Next Article in Journal
The Role of Plasmacytoid Dendritic Cells in the Immune Contexture of TP53-Mutated High-Grade Serous Ovarian Cancer
Previous Article in Journal
Exploring the Coordination of Cancer Care for Teenagers and Young Adults in England and Wales: BRIGHTLIGHT_2021 Rapid Qualitative Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CRISPR-Cas9 Genome and Double-Knockout Screening to Identify Novel Therapeutic Targets for Chemoresistance in Triple-Negative Breast Cancer

1
Department of Biomedical Informatics, College of Medicine, The Ohio State University, Columbus, OH 43210, USA
2
College of Pharmacy, The Ohio State University, Columbus, OH 43210, USA
3
Division of Medical Oncology, Department of Medicine, The Ohio State University, Columbus, OH 43210, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cancers 2025, 17(23), 3876; https://doi.org/10.3390/cancers17233876
Submission received: 22 October 2025 / Revised: 26 November 2025 / Accepted: 27 November 2025 / Published: 3 December 2025
(This article belongs to the Section Cancer Therapy)

Simple Summary

Triple-negative breast cancer (TNBC) accounts for 15 to 20% of breast cancer cases and contributes to a disproportionate 35% of breast cancer deaths. Its resistance to chemotherapy presents a significant challenge. In this study, we identified synthetic lethal targets in treating TNBC with cisplatin and doxorubicin through a genome-wide CRISPR-Cas9 screening.

Abstract

Background/Objectives: Triple-negative breast cancer (TNBC) accounts for 15 to 20% of breast cancer cases and contributes to a disproportionate 35% of breast cancer deaths. Its resistance to chemotherapy presents a significant challenge. Methods: We firstly compared transcriptomic profiles between TNBC cell lines and patient samples and inferred the MDA-MB-231 cell line as the most representative model for TNBC with poor response to chemotherapy. We then conducted a genome-wide CRISPR-Cas9 screening and RNA-seq analysis in MDA-MB-231. Results: This analysis revealed 96 and 93 genes that could re-sensitize cisplatin and doxorubicin treatment, respectively, with 19 overlapping genes. Among these genes, 28 have been studied and published previously in chemoresistance research. MCM9 was found as a new TNBC chemoresistance target. Its target drug, KPT-185, showed an additive effect with cisplatin in treating TNBC cells. In the follow-up gene combination double-knockout experiment among 65 genes selected from cell death pathways, 242 synthetic lethal gene pairs were discovered to overcome chemoresistance in TNBC. Conclusions: In this study, we identified synthetic lethal targets in treating TNBC with cisplatin and doxorubicin through a genome-wide CRISPR-Cas9 screening and gene combination double-knockout screening.

1. Introduction

Triple-negative breast cancer (TNBC) is distinguished by the absence of the estrogen receptor (ER), progesterone receptor (PR), and human epidermal growth factor receptor 2 (HER2) [1], three pivotal receptors that substantially influence the proliferation and dissemination of breast cancer cells. In most breast cancer cases, therapeutic interventions target one or more of these receptors to impede or halt cancer progression, and their absence in TNBC presents a particularly formidable challenge to treatment [2].
The conventional treatment regimen for TNBC comprises a combination of surgical intervention, radiotherapy, and chemotherapy [3]. Cisplatin and doxorubicin, two front-line chemotherapeutic agents used to manage TNBC, both function to kill cancer cells by damaging their DNA. Platinum-based cisplatin does this by forming covalent bonds with the DNA [4], and doxorubicin, an anthracycline, encompasses DNA double helix intercalation, topoisomerase II inhibition, and free radical generation [5]. However, though these chemotherapeutic drugs can effectively eradicate tumor cells, patients with TNBC have demonstrated resistance to them [6,7]. Large clinical trials have shown residual cancer in approximately half of patients with TNBC following neoadjuvant chemotherapy (NACT) [8,9], and around 40% of those with residual disease will eventually develop distant metastasis [10]. Thus, it is important to identify more effective treatment strategies for patients with TNBC who respond poorly to these chemotherapeutic drugs.
In recent years, significant progress in understanding the molecular characteristics of TNBC has led to the development of potential new targeted and personalized therapeutic options for these patients, including the use of immunotherapies, poly (ADP-ribose) polymerase (PARP) inhibitors, antibody–drug conjugates (ADCs), and targeted therapies based on specific molecular mechanisms. Such immunotherapies as atezolizumab (Tecentriq®) [11] have shown encouraging results in combination with nab-paclitaxel in advanced triple-negative breast cancer, and PARP inhibitors like olaparib (Lynparza®) [12], niraparib (Zejula), and talazoparib (Talzenna®) [13] have shown promise, particularly when combined with such chemotherapeutic agents as paclitaxel, as first- or second-line treatment in patients with metastatic TNBC. ADCs [14,15], such as sacituzumab govitecan (Trodelvy®), selectively deliver cytotoxic agents to cancer cells while minimizing damage to healthy cells, and targeted therapies based on specific molecular mechanisms [16], like phosphoinositide 3-kinase (PI3K) inhibitors, e.g., alpelisib, buparlisib, and mammalian target of rapamycin (mTOR) inhibitors, e.g., everolimus, offer new avenues for TNBC treatment. The emerging cancer treatment capitalizes on the concept of synthetic lethality (SL), which occurs when the simultaneous inhibition or disruption of two genes results in cell death [17]. The SL between BRAC and PARP stems from the mechanism whereby it controls two complementary pathways in DNA homologous recombination repair [18]. Therefore, PARP inhibitors, olaparib and talazoparib, are promising for treating breast cancer patients with BRCA1 or BRCA2 mutations [19].
The CRISPR/Cas9 system is a powerful gene editing tool that disrupts or modifies gene functions [20]. Genome-wide CRISPR screening allows researchers to examine the entire genome systematically to identify genes whose knockout or modification can lead to particular phenotypes, such as increased sensitivity to chemotherapy or SL combined with another genetic disruption [21,22]. In this paper, for the first time, we conduct genome-wide CRISPR screening using the Toronto Knockout CRISPR library (TKOV3; Version 3) to identify potential druggable targets capable of overcoming the resistance in patients with TNBC to DNA-damaging chemotherapeutic agents, such as cisplatin and doxorubicin. Specifically, we investigate SL between new druggable targets and chemotherapeutic drugs that can overcome chemoresistance. We also draw genes from the XDeathDB database [23] to create a CRISPR double-knockout library [24] for cell death. Our strategy focuses on discovering SL gene pairs that could potentially serve as drug targets for TNBC.
Prior to initiating the CRISPR-Cas9 screening experiment, our selection of an appropriate and representative cell-line model was crucial. An improperly selected cell model that is not representative of TNBC-chemoresistant patients would prevent translational target and drug discoveries from preclinical to clinical research. Indeed, many drugs with promising preclinical results failed in clinical trials because the cell-line models do not accurately represent patients [25,26]. PubMed citations indicate that 40.2% of metastatic breast cancer research employs the MDA-MB-231 cell-line model. However, Chen’s research group [27] observed less resemblance of MDA-MB-231 to basal-like metastatic breast cancer patients. In a separate study of ovarian cancer, Domcke’s team [28] compared genomic profiles of 47 ovarian cancer cell lines with ovarian cancer tumor samples and found greater resemblance of several lesser-known cell lines with high-grade serous ovarian tumor samples than other widely used cell lines. Therefore, in this paper, proper TNBC chemoresistance cell models were selected before extensive target and drug discovery experiments using CRISPR-Cas9 screening technologies.

2. Materials and Methods

2.1. Triple Negative Breast Cancer Transcriptome Data Collection

Our reported query of the Gene Expression Omnibus (GEO) [29] using the keywords “breast cancer” and “chemo” to conduct pathway and sub-pathway analyses revealed the activation of several molecular pathways associated with chemoresistance in patients with breast cancer [29]. Tumor samples were categorized as either chemo-sensitive or -resistant according to the Miller–Payne (MP) five-grade system for assessing patient resistance to chemotherapy, which serves as the basis for determining chemoresistance in accordance with the standard cancer treatment protocol. In this paper, we focus on TNBC samples from three cohorts (Table 1). We also obtained genomic profiles of 21 TNBC cell lines from 916 cancer cells in the CCLE [30]. All samples were generated using the same Affymetrix Human Genome U133 Plus 2.0 Array platform (Affymetrix, Santa Clara, CA, USA).

2.2. Bioinformatics Data Analyses of Transcriptome Between TNBC Cell Lines and Tumor Samples

2.2.1. Pre-Processing of Data for Gene Expression Profiles

We analyzed Affymetrix U133 Plus 2.0 microarray data after robust multiarray average (RMA) normalization using Expression Console™ (EC) software version 1.1 (Affymetrix). The signal intensity was logged with base 2 to stabilize variance. Probes were removed if expression levels in 80% of the samples were lower than the background noise. Affymetrix probe IDs were mapped to gene symbols based on their General Public License (GPL) platforms. When multiple probes mapped to a single gene, the median expression value was used in the analysis.

2.2.2. Bach Effect Removal

Because our samples originated from various studies, to ensure an accurate and unbiased analysis, we employed the limma package in R [34] to correct for batch effects.

2.2.3. Different Gene Expression and Hierarchical Cluster Analysis

We analyzed differential gene expression utilizing the limma package in R software (version 3.6.1) [35], adjusted nominal p-values using the Benjamini–Hochberg method, and calculated fold change by comparing mean expressions between non-response and response groups. We conducted multiple analyses for various purposes, including comparisons between non-response and response in baseline subjects (sample size: 29:15), non-response and response in post-chemotherapy subjects (sample size: 11:4), and post-chemotherapy and baseline conditions (sample size: 29:11). The hierarchical clustering analysis of TNBC clinical transcriptome data and CCLE cell line transcriptome data was performed in Heatmap in R.

2.2.4. Correlation Analysis Between TNBC Cell Lines and TNBC Patient Samples Who Were Poor Responders to Chemotherapy

Affymetrix data in the CCLE dataset include over 10,000 genes across 916 cell lines. We first calculated the standard deviation of each gene across the 916 cell lines to ascertain variation in these genes in all CCLE cancer cells and selected the top 2000 variation genes for the follow-up analysis. We then calculated Spearman correlations in these 2000 genes between a given cell line and all the TNBC poor responders, with an average of all the correlations representative of the similarity between the cancer cell and poor responder population. A higher average correlation indicates better representation.

2.2.5. ssGSEA Pathway Similarity Analysis

We obtained ssGSEA scores using R package gene set variation analysis (GSVA) [36] for the 50 MSigDB cancer hallmark gene sets [37]. Hierarchical clustering of ssGSEA scores were explored between TNBC cell lines and TNBC poor responder samples in the context of the cancer hallmark gene set.

2.2.6. Gene Ontology (GO) and KEGG Pathway Analysis

GO [38] and Kyoto Encyclopedia of Genes and Genomes (KEGG) [39] gene enrichment analyses were used in identifying and interpreting complex molecular functions, biological processes, cellular components, and signaling pathways in drug resistance genes. These analyses were performed in clusterProfiler in R [40]. The threshold of statistically significance was a p-value below 0.05.

2.3. Cell Culture

Human TNBC cell lines MDA-MB-231, MDA-MB-436, and HS-578T, along with the human embryonic kidney (HEK) cell line 293T, were obtained from the American Type Culture Collection (Manassas, VA, USA) for this study. All cell lines were cultured in Ham’s F-12K (Kaighn’s) medium, supplemented with 10% fetal bovine serum (VWR, Radnor, PA, USA), 1% GlutMax, 1% sodium pyruvate, and penicillin–streptomycin (Gibco, Waltham, MA, USA). The cell lines were incubated at 37 °C in a 5% CO2 atmosphere. All cell lines underwent authentication via short tandem repeat (STR) profiling and were tested for mycoplasma contamination every three months.

2.3.1. Cell Survival Assay Using siRNA-Mediated Gene Silencing

We used small interfering RNAs purchased from Thermo Fisher Scientific (Waltham, MA, USA) that were specifically designed to target and validate the essential genes identified in our study. Information regarding the siRNAs is detailed in the Supplementary SI Gene list. We transfected MDA-MB-231, MDA-MB-436, and HS578T cells with siRNAs using the Lipofectamine™ RNAiMAX transfection reagent kit (#13778150, Thermo Fisher Scientific) according to the manufacturer’s protocol. The cells were then seeded into 96-well plates at a density of 2.5 × 103 cells per well. After 24 h, the medium was replaced, and the cells were treated with cisplatin and doxorubicin. Following 120 h of incubation, we assessed cell viability using the alamarBlue™ HS cell viability reagent (#A50100, Thermo Fisher Scientific). The absorbance of each well was measured using a microplate reader. We determined the half-maximal inhibitory concentration (IC50) values using GraphPad Prism 7 software.

2.3.2. Cell Survival Assay Using Single Drug

MDA-MB-231 TNBC cells were seeded in 96-well plates at a density of 5000 cells per well in 90 µL medium and allowed to grow overnight. Subsequently, the cells were treated with chemotherapy drugs or molecular inhibitors in 10 µL medium (with 10% DMSO in it).
To comprehensively assess the dose–response relationship, a series of doses spanning a wide range was generated using a concentration gradient dilution scheme. The experiment is repeated three times to enhance reliability and minimize bias.
In each trial, the plates were organized with wells containing varying drug concentrations arranged in descending order, from high to no drug (medium only), within individual rows. Each concentration was replicated three times within the same row. To calculate the dose–response curve, after 120 h, cell viability was assessed following the manufacturer’s protocol using alamarBlue (#DAL1100, Thermo Fisher Scientific). This procedure involves subtracting the treatment group’s reading from the baseline and dividing it by the control group’s reading.

2.3.3. Cell Survival Assay Using Drug Combination

Cells (5000) were seeded in 80 μL of media in each well. To understand the response to a drug combination, the seeded cells underwent four different conditions: exposure to chemotherapy drugs, treatment with molecular inhibitors, a combination of chemotherapy drugs along with molecular inhibitors, and a medium-only control.
To prevent excessive motility that might overshadow the synergy effect, a low dose was chosen to specifically eliminate around 20% or fewer cells. The 20% inhibitory concentration (IC20) values were determined using measurements.
An experiment altering the order of both agents in a drug combination was conducted after an overnight culture. The first reagent was added, followed by the second reagent after 4 h. For the single-agent control, medium was added to ensure a consistent solution volume in each well.
After 120 h of incubation, cell viability was measured using alamarBlue (#DAL1100, Thermo Fisher Scientific) following the manufacturer’s protocol. As mentioned earlier, the control group without drug treatment was established. Calculation of the drug effect, including both the single drug effect and the combination group, is the same as the single drug effect parts. To identify the synergy effect in the combination group, we added the two single effects together and then compared them with the drug combination group using a t-test.

2.4. Genome-Wide CRISPR-Cas9 Screening of Chemoresistance and Data Analysis

2.4.1. Construction of TKOv3 Library

We acquired the Toronto Knockout CRISPR (TKOv3) library, which contains 71,090 sgRNAs targeting 18,049 protein-coding genes, from Addgene (Watertown, MA, USA) and applied electroporation to expand the library 1000-fold. For lentivirus production, 7.5 × 106 293T cells were seeded in 15 cm plates and prepared for transfection. Following the manufacturer’s guidelines [41], packaging vectors psPAX2, pMD2.G (Addgene), TKOv3 library plasmid, and Lipofectamine (Thermo Fisher Scientific) were mixed in Opti-MEM™ (Thermo Fisher Scientific). After 48 h of incubation, the lentivirus-containing medium was collected and stored at −80 °C.

