Open access peer-reviewed chapter

The Role of Telomeres and Telomere-associated Proteins as Components of Interactome in Cell-signaling Pathways

Written By

Ekta Khattar and Vinay Tergaonkar

Submitted: 11 May 2015 Reviewed: 08 December 2015 Published: 23 November 2016

DOI: 10.5772/62130

From the Edited Volume

Telomere - A Complex End of a Chromosome

Edited by Marcelo L. Larramendy

Chapter metrics overview

2,033 Chapter Downloads

View Full Metrics

Abstract

Telomeres represent ends of all eukaryotic chromosomes and serve specialized biological role in maintaining genomic integrity by preventing end fusions and degradation. Various protein complexes associate with telomeres to either protect them from DNA damage machinery or maintain telomere length homeostasis. These protein complex subunits cross talk with a variety of cell-signaling components to either maintain telomere integrity or perform other functions, which are either dependent or independent of telomeres and/or their telomeric role. Mutations in these protein components lead to the development of various human diseases, such as age-related disorders, which occur mainly due to telomere dysfunction or cancer development due to telomerase reactivation. This chapter focuses on the structural and functional aspects of telomeric proteins and their importance in human diseases.

Keywords

  • Telomeres
  • shelterin
  • telomerase
  • TERT
  • telomere diseases
  • cancer

1. Introduction

Human telomeres consist of TTAGGG tandem repeats, which are generally 3–15 kbp in length [1]. The distal end of telomere has a 3′ single-stranded overhang, which is also termed a G-rich strand, and it forms a higher order structure (like a lariat) named t-loop [2]. In t-loop, both strands of the chromosome are joined to an earlier point in the double-stranded DNA by the 3′ strand end invading the strand pair to further form a D-loop. Formation of the D-loop completes the t-loop, thus establishing a capping structure, which protects chromosomes from degradation and recombination [3]. Figure 1A shows a schematic representation of telomere structure. The disruption of t-loop results in telomere dysfunction and induction of DNA damage response (DDR) followed by cell cycle arrest [4]. Telomeres are bound by nucleosomes and a specialized complex known as shelterin, which is composed of six core protein subunits [5]. Shelterin determines the structure of telomeres. It is implicated in the formation of t-loops and also regulates the synthesis of telomeric DNA [6]. Additional proteins capable of interacting with shelterin proteins, such as DNA damage proteins, also play a role in maintaining telomere length and chromosomal stability [7].

Figure 1.

Schematic representation of telomere end structure (A) and Telomere associated protein complexes (B).

Telomeres shorten with replication due to two major mechanisms: (A) end-replication problem and exonuclease-mediated resection in dividing cells, and (B) damage response to reactive oxygen species in nondividing cells [8].

DNA replication involves simultaneous duplication of antiparallel DNA strands, such that replication advances in opposite directions, across a leading strand and a lagging strand. On the leading strand, daughter strand synthesis takes place continuously in the 5′–3′ direction, whereas on the lagging strand template, DNA synthesis proceeds in the 5′–3′ direction discontinuously, leading to Okazaki fragments. The leading daughter strand is completely synthesized until DNA polymerase reaches 5′ end of the leading template. However, a primer is required for DNA replication to start. At the end of replication, RNA primer occupying the 5′ end of the daughter strand is removed, and it is not possible for the overlapping strand to be replicated. Due to this, the 5′ end of each antiparallel daughter strand becomes one primer length shorter. This is referred to as the end-replication problem, which results in chromosome shortening with each subsequent cell division. Theoretically, it should result in a loss of less than 10 bp with each replication cycle; however, the rate of loss is much higher and has been calculated to be 50–200 bp per division [9]. Exonuclease activity degrading the 5′ end is another major factor, which removes the RNA primer on the lagging strand and thus also leads to the formation of 3′-end overhang structure [10]. In vitro studies have also suggested the role of oxidative stress in telomere loss [11]. Correlative and experimental studies have also suggested links between oxidative damage and telomere loss in vivo [12]. Therefore, telomere length also serves as a biological clock and marker for chronological ageing. The solution to telomere shortening is the telomerase enzyme complex, which catalyzes de novo addition of TTAGGG repeats to chromosome ends, thus preventing telomere attrition [13].

Advertisement

2. Telomere-Associated Protein Complexes (TAPs)

2.1. Shelterin

Shelterin, as the name suggests, provides shelter to the ends of linear chromosomes by repressing DNA damage-signaling responses, masking telomeres from DNA repair machinery, and regulating the length of telomeres [7, 14]. Shelterin is a highly specialized complex comprising six central components namely, TTAGGG repeat-binding factors TRF1 and TRF2, TRF1-interacting protein (TIN2), protection of telomeres protein (POT1), Pot1-interacting protein (TPP1), and repressor/activator protein (RAP1) [15]. All shelterin components are ubiquitously expressed and associated with telomeres throughout the cell division cycle. The complete abrogation of all shelterin subunits (except RAP1) in mice results in embryonic lethality, thereby implicating their essential roles in development [16].

TRF1

TRF1 was the first shelterin complex subunit to be discovered that specifically associates with double-stranded telomeric DNA, mainly as a dimer through its TRF homology domain. Recently, it has been shown that TRF1 efficiently associates with telomeric DNA in nucleosomal context and is capable of remodeling telomeric nucleosomal arrays [17]. TRF1 interacts with TIN2 in the shelterin complex as shown in Figure 1B. It also functions as a negative regulator of telomere length in telomerase-positive cells. In addition, some reports have demonstrated that it is essential for survival independent of its telomere length regulatory activity [18]. TRF1 genome-wide binding analysis revealed that it exclusively localizes at telomeres under normal conditions as well as under extreme telomere shortening unlike other shelterin members which have extra telomeric roles [19]. TRF1 has also been shown to assist Aurora-B recruitment to centromeres, thus contributing to appropriate chromosome segregation and maintenance of genomic integrity [20].

TRF2

TRF2 is highly similar to TRF1 in terms of protein sequence except that the N-terminal domain in TRF2 is acidic while that in TRF1 is rich in glycine and arginine residues (forming GAR domain). It possesses TRF homology domain, which mediates its dimerization. This N-terminal domain of TRF1 and TRF2 has been shown to regulate their ability to condense telomeric DNA [21]. TRF2 has been proposed to stabilize the t-loop by invasion of the upstream TTAGGG double-stranded region [22]. TRF2 has also been shown to bind at internal genomic regions, mainly at TTAGGG repeats referred to as interstitial telomeric sequences (ITSs) [23, 24]. Recent evidence suggests the role of TRF2 in the formation of novel chromosome end structures, which involve telomeres interacting with nontelomeric DNA, forming long-range chromosome loop that encompasses several megabases of chromatin and are known as interstitial telomeric loops (ITLs) [25]. Telomere-bound TRF2 is necessary to suppress the ataxia telangiectasia mutated (ATM)-dependent DNA damage response pathway [26] and the nonhomologous end joining (NHEJ) DNA repair pathway, thus playing a major role in protecting chromosome ends [9]. TRF2 also assists telomere replication by limiting resolvase activities leading to accurate repair of stalled forks [27]. It has been demonstrated that both TRF1 and TRF2 are modified post-translationally; however, the physiological relevance of these modifications is not yet completely understood [7].

TIN2

TIN2 (encoded by TINF2) associates with both TRF1 and TRF2, thus forming a bridge that connects the double-stranded telomeric DNA-binding proteins to those bound to single-stranded telomeric overhang region [28, 29]. A recent study by Frank et al demonstrates that TIN2 facilitates the recruitment of telomerase to telomeres [30]. In this study, the authors discovered a novel mutation in TINF2 gene (which encodes TIN2 protein) and used novel functional assays to demonstrate a direct role for TIN2 in regulating telomere length through telomerase. This role is completely independent of its role in telomere protection.

TPP1

TPP1 interacts with TIN2 and POT1 through its carboxyl terminus and central domain, respectively [31-33]. Structurally, TPP1 possesses four domains, namely, OB (oligonucleotide/oligosaccharide-binding fold), RD (POT1 recruitment domain), S/T (serine-rich region), and TID (TIN2-interacting domain). It has been demonstrated that TPP1 recruits POT1 to telomeres [34, 35]. In addition, TPP1 has been shown to contain telomerase-interacting domain, suggesting a role for TPP1 in the recruitment of telomerase to chromosome ends [36]. Interestingly, it has been shown that TPP1–POT1 interacts with N-terminal region of telomerase protein subunit TERT (telomerase reverse transcriptase), while TPP1 alone is also capable of interacting with C-terminal TERT residues through TPP1-OB domain [37]. TPP1-OB domain has been shown to contain a TEL patch (TPP1 glutamate (E)- and leucine (L)-rich patch), which mediates telomerase recruitment and processivity independent of its role in telomere capping [38]. In human stem cells, it has been further proven that TPP1 indeed mediates recruitment and activation of telomerase and telomere length homeostasis [39].

POT1

POT1 interacts with TRF1 complex via protein–protein interaction with TPP1, and this interaction is believed to affect its loading on the single-stranded TTAGGG telomeric repeat [40]. POT1 has been predicted to associate with telomeric sequences at t-loop as well as D-loop through its OB-fold domains [32]. POT1 serves a fundamental role in telomere length regulation, since it functions as a terminal transducer of telomere length control [40]. POT1–TPP1 complex has been shown to increase RAP (consecutive repeats that telomerase can add to telomere before dissociating) with minimal effect on telomerase activity [41]. Further, POT1–TPP1 has been shown to reduce the dissociation rate of telomerase from its telomeric substrate and assist in the translocation step [42].

RAP1

RAP1 associates with TRF2 and thus localizes to telomeres. As a component of the shelterin complex, it is dependent on TRF2 [43, 44]. RAP1 has been shown to be indispensable for telomere function in mouse and human systems [45, 46]. Further, RAP1 has been shown to possess extra telomeric roles that will be described in later part of this chapter.

