SlideShare ist ein Scribd-Unternehmen logo
1 von 18
Downloaden Sie, um offline zu lesen
Telomere Maintenance as a Target for Drug Discovery
Vijay Sekaran, Joana Soares, and Michael B. Jarstfer*
Division of Chemical Biology and Medicinal Chemistry, Eshelman School of Pharmacy, University of North Carolina at Chapel Hill,
Chapel Hill, North Carolina 27599, United States
ABSTRACT: The observation that the enzyme telomerase is up-regulated in 80−90% of
cancer cells isolated from primary human tumors but is absent in neighboring cells of
healthy tissue has resulted in significant efforts to validate telomerase as an anticancer drug
target and to develop effective approaches toward its inhibition. In addition to inhibitors
that target the enzymatic function of telomerase, efforts toward immunotherapy using
peptides derived from its catalytic subunit hTERT and hTERT-promoter driven gene
therapy have made significant advances. The increased level of telomerase in cancer cells
also provides a potential platform for cancer diagnostics. Telomerase inhibition leads to
disruption of a cell’s ability to maintain the very ends of the chromosomes, which are called
telomeres. Thus, the telomere itself has also attracted attention as an anticancer drug target.
In this Perspective, interdisciplinary efforts to realize the therapeutic potential of targeting
telomere maintenance with a focus on telomerase are discussed.
■ INTRODUCTION
Pathways exhibiting differential activity in cancer cells when
compared to healthy cells are attractive targets for therapeutic
intervention. Accordingly, the observation that the enzyme
telomerase is up-regulated in 80−90% of all cancer cells isolated
from primary human tumors but is absent in neighboring cells
of healthy tissue has resulted in significant efforts to validate
telomerase as an anticancer drug target and to develop effective
approaches toward its inhibition.1
In addition to inhibitors that
target the enzymatic function of telomerase, efforts toward
immunotherapy using peptides derived from the catalytic
telomerase subunit hTERT (human telomerase reverse tran-
scriptase)2
and hTERT-promoter driven gene therapy3
have
made significant advances. The increased level of telomerase
activity in cancer cells also provides a potential platform for
cancer diagnostics.4
Telomerase inhibition leads to disruption
of a cell’s ability to maintain the very ends of the chromosomes,
which are called telomeres. Thus, the telomere itself has also
attracted attention as an anticancer drug target. Here, we
discuss the interdisciplinary efforts to realize the therapeutic
potential of targeting telomere maintenance with a specific
focus on telomerase.
■ TELOMERE MAINTENANCE AND TELOMERASE IN
HUMAN CANCER CELLS
Telomeres are the physical ends of linear chromosomes, and
their importance to human health is underscored by the Nobel
Prize in Physiology or Medicine awarded to Elizabeth
Blackburn, Jack Szostak, and Carol Grieder in 2009 for their
work on telomere biology and biochemistry. Telomeres are so
important because proper telomere maintenance is an absolute
requirement for continued cellular proliferation. Functional
telomeres require a specific and repetitive DNA sequence and
several telomere-specific proteins.5,6
These telomere-associated
proteins function in two primary capacities: maintaining the
structural and functional integrity of the telomere and ensuring
complete replication of telomeric DNA. In the absence of
functional telomeres, chromosome ends are recognized as DNA
damage, and the ensuing DNA damage response generally
results in either senescence or apoptosis, depending on the cell
type.7
Mammalian telomeric DNA contains several thousand
base pairs of repetitive DNA with the sequence 5′-
TTAGGG/3′-AATCCC and a single stranded 3′ overhang of
the G-rich sequence.8,9
The canonical DNA polymerases
responsible for DNA replication cannot efficiently replicate
the very ends of linear DNA templates. As a result, proliferative
cells must employ additional pathways to overcome this end-
replication problem. In human cells, this additional activity
resides primarily in telomerase. Telomerase is a ribonucleo-
protein complex that contains several protein subunits and one
RNA subunit.10,11
Human telomerase RNA (hTER) is a highly
structured 451-nucleotide long RNA that contains several
functional domains in addition to a templating sequence used
to code for the G-rich telomeric DNA strand.12
The protein
subunit hTERT functions as a reverse transcriptase by using the
templating portion of hTER to synthesize telomeric DNA.13
In
the majority of cancer cells, hTERT is up-regulated, increasing
the telomerase activity required for continuous proliferation.8,14
In contrast, telomerase activity is down-regulated in most
normal human diploid cells, with the exception of germline and
stem cells, primarily owing to inhibition of hTERT expression
during cellular differentiation.
