Skip to main content
Erschienen in: Breast Cancer Research 1/2022

Open Access 01.12.2022 | Research

Towards resolution of the intron retention paradox in breast cancer

verfasst von: Jaynish S. Shah, Michael J. G. Milevskiy, Veronika Petrova, Amy Y. M. Au, Justin J. L. Wong, Jane E. Visvader, Ulf Schmitz, John E. J. Rasko

Erschienen in: Breast Cancer Research | Ausgabe 1/2022

Abstract

Background

After many years of neglect in the field of alternative splicing, the importance of intron retention (IR) in cancer has come into focus following landmark discoveries of aberrant IR patterns in cancer. Many solid and liquid tumours are associated with drastic increases in IR, and such patterns have been pursued as both biomarkers and therapeutic targets. Paradoxically, breast cancer (BrCa) is the only tumour type in which IR is reduced compared to adjacent normal breast tissue.

Methods

In this study, we have conducted a pan-cancer analysis of IR with emphasis on BrCa and its subtypes. We explored mechanisms that could cause aberrant and pathological IR and clarified why normal breast tissue has unusually high IR.

Results

Strikingly, we found that aberrantly decreasing IR in BrCa can be largely attributed to normal breast tissue having the highest occurrence of IR events compared to other healthy tissues. Our analyses suggest that low numbers of IR events in breast tumours are associated with poor prognosis, particularly in the luminal B subtype. Interestingly, we found that IR frequencies negatively correlate with cell proliferation in BrCa cells, i.e. rapidly dividing tumour cells have the lowest number of IR events. Aberrant RNA-binding protein expression and changes in tissue composition are among the causes of aberrantly decreasing IR in BrCa.

Conclusions

Our results suggest that IR should be considered for therapeutic manipulation in BrCa patients with aberrantly low IR levels and that further work is needed to understand the cause and impact of high IR in other tumour types.
Hinweise

Supplementary Information

The online version contains supplementary material available at https://​doi.​org/​10.​1186/​s13058-022-01593-1.
Ulf Schmitz and John E. J. Rasko should be regarded as joint senior authors

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
NR
Non-differentially retained introns
BrCa
Breast cancer
CCLE
Cancer Cell Line Encyclopedia
dIR
Differentially retained introns
ER
Oestrogen receptor
FDR
False discovery rate
HER2
Human epidermal growth factor 2
ID
Intron depth
IHC
Immunohistochemistry
IR
Intron retention
MSigDB
Molecular Signatures Database
NMD
Nonsense-mediated decay
PCA
Principal component analysis
PWM
Position weight matrix
RBP
RNA-binding protein
SE
Splice exact
SL
Splice left
SR
Splice right
TCGA
The Cancer Genome Atlas
TNBC
Triple-negative breast cancer

Background

Pre-mRNA splicing is a ubiquitous process that is crucial for the maintenance of transcriptomic complexity and gene expression regulation in eukaryotic cells [1, 2]. Perturbations to this highly calibrated system can have severe consequences and lead to diseases including cancer [36]. In this context, numerous studies describing intron retention (IR) in disease have shed light on the mechanisms leading to aberrant and pathological IR [79].
The importance of IR in cancer has been emphasized following landmark discoveries about (i) aberrant IR patterns in leukaemia [10, 11], (ii) IR as a source of neoepitopes [12], (iii) tumour suppressor gene inactivation by intronic polyadenylation [13], (iv) IR-based biomarkers [14, 15], and (v) IR as a therapeutic target [16].
IR is regulated by cis- and trans-acting modulators [2, 17, 18] facilitating cellular responses to a range of environmental stimuli [19]. Intron-retaining mRNA transcripts are often degraded via nonsense-mediated decay (NMD), thereby causing downregulation of the host gene. The burden of IR in disease is governed by perturbations to mechanisms known to regulate this form of alternative splicing, including mutations, splicing factor dysregulation, and epigenetic variations.
However, despite the cumulative evidence for the importance of IR in cancer, a systematic analysis of IR regulation in breast cancer (BrCa) and the role of aberrant IR in BrCa biology has not been conducted to date. In this study, we sought to resolve the paradox wherein breast cancer exhibits reduced IR, which is an important consequence of alternative splicing.
We analysed 615 BrCa patient transcriptomes which included four major molecular subtypes (Luminal A, Luminal B, Basal, and Her2 positive). We confirmed a consistent downregulation of IR in BrCa. However, we also observed that normal breast tissue has a significantly higher IR event frequency compared to other healthy tissues. The number of IR events correlated with survival in the luminal B BrCa subtype. Differences in IR frequencies are largely influenced by the tissue’s cellular composition as well as specific dysregulated RNA-binding proteins (RBPs).

Methods

RNA-sequencing data/patient samples

We retrieved data from nine tumour types and healthy adjacent tissue, including 615 BrCa patient samples generated by the TCGA. We used TCGA metadata to assign the samples to molecular subtypes (i.e. Luminal A, Luminal B, human epidermal growth factor receptor 2 (HER2)-enriched, and Basal-like) based on the PAM50 classification system. On one occasion (Additional file 1: Fig. S2C), we grouped samples based on the immunohistochemical (IHC) score for HER2.
Only samples for which sequencing had been performed at > 40 M read depth were selected for analysis. Moreover, only tumour types with at least 20 matched tumour/normal tissue samples were considered. RNA-seq data were downloaded as BAM files using the R/Bioconductor package TCGAbiolinks [20] and the command-line tool gdc-client v1.4.0 (github.com/NCI-GDC/gdc-client) under an approved data access application. All files were checked for integrity. Harmonized gene expression data in the form of HTseq counts (RRID:SCR_005514) [21] were downloaded using TCGAbiolinks (RRID:SCR_017683).

mRNA-sequencing and data analysis—MCF7 and MCF10A cells

Total RNA was isolated from MCF7 and MCF10A cells using TRIzol (Invitrogen). The RNA quality was assessed using RNA 6000 Nano Chips on an Agilent Bioanalyzer (Agilent Technologies) to confirm an RNA integrity score of > 7.0. mRNA-seq was performed by Macrogen (Korea; RRID:SCR_014454) using the Illumina Hi-Seq 2000 platform. RNA-seq libraries were prepared from > 1 μg of total RNA using TruSeq RNA sample prep kit (Illumina) according to the manufacturers’ instructions.