2.4.2. Genome-Wide Pooled sgRNA Screens

MDA-MB-231 cells were transduced with the TKOv3 lentivirus library at a low MOI of 0.3. After 72 h of puromycin (2 μg/mL) selection, surviving cells were considered baseline samples (T0), and 3 × 107 cells were harvested and stored at −80 °C. The remaining cells were divided into three groups (control, cisplatin treatment (1.5 µM), and doxorubicin treatment (0.026 µM)), each performed in triplicate. Following four weeks (equivalent to a single chemotherapy cycle) of treatment, 3 × 107 cells were harvested from each group. Genomic DNA was extracted using the QIAamp DNA Blood Maxi Kit (Qiagen, Hilden, Germany). Two polymerase chain reactions (PCRs) were carried out to enrich the sgRNA-targeted genomic regions and amplify the sgRNA. The resulting libraries were sequenced on a NovaSeq 6000 system (Illumina, San Diego, CA, USA), producing nearly 80 million reads per sample to achieve 600× coverage of the CRISPR library.

2.4.3. Analysis of CRISPR Screening Data

We used the MAGeCK algorithm to analyze genome-wide CRISPR/Cas9 knockout screening data [42]. Several statistics were used as quality control for CRISPR experiment sequence data, including mapping ratio (i.e., the ratio of mappable reads to the sgRNA library), missed sgRNA, Gini-index, within-time-point replicate correlation, and between-time-point correlation. sgRNA data were first normalized using a list of non-targeting control sgRNAs. Gene essentiality scores (beta-scores) were then determined between time points Tend and T0 in each condition, including control, cisplatin, and doxorubicin, using the MAGeCK maximum likelihood estimation method [43]. It provides a p-value for a gene’s impact on the cell viability phenotype. We also conducted principal component analysis using the stats R package and Pearson correlation analysis using the corrplot R package [44].

2.5. Bulk RNA Sequencing

We employed bulk RNA sequencing to analyze the transcriptome profiles of the MDA-MB-231 cells. Total RNA was extracted using the TRIzol™ reagent (Thermo Fisher Scientific, Waltham, MA, USA) according to the manufacturer’s instructions, with an additional DNase I digestion step to eliminate potential genomic DNA contamination. We used the Agilent 2100 Bioanalyzer (Agilent Technologies, Santa Clara, CA, USA) to assess the quality of RNA and the Qubit RNA Assay Kit (Thermo Fisher Scientific, Waltham, MA, USA) to evaluate concentration. Only high-quality RNA samples (RNS integrity number [RIN] ≥ 7.0) were used for library preparation. RNA sequencing libraries were prepared using the Illumina TruSeq® Stranded mRNA Library Prep Kit (Illumina, San Diego, CA, USA) following the manufacturer’s protocol. The prepared libraries were quantified by qPCR, and their quality was verified on an Agilent 2100 Bioanalyzer. Sequencing was conducted on an Illumina NovaSeq 6000 platform, generating 75 bp single end reads. Raw reads were processed to remove adapters and low-quality bases using Trimmomatic and aligned to the reference genome using STAR aligner. Quantification of gene expression levels was performed using the featureCounts tool from the Subread package (Subread v2.0.2).

2.6. CRISPR-Cas9 Gene Combination Double-Knockout Screening

2.6.1. Selection of Candidate Genes

For the construction of our CDKO experiment, we used three criteria to select candidate genes: (i) they were selected from the 149 hallmark genes identified in our previous research using the XDeathDB database [23]; (ii) they demonstrated high expression, defined as read counts above 50, in the MDA-MB-231 cell line based on bulk RNA-sequencing data; and (iii) they represented essential genes in genome-wide screening in the MDA-MB-231 cell line. As a result, we selected 65 candidate genes for the experiment.

2.6.2. Library Construction

We developed the CDKO screening library following our recently published protocol [45]. To ensure screening quality, we included three sgRNAs per gene and selected three corresponding sgRNAs in our library for each of the 65 genes. In addition to the primary gene set, we included six genes known to form well-established synthetic lethal pairs as positive control gene pairs—PARP1 and BRCA1, PARP1 and BRD4, BRD4 and CHK1, and WEE1 and HDAC1. This design led to 71 genes in total. In selecting their sgRNAs, we referred to several widely used CRISPR libraries, such as the TKO v3 [46], human genome CRISPR knockout (hGECKOv2) [21], and human kinome CRISPR knockout (KinomeKO) named Brunello. We also considered the Vienna Bioactivity CRISPR (VBC) scores [47] in selecting sgRNA. As negative controls in the pooled library, we included 17 safe sgRNAs (8% of the total) that target non-functional regions of the genome.
The CDKO library, comprising 52,900 sgRNAs targeting 5041 gene–gene pairs, was developed and amplified 1000-fold using the electroporation method. For lentivirus production, five million 293T cells were seeded in 15 cm plates and prepared for transfection. The packaging vectors psPAX2 and pMD2.G (Addgene), cell-death CDKO library plasmid, and Lipofectamine (Thermo Fisher Scientific) were combined in OptiMEM (Thermo Fisher Scientific). After 48 h of incubation, the lentivirus-containing medium was harvested and stored at −80 °C.

2.6.3. Pooled sgRNA Screening

MDA-MB-231 cells were infected with the cell-death CDKO lentiviral library at a low MOI of 0.3, and following 72 h of treatment with 5 μg/mL puromycin, the surviving cells were designated as baseline samples (T0), and 3 × 107 cells were gathered and preserved at −80 °C. The remaining cells were split into three separate groups, and after a 28-day cultivation period, 3 × 107 cells were collected from each group, Tend and T0, and genomic DNA was isolated using the QIAamp Blood Maxi Kit (Qiagen, Hilden, Germany). A series of two polymerase chain reactions (PCRs) were conducted to enrich sgRNA-targeted genomic regions and amplify the sgRNA, and the derived libraries were sequenced on a NextSeq 500 system (Illumina), generating nearly 10 million reads per sample and attaining 200× coverage of the cell-death CDKO library.

2.6.4. CDKO CRISPR Sequencing Data Analysis

To analyze the sequencing data from the CDKO CRISPR screening, we first used FastQC to obtain an overview of basic quality control metrics for the raw next-generation sequencing data, including mapping ratio (i.e., the ratio of mappable reads to the sgRNA library), missed sgRNA, the Gini-index, within-time-point replicate correlation, and between-time-point correlation.
We calculated the LFC for each sgRNA pair using MAGeCK robust rank aggregation (RRA) [42]. We calculated SL scores using multiple methods, including GEMINI score [48], Horlbeck score [49], MAGeCK score [50], median score with and without background normalization (B/NB), and sgRNA-derived score-B/NB. We proposed the last two scores in our recently developed Synthetic Lethality Knowledgebase (SLKB) [51]. For each method, we identified the top 10% of gene pairs as potential synthetic lethal pairs and then focused on gene pairs that were consistently identified across different methods, selecting the most overlapping pairs for further study.

3. Results

3.1. Hierarchical Clustering Analysis of TNBC Transcriptome Profiles Between Patients and Cell Lines

The first clustering analysis was performed between TNBC baseline transcriptome samples and CCLE TNBC cells. In comparing chemotherapy responders to non-responders, significantly upregulated or downregulated genes represented intrinsic chemoresistance mechanisms. Statistically significantly differentially expressed genes were chosen using the fold change greater than 1.5 or less than −1.5 and adjusted p-value less than 0.01 (Supplementary Table S1 DEG gene list). We then used these gene expression data to perform hierarchical clustering analysis on both TNBC patient samples and TNBC cell lines. Clustering analysis in Figure 1A revealed distinct patterns between chemotherapy responders and non-responders. Notably, some TNBC cell lines were clustered together with TNBC non-responders. It suggests these cell lines as representative models for intrinsic chemo-resistance TNBC patient samples.
The second clustering analysis was performed between TNBC transcriptome samples after chemotherapy and CCLE TNBC cells. In comparing chemotherapy responders to non-responders, significantly upregulated or downregulated genes represented acquired chemoresistance mechanisms in TNBC patients. Similarly to the first clustering analysis on the baseline TNBC transcriptome data, this cluster analysis, as shown in Figure 2A, showed clear patterns between those that did and those that did not respond to chemotherapy. Interestingly, some TNBC cell lines were again clustered with the non-responders, further supporting their value as representative models for studying chemo-resistance in TNBC.

3.2. Correlation Analysis Between TNBC Cell Lines and TNBC Non-Responders of Chemotherapy

Spearman correlation analysis was performed between the 916 CCLE cell lines and TNBC non-responders using the 2000 most varied genes. Cell lines were ranked based on their average correlation values. Figure 1B shows the correlation between TNBC cells and TNBC baseline samples that were chemotherapy non-responders, and Figure 2B displays correlations with post-chemotherapy non-responder TNBC samples. The TNBC cell HCC70 had the highest transcriptome similarity with TNBC non-responders at baseline, its Spearman rank correlation was 0.42, and its rank was 5 (Supplementary Table S2 baseline cell line correlation rank list). HCC1143 was the TNBC cell line showing the highest transcriptome similarity with TNBC non-responders following chemotherapy, its Spearman rank correlation was 0.42, and its rank was 17 (Supplementary Table S3 post cell line correlation rank list).

3.3. Pathway Similarity Analysis by Single-Sample Gene Set Enrichment Analysis (ssGSEA) Score and Overall Similarity Analysis Between TNBC Cell Lines and TNBC Chemotherapy Non-Responders

ssGSEA scores of 50 hallmark gene sets in the Molecular Signatures Database (MSigDB) were used to assess similarity between TNBC cell lines and TNBC chemotherapy non-responder samples at the pathway level. Their similarities were shown in hierarchical clustering analyses. Figure 1C shows the pathway similarity between TNBC cells and baseline TNBC non-responder samples, and Figure 2C displays the pathway similarity between TNBC cells and post-chemotherapy TNBC non-responder samples. Both cluster analyses revealed distinct clusters of TNBC cell lines and TNBC non-responder samples.
In transcriptome clustering analysis and pathway clustering analysis, a TNBC cell was labeled as 1 if it was grouped together with chemotherapy non-responder TNBC samples. In Spearman correlation analysis, a TNBC cell was denoted as 1 if it was ranked in the top 100 among 916 cancer cells. The overall similarity between TNBCs and TNBC chemotherapy non-responders was ranked by the sum of these three similarity labels. Figure 1D ranks the TNBC cells for their overall similarity with baseline TNBC non-responders, and Figure 2D shows the overall similarity with TNBC non-responders after chemotherapy. Strikingly, MDA-MB-231 and MDA-MB-157 ranked top two in both overall similarity analyses.

3.4. Genome-Wide CRISPR Screening on MDA-MB-231 Cell Lines

A genome-wide CRISPR/cas9 screening was conducted to identify genes related to cisplatin or doxorubicin resistance in the most representative cell line, MDA-MB-231, using the TKOv3 sgRNA library, which contains 70,948 sgRNAs targeting 18,053 genes. After transducing cells with the lentiviral pooled sgRNA library with lower multiplicity of infection (MOI) (0.3) and puromycin selection, the baseline sample was harvested. On Day 28, we collected triplicate samples obtained after treatment with cisplatin or doxorubicin or treatment with DMSO (dimethyl sulfoxide) (Figure 3A). After sequencing, data analysis was performed in MAGeCK. The number of missed genes for all samples ranged between 200~400 sgRNAs among the library’s total 70,948 sgRNAs (Figure 3B), i.e., less than 0.56%. An increase in the Gini index was observed from 0.04 to 0.06 on Days 0 and 28, respectively (Figure 3C). It reflected the selective adaptation of cancer cells after CRISPR-Cas9 perturbation and consequently greater unevenness in the pooled sgRNAs due to longer treatment. From the three-dimensional principal component analysis (PCA) plot (Figure 3D) and pairwise sample correlation plot (Figure 3E), samples in the same treatment group show more similarity and higher correlation values. Overall, these quality control analyses indicated a successful genome-wide CRISPR-Cas9 screening of one MDA-MB-231 cell line.
Using the MAGeCK algorithm, β scores were calculated. In the control group, a lower β score indicated a more depleted sgRNA in the final time point; in other words, gene knockout led to more cell death or was more essential to cell viability. Therefore, the β score was also called the gene essentiality score in the control group. In either cisplatin- or doxorubicin-treated groups, a lower β score indicated its corresponding gene knockout led to more cell death after chemo treatment. In other words, gene knockout re-sensitized chemo drug treatment. A lower β score within the range of −1 and +1 indicated the gene knockout did not have much influence on cell growth or cell death. In this paper, genes were selected with a β score higher than 1 in the control group but a β score less than −1 in the chemo treatment group. This means these genes were not essential in MDB-MA-231, but their CRISPR-Cas9 knockout re-sensitized low-dose chemo treatment. Figure 4A,B lay out the β scores between control and doxorubicin and cisplatin treatment, respectively. Genes in the region higher than 1 on tthhe x-axis and less than −1 on the y-axis were our selected candidate genes. We identified 96 genes that re-sensitized cisplatin treatment (Supplementary Table S4 essential gene list in cisplatin treatment) and 93 genes that re-sensitized doxorubicin treatment (Supplementary Table S5 essential gene list in doxorubicin treatment). They had 19 overlapping genes between the two lists (Supplementary Table S6 19 overlapped gene list). Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway analysis was performed for two gene lists for cisplatin treatment and doxorubicin treatment, respectively. Figure 4C depicts the 15 most enriched pathways for cisplatin, and Figure 4D compares those for doxorubicin. In both chemo treatments, most enriched pathways are associated with DNA damage/DNA repair signaling pathways.

3.5. Validation of Targeted Gene Knockout for Increased Chemo-Sensitivity in TNBC

In this study, candidate genes were validated through the literature review. In our CRISPR-Cas9 screening, there were 96 genes that re-sensitized cisplatin treatment, and 93 genes that re-sensitized doxorubicin treatment. Among these genes, 28 were investigated and published in vivo or in vitro studies on cisplatin- or doxorubicin-resistance (Table 2). After filtering out those previously studied genes, only druggable genes were considered in our follow-up analysis. That is, the candidate gene must be a viable drug target, aligning with our objective to discover drug combinations that can overcome chemoresistance in TNBC. This analysis led to several candidate genes, including ERCC1, NFE2L2, PRKCG, ATR, NEPPS, and MCM9, for testing.
We performed a cell survival assay using siRNA-mediated gene silencing to validate the novel genes identified from our CRISPR essential gene list and observed that the expression of those six genes was knocked down, but only one gene was validated in the siRNA experiment, namely MCM9. In Figure 5A–C, gene knockdown by siRNA led to an increase in cisplatin sensitivity of three TNBC cell lines, MDA-MB-231, MDA-MB-436, and HS578T.
Small molecule KPT-185 specifically targets the protein expressed by MCM9. Drug combination assays were performed using cisplatin and the MCM9 inhibitor KPT-185. After the IC20 values for cisplatin (1.5 µM) and KPT-185 (29.6 µM) were determined through cell dose–response curve assays, combination assays of IC20 doses were evaluated. In MDA-MB-231 cell, an additive effect between the MCM9 inhibitor KPT-185 and cisplatin was observed (Figure 5D).