Because telomerase activity is absent in most differentiated
human cells, telomeres erode during the cellular aging process.
This erosion allows telomere length to function like a mitotic
hourglass that regulates cellular replication and arguably
functions as a tumor suppressor mechanism.5
Cells lacking
Received: April 11, 2013
Perspective
pubs.acs.org/jmc
© XXXX American Chemical Society A dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXX
telomerase activity experience erosion of their telomeric DNA
when they divide. At a critical length, the shortened telomeres
are recognized as DNA damage and form a telomere
dysfunction induced focus (TIF). Typically, cells respond to
TIFs through senescence or apoptosis. Importantly, the
telomere-length dependence of proliferation appears to be a
result of the structure and not the length of the telomeric DNA
per se. This is evidenced in part by experiments in the de Lange
lab showing that the negative consequences of short telomeres
can be overcome by the overexpression of the telomere-binding
protein TRF2.15
It is likely that as telomeres erode in cells
lacking an active telomere maintenance pathway, the telomere
structure, predicted to be a lasso-like fold back structure termed
a t-loop,16
becomes increasingly difficult to sustain eventually
leading to TIFs and senescence or apoptosis if telomere
function is not reestablished.6
Telomere maintenance directly impacts cancer in two
specific ways. The most obvious is the absolute requirement
for telomeric DNA maintenance to offset telomere loss from
the end-replication problem, processing, and degradation. Most
cancer cells up-regulate expression of hTERT, in order to
increase telomerase activity.17
In 10−15% of cancers, a
recombination based telomere lengthening pathway called
alternative lengthening of telomeres (ALT) supplants the need
for telomerase.18
The ALT pathway is particularly prevalent in
astrocytic brain tumors and osteosarcomas.19
A second
connection between telomere biology and neoplasias is the
effect of overall length of telomeric DNA and the actions of
telomere-associated proteins on genetic stability. Telomere
erosion has been argued to serve as a tumor suppression
mechanism because it can activate the senescence pathway,
which suppresses neoplastic transformation.5
However, even in
the presence of telomere erosion, senescence is not activated in
the absence of the Rb/p16 or p53 pathways, leading to genetic
instability and increased potential for transformation. These
pivotal roles of telomere maintenance in tumorigenicity are
evidenced by the ability to inhibit tumor growth by inhibiting
telomerase in cancer cells and by the ability to convert primary
human cells to cancer cells by combining constitutive
telomerase expression with oncogene expression and/or loss
of a tumor repressor gene. For example, the simian virus 40
large-T oncoprotein together with an oncogenic allele of H-ras
and ectopic expression of hTERT has been demonstrated to
transform a variety of cell types.20
For more comprehensive
discussions on the roles of telomerase and telomere
maintenance on tumor biology, several recent reviews are
recommended.21−24
■ TARGETING TELOMERE MAINTENANCE FOR
CANCER THERAPY
When considering anticancer drug targets, several advantages
for targeting telomerase and telomere maintenance exist when
compared to other targets and pathways. Most important is the
apparently universal requirement for telomere maintenance in
cancer cells including putative cancer stem cells.25,26
In
addition, normal human cells including stem cells express
lower telomerase activity and generally maintain telomeres at
longer lengths when compared to cancer cells.27,28
These
features provide a level of specificity with telomerase inhibition.
