{(k)} $ , respectively. To determine whether the $ i $ th class member is consistent with its other class members, we compute the average distance between $ {\hat{\mathbf{R}}}^{(k)}\left[i,j\right] $ and $ {\mathbf{R}}^{(k)}\left[i,j\right] $ for all $ j=1,\dots, K $ . Namely, the grade of the $ i $ th member of the $ k $ -th class is defined by

(2.10) $$ {g}_i^{(k)}=-\frac{1}{K}\sum \limits_{j=1}^K{\left\Vert {\hat{\mathbf{R}}}^{(k)}\left[i,j\right]-{\mathbf{R}}^{(k)}\Big[i,j\Big]\right\Vert}_2. $$

We found that producing classes with 300 images, and removing 150 images with the lowest score $ {g}_i^{(k)} $ (per class) yields good results.

2.5. Expectation–maximization

After pruning out low-quality classes, and inconsistent images within each class, we are ready for the last stage of our algorithm: aligning the images within each class and averaging them to produce a high SNR output image. To this end, we apply the EM algorithm that aims to maximize the likelihood function of the observed images. The EM algorithm is applied to the raw images corresponding to each class separately (before down-sampling, phase-flipping, and so forth.)

The EM algorithm assumes that all observed images are rotated, translated, and noisy versions of a single image; this image is denoted by $ X $ and corresponds to the high SNR image we wish to estimate. The generative model of the images within a specific class is given by

(2.11) $$ {I}_i={L}_{t_i}X+{\varepsilon}_i,\hskip1em i=1,\dots, K, $$

where $ {t}_i $ encodes the unknown rotation and translation of the $ i $ th image, $ {L}_t $ is a linear rotation and translation operator (may also include the CTF), and $ {\varepsilon}_i $ is an i.i.d. Gaussian noise with variance $ {\sigma}^2 $ . Our goal is to maximize the marginalized log-likelihood, which is equal, up to a constant, to

(2.12) $$ \log \hskip0.4em p\left({I}_1,\dots, {I}_K;X\right)=\sum \limits_{i=1}^K\log \sum \limits_{t_{\mathrm{\ell}}\in \mathcal{T}}p\left({t}_{\mathrm{\ell}}\right){e}^{-\frac{1}{2{\sigma}^2}\parallel {I}_i-{L}_{t_{\mathrm{\ell}}}X\parallel }, $$

where $ \mathcal{T} $ denotes the set of possible rotations and translations. While optimizing (2.12) is a challenging non-convex problem, EM has been proven to be an effective technique for optimizing (2.12) for cryo-EM images(Reference Bendory, Bartesaghi and Singer 6 , Reference Sigworth 24 ). In particular, we used the implementation of RELION(Reference Scheres 15 ). We note that this implementation does not correct for reflections, and thus we correct for reflections before running the EM algorithm. In particular, we construct a symmetric matrix whose $ \left(i,j\right) $ entry is −1 if the $ i $ -th and $ j $ -th images are likely to be reflected, and 1 if not. This is a by-product of previous steps, described in Section 2.3, so no additional computations are required. Since this matrix is ideally a rank-one matrix (if all pair-wise estimations are consistent), we extract its leading eigenvector and round its entries into $ \pm 1 $ . This algorithm is known as the spectral algorithm for group synchronization(Reference Bendory, Edidin, Leeb and Sharon 22 , Reference Singer 25 ).

As mentioned in the introduction, a popular solution to the 2D classification task (which shares similarities with the signal enhancement problem) is to run EM on the observed images, before clustering the images. In this case, the generative model reads

(2.13) $$ {I}_i={L}_{t_i}{X}_{m_i}+{\varepsilon}_i,\hskip1em i=1,\dots, N, $$

where $ {X}_1,\dots, {X}_{N_c} $ are the class averages to be estimated. While the implementation of the EM algorithm for (2.13) follows the same lines as the EM algorithm we use, it tends to output only a few informative classes because of the “rich get richer” phenomenon(Reference Sorzano, Bilbao-Castro and Shkolnisky 16 ). We evade this pitfall by first finding the nearest neighbors of the chosen images, and running EM on each class separately to optimize (2.12). In addition, since we run multiple independent instances of the EM algorithm, they can be run in parallel, resulting in a significant acceleration.

We mention that the EM algorithm can be replaced by alternative computational strategies such as stochastic gradient descent or rotationally and translationally aligning the images, and then averaging them. The latter strategy, used by(Reference Zhao and Singer 13 ), is much faster than EM, and thus will significantly accelerate the algorithm, at the cost of lower image quality.