3.6. CRISPR-Cas9 Gene Combination Double-Knockout (CDKO) Experiment

Candidate genes were selected according to their overlap from three datasets, as shown in the Venn diagram plot in Figure 6B. These 65 genes were selected among cell death hallmark genes, expressed genes in the MDA-MB-231 cell line, and non-essential in genome-wide CRISPR-Cas9 screening. Table 3 details these 65 selected genes, including their functions and relevant cell-death modes.
A CDKO experiment was conducted to identify SL gene pairs among 65 selected genes in the MDA-MB-231 cell line (Figure 6A). Cells were transduced with the lentiviral pooled sgRNA library at a low MOI (0.3) and applied puromycin selection before collecting the baseline sample. After 28 days of cell culture, cells were harvested in triplicate samples. The sequencing results demonstrated high quality, with a mapping ratio of approximately 90% for all samples and around 10 million mapped reads per sample (Figure 6C). The number of missed genes ranged from 39 to 1000 dual-sgRNA out of a total of 52,900 sgRNAs (Table 4). The Gini index for all samples was around 0.05 (Table 4, Figure 6D). The pairwise sample correlation (Figure 6E) plots show a distinct pattern and higher correlation values of samples within the same treatment group across triplicate samples.
Five SL score calculation methods were implemented to select SL gene pairs (see Methods). Their overlaps are depicted in both the Venn diagram (Figure 7A). Among 2080 screened gene pairs, 242 gene pairs exhibited SL in more than three methods and are more likely than others to have an SL effect. An SL network based on these 242 gene pairs is shown in Figure 7B. Hub genes interacting with at least ten other genes are displayed in the histogram (Figure 7C).

4. Discussion

4.1. CRISPR-Cas9 Gene Combination Double-Knockout (CDKO) Experiment

For the first time, representative chemo-resistance TNBC cell line models were selected. It was based on three clustering/correlation analyses in the transcriptome between TNBC cell lines and TNBC chemotherapy patient responders and non-responders. The first analysis was based on a set of differentially expressed genes between chemotherapy responders and non-responders. The cluster analysis reveals TNBC cell lines that were grouped with non-responders. The second analysis started from 2000 highly varied genes among 916 cancer cell lines. Using these 2000 genes, it ranked 916 cancer cell lines in Spearman correlations with TNBC chemotherapy non-responders. The top TNBC cells were selected based on this ranking. Although these two analyses both used correlations between cell lines and TNBC non-responder patient samples, they differed in how genes were selected for correlation analysis. The first analysis selected genes from patient samples, while the second selected genes from cell lines. They are complementary to each other. The third analysis, on the other hand, focused on correlation analysis based on pathway enrichment on cancer hallmark gene sets in the Molecular Signatures Database (MSigDB) [37]. These gene sets are directly related to cancer development and progression, including cell cycle regulation, apoptosis, DNA repair, epithelial–mesenchymal transition (EMT), hypoxia, and others. Furthermore, when studying chemotherapy non-responders, transcriptome data at baseline (i.e., intrinsic chemo-resistance) and transcriptome data after chemotherapy (i.e., acquired chemo-resistance) were both considered in all three TNBC cell correlation/cluster analyses. In our view, these analyses provided the most comprehensive view of the similarities between TNBC cells and TNBC chemotherapy non-responders. We found the MDA-MB-231, MDA-MB-157, HCC1187, HCC38, and HCC1395 cell lines to be the most representative across all three methods. Three of these, MDA-MB-231, MDA-MB-157, and HCC1395, were consistently significant and therefore most representative across the three methods when we analyzed similarity with samples of TNBC poor responder samples following chemotherapy.

4.2. TNBC Chemo-Resistant Target Discovery Through Genome-Wide CRISPR-Cas9 Screening and Validation Analyses

We used genome-wide CRISPR screening to identify essential genes that could help overcome resistance to cisplatin and doxorubicin (Table 2) in triple-negative breast cancer. DNMT1, PPIA, RUNX, BCL2L1, RUNX2, NBN, GTF2H5, USP22, HSP90AB1, CDC25B, NCF1, FANCA, FANCG, and ERCC1 are well-studied genes for cisplatin treatment, some having known drug targets or inhibitors. DNMT1, for example, is involved in the DNA methylation pathway and targeted by decitabine, which is used clinically to treat other types of cancer [138]. In addition, RUNX, which regulates apoptosis and cell proliferation, has a small molecule inhibitor (AI-10-49) in preclinical development; BCL2L1, involved in the intrinsic apoptotic pathway, can be targeted by BikDD and lapatinib, both in preclinical investigation [138]; and RUNX2, which also plays a role in regulating the cell cycle and apoptosis, is targeted by BET inhibitors JQ1 and I-BET762, which are currently in Phase I/II clinical trials [139].
Other genes in our list have been implicated in chemoresistance in breast and other cancer cells. PPIA participates in the regulation of miRNA and impacts the sensitivity of breast cancer cells to doxorubicin [140] and the knockdown of RUNX, which is involved in the YAP signaling pathway, enhances sensitivity to doxorubicin in breast cancer cells [139]. In HER2- and MDM2-enriched breast cancer subtypes, NBN plays a role in doxorubicin, paclitaxel, and carboplatin resistance via its involvement in DNA repair and homologous recombination. Involved in nucleotide excision repair (NER), GTF2H5 has been shown to predict survival in high-grade serous ovarian cancer.
For doxorubicin treatment, our gene list includes ABCC1, HIST1H2BJ, ZEB2, ATM, FANCL, CDC25B, XRCC1, ACTG1, IRS1, NBN, NFE2L2, NDUFB9, CDK5, and CDCA3, some of which are reported to play a role in chemoresistance. In TNBC, for instance, ABCC1 is a drug efflux transporter implicated in resistance to doxorubicin, paclitaxel, and cisplatin [141], and the involvement of HIST1H2BJ in glutathione synthesis and copper chelation promotes resistance to doxorubicin, paclitaxel, and cisplatin [142]. ZEB2, a transcription factor, is associated with drug resistance in breast cancer cells through its regulation of the epithelial–mesenchymal transition (EMT) [143].
Some of these genes have potential drug targets or inhibitors. The ATM kinase inhibitor KU-55933, which targets ATMs involved in the DNA damage response pathway, is in preclinical development [144]; FANCL, part of the Fanconi anemia DNA repair pathway, has been targeted by small molecule inhibitors, such as curcumin, in preclinical studies [145]; and CDK5 inhibitors, like roscovitine and dinaciclib, have shown promise in preclinical studies and are in clinical trials for various cancer types [146]. CDK5 is involved in cell cycle regulation and the DNA damage response, which contribute to chemoresistance.
CDC25B and NBN are the overlapping genes between the cisplatin and doxorubicin essential gene lists. CDC25B is involved in cell cycle regulation and DNA damage response and has been targeted by thiostrepton, FDI-6, and siomycin A in preclinical studies for the treatment of platinum-resistant ovarian cancer [142]. NBN, on the other hand, is involved in DNA repair and homologous recombination and has been studied in vitro using siRNA. Though no drug targets or inhibitors have been identified for NBN, its role in DNA repair suggests it may be a potential therapeutic target in the future.
In validating some of the new targets that are druggable, we employed the MDA-MB-231 cell line used for our initial screening, as well as such other TNBC cell lines such as MDA-MB-436 and HS578T. Our results showed a significant reduction in cell viability in all three cell lines following the knockdown of MCM9, indicating an increased sensitivity to cisplatin treatment. Our study also identified an additive effect of KPT-185, an MCM9 inhibitor, when combined with cisplatin in in vitro assays. Cisplatin is a commonly used chemotherapeutic drug in TNBC. Its mode of action when entering cancer cells is to bind to DNA, causing DNA crosslinks and DNA damage, subsequently inducing apoptosis in cancer cells [4,147]. MCM9 can combine with MCM8 to form a complex, which is required for DNA damage repair caused by DNA interstrand crosslinks [148,149]. Therefore, MCM9 inhibition could sensitize TNBC cells to Cisplatin. The intersection of these methods aids the confident identification of the core set of essential genes, such as MCM9, as potential gene targets to overcome triple-negative breast cancer.
Doxorubicin and cisplatin exert cytotoxic effects through distinct mechanisms; therefore, the relatively low overlap in re-sensitizing genes is expected. In this study, our primary aim was to identify the common factors involved in overcoming chemotherapy resistance in triple-negative breast cancer. In future work, we plan to investigate drug-specific resistance mechanisms.

4.3. Discovery of Synthetic Lethal Gene Pairs in TNBC Cells

We introduced five SL gene pair as positive controls, four of which were ranked within the top 10% based on their LFC ranking scores according to the results from our screening data analysis (BRD4_PARP1, rank 9, SL score 8.56 × 10−10; HDAC1__WEE1, rank 124, SL score 3.29 × 10−5; BRD4__CHEK1, rank 172, SL score 0.00016564; and HDAC2__WEE1, rank 200, SL score 0.00027904). The presence of these positive control gene pairs in our results demonstrates the success of our experiment. HDAC1 inhibition has been shown to induce histone tail acetylation and an open chromatin structure, which promotes the expression of cell differentiation or death genes. In combination with WEE1 inhibition, impaired activity of the cell cycle checkpoint kinase and premature mitotic entry lead to an increase in DNA damage and apoptosis, making this combination a promising target for the treatment of acute myeloid leukemia (AML) [150]. Similarly, inhibition of BRD4 has been shown to induce homologous recombination deficiency and sensitize cells to inhibitors of poly(ADP-ribose) polymerases (PARPis). This combination has the potential to reverse intrinsic resistance in RAS-mutant tumors and other mechanisms of PARPi resistance, warranting clinical assessment in both PARPi-sensitive and -resistant cancers [151].
Our findings add to the ongoing efforts to discover novel synthetic lethal interactions that can be harnessed for targeted cancer therapy. The 242 synthetic lethal pairs identified in this study offer a valuable resource for future research in TNBC and other cancer types. Some of the 242 gene pairs we identified have already been reported to demonstrate synthetic lethal effects. The CDH1–HDAC1 and HDAC1–PARP1 gene pairs have been reported as synergistically lethal gene pairs and can be considered potential drug targets. Decourtye-Espiard and colleagues investigated the potential of histone deacetylase (HDAC) inhibitors as a treatment for cancers with CDH1 mutations, testing the effects of several HDAC inhibitors on gastric and breast preclinical models with and without CDH1 mutations [152], and observed greater sensitivity of CDH1-null cells than wild-type cells to pan-HDAC inhibitors, such as entinostat, pracinostat, mocetinostat, and vorinostat. This supports the notion that CDH1-HDAC1 has a strong synergistic lethal effect on cancer. In our study, we identified another potential synthetic lethal gene pair, HDAC1 and VDAC2. VDAC2 has been reported to be closely related to apoptosis. When VDAC2 is inhibited, VDAC2 releases BAK, allowing it to oligomerize and trigger mitochondrial outer membrane permeabilization (MOMP), which leads to apoptosis [153,154]. Therefore, dual loss of VDAC2 and HDAC1 is likely to enhance apoptosis, providing a plausible basis for synthetic lethality. The combination of hypomethylating agents (HMAs), PARPis, and histone deacetylases (HDACis) was hypothesized to be synergistically cytotoxic to leukemia and lymphoma cells. Valdez’s research team demonstrated that exposing AML and lymphoma cell lines to a combination of PARPi niraparib, HMA decitabine, and HDACi romidepsin or panobinostat led to a synergistic inhibition of cell proliferation up to 70%. This combination activated the ATM pathway, increased the production of reactive oxygen species, decreased mitochondrial membrane potential, and induced apoptosis [152]. This provides robust evidence that our CRISPR double-knockout experiments are indeed effective in identifying synthetic lethal gene pairs.
In the gene–gene network focusing on 242 SL interactions, we specifically selected hub genes of interest, those with the highest number of pairs (Figure 7C). Six of the nineteen hub genes were found to have related drug inhibitors, most identified in previous research for TNBC treatment (Table 5). However, though individual drugs have been studied, no research on drug combinations is reported. This provides us with ample opportunity to discover the most novel and effective drug combinations for treating triple-negative breast cancer (TNBC).
Although our study successfully identified synthetic lethal gene pairs in the MDA-MB-231 cell line, further validation of these interactions in additional TNBC cell lines and in vivo models is necessary to confirm their relevance and potential as therapeutic targets. This requires more validation experiments, such as designing a CRISPR screening library that targets only the 242 known gene pairs and assaying two-drug combinations. To understand the underlying molecular mechanisms driving synthetic lethality and determine the optimal drug combinaftions and dosages for clinical translation, detailed functional characterization of the identified gene pairs is also needed.

5. Conclusions

In conclusion, our CRISPR-based screening approach has successfully identified synthetic lethal gene pairs in a representative TNBC cell line, offering new potential drug targets for the treatment of this aggressive and challenging cancer subtype. Future studies should focus on validating these interactions in additional models and characterizing their functional roles to pave the way for the development of innovative targeted therapies for patients with TNBC.
Our study strongly supports the notion that utilizing CRISPR screening for essential cancer genes is an efficient and practical approach to identify potential drug targets to overcome resistance to chemotherapeutic drugs. By integrating multiple data analysis techniques and biological experimental analyses, we have identified several genes that hold promise as potential drug targets for combating chemoresistance. Our findings underscore the importance of leveraging advanced genetic screening tools and data-driven methods to better understand the molecular mechanisms underlying drug resistance and to develop more effective therapeutic strategies for cancer treatment.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cancers17233876/s1, Table S1: DEG gene list of Poor responder vs Good responder (Baseline); Table S2: TNBC cell line correlation score with poor responder (Baseline); Table S3: TNBC cell line correlation score with poor responder (Post Chemo); Table S4: Essential gene list in cisplatin treatment; Table S5: Essential gene list in doxorubicin treatment; Table S6: 19 overlapped gene list.

Author Contributions

Conceptualization, L.L. and S.S.; methodology, S.T., G.N. and S.S.; software, B.G. and K.F.; validation, S.S., S.L. and L.L.; formal analysis, S.S., K.F. and Y.H. (Yang Huo); investigation, S.S.; resources, J.P.; data curation, S.S. and Y.H. (Yang Huo); writing—original draft preparation, S.S.; writing—review and editing, Y.H. (Yirui Huang), L.W., L.L., S.L., L.C. and D.S.; visualization, S.S.; supervision, L.L.; project administration, L.L. and L.W.; funding acquisition, L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ADCsAntibodydrug conjugates
AMLAcute myeloid leukemia
B/NBScore with and without background normalization
CDKOCombination double-knockout
EMTEpithelialmesenchymal transition
EREstrogen receptor
GEOGene Expression Omnibus
GOGene Ontology
GSVAGene set variation analysis
hGECKOv2Human genome CRISPR knockout version 2
HDACHistone deacetylase
HER2Human epidermal growth factor receptor 2
HMAsHypomethylating agents
KEGGKyoto Encyclopedia of Genes and Genomes
KinomeKOHuman kinome CRISPR knockout
MOIMultiplicity of infection
MPMillerPayne
MSigDBMolecular Signatures Database
mTORMammalian target of rapamycin
NACTNeoadjuvant chemotherapy
NERNucleotide excision repair
PCAPrincipal component analysis
PARPPoly(ADPribose) polymerase
PARPisPoly(ADPribose) polymerase inhibitors
PCRsPolymerase chain reactions
PI3KPhosphoinositide 3kinase
PRProgesterone receptor
RMARobust multiarray average
RRARobust rank aggregation
SLSynthetic lethality
SLKBSynthetic Lethality Knowledgebase
ssGSEASinglesample Gene Set Enrichment Analysis
STRShort tandem repeat
TKOv3Toronto Knockout CRISPR library version 3
VBCVienna Bioactivity CRISPR

