Next Article in Journal
Obesity-Related Metabolic Dysfunction in Dairy Cows and Horses: Comparison to Human Metabolic Syndrome
Next Article in Special Issue
Pediatric Neuromyelitis Optica Spectrum Disorder: Case Series and Literature Review
Previous Article in Journal
Diverse Effect of Two Cytokinins, Kinetin and Benzyladenine, on Plant Development, Biotic Stress Tolerance, and Gene Expression
Previous Article in Special Issue
Approach to Neurological Channelopathies and Neurometabolic Disorders in Newborns
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Telomeres and Cancer

1
Department of Pediatrics, Tungs’ Taichung Metroharbor Hospital, Wuchi, Taichung 435, Taiwan
2
Department of Medical Research, Tungs’ Taichung Metroharbor Hospital, Wuchi, Taichung 435, Taiwan
3
Department of Rehabilitation, Jen-Teh Junior College of Medicine, Nursing and Management, Miaoli 356, Taiwan
4
Department of Life Sciences, Agricultural Biotechnology Center, National Chung Hsing University, Taichung 402, Taiwan
5
Department of Obstetrics & Gynecology, Tri-Service General Hospital, National Defense Medical Center, Taipei 11490, Taiwan
6
School of Medicine, Chung Shan Medical University, Taichung 402, Taiwan
7
Department of Pediatrics, Chung Shan Medical University Hospital, Taichung 402, Taiwan
8
Department of Functional Neurosurgery, Xiamen Humanity Hospital, Fujian Medical University, Xiamen 361015, China
9
The iEGG and Animal Biotechnology Center, and Rong Hsing Research Center for Translational Medicine, National Chung Hsing University, Taichung 402, Taiwan
10
Buddhist Tzu Chi Bioinnovation Center, Tzu Chi Foundation, Hualien 970, Taiwan
11
Bioinnovation Center, Buddhist Tzu Chi Medical Foundation, Hualien 970, Taiwan
12
Department of Neurosurgery, Buddhist Tzu Chi General Hospital, Hualien 970, Taiwan
13
Department of Medical Research, Hualien Tzu Chi Hospital, Hualien 970, Taiwan
14
Department of Pathology, Buddhist Tzu Chi General Hospital and Tzu Chi University, Hualien 970, Taiwan
*
Author to whom correspondence should be addressed.
Life 2021, 11(12), 1405; https://doi.org/10.3390/life11121405
Submission received: 4 November 2021 / Revised: 1 December 2021 / Accepted: 6 December 2021 / Published: 16 December 2021
(This article belongs to the Special Issue Research Updates in Pediatric Neuroscience)

Abstract

:
Telomeres cap the ends of eukaryotic chromosomes and are indispensable chromatin structures for genome protection and replication. Telomere length maintenance has been attributed to several functional modulators, including telomerase, the shelterin complex, and the CST complex, synergizing with DNA replication, repair, and the RNA metabolism pathway components. As dysfunctional telomere maintenance and telomerase activation are associated with several human diseases, including cancer, the molecular mechanisms behind telomere length regulation and protection need particular emphasis. Cancer cells exhibit telomerase activation, enabling replicative immortality. Telomerase reverse transcriptase (TERT) activation is involved in cancer development through diverse activities other than mediating telomere elongation. This review describes the telomere functions, the role of functional modulators, the implications in cancer development, and the future therapeutic opportunities.

1. Introduction

Cancer is notorious as it can attack any part of the body, rapidly grow beyond its usual boundaries, invade adjoining tissues, and spread to other organs, resulting in uncontrolled proliferation and eventually death. Nearly 10 million deaths were reported in 2020 from cancer, and the risk of getting cancer in a lifetime (before the age of 75 years) is 20% [1]. The most common newly-diagnosed cancers reported worldwide include breast (2.26 million), lung (2.21 million), and colorectal (1.98 million) cancers [1]. Approximately half of the newly diagnosed lung (57%) and pancreatic (52%) cancer cases in the United States are at an advanced or metastatic stage, and the majority of these patients with an early diagnosis of the disease eventually develop tumor progression [2]. The 5-year relative survival rates for advanced-stage cancers, such as lung, colorectal, liver, and pancreatic, remain low, ranging from 3% to 14% even after maximal surgical excision, radiation, chemotherapy, and hormone, immune, and targeted therapies [2]. Thus, none of the standard cancer treatments can completely cure patients at an advanced stage of the disease. Knowledge of the molecular mechanisms influencing tumor growth and invasiveness may lead to novel and effective therapies for the poor prognosis of late-stage cancers.
Cancer formation and progression is a genetic phenomenon with normal cells accruing genomic instability and thereby acquiring the ability to replicate indefinitely, which is the phenotype of immortality [3]. Telomerase, the immortality enzyme, is ubiquitous in all mammalian embryonic tissue and remains active in germs cells but is down-regulated in most somatic tissues [4]. As telomerase activity determines cellular proliferation, it must be tightly regulated to prevent the induction of carcinogenesis [5]. Telomerase reverse transcriptase (TERT), the catalytic subunit responsible for enzyme activity in telomerase, is the rate-limiting factor of human telomerase enzyme activity [5]. Two of the critical telomere-specific proteins involved in the regulation and maintenance of the telomere length are the shelterin and CST complexes [6]. Mutations in the genes encoding these complexes can result in cancers. Thus, understanding the molecular mechanisms of these proteins is fundamental given the therapeutic strategies to manage such diseases.

2. Telomeres, a Genetic Time Bomb or a Biological Clock

Human telomeres comprise a hexameric nucleotide repeat sequence (TTAGGG) that is initially double-stranded DNA (dsDNA) but ends with a single-stranded DNA (ssDNA) overhang (G’-overhang). The extended 5′ to 3′ strand contains the G-rich telomeric repeats and is referred to as the G-strand, while the 3′ to 5′ strand is defined as the C-strand [7,8]. During the cell division cycle, the eukaryotic DNA polymerase is unable to completely replicate the sequences at the chromosomal ends. This is because RNA primers attach at the lagging strand during the synthesis of Okazaki fragments, and the resulting shedding RNA leads to telomere shortening [9]. The so-called “end replication problem” results in eventual apoptosis, cellular senescence, and cell cycle arrest [10,11]. Additionally, chromosomes lacking the “capping structure” tend to get truncated and fused with other chromosomes [12]. As such, telomeres are also considered a genetic time bomb or a biological clock for cellular aging [13].
As approximately 50–200 bases are lost from the terminal sequence of chromosomes each time a cell divides [14], and more than a couple of trillion telomere sequences are in the human genome, the spatiotemporal expression of telomerase must be tightly regulated in humans. Apart from the shedding RNA and the generation of the 3′ overhang by the sequence-specific exonuclease activity to resect back the 5′end of telomeres [15], telomere shortenings can occur, irrespective of cell replication, due to accumulative oxidative stress [16], host age [17], gender [18], sex hormones [19], and lifestyle factors, such as the lack or presence of exercise [20], obesity and weight loss [21], smoking [22], and unhealthy diets [23]. However, short telomeres not only result in genomic loss [24,25], shorter lifespan [26,27] and contribute to diseases such as coronary heart disease [28], heart failure [29], osteoporosis [30], diabetes [31], but can also result in genomic instability and elevated telomerase activity, leading to a potential cancer predisposition factor [32]. Hence, proper telomere maintenance is critical for human life.

3. The Shelterin Complex

Telomeres are protected by a highly conserved mammalian nucleoprotein complex called shelterin (telosome). This nucleoprotein complex can minimize telomere fragility by enabling DNA replication at the telomeric repeats [33,34]. The shelterin complex can allow DNA to form a lasso-like structure with a telomeric loop (T-loop) and a displacement loop (D-loop) that then shields the 3′-end from DNA damage and blocks the activation of the DNA repair mechanisms, such as ataxia-telangiectasia Rad3-related (ATR)-mediated DNA damage kinase signaling and ataxia-telangiectasia mutation (ATM) kinase cascades, as well as unwanted repair reactions [35]. The shelterin complex anchors to both ssDNA and dsDNA [33,36,37,38]. The shelterin complex and other protein complexes specific to the telomere can detect and react to changes in telomere length in order to maintain the proper length of telomeres [39,40] (Figure 1). However, in cancers, mutations in the shelterin complex that cause telomere dysfunction and dysregulation are very common [41].
The shelterin complex is composed of six subunits, namely the telomere repeat-binding factor 1(TRF1), telomere repeat-binding factor 2 (TRF2), repressor activator protein 1(RAP1), TRF1-interacting nuclear factor 2(TIN2), TINT1/PTOP/PIP1(TPP1), and the protection of telomeres-1(POT1) [42,43].

3.1. TRF1

TRF1 is responsible for controlling telomeric DNA extension (Figure 1) [44]. While the over-expression of TRF1 decreases telomere length, dominant-negative TRF1 increases its length, thus suggesting that TRF1 negatively regulates the length of telomeric repeats [45]. In contrast, Karlseder et al. [46] disputed the role of TRF1 in regulating telomere length because they found that complete TRF1 deletion in mice showed no defects in the telomere length or telomere capping.TRF1 is overexpressed in several cancer types, such as renal cell carcinoma [47] and gastrointestinal tumors [48], and is upregulated in glioblastoma multiforme (GBM) [49]. Brain-specific TRF1 genetic deletion in GBM mouse models inhibited tumor initiation and progression and increased survival. In addition, deleted TRF1 increased telomeric DNA damage and reduced proliferation and stemness. TRF1 chemical inhibitors mimicked these effects in human GBM cells and also blocked tumor sphere formation [49], suggesting that TRF1 may be a promising target for developing an effective anti-cancer therapeutic strategy.

3.2. TRF2

Although TRF2 and TRF1 have similar features, they exhibit distinct functions. The binding affinity of TRF1 to TIN2 is 20-fold greater than that ofTRF2 [50]. TRF2 and TRF1 are both homodimers that can attach to telomeric dsDNA (Figure 1) [51,52,53]. TRF2 acts as a protein hub interacting with many DNA repair proteins. These proteins do contribute to human chromosomal instability syndromes typified by increased cancer and premature aging [33,54]. As a hub protein, TRF2 interacts with multiple proteins, including a specific endonuclease, the Mre11/Nbs1/Rad50 (MNR) complex, poly(ADP-ribose) polymerase 1 (PARP1) and poly(ADP-ribose) polymerase 2 (PARP2), the DNA-protein kinase complex, RAP1, the Bloom syndrome protein (BLM), and the Werner syndrome protein(WRN) helicases [55,56,57,58]. Bloom syndrome (BS) is an autosomal recessive disorder characterized by short stature, a skin rash that develops after exposure to the sun, intellectual disability, microcephaly, and increased chromosome breakage and an increased risk for cancer [59]. Hand and German [60], using diploid fibroblast cell lines derived from skin biopsies taken from five BS patients, observed a slower rate of replication fork movement compared to the normal adult controls. It is now clear that this aberration results from defective RecQ helicase function during DNA replication [61]. Loss-of-function mutations of BLM, which codes for a RecQ helicase, cause BS [62]. Patients with Werner syndrome (WS) manifest growth retardation, short stature, premature graying of hair, alopecia, wrinkling, prematurely aged faces with beaked noses, premature brain atrophy, lipodystrophy, gonad atrophy, bilateral cataracts, premature arteriosclerosis, calcinosis, type 2 diabetes, osteoporosis, telangiectasia, and malignancies [63,64]. WS is caused by mutations in the WRN gene [65]. WRN, which is coded by the WRN gene, is a DNA helicase which maintains genomic stability by participating in double-strand break (DSB) repair and inter-strand crosslink repair, as well as other DNA processing events [66].
TRF2 is also associated with the integrity of the G-strand overhang and the protection of telomeres, as it has been demonstrated that the over-expression of dominant-negative TRF2 in cells results in the loss of both the TRF2 bound at the telomeres and the G-strand overhang, the activation of the p53 damage pathway, and chromosome end fusions [53,67,68]. In addition, TRF2 is essential for T-loop assembly and maintenance of the ATM-mediated DNA damage response (DDR) suppression and repression from non-homologous end joining (NHEJ) [69]. As with the TRF1-knockout mice, TRF2-knockout mice are embryonically lethal [70]. Whereas TRF2 over-expressing mice under the 5′-regulatory region of the keratin 5 (K5) gene have increased vulnerability to spontaneous skin tumors and are sensitive to UV-induced carcinogenesis [71,72,73]. TRF2 production is often raised in human skin carcinomas [71]. Patients with xeroderma pigmentosum who have XPF, a nuclease associated with UV-damage repair, specific mutations are susceptible to TRF2-associated telomere shortening and chromosomal instability [71]. Furthermore, a lack of telomerase significantly accelerates TRF2-induced epithelial carcinogenesis, suggesting a greater chromosomal instability and an increased burden of DNA damage [74], thereby alluding to a role of TRF2 in directing telomere recombination. Similar to TRF1, some studies have reported that TRF2 is up-regulated in some human cancers, such as skin cancers, and increased TRF2 expression can promote skin tumorigenesis [54,74,75]. Bejarano et al. [76] identified a direct link between TRF1 phosphorylation by common cancer signaling pathways, telomere protection, and cancer treatment. Interestingly, the link is not only restricted to TRF1, but also to TRF2 because TRF2 is reported to be phosphorylated by the ERK1/2 kinases and interacts with Ras signaling to bypass DDR in cancer cells [77,78]. Taken together, these findings indicate that targeting common cancer signaling pathways through down-regulating the shelterin complex may be a potential anti-cancer therapeutic strategy.

3.3. RAP1

In humans, RAP1 associates with TRF2 and is enlisted to telomeres to control telomere length (Figure 1) [79]. RAP1 improves the selective binding of TRF2 to telomeric DNA [80]. Mammalian RAP1 can shield telomeres from NHEJ activities in vitro and the context of severe telomere uncapping induced by TRF2 dysfunction [81]. Loss of RAP1 function in human cells does not result in NHEJ, homology-directed repair (HDR), or a DNA damage response [82]. At the same time, mouse RAP1 has been shown to shield telomere ends by repressing HDR and preventing sister telomere recombinations [82]. RAP1 binds to extra-telomeric DNA and acts as a transcriptional regulator [82,83]. RAP1 has also been reported to associate with I kappa B (IκB) kinases and activate nuclear factor kappa B (NF-κB) [84]. The expression of RAP1 has been shown to be significantly higher in breast tumor tissues than in the adjacent non-tumor tissues [84], suggesting that RAP1 could be involved in cancer progression. Moreover, RAP1 is reported to be highly expressed in colorectal cancer tissues, and the expression levels of RAP1 are significantly correlated with poor prognosis and metastasis [85].

3.4. TIN2

TIN2 directly connects TRF1, TRF2, and TPP1 without binding POT1 or RAP1 (Figure 1) [37,86]. TIN2 regulates the telomere length [38], maintains T-loop [87], and complexes with Rap1 and TPP1 to allow cells to distinguish telomeres from sites of DNA damage [33]. The over-expression of TIN2 has been shown to prevent telomere elongation in several human cell lines, such asHT1080 (human fibrosarcoma cell line), WI-38 (fibroblast-like fetal lung cell line), U2OS (bone osteosarcoma epithelial cell line), HTB9 (bladder carcinoma cell line), C33A (cervical carcinoma cell line), MDA-452 (breast cancer cell line), and HMT-3522 (non-tumorigenic human breast epithelial cell line). In contrast, the expression of a dominant-negative TIN2 results in uncontrolled telomere elongation [88]. Chiang et al. [89] have observed that early embryonic lethality occurs in mice with TIN2 deletion, similar to the findings of TRF1- and TRF2-deficient mice [46,70]. Expression of TRF1, TRF2, and TIN2 have been detected in human gastric carcinoma [48], and down-regulation of TRF1, TRF2, and TIN2 gene expression may be vital to maintain telomeres in gastric cancers [90]. Expressions of TRF1, TRF2, and TIN2 have also been found in patients with adult T-cell leukemia [91]. Increased expression of TRF1, TRF2, and TIN2 is linked to telomere shortening during multi-step hepatocarcinogenesis [92].

3.5. TPP1

TPP1 is a protein associated with POT1 and TPP1-POT1 heterodimer binds and caps the telomeric 3′ tail to protect telomeres (Figure 2A) [93,94]. TPP1-POT1 also binds to internal sites of telomeric ssDNA [95]. TPP1 deletion results in a strong ATR-dependent DDR at telomeres, causing excess telomeric ssDNA [95,96,97]. In addition to telomere protection and maintenance of telomere integrity, TPP1 can recruit telomerase, stabilize its association with the overhang, and stimulate enzyme activity to lengthen the telomeres (Figure 2A) [98]. Furthermore, TPP1 has been reported to promote telomerase processivity in the presence of POT1 [99]. TPP1 loss causes decreased expressionof POT1 at the telomeres and reduced telomerase processing [97,100]. TPP1 interacts with the CST complex and regulates telomere elongation by telomerase [99]. Tejera et al. have found that TPP1 deletion can decrease TERT binding to the telomeres [101]. Interestingly, they have also showed that abrogation of TPP1 abolishe snet telomere elongation in the context of the nuclear reprogramming of TRF1-deficient mouse embryonic fibroblasts into induced pluripotent stem cells (iPSCs), suggesting that TPP1 has a dual role in telomere protection as well as elongation. Another report shows that the complete TPP1 abrogation results in embryonic lethality [102]. Together, TPP1 deficiency can cause telomere dysfunction phenotypes, including widespread epithelial dysplasia, defective hair follicle morphogenesis, growth, severe skin hyperpigmentation, and peri-natal death [103,104,105].

3.6. POT1

POT1, like TRF2, plays a major role in end-capping.POT1negatively regulates telomerase and helps in telomere maintenance. POT1 interacts with TPP1 and attaches to the ssDNA 3′ overhang (Figure 1), thereby repressing ATR-mediated DDR by stopping the recruitment of replication protein A (RPA) to the ssDNA [106]. A decrease in POT1 can result in an irreparable ATR pathway with significantly elongated telomeres and cell cycle arrest and is embryonically lethal [93,107,108,109]. Two POT1 orthologs, POT1a and POT1b, are present in the mouse. POT1a is necessary to fully repress a DNA damage signal at telomeres [108]. POT1b has a specific role in regulating the structure of the telomere terminus, leading to deregulation of the telomeric overhang in POT1b-deficient cells, despite the presence of POT1a [108]. POT1b regulates the amount of ssDNA at the telomeres in a telomerase-independent manner [108]. Double knockout cells for the POT1a and POT1b genes have been demonstrated to result in telomere elongation, an increase in DNA damage foci at the telomeres, endo-reduplication, and early initiation of senescence [108,109]. POT1 mRNA concentrations have been significantly linked with telomere length in colon and gastric cancer cells [108]. A known variant of POT1, with D224N mutation, disrupts POT1 binding to ssDNA telomere oligonucleotides, leading to longer and fragile telomeres, predisposing for chronic leukocyte leukemia, glioma, angiosarcoma, osteosarcoma, thyroid cancer, colorectal cancer, and cutaneous melanoma [110]. Convergently, the shelterin complex interacts with more than 300 proteins, including RING-finger- or U-box-containing proteins, functioning as ubiquitin E3 ligases or stability regulators for telomere-associated proteins, protein phosphatase catalytic and regulatory subunits (PPM1G, PHPT1, PTPN5, SAPS3, and PPP1R2), and phosphorylation-related kinases (Akt1, CAMK1D, CLK3, MAP2K3, MAP4K2, MAPK12, and PAK4) [111], indicating that the complex not only stabilizes the chromosomal ends and protects hosts from diseases, but also acts as a busy hub for the complex signaling net workflow.

4. The CST Complex

The CST complex comprises telomere-specific proteins that regulate telomere length replication and maintenance. The CST complex was initially identified in Saccharomyces cerevisiae and later in vertebrates [101,112,113].

4.1. Yeast CST Complex

Saccharomyces CST complexisa trimeric nucleoprotein complex composed of cell division control protein 13(CDC13), suppressor of CDC thirteen 1 (STN1), and telomeric pathway with STN1 (TEN1) [114]. During cell budding, the CST complex is known as the CDC13-STN1-TEN1 complex; however, fission yeast only contains STN1 and TEN1 [115]. Deletions affecting CDC13, STN1,or TEN1 make budding yeast cells unviable [114]. Therefore, the yeast CST complex is crucially important and may possess evolutionarily conserved functions in DNA replication [116]. Yeast CST complex is structurally related to the heterotrimeric replication protein A (RPA)-complex [112], which is a heterotrimeric ssDNA-binding protein complex composed of replication factor A1 (Rfa1), Rfa2, and Rfa3 (Figure 2A) [117].