3. Experimental Results

In the following experiments, we produced 3000 classes and kept the best $ {N}_c=1500 $ classes according to the method explained in Section 2.3. Each class consists of 300 images, from which only the best $ K=150 $ images were used to estimate the class average, as explained in Section 2.4. We used the EM implementation of RELION(Reference Scheres 15 ) with seven iterations. Based on the class averages, we reconstructed ab initio models using the common-lines method implemented in the ASPIRE package( 26 ). All data sets were processed using an Intel(R) Xeon(R) Gold 6252 CPU @ 2.10 GHz containing 24 cores, and a GeForce RTX 2080 Ti GPU. The run times of all stages in the process are provided in Table 1. The resolution was computed based on the Fourier Shell Correlation (FSC) criterion with cutoff of 0.5, where the reference volume was downloaded from EMDB(Reference Lawson, Patwardhan and Baker 27 ).

Table 1. Runtime.

Abbreviation: EM, expectation–maximization; sPCA, steerable principal component analysis.

3.1. EMPIAR 10028

We begin with a data set of the Plasmodium falciparum 80S ribosome bound to the anti-protozoan drug emetine, available as the entry 10028 in EMPIAR(Reference Iudin, Korir, Salavert-Torres, Kleywegt and Patwardhan 28 ) (the corresponding entry in EMDB is EMD-2660)(Reference Wong, Bai and Brown 29 ). This data set contains 105,247 images of size $ 360\times 360 $ pixels.

Figure 1 shows examples of raw data images, the corresponding class averages, and a 3D structure reconstructed using the class averages; the resolution of the reconstructed structure is 10.41 Å. The nearest neighbor’s stage took 2.5 min, and the overall process took around 75 min.

Figure 1. EMPIAR 10028. The resolution of the reconstructed structure is 10.41 Å.

3.2. EMPIAR 10073

This data set of the yeast spliceosomal U4/U6.U5 tri-snRNP is available as the entry 10073 of EMPIAR (the corresponding entry in EMDB is EMD-8012)(Reference Nguyen, Galej and Bai 30 ). This data set contains 138840 images of size $ 380\times 380 $ pixels. The nearest neighbors search took 2.5 min and producing 1500 class averages took roughly 100 min. The results are presented in Figure 2. The resolution of the reconstructed structure is 19.58 Å.

Figure 2. EMPIAR 10073. The resolution of the reconstructed structure is 19.58 Å.

3.3. EMPIAR 10081

This data set of the human HCN1 hyperpolarization-activated cyclic nucleotide-gated ion channel is available as the entry 10081 of EMPIAR (the corresponding entry in EMDB is EMD-8511)(Reference Lee and MacKinnon 31 ). This data set contains 55870 images of size $ 256\times 256 $ pixels. The nearest neighbors search took less than 3 min and producing 1500 high-quality images took roughly 1 hr. Figure 3 shows the results. The resolution of the reconstructed structure is 11.25 Å.

Figure 3. EMPIAR 10081. The resolution of the reconstructed structure is 11.25 Å.

3.4. EMPIAR 10061

This data set of the beta-galactosidase in complex with a cell-permeant inhibitor is available as the entry 10061 of EMPIAR (the corresponding entry in EMDB is EMD-2984)(Reference Bartesaghi, Merk and Banerjee 32 ). This data set contains 41123 images of size $ 768\times 768 $ pixels. The nearest neighbors search took less than 3 min and producing 1500 class averages took roughly 2.5 hr. Figure 4 shows results, where the 3D structure was reconstructed from the enhanced images using the spectral algorithm implemented in ASPIRE(Reference Rosen and Shkolnisky 33 ). Although the class averages look of good quality, the resolution of the reconstructed structure is only 22.63 Å.

Figure 4. EMPIAR 10061. The resolution of the reconstructed structure is 22.63 Å.

4. Discussion

In this paper, we have presented a new algorithm to enhance the quality of cryo-EM images, which can be used for various tasks in the computational pipeline of cryo-EM. The algorithm is based on(Reference Zhao and Singer 13 ), but extends it in several ways, which are crucial to improve its performance and to design built-in quality measures. The algorithm is computationally efficient and can be executed on large experimental data sets.

For larger data sets, the brute-force nearest neighbors search we use can be replaced by efficient randomized algorithms(Reference Jones, Osipov and Rokhlin 34 ), resulting in a better asymptotic computational complexity. However, for contemporary data sets, the running times of both approaches are comparable. Our classification is approximately invariant under in-plane rotations and reflections; see also(Reference Cahill, Iverson, Mixon and Packer 35 ) for a related approach. While it can be extended to translation invariance by explicitly considering different translations, it will significantly increase the running time. A possible alternative approach would be to employ polynomials that are approximately invariant under the group of in-plane rotations and translations (namely, the group of rigid motions SE(Reference Bendory, Hadi and Sharon 21 ).

