Skip to main content

This is a preprint.

It has not yet been peer reviewed by a journal.

The National Library of Medicine is running a pilot to include preprints that result from research funded by NIH in PMC and PubMed.

bioRxiv logoLink to bioRxiv
[Preprint]. 2023 Jan 25:2023.01.24.525447. [Version 1] doi: 10.1101/2023.01.24.525447

Investigating the Complexity of Gene Co-expression Estimation for Single-cell Data

Jiaqi Zhang 1,*, Ritambhara Singh 2
PMCID: PMC9900775  PMID: 36747724

Abstract

With the rapid advance of single-cell RNA sequencing (scRNA-seq) technology, understanding biological processes at a more refined single-cell level is becoming possible. Gene co-expression estimation is an essential step in this direction. It can annotate functionalities of unknown genes or construct the basis of gene regulatory network inference. This study thoroughly tests the existing gene co-expression estimation methods on simulation datasets with known ground truth co-expression networks. We generate these novel datasets using two simulation processes that use the parameters learned from the experimental data. We demonstrate that these simulations better capture the underlying properties of the real-world single-cell datasets than previously tested simulations for the task. Our performance results on tens of simulated and eight experimental datasets show that all methods produce estimations with a high false discovery rate potentially caused by high-sparsity levels in the data. Finally, we find that commonly used pre-processing approaches, such as normalization and imputation, do not improve the co-expression estimation. Overall, our benchmark setup contributes to the co-expression estimator development, and our study provides valuable insights for the community of single-cell data analyses.

Keywords: gene co-expression, scRNA-seq

1. Introduction

Analyzing the co-variations of gene expression across samples, also known as gene co-expression analysis, is essential for understanding complicated biological processes. For example, gene co-expression analysis can prioritize candidate disease genes (Presson et al., 2008; Van Dam et al., 2018) and molecules of disparate cellular states (Carter et al., 2004), and annotate functionalities of unknown genes (Serin et al., 2016). In previous studies on microarray or bulk RNA-seq data, gene co-expression estimation has shown superior performance for genetic module discovery (Stuart et al., 2003), gene-disease prediction (Van Dam et al., 2018; Langfelder and Horvath, 2008), and biological-processes-related genes detection (Usadel et al., 2009; Aoki et al., 2007). Recently, advances in the high throughput single-cell RNA sequencing (scRNA-seq) technology have enabled researchers to profile gene expression at a single-cell resolution (Hwang et al., 2018; Chen et al., 2019). scRNA-seq datasets allow us to investigate the gene expression of individual cells, revealing heterogeneity among cell populations. These datasets are large, generally consisting of expression for tens of thousands of cells and genes (Macosko et al., 2015; Picelli et al., 2014), thus allowing population-based gene co-expression estimation. However, these new single-cell data opportunities also present new challenges for statistical analysis (Lähnemann et al., 2020). For example, it is notoriously difficult to deal with high sparsity and heterogeneity in scRNA-seq data. Therefore, efficient and robust gene co-expression estimation models for scRNA-seq data are urgently needed in the field.

We formulate the gene co-expression estimation on scRNA-seq data as an unsupervised undirected graph structure learning task (Fig. 1). Concretely, given a tissue/organ with heterogeneous cells, scRNA-seq technique first separates it into individual cells and sequences gene expression at a single-cell level (Haque et al., 2017). The scRNA-seq data can be formalized into an expression matrix Xn×p of n cells and p genes 2, where each element Xij is a count value denoting the expression level of gene j at cell i. Given the expression matrix, we say a pair of genes are co-expressed if the genes have similar expression patterns. So the goal of gene co-expression estimation is to estimate a symmetric gene similarity matrix Wp×p with Wjk indicating the similarity score between genes j and k. We can convert the similarity matrix to an adjacency matrix G{0,1}p×p, encoding a co-expression network with nodes denoting genes and undirected edges connecting two nodes i and j if the corresponding pair of genes are significantly co-expressed, i.e., the similarity score Wij higher than a user-defined threshold. Estimating the undirected co-expression network is also a fundamental step in directed gene regulatory network (GRN) inference (Nguyen et al., 2021), in which methods can perform enrichment to determine transcript factors and target genes on the undirected network and assign directions to edges. This is suggested by the fact that co-expressed genes are likely to be involved in similar biological functions and share transcription factor binding sites (Marco et al., 2009; Kang et al., 2011; Oldham et al., 2008; Allocco et al., 2004; Yeung et al., 2004; Yin et al., 2021).

Figure 1:

Figure 1:

Illustration of gene co-expression estimation on scRNA-seq data. scRNA-seq technique records gene expression of heterogenous cells that are isolated from tissues. The scRNA-seq data is represented as a cell-by-gene matrix, with each matrix element a count value denoting the gene’s expression level at a cell. The expression matrix is then used to estimate gene co-expression, which encodes a graph structure with nodes indicating genes and edges denoting co-expression.

Previous methods have proposed various measurements for computing the gene similarity score and constructing the co-expression network. They generally fall into three categories: (1) The first group evaluates the co-expression between genes by the correlation coefficients of gene expressions. For example, (Langfelder and Horvath, 2008) proposed Weighted Correlation Network Analysis (WGCNA) to construct gene co-expression networks through weighted correlation networks. (2) The second group utilizes Mutual Information (MI) as the co-expression measurement. For example, (Butte and Kohane, 1999) estimated MI via histogram or binning technique. (Margolin et al., 2006) computed MI with differential entropy that averaged the log probability of gene marginal distributions. (Faith et al., 2007) estimated the likelihood of MI instead of calculating significance scores as in the previous works. (Meyer et al., 2007) decomposed the network inference into a series of MI-based supervised gene selection procedures. (3) Methods in the last group are based on structural models. For example, (Yin and Li, 2011; Danaher et al., 2014; Taeb et al., 2020) assumed the raw counts or log-transformed gene expression data followed Gaussian(-like) distributions and adopted Gaussian Graphical Model (GMM) (Friedman et al., 2008) to estimate the partial correlation between genes. GENIE3 (Huynh-Thu et al., 2010) decomposed the task into several regression subproblems and solved them with tree-based models. Also, (Allen and Liu, 2013) estimated gene co-expression with local Poisson graphical models.

However, these methods were mainly proposed for microarray and bulk RNA-seq data. The excess zero counts and heterogeneity in scRNA-seq data present several challenges for the gene co-expression estimation task. Some recent works (Iacono et al., 2019; Cha and Lee, 2020; Wu et al., 2021) have tried to directly apply correlation-based methods on scRNA-seq data. But the correlation coefficient is not a robust measurement for very-sparse scRNA-seq data (Sanchez-Taltavull et al., 2020). Moreover, the Gaussian or Poisson data distribution assumptions in these works may be incorrect, because recent studies (Risso et al., 2018; Townes et al., 2019; Svensson, 2020) have shown that scRNA-seq data follow negative binomial or zero-inflated distributions. To enable co-expression estimation for single-cell data, (Chan et al., 2017) proposed Partial Information Decomposition and Context (PIDC), which used partial information decomposition to explore statistical dependencies between genes in scRNA-seq data. scLink (Li and Li, 2021) revised GGM learning by explicitly modeling zeros in scRNA-seq data. Similarly, (Park et al., 2021; Choi et al., 2017) modeled excess zeros in data with (zero-inflated) negative binomial distributions. But in these works, the simulated datasets used for method evaluations are less sparse than the real-world single-cell experimental data. Two simulated datasets in (Li and Li, 2021) and (Choi et al., 2017), have around 50% and 20% zero counts, respectively, but the experimental scRNA-seq data can have more than 80% of zeros (Hicks et al., 2018; Qiu, 2020). Although these methods showed their considerable ability to recover gene co-expression networks on simulated data, the simulated datasets used in these studies did not exhibit the high sparsity level and overdispersed distribution of experimental scRNA-seq data. Evaluations on such simulations conceal the complexity of estimating co-expression networks from sparse single-cell data. Therefore, we need simulations that accurately capture real-world experimental datasets’ underlying properties to build robust methods for the co-expression estimation task.