Differential IR and gene expression analyses

IR was quantified using IRFinder v1.2.0 [17], using the Ensembl human genome (hg38, release 86; RRID:SCR_002344) as reference. The IRFinder algorithm measures 20 parameters for IR detection in each sample, including the median number of reads mapping to each nucleotide across the intron length (intron depth, ID), the ratio of nucleotides within an intron with mapped reads (coverage), the number of reads that map to the 5′ flanking exon and to another exon within the same gene (splice left, SL), the number of reads that map to the 3′ flanking exon and to another exon within the same gene (splice right, SR), the number of reads spanning the exon-exon junction (splice exact, SE) as well as the IR ratio:
$$\frac{\mathrm{ID}}{\mathrm{ID}+\mathrm{ max}(\mathrm{SL},\mathrm{ SR})}$$
Some selection criteria for IR events were chosen to minimize the chance of false positive IR calling while at the same time maintaining sufficient sensitivity to avoid too many false negative events. The following criteria were used for quantifying the number of IR events in a sample:
(1)
\(0.7\le \frac{\mathrm{SL}}{SR}\le 1.3;\)
This filter ensures that introns are flanked by constitutive exons.
 
(2)
\((\mathrm{SL}+\mathrm{SR})>10\) in ≥ 50% of samples;
Flanking exons need to be well expressed in most samples to avoid false positive IR events.
 
(3)
\(\mathrm{coverage}>0.5\) in ≥ 50% of samples;
Only introns with extensive coverage in the majority of samples are considered to prevent confounding factors that could lead to false IR calling.
 
(4)
\(\mathrm{IR}>0.05\) in ≥ 50% normal or cancer samples.
We included only samples with at least 40M reads to facilitate accurate IR quantification. We considered a 5% inclusion rate for IR to be of biological relevance.
 
The number of IR events in a sample was determined based on introns with an IR ratio > 0.1 and meeting the filtering criteria described above. Introns not meeting these criteria were not considered as being retained. Beta regression was used to identify differentially retained introns (dIR) between cancer and adjacent normal tissues using the betareg R package [22]. Since IR ratios are proportional data with values between 0 and 1, we reasoned that beta regression was best suited to model IR and identify dIRs between normal and cancer tissues. An absolute difference in the IR ratio (ΔIR = IRCancer − IRNormal) of more than 0.1 with FDR-adjusted p < 0.05 was considered significant.
Dimensionality reduction, i.e. principal component analysis (PCA), of IR profiles was performed using the package factoextra (github.com/kassambara/factoextra).
Differential gene expression between normal breast tissue and BrCa was performed using the DESeq2 package (RRID:SCR_000154) [23]. Genes with an average read count > 10 in all samples were selected for differential gene expression analysis (n = 23,072). Genes with an absolute log2 fold change > 1 and FDR-adjusted p < 0.05 were considered significant. To identify genes that were specifically differentially expressed in BrCa, we removed genes that were differentially expressed in any of the other 8 cancers and determined specificity by computing the z-score on log fold change using the log fold change observed in BrCa as reference.

Gene Ontology and RBP analyses

Gene Ontology analysis was performed using the clusterProfiler package (RRID:SCR_016884) [24]. The false discovery rate (FDR) approach was used for multiple testing correction. The list of 1542 RBPs was taken from Gerstberger et al. [25].

Survival analysis

Patient survival data were provided by the TCGA consortium. Survival analysis was performed using packages Surv and survminer (github.com/kassambara/survminer).

RNA-binding protein motif detection

RNA-binding protein (RBP) motifs in position weight matrix format (PWM) were retrieved from the ATtRACT database (version 0.99β) [26], which contains 1196 motifs corresponding to 160 human RBPs. Sequences of 100 nt were extracted from the regions flanking retained and non-retained introns and scanned for the presence of motifs using the fimo tool provided by the meme suite [27].

Results

IR in breast tumours is reduced in contrast to high occurrence in normal breast tissue

To compare IR profiles across human cancers, we retrieved transcriptomics data for nine different solid tumours and matched adjacent normal tissues from The Cancer Genome Atlas (TCGA; Fig. 1A) and quantified IR using the IRFinder algorithm, which we have previously validated [17]. Overall, we identified a total of 11,943 unique IR events, of which 917 were shared among all nine cancers analysed.
Our analyses confirmed a previous report that BrCa is the only cancer in which the number of IR events is reduced compared to normal adjacent tissue (Fig. 1B) [28]. All other cancers exhibit increased IR compared to their matched adjacent normal tissue (Fig. 1B). However, we also noticed that the number of IR events in breast cancer itself was comparable with other cancers, while normal breast tissue presented with unusually high numbers of IR events (Fig. 1B). In fact, normal breast tissue had the highest IR frequencies compared to all other normal tissues (Fig. 1C). Therefore, reduced IR in tumours, which is unique to BrCa, can be largely attributed to normal breast tissue having the highest occurrence of IR events.
We applied beta regression models to identify differentially retained introns (dIRs) and found 3024 dIRs between normal and breast cancer (Additional file 1: Fig. S1A). Of these 210 were downregulated (in 160 genes) and 69 were upregulated (in 52 genes) with a ≥ 10% difference in the IR ratio (ΔIR ≥ 0.1). Downregulated IR events in BrCa are associated with processes related to cell cycle, nuclear division, and DNA replication among others (Additional file 1: Fig. S1B).

High IR is associated with improved survival in Luminal B subtype breast cancer

Next, we explored whether the specific pattern of IR in BrCa is associated with clinical features. As shown in Additional file 1: Fig. S2A and Fig. 1D, IR patterns were distinct between BrCa vs normal tissues as well as oestrogen receptors positive (ER+) vs negative (ER) samples, respectively. The human epidermal growth factor receptor 2 (HER2) amplified molecular subtype had the lowest average number of IR events (n = 1731) compared to the other three main subtypes Luminal A (n = 2089), Luminal B (n = 2113), and Basal (n = 2018) (Fig. 1E). Strikingly, HER2-amplified tumours are associated with a 2.9-fold increased hazard ratio (p = 0.001, Additional file 1: Fig. S2B). Moreover, advanced stage tumours (Stage III) had the lowest average number of IR events (n = 1913) compared to Stage II (n = 2064) and Stage I (n = 2210) tumours (Fig. 1F). Likewise, those tumours with the highest immunohistochemical (IHC) staining score for HER2 (score: 3) exhibited the lowest average number of IR events (n = 1825) compared to score 1 (n = 2095) or score 2 (n = 2137) tumours (Additional file 1: Fig. S2C). We also found that a high number of IR events is associated with better survival in patients with the Luminal B subtype (Fig. 1G; Additional file 1: Fig. S2D). Though, Luminal B is the only subtype where high IR is associated with a survival advantage (Additional file 1: Fig. S2E). Interestingly, the trend is reversed (although not significant) in Her2 positive breast tumours that do not belong to the luminal B subtype.

