ABSTRACT
DNA replication and RNA transcription both utilize DNA as a template and therefore need to coordinate their activities. The predominant theory in the field is that in order for the replication fork to proceed, transcription machinery has to be evicted from DNA until replication is complete. If that does not occur, these machineries collide, and these collisions elicit various repair mechanisms which require displacement of one of the enzymes, often RNA polymerase, in order for replication to proceed. This model is also at the heart of the epigenetic bookmarking theory, which implies that displacement of RNA polymerase during replication requires gradual re-building of chromatin structure, which guides recruitment of transcriptional proteins and resumption of transcription. We discuss these theories but also bring to light newer data that suggest that these two processes may not be as detrimental to one another as previously thought. This includes findings suggesting that these processes can occur without fork collapse and that RNA polymerase may only be transiently displaced during DNA replication. We discuss potential mechanisms by which RNA polymerase may be retained at the replication fork and quickly rebind to DNA post-replication. These discoveries are important, not only as new evidence as to how these two processes are able to occur harmoniously but also because they have implications on how transcriptional programs are maintained through DNA replication. To this end, we also discuss the coordination of replication and transcription in light of revising the current epigenetic bookmarking theory of how the active gene status can be transmitted through S phase.
KEYWORDS: epigenetic bookmarking, nascent chromatin, RNA polymerase II, transcription memory, transcription post-replication, transcription-replication conflicts
DNA as a template for two biochemical processes
As initially described in the central dogma of molecular biology, DNA is the substrate for two critical biochemical processes: DNA replication and RNA transcription [1]. Since both of these processes require access to the DNA template, it has long been a question of what coordination is necessary in order to avoid conflicts between these two pivotal processes. Despite the theorized displacement of RNA polymerase during replication, the memory of whether a gene is active or repressed must be maintained through passage of the replisome in order for a cell to sustain its transcriptional output of lineage specific-genes during the cell cycle.
The cell cycle consists of four stages that cells pass through during a round of cellular division. Beginning in Gap 1 (G1) phase cells then move into S phase wherein all of the DNA within the nucleus is duplicated during the process of DNA replication. Cells move from S phase into Gap 2 (G2) phase, which is followed by mitosis in M phase. Cells then exit M phase back into G1. Transcription was initially believed to be ceased during the whole several hours of S phase and take place only during the gap phases [2,3]. However, replication of a relatively short region from each replicon, which may contain the entire coding region of a gene, occurs relatively fast, and it is not clear what may prevent resumption of transcription of this gene within the continuing S phase. In line with this, more recent studies have shown that transcription does occur during S phase [4,5] and even M phase [6,7], despite the structural reorganization of chromatin during these phases of the cell cycle due to passage of the replisome [8,9] and mitotic condensation [10,11], respectively. During DNA replication, this requires a mechanism by which these two processes can occur together without causing widespread issues and genomic instability. In this review, we discuss various studies and theories about how these processes can occur harmoniously. We focus primarily on eukaryotic systems, since studies in prokaryotes have been extensively discussed elsewhere, and since bacteria may have their own approach to coordinating these processes, which largely involves orienting transcriptional units to be co-directional with replication [12–15]. Eukaryotic systems, however, contain chromatinized DNA and their gene expression profiles may vary vastly depending on the particular set of genes expressed in a particular cell lineage. Taken together, this suggests that coordination of these processes is more complicated in eukaryotic systems than in bacterial systems.
DNA replication
During S phase, DNA replication initiates asynchronously from origins of replication located throughout the genome, which are bound by the origin recognition complex (ORC) [16]. The ORC then recruits CDC6, CDT1, and the replicative DNA helicase, MCM2–7. GINS and CDC45 bind to MCM2–7 (forming the CMG helicase complex) which unwinds the DNA double helix creating two replication forks which move bi-directionally from one another [17]. The CMG complex subsequently recruits the replicative DNA polymerases [18,19]. Humans have 16 DNA polymerases [20], but the three that are critical during DNA replication are as follows: DNA polymerase alpha (Pol α), DNA polymerase delta (Pol δ), and DNA polymerase epsilon (Pol ε). DNA-dependent DNA polymerases require a double-stranded template in order to catalyze the addition of nucleotides, so the Pol α/primase complex creates an RNA primer on each of the two strands [21]. In addition to an RNA primer, Pol δ and ε both require a sliding clamp processivity factor, proliferating cell nuclear antigen (PCNA), in order to efficiently synthesize DNA [22]. PCNA is a ring-shaped homotrimer that is loaded onto DNA by the clamp loading complex replication factor C (RFC), which opens the PCNA ring and encloses it around DNA [23]. The complex comprising the CMG helicase, in addition to DNA polymerases, PCNA, RFC, and other critical replication proteins is often referred to as the replisome.
DNA replication involves incorporating nucleotide triphosphates into the nascent daughter DNA strand in a 5’ to 3’ direction. Because of the antiparallel nature of the DNA double helix, this means that the two strands at a replication fork must be synthesized in different manners. The leading daughter strand is synthesized continuously, whereas the lagging daughter strand must be synthesized in short stretches called Okazaki fragments which move in the opposite direction of progression of the replication fork [24,25]. This division of labor involves different DNA polymerases for each of the two strands with the leading and lagging strands being synthesized by Pol ε and Pol δ, respectively [26,27]. Okazaki fragments in eukaryotes are ~200 nucleotide stretches of nascent DNA [17], each of which begins with an RNA primer catalyzed by Pol α/primase. Once Pol δ reaches the primer of the previous Okazaki fragment, it displaces the RNA primer creating a nucleotide flap hanging off of the DNA which is then cleaved by flap endonuclease 1 (FEN1) [28]. The two Okazaki fragments are then stitched together by DNA ligase I [29]. The coordinated synthesis of the leading and lagging daughter strands allows for faithful duplication of the entire genome within a cell.
RNA transcription
Eukaryotic cells contain three nuclear DNA-dependent RNA polymerases: RNA polymerase I (Pol I), RNA polymerase II (Pol II), and RNA polymerase III (Pol III) [30]. Pol I is responsible for transcription of the 47 S pre-ribosomal RNA (rRNA) which is cleaved to form the 18 S, 5.8 S, and 28 S rRNAs [31,32]. Pol III transcribes all transfer RNAs (tRNAs), the 5 S rRNA, and numerous other small RNAs including several small nuclear RNAs (snRNAs) and small nucleolar RNAs (snoRNAs) [28,33]. Pol II transcribes all other RNAs, including messenger RNAs (mRNAs), as well as numerous types of non-protein coding RNAs (ncRNAs) including micro RNAs (miRNAs) [34], piwi-interacting RNAs (piRNAs) [35], long noncoding RNAs (lncRNAs) [36], snRNAs, and snoRNAs [37]. Importantly, Pol I and Pol III transcribe the same genes regardless of cell lineage because expression of these genes is required by all cells in an organism. Pol II, on the other hand, plays a pivotal role in differential gene expression and must be properly directed to and maintained at target gene loci that are specific to a particular cell type [38–41]. This makes Pol II critical when it comes to understanding epigenetic inheritance because all epigenetic programs must converge on faithful restoration of Pol II at its target loci throughout the cell cycle.
The core Pol II holoenzyme is a 12-factor complex, which contains RPBs 1 through 12 (including several homodimers of these products). The largest components, RPB1 and RPB2, form a cleft containing the active enzymatic domain of the complex responsible for incorporation of ribonuclear triphosphate precursors into the growing nascent RNA strand [42,43]. Additionally, RPB1 contains a heptad-repeat in the carboxy-terminal domain (CTD) with 26 repeats in yeast and 52 repeats in humans, whose consensus sequence is Tyr1-Ser2-Pro3-Thr4-Ser5-Pro6-Ser7 [44]. Although the CTD of Pol II contains multiple residues capable of being phosphorylated, the two residues most well characterized are particularly critical for the transcription cycle: Ser5 and Ser2, whose phosphorylation corresponds to transcriptional initiation and elongation, respectively [45,46]. In order to initiate transcription, numerous general transcription factors (GTFs) must be assembled at the gene promoter in a particular order, ultimately culminating in the formation of an active transcription bubble [47–50].
In vitro assessment of replication and transcription coordination
Some understanding of what happens when replication and transcription collide is based on in vitro experiments. These studies have primarily been carried out utilizing bacterial and viral systems. The question of what happens when RNA polymerase encounters DNA replication machinery was initially addressed by the Alberts group in the early 1990s. They showed that in a reconstituted in vitro system consisting of the T4 bacteriophage replisome and the E. coli RNA polymerase, both head-on and co-directional collisions of these enzymes did not displace RNA polymerase or cause it to stop synthesizing its nascent transcript [51–53]. This observation is supported by an in vitro investigation of collisions between the B. subtilis RNA polymerase and the phage Φ29 DNA polymerase, which showed that during head-on collisions, the enzymes are able to bypass one another as long as the transcriptional complex was not stalled [54]. Using the same system, Elías-Arnanz and Salas found that co-directional collisions (with the Φ29 DNA polymerase trailing RNA polymerase) resulted in a decreased rate of replication but not displacement of either complex from DNA [55].
More recent studies utilizing a different reconstituted in vitro system showed that the E. coli replisome is able to displace the T7 bacteriophage transcriptional machinery [56,57]. Another interesting in vitro study looked at the consequence of T7 RNA polymerase encountering the ORC, MCM double hexamer, or the ORC bound to MCM-CDT1 and found that the polymerase could push these complexes down the DNA, and even frequently bypassed them [58,59]. This indicates that RNA polymerase is capable of bypassing these replication proteins.
