Next Article in Journal
Genome-Wide Characterization and Phylogenetic and Stress Response Expression Analysis of the MADS-Box Gene Family in Litchi (Litchi chinensis Sonn.)
Previous Article in Journal
Molecular and Functional Characterization of Three General Odorant-Binding Protein 2 Genes in Cydia pomonella (Lepidoptera: Tortricidae)
Previous Article in Special Issue
The Fate and Functionality of Alien tRNA Fragments in Culturing Medium and Cells of Escherichia coli
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Differential Gene Expression following DHX36/G4R1 Knockout Is Associated with G-Quadruplex Content and Cancer

by
Joseph M. Gumina
1,†,
Adam E. Richardson
1,†,
Mahmudul Hasan Shojiv
1,
Antonio E. Chambers
1,
Siara N. Sandwith
1,
Michael A. Reisinger
1,
Taylor J. Karns
1,
Tyler L. Osborne
1,
Hasna N. Alashi
1,
Quinn T. Anderson
1,
Meredith E. Sharlow
1,
Dylan C. Seiler
1,
Evan M. Rogers
1,
Anna R. Bartosik
2,
Melissa A. Smaldino
1,
James P. Vaughn
3,
Yuh-Hwa Wang
2,
Philip J. Smaldino
1,* and
Robert A. Haney
1,*
1
Department of Biology, Ball State University, Muncie, IN 47306, USA
2
School of Medicine, University of Virginia, Charlottesville, VA 22908, USA
3
NanoMedica LLC, Winston-Salem, NC 27106, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2024, 25(3), 1753; https://doi.org/10.3390/ijms25031753
Submission received: 27 November 2023 / Revised: 18 January 2024 / Accepted: 19 January 2024 / Published: 1 February 2024
(This article belongs to the Special Issue Bioinformatics of Unusual DNA and RNA Structures)

Abstract

:
G-quadruplexes (G4s) are secondary DNA and RNA structures stabilized by positive cations in a central channel formed by stacked tetrads of Hoogsteen base-paired guanines. G4s form from G-rich sequences across the genome, whose biased distribution in regulatory regions points towards a gene-regulatory role. G4s can themselves be regulated by helicases, such as DHX36 (aliases: G4R1 and RHAU), which possess the necessary activity to resolve these stable structures. G4s have been shown to both positively and negatively regulate gene expression when stabilized by ligands, or through the loss of helicase activity. Using DHX36 knockout Jurkat cell lines, we identified widespread, although often subtle, effects on gene expression that are associated with the presence or number of observed G-quadruplexes in promoters or gene regions. Genes that significantly change their expression, particularly those that show a significant increase in RNA abundance under DHX36 knockout, are associated with a range of cellular functions and processes, including numerous transcription factors and oncogenes, and are linked to several cancers. Our work highlights the direct and indirect role of DHX36 in the transcriptome of T-lymphocyte leukemia cells and the potential for DHX36 dysregulation in cancer.

1. Introduction

G-quadruplexes (G4s) are four-stranded nucleic acid secondary structures composed of four guanine bases oriented in stacks of planar arrays that form under physiological conditions [1,2]. Each participating guanine interacts with two neighboring guanines via Hoogsteen base pairing to stabilize the G4 structure. G4s are further stabilized by a monovalent cation located in a central G4 cavity [3]. G4s are highly stable structures and have been confirmed to form in vitro and in vivo upon the disruption of Watson–Crick base pairings, a necessary occurrence during transcription and replication [1,2,3,4].
Guanine-rich DNA with G4-forming potential, in particular sequences with at least three consecutive guanine nucleotides followed by short runs of unconstrained nucleotide composition, is widespread in bacterial and eukaryotic genomes [5]. There is a substantial number of possible alternative G4 structural configurations. For example, G4s can vary in loop length, location, and compositional origin. The DNA quadruplexes can have parallel or antiparallel topology in either strand of DNA, in an inter- or intra-molecular fashion, or between bases in syn- or anti- conformations [1,2,3,6,7]. G4s function in a diverse array of cellular processes, including telomere maintenance, DNA replication, recombination, transcription, and translation [2,8]. Their stable formation, coupled with their resolvability, suggests that cellular mechanisms now known to include helicase enzymes exist to regulate the formation and dissolution of G4s [9,10,11].

1.1. Evidence for a Role of G4s in Gene Regulation

There are several lines of evidence indicating the involvement of G4s in transcriptional processes and gene regulatory control. The locations of G4s appear to coincide with regions of nucleosome depletion [12,13], a feature typically associated with areas of active transcription, including promoters and enhancers. Moreover, genes in which G4 enrichment/nucleosome depletion are observed are associated with distinct cellular functions, including apoptosis and cell signaling, as well as tissue specificity [14,15]. G4s are also significantly enriched in transcription-factor binding sites and the promoter region of genes, especially in proto-oncogene promoters. More than 40% of all documented genes in the human genome contain at least one promoter G4-forming sequence, suggesting that G4 modulation may provide a significant axis of transcriptional regulation [6,16,17,18,19,20,21].

1.2. Potential Effects of G4s on Transcription

The mode of regulatory control exerted by G4s varies depending on the proposed direct, indirect, or epigenetic mechanisms [1,4,22]. The binding of transcription factors or other proteins to regulatory sequence G4s enhances the transcription of certain genes. Moreover, the binding and stabilization of G4s on the non-template strand prevents the reannealing of DNA or interferes with suppressor binding, which facilitate the initiation of repeated transcription [1,23,24]. Alternatively, the formation of G4s in regulatory regions could have negative effects on target gene transcription. For example, G4s may block the basic transcription machinery, or provide a binding site for suppressive transcription factors at a silencer element [6,23]. Furthermore, G4s may indirectly alter gene expression by acting as a competitive binding site to sequester transcription factors from other consensus sequences [25], or by binding chromatin modifying enzymes to facilitate epigenetic regulation [23]. Together, these observations exemplify the varied and context-dependent regulatory roles G4s can play in either positively or negatively regulating transcription.