Putative trans-regulators of IR in breast cancer

To confirm that differences between tumour and normal breast tissue can be observed in vitro, we sequenced the transcriptomes of cultured ER+ MCF7 cells and non-tumorigenic MCF10A cells. Indeed, we observed a similar trend as in the TCGA cohort (Fig. 2A) and found that higher IR in the breast epithelial cell line (MCF10A) was associated with reduced gene expression (Additional file 1: Fig. S3). We also analysed sequencing data (Sequence Read Archive; SRA) of other cell lines representing each of the molecular subtypes (Additional file 1: Table S1). Comparing the number of IR events, we observed a similar trend as with the TCGA tumour samples, except for HER + HCC1419 cells, which have on average more IR events (though not significant) than cell lines of other subtypes (Additional file 1: Fig. S4).
To identify potential regulators of IR, we correlated IR frequencies in the TCGA-BRCA cohort with expression values (normalized RNA-seq counts) of ~ 23,000 genes (Additional file 2: Table S2). We performed Gene Ontology (GO) enrichment analysis on the top 5% genes with the highest (r > 0.41) and lowest (r < − 0.27) correlation coefficients to identify potential positive and negative regulators of IR, respectively. We found eight GO terms that were significantly enriched in positively correlated genes (p-adj. ≤ 0.05; Fig. 2B). Intriguingly, six of these GO terms were related to RNA splicing, with DDX39B, RBM25, PRPF39, and SRSF11 being among the genes that most highly correlate with the number of IR events in each sample (Fig. 2C).
The four genes that most strongly anti-correlate with the number of IR events include the mutase PGAM1, the membrane protein encoding SURF4, the mitochondrial transmembrane transporter SLC25A5, and the mitochondrial ATP Synthase F1 Subunit Beta (ATP5F1B) (Fig. 2D). Strikingly, the top 10 most significant GO terms (out of 399) associated with genes that negatively correlate with the number of IR events correspond to mitochondrial processes and cellular energetics (Additional file 1: Fig. S5).

Highly proliferating cells have fewer IR events

Next, we investigated whether changes in IR frequencies are associated with changes to cellular states. Since the energy demand of a cell is tightly coupled with proliferation, we sought to examine a potential link between doubling times of BrCa cells and the number of IR events in 36 BrCa-related cell lines of the Cancer Cell Line Encyclopedia (CCLE). Indeed, we found that cells with a slower doubling time exhibit a higher number of IR events (Fig. 3A). While this correlation is fairly robust, it should be noted that CCLE doubling times are an error-prone surrogate for cancer cell proliferation.
To corroborate this result, we also investigated whether a common proliferation marker would inversely correlate with IR frequencies. Since immunohistochemistry (IHC) staining of Ki-67 is unavailable for the TCGA cohort, we tested whether the proliferation rate in tissues might be estimated based on MKI67 sequencing read counts. The MKI67 gene encodes the proliferation marker protein Ki-67. Using Human Protein Atlas data, we confirmed that MKI67 mRNA expression correlates with its Ki-67 protein staining intensities detected by IHC (Additional file 1: Fig. S6A). As expected, normalized MKI67 read counts were higher in all nine cancers when compared to the respective adjacent normal tissues (Additional file 1: Fig. S6B). This suggests that MKI67 read counts can be used as a proxy for IHC staining to estimate cellular proliferation rates. MKI67 expression was also found to be inversely correlated to the doubling time of 36 CCLE cell lines (Additional file 1: Fig. S6C). We observed that the number of IR events in samples of the TCGA-BRCA cohort negatively correlated with the proliferation rate (Fig. 3B). However, no correlation was observed when analysing normal and tumour samples separately (Additional file 1: Fig. S7).

The role of RNA-Binding Proteins in IR regulation

Next, we investigated genes that were specifically deregulated in BrCa. We identified a set of 150 genes that were only differentially expressed between BrCa and normal breast tissues, of which seven were RBPs (Fig. 4A). We calculated the z-score of each gene’s log fold change in BrCa versus the log fold change in other cancers in order to estimate the level of specificity of a gene being differentially expressed in BrCa only (Fig. 4B). Among the genes that are highly specifically over-expressed in BrCa are two known RBPs: ZFP36L2 and TUT4. ZFP36L2 promotes poly(A) tail removal of mRNA transcripts [30], while terminal uridylyl transferase 4 (TUT4) adds uridines to deadenylated transcripts [31]. Thus, both RBPs are mediators of mRNA decay, which could explain the observed reduction of IR transcripts in BrCa.
We also determined the frequency by which BrCa-specific genes occur in RNA-related gene sets (n = 138) in the Molecular Signatures Database (MSigDB; total ~ 5000 curated gene sets) [32]. While known RBPs such as ZFP36L2 and SNRNP25 (part of the minor U12-type spliceosome) are annotated in multiple RNA-related gene sets, other genes, that are specifically differentially expressed in BrCa, did not show any potential RNA-binding capabilities (Additional file 1: Fig. S8A).
In addition, we analysed differentially retained introns for occurrences of RBP binding motifs. We found that differentially retained introns were enriched in NELFE and SRSF9 binding sites in upstream exons (5'-up) and the 3' terminal region, respectively (Fig. 4C). Moreover, retained introns have fewer HNRNPH1 and HNRNPK binding sites in their 3' terminal region compared to non-retained introns (Fig. 4C; Additional file 1: Fig. S8B).
We conclude that RPBs are among the factors that facilitate reduced IR in BrCa by enabling efficient splicing of introns from pre-mRNA transcripts. However, RBPs specifically differentially expressed in BrCa are not among those with enriched binding motifs within and around differentially retained introns. This suggests that more complex, multifaceted regulatory mechanisms are causing the consistent reduction of IR in BrCa.

Tissue composition affects cancer IR profiles

Since the reduction in IR events in BrCa contrasts with all other cancer types analysed, we examined a possible contribution from the changing cell composition in the tumour microenvironment compared to healthy breast tissue. Gene signature-based and machine learning-based algorithms have been developed to deconvolute the cell-type composition in bulk RNA-sequencing data [33]. To compare cell environmental profiles of TCGA breast tumour samples versus healthy adjacent control samples, we used the cell-type deconvolution algorithm xCell [34], which was trained on 1,822 pure human cell-type-specific transcriptomes extracted from single-cell transcriptome profiling data. xCell analysis revealed that the breast tumour cell composition is distinct from other cancers (Additional file 1: Fig. S9). Among the most enriched cell types in BrCa are T helper cells, mesenchymal stem cells, and basophils. These predictions are supported by recent single-cell BrCa profiling studies [3537]. Normal breast tissue is enriched in endothelial cells, adipocytes, and dendritic cells (Fig. 5A).
Indeed, adipocyte and myeloid cell (M1 macrophages, basophils) enrichment are specific to normal breast tissue (Fig. 5B/C), which could explain the IR paradox in BrCa.
To determine whether cell types enriched in normal breast tissue have particularly high IR event frequencies, we retrieved RNA-sequencing data of 66 cell/tissue types from the ENCODE repository (Additional file 3: Table S3). Our analysis suggests that breast epithelial cells have the highest prevalence of IR followed by adipocytes (Fig. 5C), which could explain the drop in IR events in breast tumours.

