Biochemical Society Transactions (2017) DOI: 10.1042/BST20170097

Review Article G-quadruplex unwinding and their function in vivo

Markus Sauer and Katrin Paeschke European Research Institute for the Biology of Ageing, University of Groningen, University Medical Center Groningen, 9713 AV Groningen, The Netherlands Correspondence: Katrin Paeschke ([email protected])

The concept that G-quadruplex (G4) structures can form within DNA or RNA in vitro has been long known and extensively discussed. In recent years, accumulating evidences imply that G-quadruplex structures form in vivo. Initially, inefficient regulation of G-quad- ruplex structures was mainly associated with genome instability. However, due to the location of G-quadruplex motifs and their evolutionary conservation, different cellular func- tions of these structures have been postulated (e.g. in maintenance, DNA replica- tion, , and translation). Regardless of their function, efficient and controlled formation and unwinding are very important, because ‘mis’-regulated G-quadruplex struc- tures are detrimental for a given process, causing genome instability and diseases. Several helicases have been shown to target and regulate specificG-quadruplexstructures.This mini-review focuses on the biological consequences of G4 disruption by different helicases in vivo.

Introduction G-quadruplex (G4) structures can form in DNA and RNA molecules when harboring a specific con- sensus motif (G≥3N1-7G≥3N1-7G≥3N1-7G≥3). Recently, different in vitro and in vivo experiments show that variations in this motifs are also possible underlining the diversity of this structure (e.g. with G2 tracks or divided by a variable number of nucleotides) [1]. Four guanines are cyclically coordinated by Hoogsteen base pairing in a planar arrangement termed as G-quartet [2]. Stacking of these G-quartets results in a conformational diverse three-dimensional structure called G4 [3]. In silico analyses identified G4 motifs in various organisms and revealed that they are not randomly distributed in the genome. Instead, they are overrepresented in areas of regulatory relevance like tran- scriptional start sites and promoters, meiotic and mitotic double-strand break sites, replication hot- spots, and (reviewed in refs [4,5]). Additionally, G4 motifs and their genomic locations are evolutionarily conserved [6,7], suggesting that G4 structures potentially influence biological processes (reviewed in ref. [8,9]). On the RNA level, G4 motifs are enriched in the 50- and 30-UTRs of mRNAs as well as in many noncoding transcripts [10]. RNA-G4 structures perform functions in processes like miRNA biogenesis, telomere maintenance, and translation (reviewed in refs [11,12]). Although the existence and functions of G4 structures in vivo are controversially discussed, numer- ous experiments such as antibody staining (Sty49, BG4, and D1) [13–15] and genetic and genome- wide analyses show that G4 structures do form in vivo and bear diverse biological functions (reviewed in ref. [16]). Different studies in vitro and in vivo have shown that misregulation of G4 structures (uncontrolled formation or disruption) leads to changes in cellular processes (reviewed in refs [17,18]), resulting in genome instability and diseases. G4 structures can act as roadblocks to active rep- Received: 30 June 2017 lication as well as to the transcriptional and translational machinery, and hence need to be unwound Revised: 31 July 2017 to maintain genome and cellular output. Various G4-interacting proteins have been identified in Accepted: 10 August 2017 numerous studies (reviewed in ref. [19]). Among the identified proteins were multiple helicases that Version of Record published: specifically unwind G4 structures in vitro. To date, 15 eukaryotic helicases have been described to 22 September 2017 unwind G4 structures, differing in strand and target specificity (Table 1). They target G4 structures

© 2017 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 1 Biochemical Society Transactions (2017) DOI: 10.1042/BST20170097

during different biological processes and hence affect genome instability on many levels. For a detailed review on the mechanisms of helicases, see ref. [20]. Here, we will discuss the helicases which are able to unwind G4 structures in vivo and what the biological consequences of this regulation are.

Effects of G4 helicases during telomere maintenance G4 structures form within telomeric DNA, RNA component (TERC), internal telomeric repeats, and telomeric repeat-containing RNA (TERRA). Owing to these different locations and conformational diver- sity [1,21,22], G4 structures are able to influence telomere maintenance in multiple ways: protection from deg- radation, support of nuclear organization, and alteration of telomerase activity (reviewed in refs [9,16]). Consequently, many helicases have been implicated to affect G4 unfolding at telomeres (Figure 1). RecQ helicases are 30–50 DNA helicases present in numerous organisms. Deletion or of the human RecQ family members is associated with genetic disorders and predisposition (reviewed in ref. [23]). It has been demonstrated in vitro that most RecQ helicases can unwind G4 structures, although with different specificity and [24–27]. The human RecQ family members WRN and BLM support telo- mere replication in vivo, which is speculated to be related to G4 structure unwinding [24,28]. Without WRN, defects during lagging-strand replication and end-to-end fusions of telomeres are observed [29]. BLM, which has greater G4 unwinding activity than WRN, supports telomeric replication on the leading strand. In vivo, loss of BLM activity results in a significant enrichment of G4 structures, especially at telomeres, and altered replication speed. This slow-down of replication is further enhanced upon the addition of the G4-stabilizing ligand Phen-DC3 [30]. Interestingly, in BLM-deficient cells, such as in the disease Bloom syndrome, fragile tel- omeres are observed at sister chromatids derived from G-rich templates [31–33]. In ciliates, a RecQ-like helicase was shown to regulate G4 structures during DNA replication together with telomerase. This results in detach- ment of the telomeres from the nuclear scaffold and permits telomere elongation [34–36]. RTEL1 is another helicase, which is connected to G4 structure unwinding at telomeres. This 50–30 DNA heli- case, identified in worm, mouse, and mammals, belongs to the Rad3/XPD family and is described to be essen- tial for telomere maintenance and DNA repair [37,38]. In the absence of RTEL1, telomeres are short [38] and become fragile [31] causing the rare genetic disorder Hoyeraal-Hreidarsson syndrome [39]. This telomeric defect is even more severe upon the addition of the G4 ligand TMPyP4 [38,40]. Interestingly, no dramatic effect on telomeres is known for other family members (XPD, FANCJ, Dog-1, or ChlR1; [41]). ATRX is an SWI2/SNF2 DNA helicase/ATPase [42], which supports replication fork progression, influences telomere length, and targets G4 structures [43,44]. This led to the speculation whether these three observations are linked to each other. A recent study showed that ATRX targets R-loops in transcribed telomeric repeats. The authors suggest that ATRX averts G4 formation within these R-loops and by this prevents replication stal- ling and DNA damage at telomeres [44]. Nonetheless, additional studies in vivo are required to fully address this statement. Another DNA helicase, which is implicated in telomere biology and DNA replication (flap maturation during Okazaki fragment maturation), is Dna2. In vitro experiments showed that Dna2 can recognize and disrupt telomeric G4 structures [45]. In Dna2-deficient mouse cells, fragile telomeres and severe telomere repli- cation problems occur. This telomere phenotype is increased upon the addition of TMPyP4 or TMS, both G4 stabilization components [45]. The 30–50 RNA helicase DHX36 (RHAU or G4R1) belongs to the DEAH box family and was identified as the major source for G4 unwinding activity in HeLa cell lysate [46]. In vitro analyses confirmed resolving of DNA- and RNA-G4 structures by DHX36 [47,48]. The formation of a G4 at the 50-end disrupts an essential structure within TERC (the P1 helix) and alters telomerase activity [49,50]. DHX36 binds to this G4 structure in vitro [51]. In DHX36-knockdown HEK293T cells, telomerase activity is down-regulated and telomeres are not elongated [52,53]. This finding led to the model that DHX36 regulates telomerase function/processivity by unwinding the G4 structure within TERC. Different studies in vivo have shown that TERRA, the product of transcription at telomeres, has a major impact on telomere length regulation and telomerase function (reviewed in ref. [54]). TERRA was shown to fold into G4 structures in vitro [55,56]. So far, no helicase involved in TERRA-G4 regulation has been described. A likely candidate is DHX36, because of its presence at telomeres and its specificity to unwind RNA-G4 structures. To date, it is clear that telomere biology strongly depends on the function of helicases and that helicases, especially in the context of G4s, support telomere maintenance at multiple levels.