All of these studies, while biochemically informative, utilized reconstituted in vitro systems, which lack full cellular biological context. Furthermore, they focus on bacterial and viral enzymes, which may have no connection to how eukaryotic systems function, particularly because of the need to overcome chromatin barriers present on eukaryotic DNA. Additionally, members of the eukaryotic replication fork are continuing to be discovered [60], which could play a crucial role in resolving interactions between RNA and DNA polymerases. Thus, any in vitro system would likely lack the full complement of molecules that normally function in DNA replication. Given both the conflicting nature of these in vitro studies and their limitations, it is difficult to extrapolate their results on in vivo coordination of these processes, especially in higher eukaryotes.
Placement of origins near active gene promoters/order of replication
The S phase is often divided into early and late S phase. It has been well established that actively transcribed genes are replicated early in S phase, whereas repressed genes are replicated in late S phase [61–64]. Subsequent studies showed that origins of replication (ORIs), which fire during early S phase, are close to the promoters of actively expressed genes [65,66], as was initially demonstrated for several genes, including MYC (c-MYC) [67] and HBB (β-globin) [68]. This indicates not only that there is something about these regions that makes them more prone to firing early during replication but also that the selection of which ORIs fire and when they fire during S phase is dependent on the particular cell type [69,70].
This coordination between ORIs selected to fire and the transcriptional status of genes also indicates that there is some biological advantage to replicating these regions early in S phase. It has been shown that transcription factors associated with Pol II and Pol III can positively regulate DNA replication by their association with the budding yeast autonomously replicating sequence (ARS) [71,72]. While higher eukaryotes do not have a consensus sequence for ORC binding, data suggest that the ORC requires a nucleosome-free region (NFR) in order to efficiently load the MCM2–7 helicase complex [73]. Taken as a whole, sequences to load the ORC vary between eukaryotes, and do not appear to exist in metazoans, which likely require specific chromatin structures for ORC loading [74], a property that might be advantageous to coordinating replication with transcription in multicellular organisms. Overall, the existence of ORIs near active promoters implies a coordination between replication and transcription and that this coordination can be tailored to particular cell lineages and transcriptional profiles in higher order eukaryotes.
Transcription-replication conflict theory
Given that both transcription and replication require access to the same regions of DNA, understanding what happens when and if they come into contact with one another has been an active field of inquiry. Transcription-replication conflicts (TRCs) have been a main theory in the field and have been extensively covered in other review articles [12,75–80]. This theory holds that these processes are inherently disruptive to one another and collisions between these machineries can cause a number of outcomes including fork stalling, fork reversal, and/or formation of an R-loop [81], which is a RNA-DNA hybrid duplex, in addition to the other displaced DNA strand.
Resolution of TRCs ultimately requires removal of either Pol II or the replication machinery in order for replication to resume. It is of note, though, that these studies generally rely on stressing one or the other process in order to determine what happens when transcription and replication machinery collide due to using the same DNA template. Specifically, they use exogenous stressors including pharmacologic inhibition of replication [82,83], or transcription [84,85], or genetic disruption of the endogenous replication [86,87] or transcription [88,89] processes. This is an important consideration, because it suggests that these conflicts may not arise in cells where transcription and replication are occurring unimpaired. In addition, a recent study showed that these types of perturbations result in firing of non-canonical ORIs, which can lead to the resulting phenotype of fork stalling and/or collapse [90]. This suggests that the TRC phenotype observed may not be the result of normal transcription and replication.
Reports on the collisions between DNA and RNA polymerases have also been suggested to result in DNA replication that occurs outside of S phase of the cell cycle. Initial studies found that following DNA damage, replication could occur during mitosis (MiDAS) [91,92] and it was subsequently shown that normally cycling cells may not replicate some transcriptional start sites (TSSs) until the cell reaches G2/M phase of the cell cycle, in a process termed G-MiDAS [5]. While striking that normally dividing cells were replicating their DNA after S phase, the authors only observed this in one-fifth of their cell population, and of those cells it only occurred at approximately 400 TSSs. Thus, it does not appear likely that this is a prominent feature due to collisions between DNA and RNA polymerases even in stressed cells.
Another interesting investigation of TRCs is from a study on fragile site instability in long genes [93]. This study showed that not only are some particularly long genes transcribed through the entire S phase but that completing their transcription takes even longer than the entire length of the cell cycle in human B-lymphoblasts. This indicates that Pol II must finish transcribing these long genes following passage of the replisome. These observations led the authors to propose that there must be some mechanism to recruit Pol II bound to its immature RNA transcript back to these long genes in order to finish transcribing them, and the authors hypothesized that the immature transcript might function to recruit Pol II back to DNA [94]. Additionally, a study of the very long Ubx gene in Drosophila embryos, which have very short cell cycle, showed that its transcription begins shortly after the completion of mitosis and ends when the next mitotic phase begins [95]. Interestingly, another study also suggests that RNA may not be dissociated during DNA replication by use of an in vivo approach which knocked in the MS2 protein-binding sequence into the 3’ UTR of CCND1 in the presence of MS2-GFP expression, allowing visualization of the transcription dynamics of a single allele of a gene in real time [96]. When the gene was replicated the authors did not observe a loss of the mRNA signal during replication and instead saw the signal duplicate, suggesting that RNA, potentially in complex with Pol II, is transferred to both two daughter strands [96].
These studies imply that transcription of such long genes has to be only transiently interrupted by DNA replication. Taken together, these studies suggest that Pol II has the ability to recognize the unfinished nascent RNA transcript and complete its synthesis after passage of the replication machinery. These results imply a model of close coordination of replication and transcription, rather than TRCs, wherein Pol II must be removed from chromatin in order for replication to proceed.
Spatial separation of transcription and replication in factories
Contrary to the theory that replication and transcription machineries collide, numerous studies investigating these processes suggest that each occurs in distinct nuclear domains, termed replication factories and transcription factories, respectively [97–101]. This suggests a coordination of these two processes by physically separating them within the nucleus. The factory theory originates from studies using immunofluorescence of fixed samples where staining for transcription and replication proteins resulted in discrete foci throughout the nucleus [102–104]. However, rarely these foci may overlap, suggesting that transcription and replication may occasionally come into spatial or physical contact with one another, and that these processes may not occur entirely independently of one another [105].
These foci are termed factories because they contain multiple polymerases, on average eight for transcription factories [106] and up to 40 replication forks with numerous polymerases each within replication factories [107]. A principle component of the factory theory that distinguishes it from other conceptualizations of replication and transcription is that it posits that the enzymes are immobilized and DNA is spooled through them, as opposed to the enzymes traversing down DNA. This has long been an accepted theory for DNA replication [108–110]; however, it has only more recently become suggested for transcription [111,112]. Thus, conflicts between replication and transcription could only occur if the two are pulled together and collide.
There are several pieces of evidence supporting the factory model. For one, early data showed that components of the replisome associate with the nuclear lamin, indicating that these proteins are physically affixed to this nuclear scaffold [97,108,113]. Moreover, transcription factories have shown to remain even after the removal of chromatin, indicating that these structures are also bound to something within the nucleus apart from chromatin [113,114]. Second, in vitro studies have shown that both DNA polymerase [115] and RNA polymerase [103,116,117] are capable of spooling their DNA template through them as they catalyze their respective biochemical reactions. Thus, it is possible that replication and transcription take place within factories of immobilized enzymes, which would suggest one way to avoid direct physical collisions between the corresponding protein complexes.
Current theories of epigenetic bookmarking during S phase
Up to this point we have discussed the coordination of replication and transcription in a biochemical context, but it is important to also consider this from a conceptual standpoint. Maintenance of transcriptional programs through DNA replication is important for transmission of the status of genes as active or repressed. Since it is believed that transcription machinery leaves DNA during replication, this is thought to require epigenetic bookmarking as the replisome passes over differentially expressed regions of the genome. There are currently two main epigenetic theories during S phase. The predominant theory, is the reader-writer histone epigenetic hypothesis, where in modified histones are transferred during DNA replication from nucleosomes on the parental DNA to the newly formed nucleosomes on the two daughter strands (Figure 1(a)) [9,118]. These modified histones are able to re-recruit histone-modifying complexes via histone reader domains within the complexes, and these complexes then deposit modifications onto newly formed nucleosomes via histone writer domains. Recruitment of these complexes and propagation of histone modifications leads to the restoration of the chromatin landscape throughout the genome after DNA replication and is thought to be essential for recruitment of Pol II and other components of the transcriptional machinery to the promoters of active genes.
Figure 1.

Proposed models of epigenetic marking of active genes during DNA replication.
(a) Reader-writer epigenetic theory. During DNA replication, the replisome displaces most chromatin-bound proteins, including Pol II and chromatin-modifying complexes, such as TrxG proteins. Modified histones from parental nucleosomes are transferred semi-conservatively to nascent DNA into the newly assembled nucleosomes on daughter strands. Chromatin-modifying complexes are recruited to replicated DNA by interacting with these modified histones through “reader” protein components of these complexes. Deposition of active histone marks by “writer” protein components on all newly incorporated histones completes chromatin restoration and culminates in recruiting Pol II back to active genes.