4.2. Human CST Complex

As with yeast, the human CST complex comprises the conserved telomere maintenance component 1 (CTC1), STN1, and TEN1 (Figure 2B), and each subunit is present in the stoichiometric ratio of 1:1:1 [118,119,120]. It localizes at the chromosomal ends, preferentially to G-rich and repetitive elements [121], and can maintain telomere length [113,122]. Human CST is an RPA-like ssDNA-binding protein that has primarily been characterized as a telomere replication factor [123]. RPA is crucial for replication, repair, and recombination and is involved in multiple protein–protein interactions [124], telomere metabolism [112], and chromosome maintenance [125]. The human RPA complex comprises RPA70, RPA32, and RPA14 [117,126]. Structural analyses have demonstrated that each RPA subunit contains multiple oligonucleotide/oligosaccharide-binding domain (OB) (four in Rfa1/RPA70, one in Rfa2/RPA32, and one in Rfa3/RPA14) [126]. In yeast, the OB folds are DNA binding domains (DBD) and protein-protein interaction domains [127]. Rfa1,the largest subunit, has four OB domains: DBD-A, DBD-B, DBD-C, and DBD-F, connected by flexible linkers. Rfa2 has one OB fold (DBD-D) followed by a winged-helix (WH) domain, which participates in protein-protein interactions [126]. Rfa3, the smallest subunit, possesses one OB fold (DBD-E). Among these OB-fold domains, DBD-A, DBD-B, DBD-C, and DBD-D are the key players in RPA’s ssDNA-binding activities, while DBD-F and DBD-E have weak interactions with DNA [112,128,129]. In humans, the three RPA subunits, includingRPA70, containing four OB-fold domains (OB-A, OB-B, OB-C, and OB-F), RPA32, containing an OB-fold domain (OB-D) and a WH domain, and RPA14, the smallest subunit, containing an OB-fold domain, OB-E, form a trimerization structure. Among these OB-fold domains, the high binding affinity of RPA to ssDNA is mostly facilitated by OB-A, OB-B, OB-C, and OB-D in RPA70 and RPA32, while the OB-F and WH domains interact with its protein-binding partners [126,128,130] (Figure 3B). The OB-fold domains are connected with flexible linkers [131]. Each OB-fold domain is a five-stranded β barrel structural motif, existing in different proteins for nucleic acid recognition [132]. In addition to the recognition of ssDNA, the functions of the OB fold include the location of the binding surface, the polarity of the nucleic acid with respect to the OB fold, recognition of unusually structured nucleic acids, and a sensor of DNA damage [132]. Functionally, CST has several features:1. Binding to ssDNA with G-rich sequences [119,133,134,135,136,137]. 2. Binding to ssDNA-dsDNA junctions [138]. 3. Recognizing different specialized DNA structures at the DNA replication and breakage sites [139]. 4. Acting synergistically with ATR to maintain telomere length and genome stability [140]. 5. Stimulating the primase activity of DNA polymerase alpha (Polα) switch from RNA to DNA synthesis [141,142,143]. 6. Helping in C-strand fill-in (Figure 3B) [136,144].7. Preventing the accumulation of G4: CST can directly engage and melt G4 DNA structures and prevent the accumulation of G4 structures during unperturbed DNA replication [139]. 8. Preventing telomeric DNA damage: CST can speed up telomeric DNA replication by facilitating the restart of the stalled replication forks to prevent telomeric DNA damage [125,134]. 9. Interacting with the mini chromosome maintenance of the 2–7(MCM) complex and disrupting the binding of CDT1 to MCM, leading to decreased origin licensing [145]. 10. Interacting with acidic and nucleoplasmic DNA-binding protein (AND-1), which is an important regulator governing the assembly of the MCM complex at the replication origins during the origin licensing in human cells [146]. 11. Inhibiting telomerase (Figure 2B): The CST complex interacts with shelterin or telomeric ssDNA to terminate telomerase for telomere length homeostasis, and this reaction does not need TEN1 [138,143,144,147,148]. In addition, CST interacts with and sequesters telomeric 3′ overhangs after their elongation by telomerase, thus preventing telomere extension by telomerase [120]. Therefore, impaired CST’s functions may increase the possibilities of defective telomere replication, deregulation of telomere length, and replication fork stalling, leading to irreparable DNA damage, genome instability, and diseases [125,136] (Table 1).
Similar to CST, RPA has several functional features:1. Binding ssDNA: RPA binds and protects ssDNA from cleavage by nucleases and recruits repair proteins to initiate DNA damage responses [149]. 2. Activating the ATR signaling: Replication stress (RS) is a condition when the replication fork progression and/or DNA synthesis is stalled or slowed [150]. The RPA-coated ssDNA serves as a main activation platform for recruiting ATR-ATR-interacting protein (ATRIP) to the stalled forkin RS [151]. Activated ATR-ATRIP phosphorylates and activates Checkpoint kinase 1(CHK1), which induces cell cycle arrest to allow DNA repair, fork stabilization, or replication start [156]. 3: Activating the helicase: RPA binding stimulates the accumulation of the human DNA helicase B on chromatin in replication stress [157].4. Unwinding G4: RPA binding promotes WRN activity and multiple RPA binding makes WRN a super-helicase on G4 unwinding [152]. 5. Involvement in DNA replication, recombination, and repair: BLM forms a complex with topoisomerase IIIα, RPA, and several factors involved in functions related to DNA replication, recombination, and repair [62]. 6. Activating BLM’s bidirectional DNA unwinding [153]. 7. Modulating the fork remodeling enzyme activity: SMARCAL-1(SWI/SNF-related, matrix-associated, actin-dependent regulator of chromatin, subfamily A-like 1) is a fork-remodeling enzyme. RPA binds to ssDNA at the fork junction, creating an optimal DNA-protein substrate for SMARCAL-1. However, when the RPA binding to the ssDNA formed at the leading strand stimulates the SMARCAL1-mediated fork remodeling activity, the RPA binding at the lagging strand inhibits the SMARCAL1 activity [154]. 8. Enhancing primase: RPA enhances primase activity at forks [155] (Table 1).

4.3. CTC1

CTC1 and CDC13 are the largest subunits of the human and budding yeast CST complexes, respectively. As the CDC13 null strain cannot be generated in yeast cells [158], it is believed that CDC13 is essential for cell viability. Although STN1 and TEN1 are highly conserved [159], the genomic sequences and functions of CDC13 in yeast and that of CTC1 in humans are different [120]. Structurally, CDC13 consists of OB1, OB2, OB3, and OB4 (Figure 3A). The roles of these OBs include ssDNA binding, protein–protein interactions, DNA polymerase α-primase binding, and CDC13 homo-dimerization [160,161,162]. However, a recent study analyzing the crystal structures of the CST complex of Kluyveromyces lactis suggests that OB2 and OB4 are required for the CDC13–STN1 interaction that assembles CST in a 2:2:2, instead of1:1:1, stoichiometry [163].
Human CTC1 has OB-A, OB-B, OB-C, OB-D, OB-E, OB-F, and OB-G (Figure 3B). The C terminus of CTC1 (OB-D through OB-G) acts as a platform to assemble STN1 and TEN1 [118]. STN1 (STN1 OB and the first winged helix-turn-helix [wHTH1] domain of STN1) interacts with CTC1 at two interaction sites, CTC1 OB-G and CTC1 OB-E, respectively (Figure 3B) [118]. Structural analyses have shown that CTC1 OB-G is similar to the OB-C of RPA70 and not CDC13 [128]. The CDC13 recruitment domain (RD) contains numerous phosphorylation sites [164,165,166,167,168]. Phosphorylated CDC13 RD enhances the ever shorter telomere 1 (Est1), a component of the yeast telomerase holoenzyme binding and telomerase recruitment to telomeres [164,165]. Est1 is present in humans and a report shows that the expression of Est1 is significantly reduced in B-chronic lymphocytic leukemia [169]. Dephosphorylated CDC13 RD promotes CST complex assembly to bind and cap the ends of chromosomes [165,166,167,170]. Human CTC1 represses the elongation of telomerase by binding to telomerase-extended telomeres thus preventing telomerase activity [144]. CDC13–Est1 and POT–TPP1 are essential in directing telomerase to the chromosomal ends [171,172]. CTC1 interacts with TPP1 to compete with TPP1–POT1 for binding at the telomeric 3′ tail and sequestrate the single-stranded telomeric overhang to inhibit the telomerase extension reaction [120,144]. In humans and yeast, the CST complex prevents 3′ overhangs via boosting the fill-in synthesis [148,173]. Yeast CDC13 deficiency causes genome stability and unstable chromosomes [174]. Dyskeratosis congenita (DC) and Coats plus syndrome (CPS) are two uncommon diseases associated with mutations that affect the CST complex. CPS is an autosomal recessive, systemic disorder characterized by intrauterine growth retardation, bilateral exudative retinal telangiectasias, intracranial calcifications, intracerebral cysts, extra-neurological features, including osteopenia with a tendency of fractures and gastrointestinal bleeding, and portal hypertension [175]. Symptoms of DC include increased cancer incidence, bone marrow failure, lacy reticular pigmentation of the upper chest and/or neck and oral leukoplakia [176]. Changes that occur as a result of CTC1 and STN1 mutations include telomere DNA replication defects, genome instability, defects in interactions with Polα, chromosome breakage, and an accumulation of the ssDNA gaps of telomeric DNA [114,177,178]. Interestingly, some reports have identified shorter telomere length in the lymphocytes of subjects with CTC1 mutations [176,179], but contradictory results were also reported [180,181]. Nevertheless, further studies are crucial to clarify the roles of CST in disease pathogenesis.

4.4. STN1

Human STN1 was initially named as Polα accessory factor44 as STN1 has been shown to enhance primase and up-regulate the recruitment of Polα for lagging strand DNA replication [141,182]. Structurally, the yeast STN1 consists of an OB-5 domain and two wHTH motifs, wHTH1 and wHTH2, which may involve Polα and CDC13 binding [118]. The N-terminus of STN1 binds to TEN1, while the C-terminus associates with both CDC13 and Pol12 (the B subunit of Polα) [183,184].The STN1 and TEN1 are enlisted to telomere ends via direct association with CDC13. Both STN1 and TEN1 display relatively poor telomeric DNA-binding affinities [185]. In humans, STN1 functions as an adapter between TEN1 and CTC1 [122], and the STN1 N-terminal interacts with CTC1 OB-G and the C-terminal with CTC1 OB-E [118]. Fluorescence investigation has demonstrated that the STN1-binding sites are prone to DNA breakage in STN1 deficient cells under replication stress, leading to chromosome fragmentation [121]. The human STN1 and TEN1 can associate to form a stable complex in vitro [127,137,159,162,170,171,185,186,187,188], which may be due to the comprehensive interactions between the two C-terminal helices of the OB folds and the contacts between these domain bodies. Additionally, OB-fold functions may be related to its preference for the G-rich sequence [135,189]. CTC1 and STN1 can alone reduce telomerase activity and disrupt telomeric DNA damage signaling [138]. STN1-TEN1 forms a wide nucleic acid binding pocket on the surface of the protein complex [159]. STN1, the shelterin complex, and telomerase may recruit Polα to telomere [190]. A reduction in CTC1 or STN1 produces lengthened G-overhangs as the C-strand fill-in becomes faulty [136,177,191]. STN1mutationlike CTC1 mutation can cause CPS [179,192,193]. Depletion of human CTC1 or STN1 increases multi-telomeric signals, telomere instabilities, and chromosome breakage [191] and can result in impairing C-strand fill-in, leading to excessively long G-overhangs [125,134,136,143].

4.5. TEN1

Of the CST components, TEN1is the smallest with a single OB fold [118]. Yeast TEN1 may promote the activity of CDC13 and bind to telomeric ssDNA to enhance the DNA-binding activity of CDC13 [194]. TEN1 in humans is to stabilize the binding of CTC1-STN1to ssDNA and to support C-strand fill-in after G-strand extension by telomerase [138]. Human TEN1 attachment to CTC1 OB-G is facilitated by the OB of STN1 [118]. Human TEN1mutant strain proteins are unable promote the binding of CDC13 to telomeres in vitro, indicating that TEN1 improves the telomeric DNA-binding activity of CDC13 that then negatively affects the telomere length [138]. Knockout TEN1 cells show gradual telomere shortening comparable to that resulting from telomerase deficiency [138], indicating that TEN1 is crucial for the maintenance of telomere length. In addition to ensuring telomere stability [119], TEN1 and STN1 can rescue replication fork stalling during replication stress [122,125,195].
CDC13, STN1, and TEN1 are essential for cell viability and regulating telomere length. Subunit mutations resulting in loss-of-function can cause an accumulation of telomeric ssDNA and result in abnormal elongation of the telomeres, indicating that these three subunits are critical to the health of organisms with the CST complex [6,118,133,147]. The interactions between POT1-TPP1 and CST can significantly affect the telomere length and may result in telomere length dysregulation and cancer development, such as familial glioma [196], melanoma [197], chronic lymphocytic leukemia [198] and breast cancers [199,200], stomach cancers [199], and parathyroid cancers [201].

5. Telomerase: Breaking through the Limitation of Replication

Telomerase, the enzyme responsible for lengthening the telomeres, can extend the cellular lifespan or induce immortalization [1]. Typically, in healthy adult somatic cells, telomerase is inactive to avoid uncontrolled cellular proliferation [2], whereas in approximately 90% of human tumors, telomerase is up-regulated or reactivated to help tumor cells survive and multiply [202]. However, developing embryos, reproductive cells, activated immune cells, bone marrow, and adult stem cells show high telomerase activity [18].

5.1. Components of Telomerase

Structurally, human telomerase consists of the TERT (hTERT), the telomerase RNA template (TERC), and accessory proteins. The telomerase catalytic protein component encoded by hTERT has telomerase activity, and this activity does not rely on the other components [3]. The hTERT can wrap the chromosome to add single-stranded telomere repeats [203]. The TERC contains the template for telomere replication [7]. The accessory proteins include: 1. Dyskerin, a highly conserved nucleolar protein that catalyzes the pseudouridylation of specific residues in newly synthesized ribosomal RNAs and spliceosomal small nuclear RNAs [204]. 2. Non-histone protein 2 (NHP2) and nucleolar protein 10 (NOP10), which are both ribonucleoproteins [205]. 3. Glycine-arginine rich 1 (GAR1), which is involved in RNA metabolism [206]. 4. p23, a small but important cochaperone for the Heat shock protein 90 (HSP90) chaperoning pathway, as part of the complex with telomerase [207]. 5. Telomerase Cajal body protein 1 (TCAB1) [13] is a telomerase holoenzyme and markedly enriched in Cajal bodies (CBs). In addition to regulating the subcellular location of telomerase [208], TCAB1 facilitates the recruitment of telomerase to CBs in the S phase of the cell cycle. This recruitment is dependent on TCAB1 binding to a telomerase RNA component [209]. 6. Reptin and pontin, two ATPases, interact with TERT in the S phase of the cell cycle [210].7. Serine and arginine-rich splicing factor 11 (SRSF11) is a TERC-binding protein that influences telomerase function, subcellular localization, and biogenesis [211]. Although hTERT and TERC are sufficient to reconstitute telomerase activity in vitro, a functional holoenzyme complex assembles hTERT, TERC, and accessory proteins in vivo to extend the telomere length in rapid cell divisions for attenuating or preventing telomere erosion [212] (Figure 4A). Furthermore, TERC is up-regulated in carcinomas affecting the cervix, head and neck, lung, and ovary—possibly serving as a therapeutic target [213,214].

5.2. TERT Is Important forthe Activity of Telomerase

In cancer cells, high telomerase activity breaks through the limitation of replication and avoids activation of the DNA damage signaling pathway [215]. Telomerase activity has been detected in 42–54% of thyroid cancer cases [216,217], 86.6% of non-small cell lung cancer cases [218], more than 80% of hepatocellular carcinoma cases [219], and 76% of cervical cancer cases [220]. These data suggest the criticality of the telomerase in cancer cells acquiring immortality or progression. While telomerase activation is poorly understood, hTERT is an important factor in telomerase activation. Generally, hTERT acts as the limiting factor for controlling telomerase activity and turns on the telomere clock for the aging process in all somatic adult cells [221]. Fibroblasts are a good example as fibroblasts do not express TERT and thus demonstrate gradual telomere shortening and eventual replicative senescence. However, introduced TERT expression in fibroblasts can maintain the telomere length and immortalize the cells [222]. Furthermore, hTERT is aberrantly expressed in approximately 90% of aggressive cancers [203,223,224] and 73% of tumor cases [225]. Therefore, increased TERT expression and telomerase activity can be detected in close to 90% of human cancers [226,227], highlighting the association of the hTERT with the telomerase in the emergence of malignant and aggressive phenotypes.
hTERT is situated on the 5p15.33 chromosome, which is responsible for 40 kb of the human genome [5,228]. hTERT has 16 exons and 15 introns that can produce 22 splicing variants and these variants can be dominant or negative [229]. The promoter responsible for hTERT transcription lacks typical regulatory sequences, such as the TATA and CAAT boxes, and instead comprises a transcription start site (TSS), two E-boxes (5′-CACGTG-3′), and five GC boxes (5′-GGGCGG-3′). TSS is a single transcription start site that multi-functional transcription factors such as THF1 can bind to. E-boxes have attachment sites for MAD1 to down-regulate the transcription of hTERT. GC boxes interact with the zinc finger transcription factor SP-1 [230]. The region (Figure 4B, white rectangles) represents a 260 bp hTERT core promoter region and has multiple binding motifs for enhancing factors, such as c-MYC, SP-1, E-twenty-six (ETS) family members, and NF-kB. hTERT transcription has also been reportedly down-regulated by the transcriptional factors CTCF, MZF-2, and WT1. Therefore, the hTERT promoter is pivotal for transcriptional activity [32] (Figure 4B).

5.3. Mechanisms Involved in TERT Activation

5.3.1. TERT Promoter Is Critical in Cellular Immortality and Infinite Proliferation

In cancer cells, many mechanisms cause TERT activation. Several studies suggest that TERT is essential for malignant transformation, and transcriptional control of the TERT gene is highly regulated at various levels [231,232]. The mutation rate of the TERT promoter is more than 90% inhuman malignancies [232]. The mutation may activate telomerase, leading to infinite proliferation and infinite growth. Specifically, TERT promoter mutations may influence telomere length and affect gene expression [231]. TERT promoter mutations produce the ETS transcription factor family binding sites in multiple cancer types [232,233]. The ETS transcription factor family includes the activating GA Binding Protein Transcription Factor Subunit Alpha (GABPA) in a heterotetramer form with its counterpart, GA Binding Protein Transcription Factor Subunit Beta (GABPB), to activate TERT transcription and telomerase [233,234]. Consistently, thyroid carcinoma derived cells, which were knock-down GABPA, significantly down-regulated TERT expression [235] and GBM cells, which were knocked out of GABPB1, underwent apoptosis, and lost tumorigenic ability telomere shortening/dysfunction and proliferation/survival [234]. To sum up, inhibiting the GABPA or GABPB1 expression can lead to diminished TERT expression. These findings suggest that the TERT promoter mutations facilitate the binding of transcription factors, leading to the development of cancers.

5.3.2. C228T and C250T: Gain-of-Function Mutation

TERT transcription can be activated by point mutations at the TERT promoter, predominantly at two points (C228T and C250T) [233]. Primary tumors bearing either mutation (C228T or C250T) tend to express higher levels of TERT mRNA and telomerase activity, implying a stimulatory effect on TERT expression [225,231]. Chiba et al. created a C228T mutation in the TERT promoter region in human iPSCs and found that these cells constitutively expressed TERT and telomerase even after having undergone terminal differentiation, in contrast to the wild-type TERT promoter-bearing stem cell-derived progenies, where the TERT transcription was shut down following cellular differentiation [236]. Li et al. introduced the C228T mutation into the TERT promoter in normal human bladder stem cells, and this single event was sufficient to drive the transformation of these stem cells [237]. Thus, these findings suggest that the presence of C228T or C250T mutation has a “gain-of-function effect” that can confer cell immortality, sustain proliferation potential, and promote cell transformation by activating TERT transcription. These mutations are first found in familial and sporadic malignant melanomas [229]. These mutations were then associated with enhanced telomerase activity in cancer cells [231]. This mutation pair results in TERT activation that then triggers the telomerase to elongate telomere length [184,238] and, as a consequence, leads to the immortal, anti-senescence, and proliferative properties of tumor cells. The C228T mutation is more readily detected in cancers than the C250T mutation [225,231]. The results of C228T and C250T may enhance GABPA or GABPB1 binding, inhibiting DICER1, leading to cell proliferation, immortality, and oncogenesis.

5.3.3. Other Possible Mechanisms That Trigger TERT Activation

Several mechanisms, including TERT amplification, epigenetic changes [231], promoter methylation [239], alterations in alternative splicing of TERT pre-mRNA [214,228], chromatin remodeling, increased copy number, disruption of the telomere position effect (TPE) machinery [231], which is stronger with the longer telomere, and telomere shortening may lead to increased TERT transcription by the loss of TPE [240], which can trigger TERT activation. TPE has a silencing effect on genes located close to the telomeres [231]. TPE is also thought to be capable of regulating genes further away from the telomeres in what is called TPE over long distances (TPE-OLD) [232,241]. TERT expression is regulated by various transcription factors, such as MYC, SP-1, and E2F. AP-1, CCCTC binding factor, E2F, and estrogen response element (ERE) are associated with promoting TERT transcription [232,242,243]. Estrogen receptor α interacts with the TERT promoter and enhances the TERT mRNA output [232,242,243]. The phosphatidylinositol-3kinase/AKT pathway can phosphorylate TERT to increase the TERT functions [244,245].
TERT polymorphisms, such as rs2736100-CC genotype features, were reported more frequently than the wild-type TERT promoter in patients with cancer [246]. This genotype results in the elongation of telomeres through increasing TERT expression [247]. Although mammalian subtelomere regions are packed into constitutive heterochromatin, the detailed mechanisms of how epigenetic modifications impact telomere protection and structures are largely unknown. The heterochromatin is characterized by high levels of histone 3 lysine 9 trimethylation (H3K9me3), histone 4 at Lys20 (H4K20) trimethylation, and CpG methylation [248]. Stern et al. noted that a single nucleotide mutation in the TERT promoter led to the presence of the H3K4me2/3, which were associated with competent chromatin, and recruited the GABPA/B1 transcription factor that was transcriptionally active in several cancer cell lines [249].