2 © 2017 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society Biochemical Society Transactions (2017) DOI: 10.1042/BST20170097

Table 1 Helicases implicated to regulate G4 in vivo List of all helicases and selected features, which have been described to regulate G4 structure unwinding in vivo. Telomere DNA Name Family Processivity Specificity Organism maintenance replication Transcription Translation Genetic diseases

Pfh1 Pif1-like 50–30 DNA S. pombe ?Yes Pif1 Pif1-like 50–30 DNA S. cerevisiae ?Yes DHX36 DEAH 30–50 DNA, RNA Homo sapiens Yes Yes Yes WRN RecQ-like 30–50 DNA H. sapiens Yes Yes Yes Werner syndrome BLM RecQ-like 30–50 DNA H. sapiens Yes Yes Yes Bloom syndrome Sgs1 RecQ-like 30–50 DNA S. cerevisiae ? Only in vitro Yes FANCJ Rad3/XPD 50–30 DNA H. sapiens Yes Fanconi anemia Dog-1 Rad3/XPD 50–30 DNA C. elegans Yes RTEL1 DEAH 50–30 DNA H. sapiens Yes Yes Hoyeraal-Hreidarsson syndrome ATRX Swi2/Snf2 50–30 DNA H. sapiens Yes ATRX syndrome Dna2 Upf1-like 50–30 DNA S. cerevisiae Yes ? DDX21 DEAD 50–30 DNA, RNA H. sapiens Yes DHX9 DEAH 30–50 DNA, RNA H. sapiens Yes? Yes? XPD Rad3/XPD DNA H. sapiens Yes Xeroderma (XPB) pigmentosum eIF4A DEAD RNA H. sapiens Yes Cancer

Effects of helicases operating on G4 structures during replication G4 structures can form not only at telomeres but also at internal genomic regions [57–59]. It has been shown in human cells that G4 structure formation is stimulated during S-phase [4,14,60]. Different models have been postulated for the biological function of G4 structures during DNA replication. At origins, G4 structures have been shown to support the initiation of DNA replication (reviewed in ref. [61]). Alternatively, G4 structures might safeguard proper replication or prevent uncoupling of the leading- and lagging-strand . Contrary, G4 structures are an obstacle for the replication fork machinery and are in need of unwinding. In agreement with this is the fact that G4 structures cause genomic and epigenetic instability if they are not efficiently regulated [62,63]. So far, a few helicases have been discussed in the context of DNA replication and genome stability (Figure 1). The first helicase that has been shown to regulate G4 structures in vivo was the 50–30 DNA helicase DOG-1 of the Rad3/XPD family. Well-designed studies in Caenorhabditis elegans showed that DOG-1 is essential for genome stability at G4 motifs. Upon mutation of DOG-1, genetic alterations and deletions accumulated within these G4 motifs [64,65]. The closely related helicase FANCJ (BRIP1 or BACH1) was shown to support replication through G4 struc- tures, as demonstrated in Xenopus egg extracts [66]. Note that of FANCJ are connected to the human chromosomal instability disorder Fanconi anemia, but this is most likely not directly linked to its role in regulating G4 structures [67]. Detailed studies in vitro and in vivo revealed that FANCJ supports DNA repli- cation [68] and can unwind G4 structures in vitro [69,70]. FANCJ promotes genome stability at G4s by two independent mechanisms: FANCJ unwinds G4 structures that have been destabilized by the translesion synthe- sis REV1. Second, WRN or BLM assists FANCJ in unwinding G4 structures from both directions [71–73]. Interestingly, in the absence of FANCJ, genomic deletions at G4 motifs were identified [69]. Pif1 helicase family members have been shown to efficiently unwind G4 structures in vitro with a higher spe- cificity than other helicases [74–78]. These 50–30 DNA helicases are conserved from yeasts to human [79]. Genome-wide and molecular studies have shown that ScPif1 (Saccharomyces cerevisiae) and Pfh1 (Saccharomyces pombe) bind to G4 motifs and specifically support local DNA replication at these regions [76,80]. Two different genetic assays in yeast, one using a G-rich tandem array consisting of human minisatel- lites (CEB1) and one using inserted G4 motifs, revealed that in the absence of ScPif1, G4 motifs are unstable

© 2017 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 3 Biochemical Society Transactions (2017) DOI: 10.1042/BST20170097

Figure 1. Involvement of G-quadruplex helicases in cellular processes in vivo. Cellular processes discussed in the present paper are indicated with their cellular localization (nuclear = yellow background, cytosolic = red background). G4 structure-resolving helicases are listed according to their contribution to each process. Locations of potential G4-forming events are listed in corresponding bubbles. Top section = telomere maintenance (yellowish proteins): G4s can form in telomeric DNA, TERRA, and TERC. Middle section = DNA replication (reddish proteins): G4s can form in the leading strand, lagging strand, and at origins. Bottom left section = transcription (RNA pol II in blue, gray, and turquoise represents , respectively, exons of mRNAs): G4s can form in TSSs, promoters, and at the 30-end of . Bottom right section = translation (ribosomes in green, turquoise in mature mRNA, and translated proteins in dark gray): G4s can form in the 50-UTR, 30-UTR, and the coding sequence of mRNAs.

and gross chromosomal rearrangement increases significantly [74,78,80–82]. Pif1 binds and unwinds G4 struc- tures at the end of S-phase [80]. To date, it is not clear if ScPif1 functions on G4 structures on both replicating strands or has preferential strand specificity. Studies using CEB1 have shown that ScPif1 supports genome sta- bility only at G4 structures located on the leading strand [81], whereas other studies revealed that ScPif1 binds and regulates G4 structures on both strands [74,80]. These controversial data can be explained that under dif- ferent experimental conditions, diverse and different G4 structure conformations occurred and that other factors in addition to ScPif1 were needed to aid G4 regulation. This speculation agrees with a recent study in yeast, showing that Mms1, a ubiquitin ligase partner, supports ScPif1 binding to G4 structures located on the lagging strand [83]. The previously mentioned RecQ helicase family is also known to influence multiple steps during DNA repli- cation, to interact with multiple DNA repair proteins, and to unwind G4 structures efficiently (reviewed in ref. [84]). Nevertheless, direct evidence that RecQ helicases regulate G4 structure formation during DNA replication in vivo is missing. Many experiments have revealed that replication fork progression is negatively affected at G4 motifs and that this is enhanced in the absence of helicases. It remains elusive why G4 structures form during DNA replication. As mentioned above, replication initiation or control of replication fork progression might be positive scenarios,