To improve the quality of the nearest neighbors search, the classes are refined by analyzing the spectra of synchronization matrices. This provides a validation measure that can be computed directly from the data, which is crucial to mitigate the risk of model bias in downstream tasks, such as ab initio modeling. We have demonstrated that the enhanced images are of high quality so that they can be used to construct ab initio models. To cover all viewing angles, we randomly sample the data set. This is clearly not optimal, and we intend to study different deterministic strategies to sample the data. A successful sampling strategy may make the class sorting stage of Section 2.3 unnecessary.

Competing interests

The authors declare no competing interests exists.

Authorship contributions

Guy Sharon, Yoel Shkolnisky, and Tamir Bendory conceptualized the study, developed the models and algorithms, and wrote the software code and the manuscript.

Funding statement

T.B. is partially supported by the NSF-BSF award 2019752, the BSF grant no. 2020159, and the ISF grant no. 1924/21. Y.S. was supported by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (grant agreement 723991 - CRYOMATH) and by the NIH/NIGMS Award R01GM136780–01.

Data availability statement

Source code and user manual are available at https://github.com/TamirBendory/CryoEMSignalEnhancement.

Footnotes

1 We mention that there are many alternative algorithms for constructing ab initio models, which do not require enhanced images, such as stochastic gradient descentReference Punjani, Rubinstein, Fleet and Brubaker 9 and the method of momentsReference Levin, Bendory, Boumal, Kileel and Singer 10.

