Your privacy, your choice

We use essential cookies to make sure the site can function. We also use optional cookies for advertising, personalisation of content, usage analysis, and social media.

By accepting optional cookies, you consent to the processing of your personal data - including transfers to third parties. Some third parties are outside of the European Economic Area, with varying standards of data protection.

See our privacy policy for more information on the use of your personal data.

for further information and to change your choices.

Skip to main content

Identification and evaluation of candidate COVID-19 critical genes and medicinal drugs related to plasma cells

Abstract

The ongoing COVID-19 pandemic, caused by the SARS-CoV-2 virus, represents one of the most significant global health crises in recent history. Despite extensive research into the immune mechanisms and therapeutic options for COVID-19, there remains a paucity of studies focusing on plasma cells.

In this study, we utilized the DESeq2 package to identify differentially expressed genes (DEGs) between COVID-19 patients and controls using datasets GSE157103 and GSE152641. We employed the xCell algorithm to perform immune infiltration analyses, revealing notably elevated levels of plasma cells in COVID-19 patients compared to healthy individuals. Subsequently, we applied the Weighted Gene Co-expression Network Analysis (WGCNA) algorithm to identify COVID-19 related plasma cell module genes. Further, positive cluster biomarker genes for plasma cells were extracted from single-cell RNA sequencing data (GSE171524), leading to the identification of 122 shared genes implicated in critical biological processes such as cell cycle regulation and viral infection pathways.

We constructed a robust protein-protein interaction (PPI) network comprising 89 genes using Cytoscape, and identified 20 hub genes through cytoHubba. These genes were validated in external datasets (GSE152418 and GSE179627). Additionally, we identified three potential small molecules (GSK-1070916, BRD-K89997465, and idarubicin) that target key hub genes in the network, suggesting a novel therapeutic approach. These compounds were characterized by their ability to down-regulate AURKB, KIF11, and TOP2A effectively, as evidenced by their low free binding energies determined through computational analyses using cMAP and AutoDock.

This study marks the first comprehensive exploration of plasma cells’ role in COVID-19, offering new insights and potential therapeutic targets. It underscores the importance of a systematic approach to understanding and treating COVID-19, expanding the current body of knowledge and providing a foundation for future research.

Peer Review reports

Introduction

Infectious disease COVID-19, primarily attributed to the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), continues its relentless spread globally [1, 2]. The infection typically manifests as severe flu-like symptoms, progressing to acute respiratory distress syndrome (ARDS), pneumonia, renal failure, and, in severe cases, death [3,4,5,6]. Epidemiological evidence indicates fever, cough, and dyspnea as the predominant symptoms among affected individuals [7]. As of May 12th, 2024, the World Health Organization (WHO) reported 775,481,326 confirmed cases and 7,049,376 fatalities due to COVID-19 (https://data.who.int/dashboards/covid19/). This underscores the necessity of comprehending the disease’s characteristics and the role of pharmacological interventions in its management.

Viral invasion, immune system failure, microcirculation dysfunction, endothelial cell damage, and dysregulation of the renin-angiotensin-aldosterone system are recognized as pathophysiological hallmarks of SARS-CoV-2 infection [8, 9]. Since the emergence of COVID-19, extensive studies have focused on elucidating the molecular mechanisms and host-pathogen interactions pertinent to COVID-19 pathogenesis by analyzing biomolecular profiles in patients [10,11,12,13]. The spike protein of SARS-CoV-2 or SARS-CoV, for example, targets the ACE2 membrane protein, which possesses an enzymatic domain on human cell surfaces, facilitating viral entry [14]. Chua et al. observed that epithelial cells in COVID-19 cases exhibit a threefold increase in ACE2 expression, linked to immune cell interferon signaling [15]. Similarly, Winkler et al. reported significant inflammation, immune cell infiltration, and impaired respiratory function in infected human ACE2 transgenic mice [16]. In COVID-19 patients, a notable decrease in peripheral blood lymphocytes correlates with disease severity [17]. Improvement in lymphocyte count during treatment indicates a favorable prognosis, while a continual decrease suggests a negative outcome [18,19,20,21,22,23]. Severely affected patients exhibit a low T cell count but high inflammatory cytokine levels, primarily secreted by monocytes, macrophages, or other innate immune cells, leading to adverse patient prognoses [24, 25]. Despite the pivotal role of the immune system in COVID-19 progression, the transcriptomic alterations and key genes affecting plasma cells in the disease remain underexplored.

Weighted gene co-expression network analysis (WGCNA) is a system biology method used to analyze gene expression profiles across multiple samples [24, 25]. It clusters genes with similar expression patterns and assesses correlations between modules and specific traits [24, 25], and is widely used in studying conditions such as acute lung damage, chronic obstructive pulmonary disease, and heart failure [26,27,28]. Highly connected gene clusters, potential biomarkers, and therapeutic targets for a variety of illnesses, including viruses, can all be found using it [25, 29, 30]. Using the WGCNA algorithm on high-throughput data, we identified genes with similar expression changes. WGCNA’s criteria, rooted in biological significance, make its results more reliable than traditional methods [24, 25]. These findings enable further research, including trait associations, metabolic pathway modeling, and gene interaction network creation [26, 28, 31]. Single-cell RNA sequencing (scRNA-seq) has greatly advanced our understanding of COVID-19’s etiology and immunological characteristics [32]. It profiles transcriptomes from individual cells within a sample, enabling detailed gene expression analysis [32]. scRNA-seq has been instrumental in studying SARS-CoV-2-infected cells, compositional changes in cell subpopulations, and profiling immune signaling pathways and transcriptomes during disease progression [33, 34]. WGCNA and high-throughput technologies together offer a fresh chance to comprehend the molecular causes of COVID-19 and other infectious disorders [35, 36]. This evidence suggests a significant potential to identify crucial COVID-19 genes through integrated bioinformatics analysis.

In the face of such a highly infectious respiratory condition, developing anti-COVID-19 medications is critical due to the absence of specific treatments for COVID-19. A range of existing medications can mitigate COVID-19 symptoms. The WHO recommends nirmatrelvir and ritonavir for high-risk patients with mild to moderate COVID-19, at risk of hospitalization [37]. Nirmatrelvir inhibits SARS-CoV-2’s 3Cl protease, essential for viral replication [38,39,40], and is combined with ritonavir, a CYP3A inhibitor, to sustain effective drug concentrations [40,41,42]. However, side effects, drug interactions, and cost considerations are important. Common side effects include disturbances to the nervous and cardiovascular systems, gastrointestinal issues, and drug interactions, especially in patients with chronic conditions [43,44,45,46,47]. Hence, research into refining therapeutic strategies to alleviate symptoms and enhance effectiveness is highly promising.

Plasma cells, derived from B lymphocytes, can secrete immunoglobulins or antibodies [48, 49]. Antibodies produced against SARS-CoV-2 can neutralize the virus, reducing its pathogenicity [50]. Because plasma cells produce antibodies that specifically target the SARS-CoV-2 virus; they are essential to the immune response to COVID-19. Convalescent plasma has historically been used in treating viral diseases like Ebola and various fevers, including SARS and influenza strains [51]. The successful application of plasma in pulmonary diseases offers feasibility for COVID-19 treatment [51, 52]. Studies pertaining to COVID-19 and plasma cells have been published. Following COVID-19 or immunization against SARS-CoV-2, SARS-CoV-2 specific antibody-secreting plasma cells promoting specific normal immunity have been found in the human bone marrow. Following basic mRNA vaccination, Schulz et al. demonstrated the production of phenotypically varied SARS-CoV-2-specific plasma cells in the human bone marrow, including the development of memory characteristics [18, 53]. They also came to the conclusion that the basic mRNA vaccine against SARS-CoV-2 induces long-lasting humoral protection [53]. Long-lived bone marrow plasma cells are a reliable and necessary source of antibodies that provide protection. Turner reported that aspirates from 11 healthy people without a history of SARS-CoV-2 infection did not contain S-specific bone marrow plasma cells, indicating their possible membership in a stable compartment [54]. To sum up, in order to effectively mount an antibody response against COVID-19, plasma cells are necessary. Their function changes according to the disease’s severity, and their continued existence is essential for maintaining immunity against both vaccination and spontaneous infection.

Large-scale cytokine and chemokine release causes excessive inflammation and immune cell mobilization, which is what causes the pathogenic phenomena known as “cytokine storm” [55]. Patients with COVID-19 may have a poor prognosis and a persistent infection as a result of plasma cell malfunction [56,57,58]. As a result, the discovery of potential biomarkers linked to plasma cells during COVID-19 infection may help clarify the infection’s progression, enhance our knowledge of the molecular underpinnings of SARS-Cov-2 infection, and enhance COVID-19 diagnosis and therapy.

Herein, the purpose of this study is further to explore the role of plasma cells in COVID-19, providing new insights and novel therapeutic targets, and to highlight the significance of plasma cells in COVID-19 with regard to prognosis and treatment. It emphasizes how crucial it is to approach comprehending and treating COVID-19 methodically, adding to the body of information already in existence and laying the groundwork for further research. Therefore, we developed a workflow for identifying and evaluating candidate COVID-19 critical genes and medicinal drugs related to plasma cells in Fig. 1. In this study, we identified 122 shared genes among differentially expressed genes (DEGs), co-expressed genes, and positive cluster biomarker genes. Functional enrichment analysis annotated these genes to processes like nuclear division, organelle fission, DNA helicase and ATP hydrolysis activities, cell cycle, p53 signaling pathway, human T-cell leukemia virus 1 infection, and protein processing in the endoplasmic reticulum. We constructed a protein-protein interaction (PPI) network of 89 genes and identified 20 hub genes using the cytoHubba plugin. Finally, three potential small compounds (GSK-1070916, BRD-K89997465, and idarubicin) were identified that down-regulate key genes (AURKB, KIF11, and TOP2A) with the lowest free binding energy, using the connectivity map (cMAP) database and AutoDock modeling. Our study offers a novel perspective for identifying and evaluating plasma cell-related candidate targets and their drugs for COVID-19 patients.

Materials and methods

Data source and pre-operation

Our study compiled five independent datasets (Table 1), categorized into discovery and validation sets. The discovery datasets comprised GSE157103, GSE152641, and GSE171524, while GSE152418 and GSE179627 served as validation datasets.

GSE157103 encompasses 126 RNA-seq samples, consisting of 100 from COVID-19 patients and 26 from non-COVID-19 hospitalized individuals [59]. GSE152641, another publicly available RNA-seq dataset, includes 86 samples: 62 from patients and 24 from healthy controls [60]. Additionally, we incorporated a single-nucleus RNA-seq (snRNA-seq) dataset, GSE171524, comprising 20 samples from 19 COVID-19 decedents and 7 control donors [61].

For external validation, GSE152418 and GSE179627 were utilized. GSE152418 contains RNA-seq data of 17 COVID-19 patients and 17 healthy controls, further classified by severity status (17 healthy, 1 convalescent, 4 moderate, 8 severe, and 4 ICU) [62]. GSE179627 includes 70 RNA-seq samples from COVID-19 patients, differentiated by five immune response stages (22 uninfected, 9 asymptomatic, 13 symptomatic, 14 recovering, and 12 re-detectable positive patients) [35].

For bulk RNA-seq datasets, raw read counts and sample labels were retrieved and prepared for analysis. Gene annotation was performed using the bitr() function from the clusterProfiler (version 4.4.4) R package, converting EntrezIDs to gene symbols [63, 64]. Similarly, gene names, cell names, sparse matrices, and metadata were extracted from snRNA-seq data for subsequent analyses.

Fig. 1
figure 1

The flowchart figure for our overall methodology. The workflow of this study is mainly divided into the following two steps: Step (1) Database-based screening for hub genes associated with COVID-19 symptoms. Step (2) Verify the reliability of hub genes based on external datasets and further analyze docking results

Table 1 Summary of the datasets used in this study

Identification of DEGs

We combined the bulk RNA-seq datasets GSE157103 and GSE152641 as our discovery dataset. This amalgamation resulted in a raw read count matrix comprising 19,184 genes across 212 samples. FactoMineR (version:2.11) and factoextra (version: 1.0.7) R packages were used for performing PCA analysis and visualization. The PCA() function in FactoMineR R package was used to perform PCA analysis. All features (genes) were included in the PCA analysis. The fviz_pca_ind() function in the factoextra R package was applied for visualization. The top two principal components were selected for visualization. Given the disparate origins of the datasets, which may affect gene expression levels, we incorporated batch effect correction in our analysis (Supplementary Fig. 1). The detailed correction method is that the sample label (COVID-19 or non-COVID-19) was designated as the primary variable factor, and the batch effect was considered an additional variable factor by utilizing the “design” parameter in DESeq2 (version 1.36.0) R package for the differential gene expression (DEG) analysis [65].

The identification of DEGs is crucial in analyzing RNA-seq count data. DESeq2 (version 1.36.0), a prevalent R package, facilitates the quantification and statistical inference of RNA-seq data by constructing a negative binomial generalized linear model [65]. For count data transformation analysis, given the sample size exceeding 100, we opted for the vst() function over the rlog() function. The resulting normalized expression matrix was used as the input for heatmap generation.

Significant up-regulation was attributed to genes with an adjusted P-value less than 0.05 and a log2 fold change (log2FC) greater than 1. Conversely, significant down-regulation was assigned to genes with an adjusted P-value less than 0.05 and log2FC less than − 1. We utilized the ggplot2 (version 3.6.6) R package in RStudio for volcano plot visualization and the pheatmap (version 1.0.12) R package for depicting the top 20 up- and down-regulated genes, sorted by fold change magnitude.

Cell type enrichment analysis

Human tissue comprises various cell types. Accurately estimating the proportion of these cell types, especially immune cell types, is critical for elucidating disease mechanisms and suggesting potential treatments in the realm of precision medicine [66, 67]. The xCell algorithm is frequently used for cell type enrichment analysis from gene expression profiles [68]. It utilizes 1,822 human cell-type transcriptome datasets to generate 64 gene signatures, which are categorized into five groups: lymphoid, stem cells, myeloid, stromal cells, and others [68].

For this analysis, we converted raw read counts to fragments per kilobase million (FPKM) by normalizing for sequence depth and gene length, adhering to the xCell requirements. We employed the xCellAnalysis() function from the xCell package (version 1.1.0) on our RNA-seq dataset, yielding an xCell score matrix for 64 cell types across 212 samples.

Subsequently, we conducted comparative analyses to discern differences in xCell scores between COVID-19 patients and healthy controls. To visualize and statistically analyze these differences, we employed the ggpubr (version 0.4.0) and rstatix (version 0.7.0) R packages for generating boxplots and conducting statistical tests. We deemed cell type enrichments as significantly different if the P-value between COVID-19 and healthy controls was less than 0.05.

Weighted gene co-expression network analysis (WGCNA)

The module associated with COVID-19 and plasma cells was identified using WGCNA [24, 25]. WGCNA is a systematic biological approach designed to identify clusters of genes with similar expression patterns, often indicative of shared roles in biological processes. The WGCNA computational framework includes building a gene co-expression network, identifying modules, associating modules with clinical traits, exploring module interactions, and identifying key genes within the modules of interest.

Using WGCNA (version 1.71) in R, we sought to discover modules related to COVID-19 and plasma cells. Our analysis incorporated 162 COVID-19 and 50 healthy control samples. We categorized all samples into low and high plasma cell cohorts based on the median value of the xCell score. We input an expression matrix consisting of 19,060 genes across 212 samples into WGCNA. Only genes within the top 75% of median absolute deviation (MAD) and with a minimum MAD of 0.01 were considered. After this gene-level filtering, we processed 14,295 genes for 212 samples. Outlier samples were identified using the hclust() method. We only contained 209 samples from 212 samples by removing three outliers for the further analysis.