(b) Chromatin-modifier epigenetic theory. During DNA replication, chromatin-modifying complexes are transferred from parental DNA to newly synthesized DNA, but modified histones are displaced. Unmodified histones are incorporated into nascent DNA and are post-translationally modified by the retained chromatin-modifying complexes in active regions. Once the structure of chromatin is restored, Pol II is recruited to replicated DNA at active genes in order to restore their transcriptional status.
(c) Pol II epigenetic theory. During DNA replication, Pol II in complex with immature RNA is retained close to the replisome via protein–protein interactions with PCNA, allowing it to rebind to nascent DNA immediately following passage of DNA polymerase. This occurs both at promoters and in the gene body. While other factors are likely bound to nascent DNA during this timeframe, potentially including nucleosomes and chromatin-modifying complexes, the presence of Pol II on DNA immediately after replication alleviates the proposed necessity of any other factor in order to epigenetically bookmark active genes. Therefore, nucleosomes and chromatin-modifying complexes were not included in this model.
The second theory, which is better supported by experimental data, holds the converse, namely, that chromatin-modifying complexes are transferred during DNA replication and are able to deposit their marks on unmodified histones that are loaded onto newly synthesized DNA during replication (Figure 1(b)) [119–123]. For active genes, this means the transfer of trithorax group (TrxG) proteins to nascent DNA which can then deposit active histone marks on newly deposited histones. These findings questioning whether modified histones are necessary for transmitting epigenetic information during S phase were further supported by a study showing that modified histones in active regions are not maintained during DNA replication [124]. Current thinking in the field has shifted such that, at least in active regions, modified histones are not believed to be the mediators of epigenetic bookmarking through replication [125–127]. Based on this, active regions are marked by factor(s) other than modified histones, potentially including chromatin-modifying complexes, as suggested [119,120].
An additional important caveat to the reader-writer theory is a question of whether modified histones in active regions are the drivers of gene activation or are the result of transcription. It appears that H3K4me3, the primary modified histone associated with active gene promoters, may be the result of transcription rather than the cause of transcription [128]. Moreover, the loss of H3K4me3 does not impact transcription levels [129,130], further undercutting its putative role as an epigenetic mark. A more recent study showed that rapid depletion of SET1/COMPASS caused loss of H3K4me3 at gene promoters without impacting transcription initiation but did cause a decrease in transcription elongation [131]. These findings are in line with other evidence that suggests methylation of H3K4 and H3K79 is the result of transcription, not the driver of it [132]. The authors of this study show that the transcription elongation component PAF-1 recruits SET1/COMPASS and DOT1P, which in turn methylate H3K4 and H3K79, respectively. Taken as a whole, it appears that chromatin modifications in active regions are not likely to be marks which drive transcription, but rather are deposited as the result of transcription by Pol II. Moreover, it is well established that in very early development of Drosophila, cells lack major histone methylation marks, including H3K4me3 and H3K27me3 [119,133], yet transcription begins during early development, suggesting that these marks are not required to initiate transcription. Thus, it is questionable as to whether modified histones are capable of bookmarking active genes during DNA replication.
Epigenetic bookmarking by RNA Pol II during S phase
Both of these theories indicate that certain factors are required to bookmark chromatin during replication in order to restore the chromatin landscape, which will ultimately dictate which genes are repressed or active and therefore where Pol II should bind. Importantly, both theories assume that most proteins, including Pol II, are evicted by the replisome and must be recruited back to replicated DNA at a much later time point. This may not be the case given some of the in vitro data discussed above. Moreover, two recent studies in budding yeast have suggested that Pol II can rebind DNA shortly following replication [134,135]. The first study integrated knowledge of the order of ORI firing in budding yeast along with Pol II chromatin immunoprecipitation followed by sequencing (ChIP-seq) datasets to speculate that Pol II binds to both promoters and within gene bodies within 3 min following passage of the replisome [134]. However, given their approach, they were only able to postulate that Pol II displacement is transient within this system without fully confirming that theory. The second study in budding yeast adapted a technique called nascent chromatin avidin pull-down (NChAP) [136] by combining it with chromatin immunoprecipitation (ChIP). This approach, called ChIP-NChAP involves EdU labeling of nascent DNA followed by ChIP for a protein. The immunoprecipitated material is then subjected to click chemistry to conjugate biotin to EdU and is subsequently captured on streptavidin and analyzed by next-generation sequencing. Using ChIP-NChAP for Pol II the authors speculate that Pol II is able to quickly bind to replicated DNA [135]. Importantly, ChIP-NChAP is limited by a minimum of 20 min EdU labeling times and cannot assess events that happen immediately during passage of the replication fork. Nevertheless, both of these studies provided evidence suggesting that in budding yeast Pol II may only be removed from DNA during replication for a short period of time.
Similar results were obtained in a study in mouse embryonic stem cells, which showed that initiating Pol II S5P is absent from nascent DNA in 10 min, but rebinds DNA within 30 min after replication by chromatin occupancy after replication (ChOR-seq), an approach very similar to ChIP-NChAP [137]. A more recent study in human cells showed that total Pol II and Pol II S2P bind to DNA within 15 min of replication using ChOR-seq [138,139]. Given the inherent limitations of these techniques based on the minimal timings of EdU labeling, ChIP-NChAP and ChOR-seq are mostly informative to assess chromatin maturation from 15 to 20 min onward post-replication. Thus, they cannot be used to track proteins on DNA immediately following replication.
To assess this principal uncertainty in association of Pol II with the immediate nascent DNA, we recently investigated the timing of Pol II re-association with DNA after replication in human cells [139–141]. For this study, we utilized several techniques to assess events happening immediately following passage of the replication machinery. Using our previously developed chromatin assembly assay (CAA) [119], in which nascent DNA is EdU-labeled and subsequently conjugated with biotin and then a proximity ligation assay (PLA) is performed between biotin and a protein of interest, we found that initiating and elongating Pol II bind to DNA within 5 min of replication in human cells [140]. We then adapted our CAA approach to multiplex it, wherein two PLA reactions are performed within the same cell, in this case between biotin and Pol II in parallel with biotin and a replication protein, such as PCNA (both strands), FEN1 (lagging strand), or Pol ε (leading strand). Measuring the distance between foci in our multiplex CAA allowed us to determine that initiating and elongating Pol II bind nascent DNA within several hundred bp behind the replication fork on both the leading and lagging strands. These results were confirmed by a sequential ChIP (re-ChIP) approach [119], where Pol II ChIP material was re-immunoprecipitated using antibodies against replication proteins PCNA, FEN1, or Pol ε. These assays detected Pol II at promoters and within the gene body on both the leading and lagging strands within 100–200 bp after passage of the replication machinery. The same results were obtained with several general transcription and elongation factors, suggesting that both the initiating and elongating transcriptional complexes may quickly re-associate with nascent DNA [140].
In the same work, by using the only currently available method to experimentally assess resumption of transcription following DNA replication, a PLA-based approach called the RNA-DNA interaction assay (RDIA) [4,142], we also found that new transcription begins within 20 min after DNA replication. Surprisingly, using RDIA and adapting multiplex CAA, we also found that immature transcripts in the process of being synthesized before replication remain bound together with Pol II through DNA replication [140].
Taken as a whole, these results indicate that RNA-bound Pol II and other general transcription factors may only transiently be displaced from DNA during replication and rebind nascent chromatin throughout the gene body immediately following passage of the replisome (Figure 1(c)) [140]. Additional results of this study also suggested that new RNA synthesis may resume by completing synthesis of immature RNAs, as was implied by previous studies [93–96]. This indicates that a widely held assumption that the transcriptional complex may need some additional epigenetic marks during replication may not necessarily be the case.
Association of general and lineage specific transcription factors with nascent DNA
Recruitment of Pol II to a gene promoter involves the coordination of lineage-specific transcription factors (TFs) in addition to assembling the pre-initiation complex which consists of numerous GTFs [143]. In addition to Pol II, it is important to understand what is happening with these factors during DNA replication. Our data showed that the TATA-binding protein (TBP), a core GTF, is bound to DNA within several hundred bp of replisome passage [140]. We also detected other transcriptional components, including CDK9 and Cyclin T (components of the pTEF-b complex) as well as SPT4 and SPT5 (which form the DSIF complex) bound to nascent DNA within this timeframe [140]. Takentogether, this suggests that the core transcriptional apparatus as a whole may not be displaced during DNA replication.
Data regarding the presence of TFs on nascent DNA, however, are more limited, and whether these factors bind immediately after passage of the replication machinery remains to be fully understood. Current knowledge of these factors is limited to various DNA labeling-based approaches. For example, using nascent ChIP-seq (nasChIP-seq), an approach analogous to ChOR-seq, it was shown that the boundary element protein CTCF binds to DNA within 20 min of replication in human embryonic stem cells [144,145]. Additionally, proteomic assays, such as NCC (nascent chromatin capture) and adaptations of iPOND (isolation of proteins on nascent DNA) followed by mass spectrometry (MS), have been utilized to assess the protein composition of nascent chromatin shortly after DNA replication. Both NCC and iPOND label and capture nascent chromatin, and studies using these assays have shown that various TFs, along with some GTFs, core Pol II components and Pol II accessory factors are bound to nascent DNA labeled from 5 to 15 min [60,146–149]. This suggests that TFs and GTFs rebind DNA shortly after replication, although the precise kinetics of their binding remains to be determined.