In this paper, we test state-of-the-art gene co-expression estimation methods on two novel simulation processes, revealing the complexity of co-expression estimation for single-cell data and suggesting potential directions for improvement. First, we apply two novel simulation processes utilizing information learned from real-world single-cell data (Sec. 3.2) to produce more realistic benchmarking schemes. These schemes will assist in more robust method developments for the gene co-expression estimation task. Next, we perform a comprehensive analysis of the existing co-expression estimation methods. We benchmark nine state-of-the-art and representative gene co-expression estimation methods on ten simulated datasets with different settings and eight experimental datasets covering mouse and human cells using four different experiment protocols (Sec. 3.3 and 3.4). Our results show that existing methods give poor gene co-expression estimations and tend to have high false positive rates, thus, suggesting that robust estimation methods will need to focus on lowering the false positive rates. We also find, as suspected, that high sparsity in scRNA-seq data is the main bottleneck of co-expression estimation (Sec. 3.5). However, our implementation of several scRNA-seq data pre-processing approaches, including normalization (Stuart et al., 2019; Wolf et al., 2018; Choudhary and Satija, 2022; Borella et al., 2022), pseudo-bulk estimation (Squair et al., 2021), and imputation (Hou et al., 2020), show that while they reduce the sparsity and variance of the data, these pre-processing steps worsen the gene co-expression estimation by inducing biases (Sec. 3.6), for example, the imputation inducing some artificial correlations (Andrews and Hemberg, 2018; Hou et al., 2020). Therefore, we suggest that future works carefully choose pre-processing approaches when performing downstream estimation tasks from single-cell datasets. Previous works (Song et al., 2012; Ovens et al., 2021; Greenfield et al., 2010) tested several co-expression estimators, but they used microarray and bulk RNA-seq data. (Pratapa et al., 2020; Dibaeinia and Sinha, 2020) used scRNA-seq data in their benchmarks and found existing methods are largely inaccurate, but they focused on GRN estimation methods and did not examine the effect of different pre-processing steps. But our study gives a comprehensive investigation of co-expression estimations and is a novel supplement and extension of the earlier studies. Overall, as gene co-expression estimation is a fundamental step for downstream single-cell data analysis, our study provides valuable insights for this community that incorporate this step into their research.

2. Methods

Notations: We let bold lowercase letters denote vectors, and bold uppercase letters denote matrices. For a matrix Am×n, its i-th row and j-th column are denoted by Ai: and A:j respectively. We let AF=i,jAij2 denote the Frobenius norm of matrix A. Given a vector an, We let a2=iai2 denote its 2 norm.

We conduct a series of experiments to compare gene co-expression estimation methods on both simulated and real-world experimental scRNA-seq datasets. We first show that previously used simulations do not exhibit properties of real-world experimental data and use two recent single-cell read count simulation algorithms to generate realistic single-cell datasets (Sec. 3.1 and 3.2). Next, we compare nine bulk and single-cell gene co-expression estimators on our simulated and real-world datasets (Sec. 3.3 and 3.4). Finally, we investigate whether commonly used pre-processing approaches for scRNA-seq data could help extract robust gene co-expression networks (Sec. 3.6). We briefly introduce our experimental settings in this section (with details in the Appendix: More on Method and Results section), and Fig. 2 gives an overview of our setup.

Figure 2:

Figure 2:

The summary of our study. Part I: We compare two previous and two new simulated processes with eight experimental data of multiple species and protocols in Sec. 3.1 and 3.2. We compare cell/gene sparsity and distribution shape qualitatively using plots and quantitatively using two-sample Kolmogorov–Smirnov tests (KS test) and integrated errors. Part II: We benchmark nine gene co-expression estimators, including both bulk and single-cell models, on simulated and experimental data in Sec. 3.3 and 3.4. We test three graph structures for simulation data: hub-based, scale-free, and gene regulatory network (GRN). We take known transcription factor (TF) co-expression networks for experimental data as reference networks. We report F1-score, FDR, AUROC, and AUPRC for method performances. Part III: We test multiple pre-processing approaches, including normalization, pseudo-bulk, and imputation on scRNA-seq data in Sec. 3.6.

2.1. Data Preparation

2.1.1. Experimental Data and Pre-Processing Steps

We use eight publicly available scRNA-seq datasets in this study. We extract data of four different sequencing protocols from mouse cortex data (Ding et al., 2019) and human peripheral blood mononuclear cells (PBMC) datasets (Ding et al., 2019) downloaded from Single Cell Protal3. We filter out cells and genes with zero counts from each dataset’s raw unique molecular identifier (UMI) count matrix. The basic statistics of these eight datasets are summarized in Table 1. These datasets cover various protocols, cell library sizes, sparsity, and species. In our study, we select the top 100 and 1000 most highly variable genes (HVG) of every experimental dataset, and we conduct experiments on these subsets of data and corresponding simulations. The sparsity of 100 HVGs of each dataset are provided in Appendix Table A2.

Table 1:

Basic statistics of eight experimental datasets.

ID Species Experiment Protocol # Cells # Genes Sparsity (ratio of non-zeros) Max Count Description
#1 Mouse Cortex 1 10x Genomics 1453 21887 10.95% 4568 Mouse cortex data.
#2 Mouse Cortex 1 Smart Seq2 295 24114 22.94% 108522
#3 Mouse Cortex 2 10x Genomics 4267 22805 7.58% 3268
#4 Mouse Cortex 2 Smart Seq2 349 23974 20.48% 59321

#5 Human PBMC 1 Drop Seq 10824 22792 2.19% 316 Human PBMC.
#6 Human PBMC 1 inDrops 6184 18964 1.92% 110
#7 Human PBMC 2 Drop Seq 11578 24634 3.29% 2623
#8 Human PBMC 2 inDrops 5166 19648 3.19% 553

2.1.2. Gene Co-expression Reference Network Generation

To validate the performance of gene co-expression estimators, we need to generate gene co-expression networks to serve as the ground truth and simulate gene expression data based on it. In literature, hub genes (i.e., genes with high connectivity) are seen as highly essential and considered to play important roles in developmental processes and the regulation of transcription (Tong et al., 2004; Liu et al., 2019; Seo et al., 2009). Therefore, we generated hub-based networks as true co-expression network structures, denoted as G{1,0,1}p×p with 1 and −1 indicating positive and negative edge respectively, in all simulations. We use pcorSimulator R package to generate networks. Fig. 8 gives an example of the hub-based network. For each network G*, we obtain the network weight matrix Wp×p through the following steps:

Wij=Uniform(0.5,0.1),if Gij=1Uniform(0.1,0.5),if Gij=10,otherwise. (1)

allowing us to use W* in data simulation (see Sec. 2.1.3).

Figure 8:

Figure 8:

Examples of three gene co-expression network topologies used in simulations: hub-based network, scale-free network, and GRN.

To define the ground truth networks for real-world experimental datasets (Sec. 3.4), we use known gene-gene interaction networks as references of gene co-expression (i.e., W*). We focus on co-expression related to known transcription factors (TFs). We take the DNA-binding TF co-expression network presented in Zhou et al. (2017) as the reference for mouse datasets. Because it is constructed based on Pearson correlation, we keep edges with their absolute Pearson correlation coefficients higher than 0.2 to remove lowly co-expressed edges. For human datasets, we use the gene co-expression discovered in ChIP-seq experiments Rouillard et al. (2016). We use the gene cosine similarity matrix constructed from Pathway Common Rodchenkov et al. (2020) dataset, and select the sub-network related to human TFs Lambert et al. (2018). We keep the edges with their absolute cosine similarity higher than 0.2. The average human/mouse TF co-expression network sparsity (i.e., the number of edges divided by the number of possible gene pairs) is 11%.

2.1.3. scRNA-seq Data Simulation

Previous studies have used two representative simulations based on Gaussian and Poisson distributions to evaluate state-of-the-art gene co-expression estimation methods.

  • Zero-inflated Gaussian (ZI-Gaussian): The Gaussian-based simulation strategy in (Li and Li, 2021; Mestres et al., 2018) assumes non-zero values of the normalized gene expression matrix following a Gaussian distribution. This strategy first generates a co-expression network G following Sec. 2.1.2 and constructs a partial correlation matrix (i.e., the inverse of the covariance matrix) Σ−1 where Σij1=Wij denotes the edge weight between node i and j. Then it uses this partial correlation matrix along with gene expression means and standard deviations estimated from real scRNA-seq data to simulate a dense gene expression matrix X˜ through a multivariate Gaussian distribution. To introduce sparsity into the matrix, for each matrix element, it calculates the probability pij=expρX˜ij where ρ is a hyper-parameter controlling the level of sparsity. The final sparse simulation matrix is obtained by
    Xij=IijX˜ij with IijBernoullipij. (2)

    Following the same experiment setting in (Li and Li, 2021), we simulated four datasets with the same partial correlation matrix and different ρ{0.07,0.10,0.13,0.16} in our study. Since this simulation adds zeros on top of Gaussian samples, we name it ZI-Gaussian in our study.

  • Zero-inflated Poisson (ZI-Poisson): In addition, (Choi et al., 2017; Allen and Liu, 2013) use Poisson-based strategy. It first generates a gene expression matrix X˜ through a linear combination
    X˜=YB+E with YPoissonλtrue ,EPoissonλnoise . (3)
    Here, B is a permutation matrix based on W*. In order to have zeros, it then multiplies each element in the X˜ with a Bernoulli random variable as
    Xij=IijX˜ij with IijBernoulli(π) (4)
    where π controls the data sparsity. Detailed simulation process may refer to (Choi et al., 2017). We named this Poisson-based simulation as ZI-Poisson. Adopting the same experiment setting in (Choi et al., 2017), we generated three simulation datasets by fixing λtrue = 1.5, λnoise = 0.5, and varying π ∈ {0.8, 0.9, 1.0}.

Because ZI-Gaussian and ZI-Poisson simulations cannot exhibit characteristics of experimental scRNA-seq data (see Sec. 3.1), we apply two new simulation processes.