Many data in the relevant references prove that the network already conforms to the distribution of a scale-free network when the signed R2 is greater than 0.85 [24, 25]. Expecting to consider the recommended soft power by the script “sft$powerEstimate”, we added two criteria: (1) Draw the relationship between the soft threshold and the scale-free topology model fit, and take the soft threshold that makes scale-free topology model fit, signed R2 > 0.85 as the soft threshold. (2) Draw the relationship between the soft threshold and the mean connectivity, and take the soft threshold that makes mean connectivity < = 500 as the soft threshold. Based on the above discussion, 6 was set to the soft power. The minModuleSize parameter represents the minimum number of genes in a module. The smaller the value, the smaller the module will be retained. The mergeCutHeight indicates the distance of merging similar modules. The smaller the value, the less likely they will be merged, and the more modules are retained. The minModelSize was set to 100, and the mergeCutHeight was set to 0.25 based on the defaulted setting consistent with the WGCNA tutorial (https://rpubs.com/natmurad/WGCNA) and related published studies [69,70,71]. A module was deemed related to COVID-19 and plasma cells if it had a correlation coefficient greater than 0.3 in both the COVID-19 and high-plasma cell-enriched groups and a P-value less than 0.05. Furthermore, genes within this module were defined as module genes related to both COVID-19 and plasma cells. The relationship between module membership and gene significance was visualized using the verboseScatterplot() function.

Single-cell RNA sequencing (scRNA-seq) analysis

We analyzed positive cluster biomarker genes of plasma cells in COVID-19 samples as they are indicative of key roles in the immune microenvironment. Using Seurat (version 4.2.0), we initiated objects by introducing sparse matrices, gene names, and cell names [72,73,74,75]. Notably, the sparse matrix was first converted to a dgCMatrix format as per Seurat’s default requirement. A total of 34,546 genes across 116,313 cells were integrated into the Seurat object. In this study, we utilized the cleaned data download from the link: https://singlecell.broadinstitute.org/single_cell/study/SCP1219/columbia-university-nyp-covid-19-lung-atlas#study-download. Based on the description from Melms et al., the data was cleaned [61].

For uniformity in cell type annotation, we accessed and downloaded processed data, including metadata, from their portal website (https://singlecell.broadinstitute.org/single_cell/study/SCP1219/columbia-university-nyp-covid-19-lung-atlas). The “cell_type_intermediate” column was used to assign cell type identities. Subsequently, UMAP was employed for non-linear dimension reduction and two-dimensional visualization.

A stacked barplot visualized the distribution of 19 cell types across 26 COVID-19 samples. To compare cell ratios between COVID-19 and healthy control groups across each cell type, an inter-group t-test was conducted, setting a significance cutoff at P-value < 0.05. Positive cluster biomarker genes in plasma cells were identified using the FindMarkers() method with the parameters: only.pos = TRUE, min.pct = 0.25, logfc.threshold = 0.25, and test.use = “wilcox”. The only.pos parameter was set to TRUE because we only focus on positive markers. At the same time, 0.25 is a common parameter that can be used at the min.pct and logfc.threshold based on the associated published research [76,77,78]. The “wilcox” was set to the test method for identifying differentially expressed genes between two groups of cells using the Wilcoxon Rank Sum test by default.

Functional annotation of significant genes

Functional annotation analysis is a critical tool for unraveling the biological significance of genes [79,80,81]. It aids in understanding the functionality of gene groups, discerning gene-disease relationships, and inferring gene products and their roles in biological processes [82,83,84]. For this purpose, various gene annotation databases have been developed to annotate and store gene functions across different biological paradigms.

In this study, our gene lists were derived from differentially expressed genes (DEGs) and overlapping shared genes. We utilized the clusterProfiler package (version 4.4.4) in R to conduct functional enrichment analysis of both downregulated and upregulated genes between COVID-19 and non-COVID-19 patients. Initially, the bitr() function was employed for gene name transformation from the official gene symbol to Entrez ID. Subsequently, the top 30 Gene Ontology (GO) biological process (BP) category terms, sorted by significance level in ascending order, were visualized.

We integrated GO cellular component (CC), biological process (BP), molecular function (MF), and KEGG pathway analysis to deduce functional pathways of shared genes implicated in COVID-19, with a particular focus on plasma cells. Shared genes were defined as those appearing at least twice among DEGs, module genes, and positive cluster biomarker genes. The clusterProfiler package (version 4.4.4) facilitated the functional enrichment analysis of these shared genes [63, 64]. The Database for Annotation, Visualization, and Integrated Discovery (DAVID) was employed to provide enriched KEGG pathways of shared genes [85, 86]. The top 10 GO terms in the BP, CC, and MF categories were individually visualized, and KEGG pathways with a P-value < 0.05 were highlighted. This comprehensive analysis enables a deeper understanding of the functional implications of genes in the context of COVID-19 and plasma cells.

Construction of the PPI network

Protein-protein interactions (PPIs) are fundamental elements of cellular biochemical networks and are crucial for regulating cellular functions and signaling pathways [87,88,89]. Understanding the intricate relationships among multiple proteins via systematic analysis facilitates insights into biological signal response mechanisms and energy substance metabolism in specific physiological conditions, including disease states, as well as elucidates the functional interplay among proteins [87, 90].

To analyze PPIs, shared genes identified in our study were input into the STRING database (version 11.5, https://string-db.org/) [91]. In total, there are 19,566 proteins (nodes) in H. sapiens (https://version-11-5.string-db.org/organism/9606). We focused on retaining PPIs with a false discovery rate (FDR) of < 0.05 and a combined score of > 0.8 to ensure the relevance and reliability of the interactions based on the recommended results [92, 93]. The resulting PPI network was then visualized using Cytoscape software (version 3.9.0) [94], which is adept at depicting complex interaction networks. Within the network, node border thickness corresponds to the degree of connectivity, with thicker borders indicating higher degrees of interaction. Similarly, the edge thickness represents the interaction score, with thicker lines suggesting stronger associations. This graphical representation allows for a more intuitive understanding of the extensive and nuanced interplay between proteins in the context of COVID-19 and plasma cells, thereby providing a foundational framework for further biological and therapeutic explorations.

Identification of sub-networks and hub genes

In the construction of a protein-protein interaction (PPI) network using the STRING database, often hundreds of PPI pairs are identified. To manage this complexity and distill the network down to its most influential components, it is necessary to employ algorithms that can identify key sub-networks and central genes, known as hub genes, based on network topology.

CytoHubba, a widely acclaimed plugin for Cytoscape, is particularly adept at identifying pivotal nodes within biological networks and is thus a valuable tool for experimental biologists seeking to uncover critical regulatory networks and potential drug targets [95]. Among the various methods available in CytoHubba, the Matthews correlation coefficient (MCC) method is recommended for its superior predictive accuracy compared to other topological analysis algorithms [95].

Utilizing the MCC method, a scoring system is applied to each gene within the PPI network, facilitating the identification of the most centrally involved genes, or hub genes. The resultant sub-network, comprising the top 20 hub genes, is visually distinguished in the network diagram, with the genes colored according to their rank. This prioritization not only highlights the genes most central to the network’s structure and function but also provides a focused list of candidates for subsequent biological or therapeutic investigation. This analytical approach is particularly crucial in understanding the complex biological interactions at play in diseases such as COVID-19 and can significantly aid in identifying potential targets for further research and drug development.

Validation of hub genes

Validation of hub genes is crucial for substantiating their relevance and reliability in differentiating COVID-19 patients from non-COVID-19 individuals and across various severity stages of the disease. Initially, the expression patterns of the 20 hub genes identified in the previous step were extracted from the discovery dataset to establish a baseline comparison between COVID-19 and non-COVID-19 groups.

To ensure the robustness and applicability of these hub genes, we utilized an independent external dataset, GSE152418, which includes detailed severity clinical information. Samples within this dataset were categorized into five distinct groups: convalescent, healthy, ICU (Intensive Care Unit), moderate, and severe. We computed and compared the expression levels of the 20 hub genes across these varied groups to observe any consistent trends or significant differences that could correlate with disease severity.

Further validation was performed using another independent external dataset, GSE179627, which provided a range of immune response classifications, including asymptomatic, recovering, re-detectable positive patients, uninfected, and symptomatic groups. This diversity in immune response categories allowed for a comprehensive evaluation of the hub genes as potential biomarkers across different stages and responses to COVID-19 infection.

To quantify the predictive power of the hub genes, the R package pROC (version 1.18.4) was employed to calculate the area under the receiver operating characteristic curve (AUC) and the 95% confidence intervals (CIs). Genes with AUC values exceeding 0.65 were considered to have satisfactory discriminatory ability and were retained for further analysis based on the previous studies [96, 97]. This step is vital to ensuring that the hub genes not only show differential expression between COVID-19 and non-COVID-19 cases but also possess the ability to distinguish between various disease severities and immune responses, thereby affirming their potential utility in clinical and research settings.

Screening of candidate small molecular compounds

We incorporated the 20 hub genes into the Connectivity Map (cMAP) database [98] to explore potential small molecular compounds that may influence these genes’ expression. The cMAP database, hosting over 1.5 million expression profiles from more than 5,000 small molecule compounds and approximately 3,000 genetic perturbations across diverse cell types, provided a normalized connectivity score reflecting each compound’s effect on specific cell types and perturbations [98]. These proteins, encoded by the hub genes, were considered potential targets for novel therapeutic drugs. cMAP facilitated the prediction of small molecule medicines that might down-regulate these proteins [98]. We particularly emphasized the top ten promising small molecule substances or medications that demonstrated potential in reducing the expression of these critical genes.

In order to make the research more in-depth, we sourced structural and descriptive information on these therapeutic agents from the PubChem database [99], a leading repository for chemical data. Proteins encoded by genes with Area Under Curve (AUC) values greater than 0.75 were categorized as significant proteins associated with COVID-19, aiding the prediction of immune responses [99].

Molecular modeling simulation

Molecular docking employs a theoretical modeling approach to design drugs by analyzing the properties of receptors and how they interact with drug molecules, aiming to predict their binding affinities. Autodock Vina, version 1.1.2, is a widely recognized application in this field [100]. We collected the crystal structures of three proteins, AURKB (4AF3) [101], KIF11 (3K5E) [102], and TOP2A (4FM9), from the Protein Data Bank (PDB) in PDB format for our study [103, 104]. Correspondingly, the chemical structures for the potential inhibitory small molecules—GSK-1070916, BRD-K89997465 (chlorpromazine), and idarubicin—were sourced from the PubChem database in SDF format [99]. PyMOL, version 1.7.X, was utilized for the removal of extraneous molecules and data format transformation, while Autodock tools, version 1.5.7, were used for hydrogen addition and charge editing. Finally, LigPlus, version 2.2.8, facilitated the detailed analysis of interactions between the proteins and ligands [105, 106]. This comprehensive simulation provided a deeper insight into the potential efficacy of these compounds as COVID-19 therapeutics.

Molecular dynamics simulation

We used GROMACS (version: 2022.4) software to perform an all-atom molecular dynamics simulation analysis on the protein-ligand complex obtained by molecular docking. The protein part was parameterized using the Amber14SB force field, while the topology file of the small molecule part was generated by ACPYPE and Antechamber programs. We selected a cubic solvation box and set the shortest distance from the edge of the system to the complex to 1 nm. We selected the TIP3P water model and added appropriate amounts of sodium and chloride ions to balance the overall charge of the system.

Furthermore, the system used the steepest descent method for energy minimization, and the temperature was regulated by the NVT ensemble, and the pressure was controlled by the NPT ensemble to ensure that the temperature of the entire system was stable at 300 K and the pressure was maintained at 101.325 kPa. For each balanced system, we performed a 100ns molecular dynamics simulation at 300 K and obtained a total of 10,000 frames of simulation trajectories. We used the trajectory files obtained from the simulation to conduct an in-depth analysis of key parameters such as the root mean square deviation (RMSD), root mean square fluctuation (RMSF), radius of gyration (Rg), and the total number of hydrogen bonds between the protein and the ligand. The free energy landscape (FEL) reflects the relationship between the Gibbs free energy of the protein and the RMSD and Rg. (The presence of two low energy points (i.e., local minima) in the free energy landscape (FEL) usually means that the system has two stable conformational states.)

In addition, we take a 2 ns trajectory (i.e., 200 frames) during the equilibrium period and calculate the binding free energy using the MM-PBSA method of gmx_mmpbsa. By definition, in solution, we can write the binding free energy as

$$\:\varDelta\:{G}_{bind}={G}_{complex}-\left({G}_{free-protein}+{G}_{free-ligand}\right)$$

Then, the solvation free energy can be further decomposed into polar and nonpolar parts.

$$\:{G}_{bind}={E}_{gas}-{TS}_{gas}+{G}_{solvation}$$

In which, the solvation free energy can be further decomposed into polar and nonpolar parts.

$$\:{G}_{solvation}={G}_{polar}+{G}_{nonpolar}$$

In MM-PBSA method, the energy and entropy contributions of the gas phase are calculated according to the MM method as follows

$$\:{E}_{gas}={E}_{MM}={E}_{bond}+{E}_{angle}+{E}_{dihedral}+{E}_{vdw}+{E}_{coulomb}$$
$$\:{S}_{gas}={S}_{MM}$$

Among them, \(\:{E}_{bond}\), \(\:{E}_{angle}\), and \(\:{E}_{dihedral}\) correspond to bond, bond angle, and dihedral interactions, respectively, and \(\:{E}_{vdw}\) and\(\:\:{E}_{coulomb}\) represent van der Waals forces and Coulomb electrostatic interactions, respectively.

In the MM-PBSA method, the solvation energy contains two parts: acute solvation energy and nonpolar solvation energy. The acute solvation energy comes from the electrostatic interaction between solute and solvent molecules and is calculated using an implicit solvent model, in which the solvent is considered as a continuous medium and the corresponding Poisson-Boltzmann equation is linearized and numerically solved:

$$\:{G}_{polar}={G}_{GB}$$

The nonpolar solvation free energy can be calculated based on the empirical surface area method, so it is also called the surface solvation energy. The calculation requires knowledge of the solvent accessible surface area (A) of the molecule and uses two empirical parameters \(\:\gamma\:\) and b:

$$\:{G}_{nonpolar}={G}_{surface}=\gamma\:A+b$$

Combining the above items, the free energy equation of MM-PBSA is obtained

$$\:{G}_{bind}={E}_{MM}-{TS}_{MM}+{G}_{GB}+{G}_{surface}$$

This analysis process is similar to published analysis [107, 108].

Other statistical analysis

In this study, bioinformatics and statistical analyses were conducted using R and RStudio, version 4.2.1. We employed the VennDiagram package, version 1.7.3, to calculate and illustrate the overlap among DEGs, module genes, and positive plasma cell biomarker genes. The Student’s t-test was utilized to ascertain differences in cell ratios between COVID-19 patients and healthy controls across various cell types. Additionally, the Kruskal-Wallis test was applied to assess variations in the expression of hub genes among convalescent, healthy, ICU, moderate, and severe COVID-19 patient groups in our validation cohorts. A P-value threshold of 0.05 was adopted to determine statistical significance in all tests. This comprehensive statistical approach ensures rigorous and robust analysis of the data, aiding in reliable and insightful conclusions.

Results

DEGs between COVID-19 and non-COVID-19 patients

Upon integrating the datasets GSE157103 and GSE152641, we assembled a matrix comprising 19,184 genes across 212 samples, including 162 from COVID-19 patients and 50 from non-COVID-19 individuals. The expression profiles were adjusted for batch effects and designated as the discovery dataset.