Whether TFs are bound to nascent DNA has also been assessed through the use of various nuclease accessibility assays combined with the knowledge of where these TFs bind. These assays utilize micrococcal nuclease (MNase) in order to ascertain whether an area is protected from digestion, which would suggest it is bound by a nucleosome or TF based on the size of the protected fragment. By pairing this strategy with EdU labeling, biotinylation, and streptavidin capture, the structure of nascent chromatin has been investigated in yeast, Drosophila, and mouse ESCs [134,150–152]. These strategies gave varying results of whether factors are bound or not to nascent chromatin. These methods rely on assumptions about fragments protected from MNase digestion rather than directly assessing protein association with DNA, and they are limited by EdU labeling. Therefore, other methods, which allow assessing events occurring much closer to the replication fork, such as multiplex CAA and re-ChIP [119,140], need to be used in order to fully assess what is happening to TFs during DNA replication.
Potential mechanisms of Pol II retention on nascent DNA
In order for Pol II to rebind DNA after passage of the replisome, it likely needs to interact with component(s) of the replication machinery to facilitate its retention. We found that Pol II S5P and S2P interact with PCNA, suggesting a way to keep Pol II close to DNA during replication [140]. These results were functionally validated by affecting PCNA trimerization [140]. Another paper confirmed the existence of this interaction between PCNA and Pol II and their data suggest that it is mediated by the RPB1 component of Pol II [153]. Interestingly, other interactions between Pol II and replication proteins have been observed. The Pol II holoenzyme interacts with several MCM proteins and this interaction is mediated by the CTD of RPB1 [154,155]. Furthermore, MCM proteins were shown to interact with the Integrator complex, suggesting additional links between these helicase components and transcription proteins [89]. Interaction between MCM7 and the yeast TFIIH complex have also been reported [156]. Additionally, hyperphosphorylated Pol II interacts with the leading strand DNA polymerase, Pol ε, but not Pol δ or Pol α [157]. Taken as a whole, these two complexes interact via numerous components of each.
The mechanism of retention and transcription resumption midway through the gene is still not fully resolved. In order for the resumption of transcription to occur, as suggested for some genes [93,94], Pol II must reengage DNA and form a transcription bubble. The mechanism for this is unclear, but potentially involves recruitment of helicase(s), such as RECQ5, Senataxin (SETX), and/or RTEL1, which have all been observed to function during DNA replication. RECQ5 has been shown to be both associated with Pol II during transcription, as well as during DNA replication [158–161]. Moreover, RECQ5 has been implicated for its role in resolving TRCs, by promoting PCNA ubiquitination and potentially allowing Pol II passage [162]. The RNA-DNA helicase SETX has also been reported to be involved in transcription elongation and to associate with replication forks [163,164]. SETX is believed to play a role in allowing for replication progression by acting on R-loops formed during transcription [165,166]. Finally, the helicase RTEL1, in complex with SLX4, associates with Pol II and to bind to nascent DNA [167], suggesting that it may also play a role in facilitating Pol II function on nascent DNA. Taken as a whole, any of these helicases may function in assisting Pol II to resume transcription after DNA replication.
Conclusions
Given the focus in the field on TRCs, our recent findings that transcription may only be transiently displaced by the replisome is striking. That being said, that Pol II bound to an immature transcript is transferred during replication is the only way to explain how some transcripts, such as long ones described by Helmrich et al. [93], or transcripts synthesized during short cell cycles in early development [95] are able to be completed. Moreover, ChOR-seq for Pol II S2P in human cells found that Pol II was bound in the distal regions of genes that take more than 15 min to transcribe, further suggesting that the elongating transcriptional apparatus is likely not ejected during DNA replication [138]. These conclusions also suggest a way in which higher level eukaryotes, are able to replicate DNA in a vast number of different cell types, each of which contains different transcriptional profiles and utilizes cell-specific ORIs.
Given our recent finding that in human cells Pol II is retained in close vicinity of the replication fork and quickly re-associates following passage of the replisome, several future questions need to be addressed. Numerous studies have indicated that replication proteins including PCNA, several MCM proteins, and DNA pol ε have been shown to interact with the RNA Pol II complex, and a better understanding of the nature of these interactions is needed, particularly the specific domains making contact. This would better explain the mechanism by which RNA Pol II is able to re-bind DNA immediately following passage of the replisome. Additionally, Pol II would need to reform the transcription bubble on nascent DNA and this would likely require a helicase. As discussed, there are several candidates, RECQ5, SETX, and RTEL1, whose function could be targeted pharmacologically or genetically in order to assess what happens to RNA Pol II detection on nascent DNA.
Further work will need to be done to understand the nature of these interaction and the detailed mechanism by which Pol II is able to be transferred during DNA replication. Given that RNA polymerase travels faster than DNA polymerase different scenarios might play out depending on whether the polymerases are positioned co-directionally or head-on, since RNA polymerase would approach a different component of replication machinery depending on the direction [141]. Besides the intricacies in interactions between transcription and replication complexes, such studies would provide new insights into a previously unknown mechanism of epigenetic bookmarking during S phase.
Acknowledgments
We apologize to colleagues whose work we were unable to cite due to space considerations.
Funding Statement
This work was supported by NIH F31GM128300 to T.K.F. and NIH R01GM075141 to A.M.
Disclosure statement
No potential conflict of interest was reported by the author(s).
References
- [1].Crick F. Central dogma of molecular biology. Nature. 1970;227(5258):561–563. doi: 10.1038/227561a0 [DOI] [PubMed] [Google Scholar]
- [2].Parsons GG, Spencer CA. Mitotic repression of RNA polymerase II transcription is accompanied by release of transcription elongation complexes. Mol Cell Biol. 1997;17:5791–5802. doi: 10.1128/MCB.17.10.5791 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [3].Liang K, Woodfin AR, Slaughter BD. Mitotic transcriptional activation: clearance of actively wngaged Pol II via transcriptional elongation control in mitosis. Mol Cell. 2015;60(3):435–445. doi: 10.1016/j.molcel.2015.09.021 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [4].Petruk S, Fenstermaker TK, Black KL. Detection of RNA-DNA association by a proximity ligation-based method. Sci Rep. 2016;6(1):6–11. doi: 10.1038/srep27313 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [5].Wang J, Rojas P, Mao J et al. Persistence of RNA transcription during DNA replication delays duplication of transcription start sites until G2/M. Cell Rep. 2021;34:108759. doi: 10.1016/j.celrep.2021.108759 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [6].Palozola KC, Liu H, Nicetto D. Low-level, global transcription during mitosis and dynamic gene reactivation during mitotic exit. Cold Spring Harb Symp Quant Biol. 2017;82:197–205. doi: 10.1101/sqb.2017.82.034280 [DOI] [PubMed] [Google Scholar]
- [7].Palozola KC, Donahue G, Liu H. Mitotic transcription and waves of gene reactivation during mitotic exit. Science. 2017;358(6359):119–122. doi: 10.1126/science.aal4671 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [8].Bertoli C, Skotheim JM, de Bruin RAM. Control of cell cycle transcription during G1 and S phases. Nat Rev Mol Cell Biol. 2013;14(8):518–528. doi: 10.1038/nrm3629 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [9].Stewart-Morgan KR, Petryk N, Groth A. Chromatin replication and epigenetic cell memory. Nat Cell Biol. 2020;22(4):361–371. doi: 10.1038/s41556-020-0487-y [DOI] [PubMed] [Google Scholar]
- [10].Gottesfeld JM, Forbes DJ. Mitotic repression of the transcriptional machinery. Trends Biochem Sci. 1997;22(6):197–202. doi: 10.1016/s0968-0004(97)01045-1 [DOI] [PubMed] [Google Scholar]
- [11].Kadauke S, Blobel GA. Mitotic bookmarking by transcription factors. Epigenet Chromatin. 2013;6(1):6. doi: 10.1186/1756-8935-6-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [12].McGlynn P, Savery NJ, Dillingham MS. The conflict between DNA replication and transcription. Mol Microbiol. 2012;85:12–20. doi: 10.1111/j.1365-2958.2012.08102.x [DOI] [PubMed] [Google Scholar]