5.3.4. Telomeric Repeat-Containing RNA (TERRA)

Telomeric sequences contain TERRA, which is a very long, non-coding RNA [250]. TERRA can interact with TRF1 and TRF2 to regulate telomere length [250,251] and promote telomere shortening by inhibiting telomerase activity [252], promoting exonuclease 1 dependent resections [253], increasing the formation of euchromatin and decreasing heterochromatin formation [254]. The ability of TERRA to regulate telomere length is also related to human diseases. For instance, elevated levels of TERRA were found in ALT-positive tumors [255] and TERRA levels showed down-regulation following tumor grades in human astrocytoma cell lines [256]. These findings suggest that TERRA participates in the regulation of various physiological processes, such as telomerase activity, TERT and heterochromatinization, and disruptions to these processes can affect development and directly or indirectly induce disease. Moreover, TERRA is associated with telomeres via tethering in a DNA-RNA hybrid formation [257,258]. This structure may help to further understand the molecular interaction between DNA-RNA at the chromosome ends and to design specific telomere-targeting therapy in the future.

6. Telomerase-Based Anti-Cancer Strategy

The fundamental concept of cancer immunotherapy is based on manipulating the host immune system to attack the cancer cells. Although there are several novel cancer immunotherapy strategies, vaccine-based strategies are the most attractive and promising ones. However, it is very difficult to target tumor-associated antigens on the surface of tumor cells but not on that of the normal cells because of the heterogeneity and overlapping expression of these antigens in both cancers and healthy tissues [259]. As cancer cells lacking telomerase can undergo spontaneous remission, telomerase inhibition in most cancers may shed light on a potentially successful therapeutic strategy [260]. As telomerase is an HLA class-I antigen and can stimulate a cell-mediated immune response by inducing cytotoxic T-cells, numerous novel approaches have recently been developed to attenuate/inhibit the functions of the telomerase that impact cancer. Vaccination against telomerase is tolerable and safe and has been shown to induce excellent immunological responses associated with increased survival in several cancer types. Four well-known telomerase-based anticancer vaccines are discussed below.

6.1. GV1001

The GV1001, an HLA class II-restricted peptide vaccine, is composed of 16 amino acids (TERT611–626:EARPALLTSRLRFIPK) derived from the hTERT active site [32,261]. GV1001 was the first TERT peptide vaccine to be evaluated for treating advanced pancreatic cancer, lung carcinoma, melanoma, and liver carcinoma in clinical trials [261,262,263,264,265,266,267,268]. GM-CSF can enhance immunological response through the recruitment and maturation of dendritic cells and the activation of macrophages, neutrophils, and NK cells [269]; therefore, GV1001in combination with GM-CSF can result in a high frequency of immune responder s [32,262]. GV1001 can induce an efficient hTERT-specific T-cell activation and penetrate within tumor cells through the cell membrane [270]. Therefore, it can recognize the antigen-presenting cells that are internalized in the tumor and lymph nodes [262]. GV1001 can induce cancer cell apoptosis [271,272,273] and down-regulate heat shock proteins, hypoxia-inducible factor-1, and vascular endothelial growth factor to enhance its anti-tumor effect [270,272,274]. Although GV1001 is theoretically suitable for most cancers, a report suggests that the GV1001 vaccination is not effective in cutaneous T-cell lymphoma [275], and another report indicates that GV1001 cannot induce any specific immune responses in patients with advanced HCC [263], and the addition of GV1001 to chemotherapy (gemcitabine and capecitabine) did not show any significant clinical benefits [266]. Patients with tuberculosis or receiving tuberculin may not be suitable for GV1001 vaccination because the evoked immune response against mycobacterial peptides may be so dominant as to suppress the immune response against the hTERT peptide [264].

6.2. GX301

The GX301 vaccine contains four immunogenic peptides (hTERT540–548: ILAKFLHWL; hTERT611–626: EARPALLTSRLRFIPK; hTERT672–686: RPGLLGASVLGLDDI, and hTERT766–780: LTDLQPYMRQFVAHL) that can bind both HLA class I and II; GX301 also contains two complementary adjuvants, Montanide ISA-51 and Imiquimod. Each GX301 administration consists of four intradermal injections (a fixed hTERT peptide dose, 500 µg)—one injection for each hTERT peptide—given at the same time and followed by topical application of imiquimod [276]. Montanide can protect the degradation of the peptides by tissue proteases, enhance peptide uptake by intradermal dendritic cells, induce interferon- γ release by innate immunity cells, and increase the expression of major histocompatibility complex (MHC) by tumor cells [277]. Imiquimod can activate the Toll-like receptor-7 and receptor-8 and induce the activation and maturation of dendritic cells [278]. The immunogenicity of GX301 was demonstrated in an ex vivo study in which circulating T-cell responses to its hTERT peptides were detected in all subjects [276]. A phase I trial of GX301 has provided evidence of vaccine-specific immune response in patients with stage IV prostate and kidney cancer, and prolonged progression-free survival and overall survival were observed in patients showing a full pattern of vaccine-specific immunologic responses [276]. A phase II, randomized, parallel-group, open-label, multicenter trial (EudraCT: 2014-000095-26 and ClinicalTrials.gov Identifier: NCT02293707) has demonstrated that all the patients showed good immune responses to at least one of the peptides. The overall response was more for the multi-peptide vaccines than the single-peptide vaccines [279], suggesting that the four GX301 peptides endow a cumulative epitope pattern wide enough for inducing telomerase-specific peripheral T-cell reactivity in most individuals. A phase II, multicenter, randomized, parallel-group, open-label trial (EudraCT:2014-000095-26 and ClinicalTrials.gov Identifier:NCT02293707) was designed to comparatively analyze the safety and immunological response to GX301 regimens in castration-resistant prostate cancer patients with response/disease stability after docetaxel chemotherapy. Although the results indicate that the GX301 cancer vaccine is safe and 95% of the patients showed at least one vaccine-specific immune response, the overall survival did not differ between immunological responders and non-responders [280].

6.3. UV1

UV1 is a second-generation, multi-peptide vaccine constituted by three hTERT-derived peptides (hTERT652–665: AERLTSRVKALFSVL; hTERT660–689: ALFSVLNYERARRPGLLGASVLGLDDIHRA and hTERT691–705: RTFVLRVRAQDPPPE) [262]. In phase I and IIa trials, UV1 was administered along with GM-CSF for six months in patients with metastatic prostate cancer in combination with radiotherapy and androgen deprivation treatment (ADT). A total of 85.7% of patients showed an immune activation and 64% showed reduced levels of the prostate-specific antigen (PSA). In addition, 45% of the patients showed no evidence of the disease at the end of the trial [281]. Several checkpoint inhibitors, including Ipilimumab (anti-CTLA-4) or pembrolizumab (anti-PD-1) in melanoma patients (NCT02275416 and NCT03538314, respectively) and ipilimumab in association with nivolumab (anti-PD-L1) in patients affected by mesothelioma (NIPU trial, NCT04300244) have been singly or multiply used in combination with UV1 in clinical trials. The results showed that the treatment of UV1 together with these checkpoint inhibitors were safe and well-tolerated, and no severe allergic reactions were observed [282,283,284]. The NIPU trial is still ongoing and the primary end-point is expected to be analyzed in 2022.

6.4. Vx-001

Vx-001 is a peptide-based cancer vaccine consisting of two peptides: hTERT-derived low-affinity cryptic hTERT peptide: TERT 572 (RLFFYRKSV; ARG-Vx001) and its optimized mutant hTERT peptide: TERT 572Y (YLFFYRKSV; TYR-Vx001), which has an enhanced affinity to MHC class I molecules as the first amino acid was replaced with a tyrosine residue [285]. The antitumor efficacy and safety of Vx-001 has also been investigated in phase I/II clinical trials for different cancers, such as melanoma, bile duct cancer, breast cancer, and lung cancer. Results of these trials show that Vx-001may elicit a specific and possibly optimal cytotoxic T cell response against hTERT-expressing tumor cells and has improved clinical outcomes in clinical trials without any relevant toxicity [286,287,288,289].
Collectively, the hTERT-vaccine clinical trials indicate that these immunotherapies may represent a promising approach in cancer treatment. Apart from the TERT peptide vaccines, several novel immunotherapies, including the dendritic cell-based tumor vaccine, such as GRNVAC1 [289] and GRNVAC2 [290,291]; Tumor Antigen Presenting Cells (TAPCells) vaccines [292]; DNA vaccines such as phTERT [293], INVAC-1 [294]; adenovirus type 6 of an anticancer vaccine expressing hTERT, such astheV934/V935 vaccine [295]; gene-modified T-cell therapy, such as the use of tumor antigen-specific T-cell receptors originating from tumor-specific T cells or their clones [296,297]; the use of a chimeric antigen receptor (CAR) [298,299]; the molecules inhibiting Ras farnesylation [76], and hTERT-expressing human umbilical endothelial cells (HUVEC-TERTs) [300], may be effective without prominent toxicity.

7. Alternative Lengthening of Telomere (ALT)

7.1. ATRX and DAXX

Even with these new therapies, there have been certain cancers that can evade treatment by using an alternative lengthening of the telomere (ALT) mechanism. ALT is a telomerase-independent mechanism that uses recombination-dependent pathways to increase telomere length [301]. ALT is present in non-neoplastic tissues and in stromal, endothelial, and epithelial cells [302] and in approximately 10-15% of cancers [303], and it is common in sarcoma and glioma [304,305]. In the absence of telomerase, the ALT pathway uses a homologous recombination-based DNA replication mechanism to gain immortality. ALT activation required two chromatin-remodeling factors: the α-thalassemia X-linked intellectual disability (ATRX) and the death domain-associated protein (DAXX) [304,306]. DAXX was initially describe as a Fas death receptor binding protein [307]. ATRX is widely expressed and is a multi-functional factor involved in chromatin organization, DNA methylation, and transcriptional regulation [308]. Mutations in ATRX result in α-thalassemia ATRX syndrome, which is characterized by severe developmental delays, peculiar facial hypotonia and a characteristic mouth, intellectual impairment, genital anomalies ranging from undescended testes to ambiguous genitalia, and anemia secondary to α-thalassemia [309]. Patients with this syndrome may present long telomeres, which may be due to either improper maintenance of telomeric heterochromatin, improper resolution of replication stress at telomeres, or both by the mutation of ATRX [310].

7.2. Correlation between the Loss-of-Function of ATRX/DAXX and ALT in Cancer

The mutated ATRX gene is frequently detected in several tumors, including adrenocortical carcinoma, gliomas, GBM, neuroblastoma, and osteosarcoma [306], and pancreatic neuroendocrine tumors(panNETs), which are a group of endocrine tumors arising in the pancreas. PanNETs are among the most common neuroendocrine tumors. Functioning panNETs include insulinoma, gastrinoma, vasoactive intestinal peptide tumors (VIPoma), glucagonoma, and others that produce specific hormonal hypersecretion syndromes. Endocrine testing, imaging, and histological evidence is necessary to accurately diagnose panNETs. PanNETs may or may not cause signs or symptoms; however, as most panNETs may have malignant potential, an aggressive therapeutic approach for panNETs, including surgery, locoregional therapy, systemic therapy, and complication control, is required [311]. A report showed that 43% of panNETs contained the mutated ATRX or DAXX [312]. A correlation between the loss-of-function of ATRX/DAXX and the ALT phenotype in panNETs was found [313] and ATRX was proposed to serve as the primary suppressor of ALT [314]. Furthermore, when ATRX was reintroduced into ALT-positive ATRX-negative cell lines it was found to eliminate ALT-associated phenotypes [315,316]. Gliomas with wild-typeTERT promoters often present ATRX mutations to activate ALT [317]. A fibrosarcoma cell line (HTC75), which is telomerase-positive, can be converted to an ALT-mediated telomere elongation mechanism through TERT knockout, and the subsequent changes result in telomeric DNA damage and disruption of the ATRX/DAXX complex, indicating a negative correlation between mutations affecting TERT and ATRX/DAXX [304]. Consequently, telomeric DNA damage can reduce the compaction of telomeric chromatin, resulting in the production of altered telomeric DNA sequences. This in turn activates a telomere-specific DDR pathway [12,316], which can stimulate the homology-directed synthesis of telomeric DNA. However, cancer cells can circumvent cell death caused by an absence of telomerase or dysfunction by switching from telomerase-dependent to ALT-mediated telomere lengthening [318,319].

7.3. Targeting Telomerase Activity and the ATRX/DAXX Complex

Direct and indirect approaches to targeting telomerase activity and the ATRX/DAXX complex could prove effective. Direct approaches include immunotherapy specifically targeting TERT tumor-associated antigens, such as anti-sense oligonucleotides (e.g., Imetelstat/GRN163L) and small-molecule inhibitors (e.g., BIBR1532) and small molecule inhibitors could be used to bind telomerase and inhibit telomere elongation. Indirect techniques, such as G-quadruplex stabilizers (e.g., RHPS4, Telomestatin, TMPyP4, CX-3543/quarfloxacin), which are designed to block telomerase activity, are promising [320]. The G-rich oligos, which homolog to the telomeric overhang that forms the G4 structures, cause telomere dysregulation and a decreased proliferation rate, enhance apoptosis, and reduce expression of the TERT within melanoma cells [321].An alternative approach is based on telomere uncapping, using nucleoside analogs(e.g., 6-thio-dG) that rapidly affect telomere dysfunction, quickly triggering cancer cell death [322]. In addition, other factors, such as transcriptional, posttranscriptional, and epigenetic modifications can affect the activation or silencing of TERT; however, the effects are poorly understood in somatic, cancer, and stem cells. Epigenetic regulators, such as non-coding RNAs, histone modification, and DNA methylation, are now seen as crucial components for the regulation of telomeres and telomerase activity [323] and unlocking the epigenetic mechanisms associated with telomerase regulation could see advances in cancer diagnosis, treatment, and prognosis [324]. Convergently, a multipronged treatment strategy can maximize anti-tumor effects.

8. Conclusions

Telomeres are hexameric DNA repeats situated at the ends of human chromosomes and are associated with genome replication and protection. Telomere length regulation has been attributed to several functional modulators, telomerase, and two important protein complexes, shelterin and CST synergizing with DNA replication, repair, and RNA metabolism pathway components. The mechanisms involving telomere maintenance play a critical role in cancer development and thus form the primary targets for the development of cancer therapeutics. In the majority of cancers, tumors attain replicative immortality primarily through telomerase activation via increased TERT transcription, providing telomerase as the preferred target for drug development in cancer therapeutics.
Among numerous novel anticancer therapies, vaccine-based strategies are the most attractive and promising approach. While there is much excitement in the telomerase-against-cancer arena, many challenges and questions still remain unanswered. Although much work is needed, several clinical trials have shown the effective anticancer responses of these telomerase-based anticancer vaccines without toxicity to non-cancer cells. Furthermore, the suppression of CST, shelterin, and ALT cells may inhibit cell proliferation. Combining vaccines with the targeted manipulation of the immune response through different pharmacological approaches may improve the overall efficacy of telomerase-based anti-cancer immunotherapies in future trials.

Author Contributions

H.-C.F., F.-W.C., J.-D.T., K.-M.L., C.-M.C., C.-A.L., S.-Z.L. and H.-J.H. were involved in the conception and literature review and drafting the manuscript. H.-C.F., S.-Z.L. and H.-J.H. revised the manuscript critically for important intellectual content. F.-W.C., J.-D.T., K.-M.L., C.-M.C. and C.-A.L., provided critical questions and suggestions for the manuscripts. H.-C.F., C.-M.C., C.-A.L., S.-Z.L. and H.-J.H. conceptualized the review, supervised all aspects of the study, critically reviewed and revised the manuscript, and approved the final manuscript as submitted. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Tungs’ Taichung Metroharbor hospital research grant number TTMHH-R1100003 and TTMHH-R1100004.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable because this review did not report any data.

Acknowledgments

We thank Ching-Shiang Chi (Department of Pediatrics, Tungs’ Taichung Metroharbor Hospital, Taiwan), Shyi-Jou Chen (Department of Pediatrics, Tri-Service General Hospital, National Defense Medical Center, Taiwan), Kai-Ping Chang (Department of Pediatrics, Wei-Gong Memorial Hospital and department of Pediatrics, Taipei Veterans General Hospital, Taiwan), and Wang-Tso Lee (Department of Pediatric Neurology, National Taiwan University Children’s Hospital, Taiwan) for helpful suggestions and comments on the MS.