Analysis revealed 936 genes significantly upregulated and 125 significantly downregulated, adhering to the criteria: log2FC > 1, adjusted P-value < 0.05 for upregulation, and log2FC < -1, adjusted P-value < 0.05 for downregulation (Supplementary Table 1). We highlighted DEGs with log2FC > 2 and adjusted P-value < 0.01 in the volcano plot (Fig. 2A). The top 20 upregulated and downregulated genes were extracted from the normalized expression matrix for detailed visualization in a heatmap (Fig. 2B), indicating a predominance of upregulated genes.

Separate functional enrichment analyses were conducted for up- and down-regulated genes. Up-regulated genes are predominantly associated with organ or tissue-specific immune response, mucosal immune response, and viral life cycle regulation, among others (Fig. 2C). Conversely, down-regulated genes corresponded to pathways like leukocyte chemotaxis and various metabolic processes (Fig. 2D). Noteworthy is the overlap in immune response and signaling pathways across both gene sets, despite distinct pathway enrichments unique to each. Specifically, virus invasion pathways were exclusive to up-regulated genes, while metabolic pathways were more characteristic of down-regulated genes in GO-BP terms. This distinct pathway involvement underscores the complex interplay of biological processes in COVID-19 pathogenesis.

Fig. 2
figure 2

Visualization of DEGs and functional enrichment GO-BP terms of significantly up- and down-regulated genes between COVID-19 and non-COVID-19 patients. A Volcano plot of DEGs. The red color dots indicate up-regulated genes, the blue dots indicate down-regulated genes, and the grey dots indicate the non-significantly expressed genes in COVID-19 compared to non-COVID-19. B The heatmap shows the expression pattern of the top 20 up- and down-regulated genes. The functional enrichment in GO-BP of up- (C) and down-regulated (D) genes. The x-axis represents the gene ratio value, while the y-axis represents the GO term

The difference in plasma cell infiltration between COVID-19 and non-COVID-19 patients

Gene expression profiles normalized to gene length (FPKM) were analyzed using the xCell algorithm [68]. Cell infiltration scores for 64 cell types across 212 samples were calculated and compared, as shown in Fig. 3 and Supplementary Table 2. Notably, a substantial number of cell types exhibited significant differences between the COVID-19 and non-COVID-19 cohorts. Within the lymphoid category, B cells, various T cell subsets (including CD4 + and CD8 + populations), memory B cells, naïve B cells, NKT cells, plasma cells, Pro B cells, Th1 cells, Th2 cells, and Tregs all demonstrated significant differential enrichment between COVID-19 and non-COVID-19 patients (Fig. 3A). Among myeloid cells, aDC, basophils, cDCs, DCs, iDCs, macrophages, and neutrophils were notably differentiated between the two patient groups (Fig. 3B). Other cell types such as astrocytes and neurons also showed significant differences (Fig. 3C), as did certain stem cells including CMP, GMP, megakaryocytes, and MEP (Fig. 3D), and stromal cells like endothelial cells, MSCs, myocytes, pericytes, and smooth muscle (Fig. 3E).

Such cellular disparities align with the known pathological progression of COVID-19, where viral infection triggers an exaggerated immune response, leading to hyperinflammation or a cytokine storm, further escalating multi-organ damage and increasing mortality risk [109].

Specifically, plasma cells were markedly more prevalent in COVID-19 patients compared to non-COVID-19 individuals (P-value ~ 2e-10), highlighting them as crucial for subsequent focused analyses. This significant increase underscores the potential role of plasma cells in the immune response against SARS-CoV-2, warranting further investigation into their involvement in disease progression and potential as targets for therapeutic intervention.

Fig. 3
figure 3

Immune enumerate levels of 64 cell types between COVID-19 and non-COVID-19 patients. The 64 cell types were divided into five main categories. A Lymphoid. B Myeloids. C Others. D Stem cells. E Stromal cells. The highlighted red cell type is plasma cells. Student’s t-test was used for statistical analysis, and P < 0.05 was set as the cutoff. The boxplot shows the results from the analysis of xCell score values of 64 cell types in COVID-19 and non-COVID-19 cohorts based on the t-test statistical method. The ns, *, **, ***, and **** correspond separately to 1, 0.05, 0.01, 0.001, and 0.0001, respectively

Identification of modules and genes related to both COVID-19 and plasma cells

We constructed a co-expression network pertinent to plasma cells and COVID-19 using Weighted Gene Co-expression Network Analysis (WGCNA). This method groups genes into modules based on similar expression patterns and examines the correlation between these modules and specific traits [24, 25].

For this analysis, we established a network with a chosen soft power of 6, ensuring an R² > 0.9 for optimal network connectivity (Fig. 4A). The gene dendrogram and module color plot are illustrated in Fig. 4B. The upper panel shows the gene hierarchical cluster dendrogram, while the lower panel denotes the corresponding gene modules. This combination indicates that genes closely located in the dendrogram are grouped into the same module. A total of ten modules were identified: black (197 genes), blue (1,583 genes), brown (1,173 genes), green (472 genes), gray (657 genes), magenta (131 genes), pink (146 genes), red (214 genes), turquoise (8,928 genes), and yellow (794 genes). Notably, the black module exhibited a positive correlation with both plasma cell presence and COVID-19 status (Fig. 4C), with statistical significance (P-value < 0.05). Figure 4D depicts the correlation between module membership and gene significance within the black module. The findings indicate a strong association of the black module with plasma cells and COVID-19 (correlation = 0.32 and P-value = 4.6e-6). Consequently, genes within the black module were selected for subsequent detailed analysis due to their potential relevance to both COVID-19 and plasma cell activity (Supplementary Table 3).

Fig. 4
figure 4

WGCNA for identifying the modules and genes. A The soft thresholding power’s scale-free fit index. The correlation between soft power and R2 is demonstrated in the left panel, while the correlation between soft power and mean connectivity is shown in the right panel. B Gene and module color dendrogram. The y-axis indicates the distance between clusters. C Heatmap of the relationship between modules and clinical features of interest (plasma cells and COVID-19). D The scatter plot of the relationship between module membership and gene significance in the black module

Positive cluster biomarker genes of plasma cells in COVID-19 samples

Advancements in single-cell technology have allowed for an extensive exploration of cellular heterogeneity [110]. The cell type distribution in control and COVID-19 samples is depicted in Fig. 5A. We identified 19 distinct cell types: Airway epithelial cells, AT1, AT2, B cells, CD4 + T cells, CD8 + T cells, Cycling NK/T cells, Dendritic cells, Endothelial cells, Fibroblasts, Macrophages, Mast cells, Monocytes, Neuronal cells, NK cells, Other epithelial cells, Plasma cells, Smooth muscle, and Tregs. The cell counts for each cell type in both control and COVID-19 groups are specified in the supplementary materials.

The proportion of each cell type in the respective samples is visualized in Fig. 5B. Notably, there was a marked increase in the percentage of immune cells, particularly plasma cells (Fig. 5C), macrophages (Fig. 5D), and monocytes (Fig. 5E), in COVID-19 patients compared to controls (Supplementary Fig. 2). These alterations corroborate existing literature on the pathogenic mechanisms of COVID-19, highlighting the virus’s impact on the immune system and the subsequent hyperactive immune response, often referred to as a cytokine storm, leading to aggravated disease progression and potentially fatal outcomes [111, 112]. Consequently, 296 positive cluster biomarker genes specific to plasma cells were identified for subsequent analysis (Supplementary Table 4).

Fig. 5
figure 5

The cellular landscape of COVID-19 patients. A Major clusters and respective cell-type assignments in UMAP deduce dimension visualization of control and COVID-19 samples. B The fraction of 19 cell types in 26 individual lungs. The difference in cell compositions from two groups of patients for (C) plasma cells, (D) macrophages, and (E) monocytes. The suffix “cov” represents COVID-19, and “ctl” represents control samples

Shared genes among DEGs, module genes, and positive cluster biomarker genes

We conducted an intersection analysis to determine commonalities among DEGs, module genes, and positive cluster biomarker genes, illustrated in Fig. 6A. We prioritized genes implicated in multiple dysregulated pathways due to their potential significance in pathogenesis, designating genes appearing in at least two of the three categories as “shared genes.” Consequently, 122 shared genes were identified, including ACOXL, AURKB, BHLHA15, and others listed in Supplementary Table 5.

The shared genes were postulated to contribute to at least two of the following three dysregulated mechanisms: (1) differential expression between COVID-19 and non-COVID-19 samples; (2) association with plasma cells and COVID-19 phenotype through co-expression patterns; and (3) regulation as positive cluster biomarker genes in plasma cells.

To investigate the biological relevance of these shared genes, we conducted GO and KEGG pathway enrichment analyses, the results of which are depicted in Fig. 6B-C. The analyses revealed significant enrichment in nuclear division, organelle fission, mitotic cell cycle phase transition, DNA helicase activity, ATP hydrolysis activity, cell cycle, p53 signaling pathway, human T-cell leukemia virus 1 infection, and protein processing in the endoplasmic reticulum. These findings align with the heightened cellular and immune responses observed in combating SARS-CoV-2 infection, indicating a vigorous immune response to viral invasion [113,114,115].

Fig. 6
figure 6

Shared genes and enriched pathways. A The Venn diagram and UpSet plots show the overlapping genes among DEGs, module genes, and positive cluster biomarker genes. B The enriched GO terms of shared genes and the top 10 GO terms were selected to visualize each GO category. CC stands for cellular component, BP stands for biological process, and MF stands for molecular function. C The enriched KEGG pathways of shared genes

Analysis of PPI networks, sub-networks, and hub genes

Protein-protein interactions (PPIs) are essential for various biological functions including signaling, gene regulation, metabolism, and cell cycle control. In biomedicine, PPI network analysis is instrumental for identifying new therapeutic targets and unraveling disease mechanisms systematically. Generally, PPI network analysis employs databases like STRING for initial analysis, followed by tools like Cytoscape for network visualization and refinement [31, 116, 117].

Our analysis yielded an initial undirected network comprising 111 nodes and 4,886 edges. By applying stringent criteria, we refined this to a robust network consisting of 89 nodes and 1,283 edges, as illustrated in Fig. 7A. Using the cytoHubba plugin, we identified the top 20 hub genes: CCNA2, TOP2A, CCNB2, TPX2, BUB1, CDK1, KIF2C, RRM2, AURKB, CDCA8, KIF11, CCNB1, CDC20, UBE2C, KIF20A, PTTG1, TTK, DLGAP5, BUB1B, and PBK. A sub-network featuring these genes was subsequently delineated (Fig. 7B). An examination of the gene expression profiles of these hub genes revealed a consistent and significant upregulation in COVID-19 patients compared to non-COVID-19 individuals, as depicted in Fig. 7C. This observation underscores their potential relevance in the pathophysiology of COVID-19.

Fig. 7
figure 7

PPI network and sub-network. A The PPI network. The node indicates the shared genes, while the edge indicates the interaction obtained from the STRING database. The depth of the node and edge represents the score of interaction. The larger the interaction combined score is, the darker the color. B The sub-network is mined from network analysis. The depth of the node and edge represents the degree of the node. The larger the degree is, the darker the color. C The expression pattern of 20 hub genes of sub-network. Up-regulated illustrates that the gene is significantly up-regulated in COVID-19 compared to non-COVID-19 samples

Validation and identification of candidate targets

Identifying clinical biomarkers and developing diagnostic and therapeutic tools are critical for advancing clinical applications [118, 119]. One of the key challenges in this process is extracting potential biomarkers with high sensitivity, stability, and accuracy from the vast amounts of multi-omic data available [120, 121].We employed two external RNA-seq datasets, GSE152418 and GSE179627, to validate the robustness and potential of our identified hub genes in terms of expression levels and classification capabilities. As depicted in Fig. 8A-E and Supplementary Fig. 3, a discernible difference in expression levels among various COVID-19 patient groups was noted for the 20 hub genes (P-value < 0.05), as per external dataset 1. Moreover, these genes exhibited higher expression in ICU, moderate, and severe COVID-19 patient cohorts compared to convalescent and healthy groups, aligning with inferences drawn from the discovery dataset.

Additionally, these 20 hub genes were evaluated for their ability to classify immune responses in external dataset 2, which encompassed diverse immune statuses including asymptomatic, recovering, re-detectable positive, uninfected, and symptomatic cases (Table 2). Notably, the genes TOP2A, BUB1, KIF2C, AURKB, CDCA8, KIF11, PTTG1, and BUB1B demonstrated AUC values exceeding 0.65, thereby qualifying them as promising candidate targets (Fig. 8F). These findings suggest that these eight genes, due to their substantial classification performance, are worthy of further exploration in the context of COVID-19 treatment.

Fig. 8
figure 8

The validation and selection of target genes. The normalized expression levels of the CDCA8 (A), KIF11 (B), AURKB (C), BUB1 (D), and PTTG1 (E) genes in external dataset 1. F The ROC curves of the CDCA8, KIF11, and PTTG1 genes in by an external dataset 2

Table 2 Summary of the datasets used in this study

Docking model of candidate compounds and target proteins

Molecular docking has become an increasingly vital tool in computer-aided drug discovery, offering insights into the potential interactions between drugs and targets [122, 123]. While docking scores may not precisely predict drug inhibitory effects on proteins, the conformations derived from docking often closely align with crystal structures, making it a cost-effective approach for medicinal chemists [124, 125].

In our study, we utilized the cMAP database to identify candidate small compounds capable of down-regulating the expression of hub genes. We focused on compounds with normalized connectivity scores below zero and further refined our selection to those targeting any of the 20 hub genes. Ultimately, this process yielded candidate compounds for AURKB, KIF11, and TOP2A genes (Supplementary Tables 6–8). The leading candidates for down-regulating AURKB include GSK-1070916, KW-2449, and danusertib. For KIF11, the top candidates are BRD-K89997465, SB-743921, and ispinesib, and for TOP2A, the leading compounds are idarubicin, etoposide, and mitoxantrone.

AURKB, also known as aurora kinase B, is a serine/threonine kinase implicated in cell division. GSK-1070916 is noted as an Aurora kinase inhibitor [126,127,128]. BRD-K89997465 is known to down-regulate KIF11 expression through dopamine receptor antagonism, as indicated in the PubChem database [99]. TOP2A, or DNA topoisomerase II alpha, plays a role in DNA topology during transcription and is a known target of the anticancer drug idarubicin [129,130,131].

To assess the binding affinity between these compounds and their respective protein targets, we conducted molecular docking studies (Fig. 9A-C). The binding affinity of GSK-1070916 to AURKB is -6 kcal/mol, of BRD-K89997465 to KIF11 is -6.4 kcal/mol, and of idarubicin to TOP2A is -10 kcal/mol. Compounds with a free binding energy above − 5.5 kcal/mol are generally considered inactive [132,133,134]. Consequently, the most effective drug-protein interaction complexes are illustrated in Fig. 9.

A molecular dynamic simulation is a tool used to assess the protein-ligand system at the atomistic level and articulate the stability of the protein-ligand complex in a dynamic environment. In this study, a 100 ns simulation was performed to evaluate the complex’s stability. The RMSD of the protein and small compound drug were used to calculate the stability of the complex (Fig. 10A). The structural dynamics of the protein residues were investigated by calculating the RMSF values, which show the flexibility of protein residues in response to small compound drug binding during simulation (Fig. 10B).

Fig. 9
figure 9

The interactions between small compound drugs and proteins. A GSK-1070916 and AURKB. B BRD-K89997465 and KIF11. C Idarubicin and TOP2A

Fig. 10
figure 10

Molecular dynamic simulation of protein (KIF11) and small compound drug (BRD-K89997465). A RMSD plot for protein-small compound drug complex during 100 ns of molecular dynamics simulation. B RMSF trajectories of residues for protein-small compound drug complex

Discussion

In a matter of weeks, the COVID-19 pandemic rapidly spread to over 100 nations and territories, attracting substantial attention from the academic community and resulting in numerous studies and reports [1, 2, 135]. While some studies have proposed potential pathological mechanisms, therapeutic drugs, and target genes related to COVID-19, there remain several unresolved issues, such as drug side effects and the disease’s pathogenesis [8, 9, 37]. Particularly, there is a paucity of research on plasma cell biomarkers and therapeutic agents.