References

  1. Ades, F.; Zardavas, D.; Bozovic-Spasojevic, I.; Pugliano, L.; Fumagalli, D.; De Azambuja, E.; Viale, G.; Sotiriou, C.; Piccart, M. Luminal B breast cancer: Molecular characterization, clinical management, and future perspectives. J. Clin. Oncol. 2014, 32, 2794–2803. [Google Scholar] [CrossRef] [PubMed]
  2. Nedeljković, M.; Damjanović, A. Mechanisms of chemotherapy resistance in triple-negative breast cancer—How we can rise to the challenge. Cells 2019, 8, 957. [Google Scholar] [CrossRef] [PubMed]
  3. Yagata, H.; Kajiura, Y.; Yamauchi, H. Current strategy for triple-negative breast cancer: Appropriate combination of surgery, radiation, and chemotherapy. Breast Cancer 2011, 18, 165–173. [Google Scholar] [CrossRef]
  4. Dasari, S.; Tchounwou, P.B. Cisplatin in cancer therapy: Molecular mechanisms of action. Eur. J. Pharmacol. 2014, 740, 364–378. [Google Scholar] [CrossRef]
  5. Rivankar, S. An overview of doxorubicin formulations in cancer therapy. J. Cancer Res. Ther. 2014, 10, 853–858. [Google Scholar] [CrossRef]
  6. Keam, B.; Im, S.-A.; Kim, H.-J.; Oh, D.-Y.; Kim, J.H.; Lee, S.-H.; Chie, E.K.; Han, W.; Kim, D.-W.; Moon, W.K. Prognostic impact of clinicopathologic parameters in stage II/III breast cancer treated with neoadjuvant docetaxel and doxorubicin chemotherapy: Paradoxical features of the triple negative breast cancer. BMC Cancer 2007, 7, 203. [Google Scholar] [CrossRef]
  7. Byrski, T.; Huzarski, T.; Dent, R.; Gronwald, J.; Zuziak, D.; Cybulski, C.; Kladny, J.; Gorski, B.; Lubinski, J.; Narod, S. Response to neoadjuvant therapy with cisplatin in BRCA1-positive breast cancer patients. Breast Cancer Res. Treat. 2009, 115, 359–363. [Google Scholar] [CrossRef] [PubMed]
  8. Sikov, W.M.; Berry, D.A.; Perou, C.M.; Singh, B.; Cirrincione, C.T.; Tolaney, S.M.; Kuzma, C.S.; Pluard, T.J.; Somlo, G.; Port, E.R. Impact of the addition of carboplatin and/or bevacizumab to neoadjuvant once-per-week paclitaxel followed by dose-dense doxorubicin and cyclophosphamide on pathologic complete response rates in stage II to III triple-negative breast cancer: CALGB 40603 (Alliance). J. Clin. Oncol. 2015, 33, 13. [Google Scholar]
  9. Loibl, S.; O’Shaughnessy, J.; Untch, M.; Sikov, W.M.; Rugo, H.S.; McKee, M.D.; Huober, J.; Golshan, M.; von Minckwitz, G.; Maag, D. Addition of the PARP inhibitor veliparib plus carboplatin or carboplatin alone to standard neoadjuvant chemotherapy in triple-negative breast cancer (BrighTNess): A randomised, phase 3 trial. Lancet Oncol. 2018, 19, 497–509. [Google Scholar] [CrossRef]
  10. Schmid, P.; Cortes, J.; Pusztai, L.; McArthur, H.; Kümmel, S.; Bergh, J.; Denkert, C.; Park, Y.H.; Hui, R.; Harbeck, N. Pembrolizumab for early triple-negative breast cancer. New Engl. J. Med. 2020, 382, 810–821. [Google Scholar] [CrossRef]
  11. Schmid, P.; Adams, S.; Rugo, H.S.; Schneeweiss, A.; Barrios, C.H.; Iwata, H.; Diéras, V.; Hegg, R.; Im, S.-A.; Shaw Wright, G. Atezolizumab and nab-paclitaxel in advanced triple-negative breast cancer. New Engl. J. Med. 2018, 379, 2108–2121. [Google Scholar] [CrossRef]
  12. Dent, R.A.; Lindeman, G.J.; Clemons, M.; Wildiers, H.; Chan, A.; McCarthy, N.J.; Singer, C.F.; Lowe, E.S.; Watkins, C.L.; Carmichael, J. Phase I trial of the oral PARP inhibitor olaparib in combination with paclitaxel for first-or second-line treatment of patients with metastatic triple-negative breast cancer. Breast Cancer Res. 2013, 15, R88. [Google Scholar] [CrossRef]
  13. Geenen, J.J.; Linn, S.C.; Beijnen, J.H.; Schellens, J.H. PARP inhibitors in the treatment of triple-negative breast cancer. Clin. Pharmacokinet. 2018, 57, 427–437. [Google Scholar] [CrossRef]
  14. Nagayama, A.; Vidula, N.; Ellisen, L.; Bardia, A. Novel antibody–drug conjugates for triple negative breast cancer. Ther. Adv. Med. Oncol. 2020, 12, 1758835920915980. [Google Scholar] [CrossRef]
  15. Shao, S.; Tsai, M.-H.; Lu, J.; Yu, T.; Jin, J.; Xiao, D.; Jiang, H.; Han, M.; Wang, M.; Wang, J. Site-specific and hydrophilic ADCs through disulfide-bridged linker and branched PEG. Bioorganic Med. Chem. Lett. 2018, 28, 1363–1370. [Google Scholar] [CrossRef]
  16. Lyons, T.G. Targeted therapies for triple-negative breast cancer. Curr. Treat. Options Oncol. 2019, 20, 82. [Google Scholar] [CrossRef] [PubMed]
  17. Bradbury, A.; Hall, S.; Curtin, N.; Drew, Y. Targeting ATR as Cancer Therapy: A new era for synthetic lethality and synergistic combinations? Pharmacol. Ther. 2020, 207, 107450. [Google Scholar] [CrossRef]
  18. Cruz, C.; Castroviejo-Bermejo, M.; Gutiérrez-Enríquez, S.; Llop-Guevara, A.; Ibrahim, Y.; Gris-Oliver, A.; Bonache, S.; Morancho, B.; Bruna, A.; Rueda, O. RAD51 foci as a functional biomarker of homologous recombination repair and PARP inhibitor resistance in germline BRCA-mutated breast cancer. Ann. Oncol. 2018, 29, 1203–1210. [Google Scholar] [CrossRef]
  19. Cortesi, L.; Rugo, H.S.; Jackisch, C. An overview of PARP inhibitors for the treatment of breast cancer. Target. Oncol. 2021, 16, 255–282. [Google Scholar] [CrossRef]
  20. Ran, F.A.; Hsu, P.D.; Wright, J.; Agarwala, V.; Scott, D.A.; Zhang, F. Genome engineering using the CRISPR-Cas9 system. Nat. Protoc. 2013, 8, 2281–2308. [Google Scholar] [CrossRef]
  21. Sanjana, N.E.; Shalem, O.; Zhang, F. Improved vectors and genome-wide libraries for CRISPR screening. Nat. Methods 2014, 11, 783–784. [Google Scholar] [CrossRef]
  22. Dai, M.; Yan, G.; Wang, N.; Daliah, G.; Edick, A.M.; Poulet, S.; Boudreault, J.; Ali, S.; Burgos, S.A.; Lebrun, J.-J. In vivo genome-wide CRISPR screen reveals breast cancer vulnerabilities and synergistic mTOR/Hippo targeted combination therapy. Nat. Commun. 2021, 12, 3055. [Google Scholar] [CrossRef]
  23. Gadepalli, V.S.; Kim, H.; Liu, Y.; Han, T.; Cheng, L. XDeathDB: A visualization platform for cell death molecular interactions. Cell Death Dis. 2021, 12, 1156. [Google Scholar] [CrossRef]
  24. Tang, S.; Gökbağ, B.; Fan, K.; Shao, S.; Huo, Y.; Wu, X.; Cheng, L.; Li, L. Synthetic lethal gene pairs: Experimental approaches and predictive models. Front. Genet. 2022, 13, 961611. [Google Scholar] [CrossRef]
  25. Gillet, J.-P.; Varma, S.; Gottesman, M.M. The Clinical Relevance of Cancer Cell Lines. JNCI J. Natl. Cancer Inst. 2013, 105, 452–458. [Google Scholar] [CrossRef] [PubMed]
  26. Wilding, J.L.; Bodmer, W.F. Cancer Cell Lines for Drug Discovery and DevelopmentCancer Cell Lines for Drug Discovery and Development. Cancer Res. 2014, 74, 2377–2384. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, K.; Newbury, P.A.; Glicksberg, B.S.; Zeng, W.Z.; Paithankar, S.; Andrechek, E.R.; Chen, B. Evaluating cell lines as models for metastatic breast cancer through integrative analysis of genomic data. Nat. Commun. 2019, 10, 2138. [Google Scholar] [CrossRef] [PubMed]
  28. Domcke, S.; Sinha, R.; Levine, D.A.; Sander, C.; Schultz, N. Evaluating cell lines as tumour models by comparison of genomic profiles. Nat. Commun. 2013, 4, 2126. [Google Scholar] [CrossRef]
  29. Huo, Y.; Shao, S.; Liu, E.; Li, J.; Tian, Z.; Wu, X.; Zhang, S.; Stover, D.; Wu, H.; Cheng, L.; et al. Subpathway Analysis of Transcriptome Profiles Reveals New Molecular Mechanisms of Acquired Chemotherapy Resistance in Breast Cancer. Cancers 2022, 14, 4878. [Google Scholar] [CrossRef]
  30. Barretina, J.; Caponigro, G.; Stransky, N.; Venkatesan, K.; Margolin, A.A.; Kim, S.; Wilson, C.J.; Lehár, J.; Kryukov, G.V.; Sonkin, D.; et al. The Cancer Cell Line Encyclopedia enables predictive modelling of anticancer drug sensitivity. Nature 2012, 483, 603–607. [Google Scholar] [CrossRef]
  31. Vera-Ramirez, L.; Sanchez-Rovira, P.; Ramirez-Tortosa, C.L.; Quiles, J.L.; Ramirez-Tortosa, M.; Lorente, J.A. Transcriptional Shift Identifies a Set of Genes Driving Breast Cancer Chemoresistance. PLoS ONE 2013, 8, e53983. [Google Scholar] [CrossRef]
  32. Korde, L.A.; Lusa, L.; McShane, L.; Lebowitz, P.F.; Lukes, L.; Camphausen, K.; Parker, J.S.; Swain, S.M.; Hunter, K.; Zujewski, J.A. Gene expression pathway analysis to predict response to neoadjuvant docetaxel and capecitabine for breast cancer. Breast Cancer Res. Treat. 2010, 119, 685–699. [Google Scholar] [CrossRef]
  33. Miyake, T.; Nakayama, T.; Naoi, Y.; Yamamoto, N.; Otani, Y.; Kim, S.J.; Shimazu, K.; Shimomura, A.; Maruyama, N.; Tamaki, Y.; et al. GSTP1 expression predicts poor pathological complete response to neoadjuvant chemotherapy in ER-negative breast cancer. Cancer Sci. 2012, 103, 913–920. [Google Scholar] [CrossRef]
  34. Team R.C. R: A Language and Environment for Statistical Computing; R Foundation for Statistical Computing: Vienna, Austria, 2021. [Google Scholar]
  35. Ritchie, M.E.; Phipson, B.; Wu, D.; Hu, Y.; Law, C.W.; Shi, W.; Smyth, G.K. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 2015, 43, e47. [Google Scholar] [CrossRef] [PubMed]
  36. Hänzelmann, S.; Castelo, R.; Guinney, J. GSVA: Gene set variation analysis for microarray and RNA-Seq data. BMC Bioinform. 2013, 14, 7. [Google Scholar] [CrossRef]
  37. Liberzon, A.; Birger, C.; Thorvaldsdóttir, H.; Ghandi, M.; Mesirov, J.P.; Tamayo, P. The Molecular Signatures Database (MSigDB) Hallmark Gene Set Collection. Cell Syst. 2015, 1, 417–425. [Google Scholar] [CrossRef]
  38. Ashburner, M.; Ball, C.A.; Blake, J.A.; Botstein, D.; Butler, H.; Cherry, J.M.; Davis, A.P.; Dolinski, K.; Dwight, S.S.; Eppig, J.T.; et al. Gene ontology: Tool for the unification of biology. Nat. Genet. 2000, 25, 25–29. [Google Scholar] [CrossRef]
  39. Kanehisa, M.; Goto, S. KEGG: Kyoto Encyclopedia of Genes and Genomes. Nucleic Acids Res. 2000, 28, 27–30. [Google Scholar] [CrossRef]
  40. Yu, G.; Wang, L.-G.; Han, Y.; He, Q.-Y. clusterProfiler: An R Package for Comparing Biological Themes Among Gene Clusters. OMICS J. Integr. Biol. 2012, 16, 284–287. [Google Scholar] [CrossRef]
  41. Hart, T.; Tong, A.H.Y.; Chan, K.; Van Leeuwen, J.; Seetharaman, A.; Aregger, M.; Chandrashekhar, M.; Hustedt, N.; Seth, S.; Noonan, A.; et al. Evaluation and Design of Genome-Wide CRISPR/SpCas9 Knockout Screens. G3 Genes|Genomes|Genet. 2017, 7, 2719–2727. [Google Scholar] [CrossRef]
  42. Li, W.; Xu, H.; Xiao, T.; Cong, L.; Love, M.I.; Zhang, F.; Irizarry, R.A.; Liu, J.S.; Brown, M.; Liu, X.S. MAGeCK enables robust identification of essential genes from genome-scale CRISPR/Cas9 knockout screens. Genome Biol. 2014, 15, 554. [Google Scholar] [CrossRef] [PubMed]
  43. Li, W.; Köster, J.; Xu, H.; Chen, C.-H.; Xiao, T.; Liu, J.S.; Brown, M.; Liu, X.S. Quality control, modeling, and visualization of CRISPR screens with MAGeCK-VISPR. Genome Biol. 2015, 16, 281. [Google Scholar] [CrossRef]
  44. Ringnér, M. What is principal component analysis? Nat. Biotechnol. 2008, 26, 303–304. [Google Scholar] [CrossRef] [PubMed]
  45. Tang, S.; Wu, X.; Liu, J.; Zhang, Q.; Wang, X.; Shao, S.; Gokbag, B.; Fan, K.; Liu, X.; Li, F.; et al. Generation of dual-gRNA library for combinatorial CRISPR screening of synthetic lethal gene pairs. STAR Protoc. 2022, 3, 101556. [Google Scholar] [CrossRef]
  46. Mair, B.; Tomic, J.; Masud, S.N.; Tonge, P.; Weiss, A.; Usaj, M.; Tong, A.H.Y.; Kwan, J.J.; Brown, K.R.; Titus, E.; et al. Essential Gene Profiles for Human Pluripotent Stem Cells Identify Uncharacterized Genes and Substrate Dependencies. Cell Rep. 2019, 27, 599–615.e12. [Google Scholar] [CrossRef]
  47. Michlits, G.; Jude, J.; Hinterndorfer, M.; de Almeida, M.; Vainorius, G.; Hubmann, M.; Neumann, T.; Schleiffer, A.; Burkard, T.R.; Fellner, M.; et al. Multilayered VBC score predicts sgRNAs that efficiently generate loss-of-function alleles. Nat. Methods 2020, 17, 708–716. [Google Scholar] [CrossRef]
  48. Hao, Z.; Wu, D.; Fang, Y.; Wu, M.; Cai, R.; Li, X. Prediction of Synthetic Lethal Interactions in Human Cancers using Multi-view Graph Auto-Encoder. IEEE J. Biomed. Health Inform. 2021, 25, 4041–4051. [Google Scholar] [CrossRef]
  49. Horlbeck, M.A.; Xu, A.; Wang, M.; Bennett, N.K.; Park, C.Y.; Bogdanoff, D.; Adamson, B.; Chow, E.D.; Kampmann, M.; Peterson, T.R.; et al. Mapping the Genetic Landscape of Human Cells. Cell 2018, 174, 953–967.e22. [Google Scholar] [CrossRef]
  50. Kweon, J.; Kim, D.-E.; Jang, A.-H.; Kim, Y. CRISPR/Cas-based customization of pooled CRISPR libraries. PLoS ONE 2018, 13, e0199473. [Google Scholar] [CrossRef]
  51. Gökbağ, B.; Tang, S.; Fan, K.; Cheng, L.; Yu, L.; Zhao, Y.; Li, L. SLKB: Synthetic lethality knowledge base. Nucleic Acids Res. 2023, 52, D1418–D1428. [Google Scholar] [CrossRef] [PubMed]
  52. Westhoff, G.L.; Chen, Y.; Teng, N.N.H. Targeting FOXM1 Improves Cytotoxicity of Paclitaxel and Cisplatinum in Platinum-Resistant Ovarian Cancer. Int. J. Gynecol. Cancer 2017, 27, 1602–1609. [Google Scholar] [CrossRef]
  53. Zheng, K.; Li, Y.; Wang, S.; Wang, X.; Liao, C.; Hu, X.; Fan, L.; Kang, Q.; Zeng, Y.; Wu, X.; et al. Inhibition of autophagosome-lysosome fusion by ginsenoside Ro via the ESR2-NCF1-ROS pathway sensitizes esophageal cancer cells to 5-fluorouracil-induced cell death via the CHEK1-mediated DNA damage checkpoint. Autophagy 2016, 12, 1593–1613. [Google Scholar] [CrossRef]
  54. Kosinsky, R.L.; Helms, M.; Zerche, M.; Wohn, L.; Dyas, A.; Prokakis, E.; Kazerouni, Z.B.; Bedi, U.; Wegwitz, F.; Johnsen, S.A. USP22-dependent HSP90AB1 expression promotes resistance to HSP90 inhibition in mammary and colorectal cancer. Cell Death Dis. 2019, 10, 911. [Google Scholar] [CrossRef]
  55. Yu, J.; Qin, B.; Moyer, A.M.; Nowsheen, S.; Liu, T.; Qin, S.; Zhuang, Y.; Liu, D.; Lu, S.W.; Kalari, K.R.; et al. DNA methyltransferase expression in triple-negative breast cancer predicts sensitivity to decitabine. J. Clin. Investig. 2018, 128, 2376–2388. [Google Scholar] [CrossRef]
  56. Lang, J.-Y.; Hsu, J.L.; Meric-Bernstam, F.; Chang, C.-J.; Wang, Q.; Bao, Y.; Yamaguchi, H.; Xie, X.; Woodward, W.A.; Yu, D.; et al. BikDD Eliminates Breast Cancer Initiating Cells and Synergizes with Lapatinib for Breast Cancer Treatment. Cancer Cell 2011, 20, 341–356. [Google Scholar] [CrossRef] [PubMed]
  57. Serrano-Oviedo, L.; Nuncia-Cantarero, M.; Morcillo-Garcia, S.; Nieto-Jimenez, C.; Burgos, M.; Corrales-Sanchez, V.; Perez-Peña, J.; Győrffy, B.; Ocaña, A.; Galán-Moya, E.M. Identification of a stemness-related gene panel associated with BET inhibition in triple negative breast cancer. Cell. Oncol. 2020, 43, 431–444. [Google Scholar] [CrossRef]
  58. Cheng, Q.; Chang, J.T.; Geradts, J.; Neckers, L.M.; Haystead, T.; Spector, N.L.; Lyerly, H.K. Amplification and high-level expression of heat shock protein 90 marks aggressive phenotypes of human epidermal growth factor receptor 2 negative breast cancer. Breast Cancer Res. 2012, 14, R62. [Google Scholar] [CrossRef]
  59. Zhang, Y.; He, Y.; Lu, L.-L.; Zhou, Z.-Y.; Wan, N.-B.; Li, G.-P.; He, X.; Deng, H.-W. miRNA-192-5p impacts the sensitivity of breast cancer cells to doxorubicin via targeting peptidylprolyl isomerase A. Kaohsiung J. Med. Sci. 2019, 35, 17–23. [Google Scholar] [CrossRef]