Discussion

IR is omnipresent in vertebrate species [2, 38] and affects up to 80% of human protein-coding genes [17]. Numerous studies have highlighted the functional importance of retained introns in a wide range of biological functions including cell differentiation and development [12, 3942].
Since first reports in 2015 and subsequent confirmatory studies, BrCa has stood in stark contrast to other cancers concerning its burden of IR [28]. Dysregulation of cis- and trans-modulators can cause aberrant IR in various cancers [28]. For example, Dvinge et al. found that snRNA expression changes IR in the MCF7 cell line and to a certain degree in BrCa patient samples. They also showed that splicing factor knockdown can lead to increased IR in triple-negative BrCa (TNBC) [43]. Kim et al. found that some BrCa IR events anti-correlate with DNA methylation and that high IR levels in transcripts of migration and invasion inhibitory protein (MIIP) are associated with increased survival in European-American patients with invasive breast carcinoma [44].
We confirmed a consistent reduction of IR events in TCGA breast adenocarcinoma samples compared to adjacent normal breast tissue. While BrCa is the only cancer where this reduction is observed, IR frequencies are, in fact, comparable to those observed in other cancer types. This is due to the excessively large number of IR events in healthy breast tissue. Gascard et al. found that IR increases with differentiation state in normal human breast cells with fewer IR events in myoepithelial cells and seven times more events in luminal epithelial cells [45]. Indeed, our results suggest that an important factor in the reduction of IR events in breast tumours is the changing cell composition from adipocyte and epithelial cell-rich breast tissue to lymphocyte-infiltrated breast tumours. Adipocytes and epithelial cells have one of the highest IR frequencies in their transcriptomes compared to other cell types, while lymphocytes are known to have low IR counts [46]. Siang and co-workers have shown in this context that the RBP human antigen R (HuR), which is involved in pre-mRNA processing, is a negative regulator of adipogenesis [47]. Interestingly, Diaz-Muñoz et al. demonstrated that HuR binding to introns modulates alternative intron usage [48]. This may contribute to the high IR observed in adipocyte-rich normal breast tissue.
Aberrant IR has previously been associated with disease phenotypes and clinical outcomes. For example, IR in CMYC and SESTRIN1 genes was shown to be a reliable molecular marker separating melanoma from non-melanoma tumours [14] and Sznajder and colleagues have shown that IR can be used as a biomarker in hereditary repeat expansion diseases [15]. Despite marked differences between tumour and normal breast tissue, IR profiles in our analysis also differ between ER+ versus ER tumours. The survival advantages associated with high IR numbers in the Luminal B subtype suggest that this form of alternative splicing should be considered for therapeutic exploitation. However, the exact mechanisms whereby dynamic IR profiles lead to differences in clinical outcomes would be the subject of future studies.
The inverse relationship between IR and cell proliferation has been previously observed in the context of B cell development and T cell activation [46, 49]. Our results demonstrate that the number of IR events positively correlates with longer cancer cell doubling times and that more IR events are associated with slower cell proliferation in BrCa. Our data show that HER2 positive breast tumours have the lowest number of IR events. HER2 is known to induce cell proliferation in human cancers and is associated with poor prognosis in BrCa [50]. These results suggest that IR is a mechanism that counteracts tumour growth and would provide opportunities as therapeutic targets. Interestingly, the tumour suppressor Herstatin, expressed in healthy breast tissue [51], is a splice variant of the oncogene HER2, with a retained intron 8 [52]. Herstatin is a secreted autoinhibitor of Her2 [52], and intron 8 retention is regulated by RBPs of the HNRNP1 family (including H1, D, and A2/B1) [53]. Koedoot and co-workers have demonstrated that inhibition of cell proliferation can be achieved via splicing factor knockdown in TNBC [54].

Conclusions

In summary, our study sheds light on the unique causes and consequences of aberrant splicing in BrCa. The modulation of IR levels may offer novel opportunities for personalized BrCa treatment, especially in hormone- and chemotherapy-resistant subtypes.

Acknowledgements

The results shown here are based upon data generated by the TCGA Research Network: https://​www.​cancer.​gov/​tcga. The authors acknowledge The University of Sydney High Performance Computing service at The University of Sydney for providing resources that have contributed to the research data reported within this paper.

Declarations

Not applicable.
Not applicable.