  • NORmal-To-Anything (NORTA): The first one is NORTA based on the normal-to-anything approach (Cario and Nelson, 1997; Chen, 2001; Sekula et al., 2020). The essential idea of NORTA is transforming multivariate Gaussian samples to samples with any given marginal distributions while preserving a given covariance. Previously, Sun et al. (Sun et al., 2021) have shown that using information from experimental data helps with mimicking the properties of real single-cell data. So in our study, we use NORTA to curate data with distributions learned from experimental data. Concretely, given experimental scRNA-seq data of p genes, we first use Negative Binomial (NB) distributions to fit each gene’s expression because NB is known to fit scRNA-seq data well (Hafemeister and Satija, 2019; Svensson, 2020), and obtain p marginal distributions. Then, we generate the covariance matrix as Σ = W*. NORTA utilizes them to simulate data with these p marginal NB distributions and gene covariance encoded by Σ. Fig. 3 illustrates NORTA simulation. For every experimental dataset, we generate corresponding synthetic data with NORTA 4. In NORTA simulations, the average hub-based network sparsity is 2.3%.

  • Single-cell ExpRession of Genes In silicO (SERGIO): Another new simulation strategy is SERGIO (Dibaeinia and Sinha, 2020). Given a GRN and profiles of regulators, SERGIO models the stochasticity of transcriptions and regulators with stochastic differential equations (SDEs). Concretely, it first generates dense gene expression matrix X˜ in logarithmic scale at stationary state via SDEs. Then it adds zeros via
    X˜ij=IijX˜ij with IijBernoulliπijwith πij=11+expkX˜:jy0andy0=(100%s%) percentile ofX˜ (5)
    with two hyper-parameters k and s determining the sparsity. Finally, it convert X˜ to the expression matrix through
    Xij=PoissonexpX˜ij. (6)

    The paper (Dibaeinia and Sinha, 2020) provides a comprehensive experiment showing that SERGIO can simulate realistic scRNA-seq data. Fig. 3 provides an illustration of SERGIO simulation. In our study, we simulate five datasets with k = 6.5 and s = {1, 5, 10, 15, 20} using the GRN and regulator profiles provided by (Dibaeinia and Sinha, 2020). In SERGIO simulations, the GRN sparsity is 5.2%.

Figure 3:

Figure 3:

Illustration of NORTA and SERGIO simulations. (Top) NORTA simulation strategy. First, it fits experimental data gene expression with negative binomial distributions. It generates a network encoding the covariance matrix. Then, the fitted negative binomial distributions and covariance matrix are used to simulate scRNA-seq data. (Bottom) SERGIO simulation strategy. It first prepares GRN, regulators profile (i.e., basal production rate in each cell type), and cell type differentiation graph. Then using these data, it generates clean data based on stochastic differential equations (SDEs). Finally, it adds dropouts to the clean expression matrix and converts it to the count expression matrix.

In ZI-Gaussian and ZI-Poisson simulations, we set the number of cells as 1000. In SERGIO simulation, the number of cells produced is 2700, with nine cell types in total and 300 cells per cell type. For NORTA, the number of cells in simulations equals that in the corresponding experimental data. So in all cases, the number of cells is much larger than the number of genes. We compare the characteristics of these simulation datasets with experimental datasets in Sec. 3.1,3.2 and test methods on simulations in Sec. 3.3.

2.2. Gene Co-Expression Network Estimators and Evaluation Metrics

We have tested nine gene co-expression network estimation methods in this study. These methods are summarized in Table 2. They consist of five methods designed for microarray or bulk RNA-seq data and four methods for scRNA-seq data. These methods cover a wide range of state-of-the-art and commonly used gene co-expression estimators with different data distribution assumptions and co-expression measurements. Detailed descriptions are provided in Appendix: More on Method and Results section.

Table 2:

Summary of gene co-expression estimation methods tested in this study. The first five are bulk data methods and the rest four are single-cell data methods. “NA” means the method has no explicit assumption on the data distribution.

Name Citation Co-expression Measurement Network Estimation Hyper-parameter Assumption
Pearson (Langfelder and Horvath, 2008) Pearson correlation thresholding threshold δ NA
Spearman (Langfelder and Horvath, 2008) Spearman correlation thresholding threshold δ NA
MINet (Meyer et al., 2008) mutual information thresholding threshold δ NA
GLasso(Cov) (Friedman et al., 2008) sample covariance MLE regularization λ Gaussian
GENIE3 (Huynh-Thu et al., 2010) regression trees thresholding threshold δ NA

PIDC (Langfelder and Horvath, 2008) partial information thresholding threshold δ gamma or Gaussian
scLink (Li and Li, 2021) robust correlation copula MLE regularization λ gamma-Gaussian mixture
GLasso(scDesign2) (Sun et al., 2021) correlation regression MLE regularization λ zero-inflated negative binomial
ZILNBGM (Park et al., 2021) coefficient regression regularization λ zero-inflated negative binomial

To evaluate the performance of methods, we consider co-expression estimation a binary classification task and compare estimated co-expression structures with the reference structure. Specifically, let G* ∈ {0, 1}p×p denote the true co-expression of p genes 5 with Gij=1 representing genes i and j are co-expressed. So all 1s in G are considered positives, and 0s are negatives. G^{0,1}p×p denotes the estimated co-expression. We evaluate methods with these metrics:

F1-score:

The F1-score of estimation is computed with the upper triangle (without diagonal entries) of ground truth G* and prediction G^, because the co-expression matrix is a symmetric matrix with diagonal entries all 1s:

F1-score =2 precision 1+ recall 1 (7)

with

precision =# true positives # estimated positives , recall =# true positives # actual positives . (8)

Ratio of false positives:

To evaluate how much false positives each method produce, we computed false discovery rate (FDR) and positive predictive value (PPV), or precision, with

FDR =# false positives # estimated positives , PPV =1 FDR. (9)

All the tested methods have only one hyper-parameter to control the sparsity of estimated networks. Specifically, Pearson, Spearman, MINet, GENIE3, and PIDC require a threshold δ to filter small values. So in our experiments, we generated a sequence of 100 equally spaced δ values from 0 to the maximal value in the measurement matrix to obtain a sequence of estimations from the complete network to the empty network. Then the final estimation is selected based on the highest F1-score. On the other hand, GLasso(Cov), GLasso(scDesign2), scLink, and ZILNBGM have λ as the sparsity-regularization hyper-parameter. Similarly, we generated a sequence of 100 equally spaced λ values from 0 to 1 and picked the best estimation with the highest F1-score. Especially, ZILNBGM suggests the maximal λ as the max covariance value. So we used this setting only for ZILNBGM. We also include the area under the receiver operating characteristic curve (AUROC) and the area under the precision-recall curve (AUPRC) values. For each of the 100 trials, we compute the estimation’s precision, recall, true positive rates, and false positive rates. AUPRC and AUROC are calculated based on curves constructed from these 100 metric values.

3. Results

3.1. Existing simulations to test co-expression estimation methods do not capture real-world data properties

We first test if two previously-used simulation processes, ZI-Gaussian and ZI-Poisson, represent real-world data. We simulate ‘single-cell’ datasets with 100 genes and hub-based gene co-expression networks, following the parameter setting in (Li and Li, 2021; Choi et al., 2017). We compare the properties of these simulations with experimental datasets from Table 1. Note that for experimental data, we take 100 HVGs from tens of thousands of genes; while on simulations, we directly generate 100 genes without a selection step. The HVG selection takes informative genes, and simulated genes are naturally informative (since we are not simulating uninformative noise), so the experimental and simulated data here (and in Sec 3.2) are comparable.

We systematically compare the sparsity (i.e., the ratio of non-zero values over total values) per cell and sparsity per gene between 100 HVGs of eight experimental datasets and seven ZI-Gaussian and ZI-Poisson simulated datasets with 100 genes. Fig. 4(AB) show simulations of both processes contain much fewer zeros than the experimental data. The average sparsity per cell for all eight experimental data is 0.083. But the most sparse ZI-Gaussian simulated dataset (i.e., ρ = 0.07) has average sparsity per cell of 0.468; the most sparse ZI-Poisson simulated dataset (i.e., π = 0.8) has average sparsity per cell of 0.789. Similarly, the average gene sparsity of experimental data is 0.083, while it of the most sparse ZI-Gaussian simulation is 0.47 and the most sparse ZI-Poisson simulation is 0.79. Two-sample Kolmogorov–Smirnov tests (KS test) between every experimental data and every simulated data give p < 10−14 (for cell sparsity) and p < 10−16 (for gene sparsity). This demonstrates that the cell and gene sparsity of simulated data is significantly different from experimental data.

Figure 4:

Figure 4:

Compare properties between experimental data, ZI-Gaussian (with ρ = {0.07, 0.10, 0.13, 0.16}), and ZI-Poisson (with π = {0.8, 0.9, 1.0}) simulations. (A) Sparsity per cell. (B) Sparsity per gene. (C) Compare four experimental datasets (#1 to #4) with the most sparse ZI-Gaussian/ZI-Poisson simulations. Histograms show distributions of non-zero parts for min-max scaled data.

