Next Article in Journal
The Effect of Gender and Menstrual Phase on Serum Creatine Kinase Activity and Muscle Soreness Following Downhill Running
Next Article in Special Issue
Current Insights to Regulation and Role of Telomerase in Human Diseases
Previous Article in Journal
Inorganic Reactive Sulfur-Nitrogen Species: Intricate Release Mechanisms or Cacophony in Yellow, Blue and Red?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Treating Cancer by Targeting Telomeres and Telomerase

Department of Biomedical Sciences, College of Medicine at Rockford, University of Illinois, Rockford, IL 61107, USA
*
Author to whom correspondence should be addressed.
Antioxidants 2017, 6(1), 15; https://doi.org/10.3390/antiox6010015
Submission received: 2 December 2016 / Revised: 15 February 2017 / Accepted: 16 February 2017 / Published: 19 February 2017
(This article belongs to the Special Issue Role of Telomerase in Aging and Cancer)

Abstract

:
Telomerase is expressed in more than 85% of cancer cells. Tumor cells with metastatic potential may have a high telomerase activity, allowing cells to escape from the inhibition of cell proliferation due to shortened telomeres. Human telomerase primarily consists of two main components: hTERT, a catalytic subunit, and hTR, an RNA template whose sequence is complimentary to the telomeric 5′-dTTAGGG-3′ repeat. In humans, telomerase activity is typically restricted to renewing tissues, such as germ cells and stem cells, and is generally absent in normal cells. While hTR is constitutively expressed in most tissue types, hTERT expression levels are low enough that telomere length cannot be maintained, which sets a proliferative lifespan on normal cells. However, in the majority of cancers, telomerase maintains stable telomere length, thereby conferring cell immortality. Levels of hTERT mRNA are directly related to telomerase activity, thereby making it a more suitable therapeutic target than hTR. Recent data suggests that stabilization of telomeric G-quadruplexes may act to indirectly inhibit telomerase action by blocking hTR binding. Telomeric DNA has the propensity to spontaneously form intramolecular G-quadruplexes, four-stranded DNA secondary structures that are stabilized by the stacking of guanine residues in a planar arrangement. The functional roles of telomeric G-quadruplexes are not completely understood, but recent evidence suggests that they can stall the replication fork during DNA synthesis and inhibit telomere replication by preventing telomerase and related proteins from binding to the telomere. Long-term treatment with G-quadruplex stabilizers induces a gradual reduction in the length of the G-rich 3’ end of the telomere without a reduction of the total telomere length, suggesting that telomerase activity is inhibited. However, inhibition of telomerase, either directly or indirectly, has shown only moderate success in cancer patients. Another promising approach of targeting the telomere is the use of guanine-rich oligonucleotides (GROs) homologous to the 3’ telomere overhang sequence (T-oligos). T-oligos, particularly a specific 11-base oligonucleotide (5’-dGTTAGGGTTAG-3’) called T11, have been shown to induce DNA damage responses (DDRs) such as senescence, apoptosis, and cell cycle arrest in numerous cancer cell types with minimal or no cytostatic effects in normal, non-transformed cells. As a result, T-oligos and other GROs are being investigated as prospective anticancer therapeutics. Interestingly, the DDRs induced by T-oligos in cancer cells are similar to the effects seen after progressive telomere degradation in normal cells. The loss of telomeres is an important tumor suppressor mechanism that is commonly absent in transformed malignant cells, and hence, T-oligos have garnered significant interest as a novel strategy to combat cancer. However, little is known about their mechanism of action. In this review, we discuss the current understanding of how T-oligos exert their antiproliferative effects in cancer cells and their role in inhibition of telomerase. We also discuss the current understanding of telomerase in cancer and various therapeutic targets related to the telomeres and telomerase.

1. Introduction

Mammalian chromosome ends are capped by protective DNA structures and DNA-binding proteins, collectively known as telomeres [1]. In adult somatic cells, telomeres are composed of large noncoding sequences of approximately 1000–2000 TTAGGG tandem base pair repeats that end in a 3′ extension past the 5′ terminus [2]. In other select cells, the telomeres are maintained and can be considerably longer, such as adult human germ cells which can contain anywhere from 500–5000 repeats [3,4,5]. During each cycle of cell division telomeres are incompletely replicated, and consequently, their ends are progressively shortened [6]. When these telomeres reach a critically short length, DNA damage responses (DDRs) such as apoptosis and cellular senescence are induced. Hence, telomeres are considered to be “biological clocks”, as they limit the proliferative potential of most normal cells [2,7]. It has been demonstrated that telomeres exist in a secondary structure known as the t-loop, which is formed by the invasion of the 3’ overhang into the duplex region of the telomere and is stabilized by the shelterin complex, a grouping of six proteins that regulate telomeric stability and homeostasis (Figure 1) [2,8,9]. The shelterin complex is composed of the proteins TRF1, TRF2, POT1, TIN2, TPP1, and Rap1, which bind to the single-stranded and/or double-stranded regions of the telomere (Figure 1) [2,9,10]. The shelterin complex also plays an integral role in the regulation of telomerase and the prevention of telomere degradation by nucleases. When TRF1, TRF2, or POT1 are not functioning properly or dissociate from the telomere, the t-loop unfolds, exposing the telomere, which induces DDRs such as apoptosis and senescence [1,11,12]. It is thought that the T-loop is also stabilized by the guanine-rich (G-rich) character present in telomeres [6,12,13], allowing the telomere 3’ overhang to take on a G-quadruplex structure, a secondary structure formed from the hydrogen bonding of guanine residues in tetrad formations (Figure 1). These G-quadruplexes stack together at the D-loop, the area where the 3’ end of the telomere penetrates the duplex region [2,10]. It is thought that these G-quadruplexes have a crucial role in the blocking of the enzyme telomerase from gaining access to the telomere [1,12].
Telomerase is inappropriately expressed in roughly 85%of cancers, and the level of its activity is higher in advanced and metastatic tumors, making it a viable cancer biomarker and therapeutic target. While germ line cells, stem cells, and select others including cardiovascular cells do express detectable levels of telomerase, it is mostly quiescent, and most normal cells have minimal or no detectable telomerase activity [14,15,16,17]. Telomerase confers cellular immortality through the addition of TTAGGG tandem repeats to the 3’ end of the telomere, thereby maintaining stable telomere length, preventing the induction of DDRs by the critical shortening of telomeres, and allowing continued proliferation of neoplastic cells [7]. Telomerase is comprised of two main components: hTERT, the catalytic subunit of telomerase which catalyzes the addition of nucleotides to the 3′ overhang; and hTR, an RNA template complementary to the 3′ overhang which acts as a primer, in addition to telomerase-associated proteins hTEP1, p23, Hsp90 and dyskerin [10,18,19]. Limiting the growth potential of tumors has been the focus of chemotherapeutic intervention for decades, and because of their selective expression in neoplastic growths, telomerase and telomeres have become attractive and novel targets for the development of anticancer therapeutics [2,10].
A minority of cancers, less than 15%, maintain telomere length homeostasis through the alternative lengthening of the telomeres (ALT) pathway. ALT-positive cells are able to replenish telomeric DNA in a telomerase-independent manner through a homologous recombination-mediated replication mechanism, and are thus resistant to telomerase-based therapies [1,11]. However, the mechanism of ALT is still poorly understood, and it is possible that multiple mechanisms of ALT exist [20]. In addition, it is thought that resistance to telomerase inhibitor therapy may occur through the activation of ALT pathways in some cancers [1].