Conflicts of Interest

The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Ferlay, J.; Colombet, M.; Soerjomataram, I.; Parkin, D.M.; Piñeros, M.; Znaor, A.; Bray, F. Cancer statistics for the year 2020: An overview. Int. J. Cancer 2021, 149, 778–789. [Google Scholar] [CrossRef]
  2. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer statistics, 2021. CA Cancer J. Clin. 2021, 71, 7–33. [Google Scholar] [CrossRef]
  3. Vogelstein, B.; Kinzler, K.W. The Path to Cancer—Three Strikes and You’re out. N. Engl. J. Med. 2015, 373, 1895–1898. [Google Scholar] [CrossRef] [PubMed]
  4. Krupp, G.; Bonatz, G.; Parwaresch, R. Telomerase, immortality and cancer. Biotechnol. Annu. Rev. 2000, 6, 103–140. [Google Scholar] [CrossRef] [PubMed]
  5. Cong, Y.-S.; Wright, W.E.; Shay, J.W. Human Telomerase and Its Regulation. Microbiol. Mol. Biol. Rev. 2002, 66, 407–425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Lim, C.J.; Cech, T.R. Shaping human telomeres: From shelterin and CST complexes to telomeric chromatin organization. Nat. Rev. Mol. Cell Biol. 2021, 22, 283–298. [Google Scholar] [CrossRef]
  7. De Lange, T. How telomeres solve the end-protection problem. Science 2009, 326, 948–952. [Google Scholar] [CrossRef] [Green Version]
  8. Meyne, J.; Ratliff, R.L.; Moyzis, R.K. Conservation of the human telomere sequence (TTAGGG)n among vertebrates. Proc. Natl. Acad. Sci. USA 1989, 86, 7049–7053. [Google Scholar] [CrossRef] [Green Version]
  9. Lundblad, V. Telomere end processing: Unexpected complexity at the end game: Figure 1. Genes Dev. 2012, 26, 1123–1127. [Google Scholar] [CrossRef] [Green Version]
  10. Gong, J.; Costanzo, A.; Yang, H.-Q.; Melino, G.; Kaelin, W.G., Jr.; Levrero, M.; Wang, J.Y.J. The tyrosine kinase c-Abl regulates p73 in apoptotic response to cisplatin-induced DNA damage. Nature 1999, 399, 806–809. [Google Scholar] [CrossRef]
  11. Stiewe, T.; Putzer, B.M. P73 in apoptosis. Apoptosis 2001, 6, 447–452. [Google Scholar] [CrossRef]
  12. O’Sullivan, R.J.; Karlseder, J. Telomeres: Protecting chromosomes against genome instability. Nat. Rev. Mol. Cell Biol. 2010, 11, 171–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Lopez-Otin, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. The hallmarks of aging. Cell 2013, 153, 1194–1217. [Google Scholar] [CrossRef] [Green Version]
  14. Zhao, Y.; Abreu, E.; Kim, J.; Stadler, G.; Eskiocak, U.; Terns, M.P.; Terns, R.M.; Shay, J.W.; Wright, W.E. Processive and Distributive Extension of Human Telomeres by Telomerase under Homeostatic and Nonequilibrium Conditions. Mol. Cell 2011, 42, 297–307. [Google Scholar] [CrossRef] [Green Version]
  15. Li, B.; Reddy, S.; Comai, L. Sequence-specific processing of telomeric 3′ overhangs by the Werner syndrome protein exonuclease activity. Aging 2009, 1, 289–302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Muraki, K.; Nyhan, K.; Han, L.; Murnane, J.P. Mechanisms of telomere loss and their consequences for chromosome instability. Front. Oncol. 2012, 2, 135. [Google Scholar] [CrossRef] [Green Version]
  17. Frenck, R.W., Jr.; Blackburn, E.H.; Shannon, K.M. The rate of telomere sequence loss in human leukocytes varies with age. Proc. Natl. Acad. Sci. USA 1998, 95, 5607–5610. [Google Scholar] [CrossRef] [Green Version]
  18. Dalgård, C.; Benetos, A.; Verhulst, S.; Labat, C.; Kark, J.D.; Christensen, K.; Kimura, M.; Kyvik, K.O.; Aviv, A. Leukocyte telomere length dynamics in women and men: Menopause vs age effects. Int. J. Epidemiol. 2015, 44, 1688–1695. [Google Scholar] [CrossRef] [Green Version]
  19. Coburn, S.B.; Graubard, B.I.; Trabert, B.; McGlynn, K.A.; Cook, M.B. Associations between circulating sex steroid hormones and leukocyte telomere length in men in the National Health and Nutrition Examination Survey. Andrology 2018, 6, 542–546. [Google Scholar] [CrossRef] [PubMed]
  20. Arsenis, N.C.; You, T.; Ogawa, E.F.; Tinsley, G.M.; Zuo, L. Physical activity and telomere length: Impact of aging and potential mechanisms of action. Oncotarget 2017, 8, 45008–45019. [Google Scholar] [CrossRef] [Green Version]
  21. Welendorf, C.; Nicoletti, C.F.; Pinhel, M.A.D.S.; Noronha, N.; de Paula, B.M.F.; Nonino, C.B. Obesity, weight loss, and influence on telomere length: New insights for personalized nutrition. Nutrition 2019, 66, 115–121. [Google Scholar] [CrossRef]
  22. Salihu, H.M.; Pradhan, A.; King, L.; Paothong, A.; Nwoga, C.; Marty, P.J.; Whiteman, V. Impact of intrauterine tobacco exposure on fetal telomere length. Am. J. Obstet. Gynecol. 2015, 212, 205.e1–205.e8. [Google Scholar] [CrossRef] [PubMed]
  23. Leung, C.W.; Fung, T.T.; McEvoy, C.T.; Lin, J.; Epel, E.S. Diet Quality Indices and Leukocyte Telomere Length Among Healthy US Adults: Data from the National Health and Nutrition Examination Survey, 1999–2002. Am. J. Epidemiol. 2018, 187, 2192–2201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Calado, R.T.; Young, N.S. Telomere diseases. N. Engl. J. Med. 2009, 361, 2353–2365. [Google Scholar] [CrossRef] [PubMed]
  25. Jacobs, J.J.L. Loss of Telomere Protection: Consequences and Opportunities. Front. Oncol. 2013, 3, 88. [Google Scholar] [CrossRef] [Green Version]
  26. Farzaneh-Far, R.; Cawthon, R.M.; Na, B.; Browner, W.S.; Schiller, N.B.; Whooley, M.A. Prognostic value of leukocyte telomere length in patients with stable coronary artery disease: Data from the heart and soul study. Arterioscler. Thromb. Vasc. Biol. 2008, 28, 1379–1384. [Google Scholar] [CrossRef]
  27. Yang, Z.; Huang, X.; Jiang, H.; Zhang, Y.; Liu, H.; Qin, C.; Eisner, G.M.; Jose, P.; Rudolph, L.; Ju, Z. Short Telomeres and Prognosis of Hypertension in a Chinese Population. Hypertension 2009, 53, 639–645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Yeh, J.-K.; Lin, M.-H.; Wang, C.-Y. Telomeres as Therapeutic Targets in Heart Disease. JACC Basic Transl. Sci. 2019, 4, 855–865. [Google Scholar] [CrossRef]
  29. Sharifi-Sanjani, M.; Oyster, N.M.; Tichy, E.D.; Bedi, K.C., Jr.; Harel, O.; Margulies, K.B.; Mourkioti, F. Cardiomyocyte-Specific Telomere Shortening is a Distinct Signature of Heart Failure in Humans. J. Am. Heart Assoc. 2017, 6, e005086. [Google Scholar] [CrossRef] [Green Version]
  30. Valdes, A.M.; Richards, J.B.; Gardner, J.P.; Swaminathan, R.; Kimura, M.; Xiaobin, L.; Aviv, A.; Spector, T.D. Telomere length in leukocytes correlates with bone mineral density and is shorter in women with osteoporosis. Osteoporos. Int. 2007, 18, 1203–1210. [Google Scholar] [CrossRef]
  31. Grunnet, L.G.; Pilgaard, K.A.; Alibegovic, A.; Jensen, C.B.; Hjort, L.; Ozanne, S.; Bennett, M.; Vaag, A.A.; Brøns, C. Leukocyte telomere length is associated with elevated plasma glucose and HbA1c in young healthy men independent of birth weight. Sci. Rep. 2019, 9, 7639. [Google Scholar] [CrossRef]
  32. Jafri, M.A.; Ansari, S.A.; Alqahtani, M.H.; Shay, J.W. Roles of telomeres and telomerase in cancer, and advances in telomerase-targeted therapies. Genome Med. 2016, 8, 69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. De Lange, T. Shelterin: The protein complex that shapes and safeguards human telomeres. Genes Dev. 2005, 19, 2100–2110. [Google Scholar] [CrossRef] [Green Version]
  34. Diotti, R.; Loayza, D. Shelterin complex and associated factors at human telomeres. Nucleus 2011, 2, 119–135. [Google Scholar] [CrossRef] [Green Version]
  35. Chen, L.-Y.; Liu, D.; Songyang, Z. Telomere Maintenance through Spatial Control of Telomeric Proteins. Mol. Cell. Biol. 2007, 27, 5898–5909. [Google Scholar] [CrossRef] [Green Version]
  36. Liu, D.; O’Connor, M.S.; Qin, J.; Songyang, Z. Telosome, a Mammalian Telomere-associated Complex Formed by Multiple Telomeric Proteins. J. Biol. Chem. 2004, 279, 51338–51342. [Google Scholar] [CrossRef] [Green Version]
  37. O’Connor, M.S.; Safari, A.; Xin, H.; Liu, D.; Songyang, Z. A critical role for TPP1 and TIN2 interaction in high-order telomeric complex assembly. Proc. Natl. Acad. Sci. USA 2006, 103, 11874–11879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Ye, J.Z.; de Lange, T. Tin2 is a tankyrase 1 parp modulator in the trf1 telomere length control complex. Nat. Genet. 2004, 36, 618–623. [Google Scholar] [CrossRef] [Green Version]
  39. Li, J.S.Z.; Fusté, J.M.; Simavorian, T.; Bartocci, C.; Tsai, J.; Karlseder, J.; Denchi, E.L. TZAP: A telomere-associated protein involved in telomere length control. Science 2017, 355, 638–641. [Google Scholar] [CrossRef] [Green Version]
  40. Marcand, S.; Gilson, E.; Shore, D. A Protein-Counting Mechanism for Telomere Length Regulation in Yeast. Science 1997, 275, 986–990. [Google Scholar] [CrossRef] [PubMed]
  41. Patel, T.N.; Vasan, R.; Gupta, D.; Patel, J.; Trivedi, M. Shelterin Proteins and Cancer. Asian Pac. J. Cancer Prev. 2015, 16, 3085–3090. [Google Scholar] [CrossRef] [Green Version]
  42. Palm, W.; de Lange, T. How shelterin protects mammalian telomeres. Annu. Rev. Genet. 2008, 42, 301–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Xin, H.; Liu, D.; Songyang, Z. The telosome/shelterin complex and its functions. Genome Biol. 2008, 9, 232. [Google Scholar] [CrossRef] [PubMed]
  44. Zimmermann, M.; Kibe, T.; Kabir, S.; de Lange, T. Trf1 negotiates ttaggg repeat-associated replication problems by recruiting the blm helicase and the tpp1/pot1 repressor of atr signaling. Genes Dev. 2014, 28, 2477–2491. [Google Scholar] [CrossRef] [Green Version]
  45. Shore, D. Telomerase and telomere-binding proteins: Controlling the endgame. Trends Biochem. Sci. 1997, 22, 233–235. [Google Scholar] [CrossRef]
  46. Karlseder, J.; Kachatrian, L.; Takai, H.; Mercer, K.; Hingorani, S.; Jacks, T.; de Lange, T. Targeted Deletion Reveals an Essential Function for the Telomere Length Regulator Trf1. Mol. Cell. Biol. 2003, 23, 6533–6541. [Google Scholar] [CrossRef] [Green Version]
  47. Pal, D.; Sharma, U.; Singh, S.K.; Kakkar, N.; Prasad, R. Over-Expression of Telomere Binding Factors (TRF1 & TRF2) in Renal Cell Carcinoma and Their Inhibition by Using SiRNA Induce Apoptosis, Reduce Cell Proliferation and Migration Invitro. PLoS ONE 2015, 10, e0115651. [Google Scholar] [CrossRef] [Green Version]
  48. Hu, H.; Zhang, Y.; Zou, M.; Yang, S.; Liang, X.-Q. Expression of TRF1, TRF2, TIN2, TERT, KU70, and BRCA1 proteins is associated with telomere shortening and may contribute to multistage carcinogenesis of gastric cancer. J. Cancer Res. Clin. Oncol. 2010, 136, 1407–1414. [Google Scholar] [CrossRef] [PubMed]
  49. Bejarano, L.; Schuhmacher, A.J.; Méndez, M.; Megias, D.; Blanco-Aparicio, C.; Martínez, S.; Pastor, J.; Squatrito, M.; Blasco, M.A. Inhibition of TRF1 Telomere Protein Impairs Tumor Initiation and Progression in Glioblastoma Mouse Models and Patient-Derived Xenografts. Cancer Cell 2017, 32, 590–607.e4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Chen, Y.; Yang, Y.; van Overbeek, M.; Donigian, J.R.; Baciu, P.; de Lange, T.; Lei, M. A Shared Docking Motif in TRF1 and TRF2 Used for Differential Recruitment of Telomeric Proteins. Science 2008, 319, 1092–1096. [Google Scholar] [CrossRef] [PubMed]
  51. Bianchi, A.; Smith, S.; Chong, L.; Elias, P.; De Lange, T. TRF1 is a dimer and bends telomeric DNA. EMBO J. 1997, 16, 1785–1794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Broccoli, D.; Smogorzewska, A.; Chong, L.; de Lange, T. Human telomeres contain two distinct Myb–related proteins, TRF1 and TRF2. Nat. Genet. 1997, 17, 231–235. [Google Scholar] [CrossRef]
  53. Van Steensel, B.; Smogorzewska, A.; de Lange, T. TRF2 protects human telomeres from end-to-end fusions. Cell 1998, 92, 401–413. [Google Scholar] [CrossRef] [Green Version]
  54. Muñoz, P.; Blanco, R.; Blasco, M.A. Role of the TRF2 Telomeric Protein in Cancer and Aging. Cell Cycle 2006, 5, 718–721. [Google Scholar] [CrossRef]
  55. Dantzer, F.; Giraud-Panis, M.-J.; Jaco, I.; Amé, J.-C.; Schultz, I.; Blasco, M.; Koering, C.-E.; Gilson, E.; Murcia, J.M.-D.; de Murcia, G.; et al. Functional Interaction between Poly(ADP-Ribose) Polymerase 2 (PARP-2) and TRF2: PARP Activity Negatively Regulates TRF2. Mol. Cell. Biol. 2004, 24, 1595–1607. [Google Scholar] [CrossRef] [Green Version]
  56. Li, B.; Oestreich, S.; de Lange, T. Identification of Human Rap1: Implications for Telomere Evolution. Cell 2000, 101, 471–483. [Google Scholar] [CrossRef] [Green Version]
  57. Opresko, P.L.; von Kobbe, C.; Laine, J.-P.; Harrigan, J.; Hickson, I.D.; Bohr, V.A. Telomere-binding Protein TRF2 Binds to and Stimulates the Werner and Bloom Syndrome Helicases. J. Biol. Chem. 2002, 277, 41110–41119. [Google Scholar] [CrossRef] [Green Version]
  58. Zhu, X.D.; Kuster, B.; Mann, M.; Petrini, J.H.; de Lange, T. Cell-cycle-regulated association of rad50/mre11/nbs1 with TRF2 and human telomeres. Nat. Genet. 2000, 25, 347–352. [Google Scholar] [CrossRef] [PubMed]
  59. German, J.; Archibald, R.; Bloom, D. Chromosomal Breakage in a Rare and Probably Genetically Determined Syndrome of Man. Science 1965, 148, 506–507. [Google Scholar] [CrossRef] [PubMed]
  60. Hand, R.; German, J. A retarded rate of DNA chain growth in bloom’s syndrome. Proc. Natl. Acad. Sci. USA 1975, 72, 758–762. [Google Scholar] [CrossRef] [Green Version]
  61. Bernstein, K.A.; Gangloff, S.; Rothstein, R. The RecQ DNA Helicases in DNA Repair. Annu. Rev. Genet. 2010, 44, 393–417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Cunniff, C.; Bassetti, J.A.; Ellis, N.A. Bloom’s syndrome: Clinical spectrum, molecular pathogenesis, and cancer predisposition. Mol. Syndromol. 2017, 8, 4–23. [Google Scholar] [CrossRef] [PubMed]
  63. Goto, M.; Miller, R.W.; Ishikawa, Y.; Sugano, H. Excess of rare cancers in Werner syndrome (adult progeria). Cancer Epidemiol. Biomark. Prev. 1996, 5, 239–246. [Google Scholar]
  64. Hasty, P.; Campisi, J.; Hoeijmakers, J.; van Steeg, H.; Vijg, J. Aging and Genome Maintenance: Lessons from the Mouse? Science 2003, 299, 1355–1359. [Google Scholar] [CrossRef] [PubMed]
  65. Oshima, J. The werner syndrome protein: An update. Bioessays 2000, 22, 894–901. [Google Scholar] [CrossRef]
  66. Brosh, R.M., Jr. DNA helicases involved in DNA repair and their roles in cancer. Nat. Rev. Cancer 2013, 13, 542–558. [Google Scholar] [CrossRef]
  67. Greider, C.W. Telomeres do d-loop-t-loop. Cell 1999, 97, 419–422. [Google Scholar] [CrossRef] [Green Version]
  68. 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. [Google Scholar] [CrossRef] [Green Version]
  69. Arnoult, N.; Karlseder, J. Complex interactions between the DNA-damage response and mammalian telomeres. Nat. Struct. Mol. Biol. 2015, 22, 859–866. [Google Scholar] [CrossRef]
  70. Celli, G.B.; de Lange, T. DNA processing is not required for atm-mediated telomere damage response after TRF2 deletion. Nat. Cell Biol. 2005, 7, 712–718. [Google Scholar] [CrossRef] [PubMed]
  71. Muñoz, P.; Blanco, R.; Flores, J.M.; Blasco, M.A. XPF nuclease-dependent telomere loss and increased DNA damage in mice overexpressing TRF2 result in premature aging and cancer. Nat. Genet. 2005, 37, 1063–1071. [Google Scholar] [CrossRef]
  72. Nakane, H.; Takeuchi, S.; Yuba, S.; Saijo, M.; Nakatsu, Y.; Murai, H.; Nakatsuru, Y.; Ishikawa, T.; Hirota, S.; Kitamura, Y.; et al. High incidence of ultraviolet-B-or chemical-carcinogen-induced skin tumours in mice lacking the xeroderma pigmentosum group A gene. Nature 1995, 377, 165–168. [Google Scholar] [CrossRef]
  73. Sands, A.T.; Abuin, A.; Sanchez, A.; Conti, C.J.; Bradley, A. High susceptibility to ultraviolet-induced carcinogenesis in mice lacking XPC. Nature 1995, 377, 162–165. [Google Scholar] [CrossRef]
  74. Blanco, R.; Muñoz, P.; Flores, J.M.; Klatt, P.; Blasco, M.A. Telomerase abrogation dramatically accelerates TRF2-induced epithelial carcinogenesis. Genes Dev. 2007, 21, 206–220. [Google Scholar] [CrossRef] [Green Version]
  75. Martínez, P.; Blasco, M.A. Role of shelterin in cancer and aging. Aging Cell 2010, 9, 653–666. [Google Scholar] [CrossRef]
  76. Bejarano, L.; Bosso, G.; Louzame, J.; Serrano, R.; Gómez-Casero, E.; Martinez-Torrecuadrada, J.L.; Martínez, S.; Blanco-Aparicio, C.; Pastor, J.; Blasco, M.A. Multiple cancer pathways regulate telomere protection. EMBO Mol. Med. 2019, 11, 10292. [Google Scholar] [CrossRef] [PubMed]
  77. Biroccio, A.; Cherfils-Vicini, J.; Augereau, A.; Pinte, S.; Bauwens, S.; Ye, J.; Simonet, T.; Horard, B.; Jamet, K.; Cervera, L.; et al. TRF2 inhibits a cell-extrinsic pathway through which natural killer cells eliminate cancer cells. Nat. Cell Biol. 2013, 15, 818–828. [Google Scholar] [CrossRef] [PubMed]
  78. Picco, V.; Coste, I.; Giraud-Panis, M.J.; Renno, T.; Gilson, E.; Pages, G. Erk1/2/mapk pathway-dependent regulation of the telomeric factor TRF2. Oncotarget 2016, 7, 46615–46627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Martínez, P.; López, G.G.; Pisano, D.G.; Flores, J.M.; Blasco, M.A. A genetic interaction between RAP1 and telomerase reveals an unanticipated role for RAP1 in telomere maintenance. Aging Cell 2016, 15, 1113–1125. [Google Scholar] [CrossRef] [PubMed]
  80. Janoušková, E.; Nečasová, I.; Pavloušková, J.; Zimmermann, M.; Hluchý, M.; Marini, V.; Nováková, M.; Hofr, C. Human Rap1 modulates TRF2 attraction to telomeric DNA. Nucleic Acids Res. 2015, 43, 2691–2700. [Google Scholar] [CrossRef] [Green Version]
  81. 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 Rep. 2014, 9, 1273–1280. [Google Scholar] [CrossRef] [Green Version]
  82. Martinez, P.; Thanasoula, M.; Carlos, A.R.; Gómez-López, G.; Tejera, A.M.; Schoeftner, S.; Dominguez, O.; Pisano, D.G.; Tarsounas, M.; Blasco, M.A. Mammalian Rap1 controls telomere function and gene expression through binding to telomeric and extratelomeric sites. Nat. Cell Biol. 2010, 12, 768–780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Martínez, P.; López, G.G.; García, F.; Mercken, E.; Mitchell, S.; Flores, J.M.; de Cabo, R.; Blasco, M.A. RAP1 Protects from Obesity through Its Extratelomeric Role Regulating Gene Expression. Cell Rep. 2013, 3, 2059–2074. [Google Scholar] [CrossRef] [Green Version]