4 © 2017 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society Biochemical Society Transactions (2017) DOI: 10.1042/BST20170097

but no helicase is linked to G4 structures formed within origins of replication, yet. Potential candidates might be RECQ1 and RECQ4, which are associated with origin function [85]. Effects of helicases acting on G4 structures during transcription G4 motifs are enriched at transcriptional start sites (TSSs) and regions of genes thereby providing a decisive mechanism for transcription initiation [86]. Of great interest is the fact that especially in promoter regions of onco- and regulatory genes, G4 motifs are enriched, whereas in tumor suppressor and housekeeping genes they are depleted [86,87]. This fact provides a logical connection between G4-dependent regulation and . G4 structures are postulated to affect transcription on multiple levels: they reduce transcrip- tion by blocking polymerases, stimulate transcription by recruiting proteins, and keep the nascent strand in a single-stranded conformation by G4 structure formation on the non-transcribed strand [9,88]. In agreement with this are in cellulo and in vivo studies showing transcriptional changes upon formation of G4 structures by chemical ligands [89–92]. So far, the best described G4 structure-mediated effect on transcription comes from reports studying the pro- moter region of the proto-oncogene c-MYC [93]. Changes in c-MYC transcription levels are associated with G4 formation as observed by in vivo footprinting and other molecular analyses [58,94–98]. Although there is a lot known about G4 structure formation and its impact on c-MYC transcription, only little is known about the ‘disruption’ of these G4 structures. So far, only three proteins, among them one helicase, have been described to support unfolding of G4 structures in c-MYC: CNBP, PARP-1, and Pif1 [99–101]. Nevertheless, a detailed analysis of these proteins and changes in G4-mediated c-MYC transcription are missing. RecQ helicase family members are also assumed to regulate transcription by resolving DNA-G4 structures. The S. cerevisiae RecQ helicase Sgs1 and the above-introduced human BLM and WRN helicases can unwind G4 structures [102,103]. Gene expression changes in cells lacking these helicases are observed specifically at regions that harbor a G4 within the promoter region/TSS [90,104,105]. Two members of the Rad3/XPD family, XPD and XPB, are also proposed to affect transcription via G4 structure unfolding. XPB binds to G4 structures, whereas XPD binds and unwinds G4 structures in vitro. Interestingly, genome-wide ChIP-seq analyses of these two proteins in human cells revealed that 40% of their binding sites overlap with G4 motifs. XPD and XPB bind particularly to G4 structures at promoters of highly transcribed genes. The authors speculate that XPD and/or XPB support transcription of genes linked to dis- eases by unwinding local G4 structures [106]. Another helicase that was described to have a function at G4s in promoter regions is the RNA helicase DHX36. YY1 harbors G4 motifs in its promoter and in the 50-UTR of its mRNA. The YY1 gene encodes for a multifunctional protein with regulatory potential in tumorigenesis [107]. At both loci, G4 structures affect YY1 expression. Surprisingly, DHX36 acts only on DNA-G4 structures located in the promoter but not on putative RNA-G4 structures in the 50-UTR [108]. G4 motifs are overrepresented immediately after the 30-end of genes. This leads to a model in which G4 struc- tures assist the correct termination of transcription. To this end, G4 structures are pausing sites for the poly- merases, which ultimately result in the termination of transcription [109,110]. In this model, G4 structures can be formed by the RNA transcript [111] or by a DNA–RNA hybrid [112–114]. A good candidate for regulating this mechanism is the DEAH box helicase DHX9. DHX9 can resolve RNA-G4s nearly twice as fast as their DNA analog in vitro [115]. Furthermore, it was described as an abundant and RNA polymerase-associated protein. However, regulation of transcription by DHX9 in vivo was only demonstrated in context of other mechanisms, mainly by resolving RNA forks and R-loops resulting in fastened polymerase progression (reviewed in ref. [116]). As it becomes more obvious that G4 structures potentially influence transcription positively and negatively, it is less clear how they form or are unfolded and what are the biological consequences of G4 structure forma- tion and regulation. G4 regulation and the consequences for RNA-mediated processes Lacking a competing complementary strand, G-rich RNA is even more prone to fold G4 structures than DNA. Additionally, G4 structures have an even higher thermodynamic stability in RNA [117]. During post- transcriptional gene regulation, G4 structures can be present within the 50-or30-UTR or the coding region of mRNAs. This results in numerous potential roles for G4 during protein synthesis (reviewed in ref. [12]). Although many different RNA-G4-interacting proteins have been identified [118], only a few RNA helicases are known to regulate G4 structures. A recent study in HEK293 cells came to the conclusion that in these cells,

© 2017 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 5 Biochemical Society Transactions (2017) DOI: 10.1042/BST20170097

G4 structures are not present at the mRNA level, because they are rapidly removed by an ATP-independent mechanism [119]. Yet, there are numerous studies relating a molecular function of mRNAs with G4 structures, which will be presented below (Figure 1). DHX36 was shown to support translation of specific loci by unwinding G4 structures. In complex with Aven, DHX36 aids translation of the mRNAs MLL1 and MLL4 by binding to regions harboring G4 motifs [120]. Furthermore, experiments in mouse heart tissue showed that DHX36 knockout causes severe defects in cardiogenesis. Here, an unresolved G4 structure in the 50-UTR of the NKx2-5 mRNA prevents its translation and by this the expression of this important early transcription factor [121]. DHX36 also impedes the transla- tion of the PITX mRNA, which exhibits G4 motifs in its 30-UTR. However, this seems not to be a consequence of unwinding by DHX36 but a miRNA-mediated mechanism [122]. RNA helicases regulate gene expression not only directly during translation but also by other processes which then downstream influence protein synthesis. Upon UV-induced DNA damage, DHX36 is necessary to maintain p53 pre-mRNA 30-end processing by regulating a G4 structure within the p53 mRNA. Depletion of DHX36 results in loss of p53 protein [123]. Furthermore, DHX36 was described to be involved in mRNA turn- over and localizes to stress granules [124]. Here, it is possible that DHX36 assists in the decay of G4 motif- harboring mRNAs and prevents them from being translated. The closely related helicase DHX9 was shown to unwind G4 structures in vitro [115]. DHX9 is located in both the nucleus and the cytoplasm, and was previously described to facilitate translation of mRNAs with highly structured 50-UTRs [116]. The direct link between G4 structures, DHX9, and translation has not yet been demonstrated in vivo. DDX21 is a DEAD-box RNA helicase that was recently shown to unwind G4 structures. DDX21 acts on RNA-G4 within the 30-UTR of mRNAs, and its depletion abolishes protein translation from its target mRNAs [125]. The helicase eIF4A also belongs to the DEAD-box helicase family and is another important factor imple- mented in translational processes. It was reported that RNA-G4s cause eIF4A-dependent translation of onco- genes [126]. However, whether this helicase can directly resolve G4 structures has not been shown so far. The biology of RNA-G4 structures is less studied than its DNA counterpart. However, it is obvious that RNA-G4s have a strong impact on various steps during RNA-mediated processes (e.g. translational initiation, RNA stability, localization, and transport). Over 70 RNA helicases have been described (according to the http://www. rnahelicase.org/namingcode.htm database), but only a few have been tested on G4 substrates in vitro and in vivo. To address the question, which helicases act on RNA-G4 structures and when, is the subject of current research. Conclusion In recent years, it became clear that G4 structure formation is more complex than original anticipated. Whereas some years ago only conventional G4 motifs were discussed to fold into G4s, now less stringent motifs are shown to fold into these structures. Additionally, the topology of G4 structures themselves is highly diverse and, for the same sequence, multiple G4 conformations are possible. This diversity raises multiple open questions: have we identified all G4 unwinding helicases? How do helicases recognize different G4 structures? What is the biological mechanism how G4 structures are disrupted in vivo? Are assessor proteins involved in G4 targeting? How much impact have, in addition to helicases, other proteins in G4 disruption? If we assume that there are over 700 000 potential G4 structures in the human genome, it is unlikely that they form at once, therefore a highly interesting question is when and why and which G4 structures form in a given biological process.