  60. Zhang, Z.; Qiu, N.; Yin, J.; Zhang, J.; Liu, H.; Guo, W.; Liu, M.; Liu, T.; Chen, D.; Luo, K.; et al. SRGN crosstalks with YAP to maintain chemoresistance and stemness in breast cancer cells by modulating HDAC2 expression. Theranostics 2020, 10, 4290–4307. [Google Scholar] [CrossRef]
  61. Herok, M.; Wawrzynow, B.; Maluszek, M.J.; Olszewski, M.B.; Zylicz, A.; Zylicz, M. Chemotherapy of HER2- and MDM2-Enriched Breast Cancer Subtypes Induces Homologous Recombination DNA Repair and Chemoresistance. Cancers 2021, 13, 4501. [Google Scholar] [CrossRef]
  62. Gayarre, J.; Kamieniak, M.M.; Cazorla-Jiménez, A.; Muñoz-Repeto, I.; Borrego, S.; García-Donas, J.; Hernando, S.; Robles-Díaz, L.; García-Bueno, J.M.; Ramón y Cajal, T.; et al. The NER-related gene GTF2H5 predicts survival in high-grade serous ovarian cancer patients. J. Gynecol. Oncol. 2016, 27, e7. [Google Scholar] [CrossRef]
  63. Dai, C.-H.; Li, J.; Chen, P.; Jiang, H.-G.; Wu, M.; Chen, Y.-C. RNA interferences targeting the Fanconi anemia/BRCA pathway upstream genes reverse cisplatin resistance in drug-resistant lung cancer cells. J. Biomed. Sci. 2015, 22, 77. [Google Scholar] [CrossRef]
  64. Arora, S.; Kothandapani, A.; Tillison, K.; Kalman-Maltese, V.; Patrick, S.M. Downregulation of XPF–ERCC1 enhances cisplatin efficacy in cancer cells. DNA Repair 2010, 9, 745–753. [Google Scholar] [CrossRef]
  65. Wang, B.; Wang, M.; Zhang, W.; Xiao, T.; Chen, C.-H.; Wu, A.; Wu, F.; Traugh, N.; Wang, X.; Li, Z.J.N.p. Integrative analysis of pooled CRISPR genetic screens using MAGeCKFlute. Nat. Protoc. 2019, 14, 756. [Google Scholar] [CrossRef]
  66. Gao, X.; Wang, J.; Li, M.; Wang, J.; Lv, J.; Zhang, L.; Sun, C.; Ji, J.; Yang, W.; Zhao, Z.; et al. Berberine attenuates XRCC1-mediated base excision repair and sensitizes breast cancer cells to the chemotherapeutic drugs. J. Cell. Mol. Med. 2019, 23, 6797–6804. [Google Scholar] [CrossRef]
  67. Liu, M.-m.; Li, Z.; Han, X.-d.; Shi, J.-h.; Tu, D.-y.; Song, W.; Zhang, J.; Qiu, X.-l.; Ren, Y.; Zhen, L.-l. MiR-30e inhibits tumor growth and chemoresistance via targeting IRS1 in Breast Cancer. Sci. Rep. 2017, 7, 15929. [Google Scholar] [CrossRef]
  68. Upadhyay, A.K.; Ajay, A.K.; Singh, S.; Bhat, M.K. Cell Cycle Regulatory Protein 5 (Cdk5) is a Novel Downstream Target of ERK in Carboplatin Induced Death of Breast Cancer Cells (Supplementary Data). Curr. Cancer Drug Targets 2008, 8, 741–752. [Google Scholar] [CrossRef] [PubMed]
  69. Syu, J.-P.; Chi, J.-T.; Kung, H.-N. Nrf2 is the key to chemotherapy resistance in MCF7 breast cancer cells under hypoxia. Oncotarget 2016, 7, 14659–14672. [Google Scholar] [CrossRef]
  70. Chen, Z.; Huang, J.; Feng, Y.; Li, Z.; Jiang, Y. Profiling of specific long non-coding RNA signatures identifies ST8SIA6-AS1 AS a novel target for breast cancer. J. Gene Med. 2021, 23, e3286. [Google Scholar] [CrossRef]
  71. Lu, H.; Samanta, D.; Xiang, L.; Zhang, H.; Hu, H.; Chen, I.; Bullen, J.W.; Semenza, G.L. Chemotherapy triggers HIF-1–dependent glutathione synthesis and copper chelation that induces the breast cancer stem cell phenotype. Proc. Natl. Acad. Sci. USA 2015, 112, E4600–E4609. [Google Scholar] [CrossRef] [PubMed]
  72. Nedeljković, M.; Tanić, N.; Prvanović, M.; Milovanović, Z.; Tanić, N. Friend or foe: ABCG2, ABCC1 and ABCB1 expression in triple-negative breast cancer. Breast Cancer 2021, 28, 727–736. [Google Scholar] [CrossRef]
  73. Mohammadi, A.; Mansoori, B.; Aghapour, M.; Shirjang, S.; Nami, S.; Baradaran, B. The Urtica dioica extract enhances sensitivity of paclitaxel drug to MDA-MB-468 breast cancer cells. Biomed. Pharmacother. 2016, 83, 835–842. [Google Scholar] [CrossRef]
  74. Deng, X.; Kohanfars, M.; Hsu, H.M.; Souda, P.; Capri, J.; Whitelegge, J.P.; Chang, H.R. Combined Phosphoproteomics and Bioinformatics Strategy in Deciphering Drug Resistant Related Pathways in Triple Negative Breast Cancer. Int. J. Proteom. 2014, 2014, 390781. [Google Scholar] [CrossRef]
  75. Dou, D.; Ren, X.; Han, M.; Xu, X.; Ge, X.; Gu, Y.; Wang, X.; Zhao, S. CircUBE2D2 (hsa_circ_0005728) promotes cell proliferation, metastasis and chemoresistance in triple-negative breast cancer by regulating miR-512-3p/CDCA3 axis. Cancer Cell Int. 2020, 20, 454. [Google Scholar] [CrossRef]
  76. Zhang, Y.; Jiang, C.; Li, H.; Lv, F.; Li, X.; Qian, X.; Fu, L.; Xu, B.; Guo, X. Elevated Aurora B expression contributes to chemoresistance and poor prognosis in breast cancer. Int. J. Clin. Exp. Pathol. 2015, 8, 751–757. [Google Scholar]
  77. Lemasters, J.J. Evolution of Voltage-Dependent Anion Channel Function: From Molecular Sieve to Governator to Actuator of Ferroptosis. Front. Oncol. 2017, 7, 303. [Google Scholar] [CrossRef]
  78. Gomez-Puerto, M.C.; Folkerts, H.; Wierenga, A.T.J.; Schepers, K.; Schuringa, J.J.; Coffer, P.J.; Vellenga, E. Autophagy Proteins ATG5 and ATG7 Are Essential for the Maintenance of Human CD34+ Hematopoietic Stem-Progenitor Cells. Stem Cells 2016, 34, 1651–1663. [Google Scholar] [CrossRef]
  79. Galluzzi, L.; Vitale, I.; Aaronson, S.A.; Abrams, J.M.; Adam, D.; Agostinis, P.; Alnemri, E.S.; Altucci, L.; Amelio, I.; Andrews, D.W.; et al. Molecular mechanisms of cell death: Recommendations of the Nomenclature Committee on Cell Death 2018. Cell Death Differ. 2018, 25, 486–541. [Google Scholar] [CrossRef]
  80. Sung, Y.-J.; Kao, T.-Y.; Kuo, C.-L.; Fan, C.-C.; Cheng, A.N.; Fang, W.-C.; Chou, H.-Y.; Lo, Y.-K.; Chen, C.-H.; Jiang, S.S.; et al. Mitochondrial Lon sequesters and stabilizes p53 in the matrix to restrain apoptosis under oxidative stress via its chaperone activity. Cell Death Dis. 2018, 9, 697. [Google Scholar] [CrossRef] [PubMed]
  81. Inuzuka, H.; Shaik, S.; Onoyama, I.; Gao, D.; Tseng, A.; Maser, R.S.; Zhai, B.; Wan, L.; Gutierrez, A.; Lau, A.W.; et al. SCFFBW7 regulates cellular apoptosis by targeting MCL1 for ubiquitylation and destruction. Nature 2011, 471, 104–109. [Google Scholar] [CrossRef] [PubMed]
  82. Fan, Z.; Zheng, J.; Xue, Y.; Liu, X.; Wang, D.; Yang, C.; Ma, J.; Liu, L.; Ruan, X.; Wang, Z.; et al. NR2C2-uORF targeting UCA1-miR-627-5p-NR2C2 feedback loop to regulate the malignant behaviors of glioma cells. Cell Death Dis. 2018, 9, 1165. [Google Scholar] [CrossRef]
  83. Chai, J.; Du, C.; Wu, J.-W.; Kyin, S.; Wang, X.; Shi, Y. Structural and biochemical basis of apoptotic activation by Smac/DIABLO. Nature 2000, 406, 855–862. [Google Scholar] [CrossRef] [PubMed]
  84. Li, L.; Sun, B.; Gao, Y.; Niu, H.; Yuan, H.; Lou, H. STAT3 contributes to lysosomal-mediated cell death in a novel derivative of riccardin D-treated breast cancer cells in association with TFEB. Biochem. Pharmacol. 2018, 150, 267–279. [Google Scholar] [CrossRef] [PubMed]
  85. Han, J.-w.; Flemington, C.; Houghton, A.B.; Gu, Z.; Zambetti, G.P.; Lutz, R.J.; Zhu, L.; Chittenden, T. Expression of bbc3, a pro-apoptotic BH3-only gene, is regulated by diverse cell death and survival signals. Proc. Natl. Acad. Sci. USA 2001, 98, 11318–11323. [Google Scholar] [CrossRef] [PubMed]
  86. Zamarin, D.; García-Sastre, A.; Xiao, X.; Wang, R.; Palese, P. Influenza Virus PB1-F2 Protein Induces Cell Death through Mitochondrial ANT3 and VDAC1. PLoS Pathog. 2005, 1, e4. [Google Scholar] [CrossRef]
  87. Jiang, H.-Y.; Yang, Y.; Zhang, Y.-Y.; Xie, Z.; Zhao, X.-Y.; Sun, Y.; Kong, W.-J. The dual role of poly(ADP-ribose) polymerase-1 in modulating parthanatos and autophagy under oxidative stress in rat cochlear marginal cells of the stria vascularis. Redox Biol. 2018, 14, 361–370. [Google Scholar] [CrossRef]
  88. Cubillos-Ruiz, J.R.; Bettigole, S.E.; Glimcher, L.H. Tumorigenic and Immunosuppressive Effects of Endoplasmic Reticulum Stress in Cancer. Cell 2017, 168, 692–706. [Google Scholar] [CrossRef]
  89. Kang, M.A.; So, E.Y.; Simons, A.L.; Spitz, D.R.; Ouchi, T. DNA damage induces reactive oxygen species generation through the H2AX-Nox1/Rac1 pathway. Cell Death Dis. 2012, 3, e249. [Google Scholar] [CrossRef]
  90. Henriques, A.; Koliaraki, V.; Kollias, G. Mesenchymal MAPKAPK2/HSP27 drives intestinal carcinogenesis. Proc. Natl. Acad. Sci. USA 2018, 115, E5546–E5555. [Google Scholar] [CrossRef]
  91. Fan, Y.; Ge, N.; Wang, X.; Sun, W.; Mao, R.; Bu, W.; Creighton, C.J.; Zheng, P.; Vasudevan, S.; An, L.; et al. Amplification and over-expression of MAP3K3 gene in human breast cancer promotes formation and survival of breast cancer cells. J. Pathol. 2014, 232, 75–86. [Google Scholar] [CrossRef]
  92. Rufo, N.; Garg, A.D.; Agostinis, P. The Unfolded Protein Response in Immunogenic Cell Death and Cancer Immunotherapy. Trends Cancer 2017, 3, 643–658. [Google Scholar] [CrossRef] [PubMed]
  93. Man, S.M.; Karki, R.; Kanneganti, T.-D. Molecular mechanisms and functions of pyroptosis, inflammatory caspases and inflammasomes in infectious diseases. Immunol. Rev. 2017, 277, 61–75. [Google Scholar] [CrossRef]
  94. Newton, K. RIPK1 and RIPK3: Critical regulators of inflammation and cell death. Trends Cell Biol. 2015, 25, 347–353. [Google Scholar] [CrossRef]
  95. Berghe, T.V.; Demon, D.; Bogaert, P.; Vandendriessche, B.; Goethals, A.; Depuydt, B.; Vuylsteke, M.; Roelandt, R.; Wonterghem, E.V.; Vandenbroecke, J.; et al. Simultaneous Targeting of IL-1 and IL-18 Is Required for Protection against Inflammatory and Septic Shock. Am. J. Respir. Crit. Care Med. 2014, 189, 282–291. [Google Scholar] [CrossRef]
  96. Casson, C.N.; Yu, J.; Reyes, V.M.; Taschuk, F.O.; Yadav, A.; Copenhaver, A.M.; Nguyen, H.T.; Collman, R.G.; Shin, S. Human caspase-4 mediates noncanonical inflammasome activation against gram-negative bacterial pathogens. Proc. Natl. Acad. Sci. USA 2015, 112, 6688–6693. [Google Scholar] [CrossRef]
  97. Zheng, L.; Bidere, N.; Staudt, D.; Cubre, A.; Orenstein, J.; Chan, F.K.; Lenardo, M. Competitive Control of Independent Programs of Tumor Necrosis Factor Receptor-Induced Cell Death by TRADD and RIP1. Mol. Cell. Biol. 2006, 26, 3505–3513. [Google Scholar] [CrossRef]
  98. Lawlor, K.E.; Khan, N.; Mildenhall, A.; Gerlic, M.; Croker, B.A.; D’Cruz, A.A.; Hall, C.; Kaur Spall, S.; Anderton, H.; Masters, S.L.; et al. RIPK3 promotes cell death and NLRP3 inflammasome activation in the absence of MLKL. Nat. Commun. 2015, 6, 6282. [Google Scholar] [CrossRef] [PubMed]
  99. Guzmán, E.A. Regulated Cell Death Signaling Pathways and Marine Natural Products That Target Them. Mar. Drugs 2019, 17, 76. [Google Scholar] [CrossRef]
  100. García-Sanz, P.; Orgaz, L.; Fuentes, J.M.; Vicario, C.; Moratalla, R. Cholesterol and multilamellar bodies: Lysosomal dysfunction in GBA-Parkinson disease. Autophagy 2018, 14, 717–718. [Google Scholar] [CrossRef]
  101. Chinnaiyan, A.M.; O’Rourke, K.; Tewari, M.; Dixit, V.M. FADD, a novel death domain-containing protein, interacts with the death domain of fas and initiates apoptosis. Cell 1995, 81, 505–512. [Google Scholar] [CrossRef]
  102. Liu, C.-C.; Leclair, P.; Pedari, F.; Vieira, H.; Monajemi, M.; Sly, L.M.; Reid, G.S.; Lim, C.J. Integrins and ERp57 Coordinate to Regulate Cell Surface Calreticulin in Immunogenic Cell Death. Front. Oncol. 2019, 9, 411. [Google Scholar] [CrossRef]
  103. Bredel, M.; Scholtens, D.M.; Yadav, A.K.; Alvarez, A.A.; Renfrow, J.J.; Chandler, J.P.; Yu, I.L.Y.; Carro, M.S.; Dai, F.; Tagge, M.J.; et al. NFKBIA Deletion in Glioblastomas. N. Engl. J. Med. 2010, 364, 627–637. [Google Scholar] [CrossRef]
  104. Fujikawa, K.; Nakahara, K.; Takasugi, N.; Nishiya, T.; Ito, A.; Uchida, K.; Uehara, T. S-Nitrosylation at the active site decreases the ubiquitin-conjugating activity of ubiquitin-conjugating enzyme E2 D1 (UBE2D1), an ERAD-associated protein. Biochem. Biophys. Res. Commun. 2020, 524, 910–915. [Google Scholar] [CrossRef]
  105. Tang, D.; Kang, R.; Berghe, T.V.; Vandenabeele, P.; Kroemer, G. The molecular machinery of regulated cell death. Cell Res. 2019, 29, 347–364. [Google Scholar] [CrossRef]
  106. Shi, J.; Zhao, Y.; Wang, K.; Shi, X.; Wang, Y.; Huang, H.; Zhuang, Y.; Cai, T.; Wang, F.; Shao, F. Cleavage of GSDMD by inflammatory caspases determines pyroptotic cell death. Nature 2015, 526, 660–665. [Google Scholar] [CrossRef] [PubMed]
  107. Baines, C.P.; Kaiser, R.A.; Purcell, N.H.; Blair, N.S.; Osinska, H.; Hambleton, M.A.; Brunskill, E.W.; Sayen, M.R.; Gottlieb, R.A.; Dorn, G.W.; et al. Loss of cyclophilin D reveals a critical role for mitochondrial permeability transition in cell death. Nature 2005, 434, 658–662. [Google Scholar] [CrossRef] [PubMed]
  108. Ponder, K.G.; Boise, L.H. The prodomain of caspase-3 regulates its own removal and caspase activation. Cell Death Discov. 2019, 5, 56. [Google Scholar] [CrossRef] [PubMed]
  109. Pimentel-Muiños, F.X.; Seed, B. Regulated Commitment of TNF Receptor Signaling: A Molecular Switch for Death or Activation. Immunity 1999, 11, 783–793. [Google Scholar] [CrossRef]
  110. Fennelly, C.; Amaravadi, R.K. Lysosomal Biology in Cancer. Methods Mol. Biol. 2017, 1594, 293–308. [Google Scholar] [CrossRef]
  111. Sargeant, T.J.; Lloyd-Lewis, B.; Resemann, H.K.; Ramos-Montoya, A.; Skepper, J.; Watson, C.J. Stat3 controls cell death during mammary gland involution by regulating uptake of milk fat globules and lysosomal membrane permeabilization. Nat. Cell Biol. 2014, 16, 1057–1068. [Google Scholar] [CrossRef]
  112. Monteleone, M.; Stanley, A.C.; Chen, K.W.; Brown, D.L.; Bezbradica, J.S.; von Pein, J.B.; Holley, C.L.; Boucher, D.; Shakespear, M.R.; Kapetanovic, R.; et al. Interleukin-1β; Maturation Triggers Its Relocation to the Plasma Membrane for Gasdermin-D-Dependent and -Independent Secretion. Cell Rep. 2018, 24, 1425–1433. [Google Scholar] [CrossRef]
  113. Gröschel, M.; Basta, D.; Ernst, A.; Mazurek, B.; Szczepek, A.J. Acute Noise Exposure Is Associated with Intrinsic Apoptosis in Murine Central Auditory Pathway. Front. Neurosci. 2018, 12, 312. [Google Scholar] [CrossRef]
  114. Li, H.; Li, T.; Zhang, X. Identification of a Pyroptosis-Related Prognostic Signature Combined with Experiments in Hepatocellular Carcinoma. Front. Mol. Biosci. 2022, 9, 822503. [Google Scholar] [CrossRef]
  115. Rager, J.E. Chapter 8—The Role of Apoptosis-Associated Pathways as Responders to Contaminants and in Disease Progression. In Systems Biology in Toxicology and Environmental Health; Fry, R.C., Ed.; Academic Press: Boston, MA, USA, 2015; pp. 187–205. [Google Scholar]
  116. Kuriakose, T.; Man, S.M.; Subbarao Malireddi, R.K.; Karki, R.; Kesavardhana, S.; Place, D.E.; Neale, G.; Vogel, P.; Kanneganti, T.-D. ZBP1/DAI is an innate sensor of influenza virus triggering the NLRP3 inflammasome and programmed cell death pathways. Sci. Immunol. 2016, 1, aag2045. [Google Scholar] [CrossRef] [PubMed]