Furthermore, we compare the expression distribution shape. Because the value range differs among datasets, we first normalize each dataset by min-max scaling as

Xij=XijminijXijmaxijXijminijXij. (10)

Data distributions of all datasets have a bi-modal structure consisting of dominating zero and rest non-zero parts. Because datasets have different sparsity levels, we only compare the distribution of non-zero components in Fig. 4(C). ZI-Poisson simulations are qualitatively different from experimental data. Also, we compute integrated absolute error (IAE) as

 IAE =ih1(i)h2(i). (11)

Here h1 and h2 are 20-bins histograms of two datasets, and h1(i) is the data frequency at the i-th bin. The average IAE between ZI-Gaussian and experimental data is 0.04; between ZI-Poisson and experimental data is 1.73; between every possible pair of experimental data is 0.03. This further indicates ZI-Poisson simulations are different from experimental data.

Our results show that ZI-Gaussian and ZI-Poisson simulations cannot display properties of actual scRNA-seq data. They are either less sparse or have very different distribution shapes. Therefore, despite good performance, the reliability of methods tested on these simulations becomes less certain. Next, we perform a comprehensive benchmark of gene co-expression estimation methods by generating more realistic simulations.

3.2. Generating Simulation with Experimental Data properties to Help Validate co-expression estimation Performance

We apply two new simulation processes, namely SERGIO and NORTA, to generate more realistic data. To validate that the two new simulation processes can represent real-world experimental data, we also compare cell sparsity, gene sparsity, and distribution shapes of the normalized data between these new simulations and experimental datasets. Fig. 5(AB) and Appendix Fig. A1 qualitatively show that NORTA and SERGIO simulation datasets have the similar gene and cell sparsity as the experimental data. The average gene sparsity of NORTA simulations is 0.087, and of SERGIO simulations is 0.11, which is similar to the average gene sparsity of experimental data (0.083). The cell sparsity of NORTA simulations is 0.087, and of SERGIO simulations is 0.11, which is also similar to the experimental data average cell sparsity (0.083). Two-sample KS test between most NORTA simulation and corresponding experimental data gene sparsity gives p > 0.1, except for dataset #6 (p = 1.2 × 10−4) and #8 (p = 2.4 × 10−2). Because these two experimental datasets have genes with very high sparsity levels, e.g., 0.03% non-zeros, the NORTA generated all zeros for these genes. So the KS test gave p-values smaller than 0.1 in these cases, but overall the NORTA can accurately simulate the sparsity level of the experimental data. KS tests between SERGIO simulation and experimental data gene sparsity all give p > 0.1. These denote no significant differences between simulation gene sparsity and experimental data gene sparsity. Moreover, Fig. 5(CD) show that new simulations have similar distribution shapes with experimental data, with NORTA ISE=0.014 and SERGIO ISE=0.086. Our results demonstrate that SERGIO and NORTA better exhibit the properties of the experimental data compared to the previous simulation methods.

Figure 5:

Figure 5:

Comparison between experimental data, NORTA, and SERGIO simulations. (A) Sparsity per gene of experimental data and corresponding NORTA simulations. (B) Sparsity per gene of experimental data and SERGIO simulations. (C) Compare distribution shapes between four experimental datasets and corresponding four NORTA simulations. Histograms show distributions of non-zero parts for min-max scaled data. (D) Compare distribution shapes between four experimental datasets and SERGIO simulations. Histograms show distributions of non-zero parts for min-max scaled data. (E) Correlation matrix distance (CMD) of experimental data and corresponding NORTA simulations. The gray bar denotes CMD between experimental data correlation and true network, i.e., CMD(Correxp, W*). The red bars denotes CMD between NORTA simulation correlation and true network, i.e., CMD(Corrsim, W*). (F) Mean squared error (MSE) of experimental data and corresponding NORTA simulations. The gray bar denotes MSE between experimental data correlation and true network, i.e., MSE(Correxp, W*). The red bar denotes the MSE between NORTA simulation correlation and true network, i.e., MSE(Corrsim, W*).

Note that experimental single-cell data generally contains gene expressions of multiples cell types (Consortium et al., 2018; Han et al., 2018; Rozenblatt-Rosen et al., 2017), so we check whether the NORTA and SERGIO simulations also have such heterogeneous information via clustering. We find the first 50 principal components for each dataset through the Principal Component Analysis (PCA), and we cluster cells with the Leiden algorithm (Traag et al., 2019). Leiden algorithm is a community detection method used in many scRNA-seq clustering algorithms (Cao et al., 2019; Mohammadi et al., 2020; Stuart et al., 2019; Wolf et al., 2018) that can automatically choose the number of clusters and show superior performance on finding cell types from single-cell data (Yu et al., 2022). We visualize each dataset’s clustering result in the 2D Uniform Manifold Approximation and Projection (UMAP) (McInnes et al., 2018) embedding space. Quantitatively, we compute the silhouette score (Rousseeuw, 1987) in the UMAP space to evaluate the cohesion and separation of clusters. The silhouette score ranges from −1 to +1, with a higher value indicating dense and well-separated clusters and values around zero indicating overlapping clusters. Fig. 6 shows that the NORTA and SERGIO simulations have clusters that are somewhat well separated (Appendix Fig. A2, A3, and A4 provide clustering visualization for all the NORTA and SERGIO simulations). Also, the NORTA simulation has a similar silhouette score with its corresponding experimental data. Note that we are not matching the cell clusters between experimental and simulated data. Having a silhouette score higher than 0 and similar to that on experimental data is sufficient to demonstrate that simulations possess multi-cell-type properties and are realistic.

Figure 6:

Figure 6:

Examples of cell clustering on experimental data, NORTA simulations, and SERGIO simulations. UMAP embedding visualization for the Leiden clustering on the first 50 principal components. Silhouette score is computed in the UMAP space. (A.) Experimental dataset (#1) and its corresponding NORTA simulation. (B.) SERGIO simulations with simulation parameter s = 1 and s = 5.

Another important aspect of the simulation is imposing a given gene co-expression structure on data such that it is feasible for estimation. (Dibaeinia and Sinha, 2020) gave a comprehensive testing of SERGIO and found simulations successfully have specified co-expression patterns. So here, we only validate that NORTA simulations have information of specified gene co-expression by comparing the Pearson correlation coefficients between experimental data and corresponding simulations. Specifically, let Correxp denote Pearson correlation coefficient matrix of experimental data; Corrsim denote correlation matrix of corresponding NORTA simulations. We use mean squared error (MSE) and correlation distance matrix (CMD) (Herdin et al., 2005) to measure the similarity between two correlation matrices. Here, a smaller MSE between two correlation matrices indicates they are more similar. CMD(A, B) = 0 denotes two correlation matrices are equal up to a scaling factor and CMD(A, B) = 1 if they differ in max extents. CMD values in Fig. 5(EF) show that CMD(Corrsim, W*) < CMD(Correxp, W*) and MSE(Corrsim, W*) < MSE(Correxp, W*) in most cases. Exceptions are still datasets #6 and #8 for the reason mentioned above, having slightly higher CMD and MSE in simulations. But overall, NORTA successfully imposes the co-expression structure W* on data.

In summary, the SERGIO and NORTA are more realistic simulations and better options in future works for evaluating gene co-expression estimators.

3.3. Existing Methods Estimate Gene Co-Expression Networks from realistically Simulated Data with a high FDR

Here, we test nine state-of-the-art gene co-expression estimators on SERGIO and NORTA simulations. We compute and compare F1-score, FDR, AUROC, and AUPRC across all methods.

On NORTA simulated datasets, some methods including the state-of-the-art scRNA-seq model scLink, can have incorrect predictions with F1-score lower than 0.2 (Fig. 7(A)). In other cases, even though the F1-score is around 0.5 for Spearman, MINet, and GLasso(scDesign2), they have FDR around 0.4 to 0.5 (Fig. 7(B)), indicating the predictions have many false positives. Appendix Fig. A11 gives an example of Spearman method prediction on NORTA simulations with 100 genes. In this case, F1-score=0.43 and FDR=0.6, containing 27 true positives and 40 false positives. This denotes that although F1-score is relatively high, the predictions have many false positives. On SERGIO simulations, all methods have F1-score lower than 0.2 and FDR higher than 0.8, giving poor co-expression estimations (Fig. 7(CD)). But we still can use a few methods piratically. Suppose one has to select a co-expression estimator to give a coarse-grained estimation as the starting point of GRN construction. In that case, we recommend using Spearman, MINet, or GLasso(scDesign2) since they have better performance in our experiments. The reason for it is that Spearman correlation and mutual information are robust to outliers and noise, and GLasso(scDesign2) captures gene correlations with Gaussian copula (Sun et al., 2021).

Figure 7:

Figure 7:

Methods performances on simulated and experimental datasets. In box plots, we split the performance of five bulk RNA/microarray models and four scRNA-seq models with a vertical line. Because ZILNBGM spends several days running a single trial on experimental data, we consider it uncomparable with other methods and omit it in the plots related to experimental data. (A-B) F1-score and FDR/PPV on NORTA simulations with 100 highly variable genes. (C-D) F1-score and FDR/PPV on SERGIO simulations with 100 genes. (E-F) F1-score and FDR/PPV on mouse cortex data with 1000 highly variable genes. (G-H) F1-score and FDR/PPV on human PBMC data with 1000 highly variable genes.