- [13].Brewer BJ. When polymerases collide: replication and the transcriptional organization of the E. coli chromosome. Cell. 1988;53(5):679–686. doi: 10.1016/0092-8674(88)90086-4 [DOI] [PubMed] [Google Scholar]
- [14].Srivatsan A, Tehranchi A, MacAlpine DM, et al. Co-orientation of replication and transcription preserves genome integrity. PLoS Genet. 2010;6(1):e1000810. doi: 10.1371/journal.pgen.1000810 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [15].Lang KS, Merrikh H. The clash of macromolecular titans: replication-transcription conflicts in bacteria. Annu Rev Microbiol. 2018;72:71–88. doi: 10.1146/annurev-micro-090817-062514 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [16].Bell SP, Stillman B. ATP-dependent recognition of eukaryotic origins of DNA replication by a multiprotein complex. Nature. 1992;357(6374):128–134. doi: 10.1038/357128a0 [DOI] [PubMed] [Google Scholar]
- [17].Kriegstein HJ, Hogness DS. Mechanism of DNA replication in Drosophila chromosomes: structure of replication forks and evidence for bidirectionality. Proc Natl Acad Sci U S A. 1974;71:135–139. doi: 10.1073/pnas.71.1.135 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [18].Yeeles JTP, Deegan TD, Janska A, et al. Regulated eukaryotic DNA replication origin firing with purified proteins. Nature. 2015;519(7544):431–435. doi: 10.1038/nature14285 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [19].Seki T, Diffley JFX. Stepwise assembly of initiation proteins at budding yeast replication origins in vitro. Proc Natl Acad Sci U S A. 2000;97:14115–14120. doi: 10.1073/pnas.97.26.14115 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [20].Jain R, Aggarwal AK, Rechkoblit O. Eukaryotic DNA polymerases. Curr Opin Struct Biol. 2018;53:77–87. doi: 10.1016/j.sbi.2018.06.003 [DOI] [PubMed] [Google Scholar]
- [21].Waga S, Stillman B. Anatomy of a DNA replication fork revealed by reconstitution of SV40 DNA replication in vitro. Nature. 1994;369(6477):207–212. doi: 10.1038/369207a0 [DOI] [PubMed] [Google Scholar]
- [22].Prelich G, Tan CK, Kostura M, et al. Functional identity of proliferating cell nuclear antigen and a DNA polymerase-δ auxiliary protein. Nature. 1987;326(6112):517–520. doi: 10.1038/326517a0 [DOI] [PubMed] [Google Scholar]
- [23].Gomes XV, Schmidt SLG, Burgers PMJ. ATP utilization by yeast replication factor C. II. Multiple stepwise ATP binding events are required to load proliferating cell nuclear antigen onto primed DNA. J Biol Chem. 2001;276:34776–34783. doi: 10.1074/jbc.M011743200 [DOI] [PubMed] [Google Scholar]
- [24].Okazaki R, Okazaki T, Sakabe K, et al. Mechanism of DNA chain growth. I. Possible discontinuity and unusual secondary structure of newly synthesized chains. Proc Natl Acad Sci U S A. 1968;59:598–605. doi: 10.1073/pnas.59.2.598 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [25].Sugimoto K, Okazaki T, Okazaki R. Mechanism of DNA chain growth, II. Accumulation of newly synthesized short chains in E. coli infected with ligase-defective T4 phages. Proc Natl Acad Sci U S A. 1968;60:1356–1362. doi: 10.1073/pnas.60.4.1356 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [26].Lujan SA, Williams JS, Pursell ZF, et al. Mismatch repair balances leading and lagging strand DNA replication fidelity. PloS Genet. 2012;8(10):e1003016. doi: 10.1371/journal.pgen.1003016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [27].Lujan SA, Williams JS, Clausen AR, et al. Ribonucleotides are signals for mismatch repair of leading-strand replication errors. Mol Cell. 2013;50(3):437–443. doi: 10.1016/j.molcel.2013.03.017 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [28].Stodola JL, Burgers PM. Resolving individual steps of Okazaki-fragment maturation at a millisecond timescale. Nat Struct Mol Biol. 2016;23(5):402–408. doi: 10.1038/nsmb.3207 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [29].Turchi JJ, Bambara RA. Completion of mammalian lagging strand DNA replication using purified proteins. J Biol Chem. 1993;268:15136–15141. doi: 10.1016/s0021-9258(18)82447-4 [DOI] [PubMed] [Google Scholar]
- [30].Roeder RG, Rutter WJ. Multiple forms of DNA-dependent RNA polymerase in eukaryotic organisms. Nature. 1969;224(5216):234–237. doi: 10.1038/224234a0 [DOI] [PubMed] [Google Scholar]
- [31].Russell J, Zomerdijk JCBM. The RNA polymerase I transcription machinery. Biochem Soc Symp. 2006;73:203–216. doi: 10.1042/bss0730203 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [32].Reeder RH, Roeder RG. Ribosomal RNA synthesis in isolated nuclei. J Mol Biol. 1972;67:433–441. doi: 10.1016/0022-2836(72)90461-5 [DOI] [PubMed] [Google Scholar]
- [33].Weinmann R, Roeder RG. Role of DNA-dependent RNA polymerase 3 in the transcription of the tRNA and 5S RNA genes. Proc Natl Acad Sci USA. 1974;71(5):1790–1794. doi: 10.1073/pnas.71.5.1790 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [34].Lee Y, Kim M, Han J, et al. MicroRNA genes are transcribed by RNA polymerase II. EMBO J. 2004;23(20):4051–4060. doi: 10.1038/sj.emboj.7600385 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [35].Li XZ, Roy CK, Dong X, et al. An ancient transcription factor initiates the burst of piRNA production during early meiosis in mouse testes. Mol Cell. 2013;50(1):67–81. doi: 10.1016/j.molcel.2013.02.016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [36].Guttman M, Amit I, Garber M, et al. Chromatin signature reveals over a thousand highly conserved large non-coding RNAs in mammals. Nature. 2009;458(7235):223–227. doi: 10.1038/nature07672 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [37].Jawdekar GW, Henry RW. Transcriptional regulation of human small nuclear RNA genes. Biochim Biophys Acta–Gene Regul Mech. 2008;1779:295–305. doi: 10.1016/j.bbagrm.2008.04.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [38].Li B, Carey M, Workman JL. The role of chromatin during transcription. Cell. 2007;128(4):707–719. doi: 10.1016/j.cell.2007.01.015 [DOI] [PubMed] [Google Scholar]
- [39].Roeder RG. The eukaryotic transcriptional machinery: complexities and mechanisms unforeseen. Nat Med. 2003;9(10):1239–1244. doi: 10.1038/nm938 [DOI] [PubMed] [Google Scholar]
- [40].Cramer P. Organization and regulation of gene transcription. Nature. 2019;573(7772):45–54. doi: 10.1038/s41586-019-1517-4 [DOI] [PubMed] [Google Scholar]
- [41].Abuhashem A, Garg V, Hadjantonakis A-K. RNA polymerase II pausing in development: orchestrating transcription. Open Biol. 2022;12(1):210220. doi: 10.1098/rsob.210220 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [42].Wang D, Bushnell DA, Westover KD, et al. Structural basis of transcription: role of the trigger loop in substrate specificity and catalysis. Cell. 2006;127(5):941–954. doi: 10.1016/j.cell.2006.11.023 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [43].Cramer P, Bushnell DA, Kornberg RD. Structural basis of transcription: RNA polymerase II at 2.8 Ångstrom resolution. Science. 2001;292(5523):1863–1876. doi: 10.1126/science.1059493 [DOI] [PubMed] [Google Scholar]
- [44].Phatnani HP, Greenleaf AL. Phosphorylation and functions of the RNA polymerase II CTD. Genes Dev. 2006;20(21):2922–2936. doi: 10.1101/gad.1477006 [DOI] [PubMed] [Google Scholar]
- [45].Komarnitsky P, Cho EJ, Buratowski S. Different phosphorylated forms of RNA polymerase II and associated mRNA processing factors during transcription. Genes Dev. 2000;14(19):2452–2460. doi: 10.1101/gad.824700 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [46].Morris DP, Michelotti GA, Schwinn DA. Evidence that phosphorylation of the RNA polymerase II carboxyl-terminal repeats is similar in yeast and humans. J Biol Chem. 2005;280:31368–31377. doi: 10.1074/jbc.M501546200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [47].Matsui T, Segall J, Weil PA, et al. Multiple factors required for accurate initiation of transcription by purified RNA polymerase II. J Biol Chem. 1980;255:11992–11996. doi: 10.1016/S0021-9258(19)70232-4 [DOI] [PubMed] [Google Scholar]
- [48].Roeder RG. The complexities of eukaryotic transcription initiation: regulation of preinitiation complex assembly. Trends Biochem Sci. 1991;16:402–408. doi: 10.1016/0968-0004(91)90164-q [DOI] [PubMed] [Google Scholar]
- [49].Roeder RG. The role of general initiation factors in transcription by RNA polymerase II. Trends Biochem Sci. 1996;21:327–335. doi: 10.1016/0968-0004(96)10050-5 [DOI] [PubMed] [Google Scholar]
- [50].Buratowski S, Hahn S, Guarente L, et al. Five intermediate complexes in transcription initiation by RNA polymerase II. Cell. 1989;56(4):549–561. doi: 10.1016/0092-8674(89)90578-3 [DOI] [PubMed] [Google Scholar]
- [51].Liu B, Wong ML, Tinker RL, et al. The DNA replication fork can pass RNA polymerase without displacing the nascent transcript. Nature. 1993;366(6450):33–39. doi: 10.1038/366033a0 [DOI] [PubMed] [Google Scholar]