Abbreviations G4, G-quadruplex; TERC, telomerase RNA component; TSSs, transcriptional start sites.

Funding Work in the Paeschke Laboratory is funded by the European Research Council Stg Grant [638988-G4DSB] and an Emmy Noether Fellowship (Deutsche Forschungsgemeinschaft).

Acknowledgements We thank Stefan Juranek and Satyaprakash Pandey for careful reading of the manuscript.

Competing Interests The Authors declare that there are no competing interests associated with the manuscript.

6 © 2017 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society Biochemical Society Transactions (2017) DOI: 10.1042/BST20170097

References 1 Phan, A.T. (2010) Human telomeric G-quadruplex: structures of DNA and RNA sequences. FEBS J. 277, 1107–1117 doi:10.1111/j.1742-4658.2009. 07464.x 2 Gellert, M., Lipsett, M.N. and Davies, D.R. (1962) Helix formation by guanylic acid. Proc. Natl Acad. Sci. U.S.A. 48, 2013–2018 doi:10.1073/pnas.48. 12.2013 3 Burge, S., Parkinson, G.N., Hazel, P., Todd, A.K. and Neidle, S. (2006) Quadruplex DNA: sequence, topology and structure. Nucleic Acids Res. 34, 5402–5415 doi:10.1093/nar/gkl655 4 Murat, P. and Balasubramanian, S. (2014) Existence and consequences of G-quadruplex structures in DNA. Curr. Opin. Genet. Dev. 25,22–29 doi:10.1016/j.gde.2013.10.012 5 Huppert, J.L. (2010) Structure, location and interactions of G-quadruplexes. FEBS J. 277, 3452–3458 doi:10.1111/j.1742-4658.2010.07758.x 6 Nakken, S., Rognes, T. and Hovig, E. (2009) The disruptive positions in human G-quadruplex motifs are less polymorphic and more conserved than their neutral counterparts. Nucleic Acids Res. 37, 5749–5756 doi:10.1093/nar/gkp590 7 Capra, J.A., Paeschke, K., Singh, M. and Zakian, V.A. (2010) G-quadruplex DNA sequences are evolutionarily conserved and associated with distinct genomic features in Saccharomyces cerevisiae. PLoS Comput. Biol. 6, e1000861 doi:10.1371/journal.pcbi.1000861 8 Lipps, H.J. and Rhodes, D. (2009) G-quadruplex structures: in vivo evidence and function. Trends Cell Biol. 19, 414–422 doi:10.1016/j.tcb.2009.05.002 9 Bochman, M.L., Paeschke, K. and Zakian, V.A. (2012) DNA secondary structures: stability and function of G-quadruplex structures. Nat. Rev. Genet. 13, 770–780 doi:10.1038/nrg3296 10 Huppert, J.L., Bugaut, A., Kumari, S. and Balasubramanian, S. (2008) G-quadruplexes: the beginning and end of UTRs. Nucleic Acids Res. 36, 6260–6268 doi:10.1093/nar/gkn511 11 Agarwala, P., Pandey, S. and Maiti, S. (2015) The tale of RNA G-quadruplex. Org. Biomol. Chem. 13, 5570–5585 doi:10.1039/c4ob02681k 12 Fay, M.M., Lyons, S.M. and Ivanov, P. (2017) RNA G-quadruplexes in biology: principles and molecular mechanisms. J. Mol. Biol. 429, 2127–2147 doi:10.1016/j.jmb.2017.05.017 13 Schaffitzel, C., Berger, I., Postberg, J., Hanes, J., Lipps, H.J. and Pluckthun, A. (2001) In vitro generated antibodies specific for telomeric guanine-quadruplex DNA react with Stylonychia lemnae macronuclei. Proc. Natl Acad. Sci. U.S.A. 98, 8572–8577 doi:10.1073/pnas.141229498 14 Biffi, G., Di Antonio, M., Tannahill, D. and Balasubramanian, S. (2014) Visualization and selective chemical targeting of RNA G-quadruplex structures in the cytoplasm of human cells. Nat. Chem. 6,75–80 doi:10.1038/nchem.1805 15 Liu, H.-Y., Zhao, Q., Zhang, T.-P., Wu, Y., Xiong, Y.-X., Wang, S.-K. et al. (2016) Conformation selective antibody enables genome profiling and leads to discovery of parallel G-quadruplex in human telomeres. Cell Chem. Biol. 23, 1261–1270 doi:10.1016/j.chembiol.2016.08.013 16 Rhodes, D. and Lipps, H.J. (2015) G-quadruplexes and their regulatory roles in biology. Nucleic Acids Res. 43, 8627–8637 doi:10.1093/nar/gkv862 17 Maizels, N. (2015) G4-associated human diseases. EMBO Rep. 16, 910–922 doi:10.15252/embr.201540607 18 Cea, V., Cipolla, L. and Sabbioneda, S. (2015) Replication of structured DNA and its implication in epigenetic stability. Front. Genet. 6, 209 doi:10.3389/fgene.2015.00209 19 Brázda, V., Hároníková, L., Liao, J.C. and Fojta, M. (2014) DNA and RNA quadruplex-binding proteins. Int. J. Mol. Sci. 15, 17493–17517 doi:10.3390/ ijms151017493 20 Mendoza, O., Bourdoncle, A., Boulé, J.-B., Brosh, Jr, R.M. and Mergny, J.-L. (2016) G-quadruplexes and helicases. Nucleic Acids Res. 44, 1989–2006 doi:10.1093/nar/gkw079 21 You, H., Zeng, X., Xu, Y., Lim, C.J., Efremov, A.K., Phan, A.T. et al. (2014) Dynamics and stability of polymorphic human telomeric G-quadruplex under tension. Nucleic Acids Res. 42, 8789–8795 doi:10.1093/nar/gku581 22 Dai, J., Carver, M. and Yang, D. (2008) Polymorphism of human telomeric quadruplex structures. Biochimie 90, 1172–1183 doi:10.1016/j.biochi.2008. 02.026 23 Larsen, N.B. and Hickson, I.D. (2013) RecQ helicases: conserved guardians of genomic integrity. Adv. Exp. Med. Biol. 767, 161–184 doi:10.1007/ 978-1-4614-5037-5_8 24 Sun, H., Karow, J.K., Hickson, I.D. and Maizels, N. (1998) The Bloom’s syndrome helicase unwinds G4 DNA. J. Biol. Chem. 273, 27587–27592 doi:10.1074/jbc.273.42.27587 25 Mohaghegh, P., Karow, J.K., Brosh, Jr, R.M., Bohr, V.A. and Hickson, I.D. (2001) The Bloom’s and Werner’s syndrome proteins are DNA structure-specific helicases. Nucleic Acids Res. 29, 2843–2849 doi:10.1093/nar/29.13.2843 26 Popuri, V., Bachrati, C.Z., Muzzolini, L., Mosedale, G., Costantini, S., Giacomini, E. et al. (2008) The human RecQ helicases, BLM and RECQ1, display distinct DNA substrate specificities. J. Biol. Chem. 283, 17766–17776 doi:10.1074/jbc.M709749200 27 Budhathoki, J.B., Maleki, P., Roy, W.A., Janscak, P., Yodh, J.G. and Balci, H. (2016) A comparative study of G-quadruplex unfolding and DNA reeling activities of human RECQ5 helicase. Biophys. J. 110, 2585–2596 doi:10.1016/j.bpj.2016.05.016 28 Higa, M., Fujita, M. and Yoshida, K. (2017) DNA replication origins and fork progression at mammalian telomeres. Genes 8, 112 doi:10.3390/ genes8040112 29 Crabbe, L., Verdun, R.E., Haggblom, C.I. and Karlseder, J. (2004) Defective telomere lagging strand synthesis in cells lacking WRN helicase activity. Science 306, 1951–1953 doi:10.1126/science.1103619 30 Drosopoulos, W.C., Kosiyatrakul, S.T. and Schildkraut, C.L. (2015) BLM helicase facilitates telomere replication during leading strand synthesis of telomeres. J. Cell Biol. 210, 191–208 doi:10.1083/jcb.201410061 31 Sfeir, A., Kosiyatrakul, S.T., Hockemeyer, D., MacRae, S.L., Karlseder, J., Schildkraut, C.L. et al. (2009) Mammalian telomeres resemble fragile sites and require TRF1 for efficient replication. Cell 138,90–103 doi:10.1016/j.cell.2009.06.021 32 Zimmermann, M., Kibe, T., Kabir, S. and de Lange, T. (2014) TRF1 negotiates TTAGGG repeat-associated replication problems by recruiting the BLM helicase and the TPP1/POT1 repressor of ATR signaling. Genes Dev. 28, 2477–2491 doi:10.1101/gad.251611.114 33 Barefield, C. and Karlseder, J. (2012) The BLM helicase contributes to telomere maintenance through processing of late-replicating intermediate structures. Nucleic Acids Res. 40, 7358–7367 doi:10.1093/nar/gks407 34 Postberg, J., Tsytlonok, M., Sparvoli, D., Rhodes, D. and Lipps, H.J. (2012) A telomerase-associated RecQ protein-like helicase resolves telomeric G-quadruplex structures during replication. Gene 497, 147–154 doi:10.1016/j.gene.2012.01.068