1.3. G4 Binding and Stabilization by Ligands Can Alter Gene Expression

The treatment of yeast cells with the G4-stabilizer, N-methyl mesoporphyrin IX, leads to the upregulation of genes with potential G4-forming sequences in their promoter regions, supporting a role for G4s in the facilitation of gene expression [19]. Furthermore, multiple studies in human cells show that genes with promoters containing G4 sequences are dysregulated after G4 stabilization with ligands. Expression of the c-myc gene, which contains G4s in the promoter region, was downregulated by the ligand stabilization of G4, indicating the action of G4s in transcriptional repression [26,27], while G4 stabilization by pyridostatin significantly altered the expression of eight genes with proximate G4s relative to the control [28], with both up- and downregulation being observed.
Moreover, the genome-wide targeting of G4s in a human cell line with a G4-selective antibody showed significant G4-mediated alterations in expression, with 1767 protein-coding genes that were differentially expressed relative to control lines in a microarray analysis, with both up- and downregulation again apparent, and with affected gene promoters having higher numbers of G4-forming sequences, particularly upregulated genes [22].

1.4. Helicases Modulate G4 Effects on Expression

The unwinding of G4 structures in cells is mediated by helicase enzymes that have a potentially important role in G4-mediated gene regulation. This role for G4 helicases is supported by studies in mutant cells that lack the activity of a specific helicase. For example, alterations in gene expression were observed by microarray analysis in yeast mutants lacking activity of the RecQ family helicase Sgs1p, and coding regions with associated G4 motifs were more likely to be downregulated [19,29]. Fibroblasts from Bloom and Werner syndrome patients, lacking activity for RecQ helicases BLM or WRN, showed broad gene expression differences in comparison to control human fibroblasts, as assessed by microarray [30,31].
DHX36 (aliases: G4R1 and RHAU) is a DEAH-box helicase produced by the human DHX36 gene, which has the highest known affinity for both RNA and DNA G4s [7,10,32,33,34]. Given that this enzyme constitutes a critical G4-resolving factor in human cells, its effects on gene expression may be widespread, leading to the significant upregulation or downregulation of target genes in cells lacking DHX36 activity. Consistent with this prediction, DHX36 knockdown in human cell lines reduced the reporter and endogenous expression of the G4-containing gene, YY1 [35]. Furthermore, knockout of DHX36 led to a significant change in the abundance of mRNAs, which was dependent on the number of DHX36 binding sites on the mRNA, especially in the 5′ UTR [36]. Another study found that DHX36 associates with polysomes and the reduction in DHX36 increases the translation of repressive G4-containing upstream open reading frames (uORFs) while reducing protein production from the associated coding regions, especially impacting proto-oncogenes, transcription factors, and other epigenetic regulators [37].
Here, we describe the role of the DHX36 in modulating transcription in an immortalized human T lymphocyte Jurkat cancer cell line using RNA sequencing (RNA-seq) by exploring the functional consequences of gene dysregulation related to the loss of DHX36 with bioinformatic analyses. We further explored the association between gene expression and the presence and number of G4s in promoters and gene regions. Our results indicate that the depletion of DHX36 leads to widespread gene expression changes associated with the G4 content of differentially expressed genes.

2. Results

2.1. Data Features

A total of 160,544,634 paired-end reads remained after filtering, with paired-end read counts per replicate library ranging from 24,430,301 to 28,216,080. Mapping rates to the human reference genome were uniformly high (Table 1), and most reads were mapped to unique locations, with 84% of reads mapping to annotated exons, 11.6% to introns, and 4.4% to intergenic regions.

2.2. Differential Expression

A total of 7410 annotated genomic features showed significant differential expression between DHX36 knockout and control cell lines, with 4026 being upregulated and 3384 downregulated in knockout cell lines. Among the total features that were differentially expressed, 6705 were annotated as protein-coding genes, and 169 as long non-coding RNAs (Supplementary Table S1). Although significant, changes in expression were generally subtle, as only 455 genomic features (of which 384 were protein-coding and 29 were long non-coding RNAs) showed an upregulation of 2-fold or greater (Figure 1) in DHX36 knockout cell lines, and 675 showed a downregulation of 2-fold or greater (of which 590 were protein-coding and 23 were long non-coding RNAs). The number of genes at higher-level fold changes was limited: 67 were upregulated and 85 downregulated at a fold-change of 5, and 16 were upregulated and 19 downregulated at a fold-change of 10 (Supplementary Table S1).
A total of 269 differentially expressed genes encoded transcription factors (Supplementary Table S2), of which 173 were upregulated (17 at fold-change of 2 or greater) and 96 downregulated (22 at fold-change of at least -2) in KO cell lines. A total of 362 differentially expressed genes were categorized as oncogenes in the COSMIC database, of which 251 were upregulated (24 at fold-change of 2 or greater) and 111 downregulated (30 at fold-change of at least -2) in KO cells (Supplementary Table S2). Eleven known G4 helicases other than DHX36 [38] were found among the set of significantly upregulated genes, including DDX1, DDX5, DDX21, DDX24, DDX42, BLM, DHX9, DHX11, DNA2, MOV1, and WRN (Supplementary Table S1).