- [52].Liu B, Wong ML, Alberts B. A transcribing RNA polymerase molecule survives DNA replication without aborting its growing RNA chain. Proc Natl Acad Sci U S A. 1994;91:10660–10664. doi: 10.1073/pnas.91.22.10660 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [53].Liu B, Alberts BM. Head-on collision between a DNA replication apparatus and RNA polymerase transcription complex. Science. 1995;267(5201):1131–1137. doi: 10.1126/science.7855590 [DOI] [PubMed] [Google Scholar]
- [54].Elías-Arnanz M. Resolution of head-on collisions between the transcription machinery and bacteriophage phi29 DNA polymerase is dependent on RNA polymerase translocation. EMBO J. 1999;18(20):5675–5682. doi: 10.1093/emboj/18.20.5675 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [55].Elías-Arnanz M, Salas M. Bacteriophage phi29 DNA replication arrest caused by codirectional collisions with the transcription machinery. The EMBO Journal. 1997;16(18):5775–5783. doi: 10.1093/emboj/16.18.5775 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [56].Pomerantz RT, O’Donnell M. The replisome uses mRNA as a primer after colliding with RNA polymerase. Nature. 2008;456(7223):762–766. doi: 10.1038/nature07527 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [57].Pomerantz RT, O’Donnell M. Direct restart of a replication fork stalled by a head-on RNA polymerase. Science. 2010;327(5965):590–592. doi: 10.1126/science.1179595 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [58].Scherr MJ, Wahab SA, Remus D, et al. Mobile origin-licensing factors confer resistance to conflicts with RNA polymerase. Cell Rep. 2022;38(12):110531. doi: 10.1016/j.celrep.2022.110531 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [59].Gros J, Kumar C, Lynch G, et al. Post-licensing specification of eukaryotic replication origins by facilitated Mcm2-7 sliding along DNA. Mol Cell. 2015;60(5):797–807. doi: 10.1016/j.molcel.2015.10.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [60].Lopez-Contreras AJ, Ruppen I, Nieto-Soler M, et al. A proteomic characterization of factors enriched at nascent DNA molecules. Cell Rep. 2013;3(4):1105–1116. doi: 10.1016/j.celrep.2013.03.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [61].Iqbal MA, Chinsky J, Didamo V, et al. Replication of proto-oncogenes early during the S phase in mammalian cell lines. Nucleic Acids Res. 1987;15(1):87–103. doi: 10.1093/nar/15.1.87 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [62].Hatton KS, Dhar V, Brown EH, et al. Replication program of active and inactive multigene families in mammalian cells. Mol Cell Biol. 1988;8(5):2149–2158. doi: 10.1128/mcb.8.5.2149-2158.1988 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [63].Furst A, Brown EH, Braunstein JD, et al. α-globulin sequences are located in a region of early-replicating DNA in murine erythroleukemia cells. Proc Natl Acad Sci U S A. 1981;78:1023–1027. doi: 10.1073/pnas.78.2.1023 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [64].Goldman MA, Holmquist GP, Gray MC, et al. Replication timing of genes and middle repetitive sequences. Science. 1984;224(4650):686–692. doi: 10.1126/science.6719109 [DOI] [PubMed] [Google Scholar]
- [65].Sequeira-Mendes J, Díaz-Uriarte R, Apedaile A, et al. Transcription initiation activity sets replication origin efficiency in mammalian cells. PLoS Genet. 2009;5(4):e1000446. doi: 10.1371/journal.pgen.1000446 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [66].Cayrou C, Coulombe P, Vigneron A, et al. Genome-scale analysis of metazoan replication origins reveals their organization in specific but flexible sites defined by conserved features. Genome Res. 2011;21:1438–1449. doi: 10.1101/gr.121830.111 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [67].Iguchi-Ariga SM, Okazaki T, Itani T, et al. An initiation site of DNA replication with transcriptional enhancer activity present upstream of the c-myc gene. EMBO J. 1988;7:3135–3142. doi: 10.1002/j.1460-2075.1988.tb03180.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- [68].Kitsberg D, Selig S, Keshet I, et al. Replication structure of the human beta-globin gene domain. Nature. 1993;366(6455):588–590. doi: 10.1038/366588a0 [DOI] [PubMed] [Google Scholar]
- [69].Lunyak VV, Ezrokhi M, Smith HS, et al. Developmental changes in the sciara II/9A initiation zone for DNA replication. Mol Cell Biol. 2002;22(24):8426–8437. doi: 10.1128/MCB.22.24.8426-8437.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [70].Hiratani I, Ryba T, Itoh M, et al. Global reorganization of replication domains during embryonic stem cell differentiation. PLoS Biol. 2008;6(10):e245. doi: 10.1371/journal.pbio.0060245 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [71].Bodmer-Glavas M, Edler K, Barberis A. RNA polymerase II and III transcription factors can stimulate DNA replication by modifying origin chromatin structures. Nucleic Acids Research. 2001;29(22):4570–4580. doi: 10.1093/nar/29.22.4570 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [72].Stagljar I, Hübscher U, Barberis A. Activation of DNA replication in yeast by recruitment of the RNA polymerase II transcription complex. Biol Chem. 1999;380:525–530. doi: 10.1515/BC.1999.067 [DOI] [PubMed] [Google Scholar]
- [73].Li S, Wasserman MR, Yurieva O, et al. Nucleosome-directed replication origin licensing independent of a consensus DNA sequence. Nat Commun. 2022;13(1):4947. doi: 10.1038/s41467-022-32657-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [74].Lee CSK, Weiβ M, Hamperl S. Where and when to start: regulating DNA replication origin activity in eukaryotic genomes. Nucleus. 2023;14(1):2229642. doi: 10.1080/19491034.2023.2229642 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [75].St Germain C, Zhao H, Barlow JH. Transcription-replication collisions—A series of unfortunate events. Biomolecules. 2021;11(8):1249. doi: 10.3390/biom11081249 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [76].Brambati A, Colosio A, Zardoni L, et al. Replication and transcription on a collision course: eukaryotic regulation mechanisms and implications for DNA stability. Front Genet. 2015;6:166. doi: 10.3389/fgene.2015.00166 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [77].Knott SRV, Viggiani CJ, Aparicio OM. To promote and protect: coordinating DNA replication and transcription for genome stability. Epigenetics. 2009;4(6):362–365. doi: 10.4161/epi.4.6.9712 [DOI] [PubMed] [Google Scholar]
- [78].Bermejo R, Lai MS, Foiani M. Preventing replication stress to maintain genome stability: resolving conflicts between replication and transcription. Mol Cell. 2012;45:710–718. doi: 10.1016/j.molcel.2012.03.001 [DOI] [PubMed] [Google Scholar]
- [79].García-Muse T, Aguilera A. Transcription–replication conflicts: how they occur and how they are resolved. Nat Rev Mol Cell Biol. 2016;17(9):553–563. doi: 10.1038/nrm.2016.88 [DOI] [PubMed] [Google Scholar]
- [80].Goehring L, Huang TT, Smith DJ. Transcription–replication conflicts as a source of genome instability. Annu Rev Genet. 2023;57(1):157–179. doi: 10.1146/annurev-genet-080320-031523 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [81].Niehrs C, Luke B. Regulatory R-loops as facilitators of gene expression and genome stability. Nat Rev Mol Cell Biol. 2020;21(3):167–178. doi: 10.1038/s41580-019-0206-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [82].Poli J, Gerhold C-B, Tosi A, et al. Mec1, INO80, and the PAF1 complex cooperate to limit transcription replication conflicts through RNAPII removal during replication stress. Genes Dev. 2016;30:337–354. doi: 10.1101/gad.273813.115 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [83].Hoffman EA, McCulley A, Haarer B, et al. Break-seq reveals hydroxyurea-induced chromosome fragility as a result of unscheduled conflict between DNA replication and transcription. Genome Res. 2015;25:402–412. doi: 10.1101/gr.180497.114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [84].Shao X, Joergensen AM, Howlett NG, et al. A distinct role for recombination repair factors in an early cellular response to transcription–replication conflicts. Nucleic Acids Res. 2020;48(10):5467–5484. doi: 10.1093/nar/gkaa268 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [85].Hristova RH, Stoynov SS, Tsaneva IR, et al. Deregulated levels of RUVBL1 induce transcription-dependent replication stress. Int J Biochem Cell Biol. 2020;128:105839. doi: 10.1016/j.biocel.2020.105839 [DOI] [PubMed] [Google Scholar]
- [86].Bermejo R, Capra T, Gonzalez-Huici V, et al. Genome-organizing factors Top2 and Hmo1 prevent chromosome fragility at sites of S phase transcription. Cell. 2009;138(5):870–884. doi: 10.1016/j.cell.2009.06.022 [DOI] [PubMed] [Google Scholar]