  117. Serrano-del Valle, A.; Anel, A.; Naval, J.; Marzo, I. Immunogenic Cell Death and Immunotherapy of Multiple Myeloma. Front. Cell Dev. Biol. 2019, 7, 50. [Google Scholar] [CrossRef]
  118. Gulla, A.; Morelli, E.; Hideshima, T.; Samur, M.K.; Bianchi, G.; Fulciniti, M.; Munshi, N.; Anderson, K.C. Optimizing Mechanisms for Induction of Immunogenic Cell Death to Improve Patient Outcome in Multiple Myeloma. Blood 2018, 132, 1932. [Google Scholar] [CrossRef]
  119. König, T.; Tröder, S.E.; Bakka, K.; Korwitz, A.; Richter-Dennerlein, R.; Lampe, P.A.; Patron, M.; Mühlmeister, M.; Guerrero-Castillo, S.; Brandt, U.; et al. The m-AAA Protease Associated with Neurodegeneration Limits MCU Activity in Mitochondria. Mol. Cell 2016, 64, 148–162. [Google Scholar] [CrossRef] [PubMed]
  120. Mufti, A.R.; Burstein, E.; Duckett, C.S. XIAP: Cell death regulation meets copper homeostasis. Arch. Biochem. Biophys. 2007, 463, 168–174. [Google Scholar] [CrossRef] [PubMed]
  121. Aichberger, K.J.; Mayerhofer, M.; Krauth, M.-T.; Vales, A.; Kondo, R.; Derdak, S.; Pickl, W.F.; Selzer, E.; Deininger, M.; Druker, B.J.; et al. Low-Level Expression of Proapoptotic Bcl-2–Interacting Mediator in Leukemic Cells in Patients with Chronic Myeloid Leukemia: Role of BCR/ABL, Characterization of Underlying Signaling Pathways, and Reexpression by Novel Pharmacologic Compounds. Cancer Res. 2005, 65, 9436–9444. [Google Scholar] [CrossRef]
  122. Alvarez, M.C.; Maso, V.; Torello, C.O.; Ferro, K.P.; Saad, S.T.O. The polyphenol quercetin induces cell death in leukemia by targeting epigenetic regulators of pro-apoptotic genes. Clin. Epigenetics 2018, 10, 139. [Google Scholar] [CrossRef]
  123. Derakhshan, A.; Chen, Z.; Van Waes, C. Therapeutic Small Molecules Target Inhibitor of Apoptosis Proteins in Cancers with Deregulation of Extrinsic and Intrinsic Cell Death Pathways. Clin. Cancer Res. 2017, 23, 1379–1387. [Google Scholar] [CrossRef]
  124. D’Orsi, B.; Mateyka, J.; Prehn, J.H.M. Control of mitochondrial physiology and cell death by the Bcl-2 family proteins Bax and Bok. Neurochem. Int. 2017, 109, 162–170. [Google Scholar] [CrossRef]
  125. Campbell, K.J.; Tait, S.W.G. Targeting BCL-2 regulated apoptosis in cancer. Open Biol. 2018, 8, 180002. [Google Scholar] [CrossRef] [PubMed]
  126. Chen, H.-C.; Kanai, M.; Inoue-Yamauchi, A.; Tu, H.-C.; Huang, Y.; Ren, D.; Kim, H.; Takeda, S.; Reyna, D.E.; Chan, P.M.; et al. An interconnected hierarchical model of cell death regulation by the BCL-2 family. Nat. Cell Biol. 2015, 17, 1270–1281. [Google Scholar] [CrossRef] [PubMed]
  127. Letai, A.; Bassik, M.C.; Walensky, L.D.; Sorcinelli, M.D.; Weiler, S.; Korsmeyer, S.J. Distinct BH3 domains either sensitize or activate mitochondrial apoptosis, serving as prototype cancer therapeutics. Cancer Cell 2002, 2, 183–192. [Google Scholar] [CrossRef]
  128. Frudd, K.; Burgoyne, T.; Burgoyne, J.R. Oxidation of Atg3 and Atg7 mediates inhibition of autophagy. Nat. Commun. 2018, 9, 95. [Google Scholar] [CrossRef]
  129. Ye, X.; Zhou, X.-J.; Zhang, H. Exploring the Role of Autophagy-Related Gene 5 (ATG5) Yields Important Insights Into Autophagy in Autoimmune/Autoinflammatory Diseases. Front. Immunol. 2018, 9, 2334. [Google Scholar] [CrossRef]
  130. Durgan, J.; Florey, O. Cancer cell cannibalism: Multiple triggers emerge for entosis. Biochim. Biophys. Acta (BBA)—Mol. Cell Res. 2018, 1865, 831–841. [Google Scholar] [CrossRef]
  131. Ke, F.F.S.; Vanyai, H.K.; Cowan, A.D.; Delbridge, A.R.D.; Whitehead, L.; Grabow, S.; Czabotar, P.E.; Voss, A.K.; Strasser, A. Embryogenesis and Adult Life in the Absence of Intrinsic Apoptosis Effectors BAX, BAK, and BOK. Cell 2018, 173, 1217–1230.e1217. [Google Scholar] [CrossRef]
  132. Newton, K.; Wickliffe, K.E.; Maltzman, A.; Dugger, D.L.; Reja, R.; Zhang, Y.; Roose-Girma, M.; Modrusan, Z.; Sagolla, M.S.; Webster, J.D.; et al. Activity of caspase-8 determines plasticity between cell death pathways. Nature 2019, 575, 679–682. [Google Scholar] [CrossRef]
  133. Campbell, G.R.; Bruckman, R.S.; Chu, Y.-L.; Trout, R.N.; Spector, S.A. SMAC Mimetics Induce Autophagy-Dependent Apoptosis of HIV-1-Infected Resting Memory CD4+ T Cells. Cell Host Microbe 2018, 24, 689–702.e687. [Google Scholar] [CrossRef]
  134. Shi, X.; Yang, J.; Liu, M.; Zhang, Y.; Zhou, Z.; Luo, W.; Fung, K.-M.; Xu, C.; Bronze, M.S.; Houchen, C.W.; et al. Circular RNA ANAPC7 Inhibits Tumor Growth and Muscle Wasting via PHLPP2–AKT–TGF-β Signaling Axis in Pancreatic Cancer. Gastroenterology 2022, 162, 2004–2017.e2002. [Google Scholar] [CrossRef]
  135. Jin, L.; Williamson, A.; Banerjee, S.; Philipp, I.; Rape, M. Mechanism of Ubiquitin-Chain Formation by the Human Anaphase-Promoting Complex. Cell 2008, 133, 653–665. [Google Scholar] [CrossRef]
  136. Endo, K.; Mizuguchi, M.; Harata, A.; Itoh, G.; Tanaka, K. Nocodazole induces mitotic cell death with apoptotic-like features in Saccharomyces cerevisiae. FEBS Lett. 2010, 584, 2387–2392. [Google Scholar] [CrossRef]
  137. Waring, P.; Müllbacher, A. Cell death induced by the Fas/Fas ligand pathway and its role in pathology. Immunol. Cell Biol. 1999, 77, 312–317. [Google Scholar] [CrossRef]
  138. Sharma, S.; Kelly, T.K.; Jones, P.A. Epigenetics in cancer. Carcinogenesis 2010, 31, 27–36. [Google Scholar] [CrossRef]
  139. Filippakopoulos, P.; Qi, J.; Picaud, S.; Shen, Y.; Smith, W.B.; Fedorov, O.; Morse, E.M.; Keates, T.; Hickman, T.T.; Felletar, I. Selective inhibition of BET bromodomains. Nature 2010, 468, 1067–1073. [Google Scholar] [CrossRef]
  140. Ding, J.; Li, M.; Wan, X.; Jin, X.; Chen, S.; Yu, C.; Li, Y. Effect of miR-34a in regulating steatosis by targeting PPARα expression in nonalcoholic fatty liver disease. Sci. Rep. 2015, 5, 13729. [Google Scholar] [CrossRef]
  141. Chen, Z.S.; Tiwari, A.K. Multidrug resistance proteins (MRPs/ABCCs) in cancer chemotherapy and genetic diseases. FEBS J. 2011, 278, 3226–3245. [Google Scholar] [CrossRef]
  142. Chen, H.H.; Kuo, M.T. Role of glutathione in the regulation of Cisplatin resistance in cancer chemotherapy. Met. Based Drugs 2010, 2010, 430939. [Google Scholar] [CrossRef]
  143. Lehmann, W.; Mossmann, D.; Kleemann, J.; Mock, K.; Meisinger, C.; Brummer, T.; Herr, R.; Brabletz, S.; Stemmler, M.P.; Brabletz, T. ZEB1 turns into a transcriptional activator by interacting with YAP1 in aggressive cancer types. Nat. Commun. 2016, 7, 10498. [Google Scholar] [CrossRef]
  144. Hickson, I.; Zhao, Y.; Richardson, C.J.; Green, S.J.; Martin, N.M.; Orr, A.I.; Reaper, P.M.; Jackson, S.P.; Curtin, N.J.; Smith, G.C. Identification and characterization of a novel and specific inhibitor of the ataxia-telangiectasia mutated kinase ATM. Cancer Res. 2004, 64, 9152–9159. [Google Scholar] [CrossRef]
  145. Duan, W.; Gao, L.; Aguila, B.; Kalvala, A.; Otterson, G.A.; Villalona-Calero, M.A. Fanconi anemia repair pathway dysfunction, a potential therapeutic target in lung cancer. Front. Oncol. 2014, 4, 368. [Google Scholar] [CrossRef] [PubMed]
  146. Whittaker, S.R.; Mallinger, A.; Workman, P.; Clarke, P.A. Inhibitors of cyclin-dependent kinases as cancer therapeutics. Pharmacol. Ther. 2017, 173, 83–105. [Google Scholar] [CrossRef]
  147. Malinge, J.-M.; Giraud-Panis, M.-J.; Leng, M. Interstrand cross-links of cisplatin induce striking distortions in DNA. J. Inorg. Biochem. 1999, 77, 23–29. [Google Scholar] [CrossRef]
  148. Nishimura, K.; Ishiai, M.; Horikawa, K.; Fukagawa, T.; Takata, M.; Takisawa, H.; Kanemaki, M.T. Mcm8 and Mcm9 form a complex that functions in homologous recombination repair induced by DNA interstrand crosslinks. Mol. Cell 2012, 47, 511–522. [Google Scholar] [CrossRef] [PubMed]
  149. Helderman, N.C.; Terlouw, D.; Bonjoch, L.; Golubicki, M.; Antelo, M.; Morreau, H.; van Wezel, T.; Castellví-Bel, S.; Goldberg, Y.; Nielsen, M. Molecular functions of MCM8 and MCM9 and their associated pathologies. iScience 2023, 26, 106737. [Google Scholar] [CrossRef]
  150. Zhou, L.; Zhang, Y.; Chen, S.; Kmieciak, M.; Leng, Y.; Lin, H.; Rizzo, K.A.; Dumur, C.I.; Ferreira-Gonzalez, A.; Dai, Y.; et al. A regimen combining the Wee1 inhibitor AZD1775 with HDAC inhibitors targets human acute myeloid leukemia cells harboring various genetic mutations. Leukemia 2015, 29, 807–818. [Google Scholar] [CrossRef]
  151. Sun, C.; Yin, J.; Fang, Y.; Chen, J.; Jeong, K.J.; Chen, X.; Vellano, C.P.; Ju, Z.; Zhao, W.; Zhang, D.; et al. BRD4 Inhibition Is Synthetic Lethal with PARP Inhibitors through the Induction of Homologous Recombination Deficiency. Cancer Cell 2018, 33, 401–416.e408. [Google Scholar] [CrossRef]
  152. Valdez, B.C.; Li, Y.; Murray, D.; Liu, Y.; Nieto, Y.; Champlin, R.E.; Andersson, B.S. Combination of a hypomethylating agent and inhibitors of PARP and HDAC traps PARP1 and DNMT1 to chromatin, acetylates DNA repair proteins, down-regulates NuRD and induces apoptosis in human leukemia and lymphoma cells. Oncotarget 2018, 9, 3908. [Google Scholar] [CrossRef]
  153. Yuan, S.; Sun, R.; Shi, H.; Chapman, N.M.; Hu, H.; Guy, C.; Rankin, S.; Kc, A.; Palacios, G.; Meng, X. VDAC2 loss elicits tumour destruction and inflammation for cancer therapy. Nature 2025, 640, 1062–1071. [Google Scholar] [CrossRef] [PubMed]
  154. Chin, H.S.; Li, M.X.; Tan, I.K.; Ninnis, R.L.; Reljic, B.; Scicluna, K.; Dagley, L.F.; Sandow, J.J.; Kelly, G.L.; Samson, A.L. VDAC2 enables BAX to mediate apoptosis and limit tumor development. Nat. Commun. 2018, 9, 4976. [Google Scholar] [CrossRef]
  155. Liu, N.Q.; De Marchi, T.; Timmermans, A.M.; Beekhof, R.; Trapman-Jansen, A.M.; Foekens, R.; Look, M.P.; van Deurzen, C.H.; Span, P.N.; Sweep, F.C. Ferritin heavy chain in triple negative breast cancer: A favorable prognostic marker that relates to a cluster of differentiation 8 positive (CD8+) effector T-cell response. Mol. Cell. Proteom. 2014, 13, 1814–1827. [Google Scholar] [CrossRef]
  156. Shan, P.; Yang, F.; Qi, H.; Hu, Y.; Zhu, S.; Sun, Z.; Zhang, Z.; Wang, C.; Hou, C.; Yu, J. Alteration of MDM2 by the Small Molecule YF438 Exerts Antitumor Effects in Triple-Negative Breast CancerYF438 Exerts Promising Antitumor Effects on TNBC via MDM2. Cancer Res. 2021, 81, 4027–4040. [Google Scholar] [CrossRef]
  157. Bernard, M.M.; McConnery, J.R.; Hoskin, D.W. [10]-Gingerol, a major phenolic constituent of ginger root, induces cell cycle arrest and apoptosis in triple-negative breast cancer cells. Exp. Mol. Pathol. 2017, 102, 370–376. [Google Scholar] [CrossRef]
  158. Yan, H.; Luo, B.; Wu, X.; Guan, F.; Yu, X.; Zhao, L.; Ke, X.; Wu, J.; Yuan, J. Cisplatin induces pyroptosis via activation of MEG3/NLRP3/caspase-1/GSDMD pathway in triple-negative breast cancer. Int. J. Biol. Sci. 2021, 17, 2606. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Analysis of genomic similarity between samples of patients with triple-negative breast cancer (TNBC) (baseline) and those cell lines archived in the Cancer Cell Line Encyclopedia (CCLE). (A) Hierarchical clustering analysis. (B) Spearman’s rank correlation of 916 CCLE cell lines with TNBC samples showing a poor response to chemotherapy. (C) Heatmap of single-sample Gene Set Enrichment Analysis (ssGSEA) scores for the 50 hallmark gene sets of the Molecular Signals Database (MSigDB) across TNBC patient samples and TNBC cell lines. (D) Summary table of the most representative cell line models from these three analysis methods. Similarities among cell lines in blue with those TNBC samples responding poorly to chemotherapy showed their significant representativeness for those cells.
Figure 1. Analysis of genomic similarity between samples of patients with triple-negative breast cancer (TNBC) (baseline) and those cell lines archived in the Cancer Cell Line Encyclopedia (CCLE). (A) Hierarchical clustering analysis. (B) Spearman’s rank correlation of 916 CCLE cell lines with TNBC samples showing a poor response to chemotherapy. (C) Heatmap of single-sample Gene Set Enrichment Analysis (ssGSEA) scores for the 50 hallmark gene sets of the Molecular Signals Database (MSigDB) across TNBC patient samples and TNBC cell lines. (D) Summary table of the most representative cell line models from these three analysis methods. Similarities among cell lines in blue with those TNBC samples responding poorly to chemotherapy showed their significant representativeness for those cells.
Cancers 17 03876 g001
Figure 2. Genomic similarity analysis between samples of patients with triple-negative breast cancer (TNBC) following chemotherapy and the Cancer Cell Line Encyclopedia (CCLE) cell line model: (A) Hierarchical clustering analysis. (B) Spearman’s rank correlation of 916 CCLE cell lines with TNBC samples, showing a poor response to chemotherapy. (C) Heatmap of single-sample Gene Set Enrichment Analysis (ssGSEA) scores for the 50 hallmark gene sets of the Molecular Signals Database (MSigDB) across TNBC patient samples and TNBC cell lines. (D) Summary table of the most representative cell line models from these three analysis methods. Cell lines marked in blue indicate significant representativeness for TNBC poor responders.
Figure 2. Genomic similarity analysis between samples of patients with triple-negative breast cancer (TNBC) following chemotherapy and the Cancer Cell Line Encyclopedia (CCLE) cell line model: (A) Hierarchical clustering analysis. (B) Spearman’s rank correlation of 916 CCLE cell lines with TNBC samples, showing a poor response to chemotherapy. (C) Heatmap of single-sample Gene Set Enrichment Analysis (ssGSEA) scores for the 50 hallmark gene sets of the Molecular Signals Database (MSigDB) across TNBC patient samples and TNBC cell lines. (D) Summary table of the most representative cell line models from these three analysis methods. Cell lines marked in blue indicate significant representativeness for TNBC poor responders.
Cancers 17 03876 g002
Figure 3. Genome-wide clustered regularly interspaced short palindromic repeats (CRISPR)-Cas9 negative screens in the MDA-MB-231 cell line. (A) Schematic diagram for genome-wide CRISPR by the Toronto KnockOut CRISPR library (TKOV3; Version 3) (CIS, Cisplatin; DOX, doxorubicin). (B) The missed sgRNAs were tested on Days 0 (T0) and 28. (C) Gini index of sgRNAs on Days 0 (T0) and 28. (D) Three-dimensional PAC plot of baseline, control (Ctrl), and two-drug treatment groups. PC1 to PC3 can explain more than 80% of the total information. (E) The correlation plot among baseline, control, and two-drug treatment group, each group containing triplicate samples.
Figure 3. Genome-wide clustered regularly interspaced short palindromic repeats (CRISPR)-Cas9 negative screens in the MDA-MB-231 cell line. (A) Schematic diagram for genome-wide CRISPR by the Toronto KnockOut CRISPR library (TKOV3; Version 3) (CIS, Cisplatin; DOX, doxorubicin). (B) The missed sgRNAs were tested on Days 0 (T0) and 28. (C) Gini index of sgRNAs on Days 0 (T0) and 28. (D) Three-dimensional PAC plot of baseline, control (Ctrl), and two-drug treatment groups. PC1 to PC3 can explain more than 80% of the total information. (E) The correlation plot among baseline, control, and two-drug treatment group, each group containing triplicate samples.