In this study, we focused on identifying and assessing candidate key genes and their associated drugs to enhance COVID-19 treatment, with a specific emphasis on plasma cells. Using cell infiltration analysis conducted via the xCell method, we observed significantly higher enrichment scores for immune cells in COVID-19 samples compared to normal controls. This finding aligns with the established conclusion that as COVID-19 advances, the virus-infected immune system triggers hyperinflammation or a cytokine storm, which can lead to severe organ damage and, in some cases, death [109]. Notably, we found a remarkably elevated enrichment score for plasma cells in COVID-19, underscoring their pivotal role in the disease’s development. Consequently, we employed Weighted Gene Co-expression Network Analysis (WGCNA) to identify co-expressed genes associated with COVID-19 and plasma cells. WGCNA, with its soft power approach, enhances the biological significance of strong correlations while diminishing weak or negative correlations, thus approximating a scale-free network [24, 25]. Simultaneously, we identified positive cluster biomarker genes highly expressed in plasma cells compared to the other 18 cell-type positive cluster biomarker genes within the plasma cell sub-population in COVID-19 samples, using single-cell RNA sequencing (scRNA-seq) analysis. These positive cluster biomarker genes are known to play pivotal roles in cell type-specific biological processes, including antiviral activity and immune response [136,137,138]. In-depth investigation of these genes aids in comprehending the pathogenic mechanisms and specific therapeutic targets, fostering the development of clinically relevant strategies. For instance, we identified IGKC as a positive cluster biomarker gene of plasma cells with the most significant fold change. IGKC is associated with the humoral immune response and has links to non-small cell lung cancer (NSCLC), breast cancer, and colon cancer [139,140,141]. Another positive cluster biomarker gene, IFNG-AS1, is implicated in COVID-19 and its severity, with its expression corresponding to increased inflammation levels [142, 143]. Furthermore, IFNG-AS1 is known to promote inflammation in ulcerative colitis [143].

Our systematic bioinformatics analysis unveiled shared genes among differentially expressed genes (DEGs), module genes, and positive cluster biomarker genes. In addition, we collected and organized the clinical characteristics of these five datasets, including three discovery datasets and two validation datasets, to test whether there is a significant effect between gender and COVID-19. The chi-square test of independence evaluates whether there is an association between the categories of the two variables. In this study, chisq.test() function when the number of samples is larger than 5 and fisher.test() function when the number of samples is smaller than 5, which was performed in R language to test the effect. The results validated there is no significant impact between gender and COVID-19 (p > 0.01, Supplementary Table 9). Given that proteins seldom function in isolation but rather form specific functional structures through interactions, these functional groups and complexes can be elucidated through network analysis of these interaction relationships [144, 145]. Therefore, we constructed a protein-protein interaction (PPI) network using Cytoscape. Within this network, we identified a sub-network comprising 20 hub genes (CCNA2, TOP2A, CCNB2, TPX2, BUB1, CDK1, KIF2C, RRM2, AURKB, CDCA8, KIF11, CCNB1, CDC20, UBE2C, KIF20A, PTTG1, TTK, DLGAP5, BUB1B, and PBK) based on Maximal Clique Centrality (MCC) values from the cytoHubba plugin. To validate the robustness of these 20 hub genes, we included two external independent bulk RNA-seq datasets with detailed clinical information and immune response data. Our analysis revealed significantly distinct expression patterns for these genes in COVID-19 patients, with higher expression levels in moderate, ICU, and severe cases compared to convalescent and healthy individuals. Notably, genes such as TOP2A, BUB1, KIF2C, AURKB, CDCA8, KIF11, PTTG1, and BUB1B demonstrated satisfactory classification ability in predicting the immune response to COVID-19. For instance, BUB1, also known as budding uninhibited by benzimidazole 1, encodes a serine/threonine-protein kinase that plays a central role in mitosis and has been implicated in the pathophysiological processes of COVID-19 [117, 146].

Furthermore, we identified three essential genes (AURKB, KIF11, and TOP2A) as potential therapeutic targets for COVID-19. Aurora kinase B (AURKB) belongs to the aurora kinase subfamily of serine/threonine kinases [147, 148]. Studies have demonstrated AURKB’s excellent classification performance in COVID-19 using common algorithms [148]. Kinesin family member 11 (KIF11) encodes a motor protein from the kinesin-like protein family, known to participate in various aspects of spindle dynamics and lung-related diseases [149]. Published results indicate that KIF11 is significantly up-regulated in lung adenocarcinoma and influences the cell cycle, tumor microenvironment alteration, and tumor-infiltrating immune cell proportions [150, 151]. DNA topoisomerase II alpha (TOP2A) produces an enzyme known as DNA topoisomerase, which regulates and modifies DNA’s topological states during transcription [152]. A study by Stephenson et al. found that TOP2A expression increased with the severity of COVID-19 according to single-cell multi-omics analysis [153]. Additionally, Samy et al. proposed TOP2A as a potential drug target, with digitoxin and hydroxyquinone being effective inhibitors against COVID-19 [154]. The literature review thus validates the reliability of our results, suggesting that controlling SARS-CoV-2 replication and excessive inflammation can be achieved by inhibiting these three genes.

We also identified three potential small compounds (GSK-1070916, BRD-K89997465, and idarubicin) with the lowest free binding energy, which were evaluated and visualized using cMAP and AutoDock. GSK-1070916, an Aurora inhibitor, has been reported to affect cell cycle progression and viability [127, 155]. Given AURKB’s involvement in chromosomal segregation and cytokinesis during mitosis, its inhibition by GSK-1070916 is particularly relevant [156]. A proteomics study by Bock et al. indicated that SARS-CoV-2 infection may impact numerous proteins related to inflammatory responses and chromosomal segregation [157]. Furthermore, DEGs between mild and severe COVID-19 cases were found to be enriched in cytokinesis failure, interferon response, and host translation inhibition [158]. Moreover, cytokinesis is a significant pathological mechanism underlying liver injury in COVID-19 [159]. BRD-K89997465, as recorded in the PubChem database, down-regulates KIF11 expression through a dopamine receptor antagonist mechanism [99]. KIF11, identified as a biomarker of COVID-19 through integrated blood RNA-seq and PPI network analysis, further underscores the potential significance of BRD-K89997465 [160, 161]. Idarubicin, an anthracycline anti-cancer drug used in the treatment of adult patients with acute myeloid leukemia, has also been found to inhibit SARS-CoV-2 endoribonuclease, particularly NSP15, a hexameric endoribonuclease unique to single-stranded RNA viruses [130, 162,163,164]. Our findings thus suggest that these three candidate small compounds positively impact COVID-19 treatment, with molecular docking models supporting their efficacy.

However, it is important to acknowledge the limitations of our study. It primarily relied on retrospective analysis, necessitating further research to ascertain the precise roles of the identified genes and potential treatment drugs. Additionally, in vitro and in vivo experiments are warranted to explore the immunotherapeutic efficacy of these interventions.

In conclusion, our study identified 122 shared genes among DEGs in COVID-19 and non-COVID-19 samples, co-expressed genes associated with both COVID-19 and plasma cell infiltration, and positive cluster biomarker genes for plasma cells in COVID-19 samples. We constructed a highly reliable PPI network comprising 89 genes and identified 20 hub genes with the highest MCC values, which were validated using external datasets. Furthermore, we proposed three potential small compounds (GSK-1070916, BRD-K89997465, and idarubicin) with promising therapeutic potential by targeting AURKB, KIF11, and TOP2A. These compounds offer a potential avenue for treating COVID-19 by modulating cell cycling and the hyperimmune response.

Availability of data and materials

The datasets collected in this research are from accessible databases. The article contains the accession number and the names of the databases. The code is available at https://github.com/liuzhe93/COVID19_plasma_cells.

Data availability

The datasets collected in this research are from accessible databases. The article contains the accession number and the names of the databases. Our study includes five independent datasets (GSE157103, GSE152641, GSE171524, GSE152418, and GSE179627). They are open and accessible in the NCBI GEO database (https://www.ncbi.nlm.nih.gov/geo/) by the above accession number.

Abbreviations

ARDS:

Acute Respiratory Distress Syndrome

AT1:

Alveolar Type I cells

AT2:

Alveolar Type II cells

AUC:

Area Under the Curve

BP:

Biological Process

CC:

Cell Component

cMAP:

connectivity Map

CI:

Confidence Interval

COVID-19:

Coronavirus Disease 2019

DEG:

Differentially Expressed Gene

DAVID:

Database for Annotation, Visualization, and Integrated Discovery

GEO:

Gene Expression Omnibus

GO:

Gene Ontology

H1N1:

Influenza A

H5N1:

Avian influenza A

KEGG:

Kyoto Encyclopedia of Genes and Genomes

LLPCs:

Long-lived Plasma Cells

MAD:

Median Absolute Deviation

MCC:

Matthews Correlation Coefficient

MF:

Molecular Function

NK cell:

Natural Killer cell

NSCLC:

Non-Small Cell Lung Cancer

PPI:

Protein-Protein Interaction

SARS:

Severe Acute Respiratory Syndrome

SARS-CoV-2:

Severe Acute Respiratory Syndrome Coronavirus 2

scRNA-seq:

single-cell RNA sequencing

Treg:

T regulatory cells

WGCNA:

Weighted Gene Co-Expression Network Analysis

WHO:

World Health Organization

References

  1. Kambouris ME. Global catastrophic biological risks in the post-COVID-19 World: Time to act is now. OMICS. 2023;27:153–70. https://doi.org/10.1089/omi.2022.0178.

    Article  CAS  PubMed  Google Scholar 

  2. Kupferschmidt K, Cohen J. Will novel virus go pandemic or be contained? Science. 2020;367:610–1. https://doi.org/10.1126/science.367.6478.610.

    Article  CAS  PubMed  Google Scholar 

  3. Guery B, Poissy J, el Mansouf L, Séjourné C, Ettahar N, Lemaire X, Vuotto F, Goffard A, Behillil S, Enouf V, Caro V, Mailles A, Che D, Manuguerra J-C, Mathieu D, Fontanet A, van der Werf S. MERS-CoV study group, clinical features and viral diagnosis of two cases of infection with middle east respiratory syndrome coronavirus: a report of nosocomial transmission. Lancet. 2013;381:2265–72. https://doi.org/10.1016/S0140-6736(13)60982-4.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Chan JF-W, Yuan S, Kok K-H, To KK-W, Chu H, Yang J, Xing F, Liu J, Yip CC-Y, Poon RW-S, Tsoi H-W, Lo SK-F, Chan K-H, Poon VK-M, Chan W-M, Ip JD, Cai J-P, Cheng VC-C, Chen H, Hui CK-M, Yuen K-Y. A familial cluster of pneumonia associated with the 2019 novel coronavirus indicating person-to-person transmission: a study of a family cluster. Lancet. 2020;395:514–23. https://doi.org/10.1016/S0140-6736(20)30154-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Harrison AG, Lin T, Wang P. Mechanisms of SARS-CoV-2 transmission and Pathogenesis. Trends Immunol. 2020;41:1100–15. https://doi.org/10.1016/j.it.2020.10.004.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Wang D, Hu B, Hu C, Zhu F, Liu X, Zhang J, Wang B, Xiang H, Cheng Z, Xiong Y, Zhao Y, Li Y, Wang X, Peng Z. Clinical characteristics of 138 hospitalized patients with 2019 Novel Coronavirus-infected pneumonia in Wuhan. China JAMA. 2020;323:1061–9. https://doi.org/10.1001/jama.2020.1585.

    Article  CAS  PubMed  Google Scholar 

  7. Chen N, Zhou M, Dong X, Qu J, Gong F, Han Y, Qiu Y, Wang J, Liu Y, Wei Y, Xia J, Yu T, Zhang X, Zhang L. Epidemiological and clinical characteristics of 99 cases of 2019 novel coronavirus pneumonia in Wuhan, China: a descriptive study. Lancet. 2020;395:507–13. https://doi.org/10.1016/S0140-6736(20)30211-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Ning Q, Wu D, Wang X, Xi D, Chen T, Chen G, Wang H, Lu H, Wang M, Zhu L, Hu J, Liu T, Ma K, Han M, Luo X. The mechanism underlying extrapulmonary complications of the coronavirus disease 2019 and its therapeutic implication. Signal Transduct Target Ther. 2022;7:57. https://doi.org/10.1038/s41392-022-00907-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Lamers MM, Haagmans BL. SARS-CoV-2 pathogenesis. Nat Rev Microbiol. 2022;20:270–84. https://doi.org/10.1038/s41579-022-00713-0.

    Article  CAS  PubMed  Google Scholar 

  10. Ren X, Wang S, Chen X, Wei X, Li G, Ren S, Zhang T, Zhang X, Lu Z, You Z, Wang Z, Song N, Qin C. Multiple expression assessments of ACE2 and TMPRSS2 SARS-CoV-2 Entry molecules in the urinary tract and their associations with Clinical manifestations of COVID-19. Infect Drug Resist. 2020;13:3977–90. https://doi.org/10.2147/IDR.S270543.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Senapati S, Banerjee P, Bhagavatula S, Kushwaha PP, Kumar S. Contributions of human ACE2 and TMPRSS2 in determining host-pathogen interaction of COVID-19. J Genet. 2021;100:100. https://doi.org/10.1007/s12041-021-01262-w.

    Article  CAS  Google Scholar 

  12. Serpeloni JM, Lima Neto QA, Lucio LC, Ramão A, Carvalho de Oliveira J, Gradia DF, Malheiros D, Ferrasa A, Marchi R, Figueiredo DLA, Silva WA, Ribeiro EMSF, Cólus IMS, Cavalli LR. Genome interaction of the virus and the host genes and non-coding RNAs in SARS-CoV-2 infection. Immunobiology. 2021;226:152130. https://doi.org/10.1016/j.imbio.2021.152130.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Huang S-W, Wang S-F. SARS-CoV-2 entry related viral and host genetic variations: implications on COVID-19 severity, Immune escape, and infectivity. Int J Mol Sci. 2021;22(6):3060. https://doi.org/10.3390/ijms22063060.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. de Pondé RA. Physicochemical effects of emerging exchanges on the spike protein’s RBM of the SARS-CoV-2 Omicron subvariants BA.1-BA.5 and its influence on the biological properties and attributes developed by these subvariants. Virology. 2023;587:109850. https://doi.org/10.1016/j.virol.2023.109850.

    Article  CAS  PubMed  Google Scholar 

  15. Chua RL, Lukassen S, Trump S, Hennig BP, Wendisch D, Pott F, Debnath O, Thürmann L, Kurth F, Völker MT, Kazmierski J, Timmermann B, Twardziok S, Schneider S, Machleidt F, Müller-Redetzky H, Maier M, Krannich A, Schmidt S, Balzer F, Liebig J, Loske J, Suttorp N, Eils J, Ishaque N, Liebert UG, von Kalle C, Hocke A, Witzenrath M, Goffinet C, Drosten C, Laudi S, Lehmann I, Conrad C, Sander L-E, Eils R. COVID-19 severity correlates with airway epithelium-immune cell interactions identified by single-cell analysis. Nat Biotechnol. 2020;38:970–9. https://doi.org/10.1038/s41587-020-0602-4.