Competing interests

J.E.J.R. has received honoraria or speaker fees (GSK, Miltenyi, Takeda, Gilead, Pfizer, Spark, Novartis, Celgene, bluebird bio); Director of Pathology (Genea); equity ownership (Genea, Rarecyte); consultant (Rarecyte, Imago); and chair, Gene Technology Technical Advisory, OGTR, Australian Government. The remaining authors declare no competing financial interests.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
1.
Zurück zum Zitat Mercer TR, Gerhardt DJ, Dinger ME, Crawford J, Trapnell C, Jeddeloh JA, et al. Targeted RNA sequencing reveals the deep complexity of the human transcriptome. Nat Biotechnol. 2011;30(1):99–104.CrossRef Mercer TR, Gerhardt DJ, Dinger ME, Crawford J, Trapnell C, Jeddeloh JA, et al. Targeted RNA sequencing reveals the deep complexity of the human transcriptome. Nat Biotechnol. 2011;30(1):99–104.CrossRef
2.
Zurück zum Zitat Schmitz U, Pinello N, Jia F, Alasmari S, Ritchie W, Keightley MC, et al. Intron retention enhances gene regulatory complexity in vertebrates. Genome Biol. 2017;18(1):216.CrossRef Schmitz U, Pinello N, Jia F, Alasmari S, Ritchie W, Keightley MC, et al. Intron retention enhances gene regulatory complexity in vertebrates. Genome Biol. 2017;18(1):216.CrossRef
3.
Zurück zum Zitat Douglas AG, Wood MJ. RNA splicing: disease and therapy. Brief Funct Genomics. 2011;10(3):151–64.CrossRef Douglas AG, Wood MJ. RNA splicing: disease and therapy. Brief Funct Genomics. 2011;10(3):151–64.CrossRef
4.
Zurück zum Zitat Monteuuis G, Schmitz U, Petrova V, Kearney PS, Rasko JEJ. Holding on to junk bonds: intron retention in cancer and therapy. Can Res. 2021;81(4):779–89.CrossRef Monteuuis G, Schmitz U, Petrova V, Kearney PS, Rasko JEJ. Holding on to junk bonds: intron retention in cancer and therapy. Can Res. 2021;81(4):779–89.CrossRef
5.
Zurück zum Zitat El Marabti E, Younis I. The cancer spliceome: reprograming of alternative splicing in cancer. Front Mol Biosci. 2018;5:80.CrossRef El Marabti E, Younis I. The cancer spliceome: reprograming of alternative splicing in cancer. Front Mol Biosci. 2018;5:80.CrossRef
6.
Zurück zum Zitat Bonnal SC, Lopez-Oreja I, Valcarcel J. Roles and mechanisms of alternative splicing in cancer—implications for care. Nat Rev Clin Oncol. 2020;17(8):457–74.CrossRef Bonnal SC, Lopez-Oreja I, Valcarcel J. Roles and mechanisms of alternative splicing in cancer—implications for care. Nat Rev Clin Oncol. 2020;17(8):457–74.CrossRef
7.
Zurück zum Zitat Fish L, Navickas A, Culbertson B, Xu Y, Nguyen HCB, Zhang S, et al. Nuclear TARBP2 drives oncogenic dysregulation of RNA splicing and decay. Mol Cell. 2019;75(5):967-81.e9.CrossRef Fish L, Navickas A, Culbertson B, Xu Y, Nguyen HCB, Zhang S, et al. Nuclear TARBP2 drives oncogenic dysregulation of RNA splicing and decay. Mol Cell. 2019;75(5):967-81.e9.CrossRef
8.
Zurück zum Zitat Koh CM, Bezzi M, Low DH, Ang WX, Teo SX, Gay FP, et al. MYC regulates the core pre-mRNA splicing machinery as an essential step in lymphomagenesis. Nature. 2015;523(7558):96–100.CrossRef Koh CM, Bezzi M, Low DH, Ang WX, Teo SX, Gay FP, et al. MYC regulates the core pre-mRNA splicing machinery as an essential step in lymphomagenesis. Nature. 2015;523(7558):96–100.CrossRef
9.
Zurück zum Zitat Ziff OJ, Taha DM, Crerar H, Clarke BE, Chakrabarti AM, Kelly G, et al. Reactive astrocytes in ALS display diminished intron retention. Nucleic Acids Res. 2021;49(6):3168–84.CrossRef Ziff OJ, Taha DM, Crerar H, Clarke BE, Chakrabarti AM, Kelly G, et al. Reactive astrocytes in ALS display diminished intron retention. Nucleic Acids Res. 2021;49(6):3168–84.CrossRef
10.
Zurück zum Zitat Schmitz U, Shah JS, Dhungel BP, Monteuuis G, Luu PL, Petrova V, et al. Widespread aberrant alternative splicing despite molecular remission in chronic myeloid leukaemia patients. Cancers (Basel). 2020;12(12):3738.CrossRef Schmitz U, Shah JS, Dhungel BP, Monteuuis G, Luu PL, Petrova V, et al. Widespread aberrant alternative splicing despite molecular remission in chronic myeloid leukaemia patients. Cancers (Basel). 2020;12(12):3738.CrossRef
11.
Zurück zum Zitat Tang AD, Soulette CM, van Baren MJ, Hart K, Hrabeta-Robinson E, Wu CJ, et al. Full-length transcript characterization of SF3B1 mutation in chronic lymphocytic leukemia reveals downregulation of retained introns. Nat Commun. 2020;11(1):1438.CrossRef Tang AD, Soulette CM, van Baren MJ, Hart K, Hrabeta-Robinson E, Wu CJ, et al. Full-length transcript characterization of SF3B1 mutation in chronic lymphocytic leukemia reveals downregulation of retained introns. Nat Commun. 2020;11(1):1438.CrossRef
12.
Zurück zum Zitat Smart AC, Margolis CA, Pimentel H, He MX, Miao D, Adeegbe D, et al. Intron retention is a source of neoepitopes in cancer. Nat Biotechnol. 2018;36(11):1056–8.CrossRef Smart AC, Margolis CA, Pimentel H, He MX, Miao D, Adeegbe D, et al. Intron retention is a source of neoepitopes in cancer. Nat Biotechnol. 2018;36(11):1056–8.CrossRef
13.
Zurück zum Zitat Lee SH, Singh I, Tisdale S, Abdel-Wahab O, Leslie CS, Mayr C. Widespread intronic polyadenylation inactivates tumour suppressor genes in leukaemia. Nature. 2018;561(7721):127–31.CrossRef Lee SH, Singh I, Tisdale S, Abdel-Wahab O, Leslie CS, Mayr C. Widespread intronic polyadenylation inactivates tumour suppressor genes in leukaemia. Nature. 2018;561(7721):127–31.CrossRef
14.
Zurück zum Zitat Giannopoulou AF, Konstantakou EG, Velentzas AD, Avgeris SN, Avgeris M, Papandreou NC, et al. Gene-specific intron retention serves as molecular signature that distinguishes melanoma from non-melanoma cancer cells in Greek patients. Int J Mol Sci. 2019;20(4):937.CrossRef Giannopoulou AF, Konstantakou EG, Velentzas AD, Avgeris SN, Avgeris M, Papandreou NC, et al. Gene-specific intron retention serves as molecular signature that distinguishes melanoma from non-melanoma cancer cells in Greek patients. Int J Mol Sci. 2019;20(4):937.CrossRef
15.
Zurück zum Zitat Sznajder LJ, Thomas JD, Carrell EM, Reid T, McFarland KN, Cleary JD, et al. Intron retention induced by microsatellite expansions as a disease biomarker. Proc Natl Acad Sci USA. 2018;115(16):4234–9.CrossRef Sznajder LJ, Thomas JD, Carrell EM, Reid T, McFarland KN, Cleary JD, et al. Intron retention induced by microsatellite expansions as a disease biomarker. Proc Natl Acad Sci USA. 2018;115(16):4234–9.CrossRef
16.
Zurück zum Zitat Seiler M, Yoshimi A, Darman R, Chan B, Keaney G, Thomas M, et al. H3B-8800, an orally available small-molecule splicing modulator, induces lethality in spliceosome-mutant cancers. Nat Med. 2018;24(4):497–504.CrossRef Seiler M, Yoshimi A, Darman R, Chan B, Keaney G, Thomas M, et al. H3B-8800, an orally available small-molecule splicing modulator, induces lethality in spliceosome-mutant cancers. Nat Med. 2018;24(4):497–504.CrossRef
17.
Zurück zum Zitat Middleton R, Gao D, Thomas A, Singh B, Au A, Wong JJ, et al. IRFinder: assessing the impact of intron retention on mammalian gene expression. Genome Biol. 2017;18(1):51.CrossRef Middleton R, Gao D, Thomas A, Singh B, Au A, Wong JJ, et al. IRFinder: assessing the impact of intron retention on mammalian gene expression. Genome Biol. 2017;18(1):51.CrossRef