2.2. Telomerase complex

Telomeres in human somatic cells shorten with each replication cycle due to end-replication problem consequently leading to genomic instability and cell death. Telomerase elongates telomeres and thus maintain their replicative potential. The minimal components of telomerase are a catalytic protein subunit termed as telomerase reverse transcriptase (TERT) and an RNA template known as Terc (TR) [47]. Telomerase catalyzes the addition of TTAGGG sequences to the ends of the chromosomes during each replication cycle, thus preventing telomere attrition and maintaining genomic integrity. Telomerase activity is detected predominantly in stem cells and cancer cells [48]. In normal somatic cells, telomerase activity is almost undetectable, consequently leading to shortened telomeres, which limit their life span. Furthermore, while TR is highly expressed in all the cell types, the levels of TERT are highly regulated at transcriptional level [49]. In somatic cells, TERT is transcriptionally turned off while stem cells display high expression of TERT, resulting in high telomerase activity. Ninety percent of cancer cells reactivate TERT expression either by mutation in TERT promoter or by activating oncogenic transcription factors such as NF-κB, MYC, and β-catenin that are known to activate TERT transcription [50]. This reactivation of TERT confers cancer cells with unlimited replicative potential. Distinct from its telomeric function, TERT has also been shown to cross talk with various signaling pathways and impart several additional functions to cancer cells [51].

Terc (TR)

Mature human TR is a small noncoding RNA consisting of 451 nucleotides and serves as a template for de novo telomeric repeat synthesis by telomerase. Structurally, it is subdivided into four domains namely the pseudoknot domain, the CR4-CR5 domain, the CR7, and H/ACA domain [52]. These domains perform various functions including RNA binding, dimerization, and recruitment of telomerase to telomeres. The pseudoknot domain and the CR4/5 domain along with TERT constitute the minimal requirement for reconstituted in vitro telomerase activity [53, 54]. The CR7 and H/ACA domains are required for stability and localization of telomerase [55]. H/ACA motif consists of two hairpins connected by a short single-stranded stretch, the H box, and a terminal ACA region [56] which is found in small nucleolar and small Cajal body (CB) RNAs (snoRNAs and scaRNAs). H/ACA small nucleolar ribonucleoprotein complex (H/ACA snoRNP) catalyzes pseudouridylation of small RNAs like ribosomal RNA, which may serve to stabilize their conformation [57]. The H/ACA domain of TR is essential for assembly into ribonucleoprotein (RNP) with four member H/ACA–RNP complex which include the core heterotrimer (Dyskerin, NHP2, and NOP10) and a fourth factor, GAR1 [56]. Dyskerin represents the catalytic subunit of H/ACA complex [57]. Major function for the association of TR with dyskerin is its stabilization and nuclear retention. However, no pseudouridylation of TR has been reported [58]. Structurally, dyskerin contains a TruB domain that functions in eukaryotic ribosomal RNA processing. The TruB domain consists of two motifs, TruB I and TruB II. In addition, it has two nuclear localization (NL) signals, N-terminal and C-terminal, and the PUA, pseudouridine synthase and archaeosine transglycosylase, domain involved in RNA modification [59]. Dyskerin and NOP10 form the stable core to which GAR1 and NHP2 subsequently bind [60]. Mutations in these proteins affect the stability of TR consequently affecting telomere synthesis and result in various human diseases discussed in a later section of the chapter.

Although TR is transcribed by RNA polymerase II, it is not polyadenylated; instead, its 3′ end is formed by exonucleolytic cleavage up to the boundary element formed by the H/ACA domain, where further cleavage is prevented by dyskerin associated with RNA [61]. The tetrameric complex of dyskerin, NOP10, NHP2, and the chaperone NAF1 associates with TR cotranscriptionally and is essential for its accumulation [62]. The significance of this complex formation is highlighted by a number of telomere length-associated disorders with mutations in these factors which results in reduced levels of TR and thus telomerase activity [63].

TR also harbors specialized sequence elements in the terminal loop of its 3′ hairpin namely BIO box, which stimulates TR stability by H/ACA RNP formation, and CAB box, which it shares with the scaRNAs. CAB box is required for trafficking TR to the CB, where it receives its 2,2,7-trimethyl guanosine (TMG) cap, and NAF1 is replaced by GAR1 [57, 64, 65].

TERT

Human TERT is a large protein consisting of 1133 amino acids and it is subdivided into four distinct domains: (a) N-terminal extension (TEN), (b) TERT RNA-binding domain (TRBD), (c) the reverse transcriptase domain (RT), and (d) the C-terminal extension (CTE) [66]. TEN domain is essential for telomerase activity and functions in proper localization and correct positioning of its catalytic site on telomeric DNA [67]. The TRBD domain functions in telomerase RNP assembly as well as RNA binding [68]. The RT domain of TERT forms the catalytic center in telomerase complex and consists of seven universally conserved RT motifs [6971]. RT domain can be divided into two putative subdomains namely the fingers and the palm domains where fingers domain interacts with nucleic acid substrate while the palm domain contains the catalytic site [72]. CTE possesses the thumb domain and is sequentially not conserved among species [66]. CTE serves a critical role in catalytic activity and processivity of telomerase [73].

TERT interacts with chaperones HSP90 and p23 as well as with AAA+ ATPases pontin and reptin [74, 75]. HSP90 and p23 interact with active telomerase; however, reptin and pontin are reported to interact with a pool of TERT, which is not assembled into active complex suggesting their role in telomerase assembly [75]. In addition, reptin and pontin are known to interact with dyskerin and are necessary for H/ACA RNP assembly, which is an essential step in TR stability [75, 76].

TERT transcription and telomerase activity is highest in the S phase of the cell cycle [77], and telomerase recruitment to telomeres has been shown to be restricted to the S phase [78, 79].

TERT interacts with two TR elements. The TRBD associates with CR4/5 region of TR and RT domain of TERT associates with pseudoknot region of TR [53, 80]. The human telomerase RNP purified from HEK293T cells overexpressing TERT and TR has been shown to be a dimeric structure, which is around 28 nm in length [81]. Although many studies have suggested the existence of multiple copies of dyskerin, NOP10, NHP2, and GAR1 with human telomerase RNP, the presence of two catalytically active TERT has been controversial since its biological significance is not clearly understood [82].

Telomerase activity in cells is limited by the levels of TERT protein (reported to be around 600 molecules/cell). All the other components of telomerase RNP are abundant [77, 83]. In normal somatic cells, TERT expression is repressed epigenetically or due to lack of activating transcription factors such as MYC, NF-κB, NFAT, RAS/RAF pathway, Ets factor steroids, and HIF [84]. Thus, transcriptional reactivation of TERT represents one of the major mechanisms responsible for activating telomerase and thus achieving replicative immortality in cancers. Recently, many cancers have been reported to harbor TERT promoter mutations resulting in high TERT expression and telomerase activity [85]. This is further discussed in later sections of the chapter.

2.3. Accessory proteins/complexes/factors for proper telomere maintenance

Regulator of telomere elongation helicase 1 (RTEL1)

RTEL1 is a DNA helicase, which contains N-terminal helicase domains and a C-terminal extension [86]. TRF2 recruits RTEL1 to telomeres during S phase of cell cycle to assist t-loop disassembly. RTEL1 is required for stability, protection, and elongation of telomeres [87]. RTEL1 has also been shown to interact with proliferating cell nuclear antigen (PCNA) and this interaction is important to prevent telomere fragility [87].

CST complex

The human CST complex consisting of CTC1, STN1, and TEN1 proteins plays a role in telomere protection and DNA metabolism [88]. Each telomerase RNP is believed to add 50–60 nucleotides to most telomeres following a single initiation event [89]. The CST complex has been proposed to set the upper limit of telomere elongation by binding to telomeric single-stranded DNA (ssDNA) and displacing telomerase, once telomeric overhang has reached certain length [90]. Study reported by Chen et al showed that CST competes with POT1–TPP1 for telomeric DNA [90]. It terminates telomerase activity through primer sequestration and physical association with the POT1–TPP1 subunits, which functions as a telomerase processivity factor [91]. CST–telomeric-DNA binding increases during late S/G2 phase following telomerase activity, concurrently with telomerase turn-off. Attenuation of CST enables excessive telomerase activity, fostering telomere elongation. It is suggested that through binding of the telomerase-extended telomere, CST limits telomerase activity at individual telomeres to approximately one binding and extension event per cell cycle.

Telomerase Cajal body protein 1 (TCAB1)

TCAB1 (encoded by gene WDR79) contains a proline-rich region and WD40 motif and is localized in CBs. CBs have been shown to accumulate telomerase as well as associate with telomeres [78]. TCAB1 has been demonstrated to be a component of active telomerase and is necessary for the telomerase holoenzyme to accumulate in CB and thus regulate telomere elongation [92]. TCAB1 interacts with CAB box motif of TR and functions in telomerase assembly by driving telomerase to CB. Subsequent report by Stern et al showed that TCAB1 and CB are required for telomerase recruitment to telomeres independent of each other [93].

TERRA

TERRA functions as a negative regulator of telomere length. It is transcribed from subtelomeric regions of telomeres that consist of UUAGGG repeats. TERRA may inhibit telomerase in cis by directly binding to telomerase [94, 95].

Advertisement

3. Cross talk of telomere-associated proteins with cellular signaling pathways

3.1. TAPs and DNA damage response

When the genomic DNA undergoes any damage such as single-strand DNA breaks, double-strand breaks (DSBs), nicks, or chromosome fusions, cells activate DNA repair pathway depending on the type of damage. During this process, cells seize to grow and initiate the repair, and once the repair is completed growth resumes; otherwise, they undergo apoptosis. Telomeres can also be sensed as breaks by cellular machinery. However, telomerase and shelterin complex cap the telomeres and thus maintain telomere integrity by inhibiting DNA-damage-response pathway at telomeres. Further, some of these DNA repair proteins play an essential role in telomere maintenance by directly associating with various TAPs. Recently, these TAPs have been shown to play a role in DDR at locations distinct from telomeres. TRF1, which is predominantly localized to telomeres, has been shown to interact with proteins implicated in DNA damage response such as ATM and Mre11/Rad50/Nbs1 [96]. Further, it has been shown that TRF1 is phosphorylated by Cdk1 and this form is incapable of binding to telomeres [97]. It has been shown that this phosphorylated TRF1 in a telomere-independent way facilitates end resection and homology repair (HR), activates G2/M checkpoint, and enables cell survival following double-strand break induction [98]. TRF2 is known to prevent activation of ATM-dependent DDR pathway at telomeres [26, 99, 100]. TRF2 has also been shown to localize at DNA double-strand breaks during early stages (within 2 s) of cellular response to DSBs and leaves those sites during repair processing [101]. Further reports suggested that TRF2 may also participate in HR of the extra telomeric damaged DNA [102].