    Article  CAS  PubMed  Google Scholar 

  16. Winkler ES, Bailey AL, Kafai NM, Nair S, McCune BT, Yu J, Fox JM, Chen RE, Earnest JT, Keeler SP, Ritter JH, Kang L-I, Dort S, Robichaud A, Head R, Holtzman MJ, Diamond MS. SARS-CoV-2 infection of human ACE2-transgenic mice causes severe lung inflammation and impaired function. Nat Immunol. 2020;21:1327–35. https://doi.org/10.1038/s41590-020-0778-2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Cerapio JP, Perrier M, Pont F, Tosolini M, Laurent C, Bertani S, Fournie J-J. Single-cell RNAseq profiling of human γδ T lymphocytes in Virus-related cancers and COVID-19 disease. Viruses. 2021;13:13. https://doi.org/10.3390/v13112212.

    Article  CAS  Google Scholar 

  18. Ferreras C, Pascual-Miguel B, Mestre-Durán C, Navarro-Zapata A, Clares-Villa L, Martín-Cortázar C, De Paz R, Marcos A, Vicario JL, Balas A, García-Sánchez F, Eguizabal C, Solano C, Mora-Rillo M, Soria B, Pérez-Martínez A. SARS-CoV-2-Specific memory T lymphocytes from COVID-19 convalescent donors: identification, Biobanking, and large-scale production for adoptive cell therapy. Front Cell Dev Biol. 2021;9:620730. https://doi.org/10.3389/fcell.2021.620730.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Mahmoodpoor A, Hosseini M, Soltani-Zangbar S, Sanaie S, Aghebati-Maleki L, Saghaleini SH, Ostadi Z, Hajivalili M, Bayatmakoo Z, Haji-Fatahaliha M, Babaloo Z, Farid SS, Heris JA, Roshangar L, Rikhtegar R, Kafil HS, Yousefi M. Reduction and exhausted features of T lymphocytes under serological changes, and prognostic factors in COVID-19 progression. Mol Immunol. 2021;138:121–7. https://doi.org/10.1016/j.molimm.2021.06.001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Fenoglio D, Dentone C, Parodi A, Di Biagio A, Bozzano F, Vena A, Fabbi M, Ferrera F, Altosole T, Bruzzone B, Giacomini M, Pelosi P, De Maria A, Bassetti M, De Palma R, Filaci G. Characterization of T lymphocytes in severe COVID-19 patients. J Med Virol. 2021;93:5608–13. https://doi.org/10.1002/jmv.27037.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Rasizadeh R, Aghbash PS, Nahand JS, Entezari-Maleki T, Baghi HB. SARS-CoV-2-associated organs failure and inflammation: a focus on the role of cellular and viral microRNAs. Virol J. 2023;20:179. https://doi.org/10.1186/s12985-023-02152-6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Vályi-Nagy I, Uher F, Rákóczi É, Szekanecz Z. Adaptive immunity to viruses: what did we learn from SARS-CoV-2 infection? Int J Mol Sci. 2022;23:13951. https://doi.org/10.3390/ijms232213951.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Meidaninikjeh S, Sabouni N, Marzouni HZ, Bengar S, Khalili A, Jafari R. Monocytes and macrophages in COVID-19: friends and foes. Life Sci. 2021;269:119010. https://doi.org/10.1016/j.lfs.2020.119010

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Zhang B, Horvath S. A general framework for weighted gene co-expression network analysis. Stat Appl Genet Mol Biol. 2005;4:Article17. https://doi.org/10.2202/1544-6115.1128.

    Article  PubMed  Google Scholar 

  25. Langfelder P, Horvath S. WGCNA: an R package for weighted correlation network analysis. BMC Bioinformatics. 2008;9:559. https://doi.org/10.1186/1471-2105-9-559.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Gu C, Li C, Wang W, Yan W, Yao Y, Shi M, Han F, Shang A. Immune Infiltration and N(6)-Methyladenosine ncRNA isoform detection in Acute Lung Injury. J Oncol. 2022;2022(3922299):1. https://doi.org/10.1155/2022/3922299.

    Article  CAS  Google Scholar 

  27. Zuo Q, Wang Y, Yang D, Guo S, Li X, Dong J, Wan C, Shen Y, Wen F. Identification of hub genes and key pathways in the emphysema phenotype of COPD. Aging. 2021;13:5120–35. https://doi.org/10.18632/aging.202432.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Kong X, Sun H, Wei K, Meng L, Lv X, Liu C, Lin F, Gu X. WGCNA combined with machine learning algorithms for analyzing key genes and immune cell infiltration in heart failure due to ischemic cardiomyopathy. Front Cardiovasc Med. 2023;10:1058834. https://doi.org/10.3389/fcvm.2023.1058834.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Guo J, Yu H, Yang J, Li Y, Zhang X, Mei K, Chi S, Yuan S, Liu H, El-Ashram S, Huang S, Wen F. Weighted gene co-expression network analysis revealed host transcriptional response to H1N1 influenza a virus infection. J Infect. 2021;82:e4-7. https://doi.org/10.1016/j.jinf.2020.12.017.

    Article  CAS  PubMed  Google Scholar 

  30. Hasankhani A, Bahrami A, Sheybani N, Aria B, Hemati B, Fatehi F, Ghaem Maghami Farahani H, Javanmard G, Rezaee M, Kastelic JP, Barkema HW. Differential co-expression network analysis reveals key hub-high traffic genes as potential therapeutic targets for COVID-19 pandemic. Front Immunol. 2021;12:789317. https://doi.org/10.3389/fimmu.2021.789317.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Jing S, Tian J, Zhang Y, Chen X, Zheng S. Identification of a new pseudogenes/lncRNAs-hsa-miR-26b-5p-COL12A1 competing endogenous RNA network associated with prognosis of pancreatic cancer using bioinformatics analysis. Aging. 2020;12:19107–28. https://doi.org/10.18632/aging.103709.

    Article  CAS  PubMed  Google Scholar 

  32. Liu W, Ye X, An Z, Zhao Z. The challenges and opportunities of scRNA-seq in COVID-19 research and clinical translation. Virol J. 2022;19:217. https://doi.org/10.1186/s12985-022-01923-x.

    Article  CAS  PubMed  Google Scholar 

  33. Ren X, Wen W, Fan X, Hou W, Su B, Cai P, Li J, Liu Y, Tang F, Zhang F, Yang Y, He J, Ma W, He J, Wang P, Cao Q, Chen F, Chen Y, Cheng X, Deng G, Deng X, Ding W, Feng Y, Gan R, Guo C, Guo W, He S, Jiang C, Liang J, Li Y-M, Lin J, Ling Y, Liu H, Liu J, Liu N, Liu S-Q, Luo M, Ma Q, Song Q, Sun W, Wang G, Wang F, Wang Y, Wen X, Wu Q, Xu G, Xie X, Xiong X, Xing X, Xu H, Yin C, Yu D, Yu K, Yuan J, Zhang B, Zhang P, Zhang T, Zhao J, Zhao P, Zhou J, Zhou W, Zhong S, Zhong X, Zhang S, Zhu L, Zhu P, Zou B, Zou J, Zuo Z, Bai F, Huang X, Zhou P, Jiang Q, Huang Z, Bei J-X, Wei L, Bian X-W, Liu X, Cheng T, Li X, Zhao P, Wang F-S, Wang H, Su B, Zhang Z, Qu K, Wang X, Chen J, Jin R, Zhang Z. COVID-19 immune features revealed by a large-scale single-cell transcriptome atlas. Cell. 2021;184:1895–913.

    Article  CAS  PubMed  Google Scholar 

  34. Dudchenko O, Ordovas-Montanes J, Bingle CD. Respiratory epithelial cell types, states and fates in the era of single-cell RNA-sequencing. Biochem J. 2023;480:921–39. https://doi.org/10.1042/BCJ20220572.

    Article  CAS  PubMed  Google Scholar 

  35. Zhang J, Lin D, Li K, Ding X, Li L, Liu Y, Liu D, Lin J, Teng X, Li Y, Liu M, Shen J, Wang X, He D, Shi Y, Wang D, Xu J. Transcriptome analysis of Peripheral Blood mononuclear cells reveals distinct Immune response in asymptomatic and re-detectable positive COVID-19 patients. Front Immunol. 2021;12:716075. https://doi.org/10.3389/fimmu.2021.716075.

    Article  CAS  PubMed  Google Scholar 

  36. van Dam S, Võsa U, van der Graaf A, Franke L, de Magalhães JP. Gene co-expression analysis for functional classification and gene-disease predictions. Brief Bioinform. 2018;19:575–92. https://doi.org/10.1093/bib/bbw139.

    Article  CAS  PubMed  Google Scholar 

  37. Navitha Reddy G, Jogvanshi A, Naikwadi S, Sonti R. Nirmatrelvir and Ritonavir combination: an antiviral therapy for COVID-19. Expert Rev Anti Infect Ther. 2023;21(9):943–55. https://doi.org/10.1080/14787210.2023.2241638.

    Article  CAS  PubMed  Google Scholar 

  38. Owen DR, Allerton CMN, Anderson AS, Aschenbrenner L, Avery M, Berritt S, Boras B, Cardin RD, Carlo A, Coffman KJ, Dantonio A, Di L, Eng H, Ferre R, Gajiwala KS, Gibson SA, Greasley SE, Hurst BL, Kadar EP, Kalgutkar AS, Lee JC, Lee J, Liu W, Mason SW, Noell S, Novak JJ, Obach RS, Ogilvie K, Patel NC, Pettersson M, Rai DK, Reese MR, Sammons MF, Sathish JG, Singh RSP, Steppan CM, Stewart AE, Tuttle JB, Updyke L, Verhoest PR, Wei L, Yang Q, Zhu Y. An oral SARS-CoV-2 Mpro inhibitor clinical candidate for the treatment of COVID-19. Science. 2021;374:1586–93. https://doi.org/10.1126/science.abl4784.

    Article  CAS  PubMed  Google Scholar 

  39. Vandyck K, Deval J. Considerations for the discovery and development of 3-chymotrypsin-like cysteine protease inhibitors targeting SARS-CoV-2 infection. Curr Opin Virol. 2021;49:36–40. https://doi.org/10.1016/j.coviro.2021.04.006.

    Article  CAS  PubMed  Google Scholar 

  40. Marzi M, Vakil MK, Bahmanyar M, Zarenezhad E. Paxlovid: mechanism of action, synthesis, and in Silico Study. Biomed Res Int. 2022;2022:7341493. https://doi.org/10.1155/2022/7341493.

    Article  CAS  PubMed  Google Scholar 

  41. Tian H, Yang C, Song T, Zhou K, Wen L, Tian Y, Tang L, Xu W, Zhang X. Efficacy and safety of paxlovid (nirmatrelvir/ritonavir) in the treatment of COVID-19: an updated meta-analysis and trial sequential analysis. Rev Med Virol. 2023;33(5):e2473.

    Article  CAS  PubMed  Google Scholar 

  42. Lamb YN, Ritonavir NP. First Approval. Drugs. 2022;82:585–91. https://doi.org/10.1007/s40265-022-01692-5.

    Article  CAS  PubMed  Google Scholar 

  43. Akhvlediani T, Bernard-Valnet R, Dias SP, Eikeland R, Pfausler B, Sellner J. Infectious disease panel of the European Academy of Neurology, neurological side effects and drug interactions of antiviral compounds against SARS-CoV-2. Eur J Neurol. 2023;30(12):3904–12. https://doi.org/10.1111/ene.16017.

    Article  PubMed  Google Scholar 

  44. Hashemian SMR, Sheida A, Taghizadieh M, Memar MY, Hamblin MR, Bannazadeh Baghi H, Sadri Nahand J, Asemi Z, Mirzaei H. Paxlovid (Nirmatrelvir/Ritonavir): a new approach to Covid-19 therapy? Biomed Pharmacother. 2023;162:114367. https://doi.org/10.1016/j.biopha.2023.114367.

    Article  CAS  PubMed  Google Scholar 

  45. Loos NHC, Beijnen JH, Schinkel AH. The inhibitory and inducing effects of ritonavir on hepatic and intestinal CYP3A and other drug-handling proteins. Biomed Pharmacother. 2023;162:114636. https://doi.org/10.1016/j.biopha.2023.114636.

    Article  CAS  PubMed  Google Scholar 

  46. Nappi F, Iervolino A, Avtaar Singh SS. Molecular insights of SARS-CoV-2 antivirals Administration: a balance between Safety profiles and Impact on Cardiovascular phenotypes. Biomedicines. 2022;10(2):437. https://doi.org/10.3390/biomedicines10020437.

    Article  CAS  PubMed  Google Scholar 

  47. Haque OI, Mahar S, Hussain S, Sloane P. Pharmacokinetic interaction between verapamil and ritonavir-boosted nirmatrelvir: implications for the management of COVID-19 in patients with hypertension. BMJ Case Rep. 2023;16(1):e252677. https://doi.org/10.1136/bcr-2022-252677.

    Article  PubMed  Google Scholar 

  48. Allen HC, Sharma P. Histology, Plasma Cells. 2023.

    Google Scholar 

  49. Duan K, Liu B, Li C, Zhang H, Yu T, Qu J, Zhou M, Chen L, Meng S, Hu Y, Peng C, Yuan M, Huang J, Wang Z, Yu J, Gao X, Wang D, Yu X, Li L, Zhang J, Wu X, Li B, Xu Y, Chen W, Peng Y, Hu Y, Lin L, Liu X, Huang S, Zhou Z, Zhang L, Wang Y, Zhang Z, Deng K, Xia Z, Gong Q, Zhang W, Zheng X, Liu Y, Yang H, Zhou D, Yu D, Hou J, Shi Z, Chen S, Chen Z, Zhang X, Yang X. Effectiveness of convalescent plasma therapy in severe COVID-19 patients. Proc Natl Acad Sci U S A. 2020;117:9490–6. https://doi.org/10.1073/pnas.2004168117.

    Article  CAS  PubMed  Google Scholar 

  50. Crowe JE. Principles of broad and potent antiviral human antibodies: insights for vaccine design. Cell Host Microbe. 2017;22:193–206. https://doi.org/10.1016/j.chom.2017.07.013.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Roback JD, Guarner J. Convalescent plasma to treat COVID-19: possibilities and challenges. JAMA. 2020;323:1561–2. https://doi.org/10.1001/jama.2020.4940.

    Article  CAS  PubMed  Google Scholar 

  52. Sato S, Hosokawa T, Kawashima H. Successful treatment of plasma exchange for refractory systemic juvenile idiopathic arthritis complicated with macrophage activation syndrome and severe lung disease. Ann Rheum Dis. 2022;81:e61. https://doi.org/10.1136/annrheumdis-2020-217390.

    Article  CAS  PubMed  Google Scholar 

  53. Schulz AR, Fiebig L, Hirseland H, Diekmann L-M, Reinke S, Hardt S, Niedobitek A, Mei HE. SARS-CoV-2 specific plasma cells acquire long-lived phenotypes in human bone marrow. EBioMedicine. 2023;95:104735. https://doi.org/10.1016/j.ebiom.2023.104735.

    Article  CAS  PubMed  Google Scholar 

  54. Turner JS, Kim W, Kalaidina E, Goss CW, Rauseo AM, Schmitz AJ, Hansen L, Haile A, Klebert MK, Pusic I, O’Halloran JA, Presti RM, Ellebedy AH. SARS-CoV-2 infection induces long-lived bone marrow plasma cells in humans. Nature. 2021;595:421–5. https://doi.org/10.1038/s41586-021-03647-4.

    Article  CAS  PubMed  Google Scholar 

  55. Banji D, Alqahtani SS, Banji OJF, Machanchery S, Shoaib A. Calming the inflammatory storm in severe COVID-19 infections: role of biologics- A narrative review. Saudi Pharm J. 2021;29:213–22. https://doi.org/10.1016/j.jsps.2021.01.005.