In Appendix Fig. A7 and A8, we tested methods on NORTA and SERGIO simulated data with more genes. It shows that methods’ performance decrease with an increased number of genes (see Appendix Section A1.5). This is reasonable since having more genes indicates the problem has higher complexity. So in later sections, we only use 100-genes simulations.

We also notice that methods have high AUROC but much lower AUPRC values (Appendix Fig. A5, A6). The reason for high AUROC is that the ground-truth network is very sparse, such that the number of negative instances (i.e., absence of co-expression) is much larger than the number of positive instances (i.e., the existence of co-expression). The AUROC measurement is known to be sensitive to class imbalance, so we also use AUPRC score, which alleviates the influence of data imbalance by considering both precision and recall values. Our observations suggest that co-expression estimation studies should use multiple metrics to comprehensively evaluate their methods.

Finally, we test methods on several simulated datasets from the previous studies for the completeness of comparison. First, we compare method estimations on ZI-Gaussian and ZI-Poisson simulations generated in Sec. 3.1. On four ZI-Gaussian simulations, scLink outperforms other models with average F1-score=0.29 , average FDR=0.74, and average AUPRC=0.22 (Appendix Fig. A9). The high F1-score of scLink is reasonable since the generation of ZI-Gaussian follows the Gaussian assumptions of the scLink method. We use the same simulation used in (Li and Li, 2021) and scLink achieves a similar AUPRC value as Fig. 2(D) in (Li and Li, 2021). However, though scLink outperforms its counterparts, its predictions still report a high FDR. On the other hand, on three ZI-Poisson simulations, Appendix Fig. A10 demonstrates that most methods have relatively better performance on ZI-Poisson than on other tested simulations (including ZI-Gaussian, NORTA, and SERGIO) and experimental data. Specifically, scLink is the best with average F1-score=0.75, average FDR=0.21, and average AUPRC=0.78. We think these good estimations may be due to the low sparsity of ZI-Poisson simulations but are not feasible in real-world single-cell applications due to their inconsistency with experimental data, as shown in Fig. 4. Through these experiments, we find that existing methods either report high FDR or use less sparse simulations than real-world data, making them less applicable to experimental single-cell datasets. On the contrary, we conduct experiments on more realistic simulations with comprehensive evaluations, which will be more reliable for future testing and benchmarking.

3.4. Methods Recover Known Gene Interactions from Experimental scRNA-seq Data with a low precision

Here we test methods on four experimental datasets (#2, #4, #5, and #7 in Table 1) with 1000 HVGs 6 to evaluate how they perform on real scRNA-seq data. We compare model predictions with known TF co-expression networks discovered on bulk data (see Sec. 2.1.2). For each experimental dataset, the number of TFs and the corresponding number of known co-expression edges are listed in Appendix Table A1. Note that these TF co-expression networks are not absolute ground truth since they are obtained from bulk data instead of single cells. But they are good references for us to evaluate whether methods can recover known information. Results in Fig. 7(EH) show that all methods have less than ideal performances on detecting known TF co-expression from experimental data. All methods have average F1-score≈0.25 and average PPV< 0.2 for mouse cortex data. GLasso(Cov), GLasso(scDesign2), and scLink perform best with an average PPV of 0.191. On the human PBMC data, these methods’ average F1-score< 0.1 and average PPV< 0.12, where Spearman outperforms others, having an average PPV= 0.117.

With these tests on experimental datasets, we find that our tested methods have a low precision when recovering bulk data-based TF-TF edges via their co-expression estimation from single-cell datasets. We also note that evaluating co-expression on experimental data needs careful consideration. A missing edge in reference networks does not mean the pair of genes are not co-expressed. That edge/co-expression perhaps has not been verified by previous bulk data experimental studies. So the FDR is expected to be high, as the estimator may discover unknown gene co-expression from single-cell data. We consider an alternative metric like precision (i.e., PPV), see Eq. 8. The precision/PPV evaluates how many estimated positives are true positives. This metric can test models’ performance on recovering known gene co-expression. However, since precision=1-FDR, high FDR values lead to low precision, and all methods give precision< 0.5 in our experiments. Unfortunately, we generally have little information to validate if the false positives are newly discovered co-expression interactions or errors. Therefore, we suggest caution while choosing the evaluation metric to verify co-expression estimation for real-world datasets. Augmenting downstream validations, such as the quality of GRN inference from the co-expression matrix and the accuracy of gene enrichment analysis, might be useful. Furthermore, we continue the remaining experiments on our simulations since we know the underlying ground truth.

3.5. High Sparsity in scRNA-Seq Data Impedes the Correlation Computation leading to poor Co-Expression Estimation

We generate data with different settings to explore which property of scRNA-seq data is the bottleneck of co-expression estimation. Since NORTA is more flexible and convenient at controlling simulation properties, like the type of network topology, we explore NORTA simulation in this section.

First, we use NORTA to simulate data with 100 genes and {10, 50, 100, 150, 200} cells. We observe that using more cells helps with the estimation (Fig. 9(A)). For scRNA-seq data, as the sequencing techniques enable the profiling of more cells, we can have data with hundreds and thousands of cells. So the number of cells is not the bottleneck.

Figure 9:

Figure 9:

Methods performances on simulated data with different settings. (A) F1-scores of NORTA simulations with 100 genes and {10, 50, 100, 150, 200} cells. (B-C) F1-scores and AUPRC values of methods on NORTA simulations with three different graph structures. (D) F1-scores of NORTA simulations with different sparsity levels. (E) The absolute difference between simulation correlations and the true network at different sparsity levels.

Some works state that real-world biological networks can be scale-free networks (Khanin and Wit, 2006; Clauset et al., 2009; Albert, 2005). Additionally, using known gene regulatory networks (GRNs) in simulation is a good way to incorporate prior knowledge. So we further use scale-free networks and example GRNs from (Dibaeinia and Sinha, 2020) in NORTA simulations with 100 genes and compare them with hub-based networks. Fig. 8 gives an example of these three types of networks. Fig. 9(BC) shows the similar performance of models, denoting that the graph structure is not the important factor.

Last, we simulate NORTA data with different sparsity levels using negative binomial gene marginal distributions and adding zero counts following Splatter simulation (Zappia et al., 2017). We test methods on data with 100 genes and {10%, 20%, , 90%} sparsity. Fig. 9(D) implies that the estimation worsens as the data becomes more and more sparse. To understand how sparsity affects co-expression estimation, we compute the Pearson correlation coefficients of data at different sparsities and study how the correlation changes. Fig. 9(E) computes the absolute difference between the simulation correlation and the true network. It shows that when data gets sparser, the simulation correlation becomes more different from the true network due to increased heterogeneity in the data, resulting in difficulties in co-expression evaluation. The increase in difference involving both positive (i.e., co-expression in the true network) and negative (i.e., non-co-expression in the true network) edges in the true network is perhaps the reason for the high FDR. Our results indicate that the sparsity in data is the major bottleneck of estimation and methods that can address sparsity should be investigated for future co-expression estimation method development.

3.6. Data Imputation or Pre-Processing steps introduce More False Positives in co-expression estimation

In scRNA-seq analysis, many pre-processing and imputation approaches have been proposed to address the analytical difficulties arising from the sparsity of the data. Here, we apply a few commonly used pre-processing and imputation approaches to test whether they can improve the co-expression estimation.

First, normalization is a critical pre-processing step in scRNA-seq analysis to adjust unwanted biases and obtain information of interest (Grün and van Oudenaarden, 2015; Vallejos et al., 2017; Chen et al., 2019). We utilize three commonly-used and novel normalizing schemes: LogNorm (Satija et al., 2015; Zheng et al., 2017; Stuart et al., 2019; Butler et al., 2018), sctransform (Hafemeister and Satija, 2019; Choudhary and Satija, 2022), and PsiNorm (Borella et al., 2022). Details about these normalization methods are provided in Appendix: More on Methods and Results. After comparing method estimations from the raw expression matrix and normalized matrix, we find the normalized data does not improve and even worsens the method performance (Fig. 10(AB)) on both types of simulations. Appendix Table A2 indicates that LogNorm and PsiNorm cannot reduce the sparsity in data such that challenges of estimation remain. For sctransform, it is designed to remove variations caused by RNA sequencing depths. But there are other types of technical variations (Haque et al., 2017), such as cell cycle (Buettner et al., 2015) and transcriptional bursting (Suter et al., 2011). So the reason sctransform does not improve the estimation is potentially due to its lack of ability to address multiple types of variations.

Figure 10:

Figure 10:

Comparison of methods performances on NORTA simulations and corresponding processed data from different approaches. (A-B) F1-scores for raw and normalized simulated data from LogNorm, PsiNorm, and sctransform. (C-D) F1-scores for raw simulations and pseudo-bulks. (E-F) F1-scores for raw and imputed simulations from MAGIC, SAVER, DCA, and ALRA.