It has been shown that telomerase localizes to mitochondria and protects cancer cells from nuclear DNA damage and apoptosis [103]. Further, TERT has also been shown to function in DNA damage response pathway and in regulating histone-dependent chromatin remodeling [104]. TR knockout mice display impaired DDR in response to damaging agents; however, the effects are evident clearly in late-generation TR null mice, which show significant telomere shortening and dysfunction [105, 106].

3.2. TAPs and NF-κB pathway

NF-κB transcription factors are key mediators of various cellular, inflammatory, and development pathways [107]. NF-κB family consists of five transcription factors namely RelA (p65), RelB, c-Rel, and also includes p50 and p52 (processed from p105 and p100, respectively). Rel family proteins possess REL homology domain, which harbors DNA-binding domain, dimerization, and IκB-binding domain. NF-κB family members are held inactive in cytoplasm by IκB family of proteins. Upon stimulation, IkB proteins are phosphorylated by IκB kinase (IKK) complex and thus degraded or processed to produce active dimers, which can enter the nucleus and activate the transcription response [107]. TAFs, which have been shown to modulate NF-κB signaling, independent of their role at telomeres, include shelterin complex protein RAP1 and telomerase complex protein TERT [108, 109].

A fraction of cytosolic RAP1 associates with functional IKK complex. RAP1 increases the efficiency of IKK complex in phosphorylating p65 subunit at serine 536; however, it has no effect on the degradation of IκBa inhibitory protein, thus functioning as an adaptor in the IKK complex [110]. In line with this, it was recently shown that RAP1 regulates cytokine levels followed by fine healing of corneal injury by effective modulation of NF-κB signaling [111]. Since RAP1 is also localized in cytoplasm, there are possibilities that it might function as an adaptor in various other complexes under different stimuli.

TERT has also been shown to associate with p65 in the nuclear compartment, thus directly regulating its transcriptional response. Akiyama et al demonstrated that TERT associates with p65 to mediate its translocation to nucleus in multiple myeloma cells [112]. Recently, our group showed that TERT associates with p65 to directly affect its transcriptional output [113]. We showed that TR-knockout mice are more resistant to inflammatory agent lipopolysaccharide (LPS), which majorly functions by activating NF-κB pathway. Ectopic expression of TERT led to increased proliferation of cancer cells as well as xenograft model, which could be abrogated by inhibiting p65. Subsequently, another group showed that TERT regulates matrix metalloproteinase (MMP) expression independently of telomerase activity via activation of NF-κB-dependent transcription [114].

3.3. TAPs and WNT/β-catenin pathway

WNT pathway plays a key role in development processes like cell-fate determination, progenitor cell proliferation, and cell polarity [115]. In canonical WNT pathway, WNT ligand binds to its receptor leading to stabilization of β-catenin in the cytoplasm. Stabilized β-catenin then enters nucleus to activate transcription through its interaction with TCF/Lef family members. It further recruits chromatin remodelers like BRG1 to facilitate transcription.

TERT is the only TAP shown to modulate WNT pathway independent of telomeres and telomerase catalytic activity. The first evidence demonstrating a link between TERT and WNT pathway came from Choi et al who showed that knock in of catalytically inactive TERT in hair follicle stem cells led to their proliferation [116]. Changes in gene expression as analyzed by microarray, revealed differential expression of genes involved in development/morphogenesis, signal transduction, and cytoskeleton/cell adhesion signaling pathways. Modulated gene expression pattern strongly correlated with transcriptional program of MYC and WNT, suggesting existence of a potential association of TERT with the WNT and MYC pathways. Subsequently, Park et al demonstrated the first evidence of the direct regulation of Wnt/β-catenin signaling by telomerase in mouse embryonic stem cells and Xenopus laevis embryos. The study reported that TERT functions as a cofactor in the β-catenin transcriptional complex through its interaction with Brg1, a chromatin-remodeling factor [117]. Ectopic expression of TERT or catalytically inactive TERT led to the activation of WNT-dependent reporters in vitro and in vivo, while chromatin immunoprecipitation assays uncovered TERT localization at WNT target gene promoters. It was shown that TERT null mice display partially penetrant homeotic transformation of vertebrae, due to the loss of 13th rib of one or both the vertebrae. Recently, another group reported that TERT forms a complex with Brg1, together with nucleostemin (NS), a nucleolar GTP-binding protein and/or its family member GNL3L and is essential for maintenance of the tumor-initiating cell phenotype in human cancer cells [118].

3.4. TAPs and MYC

Recently, we reported that TERT regulates MYC transcription [119]. It is well known that MYC directly regulates TERT transcription [120]. However, our study illuminated the existence of a feed-forward loop between TERT and MYC in MYC-driven cancers such as lymphomas. Using genetic and biochemical approaches, we showed that the absence of TERT delayed MYC-dependent lymphomagenesis and strikingly, this effect was not observed when the RNA component of telomerase, TR was removed. Using in vivo and in vitro approaches, we established that TERT stabilizes MYC and thus results in increased MYC-dependent transcriptional output. Furthermore, we showed that this effect of TERT on MYC stability was independent of its catalytic activity. Mechanistically, we showed that TERT associates with MYC, preventing its proteasomal degradation, thus stabilizing its protein levels [119].

3.5. TAPs and mitochondria

Among TAPs, TERT and TIN2 have been shown to be involved in regulating mitochondrial activity. It has been shown that TERT translocates to mitochondria under certain stress conditions [121123]. Mitochondrial TERT binds to and protects mitochondrial DNA from hydrogen peroxide-induced oxidative damage [103, 124]. Overexpression and knockdown studies involving TERT in cancer cells have shown that the role of TERT in mitochondrial pathway of apoptosis is independent of its catalytic activity [122, 125]. Interestingly, it has also been shown that TERT functions as a reverse transcriptase in mitochondria using mitochondrial tRNA as a template [126]. Furthermore, it has been shown that TERT can interact with RNA component of mitochondrial RNA-processing endoribonuclease (RMRP) to form a complex similar to RNA-dependent RNA polymerase (RdRP). This complex then affects gene silencing at the post-transcriptional level [126].

TIN2 has also been shown to localize to mitochondria, where it results in altered mitochondrial structure. The group showed that the reduction of TIN2 levels led to augmented mitochondrial oxidative phosphorylation and reduced aerobic glycolysis in cancer cells [127].

3.6. TAPs and miscellaneous associations

Apart from the above-described associations of TAPs with cellular machinery, there are various reports about many more interacting partners. TRF2 has been shown to function as a transcriptional activator by directly binding to promoter of the angiogenic tyrosine kinase platelet-derived growth factor receptor β (PDGFRβ). This study highlighted the angiogenic role of TRF2 uncoupled from its telomere-capping role [128]. Telomerase was shown to regulate rDNA transcription by directly associating with RNA polymerase I upon hyperproliferative stimuli such as during liver regeneration and Ras-induced hyperproliferation [129].

Figure 2 summarizes the role of telomere-interacting proteins in cross talk with cellular signaling pathways.

Figure 2.

Schematic representation showing Telomere associated proteins interacting with several cell-signaling pathways.

Advertisement

4. Telomeres and TAPs in human diseases: Telomeropathies

Telomeres shorten with each cell division. When telomeres become excessively short, they lose their protective role and activate a DNA damaging signal response resulting in genomic instability, cell cycle arrest, and senescence. TAPs play an essential role in maintaining telomere length, and genetic mutations affecting their activity can result in telomere dysfunction. This manifests into a wide variety of diseases collectively named as “telomeropathies” or “telomere syndromes”, which exhibit impaired telomere maintenance.

4.1. Telomere-shortening syndromes

Inherited mutations, which hamper telomerase or telomere maintenance genes, result in progressive shortening of telomeres. Telomere shortening has major impact on highly proliferating tissues, such as bone marrow, where stem cells reach senescence stage and organ failure might ensue. Clinical conditions associated with shortened telomeres may be very different. This may be partly due to genetic anticipation since telomere length is inherited [63].

4.1.1. Dyskeratosis congenita

Dyskeratosis congenita (DC) arises primarily due to bone marrow failure and is associated with a diagnostic triad of oral leukoplakia, skin hyperpigmentation, nail dystrophy, and other manifestations. Dyskerin (encoded by DKC1), which is an essential component of telomerase enzyme in vivo, was the first gene identified as a cause of DC, and was thus named after this syndrome. DC is a heterogeneous disease showing all modes of inheritance. To date, 11 genes have been associated with DC. These include genes encoding products involved in telomere elongating enzyme, telomerase components (TERT and TR), telomerase stability (dyskerin, NOP10, NHP2), telomerase recruitment (TIN2 and TPP1), telomerase trafficking (TCAB1), telomerase docking (CTC1), and telomere replication (RTEL1) [130].

Figure 3 shows schematic representation of telomere-interacting proteins with domains and positions of reported germ-line mutations, which result in various forms of DC.

Figure 3.

Schematic representation showing protein structure and localization of reported mutations in telomere associated proteins. The information is adapted from Espinoza et al [144].

Recently poly(A)-specific ribonuclease (PARN) gene mutations have been reported in a subgroup of patients with DC wherein PARN deficiency results in reduced stability of several key TAPs (dyskerin, TRF1, RTEL1, and TR) and specifically leads to telomere attrition [131]. Almost all modes of inheritance have been reported in DC, which include autosomal recessive, autosomal dominant, and X-linked. Based on functional relevance of mutated proteins and their penetrance, clinically diverse variant manifestations of DC are reported.

Calado et al reported a study of five families who were suffering from liver disease (familial liver cirrhosis) in combination with hematologic and autoimmune disorders [132]. They analyzed the mutations associated with the disease and found rare loss of function mutations in TERT or TR (3.7% vs 0.8%). Hoyeraal–Hreidarsson syndrome (HHS) is associated with intrauterine growth retardation (IUGR), microcephaly, cerebellar hypoplasia, and thrombocytopenia along with various nonspecific enteropathies. HHS patients are also found to harbor DC mucocutaneous triad in adulthood. Detailed analysis revealed mutations in DKC1, TIN2 along with some cases showing autosomal recessive mutations in TERT, NHP2, NOP10, TPP1, and RTEL1 genes [133, 134]. Revesz syndrome (RS) is associated with various disease manifestations mainly bilateral exudative retinopathy. Other symptoms reported include IUGR, intracranial calcifications, developmental delay, and nail dystrophy in different cases, which were highly overlapping with DC symptoms. It was discovered that RS patients have short telomeres and harbor germ-line mutations in TINF2 gene [135]. Coats plus syndrome (CCS) is a rare recessive disorder that is characterized by intracranial calcifications, hematological abnormalities, and retinal vascular defects. CCS patients display shortened telomeres indicating telomere dysfunction as a major cause. Missense mutations in CTC1 gene whose protein is a part of CST complex has been reported to occur in CCS patients [136]. HHS, RS, and CCS represent severe forms of DC.