    Article  CAS  PubMed  Google Scholar 

  56. Sosa-Hernández VA, Torres-Ruíz J, Cervantes-Díaz R, Romero-Ramírez S, Páez-Franco JC, Meza-Sánchez DE, Juárez-Vega G, Pérez-Fragoso A, Ortiz-Navarrete V, Ponce-de-León A, Llorente L, Berrón-Ruiz L, Mejía-Domínguez NR, Gómez-Martín D, Maravillas-Montero JL. B cell subsets as severity-associated signatures in COVID-19 patients. Front Immunol. 2020;11:611004. https://doi.org/10.3389/fimmu.2020.611004.

    Article  CAS  PubMed  Google Scholar 

  57. Giannotta G, Giannotta N. mRNA. COVID-19 vaccines and long-lived plasma cells: a complicated relationship. Vaccines (Basel). 2021;9(12):1503. https://doi.org/10.3390/vaccines9121503.

    Article  CAS  PubMed  Google Scholar 

  58. Tehrani ZR, Habibzadeh P, Flinko R, Chen H, Abbasi A, Yared JA, Ciupe SM, Lewis GK, Sajadi MM. Deficient generation of spike-specific long-lived plasma cells in the bone marrow after severe acute respiratory syndrome coronavirus 2 infection. J Infect Dis. 2024;230(1):e30–3. https://doi.org/10.1093/infdis/jiad603.

    Article  PubMed  Google Scholar 

  59. Overmyer KA, Shishkova E, Miller IJ, Balnis J, Bernstein MN, Peters-Clarke TM, Meyer JG, Quan Q, Muehlbauer LK, Trujillo EA, He Y, Chopra A, Chieng HC, Tiwari A, Judson MA, Paulson B, Brademan DR, Zhu Y, Serrano LR, Linke V, Drake LA, Adam AP, Schwartz BS, Singer HA, Swanson S, Mosher DF, Stewart R, Coon JJ, Jaitovich A. Large-scale multi-omic analysis of COVID-19 severity. Cell Syst. 2021;12:23-e407. https://doi.org/10.1016/j.cels.2020.10.003.

    Article  CAS  PubMed  Google Scholar 

  60. Thair SA, He YD, Hasin-Brumshtein Y, Sakaram S, Pandya R, Toh J, Rawling D, Remmel M, Coyle S, Dalekos GN, Koutsodimitropoulos I, Vlachogianni G, Gkeka E, Karakike E, Damoraki G, Antonakos N, Khatri P, Giamarellos-Bourboulis EJ, Sweeney TE. Transcriptomic similarities and differences in host response between SARS-CoV-2 and other viral infections. IScience. 2021;24:101947. https://doi.org/10.1016/j.isci.2020.101947.

    Article  CAS  PubMed  Google Scholar 

  61. Melms JC, Biermann J, Huang H, Wang Y, Nair A, Tagore S, Katsyv I, Rendeiro AF, Amin AD, Schapiro D, Frangieh CJ, Luoma AM, Filliol A, Fang Y, Ravichandran H, Clausi MG, Alba GA, Rogava M, Chen SW, Ho P, Montoro DT, Kornberg AE, Han AS, Bakhoum MF, Anandasabapathy N, Suárez-Fariñas M, Bakhoum SF, Bram Y, Borczuk A, Guo XV, Lefkowitch JH, Marboe C, Lagana SM, Del Portillo A, Tsai EJ, Zorn E, Markowitz GS, Schwabe RF, Schwartz RE, Elemento O, Saqi A, Hibshoosh H, Que J, Izar B. A molecular single-cell lung atlas of lethal COVID-19. Nature. 2021;595:114–9. https://doi.org/10.1038/s41586-021-03569-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Arunachalam PS, Wimmers F, Mok CKP, Perera RAPM, Scott M, Hagan T, Sigal N, Feng Y, Bristow L, Tak-Yin Tsang O, Wagh D, Coller J, Pellegrini KL, Kazmin D, Alaaeddine G, Leung WS, Chan JMC, Chik TSH, Choi CYC, Huerta C, Paine M, McCullough H, Lv E, Anderson S, Edupuganti AA, Upadhyay SE, Bosinger HT, Maecker P, Khatri N, Rouphael M, Peiris B, Pulendran. Systems biological assessment of immunity to mild versus severe COVID-19 infection in humans. Science. 2020;369:1210–20. https://doi.org/10.1126/science.abc6261.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Yu G, Wang L-G, Han Y, He Q-Y. clusterProfiler: an R package for comparing biological themes among gene clusters. OMICS. 2012;16:284–7. https://doi.org/10.1089/omi.2011.0118.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Wu T, Hu E, Xu S, Chen M, Guo P, Dai Z, Feng T, Zhou L, Tang W, Zhan L, Fu X, Liu S, Bo X, Yu G. clusterProfiler 4.0: A universal enrichment tool for interpreting omics data. Innovation (Camb). 2021;2(3):100141. https://doi.org/10.1016/j.xinn.2021.100141.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Love MI, Huber W, Anders S. Moderated estimation of Fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014;15:550. https://doi.org/10.1186/s13059-014-0550-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Galon J, Costes A, Sanchez-Cabo F, Kirilovsky A, Mlecnik B, Lagorce-Pagès C, Tosolini M, Camus M, Berger A, Wind P, Zinzindohoué F, Bruneval P, Cugnenc P-H, Trajanoski Z, Fridman W-H, Pagès F. Type, density, and location of immune cells within human colorectal tumors predict clinical outcome. Science. 2006;313:1960–4. https://doi.org/10.1126/science.1129139.

    Article  CAS  PubMed  Google Scholar 

  67. Hanahan D, Coussens LM. Accessories to the crime: functions of cells recruited to the tumor microenvironment. Cancer Cell. 2012;21:309–22. https://doi.org/10.1016/j.ccr.2012.02.022.

    Article  CAS  PubMed  Google Scholar 

  68. Aran D, Hu Z, Butte AJ. xCell: digitally portraying the tissue cellular heterogeneity landscape. Genome Biol. 2017;18:220. https://doi.org/10.1186/s13059-017-1349-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Wang S, Yang C, Pan C, Feng X, Lei Z, Huang J, Wei X, Li F, Ma Y. Identification of key genes and functional enrichment pathways involved in fat deposition in Xinyang buffalo by WGCNA. Gene. 2022;818:146225. https://doi.org/10.1016/j.gene.2022.146225.

    Article  CAS  PubMed  Google Scholar 

  70. Shahan R, Zawora C, Wight H, Sittmann J, Wang W, Mount SM, Liu Z. Consensus Coexpression Network Analysis Identifies Key Regulators of Flower and Fruit Development in Wild Strawberry. Plant Physiol. 2018;178:202–16. https://doi.org/10.1104/pp.18.00086.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Zhang Y, Luo J, Liu Z, Liu X, Ma Y, Zhang B, Chen Y, Li X, Feng Z, Yang N, Feng D, Wang L, Song X. Identification of hub genes in colorectal cancer based on weighted gene co-expression network analysis and clinical data from the Cancer Genome Atlas. Biosci Rep. 2021;41:41. https://doi.org/10.1042/BSR20211280.

    Article  CAS  Google Scholar 

  72. Hao Y, Hao S, Andersen-Nissen E, Mauck WM, Zheng S, Butler A, Lee MJ, Wilk AJ, Darby C, Zager M, Hoffman P, Stoeckius M, Papalexi E, Mimitou EP, Jain J, Srivastava A, Stuart T, Fleming LM, Yeung B, Rogers AJ, McElrath JM, Blish CA, Gottardo R, Smibert P, Satija R. Integrated analysis of multimodal single-cell data. Cell. 2021;184:3573-e358729. https://doi.org/10.1016/j.cell.2021.04.048.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Stuart T, Butler A, Hoffman P, Hafemeister C, Papalexi E, Mauck WM, Hao Y, Stoeckius M, Smibert P, Satija R. Comprehensive Integration of single-cell data. Cel. 2019;177:1888-1902.e21. https://doi.org/10.1016/j.cell.2019.05.031.

    Article  CAS  Google Scholar 

  74. Butler A, Hoffman P, Smibert P, Papalexi E, Satija R. Integrating single-cell transcriptomic data across different conditions, technologies, and species. Nat Biotechnol. 2018;36:411–20. https://doi.org/10.1038/nbt.4096.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Satija R, Farrell JA, Gennert D, Schier AF, Regev A. Spatial reconstruction of single-cell gene expression data. Nat Biotechnol. 2015;33:495–502. https://doi.org/10.1038/nbt.3192.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Fei X, Liu J, Xu J, Jing H, Cai Z, Yan J, Wu Z, Li H, Wang Z, Shen Y. Integrating spatial transcriptomics and single-cell RNA-sequencing reveals the alterations in epithelial cells during nodular formation in benign prostatic hyperplasia. J Transl Med. 2024;22:380. https://doi.org/10.1186/s12967-024-05212-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Xia G, Zhu S, Liu Y, Pan J, Wang X, Shen C, Du A, Xu C. Transcriptomic profiling and regulatory pathways of cardiac resident macrophages in aging. Cell Mol Life Sci. 2024;81:220. https://doi.org/10.1007/s00018-024-05235-x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Pathak E, Atri N, Mishra R. Single-cell transcriptome analysis reveals the role of pancreatic secretome in COVID-19 Associated multi-organ dysfunctions. Interdiscip Sci. 2022;14:863–78. https://doi.org/10.1007/s12539-022-00513-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Chen J-W, Shrestha L, Green G, Leier A, Marquez-Lago TT. The hitchhikers’ guide to RNA sequencing and functional analysis. Brief Bioinform. 2023;24(1):bbac529. https://doi.org/10.1093/bib/bbac529.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Rosati D, Palmieri M, Brunelli G, Morrione A, Iannelli F, Frullanti E, Giordano A. Differential gene expression analysis pipelines and bioinformatic tools for the identification of specific biomarkers: a review. Comput Struct Biotechnol J. 2024;23:1154–68. https://doi.org/10.1016/j.csbj.2024.02.018.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. WASWA Munawar, MH Elias, FH Addnan, P Hassandarvish, S AbuBakar, N Roslan. Gene expression profiling of host lipid metabolism in SARS-CoV-2 infected patients: a systematic review and integrated bioinformatics analysis., BMC Infect Dis. 2024;24(1):124. https://doi.org/10.1186/s12879-024-08983-0.

  82. Li J, Zhang Y, Bai K-X, Qi X-J, Zhao Y, Bu H. Bioinformatics screening of gene expression profile and diagnostic application of meningeal carcinoma. Eur Rev Med Pharmacol Sci. 2023;27:9601–13. https://doi.org/10.26355/eurrev_202310_34132.

    Article  CAS  PubMed  Google Scholar 

  83. Zhang Z, Singh SP. A study on the analysis of important gene networks and pathways involved in progression of endometriosis to ovarian endometrioma cyst. Appl Biochem Biotechnol. 2023;196(7):4352–65. https://doi.org/10.1007/s12010-023-04778-2.

    Article  CAS  PubMed  Google Scholar 

  84. Li L, Hu Y, Li X, Ju B. A comprehensive analysis of the KLRB1 expression and its clinical implication in testicular germ cell tumors: a review. Medicine. 2024;103:e37688. https://doi.org/10.1097/MD.0000000000037688.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Sherman BT, Hao M, Qiu J, Jiao X, Baseler MW, Lane HC, Imamichi T, Chang W. DAVID: a web server for functional enrichment analysis and functional annotation of gene lists (2021 update). Nucleic Acids Res. 2022;50:W216–21. https://doi.org/10.1093/nar/gkac194.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Huang DW, Sherman BT, Lempicki RA. Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat Protoc. 2009;4:44–57. https://doi.org/10.1038/nprot.2008.211.

    Article  CAS  PubMed  Google Scholar 

  87. Lu H, Zhou Q, He J, Jiang Z, Peng C, Tong R, Shi J. Recent advances in the development of protein-protein interactions modulators: mechanisms and clinical trials. Signal Transduct Target Ther. 2020;5:213. https://doi.org/10.1038/s41392-020-00315-3.

    Article  PubMed  PubMed Central  Google Scholar 

  88. De Las Rivas J, Fontanillo C. Protein-protein interaction networks: unraveling the wiring of molecular machines within the cell. Brief Funct Genomics. 2012;11:489–96. https://doi.org/10.1093/bfgp/els036.

    Article  CAS  Google Scholar 

  89. Zheng Q, Lin R, Chen Y, Lv Q, Zhang J, Zhai J, Xu W, Wang W. SARS-CoV-2 induces cytokine storm hyperinflammatory responses in RA patients through pyroptosis. Front Immunol. 2022;13:1058884. https://doi.org/10.3389/fimmu.2022.1058884.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Hu Y, Yang Y, Fang Z, Hu Y-S, Zhang L, Wang J. Detecting pathway relationship in the context of human protein-protein interaction network and its application to Parkinson’s disease. Methods. 2017;131:93–103. https://doi.org/10.1016/j.ymeth.2017.08.001.

    Article  CAS  PubMed  Google Scholar 

  91. Szklarczyk D, Franceschini A, Wyder S, Forslund K, Heller D, Huerta-Cepas J, Simonovic M, Roth A, Santos A, Tsafou KP, Kuhn M, Bork P, Jensen LJ, von Mering C. STRING v10: protein-protein interaction networks, integrated over the tree of life. Nucleic Acids Res. 2015;43:D447-52. https://doi.org/10.1093/nar/gku1003.

    Article  CAS  PubMed  Google Scholar 

  92. Zhang Q, He M, Wang J, Liu S, Cheng H, Cheng Y. Predicting of disease genes for gestational diabetes mellitus based on network and functional consistency. Eur J Obstet Gynecol Reprod Biol. 2015;186:91–6. https://doi.org/10.1016/j.ejogrb.2014.12.016.

    Article  CAS  PubMed  Google Scholar 

  93. Bozhilova LV, Whitmore AV, Wray J, Reinert G, Deane CM. Measuring rank robustness in scored protein interaction networks. BMC Bioinformatics. 2019;20:446. https://doi.org/10.1186/s12859-019-3036-6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Shannon P, Markiel A, Ozier O, Baliga NS, Wang JT, Ramage D, Amin N, Schwikowski B, Ideker T. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 2003;13:2498–504. https://doi.org/10.1101/gr.1239303.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Chin C-H, Chen S-H, Wu H-H, Ho C-W, Ko M-T, Lin C-Y. cytoHubba: identifying hub objects and sub-networks from complex interactome. BMC Syst Biol. 2014;8 Suppl 4(Suppl 4):S11. https://doi.org/10.1186/1752-0509-8-S4-S11.

    Article  PubMed  Google Scholar 

  96. Trifonova OP, Lokhov PG, Archakov AI. Metabolic profiling of human blood. Biomed Khim. 2014;60:281–94. https://doi.org/10.18097/pbmc20146003281.

    Article  CAS  PubMed  Google Scholar 

  97. Chen Y, Liu H, Ning S, Wei C, Li J, Wei W, Zhang L. The high ratio of the plasma miR-96/miR-99b correlated with poor prognosis in patients with metastatic colorectal Cancer. Front Mol Biosci. 2021;8:799060. https://doi.org/10.3389/fmolb.2021.799060.

    Article  CAS  PubMed  Google Scholar 

  98. Subramanian A, Narayan R, Corsello SM, Peck DD, Natoli TE, Lu X, Gould J, Davis JF, Tubelli AA, Asiedu JK, Lahr DL, Hirschman JE, Liu Z, Donahue M, Julian B, Khan M, Wadden D, Smith IC, Lam D, Liberzon A, Toder C, Bagul M, Orzechowski M, Enache OM, Piccioni F, Johnson SA, Lyons NJ, Berger AH, Shamji AF, Brooks AN, Vrcic A, Flynn C, Rosains J, Takeda DY, Hu R, Davison D, Lamb J, Ardlie K, Hogstrom L, Greenside P, Gray NS, Clemons PA, Silver S, Wu X, Zhao W-N, Read-Button W, Wu X, Haggarty SJ, Ronco LV, Boehm JS, Schreiber SL, Doench JG, Bittker JA, Root DE, Wong B, Golub TR. A next generation connectivity map: L1000 platform and the first 1,000,000 profiles. Cell. 2017;171:1437–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Kim S, Chen J, Cheng T, Gindulyte A, He J, He S, Li Q, Shoemaker BA, Thiessen PA, Yu B, Zaslavsky L, Zhang J, Bolton EE. PubChem 2023 update. Nucleic Acids Res. 2023;51:D1373-80. https://doi.org/10.1093/nar/gkac956.