Telomerase activity itself was validated as an anticancer drug
target in 1999 when two independent studies reported that
overexpressing dominant-negative-hTERT mutants resulted in
telomerase inhibition, telomere shortening, morphological
changes, growth arrest, and apoptosis in a variety of tumor
cell lines.29,30
Since that time, several strategies for targeting
telomere maintenance and telomerase-positive cancer cells have
been explored (Figure 1). The initial approach explored was the
direct inhibition of telomerase’s enzymatic activity. As
validation of the approach, the telomerase inhibitor imetelstat
(1, GRN163L) has been explored in anticancer therapy in
several clinical trials.31
Direct targeting of the telomere itself has
also been accomplished with specific DNA binding agents that
bind to and stabilize the G-quadruplex structures formed by
telomeric DNA.32,33
The high activity of the hTERT promoter
in cancer cells has resulted in the exploration of hTERT-
promoter driven expression of cancer killing genes and
viruses,34
and one clinical trial for hTERT-driven replication
of an oncolytic virus is underway.35
Immunotherapy targeting
hTERT-derived peptides has also been explored leading to
several clinical trials.36
Each of these platforms will be discussed
further below.
■ CONSEQUENCE OF TELOMERASE INHIBITION
AND TELOMERE DISRUPTION
If telomere maintenance is to be considered a viable anticancer
drug discovery platform, it is paramount to understand the
consequences of disrupting telomere maintenance for both
healthy and cancerous cells. The simplest and best explored
telomere-targeted anticancer drug platform is the inhibition of
telomerase enzymatic activity. When telomerase activity is
inhibited, DNA replication attending cellular proliferation
results in telomeric DNA erosion.29,37
Erosion of telomeric
DNA eventually results in dysfunctional telomeres that initiate
TIF formation.5,6
Generally, proliferating human cells lose 50−
200 base pairs of telomeric DNA during each cell cycle in the
absence of telomerase activity.38
Therefore, a lag between the
phenotypic consequences of telomerase inhibition and the
initiation of telomerase inhibition therapy is likely to occur. The
duration of this phenotypic lag depends on both the initial
length of telomeric DNA, which ranges from 5000 to 15 000
base pairs in human cancer cells, and the rate of erosion.
Importantly, it is suspected that the length of a few telomeres,
Figure 1. Many ways to target the overexpression of telomerase in
cancer cells have been developed.
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXB
not the average length of all telomeres, dictates the cellular
response to telomerase inhibition.39
This suggests that some
tumor types may be especially sensitive to telomerase
inhibition. Once telomeres erode to a critically short length,
they cannot function properly,40
perhaps because they cannot
be folded into the proper protected state, presumably a t-
loop.16
In most cancer cells, the loss of telomere function
results in apoptosis, whereas in normal cells, telomere loss
results in cell cycle arrest. Notably, telomerase inhibition has
also been observed to promote the ALT phenotype in tumors
concomitant with up-regulation of the PGC-1β, a master
regulator of mitochondrial biogenesis.41
This report offers
critical insight into the mechanisms of adaptation that may be
in play when tumors are subjected to telomerase inhibition.
Inhibition of telomerase by some methods, for example,
antisense oligonucleotides targeting the hTERT mRNA,42
can
result in immediate effects that are not preceded by the
phenotypic lag characteristic of telomerase activity inhibitors
and are not coupled to erosion of telomeric DNA. While a
precise explanation remains missing, one intriguing possibility is
that hTERT has several noncanonical properties that are
separate from its enzymatic activity, and inhibition of these may
have immediate effects.43
Proposed additional activities of
telomerase include telomere capping, inhibition of apoptosis,
activation of the Wnt pathway, and activation of the NFκB
pathway.44,45
Thus, down-regulation of hTERT can potentially
lead to immediate antiproliferative effects. Understanding why
some cells are immediately impacted by the loss of hTERT and
others are not remains to be determined. Perhaps some cancer
cells become oncogenically addicted to hTERT and require
hTERT to maintain elevated growth signaling.