About 10% of DC patients develop cancer at a very young age. Various DC families display an increased incidence of acute myeloid leukemia and myelodysplastic syndrome [137]. Spontaneous reversion to the functional TR allele in hematopoietic stem cells of haploinsufficient DC patients has been observed predisposing them to hematological disorders. The mechanism behind high cancer incidence, in spite of short telomeres that should have cancer-protecting effect, remains largely unexplained. The only proposed mechanism is genomic instability due to fusion of chromosome ends by NHEJ as has been observed in mutation carriers and in TR-knockout mice [138].

4.1.2. Pulmonary fibrosis

Idiopathic pulmonary fibrosis (IPF) disease is characterized by progressive lung scarring and fibrotic changes. The disease is associated with abnormal telomere maintenance and is an attenuated form where fibrosis develops with cumulative age-related changes. This disease arises from mutations in genes encoding TERT and TR leading to reduced telomerase activity and subsequently shorter telomeres, resulting in impaired growth of lung stem cells [139]. Surprisingly, short telomeres have been detected in IPF patients with intact telomerase genes, indicating that IPF may develop in people who have short telomere lengths [140]. This study also showed the development of liver cirrhosis in 3% of sporadic IPF patients, demonstrating a complication of telomere-mediated disease outside the lung even in the absence of telomerase mutations. Also, increased incidence of insulin-dependent diabetes is detected in IPF patients [141]. Short telomeres have been shown to cause insulin secretion defects and glucose intolerance in telomerase-deficient mice [142].

4.1.3. Bone marrow failure

Many bone marrow failure disease cases have been linked to telomere biology. Mutations in telomeric proteins can lead to accelerated telomere attrition in hematopoietic compartment leading to bone marrow failure. The most common gene associated with bone marrow failure is TERT, which generally harbors point mutations in its gene [143, 144].

4.2. Role of TAPs in cancer

The role of TAPs in cancer development is well known. People with long telomeres are at a greater genetic risk of developing cancers [145]. Thus, examining the role of telomere proteins in cancer holds immense prognostic, diagnostic, and therapeutic value.

4.2.1. Shelterin proteins and cancer

The shelterin complex member POT1 was found to be somatically inactivated in chronic lymphocytic leukemia where it led to telomere deprotection and length extension [146]. Recently, two studies reported occurrence of rare, germ-line variants in POT1, making them susceptible to the development of familial melanoma [145]. In these cohorts, carrier individuals displayed significantly longer and more fragile telomeres than controls, and in some cases developed cancer in other tissues along with melanoma. Molecular and functional analysis showed that some of the variants abrogate the binding of POT1 to ssDNA, thus raising the possibility that carriers are predisposed to malignancy via telomere uncapping and a more permissive extension of chromosome ends. However, the exact biological mechanism needs further investigation. Mutation in RAP1, another shelterin protein member was reported in a melanoma cohort. RAP1 is involved in negative regulation of telomere length and functions by repressing homology-directed repair [147]. Mutations were reported to occur in TRF1-interacting region of RAP1. This loss of interaction with shelterin may increase the risk of cancer development.

Germ-line mutations affecting other proteins that interact with shelterin complex members and increase cancer risk have also been reported. For example, ku80, which interacts with RAP1 and PARP1, which in turn interacts with TRF2, has been found to be associated with diffuse large B-cell lymphomas.

4.2.2. Telomerase and cancer

Telomerase activity is essential for immortalization. Thus, targeting telomerase activity represents an attractive approach for both cancer diagnosis and treatment [148, 149]. As described previously, TERT is the limiting factor for telomerase activity. Therefore, its reactivation mechanisms hold great significance in understanding the development of cancer and thus designing targeted therapies.

Two hot-spot mutations in the TERT promoter, -228 C>T and -250 C>T, were recently reported to occur at high frequency in several solid tumors, for example: melanoma, gliomas, carcinoma of bladder, urothelial cancer, thyroid and squamous cell carcinoma of the tongue, as well as in liposarcomas and hepatocellular carcinomas, which have relatively low rates of self-renewal [85, 150153]. It was recently shown that TERT promoter mutations create novel binding sites for GABP, which belongs to Ets family of transcription factors [154]. These mutations have strong clinical implications with worse prognosis and poor survival, and thus may represent a novel therapeutic target [153].

TERT promoter mutation in skin cancers

Stem cells differentiate into normal somatic cells and as a consequence repress TERT transcription. Upon subsequent cell division, progressive telomere shortening occurs due to lack of telomerase activity. This acts as a barrier for tumor development and progression. Skin epidermal cells are highly differentiated cells, possess short telomeres, and are thus capable of undergoing limited proliferation [155]. However, in melanoma, increased telomerase activity is reported and this has been associated with high proliferation rate and early metastasis [156, 157].

High frequency of TERT promoter mutations has been reported in familial and sporadic melanoma (about 29–73%) [150, 151]. In primary cutaneous melanoma, TERT promoter mutations were found to be associated with BRAFV600E mutations, worse prognostic features, and shorter disease-free and overall survival [158, 159]. TERT promoter mutations have also been reported to be common in nonmelanoma skin cancer ranging from 39 to 74% in sporadic basal cell carcinoma and up to 50% mutation frequency in squamous cell carcinoma [158, 160, 161]. Various studies have assessed the association between telomere length and risk of developing skin cancer [162]. Some reports suggest no association between telomere length in peripheral blood leukocytes (PBL) and risk of nonmelanoma skin cancer [163]. On the contrary, other authors have reported that longer telomeres in PBL are protective for certain skin cancer types [162].

TERT promoter mutations in central nervous system (CNS) tumors

Within CNS tumors, gliomas have been shown to possess the highest frequency of TERT promoter mutations, while medulloblastoma and meningioma show lower frequencies [164]. Within gliomas, the percentage of cases with TERT promoter mutations varies depending on the histopathological type of tumor. TERT promoter mutations are reported in a large number of cases of glioblastoma multiforme (GBM), which is the most frequent and aggressive form of glioma, and in oligodendrogliomas, in contrast to astrocytoma and ependymoma, where only a small percentage of the tumors possess such mutations [159, 164]. Furthermore, the frequency of TERT promoter mutations in oligoastrocytomas, gliomas with a mixed origin, is in between that of oligodendrogliomas and astrocytomas [152]. These findings are consistent with the reported data on telomerase activity in gliomas, which is significantly higher in GBM (50–89%) and oligodendrogliomas (75–100%) than in astrocytomas (0–45%) [165167].

Some studies also reported an association between single-nucleotide polymorphisms (SNPs) in the TERT gene and an increased risk of glioma development [168, 169].

TERT promoter mutations in other cancers

Telomerase role in bladder carcinoma (BC) has been reported. In majority of BC tumor samples, telomerase activity was detected, while it was absent in the respective normal parallel samples [170, 171]. In some reports, telomerase activity was associated with lower grade and lower stage BC [170, 172]. Other studies showed that both telomerase activity and telomerase expression are associated with more advanced and higher grade of cancers [171, 173]. Preliminary evidence obtained in cell lines suggests that BC might have TERT promoter mutations [150]. TERT promoter mutations are also frequently detected in BC cell lines, with a prevalence ranging from 47 to 85%. These results have clearly shown that TERT promoter mutations represent one of the most common genetic events, perhaps the most frequent, in BC [85].

TERT promoter mutations also occur at high frequency in other cancer types, for example: hepatocellular carcinoma (56%), several soft tissue tumors histotypes (e.g., 93% in atypical fibroxanthoma, 79% in myxoid liposarcoma, and 76% in pleomorphic dermal sarcoma) and carcinoma of the renal pelvis (64%). Tumor histotypes with intermediate frequencies of TERT promoter mutations comprise laryngeal carcinoma (27%) and clear cell carcinoma of the ovary (16%) [174]. TERT promoter mutations are not frequently found in leukemias and colorectal cancers [174].

The high prevalence of TERT promoter mutations suggests the significance of telomere maintenance in cancers. Clinically, TERT promoter mutations represent a potential biomarker in cancer prognosis. Furthermore, TERT promoter mutations also serve as an attractive therapeutic target since they occur specifically in cancer cells and are absent in surrounding healthy tissues.

Advertisement

5. Conclusion

Telomeres are organized into highly specialized structures at chromosome ends. Telomerase and shelterin plays a role in telomere homeostasis. Along with telomere maintenance, telomere-associated proteins also play a significant role in various cell-signaling pathways. The significance and implication of telomerase and shelterin in human diseases have also been firmly established in various models of degenerative diseases. In cancer, telomerase dysfunction has been identified as a critical step for immortalization, although the underlying mechanisms are unclear. The recent identification of telomerase promoter mutations has stimulated research, following which numerous studies have reported similar alterations in various cancer models. In several relevant cancer types, telomerase promoter mutations seem to represent a new biomarker for prognosis with potential applications in presurgical diagnosis and in the monitoring of patients. Mechanisms regulating telomerase promoter mutations also hold immense therapeutic value since they occur specifically in cancers.