18.
Zurück zum Zitat Petrova V, Song R, Nordström KJV, Walter J, Wong JJL, Armstrong NJ, et al. Increased chromatin accessibility facilitates intron retention in specific cell differentiation states. Nucleic Acids Res. 2022;50(20);11563–79.CrossRef Petrova V, Song R, Nordström KJV, Walter J, Wong JJL, Armstrong NJ, et al. Increased chromatin accessibility facilitates intron retention in specific cell differentiation states. Nucleic Acids Res. 2022;50(20);11563–79.CrossRef
19.
Zurück zum Zitat Monteuuis G, Wong JJL, Bailey CG, Schmitz U, Rasko JEJ. The changing paradigm of intron retention: regulation, ramifications and recipes. Nucleic Acids Res. 2019;47(22):11497–513. Monteuuis G, Wong JJL, Bailey CG, Schmitz U, Rasko JEJ. The changing paradigm of intron retention: regulation, ramifications and recipes. Nucleic Acids Res. 2019;47(22):11497–513.
20.
Zurück zum Zitat Colaprico A, Silva TC, Olsen C, Garofano L, Cava C, Garolini D, et al. TCGAbiolinks: an R/Bioconductor package for integrative analysis of TCGA data. Nucleic Acids Res. 2016;44(8):e71.CrossRef Colaprico A, Silva TC, Olsen C, Garofano L, Cava C, Garolini D, et al. TCGAbiolinks: an R/Bioconductor package for integrative analysis of TCGA data. Nucleic Acids Res. 2016;44(8):e71.CrossRef
21.
Zurück zum Zitat Anders S, Pyl PT, Huber W. HTSeq—a Python framework to work with high-throughput sequencing data. Bioinformatics. 2015;31(2):166–9.CrossRef Anders S, Pyl PT, Huber W. HTSeq—a Python framework to work with high-throughput sequencing data. Bioinformatics. 2015;31(2):166–9.CrossRef
22.
Zurück zum Zitat Cribari-Neto F, Zeileis A. Beta regression in R. J Stat Softw. 2010;34(2):1–24.CrossRef Cribari-Neto F, Zeileis A. Beta regression in R. J Stat Softw. 2010;34(2):1–24.CrossRef
23.
Zurück zum Zitat Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014;15(12):550.CrossRef Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014;15(12):550.CrossRef
24.
Zurück zum Zitat Yu G, Wang LG, Han Y, He QY. clusterProfiler: an R package for comparing biological themes among gene clusters. OMICS. 2012;16(5):284–7.CrossRef Yu G, Wang LG, Han Y, He QY. clusterProfiler: an R package for comparing biological themes among gene clusters. OMICS. 2012;16(5):284–7.CrossRef
25.
Zurück zum Zitat Gerstberger S, Hafner M, Tuschl T. A census of human RNA-binding proteins. Nat Rev Genet. 2014;15(12):829–45.CrossRef Gerstberger S, Hafner M, Tuschl T. A census of human RNA-binding proteins. Nat Rev Genet. 2014;15(12):829–45.CrossRef
27.
Zurück zum Zitat Bailey TL, Boden M, Buske FA, Frith M, Grant CE, Clementi L, et al. MEME SUITE: tools for motif discovery and searching. Nucleic Acids Res. 2009;37(Web Server issue):W202–8.CrossRef Bailey TL, Boden M, Buske FA, Frith M, Grant CE, Clementi L, et al. MEME SUITE: tools for motif discovery and searching. Nucleic Acids Res. 2009;37(Web Server issue):W202–8.CrossRef
28.
Zurück zum Zitat Dvinge H, Bradley RK. Widespread intron retention diversifies most cancer transcriptomes. Genome Med. 2015;7(1):45.CrossRef Dvinge H, Bradley RK. Widespread intron retention diversifies most cancer transcriptomes. Genome Med. 2015;7(1):45.CrossRef
29.
Zurück zum Zitat Audic S, Claverie JM. The significance of digital gene expression profiles. Genome Res. 1997;7(10):986–95.CrossRef Audic S, Claverie JM. The significance of digital gene expression profiles. Genome Res. 1997;7(10):986–95.CrossRef
30.
Zurück zum Zitat Hudson BP, Martinez-Yamout MA, Dyson HJ, Wright PE. Recognition of the mRNA AU-rich element by the zinc finger domain of TIS11d. Nat Struct Mol Biol. 2004;11(3):257–64.CrossRef Hudson BP, Martinez-Yamout MA, Dyson HJ, Wright PE. Recognition of the mRNA AU-rich element by the zinc finger domain of TIS11d. Nat Struct Mol Biol. 2004;11(3):257–64.CrossRef
31.
Zurück zum Zitat Lim J, Ha M, Chang H, Kwon SC, Simanshu DK, Patel DJ, et al. Uridylation by TUT4 and TUT7 marks mRNA for degradation. Cell. 2014;159(6):1365–76.CrossRef Lim J, Ha M, Chang H, Kwon SC, Simanshu DK, Patel DJ, et al. Uridylation by TUT4 and TUT7 marks mRNA for degradation. Cell. 2014;159(6):1365–76.CrossRef
32.
Zurück zum Zitat Subramanian A, Tamayo P, Mootha VK, Mukherjee S, Ebert BL, Gillette MA, et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc Natl Acad Sci USA. 2005;102(43):15545–50.CrossRef Subramanian A, Tamayo P, Mootha VK, Mukherjee S, Ebert BL, Gillette MA, et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc Natl Acad Sci USA. 2005;102(43):15545–50.CrossRef
33.
Zurück zum Zitat Avila Cobos F, Alquicira-Hernandez J, Powell JE, Mestdagh P, De Preter K. Benchmarking of cell type deconvolution pipelines for transcriptomics data. Nat Commun. 2020;11(1):5650.CrossRef Avila Cobos F, Alquicira-Hernandez J, Powell JE, Mestdagh P, De Preter K. Benchmarking of cell type deconvolution pipelines for transcriptomics data. Nat Commun. 2020;11(1):5650.CrossRef
34.
Zurück zum Zitat Aran D, Hu Z, Butte AJ. xCell: digitally portraying the tissue cellular heterogeneity landscape. Genome Biol. 2017;18(1):220.CrossRef Aran D, Hu Z, Butte AJ. xCell: digitally portraying the tissue cellular heterogeneity landscape. Genome Biol. 2017;18(1):220.CrossRef
35.
Zurück zum Zitat Wagner J, Rapsomaniki MA, Chevrier S, Anzeneder T, Langwieder C, Dykgers A, et al. A single-cell atlas of the tumor and immune ecosystem of human breast cancer. Cell. 2019;177(5):1330-45.e18.CrossRef Wagner J, Rapsomaniki MA, Chevrier S, Anzeneder T, Langwieder C, Dykgers A, et al. A single-cell atlas of the tumor and immune ecosystem of human breast cancer. Cell. 2019;177(5):1330-45.e18.CrossRef
36.
Zurück zum Zitat Jackson HW, Fischer JR, Zanotelli VR, Ali HR, Mechera R, Soysal SD, et al. The single-cell pathology landscape of breast cancer. Nature. 2020;578(7796):615–20.CrossRef Jackson HW, Fischer JR, Zanotelli VR, Ali HR, Mechera R, Soysal SD, et al. The single-cell pathology landscape of breast cancer. Nature. 2020;578(7796):615–20.CrossRef
37.
Zurück zum Zitat Pal B, Chen Y, Vaillant F, Capaldo BD, Joyce R, Song X, et al. A single-cell RNA expression atlas of normal, preneoplastic and tumorigenic states in the human breast. EMBO J. 2021;40(11):e107333.CrossRef Pal B, Chen Y, Vaillant F, Capaldo BD, Joyce R, Song X, et al. A single-cell RNA expression atlas of normal, preneoplastic and tumorigenic states in the human breast. EMBO J. 2021;40(11):e107333.CrossRef
38.
Zurück zum Zitat Braunschweig U, Barbosa-Morais NL, Pan Q, Nachman EN, Alipanahi B, Gonatopoulos-Pournatzis T, et al. Widespread intron retention in mammals functionally tunes transcriptomes. Genome Res. 2014;24(11):1774–86.CrossRef Braunschweig U, Barbosa-Morais NL, Pan Q, Nachman EN, Alipanahi B, Gonatopoulos-Pournatzis T, et al. Widespread intron retention in mammals functionally tunes transcriptomes. Genome Res. 2014;24(11):1774–86.CrossRef
39.