Currently, one of the most studied small-molecule
telomerase inhibitors are G-quadruplex stabilizers.46
This class
of molecules can impact telomere functions in at least two
primary mechanisms. One is by inhibiting the enzymatic
activity of telomerase by preventing telomerase from binding to
its primer or by interacting with the nascent DNA telomerase
product. The second is to dissociate or block telomere-binding
proteins from the telomere, disrupting the formation of the
protective cap. As expected, the former mechanism results in a
delayed but highly telomerase-expression-specific effect while
the second mechanism is more immediate but may suffer from
reduced specificity because formation of proper telomere
structure is required in all cells including normal somatic and
stem cells. Early evidence suggests that the therapeutic index
for telomere disrupters is surprisingly high at least in a mouse
model.47,48
One potential reason that telomere disruption may
be well tolerated in normal cells is that the DNA damage
response at the telomere may be transient and reversible.
Evidence for this hypothesis comes from experiments using a
ligand-stabilized version of the telomere binding protein
Protection of Telomeres 1 (POT1). POT1 is essential for
telomere protection, and the ligand-stabilized POT1 allowed
POT1 to be inhibited transiently and reversibly. In cancer cells,
POT1 inhibition resulted in cell death. But in normal cells,
POT1 inhibition resulted in arrest that was reversed by
reactivating POT1.49
■ SMALL-MOLECULE TELOMERASE INHIBITORS
Once the importance of telomerase activity for cancer cell
growth was recognized, several academic and pharmaceutical
groups initiated programs to discover small molecule
telomerase inhibitors. Approaches toward the identification of
effective telomerase inhibitors have included repurposing of
viral reverse transcriptase inhibitors, screening of natural
products, high throughput screens of diverse compound
libraries, and more recently, the beginnings of structure-based
discovery and optimization. Despite these efforts, small
molecule telomerase inhibitors have yet to reach the clinic. In
this section, the varieties of approaches toward direct
telomerase activity inhibition with small molecules will be
discussed.
■ NUCLEOSIDE ANALOGUES
Nucleoside analogues developed as viral reverse transcriptase
inhibitors have been investigated as telomerase inhibitors
(Figure 2). Nucleoside analogues can hypothetically affect the
telomere through several mechanisms, including telomerase
inhibition and chain termination that can lead to telomere
shortening and perturbation of telomere sequence that can lead
to disruption of telomere function. The chain terminator 3′-
azido-2′,3′-dideoxythymidine (2, AZT), the first nucleoside
reverse transcriptase inhibitor approved for the treatment of
HIV-1, inhibits telomerase in vitro and in vivo, giving rise to
telomere shortening.50,51
In cultured cells, treatment with 2
resulted in telomere shortening, cell-cycle perturbation, and
increased p14ARF expression,52
and 2 exhibited synergistic
effects when combined with other chemotherapeutic
drugs.53−55
Several nucleoside HIV reverse transcriptase
inhibitors including stavudine, tenofovir, didanosine, and
abacavir also inhibit telomerase activity and lead to telomere
erosion in cultured cells, but non-nucleoside HIV RT inhibitors
did not inhibit telomerase in these studies.56
HIV patients have
been taking these and other nucleoside reverse transcriptase
inhibitors for decades, leading to possible effects from long-
term telomerase inhibition. This potential side effect has been
investigated in a small patient population, and it was found that
telomere length was inversely correlated with total time
patients were exposed to nucleoside reverse transcriptase
inhibitors.57
Although other nucleoside triphosphate analogues,
including arabinofuranylguanosine and 6-thio-7-deaza-2′-deox-
yguanosine 5′-triphosphate, ddGTP, ddATP, and ddTTP,
produced substantial inhibition of telomerase and telomere
shortening, they have not been further tested in cell cultures or
in animal models and are likely to be nonspecific, thus limiting
Figure 2. Representative nucleoside analogues identified as telomerase inhibitors.