© 2017 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 7 Biochemical Society Transactions (2017) DOI: 10.1042/BST20170097

35 Paeschke, K., Simonsson, T., Postberg, J., Rhodes, D. and Lipps, H.J. (2005) Telomere end-binding proteins control the formation of G-quadruplex DNA structures in vivo. Nat. Struct. Mol. Biol. 12, 847–854 doi:10.1038/nsmb982 36 Paeschke, K., Juranek, S., Simonsson, T., Hempel, A., Rhodes, D. and Lipps, H.J. (2008) Telomerase recruitment by the telomere end binding protein-β facilitates G-quadruplex DNA unfolding in ciliates. Nat. Struct. Mol. Biol. 15, 598–604 doi:10.1038/nsmb.1422 37 Barber, L.J., Youds, J.L., Ward, J.D., McIlwraith, M.J., O’Neil, N.J., Petalcorin, M.I. et al. (2008) RTEL1 maintains genomic stability by suppressing homologous recombination. Cell 135, 261–271 doi:10.1016/j.cell.2008.08.016 38 Ding, H., Schertzer, M., Wu, X., Gertsenstein, M., Selig, S., Kammori, M. et al. (2004) Regulation of murine telomere length by Rtel: an essential gene encoding a helicase-like protein. Cell 117, 873–886 doi:10.1016/j.cell.2004.05.026 39 Le Guen, T., Jullien, L., Touzot, F., Schertzer, M., Gaillard, L., Perderiset, M. et al. (2013) Human RTEL1 deficiency causes Hoyeraal–Hreidarsson syndrome with short telomeres and genome instability. Hum. Mol. Genet. 22, 3239–3249 doi:10.1093/hmg/ddt178 40 Vannier, J.-B., Pavicic-Kaltenbrunner, V., Petalcorin, M.I.R., Ding, H. and Boulton, S.J. (2012) RTEL1 dismantles T loops and counteracts telomeric G4-DNA to maintain telomere integrity. Cell 149, 795–806 doi:10.1016/j.cell.2012.03.030 41 White, M.F. (2009) Structure, function and evolution of the XPD family of iron–sulfur-containing 50→30 DNA helicases. Biochem. Soc. Trans. 37, 547–551 doi:10.1042/BST0370547 42 Picketts, D.J., Higgs, D.R., Bachoo, S., Blake, D.J., Quarrell, O.W. and Gibbons, R.J. (1996) ATRX encodes a novel member of the SNF2 family of proteins: mutations point to a common mechanism underlying the ATR-X syndrome. Hum. Mol. Genet. 5, 1899–1907 doi:10.1093/hmg/5.12.1899 43 Law, M.J., Lower, K.M., Voon, H.P.J., Hughes, J.R., Garrick, D., Viprakasit, V. et al. (2010) ATR-X syndrome protein targets tandem repeats and influences allele-specific expression in a size-dependent manner. Cell 143, 367–378 doi:10.1016/j.cell.2010.09.023 44 Nguyen, D.T., Voon, H.P.J., Xella, B., Scott, C., Clynes, D., Babbs, C. et al. (2017) The chromatin remodelling factor ATRX suppresses R-loops in transcribed telomeric repeats. EMBO Rep. 18, 914–928 doi:10.15252/embr.201643078 45 Lin, W., Sampathi, S., Dai, H., Liu, C., Zhou, M., Hu, J. et al. (2013) Mammalian DNA2 helicase/nuclease cleaves G-quadruplex DNA and is required for telomere integrity. EMBO J. 32, 1425–1439 doi:10.1038/emboj.2013.88 46 Vaughn, J.P., Creacy, S.D., Routh, E.D., Joyner-Butt, C., Jenkins, G.S., Pauli, S. et al. (2005) The DEXH protein product of the DHX36 gene is the major source of tetramolecular quadruplex G4-DNA resolving activity in HeLa cell lysates. J. Biol. Chem. 280, 38117–38120 doi:10.1074/jbc. C500348200 47 Creacy, S.D., Routh, E.D., Iwamoto, F., Nagamine, Y., Akman, S.A. and Vaughn, J.P. (2008) G4 resolvase 1 binds both DNA and RNA tetramolecular quadruplex with high affinity and is the major source of tetramolecular quadruplex G4-DNA and G4-RNA resolving activity in HeLa cell lysates. J. Biol. Chem. 283, 34626–34634 doi:10.1074/jbc.M806277200 48 Lattmann, S., Giri, B., Vaughn, J.P., Akman, S.A. and Nagamine, Y. (2010) Role of the amino terminal RHAU-specific motif in the recognition and resolution of guanine quadruplex-RNA by the DEAH-box RNA helicase RHAU. Nucleic Acids Res. 38, 6219–6233 doi:10.1093/nar/gkq372 49 Gros, J., Guédin, A., Mergny, J.-L. and Lacroix, L. (2008) G-quadruplex formation interferes with P1 helix formation in the RNA component of telomerase hTERC. ChemBioChem 9, 2075–2079 doi:10.1002/cbic.200800300 50 Li, X., Nishizuka, H., Tsutsumi, K., Imai, Y., Kurihara, Y. and Uesugi, S. (2007) Structure, interactions and effects on activity of the 50-terminal region of human telomerase RNA. J. Biochem. 141, 755–765 doi:10.1093/jb/mvm081 51 Lattmann, S., Stadler, M.B., Vaughn, J.P., Akman, S.A. and Nagamine, Y. (2011) The DEAH-box RNA helicase RHAU binds an intramolecular RNA G-quadruplex in TERC and associates with telomerase holoenzyme. Nucleic Acids Res. 39, 9390–9404 doi:10.1093/nar/gkr630 52 Booy, E.P., Meier, M., Okun, N., Novakowski, S.K., Xiong, S., Stetefeld, J. et al. (2012) The RNA helicase RHAU (DHX36) unwinds a G4-quadruplex in human telomerase RNA and promotes the formation of the P1 helix template boundary. Nucleic Acids Res. 40, 4110–4124 doi:10.1093/nar/gkr1306 53 Booy, E.P., McRae, E.K. and McKenna, S.A. (2015) Biochemical characterization of G4 quadruplex telomerase RNA unwinding by the RNA helicase RHAU. Methods Mol. Biol. 1259, 125–135 doi:10.1007/978-1-4939-2214-7_9 54 Cusanelli, E. and Chartrand, P. (2015) Telomeric repeat-containing RNA TERRA: a noncoding RNA connecting telomere biology to genome integrity. Front. Genet. 6, 143 doi:10.3389/fgene.2015.00143 55 Martadinata, H. and Phan, A.T. (2013) Structure of human telomeric RNA (TERRA): stacking of two G-quadruplex blocks in K+ solution. Biochemistry 52, 2176–2183 doi:10.1021/bi301606u 56 Xu, Y., Suzuki, Y., Ito, K. and Komiyama, M. (2010) Telomeric repeat-containing RNA structure in living cells. Proc. Natl Acad. Sci. U.S.A. 107, 14579–14584 doi:10.1073/pnas.1001177107 57 Chambers, V.S., Marsico, G., Boutell, J.M., Di Antonio, M., Smith, G.P. and Balasubramanian, S. (2015) High-throughput sequencing of DNA G-quadruplex structures in the human genome. Nat. Biotechnol. 33, 877–881 doi:10.1038/nbt.3295 58 Hänsel-Hertsch, R., Beraldi, D., Lensing, S.V., Marsico, G., Zyner, K., Parry, A. et al. (2016) G-quadruplex structures mark human regulatory chromatin. Nat. Genet. 48, 1267–1272 doi:10.1038/ng.3662 59 Lam, E.Y.N., Beraldi, D., Tannahill, D. and Balasubramanian, S. (2013) G-quadruplex structures are stable and detectable in human genomic DNA. Nat. Commun. 4, 1796 doi:10.1038/ncomms2792 60 Biffi, G., Tannahill, D., McCafferty, J. and Balasubramanian, S. (2013) Quantitative visualization of DNA G-quadruplex structures in human cells. Nat. Chem. 5, 182–186 doi:10.1038/nchem.1548 61 Valton, A.-L. and Prioleau, M.-N. (2016) G-quadruplexes in DNA replication: a problem or a necessity? Trends Genet. 32, 697–706 doi:10.1016/j.tig. 2016.09.004 62 Bharti, S.K., Sommers, J.A., George, F., Kuper, J., Hamon, F., Shin-ya, K. et al. (2013) Specialization among iron-sulfur cluster helicases to resolve G-quadruplex DNA structures that threaten genomic stability. J. Biol. Chem. 288, 28217–28229 doi:10.1074/jbc.M113.496463 63 Wickramasinghe, C.M., Arzouk, H., Frey, A., Maiter, A. and Sale, J.E. (2015) Contributions of the specialised DNA polymerases to replication of structured DNA. DNA Repair 29,83–90 doi:10.1016/j.dnarep.2015.01.004 64 Cheung, I., Schertzer, M., Rose, A. and Lansdorp, P.M. (2002) Disruption of dog-1 in Caenorhabditis elegans triggers deletions upstream of guanine-rich DNA. Nat. Genet. 31, 405–409 doi:10.1038/ng928 65 Kruisselbrink, E., Guryev, V., Brouwer, K., Pontier, D.B., Cuppen, E. and Tijsterman, M. (2008) Mutagenic capacity of endogenous G4 DNA underlies genome instability in FANCJ-defective C. elegans. Curr. Biol. 18, 900–905 doi:10.1016/j.cub.2008.05.013