    Article  PubMed  Google Scholar 

  100. Trott O, Olson AJ. AutoDock Vina: improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J Comput Chem. 2010;31:455–61. https://doi.org/10.1002/jcc.21334.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Elkins JM, Santaguida S, Musacchio A, Knapp S. Crystal structure of human aurora B in complex with INCENP and VX-680. J Med Chem. 2012;55:7841–8. https://doi.org/10.1021/jm3008954.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Wendorff TJ, Schmidt BH, Heslop P, Austin CA, Berger JM. The structure of DNA-bound human topoisomerase II alpha: conformational mechanisms for coordinating inter-subunit interactions with DNA cleavage. J Mol Biol. 2012;424:109–24. https://doi.org/10.1016/j.jmb.2012.07.014.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Berman HM, Westbrook J, Feng Z, Gilliland G, Bhat TN, Weissig H, Shindyalov IN, Bourne PE. The protein data bank. Nucleic Acids Res. 2000;28:235–42. https://doi.org/10.1093/nar/28.1.235.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Burley SK, Bhikadiya C, Bi C, Bittrich S, Chao H, Chen L, Craig PA, Crichlow GV, Dalenberg K, Duarte JM, Dutta S, Fayazi M, Feng Z, Flatt JW, Ganesan S, Ghosh S, Goodsell DS, Green RK, Guranovic V, Henry J, Hudson BP, Khokhriakov I, Lawson CL, Liang Y, Lowe R, Peisach E, Persikova I, Piehl DW, Rose Y, Sali A, Segura J, Sekharan M, Shao C, Vallat B, Voigt M, Webb B, Westbrook JD, Whetstone S, Young JY, Zalevsky A, Zardecki C. RCSB Protein Data Bank (RCSB.org): delivery of experimentally-determined PDB structures alongside one million computed structure models of proteins from artificial intelligence/machine learning. Nucleic Acids Res. 2023;51:D488-508. https://doi.org/10.1093/nar/gkac1077.

    Article  CAS  PubMed  Google Scholar 

  105. Laskowski RA, Swindells MB. LigPlot+: multiple ligand-protein interaction diagrams for drug discovery. J Chem Inf Model. 2011;51:2778–86. https://doi.org/10.1021/ci200227u.

    Article  CAS  PubMed  Google Scholar 

  106. Wallace AC, Laskowski RA, Thornton JM. LIGPLOT: a program to generate schematic diagrams of protein-ligand interactions. Protein Eng. 1995;8:127–34. https://doi.org/10.1093/protein/8.2.127.

    Article  CAS  PubMed  Google Scholar 

  107. Wang Y, Zu G, Yu Y, Tang J, Han T, Zhang C. Curcumin’s mechanism of action against ischemic stroke: a network pharmacology and molecular dynamics study. PLoS ONE. 2023;18:e0280112. https://doi.org/10.1371/journal.pone.0280112.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Liu L, Jiao Y, Yang M, Wu L, Long G, Hu W. Network Pharmacology, Molecular Docking and Molecular Dynamics to explore the potential immunomodulatory mechanisms of deer Antler. Int J Mol Sci. 2023;24:24. https://doi.org/10.3390/ijms241210370.

    Article  CAS  Google Scholar 

  109. Yazdanpanah F, Hamblin MR, Rezaei N. The immune system and COVID-19: friend or foe? Life Sci. 2020;256:117900. https://doi.org/10.1016/j.lfs.2020.117900.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Yuan G-C, Cai L, Elowitz M, Enver T, Fan G, Guo G, Irizarry R, Kharchenko P, Kim J, Orkin S, Quackenbush J, Saadatpour A, Schroeder T, Shivdasani R, Tirosh I. Challenges and emerging directions in single-cell analysis. Genome Biol. 2017;18:84. https://doi.org/10.1186/s13059-017-1218-y.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Raje N, Anderson K, Einsele H, Efebera Y, Gay F, Hammond SP, Lesokhin AM, Lonial S, Ludwig H, Moreau P, Patel K, Ramasamy K, Mateos M-V. Monitoring, prophylaxis, and treatment of infections in patients with MM receiving bispecific antibody therapy: consensus recommendations from an expert panel. Blood Cancer J. 2023;13:116. https://doi.org/10.1038/s41408-023-00879-7.

    Article  PubMed  PubMed Central  Google Scholar 

  112. N.S. Brunetti, G.G. Davanzo, D. de Moraes, A.J.R. Ferrari, G.F. Souza, S.P. Muraro, T.L. Knittel, V.O. Boldrini, L.B. Monteiro, J.V. Virgílio-da-Silva, G.S. Profeta, N.S. Wassano, L. Nunes Santos, V.C. Carregari, A.H.S. Dias, F.P. Veras, L.A. Tavares, J. Forato, I.M.S. Castro, L.C. Silva-Costa, A.C. Palma, E. Mansour, R.G. Ulaf, A.F. Bernardes, T.A. Nunes, L.C. Ribeiro, M. V Agrela, M.L. Moretti, L.I. Buscaratti, F. Crunfli, R.G. Ludwig, J.A. Gerhardt, N. Munhoz-Alves, A.M. Marques, R. Sesti-Costa, M.R. Amorim, D.A. Toledo-Teixeira, P.L. Parise, M.C. Martini, K. Bispos-Dos-Santos, C.L. Simeoni, F. Granja, V.C. Silvestrini, E.B. de Oliveira, V.M. Faca, M. Carvalho, B.G. Castelucci, A.B. Pereira, L.D. Coimbra, M.M.G. Dias, P.B. Rodrigues, A.B.S.P. Gomes, F.B. Pereira, L.M.B. Santos, L.-M. Bloyet, S. Stumpf, M.C. Pontelli, S. Whelan, A.C. Sposito, R.F. Carvalho, A.S. Vieira, M.A.R. Vinolo, A. Damasio, L. Velloso, A.C.M. Figueira, L.L.P. da Silva, T.M. Cunha, H.I. Nakaya, H. Marques-Souza, R.E. Marques, D. Martins-de-Souza, M.S. Skaf, J.L. Proenca-Modena, P.M.M. Moraes-Vieira, M.A. Mori, A.S. Farias. SARS-CoV-2 uses CD4 to infect T helper lymphocytes. Elife 2023;12. https://doi.org/10.7554/eLife.84790.

  113. Santos AF, Póvoa P, Paixão P, Mendonça A, Taborda-Barata L. Changes in Glycolytic Pathway in SARS-COV 2 infection and their importance in understanding the severity of COVID-19. Front Chem. 2021;9:685196. https://doi.org/10.3389/fchem.2021.685196.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Alahdal M, Elkord E. Promising use of immune cell-derived exosomes in the treatment of SARS-CoV-2 infections. Clin Transl Med. 2022;12:e1026. https://doi.org/10.1002/ctm2.1026.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Brodin P. SARS-CoV-2 infections in children: understanding diverse outcomes. Immunity. 2022;55:201–9. https://doi.org/10.1016/j.immuni.2022.01.014.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Dai W, Guo C, Wang Y, Li Y, Xie R, Wu J, Yao B, Xie D, He L, Li Y, Huang H, Wang Y, Liu S. Identification of hub genes and pathways in lung metastatic colorectal cancer. BMC Cancer. 2023;23:323. https://doi.org/10.1186/s12885-023-10792-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Jin Q, Li W, Yu W, Zeng M, Liu J, Xu P. Analysis and identification of potential type II helper T cell (Th2)-Related key genes and therapeutic agents for COVID-19. Comput Biol Med. 2022;150:106134. https://doi.org/10.1016/j.compbiomed.2022.106134.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Sawyers CL. The cancer biomarker problem. Nature. 2008;452:548–52. https://doi.org/10.1038/nature06913.

    Article  CAS  PubMed  Google Scholar 

  119. Goossens N, Nakagawa S, Sun X, Hoshida Y. Cancer biomarker discovery and validation. Transl Cancer Res. 2015;4:256–69. https://doi.org/10.3978/j.issn.2218-676X.2015.06.04.

    Article  CAS  PubMed  Google Scholar 

  120. Bensalah K, Montorsi F, Shariat SF. Challenges of cancer biomarker profiling. Eur Urol. 2007;52:1601–9. https://doi.org/10.1016/j.eururo.2007.09.036.

    Article  PubMed  Google Scholar 

  121. Mäbert K, Cojoc M, Peitzsch C, Kurth I, Souchelnytskyi S, Dubrovska A. Cancer biomarker discovery: current status and future perspectives. Int J Radiat Biol. 2014;90:659–77. https://doi.org/10.3109/09553002.2014.892229.

    Article  CAS  PubMed  Google Scholar 

  122. Gurung AB, Ali MA, Lee J, Farah MA, Al-Anazi KM. An updated review of computer-aided drug design and its application to COVID-19. Biomed Res Int. 2021;2021:8853056. https://doi.org/10.1155/2021/8853056.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Basak SC. Recent developments and future directions at current computer aided drug design. Curr Comput Aided Drug Des. 2013;9:1. https://doi.org/10.2174/1573409911309010001.

    Article  CAS  PubMed  Google Scholar 

  124. Sabe VT, Ntombela T, Jhamba LA, Maguire GEM, Govender T, Naicker T, Kruger HG. Current trends in computer aided drug design and a highlight of drugs discovered via computational techniques: a review. Eur J Med Chem. 2021;224:113705. https://doi.org/10.1016/j.ejmech.2021.113705.

    Article  CAS  PubMed  Google Scholar 

  125. Meng X-Y, Zhang H-X, Mezei M, Cui M. Molecular docking: a powerful approach for structure-based drug discovery. Curr Comput Aided Drug Des. 2011;7:146–57. https://doi.org/10.2174/157340911795677602.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Wu Z-X, Yang Y, Wang J-Q, Zhou W-M, Chen J, Fu Y-G, Patel K, Chen Z-S, Zhang J-Y. Elevated ABCB1 expression confers acquired resistance to Aurora kinase inhibitor GSK-1070916 in Cancer cells. Front Pharmacol. 2020;11:615824. https://doi.org/10.3389/fphar.2020.615824.

    Article  CAS  PubMed  Google Scholar 

  127. Uitdehaag JCM, de Roos JADM, van Doornmalen AM, Prinsen MBW, Spijkers-Hagelstein JAP, de Vetter JRF, de Man J, Buijsman RC, Zaman GJR. Selective targeting of CTNBB1-, KRAS- or MYC-Driven cell growth by combinations of existing drugs. PLoS ONE. 2015;10:e0125021. https://doi.org/10.1371/journal.pone.0125021.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Chhajer R, Bhattacharyya A, Didwania N, Shadab M, Das N, Palit P, Vaidya T, Ali N. Leishmania Donovani Aurora kinase: a promising therapeutic target against visceral leishmaniasis. Biochim Biophys Acta. 2016;1860:1973–88. https://doi.org/10.1016/j.bbagen.2016.06.005.

    Article  CAS  PubMed  Google Scholar 

  129. Duan Y, Fu H, Huang J, Yin N, Liu L, Liu X. TOP2A deficiency leads to human recurrent spontaneous abortion and growth retardation of mouse pre-implantation embryos. Mol Med. 2022;28:165. https://doi.org/10.1186/s10020-022-00592-4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. de Campos-Nebel M, Larripa I, González-Cid M. Topoisomerase II-mediated DNA damage is differently repaired during the cell cycle by non-homologous end joining and homologous recombination. PLoS ONE. 2010;5:5. https://doi.org/10.1371/journal.pone.0012541.

    Article  CAS  Google Scholar 

  131. Kadia TM, Yang H, Ferrajoli A, Maddipotti S, Schroeder C, Madden TL, Holleran JL, Egorin MJ, Ravandi F, Thomas DA, Newsome W, Sanchez-Gonzalez B, Zwiebel JA, Espinoza-Delgado I, Kantarjian HM, Garcia-Manero G. A phase I study of vorinostat in combination with idarubicin in relapsed or refractory leukaemia. Br J Haematol. 2010;150:72–82. https://doi.org/10.1111/j.1365-2141.2010.08211.x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Eleftheriou P, Amanatidou D, Petrou A, Geronikaki A. Silico evaluation of the effectivity of approved protease inhibitors against the Main protease of the Novel SARS-CoV-2 Virus. Molecules. 2020;25:25. https://doi.org/10.3390/molecules25112529.

    Article  CAS  Google Scholar 

  133. Ganou CA, Eleftheriou PT, Theodosis-Nobelos P, Fesatidou M, Geronikaki AA, Lialiaris T, Rekka EA. Docking analysis targeted to the whole enzyme: an application to the prediction of inhibition of PTP1B by thiomorpholine and thiazolyl derivatives. SAR QSAR Environ Res. 2018;29:133–49. https://doi.org/10.1080/1062936X.2017.1414874.

    Article  CAS  PubMed  Google Scholar 

  134. Eleftheriou P, Petrou A, Geronikaki A, Liaras K, Dirnali S, Anna M. Prediction of enzyme inhibition and mode of inhibitory action based on calculation of distances between hydrogen bond donor/acceptor groups of the molecule and docking analysis: an application on the discovery of novel effective PTP1B inhibitors. SAR QSAR Environ Res. 2015;26:557–76. https://doi.org/10.1080/1062936X.2015.1074939.

    Article  CAS  PubMed  Google Scholar 

  135. Remuzzi A, Remuzzi G. COVID-19 and Italy: what next? Lancet. 2020;395:1225–8. https://doi.org/10.1016/S0140-6736(20)30627-9.

    Article  CAS  PubMed  Google Scholar 

  136. Zhao J, Jaffe A, Li H, Lindenbaum O, Sefik E, Jackson R, Cheng X, Flavell RA, Kluger Y. Detection of differentially abundant cell subpopulations in scRNA-seq data. Proc Natl Acad Sci U S A. 2021;118:118. https://doi.org/10.1073/pnas.2100293118.

    Article  CAS  Google Scholar 

  137. Wang G, Xiong Z, Yang F, Zheng X, Zong W, Li R, Bao Y. Identification of COVID-19-associated DNA methylation variations by integrating methylation array and scRNA-Seq data at cell-type resolution. Genes (Basel). 2022;13:1109. https://doi.org/10.3390/genes13071109.

    Article  CAS  PubMed  Google Scholar 

  138. Xu K, Zhang W, Wang C, Hu L, Wang R, Wang C, Tang L, Zhou G, Zou B, Xie H, Tang J, Guan X. Integrative analyses of scRNA-seq and scATAC-seq reveal CXCL14 as a key regulator of lymph node metastasis in breast cancer. Hum Mol Genet. 2021;30:370–80. https://doi.org/10.1093/hmg/ddab042.

    Article  CAS  PubMed  Google Scholar 

  139. Lohr M, Edlund K, Botling J, Hammad S, Hellwig B, Othman A, Berglund A, Lambe M, Holmberg L, Ekman S, Bergqvist M, Pontén F, Cadenas C, Marchan R, Hengstler JG, Rahnenführer J, Micke P. The prognostic relevance of tumour-infiltrating plasma cells and immunoglobulin kappa C indicates an important role of the humoral immune response in non-small cell lung cancer. Cancer Lett. 2013;333:222–8. https://doi.org/10.1016/j.canlet.2013.01.036.