One other way to reduce variations in scRNA-seq data is calculating the pseudo-bulk signal (Squair et al., 2021), which is obtained from averaging cell expressions for cells in the same cluster. For each simulation dataset, we apply UMAP to find the low-dimensional representation and then cluster cells with Leiden clustering. The pseudo-bulk expression is obtained by averaging cell expressions in each cluster. Fig. 10(CD) show that the pseudo-bulk cannot improve method predictions and might break down the method, e.g., for MINet. We hypothesize that clustering may induce biases, and constructing the pseudo-bulk results in statistical information loss, which affects the co-expression estimation.

To solve the problem with sparsity in scRNA-seq data, imputation (Hou et al., 2020) is a common solution. In scRNA-seq data, observed zeros can be either biological zeros denoting non-expression of genes or a technical limitation in quantifying the small number of molecules. To address the challenges induced by the sparsity, some recent works have developed several imputation methods to impute observed zeros. They were shown to improve downstream single-cell data analysis (Hou et al., 2020), such as differential gene expression analysis and cell clustering. Here, we test whether the imputation can improve gene co-expression estimation. We apply all methods on simulation data imputed with four commonly used state-of-the-art algorithms, MAGIC(Van Dijk et al., 2018), SAVER(Huang et al., 2018), DCA (Eraslan et al., 2019), and ALRA (Linderman et al., 2022). Appendix Table A2 shows the imputation has reduced data sparsity. But results indicate that imputed data worsen the method predictions, producing higher FDR scores (Fig. 10(EF) and Appendix Fig. A14). The imputation step has been reported to induce artificial correlations in data (Andrews and Hemberg, 2018; Hou et al., 2020) and thus might explain the observed reduced estimation performance.

In addition, simply removing lowly-expressed cells seems useful. So for each NORTA simulated dataset, we remove cells with library size smaller than t% × K, t ∈ {1%, 5%, 10%, 15%, 20%}, in which K is the maximal library cell of all cells. Appendix Fig. A18 shows that removing lowly-expressed cells doesn’t improve performance either. This is reasonable because Appendix Fig. A1 shows that all cells are sparse, so simply removing cells will not reduce data sparsity.

Overall, we have found that these commonly used pre-processing approaches do not improve the gene co-expression estimation and might further introduce more false positives, resulting in estimations of lower quality.

4. Discussion and Conclusion

In this study, we test several gene co-expression estimators on scRNA-seq data. Experiment results show that existing methods tend to produce estimations with high FDR. The choice of realistic underlying assumptions and testing on simulations that represent properties of experimental data are keys to developing robust estimation methods. We find that methods like GLasso and GENIE3 that are proposed for bulk data do not perform well for scRNA-seq data because scRNA-seq data distribution is significantly different. We also show that previously used simulations, ZI-Gaussian and ZI-Poisson, have much fewer zero counts and different distribution shapes compared to real-world experimental data. Therefore, we propose the use of two types of new simulated datasets, NORTA and SERGIO, that exhibit characteristics of experimental data. The low performance of existing methods on these simulations needs to be considered when applying them to real-world single-cell datasets. Especially a co-expression estimation with a low FDR is required in the field.

We also test different pre-processing steps of scRNA-seq datasets to see whether they can improve the estimation. We find that the widely-used normalization step can reduce data variations, however, the high-level of sparsity in the data continues to make co-expression estimation challenging. Next, we use pseudo-bulk to reduce data sparsity by averaging cell expressions through clustering. But it obtains a worse estimation compared to raw data in our experiments. We hypothesize that the clustering may induce biases in the data. Furthermore, constructing the pseudo-bulk reduces the number of samples and results in statistical information loss, which negatively affects the analysis. Finally, we show that scRNA-seq imputation methods also worsen the methods’ performance by potentially inducing artificial correlations in data. Overall, the pre-processing steps commonly used on scRNA-data do not help with co-expression estimation due to various limitations and biases related to the complexity of single-cell datasets.

While our study shows that none of the existing methods provide correct estimations, we believe a few methods can still be used practically as first steps for downstream improvement or analysis. Based on our experiments, we recommend using Spearman, MINet, or GLasso with scDesign2 transformation to measure gene co-expression since they have better performance in our experiments.

Unsurprisingly, we find that the main challenge of scRNA-seq data is its high sparsity. This makes robust analysis of the data a difficult task. We suggest that, in future works, the improvement on gene co-expression estimation should be made upon model robustness against sparsity and FDR control. Another potential solution might be adding more data from other modalities to help with the estimation. Due to the high sparsity and variations, the scRNA-seq data itself may not provide enough information for promising analysis, so integrating other data, for example scATAC-seq data, perhaps is important for more comprehensive and confident analyses.

In summary, our tested simulations will be helpful in gene co-expression analysis, especially for benchmarking and validating methods. We think our study provides valuable insights for scRNA-seq data analyses, and our benchmark setup can be a resource for the community as we work towards developing more accurate gene co-expression estimators.

Supplementary Material

Supplement 1

Acknowledgements

We thank Pinar Demetci and Tassallah Amina Abdullahi for their useful comments on this study.

Funding

This work is supported by National Institute of Health (NIH) grant R35 HG011939.

Footnotes

Code Availability

All the codes and data used in this study are available at https://github.com/rsinghlab/scRNAseq_Coexpression_Benchmark.

2

denotes the non-negative integers set.

3

https://singlecell.broadinstitute.org/single_cell. Mouse cortex data study number is SCP425; human PBMC data study number is SCP424.

4

We use the rnorta function in SimCorMultRes (Touloumis, 2016) R packages in experiment implementations.

5

In the network generation (Sec. 2.1.2), the G* can have negative co-expression. But here, to simplify the evaluation, we only focus on the existence or absence of co-expression.

6

Here, we use 1000 HVGs instead of 100 HVGs because the 100 HVGs contains only a few TFs such that the reference network will be too small.

Contributor Information

Jiaqi Zhang, Department of Computer Science, Brown University.

Ritambhara Singh, Department of Computer Science, Center for Computational Molecular Biology, Brown University.