- [87].Schwab RA, Nieminuszczy J, Shah F, et al. The fanconi anemia pathway maintains genome stability by coordinating replication and transcription. Mol Cell. 2015;60(3):351–361. doi: 10.1016/j.molcel.2015.09.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [88].Felipe-Abrio I, Lafuente-Barquero J, García-Rubio ML, et al. RNA polymerase II contributes to preventing transcription-mediated replication fork stalls. EMBO J. 2015;34:236–250. doi: 10.15252/embj.201488544 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [89].Bhowmick R, Mehta KPM, Lerdrup M, et al. Integrator facilitates RNAPII removal to prevent transcription-replication collisions and genome instability. Mol Cell. 2023;83(13):2357–2366.e8. doi: 10.1016/j.molcel.2023.05.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [90].Chen Y-H, Keegan S, Kahli M, et al. Transcription shapes DNA replication initiation and termination in human cells. Nat Struct Mol Biol. 2019;26(1):67–77. doi: 10.1038/s41594-018-0171-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [91].Minocherhomji S, Ying S, Bjerregaard VA, et al. Replication stress activates DNA repair synthesis in mitosis. Nature. 2015;528(7581):286–290. doi: 10.1038/nature16139 [DOI] [PubMed] [Google Scholar]
- [92].Bhowmick R, Minocherhomji S, Hickson ID. RAD52 facilitates mitotic DNA synthesis following replication stress. Mol Cell. 2016;64:1117–1126. doi: 10.1016/j.molcel.2016.10.037 [DOI] [PubMed] [Google Scholar]
- [93].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. doi: 10.1016/j.molcel.2011.10.013 [DOI] [PubMed] [Google Scholar]
- [94].Helmrich A, Ballarino M, Nudler E, et al. Transcription-replication encounters, consequences and genomic instability. Nat Struct Mol Biol. 2013;20(4):412–418. doi: 10.1038/nsmb.2543 [DOI] [PubMed] [Google Scholar]
- [95].Shermoen AW, O’Farrell PH. Progression of the cell cycle through mitosis leads to abortion of nascent transcripts. Cell. 1991;67(2):303–310. doi: 10.1016/0092-8674(91)90182-x [DOI] [PMC free article] [PubMed] [Google Scholar]
- [96].Yunger S, Rosenfeld L, Garini Y, et al. Single-allele analysis of transcription kinetics in living mammalian cells. Nat Methods. 2010;7(8):631–633. doi: 10.1038/nmeth.1482 [DOI] [PubMed] [Google Scholar]
- [97].Anachkova B, Djeliova V, Russev G. Nuclear matrix support of DNA replication. J Cell Biochem. 2005;96:951–961. doi: 10.1002/jcb.20610 [DOI] [PubMed] [Google Scholar]
- [98].Sutherland H, Bickmore WA. Transcription factories: gene expression in unions? Nat Rev Genet. 2009;10(7):457–466. doi: 10.1038/nrg2592 [DOI] [PubMed] [Google Scholar]
- [99].Natsume T, Tanaka TU. Spatial regulation and organization of DNA replication within the nucleus. Chromosome Res. 2010;18(1):7–17. doi: 10.1007/s10577-009-9088-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [100].Razin SV, Gavrilov AA, Pichugin A, et al. Transcription factories in the context of the nuclear and genome organization. Nucleic Acids Res. 2011;39:9085–9092. doi: 10.1093/nar/gkr683 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [101].Rieder D, Trajanoski Z, McNally JG. Transcription factories. Front Genet. 2012;3:620–622. doi: 10.3389/fgene.2012.00221 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [102].Jackson DA, Iborra FJ, Manders EMM, et al. Numbers and organization of RNA polymerases, nascent transcripts, and transcription units in HeLa nuclei. Mol Biol Cell. 1998;9(6):1523–1536. doi: 10.1091/mbc.9.6.1523 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [103].Iborra FJ, Pombo A, McManus J, et al. The topology of transcription by immobilized polymerases. Exp Cell Res. 1996;229:167–173. doi: 10.1006/excr.1996.0355 [DOI] [PubMed] [Google Scholar]
- [104].Hozák P, Hassan AB, Jackson DA, et al. Visualization of replication factories attached to a nucleoskeleton. Cell. 1993;73(2):361–373. doi: 10.1016/0092-8674(93)90235-i [DOI] [PubMed] [Google Scholar]
- [105].Hassan AB, Errington RJ, White NS, et al. Replication and transcription sites are colocalized in human cells. J Cell Sci. 1994;107:425–434. doi: 10.1242/jcs.107.2.425 [DOI] [PubMed] [Google Scholar]
- [106].Martin S, Pombo A. Transcription factories: quantitative studies of nanostructures in the mammalian nucleus. Chromosome Res. 2003;11:461–470. doi: 10.1023/a:1024926710797 [DOI] [PubMed] [Google Scholar]
- [107].Hozák P, Cook PR. Replication factories. Trends Cell Biol. 1994;4:48–52. doi: 10.1016/0962-8924(94)90009-4 [DOI] [PubMed] [Google Scholar]
- [108].Jackson DA, Cook PR. Replication occurs at a nucleoskeleton. EMBO J. 1986;5:1403–1410. doi: 10.1002/j.1460-2075.1986.tb04374.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- [109].Pardoll DM, Vogelstein B, Coffey DS. A fixed site of DNA replication in eucaryotic cells. Cell. 1980;19(2):527–536. doi: 10.1016/0092-8674(80)90527-9 [DOI] [PubMed] [Google Scholar]
- [110].Comings DE, Kakefuda T. Initiation of deoxyribonucleic acid replication at the nuclear membrane in human cells. J Mol Biol. 1968;33:225–229. doi: 10.1016/0022-2836(68)90290-8 [DOI] [PubMed] [Google Scholar]
- [111].Papantonis A, Cook PR. Fixing the model for transcription. Transcription. 2011;2(1):41–44. doi: 10.4161/trns.2.1.14275 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [112].Papantonis A, Larkin JD, Wada Y, et al. Active RNA polymerases: mobile or immobile molecular machines? PLoS Biol. 2010;8(7):e1000419. doi: 10.1371/journal.pbio.1000419 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [113].Jackson DA, McCready SJ, Cook PR. Replication and transcription depend on attachment of DNA to the nuclear cage. J of Cell Science. 1984;1(Supplement_1):59–79. doi: 10.1242/jcs.1984.supplement_1.5 [DOI] [PubMed] [Google Scholar]
- [114].Jackson DA, Hassan AB, Errington RJ, et al. Visualization of focal sites of transcription within human nuclei. EMBO J. 1993;12:1059–1065. doi: 10.1002/j.1460-2075.1993.tb05747.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- [115].Wessel R, Schweizer J, Stahl H. Simian virus 40 T-antigen DNA helicase is a hexamer which forms a binary complex during bidirectional unwinding from the viral origin of DNA replication. J Virol. 1992;66(2):804–815. doi: 10.1128/JVI.66.2.804-815.1992 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [116].Cook PR, Gove F. Transcription by an immobilized RNA polymerase from bacteriophage T7 and the topology of transcription. Nucl Acids Res. 1992;20(14):3591–3598. doi: 10.1093/nar/20.14.3591 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [117].Yin H, Wang MD, Svoboda K, et al. Transcription against an applied force. Science. 1995;270(5242):1653–1657. doi: 10.1126/science.270.5242.1653 [DOI] [PubMed] [Google Scholar]
- [118].Almouzni G, Cedar H. Maintenance of epigenetic information. Cold Spring Harb Perspect Biol. 2016;8(5):a019372. doi: 10.1101/cshperspect.a019372 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [119].Petruk S, Sedkov Y, Johnston DM, et al. TrxG and PcG proteins but not methylated histones remain associated with DNA through replication. Cell. 2012;150(5):922–933. doi: 10.1016/j.cell.2012.06.046 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [120].Petruk S, Black KL, Kovermann SK. Stepwise histone modifications are mediated by multiple enzymes that rapidly associate with nascent DNA during replication. Nat Commun. 2013;4(1):2841. doi: 10.1038/ncomms3841 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [121].Abmayr SM, Workman JL. Holding on through DNA replication: histone modification or modifier? Cell. 2012;150(5):875–877. doi: 10.1016/j.cell.2012.08.006 [DOI] [PubMed] [Google Scholar]
- [122].Escobar TM, Yu J-R, Liu S, et al. Inheritance of repressed chromatin domains during S phase requires the histone chaperone NPM1. Sci Adv. 2022;8(17):eabm3945. doi: 10.1126/sciadv.abm3945 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [123].Estève P-O, Chin HG, Smallwood A, et al. Direct interaction between DNMT1 and G9a coordinates DNA and histone methylation during replication. Genes Dev. 2006;20:3089–3103. doi: 10.1101/gad.1463706 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [124].Escobar TM, Oksuz O, Saldaña-Meyer R, et al. Active and repressed chromatin domains exhibit distinct nucleosome segregation during DNA replication. Cell. 2019;179(4):953–963.e11. doi: 10.1016/j.cell.2019.10.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [125].Escobar TM, Loyola A, Reinberg D. Parental nucleosome segregation and the inheritance of cellular identity. Nat Rev Genet. 2021;22(6):379–392. doi: 10.1038/s41576-020-00312-w [DOI] [PMC free article] [PubMed] [Google Scholar]
- [126].Henikoff S, Shilatifard A. Histone modification: Cause or cog? Trends in Genetics. 2011;27(10):389–396. doi: 10.1016/j.tig.2011.06.006 [DOI] [PubMed] [Google Scholar]
- [127].Reinberg D, Vales LD. Chromatin domains rich in inheritance. Science. 2018;361(6397):33–34. doi: 10.1126/science.aat7871 [DOI] [PubMed] [Google Scholar]
- [128].Howe FS, Fischl H, Murray SC, et al. Is H3K4me3 instructive for transcription activation? BioEssays. 2017;39(1):1–12. doi: 10.1002/bies.201600095 [DOI] [PubMed] [Google Scholar]