2.3. Functional Genomics

A total of 1280 gene ontology (GO) terms were significantly enriched in knockout line upregulated genes, of which 916 belonged to the biological process category, 157 to the molecular function category, and 207 to the cellular component category (Supplementary Table S3). The most enriched biological process terms related to the processing and transport of RNA, and the most enriched molecular function terms related to nucleic acid and protein binding and modification, helicase activity, and transcriptional coactivation, while enriched cellular component terms included a number related to chromatin and chromosomal structures (Figure 2). In contrast, only 63 GO terms were significantly enriched in DHX36 knockout-cell-line-downregulated genes, of which 18 belonged to the biological process category and 45 to the cellular component category (Supplementary Table S3), including a number of terms related to cell–cell junctions and vesicle organization and transport. A total of 34 significantly enriched KEGG categories were found for genes that were upregulated in knockout cell lines (Figure 3; Supplementary Table S4), including those related to cell signaling, RNA processing, DNA repair, and several cancers. Only four significant KEGG categories were found for KO-downregulated genes, including lysosome, arrhythmogenic right ventricular cardiomyopathy, axon guidance, and cell adhesion pathways.

2.4. Disease Associations of Differentially Expressed Genes

While there were no enriched disease ontology terms for knockout-line-downregulated genes, there were nine for knockout-line-upregulated genes, of which seven were related to ocular or retinal cancers, breast and ovarian cancer, or leukemia (Table 2, Figure 4). The other two significant disease ontology terms were for autosomal dominant disorders in general and the neurodegenerative disorder ataxia telangiectasia.

2.5. Differential Expression on DHX36 Knockout Is Associated with G-Quadruplex Content in Promoters

We defined 48,149 promoters as regions directly upstream of the annotated transcription start site for each protein-coding or non-coding gene. A total of 409,365 G4s were consistently identified by G4-Seq in both K+ ion and pyridostatin stabilization experiments [39], promoters from a total of 11,358 genes (23.6%) contained at least one observed G4 (OQ), and a total of 17,565 promoter-located OQs were identified for an average of 1.55 OQs per OQ-containing promoter.
From the 7410 differentially expressed genes, 2682 showed at least one OQ present in the defined promoter region (36.2% of DE gene promoters), with a total of 4420 OQs within the 2682 promoters, for an average of 1.65 OQs per promoter. Promoters of differentially expressed genes were significantly more likely to harbor an OQ (hypergeometric test, p = 2.71 × 10−157) than the genomic background (Figure 5), and had significantly more OQs per promoter (mean = 1.65) than the 8676 non-differentially expressed genes (Figure 6), with at least one promoter OQ (mean = 1.51) being obtained by Wilcoxon Rank-Sum test (W = 12,577,116, p = 5.74 × 10−14). As only 1130 differentially expressed genes were found with a fold-change of two or greater (2FC) between treatment and control, we found that 433 promoters (38.3%) harbored at least one OQ, and hence 2FC DE gene promoters were significantly more likely to harbor an OQ (hypergeometric test, p = 3.63 × 10−29) than the genomic background, and contained a total of 727 OQs for an average of 1.68 OQs per promoter, significantly more (W = 2,585,740, p = 9.70 × 10−5) than the 10,925 promoters for genes that were not considered to be differentially expressed (1.54 OQs per promoter on average) under this criterion. These data suggest that DHX36 knockout preferentially impacts the RNA abundance of genes with G4-rich promoters.

2.6. Differential Expression on DHX36 Knockout Is Associated with G-Quadruplex Content in Gene Regions

OQs were found within the gene intervals of 27,091 (56.3%) of 48,149 coding and non-coding genes, for a total of 280,899 gene-located OQs, an average of 10.37 OQs per OQ-containing gene. For DE genes between DHX36 knockout and control cells, 6779 of 7410 had at least one OQ within the gene (91.5%) for a total of 96,482 OQs and an average of 14.2 OQs per OQ-containing gene. Differentially expressed genes were significantly more likely to contain an OQ (hypergeometric test, p = 1.72 × 10−1131) than the genomic background, (Figure 5) and had significantly more OQs per gene (mean = 14.2) than the 20,312 non-differentially expressed genes. Figure 6 shows that at least one OQ (mean = 9.08) was obtained via the Wilcoxon Rank-Sum test (W = 86,042,226, p < 2.2 × 10−16), although DE genes with OQs (mean length = 81,209 bp) were also significantly longer than non-DE genes (mean length = 70,181) with OQs (W = 249,715,340, p < 2.2 × 10−16). These data suggest that DHX36 knockout preferentially impacts RNA abundance for genes with G4-rich gene regions.

3. Discussion

The association of G4s with gene regulatory regions, particularly promoters, indicates a role for G4s in transcriptional control [16,17,18,19]. The stabilization of G4s has been shown, in some cases, to have negative effects on the level of transcription [26,27], suggesting that their unwinding helicases play a role in gene expression, as the formation and processing of G4 DNA structures may serve as a topological gene-regulatory mechanism. Consistent with this, our results show that the loss of a G4-helicase, DHX36, in Jurkat cells has widespread effects on gene expression, as also seen in other studies [36], but with thousands of genes showing increased or decreased expression in KO cells. Our results indicated that the effect sizes were generally small, as the majority of genes showed a fold-change of less than two, and few genes had fold-changes that exceeded two. Given the upregulation of 11 other G4 helicases (Supplementary Table S1: in bold) known from the human genome [38], it is possible that compensatory effects are acting to attenuate large changes in the expression of G4-regulated genes under DHX36 knockout.
While the effects are generally subtle, DHX36 knockout may impact gene expression through multiple mechanisms. Previous studies showed that the presence of functionally relevant G4s [27], or an enrichment of G4-forming motifs or the observed G4s in promoter regions, may be associated with impacts on transcription levels. However, based on OQ locations discerned from G4-Seq data, the promoters in our DE gene set were enriched for the presence of at least one OQ and had a few but significantly more OQs in their promoters than genes that did not change their expression in KO cells relative to the wild-type. However, most of the dysregulated genes lacked G4s in their promoter regions. Although gene-level expression analysis, as performed here, is more robust and less likely to lead to false positives [40], genes may contain alternative transcripts with distinct promoters, which are not considered in this analysis. Enhanced replication could make transcript-level DE analysis feasible and reveal additional promoter OQs linked to changes in gene expression. It is also possible that some of these genes were indirectly affected by being downstream of a transcription factor whose dosage was affected by DHX36 knockout, as hundreds of transcription factors are among the differentially expressed genes, potentially leading to a cascade of effects on gene expression.
Alternatively, co-transcriptional or post-transcriptional mechanisms might be involved in the regulation of gene expression modulated by G4s and the DHX36 helicase. Previous studies found the highest number of observed quadruplexes in splice sites and 5′ UTRs, where they might have a post-transcriptional regulatory influence [39]. In one study, DHX36 knockout enhanced mRNA abundance via post-transcriptional mechanisms only, and by enhancing mRNA stability via G4s in UTRs and in the coding sequences in HEK293 cells [36]. Another study in HeLa cells found DHX36 to be associated with polysomes (i.e., actively translated mRNAs) and that DHX36 depletion increased the translation of G4-containing repressive upstream ORFs, thereby reducing coding region translation [37]. Differences in these studies are likely reflective of the different adherent cell lines utilized (HEK293 vs. Hela). Indeed, we did find a marked excess of differentially expressed genes with OQs within the gene region itself, and DE genes also had significantly more OQs than non-differentially expressed genes under helicase knockout in a non-adherent Jurkat cell line. Both factors may be influenced by the fact that genes that were differentially expressed due to loss of DHX36 activity were significantly longer than those that did not change expression.
While we did not collect data on RNA G4 prevalence in KO cells, Guo and Bartel [41] found no increase in RNA G4 formation after DHX36 deletion in mouse embryonic fibroblasts. There are many alternative G4 helicases in the human genome, including eight other DEAH box helicases [11,38], and eleven G4 helicases were upregulated in DHX36 KO cells, so it is not entirely clear whether DHX36 gene expression influences are modulated by the enhanced stability of RNA G4 structures. However, the association of RNA G4s with changes in gene expression following the knockout of an important helicase in human cells suggests a change in G4 state [36,37].
From a functional perspective, genes downregulated by DHX36 knockout seem to lack the functional coherence of upregulated genes, as few GO, DO, or KEGG terms were significantly enriched in the downregulated set. In contrast, the upregulated gene set was significantly enriched for many GO, KEGG, and DO terms, including several cancers. These cancer associations were for genes that were upregulated in DHX36 KO cells, which included many oncogenes in which G4 stability may be enhanced, indicating that DHX36 activity could typically act to suppress their expression under normal conditions. This contrasts with some single oncogene studies [26,27] in which the stabilization of G4 by other means led to oncogene downregulation. Therefore, the gene regulatory mechanisms of DHX36 knockout are areas for future elucidation, as is determining whether other aspects of cellular function are also affected.

4. Materials and Methods

The creation of DHX36 knockout Jurkat cell lines was achieved with CRISPR/Cas9-mediated gene editing and validated with PCR, qPCR, and Western blot, as previously described [42]. Cells were cultured in 10 cm cell plates in DMEM supplemented with 10% (v/v) bovine growth serum and 1% (v/v) antibacterial–antimycotic. For each replicate, RNA was extracted from a separate plate of cells. The extraction of RNA was performed, and libraries for Illumina sequencing were produced with the NEB Next Ultra RNA Library Prep Kit (New England Biolabs, Ipswich, MA, USA) according to the manufacturer’s instructions, for three replicates from knockout and control cell lines. Paired-end reads of 150 bp length were generated from each library by Novogene Corporation Inc. (Sacramento, CA, USA). Raw reads were filtered to remove adapter sequence, reads with more than 10% ambiguous base calls, or reads in which the quality score was less than 5 for more than 50% of the bases in the read.
STAR 2.5 [43] was used to align cleaned reads to the human GRCh37 reference genome sequence, and HT-Seq 0.6.1 [44] was used to estimate read counts for annotated genes. A differential expression analysis of read counts was performed with DESeq2 [45] with the significance criterion represented by an adjusted p-value of <0.05. Transcription factors among differentially expressed genes were identified by comparison with all annotated entries in the TFCat database of human and mouse transcription factors [46], while oncogenes were identified using the COSMIC database [47]. Enrichment for gene and disease ontology terms, and for the Kyoto Encyclopedia of Genes and Genomes (KEGG) categories corresponding to pathways, processes, or phenotypes, were performed using ClusterProfiler with a p-value adjusted by Benjamini–Hochberg correction for multiple tests of <0.05 [48].
We defined promoters as regions 1000 bp directly upstream of the transcription start site (one per gene) of protein-coding (21,514) and non-coding (25,117) genes on autosomal (1–22) or sex chromosomes (X, Y) in the GRCh37 reference genome annotation. G4 locations were obtained as the set common to G4-Seq experiments, which stabilized and identified observed G4s (OQs) across the genome with both K+ and pyridostatin [39], and OQs localized to promoters or genes were identified by comparing OQ coordinates to promoter or gene coordinates. Only OQs whose coordinates were entirely within the defined promoter or gene region were considered.
The enrichment of promoter OQs in differentially expressed genes relative to the genomic background (all coding and non-coding genes in the Ensembl annotation) was tested using a hypergeometric test. The average number of OQs per promoter for differentially expressed, and non-differentially expressed genes were compared by Wilcoxon Rank-Sum test given the skewed distribution of the number of OQs per promoter.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25031753/s1.

Author Contributions

Conceptualization, P.J.S., R.A.H., M.A.S., J.P.V., Y.-H.W., T.J.K., S.N.S., A.E.C., A.E.R. and J.M.G.; methodology, P.J.S., R.A.H., A.E.C., A.E.R., M.H.S. and J.M.G.; software, R.A.H., M.H.S. and J.M.G.; validation, P.J.S., R.A.H., Y.-H.W. and A.R.B.; formal analysis, P.J.S., R.A.H., A.E.R., M.H.S. and J.M.G.; investigation, P.J.S., R.A.H., E.M.R., D.C.S., M.E.S., Q.T.A., H.N.A., T.L.O., T.J.K., M.A.R., S.N.S., A.E.C., A.E.R., M.H.S. and J.M.G.; resources, P.J.S., R.A.H. and J.P.V.; data curation, R.A.H. and J.M.G.; writing—original draft preparation, P.J.S., R.A.H. and J.M.G.; writing—review and editing, P.J.S., R.A.H., J.P.V., M.A.S., Y.-H.W., A.R.B., E.M.R., D.C.S., M.E.S., Q.T.A., H.N.A., T.L.O., T.J.K., M.A.R., S.N.S., A.E.C., A.E.R. and J.M.G.; visualization, P.J.S., R.A.H., A.E.R. and J.M.G.; supervision, P.J.S., R.A.H. and Y.-H.W.; project administration, P.J.S., R.A.H. and Y.-H.W.; funding acquisition, P.J.S., R.A.H., M.A.S., J.P.V., Y.-H.W. and A.E.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by NIH-NCI (R15CA252996-01) to P.J.S. and R.A.H., and NIH-NIA (1R15AG067291-01) to P.J.S., NIH-NIGMS (RO1GM101192) to Y.-H.W a Ball State Student ASPIRE grant to A.E.R., The Miami Foundation for Cancer Research grants to P.J.S. and Ball State University start-up funds to P.J.S. and R.A.H.

Informed Consent Statement

Not applicable.

Data Availability Statement

Raw sequencing reads are available at NCBI’s Short Read Archive under accession number SRRXXXXXX.

Acknowledgments

We thank the Ball State University Department of Biology for supportive research environment and the generous sharing of cell culture facilities.

Conflicts of Interest

Author James P. Vaughn was employed by the company NanoMedica LLC. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Eddy, J.; Maizels, N. Gene function correlates with potential for G4 DNA formation in the human genome. Nucleic Acids Res. 2006, 34, 3887–3896. [Google Scholar] [CrossRef]
  2. Maizels, N.; Gray, L.T. The G4 genome. PLoS Genet. 2013, 9, e1003468. [Google Scholar] [CrossRef]
  3. Mendoza, O.; Bourdoncle, A.; Boule, J.B.; Brosh, R.M., Jr.; Mergny, J.L. G-quadruplexes and helicases. Nucleic Acids Res. 2016, 44, 1989–2006. [Google Scholar] [CrossRef]
  4. Maizels, N. G4 motifs in human genes. Ann. N. Y. Acad. Sci. 2012, 1267, 53–60. [Google Scholar] [CrossRef] [PubMed]
  5. Yadav, P.; Kim, N.; Kumari, M.; Verma, S.; Sharma, T.K.; Yadav, V.; Kumar, A. G-Quadruplex Structures in Bacteria: Biological Relevance and Potential as an Antimicrobial Target. J. Bacteriol. 2021, 203. [Google Scholar] [CrossRef]
  6. Balasubramanian, S.; Hurley, L.H.; Neidle, S. Targeting G-quadruplexes in gene promoters: A novel anticancer strategy? Nat. Rev. Drug Discov. 2011, 10, 261–275. [Google Scholar] [CrossRef]
  7. Smaldino, P.J.; Routh, E.D.; Kim, J.H.; Giri, B.; Creacy, S.D.; Hantgan, R.R.; Akman, S.A.; Vaughn, J.P. Mutational Dissection of Telomeric DNA Binding Requirements of G4 Resolvase 1 Shows that G4-Structure and Certain 3′-Tail Sequences Are Sufficient for Tight and Complete Binding. PLoS ONE 2015, 10, e0132668. [Google Scholar] [CrossRef]
  8. Bochman, M.L.; Paeschke, K.; Zakian, V.A. DNA secondary structures: Stability and function of G-quadruplex structures. Nat. Rev. Genet. 2012, 13, 770–780. [Google Scholar] [CrossRef] [PubMed]
  9. Harrington, C.; Lan, Y.; Akman, S.A. The identification and characterization of a G4-DNA resolvase activity. J. Biol. Chem. 1997, 272, 24631–24636. [Google Scholar] [CrossRef]
  10. Vaughn, J.P.; Creacy, S.D.; Routh, E.D.; Joyner-Butt, C.; Jenkins, G.S.; Pauli, S.; Nagamine, Y.; Akman, S.A. The DEXH protein product of the DHX36 gene is the major source of tetramolecular quadruplex G4-DNA resolving activity in HeLa cell lysates. J. Biol. Chem. 2005, 280, 38117–38120. [Google Scholar] [CrossRef] [PubMed]
  11. Sauer, M.; Paeschke, K. G-quadruplex unwinding helicases and their function in vivo. Biochem. Soc. Trans. 2017, 45, 1173–1182. [Google Scholar] [CrossRef]
  12. Halder, K.; Halder, R.; Chowdhury, S. Genome-wide analysis predicts DNA structural motifs as nucleosome exclusion signals. Mol. Biosyst. 2009, 5, 1703–1712. [Google Scholar] [CrossRef]
  13. Wong, H.M.; Huppert, J.L. Stable G-quadruplexes are found outside nucleosome-bound regions. Mol. Biosyst. 2009, 5, 1713–1719. [Google Scholar] [CrossRef]
  14. Andersson, R.; Sandelin, A. Determinants of enhancer and promoter activities of regulatory elements. Nat. Rev. Genet. 2020, 21, 71–87. [Google Scholar] [CrossRef]
  15. Zheng, K.W.; Zhang, J.Y.; He, Y.D.; Gong, J.Y.; Wen, C.J.; Chen, J.N.; Hao, Y.H.; Zhao, Y.; Tan, Z. Detection of genomic G-quadruplexes in living cells using a small artificial protein. Nucleic Acids Res. 2020, 48, 11706–11720. [Google Scholar] [CrossRef]
  16. Huppert, J.L.; Balasubramanian, S. Prevalence of quadruplexes in the human genome. Nucleic Acids Res. 2005, 33, 2908–2916. [Google Scholar] [CrossRef]
  17. Huppert, J.L.; Balasubramanian, S. G-quadruplexes in promoters throughout the human genome. Nucleic Acids Res. 2007, 35, 406–413. [Google Scholar] [CrossRef]
  18. Rawal, P.; Kummarasetti, V.B.; Ravindran, J.; Kumar, N.; Halder, K.; Sharma, R.; Mukerji, M.; Das, S.K.; Chowdhury, S. Genome-wide prediction of G4 DNA as regulatory motifs: Role in Escherichia coli global regulation. Genome Res. 2006, 16, 644–655. [Google Scholar] [CrossRef] [PubMed]
  19. Hershman, S.G.; Chen, Q.; Lee, J.Y.; Kozak, M.L.; Yue, P.; Wang, L.S.; Johnson, F.B. Genomic distribution and functional analyses of potential G-quadruplex-forming sequences in Saccharomyces cerevisiae. Nucleic Acids Res. 2008, 36, 144–156. [Google Scholar] [CrossRef] [PubMed]
  20. Todd, A.K.; Neidle, S. The relationship of potential G-quadruplex sequences in cis-upstream regions of the human genome to SP1-binding elements. Nucleic Acids Res. 2008, 36, 2700–2704. [Google Scholar] [CrossRef] [PubMed]
  21. Du, X.; Gertz, E.M.; Wojtowicz, D.; Zhabinskaya, D.; Levens, D.; Benham, C.J.; Schäffer, A.A.; Przytycka, T.M. Potential non-B DNA regions in the human genome are associated with higher rates of nucleotide mutation and expression variation. Nucleic Acids Res. 2014, 42, 12367–12379. [Google Scholar] [CrossRef] [PubMed]
  22. Fernando, H.; Sewitz, S.; Darot, J.; Tavaré, S.; Huppert, J.L.; Balasubramanian, S. Genome-wide analysis of a G-quadruplex-specific single-chain antibody that regulates gene expression. Nucleic Acids Res. 2009, 37, 6716–6722. [Google Scholar] [CrossRef] [PubMed]
  23. Varshney, D.; Spiegel, J.; Zyner, K.; Tannahill, D.; Balasubramanian, S. The regulation and functions of DNA and RNA G-quadruplexes. Nat. Rev. Mol. Cell Biol. 2020, 21, 459–474. [Google Scholar] [CrossRef] [PubMed]
  24. Shu, H.; Zhang, R.; Xiao, K.; Yang, J.; Sun, X. G-Quadruplex-Binding Proteins: Promising Targets for Drug Design. Biomolecules 2022, 12, 648. [Google Scholar] [CrossRef] [PubMed]
  25. Spiegel, J.; Cuesta, S.M.; Adhikari, S.; Hänsel-Hertsch, R.; Tannahill, D.; Balasubramanian, S. G-quadruplexes are transcription factor binding hubs in human chromatin. Genome Biol. 2021, 22, 117. [Google Scholar] [CrossRef]
  26. Hurley, L.H.; Von Hoff, D.D.; Siddiqui-Jain, A.; Yang, D. Drug targeting of the c-MYC promoter to repress gene expression via a G-quadruplex silencer element. Semin. Oncol. 2006, 33, 498–512. [Google Scholar] [CrossRef]
  27. Siddiqui-Jain, A.; Grand, C.L.; Bearss, D.J.; Hurley, L.H. Direct evidence for a G-quadruplex in a promoter region and its targeting with a small molecule to repress c-MYC transcription. Proc. Natl. Acad. Sci. USA 2002, 99, 11593–11598. [Google Scholar] [CrossRef]
  28. Lam, E.Y.; Beraldi, D.; Tannahill, D.; Balasubramanian, S. G-quadruplex structures are stable and detectable in human genomic DNA. Nat. Commun. 2013, 4, 1796. [Google Scholar] [CrossRef]
  29. Fry, R.C.; Sambandan, T.G.; Rha, C. DNA damage and stress transcripts in Saccharomyces cerevisiae mutant sgs1. Mech. Ageing Dev. 2003, 124, 839–846. [Google Scholar] [CrossRef] [PubMed]
  30. Nguyen, G.H.; Tang, W.; Robles, A.I.; Beyer, R.P.; Gray, L.T.; Welsh, J.A.; Schetter, A.J.; Kumamoto, K.; Wang, X.W.; Hickson, I.D.; et al. Regulation of gene expression by the BLM helicase correlates with the presence of G-quadruplex DNA motifs. Proc. Natl. Acad. Sci. USA 2014, 111, 9905–9910. [Google Scholar] [CrossRef] [PubMed]
  31. Cheung, H.H.; Liu, X.; Canterel-Thouennon, L.; Li, L.; Edmonson, C.; Rennert, O.M. Telomerase protects werner syndrome lineage-specific stem cells from premature aging. Stem Cell Rep. 2014, 2, 534–546. [Google Scholar] [CrossRef] [PubMed]
  32. Creacy, S.D.; Routh, E.D.; Iwamoto, F.; Nagamine, Y.; Akman, S.A.; Vaughn, J.P. G4 resolvase 1 binds both DNA and RNA tetramolecular quadruplex with high affinity and is the major source of tetramolecular quadruplex G4-DNA and G4-RNA resolving activity in HeLa cell lysates. J. Biol. Chem. 2008, 283, 34626–34634. [Google Scholar] [CrossRef]
  33. Giri, B.; Smaldino, P.J.; Thys, R.G.; Creacy, S.D.; Routh, E.D.; Hantgan, R.R.; Lattmann, S.; Nagamine, Y.; Akman, S.A.; Vaughn, J.P. G4 resolvase 1 tightly binds and unwinds unimolecular G4-DNA. Nucleic Acids Res. 2011, 39, 7161–7178. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, M.C.; Tippana, R.; Demeshkina, N.A.; Murat, P.; Balasubramanian, S.; Myong, S.; Ferre-D’Amare, A.R. Structural basis of G-quadruplex unfolding by the DEAH/RHA helicase DHX36. Nature 2018, 558, 465–469. [Google Scholar] [CrossRef] [PubMed]
  35. Huang, W.; Smaldino, P.J.; Zhang, Q.; Miller, L.D.; Cao, P.; Stadelman, K.; Wan, M.; Giri, B.; Lei, M.; Nagamine, Y.; et al. Yin Yang 1 contains G-quadruplex structures in its promoter and 5′-UTR and its expression is modulated by G4 resolvase 1. Nucleic Acids Res. 2012, 40, 1033–1049. [Google Scholar] [CrossRef] [PubMed]
  36. Sauer, M.; Juranek, S.A.; Marks, J.; De Magis, A.; Kazemier, H.G.; Hilbig, D.; Benhalevy, D.; Wang, X.; Hafner, M.; Paeschke, K. DHX36 prevents the accumulation of translationally inactive mRNAs with G4-structures in untranslated regions. Nat. Commun. 2019, 10, 2421. [Google Scholar] [CrossRef]
  37. Murat, P.; Marsico, G.; Herdy, B.; Ghanbarian, A.T.; Portella, G.; Balasubramanian, S. RNA G-quadruplexes at upstream open reading frames cause DHX36- and DHX9-dependent translation of human mRNAs. Genome Biol. 2018, 19, 229. [Google Scholar] [CrossRef]
  38. Liu, Y.; Zhu, X.; Wang, K.; Zhang, B.; Qiu, S. Corrigendum: The Cellular Functions and Molecular Mechanisms of G-Quadruplex Unwinding Helicases in Humans. Front. Mol. Biosci. 2022, 8, 833430. [Google Scholar] [CrossRef]
  39. Chambers, V.S.; Marsico, G.; Boutell, J.M.; Di Antonio, M.; Smith, G.P.; Balasubramanian, S. High-throughput sequencing of DNA G-quadruplex structures in the human genome. Nat. Biotechnol. 2015, 33, 877–881. [Google Scholar] [CrossRef]
  40. Soneson, C.; Love, M.I.; Robinson, M.D. Differential analyses for RNA-seq: Transcript-level estimates improve gene-level inferences. F1000Res 2015, 4, 1521. [Google Scholar] [CrossRef]
  41. Guo, J.U.; Bartel, D.P. RNA G-quadruplexes are globally unfolded in eukaryotic cells and depleted in bacteria. Science 2016, 353, aaf5371. [Google Scholar] [CrossRef]
  42. Tseng, Y.J.; Sandwith, S.N.; Green, K.M.; Chambers, A.E.; Krans, A.; Raimer, H.M.; Sharlow, M.E.; Reisinger, M.A.; Richardson, A.E.; Routh, E.D.; et al. The RNA helicase DHX36-G4R1 modulates C9orf72 GGGGCC hexanucleotide repeat-associated translation. J. Biol. Chem. 2021, 297, 100914. [Google Scholar] [CrossRef]
  43. Dobin, A.; Davis, C.A.; Schlesinger, F.; Drenkow, J.; Zaleski, C.; Jha, S.; Batut, P.; Chaisson, M.; Gingeras, T.R. STAR: Ultrafast universal RNA-seq aligner. Bioinformatics 2013, 29, 15–21. [Google Scholar] [CrossRef]
  44. Anders, S.; Pyl, P.T.; Huber, W. HTSeq--a Python framework to work with high-throughput sequencing data. Bioinformatics 2015, 31, 166–169. [Google Scholar] [CrossRef] [PubMed]
  45. Anders, S.; Huber, W. Differential expression analysis for sequence count data. Genome Biol. 2010, 11, R106. [Google Scholar] [CrossRef] [PubMed]
  46. Fulton, D.L.; Sundararajan, S.; Badis, G.; Hughes, T.R.; Wasserman, W.W.; Roach, J.C.; Sladek, R. TFCat: The curated catalog of mouse and human transcription factors. Genome Biol. 2009, 10, R29. [Google Scholar] [CrossRef] [PubMed]
  47. Tate, J.G.; Bamford, S.; Jubb, H.C.; Sondka, Z.; Beare, D.M.; Bindal, N.; Boutselakis, H.; Cole, C.G.; Creatore, C.; Dawson, E.; et al. COSMIC: The Catalogue Of Somatic Mutations In Cancer. Nucleic Acids Res. 2019, 47, D941–D947. [Google Scholar] [CrossRef] [PubMed]
  48. Yu, G.; Wang, L.G.; Han, Y.; He, Q.Y. clusterProfiler: An R package for comparing biological themes among gene clusters. Omics 2012, 16, 284–287. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Distributions of effect sizes for genomic features that are upregulated (A) or downregulated (B) in DHX36 knockout cell lines relative to control.
Figure 1. Distributions of effect sizes for genomic features that are upregulated (A) or downregulated (B) in DHX36 knockout cell lines relative to control.
Ijms 25 01753 g001
Figure 2. Top 10 most significantly enriched (by adjusted p-value) GO terms among three categories for genes upregulated in DHX36 knockout cell lines. BP = biological process; MF = molecular function; CC = cellular component.
Figure 2. Top 10 most significantly enriched (by adjusted p-value) GO terms among three categories for genes upregulated in DHX36 knockout cell lines. BP = biological process; MF = molecular function; CC = cellular component.
Ijms 25 01753 g002
Figure 3. Significantly enriched KEGG categories for genes upregulated in knockout versus control cells. Gene Ratio = fraction of differentially expressed genes matching KEGG description. padj = Benjamini–Hochberg-adjusted p-value.
Figure 3. Significantly enriched KEGG categories for genes upregulated in knockout versus control cells. Gene Ratio = fraction of differentially expressed genes matching KEGG description. padj = Benjamini–Hochberg-adjusted p-value.
Ijms 25 01753 g003
Figure 4. Significantly enriched disease ontology (DO) categories for genes upregulated in knockout versus control cells. Gene Ratio = fraction of differentially expressed genes matching disease ontology description. padj = Benjamini–Hochberg-adjusted p-value.
Figure 4. Significantly enriched disease ontology (DO) categories for genes upregulated in knockout versus control cells. Gene Ratio = fraction of differentially expressed genes matching disease ontology description. padj = Benjamini–Hochberg-adjusted p-value.
Ijms 25 01753 g004
Figure 5. Differentially expressed promoters and genes following DHX36 KO are more likely to contain OQs compared to the genomic background. (A) Percentage of gene promoters with OQ. (B) Percentage of gene regions with OQs.
Figure 5. Differentially expressed promoters and genes following DHX36 KO are more likely to contain OQs compared to the genomic background. (A) Percentage of gene promoters with OQ. (B) Percentage of gene regions with OQs.
Ijms 25 01753 g005
Figure 6. Differentially expressed promoters and genes following DHX36 KO have significantly more OQs per region than non-differentially expressed genes. (A) Mean number of G4s per promoter for differentially and non-differentially expressed genes. (B) Mean number of G4s per gene, for differentially and non-differentially expressed genes.
Figure 6. Differentially expressed promoters and genes following DHX36 KO have significantly more OQs per region than non-differentially expressed genes. (A) Mean number of G4s per promoter for differentially and non-differentially expressed genes. (B) Mean number of G4s per gene, for differentially and non-differentially expressed genes.
Ijms 25 01753 g006
Table 1. For each replicate of control and treatment cell lines, the number of paired-end reads sequenced, the number passing quality control filters, the percent of bases with phred scores of at least 20, the percentage of reads that map to unique genomic locations, and the percentage of reads that map to multiple genomic locations are shown.
Table 1. For each replicate of control and treatment cell lines, the number of paired-end reads sequenced, the number passing quality control filters, the percent of bases with phred scores of at least 20, the percentage of reads that map to unique genomic locations, and the percentage of reads that map to multiple genomic locations are shown.
Sample NameTreatmentRaw ReadsFiltered Reads% Q20% Uniquely Mapped% Multi-Mapping
A11_1DHX36 KO25,210,87424,430,30198.1193.582.24
A11_2DHX36 KO28,718,45227,622,29998.0593.262.26
A11_3DHX36 KO28,926,80128,216,08098.1693.872.14
A28_1Wild-type27,734,05427,130,31598.1293.592.14
A28_2Wild-type26,521,78425,939,49698.1393.872.14
A28_3Wild-type28,081,20927,206,14398.2393.562.23
Table 2. Significantly enriched disease ontology (DO) terms for genes upregulated in DHX36 knockout versus control cell lines. Terms with an adjusted p-value less than 0.05 after Benjamini-Hochberg correction for multiple tests were considered significantly enriched.
Table 2. Significantly enriched disease ontology (DO) terms for genes upregulated in DHX36 knockout versus control cell lines. Terms with an adjusted p-value less than 0.05 after Benjamini-Hochberg correction for multiple tests were considered significantly enriched.
IDDescriptionGene RatioBackground RatioRaw p-ValueAdjusted p-Value
DOID:4645retinal cancer54/1667113/74052.29 × 10−96.79 × 10−7
DOID:5683hereditary breast ovarian cancer86/1667214/74053.04 × 10−96.79 × 10−7
DOID:768retinoblastoma53/1667111/74053.34 × 10−96.79 × 10−7
DOID:771retinal cell cancer53/1667111/74053.34 × 10−96.79 × 10−7
DOID:0060116sensory system cancer57/1667134/74051.49 × 10−72.01 × 10−5
DOID:2174ocular cancer57/1667134/74051.49 × 10−72.01 × 10−5
DOID:0050736autosomal dominant disease123/1667394/74052.56 × 10−50.00297116
DOID:12704ataxia telangiectasia19/166737/74050.00011420.01159172
DOID:1036ataxia telangiectasia69/1667209/74050.000267250.02411196
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

Gumina, J.M.; Richardson, A.E.; Shojiv, M.H.; Chambers, A.E.; Sandwith, S.N.; Reisinger, M.A.; Karns, T.J.; Osborne, T.L.; Alashi, H.N.; Anderson, Q.T.; et al. Differential Gene Expression following DHX36/G4R1 Knockout Is Associated with G-Quadruplex Content and Cancer. Int. J. Mol. Sci. 2024, 25, 1753. https://doi.org/10.3390/ijms25031753

AMA Style

Gumina JM, Richardson AE, Shojiv MH, Chambers AE, Sandwith SN, Reisinger MA, Karns TJ, Osborne TL, Alashi HN, Anderson QT, et al. Differential Gene Expression following DHX36/G4R1 Knockout Is Associated with G-Quadruplex Content and Cancer. International Journal of Molecular Sciences. 2024; 25(3):1753. https://doi.org/10.3390/ijms25031753

Chicago/Turabian Style

Gumina, Joseph M., Adam E. Richardson, Mahmudul Hasan Shojiv, Antonio E. Chambers, Siara N. Sandwith, Michael A. Reisinger, Taylor J. Karns, Tyler L. Osborne, Hasna N. Alashi, Quinn T. Anderson, and et al. 2024. "Differential Gene Expression following DHX36/G4R1 Knockout Is Associated with G-Quadruplex Content and Cancer" International Journal of Molecular Sciences 25, no. 3: 1753. https://doi.org/10.3390/ijms25031753

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

Article Metrics

Back to TopTop