  84. Teo, H.; Ghosh, S.; Luesch, H.; Ghosh, A.; Wong, E.T.; Malik, N.; Orth, A.; de Jesus, P.; Perry, A.S.; Oliver, J.D.; et al. Telomere-independent rap1 is an ikk adaptor and regulates nf-kappab-dependent gene expression. Nat. Cell Biol. 2010, 12, 758–767. [Google Scholar] [CrossRef]
  85. Yang, Y.; Ye, C.; Wang, L.; An, G.; Tian, Z.; Meng, L.; Qu, L.; Lian, S.; Shou, C. Repressor activator protein 1–promoted colorectal cell migration is associated with the regulation of Vimentin. Tumor Biol. 2017, 39, 1010428317695034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Feuerhahn, S.; Chen, L.-Y.; Luke, B.; Porro, A. No DDRama at chromosome ends: TRF2 takes centre stage. Trends Biochem. Sci. 2015, 40, 275–285. [Google Scholar] [CrossRef] [PubMed]
  87. Kuimov, A.N. Polypeptide Components of Telomere Nucleoprotein Complex. Biochemistry 2004, 69, 117–129. [Google Scholar] [CrossRef]
  88. Kim, S.-H.; Kaminker, P.; Campisi, J. TIN2, a new regulator of telomere length in human cells. Nat. Genet. 1999, 23, 405–412. [Google Scholar] [CrossRef] [Green Version]
  89. Chiang, Y.J.; Kim, S.-H.; Tessarollo, L.; Campisi, J.; Hodes, R.J. Telomere-Associated Protein TIN2 Is Essential for Early Embryonic Development through a Telomerase-Independent Pathway. Mol. Cell. Biol. 2004, 24, 6631–6634. [Google Scholar] [CrossRef] [Green Version]
  90. Yamada, M.; Tsuji, N.; Nakamura, M.; Moriai, R.; Kobayashi, D.; Yagihashi, A.; Watanabe, N. Down-regulation of TRF1, TRF2 and TIN2 genes is important to maintain telomeric DNA for gastric cancers. Anticancer Res. 2003, 22, 3303–3307. [Google Scholar]
  91. Bellon, M.; Datta, A.; Brown, M.; Pouliquen, J.-F.; Couppie, P.; Kazanji, M.; Nicot, C. Increased expression of telomere length regulating factors TRF1, TRF2 and TIN2 in patients with adult T-cell leukemia. Int. J. Cancer 2006, 119, 2090–2097. [Google Scholar] [CrossRef]
  92. Oh, B.-K.; Kim, Y.-J.; Park, C.; Park, Y.N. Up-Regulation of Telomere-Binding Proteins, TRF1, TRF2, and TIN2 Is Related to Telomere Shortening during Human Multistep Hepatocarcinogenesis. Am. J. Pathol. 2005, 166, 73–80. [Google Scholar] [CrossRef] [Green Version]
  93. Baumann, P.; Price, C. Pot1 and telomere maintenance. FEBS Lett. 2010, 584, 3779–3784. [Google Scholar] [CrossRef] [Green Version]
  94. Loayza, D.; de Lange, T. POT1 as a terminal transducer of TRF1 telomere length control. Nature 2003, 423, 1013–1018. [Google Scholar] [CrossRef] [PubMed]
  95. Deng, Y.; Guo, X.; Ferguson, D.O.; Chang, S. Multiple roles for MRE11 at uncapped telomeres. Nature 2009, 460, 914–918. [Google Scholar] [CrossRef] [PubMed]
  96. Guo, X.; Deng, Y.; Lin, Y.; Cosme-Blanco, W.; Chan, S.; He, H.; Yuan, G.; Brown, E.J.; Chang, S. Dysfunctional telomeres activate an ATM-ATR-dependent DNA damage response to suppress tumorigenesis. EMBO J. 2007, 26, 4709–4719. [Google Scholar] [CrossRef] [PubMed]
  97. Kibe, T.; Zimmermann, M.; De Lange, T. TPP1 Blocks an ATR-Mediated Resection Mechanism at Telomeres. Mol. Cell 2017, 66, 300. [Google Scholar] [CrossRef] [Green Version]
  98. Abreu, E.; Aritonovska, E.; Reichenbach, P.; Cristofari, G.; Culp, B.; Terns, R.M.; Lingner, J.; Terns, M.P. TIN2-Tethered TPP1 Recruits Human Telomerase to Telomeres In Vivo. Mol. Cell. Biol. 2010, 30, 2971–2982. [Google Scholar] [CrossRef] [Green Version]
  99. Xin, H.; Liu, D.; Wan, M.; Safari, A.; Kim, H.; Sun, W.; O’Connor, M.S.; Songyang, Z. Tpp1 is a homologue of ciliate tebp-beta and interacts with pot1 to recruit telomerase. Nature 2007, 445, 559–562. [Google Scholar] [CrossRef]
  100. Wang, Y.; Wang, X.; Flores, E.R.; Yu, J.; Chang, S. Dysfunctional telomeres induce p53-dependent and independent apoptosis to compromise cellular proliferation and inhibit tumor formation. Aging Cell 2016, 15, 646–660. [Google Scholar] [CrossRef]
  101. Tejera, A.M.; Stagno d’Alcontres, M.; Thanasoula, M.; Marion, R.M.; Martinez, P.; Liao, C.; Flores, J.M.; Tarsounas, M.; Blasco, M.A. Tpp1 is required for tert recruitment, telomere elongation during nuclear reprogramming, and normal skin development in mice. Dev. Cell 2010, 18, 775–789. [Google Scholar] [CrossRef] [Green Version]
  102. Kibe, T.; Osawa, G.A.; Keegan, C.E.; de Lange, T. Telomere Protection by TPP1 Is Mediated by POT1a and POT1b. Mol. Cell. Biol. 2010, 30, 1059–1066. [Google Scholar] [CrossRef] [Green Version]
  103. Feldser, D.M.; Greider, C.W. Short Telomeres Limit Tumor Progression In Vivo by Inducing Senescence. Cancer Cell 2007, 11, 461–469. [Google Scholar] [CrossRef] [Green Version]
  104. Martínez, P.; Thanasoula, M.; Muñoz, P.; Liao, C.; Tejera, A.; McNees, C.; Flores, J.M.; Fernández-Capetillo, O.; Tarsounas, M.; Blasco, M.A. Increased telomere fragility and fusions resulting from TRF1 deficiency lead to degenerative pathologies and increased cancer in mice. Genes Dev. 2009, 23, 2060–2075. [Google Scholar] [CrossRef] [Green Version]
  105. Stout, G.J.; Blasco, M.A. Genetic dissection of the mechanisms underlying telomere-associated diseases: Impact of the TRF2 telomeric protein on mouse epidermal stem cells. Dis. Model. Mech. 2009, 2, 139–156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Denchi, E.L.; de Lange, T. Protection of telomeres through independent control of ATM and ATR by TRF2 and POT1. Nature 2007, 448, 1068–1071. [Google Scholar] [CrossRef] [PubMed]
  107. Glousker, G.; Briod, A.; Quadroni, M.; Lingner, J. Human shelterin protein POT 1 prevents severe telomere instability induced by homology-directed DNA repair. EMBO J. 2020, 39, e104500. [Google Scholar] [CrossRef] [PubMed]
  108. Hockemeyer, D.; Daniels, J.-P.; Takai, H.; De Lange, T. Recent Expansion of the Telomeric Complex in Rodents: Two Distinct POT1 Proteins Protect Mouse Telomeres. Cell 2006, 126, 63–77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Wu, L.; Multani, A.S.; He, H.; Cosme-Blanco, W.; Deng, Y.; Deng, J.M.; Bachilo, O.; Pathak, S.; Tahara, H.; Bailey, S.M.; et al. Pot1 Deficiency Initiates DNA Damage Checkpoint Activation and Aberrant Homologous Recombination at Telomeres. Cell 2006, 126, 49–62. [Google Scholar] [CrossRef] [Green Version]
  110. Nathan, V.; Johansson, P.A.; Palmer, J.M.; Hamilton, H.R.; Howlie, M.; Brooks, K.M.; Hayward, N.K.; Pritchard, A.L. A rare missense variant in protection of telomeres 1 (POT1) predisposes to a range of haematological malignancies. Br. J. Haematol. 2021, 192, e57–e60. [Google Scholar] [CrossRef]
  111. Lee, O.-H.; Kim, H.; He, Q.; Baek, H.J.; Yang, D.; Chen, L.-Y.; Liang, J.; Chae, H.K.; Safari, A.; Liu, D.; et al. Genome-wide YFP Fluorescence Complementation Screen Identifies New Regulators for Telomere Signaling in Human Cells. Mol. Cell. Proteom. 2011, 10, S1–S11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Gao, H.; Cervantes, R.B.; Mandell, E.K.; Otero, J.H.; Lundblad, V. RPA-like proteins mediate yeast telomere function. Nat. Struct. Mol. Biol. 2007, 14, 208–214. [Google Scholar] [CrossRef] [PubMed]
  113. Wellinger, R.J. The CST Complex and Telomere Maintenance: The Exception Becomes the Rule. Mol. Cell 2009, 36, 168–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Chen, L.-Y.; Lingner, J. CST for the grand finale of telomere replication. Nucleus 2013, 4, 277–282. [Google Scholar] [CrossRef] [Green Version]
  115. Martín, V.; Du, L.-L.; Rozenzhak, S.; Russell, P. Protection of telomeres by a conserved Stn1 Ten1 complex. Proc. Natl. Acad. Sci. USA 2007, 104, 14038–14043. [Google Scholar] [CrossRef] [Green Version]
  116. Chen, H.-W.; Xue, J.; Churikov, D.; Hass, E.P.; Shi, S.; Lemon, L.D.; Luciano, P.; Bertuch, A.A.; Zappulla, D.C.; Géli, V.; et al. Structural Insights into Yeast Telomerase Recruitment to Telomeres. Cell 2018, 172, 331–343.e13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Wold, M.S. Replication Protein A: A Heterotrimeric, Single-Stranded DNA-Binding Protein Required for Eukaryotic DNA Metabolism. Annu. Rev. Biochem. 1997, 66, 61–92. [Google Scholar] [CrossRef]
  118. Lim, C.J.; Barbour, A.T.; Zaug, A.J.; Goodrich, K.J.; McKay, A.E.; Wuttke, D.S.; Cech, T.R. The structure of human CST reveals a decameric assembly bound to telomeric DNA. Science 2020, 368, 1081–1085. [Google Scholar] [CrossRef]
  119. 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. Mol. Cell 2009, 36, 193–206. [Google Scholar] [CrossRef]
  120. Rice, C.; Skordalakes, E. Structure and function of the telomeric CST complex. Comput. Struct. Biotechnol. J. 2016, 14, 161–167. [Google Scholar] [CrossRef] [Green Version]
  121. Chastain, M.; Zhou, Q.; Shiva, O.; Fadri-Moskwik, M.; Whitmore, L.; Jia, P.; Dai, X.; Huang, C.; Ye, P.; Chai, W. Human CST Facilitates Genome-wide RAD51 Recruitment to GC-Rich Repetitive Sequences in Response to Replication Stress. Cell Rep. 2016, 16, 1300–1314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Price, C.; Boltz, K.A.; Chaiken, M.F.; Stewart, J.A.; Beilstein, M.A.; Shippen, D.E. Evolution of CST function in telomere maintenance. Cell Cycle 2010, 9, 3177–3185. [Google Scholar] [CrossRef] [Green Version]
  123. Stewart, J.A.; Wang, Y.; Ackerson, S.M.; Schuck, P.L. Emerging roles of cst in maintaining genome stability and human disease. Front. Biosci. 2018, 23, 1564–1586. [Google Scholar] [CrossRef] [Green Version]
  124. Maréchal, A.; Zou, L. RPA-coated single-stranded DNA as a platform for post-translational modifications in the DNA damage response. Cell Res. 2015, 25, 9–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Stewart, J.A.; Wang, F.; Chaiken, M.F.; Kasbek, C.; Chastain, P.D., 2nd; Wright, W.E.; Price, C.M. Human CST promotes telomere duplex replication and general replication restart after fork stalling. EMBO J. 2012, 31, 3537–3549. [Google Scholar] [CrossRef] [Green Version]
  126. Fanning, E.; Klimovich, V.; Nager, A.R. A dynamic model for replication protein A (RPA) function in DNA processing pathways. Nucleic Acids Res. 2006, 34, 4126–4137. [Google Scholar] [CrossRef] [Green Version]
  127. Flynn, R.L.; Zou, L. Oligonucleotide/oligosaccharide-binding fold proteins: A growing family of genome guardians. Crit. Rev. Biochem. Mol. Biol. 2010, 45, 266–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Bochkareva, E.; Korolev, S.; Lees-Miller, S.P.; Bochkarev, A. Structure of the RPA trimerization core and its role in the multistep DNA-binding mechanism of RPA. EMBO J. 2002, 21, 1855–1863. [Google Scholar] [CrossRef] [Green Version]
  129. Salas, T.R.; Petruseva, I.; Lavrik, O.; Saintomé, C. Evidence for direct contact between the RPA3 subunit of the human replication protein A and single-stranded DNA. Nucleic Acids Res. 2008, 37, 38–46. [Google Scholar] [CrossRef] [Green Version]
  130. Kim, C.; Paulus, B.F.; Wold, M. Interactions of human replication protein A with oligonucleotides. Biochemistry 1994, 33, 14197–14206. [Google Scholar] [CrossRef]
  131. Yates, L.A.; Aramayo, R.J.; Pokhrel, N.; Caldwell, C.; Kaplan, J.A.; Perera, R.; Spies, M.; Antony, E.; Zhang, X. A structural and dynamic model for the assembly of Replication Protein A on single-stranded DNA. Nat. Commun. 2018, 9, 1–14. [Google Scholar] [CrossRef]
  132. Theobald, D.L.; Mitton-Fry, R.M.; Wuttke, D.S. Nucleic Acid Recognition by OB-Fold Proteins. Annu. Rev. Biophys. Biomol. Struct. 2003, 32, 115–133. [Google Scholar] [CrossRef] [Green Version]
  133. Gu, P.; Jia, S.; Takasugi, T.; Smith, E.; Nandakumar, J.; Hendrickson, E.; Chang, S. CTC1-STN1 coordinates G- and C-strand synthesis to regulate telomere length. Aging Cell 2018, 17, e12783. [Google Scholar] [CrossRef]
  134. Gu, P.; Min, J.-N.; Wang, Y.; Huang, C.; Peng, T.; Chai, W.; Chang, S. CTC1 deletion results in defective telomere replication, leading to catastrophic telomere loss and stem cell exhaustion. EMBO J. 2012, 31, 2309–2321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Hom, R.A.; Wuttke, D.S. Human CST Prefers G-Rich but Not Necessarily Telomeric Sequences. Biochemistry 2017, 56, 4210–4218. [Google Scholar] [CrossRef] [PubMed]
  136. Huang, C.; Dai, X.; Chai, W. Human Stn1 protects telomere integrity by promoting efficient lagging-strand synthesis at telomeres and mediating C-strand fill-in. Cell Res. 2012, 22, 1681–1695. [Google Scholar] [CrossRef] [Green Version]
  137. Huang, C.; Jia, P.; Chastain, M.; Shiva, O.; Chai, W. The human ctc1/stn1/ten1 complex regulates telomere maintenance in alt cancer cells. Exp. Cell Res. 2017, 355, 95–104. [Google Scholar] [CrossRef] [Green Version]
  138. Feng, X.; Hsu, S.-J.; Bhattacharjee, A.; Wang, Y.; Diao, J.; Price, C.M. CTC1-STN1 terminates telomerase while STN1-TEN1 enables C-strand synthesis during telomere replication in colon cancer cells. Nat. Commun. 2018, 9, 2827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Bhattacharjee, A.; Wang, Y.; Diao, J.; Price, C.M. Dynamic DNA binding, junction recognition and G4 melting activity underlie the telomeric and genome-wide roles of human CST. Nucleic Acids Res. 2017, 45, 12311–12324. [Google Scholar] [CrossRef] [Green Version]
  140. Boltz, K.A.; Leehy, K.; Song, X.; Nelson, A.; Shippen, D.E. ATR cooperates with CTC1 and STN1 to maintain telomeres and genome integrity in Arabidopsis. Mol. Biol. Cell 2012, 23, 1558–1568. [Google Scholar] [CrossRef]
  141. Casteel, D.E.; Zhuang, S.; Zeng, Y.; Perrino, F.W.; Boss, G.R.; Goulian, M.; Pilz, R.B. A DNA polymerase-α·primase cofactor with homology to replication protein a-32 regulates DNA replication in mammalian cells. J. Biol. Chem. 2009, 284, 5807–5818. [Google Scholar] [CrossRef] [Green Version]
  142. Ganduri, S.; Lue, N.F. STN1–POLA2 interaction provides a basis for primase-pol α stimulation by human STN1. Nucleic Acids Res. 2017, 45, 9455–9466. [Google Scholar] [CrossRef] [Green Version]
  143. Wang, F.; Stewart, J.A.; Kasbek, C.; Zhao, Y.; Wright, W.E.; Price, C.M. Human CST Has Independent Functions during Telomere Duplex Replication and C-Strand Fill-In. Cell Rep. 2012, 2, 1096–1103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Chen, L.-Y.; Redon, S.; Lingner, J. The human CST complex is a terminator of telomerase activity. Nature 2012, 488, 540–544. [Google Scholar] [CrossRef]
  145. Wang, Y.; Brady, K.S.; Caiello, B.P.; Ackerson, S.M.; Stewart, J.A. Human cst suppresses origin licensing and promotes and-1/ctf4 chromatin association. Life Sci. Alliance 2019, 2. [Google Scholar] [CrossRef] [Green Version]
  146. Li, Y.; Xiao, H.; de Renty, C.; Jaramillo-Lambert, A.; Han, Z.; DePamphilis, M.L.; Brown, K.; Zhu, W. The Involvement of Acidic Nucleoplasmic DNA-binding Protein (And-1) in the Regulation of Prereplicative Complex (pre-RC) Assembly in Human Cells. J. Biol. Chem. 2012, 287, 42469–42479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Feng, X.; Hsu, S.-J.; Kasbek, C.; Chaiken, M.; Price, C.M. CTC1-mediated C-strand fill-in is an essential step in telomere length maintenance. Nucleic Acids Res. 2017, 45, 4281–4293. [Google Scholar] [CrossRef] [Green Version]
  148. Wu, P.; Takai, H.; de Lange, T. Telomeric 3′ Overhangs Derive from Resection by Exo1 and Apollo and Fill-In by POT1b-Associated CST. Cell 2012, 150, 39–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Feng, S.; Zhao, Y.; Xu, Y.; Ning, S.; Huo, W.; Hou, M.; Gao, G.; Ji, J.; Guo, R.; Xu, D. Ewing Tumor-associated Antigen 1 Interacts with Replication Protein A to Promote Restart of Stalled Replication Forks. J. Biol. Chem. 2016, 291, 21956–21962. [Google Scholar] [CrossRef] [Green Version]
  150. Zeman, M.K.; Cimprich, K.A. Causes and consequences of replication stress. Nat. Cell Biol. 2014, 16, 2–9. [Google Scholar] [CrossRef] [Green Version]
  151. Cimprich, K.A.; Cortez, D. ATR: An essential regulator of genome integrity. Nat. Rev. Mol. Cell Biol. 2008, 9, 616–627. [Google Scholar] [CrossRef] [Green Version]
  152. Patil, M.; Pabla, N.; Dong, Z. Checkpoint kinase 1 in DNA damage response and cell cycle regulation. Cell. Mol. Life Sci. 2013, 70, 4009–4021. [Google Scholar] [CrossRef] [Green Version]
  153. Guler, G.D.; Liu, H.; Vaithiyalingam, S.; Arnett, D.R.; Kremmer, E.; Chazin, W.J.; Fanning, E. Human DNA Helicase B (HDHB) Binds to Replication Protein A and Facilitates Cellular Recovery from Replication Stress. J. Biol. Chem. 2012, 287, 6469–6481. [Google Scholar] [CrossRef] [Green Version]
  154. Lee, M.; Shin, S.; Uhm, H.; Hong, H.; Kirk, J.; Hyun, K.; Kulikowicz, T.; Kim, J.; Ahn, B.; Bohr, V.A.; et al. Multiple RPAs make WRN syndrome protein a superhelicase. Nucleic Acids Res. 2018, 46, 4689–4698. [Google Scholar] [CrossRef]
  155. Qin, Z.; Bi, L.; Hou, X.-M.; Zhang, S.; Zhang, X.; Lu, Y.; Li, M.; Modesti, M.; Xi, X.-G.; Sun, B. Human RPA activates BLM’s bidirectional DNA unwinding from a nick. eLife 2020, 9, 9. [Google Scholar] [CrossRef] [PubMed]
  156. Bhat, K.; Bétous, R.; Cortez, D. High-affinity DNA-binding Domains of Replication Protein A (RPA) Direct SMARCAL1-dependent Replication Fork Remodeling. J. Biol. Chem. 2015, 290, 4110–4117. [Google Scholar] [CrossRef] [Green Version]
  157. Martínez-Jiménez, M.I.; Lahera, A.; Blanco, L. Human PrimPol activity is enhanced by RPA. Sci. Rep. 2017, 7, 783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Garvik, B.; Carson, M.; Hartwell, L. Single-stranded DNA arising at telomeres in cdc13 mutants may constitute a specific signal for the RAD9 checkpoint. Mol. Cell. Biol. 1995, 15, 6128–6138. [Google Scholar] [CrossRef] [Green Version]
  159. Bryan, C.; Rice, C.; Harkisheimer, M.; Schultz, D.C.; Skordalakes, E. Structure of the Human Telomeric Stn1-Ten1 Capping Complex. PLoS ONE 2013, 8, e66756. [Google Scholar] [CrossRef] [PubMed]
  160. Hughes, T.R.; Weilbaecher, R.G.; Walterscheid, M.; Lundblad, V. Identification of the single-strand telomeric DNA binding domain of the Saccharomyces cerevisiae Cdc13 protein. Proc. Natl. Acad. Sci. USA 2000, 97, 6457–6462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Lewis, K.A.; Pfaff, D.A.; Earley, J.N.; Altschuler, S.E.; Wuttke, D.S. The tenacious recognition of yeast telomere sequence by Cdc13 is fully exerted by a single OB-fold domain. Nucleic Acids Res. 2014, 42, 475–484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Sun, J.; Yang, Y.; Wan, K.; Mao, N.; Yu, T.-Y.; Lin, Y.-C.; DeZwaan, D.C.; Freeman, B.C.; Lin, J.-J.; Lue, N.F.; et al. Structural bases of dimerization of yeast telomere protein Cdc13 and its interaction with the catalytic subunit of DNA polymerase α. Cell Res. 2010, 21, 258–274. [Google Scholar] [CrossRef] [Green Version]
  163. Ge, Y.; Wu, Z.; Chen, H.; Zhong, Q.; Shi, S.; Li, G.; Wu, J.; Lei, M. Structural insights into telomere protection and homeostasis regulation by yeast CST complex. Nat. Struct. Mol. Biol. 2020, 27, 752–762. [Google Scholar] [CrossRef] [PubMed]
  164. Li, S.; Makovets, S.; Matsuguchi, T.; Blethrow, J.D.; Shokat, K.M.; Blackburn, E.H. Cdk1-Dependent Phosphorylation of Cdc13 Coordinates Telomere Elongation during Cell-Cycle Progression. Cell 2009, 136, 50–61. [Google Scholar] [CrossRef] [Green Version]
  165. Liu, C.-C.; Gopalakrishnan, V.; Poon, L.-F.; Yan, T.; Li, S. Cdk1 Regulates the Temporal Recruitment of Telomerase and Cdc13-Stn1-Ten1 Complex for Telomere Replication. Mol. Cell. Biol. 2014, 34, 57–70. [Google Scholar] [CrossRef] [Green Version]
  166. Tseng, S.-F.; Shen, Z.-J.; Tsai, H.-J.; Lin, Y.-H.; Teng, S.-C. Rapid Cdc13 turnover and telomere length homeostasis are controlled by Cdk1-mediated phosphorylation of Cdc13. Nucleic Acids Res. 2009, 37, 3602–3611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Wu, Y.; DiMaggio, P.A., Jr.; Perlman, D.H.; Zakian, V.A.; Garcia, B.A. Novel Phosphorylation Sites in the S. cerevisiae Cdc13 Protein Reveal New Targets for Telomere Length Regulation. J. Proteome Res. 2013, 12, 316–327. [Google Scholar] [CrossRef] [Green Version]
  168. Zhang, W.; Durocher, D. De novo telomere formation is suppressed by the Mec1-dependent inhibition of Cdc13 accumulation at DNA breaks. Genes Dev. 2010, 24, 502–515. [Google Scholar] [CrossRef] [Green Version]
  169. Poncet, D.; Belleville, A.; de Roodenbeke, C.T.; de Climens, A.R.; Ben Simon, E.; Merle-Beral, H.; Callet-Bauchu, E.; Salles, G.; Sabatier, L.; Delic, J.; et al. Changes in the expression of telomere maintenance genes suggest global telomere dysfunction in B-chronic lymphocytic leukemia. Blood 2008, 111, 2388–2391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Chandra, A.; Hughes, T.R.; Nugent, C.I.; Lundblad, V. Cdc13 both positively and negatively regulates telomere replication. Genes Dev. 2001, 15, 404–414. [Google Scholar] [CrossRef] [Green Version]
  171. Evans, S.K.; Lundblad, V. Positive and negative regulation of telomerase access to the telomere. J. Cell Sci. 2000, 113, 3357–3364. [Google Scholar] [CrossRef]
  172. Zhong, F.; Batista, L.; Freund, A.; Pech, M.F.; Venteicher, A.; Artandi, S.E. TPP1 OB-Fold Domain Controls Telomere Maintenance by Recruiting Telomerase to Chromosome Ends. Cell 2012, 150, 481–494. [Google Scholar] [CrossRef] [Green Version]
  173. Giraud-Panis, M.-J.; Teixeira, M.T.; Géli, V.; Gilson, E. CST Meets Shelterin to Keep Telomeres in Check. Mol. Cell 2010, 39, 665–676. [Google Scholar] [CrossRef] [PubMed]
  174. Langston, R.E.; Palazzola, D.; Bonnell, E.; Wellinger, R.J.; Weinert, T. Loss of Cdc13 causes genome instability by a deficiency in replication-dependent telomere capping. PLoS Genet. 2020, 16, e1008733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Han, E.; Patel, N.A.; Yannuzzi, N.A.; Laura, D.M.; Fan, K.C.; Negron, C.I.; Prakhunhungsit, S.; Thorson, W.L.; Berrocal, A.M. A unique case of coats plus syndrome and dyskeratosis congenita in a patient with CTC1 mutations. Ophthalmic Genet. 2020, 41, 363–367. [Google Scholar] [CrossRef]
  176. Bs, R.B.K.; Bs, K.E.G.; Usmani, G.N.; Asdourian, G.K.; Williams, D.A.; Hofmann, I.; Agarwal, S. CTC1 Mutations in a patient with dyskeratosis congenita. Pediatr. Blood Cancer 2012, 59, 311–314. [Google Scholar] [CrossRef] [Green Version]
  177. Dai, X.; Huang, C.; Bhusari, A.; Sampathi, S.; Schubert, K.; Chai, W. Molecular steps of G-overhang generation at human telomeres and its function in chromosome end protection. EMBO J. 2010, 29, 2788–2801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Wang, Y.; Chai, W. Pathogenic CTC1 mutations cause global genome instabilities under replication stress. Nucleic Acids Res. 2018, 46, 3981–3992. [Google Scholar] [CrossRef] [PubMed]
  179. Anderson, B.H.; Kasher, P.; Mayer, J.; Szynkiewicz, M.; Jenkinson, E.M.; Bhaskar, S.S.; Urquhart, J.; Daly, S.B.; Dickerson, J.E.; O’Sullivan, J.; et al. Mutations in CTC1, encoding conserved telomere maintenance component 1, cause Coats plus. Nat. Genet. 2012, 44, 338–342. [Google Scholar] [CrossRef]
  180. Polvi, A.; Linnankivi, T.; Kivelä, T.; Herva, R.; Keating, J.P.; Mäkitie, O.; Pareyson, D.; Vainionpää, L.; Lahtinen, J.; Hovatta, I.; et al. Mutations in CTC1, Encoding the CTS Telomere Maintenance Complex Component 1, Cause Cerebroretinal Microangiopathy with Calcifications and Cysts. Am. J. Hum. Genet. 2012, 90, 540–549. [Google Scholar] [CrossRef] [Green Version]
  181. Walne, A.J.; Bhagat, T.; Kirwan, M.; Gitiaux, C.; Desguerre, I.; Leonard, N.; Nogales, E.; Vulliamy, T.; Dokal, I.S. Mutations in the telomere capping complex in bone marrow failure and related syndromes. Haematologica 2013, 98, 334–338. [Google Scholar] [CrossRef] [Green Version]
  182. Goulian, M.; Heard, C.J. The mechanism of action of an accessory protein for DNA polymerase alpha/primase. J. Biol. Chem. 1990, 265, 13231–13239. [Google Scholar] [CrossRef]
  183. Grossi, S.; Puglisi, A.; Dmitriev, P.V.; Lopes, M.; Shore, D. Pol12, the B subunit of DNA polymerase α, functions in both telomere capping and length regulation. Genes Dev. 2004, 18, 992–1006. [Google Scholar] [CrossRef] [Green Version]
  184. Petreaca, R.C.; Chiu, H.-C.; Eckelhoefer, H.A.; Chuang, C.; Xu, L.; Nugent, C.I. Chromosome end protection plasticity revealed by Stn1p and Ten1p bypass of Cdc13p. Nat. Cell Biol. 2006, 8, 748–755. [Google Scholar] [CrossRef] [PubMed]
  185. Pennock, E.; Buckley, K.; Lundblad, V. Cdc13 Delivers Separate Complexes to the Telomere for End Protection and Replication. Cell 2001, 104, 387–396. [Google Scholar] [CrossRef] [Green Version]
  186. Mason, M.; Wanat, J.J.; Harper, S.; Schultz, D.C.; Speicher, D.W.; Johnson, F.B.; Skordalakes, E. Cdc13 OB2 Dimerization Required for Productive Stn1 Binding and Efficient Telomere Maintenance. Struct. 2013, 21, 109–120. [Google Scholar] [CrossRef] [Green Version]
  187. Mitchell, M.T.; Smith, J.S.; Mason, M.; Harper, S.; Speicher, D.W.; Johnson, F.B.; Skordalakes, E. Cdc13 N-terminal Dimerization, DNA Binding, and Telomere Length Regulation. Mol Cell Biol 2010, 30, 5325–5334. [Google Scholar] [CrossRef] [Green Version]
  188. Lue, N.F.; Chan, J.; Wright, W.E.; Hurwitz, J. The CDC13-STN1-TEN1 Complex Stimulates Pol Alpha Activity by Promoting RNA Priming and Primase-to-Polymerase Switch. Nat. Commun. 2014, 5, 5762. [Google Scholar] [CrossRef] [Green Version]
  189. Bhattacharjee, A.; Stewart, J.; Chaiken, M.; Price, C.M. STN1 OB Fold Mutation Alters DNA Binding and Affects Selective Aspects of CST Function. PLoS Genet. 2016, 12, e1006342. [Google Scholar] [CrossRef] [Green Version]
  190. Diotti, R.; Kalan, S.; Matveyenko, A.; Loayza, D. DNA-Directed Polymerase Subunits Play a Vital Role in Human Telomeric Overhang Processing. Mol. Cancer Res. 2014, 13, 402–410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Surovtseva, Y.V.; Churikov, D.; Boltz, K.A.; Song, X.; Lamb, J.C.; Warrington, R.; Leehy, K.; Heacock, M.; Price, C.M.; Shippen, D.E. Conserved Telomere Maintenance Component 1 Interacts with STN1 and Maintains Chromosome Ends in Higher Eukaryotes. Mol. Cell 2009, 36, 207–218. [Google Scholar] [CrossRef] [Green Version]
  192. Passi, G.R.; Shamim, U.; Rathore, S.; Joshi, A.; Mathur, A.; Parveen, S.; Sharma, P.; Crow, Y.J.; Faruq, M. An Indian child with Coats plus syndrome due to mutations in STN1. Am. J. Med. Genet. Part A 2020, 182, 2139–2144. [Google Scholar] [CrossRef]
  193. Simon, A.J.; Lev, A.; Zhang, Y.; Weiss, B.; Rylova, A.; Eyal, E.; Kol, N.; Barel, O.; Cesarkas, K.; Soudack, M.; et al. Mutations in STN1 cause Coats plus syndrome and are associated with genomic and telomere defects. J. Exp. Med. 2016, 213, 1429–1440. [Google Scholar] [CrossRef] [PubMed]
  194. Qian, W.; Wang, J.; Jin, N.-N.; Fu, X.-H.; Lin, Y.-C.; Lin, J.-J.; Zhou, J.-Q. Ten1p promotes the telomeric DNA-binding activity of Cdc13p: Implication for its function in telomere length regulation. Cell Res. 2009, 19, 849–863. [Google Scholar] [CrossRef] [Green Version]
  195. Wang, F.; Stewart, J.; Price, C.M. Human cst abundance determines recovery from diverse forms of DNA damage and replication stress. Cell Cycle 2014, 13, 3488–3498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Bagcchi, S. POT1: A genetic link for familial glioma. Lancet Oncol. 2015, 16, e12. [Google Scholar] [CrossRef]
  197. Robles-Espinoza, C.D.; Harland, M.; Ramsay, A.J.; Aoude, L.G.; Quesada, V.; Ding, Z.; Pooley, K.A.; Pritchard, A.L.; Tiffen, J.C.; Petljak, M.; et al. POT1 loss-of-function variants predispose to familial melanoma. Nat. Genet. 2014, 46, 478–481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Ramsay, A.J.; Quesada, V.; Foronda, M.; Conde, L.; Martínez-Trillos, A.; Villamor, N.; Rodríguez, D.; Kwarciak, A.; Garabaya, C.; Gallardo, M.; et al. POT1 mutations cause telomere dysfunction in chronic lymphocytic leukemia. Nat. Genet. 2013, 45, 526–530. [Google Scholar] [CrossRef]
  199. Calvete, O.; Martinez, P.; Garcia-Pavia, P.; Benitez-Buelga, C.; Paumard-Hernandez, B.; Fernandez, V.; Dominguez, F.; Salas, C.; Romero-Laorden, N.; Garcia-Donas, J.; et al. A mutation in the POT1 gene is responsible for cardiac angiosarcoma in TP53-negative Li–Fraumeni-like families. Nat. Commun. 2015, 6, 8383. [Google Scholar] [CrossRef] [Green Version]
  200. Shen, J.; Gammon, M.D.; Wu, H.-C.; Terry, M.B.; Wang, Q.; Bradshaw, P.T.; Teitelbaum, S.L.; Neugut, A.I.; Santella, R.M. Multiple Genetic Variants in Telomere Pathway Genes and Breast Cancer Risk. Cancer Epidemiol. Biomark. Prev. 2010, 19, 219–228. [Google Scholar] [CrossRef] [Green Version]
  201. Richard, M.A.; Lupo, P.J.; Morton, L.M.; Yasui, Y.A.; Sapkota, Y.A.; Arnold, M.A.; Aubert, G.; Neglia, J.P.; Turcotte, L.M.; Leisenring, W.M.; et al. Genetic variation in POT1 and risk of thyroid subsequent malignant neoplasm: A report from the Childhood Cancer Survivor Study. PLoS ONE 2020, 15, e0228887. [Google Scholar] [CrossRef] [Green Version]
  202. Hiyama, E.; Hiyama, K. Telomere and telomerase in stem cells. Br. J. Cancer 2007, 96, 1020–1024. [Google Scholar] [CrossRef] [Green Version]
  203. Shay, J.W.; Wright, W.E. Telomeres and telomerase: Three decades of progress. Nat. Rev. Genet. 2019, 20, 299–309. [Google Scholar] [CrossRef]
  204. Meier, U.T. The many facets of h/aca ribonucleoproteins. Chromosoma 2005, 114, 1–14. [Google Scholar] [CrossRef] [Green Version]
  205. Trybek, T.; Kowalik, A.; Góźdź, S.; Kowalska, A. Telomeres and telomerase in oncogenesis. Oncol. Lett. 2020, 20, 1015–1027. [Google Scholar] [CrossRef] [PubMed]
  206. Girard, J.P.; Lehtonen, H.; Caizergues-Ferrer, M.; Amalric, F.; Tollervey, D.; Lapeyre, B. Gar1 is an essential small nucleolar rnp protein required for pre-rrna processing in yeast. EMBO J. 1992, 11, 673–682. [Google Scholar] [CrossRef]
  207. Holt, S.E.; Aisner, D.L.; Baur, J.; Tesmer, V.M.; Dy, M.; Ouellette, M.; Trager, J.B.; Morin, G.B.; Toft, D.O.; Shay, J.W.; et al. Functional requirement of p23 and Hsp90 in telomerase complexes. Genes Dev. 1999, 13, 817–826. [Google Scholar] [CrossRef]
  208. Zhong, F.; Savage, S.A.; Shkreli, M.; Giri, N.; Jessop, L.; Myers, T.; Chen, R.; Alter, B.P.; Artandi, S.E. Disruption of telomerase trafficking by TCAB1 mutation causes dyskeratosis congenita. Genes Dev. 2011, 25, 11–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Venteicher, A.; Artandi, S.E. TCAB1: Driving telomerase to Cajal bodies. Cell Cycle 2009, 8, 1329–1331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Mao, Y.-Q.; Houry, W.A. The Role of Pontin and Reptin in Cellular Physiology and Cancer Etiology. Front. Mol. Biosci. 2017, 4, 58. [Google Scholar] [CrossRef] [Green Version]
  211. Lee, J.H.; Jeong, S.A.; Khadka, P.; Hong, J.; Chung, I.K. Involvement of SRSF11 in cell cycle-specific recruitment of telomerase to telomeres at nuclear speckles. Nucleic Acids Res. 2015, 43, 8435–8451. [Google Scholar] [CrossRef] [Green Version]
  212. Schmidt, J.C.; Cech, T.R. Human telomerase: Biogenesis, trafficking, recruitment, and activation. Genes Dev. 2015, 29, 1095–1105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Vinagre, J.; Almeida, A.; Pópulo, 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. Nat. Commun. 2013, 4, 2185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Wong, M.S.; Wright, W.E.; Shay, J.W. Alternative splicing regulation of telomerase: A new paradigm? Trends Genet. 2014, 30, 430–438. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Gobbini, E.; Trovesi, C.; Cassani, C.; Longhese, M.P. Telomere uncapping at the crossroad between cell cycle arrest and carcinogenesis. Mol. Cell. Oncol. 2014, 1, e29901. [Google Scholar] [CrossRef] [Green Version]
  216. Bornstein-Quevedo, L.; García-Hernández, M.L.; Camacho-Arroyo, I.; Herrera, M.F.; Angeles, A.A.; Treviño, O.G.; Gamboa-Domínguez, A. Telomerase Activity in Well-Differentiated Papillary Thyroid Carcinoma Correlates with Advanced Clinical Stage of the Disease. Endocr. Pathol. 2003, 14, 213–220. [Google Scholar] [CrossRef]
  217. Wang, N.; Xu, D.; Sofiadis, A.; Höög, A.; Vukojević, V.; Bäckdahl, M.; Zedenius, J.; Larsson, C. Telomerase-dependent and independent telomere maintenance and its clinical implications in medullary thyroid carcinoma. J. Clin. Endocrinol. Metab. 2014, 99, E1571–E1579. [Google Scholar] [CrossRef] [Green Version]
  218. Fernandez-Marcelo, T.; Gómez, A.; Pascua, I.; De Juan, C.; Head, J.; Hernando, F.; Jarabo, J.-R.; Calatayud, J.; Torres-García, A.-J.; Iniesta, P. Telomere length and telomerase activity in non-small cell lung cancer prognosis: Clinical usefulness of a specific telomere status. J. Exp. Clin. Cancer Res. 2015, 34, 78. [Google Scholar] [CrossRef] [Green Version]
  219. Satyanarayana, A.; Manns, M.P.; Rudolph, K.L. Telomeres and telomerase: A dual role in hepatocarcinogenesis. Hepatology 2004, 40, 276–283. [Google Scholar] [CrossRef]
  220. Molano, M.; Martín, D.C.; Moreno-Acosta, P.; Hernández, G.; Cornall, A.; Buitrago, O.; Gamboa, O.; Garland, S.; Tabrizi, S.; Muñoz, N. Telomerase activity in cervical scrapes of women with high-grade cervical disease: A nested case-control study. Oncol. Lett. 2017, 15, 354–360. [Google Scholar] [CrossRef] [Green Version]
  221. Phillips, H.S.; Kharbanda, S.; Chen, R.; Forrest, W.F.; Soriano, R.H.; Wu, T.D.; Misra, A.; Nigro, J.M.; Colman, H.; Soroceanu, L.; et al. Molecular subclasses of high-grade glioma predict prognosis, delineate a pattern of disease progression, and resemble stages in neurogenesis. Cancer Cell 2006, 9, 157–173. [Google Scholar] [CrossRef] [Green Version]
  222. Bodnar, A.G.; Ouellette, M.; Frolkis, M.; Holt, S.E.; Chiu, C.-P.; Morin, G.B.; Harley, C.B.; Shay, J.W.; Lichtsteiner, S.; Wright, W.E. Extension of Life-Span by Introduction of Telomerase into Normal Human Cells. Science 1998, 279, 349–352. [Google Scholar] [CrossRef] [Green Version]
  223. Castelo-Branco, P.; Choufani, S.; Mack, S.C.; Gallagher, D.; Zhang, C.; Lipman, T.; Zhukova, N.; Walker, E.J.; Martin, D.; Merino, D.; et al. Methylation of the TERT promoter and risk stratification of childhood brain tumours: An integrative genomic and molecular study. Lancet Oncol. 2013, 14, 534–542. [Google Scholar] [CrossRef]
  224. Kim, N.W.; Piatyszek, M.A.; Prowse, K.R.; Harley, C.B.; West, M.D.; Ho, P.L.; Coviello, G.M.; Wright, W.E.; Weinrich, S.L.; Shay, J.W. Specific association of human telomerase activity with immortal cells and cancer. Science 1994, 266, 2011–2015. [Google Scholar] [CrossRef]
  225. Barthel, F.P.; Wei, W.; Tang, M.; Martinez-Ledesma, E.; Hu, X.; Amin, S.B.; Akdemir, K.C.; Seth, S.; Song, X.; Wang, Q.; et al. Systematic analysis of telomere length and somatic alterations in 31 cancer types. Nat. Genet. 2017, 49, 349–357. [Google Scholar] [CrossRef]
  226. Bernardes de Jesus, B.; Blasco, M.A. Telomerase at the intersection of cancer and aging. Trends Genet. 2013, 29, 513–520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Xu, Y.; Goldkorn, A. Telomere and Telomerase Therapeutics in Cancer. Genes 2016, 7, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Cong, Y.-S.; Wen, J.; Bacchetti, S. The human telomerase catalytic subunit hTERT: Organization of the gene and characterization of the promoter. Hum. Mol. Genet. 1999, 8, 137–142. [Google Scholar] [CrossRef] [PubMed]
  229. Horn, S.; Figl, A.; Rachakonda, P.S.; 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Prasad, R.; Pal, D.; Mohammad, W. Therapeutic Targets in Telomerase and Telomere Biology of Cancers. Indian J. Clin. Biochem. 2020, 35, 135–146. [Google Scholar] [CrossRef]
  231. Liu, T.; Yuan, X.; Xu, D. Cancer-Specific Telomerase Reverse Transcriptase (TERT) Promoter Mutations: Biological and Clinical Implications. Genes 2016, 7, 38. [Google Scholar] [CrossRef] [PubMed]
  232. Yuan, X.; Larsson, C.; Xu, D. Mechanisms underlying the activation of TERT transcription and telomerase activity in human cancer: Old actors and new players. Oncogene 2019, 38, 6172–6183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Bell, R.J.A.; Rube, H.T.; Kreig, A.; Mancini, A.; Fouse, S.D.; Nagarajan, R.P.; Choi, S.; Hong, C.; He, D.; Pekmezci, M.; et al. The transcription factor GABP selectively binds and activates the mutant TERT promoter in cancer. Science 2015, 348, 1036–1039. [Google Scholar] [CrossRef] [Green Version]
  234. Mancini, A.; Xavier-Magalhaes, A.; Woods, W.S.; Nguyen, K.T.; Amen, A.M.; Hayes, J.L.; Fellmann, C.; Gapinske, M.; McKinney, A.M.; Hong, C.; et al. Disruption of the beta1l isoform of gabp reverses glioblastoma replicative immortality in a tert promoter mutation-dependent manner. Cancer Cell 2018, 34, 513–528.e518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Yuan, X.; Mu, N.; Wang, N.; Strååt, K.; Sofiadis, A.; Guo, Y.; Stenman, A.; Li, K.; Cheng, G.; Zhang, L.; et al. GABPA inhibits invasion/metastasis in papillary thyroid carcinoma by regulating DICER1 expression. Oncogene 2018, 38, 965–979. [Google Scholar] [CrossRef]
  236. Chiba, K.; Johnson, J.Z.; Vogan, J.M.; Wagner, T.; Boyle, J.M.; Hockemeyer, D. Cancer-associated TERT promoter mutations abrogate telomerase silencing. eLife 2015, 4, e07918. [Google Scholar] [CrossRef] [PubMed]
  237. Li, C.; Wu, S.; Wang, H.; Bi, X.; Yang, Z.; Du, Y.; He, L.; Cai, Z.; Wang, J.; Fan, Z. The C228T mutation of TERT promoter frequently occurs in bladder cancer stem cells and contributes to tumorigenesis of bladder cancer. Oncotarget 2015, 6, 19542–19551. [Google Scholar] [CrossRef] [Green Version]
  238. Wang, M.-J.; Lin, Y.-C.; Pang, T.-L.; Lee, J.-M.; Chou, C.-C.; Lin, J.-J. Telomere-binding and Stn1p-interacting activities are required for the essential function of Saccharomyces cerevisiae Cdc13p. Nucleic Acids Res. 2000, 28, 4733–4741. [Google Scholar] [CrossRef] [Green Version]
  239. Bougel, S.; Lhermitte, B.; Gallagher, G.; De Flaugergues, J.-C.; Janzer, R.C.; Benhattar, J. Methylation of the hTERT Promoter: A Novel Cancer Biomarker for Leptomeningeal Metastasis Detection in Cerebrospinal Fluids. Clin. Cancer Res. 2013, 19, 2216–2223. [Google Scholar] [CrossRef] [Green Version]
  240. Fernandes, S.G.; Dsouza, R.; Pandya, G.; Kirtonia, A.; Tergaonkar, V.; Lee, S.Y.; Garg, M.; Khattar, E. Role of Telomeres and Telomeric Proteins in Human Malignancies and Their Therapeutic Potential. Cancers 2020, 12, 1901. [Google Scholar] [CrossRef] [PubMed]
  241. Kim, W.; Ludlow, A.T.; Min, J.; Robin, J.D.; Stadler, G.; Mender, I.; Lai, T.-P.; Zhang, N.; Wright, W.E.; Shay, J.W. Regulation of the Human Telomerase Gene TERT by Telomere Position Effect-Over Long Distances (TPE-OLD): Implications for Aging and Cancer. PLoS Biol. 2016, 14, e2000016. [Google Scholar] [CrossRef] [Green Version]
  242. Li, Y.; Liu, L.; Tollefsbol, T.O. Glucose restriction can extend normal cell lifespan and impair precancerous cell growth through epigenetic control of hTERT and p16 expression. FASEB J. 2010, 24, 1442–1453. [Google Scholar] [CrossRef] [Green Version]
  243. Xu, D.; Dwyer, J.; Li, H.; Duan, W.; Liu, J.-P. Ets2 Maintains hTERT Gene Expression and Breast Cancer Cell Proliferation by Interacting with c-Myc. J. Biol. Chem. 2008, 283, 23567–23580. [Google Scholar] [CrossRef] [Green Version]
  244. Kang, S.S.; Kwon, T.; Kwon, D.Y.; Do, S.I. Akt Protein Kinase Enhances Human Telomerase Activity through Phosphorylation of Telomerase Reverse Transcriptase Subunit. J. Biol. Chem. 1999, 274, 13085–13090. [Google Scholar] [CrossRef] [Green Version]
  245. Kimura, A.; Ohmichi, M.; Kawagoe, J.; Kyo, S.; Mabuchi, S.; Takahashi, T.; Ohshima, C.; Arimoto-Ishida, E.; Nishio, Y.; Inoue, M.; et al. Induction of hTERT expression and phosphorylation by estrogen via Akt cascade in human ovarian cancer cell lines. Oncogene 2004, 23, 4505–4515. [Google Scholar] [CrossRef] [Green Version]
  246. Yuan, X.; Cheng, G.; Yu, J.; Zheng, S.; Sun, C.; Sun, Q.; Li, K.; Lin, Z.; Liu, T.; Li, P.; et al. The TERT promoter mutation incidence is modified by germline TERT rs2736098 and rs2736100 polymorphisms in hepatocellular carcinoma. Oncotarget 2017, 8, 23120–23129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Dahlström, J.; Liu, T.; Yuan, X.; Saft, L.; Ghaderi, M.; Bin Wei, Y.; Lavebratt, C.; Li, P.; Zheng, C.; Björkholm, M.; et al. TERT rs2736100 genotypes are associated with differential risk of myeloproliferative neoplasms in Swedish and Chinese male patient populations. Ann. Hematol. 2016, 95, 1825–1832. [Google Scholar] [CrossRef] [Green Version]
  248. Blasco, M.A. The epigenetic regulation of mammalian telomeres. Nat. Rev. Genet. 2007, 8, 299–309. [Google Scholar] [CrossRef] [PubMed]
  249. Stern, J.L.; Theodorescu, D.; Vogelstein, B.; Papadopoulos, N.; Cech, T.R. Mutation of the TERT promoter, switch to active chromatin, and monoallelic TERT expression in multiple cancers. Genes Dev. 2015, 29, 2219–2224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Scheibe, M.; Arnoult, N.; Kappei, D.; Buchholz, F.; Decottignies, A.; Butter, F.; Mann, M. Quantitative interaction screen of telomeric repeat-containing RNA reveals novel TERRA regulators. Genome Res. 2013, 23, 2149–2157. [Google Scholar] [CrossRef] [Green Version]
  251. Biffi, G.; Tannahill, D.; Balasubramanian, S. An Intramolecular G-Quadruplex Structure Is Required for Binding of Telomeric Repeat-Containing RNA to the Telomeric Protein TRF2. J. Am. Chem. Soc. 2012, 134, 11974–11976. [Google Scholar] [CrossRef]
  252. Chow, T.T.; Shi, X.; Wei, J.-H.; Guan, J.; Stadler, G.; Huang, B.; Blackburn, E.H. Local enrichment of HP1alpha at telomeres alters their structure and regulation of telomere protection. Nat. Commun. 2018, 9, 3583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Pfeiffer, V.; Lingner, J. TERRA Promotes Telomere Shortening through Exonuclease 1–Mediated Resection of Chromosome Ends. PLoS Genet. 2012, 8, e1002747. [Google Scholar] [CrossRef] [Green Version]
  254. Wang, C.; Zhao, L.; Lu, S. Role of TERRA in the Regulation of Telomere Length. Int. J. Biol. Sci. 2015, 11, 316–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Lovejoy, C.A.; Li, W.; Reisenweber, S.; Thongthip, S.; Bruno, J.; De Lange, T.; De, S.; Petrini, J.H.; Sung, P.A.; Jasin, M.; et al. Loss of ATRX, Genome Instability, and an Altered DNA Damage Response Are Hallmarks of the Alternative Lengthening of Telomeres Pathway. PLoS Genet. 2012, 8, e1002772. [Google Scholar] [CrossRef] [PubMed]
  256. Sampl, S.; Pramhas, S.; Stern, C.; Preusser, M.; Marosi, C.; Holzmann, K. Expression of Telomeres in Astrocytoma WHO Grade 2 to 4: TERRA Level Correlates with Telomere Length, Telomerase Activity, and Advanced Clinical Grade. Transl. Oncol. 2012, 5, 56–65. [Google Scholar] [CrossRef] [Green Version]
  257. Luke, B.; Panza, A.; Redon, S.; Iglesias, N.; Li, Z.; Lingner, J. The Rat1p 5′ to 3′ Exonuclease Degrades Telomeric Repeat-Containing RNA and Promotes Telomere Elongation in Saccharomyces cerevisiae. Mol. Cell 2008, 32, 465–477. [Google Scholar] [CrossRef]
  258. Xu, Y.; Kimura, T.; Komiyama, M. Human telomere RNA and DNA form an intermolecular G-quadruplex. Nucleic Acids Symp. Ser. 2008, 52, 169–170. [Google Scholar] [CrossRef] [Green Version]
  259. Lucibello, F.; Menegatti, S.; Menger, L. Methods to edit T cells for cancer immunotherapy. Methods Enzymol. 2020, 631, 107–135. [Google Scholar] [CrossRef]
  260. Borah, S.; Xi, L.; Zaug, A.J.; Powell, N.M.; Dancik, G.M.; Cohen, S.B.; Costello, J.C.; Theodorescu, D.; Cech, T.R. TERTpromoter mutations and telomerase reactivation in urothelial cancer. Science 2015, 347, 1006–1010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Inderberg-Suso, E.-M.; Trachsel, S.; Lislerud, K.; Rasmussen, A.-M.; Gaudernack, G. Widespread CD4+ T-cell reactivity to novel hTERT epitopes following vaccination of cancer patients with a single hTERT peptide GV1001. OncoImmunology 2012, 1, 670–686. [Google Scholar] [CrossRef] [Green Version]
  262. Brunsvig, P.F.; Aamdal, S.; Gjertsen, M.K.; Kvalheim, G.; Markowski-Grimsrud, C.J.; Sve, I.; Dyrhaug, M.; Trachsel, S.; Møller, M.; Eriksen, J.A.; et al. Telomerase peptide vaccination: A phase I/II study in patients with non-small cell lung cancer. Cancer Immunol. Immunother. 2006, 55, 1553–1564. [Google Scholar] [CrossRef]
  263. Greten, T.F.; Forner, A.; Korangy, F.; N’Kontchou, G.; Barget, N.; Ayuso, C.; Ormandy, L.A.; Manns, M.P.; Beaugrand, M.; Bruix, J. A phase II open label trial evaluating safety and efficacy of a telomerase peptide vaccination in patients with advanced hepatocellular carcinoma. BMC Cancer 2010, 10, 209. [Google Scholar] [CrossRef] [Green Version]
  264. Hunger, R.E.; Lang, K.K.; Markowski, C.J.; Trachsel, S.; Møller, M.; Eriksen, J.A.; Rasmussen, A.-M.; Braathen, L.R.; Gaudernack, G. Vaccination of patients with cutaneous melanoma with telomerase-specific peptides. Cancer Immunol. Immunother. 2011, 60, 1553–1564. [Google Scholar] [CrossRef] [Green Version]
  265. Kyte, J.A.; Gaudernack, G.; Dueland, S.; Trachsel, S.; Julsrud, L.; Aamdal, S. Telomerase Peptide Vaccination Combined with Temozolomide: A Clinical Trial in Stage IV Melanoma Patients. Clin. Cancer Res. 2011, 17, 4568–4580. [Google Scholar] [CrossRef] [Green Version]
  266. Middleton, G.; Silcocks, P.; Cox, T.; Valle, J.; Wadsley, J.; Propper, D.; Coxon, F.; Ross, P.; Madhusudan, S.; Roques, T.; et al. Gemcitabine and capecitabine with or without telomerase peptide vaccine GV1001 in patients with locally advanced or metastatic pancreatic cancer (TeloVac): An open-label, randomised, phase 3 trial. Lancet Oncol. 2014, 15, 829–840. [Google Scholar] [CrossRef]
  267. Mizukoshi, E.; Kaneko, S. Telomerase-Targeted Cancer Immunotherapy. Int. J. Mol. Sci. 2019, 20, 1823. [Google Scholar] [CrossRef] [Green Version]
  268. Staff, C.; Mozaffari, F.; Frodin, J.-E.; Mellstedt, H.; Liljefors, M.G. Telomerase (GV1001) vaccination together with gemcitabine in advanced pancreatic cancer patients. Int. J. Oncol. 2014, 45, 1293–1303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  269. Khong, H.; Overwijk, W.W. Adjuvants for peptide-based cancer vaccines. J. Immunother. Cancer 2016, 4, 56. [Google Scholar] [CrossRef] [Green Version]
  270. Kim, B.-K.; Kim, B.-R.; Lee, H.-J.; Lee, S.-A.; Kim, B.-J.; Kim, H.; Won, Y.-S.; Shon, W.-J.; Lee, N.-R.; Inn, K.-S.; et al. Tumor-suppressive effect of a telomerase-derived peptide by inhibiting hypoxia-induced HIF-1α-VEGF signaling axis. Biomaterials 2014, 35, 2924–2933. [Google Scholar] [CrossRef] [PubMed]
  271. Kim, C.; Lee, S.G.; Yang, W.M.; Arfuso, F.; Um, J.Y.; Kumar, A.P.; Bian, J.; Sethi, G.; Ahn, K.S. Formononetin-induced oxidative stress abrogates the activation of stat3/5 signaling axis and suppresses the tumor growth in multiple myeloma preclinical model. Cancer Lett. 2018, 431, 123–141. [Google Scholar] [CrossRef] [PubMed]
  272. Kim, G.E.; Jung, A.R.; Kim, M.Y.; Lee, J.B.; Im, J.H.; Lee, K.W.; Park, Y.H.; Lee, J.Y. GV1001 Induces Apoptosis by Reducing Angiogenesis in Renal Cell Carcinoma Cells Both In Vitro and In Vivo. Urology 2018, 113, 129–137. [Google Scholar] [CrossRef] [PubMed]
  273. Park, Y.H.; Jung, A.R.; Kim, G.E.; Kim, M.Y.; Sung, J.W.; Shin, D.; Cho, H.J.; Ha, U.S.; Hong, S.H.; Kim, S.W.; et al. Gv1001 inhibits cell viability and induces apoptosis in castration-resistant prostate cancer cells through the akt/nf-kappab/vegf pathway. J. Cancer 2019, 10, 6269–6277. [Google Scholar] [CrossRef]
  274. Kim, H.; Seo, E.-H.; Lee, S.-H.; Kim, B.-J. The Telomerase-Derived Anticancer Peptide Vaccine GV1001 as an Extracellular Heat Shock Protein-Mediated Cell-Penetrating Peptide. Int. J. Mol. Sci. 2016, 17, 2054. [Google Scholar] [CrossRef] [Green Version]
  275. Schlapbach, C.; Yerly, D.; Daubner, B.; Yawalkar, N.; Hunger, R.E. Telomerase-specific GV1001 peptide vaccination fails to induce objective tumor response in patients with cutaneous T cell lymphoma. J. Dermatol. Sci. 2011, 62, 75–83. [Google Scholar] [CrossRef] [PubMed]
  276. Fenoglio, D.; Traverso, P.; Parodi, A.; Tomasello, L.; Negrini, S.; Kalli, F.; Battaglia, F.; Ferrera, F.; Sciallero, M.S.; Murdaca, G.; et al. A multi-peptide, dual-adjuvant telomerase vaccine (GX301) is highly immunogenic in patients with prostate and renal cancer. Cancer Immunol. Immunother. 2013, 62, 1041–1052. [Google Scholar] [CrossRef]
  277. Aucouturier, J.; Dupuis, L.; Deville, S.; Ascarateil, S.; Ganne, V. Montanide ISA 720 and 51: A new generation of water in oil emulsions as adjuvants for human vaccines. Expert Rev. Vaccines 2002, 1, 111–118. [Google Scholar] [CrossRef]
  278. Johnston, D.; Bystryn, J.-C. Topical imiquimod is a potent adjuvant to a weakly-immunogenic protein prototype vaccine. Vaccine 2006, 24, 1958–1965. [Google Scholar] [CrossRef]
  279. Fenoglio, D.; Parodi, A.; Lavieri, R.; Kalli, F.; Ferrera, F.; Tagliamacco, A.; Guastalla, A.; Lamperti, M.G.; Giacomini, M.; Filaci, G. Immunogenicity of GX301 cancer vaccine: Four (telomerase peptides) are better than one. Hum. Vaccines Immunother. 2015, 11, 838–850. [Google Scholar] [CrossRef] [Green Version]
  280. Filaci, G.; Fenoglio, D.; Nolè, F.; Zanardi, E.; Tomasello, L.; Aglietta, M.; Del Conte, G.; Carles, J.; Morales-Barrera, R.; Guglielmini, P.; et al. Telomerase-based GX301 cancer vaccine in patients with metastatic castration-resistant prostate cancer: A randomized phase II trial. Cancer Immunol. Immunother. 2021, 70, 1–14. [Google Scholar] [CrossRef]
  281. Van der Burg, S.H. Correlates of immune and clinical activity of novel cancer vaccines. Semin. Immunol. 2018, 39, 119–136. [Google Scholar] [CrossRef] [PubMed]
  282. Ellingsen, E.B.; Aamdal, E.; Inderberg, E.M.; Rasch, W.; Brunsvig, P.; Aamdal, S.; Hovig, E.; Nyakas, M.; Guren, T.K.; Gaudernack, G. A phase I/IIa clinical trial investigating the therapeutic cancer vaccine UV1 in combination with ipilimumab in patients with malignant melanoma: Four-year survival update. J. Clin. Oncol. 2020, 38, 62. [Google Scholar] [CrossRef]
  283. Haakensen, V.D.; Nowak, A.K.; Ellingsen, E.B.; Farooqi, S.J.; Bjaanæs, M.M.; Horndalsveen, H.; Mcculloch, T.; Grundberg, O.; Cedres, S.M.; Helland, Å. NIPU: A randomised, open-label, phase II study evaluating nivolumab and ipilimumab combined with UV1 vaccination as second line treatment in patients with malignant mesothelioma. J. Transl. Med. 2021, 19, 1–9. [Google Scholar] [CrossRef] [PubMed]
  284. Zakharia, Y.; O’Day, S.; Rasch, W.; Milhem, M.M. A phase I clinical trial investigating the telomerase vaccine UV1 in combination with pembrolizumab in patients with advanced melanoma. J. Clin. Oncol. 2021, 39, 2620. [Google Scholar] [CrossRef]
  285. Vetsika, E.-K.; Papadimitraki, E.; Aggouraki, D.; Konsolakis, G.; Mela, M.-E.; Kotsakis, A.; Christou, S.; Patramani, S.; Alefantinou, M.; Kaskara, A.; et al. Sequential Administration of the Native TERT572 Cryptic Peptide Enhances the Immune Response Initiated by its Optimized Variant TERT572Y in Cancer Patients. J. Immunother. 2011, 34, 641–650. [Google Scholar] [CrossRef]
  286. Bolonaki, I.; Kotsakis, A.; Papadimitraki, E.; Aggouraki, D.; Konsolakis, G.; Vagia, A.; Christophylakis, C.; Nikoloudi, I.; Magganas, E.; Galanis, A.; et al. Vaccination of Patients with Advanced Non–Small-Cell Lung Cancer With an Optimized Cryptic Human Telomerase Reverse Transcriptase Peptide. J. Clin. Oncol. 2007, 25, 2727–2734. [Google Scholar] [CrossRef] [PubMed]
  287. Kotsakis, A.; Papadimitraki, E.; Vetsika, E.K.; Aggouraki, D.; Dermitzaki, E.K.; Hatzidaki, D.; Kentepozidis, N.; Mavroudis, D.; Georgoulias, V. A phase II trial evaluating the clinical and immunologic response of HLA-A2+ non-small cell lung cancer patients vaccinated with an hTERT cryptic peptide. Lung Cancer 2014, 86, 59–66. [Google Scholar] [CrossRef]
  288. Kotsakis, A.; Vetsika, E.-K.; Christou, S.; Hatzidaki, D.; Vardakis, N.; Aggouraki, D.; Konsolakis, G.; Georgoulias, V.; Christophyllakis, C.; Cordopatis, P.; et al. Clinical outcome of patients with various advanced cancer types vaccinated with an optimized cryptic human telomerase reverse transcriptase (TERT) peptide: Results of an expanded phase II study. Ann. Oncol. 2012, 23, 442–449. [Google Scholar] [CrossRef] [PubMed]
  289. Vetsika, E.-K.; Konsolakis, G.; Aggouraki, D.; Kotsakis, A.; Papadimitraki, E.; Christou, S.; Menez-Jamet, J.; Kosmatopoulos, K.; Georgoulias, V.; Mavroudis, D. Immunological responses in cancer patients after vaccination with the therapeutic telomerase-specific vaccine Vx-001. Cancer Immunol. Immunother. 2012, 61, 157–168. [Google Scholar] [CrossRef]
  290. Brower, V. Telomerase-Based Therapies Emerging Slowly. J. Natl. Cancer Inst. 2010, 102, 520–521. [Google Scholar] [CrossRef]
  291. Ruden, M.; Puri, N. Novel anticancer therapeutics targeting telomerase. Cancer Treat. Rev. 2013, 39, 444–456. [Google Scholar] [CrossRef] [PubMed]
  292. Salazar-Onfray, F.; Pereda, C.; Reyes, D.; López, M.N. TAPCells, the Chilean dendritic cell vaccine against melanoma and prostate cancer. Biol. Res. 2013, 46, 431–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Yan, J.; Pankhong, P.; Shin, T.H.; Obeng-Adjei, N.; Morrow, M.P.; Walters, J.N.; Khan, A.S.; Sardesai, N.; Weiner, D.B. Highly Optimized DNA Vaccine Targeting Human Telomerase Reverse Transcriptase Stimulates Potent Antitumor Immunity. Cancer Immunol. Res. 2013, 1, 179–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Thalmensi, J.; Pliquet, E.; Liard, C.; Escande, M.; Bestetti, T.; Julithe, M.; Kostrzak, A.; Pailhes-Jimenez, A.-S.; Bourges, E.; Loustau, M.; et al. Anticancer DNA vaccine based on human telomerase reverse transcriptase generates a strong and specific T cell immune response. OncoImmunology 2016, 5, e1083670. [Google Scholar] [CrossRef] [Green Version]
  295. Aurisicchio, L.; Fridman, A.; Mauro, D.; Sheloditna, R.; Chiappori, A.; Bagchi, A.; Ciliberto, G. Safety, tolerability and immunogenicity of v934/v935 htert vaccination in cancer patients with selected solid tumors: A phase I study. J. Transl. Med. 2020, 18, 39. [Google Scholar] [CrossRef] [Green Version]
  296. Gangat, A.A.; Te, I.; Kao, Y.-J. Steady States of Infinite-Size Dissipative Quantum Chains via Imaginary Time Evolution. Phys. Rev. Lett. 2017, 119, 010501. [Google Scholar] [CrossRef] [Green Version]
  297. Kageyama, S.; Ikeda, H.; Miyahara, Y.; Imai, N.; Ishihara, M.; Saito, K.; Sugino, S.; Ueda, S.; Ishikawa, T.; Kokura, S.; et al. Adoptive Transfer of MAGE-A4 T-cell Receptor Gene-Transduced Lymphocytes in Patients with Recurrent Esophageal Cancer. Clin. Cancer Res. 2015, 21, 2268–2277. [Google Scholar] [CrossRef] [Green Version]
  298. Jackson, H.J.; Rafiq, S.; Brentjens, R.J. Driving CAR T-cells forward. Nat. Rev. Clin. Oncol. 2016, 13, 370–383. [Google Scholar] [CrossRef] [PubMed]
  299. Neelapu, S.S.; Locke, F.L.; Bartlett, N.L.; Lekakis, L.J.; Miklos, D.B.; Jacobson, C.A.; Braunschweig, I.; Oluwole, O.O.; Siddiqi, T.; Lin, Y.; et al. Axicabtagene Ciloleucel CAR T-Cell Therapy in Refractory Large B-Cell Lymphoma. N. Engl. J. Med. 2017, 377, 2531–2544. [Google Scholar] [CrossRef]
  300. Mu, X.; Sang, Y.; Fang, C.; Shao, B.; Yang, L.; Yao, K.; Zhao, X.; Gou, J.; Wei, Y.; Yi, T.; et al. Immunotherapy of tumors with human telomerase reverse transcriptase immortalized human umbilical vein endothelial cells. Int. J. Oncol. 2015, 47, 1901–1911. [Google Scholar] [CrossRef]
  301. Cesare, A.; Reddel, R. Alternative lengthening of telomeres: Models, mechanisms and implications. Nat. Rev. Genet. 2010, 11, 319–330. [Google Scholar] [CrossRef] [PubMed]
  302. Slatter, T.L.; Tan, X.; Yuen, Y.C.; Gunningham, S.; Ma, S.S.; Daly, E.; Packer, S.; Devenish, C.; Royds, J.A.; Hung, N.A. The alternative lengthening of telomeres pathway may operate in non-neoplastic human cells. J. Pathol. 2011, 226, 509–518. [Google Scholar] [CrossRef] [PubMed]
  303. Dilley, R.L.; Greenberg, R.A. ALTernative Telomere Maintenance and Cancer. Trends Cancer 2015, 1, 145–156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Fan, H.-C.; Chen, C.-M.; Chi, C.-S.; Tsai, J.-D.; Chiang, K.-L.; Chang, Y.-K.; Lin, S.-Z.; Harn, H.-J. Targeting Telomerase and ATRX/DAXX Inducing Tumor Senescence and Apoptosis in the Malignant Glioma. Int. J. Mol. Sci. 2019, 20, 200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Sommer, A.; Royle, N.J. ALT: A Multi-Faceted Phenomenon. Genes 2020, 11, 133. [Google Scholar] [CrossRef] [Green Version]
  306. Dyer, M.A.; Qadeer, Z.; Valle-Garcia, D.; Bernstein, E. ATRX and DAXX: Mechanisms and Mutations. Cold Spring Harb. Perspect. Med. 2017, 7, a026567. [Google Scholar] [CrossRef]
  307. Yang, X.; Khosravi-Far, R.; Chang, H.Y.; Baltimore, D. Daxx, a Novel Fas-Binding Protein That Activates JNK and Apoptosis. Cell 1997, 89, 1067–1076. [Google Scholar] [CrossRef] [Green Version]
  308. Gibbons, R.J.; McDowell, T.L.; Raman, S.; O’Rourke, D.M.; Garrick, D.; Ayyub, H.; Higgs, D.R. Mutations in ATRX, encoding a SWI/SNF-like protein, cause diverse changes in the pattern of DNA methylation. Nat. Genet. 2000, 24, 368–371. [Google Scholar] [CrossRef] [Green Version]
  309. Gibbons, R. Alpha thalassaemia-mental retardation, X linked. Orphanet J. Rare Dis. 2006, 1, 15. [Google Scholar] [CrossRef] [Green Version]
  310. Brosnan-Cashman, J.A.; Yuan, M.; Graham, M.K.; Rizzo, A.J.; Myers, K.M.; Davis, C.; Zhang, R.; Esopi, D.M.; Raabe, E.H.; Eberhart, C.G.; et al. ATRX loss induces multiple hallmarks of the alternative lengthening of telomeres (ALT) phenotype in human glioma cell lines in a cell line-specific manner. PLoS ONE 2018, 13, e0204159. [Google Scholar] [CrossRef] [Green Version]
  311. Ro, C.; Chai, W.; Yu, V.E.; Yu, R. Pancreatic neuroendocrine tumors: Biology, diagnosis, and treatment. Chin. J. Cancer 2013, 32, 312–324. [Google Scholar] [CrossRef] [Green Version]
  312. Jiao, Y.; Shi, C.; Edil, B.H.; de Wilde, R.F.; Klimstra, D.S.; Maitra, A.; Schulick, R.D.; Tang, L.H.; Wolfgang, C.L.; Choti, M.A.; et al. DAXX/ATRX, MEN1, and mTOR Pathway Genes Are Frequently Altered in Pancreatic Neuroendocrine Tumors. Science 2011, 331, 1199–1203. [Google Scholar] [CrossRef] [Green Version]
  313. Heaphy, C.M.; de Wilde, R.F.; Jiao, Y.; Klein, A.P.; Edil, B.H.; Shi, C.; Bettegowda, C.; Rodriguez, F.J.; Eberhart, C.G.; Hebbar, S.; et al. Altered Telomeres in Tumors with ATRX and DAXX Mutations. Science 2011, 333, 425. [Google Scholar] [CrossRef] [Green Version]
  314. Pekmezci, M.; Rice, T.; Molinaro, A.M.; Walsh, K.; Decker, P.A.; Hansen, H.; Sicotte, H.; Kollmeyer, T.M.; McCoy, L.S.; Sarkar, G.; et al. Adult infiltrating gliomas with WHO 2016 integrated diagnosis: Additional prognostic roles of ATRX and TERT. Acta Neuropathol. 2017, 133, 1001–1016. [Google Scholar] [CrossRef] [PubMed]
  315. Clynes, D.; Higgs, D.; Gibbons, R. The chromatin remodeller ATRX: A repeat offender in human disease. Trends Biochem. Sci. 2013, 38, 461–466. [Google Scholar] [CrossRef]
  316. Napier, C.E.; Huschtscha, L.I.; Harvey, A.; Bower, K.; Noble, J.R.; Hendrickson, E.A.; Reddel, R.R. ATRX represses alternative lengthening of telomeres. Oncotarget 2015, 6, 16543–16558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  317. Killela, P.J.; Reitman, Z.J.; Jiao, Y.; Bettegowda, C.; Agrawal, N.; Diaz, L.A., Jr.; Friedman, A.H.; Friedman, H.; Gallia, G.L.; Giovanella, B.C.; et al. TERT promoter mutations occur frequently in gliomas and a subset of tumors derived from cells with low rates of self-renewal. Proc. Natl. Acad. Sci. USA 2013, 110, 6021–6026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Flynn, R.L.; Cox, K.E.; Jeitany, M.; Wakimoto, H.; Bryll, A.R.; Ganem, N.J.; Bersani, F.; Pineda, J.R.; Suvà, M.L.; Benes, C.H.; et al. Alternative lengthening of telomeres renders cancer cells hypersensitive to ATR inhibitors. Science 2015, 347, 273–277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  319. Shay, J.W.; Reddel, R.R.; Wright, W.E. Cancer and Telomeres—An ALTernative to Telomerase. Science 2012, 336, 1388–1390. [Google Scholar] [CrossRef]
  320. Kosiol, N.; Juranek, S.; Brossart, P.; Heine, A.; Paeschke, K. G-quadruplexes: A promising target for cancer therapy. Mol. Cancer 2021, 20, 1–18. [Google Scholar] [CrossRef] [PubMed]
  321. Chhabra, G.; Wojdyla, L.; Frakes, M.; Schrank, Z.; Leviskas, B.; Ivancich, M.; Vinay, P.; Ganapathy, R.; Ramirez, B.E.; Puri, N. Mechanism of Action of G-Quadruplex–Forming Oligonucleotide Homologous to the Telomere Overhang in Melanoma. J. Investig. Dermatol. 2018, 138, 903–910. [Google Scholar] [CrossRef] [PubMed]
  322. Mender, I.; Gryaznov, S.; Dikmen, Z.G.; Wright, W.E.; Shay, J.W. Induction of telomere dysfunction mediated by the telomerase substrate precursor 6-thio-2′-deoxyguanosine. Cancer Discov. 2014, 5, 82–95. [Google Scholar] [CrossRef] [Green Version]
  323. Naderlinger, E.; Holzmann, K. Epigenetic Regulation of Telomere Maintenance for Therapeutic Interventions in Gliomas. Genes 2017, 8, 145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Dogan, F.; Forsyth, N. Telomerase Regulation: A Role for Epigenetics. Cancers 2021, 13, 1213. [Google Scholar] [CrossRef] [PubMed]
Figure 1. A graphic presentation of telomeric DNA and the proteins that form the shelterin complex. Telomeres are capping structures and are situated at the ends of linear chromosomes. Telomeric DNA, TTAGGG at the chromosome ends, and the complementary DNA strand sequence AATCCC form an extended region of dsDNA ending with a ssDNA G-rich overhang. The 3′ G-rich overhang enables telomeric DNA to form a secondary structure in which the 3′ single-stranded overhang folds back and displaces a strand in the homologous dsDNA TTAGGG region, to create a D-loop that protects the 3′-end from being identified as damaged DNA, thereby preventing the activation of the ataxia-telangiectasia mutation and Rad3-related (ATM/ATR) damage response pathways. The shelterin complex comprises six telomeric proteins: TRF1, TRF2, RAP1, TIN2, POT1, and TPP1. The complex enables the telomeric 3′-overhang/G-tail to fold into a lasso-like structure with a telomeric loop (T-loop) that protects the 3′-end from being recognized for DNA damage and blocks the DNA damage response.
Figure 1. A graphic presentation of telomeric DNA and the proteins that form the shelterin complex. Telomeres are capping structures and are situated at the ends of linear chromosomes. Telomeric DNA, TTAGGG at the chromosome ends, and the complementary DNA strand sequence AATCCC form an extended region of dsDNA ending with a ssDNA G-rich overhang. The 3′ G-rich overhang enables telomeric DNA to form a secondary structure in which the 3′ single-stranded overhang folds back and displaces a strand in the homologous dsDNA TTAGGG region, to create a D-loop that protects the 3′-end from being identified as damaged DNA, thereby preventing the activation of the ataxia-telangiectasia mutation and Rad3-related (ATM/ATR) damage response pathways. The shelterin complex comprises six telomeric proteins: TRF1, TRF2, RAP1, TIN2, POT1, and TPP1. The complex enables the telomeric 3′-overhang/G-tail to fold into a lasso-like structure with a telomeric loop (T-loop) that protects the 3′-end from being recognized for DNA damage and blocks the DNA damage response.
Life 11 01405 g001
Figure 2. Interaction of the shelterin complex, CST complex and telomerase to maintain telomere length. (A) The shelterin complex is essential for telomere protection and for regulating telomere elongation. TIN2, RAP1, and TRF1/2 subunits of the shelterin complex associate with telomeric dsDNA, while POT1 and TPP1 bind telomeric ssDNA and are responsible for recruiting telomerase to the telomeres. The shelterin complex also stimulates extension of the G-overhang by telomerase. (B) The CST complex prevents telomerase from engaging the G-overhang and facilitates the C-strand fill-in. The CST complex has three components—conserved telomere protection component 1 (CTC1), suppressor of cdc thirteen 1 (STN1) and telomeric pathway with STN1 (TEN1)—which are thought to function in part in telomere lagging-strand synthesis. The human telomerase consists of the hTERT, the TERC, and accessory proteins. The hTERT can wrap the chromosome to add single-stranded telomere repeats. The TERC contains the template for telomere replication. When an ongoing extension of a stranded DNA is finished, telomerase activity is terminated at the ssDNA overhang by the CST complex, which also activates the C-strand fill-in: the CST complex recruits DNA polymerase alpha (Polα) for lagging strand synthesis of the telomeric C-strand to convert the newly synthesized G-overhang into double-stranded DNA.
Figure 2. Interaction of the shelterin complex, CST complex and telomerase to maintain telomere length. (A) The shelterin complex is essential for telomere protection and for regulating telomere elongation. TIN2, RAP1, and TRF1/2 subunits of the shelterin complex associate with telomeric dsDNA, while POT1 and TPP1 bind telomeric ssDNA and are responsible for recruiting telomerase to the telomeres. The shelterin complex also stimulates extension of the G-overhang by telomerase. (B) The CST complex prevents telomerase from engaging the G-overhang and facilitates the C-strand fill-in. The CST complex has three components—conserved telomere protection component 1 (CTC1), suppressor of cdc thirteen 1 (STN1) and telomeric pathway with STN1 (TEN1)—which are thought to function in part in telomere lagging-strand synthesis. The human telomerase consists of the hTERT, the TERC, and accessory proteins. The hTERT can wrap the chromosome to add single-stranded telomere repeats. The TERC contains the template for telomere replication. When an ongoing extension of a stranded DNA is finished, telomerase activity is terminated at the ssDNA overhang by the CST complex, which also activates the C-strand fill-in: the CST complex recruits DNA polymerase alpha (Polα) for lagging strand synthesis of the telomeric C-strand to convert the newly synthesized G-overhang into double-stranded DNA.
Life 11 01405 g002
Figure 3. Comparison of (A) S. cerevisiae Rfa and CST with (B) human RPA and CST. Domain structures of Rfa and CST. DBD: DNA-binding domain;OB: OB-fold domain; RD: recruitment domain; TR2: the single RAD51-binding domain; WH: winged helix domain; wHTH: winged helix-turn-helix domain.
Figure 3. Comparison of (A) S. cerevisiae Rfa and CST with (B) human RPA and CST. Domain structures of Rfa and CST. DBD: DNA-binding domain;OB: OB-fold domain; RD: recruitment domain; TR2: the single RAD51-binding domain; WH: winged helix domain; wHTH: winged helix-turn-helix domain.
Life 11 01405 g003
Figure 4. Schematic structure of telomerase. (A)Components of telomerase consist of the TERT (hTERT), the telomerase RNA template (TERC), and accessory proteins, including dyskerin, GAR1, NHP2, and NOP10. HSP 90, p23, pontin, reptin, and serine, SRSF11, and TCAB1. (B) The hTERT gene is situated on 5p15.33 which is responsible for 40 kb of human genome. The hTERT gene promoter contains five GC boxes (5′-GGGCGG-3′), two E-boxes (5′-CACGTG-3′), and one TSS. GC boxes are interacted with SP-1; E-boxes also have binding sites to MAD1; TSS binds THF1. Point mutations at the TERT promoter, predominantly at two points (C228T and C250T) generate new ETS/ternary complex (ETS/TCF) binding sites for transcription factors (TF). Increasing the expression of TFs such as c-MYC, ETS, NF-kB, and SP-1 results in binding to their particular sites and can up-regulate hTERT transcription. Binding of down-regulating transcription factors, such as WT1, CTCF, and MZF2, down-regulate TERT transcription. CTCF: CCCTC-binding factor; hTERT: human telomerase reverse transcriptase; NF-kB: nuclear factor kappa-light-chain-enhancer of activated B; TSS: transcription start site; WT1: Wilms tumor protein1; MZF2: myeloid zinc finger protein 2; GAR1: glycine-arginine rich 1; NHP2: non-histone protein 2; NOP10: nucleolar protein 10; HSP90: heat shock protein 90; TCAB1: telomerase Cajal body protein 1; SRSF11: serine and arginine-rich splicing factor 11.
Figure 4. Schematic structure of telomerase. (A)Components of telomerase consist of the TERT (hTERT), the telomerase RNA template (TERC), and accessory proteins, including dyskerin, GAR1, NHP2, and NOP10. HSP 90, p23, pontin, reptin, and serine, SRSF11, and TCAB1. (B) The hTERT gene is situated on 5p15.33 which is responsible for 40 kb of human genome. The hTERT gene promoter contains five GC boxes (5′-GGGCGG-3′), two E-boxes (5′-CACGTG-3′), and one TSS. GC boxes are interacted with SP-1; E-boxes also have binding sites to MAD1; TSS binds THF1. Point mutations at the TERT promoter, predominantly at two points (C228T and C250T) generate new ETS/ternary complex (ETS/TCF) binding sites for transcription factors (TF). Increasing the expression of TFs such as c-MYC, ETS, NF-kB, and SP-1 results in binding to their particular sites and can up-regulate hTERT transcription. Binding of down-regulating transcription factors, such as WT1, CTCF, and MZF2, down-regulate TERT transcription. CTCF: CCCTC-binding factor; hTERT: human telomerase reverse transcriptase; NF-kB: nuclear factor kappa-light-chain-enhancer of activated B; TSS: transcription start site; WT1: Wilms tumor protein1; MZF2: myeloid zinc finger protein 2; GAR1: glycine-arginine rich 1; NHP2: non-histone protein 2; NOP10: nucleolar protein 10; HSP90: heat shock protein 90; TCAB1: telomerase Cajal body protein 1; SRSF11: serine and arginine-rich splicing factor 11.
Life 11 01405 g004
Table 1. Comparison of CST and RPA.
Table 1. Comparison of CST and RPA.
CST.ComponentaaOBwHwHTH1FunctionsReferences
  • Binding to ssDNA.
[119,133,134,135,136,137]
CTC11217700
2.
Binding to ssDNA-dsDNA junctions.
[138]
STN1368102
3.
Recognize different specialized DNA structures at DNA replication and breakage sites.
[139]
TEN1123100
4.
Acting synergistically with ATR to maintain telomere length and genome stability.
[140]
5.
Stimulating Polα.
[141,142,143]
6.
Helping in C-strand fill-in.
[136,144]
7.
Preventing the accumulation of G4.
[139]
8.
Preventing telomeric DNA damage.
[125,134]
9.
Interacting with the MCM and disrupting binding of CDT1 to MCM, leading to decreased origin licensing.
[145]
10.
Interacting with AND-1.
[146]
11.
Inhibiting telomerase.
[138,143,144,147,148]
RPA
1.
Binding ssDNA.
[149]
RPA70616400
2.
Activating the ATR signaling.
[150]
RPA32270110
3.
Activating the helicase.
[151]
RPA14121100
4.
Unwinding G4.
[152]
5.
Involved in DNA replication, recombination, and repair.
[62]
6.
Activating BLM’s bidirectional DNA unwinding.
[153]
7.
Modulating the fork remodeling enzyme activity.
[154]
8.
Enhancing primase.
[155]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fan, H.-C.; Chang, F.-W.; Tsai, J.-D.; Lin, K.-M.; Chen, C.-M.; Lin, S.-Z.; Liu, C.-A.; Harn, H.-J. Telomeres and Cancer. Life 2021, 11, 1405. https://doi.org/10.3390/life11121405

AMA Style

Fan H-C, Chang F-W, Tsai J-D, Lin K-M, Chen C-M, Lin S-Z, Liu C-A, Harn H-J. Telomeres and Cancer. Life. 2021; 11(12):1405. https://doi.org/10.3390/life11121405

Chicago/Turabian Style

Fan, Hueng-Chuen, Fung-Wei Chang, Jeng-Dau Tsai, Kao-Min Lin, Chuan-Mu Chen, Shinn-Zong Lin, Ching-Ann Liu, and Horng-Jyh Harn. 2021. "Telomeres and Cancer" Life 11, no. 12: 1405. https://doi.org/10.3390/life11121405

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

Article Metrics

Back to TopTop