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXC
their clinical utility.51,58
As telomerase structural biology
increases, it may be possible to rationally design nucleoside
analogues with the requisite telomerase specificity for this
platform to be viable.
■ NON-NUCLEOSIDE ANALOGUES
Non-nucleoside, small-molecule telomerase inhibitors have
been identified by a variety of approaches. High throughput
screening of chemical libraries has been used to identify
telomerase inhibitors (Figure 3). The isothiazolone derivative
2-[3-(trifluoromethyl)phenyl]isothiazolin-3-one (3, IC50 = 1
μM) was identified from a screen of a 16 000-member, chemical
library using rat telomerase (Figure 3).59
The reducing agent
dithiothreitol was found to protect telomerase from inhibition
by 3, suggesting that inhibition resulted from interaction
between the isothiazolone moiety of 3 with the sulfhydryl
group of key cysteine residue(s).59
The nitrostyrene derivative
3-(3,5-dichlorophenoxy)nitrostyrene (4, IC50 = 0.4 μM)60
and
the quinoxaline derivative 2,3,7-trichloro-5-nitroquinoxaline
(5),61
two noncompetitive telomerase inhibitors, were also
identified by screening chemical libraries. At low concentrations
(1 μM), 4 was not acutely cytotoxic to HeLa cells, but chronic
exposure led to progressive telomere erosion followed by the
induction of senescence consistent with telomerase inhibition.60
Long-term treatment of the breast cancer cell line MCF7 with 5
(1 μM) also resulted in telomere erosion, chromosome
abnormalities, and senescence.61
The crystal structure of TERT from the beetle Tribolium
castaneum62
has been used in limited structure based design of
telomerase inhibitors. Because 2H-pyrazole and dihydropyr-
azole derivatives have been shown to have strong anticancer,
antmalarial, antiviral, and anti-inflammatory activity, they were
selected as a starting scaffold for telomerase inhibition.63−65
Aryl-2H-pyrazoles with 1,3-benzodioxoles and electron with-
drawing aryls groups and sulfonyl-4,5-dihydropyrazoles dis-
played telomerase inhibition with IC50 values ranging from 0.8
to 10 μM in a variety of cancer cells (Figure 4).63−65
Since
previous studies have shown that tetracyclic coumarins have
notable anti-HIV activities and that the HIV and telomerase
reverse transcriptase proteins display structural homology,
dihydropyrazolepiperidine derivatives containing coumarin
moieties were also synthesized.63,66
These molecules showed
anticancer activity against gastric cancer and the most potent
inhibited telomerase in crude cell extracts with IC50 values in
the 1−2 μM range.63,66
Results from docking studies were used
to suggest that these molecules bind the hTERT active site
through key intermolecular hydrogen bonding and π−cation
interactions with invariant residues Lys189, Asp254, and
Gln308 and by projecting aryl and dihyrdopyrazole rings into
the hydrophobic dNTP-binding pocket.63,65,66
The π−cation
interaction is consistent with the structure−activity relationship
(SAR) trend in which electron withdrawing groups on the
aromatic ring attenuate inhibition, whereas electron donating
groups increased inhibition. 3D quantitative SAR (QSAR) was
used to rationalize the proposed binding mechanism, and a
correlation coefficient of r2
= 0.966 was found for predicted and
observed IC50 for a small, chemically similar test set.65
Other
telomerase inhibitors proposed to bind telomerase using
docking studies include 2-chloropyridines and 2-aminomethyl-
5-(quinolin-2-yl)-1,3,4-oxadiazole-2(3H)-thionequinoline de-
rivatives, although these studies did not offer strong validation
of the binding models.67,68
Several natural products have been identified as telomerase
inhibitors as well (Figure 5). Helenalin (6), a natural
sesquiterpene lactone, was found to inhibit telomerase
apparently by reaction with a cysteine residue similar to 3.69
o-Quinone containing compounds, for example, tanshinone 2A
(7), a plant diterpene known to exhibit diverse pharmacological
activities, also inhibit telomerase (IC50 = 3 μM). The quinone
was fully responsible for inhibition that was time- and DTT-
dependent. It was also found that 7 undergoes redox cycling to
produce hydrogen peroxide. The data suggest that quinone
redox cycling inhibits telomerase, perhaps by oxidation of
conserved cysteine residues.70
Another apparently reversible
inhibitor is 8 (U-73122), commonly used as an inhibitor of
Figure 3. Examples of small-molecule telomerase inhibitors identified
by screening small molecule libraries.