8 © 2017 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society Biochemical Society Transactions (2017) DOI: 10.1042/BST20170097

66 Castillo Bosch, P., Segura-Bayona, S., Koole, W., van Heteren, J.T., Dewar, J.M., Tijsterman, M. et al. (2014) FANCJ promotes DNA synthesis through G-quadruplex structures. EMBO J. 33, 2521–2533 doi:10.15252/embj.201488663 67 Brosh, Jr, R.M. and Cantor, S.B. (2014) Molecular and cellular functions of the FANCJ DNA helicase defective in cancer and in Fanconi anemia. Front. Genet. 5, 372 doi:10.3389/fgene.2014.00372 68 Hiom, K. (2010) FANCJ: solving problems in DNA replication. DNA Repair 9, 250–256 doi:10.1016/j.dnarep.2010.01.005 69 London, T.B.C., Barber, L.J., Mosedale, G., Kelly, G.P., Balasubramanian, S., Hickson, I.D. et al. (2008) FANCJ is a structure-specific DNA helicase associated with the maintenance of genomic G/C tracts. J. Biol. Chem. 283, 36132–36139 doi:10.1074/jbc.M808152200 70 Wu, Y., Shin-ya, K. and Brosh, Jr., R.M. (2008) FANCJ helicase defective in Fanconia anemia and breast cancer unwinds G-quadruplex DNA to defend genomic stability. Mol. Cell. Biol. 28, 4116–4128 doi:10.1128/MCB.02210-07 71 Sarkies, P., Murat, P., Phillips, L.G., Patel, K.J., Balasubramanian, S. and Sale, J.E. (2012) FANCJ coordinates two pathways that maintain epigenetic stability at G-quadruplex DNA. Nucleic Acids Res. 40, 1485–1498 doi:10.1093/nar/gkr868 72 Eddy, S., Ketkar, A., Zafar, M.K., Maddukuri, L., Choi, J.-Y. and Eoff, R.L. (2014) Human Rev1 polymerase disrupts G-quadruplex DNA. Nucleic Acids Res. 42, 3272–3285 doi:10.1093/nar/gkt1314 73 Suhasini, A.N. and Brosh, Jr, R.M. (2012) Fanconi anemia and Bloom’s syndrome crosstalk through FANCJ–BLM helicase interaction. Trends Genet. 28, 7–13 doi:10.1016/j.tig.2011.09.003 74 Paeschke, K., Bochman, M.L., Garcia, P.D., Cejka, P., Friedman, K.L., Kowalczykowski, S.C. et al. (2013) Pif1 family helicases suppress genome instability at G-quadruplex motifs. Nature 497, 458–462 doi:10.1038/nature12149 75 Sanders, C.M. (2010) Human Pif1 helicase is a G-quadruplex DNA-binding protein with G-quadruplex DNA-unwinding activity. Biochem. J. 430, 119–128 doi:10.1042/BJ20100612 76 Sabouri, N., Capra, J.A. and Zakian, V.A. (2014) The essential Schizosaccharomyces pombe Pfh1 DNA helicase promotes fork movement past G-quadruplex motifs to prevent DNA damage. BMC Biol. 12, 101 doi:10.1186/s12915-014-0101-5 77 Wallgren, M., Mohammad, J.B., Yan, K.P., Pourbozorgi-Langroudi, P., Ebrahimi, M. and Sabouri, N. (2016) 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. 44, 6213–6231 doi:10.1093/nar/gkw349 78 Ribeyre, C., Lopes, J., Boulé, J.-B., Piazza, P., Guédin, A., Zakian, V.A. et al. (2009) The yeast Pif1 helicase prevents genomic instability causedby G-quadruplex-forming CEB1 sequences in vivo. PLoS Genet. 5, e1000475 doi:10.1371/journal.pgen.1000475 79 Bochman, M.L., Sabouri, N. and Zakian, V.A. (2010) Unwinding the functions of the Pif1 family helicases. DNA Repair 9, 237–249 doi:10.1016/j. dnarep.2010.01.008 80 Paeschke, K., Capra, J.A. and Zakian, V.A. (2011) DNA replication through G-quadruplex motifs is promoted by the Saccharomyces cerevisiae Pif1 DNA helicase. Cell 145, 678–691 doi:10.1016/j.cell.2011.04.015 81 Lopes, J., Piazza, A., Bermejo, R., Kriegsman, B., Colosio, A., Teulade-Fichou, M.-P. et al. (2011) G-quadruplex-induced instability during leading-strand replication. EMBO J. 30, 4033–4046 doi:10.1038/emboj.2011.316 82 Piazza, A., Serero, A., Boulé, J.-B., Legoix-Né, P., Lopes, J. and Nicolas, A. (2012) Stimulation of gross chromosomal rearrangements by the human CEB1 and CEB25 minisatellites in Saccharomyces cerevisiae depends on G-quadruplexes or Cdc13. PLoS Genet. 8, e1003033 doi:10.1371/journal. pgen.1003033 83 Wanzek, K., Schwindt, E., Capra, J.A. and Paeschke, K. (2017) Mms1 binds to G-rich regions in Saccharomyces cerevisiae and influences replication and genome stability. Nucleic Acids Res. 45, 7796–7806 doi:10.1093/nar/gkx467 84 Bernstein, K.A., Gangloff, S. and Rothstein, R. (2010) The RecQ DNA helicases in DNA repair. Annu. Rev. Genet. 