Zurück zum Zitat Edwards CR, Ritchie W, Wong JJ, Schmitz U, Middleton R, An X, et al. A dynamic intron retention program in the mammalian megakaryocyte and erythrocyte lineages. Blood. 2016;127(17):e24–34.CrossRef Edwards CR, Ritchie W, Wong JJ, Schmitz U, Middleton R, An X, et al. A dynamic intron retention program in the mammalian megakaryocyte and erythrocyte lineages. Blood. 2016;127(17):e24–34.CrossRef
40.
Zurück zum Zitat Green ID, Pinello N, Song R, Lee Q, Halstead JM, Kwok CT, et al. Macrophage development and activation involve coordinated intron retention in key inflammatory regulators. Nucleic Acids Res. 2020;48(12):6513–29.CrossRef Green ID, Pinello N, Song R, Lee Q, Halstead JM, Kwok CT, et al. Macrophage development and activation involve coordinated intron retention in key inflammatory regulators. Nucleic Acids Res. 2020;48(12):6513–29.CrossRef
41.
Zurück zum Zitat Wong JJ, Ritchie W, Ebner OA, Selbach M, Wong JW, Huang Y, et al. Orchestrated intron retention regulates normal granulocyte differentiation. Cell. 2013;154(3):583–95.CrossRef Wong JJ, Ritchie W, Ebner OA, Selbach M, Wong JW, Huang Y, et al. Orchestrated intron retention regulates normal granulocyte differentiation. Cell. 2013;154(3):583–95.CrossRef
42.
Zurück zum Zitat Yeom KH, Pan Z, Lin CH, Lim HY, Xiao W, Xing Y, et al. Tracking pre-mRNA maturation across subcellular compartments identifies developmental gene regulation through intron retention and nuclear anchoring. Genome Res. 2021;31(6):1106–19.CrossRef Yeom KH, Pan Z, Lin CH, Lim HY, Xiao W, Xing Y, et al. Tracking pre-mRNA maturation across subcellular compartments identifies developmental gene regulation through intron retention and nuclear anchoring. Genome Res. 2021;31(6):1106–19.CrossRef
43.
Zurück zum Zitat Dvinge H, Guenthoer J, Porter PL, Bradley RK. RNA components of the spliceosome regulate tissue- and cancer-specific alternative splicing. Genome Res. 2019;29(10):1591–604.CrossRef Dvinge H, Guenthoer J, Porter PL, Bradley RK. RNA components of the spliceosome regulate tissue- and cancer-specific alternative splicing. Genome Res. 2019;29(10):1591–604.CrossRef
44.
Zurück zum Zitat Kim D, Shivakumar M, Han S, Sinclair MS, Lee YJ, Zheng Y, et al. Population-dependent Intron retention and DNA methylation in breast cancer. Mol Cancer Res. 2018;16(3):461–9.CrossRef Kim D, Shivakumar M, Han S, Sinclair MS, Lee YJ, Zheng Y, et al. Population-dependent Intron retention and DNA methylation in breast cancer. Mol Cancer Res. 2018;16(3):461–9.CrossRef
45.
Zurück zum Zitat Gascard P, Bilenky M, Sigaroudinia M, Zhao J, Li L, Carles A, et al. Epigenetic and transcriptional determinants of the human breast. Nat Commun. 2015;6:6351.CrossRef Gascard P, Bilenky M, Sigaroudinia M, Zhao J, Li L, Carles A, et al. Epigenetic and transcriptional determinants of the human breast. Nat Commun. 2015;6:6351.CrossRef
46.
Zurück zum Zitat Ni T, Yang W, Han M, Zhang Y, Shen T, Nie H, et al. Global intron retention mediated gene regulation during CD4+ T cell activation. Nucleic Acids Res. 2016;44(14):6817–29.CrossRef Ni T, Yang W, Han M, Zhang Y, Shen T, Nie H, et al. Global intron retention mediated gene regulation during CD4+ T cell activation. Nucleic Acids Res. 2016;44(14):6817–29.CrossRef
47.
Zurück zum Zitat Siang DTC, Lim YC, Kyaw AMM, Win KN, Chia SY, Degirmenci U, et al. The RNA-binding protein HuR is a negative regulator in adipogenesis. Nat Commun. 2020;11(1):213.CrossRef Siang DTC, Lim YC, Kyaw AMM, Win KN, Chia SY, Degirmenci U, et al. The RNA-binding protein HuR is a negative regulator in adipogenesis. Nat Commun. 2020;11(1):213.CrossRef
48.
Zurück zum Zitat Diaz-Munoz MD, Bell SE, Fairfax K, Monzon-Casanova E, Cunningham AF, Gonzalez-Porta M, et al. The RNA-binding protein HuR is essential for the B cell antibody response. Nat Immunol. 2015;16(4):415–25.CrossRef Diaz-Munoz MD, Bell SE, Fairfax K, Monzon-Casanova E, Cunningham AF, Gonzalez-Porta M, et al. The RNA-binding protein HuR is essential for the B cell antibody response. Nat Immunol. 2015;16(4):415–25.CrossRef
49.
Zurück zum Zitat Ullrich S, Guigo R. Dynamic changes in intron retention are tightly associated with regulation of splicing factors and proliferative activity during B-cell development. Nucleic Acids Res. 2020;48(3):1327–40.CrossRef Ullrich S, Guigo R. Dynamic changes in intron retention are tightly associated with regulation of splicing factors and proliferative activity during B-cell development. Nucleic Acids Res. 2020;48(3):1327–40.CrossRef
50.
Zurück zum Zitat Iqbal N, Iqbal N. Human Epidermal Growth Factor Receptor 2 (HER2) in cancers: overexpression and therapeutic implications. Mol Biol Int. 2014;2014:852748.CrossRef Iqbal N, Iqbal N. Human Epidermal Growth Factor Receptor 2 (HER2) in cancers: overexpression and therapeutic implications. Mol Biol Int. 2014;2014:852748.CrossRef
51.
Zurück zum Zitat Koletsa T, Kostopoulos I, Charalambous E, Christoforidou B, Nenopoulou E, Kotoula V. A splice variant of HER2 corresponding to Herstatin is expressed in the noncancerous breast and in breast carcinomas. Neoplasia. 2008;10(7):687–96.CrossRef Koletsa T, Kostopoulos I, Charalambous E, Christoforidou B, Nenopoulou E, Kotoula V. A splice variant of HER2 corresponding to Herstatin is expressed in the noncancerous breast and in breast carcinomas. Neoplasia. 2008;10(7):687–96.CrossRef
52.
Zurück zum Zitat Doherty JK, Bond C, Jardim A, Adelman JP, Clinton GM. The HER-2/neu receptor tyrosine kinase gene encodes a secreted autoinhibitor. Proc Natl Acad Sci USA. 1999;96(19):10869–74.CrossRef Doherty JK, Bond C, Jardim A, Adelman JP, Clinton GM. The HER-2/neu receptor tyrosine kinase gene encodes a secreted autoinhibitor. Proc Natl Acad Sci USA. 1999;96(19):10869–74.CrossRef
53.
Zurück zum Zitat Silipo M, Gautrey H, Satam S, Lennard T, Tyson-Capper A. How is Herstatin, a tumor suppressor splice variant of the oncogene HER2, regulated? RNA Biol. 2017;14(5):536–43.CrossRef Silipo M, Gautrey H, Satam S, Lennard T, Tyson-Capper A. How is Herstatin, a tumor suppressor splice variant of the oncogene HER2, regulated? RNA Biol. 2017;14(5):536–43.CrossRef
54.
Zurück zum Zitat Koedoot E, van Steijn E, Vermeer M, Gonzalez-Prieto R, Vertegaal ACO, Martens JWM, et al. Splicing factors control triple-negative breast cancer cell mitosis through SUN2 interaction and sororin intron retention. J Exp Clin Cancer Res. 2021;40(1):82.CrossRef Koedoot E, van Steijn E, Vermeer M, Gonzalez-Prieto R, Vertegaal ACO, Martens JWM, et al. Splicing factors control triple-negative breast cancer cell mitosis through SUN2 interaction and sororin intron retention. J Exp Clin Cancer Res. 2021;40(1):82.CrossRef
Metadaten
Titel
Towards resolution of the intron retention paradox in breast cancer
verfasst von
Jaynish S. Shah
Michael J. G. Milevskiy
Veronika Petrova
Amy Y. M. Au
Justin J. L. Wong
Jane E. Visvader
Ulf Schmitz
John E. J. Rasko
Publikationsdatum
01.12.2022
Verlag
BioMed Central
Erschienen in
Breast Cancer Research / Ausgabe 1/2022
Elektronische ISSN: 1465-542X
DOI
https://doi.org/10.1186/s13058-022-01593-1