Figure 4. Telomerase inhibitors reported to bind the telomerase active site based on docking studies.
Figure 5. Natural product and natural product derivatives identified as
telomerase inhibitors.
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXD
phospholipase C, which was reported to be a potent (IC50 = 0.2
μM) telomerase inhibitor.71
Inhibition was shown to be
dependent on the pyrrole-2,5-dione functional group, suggest-
ing potential alkylation of telomerase. β-Rubromycin (9), a
quinone antibiotic, inhibited telomerase with an IC50 of 3 μM.72
Mechanistically, 9 appears to be a competitive inhibitor versus
the telomerase primer, it also inhibits retroviral reverse
transcriptases, mammalian DNA polymerases, and terminal
deoxynucleotidyl transferase,73
indicating that rubromycins are
not telomerase specific.
The search for more biologically effective telomerase
inhibitors has encouraged innovative yeast-based screens.
Natural products chrolactomycin (10) and UCS1025A (11)
were identified as direct telomerase inhibitors using a forward
chemical genetics screen in a yeast system (Figure 6).74
A yeast
strain with artificially shortened telomeres was used to screen a
microbial metabolite library. Since telomeres were shortened,
the strain was more sensitive to telomerase inhibition than wild
type. Therefore, compounds with selective activity in the
telomere shortened strain were more likely to be telomerase
inhibitors. With this assay, two new telomerase inhibitors 10
(IC50 = 0.5 μM) and 11 (IC50 = 1.3 μM) were identified. In
addition, the Hsp90 inhibitor radicicol was also identified as a
telomerase maintenance inhibitor in screen.74
The synthesis of
11 has been accomplished, but structure−activity relationships
of analogues have not been reported.75,76
More recently, an
alternative yeast based assay was developed using recombinant
human telomerase. Expression of human telomerase activity in
yeast results in a growth delay; therefore, potential inhibitors of
human telomerase could rescue the transfected cells. By use of
this screen, several potential telomerase inhibitors with
moderate potency (IC50 values of ∼1−10 μM) were identified
to confirm the screening platform.77
The most potent
compound identified in the screen was the dichlorothiophene
12 (CD11359).
One promising class of small molecule telomerase inhibitors
is the potent enoylaminobenzoic acids represented by 13
(BIBR1532, IC50 = 0.093 μM) and 14 (BIBR1591, IC50 = 0.47
μM) that function as selective mixed-type, noncompetitive
telomerase inhibitors and were developed from systematic
structure−activity correlations (Figure 7).37
Limited structure−
activity relationship data are publically available and show that
the α−β unsaturated amide is essential and modifications to the
naphthyl group are not well tolerated.78
The mechanism of
inhibition by 13 is distinct from those by other non-nucleoside
compounds or antisense oligonucleotides, as it does not cause
chain termination events but rather inhibits the formation of
long reaction products. Compound 13 seems to act allosteri-
cally to impair the unique ability of telomerase to catalyze
addition of multiple telomeric repeats onto a DNA primer.79
Acute treatment with high concentrations of 13 caused direct
cytotoxic effects on malignant cells of the hematopoietic
system, resulting in end to end fusions, phosphorylated p53,
and decreased expression of TRF2, consistent with a telomere-
associated effect.80
Compound 13 is also capable of sensitizing
drug-resistant cells to chemotherapeutic agents.81
Chronic
Figure 6. Small molecules identified as telomerase inhibitors utilizing
yeast-based assays.