44, 393–417 doi:10.1146/ annurev-genet-102209-163602 85 Thangavel, S., Mendoza-Maldonado, R., Tissino, E., Sidorova, J.M., Yin, J., Wang, W. et al. (2010) Human RECQ1 and RECQ4 helicases play distinct roles in DNA replication initiation. Mol. Cell. Biol. 30, 1382–1396 doi:10.1128/MCB.01290-09 86 Huppert, J.L. and Balasubramanian, S. (2007) G-quadruplexes in promoters throughout the human genome. Nucleic Acids Res. 35, 406–413 doi:10.1093/nar/gkl1057 87 Eddy, J. and Maizels, N. (2006) Gene function correlates with potential for G4 DNA formation in the human genome. Nucleic Acids Res. 34, 3887–3896 doi:10.1093/nar/gkl529 88 Agarwal, T., Roy, S., Kumar, S., Chakraborty, T.K. and Maiti, S. (2014) In the sense of transcription regulation by G-quadruplexes: asymmetric effects in sense and antisense strands. Biochemistry 53, 3711–3718 doi:10.1021/bi401451q 89 Agarwal, T., Lalwani, M.K., Kumar, S., Roy, S., Chakraborty, T.K., Sivasubbu, S. et al. (2014) Morphological effects of G-quadruplex stabilization using a small molecule in zebrafish. Biochemistry 53, 1117–1124 doi:10.1021/bi4009352 90 Hershman, S.G., Chen, Q., Lee, J.Y., Kozak, M.L., Yue, P., Wang, L.-S. et al. (2008) Genomic distribution and functional analyses of potential G-quadruplex-forming sequences in Saccharomyces cerevisiae. Nucleic Acids Res. 36, 144–156 doi:10.1093/nar/gkm986 91 Balasubramanian, S. and Neidle, S. (2009) G-quadruplex nucleic acids as therapeutic targets. Curr. Opin. Chem. Biol. 13, 345–353 doi:10.1016/j. cbpa.2009.04.637 92 Verma, A., Halder, K., Halder, R., Yadav, V.K., Rawal, P., Thakur, R.K. et al. (2008) Genome-wide computational and expression analyses reveal G-quadruplex DNA motifs as conserved cis-regulatory elements in human and related species. J. Med. Chem. 51, 5641–5649 doi:10.1021/jm800448a 93 Simonsson, T., Pecinka, P. and Kubista, M. (1998) DNA tetraplex formation in the control region of c-myc. Nucleic Acids Res. 26, 1167–1172 doi:10.1093/nar/26.5.1167 94 Michelotti, G.A., Michelotti, E.F., Pullner, A., Duncan, R.C., Eick, D. and Levens, D. (1996) Multiple single-stranded cis elements are associated with activated chromatin of the human c-myc gene in vivo. Mol. Cell. Biol. 16, 2656–2669 doi:10.1128/MCB.16.6.2656 95 Grand, C.L., Han, H., Muñoz, R.M., Weitman, S., Von Hoff, D.D., Hurley, L.H. et al. (2002) The cationic porphyrin TMPyP4 down-regulates c-MYC and human telomerase expression and inhibits tumor growth in vivo. Mol. Cancer Ther. 1, 565–573 PMID:12479216 96 Ou, T.-M., Lu, Y.-J., Zhang, C., Huang, Z.-S., Wang, X.-D., Tan, J.-H. et al. (2007) Stabilization of G-quadruplex DNA and down-regulation of oncogene c-myc by quindoline derivatives. J. Med. Chem. 50, 1465–1474 doi:10.1021/jm0610088

© 2017 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 9 Biochemical Society Transactions (2017) DOI: 10.1042/BST20170097

97 Siddiqui-Jain, A., Grand, C.L., Bearss, D.J. and Hurley, L.H. (2002) Direct evidence for a G-quadruplex in a promoter region and its targeting with a small molecule to repress c-MYC transcription. Proc. Natl Acad. Sci. U.S.A. 99, 11593–11598 doi:10.1073/pnas.182256799 98 Sun, D. and Hurley, L.H. (2009) The importance of negative superhelicity in inducing the formation of G-quadruplex and i-motif structures in the c-Myc promoter: implications for drug targeting and control of gene expression. J. Med. Chem. 52, 2863–2874 doi:10.1021/jm900055s 99 Byrd, A.K. and Raney, K.D. (2015) A parallel quadruplex DNA is bound tightly but unfolded slowly by Pif1 helicase. J. Biol. Chem. 290, 6482–6494 doi:10.1074/jbc.M114.630749 100 Fekete, A., Kenesi, E., Hunyadi-Gulyas, E., Durgo, H., Berko, B., Dunai, Z.A. et al. (2012) The guanine-quadruplex structure in the human c-myc gene’s promoter is converted into B-DNA form by the human poly(ADP-ribose)polymerase-1. PLoS ONE 7, e42690 doi:10.1371/journal.pone.0042690 101 Benhalevy, D., Gupta, S.K., Danan, C.H., Ghosal, S., Sun, H.-W., Kazemier, H.G. et al. (2017) The human CCHC-type zinc finger nucleic acid-binding protein binds G-rich elements in target mRNA coding sequences and promotes translation. Cell Rep. 18, 2979–2990 doi:10.1016/j.celrep.2017.02.080 102 Huber, M.D., Lee, D.C. and Maizels, N. (2002) G4 DNA unwinding by BLM and Sgs1p: substrate specificity and substrate-specific inhibition. Nucleic Acids Res. 30, 3954–3961 doi:10.1093/nar/gkf530 103 Fry, M. and Loeb, L.A. (1999) Human Werner syndrome DNA helicase unwinds tetrahelical structures of the fragile X syndrome repeat sequence d