Weitere Artikel der Ausgabe 1/2022

Breast Cancer Research 1/2022 Zur Ausgabe

Erhebliches Risiko für Kehlkopfkrebs bei mäßiger Dysplasie

29.05.2024 Larynxkarzinom Nachrichten

Fast ein Viertel der Personen mit mäßig dysplastischen Stimmlippenläsionen entwickelt einen Kehlkopftumor. Solche Personen benötigen daher eine besonders enge ärztliche Überwachung.

15% bedauern gewählte Blasenkrebs-Therapie

29.05.2024 Urothelkarzinom Nachrichten

Ob Patienten und Patientinnen mit neu diagnostiziertem Blasenkrebs ein Jahr später Bedauern über die Therapieentscheidung empfinden, wird einer Studie aus England zufolge von der Radikalität und dem Erfolg des Eingriffs beeinflusst.

Erhöhtes Risiko fürs Herz unter Checkpointhemmer-Therapie

28.05.2024 Nebenwirkungen der Krebstherapie Nachrichten

Kardiotoxische Nebenwirkungen einer Therapie mit Immuncheckpointhemmern mögen selten sein – wenn sie aber auftreten, wird es für Patienten oft lebensgefährlich. Voruntersuchung und Monitoring sind daher obligat.

Costims – das nächste heiße Ding in der Krebstherapie?

28.05.2024 Onkologische Immuntherapie Nachrichten

„Kalte“ Tumoren werden heiß – CD28-kostimulatorische Antikörper sollen dies ermöglichen. Am besten könnten diese in Kombination mit BiTEs und Checkpointhemmern wirken. Erste klinische Studien laufen bereits.

Update Onkologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.