Figure 7. Treatment with 14 reversibly inhibits telomere length maintenance. (A) Structures of 13 and 14. (B) Chronic treatment of NCI-H460
cells with 14 results in growth arrest that can be reversed after ending the treatment. Open circles are untreated cells. Filled triangles are cells treated
with 14. Treatment was halted at day 220; note the return to proliferative growth. (C) Southern blot analysis of average telomere length of
telomerase inhibitor treated cells and cells released from telomerase inhibition. Treatment with 14 resulted in loss of telomere length and exhibited
∼1.5 kb of telomeric DNA (lanes 1 and 2). Removal of 14 resulted in telomere elongation to ∼3 kb of telomeric DNA (lane 3). Untreated cells are
shown in lane 4 for comparison and have 3 kb of telomeric DNA. Parts B and C are reprinted by permission from Macmillan Publishers Ltd.: The
EMBO Journal (http://www.nature.com/emboj/index.html) (Damm, K.; Hemmann, U.; Garin-Chesa, P.; Hauel, N.; Kauffmann, I.; Priepke, H.;
Niestroj, C.; Daiber, C.; Enenkel, B.; Guilliard, B.; Lauritsch, I.; Muller, E.; Pascolo, E.; Sauter, G.; Pantic, M.; Martens, U. M.; Wenz, C.; Lingner, J.;
Kraut, N.; Rettig, W. J.; Schnapp, A. EMBO J. 2001, 20, 6958−6968),37
copyright 2001.
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXE
exposure to lower doses of 14 resulted in progressive erosion of
telomeric DNA leading to senescence.37
However, the lag
phase between telomerase inhibition and the impact on
proliferative capacity under the chronic treatment condition is
a major concern,37
as is the apparently high protein binding
evidenced by the significantly decreased potency in crude cell
extracts compared to purified telomerase.78
More than 120
days, approximately 135 population doublings, was necessary
for treatment with 13 to inhibit cell proliferation of NCI-H460
lung cancer.82
Interestingly, washout of 14 in telomerase
positive cells allowed recovery of telomere length (Figure 7).37
■ SMALL MOLECULES THAT TARGET hTER AND THE
hTER-TELOMERIC DNA HETERODUPLEX
In addition to targeting the enzymatic activity of telomerase,
exploratory efforts in identifying ligands that target hTER
suggest that RNA-binding ligands may be a productive platform
for targeting telomerase. In two separate reports, oligonucleo-
tides were used to identify regions of hTER that can be
specifically targeted to prevent proper telomerase assem-
blage.83,84
Specifically, the CR4-CR5 and pseudoknot domains
of hTER are susceptible. Detailed analysis demonstrated that
discrete domains of hTER are blocked without losing the ability
of hTERT to associate with hTER.84
This suggests that the
CR4-CR5 and pseudoknot domain targeting oligonucleotides
effectively cause formation of a dysfunctional telomerase
complex. Subsequently, small molecules that bind hTER have
been identified,85−87
and several of these including amino-
glycosides and the bisbenzimide Heochst 33258 have been
shown to inhibit telomerase,86
presumably by binding hTER
and perturbing proper assemblage of the holoenzyme complex
(Figure 8). Efforts to target the telomerase template RNA−
DNA primer/nascent product heteroduplex have also been
reported.88,89
In these studies, several intercalators were
identified that inhibit telomerase and bind preferentially to
the RNA−DNA heteroduplex as opposed to G-quadruplex
DNA. Advances in this area include second generation
intercalators and bis-intercalators that are fairly potent
telomerase inhibitors in cell free assays. A small library of bis-
ethidium derivatives with varying linkers was screened.