(CGG)n. J. Biol. Chem. 274, 12797–12802 doi:10.1074/jbc.274.18.12797 104 Johnson, J.E., Cao, K., Ryvkin, P., Wang, L.-S. and Johnson, F.B. (2010) Altered gene expression in the Werner and Bloom syndromes is associated with sequences having G-quadruplex forming potential. Nucleic Acids Res. 38, 1114–1122 doi:10.1093/nar/gkp1103 105 Nguyen, G.H., Tang, W., Robles, A.I., Beyer, R.P., Gray, L.T., Welsh, J.A. et al. (2014) Regulation of gene expression by the BLM helicase correlates with the presence of G-quadruplex DNA motifs. Proc. Natl Acad. Sci. U.S.A. 111, 9905–9910 doi:10.1073/pnas.1404807111 106 Gray, L.T., Vallur, A.C., Eddy, J. and Maizels, N. (2014) G quadruplexes are genomewide targets of transcriptional helicases XPB and XPD. Nat. Chem. Biol. 10, 313–318 doi:10.1038/nchembio.1475 107 Wang, C.-C., Chen, J.J.W. and Yang, P.-C. (2006) Multifunctional transcription factor YY1: a therapeutic target in human cancer? Expert Opin. Ther. Targets 10, 253–266 doi:10.1517/14728222.10.2.253 108 Huang, W., Smaldino, P.J., Zhang, Q., Miller, L.D., Cao, P., Stadelman, K. et al. (2012) Yin Yang 1 contains G-quadruplex structures in its promoter and 50-UTR and its expression is modulated by G4 resolvase 1. Nucleic Acids Res. 40, 1033–1049 doi:10.1093/nar/gkr849 109 Yonaha, M. and Proudfoot, N.J. (1999) Specific transcriptional pausing activates polyadenylation in a coupled in vitro system. Mol. Cell 3, 593–600 doi:10.1016/S1097-2765(00)80352-4 110 Gromak, N., West, S. and Proudfoot, N.J. (2006) Pause sites promote transcriptional termination of mammalian RNA polymerase II. Mol. Cell. Biol. 26, 3986–3996 doi:10.1128/MCB.26.10.3986-3996.2006 111 Wanrooij, P.H., Uhler, J.P., Simonsson, T., Falkenberg, M. and Gustafsson, C.M. (2010) G-quadruplex structures in RNA stimulate mitochondrial transcription termination and primer formation. Proc. Natl Acad. Sci. U.S.A. 107, 16072–16077 doi:10.1073/pnas.1006026107 112 Zheng, K.-W., Wu, R.-Y., He, Y.-D., Xiao, S., Zhang, J.-Y., Liu, J.-Q. et al. (2014) A competitive formation of DNA:RNA hybrid G-quadruplex is responsible to the mitochondrial transcription termination at the DNA replication priming site. Nucleic Acids Res. 42, 10832–10844 doi:10.1093/nar/ gku764 113 Duquette, M.L., Handa, P., Vincent, J.A., Taylor, A.F. and Maizels, N. (2004) Intracellular transcription of G-rich DNAs induces formation of G-loops, novel structures containing G4 DNA. Genes Dev. 18, 1618–1629 doi:10.1101/gad.1200804 114 Wu, R.-Y., Zheng, K.-W., Zhang, J.-Y., Hao, Y.-H. and Tan, Z. (2015) Formation of DNA:RNA hybrid G-quadruplex in bacterial cells and its dominance over the intramolecular DNA G-quadruplex in mediating transcription termination. Angew. Chem. Int. Ed. Engl. 54, 2447–2451 doi:10.1002/anie. 201408719 115 Chakraborty, P. and Grosse, F. (2011) Human DHX9 helicase preferentially unwinds RNA-containing displacement loops (R-loops) and G-quadruplexes. DNA Repair 10, 654–665 doi:10.1016/j.dnarep.2011.04.013 116 Lee, T. and Pelletier, J. (2016) The biology of DHX9 and its potential as a therapeutic target. Oncotarget 7, 42716–42739 doi:10.18632/oncotarget. 8446 117 Joachimi, A., Benz, A. and Hartig, J.S. (2009) A comparison of DNA and RNA quadruplex structures and stabilities. Bioorg. Med. Chem. 17, 6811–6815 doi:10.1016/j.bmc.2009.08.043 118 von Hacht, A., Seifert, O., Menger, M., Schutze, T., Arora, A., Konthur, Z. et al. (2014) Identification and characterization of RNA guanine-quadruplex binding proteins. Nucleic Acids Res. 42, 6630–6644 doi:10.1093/nar/gku290 119 Guo, J.U. and Bartel, D.P. (2016) RNA G-quadruplexes are globally unfolded in eukaryotic cells and depleted in bacteria. Science 353, aaf5371 doi:10.1126/science.aaf5371 120 Thandapani, P., Song, J., Gandin, V., Cai, Y., Rouleau, S.G., Garant, J.-M. et al. (2015) Aven recognition of RNA G-quadruplexes regulates translation of the mixed lineage leukemia protooncogenes. eLife 4, e06234 doi:10.7554/eLife.06234 121 Nie, J., Jiang, M., Zhang, X., Tang, H., Jin, H., Huang, X. et al. (2015) Post-transcriptional regulation of Nkx2-5 by RHAU in heart development. Cell Rep. 13, 723–732 doi:10.1016/j.celrep.2015.09.043 122 Booy, E.P., Howard, R., Marushchak, O., Ariyo, E.O., Meier, M., Novakowski, S.K. et al. (2014) The RNA helicase RHAU (DHX36) suppresses expression of the transcription factor PITX1. Nucleic Acids Res. 42, 3346–3361 doi:10.1093/nar/gkt1340 123 Newman, M., Sfaxi, R., Saha, A., Monchaud, D., Teulade-Fichou, M.P. and Vagner, S. (2016) The G-quadruplex-specific RNA helicase DHX36 regulates p53 pre-mRNA 30-end processing following UV-induced DNA damage. J. Mol. Biol. PMID:27940037 doi:10.1016/j.jmb.2016.11.033 124 Chalupníková, K., Lattmann, S., Selak, N., Iwamoto, F., Fujiki, Y. and Nagamine, Y. (2008) Recruitment of the RNA helicase RHAU to stress granules via a unique RNA-binding . J. Biol. Chem. 283, 35186–35198 doi:10.1074/jbc.M804857200 125 McRae, E.K.S., Booy, E.P., Moya-Torres, A., Ezzati, P., Stetefeld, J. and McKenna, S.A. (2017) Human DDX21 binds and unwinds RNA guanine quadruplexes. Nucleic Acids Res. 45, 6656–6668 doi:10.1093/nar/gkx380 126 Wolfe, A.L., Singh, K., Zhong, Y., Drewe, P., Rajasekhar, V.K., Sanghvi, V.R. et al. (2014) RNA G-quadruplexes cause eIF4A-dependent oncogene translation in cancer. Nature 513,65–70 doi:10.1038/nature13485

10 © 2017 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society