Cancers 17 03876 g003
Figure 4. Gene essentiality scores (β scores) reported using the MAGeCK–maximum-likelihood estimation algorithm in cisplatin treatment (A) and doxorubicin treatment (B). Essential gene enrichment pathways in cisplatin treatment (C) and doxorubicin treatment (D) by analysis using the Kyoto Encyclopedia of Genes and Genomes (KEGG).
Figure 4. Gene essentiality scores (β scores) reported using the MAGeCK–maximum-likelihood estimation algorithm in cisplatin treatment (A) and doxorubicin treatment (B). Essential gene enrichment pathways in cisplatin treatment (C) and doxorubicin treatment (D) by analysis using the Kyoto Encyclopedia of Genes and Genomes (KEGG).
Cancers 17 03876 g004
Figure 5. Cell growth inhibition of MDA-MB-231, MDA-MB-436, and HS578T transfected siRNA MCM9, followed by treatment with four serially diluted cisplatin doses for 120 h in (AC). (D) Drug combination study between cisplatin and MCM9 inhibitor KPT-185.
Figure 5. Cell growth inhibition of MDA-MB-231, MDA-MB-436, and HS578T transfected siRNA MCM9, followed by treatment with four serially diluted cisplatin doses for 120 h in (AC). (D) Drug combination study between cisplatin and MCM9 inhibitor KPT-185.
Cancers 17 03876 g005
Figure 6. Cell-death clustered regularly interspaced short palindromic repeats (CRISPR)-Cas9 double-knockout screens in the MDA-MB-231 cell line. (A) Schematic diagram for genome-wide CRISPR by the cell-death CDKO library. (B) Candidate gene selection for constructing the cell-death CDKO library. (C) Read counts Mapping ratio of plasmid, days 0 (T0), and days 28 (Tend) samples. (D) The Gini index of sgRNAs on plasmid, days 0 (T0), and days 28 (Trend) samples. (E) Correlation plot between the baseline Day 0 (T0) and end date Day 28 (Tend) groups, each group containing triplicate samples.
Figure 6. Cell-death clustered regularly interspaced short palindromic repeats (CRISPR)-Cas9 double-knockout screens in the MDA-MB-231 cell line. (A) Schematic diagram for genome-wide CRISPR by the cell-death CDKO library. (B) Candidate gene selection for constructing the cell-death CDKO library. (C) Read counts Mapping ratio of plasmid, days 0 (T0), and days 28 (Tend) samples. (D) The Gini index of sgRNAs on plasmid, days 0 (T0), and days 28 (Trend) samples. (E) Correlation plot between the baseline Day 0 (T0) and end date Day 28 (Tend) groups, each group containing triplicate samples.
Cancers 17 03876 g006
Figure 7. Identification of synergistic lethal (SL) gene pairs. (A) Venn plot of seven methods for scoring SL pairs. (B) Gene network by 242 SL plots. (C) Histogram displaying the frequency distribution of top hub genes in 242 SL gene pairs.
Figure 7. Identification of synergistic lethal (SL) gene pairs. (A) Venn plot of seven methods for scoring SL pairs. (B) Gene network by 242 SL plots. (C) Histogram displaying the frequency distribution of top hub genes in 242 SL gene pairs.
Cancers 17 03876 g007
Table 1. Genomic dataset in study of chemoresistance in breast cancer and genomic profiles of Cancer Cell Line Encyclopedia (CCLE) cell lines.
Table 1. Genomic dataset in study of chemoresistance in breast cancer and genomic profiles of Cancer Cell Line Encyclopedia (CCLE) cell lines.
DatasetSample SizePlatformSourcePublicationGenes and Pathways Discovered in Breast Cancer Chemoresistance
GSE2884461Affymetrix (Affymetrix, Santa Clara, CA, USA)PatientLaura et al., 2013 [31]chemoresistance-associated gene shows enrichment in Wnt, HIF1, p53, and Rho GTPases signaling pathways
GSE1872861Affymetrix (Affymetrix, Santa Clara, CA, USA)PatientKorde et al., 2010 [32]MAP-2, MACF1, VEGF-B, EGFR showed upregulation in poor responder after chemotherapy
GSE32646115Affymetrix (Affymetrix, Santa Clara, CA, USA)PatientMiyake et al., 2012 [33]GSTP1 expression predicts poor response to neoadjuvant chemotherapy in patient with ER-negative breast cancer
GSE36133 (CCLE)917 (21 TNBC)Affymetrix (Affymetrix, Santa Clara, CA, USA)Cell lineBarretina et al., 2010 [30].
Table 2. Well-studied genes in the essential gene list for doxorubicin and cisplatin treatment. BET, bromodomain and extra-terminal motif proteins BRD2, BRD3, BRD4, and BRDT; ROS, reactive oxygen species.
Table 2. Well-studied genes in the essential gene list for doxorubicin and cisplatin treatment. BET, bromodomain and extra-terminal motif proteins BRD2, BRD3, BRD4, and BRDT; ROS, reactive oxygen species.
GenePathwayValidation MethodInhibitorDrug Development StatusChemotherapy DrugCancer Type
CDC25B [52]Cell cycle regulation,
DNA damage response
In vitro (siRNA),
In vivo (xenograft model)
Thiostrepton, FDI-6, Siomycin APreclinicalPaclitaxel, CisplatinPlatinum-resistant
ovarian cancer
NCF1 [53]Autophagy, ROS productionIn vitro (siRNA)Ginsenoside RoPreclinical5-fluorouracilEsophageal cancer
USP22, HSP90AB1 [54]HSP90 regulation and ubiquitin pathwayIn vitro (siRNA),
In vivo (xenograft model)
Ganetespib, AT13387Phase II trialsIrinotecanMammary and
colorectal cancer
DNMT1 [55]DNA methylationIn vitro (siRNA),
In vivo (mice model, xenograft with gene knockdown)
DecitabineClinical
(used for other cancer types)
DecitabineTriple-negative
breast cancer
BCL2L1 [56]ApoptosisIn vitro (cell lines),
In vivo (mice model)
BikDD, LapatinibPreclinicalDoxorubicinbreast cancer
RUNX2 [57]BET inhibitionIn vitro (siRNA),
In vivo (xenograft model, CRISPR knockout)
BET inhibitors: JQ1, I-BET762Preclinical, Phase I/IICisplatin, TaxanesTriple-negative
breast cancer
HSP90 [58]Chaperone protein functionIn vitro (siRNA),
In vivo (xenograft model)
17-AAG, PU-H71Phase II/III trialsDoxorubicinHER2-negative
breast cancer
PPIA [59]miRNA regulationIn vitro (miRNA-192-5p mimic)--DoxorubicinBreast cancer
RUNX1 [60]YAP signaling pathwayIn vitro (shRNA knockdown),
In vivo (xenograft)
--DoxorubicinBreast cancer
NBN [61] DNA repair, homologous recombinationIn vitro (siRNA)--Doxorubicin, Paclitaxel, CarboplatinHER2- and MDM2-enriched breast cancer subtypes
GTF2H5 [62]Nucleotide excision repair (NER)In vitro--Carboplatin, PaclitaxelHigh-grade serous ovarian cancer
FANCA, FANCG [63]DNA damage repair, Fanconi anemia/BRCA pathwayIn vitro (siRNA)--CisplatinDrug-resistant lung cancer
ERCC1 [64]Nucleotide excision repairIn vitro (siRNA),
In vivo (xenograft model)
--CisplatinVarious cancer types
XRCC1 [65]DNA repairIn vitro (siRNA)TriptolidePreclinicalCisplatinTriple-negative breast cancer
XRCC1 [66]Base excision repairIn vitro (siRNA)BerberinePreclinicalEpirubicin, Doxorubicin, Cyclophosphamide, 5-fluorouracil, Docetaxel, CisplatinBreast cancer
IRS1 [67]PI3K-AKT-mTOR signalingIn vitro (miRNA
and inhibitor)
Y-29794PreclinicalPaclitaxel, Carboplatin, Gemcitabine, Doxorubicin, CisplatinTriple-negative breast cancer
Cdk5 [68]Cell cycle regulation, carboplatin-induced cell deathIn vitro (siRNA)--CarboplatinBreast cancer
FANCL [63]Fanconi anemia/BRCA pathwayIn vitro (siRNA)--CisplatinLung cancer
NFE2L2 [69]Chemotherapy resistance, hypoxia responseIn vitro (siRNA, hypoxia exposure)--Cisplatin, Doxorubicin, and EtoposideBreast cancer
NBN [61]Homologous recombination DNA repairIn vitro (immuno-fluorescence, Western blot)--Docetaxel, Doxorubicin, and CyclophosphamideBreast cancer
HIST1H2BJ [70,71]Glutathione synthesis, copper chelationIn vitro (siRNA),
In vivo (mice)
--Doxorubicin, Paclitaxel,
5-fluorouracil
Breast cancer
ABCC1 [72]Drug efflux transportersIn vitro (siRNA)--Doxorubicin, Paclitaxel, CisplatinTriple-negative breast cancer
ZEB2 [73]ATM activationIn vitro (siRNA)--Doxorubicin, Paclitaxel, CisplatinBreast cancer
CDK5 [74]Drug resistance-related pathwaysIn vitro (siRNA)--Paclitaxel, Cisplatin, and DoxorubicinTriple-negative breast cancer
CDCA3 [75]Cell proliferation, metastasis, chemoresistanceIn vitro (siRNA,
RT-qPCR)
--Paclitaxel, Cisplatin, and DoxorubicinTriple-negative breast cancer
CDC25B [52]Cell cycle regulationIn vitro (siRNA)--Paclitaxel, CisplatinPlatinum-resistant ovarian cancer
ATM [76]Cell cycle regulationIn vitro (siRNA),
In vivo (xenograft mice)
--TaxanesBreast cancer
Table 3. Candidate genes for developing cell-death CRISPR double-knockout (CDKO) library.
Table 3. Candidate genes for developing cell-death CRISPR double-knockout (CDKO) library.
GenesCell_Death_ModeSynonymsReference
VDAC3FerroptosisVDAC-3, HD-VDAC3, HVDACLemasters 2017 [77]
VDAC2FerroptosisVDAC-2, HVDAC2, PORLemasters 2017 [77]
ATG7AutophagyUbiquitin-activating enzyme E1-like protein, ubiquitin-like modifier-activating enzyme ATG7Gomez-Puerto et al., 2016 [78]
UBE2E1Mitotic_CDUBCH6Galluzzi et al., 2018 [79]
TP53MPTTumor protein 53, P53Sung et al., 2018 [80]
MCL1ApoptosisTM, EAT, MCL1L1Inuzuka et al., 2011 [81]
NR2C2ApoptosisTAK1, TR4Fan et al., 2018 [82]
DIABLOApoptosisSMAC, DFNA64Chai et al., 2000 [83]
STAT3ParthanatosSignal transducer and activator of transcription 3 (acute-phase response factor), DNA-binding protein APRFLi et al., 2018 [84]
BBC3ApoptosisPUMA, JFY1Han et al., 2001 [85]
VDAC1MPTPORIN, VDAC-1Zamarin et al., 2005 [86]
PARP1ParthanatosPoly [ADP-Ribose] Polymerase 1, Poly [ADP-Ribose] Synthase 1, EC 2.4.2.30, ADPRT 1, PARP-1Jiang et al., 2018 [87]
EIF2AK3ParthanatosPERK, PEK, HsPEKCubillos-Ruiz et al., 2017 [88]
NOXA1Apoptosisp51NOX, NY-CO-31Kang et al., 2012 [89]
MAPKAPK2ApoptosisMK2, MK-2, MAPKAP-K2Henriques et al., 2018 [90]
MAP3K3EfferocytosisMAPKKK3, MEKK3Fan et al., 2014 [91]
ERN1ParthanatosInositol-requiring protein 1, Inositol-requiring enzyme 1Rufo et al., 2017 [92]
CASP1PyroptosisInflammasome (Nalp3, Asc, Casp1)Man et al., 2017 [93]
RIPK1ApoptosisIMD57, RIP, RIP1Newton 2015 [94]
TICAM1ApoptosisIIAE6, TRIF, MyD88-3Galluzzi et al., 2018 [79]
IL18PyroptosisIFIF, IL-18, IL1F4Berghe et al., 2014 [95]
CASP4PyroptosisICEREL-II, ICH-2Casson et al., 2015 [96]
TRADDApoptosisHs. 89862Zheng et al., 2006 [97]
MLKLNecroptosishMLKLLawlor et al. [98]
HK1ParthanatosHK1-Tb, HK1-Tc, HMSNR, HXK1Guzmán 2019 [99]
GBAAutophagyGLCM_HUMAN, GLUCGarcía-Sanz et al., 2018 [100]
FADDApoptosisGIG3, MORTChinnaiyan et al., 1995 [101]
RAPGEF3MPTEPAC1, HSU79275, CAP-GEFIGalluzzi et al., 2018 [79]
PDIA3ParthanatosEndoplasmic reticulum resident protein 60, protein disulfide isomerase-associated 3Liur et al., 2019 [102]
NFKBIAEfferocytosisEDAID2, IKBA, MAD-3Bredel et al., 2010 [103]
UBE2D1Mitotic_CDE1 (17)KB1, SFT, UBC4/5Fujikawa et al., 2020 [104]
CDH1ParthanatosEpithelial cadherin, CAM 120/80, CDHE, UVOTang et al., 2019 [105]
GSDMDPyroptosisDF5L, DFNA5LShi et al., 2015 [106]
PPIFApoptosisCYP3, CypD, CyP-MBaines et al., 2005 [107]
CASP3ApoptosisCPP32, CPP32BPonder et al., 2019 [108]
TNFRSF1BEfferocytosisCD120b, TBPII, TNF-R-II, TNFR2Pimentel-Muiños et al., 1999 [109]
TNFRSF1AApoptosisCD120a, FPF, TBP1, TNF-R, R55Galluzzi et al., 2018 [79]
LAMP1ParthanatosCD107 antigen-like family member A, LAMP-1, lysosomal-associated membrane protein 1Fennelly et al., 2017 [110]
CTSLLysosomal_CDCATL1, MEP, CTSLSargeant et al., 2014 [111]
IL1BPyroptosisCatabolin, IL-1 beta, IL1F2, pro-interleukin-1-betaMonteleone et al., 2018 [112]
CASP6ApoptosisCASP3/6/7, caspase 3, 6, 7, caspase-3, -6, and -7Gröschel et al., 2018 [113]
SCAF11PyroptosisCASP11, SFRS2IP, SIP1, SRRP129Li et al., 2022 [114]
CASP7ApoptosisCASP-7, CMH-1, ICE-LAP3Rager 2015 [115]
MAVSPyroptosisCARDIF, IPS-1, IPS1, VISAKuriakose et al., 2016 [116]
ATF6ParthanatosCAMP-dependent transcription factor ATF-6 alpha, activating transcription factor 6 alpha, ATF6-AlphSerrano-del Valle et al., 2019 [117]
CALRParthanatosCalregulin, CRP55, ERp60, HACBP, Grp60Gullai et al., 2018 [118]
MCUMPTC10orf42, CCDC109A, HsMCUKönig et al., 2016 [119]
XIAPApoptosisBIRC4, API3, IPA-3, XLP2, hIAP-3Mufti et al., 2007 [120]
BCL2L11ApoptosisBIM, BAM, BOD,Alvarez et al., 2018 [121,122]
BIDApoptosisBID isoform L(2), BID isoform Si6, FP497Derakhshan et al., 2017 [123]
BOKApoptosisBCL2L9D’Orsi et al., 2017 [124]
BCL2ApoptosisBCL2, apoptosis regulator, B-cell CLL/Lymphoma, PPP1R50, Bcl-2Campbell et al., 2018 [125]
BCL2L1ApoptosisBCL-XL, BCLX, BCL2LChen et al., 2015 [126]
BADApoptosisBBC2, BCL2L8Letai et al., 2002 [127]
ATG3AutophagyAutophagy-related protein 3, APG3L, HApg, APG3Frudd et al., 2018 [128]
ATG5AutophagyAutophagy protein 5, ATG5 autophagy-related 5 homologYe et al., 2018 [129]
RHOAParthanatosARH12, RHO12, ARHADurgan et al., 2018 [130]
BAXApoptosisApoptosis regulator BAX, Bcl-2-like protein 4, Bcl2-L-4Ke et al., 2018 [131]
CASP8ApoptosisAPLS2B, CAP4, FLICENewton et al., 2019 [132]
BIRC2AutophagyAPI1, MIHB, baculoviral IAP repeat-containing 2Campbell et al., 2018 [133]
ANAPC7Mitotic_CDAPC7Shi et al., 2022 [134]
ANAPC10Mitotic_CDAPC10, DOC1Jin et al., 2008 [135]
CDC26Mitotic_CDANAPC12, APC12Endo et al., 2010 [136]
FASApoptosisALPS1A, APT1, CD95, APO-1Waring et al., 1999 [137]
Table 4. CRISPR double-knockout (CDKO) sequencing statistics.
Table 4. CRISPR double-knockout (CDKO) sequencing statistics.
SampleReadsMappedMapped
%
Zero_
Counts
Zero_
Counts%
Gini_
Index
Above_
Threshold
Above_
Threshold%
Library/Plasmid9,272,5498,159,44888.001410.270.0548,92192.48
T0_112,431,41911,383,72091.578921.690.0647,10189.04
T0_215,647,05114,281,48591.277221.360.0647,66990.11
T0_315,104,60113,610,18190.1113712.590.0547,54789.88
Tend_110,004,4989,114,02791.101090.210.0550,65895.76
Tend_215,386,44713,994,09490.95480.090.0551,87598.06
Tend_315,837,76214,264,60390.07390.070.0551,86698.05
Table 5. Hub genes with targeting inhibitor.
Table 5. Hub genes with targeting inhibitor.
NamePathwayDrug Bank IDInhibitor NamePrevious Study in TNBCInhibitors of Gene Pairs
FTH1AutophagyDB00852Pseudo-ephedrinePrognostic marker [155]YF438 (HDAC1)
HDAC1Cell cycleDB02546YF438anti-TNBC activity [156]Pseudoephedrine (FTH1), Fostamatinib (MAP3K3), Aluminum monostearate (VDAC2), Bryostatin 1 (DIABLO)
MAP3K3Effero-cytosisDB12010Fostamatinib--YF438 (HDAC1), Aluminum monostearate (VDAC2)
DIABLOApoptosisDB11752Bryostatin 1anti-TNBC activity [157]YF438 (HDAC1), Aluminum monostearate (VDAC2)
VDAC2FerroptosisDB01375Aluminum monostearate--Bryostatin 1 (DIABLO), YF438 (HDAC1),
Fostamatinib (MAP3K3)
CASP1PyroptosisDB00945Acetylsalicylic acidanti-TNBC activity [158]YF438 (HDAC1), Aluminum monostearate (VDAC2), Fostamatinib (MAP3K3)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shao, S.; Li, S.; Huo, Y.; Tang, S.; Gökbağ, B.; Fan, K.; Huang, Y.; Wang, L.; Nagy, G.; Parvin, J.; et al. CRISPR-Cas9 Genome and Double-Knockout Screening to Identify Novel Therapeutic Targets for Chemoresistance in Triple-Negative Breast Cancer. Cancers 2025, 17, 3876. https://doi.org/10.3390/cancers17233876