References

Bai, X-C, McMullan, G & Scheres, SHW (2015) How cryo-EM is revolutionizing structural biology. Trends Biochem Sci 40(1), 4957.CrossRefGoogle ScholarPubMed
Frank, J (2006) Three-Dimensional Electron Microscopy of Macromolecular Assemblies: Visualization of Biological Molecules in their Native State. Oxford: Oxford University Press.CrossRefGoogle Scholar
Kühlbrandt, W (2014) The resolution revolution. Science 343(6178), 14431444.CrossRefGoogle ScholarPubMed
Lyumkis, D (2019) Challenges and opportunities in cryo-EM single-particle analysis. J Biol Chem 294(13), 51815197.CrossRefGoogle ScholarPubMed
Murata, K & Wolf, M (2018) Cryo-electron microscopy for structural analysis of dynamic biological macromolecules. Biochim Biophys Acta - Gen Subj 1862(2), 324334.CrossRefGoogle ScholarPubMed
Bendory, T, Bartesaghi, A & Singer, A (2020) Single-particle cryo-electron microscopy: mathematical theory, computational challenges, and opportunities. IEEE Signal Process Mag 37(2), 5876.CrossRefGoogle ScholarPubMed
Greenberg, I & Shkolnisky, Y (2017) Common lines modeling for reference free ab-initio reconstruction in cryo-EM. J Struct Biol 200(2), 106117.CrossRefGoogle ScholarPubMed
Singer, A & Shkolnisky, Y (2011) Three-dimensional structure determination from common lines in cryo-EM by eigenvectors and semidefinite programming. SIAM J Imaging Sci 4(2), 543572.CrossRefGoogle ScholarPubMed
Punjani, A, Rubinstein, JL, Fleet, DJ & Brubaker, MA (2017) cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination. Nat Methods 14(3), 290296.CrossRefGoogle ScholarPubMed
Levin, E, Bendory, T, Boumal, N, Kileel, J & Singer, A (2018) 3D ab initio modeling in cryo-EM by autocorrelation analysis. In 2018 IEEE 15th International Symposium on Biomedical Imaging (ISBI 2018), pp. 15691573. IEEE.CrossRefGoogle Scholar
Singer, A & Sigworth, FJ (2020) Computational methods for single-particle electron cryomicroscopy. Annu Rev Biomed Eng 3, 163.Google ScholarPubMed
Zhao, Z, Shkolnisky, Y & Singer, A (2016) Fast steerable principal component analysis. IEEE Trans Comput Imaging 2(1), 112.CrossRefGoogle ScholarPubMed
Zhao, Z & Singer, A (2014) Rotationally invariant image representation for viewing direction classification in cryo-EM. J Struct Biol 186(1), 153166.CrossRefGoogle ScholarPubMed
Bendory, T, Boumal, N, Ma, C, Zhao, Z & Singer, A (2017) Bispectrum inversion with application to multireference alignment. IEEE Trans Signal Process 66(4), 10371050.CrossRefGoogle ScholarPubMed
Scheres, SHW (2012) RELION: implementation of a Bayesian approach to cryo-EM structure determination. J Struct Biol 180(3), 519530.CrossRefGoogle ScholarPubMed
Sorzano, COS, Bilbao-Castro, JR, Shkolnisky, Y, et al. (2010) A clustering approach to multireference alignment of single-particle projections in electron microscopy. J Struct Biol 171(2), 197206.CrossRefGoogle ScholarPubMed
Bhamre, T, Zhang, T & Singer, A (2016) Denoising and covariance estimation of single particle cryo-EM images. J Struct Biol 195(1), 7281.CrossRefGoogle ScholarPubMed
Bepler, T, Kelley, K, Noble, AJ & Berger, B (2020) Topaz-denoise: general deep denoising models for cryoEM and cryoET. Nat Commun 11(1), 112.CrossRefGoogle ScholarPubMed
Li, H, Zhang, H, Wan, X, Yang, Z, Li, C, Li, J, Han, R, Zhu, P, & Zhang, F (2022) Noise-Transfer2Clean: denoising cryo-EM images based on noise modeling and transfer. Bioinformatics 38(7), 20222029.CrossRefGoogle ScholarPubMed
Palovcak, E, Asarnow, D, Campbell, MG, Yu, Z & Cheng, Y (2020) Enhancing the signal-to-noise ratio and generating contrast for cryo-EM images with convolutional neural networks. IUCrJ 7(6), 11421150.CrossRefGoogle ScholarPubMed
Bendory, T, Hadi, I & Sharon, N (2022) Compactification of the rigid motions group in image processing. SIAM J Imaging Sci 15(3), 10411078.CrossRefGoogle Scholar
Bendory, T, Edidin, D, Leeb, W & Sharon, N (2022) Dihedral multi-reference alignment. IEEE Trans Inf Theory 68(5), 34893499.CrossRefGoogle Scholar
Henderson, R (2013) Avoiding the pitfalls of single particle cryo-electron microscopy: Einstein from noise. Proc Natl Acad Sci 110(45), 1803718041.CrossRefGoogle ScholarPubMed
Sigworth, FJ (1998) A maximum-likelihood approach to single-particle image refinement. J Struct Biol 122(3), 328339.CrossRefGoogle ScholarPubMed
Singer, A (2011) Angular synchronization by eigenvectors and semidefinite programming. Appl Comput Harmon Anal 30(1), 2036.CrossRefGoogle ScholarPubMed
ASPIRE: algorithms for single particle reconstruction software package. http://spr.math.princeton.edu/.Google Scholar
Lawson, CL, Patwardhan, A, Baker, ML, et al. (2016) EMDataBank unified data resource for 3DEM. Nucl Acid Res 44(D1), D396D403.CrossRefGoogle ScholarPubMed
Iudin, A, Korir, PK, Salavert-Torres, J, Kleywegt, GJ & Patwardhan, A (2016) EMPIAR: a public archive for raw electron microscopy image data. Nat Methods 13(5), 387388.CrossRefGoogle ScholarPubMed
Wong, W, Bai, X-c, Brown, A, et al. (2014) Cryo-EM structure of the plasmodium falciparum 80S ribosome bound to the anti-protozoan drug emetine. eLife 3, e03080.CrossRefGoogle Scholar
Nguyen, THD, Galej, WP, Bai, X-c, et al. (2016) Cryo-EM structure of the yeast U4/U6. U5 tri-snRNP at 3.7 Å resolution. Nature 530(7590), 298302.CrossRefGoogle ScholarPubMed
Lee, C-H & MacKinnon, R (2017) Structures of the human HCN1 hyperpolarization-activated channel. Cell 168(1–2), 111120.CrossRefGoogle ScholarPubMed
Bartesaghi, A, Merk, A, Banerjee, S, et al. (2015) 2.2 Å resolution cryo-EM structure of β-galactosidase in complex with a cell-permeant inhibitor. Science 348(6239), 11471151.CrossRefGoogle ScholarPubMed
Rosen, E & Shkolnisky, Y (2020) Common lines ab initio reconstruction of D_2-symmetric molecules in cryo-electron microscopy. SIAM J Imaging Sci 13(4), 18981944.CrossRefGoogle Scholar
Jones, PW, Osipov, A & Rokhlin, V (2011) Randomized approximate nearest neighbors algorithm. Proc Natl Acad Sci 108(38), 1567915686.CrossRefGoogle ScholarPubMed
Cahill, J, Iverson, JW, Mixon, DG & Packer, D (2022) Group-invariant max filtering. Preprint, arXiv:2205.14039.Google Scholar
Figure 0

Table 1. Runtime.

Figure 1

Figure 1. EMPIAR 10028. The resolution of the reconstructed structure is 10.41 Å.

Figure 2

Figure 2. EMPIAR 10073. The resolution of the reconstructed structure is 19.58 Å.

Figure 3

Figure 3. EMPIAR 10081. The resolution of the reconstructed structure is 11.25 Å.

Figure 4

Figure 4. EMPIAR 10061. The resolution of the reconstructed structure is 22.63 Å.