References

  1. Albert R. (2005). Scale-free networks in cell biology. Journal of cell science, 118(21):4947–4957. [DOI] [PubMed] [Google Scholar]
  2. Allen G. I. and Liu Z. (2013). A local poisson graphical model for inferring networks from sequencing data. IEEE transactions on nanobioscience, 12(3):189–198. [DOI] [PubMed] [Google Scholar]
  3. Allocco D. J., Kohane I. S., and Butte A. J. (2004). Quantifying the relationship between co-expression, co-regulation and gene function. BMC bioinformatics, 5(1):1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Andrews T. S. and Hemberg M. (2018). False signals induced by single-cell imputation. F1000Research, 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Aoki K., Ogata Y., and Shibata D. (2007). Approaches for extracting practical information from gene co-expression networks in plant biology. Plant and Cell Physiology, 48(3):381–390. [DOI] [PubMed] [Google Scholar]
  6. Borella M., Martello G., Risso D., and Romualdi C. (2022). Psinorm: a scalable normalization for single-cell rna-seq data. Bioinformatics, 38(1):164–172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Buettner F., Natarajan K. N., Casale F. P., Proserpio V., Scialdone A., Theis F. J., Teichmann S. A., Marioni J. C., and Stegle O. (2015). Computational analysis of cell-to-cell heterogeneity in single-cell rna-sequencing data reveals hidden subpopulations of cells. Nature biotechnology, 33(2):155–160. [DOI] [PubMed] [Google Scholar]
  8. Butler A., Hoffman P., Smibert P., Papalexi E., and Satija R. (2018). Integrating single-cell transcriptomic data across different conditions, technologies, and species. Nature biotechnology, 36(5):411–420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Butte A. J. and Kohane I. S. (1999). Mutual information relevance networks: functional genomic clustering using pairwise entropy measurements. In Biocomputing 2000, pages 418–429. World Scientific. [DOI] [PubMed] [Google Scholar]
  10. Cao J., Spielmann M., Qiu X., Huang X., Ibrahim D. M., Hill A. J., Zhang F., Mundlos S., Christiansen L., Steemers F. J., et al. (2019). The single-cell transcriptional landscape of mammalian organogenesis. Nature, 566(7745):496–502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Cario M. C. and Nelson B. L. (1997). Modeling and generating random vectors with arbitrary marginal distributions and correlation matrix. Technical report, Citeseer. [Google Scholar]
  12. Carter S. L., Brechbühler C. M., Griffin M., and Bond A. T. (2004). Gene co-expression network topology provides a framework for molecular characterization of cellular state. Bioinformatics, 20(14):2242–2250. [DOI] [PubMed] [Google Scholar]
  13. Cha J. and Lee I. (2020). Single-cell network biology for resolving cellular heterogeneity in human diseases. Experimental & molecular medicine, 52(11):1798–1808. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Chan T. E., Stumpf M. P., and Babtie A. C. (2017). Gene regulatory network inference from single-cell data using multivariate information measures. Cell systems, 5(3):251–267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Chen G., Ning B., and Shi T. (2019). Single-cell rna-seq technologies and related computational data analysis. Frontiers in genetics, page 317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Chen H. (2001). Initialization for norta: Generation of random vectors with specified marginals and correlations. INFORMS Journal on Computing, 13(4):312–331. [Google Scholar]
  17. Choi H., Gim J., Won S., Kim Y. J., Kwon S., and Park C. (2017). Network analysis for count data with excess zeros. BMC genetics, 18(1):1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Choudhary S. and Satija R. (2022). Comparison and evaluation of statistical error models for scrna-seq. Genome biology, 23(1):1–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Clauset A., Shalizi C. R., and Newman M. E. (2009). Power-law distributions in empirical data. SIAM review, 51(4):661–703. [Google Scholar]
  20. Consortium T. M. et al. (2018). Single-cell transcriptomics of 20 mouse organs creates a tabula muris. Nature, 562(7727):367–372. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Danaher P., Wang P., and Witten D. M. (2014). The joint graphical lasso for inverse covariance estimation across multiple classes. Journal of the Royal Statistical Society: Series B (Statistical Methodology), 76(2):373–397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Dibaeinia P. and Sinha S. (2020). Sergio: a single-cell expression simulator guided by gene regulatory networks. Cell systems, 11(3):252–271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Ding J., Adiconis X., Simmons S. K., Kowalczyk M. S., Hession C. C., Marjanovic N. D., Hughes T. K., Wadsworth M. H., Burks T., Nguyen L. T., et al. (2019). Systematic comparative analysis of single cell rna-sequencing methods. BioRxiv, page 632216. [Google Scholar]
  24. Eraslan G., Simon L. M., Mircea M., Mueller N. S., and Theis F. J. (2019). Single-cell rna-seq denoising using a deep count autoencoder. Nature communications, 10(1):1–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Faith J. J., Hayete B., Thaden J. T., Mogno I., Wierzbowski J., Cottarel G., Kasif S., Collins J. J., and Gardner T. S. (2007). Large-scale mapping and validation of escherichia coli transcriptional regulation from a compendium of expression profiles. PLoS biology, 5(1):e8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Friedman J., Hastie T., and Tibshirani R. (2008). Sparse inverse covariance estimation with the graphical lasso. Biostatistics, 9(3):432–441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Greenfield A., Madar A., Ostrer H., and Bonneau R. (2010). Dream4: Combining genetic and dynamic information to identify biological networks and dynamical models. PloS one, 5(10):e13397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Grün D. and van Oudenaarden A. (2015). Design and analysis of single-cell sequencing experiments. Cell, 163(4):799–810. [DOI] [PubMed] [Google Scholar]
  29. Hafemeister C. and Satija R. (2019). Normalization and variance stabilization of single-cell rna-seq data using regularized negative binomial regression. Genome biology, 20(1):1–15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Han X., Wang R., Zhou Y., Fei L., Sun H., Lai S., Saadatpour A., Zhou Z., Chen H., Ye F., et al. (2018). Mapping the mouse cell atlas by microwell-seq. Cell, 172(5):1091–1107. [DOI] [PubMed] [Google Scholar]
  31. Haque A., Engel J., Teichmann S. A., and Lönnberg T. (2017). A practical guide to single-cell rna-sequencing for biomedical research and clinical applications. Genome medicine, 9(1):1–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Herdin M., Czink N., Ozcelik H., and Bonek E. (2005). Correlation matrix distance, a meaningful measure for evaluation of non-stationary mimo channels. In 2005 IEEE 61st Vehicular Technology Conference, volume 1, pages 136–140. IEEE. [Google Scholar]
  33. Hicks S. C., Townes F. W., Teng M., and Irizarry R. A. (2018). Missing data and technical variability in single-cell rna-sequencing experiments. Biostatistics, 19(4):562–578. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Hou W., Ji Z., Ji H., and Hicks S. C. (2020). A systematic evaluation of single-cell rna-sequencing imputation methods. Genome biology, 21(1):1–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Hsieh C.-J., Dhillon I., Ravikumar P., and Sustik M. (2011). Sparse inverse covariance matrix estimation using quadratic approximation. Advances in neural information processing systems, 24. [Google Scholar]
  36. Huang M., Wang J., Torre E., Dueck H., Shaffer S., Bonasio R., Murray J. I., Raj A., Li M., and Zhang N. R. (2018). Saver: gene expression recovery for single-cell rna sequencing. Nature methods, 15(7):539–542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Huynh-Thu V. A., Irrthum A., Wehenkel L., and Geurts P. (2010). Inferring regulatory networks from expression data using tree-based methods. PloS one, 5(9):e12776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Hwang B., Lee J. H., and Bang D. (2018). Single-cell rna sequencing technologies and bioinformatics pipelines. Experimental & molecular medicine, 50(8):1–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Iacono G., Massoni-Badosa R., and Heyn H. (2019). Single-cell transcriptomics unveils gene regulatory network plasticity. Genome biology, 20(1):1–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Kang H. J., Kawasawa Y. I., Cheng F., Zhu Y., Xu X., Li M., Sousa A. M., Pletikos M., Meyer K. A., Sedmak G., et al. (2011). Spatio-temporal transcriptome of the human brain. Nature, 478(7370):483–489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Khanin R. and Wit E. (2006). How scale-free are biological networks. Journal of computational biology, 13(3):810–818. [DOI] [PubMed] [Google Scholar]
  42. Lähnemann D., Köster J., Szczurek E., McCarthy D. J., Hicks S. C., Robinson M. D., Vallejos C. A., Campbell K. R., Beerenwinkel N., Mahfouz A., et al. (2020). Eleven grand challenges in single-cell data science. Genome biology, 21(1):1–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Lambert S. A., Jolma A., Campitelli L. F., Das P. K., Yin Y., Albu M., Chen X., Taipale J., Hughes T. R., and Weirauch M. T. (2018). The human transcription factors. Cell, 172(4):650–665. [DOI] [PubMed] [Google Scholar]
  44. Langfelder P. and Horvath S. (2008). Wgcna: an r package for weighted correlation network analysis. BMC bioinformatics, 9(1):1–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Li W. V. and Li Y. (2021). sclink: Inferring sparse gene co-expression networks from single-cell expression data. Genomics, proteomics & bioinformatics, 19(3):475–492. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Linderman G. C., Zhao J., Roulis M., Bielecki P., Flavell R. A., Nadler B., and Kluger Y. (2022). Zero-preserving imputation of single-cell rna-seq data. Nature communications, 13(1):1–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Liu Y., Gu H.-Y., Zhu J., Niu Y.-M., Zhang C., and Guo G.-L. (2019). Identification of hub genes and key pathways associated with bipolar disorder based on weighted gene co-expression network analysis. Frontiers in physiology, 10:1081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Macosko E. Z., Basu A., Satija R., Nemesh J., Shekhar K., Goldman M., Tirosh I., Bialas A. R., Kamitaki N., Martersteck E. M., et al. (2015). Highly parallel genome-wide expression profiling of individual cells using nanoliter droplets. Cell, 161(5):1202–1214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Marco A., Konikoff C., Karr T. L., and Kumar S. (2009). Relationship between gene co-expression and sharing of transcription factor binding sites in drosophila melanogaster. Bioinformatics, 25(19):2473–2477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Margolin A. A., Nemenman I., Basso K., Wiggins C., Stolovitzky G., Favera R. D., and Califano A. (2006). Aracne: an algorithm for the reconstruction of gene regulatory networks in a mammalian cellular context. In BMC bioinformatics, volume 7, pages 1–15. BioMed Central. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. McInnes L., Healy J., and Melville J. (2018). Umap: Uniform manifold approximation and projection for dimension reduction. arXiv preprint arXiv:1802.03426. [Google Scholar]