    Article  CAS  PubMed  Google Scholar 

  140. Schmidt M, Böhm D, von Törne C, Steiner E, Puhl A, Pilch H, Lehr H-A, Hengstler JG, Kölbl H, Gehrmann M. The humoral immune system has a key prognostic impact in node-negative breast cancer. Cancer Res. 2008;68:5405–13. https://doi.org/10.1158/0008-5472.CAN-07-5206.

    Article  CAS  PubMed  Google Scholar 

  141. Fristedt R, Borg D, Hedner C, Berntsson J, Nodin B, Eberhard J, Micke P, Jirström K. Prognostic impact of tumour-associated B cells and plasma cells in oesophageal and gastric adenocarcinoma. J Gastrointest Oncol. 2016;7:848–59. https://doi.org/10.21037/jgo.2016.11.07.

    Article  PubMed  Google Scholar 

  142. Xu Y, Shao B. Circulating lncRNA IFNG-AS1 expression correlates with increased disease risk, higher disease severity and elevated inflammation in patients with coronary artery disease. J Clin Lab Anal. 2018;32:e22452. https://doi.org/10.1002/jcla.22452.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Padua D, Mahurkar-Joshi S, Law IKM, Polytarchou C, Vu JP, Pisegna JR, Shih D, Iliopoulos D, Pothoulakis C. A long noncoding RNA signature for ulcerative colitis identifies IFNG-AS1 as an enhancer of inflammation. Am J Physiol Gastrointest Liver Physiol. 2016;311:G446-457. https://doi.org/10.1152/ajpgi.00212.2016.

    Article  PubMed  PubMed Central  Google Scholar 

  144. Burke DF, Bryant P, Barrio-Hernandez I, Memon D, Pozzati G, Shenoy A, Zhu W, Dunham AS, Albanese P, Keller A, Scheltema RA, Bruce JE, Leitner A, Kundrotas P, Beltrao P, Elofsson A. Nat Struct Mol Biol. 2023;30:216–25. https://doi.org/10.1038/s41594-022-00910-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Spirin V, Mirny LA. Protein complexes and functional modules in molecular networks. Proc Natl Acad Sci U S A. 2003;100:12123–8. https://doi.org/10.1073/pnas.2032324100.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Agrawal P, Sambaturu N, Olgun G, Hannenhalli S, Path-Based A. Analysis of infected cell line and COVID-19 patient transcriptome reveals novel potential targets and drugs against SARS-CoV-2. Front Immunol. 2022;13:918817. https://doi.org/10.3389/fimmu.2022.918817.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Roshan P, Kuppa S, Mattice JR, Kaushik V, Chadda R, Pokhrel N, Tumala BR, Biswas A, Bothner B, Antony E, Origanti S. An Aurora B-RPA signaling axis secures chromosome segregation fidelity. Nat Commun. 2023;14:3008. https://doi.org/10.1038/s41467-023-38711-2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Kirsch-Volders M, Fenech M. Inflammatory cytokine storms severity may be fueled by interactions of micronuclei and RNA viruses such as COVID-19 virus SARS-CoV-2. A hypothesis. Mutat Res Rev Mutat Res. 2021;788:108395. https://doi.org/10.1016/j.mrrev.2021.108395.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Guo X, Zhou L, Wu Y, Li J. KIF11 as a potential Pan-cancer Immunological Biomarker encompassing the Disease Staging, Prognoses, Tumor Microenvironment, and therapeutic responses. Oxid Med Cell Longev. 2022;2022(2764940):1. https://doi.org/10.1155/2022/2764940.

    Article  CAS  Google Scholar 

  150. Li Z, Yu B, Qi F, Li F. KIF11 serves as an independent prognostic factor and therapeutic target for patients with lung adenocarcinoma. Front Oncol. 2021;11:670218. https://doi.org/10.3389/fonc.2021.670218.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Tabassum G, Singh P, Gurung R, Hakami MA, Alkhorayef N, Alsaiari AA, Alqahtani LS, Hasan MR, Rashid S, Kumar A, Dev K, Dohare R. Investigating the role of Kinesin family in lung adenocarcinoma via integrated bioinformatics approach. Sci Rep. 2023;13:9859. https://doi.org/10.1038/s41598-023-36842-6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Liu Y, Yu K, Zhang K, Niu M, Chen Q, Liu Y, Wang L, Zhang N, Li W, Zhong X, Li G, Wu S, Zhang J, Liu Y. O-GlcNAcylation promotes topoisomerase IIα catalytic activity in breast cancer chemoresistance. EMBO Rep. 2023;24:e56458. https://doi.org/10.15252/embr.202256458.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Stephenson E, Reynolds G, Botting RA, Calero-Nieto FJ, Morgan MD, Tuong ZK, Bach K, Sungnak W, Worlock KB, Yoshida M, Kumasaka N, Kania K, Engelbert J, Olabi B, Spegarova JS, Wilson NK, Mende N, Jardine L, Gardner LCS, Goh I, Horsfall D, McGrath J, Webb S, Mather MW, Lindeboom RGH, Dann E, Huang N, Polanski K, Prigmore E, Gothe F, Scott J, Payne RP, Baker KF, Hanrath AT, Schim van der Loeff ICD, Barr AS, Sanchez-Gonzalez A, Bergamaschi L, Mescia F, Barnes JL, Kilich E, de Wilton A, Saigal A, Saleh A, Janes SM, Smith CM, Gopee N, Wilson C, Coupland P, Coxhead JM, Kiselev VY, van Dongen S, Bacardit J, King HW, Rostron AJ, Simpson AJ, Hambleton S, Laurenti E, Lyons PA, Meyer KB, Nikolić MZ, Duncan CJA, Smith KGC, Teichmann SA, Clatworthy MR, Marioni JC, Göttgens B, Haniffa M, Cambridge Institute of Therapeutic Immunology and Infectious Disease-National Institute of Health Research (CITIID-NIHR) COVID-19 BioResource Collaboration. Single-cell multi-omics analysis of the immune response in COVID-19. Nat Med. 2021;27:904–16. https://doi.org/10.1038/s41591-021-01329-2.

    Article  CAS  PubMed  Google Scholar 

  154. Samy A, Maher MA, Abdelsalam NA, Badr E. SARS-CoV-2 potential drugs, drug targets, and biomarkers: a viral-host interaction network-based analysis. Sci Rep. 2022;12:11934. https://doi.org/10.1038/s41598-022-15898-w.

    Article  CAS  PubMed  Google Scholar 

  155. Adams ND, Adams JL, Burgess JL, Chaudhari AM, Copeland RA, Donatelli CA, Drewry DH, Fisher KE, Hamajima T, Hardwicke MA, Huffman WF, Koretke-Brown KK, Lai ZV, McDonald OB, Nakamura H, Newlander KA, Oleykowski CA, Parrish CA, Patrick DR, Plant R, Sarpong MA, Sasaki K, Schmidt SJ, Silva DJ, Sutton D, Tang J, Thompson CS, Tummino PJ, Wang JC, Xiang H, Yang J, Dhanak D. Discovery of GSK1070916, a potent and selective inhibitor of Aurora B/C kinase. J Med Chem. 2010;53:3973–4001. https://doi.org/10.1021/jm901870q.

    Article  CAS  PubMed  Google Scholar 

  156. Kallio MJ, McCleland ML, Stukenberg PT, Gorbsky GJ. Inhibition of aurora B kinase blocks chromosome segregation, overrides the spindle checkpoint, and perturbs microtubule dynamics in mitosis. Curr Biol. 2002;12:900–5. https://doi.org/10.1016/s0960-9822(02)00887-4.

    Article  CAS  PubMed  Google Scholar 

  157. Bock J-O, Ortea I. Re-analysis of SARS-CoV-2-infected host cell proteomics time-course data by impact pathway analysis and network analysis: a potential link with inflammatory response. Aging. 2020;12:11277–86. https://doi.org/10.18632/aging.103524.

    Article  CAS  PubMed  Google Scholar 

  158. de Cavalcante LT, da Fonseca GC, Amado Leon LA, Salvio AL, Brustolini OJ, Gerber AL, de Guimarães AP, Marques CAB, Fernandes RA, Ramos Filho CHF, Kader RL, Pimentel Amaro M, da Costa JP, Gonçalves S, Vieira Alves-Leon ATR, Vasconcelos. Buffy Coat Transcriptomic Analysis reveals alterations in host cell protein synthesis and cell cycle in severe COVID-19 patients. Int J Mol Sci. 2022;23. https://doi.org/10.3390/ijms232113588.

  159. Su Y-J, Chang C-W, Chen M-J, Lai Y-C. Impact of COVID-19 on liver. World J Clin Cases. 2021;9:7998–8007. https://doi.org/10.12998/wjcc.v9.i27.7998.

    Article  PubMed  Google Scholar 

  160. Ghosh N, Saha I, Plewczynski D. Unveiling the biomarkers of Cancer and COVID-19 and their regulations in different organs by integrating RNA-Seq expression and protein-protein interactions. ACS Omega. 2022;7:43589–602. https://doi.org/10.1021/acsomega.2c04389.

    Article  CAS  PubMed  Google Scholar 

  161. Noor F, Ashfaq UA, Bakar A, Ul Haq W, Allemailem KS, Alharbi BF, Al-Megrin WAI. Tahir Ul Qamar, discovering common pathogenic processes between COVID-19 and HFRS by integrating RNA-seq differential expression analysis with machine learning. Front Microbiol. 2023;14:1175844. https://doi.org/10.3389/fmicb.2023.1175844.

    Article  PubMed  Google Scholar 

  162. Dean EA, Brown RA, Kaur P, Casaus DV. Viral clearance with Neutrophil Recovery in a patient with active COVID-19 infection and refractory Acute Myeloid Leukemia who underwent successful reinduction with Cytarabine/Idarubicin. Case Rep Oncol. 2022;15:705–12. https://doi.org/10.1159/000525766.

    Article  CAS  PubMed  Google Scholar 

  163. Al-Rashedi NAM, Munahi MG, Ah L, ALObaidi. Prediction of potential inhibitors against SARS-CoV-2 endoribonuclease: RNA immunity sensing. J Biomol Struct Dyn. 2022;40:4879–92. https://doi.org/10.1080/07391102.2020.1863265.

    Article  CAS  PubMed  Google Scholar 

  164. Chandra A, Gurjar V, Qamar I, Singh N. Identification of potential inhibitors of SARS-COV-2 endoribonuclease (EndoU) from FDA approved drugs: a drug repurposing approach to find therapeutics for COVID-19. J Biomol Struct Dyn. 2021;39:4201–11. https://doi.org/10.1080/07391102.2020.1775127.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors would like to express our appreciation to the computational clusters source supported by the Department of Computer Science, City University of Hong Kong. This research was substantially sponsored by the research project (Grant No. 32170654 and Grant No. 32000464) supported by the National Natural Science Foundation of China and was substantially supported by the Shenzhen Research Institute, City University of Hong Kong. The work described in this paper was substantially supported by the grant from the Research Grants Council of the Hong Kong Special Administrative Region [CityU 11203723]. The work described in this paper was partially supported by the grants from City University of Hong Kong (2021SIRG036, CityU 9667265, CityU 11203221) and Innovation and Technology Commission (ITB/FBL/9037/22/S).

Conflict of interest disclosure

All authors disclosed no relevant relationships. The authors reported no potential conflicts of interest.

Patient consent statement

Not applicable.

Permission to reproduce material from other sources

Not applicable.

Clinical trial registration

Not applicable.

Funding

This research was substantially sponsored by the research project (Grant No. 32170654 and Grant No. 32000464) supported by the National Natural Science Foundation of China and was substantially supported by the Shenzhen Research Institute, City University of Hong Kong. The work described in this paper was substantially supported by the grant from the Research Grants Council of the Hong Kong Special Administrative Region [CityU 11203723]. The work described in this paper was partially supported by the grants from City University of Hong Kong (2021SIRG036, CityU 9667265, CityU 11203221) and Innovation and Technology Commission (ITB/FBL/9037/22/S).

Author information

Authors and Affiliations

Authors

Contributions

Zhe Liu: Data curation, Formal analysis, Investigation, Methodology, and Roles/Writing - original draft. Olutomilayo Olayemi Petinrin: Software, Formal analysis, and Data curation. Nanjun Chen: Data curation, Formal analysis, Methodology, Writing – review and editing. Muhammad Toseef: Validation, and Data curation.: Resources, and Writing - review and editing. Fang Liu: Formal analysis, Investigation, Methodology. Zhongxu Zhu: Conceptualization, Project administration, Resources, Supervision, and Writing – review and editing. Furong Qi: Conceptualization, Project administration, Resources, Supervision, and Writing - review and editing. Ka-Chun Wong: Conceptualization, Funding acquisition, Project administration, Resources, Supervision, and Writing - review and editing.

Corresponding authors

Correspondence to Zhongxu Zhu, Furong Qi or Ka-Chun Wong.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Conflict of interest

All authors disclosed no relevant relationships. The authors reported no potential conflicts of interest.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

12879_2024_10000_MOESM1_ESM.pdf

Additional file 1: Supplementary Figure 1. The PCA plot before and after batch effect correction. (A) Before removing the batch effect. (B) After removing the batch effect.

12879_2024_10000_MOESM2_ESM.pdf

Additional file 2: Supplementary Figure 2. The inter-group t-test of 16 cell type proportions between COVID-19 and normal samples. (A) Airway epithelial cells. (B) CD8+ T cells. (C) AT1. (D) NK cells. (E) Smooth muscle. (F) AT2. (G) Fibroblasts. (H) CD4+ T cells. (I) Dendritic cells. (J) Cycling NK/T cells. (K) Endothelial cells. (L) B cells. (M) Other epithelial cells. (N) Mast cells. (O) Neuronal cells. (P) Tregs.

12879_2024_10000_MOESM3_ESM.pdf

Additional file 3: Supplementary Figure 3. The normalized expression levels of 15 hub genes across different COVID-19 cohorts, convalescent, healthy, ICU, moderate, and severe. The 15 hub genes areBUB1B (A), CCNA2 (B), CCNB1 (C), CCNB2 (D), CDC20 (E), CDK1 (F), DLGAP5 (G), KIF2C (H), KIF20A (I), PBK (J), RRM2 (K), TOP2A (L), TPX2 (M), TTK (N), and UBE2C (O), respectively.

12879_2024_10000_MOESM4_ESM.xlsx

Additional file 4: Supplementary Table 1. The DEG list and normalized expressed values between COVID-19 and non-COVID-19 samples.

12879_2024_10000_MOESM5_ESM.xlsx

Additional file 5: Supplementary Table 2. The immune cell infiltration score of 64 immune cell types across 212 COVID-19 samples.

Additional file 6: Supplementary Table 3. The module gene list.

Additional file 7: Supplementary Table 4. The positive cluster biomarker gene list.

12879_2024_10000_MOESM8_ESM.xlsx

Additional file 8: Supplementary Table 5. The 122 shared genes among DEGs, module genes, and positive cluster biomarker genes.

12879_2024_10000_MOESM9_ESM.xlsx

Additional file 9: Supplementary Table 6. The top ten potential small molecule compounds or drugs that down-regulate AURKB expression.

12879_2024_10000_MOESM10_ESM.xlsx

Additional file 10: Supplementary Table 7. The top ten potential small molecule compounds or drugs that down-regulate KIF11 expression.

Additional file 11: Supplementary Table 9. The correlation analysis between gender and COVID-19.

Additional file 12: Supplementary Table 9. The correlation analysis between gender and COVID-19.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, Z., Petinrin, O.O., Chen, N. et al. Identification and evaluation of candidate COVID-19 critical genes and medicinal drugs related to plasma cells. BMC Infect Dis 24, 1099 (2024). https://doi.org/10.1186/s12879-024-10000-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12879-024-10000-3

Keywords