2. Current Therapies Related to Telomeres and Telomerase

Due to its over-expression in the majority of cancers, and minimal or nonexistent expression in most somatic cells, telomerase is a unique cancer biomarker. Thus, telomerase and other telomere components are attractive targets for developing therapeutics. Many telomere-based therapies are currently under investigation, including telomerase inhibitors, tankyrase inhibitors, and guanine-rich oligonucleotides (GROs) [2,10].
Currently there is a multitude of drugs designed to directly inhibit telomerase. GRN163L, also called Imetelstat, has been one of the most widely developed and is arguably the most successful [21]. GRN163L is a 13-mer oligonucleotide that acts as a direct telomerase inhibitor by antagonistically binding to the RNA template of telomerase (hTR) [21,22]. Preclinical studies utilizing GRN163L have shown effective inhibition of telomerase. Treatment on breast cancer cells demonstrated a reduction in cell tumorigenicity and invasiveness [23,24]. However, the effects of long-term treatment with telomerase inhibitors have not yet been investigated and there is currently no clear information on their effects on normal cells that transiently express telomerase, such as germ cells. Studies by Herbert et al. with modified oligonucleotides as telomerase inhibitors, using immortalized breast cancer cells, indicate that telomere shortening is reversible and may have limited side effects on stem cells [25]. Furthermore, several cancers maintain very short telomeres that are maintained by telomerase, and evidence suggests that increased tumorigenicity occurs due to dysfunction of these telomeres [1,19,26]. These shortened telomeres of tumors may erode to a critical length before irreversible harm occurs to telomerase-positive stem cells [25].
Indirect inhibition of telomerase can occur via the silencing of tankyrase 1 (TNKS1), a protein that PARsylates TRF1 during the S-phase of cell division and displaces it from the telomere, which is required for telomerase activity [27,28,29]. Depletion of TNKS1 with siRNA knockdown results in telomere uncapping and increased sensitivity to ionizing radiation, which could in turn be potentially useful to cancer patients [30]. shRNA against tankyrase 1 has been shown to reduce cell viability in cancers [31]. Furthermore, studies have demonstrated enhanced telomerase inhibition through the inhibition of tankyrase 1 in conjunction with telomerase inhibitors such as MST-312 [31,32]. Combinatorial therapy using tankyrase 1 inhibitors and telomerase inhibitors may have an even greater inhibitory effect on telomerase and thus may be an effective anticancer therapeutic.
Another strategy aimed at overcoming telomerase-related immortality is the increased stabilization and formation of endogenous telomeric G-quadruplex secondary structures within the telomere. G-quadruplexes (G4) form naturally in telomeric regions and it is thought that the steric hindrance at these regions may disrupt telomerase activity in these regions [33]. The stabilization of G-quadruplexes can prevent telomerase from accessing and elongating the telomere, thus impeding the progression of the replication fork through telomeric tracts [1,2,34]. However, telomerase has no role in resolving G4-structures along the telomeric repeats since they are resolved by the DNA helicases [34]. As such, the use of G-quadruplex–stabilizing ligands has potential applications for the development of treatments for malignant and progressive cancers. These agents bind with high affinity to the 3’ single-stranded region of the telomere, facilitating formation and stabilization of the G-quadruplexes. Some of the most promising G-quadruplex–stabilizing ligands include telomestatin, BRACO-19, and RHPS4, and studies utilizing these agents have demonstrated increased G-quadruplex stability and upregulation of DDRs in cancer cells [33,35,36,37,38]. Long-term treatment with these agents has also demonstrated a reduction in 3’ overhang length without a reduction in the overall telomere length, which suggests the inhibition of telomerase [39].
Recent studies have suggested that the reverse transcriptase catalytic subunit (TERT) of telomerase mediates other oncogenic activity in addition to its role in telomere elongation [40,41]. TERT upregulation or reactivation seems to be associated with several “hallmarks of cancer” including increased mitochondrial activity, DDR signaling, and WNT/β-catenin signaling [42,43,44]. More recently, it was shown that TERT regulates MYC-driven oncogenesis independently of its reverse transcriptase activity [13]. TERT-null mice showed a delayed development of MYC-induced lymphomagenesis, while Terc-null mice did not [45]. Studies after ectopic expression of TERT showed enhanced translation in cells to regulate cancer cell proliferation independent of telomere length, suggesting that the effects of TERT could be telomere-independent [46]. Additionally, it has been known that TERT is a major component and limiting factor for telomerase activity [41,47]. Based on these findings, we suggest that a human TERT (hTERT) inhibitor may be a more efficient therapeutic than current hTR inhibitors since, it would prohibit telomere elongation in addition to halting TERT’s non-canonical effects.
TERT expression is often increased in cancer due to point mutations in TERT’s promoter region, resulting in the recruitment of multiple different potential transcriptions factors that upregulate TERT and thus increase telomerase activity [48,49,50,51]. TERT expression has been demonstrated to be dependent on MAPK pathway activation which is mediated by transcription factors of the ETS (E26 transformation-specific) family in melanoma cells [51]. Examination of TERT reactivation through the recruitment of ETS transcription factors demonstrated that the process is dependent on non-canonical NF-κB signaling [48]. Given TERT’s potential role in multiple oncogenic proliferative processes and its association with ETS, future studies that target inhibition of mutant TERT promoters appear to be a promising strategy.
Current therapeutic agents that promote telomere attrition through direct or indirect inhibition of telomerase have demonstrated limited improvement in cancer patient prognosis [13,21]. Although the inhibition of telomerase may strip some cancers of their immortality, cancers are still viable and largely unaffected by the loss of telomerase. Furthermore, some cancers develop resistance to these inhibitors by means of ALT, rendering them totally ineffective [52]. Targeting hTERT, specifically via the inhibition of TERT promoter- or mutant promoter-mediated transcription, may be a far more effective strategy [49,53]. Additionally, a potentially promising telomere-related therapy is the use of guanine-rich oligonucleotides (GROs), specifically those that are homologous to the telomere, to disrupt the telomere [1].