References

  1. 1. Moyzis RK, Buckingham JM, Cram LS, Dani M, Deaven LL, Jones MD, Meyne J, Ratliff RL, Wu JR: A highly conserved repetitive DNA sequence, (TTAGGG)n, present at the telomeres of human chromosomes. Proceedings of the National Academy of Sciences of the United States of America 1988, 85:6622–6626.
  2. 2. Griffith JD, Comeau L, Rosenfield S, Stansel RM, Bianchi A, Moss H, de Lange T: Mammalian telomeres end in a large duplex loop. Cell 1999, 97:503–514.
  3. 3. Greider CW: Telomeres do D-loop-T-loop. Cell 1999, 97:419–422.
  4. 4. d'Adda di Fagagna F, Reaper PM, Clay-Farrace L, Fiegler H, Carr P, Von Zglinicki T, Saretzki G, Carter NP, Jackson SP: A DNA damage checkpoint response in telomere-initiated senescence. Nature 2003, 426:194–198.
  5. 5. Diotti R, Loayza D: Shelterin complex and associated factors at human telomeres. Nucleus 2011, 2:119–135.
  6. 6. de Lange T: Shelterin: the protein complex that shapes and safeguards human telomeres. Genes & Development 2005, 19:2100–2110.
  7. 7. Palm W, de Lange T: How shelterin protects mammalian telomeres. Annual Review of Genetics 2008, 42:301–334.
  8. 8. Passos JF, von Zglinicki T: Mitochondria, telomeres and cell senescence. Experimental Gerontology 2005, 40:466–472.
  9. 9. Takai H, Smogorzewska A, de Lange T: DNA damage foci at dysfunctional telomeres. Current Biology 2003, 13:1549–1556.
  10. 10. Wai LK: Telomeres, telomerase, and tumorigenesis – a review. MedGenMed: Medscape General Medicine 2004, 6:19.
  11. 11. Richter T, von Zglinicki T: A continuous correlation between oxidative stress and telomere shortening in fibroblasts. Experimental Gerontology 2007, 42:1039–1042.
  12. 12. Cattan V, Mercier N, Gardner JP, Regnault V, Labat C, Maki-Jouppila J, Nzietchueng R, Benetos A, Kimura M, Aviv A, et al: Chronic oxidative stress induces a tissue-specific reduction in telomere length in CAST/Ei mice. Free Radical Biology & Medicine 2008, 44:1592–1598.
  13. 13. Greider CW, Blackburn EH: Identification of a specific telomere terminal transferase activity in tetrahymena extracts. Cell 1985, 43:405–413.
  14. 14. Sfeir A, de Lange T: Removal of shelterin reveals the telomere end-protection problem. Science 2012, 336:593–597.
  15. 15. Xin H, Liu D, Songyang Z: The telosome/shelterin complex and its functions. Genome Biology 2008, 9:232.
  16. 16. Martinez P, Blasco MA: Telomeric and extra-telomeric roles for telomerase and the telomere-binding proteins. Nature Reviews Cancer 2011, 11:161–176.
  17. 17. Galati A, Micheli E, Alicata C, Ingegnere T, Cicconi A, Pusch MC, Giraud-Panis MJ, Gilson E, Cacchione S: TRF1 and TRF2 binding to telomeres is modulated by nucleosomal organization. Nucleic Acids Research 2015, 43:5824–5837.
  18. 18. van Steensel B, de Lange T: Control of telomere length by the human telomeric protein TRF1. Nature 1997, 385:740–743.
  19. 19. Garrobo I, Marion RM, Dominguez O, Pisano DG, Blasco MA: Genome-wide analysis of in vivo TRF1 binding to chromatin restricts its location exclusively to telomeric repeats. Cell Cycle 2014, 13:3742–3749.
  20. 20. Ohishi T, Muramatsu Y, Yoshida H, Seimiya H: TRF1 ensures the centromeric function of Aurora-B and proper chromosome segregation. Molecular and Cellular Biology 2014, 34:2464–2478.
  21. 21. Poulet A, Pisano S, Faivre-Moskalenko C, Pei B, Tauran Y, Haftek-Terreau Z, Brunet F, Le Bihan YV, Ledu MH, Montel F, et al: The N-terminal domains of TRF1 and TRF2 regulate their ability to condense telomeric DNA. Nucleic Acids Research 2012, 40:2566–2576.
  22. 22. Doksani Y, Wu JY, de Lange T, Zhuang X: Super-resolution fluorescence imaging of telomeres reveals TRF2-dependent T-loop formation. Cell 2013, 155:345–356.
  23. 23. Simonet T, Zaragosi LE, Philippe C, Lebrigand K, Schouteden C, Augereau A, Bauwens S, Ye J, Santagostino M, Giulotto E, et al: The human TTAGGG repeat factors 1 and 2 bind to a subset of interstitial telomeric sequences and satellite repeats. Cell Research 2011, 21:1028–1038.
  24. 24. Yang D, Xiong Y, Kim H, He Q, Li Y, Chen R, Songyang Z: Human telomeric proteins occupy selective interstitial sites. Cell Research 2011, 21:1013–1027.
  25. 25. Wood AM, Rendtlew Danielsen JM, Lucas CA, Rice EL, Scalzo D, Shimi T, Goldman RD, Smith ED, Le Beau MM, Kosak ST: TRF2 and lamin A/C interact to facilitate the functional organization of chromosome ends. Nature Communications 2014, 5:5467.
  26. 26. Celli GB, de Lange T: DNA processing is not required for ATM-mediated telomere damage response after TRF2 deletion. Nature Cell Biology 2005, 7:712–718.
  27. 27. Saint-Leger A, Koelblen M, Civitelli L, Bah A, Djerbi N, Giraud-Panis MJ, Londono-Vallejo A, Ascenzioni F, Gilson E: The basic N-terminal domain of TRF2 limits recombination endonuclease action at human telomeres. Cell Cycle 2014, 13:2469–2474.
  28. 28. Kim SH, Kaminker P, Campisi J: TIN2, a new regulator of telomere length in human cells. Nature Genetics 1999, 23:405–412.
  29. 29. Ye JZ, Donigian JR, van Overbeek M, Loayza D, Luo Y, Krutchinsky AN, Chait BT, de Lange T: TIN2 binds TRF1 and TRF2 simultaneously and stabilizes the TRF2 complex on telomeres. The Journal of Biological Chemistry 2004, 279:47264–47271.
  30. 30. Frank AK, Tran DC, Qu RW, Stohr BA, Segal DJ, Xu L: The shelterin TIN2 subunit mediates recruitment of telomerase to telomeres. PLoS Genetics 2015, 11:e1005410.
  31. 31. Hockemeyer D, Palm W, Else T, Daniels JP, Takai KK, Ye JZ, Keegan CE, de Lange T, Hammer GD: Telomere protection by mammalian Pot1 requires interaction with Tpp1. Nature Structural & Molecular Biology 2007, 14:754–761.
  32. 32. Loayza D, Parsons H, Donigian J, Hoke K, de Lange T: DNA binding features of human POT1: a nonamer 5'-TAGGGTTAG-3' minimal binding site, sequence specificity, and internal binding to multimeric sites. The Journal of Biological Chemistry 2004, 279:13241–13248.
  33. 33. Liu D, Safari A, O'Connor MS, Chan DW, Laegeler A, Qin J, Songyang Z: PTOP interacts with POT1 and regulates its localization to telomeres. Nature Cell Biology 2004, 6:673–680.
  34. 34. Chen LY, Liu D, Songyang Z: Telomere maintenance through spatial control of telomeric proteins. Molecular and Cellular Biology 2007, 27:5898–5909.
  35. 35. Kibe T, Osawa GA, Keegan CE, de Lange T: Telomere protection by TPP1 is mediated by POT1a and POT1b. Molecular and Cellular Biology 2010, 30:1059–1066.
  36. 36. Xin H, Liu D, Wan M, Safari A, Kim H, Sun W, O'Connor MS, Songyang Z: TPP1 is a homologue of ciliate TEBP-beta and interacts with POT1 to recruit telomerase. Nature 2007, 445:559–562.
  37. 37. Zhong FL, Batista LF, Freund A, Pech MF, Venteicher AS, Artandi SE: TPP1 OB-fold domain controls telomere maintenance by recruiting telomerase to chromosome ends. Cell 2012, 150:481–494.
  38. 38. Nandakumar J, Bell CF, Weidenfeld I, Zaug AJ, Leinwand LA, Cech TR: The TEL patch of telomere protein TPP1 mediates telomerase recruitment and processivity. Nature 2012, 492:285–289.
  39. 39. Sexton AN, Regalado SG, Lai CS, Cost GJ, O'Neil CM, Urnov FD, Gregory PD, Jaenisch R, Collins K, Hockemeyer D: Genetic and molecular identification of three human TPP1 functions in telomerase action: recruitment, activation, and homeostasis set point regulation. Genes & Development 2014, 28:1885–1899.
  40. 40. Loayza D, De Lange T: POT1 as a terminal transducer of TRF1 telomere length control. Nature 2003, 423:1013–1018.
  41. 41. Wang F, Podell ER, Zaug AJ, Yang Y, Baciu P, Cech TR, Lei M: The POT1-TPP1 telomere complex is a telomerase processivity factor. Nature 2007, 445:506–510.
  42. 42. Latrick CM, Cech TR: POT1-TPP1 enhances telomerase processivity by slowing primer dissociation and aiding translocation. The EMBO Journal 2010, 29:924–933.
  43. 43. Li B, Oestreich S, de Lange T: Identification of human Rap1: implications for telomere evolution. Cell 2000, 101:471–483.
  44. 44. Takai KK, Hooper S, Blackwood S, Gandhi R, de Lange T: In vivo stoichiometry of shelterin components. The Journal of Biological Chemistry 2010, 285:1457–1467.
  45. 45. Kabir S, Hockemeyer D, de Lange T: TALEN gene knockouts reveal no requirement for the conserved human shelterin protein Rap1 in telomere protection and length regulation. Cell Reports 2014, 9:1273–1280.
  46. 46. Martinez P, Thanasoula M, Carlos AR, Gomez-Lopez G, Tejera AM, Schoeftner S, Dominguez O, Pisano DG, Tarsounas M, Blasco MA: Mammalian Rap1 controls telomere function and gene expression through binding to telomeric and extratelomeric sites. Nature Cell Biology 2010, 12:768–780.
  47. 47. Gilson E, Geli V: How telomeres are replicated. Nature Reviews Molecular Cell Biology 2007, 8:825–838.
  48. 48. Blasco MA: Telomeres and human disease: ageing, cancer and beyond. Nature Reviews Genetics 2005, 6:611–622.
  49. 49. Liu Y, Snow BE, Hande MP, Yeung D, Erdmann NJ, Wakeham A, Itie A, Siderovski DP, Lansdorp PM, Robinson MO, et al: The telomerase reverse transcriptase is limiting and necessary for telomerase function in vivo. Current Biology: CB 2000, 10:1459–1462.
  50. 50. Li Y, Tergaonkar V: Noncanonical functions of telomerase: implications in telomerase-targeted cancer therapies. Cancer Research 2014, 74:1639–1644.
  51. 51. Low KC, Tergaonkar V: Telomerase: central regulator of all of the hallmarks of cancer. Trends in Biochemical Sciences 2013, 38:426–434.
  52. 52. Theimer CA, Feigon J: Structure and function of telomerase RNA. Current Opinion in Structural Biology 2006, 16:307–318.
  53. 53. Mitchell JR, Collins K: Human telomerase activation requires two independent interactions between telomerase RNA and telomerase reverse transcriptase. Molecular Cell 2000, 6:361–371.
  54. 54. Autexier C, Pruzan R, Funk WD, Greider CW: Reconstitution of human telomerase activity and identification of a minimal functional region of the human telomerase RNA. The EMBO Journal 1996, 15:5928–5935.
  55. 55. Marrone A, Dokal I: Dyskeratosis congenita: molecular insights into telomerase function, ageing and cancer. Expert Reviews in Molecular Medicine 2004, 6:1–23.
  56. 56. Egan ED, Collins K: Biogenesis of telomerase ribonucleoproteins. RNA 2012, 18:1747–1759.
  57. 57. Kiss T, Fayet-Lebaron E, Jady BE: Box H/ACA small ribonucleoproteins. Molecular Cell 2010, 37:597–606.
  58. 58. Kittur N, Darzacq X, Roy S, Singer RH, Meier UT: Dynamic association and localization of human H/ACA RNP proteins. RNA 2006, 12:2057–2062.
  59. 59. Vulliamy TJ, Marrone A, Knight SW, Walne A, Mason PJ, Dokal I: Mutations in dyskeratosis congenita: their impact on telomere length and the diversity of clinical presentation. Blood 2006, 107:2680–2685.
  60. 60. Hamma T, Ferre-D'Amare AR: The box H/ACA ribonucleoprotein complex: interplay of RNA and protein structures in post-transcriptional RNA modification. The Journal of Biological Chemistry 2010, 285:805–809.
  61. 61. Fu D, Collins K: Distinct biogenesis pathways for human telomerase RNA and H/ACA small nucleolar RNAs. Molecular Cell 2003, 11:1361–1372.
  62. 62. Darzacq X, Kittur N, Roy S, Shav-Tal Y, Singer RH, Meier UT: Stepwise RNP assembly at the site of H/ACA RNA transcription in human cells. The Journal of Cell Biology 2006, 173:207–218.
  63. 63. Armanios M, Blackburn EH: The telomere syndromes. Nature Reviews Genetics 2012, 13:693–704.
  64. 64. Jady BE, Bertrand E, Kiss T: Human telomerase RNA and box H/ACA scaRNAs share a common Cajal body-specific localization signal. The Journal of Cell Biology 2004, 164:647–652.
  65. 65. Girard C, Verheggen C, Neel H, Cammas A, Vagner S, Soret J, Bertrand E, Bordonne R: Characterization of a short isoform of human Tgs1 hypermethylase associating with small nucleolar ribonucleoprotein core proteins and produced by limited proteolytic processing. The Journal of Biological Chemistry 2008, 283:2060–2069.
  66. 66. Mason M, Schuller A, Skordalakes E: Telomerase structure function. Current Opinion in Structural Biology 2011, 21:92–100.
  67. 67. Schmidt JC, Dalby AB, Cech TR: Identification of human TERT elements necessary for telomerase recruitment to telomeres. eLife 2014, 3.
  68. 68. Bley CJ, Qi X, Rand DP, Borges CR, Nelson RW, Chen JJ: RNA-protein binding interface in the telomerase ribonucleoprotein. Proceedings of the National Academy of Sciences of the United States of America 2011, 108:20333–20338.
  69. 69. Nakamura TM, Morin GB, Chapman KB, Weinrich SL, Andrews WH, Lingner J, Harley CB, Cech TR: Telomerase catalytic subunit homologs from fission yeast and human. Science 1997, 277:955–959.
  70. 70. Lingner J, Hughes TR, Shevchenko A, Mann M, Lundblad V, Cech TR: Reverse transcriptase motifs in the catalytic subunit of telomerase. Science 1997, 276:561–567.
  71. 71. Xiong Y, Eickbush TH: Origin and evolution of retroelements based upon their reverse transcriptase sequences. The EMBO Journal 1990, 9:3353–3362.
  72. 72. Steitz TA: DNA polymerases: structural diversity and common mechanisms. The Journal of Biological Chemistry 1999, 274:17395–17398.
  73. 73. Huard S, Moriarty TJ, Autexier C: The C terminus of the human telomerase reverse transcriptase is a determinant of enzyme processivity. Nucleic Acids Research 2003, 31:4059–4070.
  74. 74. Holt SE, Aisner DL, Baur J, Tesmer VM, Dy M, Ouellette M, Trager JB, Morin GB, Toft DO, Shay JW, et al: Functional requirement of p23 and Hsp90 in telomerase complexes. Genes & Development 1999, 13:817–826.
  75. 75. Venteicher AS, Meng Z, Mason PJ, Veenstra TD, Artandi SE: Identification of ATPases pontin and reptin as telomerase components essential for holoenzyme assembly. Cell 2008, 132:945–957.
  76. 76. King TH, Decatur WA, Bertrand E, Maxwell ES, Fournier MJ: A well-connected and conserved nucleoplasmic helicase is required for production of box C/D and H/ACA snoRNAs and localization of snoRNP proteins. Molecular and Cellular Biology 2001, 21:7731–7746.
  77. 77. Xi L, Cech TR: Inventory of telomerase components in human cells reveals multiple subpopulations of hTR and hTERT. Nucleic Acids Research 2014, 42:8565–8577.
  78. 78. Jady BE, Richard P, Bertrand E, Kiss T: Cell cycle-dependent recruitment of telomerase RNA and Cajal bodies to human telomeres. Molecular Biology of the Cell 2006, 17:944–954.
  79. 79. Tomlinson RL, Ziegler TD, Supakorndej T, Terns RM, Terns MP: Cell cycle-regulated trafficking of human telomerase to telomeres. Molecular Biology of the Cell 2006, 17:955–965.
  80. 80. Robart AR, Collins K: Human telomerase domain interactions capture DNA for TEN domain-dependent processive elongation. Molecular Cell 2011, 42:308–318.
  81. 81. Sauerwald A, Sandin S, Cristofari G, Scheres SH, Lingner J, Rhodes D: Structure of active dimeric human telomerase. Nature Structural & Molecular Biology 2013, 20:454–460.
  82. 82. Egan ED, Collins K: Specificity and stoichiometry of subunit interactions in the human telomerase holoenzyme assembled in vivo. Molecular and Cellular Biology 2010, 30:2775–2786.
  83. 83. Akincilar SC, Low KC, Liu CY, Yan TD, Oji A, Ikawa M, Li S, Tergaonkar V: Quantitative assessment of telomerase components in cancer cell lines. FEBS Letters 2015, 589:974–984.
  84. 84. Gladych M, Wojtyla A, Rubis B: Human telomerase expression regulation. Biochemistry and Cell Biology = Biochimie et biologie cellulaire 2011, 89:359–376.
  85. 85. Vinagre J, Pinto V, Celestino R, Reis M, Populo H, Boaventura P, Melo M, Catarino T, Lima J, Lopes JM, et al: Telomerase promoter mutations in cancer: an emerging molecular biomarker? Virchows Archiv: An International Journal of Pathology 2014, 465:119–133.
  86. 86. Faure G, Revy P, Schertzer M, Londono-Vallejo A, Callebaut I: The C-terminal extension of human RTEL1, mutated in Hoyeraal–Hreidarsson syndrome, contains harmonin-N-like domains. Proteins 2014, 82:897–903.
  87. 87. Sarek G, Vannier JB, Panier S, Petrini JH, Boulton SJ: TRF2 recruits RTEL1 to telomeres in S phase to promote t-loop unwinding. Molecular Cell 2015, 57:622–635.
  88. 88. Miyake Y, Nakamura M, Nabetani A, Shimamura S, Tamura M, Yonehara S, Saito M, Ishikawa F: RPA-like mammalian Ctc1-Stn1-Ten1 complex binds to single-stranded DNA and protects telomeres independently of the Pot1 pathway. Molecular Cell 2009, 36:193–206.
  89. 89. Zhao Y, Sfeir AJ, Zou Y, Buseman CM, Chow TT, Shay JW, Wright WE: Telomere extension occurs at most chromosome ends and is uncoupled from fill-in in human cancer cells. Cell 2009, 138:463–475.
  90. 90. Chen LY, Redon S, Lingner J: The human CST complex is a terminator of telomerase activity. Nature 2012, 488:540–544.
  91. 91. Schmidt JC, Cech TR: Human telomerase: biogenesis, trafficking, recruitment, and activation. Genes & Development 2015, 29:1095–1105.
  92. 92. Venteicher AS, Artandi SE: TCAB1: driving telomerase to Cajal bodies. Cell Cycle 2009, 8:1329–1331.
  93. 93. Stern JL, Zyner KG, Pickett HA, Cohen SB, Bryan TM: Telomerase recruitment requires both TCAB1 and Cajal bodies independently. Molecular and Cellular Biology 2012, 32:2384–2395.
  94. 94. Schoeftner S, Blasco MA: Developmentally regulated transcription of mammalian telomeres by DNA-dependent RNA polymerase II. Nature Cell Biology 2008, 10:228–236.
  95. 95. Redon S, Reichenbach P, Lingner J: The non-coding RNA TERRA is a natural ligand and direct inhibitor of human telomerase. Nucleic Acids Research 2010, 38:5797–5806.
  96. 96. Wu Y, Xiao S, Zhu XD: MRE11-RAD50-NBS1 and ATM function as co-mediators of TRF1 in telomere length control. Nature Structural & Molecular Biology 2007, 14:832–840.
  97. 97. McKerlie M, Zhu XD: Cyclin B-dependent kinase 1 regulates human TRF1 to modulate the resolution of sister telomeres. Nature Communications 2011, 2:371.
  98. 98. McKerlie M, Walker JR, Mitchell TR, Wilson FR, Zhu XD: Phosphorylated (pT371)TRF1 is recruited to sites of DNA damage to facilitate homologous recombination and checkpoint activation. Nucleic Acids Research 2013, 41:10268–10282.
  99. 99. Karlseder J, Broccoli D, Dai Y, Hardy S, de Lange T: p53- and ATM-dependent apoptosis induced by telomeres lacking TRF2. Science 1999, 283:1321–1325.
  100. 100. Denchi EL, de Lange T: Protection of telomeres through independent control of ATM and ATR by TRF2 and POT1. Nature 2007, 448:1068–1071.
  101. 101. Bradshaw PS, Stavropoulos DJ, Meyn MS: Human telomeric protein TRF2 associates with genomic double-strand breaks as an early response to DNA damage. Nature Genetics 2005, 37:193–197.
  102. 102. Fouche N, Cesare AJ, Willcox S, Ozgur S, Compton SA, Griffith JD: The basic domain of TRF2 directs binding to DNA junctions irrespective of the presence of TTAGGG repeats. The Journal of Biological Chemistry 2006, 281:37486–37495.
  103. 103. Singhapol C, Pal D, Czapiewski R, Porika M, Nelson G, Saretzki GC: Mitochondrial telomerase protects cancer cells from nuclear DNA damage and apoptosis. PloS One 2013, 8:e52989.
  104. 104. Masutomi K, Possemato R, Wong JM, Currier JL, Tothova Z, Manola JB, Ganesan S, Lansdorp PM, Collins K, Hahn WC: The telomerase reverse transcriptase regulates chromatin state and DNA damage responses. Proceedings of the National Academy of Sciences of the United States of America 2005, 102:8222–8227.
  105. 105. Goytisolo FA, Samper E, Martin-Caballero J, Finnon P, Herrera E, Flores JM, Bouffler SD, Blasco MA: Short telomeres result in organismal hypersensitivity to ionizing radiation in mammals. The Journal of Experimental Medicine 2000, 192:1625–1636.
  106. 106. Wong KK, Chang S, Weiler SR, Ganesan S, Chaudhuri J, Zhu C, Artandi SE, Rudolph KL, Gottlieb GJ, Chin L, et al: Telomere dysfunction impairs DNA repair and enhances sensitivity to ionizing radiation. Nature Genetics 2000, 26:85–88.
  107. 107. Hayden MS, Ghosh S: Shared principles in NF-kappaB signaling. Cell 2008, 132:344–362.
  108. 108. Ghosh AS, Tergaonkar V: Telomeres and inflammation: Rap1 joins the ends? Cell Cycle 2010, 9:3834–3835.
  109. 109. Tergaonkar V: Catalytically inactive telomerase in oncogenesis. Oncotarget 2015, 6:14725–14726.
  110. 110. Teo H, Ghosh S, Luesch H, Ghosh A, Wong ET, Malik N, Orth A, de Jesus P, Perry AS, Oliver JD, et al: Telomere-independent Rap1 is an IKK adaptor and regulates NF-kappaB-dependent gene expression. Nature Cell Biology 2010, 12:758–767.
  111. 111. Poon MW, Yan L, Jiang D, Qin P, Tse HF, Wong IY, Wong DS, Tergaonkar V, Lian Q: Inhibition of RAP1 enhances corneal recovery following alkali injury. Investigative Ophthalmology & Visual Science 2015, 56:711–721.
  112. 112. Akiyama M, Hideshima T, Hayashi T, Tai YT, Mitsiades CS, Mitsiades N, Chauhan D, Richardson P, Munshi NC, Anderson KC: Nuclear factor-kappaB p65 mediates tumor necrosis factor alpha-induced nuclear translocation of telomerase reverse transcriptase protein. Cancer Research 2003, 63:18–21.
  113. 113. Ghosh A, Saginc G, Leow SC, Khattar E, Shin EM, Yan TD, Wong M, Zhang Z, Li G, Sung WK, et al: Telomerase directly regulates NF-kappaB-dependent transcription. Nature Cell Biology 2012, 14:1270–1281.
  114. 114. Ding D, Xi P, Zhou J, Wang M, Cong YS: Human telomerase reverse transcriptase regulates MMP expression independently of telomerase activity via NF-kappaB-dependent transcription. FASEB Journal: official publication of the Federation of American Societies for Experimental Biology 2013, 27:4375–4383.
  115. 115. Clevers H, Nusse R: Wnt/beta-catenin signaling and disease. Cell 2012, 149:1192–1205.
  116. 116. Choi J, Southworth LK, Sarin KY, Venteicher AS, Ma W, Chang W, Cheung P, Jun S, Artandi MK, Shah N, et al: TERT promotes epithelial proliferation through transcriptional control of a Myc- and Wnt-related developmental program. PLoS Genetics 2008, 4:e10.
  117. 117. Park JI, Venteicher AS, Hong JY, Choi J, Jun S, Shkreli M, Chang W, Meng Z, Cheung P, Ji H, et al: Telomerase modulates Wnt signalling by association with target gene chromatin. Nature 2009, 460:66–72.
  118. 118. Okamoto N, Yasukawa M, Nguyen C, Kasim V, Maida Y, Possemato R, Shibata T, Ligon KL, Fukami K, Hahn WC, Masutomi K: Maintenance of tumor initiating cells of defined genetic composition by nucleostemin. Proceedings of the National Academy of Sciences of the United States of America 2011, 108:20388–20393.
  119. 119. Koh CM, Khattar E, Leow SC, Liu CY, Muller J, Ang WX, Li Y, Franzoso G, Li S, Guccione E, Tergaonkar V: Telomerase regulates MYC-driven oncogenesis independent of its reverse transcriptase activity. The Journal of Clinical Investigation 2015, 125:2109–2122.
  120. 120. Wu KJ, Grandori C, Amacker M, Simon-Vermot N, Polack A, Lingner J, Dalla-Favera R: Direct activation of TERT transcription by c-MYC. Nature Genetics 1999, 21:220–224.
  121. 121. Mondello C, Bottone MG, Noriki S, Soldani C, Pellicciari C, Scovassi AI: Oxidative stress response in telomerase-immortalized fibroblasts from a centenarian. Annals of the New York Academy of Sciences 2006, 1091:94–101.
  122. 122. Lee J, Sung YH, Cheong C, Choi YS, Jeon HK, Sun W, Hahn WC, Ishikawa F, Lee HW: TERT promotes cellular and organismal survival independently of telomerase activity. Oncogene 2008, 27:3754–3760.
  123. 123. Santos JH, Meyer JN, Skorvaga M, Annab LA, Van Houten B: Mitochondrial hTERT exacerbates free-radical-mediated mtDNA damage. Aging Cell 2004, 3:399–411.
  124. 124. Haendeler J, Drose S, Buchner N, Jakob S, Altschmied J, Goy C, Spyridopoulos I, Zeiher AM, Brandt U, Dimmeler S: Mitochondrial telomerase reverse transcriptase binds to and protects mitochondrial DNA and function from damage. Arteriosclerosis, Thrombosis, and Vascular Biology 2009, 29:929–935.
  125. 125. Massard C, Zermati Y, Pauleau AL, Larochette N, Metivier D, Sabatier L, Kroemer G, Soria JC: hTERT: a novel endogenous inhibitor of the mitochondrial cell death pathway. Oncogene 2006, 25:4505–4514.
  126. 126. Sharma NK, Reyes A, Green P, Caron MJ, Bonini MG, Gordon DM, Holt IJ, Santos JH: Human telomerase acts as a hTR-independent reverse transcriptase in mitochondria. Nucleic Acids Research 2012, 40:712–725.
  127. 127. Chen LY, Zhang Y, Zhang Q, Li H, Luo Z, Fang H, Kim SH, Qin L, Yotnda P, Xu J, et al: Mitochondrial localization of telomeric protein TIN2 links telomere regulation to metabolic control. Molecular Cell 2012, 47:839–850.
  128. 128. El Mai M, Wagner KD, Michiels JF, Ambrosetti D, Borderie A, Destree S, Renault V, Djerbi N, Giraud-Panis MJ, Gilson E, et al: The telomeric protein TRF2 regulates angiogenesis by binding and activating the PDGFRbeta promoter. Cell Reports 2014, 9:1047–1060.
  129. 129. Gonzalez OG, Assfalg R, Koch S, Schelling A, Meena JK, Kraus J, Lechel A, Katz SF, Benes V, Scharffetter-Kochanek K, et al: Telomerase stimulates ribosomal DNA transcription under hyperproliferative conditions. Nature Communications 2014, 5:4599.
  130. 130. Holohan B, Wright WE, Shay JW: Cell biology of disease: Telomeropathies: an emerging spectrum disorder. The Journal of Cell Biology 2014, 205:289–299.
  131. 131. Tummala H, Walne A, Collopy L, Cardoso S, de la Fuente J, Lawson S, Powell J, Cooper N, Foster A, Mohammed S, et al: Poly(A)-specific ribonuclease deficiency impacts telomere biology and causes dyskeratosis congenita. The Journal of Clinical Investigation 2015, 125:2151–2160.
  132. 132. Calado RT, Regal JA, Kleiner DE, Schrump DS, Peterson NR, Pons V, Chanock SJ, Lansdorp PM, Young NS: A spectrum of severe familial liver disorders associate with telomerase mutations. PloS One 2009, 4:e7926.
  133. 133. Le Guen T, Jullien L, Touzot F, Schertzer M, Gaillard L, Perderiset M, Carpentier W, Nitschke P, Picard C, Couillault G, et al: Human RTEL1 deficiency causes Hoyeraal-Hreidarsson syndrome with short telomeres and genome instability. Human Molecular Genetics 2013, 22:3239–3249.
  134. 134. Kocak H, Ballew BJ, Bisht K, Eggebeen R, Hicks BD, Suman S, O'Neil A, Giri N, Maillard I, Alter BP, et al: Hoyeraal-Hreidarsson syndrome caused by a germline mutation in the TEL patch of the telomere protein TPP1. Genes & Development 2014, 28:2090–2102.
  135. 135. Sasa GS, Ribes-Zamora A, Nelson ND, Bertuch AA: Three novel truncating TINF2 mutations causing severe dyskeratosis congenita in early childhood. Clinical Genetics 2012, 81:470–478.
  136. 136. Gu P, Chang S: Functional characterization of human CTC1 mutations reveals novel mechanisms responsible for the pathogenesis of the telomere disease Coats plus. Aging Cell 2013, 12:1100–1109.
  137. 137. Alter BP, Giri N, Savage SA, Peters JA, Loud JT, Leathwood L, Carr AG, Greene MH, Rosenberg PS: Malignancies and survival patterns in the National Cancer Institute inherited bone marrow failure syndromes cohort study. British Journal of Haematology 2010, 150:179–188.
  138. 138. Vulliamy T, Marrone A, Goldman F, Dearlove A, Bessler M, Mason PJ, Dokal I: The RNA component of telomerase is mutated in autosomal dominant dyskeratosis congenita. Nature 2001, 413:432–435.
  139. 139. Traila D, Mladinescu OF, Oancea C, Tudorache V: Short telomeres in pulmonary fibrosis: from genetics to clinical significance. Pneumologia 2015, 64:8, 11–13.
  140. 140. Alder JK, Chen JJ, Lancaster L, Danoff S, Su SC, Cogan JD, Vulto I, Xie M, Qi X, Tuder RM, et al: Short telomeres are a risk factor for idiopathic pulmonary fibrosis. Proceedings of the National Academy of Sciences of the United States of America 2008, 105:13051–13056.
  141. 141. Gribbin J, Hubbard R, Smith C: Role of diabetes mellitus and gastro-oesophageal reflux in the aetiology of idiopathic pulmonary fibrosis. Respiratory Medicine 2009, 103:927–931.
  142. 142. Guo N, Parry EM, Li LS, Kembou F, Lauder N, Hussain MA, Berggren PO, Armanios M: Short telomeres compromise beta-cell signaling and survival. PloS One 2011, 6:e17858.
  143. 143. Yamaguchi H, Calado RT, Ly H, Kajigaya S, Baerlocher GM, Chanock SJ, Lansdorp PM, Young NS: Mutations in TERT, the gene for telomerase reverse transcriptase, in aplastic anemia. The New England Journal of Medicine 2005, 352:1413–1424.
  144. 144. Calado RT, Regal JA, Hills M, Yewdell WT, Dalmazzo LF, Zago MA, Lansdorp PM, Hogge D, Chanock SJ, Estey EH, et al: Constitutional hypomorphic telomerase mutations in patients with acute myeloid leukemia. Proceedings of the National Academy of Sciences of the United States of America 2009, 106:1187–1192.
  145. 145. Robles-Espinoza CD, Velasco-Herrera Mdel C, Hayward NK, Adams DJ: Telomere-regulating genes and the telomere interactome in familial cancers. Molecular Cancer Research : MCR 2015, 13:211–222.
  146. 146. Ramsay AJ, Quesada V, Foronda M, Conde L, Martinez-Trillos A, Villamor N, Rodriguez D, Kwarciak A, Garabaya C, Gallardo M, et al: POT1 mutations cause telomere dysfunction in chronic lymphocytic leukemia. Nature Genetics 2013, 45:526–530.
  147. 147. Aoude LG, Pritchard AL, Robles-Espinoza CD, Wadt K, Harland M, Choi J, Gartside M, Quesada V, Johansson P, Palmer JM, et al: Nonsense mutations in the shelterin complex genes ACD and TERF2IP in familial melanoma. Journal of the National Cancer Institute 2015, 107.
  148. 148. Artandi SE, DePinho RA: Telomeres and telomerase in cancer. Carcinogenesis 2010, 31:9–18.
  149. 149. Donate LE, Blasco MA: Telomeres in cancer and ageing. Philosophical transactions of the Royal Society of London Series B, Biological Sciences 2011, 366:76–84.
  150. 150. Huang FW, Hodis E, Xu MJ, Kryukov GV, Chin L, Garraway LA: Highly recurrent TERT promoter mutations in human melanoma. Science 2013, 339:957–959.
  151. 151. Horn S, Figl A, Rachakonda PS, Fischer C, Sucker A, Gast A, Kadel S, Moll I, Nagore E, Hemminki K, et al: TERT promoter mutations in familial and sporadic melanoma. Science 2013, 339:959–961.
  152. 152. Killela PJ, Reitman ZJ, Jiao Y, Bettegowda C, Agrawal N, Diaz LA, Jr., Friedman AH, Friedman H, Gallia GL, Giovanella BC, et al: TERT promoter mutations occur frequently in gliomas and a subset of tumors derived from cells with low rates of self-renewal. Proceedings of the National Academy of Sciences of the United States of America 2013, 110:6021–6026.
  153. 153. Liu X, Wu G, Shan Y, Hartmann C, von Deimling A, Xing M: Highly prevalent TERT promoter mutations in bladder cancer and glioblastoma. Cell Cycle 2013, 12:1637–1638.
  154. 154. Bell RJ, Rube HT, Kreig A, Mancini A, Fouse SD, Nagarajan RP, Choi S, Hong C, He D, Pekmezci M, et al: Cancer. The transcription factor GABP selectively binds and activates the mutant TERT promoter in cancer. Science 2015, 348:1036–1039.
  155. 155. Buckingham EM, Klingelhutz AJ: The role of telomeres in the ageing of human skin. Experimental Dermatology 2011, 20:297–302.
  156. 156. Ramirez RD, D'Atri S, Pagani E, Faraggiana T, Lacal PM, Taylor RS, Shay JW: Progressive increase in telomerase activity from benign melanocytic conditions to malignant melanoma. Neoplasia 1999, 1:42–49.
  157. 157. Rudolph P, Schubert C, Tamm S, Heidorn K, Hauschild A, Michalska I, Majewski S, Krupp G, Jablonska S, Parwaresch R: Telomerase activity in melanocytic lesions: a potential marker of tumor biology. The American Journal of Pathology 2000, 156:1425–1432.
  158. 158. Populo H, Boaventura P, Vinagre J, Batista R, Mendes A, Caldas R, Pardal J, Azevedo F, Honavar M, Guimaraes I, et al: TERT promoter mutations in skin cancer: the effects of sun exposure and X-irradiation. The Journal of Investigative Dermatology 2014, 134:2251–2257.
  159. 159. Vinagre J, Almeida A, Populo H, Batista R, Lyra J, Pinto V, Coelho R, Celestino R, Prazeres H, Lima L, et al: Frequency of TERT promoter mutations in human cancers. Nature Communications 2013, 4:2185.
  160. 160. Griewank KG, Murali R, Schilling B, Schimming T, Moller I, Moll I, Schwamborn M, Sucker A, Zimmer L, Schadendorf D, et al: TERT promoter mutations are frequent in cutaneous basal cell carcinoma and squamous cell carcinoma. PloS One 2013, 8:e80354.
  161. 161. Scott GA, Laughlin TS, Rothberg PG: Mutations of the TERT promoter are common in basal cell carcinoma and squamous cell carcinoma. Modern Pathology: An Official Journal of the United States and Canadian Academy of Pathology, Inc 2014, 27:516–523.
  162. 162. Anic GM, Sondak VK, Messina JL, Fenske NA, Zager JS, Cherpelis BS, Lee JH, Fulp WJ, Epling-Burnette PK, Park JY, Rollison DE: Telomere length and risk of melanoma, squamous cell carcinoma, and basal cell carcinoma. Cancer Epidemiology 2013, 37:434–439.
  163. 163. Liang G, Qureshi AA, Guo Q, De Vivo I, Han J: No association between telomere length in peripheral blood leukocytes and the risk of nonmelanoma skin cancer. Cancer Epidemiology, Biomarkers & Prevention: a publication of the American Association for Cancer Research, cosponsored by the American Society of Preventive Oncology 2011, 20:1043–1045.
  164. 164. Koelsche C, Sahm F, Capper D, Reuss D, Sturm D, Jones DT, Kool M, Northcott PA, Wiestler B, Bohmer K, et al: Distribution of TERT promoter mutations in pediatric and adult tumors of the nervous system. Acta Neuropathologica 2013, 126:907–915.
  165. 165. Hiraga S, Ohnishi T, Izumoto S, Miyahara E, Kanemura Y, Matsumura H, Arita N: Telomerase activity and alterations in telomere length in human brain tumors. Cancer Research 1998, 58:2117–2125.
  166. 166. Langford LA, Piatyszek MA, Xu R, Schold SC, Jr., Shay JW: Telomerase activity in human brain tumours. Lancet 1995, 346:1267–1268.
  167. 167. Sano T, Asai A, Mishima K, Fujimaki T, Kirino T: Telomerase activity in 144 brain tumours. British Journal of Cancer 1998, 77:1633–1637.
  168. 168. Shete S, Hosking FJ, Robertson LB, Dobbins SE, Sanson M, Malmer B, Simon M, Marie Y, Boisselier B, Delattre JY, et al: Genome-wide association study identifies five susceptibility loci for glioma. Nature Genetics 2009, 41:899–904.
  169. 169. Zhao Y, Chen G, Song X, Chen H, Mao Y, Lu D: Fine-mapping of a region of chromosome 5p15.33 (TERT-CLPTM1L) suggests a novel locus in TERT and a CLPTM1L haplotype are associated with glioma susceptibility in a Chinese population. International Journal of Cancer: Journal international du cancer 2012, 131:1569–1576.
  170. 170. Okumura A, Mizuno I, Nagakawa O, Fuse H: Telomerase activity is correlated with lower grade and lower stage bladder carcinomas. International journal of urology : official journal of the Japanese Urological Association 2004, 11:1082-1086.
  171. 171. Lin Y, Miyamoto H, Fujinami K, Uemura H, Hosaka M, Iwasaki Y, Kubota Y: Telomerase activity in human bladder cancer. Clinical Cancer Research: An Official Journal of the American Association for Cancer Research 1996, 2:929–932.
  172. 172. Morii A, Komiya A, Okumura A, Fuse H: Telomerase activity in bladder cancer tissue. Experimental and Therapeutic Medicine 2010, 1:85–88.
  173. 173. Yang CH, Hung WC, Wang SL, Kang WY, Chen WT, Huang YC, Su YC, Chai CY: Immunoexpression and prognostic role of hTERT and cyclin D1 in urothelial carcinoma. APMIS: acta pathologica, microbiologica, et immunologica Scandinavica 2008, 116:309–316.
  174. 174. Heidenreich B, Rachakonda PS, Hemminki K, Kumar R: TERT promoter mutations in cancer development. Current Opinion in Genetics & Development 2014, 24:30–37.

Written By

Ekta Khattar and Vinay Tergaonkar

Submitted: 11 May 2015 Reviewed: 08 December 2015 Published: 23 November 2016