- [129].Clouaire T, Webb S, Skene P, et al. Cfp1 integrates both CpG content and gene activity for accurate H3K4me3 deposition in embryonic stem cells. Genes Dev. 2012;26:1714–1728. doi: 10.1101/gad.194209.112 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [130].Margaritis T, Oreal V, Brabers N, et al. Two distinct repressive mechanisms for histone 3 lysine 4 methylation through promoting 3'-end antisense transcription. PLoS Genet. 2012;8(9):e1002952. doi: 10.1371/journal.pgen.1002952 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [131].Wang H, Fan Z, Shliaha PV, et al. H3K4me3 regulates RNA polymerase II promoter-proximal pause-release. Nature. 2023;615(7951):339–348. doi: 10.1038/s41586-023-05780-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [132].Krogan NJ, Dover J, Wood A, et al. The Paf1 complex is required for histone H3 methylation by COMPASS and Dot1p: linking transcriptional elongation to histone methylation. Mol Cell. 2003;11(3):721–729. doi: 10.1016/s1097-2765(03)00091-1 [DOI] [PubMed] [Google Scholar]
- [133].Chen K, Johnston J, Shao W, et al. A global change in RNA polymerase II pausing during the Drosophila midblastula transition. Elife. 2013;2:e00861. doi: 10.7554/eLife.00861 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [134].Bar-Ziv R, Brodsky S, Chapal M, et al. Transcription factor binding to replicated DNA. Cell Rep. 2020;30(12):3989–3995.e4. doi: 10.1016/j.celrep.2020.02.114 [DOI] [PubMed] [Google Scholar]
- [135].Ziane R, Camasses A, Radman-Livaja M. The asymmetric distribution of RNA polymerase II and nucleosomes on replicated daughter genomes is caused by differences in replication timing between the lagging and the leading strand. Genome Res. 2022;32(2):337–356. doi: 10.1101/gr.275387.121 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [136].Vasseur P, Tonazzini S, Ziane R, et al. Dynamics of nucleosome positioning maturation following genomic replication. Cell Rep. 2016;16:2651–2665. doi: 10.1016/j.celrep.2016.07.083 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [137].Stewart-Morgan KR, Reverón-Gómez N, Groth A. Transcription restart establishes chromatin accessibility after DNA replication. Molecular Cell. 2019;75(2):284–297.e6. doi: 10.1016/j.molcel.2019.04.033 [DOI] [PubMed] [Google Scholar]
- [138].Bruno F, Coronel-Guisado C, González-Aguilera, C. Collisions of RNA polymerases behind the replication fork promote alternative RNA splicing in newly replicated chromatin. Mol Cell. 2023;84(2):221–233. doi: 10.1016/j.molcel.2023.11.036 [DOI] [PubMed] [Google Scholar]
- [139].Werner M, Hamperl S. A quick restart: RNA polymerase jumping onto post-replicative chromatin. Mol Cell. 2024;84:186–188. doi: 10.1016/j.molcel.2023.12.029 [DOI] [PubMed] [Google Scholar]
- [140].Fenstermaker TK, Petruk S, Kovermann SK, et al. RNA polymerase II associates with active genes during DNA replication. Nature. 2023;620(7973):426–433. doi: 10.1038/s41586-023-06341-9 [DOI] [PubMed] [Google Scholar]
- [141].Yakoub G, Luijsterburg MS. Transcription and replication: walking a genomic tightrope hand-in-hand. Trends Biochem Sci. 2023;48:1012–1013. doi: 10.1016/j.tibs.2023.09.007 [DOI] [PubMed] [Google Scholar]
- [142].Fenstermaker TK, Sun G, Mazo A, et al. A proximity ligation-based method to detect RNA-DNA association. Methods Mol Biol. 2019;2008:121–129. doi: 10.1007/978-1-4939-9537-0_10 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [143].Malik S, Roeder RG. Regulation of the RNA polymerase II pre-initiation complex by its associated coactivators. Nat Rev Genet. 2023;24(11):767–782. doi: 10.1038/s41576-023-00630-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [144].Xu C, Corces VG. Nascent DNA methylome mapping reveals inheritance of hemimethylation at CTCF/cohesin sites. Science. 2018;359(6380):1166–1170. doi: 10.1126/science.aan5480 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [145].Xu C, Corces VG. Genome-wide mapping of protein-DNA interactions on nascent chromatin. Methods Mol Biol. 2018;1766:231–238. doi: 10.1007/978-1-4939-7768-0_13 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [146].Alabert C, Bukowski-Wills J-C, Lee S-B, et al. Nascent chromatin capture proteomics determines chromatin dynamics during DNA replication and identifies unknown fork components. Nat Cell Biol. 2014;16(3):281–291. doi: 10.1038/ncb2918 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [147].Dungrawala H, Cortez D. Purification of proteins on newly synthesized DNA using iPOND. Methods Mol Biol. 2015;1228:123–131. doi: 10.1007/978-1-4939-1680-1_10 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [148].Alvarez V, Bandau S, Jiang H, et al. Proteomic profiling reveals distinct phases to the restoration of chromatin following DNA replication. Cell Rep. 2023;42:111996. doi: 10.1016/j.celrep.2023.111996 [DOI] [PubMed] [Google Scholar]
- [149].Aranda S, Rutishauser D, Ernfors P. Identification of a large protein network involved in epigenetic transmission in replicating DNA of embryonic stem cells. Nucleic Acids Res. 2014;42:6972–6986. doi: 10.1093/nar/gku374 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [150].Gutiérrez MP, MacAlpine HK, MacAlpine DM. Nascent chromatin occupancy profiling reveals locus- and factor-specific chromatin maturation dynamics behind the DNA replication fork. Genome Res. 2019;29(7):1123–1133. doi: 10.1101/gr.243386.118 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [151].Ramachandran S, Henikoff S. Transcriptional regulators compete with nucleosomes post-replication. Cell. 2016;165(3):580–592. doi: 10.1016/j.cell.2016.02.062 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [152].Owens N, Papadopoulou T, Festuccia N, et al. CTCF confers local nucleosome resiliency after DNA replication and during mitosis. Elife. 2019;8:563619. doi: 10.7554/eLife.47898 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [153].Gu L, Li M, Li CM. Small molecule targeting of transcription-replication conflict for selective chemotherapy. Cell Chem Biol. 2023;30(10):1235–1247. doi: 10.1016/j.chembiol.2023.07.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [154].Yankulov K, Todorov I, Romanowski P, et al. MCM proteins are associated with RNA polymerase II holoenzyme. Mol Cell Biol. 1999;19:6154–6163. doi: 10.1128/MCB.19.9.6154 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [155].Holland L, Gauthier L, Bell-Rogers P, et al. Distinct parts of minichromosome maintenance protein 2 associate with histone H3/H4 and RNA polymerase II holoenzyme. Eur J Biochem. 2002;269(21):5192–5202. doi: 10.1046/j.1432-1033.2002.03224.x [DOI] [PubMed] [Google Scholar]
- [156].Wang Y, Xu F, Hall FL. The MAT1 cyclin-dependent kinase-activating kinase (CAK) assembly/targeting factor interacts physically with the MCM7 DNA licensing factor. FEBS Lett. 2000;484:17–21. doi: 10.1016/s0014-5793(00)02117-7 [DOI] [PubMed] [Google Scholar]
- [157].Rytkönen AK, Hillukkala T, Vaara M, et al. DNA polymerase ε associates with the elongating form of RNA polymerase II and nascent transcripts. FEBS J. 2006;273(24):5535–5549. doi: 10.1111/j.1742-4658.2006.05544.x [DOI] [PubMed] [Google Scholar]
- [158].Aygün O, Svejstrup J, Liu Y. A RECQ5–RNA polymerase II association identified by targeted proteomic analysis of human chromatin. Proc Natl Aca Sci USA. 2008;105(25):8580–8584. doi: 10.1073/pnas.0804424105 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [159].Urban V, Dobrovolna J, Janscak P. Distinct functions of human RecQ helicases during DNA replication. Biophys Chem. 2017;225:20–26. doi: 10.1016/j.bpc.2016.11.005 [DOI] [PubMed] [Google Scholar]
- [160].Andrs M, Hasanova Z, Oravetzova A, et al. RECQ5: a mysterious helicase at the interface of DNA replication and transcription. Genes (Basel). 2020;11(2):11. doi: 10.3390/genes11020232 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [161].Kanagaraj R, Huehn D, MacKellar A, et al. RECQ5 helicase associates with the C-terminal repeat domain of RNA polymerase II during productive elongation phase of transcription. Nucleic Acids Res. 2010;38(22):8131–8140. doi: 10.1093/nar/gkq697 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [162].Urban V, Dobrovolna J, Hühn D, et al. RECQ5 helicase promotes resolution of conflicts between replication and transcription in human cells. J Cell Bio. 2016;214:401–415. doi: 10.1083/jcb.201507099 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [163].Hasanova Z, Klapstova V, Porrua O, et al. Human senataxin is a bona fide R-loop resolving enzyme and transcription termination factor. Nucleic Acids Res. 2023:1–20. doi: 10.1093/nar/gkad092 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [164].Groh M, Albulescu LO, Cristini A, et al. Senataxin: genome guardian at the interface of transcription and neurodegeneration. J Mol Biol. 2017;429:3181–3195. doi: 10.1016/j.jmb.2016.10.021 [DOI] [PubMed] [Google Scholar]
- [165].Alzu A, Bermejo R, Begnis M, et al. Senataxin associates with replication forks to protect fork integrity across RNA-polymerase-II-transcribed genes. Cell. 2012;151(4):835–846. doi: 10.1016/j.cell.2012.09.041 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [166].Zardoni L, Nardini E, Brambati A, et al. 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. doi: 10.1093/nar/gkab1146 [DOI] [PMC free article] [PubMed] [Google Scholar]
- [167].Takedachi A, Despras E, Scaglione S, et al. SLX4 interacts with RTEL1 to prevent transcription-mediated DNA replication perturbations. Nat Struct Mol Biol. 2020;27(5):438–449. doi: 10.1038/s41594-020-0419-3 [DOI] [PubMed] [Google Scholar]