3. Telomere Homolog Oligonucleotides

Commonly known as T-oligos, telomere homolog oligonucleotides are known to have potent anticancer effects when administered to several malignant cell types, both in vitro and in vivo. Our lab has studied a number of these oligonucleotides in several different cancer cell lines, and our current research centers around one particular 11-base oligonucleotide (5′-dGTTAGGGTTAG-3′) called T11 [1,2]. T-oligos accumulate in the nucleus and rapidly induce DDRs mediated by ATM, p53, E2F1, cdk2, and p95/NBS1 and their downstream targets, resulting in cell cycle arrest, senescence, apoptosis, and differentiation [54,55,56,57,58].
In vitro studies demonstrate that T11 is highly effective in reducing viability and growth of several cancers including melanoma, lung, prostate, ovarian, breast, and colorectal [1,2,56,57,58,59,60,61,62,63]. Treatment with T11 demonstrated upregulation of several tumor differentiation markers in colorectal cancer, which have roles in inhibiting proliferation and are lost in poorly differentiated cancers [61]. T11 treatment also induces the upregulation of several melanoma differentiation proteins which are currently the targets of melanoma vaccine therapies [63]. Combination treatment of T11 with a tyrosine kinase inhibitor or histone deacetylase inhibitor that are currently used clinically, have demonstrated additive inhibition of cancer cell cellular growth [61,62]. Furthermore, when combined with ionizing radiation treatment, T11 increased radiosensitivity and synergistically inhibited cell and tumor growth both in vitro and in vivo, respectively [59].
T11 also stimulates various anti-cancer responses in vivo, such as reduction of tumor burden and metastatic potential in mice, with no detectable toxicity [63,64,65,66,67]. Our lab has shown that tumor volume in mice with NSCLC H358 and SW1573 lung cancer xenografts was reduced by 80% and 88%, respectively, after treatment with T11 for seven weeks [64]. Another in vivo study on melanoma reported that, when compared with controls, T11 reduced the average number of metastases by 90%–95% and tumor volume by 84%–88% [63]. T11 also demonstrated its ability to elicit its anti-tumor effects by inducing senescence and inhibiting angiogenesis in melanoma and lung cancer [63,64,65].
Although the DDRs initiated by T11 have been intensely studied and are mostly elucidated, the process by which T11 initially activates these DDRs is not completely understood [1,63,68]. It has been proposed that T11 mimics the physiological signal of telomere exposure [57,63], since it elicits DDRs similar to those induced by telomere dysfunction after ectopic expression of a non-functional TRF2 [69,70]. Studies have shown that T11 upregulates the shelterin protein TRF2 [71]. Recently, preliminary studies from our laboratory showed that TRF2 and POT1 are upregulated after treatment with T11, and immunofluorescence studies performed in our lab showed possible co-localization of these proteins with T11.
From these findings we are able to conclude that there are two potential modes of T-oligo action. We have designated the first mode as the shelterin dissociation model (SDM) and the other we refer to as the exposed telomere mimicry model (ETM). The SDM hypothesizes that the introduction of T11 into the nucleus displaces shelterin proteins from the telomere, thereby critically compromising the telomere secondary structure. It is thought that at least some shelterin proteins bind T11, resulting in the opening of the T-loop causing DDR signaling (Figure 2). The ETM model hypothesizes that T11 accumulates in the nucleus and is recognized as an exposed or damaged telomere, thus initiating DDRs identical to those that occur under normal physiological conditions when telomeres are critically shortened (Figure 3). It is also possible that both potential modes are occurring simultaneously and that T11’s effectiveness is due to both its ability to disrupt the shelterin complex and mimic telomere exposure. In summary, in the exposed telomere mimicry model exogenous T11 mimics the exposed telomere overhang, resulting in initiating DDRs in cancer cells. In the shelterin dissociation model, the presence of T11 in the nucleus causes the disruption of the telomere and the shelterin complex since T11 competes with telomeric DNA for the binding of shelterin proteins, resulting in telomere overhang exposure and initiation of DDRs. Since the activity of telomerase is independent of T11, it will not be modulated in either model [72,73]. Further studies are needed to improve our understanding of these mechanisms.
Interestingly, T11 and similar GROs have been shown to have little or no antiproliferative effect on normal cells studied and had no detectable toxic effects in mice [2,63,67]. It is important to note that T11 and other GROs have only been tested in a number of non-malignant tissues and treatment times vary between three to six days [60,63,64]. Furthermore, murine telomeres are five to 10 times longer and have constitutive expression of telomerase compared to humans, resulting in reduced species longevity in mice [74,75]. Laboratory mice and humans also have differences in the sequence homology of telomeric proteins TRF1 and POT1 [76,77,78]. This indicates that T-oligo–initiated DDRs are specific to cancer cells and may be due to changes in telomere physiology that are common among cancers. Silencing hTERT in oral cancer cells is accompanied by caspase-9/-3 cleavage, DNA damage responses, and decreased cell viability [79]. It may be possible that T11’s antiproliferative response is in some way hTERT-dependent or hTERT-mediated due to its homology to the telomere, and therefore selective to cells that express the telomerase enzyme. Further studies should be done to elucidate the role hTERT plays in T11-induced DDRs in addition to analysis of T11’s effect on non-cancerous telomerase positive cells. The effects of T-oligo on cells that exhibit ALT are also poorly understood, and should be a subject of future studies. However, one study conducted using ALT-positive U20S osteosarcoma cells demonstrated a function for WRN in T-oligo–induced responses [80]. WRN is a protein which has exonuclease and helicase activity and is defective in Werner Syndrome patients [80]. Downregulation of WRN using siRNA decreased phosphorylation of γH2AX, p53 and ATM in T-oligo–treated osteosarcoma cells [80]. Though this study demonstrates that T-oligo induces DNA damage responses in ALT-positive cells, further investigation into this area is a necessity.
Despite its effectiveness, T11 has limited stability in vitro and in vivo due to degradation by nucleases [66]. To enhance its stability and delivery, our lab has recently complexed T11 with a cationic helical polypeptide, PVBLG-8 (PVBLG) [66]. The helical structure of PVBLG allows it to remain stable across a wide range of pH values, temperatures, and salt concentrations in the presence of denaturing agents, making it an excellent delivery vehicle [66,81]. It has been shown to be a highly efficient vehicle for plasmid delivery and siRNA in many cell lines [81,82]. PVBLG is cationic and self-assembles with negatively charged oligonucleotides such as T11, neutralizing their charges and stabilizing them for optimal delivery [66]. When complexed with PVBLG, T11’s cellular uptake was improved on a log scale and its ability to inhibit cellular growth and reduce tumor burden in mice was significantly enhanced [66]. Melanoma tumors grown on the flanks of SCID mice and subsequently treated with T11 in the presence or absence of PVBLG showed a three-fold and nine-fold reduction in tumor size, respectively [66]. However, since PVBLG is currently difficult to synthesize, a more efficient yet equally effective alternative is needed in order for it to be produced on a large scale for cancer treatment.
Novel lipid nano-particles are being researched by our lab as a potential alternative for PVBLG. In theory, lipid nanoparticles will form a liposome around T11, thereby increasing its cellular uptake and minimizing degradation. However, this delivery system has had limited success as many of the current lipids have high degrees of toxicity [83,84]. In addition, it was recently discovered that many guanine-rich oligonucleotides, such as T11, are able to form G-quadruplexes [57]. We suggest that the additional secondary structure of a G-quadruplexed T11 may confer added stability in serum, thus increasing its efficacy which is supported by preliminary studies in our laboratory.

4. Conclusions

Molecularly targeted therapies are a cornerstone for the treatment of aggressive cancers [85]. Due to the role that telomeres play in the life cycle of cancers, telomerase and telomere-based therapeutics may be able to replace conventional treatments in the future. Undesirable side effects that conventional therapies elicit may be reduced or eliminated due to the specificity of these treatments. There is a small percentage of cells that express telomerase, such as germ cells and stem cells. Extended exposure to telomerase-targeted therapies may potentially inhibit growth of these stem cells. Normal-tissue stem cells are telomerase-competent. However, telomerase is mostly silent with minimal activation in stem cells, while cancer cells are almost universally telomerase-expressing and active [16,17]. Currently, the effects of telomerase inhibition on stem cells with telomerase activity are unclear and require further study [2].
In this review, we described the two current proposed mechanisms of action of T11 and other related short-stranded telomere-homologous oligonucleotides (Figure 2 and Figure 3). However, further research is needed to investigate both these models. It is also critical to determine why the antiproliferative responses of T11 appear to be limited only to malignant cells [1,2,68], which would aide researchers in progressing this therapy into clinical trials. Additionally, exploring T11’s mechanism of action may lead to the discovery of novel potential targets for future therapies. T11 and other T-oligos show promise in preclinical studies; however, there is a hindrance in their delivery in vivo due to their degradation by serum and intracellular nucleases [66]. Stabilization of oligonucleotides by nanocomplex formation with cationic polypeptides is a potential solution, since they have excellent bioavailability and confer enhanced gene delivery with minimal toxicity [66]. Due to its high activity in the majority of cancers and low expression in somatic cells, telomerase and the telomere as a whole are very promising targets [2,10]. Continued studies on telomeres may provide promising avenues of research in the development of novel anticancer therapeutics. G-quadruplex stabilizers could provide a viable means of blocking telomerase [1]. These therapeutics may prove to be more effective than direct or indirect inhibition of telomerase in decreasing cancer cell survival due to their ability to form telomeric G-quadruplexes in situ. Another novel approach of inhibition of telomerase activity in cancer cells relies on the inhibition of tankyrase-1, which shortens telomeres in cancer cells and initiates DDRs [2]. Studies have demonstrated that the suppression of telomerase activity by these inhibitors can eliminate resistance to telomerase inhibition and may be used in combinatorial clinical therapies [31]. In the future, combination therapies of telomerase inhibitors and standard-of-care or traditional therapies may be the most effective way to target telomerase-positive tumors [2]. These modalities could include combination therapy with targets of the non-canonical variety which can eliminate resistance to telomerase inhibition and may be used in conjunction in clinical therapies [40].

Author Contributions

M.I., Z.S., L.W., B.L., A.K., A.S. and N.P. wrote and designed the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Crees, Z.; Girard, J.; Rios, Z.; Botting, G.M.; Harrington, K.; Shearrow, C.; Wojdyla, L.; Stone, A.L.; Uppada, S.B.; Devito, J.T.; et al. Oligonucleotides and g-quadruplex stabilizers: Targeting telomeres and telomerase in cancer therapy. Curr. Pharm. Des. 2014, 20, 6422–6437. [Google Scholar] [CrossRef] [PubMed]
  2. Ruden, M.; Puri, N. Novel anticancer therapeutics targeting telomerase. Cancer Treat. Rev. 2013, 39, 444–456. [Google Scholar] [CrossRef] [PubMed]
  3. Thilagavathi, J.; Venkatesh, S.; Dada, R. Telomere length in reproduction. Andrologia 2013, 45, 289–304. [Google Scholar] [CrossRef] [PubMed]
  4. Reig-Viader, R.; Garcia-Caldes, M.; Ruiz-Herrera, A. Telomere homeostasis in mammalian germ cells: A review. Chromosoma 2016, 125, 337–351. [Google Scholar] [CrossRef] [PubMed]
  5. Antunes, D.M.; Kalmbach, K.H.; Wang, F.; Dracxler, R.C.; Seth-Smith, M.L.; Kramer, Y.; Buldo-Licciardi, J.; Kohlrausch, F.B.; Keefe, D.L. A single-cell assay for telomere DNA content shows increasing telomere length heterogeneity, as well as increasing mean telomere length in human spermatozoa with advancing age. J. Assist. Reprod. Genet. 2015, 32, 1685–1690. [Google Scholar] [CrossRef] [PubMed]
  6. 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]
  7. De Lange, T. Protection of mammalian telomeres. Oncogene 2002, 21, 532–540. [Google Scholar] [CrossRef] [PubMed]
  8. Dimitrova, N.; de Lange, T. Cell cycle-dependent role of MRN at dysfunctional telomeres: ATM signaling-dependent induction of nonhomologous end joining (NHEJ) in G1 and resection-mediated inhibition of NHEJ in G2. Mol. Cell. Biol. 2009, 29, 5552–5563. [Google Scholar] [CrossRef] [PubMed]
  9. Ribes-Zamora, A.; Indiviglio, S.M.; Mihalek, I.; Williams, C.L.; Bertuch, A.A. TRF2 interaction with Ku heterotetramerization interface gives insight into c-NHEJ prevention at human telomeres. Cell Rep. 2013, 5, 194–206. [Google Scholar] [CrossRef] [PubMed]
  10. 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]
  11. Dunham, M.A.; Neumann, A.A.; Fasching, C.L.; Reddel, R.R. Telomere maintenance by recombination in human cells. Nat. Genet. 2000, 26, 447–450. [Google Scholar] [PubMed]
  12. Luke-Glaser, S.; Poschke, H.; Luke, B. Getting in (and out of) the loop: Regulating higher order telomere structures. Front. Oncol. 2012, 2, 180. [Google Scholar] [CrossRef] [PubMed]
  13. Xu, Y.; Goldkorn, A. Telomere and telomerase therapeutics in cancer. Genes 2016, 7, 22. [Google Scholar] [CrossRef] [PubMed]
  14. Zurek, M.; Altschmied, J.; Kohlgruber, S.; Ale-Agha, N.; Haendeler, J. Role of telomerase in the cardiovascular system. Genes 2016, 7, 29. [Google Scholar] [CrossRef] [PubMed]
  15. Pech, M.F.; Garbuzov, A.; Hasegawa, K.; Sukhwani, M.; Zhang, R.J.; Benayoun, B.A.; Brockman, S.A.; Lin, S.; Brunet, A.; Orwig, K.E.; et al. High telomerase is a hallmark of undifferentiated spermatogonia and is required for maintenance of male germline stem cells. Genes Dev. 2015, 29, 2420–2434. [Google Scholar] [CrossRef] [PubMed]
  16. Collins, K.; Mitchell, J.R. Telomerase in the human organism. Oncogene 2002, 21, 564–579. [Google Scholar] [CrossRef] [PubMed]
  17. Shay, J.W.; Wright, W.E. Telomeres and telomerase in normal and cancer stem cells. FEBS Lett. 2010, 584, 3819–3825. [Google Scholar] [CrossRef] [PubMed]
  18. Shigeishi, H.; Sugiyama, M.; Tahara, H.; Ono, S.; Kumar Bhawal, U.; Okura, M.; Kogo, M.; Shinohara, M.; Shindoh, M.; Shintani, S.; et al. Increased telomerase activity and htert expression in human salivary gland carcinomas. Oncol. Lett. 2011, 2, 845–850. [Google Scholar] [CrossRef] [PubMed]
  19. Poncet, D.; Belleville, A.; t’kint de Roodenbeke, C.; Roborel de Climens, A.; 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]
  20. Cesare, A.J.; Reddel, R.R. Alternative lengthening of telomeres: Models, mechanisms and implications. Nat. Rev. Genet. 2010, 11, 319–330. [Google Scholar] [CrossRef] [PubMed]
  21. Chiappori, A.A.; Kolevska, T.; Spigel, D.R.; Hager, S.; Rarick, M.; Gadgeel, S.; Blais, N.; Von Pawel, J.; Hart, L.; Reck, M.; et al. A randomized phase II study of the telomerase inhibitor imetelstat as maintenance therapy for advanced non-small-cell lung cancer. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2015, 26, 354–362. [Google Scholar] [CrossRef] [PubMed]
  22. Asai, A.; Oshima, Y.; Yamamoto, Y.; Uochi, T.A.; Kusaka, H.; Akinaga, S.; Yamashita, Y.; Pongracz, K.; Pruzan, R.; Wunder, E.; et al. A novel telomerase template antagonist (GRN163) as a potential anticancer agent. Cancer Res. 2003, 63, 3931–3939. [Google Scholar] [PubMed]
  23. Gellert, G.C.; Dikmen, Z.G.; Wright, W.E.; Gryaznov, S.; Shay, J.W. Effects of a novel telomerase inhibitor, GRN163L, in human breast cancer. Breast Cancer Res. Treat. 2006, 96, 73–81. [Google Scholar] [CrossRef] [PubMed]
  24. Hochreiter, A.E.; Xiao, H.; Goldblatt, E.M.; Gryaznov, S.M.; Miller, K.D.; Badve, S.; Sledge, G.W.; Herbert, B.S. Telomerase template antagonist GRN163L disrupts telomere maintenance, tumor growth, and metastasis of breast cancer. Clin. Cancer Res. 2006, 12, 3184–3192. [Google Scholar] [CrossRef]
  25. Herbert, B.; Pitts, A.E.; Baker, S.I.; Hamilton, S.E.; Wright, W.E.; Shay, J.W.; Corey, D.R. Inhibition of human telomerase in immortal human cells leads to progressive telomere shortening and cell death. Proc. Natl. Acad. Sci. USA 1999, 96, 14276–14281. [Google Scholar] [CrossRef] [PubMed]
  26. Lin, S.; Wei, J.; Wunderlich, M.; Chou, F.S.; Mulloy, J.C. Immortalization of human ae pre-leukemia cells by htert allows leukemic transformation. Oncotarget 2016, 7, 55939–55950. [Google Scholar] [CrossRef] [PubMed]
  27. Cook, B.D.; Dynek, J.N.; Chang, W.; Shostak, G.; Smith, S. Role for the related poly(ADP-Ribose) polymerases tankyrase 1 and 2 at human telomeres. Mol. Cell. Biol. 2002, 22, 332–342. [Google Scholar] [CrossRef] [PubMed]
  28. Riffell, J.L.; Lord, C.J.; Ashworth, A. Tankyrase-targeted therapeutics: Expanding opportunities in the parp family. Nat. Rev. Drug Discov. 2012, 11, 923–936. [Google Scholar] [CrossRef] [PubMed]
  29. Waaler, J.; Machon, O.; Tumova, L.; Dinh, H.; Korinek, V.; Wilson, S.R.; Paulsen, J.E.; Pedersen, N.M.; Eide, T.J.; Machonova, O.; et al. A novel tankyrase inhibitor decreases canonical wnt signaling in colon carcinoma cells and reduces tumor growth in conditional APC mutant mice. Cancer Res. 2012, 72, 2822–2832. [Google Scholar] [CrossRef] [PubMed]
  30. Dregalla, R.C.; Zhou, J.; Idate, R.R.; Battaglia, C.L.; Liber, H.L.; Bailey, S.M. Regulatory roles of tankyrase 1 at telomeres and in DNA repair: Suppression of T-SCE and stabilization of DNA-PKcs. Aging 2010, 2, 691–708. [Google Scholar] [CrossRef] [PubMed]
  31. Seimiya, H.; Muramatsu, Y.; Ohishi, T.; Tsuruo, T. Tankyrase 1 as a target for telomere-directed molecular cancer therapeutics. Cancer Cell 2005, 7, 25–37. [Google Scholar] [CrossRef] [PubMed]
  32. Fatemi, A.; Safa, M.; Kazemi, A. MST-312 induces G2/M cell cycle arrest and apoptosis in APL cells through inhibition of telomerase activity and suppression of NF-κB pathway. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2015, 36, 8425–8437. [Google Scholar] [CrossRef] [PubMed]
  33. Tauchi, T.; Shin-ya, K.; Sashida, G.; Sumi, M.; Okabe, S.; Ohyashiki, J.H.; Ohyashiki, K. Telomerase inhibition with a novel G-quadruplex-interactive agent, telomestatin: In vitro and in vivo studies in acute leukemia. Oncogene 2006, 25, 5719–5725. [Google Scholar] [CrossRef] [PubMed]
  34. Maestroni, L.; Matmati, S.; Coulon, S. Solving the telomere replication problem. Genes 2017, 8, 55. [Google Scholar] [CrossRef] [PubMed]
  35. Burger, A.M.; Dai, F.; Schultes, C.M.; Reszka, A.P.; Moore, M.J.; Double, J.A.; Neidle, S. The G-quadruplex-interactive molecule BRACO-19 inhibits tumor growth, consistent with telomere targeting and interference with telomerase function. Cancer Res. 2005, 65, 1489–1496. [Google Scholar] [CrossRef] [PubMed]
  36. Gunaratnam, M.; Greciano, O.; Martins, C.; Reszka, A.P.; Schultes, C.M.; Morjani, H.; Riou, J.F.; Neidle, S. Mechanism of acridine-based telomerase inhibition and telomere shortening. Biochem. Pharmacol. 2007, 74, 679–689. [Google Scholar] [CrossRef] [PubMed]
  37. Cookson, J.C.; Dai, F.; Smith, V.; Heald, R.A.; Laughton, C.A.; Stevens, M.F.; Burger, A.M. Pharmacodynamics of the G-quadruplex-stabilizing telomerase inhibitor 3,11-difluoro-6,8,13-trimethyl-8H-quino[4,3,2-kl]acridinium methosulfate (RHPS4) in vitro: Activity in human tumor cells correlates with telomere length and can be enhanced, or antagonized, with cytotoxic agents. Mol. Pharmacol. 2005, 68, 1551–1558. [Google Scholar] [PubMed]
  38. Salvati, E.; Leonetti, C.; Rizzo, A.; Scarsella, M.; Mottolese, M.; Galati, R.; Sperduti, I.; Stevens, M.F.; D’Incalci, M.; Blasco, M.; et al. Telomere damage induced by the G-quadruplex ligand RHPS4 has an antitumor effect. J. Clin. Investig. 2007, 117, 3236–3247. [Google Scholar] [CrossRef] [PubMed]
  39. Tahara, H.; Shin-Ya, K.; Seimiya, H.; Yamada, H.; Tsuruo, T.; Ide, T. G-Quadruplex stabilization by telomestatin induces TRF2 protein dissociation from telomeres and anaphase bridge formation accompanied by loss of the 3’ telomeric overhang in cancer cells. Oncogene 2006, 25, 1955–1966. [Google Scholar] [CrossRef] [PubMed]
  40. Li, Y.; Tergaonkar, V. Noncanonical functions of telomerase: Implications in telomerase-targeted cancer therapies. Cancer Res. 2014, 74, 1639–1644. [Google Scholar] [CrossRef] [PubMed]
  41. Low, K.C.; Tergaonkar, V. Telomerase: Central regulator of all of the hallmarks of cancer. Trends Biochem. Sci. 2013, 38, 426–434. [Google Scholar] [CrossRef] [PubMed]
  42. Indran, I.R.; Hande, M.P.; Pervaiz, S. hTERT overexpression alleviates intracellular ROS production, improves mitochondrial function, and inhibits ROS-mediated apoptosis in cancer cells. Cancer Res. 2011, 71, 266–276. [Google Scholar] [CrossRef] [PubMed]
  43. Masutomi, K.; Possemato, R.; Wong, J.M.; Currier, J.L.; Tothova, Z.; Manola, J.B.; Ganesan, S.; Lansdorp, P.M.; Collins, K.; Hahn, W.C. The telomerase reverse transcriptase regulates chromatin state and DNA damage responses. Proc. Natl. Acad. Sci. USA 2005, 102, 8222–8227. [Google Scholar] [CrossRef] [PubMed]
  44. Park, J.I.; Venteicher, A.S.; Hong, J.Y.; Choi, J.; Jun, S.; Shkreli, M.; Chang, W.; Meng, Z.; Cheung, P.; Ji, H.; et al. Telomerase modulates Wnt signalling by association with target gene chromatin. Nature 2009, 460, 66–72. [Google Scholar] [CrossRef] [PubMed]
  45. Koh, C.M.; Khattar, E.; Leow, S.C.; Liu, C.Y.; Muller, J.; Ang, W.X.; Li, Y.; Franzoso, G.; Li, S.; Guccione, E.; et al. Telomerase regulates MYC-driven oncogenesis independent of its reverse transcriptase activity. J. Clin. Investig. 2015, 125, 2109–2122. [Google Scholar] [CrossRef] [PubMed]
  46. Khattar, E.; Kumar, P.; Liu, C.Y.; Akincilar, S.C.; Raju, A.; Lakshmanan, M.; Maury, J.J.; Qiang, Y.; Li, S.; Tan, E.Y.; et al. Telomerase reverse transcriptase promotes cancer cell proliferation by augmenting tRNA expression. J. Clin. Investig. 2016, 126, 4045–4060. [Google Scholar] [CrossRef] [PubMed]
  47. Hahn, W.C.; Stewart, S.A.; Brooks, M.W.; York, S.G.; Eaton, E.; Kurachi, A.; Beijersbergen, R.L.; Knoll, J.H.; Meyerson, M.; Weinberg, R.A. Inhibition of telomerase limits the growth of human cancer cells. Nat. Med. 1999, 5, 1164–1170. [Google Scholar] [PubMed]
  48. Li, Y.; Zhou, Q.L.; Sun, W.; Chandrasekharan, P.; Cheng, H.S.; Ying, Z.; Lakshmanan, M.; Raju, A.; Tenen, D.G.; Cheng, S.Y.; et al. Non-canonical NF-kappaB signalling and ETS1/2 cooperatively drive C250T mutant TERT promoter activation. Nat. Cell biol. 2015, 17, 1327–1338. [Google Scholar] [CrossRef] [PubMed]
  49. Akincilar, S.C.; Khattar, E.; Boon, P.L.; Unal, B.; Fullwood, M.J.; Tergaonkar, V. Long-range chromatin interactions drive mutant tert promoter activation. Cancer Discov. 2016, 6, 1276–1291. [Google Scholar] [CrossRef] [PubMed]
  50. Blackburn, E.H. Telomeres and telomerase: Their mechanisms of action and the effects of altering their functions. FEBS Lett. 2005, 579, 859–862. [Google Scholar] [CrossRef] [PubMed]
  51. Vallarelli, A.F.; Rachakonda, P.S.; Andre, J.; Heidenreich, B.; Riffaud, L.; Bensussan, A.; Kumar, R.; Dumaz, N. TERT promoter mutations in melanoma render TERT expression dependent on MAPK pathway activation. Oncotarget 2016, 7, 53127–53136. [Google Scholar] [CrossRef] [PubMed]
  52. Queisser, A.; Heeg, S.; Thaler, M.; von Werder, A.; Opitz, O.G. Inhibition of telomerase induces alternative lengthening of telomeres during human esophageal carcinogenesis. Cancer Genet. 2013, 206, 374–386. [Google Scholar] [CrossRef] [PubMed]
  53. Kyo, S.; Takakura, M.; Fujiwara, T.; Inoue, M. Understanding and exploiting htert promoter regulation for diagnosis and treatment of human cancers. Cancer Sci. 2008, 99, 1528–1538. [Google Scholar] [CrossRef] [PubMed]
  54. Li, G.Z.; Eller, M.S.; Firoozabadi, R.; Gilchrest, B.A. Evidence that exposure of the telomere 3’ overhang sequence induces senescence. Proc. Natl. Acad. Sci. USA 2003, 100, 527–531. [Google Scholar] [CrossRef] [PubMed]
  55. Saretzki, G.; Sitte, N.; Merkel, U.; Wurm, R.E.; von Zglinicki, T. Telomere shortening triggers a p53-dependent cell cycle arrest via accumulation of G-rich single stranded DNA fragments. Oncogene 1999, 18, 5148–5158. [Google Scholar] [CrossRef] [PubMed]
  56. Gnanasekar, M.; Thirugnanam, S.; Zheng, G.; Chen, A.; Ramaswamy, K. T-oligo induces apoptosis in advanced prostate cancer cells. Oligonucleotides 2009, 19, 287–292. [Google Scholar] [CrossRef] [PubMed]
  57. Rankin, A.M.; Forman, L.; Sarkar, S.; Faller, D.V. Enhanced cytotoxicity from deoxyguanosine-enriched T-oligo in prostate cancer cells. Nucleic Acid Ther. 2013, 23, 311–321. [Google Scholar] [CrossRef] [PubMed]
  58. Bates, P.J.; Laber, D.A.; Miller, D.M.; Thomas, S.D.; Trent, J.O. Discovery and development of the G-rich oligonucleotide AS1411 as a novel treatment for cancer. Exp. Mol. Pathol. 2009, 86, 151–164. [Google Scholar] [CrossRef] [PubMed]
  59. Weng, D.; Cunin, M.C.; Song, B.; Price, B.D.; Eller, M.S.; Gilchrest, B.A.; Calderwood, S.K.; Gong, J. Radiosensitization of mammary carcinoma cells by telomere homolog oligonucleotide pretreatment. Breast Cancer Res. 2010, 12, R71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Longe, H.O.; Romesser, P.B.; Rankin, A.M.; Faller, D.V.; Eller, M.S.; Gilchrest, B.A.; Denis, G.V. Telomere homolog oligonucleotides induce apoptosis in malignant but not in normal lymphoid cells: Mechanism and therapeutic potential. Int. J. Cancer 2009, 124, 473–482. [Google Scholar] [CrossRef] [PubMed]
  61. Wojdyla, L.; Stone, A.L.; Sethakorn, N.; Uppada, S.B.; Devito, J.T.; Bissonnette, M.; Puri, N. T-oligo as an anticancer agent in colorectal cancer. Biochem. Biophys. Res. Commun. 2014, 446, 596–601. [Google Scholar] [CrossRef] [PubMed]
  62. Sarkar, S.; Faller, D.V. Telomere-homologous g-rich oligonucleotides sensitize human ovarian cancer cells to trail-induced growth inhibition and apoptosis. Nucleic Acid Ther. 2013, 23, 167–174. [Google Scholar] [CrossRef] [PubMed]
  63. Puri, N.; Eller, M.S.; Byers, H.R.; Dykstra, S.; Kubera, J.; Gilchrest, B.A. Telomere-based DNA damage responses: A new approach to melanoma. FASEB J. 2004, 18, 1373–1381. [Google Scholar] [CrossRef] [PubMed]
  64. Puri, N.; Pitman, R.T.; Mulnix, R.E.; Erickson, T.; Iness, A.N.; Vitali, C.; Zhao, Y.; Salgia, R. Non-small cell lung cancer is susceptible to induction of DNA damage responses and inhibition of angiogenesis by telomere overhang oligonucleotides. Cancer Lett. 2014, 343, 14–23. [Google Scholar] [CrossRef] [PubMed]
  65. Coleman, C.; Levine, D.; Kishore, R.; Qin, G.; Thorne, T.; Lambers, E.; Sasi, S.P.; Yaar, M.; Gilchrest, B.A.; Goukassian, D.A. Inhibition of melanoma angiogenesis by telomere homolog oligonucleotides. J. Oncol. 2010, 2010, 928628. [Google Scholar] [CrossRef] [PubMed]
  66. Uppada, S.B.; Erickson, T.; Wojdyla, L.; Moravec, D.N.; Song, Z.; Cheng, J.; Puri, N. Novel delivery system for T-oligo using a nanocomplex formed with an alpha helical peptide for melanoma therapy. Int. J. Nanomed. 2014, 9, 43–53. [Google Scholar]
  67. Yaar, M.; Eller, M.S.; Panova, I.; Kubera, J.; Wee, L.H.; Cowan, K.H.; Gilchrest, B.A. Telomeric DNA induces apoptosis and senescence of human breast carcinoma cells. Breast Cancer Res. 2007, 9, R13. [Google Scholar] [CrossRef] [PubMed]
  68. Rankin, A.M.; Faller, D.V.; Spanjaard, R.A. Telomerase inhibitors and ‘T-oligo’ as cancer therapeutics: Contrasting molecular mechanisms of cytotoxicity. Anti-Cancer Drugs 2008, 19, 329–338. [Google Scholar] [CrossRef] [PubMed]
  69. Karlseder, J.; Smogorzewska, A.; de Lange, T. Senescence induced by altered telomere state, not telomere loss. Science 2002, 295, 2446–2449. [Google Scholar] [CrossRef] [PubMed]
  70. 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]
  71. Pitman, R.T.; Wojdyla, L.; Puri, N. Mechanism of DNA damage responses induced by exposure to an oligonucleotide homologous to the telomere overhang in melanoma. Oncotarget 2013, 4, 761–771. [Google Scholar] [CrossRef] [PubMed]
  72. Eller, M.S.; Li, G.Z.; Firoozabadi, R.; Puri, N.; Gilchrest, B.A. Induction of a p95/Nbs1-mediated s phase checkpoint by telomere 3’ overhang specific DNA. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2003, 17, 152–162. [Google Scholar] [CrossRef] [PubMed]
  73. Page, T.J.; Mata, J.E.; Bridge, J.A.; Siebler, J.C.; Neff, J.R.; Iversen, P.L. The cytotoxic effects of single-stranded telomere mimics on OMA-BL1 cells. Exp. Cell Res. 1999, 252, 41–49. [Google Scholar] [CrossRef] [PubMed]
  74. Calado, R.T.; Dumitriu, B. Telomere dynamics in mice and humans. Semin. Hematol. 2013, 50, 165–174. [Google Scholar] [CrossRef] [PubMed]
  75. Gomes, N.M.; Ryder, O.A.; Houck, M.L.; Charter, S.J.; Walker, W.; Forsyth, N.R.; Austad, S.N.; Venditti, C.; Pagel, M.; Shay, J.W.; et al. Comparative biology of mammalian telomeres: Hypotheses on ancestral states and the roles of telomeres in longevity determination. Aging Cell 2011, 10, 761–768. [Google Scholar] [CrossRef] [PubMed]
  76. Kim, S.H.; Parrinello, S.; Kim, J.; Campisi, J. Mus musculus and mus spretus homologues of the human telomere-associated protein TIN2. Genomics 2003, 81, 422–432. [Google Scholar] [CrossRef]
  77. Broccoli, D.; Chong, L.; Oelmann, S.; Fernald, A.A.; Marziliano, N.; van Steensel, B.; Kipling, D.; Le Beau, M.M.; de Lange, T. Comparison of the human and mouse genes encoding the telomeric protein, TRF1: Chromosomal localization, expression and conserved protein domains. Hum. Mol. Genet. 1997, 6, 69–76. [Google Scholar] [CrossRef] [PubMed]
  78. Palm, W.; Hockemeyer, D.; Kibe, T.; de Lange, T. Functional dissection of human and mouse POT1 proteins. Mol. Cell. Biol. 2009, 29, 471–482. [Google Scholar] [CrossRef] [PubMed]
  79. Tian, X.; Dai, S.; Sun, J.; Jiang, S.; Sui, C.; Meng, F.; Li, Y.; Fu, L.; Jiang, T.; Wang, Y.; et al. Bufalin induces mitochondria-dependent apoptosis in pancreatic and oral cancer cells by downregulating hTERT expression via activation of the JNK/p38 pathway. Evide-Based Complement. Altern. Med. 2015, 2015, 546210. [Google Scholar] [CrossRef] [PubMed]
  80. Eller, M.S.; Liao, X.; Liu, S.; Hanna, K.; Backvall, H.; Opresko, P.L.; Bohr, V.A.; Gilchrest, B.A. A role for WRN in telomere-based DNA damage responses. Proc. Natl. Acad. Sci. USA 2006, 103, 15073–15078. [Google Scholar] [CrossRef] [PubMed]
  81. Lu, H.; Wang, J.; Bai, Y.; Lang, J.W.; Liu, S.; Lin, Y.; Cheng, J. Ionic polypeptides with unusual helical stability. Nat. Commun. 2011, 2, 206. [Google Scholar] [CrossRef] [PubMed]
  82. Yen, J.; Zhang, Y.; Gabrielson, N.P.; Yin, L.; Guan, L.; Chaudhury, I.; Lu, H.; Wang, F.; Cheng, J. Cationic, helical polypeptide-based gene delivery for IMR-90 fibroblasts and human embryonic stem cells. Biomater. Sci. 2013, 1, 719–727. [Google Scholar] [CrossRef] [PubMed]
  83. Wan, C.; Allen, T.M.; Cullis, P.R. Lipid nanoparticle delivery systems for siRNA-based therapeutics. Drug Deliv. Transl. Res. 2014, 4, 74–83. [Google Scholar] [CrossRef] [PubMed]
  84. Wissing, S.A.; Kayser, O.; Muller, R.H. Solid lipid nanoparticles for parenteral drug delivery. Adv. Drug Deliv. Rev. 2004, 56, 1257–1272. [Google Scholar] [CrossRef] [PubMed]
  85. Rajanna, S.; Rastogi, I.; Wojdyla, L.; Furo, H.; Kulesza, A.; Lin, L.; Sheu, B.; Frakes, M.; Ivanovich, M.; Puri, N. Current molecularly targeting therapies in NSCLC and melanoma. Anti-Cancer Agents Med. Chem. 2015, 15, 856–868. [Google Scholar] [CrossRef]
Figure 1. (a) Telomeric DNA has the ability to fold over itself, forming what is called the T-loop. Furthermore, the 3’ single-stranded overhang can tuck under itself, forming what is called the D-loop. Additionally, there are a group of regulatory proteins attached to the telomere at several locations, called the shelterin complex, which maintain telomere homeostasis; (b) Telomeric DNA forms intramolecular G-quadruplexes.
Figure 1. (a) Telomeric DNA has the ability to fold over itself, forming what is called the T-loop. Furthermore, the 3’ single-stranded overhang can tuck under itself, forming what is called the D-loop. Additionally, there are a group of regulatory proteins attached to the telomere at several locations, called the shelterin complex, which maintain telomere homeostasis; (b) Telomeric DNA forms intramolecular G-quadruplexes.
Antioxidants 06 00015 g001
Figure 2. (a) The shelterin dissociation model hypothesizes that the presence of T11 in the nucleus causes the disruption of the telomere, particularly the shelterin complex. This model suggests that T11 competes with telomeric DNA for the binding of shelterin proteins; (b) Key proteins of the shelterin complex are displaced from the telomere and/or bind to T11. This disruption leaves the telomere exposed and initiates DDRs.
Figure 2. (a) The shelterin dissociation model hypothesizes that the presence of T11 in the nucleus causes the disruption of the telomere, particularly the shelterin complex. This model suggests that T11 competes with telomeric DNA for the binding of shelterin proteins; (b) Key proteins of the shelterin complex are displaced from the telomere and/or bind to T11. This disruption leaves the telomere exposed and initiates DDRs.
Antioxidants 06 00015 g002
Figure 3. (a) The exposed telomere mimicry model hypothesizes that exogenous T11 mimics the exposed endogenous telomere overhang; (b) The cell responds to this mimicry of the telomere overhang exposure by initiating DDRs in cancer cells and triggering upregulation of shelterin proteins in attempt to overcome apparent telomere exposure.
Figure 3. (a) The exposed telomere mimicry model hypothesizes that exogenous T11 mimics the exposed endogenous telomere overhang; (b) The cell responds to this mimicry of the telomere overhang exposure by initiating DDRs in cancer cells and triggering upregulation of shelterin proteins in attempt to overcome apparent telomere exposure.
Antioxidants 06 00015 g003

Share and Cite

MDPI and ACS Style

Ivancich, M.; Schrank, Z.; Wojdyla, L.; Leviskas, B.; Kuckovic, A.; Sanjali, A.; Puri, N. Treating Cancer by Targeting Telomeres and Telomerase. Antioxidants 2017, 6, 15. https://doi.org/10.3390/antiox6010015

AMA Style

Ivancich M, Schrank Z, Wojdyla L, Leviskas B, Kuckovic A, Sanjali A, Puri N. Treating Cancer by Targeting Telomeres and Telomerase. Antioxidants. 2017; 6(1):15. https://doi.org/10.3390/antiox6010015

Chicago/Turabian Style

Ivancich, Marko, Zachary Schrank, Luke Wojdyla, Brandon Leviskas, Adijan Kuckovic, Ankita Sanjali, and Neelu Puri. 2017. "Treating Cancer by Targeting Telomeres and Telomerase" Antioxidants 6, no. 1: 15. https://doi.org/10.3390/antiox6010015

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