  52. Mestres A. C., Bochkina N., and Mayer C. (2018). Selection of the regularization parameter in graphical models using network characteristics. Journal of Computational and Graphical Statistics, 27(2):323–333. [Google Scholar]
  53. Meyer P. E., Kontos K., Lafitte F., and Bontempi G. (2007). Information-theoretic inference of large transcriptional regulatory networks. EURASIP journal on bioinformatics and systems biology, 2007:1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Meyer P. E., Lafitte F., and Bontempi G. (2008). minet: Ar/bioconductor package for inferring large transcriptional networks using mutual information. BMC bioinformatics, 9(1):1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Mohammadi S., Davila-Velderrain J., and Kellis M. (2020). A multiresolution framework to characterize single-cell state landscapes. Nature communications, 11(1):1–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Nguyen H., Tran D., Tran B., Pehlivan B., and Nguyen T. (2021). A comprehensive survey of regulatory network inference methods using single cell rna sequencing data. Briefings in bioinformatics, 22(3):bbaa190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Oldham M. C., Konopka G., Iwamoto K., Langfelder P., Kato T., Horvath S., and Geschwind D. H. (2008). Functional organization of the transcriptome in human brain. Nature neuroscience, 11(11):1271–1282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Ovens K., Eames B. F., and McQuillan I. (2021). Comparative analyses of gene co-expression networks: Implementations and applications in the study of evolution. Frontiers in Genetics, 12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Park B., Choi H., and Park C. (2021). Negative binomial graphical model with excess zeros. Statistical Analysis and Data Mining: The ASA Data Science Journal, 14(5):449–465. [Google Scholar]
  60. Pedregosa F., Varoquaux G., Gramfort A., Michel V., Thirion B., Grisel O., Blondel M., Prettenhofer P., Weiss R., Dubourg V., Vanderplas J., Passos A., Cournapeau D., Brucher M., Perrot M., and Duchesnay E. (2011). Scikit-learn: Machine learning in Python. Journal of Machine Learning Research, 12:2825–2830. [Google Scholar]
  61. Picelli S., Faridani O. R., Björklund Å. K., Winberg G., Sagasser S., and Sandberg R. (2014). Full-length rna-seq from single cells using smart-seq2. Nature protocols, 9(1):171–181. [DOI] [PubMed] [Google Scholar]
  62. Pratapa A., Jalihal A. P., Law J. N., Bharadwaj A., and Murali T. (2020). Benchmarking algorithms for gene regulatory network inference from single-cell transcriptomic data. Nature methods, 17(2):147–154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Presson A. P., Sobel E. M., Papp J. C., Suarez C. J., Whistler T., Rajeevan M. S., Vernon S. D., and Horvath S. (2008). Integrated weighted gene co-expression network analysis with an application to chronic fatigue syndrome. BMC systems biology, 2(1):1–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Qiu P. (2020). Embracing the dropouts in single-cell rna-seq analysis. Nature communications, 11(1):1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Risso D., Perraudeau F., Gribkova S., Dudoit S., and Vert J.-P. (2018). A general and flexible method for signal extraction from single-cell rna-seq data. Nature communications, 9(1):1–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Rodchenkov I., Babur O., Luna A., Aksoy B. A., Wong J. V., Fong D., Franz M., Siper M. C., Cheung M., Wrana M., et al. (2020). Pathway commons 2019 update: integration, analysis and exploration of pathway data. Nucleic acids research, 48(D1):D489–D497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Rouillard A. D., Gundersen G. W., Fernandez N. F., Wang Z., Monteiro C. D., McDermott M. G., and Ma’ayan A. (2016). The harmonizome: a collection of processed datasets gathered to serve and mine knowledge about genes and proteins. Database, 2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Rousseeuw P. J. (1987). Silhouettes: a graphical aid to the interpretation and validation of cluster analysis. Journal of computational and applied mathematics, 20:53–65. [Google Scholar]
  69. Rozenblatt-Rosen O., Stubbington M. J., Regev A., and Teichmann S. A. (2017). The human cell atlas: from vision to reality. Nature, 550(7677):451–453. [DOI] [PubMed] [Google Scholar]
  70. Sanchez-Taltavull D., Perkins T. J., Dommann N., Melin N., Keogh A., Candinas D., Stroka D., and Beldi G. (2020). Bayesian correlation is a robust gene similarity measure for single-cell rna-seq data. NAR genomics and bioinformatics, 2(1):lqaa002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Satija R., Farrell J. A., Gennert D., Schier A. F., and Regev A. (2015). Spatial reconstruction of single-cell gene expression data. Nature biotechnology, 33(5):495–502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Sekula M., Gaskins J., and Datta S. (2020). A sparse bayesian factor model for the construction of gene co-expression networks from single-cell rna sequencing count data. BMC bioinformatics, 21(1):1–19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Seo C. H., Kim J.-R., Kim M.-S., and Cho K.-H. (2009). Hub genes with positive feedbacks function as master switches in developmental gene regulatory networks. Bioinformatics, 25(15):1898–1904. [DOI] [PubMed] [Google Scholar]
  74. Serin E. A., Nijveen H., Hilhorst H. W., and Ligterink W. (2016). Learning from co-expression networks: possibilities and challenges. Frontiers in plant science, 7:444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Song L., Langfelder P., and Horvath S. (2012). Comparison of co-expression measures: mutual information, correlation, and model based indices. BMC bioinformatics, 13(1):1–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Squair J. W., Gautier M., Kathe C., Anderson M. A., James N. D., Hutson T. H., Hudelle R., Qaiser T., Matson K. J., Barraud Q., et al. (2021). Confronting false discoveries in single-cell differential expression. Nature communications, 12(1):1–15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Stuart J. M., Segal E., Koller D., and Kim S. K. (2003). A gene-coexpression network for global discovery of conserved genetic modules. science, 302(5643):249–255. [DOI] [PubMed] [Google Scholar]
  78. Stuart T., Butler A., Hoffman P., Hafemeister C., Papalexi E., Mauck III W. M., Hao Y., Stoeckius M., Smibert P., and Satija R. (2019). Comprehensive integration of single-cell data. Cell, 177(7):1888–1902. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Sun T., Song D., Li W. V., and Li J. J. (2021). scdesign2: a transparent simulator that generates high-fidelity single-cell gene expression count data with gene correlations captured. Genome biology, 22(1):1–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Suter D. M., Molina N., Gatfield D., Schneider K., Schibler U., and Naef F. (2011). Mammalian genes are transcribed with widely different bursting kinetics. science, 332(6028):472–474. [DOI] [PubMed] [Google Scholar]
  81. Svensson V. (2020). Droplet scrna-seq is not zero-inflated. Nature Biotechnology, 38(2):147–150. [DOI] [PubMed] [Google Scholar]
  82. Taeb A., Shah P., and Chandrasekaran V. (2020). Learning exponential family graphical models with latent variables using regularized conditional likelihood. arXiv preprint arXiv:2010.09386. [Google Scholar]
  83. Tong A. H. Y., Lesage G., Bader G. D., Ding H., Xu H., Xin X., Young J., Berriz G. F., Brost R. L., Chang M., et al. (2004). Global mapping of the yeast genetic interaction network. science, 303(5659):808–813. [DOI] [PubMed] [Google Scholar]
  84. Touloumis A. (2016). Simulating correlated binary and multinomial responses under marginal model specification: The simcormultres package. The R Journal, 8(2):79–91. [Google Scholar]
  85. Townes F. W., Hicks S. C., Aryee M. J., and Irizarry R. A. (2019). Feature selection and dimension reduction for single-cell rna-seq based on a multinomial model. Genome biology, 20(1):1–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Traag V. A., Waltman L., and Van Eck N. J. (2019). From louvain to leiden: guaranteeing well-connected communities. Scientific reports, 9(1):1–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Usadel B., Obayashi T., Mutwil M., Giorgi F. M., Bassel G. W., Tanimoto M., Chow A., Steinhauser D., Persson S., and Provart N. J. (2009). Co-expression tools for plant biology: opportunities for hypothesis generation and caveats. Plant, cell & environment, 32(12):1633–1651. [DOI] [PubMed] [Google Scholar]
  88. Vallejos C. A., Risso D., Scialdone A., Dudoit S., and Marioni J. C. (2017). Normalizing single-cell rna sequencing data: challenges and opportunities. Nature methods, 14(6):565–571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Van Dam S., Vosa U., van der Graaf A., Franke L., and de Magalhaes J. P. (2018). Gene co-expression analysis for functional classification and gene–disease predictions. Briefings in bioinformatics, 19(4):575–592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Van Dijk D., Sharma R., Nainys J., Yim K., Kathail P., Carr A. J., Burdziak C., Moon K. R., Chaffer C. L., Pattabiraman D., et al. (2018). Recovering gene interactions from single-cell data using data diffusion. Cell, 174(3):716–729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Wolf F. A., Angerer P., and Theis F. J. (2018). Scanpy: large-scale single-cell gene expression data analysis. Genome biology, 19(1):1–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Wu C.-L., Dicks A., Steward N., Tang R., Katz D. B., Choi Y.-R., and Guilak F. (2021). Single cell transcriptomic analysis of human pluripotent stem cell chondrogenesis. Nature communications, 12(1):1–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Yeung K. Y., Medvedovic M., and Bumgarner R. E. (2004). From co-expression to co-regulation: how many microarray experiments do we need? Genome biology, 5(7):1–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Yin J. and Li H. (2011). A sparse conditional gaussian graphical model for analysis of genetical genomics data. The annals of applied statistics, 5(4):2630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Yin W., Mendoza L., Monzon-Sandoval J., Urrutia A. O., and Gutierrez H. (2021). Emergence of co-expression in gene regulatory networks. PloS one, 16(4):e0247671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Yu L., Cao Y., Yang J. Y., and Yang P. (2022). Benchmarking clustering algorithms on estimating the number of cell types from single-cell rna-sequencing data. Genome biology, 23(1):1–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Zappia L., Phipson B., and Oshlack A. (2017). Splatter: simulation of single-cell rna sequencing data. Genome biology, 18(1):1–15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Zheng G. X., Terry J. M., Belgrader P., Ryvkin P., Bent Z. W., Wilson R., Ziraldo S. B., Wheeler T. D., McDermott G. P., Zhu J., et al. (2017). Massively parallel digital transcriptional profiling of single cells. Nature communications, 8(1):1–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Zhou Q., Liu M., Xia X., Gong T., Feng J., Liu W., Liu Y., Zhen B., Wang Y., Ding C., et al. (2017). A mouse tissue transcription factor atlas. Nature communications, 8(1):1–15. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplement 1

Articles from bioRxiv are provided here courtesy of Cold Spring Harbor Laboratory Preprints

RESOURCES