AMA Style

Shao S, Li S, Huo Y, Tang S, Gökbağ B, Fan K, Huang Y, Wang L, Nagy G, Parvin J, et al. CRISPR-Cas9 Genome and Double-Knockout Screening to Identify Novel Therapeutic Targets for Chemoresistance in Triple-Negative Breast Cancer. Cancers. 2025; 17(23):3876. https://doi.org/10.3390/cancers17233876

Chicago/Turabian Style

Shao, Shuai, Shangjia Li, Yang Huo, Shan Tang, Birkan Gökbağ, Kunjie Fan, Yirui Huang, Lingling Wang, Gregory Nagy, Jeffrey Parvin, and et al. 2025. "CRISPR-Cas9 Genome and Double-Knockout Screening to Identify Novel Therapeutic Targets for Chemoresistance in Triple-Negative Breast Cancer" Cancers 17, no. 23: 3876. https://doi.org/10.3390/cancers17233876

APA Style

Shao, S., Li, S., Huo, Y., Tang, S., Gökbağ, B., Fan, K., Huang, Y., Wang, L., Nagy, G., Parvin, J., Stover, D., Cheng, L., & Li, L. (2025). CRISPR-Cas9 Genome and Double-Knockout Screening to Identify Novel Therapeutic Targets for Chemoresistance in Triple-Negative Breast Cancer. Cancers, 17(23), 3876. https://doi.org/10.3390/cancers17233876

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop