Next Article in Journal
The “Bagno dell’Acqua” Lake as a Novel Mars-like Analogue: Prebiotic Syntheses of PNA and RNA Building Blocks and Oligomers
Previous Article in Journal
Characteristics and Expression Profiles of Identified WRKY Genes in Barley Landraces Under Cold Stress
Previous Article in Special Issue
Insights on SNPs of Human Activation-Induced Cytidine Deaminase AID
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Participants in Transcription–Replication Conflict and Their Role in Formation and Resolution of R-Loops

by
Anastasiia T. Davletgildeeva
1,* and
Nikita A. Kuznetsov
1,2
1
Institute of Chemical Biology and Fundamental Medicine, Siberian Branch of Russian Academy of Sciences, Novosibirsk 630090, Russia
2
Department of Natural Sciences, Novosibirsk State University, Novosibirsk 630090, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(14), 6951; https://doi.org/10.3390/ijms26146951
Submission received: 9 June 2025 / Revised: 16 July 2025 / Accepted: 16 July 2025 / Published: 19 July 2025

Abstract

The DNA of all living organisms is a common matrix for both replication and transcription processes. This sometimes leads to inevitable collisions between DNA replication and transcription machinery. There is plethora of evidence demonstrating that such collisions (or TRCs) are one of the most common and significant reasons for genomic instability. One of the key outcomes of TRCs is the accumulation of non-canonical DNA secondary structures, including R-loops. R-loops are three-stranded DNA–RNA hybrids with a displaced third single-stranded DNA fragment. Although R-loops are thought to play several functional roles in biological processes, an imbalance in their metabolism has been proven to have severe consequences. In this review, we attempt to summarize the current knowledge of the participants in the process of R-loop regulation in cells, with an emphasis on eukaryotic systems. We also touch upon the conditions favoring TRCs and the possible ways of dealing with these conflicts.

1. Introduction

The effective firing of replication origins and the ongoing progression of the replication fork play crucial roles in the genome duplication process, which is essential to every living cell. This process is important for normal cell cycle progression and to avoid deleterious DNA damage and cell death [1]. Nevertheless, the progression of the replication fork can be disrupted if it meets unrepaired DNA lesions, non-canonical (non-B) DNA secondary structures, or various nonhistone proteins that are bound to DNA [2]. If the replication fork stalls, replication checkpoint factors promote its stabilization to ensure the association of the replisome with the template and allow for the possibility of resuming the replication process, as its impairment is a major issue for dividing cells [3].
Considering the fact that DNA in a cell is not only a substrate for replication, but also for transcription, it is natural that the interference of these two processes, which are localized at the same genome region, is a prominent example of an endogenous threat to optimal replication progression [4,5]. A rather common and telling fact indicating that transcription is a source of replicative stress is that the active transcription process leads to an increase in spontaneous mitotic recombination. It is widely accepted that the passage of transcription leads to the most detrimental DNA lesions, double-strand breaks (DSBs), which are repaired via recombination [6]. Indeed, during transcription, the positive supercoiling of processed DNA occurs before this process; however, after the RNA polymerase (RNAP) has passed through, DNA with negative supercoiling remains [2,7]. These topological changes in DNA can accumulate in regions of highly transcribed genes, disrupting any further passage of the replication fork through these regions [2,8]. Topoisomerase I (Top1) is the main enzyme responsible for the resolution of negative supercoiled DNA associated with transcription [9]. In contrast, topoisomerase II (Top2) is needed to relax the positive supercoils that are concentrated before the elongating RNAP [10]. When the activities of these two enzymes are abolished, highly transcribed regions of DNA tend to accumulate torsional stress, which can subsequently lead to stretches of single-stranded DNA (ssDNA) behind the RNAP [11,12]. The resulting persistence of ssDNA regions may lead to a burst in DNA damage levels due to the higher susceptibility of such regions to genotoxic agents, or it could give rise to the formation of different non-B DNA structures [2].
DNA molecules are capable of adopting several various conformational forms besides the classic B structure, including hairpins, left-handed Z-DNA, G-quadruplexes, triplex DNA (or H-DNA), and RNA–DNA hybrids. Negative supercoiling and the unwinding of the DNA duplex that accompany replication or transcription processes ensure the formation of such structures along with a suitable DNA sequence [2]. A number of the negative effects of such non-B structures on the genome’s stability are due to their interference with replication fork passage [3,13]. Additionally, the occurrence of these structures within gene bodies can provoke RNAP stalling and the termination of transcription elongation [14]. RNAP stalling itself acts as a signal for transcription-coupled nucleotide excision repair (TC-NER) in cells. Thus, although nucleotide excision repair (NER) usually proceeds with low error levels, the enrollment of the ssDNA intermediate, which is sensitive to damaging agents, and gap-filling DNA synthesis could be a path for introducing new errors [2,15].
Transcription–replication collisions or conflicts (TRCs) are considered to be one of the major sources of the replication stress associated with transcriptional dysregulation [16]. The type, and consequently severity and harmfulness of TRCs depends on the direction in which the collision of transcription and replication processes occurs. When the replication fork moves in the same direction as the transcription machinery and they use the same DNA strand as a template, such a conflict is called co-directional TRC (CD-TRC, Figure 1), which is considered to occur at a lower cost to the cell and be more easily resolved [17]. Since the velocity of the passage of E. coli replication forks is 10 times higher than the transcription machinery, their co-directional collision is inevitable [18]. Nevertheless, bacterial replication forks were demonstrated in vitro to be able to pass RNAP without displacing it and a transcript during CD-TRCs [19,20]. When replication and transcription processes occur in opposite directions and collide with each other, the conflict is called a head-on TRC (HO-TRC, Figure 1) [16,21,22], and the detrimental consequences of this type of TRC have long been recognized [23]. It was demonstrated that DNA polymerase is unable to bypass RNAP and is forced to pause during HO-TRCs [23], while it is only slowed down during CD-TRCs [24]. However, RNAPs appear to be able to resume RNA synthesis in either type of TRC. Although it is often forced to dissemble during TRC resolution, replication restart proteins, which have been found to be essential in many procaryotes, rebuild the disrupted replisome at TRC sites [25]. In bacterial systems, the replisome has been shown to be able to displace RNAP from DNA in both CD-TRCs and HO-TRCs, although in the latter case, it requires the help of the elongation factor Mfd [26,27]. This translocase protein, which is involved in NER, recognizes stalled RNAPs and displaces them from DNA in an ATP-dependent manner [28,29].

2. Dealing with TRCs

The fact that transcription tends to be a constant threat to the normal progression of replication, especially when these two processes collide, suggests that there must be various factors and mechanisms that promote the progression of a stalled replication fork and maintain genomic stability in such cases [2].
One of the general characteristics that indicates a greater level of danger from HO-TRCs for the genome’s stability is the prevailing co-directional orientation of actively transcribed genes. Indeed, bacterial genomes were demonstrated to be mostly organized to orient vital and highly transcribed genes in the same direction as replication, although it necessarily favors evaluation during CD-TRC events [31]. Numerous studies have demonstrated the same overall orientation of the genome for eukaryotes as well [32,33,34,35,36]. Thus, it has been shown that the transcription start sites of many long and highly transcribed genes coincide with replication initiation sites, while the transcription end sites of these genes often overlap with their replication termination sites [34,37,38]. Among the most problematic for the eucaryotic replication of actively transcribed genes are ribosomal DNAs (rDNAs). About 150 rDNA repeats are located at the S. cerevisiae rDNA locus, each consisting of 35S (RNAP I–transcribed) and 5S (RNAP III–transcribed) pre-rRNA genes [39,40]. These loci are also equipped with special sites that act as barriers to the passage of the replication fork downstream of the 35S gene [41] due to the binding of particular proteins to them, such as Fob1 known to create a unidirectional block to replication forks [42]. This allows for the blocking of the active passage of the replication fork at certain sites by slowing it down, thus avoiding the occurrence of HO-TRC events [43].
At the same time, there might be a reason for the existence of head-on-oriented genes. It has been suggested that head-on genes must undergo elevated rates of TRC-induced mutagenesis, so it is possible that this orientation does not establish by chance and that selection is aimed at achieving high variability in these genes [44,45]. CD-TRCs are able to provoke the disruption of replication at certain sites and even cause DSB emergence [46,47]. In eukaryotes, it was demonstrated that HO-TRCs at RNAP II-transcribed genes can induce recombination [48]. Moreover, transcription-induced stress, especially in the case of Top1 dysfunction, can provoke the elevation and recombination of S-phase-dependent DNA breaks [49,50]. As long as the transcription of long genes can take several cell cycles, avoiding TRCs at such genes can be quite difficult, which is further complicated by the fact that some long genes may overlap with hotspots for chromosomal instability (or so-called common fragile sites), which replicate late in the S-phase [51,52].
Topological stress leads to HO-TRCs causing more harm to cells than CO-TRCs, meaning that topoisomerases play a key role in HO-TRC-induced stress prevention. It was demonstrated by Lang and Meriikh that the prevention of HO-TRCs specifically requires the relaxation of positive supercoiling by type II topoisomerases DNA gyrase and Topo IV in Bacillus subtilis. Moreover, it was shown that DNA gyrase and Topo IV preferentially associate with head-on-oriented genes. At the same time, it seems that the subsequential introduction of negative supercoiling by DNA gyrase is responsible for the formation of R-loops [53]. An increase in the rate of positive supercoiling of DNA resulting from HO-TRCs was found to stall the progress of transcription and replication machinery [5,53,54,55].
The results of several in vitro studies demonstrate that the blockade of the replisome resulting from HO-TRC events does not necessarily lead to its disruption, and replication can be resumed in the presence of certain helping factors, such as T4 Dda helicase and Mfd translocase [23,26]. To displace proteins interfering with the advancement of the replication fork, E. coli utilize accessory helicases, such as Rep, UvrD, and DinG [56], while for B. subtilis helicase, PcrA was demonstrated to assist replication forks with progressing through several genes placed in a head-on orientation to replication [57]. The recruitment of various accessory DNA helicases to resolve the progression of replication forks has also been demonstrated in eukaryotic cells. One of the most-studied examples is the helicase Rrm3 of S. cerevisiae, which is a part of the replisome complex and was demonstrated to facilitate the passage of the replication fork through the protein-bound regions of DNA [58,59].

3. TRC-Induced Non-B DNA Structures

3.1. G-Quadruplexes

G-quadruplexes are the result of the folding of the G-rich DNA motifs into tetraplex structures, with stacked groups of the four guanines being stabilized in one planar orientation via Hoogsteen hydrogen bond pairing [60] (Figure 1). The appearance of G-quadruplexes on the replication or transcription pathway can become a serious obstacle to the passage of the replication forks or RNAPs [61,62]. Indeed, it is known that the promoter regions and 5′-untranslated regions of highly transcribed DNA genes are enriched in sequences that are prone to forming G-quadruplex structures [63]. For example, eukaryotic rDNA consisting of a series of repeating pre-rRNA genes is predicted to have high potency in non-B structure formation, specifically G-quadruplexes [64,65]. Several pieces of evidence were found to indicate that specialized helicases resolving G-quadruplexes participate in the facilitation of replication and transcription [66,67,68]. Thus, the highly conserved fission yeast Pfh1 helicase, which is homologous to S. cerevisiae Rrm3 [69,70], was demonstrated to have a preference for binding rDNA and being able to unwind model rDNA G-quadruplex structures [71]. The ability to unwind G-quadruplex structures has been demonstrated for several human helicases as well, such as the Bloom syndrome protein (BLM), Werner syndrome helicase (WRN), FANCJ, RTEL1, and PIF1; however, most of these experiments were conducted in vitro, so the real role of these enzymes in facilitating replication passage through G-quadruplex obstacles remains to be unraveled [67]. For example, it was demonstrated by Isik et al. that FANCJ helicase participates in the restart of R-loops–stalled forks though its interaction with the mismatch repair (MMR) factors MutSβ and MLH1, components of the MutLβ complex and several other complexes participating in MMR [8].
In addition to the specialized helicases that are capable of resolving G-quadruplex structures along the replication fork passage, polymerases that carry out trans-lesion synthesis (TLS) are also capable of dealing with such obstacles in the cell [60]. DNA replication machinery can utilize DNA-damage-tolerant TLS to bypass some lesions, and thus avoid malicious replication stops. It has been demonstrated that this mechanism can be carried out in eukaryotic cells to pass through G-quadruplex regions during replication; in this case, TLS is achieved by the work of the polymerases Rev1 [72], η [73,74], κ [75], and θ [76].

3.2. R-Loops

The first demonstration of co-transcriptional R-loops being a threat to genomic stability through recombination intensification was the emergence of the hyper-recombination phenotype of the S. cerevisiae mutants of the THO/TREX, a conserved eukaryotic protein complex which plays important role in transcription and mRNA metabolism [77,78,79]. It was shown that this phenotype also depends on the capacity of the nascent RNA molecule to form an R-loop behind the elongating DNA polymerase [77].
R-loops are structures consisting of a DNA–RNA hybrid and the displaced ssDNA (Figure 1) and have several different ways of forming in cells, including TRCs [80,81,82,83]. R-loops were first discovered in bacteria and were shown to initiate DNA replication [84]. Today, these structures are known to exist in the cells of many living organisms, including mammals, and they participate in a variety of processes, including DNA replication, the regulation of gene expression, and immunoglobulin class-switch recombination; however, they also pose a serious threat to genomic stability [82,85,86,87]. Given that R-loops appear to have great functional significance in certain conditions, they are usually divided into physiological and pathological, that is, those that are formed during specialized processes and impacted by specific factors, and those that are formed spontaneously or by mistake and cause disturbances in the genetic stability of organisms [88]. It was suggested to separate RNAPII-promoted R-loops into two classes: (1) promotor-paused R-loops, which are short R-loops that form at a high frequency during promoter–proximal pausing by RNA polymerase; and (2) the considerably less common long elongation-associated R-loops, which occur throughout gene bodies [89]. The authors further suggested that Class 1 R-loops are responsible for most cases of R-loop-induced genome instability.
Persistent R-loops are one of the reasons for ongoing DNA damage and transcription-associated replication stress. The formation of R-loops can lead to DNA damage and genetic instability in several ways: First, the existence of very long R-loops leads to the formation of extended ssDNA regions, which are more sensitive to the occurrence of various forms of damage than double-stranded DNA (dsDNA). This is effectively demonstrated by the fact that in E. coli, the DNA strand that is not currently transcribed (the ssDNA fragment in the R-loop) is more often a source of mutation during transcription than the transcribed one (the DNA that is hybridized with RNA) [90]. Second, it was shown that reducing the factors participating in the prevention of R-loop formation or their resolution leads to the active transformation of these structures into DNA DSBs by the NER endonucleases XPF (ERCC1) and XPG (ERCC5), which are named after the rare autosomal recessive congenital syndrome xeroderma pigmentosum (XP) [91]. Third, the emergence of R-loops, as well as other secondary DNA structures and DNA–protein complexes, along the pathway of RNAPs can lead to transcription arrest [92]. Finally, the emergence of persistent R-loops in the replication machinery passage may end up blocking its progression and causing the replication fork to collapse, leading to subsequent DSB formation [85,93]. HO-TRC-induced R-loops may lead to replication fork reversal (the conversion of replication forks into four-way junctions via strand exchange reactions). This process involves replisome disruption and the protection of the newly formed DNA duplex with RAD51 filaments, stabilized by breast cancer type 2 susceptibility gene protein (BRCA2) [94,95,96]. Despite the fact that in general replication fork reversal is a protective cellular mechanism, it can have detrimental consequences if not properly regulated. Reversal that is not properly resolved, or excessive fork reversal can result in DNA degradation by nucleases, leading to DNA break emergence, genomic instability, and potentially cell death [97,98,99].
In fact, the details of what happens when the replication fork encounters R-loops remain murky. In the case of eukaryotes, this is largely due to the plasticity of their replicative apparatus, which complicates the study of the influence of R-loops on the possibility of resolving TRCs. In their work, Hamper et al. developed an episomal system to study TRCs and demonstrated the main difference between the cellular responses toward CD- and HO-TRC-induced R-loops [30]. First, the level of R-loop accumulation itself was shown to be dependent on the type of TRC, with HO-TRCs leading to R-loop formation and CO-TRCs resulting in R-loops being resolved [30]. It also appeared that CD-TRCs induce serine protein kinase ATM activation, while HO-TRCs are ATR-inducing. Ataxia telangiectasia-mutated (ATM) and RAD3-related DNA damage response (DDR) kinases (ATRs) are typically activated in response to stretches of replication protein A (RPA)-coated ssDNA at stalled replication fork sites, while ATM is mostly activated by DSBs [100]. The recent work by Zhand et al. linked the activation of ATR by HO-TRC-induced R-loops to the RNA-editing enzyme ADAR1 and its interaction with TOPBP1 on R-loops. It was also assumed that ADAR1 further recruits specialized DHX9 and DDX21 helicases to unwind R-loops [101].

3.2.1. Factors Promoting R-Loop Formation

Despite the fact that R-loops are ubiquitously formed in the cells of living organisms, little is known about the factors promoting their formation. Generally, it is thought that R-loops are mostly formed co-transcriptionally (in a cis manner) due to the hybridization of the newly synthetized nascent messenger RNA and the DNA template [85,102]. There is also a possibility of R-loops being formed in a trans manner when RNA hybridizes with a complementary DNA sequence far away from the original site of its transcription [103]; recombination factors, such as RAD51 and RAD52, as well as the RAD51 bacterial homolog RecA, may play a role in this trans R-loop formation [87,104].
There is presumably a direct link between the formation of G-quadruplex structures and R-loops. R-loops are formed at the promoter and terminator sites of genes [105,106,107] and are particularly common where there is an imbalance between the concentrations of C and G on complementary strands [105,108]. Such sites are known to be prone to the formation of G-quadruplex structures, while it has also been shown that the transcription of G-rich sites does indeed lead to the accumulation of R-loops [83,109,110]. For example, it was demonstrated that the helicase DDX1 is able to convert G-quadruplexes into R-loops, which promotes IgH class-switch recombination [111]. Apparently, the most significant impacts on the cell are caused by TRCs causing RNAP arrest, since this can lead to the blockage of the replication fork and a collision with moving RNAP, as it does not necessarily result in the stalling of the replication machinery [112,113]. As was mentioned above, the occurrence of different non-B DNA structures, including R-loops, before the replisome passage can easily result in RNAP stalling [14]. On the other hand, the extended pausing of RNAP on transcription sites can stabilize R-loops [114]. It is therefore hard to distinguish between the harmful effects of R-loops and RNAPs and the detrimental consequences of HO-TRCs. In Drosophila melanogaster, RNAP II abundance appears to be a more significant R-loop-inducing factor than even the properties of R-loop-forming DNA (GC-rich sequences). Global R-loop induction was demonstrated to be strongly coupled with RNAP II pausing [115]. It was also shown that RNAP II tends to accumulate with R-loops at HO-TRC sites and acts as the main obstacle to replication fork progression [116].
In a comprehensive study [117], more than 800 proteins that bind to DNA–RNA hybrids (particularly in R-loop structures) were identified, and more than 300 of these preferred hybrid structures to dsDNA, including known R-loops interactors, such as RNaseH 1 and DDX5. Such studies could provide valuable insights to help identify which of these proteins play important roles in both R-loop formation and resolution. In this review, we have tried to compile the available information on certain protein factors that supposably promote R-loop formation, which are all listed in Table 1.
In their study [114], Chakraborty et al. drew a line between R-loop formation and the role of certain splicing factors. Apparently, DEXH-box RNA helicase DHX9 (also known as RNA helicase A, RHA), which is known to participate in the assembly of splicing factors onto nascent RNA, promotes the formation of R-loops in cells where splicing factors are absent. On the other hand, in other studies, the ability of DHX9 to unwind DNA–RNA hybrids and G-quadruplex structures has instead been shown to be associated with its putative role in preventing the formation of R-loops [118,119]. It is possible that specialized helicases’ activity may have opposite consequences for R-loop persistence under different conditions. Their possible role in controlling DHX9 action might be facilitated by the factors attracting these helicases to R-loops, such as polycytosine (poly(C))-binding protein 1 (PCBP1), which was demonstrated to modulate transcription by regulating the accumulation and activity of DHX9 [120]. It was also shown that DHX9 is phosphorylated at Ser321 by HO-TRC-activated ATR, which facilitates the interaction of this helicase with BRCA1 and RPA and causes its association with R-loop sites [121].
Table 1. Protein factors promoting R-loop formation.
Table 1. Protein factors promoting R-loop formation.
ProteinProtein Function (Known Role in The Organism)Possible Role in R-Loops FormationRef.
RAD51 (RecA *)Plays an important role in homologous strand exchange, a key step in DNA repair, through homologous recombination (HR). Coats DNA-forming nucleoprotein filaments. Interacts with many partners. Participates in the Fanconi anemia (FA) pathway.Possibly involved in the formation of trans R-loops.[87,104]
RAD52Involved in DSB repair. Plays a key role in genetic recombination and DNA repair, promoting the annealing of complementary single-stranded DNA and stimulating RAD51 recombinase.
RNAP II (RNAP)DNA-dependent RNA polymerase which synthesizes mRNA precursors and several functional non-coding RNAs.Tends to accumulate together with R-loops at HO-TRC sites and acts as a main obstacle to replication fork progression. Extended pausing can stabilize R-loops. RNAP was demonstrated to partially protect short R-loop segments from RNase H1-facilitated cleavage.[114,116,122]
DHX9 (also known as RHA)RNA helicase A. Participates in the assembly of splicing factors onto nascent RNA.Promotes the formation of R-loops in cells where splicing factors are absent.[123]
PRC2Polycomb repressive complex. Exhibits histone methyltransferase activity and primarily methylates histone H3 on lysine 27.Opens DNA bubbles and induces the formation of RNA–DNA hybrids.[124]
RPABinds and stabilizes ssDNA intermediates that form during DNA replication or upon DNA stress.Supposably promotes R-loop formation by binding to RNA.[125]
TET1DNA dioxygenase. Catalyzes oxidation of epigenetic 5-methylcytosine to 5-hydroxymethylcytosine.Catalytic activity in the region of transcribed genes leads to preferential formation of R-loops therein.[86,126]
* Procaryotic homolog.
In their recent work, Alecki et al. demonstrated that D. melanogaster Polycomb repressive complex 2 (PRC2) opens DNA bubbles and induces the formation of RNA–DNA hybrids, which are essential components of R-loops [124]. This conclusion was supported by the observation that the activity of RING1B, a core component of PRC1, is crucial for the formation of nascent RNA transcripts and R-loops at the estrogen receptor alpha (Erα) target genes [127]. In contrast, Sanchez et al. demonstrated that the human Polycomb group proteins BMI1 and RNF2 suppressed TRCs, presumably through the regulation of RNAP II elongation, thus protecting common fragile sites from breakage [128].
It is interesting that the role of RPA, which is generally considered to preferentially bind ssDNA, in R-loop formation was also suggested. It was shown that RPA is not only able to bind RNA with considerably high efficiency, but also tends to associate with R-loops in vivo [125]. The authors of this work suggest that RPA promotes R-loop formation by binding to RNA. Interestingly, Mazina et al. demonstrated the ability of human DNA polymerases to initiate DNA synthesis by utilizing RPA-generated R-loops, thus reproducing replication restart in vivo. Moreover, it was suggested that RPA, localizing at the R-loop sites in SETX-deficient cells, prevents R-loop-induced DSB formation [129].
An interesting link between DNA methylation and R-loops has been demonstrated in several studies, suggesting new protein candidates for promoting R-loop formation. It was demonstrated that TET1, a member of the ten-eleven translocation (TET) DNA dioxygenases family, can be recruited to CpG island promoters through interactions with growth arrest and DNA damage protein 45A (GADD45A), which specifically binds to R-loops [86]. Sabino et al. demonstrated that the emergence of 5-methylcytosine (m5C)—which is subsequently transformed into 5-hydroxymethylcytosine (hm5C) via the dioxygenase activity of TET—in the body of transcribed genes favors R-loop formation [126]. Moreover, the depletion of TET in cells leads to a decrease in R-loop levels. TET DNA dioxygenases are key participants in the modulation of epigenetic methylation levels in DNA [130,131,132]. Another essential activity in this process is carried out by the enzymes belonging to the DNA methyltransferase family (DNMT) [133,134,135]. In their recent work, Shih et al. discovered that abolishing DNMT3B activity in cells leads to an increase in the XPG/XPF-mediated conversion of R-loops into DSBs [136]. The authors suggested that the de novo methyltransferase DNMT3B somehow protects R-loops from elevated processing by XPG and XPF. These observations reveal new facets of the relationship between R-loop-induced replication stress and the epigenetic modulation of transcription.

3.2.2. Factors Facilitating R-Loop Suppression

In contrast, the mechanisms that play a part in resolving R-loops, or preventing their formation, are intensively studied [102]. In general, the different factors protecting genomic stability from undesired R-loop propagation can be divided into three partially overlapping groups: (1) the factors preventing R-loop formation (Figure 2), (2) the factors resolving already formed R-loops (Figure 3 and Figure 4), and (3) the factors removing R-loops indirectly by restarting stalled replication forks and repairing DNA damage (Figure 5) [137]. The first group can be conditionally classified as all the factors that are responsible for DNA and chromatin structure stabilization (including topoisomerases, DNA helicases, and nucleosome assembly factors). The second group mainly consists of transcription and RNA metabolism factors (such as RNA-coating proteins, RNA helicases, and RNase H). Finally, the third group includes a range of protein factors that is involved in DNA metabolism and repair (such as the oncosuppressors BRCA1 and 2, the Fanconi anemia (FA) complex, ATR, and XPG) [137]. All the proteins that participate in different aspects of the protection of cells from R-loop propagation and are described in this review are listed in Table 2.
Factors Preventing R-Loop Formation
  • RNA-coating proteins
One of the possible ways for cells to prevent R-loop formation is thought to be the coating of nascent RNA with specialized proteins that are involved in processing and exporting newly synthetized RNA molecules [77,214], as was hypothesized for RPA for example [129]. In their recent work, Pan et al. demonstrated that the cohesins SA1 and SA2, the key components of the cohesin complex, play important roles in 3D chromatin organization, colocalizing with several R-loop sites (Figure 2) [141]. Considering the fact that SA1 and SA2 bind strongly to RNA, especially preferring DNA–RNA hybrids; this observation suggests that these proteins have some role in R-loop metabolism, although there is currently no direct evidence that they resolve R-loops.
  • Chromatin structure
There is also growing evidence that the chromatin structure itself may influence the prevention of pathological R-loop formation (Figure 2). Garcia-Pichardo et al. found several viable S. cerevisiae histone H3 and H4 mutants, characterized by increased levels of R-loops [215]. Additionally, Zhou et al. suggested that the loss of H3K9 dimethylation (H3K9me2), associated with transcriptional repression, provokes the accumulation of R-loops at the rDNA locus [216,217,218]. To ensure rescue from a situation in which TRCs can lead to the formation of stable R-loops and the stalling of replication forks, cells can switch between different types of chromatin modification, such as the change from crotonylation to the ubiquitination of H2A at Lys119 [219]. Thus, the H2AK119ub histone mark correlates with the repression of transcription [220] and was shown to promote the dissociation of RNAP II at reversive replication forks, preventing the transcription of genes located near the stalled replication forks, and thus preventing TRCs [219]. Recently, Bayona-Feliu et al. demonstrated that the depletion of the key chromatin factors INO80, SMARCA5, and MTA2 results in increases in TRCs, replication fork stalling, and R-loop-mediated DNA damage [142]. Overall, the association of various changes in histone functionality, leading to the disruption of genome heterochromatinization, with the increased accumulation of R-loops is in accordance with the tendency of more exposed DNA regions to favor DNA–RNA hybrid formation [221,222].
  • Topoisomerases
Given that the formation of R-loops requires the transcription process to occur and is facilitated by the negative supercoiled topology of DNA that is generated through the passage of RNAP, the prevention of R-loop formation can also be facilitated by the work of specialized topoisomerases. However, R-loop formation that is caused by the excessive presence of gyrase over that of Top1 leads to an increase in the hypernegative supercoiling of DNA [84,223]. It was shown that Top1, a ubiquitous member of the Type IB subfamily of topoisomerases, participates in preventing R-loop formation by resolving the accumulation of local negative supercoils on transcribed regions (Figure 2) [50,144]. Top1 is also required to prevent HO-TRCs by pausing the replication fork at the terminators of highly expressed genes containing R-loops [224]. Recently, it was demonstrated that a much-less-commonly studied member of the Type IA subfamily of topoisomerases, Topoisomerase III beta (TOP3B), also plays an important role in the suppression of R-loop accumulation [143,145,146]. Saha et al. proposed that TOP3B suppresses R-loop propagation in tandem with DDX5 helicase [225].
  • PrimPol
Among the different sequences that are prone to forming various non-B DNA secondary structures, a major proportion of the range of tandem repeats are also known as microsatellites [226]. For example, a long tract of polypurine–polypyrimidine (GAA)n repeats can form H-DNA, a triplex non-B DNA structure, which is associated with the inherited neurodegenerative disorder Friedreich’s ataxia. This is very probably due to the tendency of such structures to efficiently block replication [151,227,228]. Moreover, such triplet tandem repeats are prone to R-loop formation [229]. It is interesting that considerably shorter repetitive tracts, which are distributed throughout the human genome [230,231], apparently provoke R-loop formation as well [151]. Šviković et al. demonstrated that the impediment of (GAA)10 repeat-induced R-loop formation and replication can be resolved by including PrimPol in the processive replication of this sequence [151]. The DNA polymerase called Primase-Polymerase (PrimPol) was identified over a decade ago and has been assigned with a role in DNA damage tolerance in eukaryotes [232,233,234]. Schiavone et al. demonstrated that PrimPol, although unable to directly replicate G-quadruplex structures, facilitates their bypassing by repriming downstream of these structures (Figure 2) [235]. These observations suggest that PrimPol has a role in TRC resolution.
Factors Resolving R-Loops
  • RNase H and others
The primary factors protecting genetic stability against the detrimental propagation of R-loops are currently thought to be RNA-cleaving enzymes, specifically RNase H (Figure 3) [236]. Monomeric RNase H1 and heterotrimeric RNase H2 are highly conserved among living organisms and were demonstrated to specifically degrade RNA in DNA–RNA hybrids [237,238]. The preference of these enzymes toward DNA–RNA hybrids is facilitated by their specialized hybrid-binding domain, and the RNase H domain ensures the degradation of RNA strands [237]. In E. coli, the interaction of RNase H1 with ssDNA-binding protein (SSB) binds RNase to DNA replication sites [239]. In eukaryotes, RPA has also been shown to be a key partner of RNase H1, not only colocalizing with this enzyme on R-loops, but also stimulating its nuclease activity [138,139]. RNase H1 and H2 both were demonstrated to disintegrate R-loops, but RNase H2 can also remove mis-incorporated ribonucleotides from DNA [153,155], as well as Okazaki primers during lagging-strand DNA synthesis [240].
Figure 3. Factors directly resolving R-loops. RNA-cleaving enzymes include RNase H1 and H2, ribonuclease DICER, RNA exonuclease 5 (REXO5), and flap endonuclease 1 (FEN1). Examples of DNA-RNA helicases include senataxin (SETX), members of RecQ-like family helicases (Bloom syndrome protein (BLM), Werner syndrome helicase (WRN), and RECQ5) and DEAD/DEXH-box RNA helicases (such as DDX1, DDX5, DDX19, DDX21, DDX37, DDX50, and DHX9).
Figure 3. Factors directly resolving R-loops. RNA-cleaving enzymes include RNase H1 and H2, ribonuclease DICER, RNA exonuclease 5 (REXO5), and flap endonuclease 1 (FEN1). Examples of DNA-RNA helicases include senataxin (SETX), members of RecQ-like family helicases (Bloom syndrome protein (BLM), Werner syndrome helicase (WRN), and RECQ5) and DEAD/DEXH-box RNA helicases (such as DDX1, DDX5, DDX19, DDX21, DDX37, DDX50, and DHX9).
Ijms 26 06951 g003
It seems that RNase H2 plays a pivotal role in controlling co-transcriptional R-loop accumulation, as it has been shown that this nuclease binds directly to RNAP II (Figure 3) [154]. Nevertheless, Amon end Koshland demonstrated that an increase in the number of RAD52 foci, which is interpreted as a signal for R-loop-associated damage and homologue recombination (HR), is only observed in cases of diminished activity of both RNase H enzymes [204]. It was also demonstrated that another DNA repair factor, RAD52, plays an important role in R-loop suppression, as its deficiency causes R-loop accumulation, leading to DNA damage [205]. It is thought that RNase H1 is especially important for mitochondrial DNA replication and resolving R-loops [241,242]. In addition, RNase H1 and H2 seem to have different degrees of cell cycle dependence, allowing them to control R-loop accumulation under different conditions. RNase H2 appears to be restricted to the G2/M phase of the cell cycle, whereas RNase H1 is cell cycle-independent and is activated in response to high R-loop loads [243]. It was also demonstrated that E. coli RNase H1′s R-loop-degrading activity depends on the structural features of the particular R-loop in vitro, as well as the presence of RNAP [122]. It appears that RNAP can partially protect short R-loop segments from RNase H1-facilitated cleavage, but it is forced to dissociate if the length of the uncovered hybrid duplex segment is sufficient to initiate RNase H1 activity, at least based on the in vitro experimental results.
RNase H3, which is closely related to RNase H2 in its structure and found in some Archaea and bacteria [244], is also able to efficiently remove RNA strands from HO-TRC-induced R-loops [245].
With the details of the pathways of R-loop formation, signaling, and resolution being far from completely understood, new studies make various contributions to this field. One of the latest and most interesting propositions about the recruitment of RNase H1 to R-loop sites is that this occurs through RNA methylation facilitated by methyltransferase-like 3 (METTL3) [210]. N6-methyladenosine (m6A) RNA methylation is part of several important processes relating to RNA functioning. The establishment of this mark is facilitated by the methyltransferase complex of METTL3, METTL14, and the Wilms tumor 1-associated protein (WTAP), a mammalian splicing factor [209,211,212]. METTL3 is thought to have a special role in m6A RNA methylation in response to DNA damage [246]. Abakir et al. proposed that m6A RNA methylation plays a role in R-loop removal under physiological conditions [207]. Kang et al. added to this observation a potential mechanism for METTL3 enrollment into R-loop formation regions. They demonstrated that one of the transcriptional regulators, the tonicity-responsive enhancer-binding protein (TonEBP), recruits METTL3 for the m6A RNA methylation of RNA strands from an R-loop, and that this methylation, in turn, facilitates the recruitment of RNase H1 to degrade this R-loop [210]. In contrast, according to Hao et al., DDX21 helicase, which is involved in R-loop unwinding [178], can also attract METTL3 at R-loop sites for their further methylation and degradation [208].
A role in R-loop degradation was recently assigned to a different ribonuclease than RNase H type: DICER. Despite the fact that DICER ribonuclease is mostly known for its role in the formation of small regulatory RNAs and microRNAs in the cytoplasm [247,248,249], it can also function in the nucleus. It was demonstrated that DICER participates in resolving R-loops by specifically cleaving the RNA within R-loops [156].
Fairly recently, Lee et al. identified a novel RNA exonuclease 5 (REXO5/LOC81691) due to its elevated mRNA expression level in chronic myeloid leukemia patients [157]. This novel enzyme was proposed to degrade R-loops, as its knockout cells demonstrated elevated levels of R-loops and DNA damage. This suggestion was confirmed in latter experiments, in which the abilities of REXO5 to bind to R-loop structures and degrade RNA within R-loops were demonstrated.
Another interesting example of an enzyme which has shown itself to be a potential tool for RNA degradation in R-loops is flap endonuclease 1 (FEN1). FEN1 plays an important role in DNA lagging strand maturation, as well as in a long-patch variant of base excision repair (BER) [206,250,251]. It was recently demonstrated that FEN1, besides performing other essential functions facilitated through its RNA- and DNA-degrading activity, is able to process RNA to resolve R-loops [159,160]. Moreover, FEN1 was shown to be recruited to R-loops in cells, and this process was significantly fueled by oxidative DNA damage [158]. The role of the BER pathway in the resolution of R-loops and TRCs was also suggested in another recent work [252].
Additionally, a role in R-loop processing was suggested for another enzyme that is involved in DNA repair. Poly [ADP-ribose] polymerase 1 (PARP1) is an important part of several cellular processes besides DNA repair. This enzyme exhibits ADP-ribose transferase activity, catalyzing the transfer of ADP-ribose units onto itself, as well as other protein targets and DNA ends [253,254,255]. Laspata et al. demonstrated that PARP1 is not only able to bind R-loops in vitro, but also shows a tendency to accumulate at R-loop sites in cells [206]. Considering the important role of this protein in facilitating DNA repair, this observation suggests the involvement of new players in the task of dealing with R-loops.
  • Helicases
Although helicases are currently thought to play a secondary role in R-loop resolution compared with RNases H, a growing body of evidence is suggesting that different helicases are required for the efficient resolution of TRCs and to reduce the burden of R-loop accumulation in cells (Figure 3) [256]. Thus, it was shown that two S. cerevisiae Pif1 family DNA helicases, Pif1 and Rrm3, resolve R-loops at the tRNA gene (tDNA) regions [162], thereby preventing the arrest of the replisome at these genes during HO-TRCs [161].
A putative role in R-loop resolution is also attributed to a subfamily of proteins that contain a conserved DEAxQ-like domain with RNA/DNA helicase activity [257], namely senataxin (SETX) and Aquarius (AQR) proteins [91,107,258]. These proteins belong to the same helicase superfamily, SF1, which also includes Pif1 and RecQ-like family helicases [257,259]. First, it was shown by Mischo et al. that Sen1, an essential component of the Nrd1-Nab3-Sen1 (NRD) complex of S. cerevisiae, plays an important role in transcription-associated genetic instability, not only by participating in transcription termination [260], but also by restricting the accumulation of R-loops [258]. The importance of the human homolog of Sen1 (SETX) in resolving R-loops was also demonstrated [107,261]. SETX was shown to unwind R-loops in the promotors of the RNAP II-transcribed gene regions and is probably recruited to transcription pause sites through its interaction with BRCA1 [168,169]. The mechanism of resolution of TRCs via the helicase activity of SETX seems to be associated with the promotion of replication restart at R-loop formation sites through the MUS81–LIG4–ELL pathway, which involves the MUS81-mediated cleavage of the leading chain of the stalled fork, the DNA ligase IV (LIG4)/XRCC4 complex-facilitated religation of the fork, and RNAP II passage provided by the elongation factor ELL (Figure 4) [95,170]. It is worth noting that an MUS81-dependant mechanism of R-loop resolution was also demonstrated for DDX17 helicase [171]. Additionally, Zhao et al. demonstrated in their work that SETX acts redundantly with RNase H2, considering that the double deletion of these two enzymes leads to more severe consequences than the individual suppression of each of them [262].
The RecQ-like helicases family, which is named after E. coli RecQ [263,264], in humans is represented by five enzymes: RECQ1, BLM, WRN, RECQ4, and RECQ5 [265]. All the helicases in this family possess a DEAD/DEAH-box helicase-conserved C-terminal domain and can unwind a range of DNA structures, including G-quadruplexes and forked DNA duplexes [264]. It was recently shown that RECQ1, along with RECQ5, is involved in a multistep process of reactivating R-loop-induced stalled replication forks [95]. RECQ5 is known to be involved in the prevention of replication fork stalling on genes that are transcribed by RNAP I and RNAP II [150]. This helicase was demonstrated to play an important role in the resolution of TRCs, particularly by removing RAD51 from the stalled replication fork to facilitate MUS81 endonuclease’s cleavage of the fork (Figure 4) [95,147,148,149]. At the same time, the observation that in cells that are deficient in RECQ5, RAD18- and BRCA1-dependent RAD51 foci are accumulated, and that the localization of this accumulation indicates a rise in unresolved replication intermediates, indicates that RECQ5 may exert a significant effect on cells recovering from TRCs through putative mechanisms [150]. There is a possibility that RECQ5 works on G-quadruplex structures, as it was demonstrated that this helicase is able to destabilize such structures in vitro [266]. Another homolog of RECQ5, BLM, was shown to participate in R-loop suppression. It was demonstrated that BLM is not only able to unwind DNA–RNA hybrids in vitro [267], but also directly resolves R-loops in cells [164]. WRN, which is involved in the modulation of RNAP I/II-dependent gene transcription, was also demonstrated to participate in the suppression of R-loop accumulation [163,165,166,167].
Figure 4. The MUS81-dependent mechanism of resolving R-loops. The helicase REQ5 removes RAD51 from the stalled replication fork to facilitate MUS81 endonuclease’s cleavage of the leading chain of the stalled replication fork. The helicase SETX or DDX17 directly unwinds the R-loop. The further religation of the fork is facilitated through the participation of the DNA ligase IV (LIG4)/XRCC4 complex, and RNAP II’s passage is provided by the elongation factor ELL. Based on [95,170].
Figure 4. The MUS81-dependent mechanism of resolving R-loops. The helicase REQ5 removes RAD51 from the stalled replication fork to facilitate MUS81 endonuclease’s cleavage of the leading chain of the stalled replication fork. The helicase SETX or DDX17 directly unwinds the R-loop. The further religation of the fork is facilitated through the participation of the DNA ligase IV (LIG4)/XRCC4 complex, and RNAP II’s passage is provided by the elongation factor ELL. Based on [95,170].
Ijms 26 06951 g004
One of the possible mechanisms for recruiting specialized helicases, namely DEAD/DEXH-box RNA helicases (such as DDX1, DDX5, DDX19, DDX21, DDX37, DDX50, and DHX9), to resolve undesirable R-loops is through the activity of the ATAD5 tumor suppressor [119,172,173,174,175,176,177,178]. ATAD5 plays an important role in maintaining genomic stability, mainly due to its function as a PCNA (the eukaryotic sliding clamp for replicative polymerases) unloader at the end of DNA synthesis during normal DNA replication [268,269,270]. Lee et al. demonstrated that ATAD5 increases the abundance of DEAD/DEXH-box RNA helicases at replication fork sites, thus participating in R-loop resolution [175].
Recently, it was demonstrated that the HELQ helicase belonging to the superfamily 2 (SF2) (unlike SETX and RECQ5, which are discussed above) is recruited to R-loops through its interaction with RPA [140]. It was shown to unwind R-loops in vitro as well as in cells, and more intriguingly to interact with nuclear 5’-to-3’ exoribonuclease XRN2, supposably coordinating the unwinding and degradation of R-loops.
A rather curious example of the enzymes that are involved in R-loop resolution is nucleolar N-acetyltransferase 10 (NAT10). This enzyme, besides its lysine and/or cytidine acetyltransferase activity, also harbors an RecD helicase domain (RHD) [179,180,181,182]. Su et al. demonstrated the ability of NAT10 to resolve R-loops in vitro by utilizing its RHD [182]. Moreover, the in vivo participation of NAT10 in controlling the propagation of R-loops was also dependent on its acetyltransferase activity. NAT10 was demonstrated to acetylate the DEAD-box RNA helicase DDX21 at K236 and K573, thus enhancing its helicase activity towards nucleolar R-loops [182]. It was found that NAT10 can facilitate R-loop resolution through different pathways, which highlights its role in protecting the cell from harmful TRCs.
Factors Providing Indirect R-Loop Removal
  • DNA repair complexes
Given that pathological R-loop propagation is clearly connected to the significant growth of DSBs [85,91,93], it is fair to suppose that some pathway cell activities addressing persistent R-loops are made possible by the proteins that are involved in DSB formation and repair. Indeed, it was demonstrated that the deletion of RAD50, an important member of the MRE11-RAD50-NBS1 (MRN) complex, results in increasing R-loop abundance in long coding genes [190]. The MRN complex is a highly conserved protein complex that plays an important role in the maintenance of genomic stability through its functions in early DNA damage signaling and the processing of DNA ends in the DSB repair pathway [191,192,193]. Chang et al. concluded in their study that the role of MRN in the suppression of R-loop formation at TRCs is provided through the recruitment of FA pathway complexes to these sites (Figure 5) and not through the nuclease activity of MRE11 [190,193]. This is in a good agreement with evidence showing that the homologues yeast complex MRX (Mre11-Rad50-Xrs2) is important for maintaining the stability of replisomes at TRCs with stalled replication forks in Sen1-deficient cells [271].
Figure 5. Several factors have a complex indirect influence on R-loop resolution. The MRE11-RAD50-NBS1 (MRN) complex suppresses R-loop formation through the recruitment of Fanconi anemia (FA) pathway complexes to these sites. Replication protein A (RPA) coats DNA-forming nucleoprotein filaments and signals to activate ataxia telangiectasia-mutated (ATM) and RAD3-related DNA damage response (DDR) kinases (ATRs), as well as attracting RNA-degrading RNase H1 and HELQ helicase to R-loops. ATR interacts with the RNA-editing enzyme ADAR1, which further recruits the specialized DHX9 and DDX21 helicases to unwind R-loops. ATR also phosphorylates DHX9, facilitating the interaction of this helicase with BRCA1 and RPA, and provides its association with R-loop sites. BRCA1 can associate with SETX helicase, thus unwinding R-loops. BRCA2 targets RAD51 to ssDNA rather than dsDNA, thus enabling RAD51 to displace RPA from ssDNA, and recruits RNase H2.
Figure 5. Several factors have a complex indirect influence on R-loop resolution. The MRE11-RAD50-NBS1 (MRN) complex suppresses R-loop formation through the recruitment of Fanconi anemia (FA) pathway complexes to these sites. Replication protein A (RPA) coats DNA-forming nucleoprotein filaments and signals to activate ataxia telangiectasia-mutated (ATM) and RAD3-related DNA damage response (DDR) kinases (ATRs), as well as attracting RNA-degrading RNase H1 and HELQ helicase to R-loops. ATR interacts with the RNA-editing enzyme ADAR1, which further recruits the specialized DHX9 and DDX21 helicases to unwind R-loops. ATR also phosphorylates DHX9, facilitating the interaction of this helicase with BRCA1 and RPA, and provides its association with R-loop sites. BRCA1 can associate with SETX helicase, thus unwinding R-loops. BRCA2 targets RAD51 to ssDNA rather than dsDNA, thus enabling RAD51 to displace RPA from ssDNA, and recruits RNase H2.
Ijms 26 06951 g005
It seems natural that the types of cell that go through rapid proliferation will encounter TRCs more often. Thus, it was demonstrated that primordial germ cells (PGCs) are subjected to high levels of TRC-induced replication stress, while nevertheless preserving their genetic integrity with excellent efficiency [272]. It was discovered that one of the main factors for PGCs to cope effectively with TRCs is the presence of a functional replication-coupled FA pathway. FA proteins are represented by an FA core complex, the FANCI-FANCD2 complex, as well as downstream DNA repair proteins, including BRCA1 and 2, RAD51, XRCC2, and 9 [273]. Yang et al. showed that when the key step of FA pathway activation, the monoubiquitination of FANCD2, is disabled, rises in the number of stalled replication forks and R-loop accumulation occur [272,274]. This observation is in good agreement with the above statements on the role of BRCA1 and 2 in the suppression of pathological R-loop propagation.
BRCA1 is a widely known tumor suppressor gene, whose protein product participates, through its interactions with different partners, in a variety of cellular processes, including DNA damage repair, replication fork protection, transcription, and cell cycle regulation [275,276,277]. It was demonstrated for mammalian and yeast cells that BRCA1 participates in transcription activation [278] and is able to interact with RNAP II through DHX9 [198,200]. The recent data has linked BRCA1 and R-loop resolution. It seems that BRCA1 participates in R-loop suppression during RNAP II stalling [195]. Moreover, it was shown that BRCA1 is able to associate with Sen1/SETX, thus probably facilitating R-loop resolution through this interaction (Figure 5) [194,196,197,199]. The connection between the role of BRCA1 (as well as BRCA2) in R-loop resolution and DHX9 helicase was highlighted in a recent study conducted by Patel et al., where deficiency in RNF168, an E3 ubiquitin ligase and a DSB responder that directly ubiquitylates DHX9, resulted in the accumulation of R-loops in BRCA1/2-deficient breast and ovarian cancer cells [279]. On the other hand, it was shown that BRCA1 participates in the activation of the persistent R-loop-induced HR pathway, involving this protein in the development of a negative cellular scenario that is caused by the formation of R-loops [280].
Although BRCA2 is also a known tumor suppressor protein, mutations of which play a critical role in the development of high-grade serous ovarian carcinoma, the mechanisms by which this protein contributes to early oncogenesis prevention continue to be debated [1]. Interestingly, a series of recent studies have linked this tumor suppressor to the transcriptional regulation and recruitment of RNase H2 to DNA damage sites (Figure 5) [202,203], as well as to the R-loop resolution process itself [1,201]. Goehring et al. demonstrated that the preferential occurrence of the newly fired replication origins (dormant origin) near the transcription termination sites of active genes during global replication stress leads to the accumulation of HO-TRCs in BRCA2-deficient cells [1].
  • Transcription factors
One of the most efficient strategies to deal with the detrimental consequences of HO-TRCs is to control the available amount of arrested RNAPs. It was first demonstrated that transcription termination factors could participate in TRC resolution through the observation that the chemical inhibition of the bacterial transcription termination factor Rho results in a rise in replication-associated DSBs [281]. Later, it was demonstrated that the depletion of Rho leads to persistent R-loop formation [213]. It was also shown that another transcription termination factor, Rat1/XRN2, facilitates transcription through R-loops, presumably by degrading an RNA strand with its 5’-3’-exonuclease activity and inducing the premature termination of arrested RNAP II [282,283]. The presence of bacterial transcription elongation factors, such as DskA, Mfd, and GreA/B, was also demonstrated to be important for resolving TRCs in vivo [284,285].
CDK12 was shown to be one of the key mediators of transcriptional elongation in eukaryotes, and recent papers revealed the connection between CDK12 and R-loops and TRCs [286,287]. It was demonstrated that the knockout of oncogene CDK12 in murine prostate cancer cells leads to the increased formation of R-loops through the upregulation of the androgen receptor (AR) and its coactivator FOXA1 [288].

4. Conclusions and Future Perspectives

R-loops are ubiquitous structures that can be found in many living organisms; they are known to participate in a variety of processes, including DNA replication, the regulation of gene expression, and immunoglobulin class-switch recombination, but a possible imbalance in R-loop metabolism is recognized as a serious threat to genomic stability. In recent decades, HO-TRC-induced R-loop formation has received significant attention from the scientific community. Its role in detrimental events, such as replication fork disruption and a rise in DSB levels, as well as the resulting elevated mutagenesis, has been demonstrated in several studies. The normal level of R-loops in a cell is regulated by various factors and processes that control their formation and resolution [236]. However, little is known about the factors facilitating R-loop formation. Along with the existence of particular sequences in DNA, prone to forming R-loops, and the key role of RNAP pausing in TRCs, there is extensive evidence for the participation of certain protein factors, such as RAD51 and RAD52 [87,104], DHX9 [123], and PRC2 [124], in R-loop formation. However, even these scarce examples need to be further supported by new research, taking the existing contradictions in the available data into account.
As soon as the negative supercoiling of DNA is favorable for R-loop formation, Top1 relaxes this type of supercoiling, thus playing an important role in the prevention of R-loop formation [50,144,224]. The level of R-loops increases significantly if the functioning of factors that are responsible for the export of mRNAs, such as the THO/TREX complex, is disrupted in the cell [77,78,79,289]. If persistent R-loops are formed, RNase H1, 2, and 3 are the enzymes which specialize in the enzymatic degradation of these structures [154,241,242,243], although growing interest in this area of research is bearing fruit in the form of new candidates for this role [156,157,158]. On the other hand, R-loops can be resolved by specialized helicases that are recruited to transcription and replication machineries [55,162,258,290]. The important roles of numerous factors which are also involved in various DNA repair pathways and epigenetic methylation maintenance, also becomes undeniable [126,136,195,272,274]. There is growing evidence that it is the disruption of R-loop metabolism that subsequently leads to the replicative stress and genomic instability associated with such structures [55,291,292].

Author Contributions

Conceptualization, A.T.D. and N.A.K.; formal analysis, A.T.D.; investigation, A.T.D.; resources, A.T.D. and N.A.K.; writing—original draft preparation, A.T.D. and N.A.K.; writing—review and editing, A.T.D. and N.A.K.; project administration, N.A.K.; funding acquisition, N.A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation, grant No. 23-44-00064. Partial support by Russian state-funded project No. 121031300041-4 for the routine maintenance of equipment is also acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
hm5C5-hydroxymethylcytosine
m5C5-methylcytosine
AQRAquarius protein
ATRAtaxia telangiectasia-mutated (ATM) and RAD3-related DDR kinase
BERBase excision repair
BLMBloom syndrome protein
BRCA1/2Breast cancer type 1/2 susceptibility gene protein
CD-TRCCo-directional TRC
DSBsDouble-strand breaks
dsDNADouble-stranded DNA
dsRNADouble-stranded RNA
GADD45ADNA damage protein 45A
DDRDNA damage response
LIG4DNA ligase IV
DNMTDNA methyltransferase
ErαEstrogen receptor alpha
FAFanconi anemia
FEN1Flap endonuclease 1
H3K9me2H3K9 dimethylation
HO-TRCHead-on TRC
HRHomologous recombination
METTL3Methyltransferase-like 3
MMRMismatch repair
m6AN6-methyladenosine
NAT10N-acetyltransferase 10
NERNucleotide excision repair
PARP1Poly [ADP-ribose] polymerase 1
PRC2Polycomb repressive complex 2
PCBP1Polycytosine (poly(C))-binding protein 1
PGCsPrimordial germ cells
RHDRecD helicase domain
RPAReplication protein A
rDNARibosomal DNA
RNAPRNA polymerase
SETXSenataxin
ssDNASingle-stranded DNA
SSBSingle-stranded DNA-binding protein
SF2Superfamily 2 helicases
TETTen-eleven translocation DNA dioxygenase
TonEBPTonicity-responsive enhancer-binding protein
Top1Topoisomerase I
Top2Topoisomerase II
TOP3BTopoisomerase III beta
TC-NERTranscription-coupled nucleotide excision repair
TRCsTranscription–replication collisions/conflicts
TLSTrans-lesion synthesis
WRNWerner syndrome helicase
WTAPWilms tumor 1-associated protein

References

  1. Goehring, L.; Keegan, S.; Lahiri, S.; Xia, W.; Kong, M.; Jimenez-Sainz, J.; Gupta, D.; Drapkin, R.; Jensen, R.B.; Smith, D.J.; et al. Dormant Origin Firing Promotes Head-on Transcription-Replication Conflicts at Transcription Termination Sites in Response to BRCA2 Deficiency. Nat. Commun. 2024, 15, 4716. [Google Scholar] [CrossRef]
  2. Gaillard, H.; Aguilera, A. Transcription as a Threat to Genome Integrity. Annu. Rev. Biochem. 2016, 85, 291–317. [Google Scholar] [CrossRef]
  3. Aguilera, A.; García-Muse, T. Causes of Genome Instability. Annu. Rev. Genet. 2013, 47, 1–32. [Google Scholar] [CrossRef] [PubMed]
  4. Deshpande, A.M.; Newlon, C.S. DNA Replication Fork Pause Sites Dependent on Transcription. Science 1996, 272, 1030–1033. [Google Scholar] [CrossRef] [PubMed]
  5. French, S. Consequences of Replication Fork Movement through Transcription Units in Vivo. Science 1992, 258, 1362–1365. [Google Scholar] [CrossRef]
  6. Aguilera, A.; Gaillard, H. Transcription and Recombination: When RNA Meets DNA. Cold Spring Harb. Perspect. Biol. 2014, 6, a016543. [Google Scholar] [CrossRef]
  7. Liu, L.F.; Wang, J.C. Supercoiling of the DNA Template during Transcription. Proc. Natl. Acad. Sci. USA 1987, 84, 7024–7027. [Google Scholar] [CrossRef]
  8. Isik, E.; Shukla, K.; Pospisilova, M.; König, C.; Andrs, M.; Rao, S.; Rosano, V.; Dobrovolna, J.; Krejci, L.; Janscak, P. MutSβ-MutLβ-FANCJ Axis Mediates the Restart of DNA Replication after Fork Stalling at Cotranscriptional G4/R-Loops. Sci. Adv. 2024, 10, eadk2685. [Google Scholar] [CrossRef]
  9. Brill, S.J.; Sternglanz, R. Transcription-Dependent DNA Supercoiling in Yeast DNA Topoisomerase Mutants. Cell 1988, 54, 403–411. [Google Scholar] [CrossRef]
  10. Sperling, A.S.; Jeong, K.S.; Kitada, T.; Grunstein, M. Topoisomerase II Binds Nucleosome-Free DNA and Acts Redundantly with Topoisomerase I to Enhance Recruitment of RNA Pol II in Budding Yeast. Proc. Natl. Acad. Sci. USA 2011, 108, 12693–12698. [Google Scholar] [CrossRef]
  11. Christman, M.F.; Dietrich, F.S.; Fink, G.R. Mitotic Recombination in the rDNA of S. Cerevisiae Is Suppressed by the Combined Action of DNA Topoisomerases I and II. Cell 1988, 55, 413–425. [Google Scholar] [CrossRef]
  12. García-Rubio, M.L.; Aguilera, A. Topological Constraints Impair RNA Polymerase II Transcription and Causes Instability of Plasmid-Borne Convergent Genes. Nucleic Acids Res. 2012, 40, 1050–1064. [Google Scholar] [CrossRef]
  13. Duardo, R.C.; Guerra, F.; Pepe, S.; Capranico, G. Non-B DNA Structures as a Booster of Genome Instability. Biochimie 2023, 214, 176–192. [Google Scholar] [CrossRef] [PubMed]
  14. Belotserkovskii, B.P.; Mirkin, S.M.; Hanawalt, P.C. DNA Sequences That Interfere with Transcription: Implications for Genome Function and Stability. Chem. Rev. 2013, 113, 8620–8637. [Google Scholar] [CrossRef] [PubMed]
  15. Hanawalt, P.C.; Spivak, G. Transcription-Coupled DNA Repair: Two Decades of Progress and Surprises. Nat. Rev. Mol. Cell Biol. 2008, 9, 958–970. [Google Scholar] [CrossRef]
  16. Hamperl, S.; Cimprich, K.A. Conflict Resolution in the Genome: How Transcription and Replication Make It Work. Cell 2016, 167, 1455–1467. [Google Scholar] [CrossRef]
  17. Gómez-González, B.; Aguilera, A. Transcription-Mediated Replication Hindrance: A Major Driver of Genome Instability. Genes Dev. 2019, 33, 1008–1026. [Google Scholar] [CrossRef]
  18. Breier, A.M.; Weier, H.U.G.; Cozzarelli, N.R. Independence of Replisomes in Escherichia Coli Chromosomal Replication. Proc. Natl. Acad. Sci. USA 2005, 102, 3942–3947. [Google Scholar] [CrossRef]
  19. Liu, B.; Wong, M.L.; Alberts, B. A Transcribing RNA Polymerase Molecule Survives DNA Replication without Aborting Its Growing RNA Chain. Proc. Natl. Acad. Sci. USA 1994, 91, 10660–10664. [Google Scholar] [CrossRef]
  20. Liu, B.; Lie Wong, M.; Tinker, R.L.; Geiduschek, E.P.; Alberts, B.M. The DNA Replication Fork Can Pass RNA Polymerase without Displacing the Nascent Transcript. Nature 1993, 366, 33–39. [Google Scholar] [CrossRef]
  21. Kumar, C.; Batra, S.; Griffith, J.D.; Remus, D. The Interplay of RNA:DNA Hybrid Structure and G-Quadruplexes Determines the Outcome of R-Loop-Replisome Collisions. Elife 2021, 10, e72286. [Google Scholar] [CrossRef]
  22. Stoy, H.; Zwicky, K.; Kuster, D.; Lang, K.S.; Krietsch, J.; Crossley, M.P.; Schmid, J.A.; Cimprich, K.A.; Merrikh, H.; Lopes, M. Direct Visualization of Transcription-Replication Conflicts Reveals Post-Replicative DNA:RNA Hybrids. Nat. Struct. Mol. Biol. 2023, 30, 348–359. [Google Scholar] [CrossRef]
  23. Liu, B.; Alberts, B.M. Head-on Collision between a DNA Replication Apparatus and RNA Polymerase Transcription Complex. Science 1995, 267, 1131–1137. [Google Scholar] [CrossRef] [PubMed]
  24. Bedinger, P.; Hochstrasser, M.; Jongeneel, C.V.; Alberts, B.M. Properties of the T4 Bacteriophage DNA Replication Apparatus: The T4 Dda DNA Helicase Is Required to Pass a Bound RNA Polymerase Molecule. Cell 1983, 34, 115–123. [Google Scholar] [CrossRef] [PubMed]
  25. Browning, K.R.; Merrikh, H. Replication-Transcription Conflicts: A Perpetual War on the Chromosome. Annu. Rev. Biochem. 2024, 93, 21–46. [Google Scholar] [CrossRef] [PubMed]
  26. Pomerantz, R.T.; O’Donnell, M. Direct Restart of a Replication Fork Stalled by a Head-on RNA Polymerase. Science 2010, 327, 590–592. [Google Scholar] [CrossRef]
  27. Pomerantz, R.T.; O’Donnell, M. The Replisome Uses mRNA as a Primer after Colliding with RNA Polymerase. Nature 2008, 456, 762–766. [Google Scholar] [CrossRef]
  28. Fan, J.; Leroux-Coyau, M.; Savery, N.J.; Strick, T.R. Reconstruction of Bacterial Transcription-Coupled Repair at Single-Molecule Resolution. Nature 2016, 536, 234–237. [Google Scholar] [CrossRef]
  29. Mellon, I.; Hanawalt, P.C. Induction of the Escherichia Coli Lactose Operon Selectively Increases Repair of Its Transcribed DNA Strand. Nature 1989, 342, 95–98. [Google Scholar] [CrossRef]
  30. Hamperl, S.; Bocek, M.J.; Saldivar, J.C.; Swigut, T.; Cimprich, K.A. Transcription-Replication Conflict Orientation Modulates R-Loop Levels and Activates Distinct DNA Damage Responses. Cell 2017, 170, 774–786.e19. [Google Scholar] [CrossRef]
  31. Rocha, E.P. Evolutionary Patterns in Prokaryotic Genomes. Curr. Opin. Microbiol. 2008, 11, 454–460. [Google Scholar] [CrossRef]
  32. Akerman, I.; Kasaai, B.; Bazarova, A.; Sang, P.B.; Peiffer, I.; Artufel, M.; Derelle, R.; Smith, G.; Rodriguez-Martinez, M.; Romano, M.; et al. A Predictable Conserved DNA Base Composition Signature Defines Human Core DNA Replication Origins. Nat. Commun. 2020, 11, 4826. [Google Scholar] [CrossRef] [PubMed]
  33. Barlow, J.H.; Nussenzweig, A. Replication Initiation and Genome Instability: A Crossroads for DNA and RNA Synthesis. Cell. Mol. Life Sci. 2014, 71, 4545–4559. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, Y.-H.; Keegan, S.; Kahli, M.; Tonzi, P.; Fenyö, D.; Huang, T.T.; Smith, D.J. Transcription Shapes DNA Replication Initiation and Termination in Human Cells. Nat. Struct. Mol. Biol. 2019, 26, 67–77. [Google Scholar] [CrossRef] [PubMed]
  35. Guilbaud, G.; Murat, P.; Wilkes, H.S.; Lerner, L.K.; Sale, J.E.; Krude, T. Determination of Human DNA Replication Origin Position and Efficiency Reveals Principles of Initiation Zone Organisation. Nucleic Acids Res. 2022, 50, 7436–7450. [Google Scholar] [CrossRef]
  36. Prioleau, M.-N.; MacAlpine, D.M. DNA Replication Origins-Where Do We Begin? Genes Dev. 2016, 30, 1683–1697. [Google Scholar] [CrossRef]
  37. Koyanagi, E.; Kakimoto, Y.; Minamisawa, T.; Yoshifuji, F.; Natsume, T.; Higashitani, A.; Ogi, T.; Carr, A.M.; Kanemaki, M.T.; Daigaku, Y. Global Landscape of Replicative DNA Polymerase Usage in the Human Genome. Nat. Commun. 2022, 13, 7221. [Google Scholar] [CrossRef]
  38. Petryk, N.; Kahli, M.; d’Aubenton-Carafa, Y.; Jaszczyszyn, Y.; Shen, Y.; Silvain, M.; Thermes, C.; Chen, C.-L.; Hyrien, O. Replication Landscape of the Human Genome. Nat. Commun. 2016, 7, 10208. [Google Scholar] [CrossRef]
  39. Ide, S.; Miyazaki, T.; Maki, H.; Kobayashi, T. Abundance of Ribosomal RNA Gene Copies Maintains Genome Integrity. Science 2010, 327, 693–696. [Google Scholar] [CrossRef]
  40. Kobayashi, T.; Heck, D.J.; Nomura, M.; Horiuchi, T. Expansion and Contraction of Ribosomal DNA Repeats in Saccharomyces Cerevisiae: Requirement of Replication Fork Blocking (Fob1) Protein and the Role of RNA Polymerase I. Genes Dev. 1998, 12, 3821–3830. [Google Scholar] [CrossRef]
  41. Brewer, B.J.; Fangman, W.L. A Replication Fork Barrier at the 3’ End of Yeast Ribosomal RNA Genes. Cell 1988, 55, 637–643. [Google Scholar] [CrossRef]
  42. Defossez, P.-A.; Prusty, R.; Kaeberlein, M.; Lin, S.-J.; Ferrigno, P.; Silver, P.A.; Keil, R.L.; Guarente, L. Elimination of Replication Block Protein Fob1 Extends the Life Span of Yeast Mother Cells. Mol. Cell 1999, 3, 447–455. [Google Scholar] [CrossRef]
  43. Takeuchi, Y.; Horiuchi, T.; Kobayashi, T. Transcription-Dependent Recombination and the Role of Fork Collision in Yeast rDNA. Genes Dev. 2003, 17, 1497–1506. [Google Scholar] [CrossRef]
  44. Merrikh, C.N.; Merrikh, H. Gene Inversion Potentiates Bacterial Evolvability and Virulence. Nat. Commun. 2018, 9, 4662. [Google Scholar] [CrossRef]
  45. Paul, S.; Million-Weaver, S.; Chattopadhyay, S.; Sokurenko, E.; Merrikh, H. Accelerated Gene Evolution through Replication-Transcription Conflicts. Nature 2013, 495, 512–515. [Google Scholar] [CrossRef]
  46. Dutta, D.; Shatalin, K.; Epshtein, V.; Gottesman, M.E.; Nudler, E. Linking RNA Polymerase Backtracking to Genome Instability in E. Coli. Cell 2011, 146, 533–543. [Google Scholar] [CrossRef]
  47. Merrikh, H.; Machón, C.; Grainger, W.H.; Grossman, A.D.; Soultanas, P. Co-Directional Replication-Transcription Conflicts Lead to Replication Restart. Nature 2011, 470, 554–557. [Google Scholar] [CrossRef]
  48. Prado, F.; Aguilera, A. Impairment of Replication Fork Progression Mediates RNA polII Transcription-Associated Recombination. EMBO J. 2005, 24, 1267–1276. [Google Scholar] [CrossRef] [PubMed]
  49. Gottipati, P.; Cassel, T.N.; Savolainen, L.; Helleday, T. Transcription-Associated Recombination Is Dependent on Replication in Mammalian Cells. Mol. Cell. Biol. 2008, 28, 154–164. [Google Scholar] [CrossRef] [PubMed]
  50. Tuduri, S.; Crabbé, L.; Conti, C.; Tourrière, H.; Holtgreve-Grez, H.; Jauch, A.; Pantesco, V.; De Vos, J.; Thomas, A.; Theillet, C.; et al. Topoisomerase I Suppresses Genomic Instability by Preventing Interference between Replication and Transcription. Nat. Cell Biol. 2009, 11, 1315–1324. [Google Scholar] [CrossRef] [PubMed]
  51. Helmrich, A.; Ballarino, M.; Tora, L. Collisions between Replication and Transcription Complexes Cause Common Fragile Site Instability at the Longest Human Genes. Mol. Cell 2011, 44, 966–977. [Google Scholar] [CrossRef]
  52. Le Tallec, B.; Koundrioukoff, S.; Wilhelm, T.; Letessier, A.; Brison, O.; Debatisse, M. Updating the Mechanisms of Common Fragile Site Instability: How to Reconcile the Different Views? Cell. Mol. Life Sci. 2014, 71, 4489–4494. [Google Scholar] [CrossRef] [PubMed]
  53. Lang, K.S.; Merrikh, H. Topological Stress Is Responsible for the Detrimental Outcomes of Head-on Replication-Transcription Conflicts. Cell Rep. 2021, 34, 108797. [Google Scholar] [CrossRef] [PubMed]
  54. Brewer, B.J. When Polymerases Collide: Replication and the Transcriptional Organization of the E. coli Chromosome. Cell 1988, 53, 679–686. [Google Scholar] [CrossRef] [PubMed]
  55. García-Muse, T.; Aguilera, A. Transcription-Replication Conflicts: How They Occur and How They Are Resolved. Nat. Rev. Mol. Cell Biol. 2016, 17, 553–563. [Google Scholar] [CrossRef]
  56. Boubakri, H.; de Septenville, A.L.; Viguera, E.; Michel, B. The Helicases DinG, Rep and UvrD Cooperate to Promote Replication across Transcription Units in Vivo. EMBO J. 2010, 29, 145–157. [Google Scholar] [CrossRef]
  57. Merrikh, C.N.; Brewer, B.J.; Merrikh, H. The B. Subtilis Accessory Helicase PcrA Facilitates DNA Replication through Transcription Units. PLoS Genet. 2015, 11, e1005289. [Google Scholar] [CrossRef]
  58. Azvolinsky, A.; Dunaway, S.; Torres, J.Z.; Bessler, J.B.; Zakian, V.A. The S. Cerevisiae Rrm3p DNA Helicase Moves with the Replication Fork and Affects Replication of All Yeast Chromosomes. Genes Dev. 2006, 20, 3104–3116. [Google Scholar] [CrossRef]
  59. Ivessa, A.S.; Lenzmeier, B.A.; Bessler, J.B.; Goudsouzian, L.K.; Schnakenberg, S.L.; Zakian, V.A. The Saccharomyces Cerevisiae Helicase Rrm3p Facilitates Replication Past Nonhistone Protein-DNA Complexes. Mol. Cell 2003, 12, 1525–1536. [Google Scholar] [CrossRef]
  60. Estep, K.N.; Butler, T.J.; Ding, J.; Brosh, R.M. G4-Interacting DNA Helicases and Polymerases: Potential Therapeutic Targets. Curr. Med. Chem. 2019, 26, 2881–2897. [Google Scholar] [CrossRef]
  61. Datta, A.; Pollock, K.J.; Kormuth, K.A.; Brosh, R.M. G-Quadruplex Assembly by Ribosomal DNA: Emerging Roles in Disease Pathogenesis and Cancer Biology. Cytogenet. Genome Res. 2021, 161, 285–296. [Google Scholar] [CrossRef]
  62. Kamath-Loeb, A.S.; Loeb, L.A.; Johansson, E.; Burgers, P.M.; Fry, M. Interactions between the Werner Syndrome Helicase and DNA Polymerase Delta Specifically Facilitate Copying of Tetraplex and Hairpin Structures of the d(CGG)n Trinucleotide Repeat Sequence. J. Biol. Chem. 2001, 276, 16439–16446. [Google Scholar] [CrossRef] [PubMed]
  63. Hänsel-Hertsch, R.; Beraldi, D.; Lensing, S.V.; Marsico, G.; Zyner, K.; Parry, A.; Di Antonio, M.; Pike, J.; Kimura, H.; Narita, M.; et al. G-Quadruplex Structures Mark Human Regulatory Chromatin. Nat. Genet. 2016, 48, 1267–1272. [Google Scholar] [CrossRef] [PubMed]
  64. Hanakahi, L.A.; Sun, H.; Maizels, N. High Affinity Interactions of Nucleolin with G-G-Paired rDNA. J. Biol. Chem. 1999, 274, 15908–15912. [Google Scholar] [CrossRef] [PubMed]
  65. 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]
  66. Bharti, S.K.; Awate, S.; Banerjee, T.; Brosh, R.M. Getting Ready for the Dance: FANCJ Irons Out DNA Wrinkles. Genes 2016, 7, 31. [Google Scholar] [CrossRef]
  67. Mendoza, O.; Bourdoncle, A.; Boulé, J.-B.; Brosh, R.M.; Mergny, J.-L. G-Quadruplexes and Helicases. Nucleic Acids Res. 2016, 44, 1989–2006. [Google Scholar] [CrossRef]
  68. Sauer, M.; Paeschke, K. G-Quadruplex Unwinding Helicases and Their Function in Vivo. Biochem. Soc. Trans. 2017, 45, 1173–1182. [Google Scholar] [CrossRef]
  69. Sabouri, N.; McDonald, K.R.; Webb, C.J.; Cristea, I.M.; Zakian, V.A. DNA Replication through Hard-to-Replicate Sites, Including Both Highly Transcribed RNA Pol II and Pol III Genes, Requires the S. Pombe Pfh1 Helicase. Genes Dev. 2012, 26, 581–593. [Google Scholar] [CrossRef]
  70. Steinacher, R.; Osman, F.; Dalgaard, J.Z.; Lorenz, A.; Whitby, M.C. The DNA Helicase Pfh1 Promotes Fork Merging at Replication Termination Sites to Ensure Genome Stability. Genes Dev. 2012, 26, 594–602. [Google Scholar] [CrossRef]
  71. Wallgren, M.; Mohammad, J.B.; Yan, K.-P.; Pourbozorgi-Langroudi, P.; Ebrahimi, M.; Sabouri, N. G-Rich Telomeric and Ribosomal DNA Sequences from the Fission Yeast Genome Form Stable G-Quadruplex DNA Structures in Vitro and Are Unwound by the Pfh1 DNA Helicase. Nucleic Acids Res. 2016, 44, 6213–6231. [Google Scholar] [CrossRef]
  72. Eddy, S.; Ketkar, A.; Zafar, M.K.; Maddukuri, L.; Choi, J.-Y.; Eoff, R.L. Human Rev1 Polymerase Disrupts G-Quadruplex DNA. Nucleic Acids Res. 2014, 42, 3272–3285. [Google Scholar] [CrossRef]
  73. Bétous, R.; Rey, L.; Wang, G.; Pillaire, M.-J.; Puget, N.; Selves, J.; Biard, D.S.F.; Shin-ya, K.; Vasquez, K.M.; Cazaux, C.; et al. Role of TLS DNA Polymerases Eta and Kappa in Processing Naturally Occurring Structured DNA in Human Cells. Mol. Carcinog. 2009, 48, 369–378. [Google Scholar] [CrossRef] [PubMed]
  74. Eddy, S.; Maddukuri, L.; Ketkar, A.; Zafar, M.K.; Henninger, E.E.; Pursell, Z.F.; Eoff, R.L. Evidence for the Kinetic Partitioning of Polymerase Activity on G-Quadruplex DNA. Biochemistry 2015, 54, 3218–3230. [Google Scholar] [CrossRef] [PubMed]
  75. Eddy, S.; Tillman, M.; Maddukuri, L.; Ketkar, A.; Zafar, M.K.; Eoff, R.L. Human Translesion Polymerase κ Exhibits Enhanced Activity and Reduced Fidelity Two Nucleotides from G-Quadruplex DNA. Biochemistry 2016, 55, 5218–5229. [Google Scholar] [CrossRef]
  76. Koole, W.; van Schendel, R.; Karambelas, A.E.; van Heteren, J.T.; Okihara, K.L.; Tijsterman, M. A Polymerase Theta-Dependent Repair Pathway Suppresses Extensive Genomic Instability at Endogenous G4 DNA Sites. Nat. Commun. 2014, 5, 3216. [Google Scholar] [CrossRef]
  77. Huertas, P.; Aguilera, A. Cotranscriptionally Formed DNA:RNA Hybrids Mediate Transcription Elongation Impairment and Transcription-Associated Recombination. Mol. Cell 2003, 12, 711–721. [Google Scholar] [CrossRef]
  78. Piruat, J.I.; Aguilera, A. A Novel Yeast Gene, THO2, Is Involved in RNA Pol II Transcription and Provides New Evidence for Transcriptional Elongation-Associated Recombination. EMBO J. 1998, 17, 4859–4872. [Google Scholar] [CrossRef]
  79. Prado, F.; Piruat, J.I.; Aguilera, A. Recombination between DNA Repeats in Yeast Hpr1delta Cells Is Linked to Transcription Elongation. EMBO J. 1997, 16, 2826–2835. [Google Scholar] [CrossRef]
  80. Crossley, M.P.; Bocek, M.; Cimprich, K.A. R-Loops as Cellular Regulators and Genomic Threats. Mol. Cell 2019, 73, 398–411. [Google Scholar] [CrossRef]
  81. Marnef, A.; Legube, G. R-Loops as Janus-Faced Modulators of DNA Repair. Nat. Cell Biol. 2021, 23, 305–313. [Google Scholar] [CrossRef]
  82. Skourti-Stathaki, K.; Proudfoot, N.J. A Double-Edged Sword: R Loops as Threats to Genome Integrity and Powerful Regulators of Gene Expression. Genes Dev. 2014, 28, 1384–1396. [Google Scholar] [CrossRef]
  83. Sollier, J.; Cimprich, K.A. Breaking Bad: R-Loops and Genome Integrity. Trends Cell Biol. 2015, 25, 514–522. [Google Scholar] [CrossRef] [PubMed]
  84. Drolet, M.; Bi, X.; Liu, L.F. Hypernegative Supercoiling of the DNA Template during Transcription Elongation in Vitro. J. Biol. Chem. 1994, 269, 2068–2074. [Google Scholar] [CrossRef] [PubMed]
  85. Aguilera, A.; García-Muse, T. R Loops: From Transcription Byproducts to Threats to Genome Stability. Mol. Cell 2012, 46, 115–124. [Google Scholar] [CrossRef] [PubMed]
  86. Arab, K.; Karaulanov, E.; Musheev, M.; Trnka, P.; Schäfer, A.; Grummt, I.; Niehrs, C. GADD45A Binds R-Loops and Recruits TET1 to CpG Island Promoters. Nat. Genet. 2019, 51, 217–223. [Google Scholar] [CrossRef]
  87. Keskin, H.; Shen, Y.; Huang, F.; Patel, M.; Yang, T.; Ashley, K.; Mazin, A.V.; Storici, F. Transcript-RNA-Templated DNA Recombination and Repair. Nature 2014, 515, 436–439. [Google Scholar] [CrossRef]
  88. García-Muse, T.; Aguilera, A. R Loops: From Physiological to Pathological Roles. Cell 2019, 179, 604–618. [Google Scholar] [CrossRef]
  89. Castillo-Guzman, D.; Chédin, F. Defining R-Loop Classes and Their Contributions to Genome Instability. DNA Repair 2021, 106, 103182. [Google Scholar] [CrossRef]
  90. Beletskii, A.; Bhagwat, A.S. Transcription-Induced Mutations: Increase in C to T Mutations in the Nontranscribed Strand during Transcription in Escherichia Coli. Proc. Natl. Acad. Sci. USA 1996, 93, 13919–13924. [Google Scholar] [CrossRef]
  91. Sollier, J.; Stork, C.T.; García-Rubio, M.L.; Paulsen, R.D.; Aguilera, A.; Cimprich, K.A. Transcription-Coupled Nucleotide Excision Repair Factors Promote R-Loop-Induced Genome Instability. Mol. Cell 2014, 56, 777–785. [Google Scholar] [CrossRef]
  92. Gaillard, H.; Herrera-Moyano, E.; Aguilera, A. Transcription-Associated Genome Instability. Chem. Rev. 2013, 113, 8638–8661. [Google Scholar] [CrossRef]
  93. Gan, W.; Guan, Z.; Liu, J.; Gui, T.; Shen, K.; Manley, J.L.; Li, X. R-Loop-Mediated Genomic Instability Is Caused by Impairment of Replication Fork Progression. Genes Dev. 2011, 25, 2041–2056. [Google Scholar] [CrossRef]
  94. Berti, M.; Cortez, D.; Lopes, M. The Plasticity of DNA Replication Forks in Response to Clinically Relevant Genotoxic Stress. Nat. Rev. Mol. Cell Biol. 2020, 21, 633–651. [Google Scholar] [CrossRef]
  95. Chappidi, N.; Nascakova, Z.; Boleslavska, B.; Zellweger, R.; Isik, E.; Andrs, M.; Menon, S.; Dobrovolna, J.; Balbo Pogliano, C.; Matos, J.; et al. Fork Cleavage-Religation Cycle and Active Transcription Mediate Replication Restart after Fork Stalling at Co-Transcriptional R-Loops. Mol. Cell 2020, 77, 528–541.e8. [Google Scholar] [CrossRef] [PubMed]
  96. Mijic, S.; Zellweger, R.; Chappidi, N.; Berti, M.; Jacobs, K.; Mutreja, K.; Ursich, S.; Ray Chaudhuri, A.; Nussenzweig, A.; Janscak, P.; et al. Replication Fork Reversal Triggers Fork Degradation in BRCA2-Defective Cells. Nat. Commun. 2017, 8, 859. [Google Scholar] [CrossRef] [PubMed]
  97. Adolph, M.B.; Cortez, D. Mechanisms and Regulation of Replication Fork Reversal. DNA Repair 2024, 141, 103731. [Google Scholar] [CrossRef] [PubMed]
  98. Saxena, S.; Zou, L. Hallmarks of DNA Replication Stress. Mol. Cell 2022, 82, 2298–2314. [Google Scholar] [CrossRef]
  99. Liao, H.; Ji, F.; Helleday, T.; Ying, S. Mechanisms for Stalled Replication Fork Stabilization: New Targets for Synthetic Lethality Strategies in Cancer Treatments. EMBO Rep. 2018, 19, e46263. [Google Scholar] [CrossRef]
  100. Cimprich, K.A.; Cortez, D. ATR: An Essential Regulator of Genome Integrity. Nat. Rev. Mol. Cell Biol. 2008, 9, 616–627. [Google Scholar] [CrossRef]
  101. Zhang, B.; Li, Y.; Zhang, J.; Wang, Y.; Liang, C.; Lu, T.; Zhang, C.; Liu, L.; Qin, Y.; He, J.; et al. ADAR1 Links R-Loop Homeostasis to ATR Activation in Replication Stress Response. Nucleic Acids Res. 2023, 51, 11668–11687. [Google Scholar] [CrossRef] [PubMed]
  102. Hamperl, S.; Cimprich, K.A. The Contribution of Co-Transcriptional RNA:DNA Hybrid Structures to DNA Damage and Genome Instability. DNA Repair 2014, 19, 84–94. [Google Scholar] [CrossRef] [PubMed]
  103. Wahba, L.; Gore, S.K.; Koshland, D. The Homologous Recombination Machinery Modulates the Formation of RNA-DNA Hybrids and Associated Chromosome Instability. Elife 2013, 2, e00505. [Google Scholar] [CrossRef] [PubMed]
  104. Kasahara, M.; Clikeman, J.A.; Bates, D.B.; Kogoma, T. RecA Protein-Dependent R-Loop Formation in Vitro. Genes Dev. 2000, 14, 360–365. [Google Scholar] [CrossRef]
  105. Ginno, P.A.; Lott, P.L.; Christensen, H.C.; Korf, I.; Chédin, F. R-Loop Formation Is a Distinctive Characteristic of Unmethylated Human CpG Island Promoters. Mol. Cell 2012, 45, 814–825. [Google Scholar] [CrossRef]
  106. Skourti-Stathaki, K.; Kamieniarz-Gdula, K.; Proudfoot, N.J. R-Loops Induce Repressive Chromatin Marks over Mammalian Gene Terminators. Nature 2014, 516, 436–439. [Google Scholar] [CrossRef]
  107. Skourti-Stathaki, K.; Proudfoot, N.J.; Gromak, N. Human Senataxin Resolves RNA/DNA Hybrids Formed at Transcriptional Pause Sites to Promote Xrn2-Dependent Termination. Mol. Cell 2011, 42, 794–805. [Google Scholar] [CrossRef]
  108. Ginno, P.A.; Lim, Y.W.; Lott, P.L.; Korf, I.; Chédin, F. GC Skew at the 5’ and 3’ Ends of Human Genes Links R-Loop Formation to Epigenetic Regulation and Transcription Termination. Genome Res. 2013, 23, 1590–1600. [Google Scholar] [CrossRef]
  109. Duquette, M.L.; Pham, P.; Goodman, M.F.; Maizels, N. AID Binds to Transcription-Induced Structures in c-MYC That Map to Regions Associated with Translocation and Hypermutation. Oncogene 2005, 24, 5791–5798. [Google Scholar] [CrossRef]
  110. Duquette, M.L.; Handa, P.; Vincent, J.A.; Taylor, A.F.; Maizels, N. Intracellular Transcription of G-Rich DNAs Induces Formation of G-Loops, Novel Structures Containing G4 DNA. Genes Dev. 2004, 18, 1618–1629. [Google Scholar] [CrossRef]
  111. Ribeiro de Almeida, C.; Dhir, S.; Dhir, A.; Moghaddam, A.E.; Sattentau, Q.; Meinhart, A.; Proudfoot, N.J. RNA Helicase DDX1 Converts RNA G-Quadruplex Structures into R-Loops to Promote IgH Class Switch Recombination. Mol. Cell 2018, 70, 650–662.e8. [Google Scholar] [CrossRef]
  112. Elías-Arnanz, M.; Salas, M. Resolution of Head-on Collisions between the Transcription Machinery and Bacteriophage Φ29 DNA Polymerase Is Dependent on RNA Polymerase Translocation. EMBO J. 1999, 18, 5675–5682. [Google Scholar] [CrossRef]
  113. Elías-Arnanz, M.; Salas, M. Bacteriophage Φ29 DNA Replication Arrest Caused by Codirectional Collisions with the Transcription Machinery. EMBO J. 1997, 16, 5775–5783. [Google Scholar] [CrossRef]
  114. Belotserkovskii, B.P.; Shin, J.H.S.; Hanawalt, P.C. Strong Transcription Blockage Mediated by R-Loop Formation within a G-Rich Homopurine-Homopyrimidine Sequence Localized in the Vicinity of the Promoter. Nucleic Acids Res. 2017, 45, 6589–6599. [Google Scholar] [CrossRef] [PubMed]
  115. Zhang, X.; Liang, S.-B.; Yi, Z.; Qiao, Z.; Xu, B.; Geng, H.; Wang, H.; Yin, X.; Tang, M.; Ge, W.; et al. Global Coupling of R-Loop Dynamics with RNA Polymerase II Modulates Gene Expression and Early Development of Drosophila. Nucleic Acids Res. 2024, 52, 13110–13127. [Google Scholar] [CrossRef] [PubMed]
  116. Zardoni, L.; Nardini, E.; Brambati, A.; Lucca, C.; Choudhary, R.; Loperfido, F.; Sabbioneda, S.; Liberi, G. Elongating RNA Polymerase II and RNA:DNA Hybrids Hinder Fork Progression and Gene Expression at Sites of Head-on Replication-Transcription Collisions. Nucleic Acids Res. 2021, 49, 12769–12784. [Google Scholar] [CrossRef] [PubMed]
  117. Wang, I.X.; Grunseich, C.; Fox, J.; Burdick, J.; Zhu, Z.; Ravazian, N.; Hafner, M.; Cheung, V.G. Human Proteins That Interact with RNA/DNA Hybrids. Genome Res. 2018, 28, 1405–1414. [Google Scholar] [CrossRef]
  118. Chakraborty, P.; Grosse, F. Human DHX9 Helicase Preferentially Unwinds RNA-Containing Displacement Loops (R-Loops) and G-Quadruplexes. DNA Repair 2011, 10, 654–665. [Google Scholar] [CrossRef]
  119. Cristini, A.; Groh, M.; Kristiansen, M.S.; Gromak, N. RNA/DNA Hybrid Interactome Identifies DXH9 as a Molecular Player in Transcriptional Termination and R-Loop-Associated DNA Damage. Cell Rep. 2018, 23, 1891–1905. [Google Scholar] [CrossRef]
  120. Karam, J.A.Q.; Fréreux, C.; Mohanty, B.K.; Dalton, A.C.; Dincman, T.A.; Palanisamy, V.; Howley, B.V.; Howe, P.H. The RNA-Binding Protein PCBP1 Modulates Transcription by Recruiting the G-Quadruplex-Specific Helicase DHX9. J. Biol. Chem. 2024, 300, 107830. [Google Scholar] [CrossRef]
  121. Liu, M.-Y.; Lin, K.-R.; Chien, Y.-L.; Yang, B.-Z.; Tsui, L.-Y.; Chu, H.-P.C.; Wu, C.-S.P. ATR Phosphorylates DHX9 at Serine 321 to Suppress R-Loop Accumulation upon Genotoxic Stress. Nucleic Acids Res. 2024, 52, 204–222. [Google Scholar] [CrossRef]
  122. Kuznetsova, A.A.; Kosarev, I.A.; Timofeyeva, N.A.; Novopashina, D.S.; Kuznetsov, N.A. Kinetic Features of Degradation of R-Loops by RNase H1 from Escherichia Coli. Int. J. Mol. Sci. 2024, 25, 12263. [Google Scholar] [CrossRef]
  123. Chakraborty, P.; Huang, J.T.J.; Hiom, K. DHX9 Helicase Promotes R-Loop Formation in Cells with Impaired RNA Splicing. Nat. Commun. 2018, 9, 4346. [Google Scholar] [CrossRef]
  124. Alecki, C.; Chiwara, V.; Sanz, L.A.; Grau, D.; Arias Pérez, O.; Boulier, E.L.; Armache, K.-J.; Chédin, F.; Francis, N.J. RNA-DNA Strand Exchange by the Drosophila Polycomb Complex PRC2. Nat. Commun. 2020, 11, 1781. [Google Scholar] [CrossRef] [PubMed]
  125. Mazina, O.M.; Somarowthu, S.; Kadyrova, L.Y.; Baranovskiy, A.G.; Tahirov, T.H.; Kadyrov, F.A.; Mazin, A.V. Replication Protein A Binds RNA and Promotes R-Loop Formation. J. Biol. Chem. 2020, 295, 14203–14213. [Google Scholar] [CrossRef] [PubMed]
  126. Sabino, J.C.; de Almeida, M.R.; Abreu, P.L.; Ferreira, A.M.; Caldas, P.; Domingues, M.M.; Santos, N.C.; Azzalin, C.M.; Grosso, A.R.; de Almeida, S.F. Epigenetic Reprogramming by TET Enzymes Impacts Co-Transcriptional R-Loops. Elife 2022, 11, e69476. [Google Scholar] [CrossRef] [PubMed]
  127. Zhang, Y.; Liu, T.; Yuan, F.; Garcia-Martinez, L.; Lee, K.D.; Stransky, S.; Sidoli, S.; Verdun, R.E.; Zhang, Y.; Wang, Z.; et al. The Polycomb Protein RING1B Enables Estrogen-Mediated Gene Expression by Promoting Enhancer-Promoter Interaction and R-Loop Formation. Nucleic Acids Res. 2021, 49, 9768–9782. [Google Scholar] [CrossRef]
  128. Sanchez, A.; de Vivo, A.; Tonzi, P.; Kim, J.; Huang, T.T.; Kee, Y. Transcription-Replication Conflicts as a Source of Common Fragile Site Instability Caused by BMI1-RNF2 Deficiency. PLoS Genet. 2020, 16, e1008524. [Google Scholar] [CrossRef]
  129. Feng, S.; Manley, J.L. Replication Protein A Associates with Nucleolar R Loops and Regulates rRNA Transcription and Nucleolar Morphology. Genes Dev. 2021, 35, 1579–1594. [Google Scholar] [CrossRef]
  130. Davletgildeeva, A.T.; Kuznetsov, N.A. The Role of DNMT Methyltransferases and TET Dioxygenases in the Maintenance of the DNA Methylation Level. Biomolecules 2024, 14, 1117. [Google Scholar] [CrossRef]
  131. Ito, S.; Shen, L.; Dai, Q.; Wu, S.C.; Collins, L.B.; Swenberg, J.A.; He, C.; Zhang, Y. Tet Proteins Can Convert 5-Methylcytosine to 5-Formylcytosine and 5-Carboxylcytosine. Science 2011, 333, 1300–1303. [Google Scholar] [CrossRef]
  132. Pastor, W.A.; Aravind, L.; Rao, A. TETonic Shift: Biological Roles of TET Proteins in DNA Demethylation and Transcription. Nat. Rev. Mol. Cell Biol. 2013, 14, 341–356. [Google Scholar] [CrossRef]
  133. Hirasawa, R.; Chiba, H.; Kaneda, M.; Tajima, S.; Li, E.; Jaenisch, R.; Sasaki, H. Maternal and Zygotic Dnmt1 Are Necessary and Sufficient for the Maintenance of DNA Methylation Imprints during Preimplantation Development. Genes Dev. 2008, 22, 1607–1616. [Google Scholar] [CrossRef]
  134. Kaneda, M.; Okano, M.; Hata, K.; Sado, T.; Tsujimoto, N.; Li, E.; Sasaki, H. Essential Role for de Novo DNA Methyltransferase Dnmt3a in Paternal and Maternal Imprinting. Nature 2004, 429, 900–903. [Google Scholar] [CrossRef]
  135. Okano, M.; Bell, D.W.; Haber, D.A.; Li, E. DNA Methyltransferases Dnmt3a and Dnmt3b Are Essential for De Novo Methylation and Mammalian Development. Cell 1999, 99, 247–257. [Google Scholar] [CrossRef]
  136. Shih, H.-T.; Chen, W.-Y.; Wang, H.-Y.; Chao, T.; Huang, H.-D.; Chou, C.-H.; Chang, Z.-F. DNMT3b Protects Centromere Integrity by Restricting R-Loop-Mediated DNA Damage. Cell Death Dis. 2022, 13, 546. [Google Scholar] [CrossRef]
  137. Luna, R.; Gómez-González, B.; Aguilera, A. RNA Biogenesis and RNA Metabolism Factors as R-Loop Suppressors: A Hidden Role in Genome Integrity. Genes Dev. 2024, 38, 504–527. [Google Scholar] [CrossRef]
  138. Li, Y.; Liu, C.; Jia, X.; Bi, L.; Ren, Z.; Zhao, Y.; Zhang, X.; Guo, L.; Bao, Y.; Liu, C.; et al. RPA Transforms RNase H1 to a Bidirectional Exoribonuclease for Processive RNA-DNA Hybrid Cleavage. Nat. Commun. 2024, 15, 7464. [Google Scholar] [CrossRef]
  139. Nguyen, H.D.; Yadav, T.; Giri, S.; Saez, B.; Graubert, T.A.; Zou, L. Functions of Replication Protein A as a Sensor of R Loops and a Regulator of RNaseH1. Mol. Cell 2017, 65, 832–847.e4. [Google Scholar] [CrossRef]
  140. Pan, J.M.; Betts, H.; Cubbon, A.; He, L.; Bolt, E.L.; Soultanas, P. The Human HELQ Helicase and XRN2 Exoribonuclease Cooperate in R-Loop Resolution. Open Biol. 2025, 15, 240112. [Google Scholar] [CrossRef]
  141. Pan, H.; Jin, M.; Ghadiyaram, A.; Kaur, P.; Miller, H.E.; Ta, H.M.; Liu, M.; Fan, Y.; Mahn, C.; Gorthi, A.; et al. Cohesin SA1 and SA2 Are RNA Binding Proteins That Localize to RNA Containing Regions on DNA. Nucleic Acids Res. 2020, 48, 5639–5655. [Google Scholar] [CrossRef]
  142. Bayona-Feliu, A.; Herrera-Moyano, E.; Badra-Fajardo, N.; Galván-Femenía, I.; Soler-Oliva, M.E.; Aguilera, A. The Chromatin Network Helps Prevent Cancer-Associated Mutagenesis at Transcription-Replication Conflicts. Nat. Commun. 2023, 14, 6890. [Google Scholar] [CrossRef]
  143. Ahmad, M.; Xu, D.; Wang, W. Type IA Topoisomerases Can Be “Magicians” for Both DNA and RNA in All Domains of Life. RNA Biol. 2017, 14, 854–864. [Google Scholar] [CrossRef]
  144. El Hage, A.; French, S.L.; Beyer, A.L.; Tollervey, D. Loss of Topoisomerase I Leads to R-Loop-Mediated Transcriptional Blocks during Ribosomal RNA Synthesis. Genes Dev. 2010, 24, 1546–1558. [Google Scholar] [CrossRef]
  145. Yang, Y.; McBride, K.M.; Hensley, S.; Lu, Y.; Chedin, F.; Bedford, M.T. Arginine Methylation Facilitates the Recruitment of TOP3B to Chromatin to Prevent R Loop Accumulation. Mol. Cell 2014, 53, 484–497. [Google Scholar] [CrossRef]
  146. Zhang, T.; Wallis, M.; Petrovic, V.; Challis, J.; Kalitsis, P.; Hudson, D.F. Loss of TOP3B Leads to Increased R-Loop Formation and Genome Instability. Open Biol. 2019, 9, 190222. [Google Scholar] [CrossRef]
  147. Andrs, M.; Hasanova, Z.; Oravetzova, A.; Dobrovolna, J.; Janscak, P. RECQ5: A Mysterious Helicase at the Interface of DNA Replication and Transcription. Genes 2020, 11, 232. [Google Scholar] [CrossRef]
  148. Di Marco, S.; Hasanova, Z.; Kanagaraj, R.; Chappidi, N.; Altmannova, V.; Menon, S.; Sedlackova, H.; Langhoff, J.; Surendranath, K.; Hühn, D.; et al. RECQ5 Helicase Cooperates with MUS81 Endonuclease in Processing Stalled Replication Forks at Common Fragile Sites during Mitosis. Mol. Cell 2017, 66, 658–671.e8. [Google Scholar] [CrossRef]
  149. Hamadeh, Z.; Lansdorp, P. RECQL5 at the Intersection of Replication and Transcription. Front. Cell Dev. Biol. 2020, 8, 324. [Google Scholar] [CrossRef]
  150. Urban, V.; Dobrovolna, J.; Hühn, D.; Fryzelkova, J.; Bartek, J.; Janscak, P. RECQ5 Helicase Promotes Resolution of Conflicts between Replication and Transcription in Human Cells. J. Cell. Biol. 2016, 214, 401–415. [Google Scholar] [CrossRef]
  151. Šviković, S.; Crisp, A.; Tan-Wong, S.M.; Guilliam, T.A.; Doherty, A.J.; Proudfoot, N.J.; Guilbaud, G.; Sale, J.E. R-Loop Formation during S Phase Is Restricted by PrimPol-Mediated Repriming. EMBO J. 2019, 38, e99793. [Google Scholar] [CrossRef]
  152. Morales, J.C.; Richard, P.; Patidar, P.L.; Motea, E.A.; Dang, T.T.; Manley, J.L.; Boothman, D.A. XRN2 Links Transcription Termination to DNA Damage and Replication Stress. PLoS Genet. 2016, 12, e1006107. [Google Scholar] [CrossRef]
  153. Cornelio, D.A.; Sedam, H.N.C.; Ferrarezi, J.A.; Sampaio, N.M.V.; Argueso, J.L. Both R-Loop Removal and Ribonucleotide Excision Repair Activities of RNase H2 Contribute Substantially to Chromosome Stability. DNA Repair 2017, 52, 110–114. [Google Scholar] [CrossRef]
  154. Cristini, A.; Tellier, M.; Constantinescu, F.; Accalai, C.; Albulescu, L.O.; Heiringhoff, R.; Bery, N.; Sordet, O.; Murphy, S.; Gromak, N. RNase H2, Mutated in Aicardi-Goutières Syndrome, Resolves Co-Transcriptional R-Loops to Prevent DNA Breaks and Inflammation. Nat. Commun. 2022, 13, 2961. [Google Scholar] [CrossRef]
  155. Hiller, B.; Achleitner, M.; Glage, S.; Naumann, R.; Behrendt, R.; Roers, A. Mammalian RNase H2 Removes Ribonucleotides from DNA to Maintain Genome Integrity. J. Exp. Med. 2012, 209, 1419–1426. [Google Scholar] [CrossRef]
  156. Camino, L.P.; Dutta, A.; Barroso, S.; Pérez-Calero, C.; Katz, J.N.; García-Rubio, M.; Sung, P.; Gómez-González, B.; Aguilera, A. DICER Ribonuclease Removes Harmful R-Loops. Mol. Cell 2023, 83, 3707–3719.e5. [Google Scholar] [CrossRef]
  157. Lee, Y.J.; Lee, S.Y.; Kim, S.; Kim, S.-H.; Lee, S.H.; Park, S.; Kim, J.J.; Kim, D.-W.; Kim, H. REXO5 Promotes Genomic Integrity through Regulating R-Loop Using Its Exonuclease Activity. Leukemia 2024, 38, 2150–2161. [Google Scholar] [CrossRef]
  158. Laverde, E.E.; Polyzos, A.A.; Tsegay, P.P.; Shaver, M.; Hutcheson, J.D.; Balakrishnan, L.; McMurray, C.T.; Liu, Y. Flap Endonuclease 1 Endonucleolytically Processes RNA to Resolve R-Loops through DNA Base Excision Repair. Genes 2022, 14, 98. [Google Scholar] [CrossRef]
  159. Laverde, E.E.; Lai, Y.; Leng, F.; Balakrishnan, L.; Freudenreich, C.H.; Liu, Y. R-Loops Promote Trinucleotide Repeat Deletion through DNA Base Excision Repair Enzymatic Activities. J. Biol. Chem. 2020, 295, 13902–13913. [Google Scholar] [CrossRef]
  160. Teasley, D.C.; Parajuli, S.; Nguyen, M.; Moore, H.R.; Alspach, E.; Lock, Y.J.; Honaker, Y.; Saharia, A.; Piwnica-Worms, H.; Stewart, S.A. Flap Endonuclease 1 Limits Telomere Fragility on the Leading Strand. J. Biol. Chem. 2015, 290, 15133–15145. [Google Scholar] [CrossRef]
  161. Osmundson, J.S.; Kumar, J.; Yeung, R.; Smith, D.J. Pif1-Family Helicases Cooperatively Suppress Widespread Replication-Fork Arrest at tRNA Genes. Nat. Struct. Mol. Biol. 2017, 24, 162–170. [Google Scholar] [CrossRef] [PubMed]
  162. Tran, P.L.T.; Pohl, T.J.; Chen, C.-F.; Chan, A.; Pott, S.; Zakian, V.A. PIF1 Family DNA Helicases Suppress R-Loop Mediated Genome Instability at tRNA Genes. Nat. Commun. 2017, 8, 15025. [Google Scholar] [CrossRef] [PubMed]
  163. Balajee, A.S.; Machwe, A.; May, A.; Gray, M.D.; Oshima, J.; Martin, G.M.; Nehlin, J.O.; Brosh, R.; Orren, D.K.; Bohr, V.A. The Werner Syndrome Protein Is Involved in RNA Polymerase II Transcription. Mol. Biol. Cell. 1999, 10, 2655–2668. [Google Scholar] [CrossRef] [PubMed]
  164. Chang, E.Y.-C.; Novoa, C.A.; Aristizabal, M.J.; Coulombe, Y.; Segovia, R.; Chaturvedi, R.; Shen, Y.; Keong, C.; Tam, A.S.; Jones, S.J.M.; et al. RECQ-like Helicases Sgs1 and BLM Regulate R-Loop-Associated Genome Instability. J. Cell. Biol. 2017, 216, 3991–4005. [Google Scholar] [CrossRef]
  165. Gray, M.D.; Wang, L.; Youssoufian, H.; Martin, G.M.; Oshima, J. Werner Helicase Is Localized to Transcriptionally Active Nucleoli of Cycling Cells. Exp. Cell Res. 1998, 242, 487–494. [Google Scholar] [CrossRef]
  166. Marabitti, V.; Lillo, G.; Malacaria, E.; Palermo, V.; Sanchez, M.; Pichierri, P.; Franchitto, A. ATM Pathway Activation Limits R-Loop-Associated Genomic Instability in Werner Syndrome Cells. Nucleic Acids Res. 2019, 47, 3485–3502. [Google Scholar] [CrossRef]
  167. Shiratori, M.; Suzuki, T.; Itoh, C.; Goto, M.; Furuichi, Y.; Matsumoto, T. WRN Helicase Accelerates the Transcription of Ribosomal RNA as a Component of an RNA Polymerase I-Associated Complex. Oncogene 2002, 21, 2447–2454. [Google Scholar] [CrossRef]
  168. Hatchi, E.; Skourti-Stathaki, K.; Ventz, S.; Pinello, L.; Yen, A.; Kamieniarz-Gdula, K.; Dimitrov, S.; Pathania, S.; McKinney, K.M.; Eaton, M.L.; et al. BRCA1 Recruitment to Transcriptional Pause Sites Is Required for R-Loop-Driven DNA Damage Repair. Mol. Cell 2015, 57, 636–647. [Google Scholar] [CrossRef]
  169. Kanagaraj, R.; Mitter, R.; Kantidakis, T.; Edwards, M.M.; Benitez, A.; Chakravarty, P.; Fu, B.; Becherel, O.; Yang, F.; Lavin, M.F.; et al. Integrated Genome and Transcriptome Analyses Reveal the Mechanism of Genome Instability in Ataxia with Oculomotor Apraxia 2. Proc. Natl. Acad. Sci. USA 2022, 119, e2114314119. [Google Scholar] [CrossRef]
  170. Rao, S.; Andrs, M.; Shukla, K.; Isik, E.; König, C.; Schneider, S.; Bauer, M.; Rosano, V.; Prokes, J.; Müller, A.; et al. Senataxin RNA/DNA Helicase Promotes Replication Restart at Co-Transcriptional R-Loops to Prevent MUS81-Dependent Fork Degradation. Nucleic Acids Res. 2024, 52, 10355–10369. [Google Scholar] [CrossRef]
  171. Boleslavska, B.; Oravetzova, A.; Shukla, K.; Nascakova, Z.; Ibini, O.N.; Hasanova, Z.; Andrs, M.; Kanagaraj, R.; Dobrovolna, J.; Janscak, P. DDX17 Helicase Promotes Resolution of R-Loop-Mediated Transcription-Replication Conflicts in Human Cells. Nucleic Acids Res. 2022, 50, 12274–12290. [Google Scholar] [CrossRef] [PubMed]
  172. de Amorim, J.L.; Leung, S.W.; Haji-Seyed-Javadi, R.; Hou, Y.; Yu, D.S.; Ghalei, H.; Khoshnevis, S.; Yao, B.; Corbett, A.H. The Putative RNA Helicase DDX1 Associates with the Nuclear RNA Exosome and Modulates RNA/DNA Hybrids (R-Loops). J. Biol. Chem. 2024, 300, 105646. [Google Scholar] [CrossRef] [PubMed]
  173. Hernández-Reyes, Y.; Fonseca-Rodríguez, C.; Freire, R.; Smits, V.A.J. DDX37 and DDX50 Maintain Genome Stability by Preventing Transcription-Dependent R-Loop Formation. J. Mol. Biol. 2025, 437, 169061. [Google Scholar] [CrossRef] [PubMed]
  174. Hodroj, D.; Recolin, B.; Serhal, K.; Martinez, S.; Tsanov, N.; Abou Merhi, R.; Maiorano, D. An ATR-Dependent Function for the Ddx19 RNA Helicase in Nuclear R-Loop Metabolism. EMBO J. 2017, 36, 1182–1198. [Google Scholar] [CrossRef]
  175. Kim, S.; Kang, N.; Park, S.H.; Wells, J.; Hwang, T.; Ryu, E.; Kim, B.-G.; Hwang, S.; Kim, S.-J.; Kang, S.; et al. ATAD5 Restricts R-Loop Formation through PCNA Unloading and RNA Helicase Maintenance at the Replication Fork. Nucleic Acids Res. 2020, 48, 7218–7238. [Google Scholar] [CrossRef]
  176. Li, L.; Germain, D.R.; Poon, H.-Y.; Hildebrandt, M.R.; Monckton, E.A.; McDonald, D.; Hendzel, M.J.; Godbout, R. DEAD Box 1 Facilitates Removal of RNA and Homologous Recombination at DNA Double-Strand Breaks. Mol. Cell. Biol. 2016, 36, 2794–2810. [Google Scholar] [CrossRef]
  177. Mersaoui, S.Y.; Yu, Z.; Coulombe, Y.; Karam, M.; Busatto, F.F.; Masson, J.-Y.; Richard, S. Arginine Methylation of the DDX5 Helicase RGG/RG Motif by PRMT5 Regulates Resolution of RNA: DNA Hybrids. EMBO J. 2019, 38, e100986. [Google Scholar] [CrossRef]
  178. Song, C.; Hotz-Wagenblatt, A.; Voit, R.; Grummt, I. SIRT7 and the DEAD-Box Helicase DDX21 Cooperate to Resolve Genomic R Loops and Safeguard Genome Stability. Genes Dev. 2017, 31, 1370–1381. [Google Scholar] [CrossRef]
  179. Chimnaronk, S.; Suzuki, T.; Manita, T.; Ikeuchi, Y.; Yao, M.; Suzuki, T.; Tanaka, I. RNA Helicase Module in an Acetyltransferase That Modifies a Specific tRNA Anticodon. EMBO J. 2009, 28, 1362–1373. [Google Scholar] [CrossRef]
  180. Ito, S.; Horikawa, S.; Suzuki, T.; Kawauchi, H.; Tanaka, Y.; Suzuki, T.; Suzuki, T. Human NAT10 Is an ATP-Dependent RNA Acetyltransferase Responsible for N4-Acetylcytidine Formation in 18 S Ribosomal RNA (rRNA). J. Biol. Chem. 2014, 289, 35724–35730. [Google Scholar] [CrossRef]
  181. Kong, R.; Zhang, L.; Hu, L.; Peng, Q.; Han, W.; Du, X.; Ke, Y. hALP, a Novel Transcriptional U Three Protein (t-UTP), Activates RNA Polymerase I Transcription by Binding and Acetylating the Upstream Binding Factor (UBF). J. Biol. Chem. 2011, 286, 7139–7148. [Google Scholar] [CrossRef]
  182. Su, K.; Zhao, Z.; Wang, Y.; Sun, S.; Liu, X.; Zhang, C.; Jiang, Y.; Du, X. NAT10 Resolves Harmful Nucleolar R-Loops Depending on Its Helicase Domain and Acetylation of DDX21. Cell Commun. Signal. 2024, 22, 490. [Google Scholar] [CrossRef] [PubMed]
  183. Bao, S.; Tibbetts, R.S.; Brumbaugh, K.M.; Fang, Y.; Richardson, D.A.; Ali, A.; Chen, S.M.; Abraham, R.T.; Wang, X.F. ATR/ATM-Mediated Phosphorylation of Human Rad17 Is Required for Genotoxic Stress Responses. Nature 2001, 411, 969–974. [Google Scholar] [CrossRef] [PubMed]
  184. Hammond, E.M.; Denko, N.C.; Dorie, M.J.; Abraham, R.T.; Giaccia, A.J. Hypoxia Links ATR and P53 through Replication Arrest. Mol. Cell. Biol. 2002, 22, 1834–1843. [Google Scholar] [CrossRef] [PubMed]
  185. Kumar, A.; Mazzanti, M.; Mistrik, M.; Kosar, M.; Beznoussenko, G.V.; Mironov, A.A.; Garrè, M.; Parazzoli, D.; Shivashankar, G.V.; Scita, G.; et al. ATR Mediates a Checkpoint at the Nuclear Envelope in Response to Mechanical Stress. Cell 2014, 158, 633–646. [Google Scholar] [CrossRef]
  186. Liu, S.; Shiotani, B.; Lahiri, M.; Maréchal, A.; Tse, A.; Leung, C.C.Y.; Glover, J.N.M.; Yang, X.H.; Zou, L. ATR Autophosphorylation as a Molecular Switch for Checkpoint Activation. Mol. Cell 2011, 43, 192–202. [Google Scholar] [CrossRef]
  187. Peterson, S.E.; Li, Y.; Wu-Baer, F.; Chait, B.T.; Baer, R.; Yan, H.; Gottesman, M.E.; Gautier, J. Activation of DSB Processing Requires Phosphorylation of CtIP by ATR. Mol. Cell 2013, 49, 657–667. [Google Scholar] [CrossRef]
  188. Tibbetts, R.S.; Cortez, D.; Brumbaugh, K.M.; Scully, R.; Livingston, D.; Elledge, S.J.; Abraham, R.T. Functional Interactions between BRCA1 and the Checkpoint Kinase ATR during Genotoxic Stress. Genes Dev. 2000, 14, 2989–3002. [Google Scholar] [CrossRef]
  189. Tibbetts, R.S.; Brumbaugh, K.M.; Williams, J.M.; Sarkaria, J.N.; Cliby, W.A.; Shieh, S.Y.; Taya, Y.; Prives, C.; Abraham, R.T. A Role for ATR in the DNA Damage-Induced Phosphorylation of P53. Genes Dev. 1999, 13, 152–157. [Google Scholar] [CrossRef]
  190. Chang, E.Y.-C.; Tsai, S.; Aristizabal, M.J.; Wells, J.P.; Coulombe, Y.; Busatto, F.F.; Chan, Y.A.; Kumar, A.; Dan Zhu, Y.; Wang, A.Y.-H.; et al. MRE11-RAD50-NBS1 Promotes Fanconi Anemia R-Loop Suppression at Transcription-Replication Conflicts. Nat. Commun. 2019, 10, 4265. [Google Scholar] [CrossRef]
  191. He, Y.J.; Meghani, K.; Caron, M.-C.; Yang, C.; Ronato, D.A.; Bian, J.; Sharma, A.; Moore, J.; Niraj, J.; Detappe, A.; et al. DYNLL1 Binds to MRE11 to Limit DNA End Resection in BRCA1-Deficient Cells. Nature 2018, 563, 522–526. [Google Scholar] [CrossRef]
  192. Paull, T.T. 20 Years of Mre11 Biology: No End in Sight. Mol. Cell 2018, 71, 419–427. [Google Scholar] [CrossRef]
  193. Shibata, A.; Moiani, D.; Arvai, A.S.; Perry, J.; Harding, S.M.; Genois, M.-M.; Maity, R.; van Rossum-Fikkert, S.; Kertokalio, A.; Romoli, F.; et al. DNA Double-Strand Break Repair Pathway Choice Is Directed by Distinct MRE11 Nuclease Activities. Mol. Cell 2014, 53, 7–18. [Google Scholar] [CrossRef]
  194. Groh, M.; Albulescu, L.O.; Cristini, A.; Gromak, N. Senataxin: Genome Guardian at the Interface of Transcription and Neurodegeneration. J. Mol. Biol. 2017, 429, 3181–3195. [Google Scholar] [CrossRef]
  195. Herold, S.; Kalb, J.; Büchel, G.; Ade, C.P.; Baluapuri, A.; Xu, J.; Koster, J.; Solvie, D.; Carstensen, A.; Klotz, C.; et al. Recruitment of BRCA1 Limits MYCN-Driven Accumulation of Stalled RNA Polymerase. Nature 2019, 567, 545–549. [Google Scholar] [CrossRef]
  196. Hill, S.J.; Rolland, T.; Adelmant, G.; Xia, X.; Owen, M.S.; Dricot, A.; Zack, T.I.; Sahni, N.; Jacob, Y.; Hao, T.; et al. Systematic Screening Reveals a Role for BRCA1 in the Response to Transcription-Associated DNA Damage. Genes Dev. 2014, 28, 1957–1975. [Google Scholar] [CrossRef]
  197. Martin-Tumasz, S.; Brow, D.A. Saccharomyces Cerevisiae Sen1 Helicase Domain Exhibits 5’- to 3’-Helicase Activity with a Preference for Translocation on DNA Rather than RNA. J. Biol. Chem. 2015, 290, 22880–22889. [Google Scholar] [CrossRef] [PubMed]
  198. Monteiro, A.N.; August, A.; Hanafusa, H. Evidence for a Transcriptional Activation Function of BRCA1 C-Terminal Region. Proc. Natl. Acad. Sci. USA 1996, 93, 13595–13599. [Google Scholar] [CrossRef] [PubMed]
  199. San Martin Alonso, M.; Noordermeer, S.M. Untangling the Crosstalk between BRCA1 and R-Loops during DNA Repair. Nucleic Acids Res. 2021, 49, 4848–4863. [Google Scholar] [CrossRef]
  200. Scully, R.; Anderson, S.F.; Chao, D.M.; Wei, W.; Ye, L.; Young, R.A.; Livingston, D.M.; Parvin, J.D. BRCA1 Is a Component of the RNA Polymerase II Holoenzyme. Proc. Natl. Acad. Sci. USA 1997, 94, 5605–5610. [Google Scholar] [CrossRef]
  201. Bhatia, V.; Barroso, S.I.; García-Rubio, M.L.; Tumini, E.; Herrera-Moyano, E.; Aguilera, A. BRCA2 Prevents R-Loop Accumulation and Associates with TREX-2 mRNA Export Factor PCID2. Nature 2014, 511, 362–365. [Google Scholar] [CrossRef]
  202. D’Alessandro, G.; Whelan, D.R.; Howard, S.M.; Vitelli, V.; Renaudin, X.; Adamowicz, M.; Iannelli, F.; Jones-Weinert, C.W.; Lee, M.; Matti, V.; et al. BRCA2 Controls DNA:RNA Hybrid Level at DSBs by Mediating RNase H2 Recruitment. Nat. Commun. 2018, 9, 5376. [Google Scholar] [CrossRef] [PubMed]
  203. Shivji, M.K.K.; Renaudin, X.; Williams, Ç.H.; Venkitaraman, A.R. BRCA2 Regulates Transcription Elongation by RNA Polymerase II to Prevent R-Loop Accumulation. Cell Rep. 2018, 22, 1031–1039. [Google Scholar] [CrossRef] [PubMed]
  204. Amon, J.D.; Koshland, D. RNase H Enables Efficient Repair of R-Loop Induced DNA Damage. Elife 2016, 5, e20533. [Google Scholar] [CrossRef] [PubMed]
  205. Jalan, M.; Sharma, A.; Pei, X.; Weinhold, N.; Buechelmaier, E.S.; Zhu, Y.; Ahmed-Seghir, S.; Ratnakumar, A.; Di Bona, M.; McDermott, N.; et al. RAD52 Resolves Transcription-Replication Conflicts to Mitigate R-Loop Induced Genome Instability. Nat. Commun. 2024, 15, 7776. [Google Scholar] [CrossRef]
  206. Laspata, N.; Kaur, P.; Mersaoui, S.Y.; Muoio, D.; Liu, Z.S.; Bannister, M.H.; Nguyen, H.D.; Curry, C.; Pascal, J.M.; Poirier, G.G.; et al. PARP1 Associates with R-Loops to Promote Their Resolution and Genome Stability. Nucleic Acids Res. 2023, 51, 2215–2237. [Google Scholar] [CrossRef]
  207. Abakir, A.; Giles, T.C.; Cristini, A.; Foster, J.M.; Dai, N.; Starczak, M.; Rubio-Roldan, A.; Li, M.; Eleftheriou, M.; Crutchley, J.; et al. N6-Methyladenosine Regulates the Stability of RNA:DNA Hybrids in Human Cells. Nat. Genet. 2020, 52, 48–55. [Google Scholar] [CrossRef]
  208. Hao, J.-D.; Liu, Q.-L.; Liu, M.-X.; Yang, X.; Wang, L.-M.; Su, S.-Y.; Xiao, W.; Zhang, M.-Q.; Zhang, Y.-C.; Zhang, L.; et al. DDX21 Mediates Co-Transcriptional RNA m6A Modification to Promote Transcription Termination and Genome Stability. Mol. Cell 2024, 84, 1711–1726.e11. [Google Scholar] [CrossRef]
  209. Horiuchi, K.; Kawamura, T.; Iwanari, H.; Ohashi, R.; Naito, M.; Kodama, T.; Hamakubo, T. Identification of Wilms’ Tumor 1-Associating Protein Complex and Its Role in Alternative Splicing and the Cell Cycle. J. Biol. Chem. 2013, 288, 33292–33302. [Google Scholar] [CrossRef]
  210. Kang, H.J.; Cheon, N.Y.; Park, H.; Jeong, G.W.; Ye, B.J.; Yoo, E.J.; Lee, J.H.; Hur, J.-H.; Lee, E.-A.; Kim, H.; et al. TonEBP Recognizes R-Loops and Initiates m6A RNA Methylation for R-Loop Resolution. Nucleic Acids Res. 2021, 49, 269–284. [Google Scholar] [CrossRef]
  211. Liu, J.; Yue, Y.; Han, D.; Wang, X.; Fu, Y.; Zhang, L.; Jia, G.; Yu, M.; Lu, Z.; Deng, X.; et al. A METTL3-METTL14 Complex Mediates Mammalian Nuclear RNA N6-Adenosine Methylation. Nat. Chem. Biol. 2014, 10, 93–95. [Google Scholar] [CrossRef] [PubMed]
  212. Ping, X.-L.; Sun, B.-F.; Wang, L.; Xiao, W.; Yang, X.; Wang, W.-J.; Adhikari, S.; Shi, Y.; Lv, Y.; Chen, Y.-S.; et al. Mammalian WTAP Is a Regulatory Subunit of the RNA N6-Methyladenosine Methyltransferase. Cell Res. 2014, 24, 177–189. [Google Scholar] [CrossRef] [PubMed]
  213. Leela, J.K.; Syeda, A.H.; Anupama, K.; Gowrishankar, J. Rho-Dependent Transcription Termination Is Essential to Prevent Excessive Genome-Wide R-Loops in Escherichia Coli. Proc. Natl. Acad. Sci. USA 2013, 110, 258–263. [Google Scholar] [CrossRef] [PubMed]
  214. Li, X.; Manley, J.L. Inactivation of the SR Protein Splicing Factor ASF/SF2 Results in Genomic Instability. Cell 2005, 122, 365–378. [Google Scholar] [CrossRef]
  215. García-Pichardo, D.; Cañas, J.C.; García-Rubio, M.L.; Gómez-González, B.; Rondón, A.G.; Aguilera, A. Histone Mutants Separate R Loop Formation from Genome Instability Induction. Mol. Cell 2017, 66, 597–609.e5. [Google Scholar] [CrossRef]
  216. Berger, S.L. The Complex Language of Chromatin Regulation during Transcription. Nature 2007, 447, 407–412. [Google Scholar] [CrossRef]
  217. Kubicek, S.; O’Sullivan, R.J.; August, E.M.; Hickey, E.R.; Zhang, Q.; Teodoro, M.L.; Rea, S.; Mechtler, K.; Kowalski, J.A.; Homon, C.A.; et al. Reversal of H3K9me2 by a Small-Molecule Inhibitor for the G9a Histone Methyltransferase. Mol. Cell 2007, 25, 473–481. [Google Scholar] [CrossRef]
  218. Zhou, H.; Li, L.; Wang, Q.; Hu, Y.; Zhao, W.; Gautam, M.; Li, L. H3K9 Demethylation-Induced R-Loop Accumulation Is Linked to Disorganized Nucleoli. Front. Genet. 2020, 11, 43. [Google Scholar] [CrossRef]
  219. Hao, S.; Wang, Y.; Zhao, Y.; Gao, W.; Cui, W.; Li, Y.; Cui, J.; Liu, Y.; Lin, L.; Xu, X.; et al. Dynamic Switching of Crotonylation to Ubiquitination of H2A at Lysine 119 Attenuates Transcription-Replication Conflicts Caused by Replication Stress. Nucleic Acids Res. 2022, 50, 9873–9892. [Google Scholar] [CrossRef]
  220. Wang, H.; Wang, L.; Erdjument-Bromage, H.; Vidal, M.; Tempst, P.; Jones, R.S.; Zhang, Y. Role of Histone H2A Ubiquitination in Polycomb Silencing. Nature 2004, 431, 873–878. [Google Scholar] [CrossRef]
  221. Bayona-Feliu, A.; Casas-Lamesa, A.; Reina, O.; Bernués, J.; Azorín, F. Linker Histone H1 Prevents R-Loop Accumulation and Genome Instability in Heterochromatin. Nat. Commun. 2017, 8, 283. [Google Scholar] [CrossRef]
  222. Zeller, P.; Padeken, J.; van Schendel, R.; Kalck, V.; Tijsterman, M.; Gasser, S.M. Histone H3K9 Methylation Is Dispensable for Caenorhabditis Elegans Development but Suppresses RNA:DNA Hybrid-Associated Repeat Instability. Nat. Genet. 2016, 48, 1385–1395. [Google Scholar] [CrossRef]
  223. Phoenix, P.; Raymond, M.A.; Massé, E.; Drolet, M. Roles of DNA Topoisomerases in the Regulation of R-Loop Formation in Vitro. J. Biol. Chem. 1997, 272, 1473–1479. [Google Scholar] [CrossRef]
  224. Promonet, A.; Padioleau, I.; Liu, Y.; Sanz, L.; Biernacka, A.; Schmitz, A.-L.; Skrzypczak, M.; Sarrazin, A.; Mettling, C.; Rowicka, M.; et al. Topoisomerase 1 Prevents Replication Stress at R-Loop-Enriched Transcription Termination Sites. Nat. Commun. 2020, 11, 3940. [Google Scholar] [CrossRef] [PubMed]
  225. Saha, S.; Yang, X.; Huang, S.-Y.N.; Agama, K.; Baechler, S.A.; Sun, Y.; Zhang, H.; Saha, L.K.; Su, S.; Jenkins, L.M.; et al. Resolution of R-Loops by Topoisomerase III-β (TOP3B) in Coordination with the DEAD-Box Helicase DDX5. Cell Rep. 2022, 40, 111067. [Google Scholar] [CrossRef] [PubMed]
  226. Mirkin, E.V.; Mirkin, S.M. Replication Fork Stalling at Natural Impediments. Microbiol. Mol. Biol. Rev. 2007, 71, 13–35. [Google Scholar] [CrossRef] [PubMed]
  227. Campuzano, V.; Montermini, L.; Moltò, M.D.; Pianese, L.; Cossée, M.; Cavalcanti, F.; Monros, E.; Rodius, F.; Duclos, F.; Monticelli, A.; et al. Friedreich’s Ataxia: Autosomal Recessive Disease Caused by an Intronic GAA Triplet Repeat Expansion. Science 1996, 271, 1423–1427. [Google Scholar] [CrossRef]
  228. Frank-Kamenetskii, M.D.; Mirkin, S.M. Triplex DNA Structures. Annu. Rev. Biochem. 1995, 64, 65–95. [Google Scholar] [CrossRef]
  229. Groh, M.; Lufino, M.M.P.; Wade-Martins, R.; Gromak, N. R-Loops Associated with Triplet Repeat Expansions Promote Gene Silencing in Friedreich Ataxia and Fragile X Syndrome. PLoS Genet. 2014, 10, e1004318. [Google Scholar] [CrossRef]
  230. Clark, R.M.; Bhaskar, S.S.; Miyahara, M.; Dalgliesh, G.L.; Bidichandani, S.I. Expansion of GAA Trinucleotide Repeats in Mammals. Genomics 2006, 87, 57–67. [Google Scholar] [CrossRef]
  231. Willems, T.; Gymrek, M.; Highnam, G.; 1000 Genomes Project Consortium; Mittelman, D.; Erlich, Y. The Landscape of Human STR Variation. Genome Res. 2014, 24, 1894–1904. [Google Scholar] [CrossRef]
  232. Bianchi, J.; Rudd, S.G.; Jozwiakowski, S.K.; Bailey, L.J.; Soura, V.; Taylor, E.; Stevanovic, I.; Green, A.J.; Stracker, T.H.; Lindsay, H.D.; et al. PrimPol Bypasses UV Photoproducts during Eukaryotic Chromosomal DNA Replication. Mol. Cell 2013, 52, 566–573. [Google Scholar] [CrossRef]
  233. García-Gómez, S.; Reyes, A.; Martínez-Jiménez, M.I.; Chocrón, E.S.; Mourón, S.; Terrados, G.; Powell, C.; Salido, E.; Méndez, J.; Holt, I.J.; et al. PrimPol, an Archaic Primase/Polymerase Operating in Human Cells. Mol. Cell 2013, 52, 541–553. [Google Scholar] [CrossRef] [PubMed]
  234. Wan, L.; Lou, J.; Xia, Y.; Su, B.; Liu, T.; Cui, J.; Sun, Y.; Lou, H.; Huang, J. hPrimpol1/CCDC111 Is a Human DNA Primase-Polymerase Required for the Maintenance of Genome Integrity. EMBO Rep. 2013, 14, 1104–1112. [Google Scholar] [CrossRef] [PubMed]
  235. Schiavone, D.; Jozwiakowski, S.K.; Romanello, M.; Guilbaud, G.; Guilliam, T.A.; Bailey, L.J.; Sale, J.E.; Doherty, A.J. PrimPol Is Required for Replicative Tolerance of G Quadruplexes in Vertebrate Cells. Mol. Cell 2016, 61, 161–169. [Google Scholar] [CrossRef] [PubMed]
  236. Costantino, L.; Koshland, D. The Yin and Yang of R-Loop Biology. Curr. Opin. Cell Biol. 2015, 34, 39–45. [Google Scholar] [CrossRef]
  237. Cerritelli, S.M.; Crouch, R.J. Ribonuclease H: The Enzymes in Eukaryotes. FEBS J. 2009, 276, 1494–1505. [Google Scholar] [CrossRef]
  238. Wahba, L.; Amon, J.D.; Koshland, D.; Vuica-Ross, M. RNase H and Multiple RNA Biogenesis Factors Cooperate to Prevent RNA:DNA Hybrids from Generating Genome Instability. Mol. Cell 2011, 44, 978–988. [Google Scholar] [CrossRef]
  239. Wolak, C.; Ma, H.J.; Soubry, N.; Sandler, S.J.; Reyes-Lamothe, R.; Keck, J.L. Interaction with Single-Stranded DNA-Binding Protein Localizes Ribonuclease HI to DNA Replication Forks and Facilitates R-Loop Removal. Mol. Microbiol. 2020, 114, 495–509. [Google Scholar] [CrossRef]
  240. Qiu, J.; Qian, Y.; Frank, P.; Wintersberger, U.; Shen, B. Saccharomyces Cerevisiae RNase H(35) Functions in RNA Primer Removal during Lagging-Strand DNA Synthesis, Most Efficiently in Cooperation with Rad27 Nuclease. Mol. Cell. Biol. 1999, 19, 8361–8371. [Google Scholar] [CrossRef]
  241. Holt, I.J. The Jekyll and Hyde Character of RNase H1 and Its Multiple Roles in Mitochondrial DNA Metabolism. DNA Repair 2019, 84, 102630. [Google Scholar] [CrossRef]
  242. Lima, W.F.; Murray, H.M.; Damle, S.S.; Hart, C.E.; Hung, G.; De Hoyos, C.L.; Liang, X.-H.; Crooke, S.T. Viable RNaseH1 Knockout Mice Show RNaseH1 Is Essential for R Loop Processing, Mitochondrial and Liver Function. Nucleic Acids Res. 2016, 44, 5299–5312. [Google Scholar] [CrossRef]
  243. Lockhart, A.; Pires, V.B.; Bento, F.; Kellner, V.; Luke-Glaser, S.; Yakoub, G.; Ulrich, H.D.; Luke, B. RNase H1 and H2 Are Differentially Regulated to Process RNA-DNA Hybrids. Cell Rep. 2019, 29, 2890–2900.e5. [Google Scholar] [CrossRef]
  244. Hyjek, M.; Figiel, M.; Nowotny, M. RNases H: Structure and Mechanism. DNA Repair 2019, 84, 102672. [Google Scholar] [CrossRef]
  245. Lang, K.S.; Hall, A.N.; Merrikh, C.N.; Ragheb, M.; Tabakh, H.; Pollock, A.J.; Woodward, J.J.; Dreifus, J.E.; Merrikh, H. Replication-Transcription Conflicts Generate R-Loops That Orchestrate Bacterial Stress Survival and Pathogenesis. Cell 2017, 170, 787–799.e18. [Google Scholar] [CrossRef]
  246. Xiang, Y.; Laurent, B.; Hsu, C.-H.; Nachtergaele, S.; Lu, Z.; Sheng, W.; Xu, C.; Chen, H.; Ouyang, J.; Wang, S.; et al. RNA m6A Methylation Regulates the Ultraviolet-Induced DNA Damage Response. Nature 2017, 543, 573–576. [Google Scholar] [CrossRef]
  247. Bernstein, E.; Caudy, A.A.; Hammond, S.M.; Hannon, G.J. Role for a Bidentate Ribonuclease in the Initiation Step of RNA Interference. Nature 2001, 409, 363–366. [Google Scholar] [CrossRef] [PubMed]
  248. Ciechanowska, K.; Pokornowska, M.; Kurzyńska-Kokorniak, A. Genetic Insight into the Domain Structure and Functions of Dicer-Type Ribonucleases. Int. J. Mol. Sci. 2021, 22, 616. [Google Scholar] [CrossRef] [PubMed]
  249. Kurzynska-Kokorniak, A.; Koralewska, N.; Pokornowska, M.; Urbanowicz, A.; Tworak, A.; Mickiewicz, A.; Figlerowicz, M. The Many Faces of Dicer: The Complexity of the Mechanisms Regulating Dicer Gene Expression and Enzyme Activities. Nucleic Acids Res. 2015, 43, 4365–4380. [Google Scholar] [CrossRef] [PubMed]
  250. Balakrishnan, L.; Bambara, R.A. Flap Endonuclease 1. Annu. Rev. Biochem. 2013, 82, 119–138. [Google Scholar] [CrossRef]
  251. Liu, Y.; Kao, H.-I.; Bambara, R.A. Flap Endonuclease 1: A Central Component of DNA Metabolism. Annu. Rev. Biochem. 2004, 73, 589–615. [Google Scholar] [CrossRef]
  252. Meng, F.; Li, T.; Singh, A.K.; Wang, Y.; Attiyeh, M.; Kohram, F.; Feng, Q.; Li, Y.R.; Shen, B.; Williams, T.; et al. Base-Excision Repair Pathway Regulates Transcription-Replication Conflicts in Pancreatic Ductal Adenocarcinoma. Cell Rep. 2024, 43, 114820. [Google Scholar] [CrossRef] [PubMed]
  253. Groslambert, J.; Prokhorova, E.; Ahel, I. ADP-Ribosylation of DNA and RNA. DNA Repair 2021, 105, 103144. [Google Scholar] [CrossRef] [PubMed]
  254. Matta, E.; Kiribayeva, A.; Khassenov, B.; Matkarimov, B.T.; Ishchenko, A.A. Insight into DNA Substrate Specificity of PARP1-Catalysed DNA Poly(ADP-Ribosyl)Ation. Sci. Rep. 2020, 10, 3699. [Google Scholar] [CrossRef]
  255. Ziegler, R.G.; Weinstein, S.J.; Fears, T.R. Nutritional and Genetic Inefficiencies in One-Carbon Metabolism and Cervical Cancer Risk. J. Nutr. 2002, 132, 2345S–2349S. [Google Scholar] [CrossRef]
  256. Pohl, T.J.; Zakian, V.A. Pif1 Family DNA Helicases: A Helpmate to RNase H? DNA Repair 2019, 84, 102633. [Google Scholar] [CrossRef] [PubMed]
  257. Fairman-Williams, M.E.; Guenther, U.-P.; Jankowsky, E. SF1 and SF2 Helicases: Family Matters. Curr. Opin. Struct. Biol. 2010, 20, 313–324. [Google Scholar] [CrossRef]
  258. Mischo, H.E.; Gómez-González, B.; Grzechnik, P.; Rondón, A.G.; Wei, W.; Steinmetz, L.; Aguilera, A.; Proudfoot, N.J. Yeast Sen1 Helicase Protects the Genome from Transcription-Associated Instability. Mol. Cell 2011, 41, 21–32. [Google Scholar] [CrossRef]
  259. Bochman, M.L.; Sabouri, N.; Zakian, V.A. Unwinding the Functions of the Pif1 Family Helicases. DNA Repair 2010, 9, 237–249. [Google Scholar] [CrossRef]
  260. Rivosecchi, J.; Larochelle, M.; Teste, C.; Grenier, F.; Malapert, A.; Ricci, E.P.; Bernard, P.; Bachand, F.; Vanoosthuyse, V. Senataxin Homologue Sen1 Is Required for Efficient Termination of RNA Polymerase III Transcription. EMBO J. 2019, 38, e101955. [Google Scholar] [CrossRef]
  261. Sberna, S.; Filipuzzi, M.; Bianchi, N.; Croci, O.; Fardella, F.; Soriani, C.; Rohban, S.; Carnevali, S.; Albertini, A.A.; Crosetto, N.; et al. Senataxin Prevents Replicative Stress Induced by the Myc Oncogene. Cell Death Dis. 2025, 16, 187. [Google Scholar] [CrossRef]
  262. Zhao, H.; Hartono, S.R.; de Vera, K.M.F.; Yu, Z.; Satchi, K.; Zhao, T.; Sciammas, R.; Sanz, L.; Chédin, F.; Barlow, J. Senataxin and RNase H2 Act Redundantly to Suppress Genome Instability during Class Switch Recombination. Elife 2022, 11, e78917. [Google Scholar] [CrossRef]
  263. Cobb, J.A.; Bjergbaek, L. RecQ Helicases: Lessons from Model Organisms. Nucleic Acids Res. 2006, 34, 4106–4114. [Google Scholar] [CrossRef] [PubMed]
  264. Croteau, D.L.; Popuri, V.; Opresko, P.L.; Bohr, V.A. Human RecQ Helicases in DNA Repair, Recombination, and Replication. Annu. Rev. Biochem. 2014, 83, 519–552. [Google Scholar] [CrossRef] [PubMed]
  265. Lu, H.; Davis, A.J. Human RecQ Helicases in DNA Double-Strand Break Repair. Front. Cell Dev. Biol. 2021, 9, 640755. [Google Scholar] [CrossRef] [PubMed]
  266. Budhathoki, J.B.; Maleki, P.; Roy, W.A.; Janscak, P.; Yodh, J.G.; Balci, H. A Comparative Study of G-Quadruplex Unfolding and DNA Reeling Activities of Human RECQ5 Helicase. Biophys J. 2016, 110, 2585–2596. [Google Scholar] [CrossRef]
  267. Popuri, V.; Bachrati, C.Z.; Muzzolini, L.; Mosedale, G.; Costantini, S.; Giacomini, E.; Hickson, I.D.; Vindigni, A. The Human RecQ Helicases, BLM and RECQ1, Display Distinct DNA Substrate Specificities. J. Biol. Chem. 2008, 283, 17766–17776. [Google Scholar] [CrossRef]
  268. Kang, M.-S.; Ryu, E.; Lee, S.-W.; Park, J.; Ha, N.Y.; Ra, J.S.; Kim, Y.J.; Kim, J.; Abdel-Rahman, M.; Park, S.H.; et al. Regulation of PCNA Cycling on Replicating DNA by RFC and RFC-like Complexes. Nat. Commun. 2019, 10, 2420. [Google Scholar] [CrossRef]
  269. Kubota, T.; Nishimura, K.; Kanemaki, M.T.; Donaldson, A.D. The Elg1 Replication Factor C-like Complex Functions in PCNA Unloading during DNA Replication. Mol. Cell 2013, 50, 273–280. [Google Scholar] [CrossRef]
  270. Lee, K.; Fu, H.; Aladjem, M.I.; Myung, K. ATAD5 Regulates the Lifespan of DNA Replication Factories by Modulating PCNA Level on the Chromatin. J. Cell. Biol. 2013, 200, 31–44. [Google Scholar] [CrossRef]
  271. Brambati, A.; Zardoni, L.; Achar, Y.J.; Piccini, D.; Galanti, L.; Colosio, A.; Foiani, M.; Liberi, G. Dormant Origins and Fork Protection Mechanisms Rescue Sister Forks Arrested by Transcription. Nucleic Acids Res. 2018, 46, 1227–1239. [Google Scholar] [CrossRef]
  272. Yang, Y.; Xu, W.; Gao, F.; Wen, C.; Zhao, S.; Yu, Y.; Jiao, W.; Mi, X.; Qin, Y.; Chen, Z.-J.; et al. Transcription-Replication Conflicts in Primordial Germ Cells Necessitate the Fanconi Anemia Pathway to Safeguard Genome Stability. Proc. Natl. Acad. Sci. USA 2022, 119, e2203208119. [Google Scholar] [CrossRef] [PubMed]
  273. Che, R.; Zhang, J.; Nepal, M.; Han, B.; Fei, P. Multifaceted Fanconi Anemia Signaling. Trends Genet. 2018, 34, 171–183. [Google Scholar] [CrossRef] [PubMed]
  274. Olazabal-Herrero, A.; He, B.; Kwon, Y.; Gupta, A.K.; Dutta, A.; Huang, Y.; Boddu, P.; Liang, Z.; Liang, F.; Teng, Y.; et al. The FANCI/FANCD2 Complex Links DNA Damage Response to R-Loop Regulation through SRSF1-Mediated mRNA Export. Cell Rep. 2024, 43, 113610. [Google Scholar] [CrossRef] [PubMed]
  275. Christou, C.M.; Kyriacou, K. BRCA1 and Its Network of Interacting Partners. Biology 2013, 2, 40–63. [Google Scholar] [CrossRef]
  276. Hall, J.M.; Lee, M.K.; Newman, B.; Morrow, J.E.; Anderson, L.A.; Huey, B.; King, M.C. Linkage of Early-Onset Familial Breast Cancer to Chromosome 17q21. Science 1990, 250, 1684–1689. [Google Scholar] [CrossRef]
  277. Smith, S.A.; Easton, D.F.; Evans, D.G.; Ponder, B.A. Allele Losses in the Region 17q12-21 in Familial Breast and Ovarian Cancer Involve the Wild-Type Chromosome. Nat. Genet. 1992, 2, 128–131. [Google Scholar] [CrossRef]
  278. Anderson, S.F.; Schlegel, B.P.; Nakajima, T.; Wolpin, E.S.; Parvin, J.D. BRCA1 Protein Is Linked to the RNA Polymerase II Holoenzyme Complex via RNA Helicase A. Nat. Genet. 1998, 19, 254–256. [Google Scholar] [CrossRef]
  279. Patel, P.S.; Abraham, K.J.; Guturi, K.K.N.; Halaby, M.-J.; Khan, Z.; Palomero, L.; Ho, B.; Duan, S.; St-Germain, J.; Algouneh, A.; et al. RNF168 Regulates R-Loop Resolution and Genomic Stability in BRCA1/2-Deficient Tumors. J. Clin. Investig. 2021, 131, e140105. [Google Scholar] [CrossRef]
  280. Yasuhara, T.; Kato, R.; Hagiwara, Y.; Shiotani, B.; Yamauchi, M.; Nakada, S.; Shibata, A.; Miyagawa, K. Human Rad52 Promotes XPG-Mediated R-Loop Processing to Initiate Transcription-Associated Homologous Recombination Repair. Cell 2018, 175, 558–570.e11. [Google Scholar] [CrossRef]
  281. Washburn, R.S.; Gottesman, M.E. Transcription Termination Maintains Chromosome Integrity. Proc. Natl. Acad. Sci. USA 2011, 108, 792–797. [Google Scholar] [CrossRef] [PubMed]
  282. Kim, M.; Krogan, N.J.; Vasiljeva, L.; Rando, O.J.; Nedea, E.; Greenblatt, J.F.; Buratowski, S. The Yeast Rat1 Exonuclease Promotes Transcription Termination by RNA Polymerase II. Nature 2004, 432, 517–522. [Google Scholar] [CrossRef] [PubMed]
  283. Mérida-Cerro, J.A.; Maraver-Cárdenas, P.; Rondón, A.G.; Aguilera, A. Rat1 Promotes Premature Transcription Termination at R-Loops. Nucleic Acids Res. 2024, 52, 3623–3635. [Google Scholar] [CrossRef] [PubMed]
  284. Tehranchi, A.K.; Blankschien, M.D.; Zhang, Y.; Halliday, J.A.; Srivatsan, A.; Peng, J.; Herman, C.; Wang, J.D. The Transcription Factor DksA Prevents Conflicts between DNA Replication and Transcription Machinery. Cell 2010, 141, 595–605. [Google Scholar] [CrossRef]
  285. Trautinger, B.W.; Jaktaji, R.P.; Rusakova, E.; Lloyd, R.G. RNA Polymerase Modulators and DNA Repair Activities Resolve Conflicts between DNA Replication and Transcription. Mol. Cell 2005, 19, 247–258. [Google Scholar] [CrossRef]
  286. Jiang, C.; Hong, Z.; Liu, S.; Hong, Z.; Dai, B. Roles of CDK12 Mutations in PCa Development and Treatment. Biochim. Biophys. Acta (BBA)-Rev. Cancer 2025, 1880, 189247. [Google Scholar] [CrossRef]
  287. Milano, L.; Gautam, A.; Caldecott, K.W. DNA Damage and Transcription Stress. Mol. Cell 2024, 84, 70–79. [Google Scholar] [CrossRef]
  288. Tien, J.C.-Y.; Luo, J.; Chang, Y.; Zhang, Y.; Cheng, Y.; Wang, X.; Yang, J.; Mannan, R.; Mahapatra, S.; Shah, P.; et al. CDK12 Loss Drives Prostate Cancer Progression, Transcription-Replication Conflicts, and Synthetic Lethality with Paralog CDK13. Cell Rep. Med. 2024, 5, 101758. [Google Scholar] [CrossRef]
  289. Gómez-González, B.; García-Rubio, M.; Bermejo, R.; Gaillard, H.; Shirahige, K.; Marín, A.; Foiani, M.; Aguilera, A. Genome-Wide Function of THO/TREX in Active Genes Prevents R-Loop-Dependent Replication Obstacles. EMBO J. 2011, 30, 3106–3119. [Google Scholar] [CrossRef]
  290. Yang, Z.; Li, M.; Sun, Q. RHON1 Co-Transcriptionally Resolves R-Loops for Arabidopsis Chloroplast Genome Maintenance. Cell Rep. 2020, 30, 243–256.e5. [Google Scholar] [CrossRef]
  291. Alzu, A.; Bermejo, R.; Begnis, M.; Lucca, C.; Piccini, D.; Carotenuto, W.; Saponaro, M.; Brambati, A.; Cocito, A.; Foiani, M.; et al. Senataxin Associates with Replication Forks to Protect Fork Integrity across RNA-Polymerase-II-Transcribed Genes. Cell 2012, 151, 835–846. [Google Scholar] [CrossRef]
  292. Chang, E.Y.-C.; Stirling, P.C. Replication Fork Protection Factors Controlling R-Loop Bypass and Suppression. Genes 2017, 8, 33. [Google Scholar] [CrossRef]
Figure 1. Types of TRCs and R-loop formation. HO-TRC (left) is head-on transcription–replication conflict, while CD-TRC (right) is co-directional transcription–replication conflict. RNAP is RNA polymerase. Replisome was drawn schematically based on [30].
Figure 1. Types of TRCs and R-loop formation. HO-TRC (left) is head-on transcription–replication conflict, while CD-TRC (right) is co-directional transcription–replication conflict. RNAP is RNA polymerase. Replisome was drawn schematically based on [30].
Ijms 26 06951 g001
Figure 2. The factors preventing R-loop formation. RNAP is RNA polymerase. Primase-Polymerase (PrimPol) participates in the replication process of sequences that are prone to R-loop formation. The coating proteins cohesin SA1/2 supposably participate in the protection of newly synthesized RNA. Topoisomerase Top1 resolves the accumulation of local negative supercoils on transcribed regions. The chromatin structure, including marks such as H3K9 dimethylation (H3K9me2) and H2AK119 ubiquitination (H2AK119ub) and the enrolment of chromatin remodeling factors (INO80, SMARCA5, and MTA2), also plays a role in preventing R-loop formation.
Figure 2. The factors preventing R-loop formation. RNAP is RNA polymerase. Primase-Polymerase (PrimPol) participates in the replication process of sequences that are prone to R-loop formation. The coating proteins cohesin SA1/2 supposably participate in the protection of newly synthesized RNA. Topoisomerase Top1 resolves the accumulation of local negative supercoils on transcribed regions. The chromatin structure, including marks such as H3K9 dimethylation (H3K9me2) and H2AK119 ubiquitination (H2AK119ub) and the enrolment of chromatin remodeling factors (INO80, SMARCA5, and MTA2), also plays a role in preventing R-loop formation.
Ijms 26 06951 g002
Table 2. Protein factors involved in suppression of pathological R-loop accumulation.
Table 2. Protein factors involved in suppression of pathological R-loop accumulation.
ProteinProtein Function (Known Role in The Organism)Role in R-Loop SuppressionRef.
Prevention of R-loop formation
RPABinds and stabilizes ssDNA intermediates that form during DNA replication or following DNA stress.Protects the nascent RNA by coating it. Allows human DNA polymerases to initiate DNA synthesis utilizing RPA-generated R-loops, thus reproducing replication restart in vivo. Prevents R-loop-induced DSB formation in SETX-deficient cells. Attracts RNase H1 to the regions of the formed R-loops and enhances its nuclease activity. Recruits HELQ helicase to R-loops. Signals ATR activation.[100,125,129,138,139,140]
SA1, SA2Components of the cohesin complex that play an important role in 3D chromatin organization. Strongly bind to RNA, especially DNA–RNA hybrids.Protect the nascent RNA by coating it.[141]
INO80ATP-ase of the chromatin remodeling complex.Suppression of TRCs due to heterochromatinization.[142]
SMARCA5Helicase that possesses intrinsic ATP-dependent nucleosome-remodeling activity.
MTA2Acts as a component of the histone deacetylase NuRD complex, which participates in the remodeling of chromatin.
Fob1Nucleolar protein that binds to the rDNA replication fork barrier site. Required for replication fork blocking.Suppression of TRCs due to replication blockage.[43]
Topo IV *, DNA gyrase *, Top1 *, TOP1 TOP3BTopoisomerases.Prevention of topological stress accumulation in DNA.[50,53,143,144,145,146]
T4 Dda *, Rep *, UvrD *, DinG *, PcrA *, Rrm3, Mfd *Helicases.Prevent TRCs from occurring by displacing proteins that may be an obstacle to the passage of the replication fork and RNAP from DNA. Mfd is an elongation factor that displaces RNAP from the replisome pathway.[23,26,56,57,58,59]
RECQ5Helicase.Plays an important role in the resolution of TRCs, particularly by removing RAD51 from the stalled replication fork to facilitate MUS81 endonuclease’s cleavage of the fork.[95,147,148,149,150]
PrimPolDNA polymerase called Primase-Polymerase. Plays a role in DNA damage tolerance in eukaryotes.The participation of PrimPol in the replication process of sequences that are prone to R-loop formation reduces their formation.[151]
XRN25’-3’ exoribonuclease is implicated in transcription termination.Supposably prevents formation of R-loop-degrading downstream RNA containing a 5′ monophosphate as part of the termination process for most RNAP II transcripts.[152]
Direct resolution of R-loops
RNase HEndonucleases that specifically degrade RNA in DNA–RNA hybrids.Direct nuclease digestion of R-loops. RNase H2 interacts with RNAP II.[153,154,155]
DICERDouble-stranded RNA (dsRNA) endoribonuclease playing a central role in short dsRNA-mediated post-transcriptional gene silencing.Specifically cleaving RNA within R-loops.[156]
REXO5RNA exonuclease.[157]
FEN1Structure-specific nuclease with 5′-flap endonuclease and 5′-3′ exonuclease activities involved in DNA replication and repair.[158,159,160]
Pif1, Rrm3Specific helicases.Directly resolve R-loops and/or G-quadruplexes.[161,162]
BLM, WRN, RTEL1, PIF1[67,163,164,165,166,167]
FANCJDirectly resolve R-loops through interacting with MutSβ and MLH1, the components of the mismatch repair complex MutLβ.[8]
SETX (Sen1)RNA/DNA helicase involved in diverse aspects of RNA metabolism and genomic integrity.The mechanism of resolution of TRCs via the helicase activity of SETX seems to be associated with the promotion of replication restart at R-loop formation sites through the MUS81–LIG4–ELL pathway, involving MUS81-mediated cleavage of the leading chain of the stalled fork, DNA ligase IV (LIG4)/XRCC4 complex-facilitated religation of the fork, and RNAP II passage provided by the elongation factor ELL.[95,168,169,170]
DDX1, DDX5, DHX9, DDX17, DDX19, DDX21, DDX37, DDX50DEAD/DEXH-box RNA helicases.Directly resolve R-loops.[101,119,171,172,173,174,175,176,177,178]
HELQHelicase belonging to the superfamily 2 (SF2).Unwinds R-loops in vitro as well as in cells. Interacts with nuclear 5’ to 3’ exoribonuclease, a transcription termination factor of Rat1/XRN2, and supposably coordinating the unwinding and degradation of R-loops.[140]
NAT10RNA cytidine acetyltransferase that catalyzes the modification of N4-acetylcytidine on mRNAs, 18S rRNA, and tRNAs.Resolves R-loops through its RecD helicase domain activity. Also acetylates DDX21 at K236 and K573, thus enhancing its helicase activity towards nucleolar R-loops.[179,180,181,182]
Indirect impact on R-loop resolution
ATAD5Tumor suppressor. Functions as a PCNA (the eukaryotic sliding clamp for replicative polymerases) unloader.Increases the abundance of DEAD/DEXH-box RNA helicases at replication fork sites, thus participating in R-loop resolution.[119,172,173,174,175,176,177,178]
ATRSerine/threonine protein kinase which activates checkpoint signaling upon genotoxic stresses such as ionizing radiation, ultraviolet light, or DNA replication stalling, thereby acting as a DNA damage sensor.Activated in response to HO-TRCs. Phosphorylates BRCA1, CHEK1, MCM2, RAD17, RBBP8, RPA2, SMC1, DHX9, and p53/TP53, which collectively inhibit DNA replication and mitosis and promotes DNA repair, recombination, and apoptosis. The phosphorylation of DHX9 by ATR facilitates its interaction with BRCA1 and RPA, leading to its accumulation at R-loops.[100,121,183,184,185,186,187,188,189]
ADAR1Catalyzes the hydrolytic deamination of adenosine to inosine in double-stranded RNA (dsRNA), referred to as A-to-I RNA editing.Attracts ATR to R-loops through its interaction with TOPBP1 (scaffold protein that acts as a key protein–protein adapter in DNA replication and DNA repair and promotes the loading of RAD51). Is also suggested to attract DHX9 and DDX21 helicases to R-loops.[101]
MRN complex (MRE11-RAD50-NBS1)Plays an important role in early DNA damage signaling and the processing of DNA ends at DSBs.Plays a role in suppressing R-loops through the recruitment of FA pathway complexes to these sites.[190,191,192,193]
BRCA1Tumor suppressor, participates in a variety of cellular processes, including DNA damage repair, replication fork protection, transcription, and cell cycle regulation, through its interactions with different partners. Participates in the FA pathway.Participates in transcription activation and R-loop resolution and is able to interact with RNAP II through DHX9. Attracts SETX to R-loops. Was demonstrated to participate in the activation of the persistent R-loop-induced homologue recombination pathway.[194,195,196,197,198,199,200]
BRCA2Involved in double-strand break repair and homologous recombination. Participates in the FA pathway.Acts by targeting RAD51 to ssDNA over dsDNA, enabling RAD51 to displace replication protein A (RPA) from ssDNA and stabilizing RAD51-ssDNA filaments by blocking ATP hydrolysis. Was also shown to recruit RNase H2.[1,94,95,201,202,203]
RAD52Involved in double-strand break repair. Plays a central role in genetic recombination and DNA repair by promoting the annealing of complementary ssDNA and stimulating RAD51 recombinase.Signaling of R-loop-associated damage and homologous recombination.[204,205]
PARP1Poly-ADP-ribosyltransferase that mediates poly-ADP-ribosylation of proteins and plays a key role in DNA repair.Exhibits a tendency to accumulate at R-loop sites in cells, possibly attracting other DNA repair enzymes.[206]
XPF (ERCC1) and XPG (ERCC5)NER endonucleases.Attracted to R-loops and initiate DSBs at their ssDNA sites.[136]
METTL3, METTL14 and Wilms tumor 1-associated protein (WTAP)N6-methyltransferase complex that methylates adenosine residues at the N6 position of some RNAs and regulates various processes, such as the circadian clock, response to DNA damage, and primary miRNA processing.Attracts RNase H1 through m6A methylation of RNA. Possibly attracted to R-loops by DDX21 helicase.[207,208,209,210,211,212]
Rho *Transcription termination factor.Prevents R-loop formation, possibly through transcriptional arrest.[213]
* Procaryotic homolog.
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

Davletgildeeva, A.T.; Kuznetsov, N.A. Participants in Transcription–Replication Conflict and Their Role in Formation and Resolution of R-Loops. Int. J. Mol. Sci. 2025, 26, 6951. https://doi.org/10.3390/ijms26146951

AMA Style

Davletgildeeva AT, Kuznetsov NA. Participants in Transcription–Replication Conflict and Their Role in Formation and Resolution of R-Loops. International Journal of Molecular Sciences. 2025; 26(14):6951. https://doi.org/10.3390/ijms26146951

Chicago/Turabian Style

Davletgildeeva, Anastasiia T., and Nikita A. Kuznetsov. 2025. "Participants in Transcription–Replication Conflict and Their Role in Formation and Resolution of R-Loops" International Journal of Molecular Sciences 26, no. 14: 6951. https://doi.org/10.3390/ijms26146951

APA Style

Davletgildeeva, A. T., & Kuznetsov, N. A. (2025). Participants in Transcription–Replication Conflict and Their Role in Formation and Resolution of R-Loops. International Journal of Molecular Sciences, 26(14), 6951. https://doi.org/10.3390/ijms26146951

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