Surprisingly, the library exhibited a narrow range of moderately
potent telomerase inhibition without significant improvement
over the parent monointercalator.88
Data for cell based assays
using small molecule targeting of hTER or the hTER/DNA
heteroduplex have not yet been reported, so it is still too early
to tell if this approach will be clinical viable.
■ TELOMERASE INHIBITION BY INDIRECT MEANS
hTERT Transcription Regulators. Cellular telomerase
activity is regulated at various levels including transcription,
mRNA splicing, post-transcriptional and post-translational
modifications of hTER and hTERT, transport and subcellular
localization, assembly of active telomerase ribonucleoprotein,
and telomeric DNA accessibility. Transcriptional regulation of
hTERT is thought to be the major mechanism of telomerase
regulation.90
The hTERT promoter contains binding sites for
several transcription factors, suggesting that the regulation of
hTERT expression may be subject to multiple levels of control
by various factors and dependent on cellular contexts.91
A
primary activator of hTERT is the c-Myc oncogene, a known
regulator of cell proliferation, growth, and apoptosis.92
Targets
of the Myc family are involved in various aspects of cellular
functions, including cell cycle, growth, differentiation, and life
span. c-Myc expression correlates with hTERT, and both are
Figure 8. Ligands that inhibit telomerase by binding to the RNA subunit hTER (Hoechst 33258 and neomycin) and the telomerase DNA/RNA
heteroduplex (bis-ethidium compounds).
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXF
up-regulated in highly proliferative and immortal cells but
down-regulated during terminal differentiation and in non-
proliferative cells. Therefore, it is not surprising that inhibitors
of c-Myc, butein,93
gamboic acid,94
and genistein,95
also inhibit
telomerase. Additionally, genistein treatment resulted in down-
regulation of the protein kinase Akt and thus hTERT
phosphorylation, indicating that genistein can inhibit telomer-
ase by at least two mechansisms: decreased hTERT tran-
scription and decreased posttranslational activation of
hTERT.95
Second generation tyrosine kinase inhibitors and the lipid-
derived signaling molecule ceramide have been shown to inhibit
hTERT promoter activity by altering the functions of Sp1 and
Sp3.96,97
Sp1 has been shown to collaborate with c-Myc in
order to initiate hTERT transcription.98
Imatinib, dasatinib, and
nilotininb decrease telomerase activity and hTERT expression
by decreasing Sp1 nuclear levels and activation and reducing
the nuclear translocation of hTERT via suppression of Akt
phosphorylation.96
Agents that act on hormonal pathways have
also been linked to telomerase activity regulation. For example,
estrogen activates telomerase in hormone-sensitive tissues such
as mammary and ovary epithelial cells,99,100
while the selective
estrogen receptor modulator tamoxifen reduces telomerase
activity in cells where it functions as an estrogen antago-
nist.101,102
Progesterone also has an antagonistic effect on
estrogen-induced hTERT expression.103
These observations
highlight the fact that the polypharmacological effects of many
anticancer drugs are anticipated to include inhibition of
telomere maintenance as a mechanism of their anticancer
effects.
Negative Regulators of hTERT Transcription. The lack
of telomerase activity in somatic cells appears to be a
consequence of transcriptional repression of the hTERT
expression. The loss of transcriptional repression results in
the up-regulation of hTERT expression and telomerase activity
as seen during carcinogenesis and cellular immortalization.
Several transcriptional repressors of hTERT have been
identified. One such negative regulator of hTERT transcription
is Mad1. Mad1, c-Myc, and Max are part of a family of
transcription factors that bind to E-box-containing promoters
to either down-regulate or up-regulate gene expression. Mad1 is
up-regulated in telomerase negative somatic cells, c-Myc is up-
regulated in cancer cell lines, and Max is expressed
ubiquitously.104
Consistent with its role in regulating
telomerase, overexpression of Mad1 resulted in decreased
hTERT promoter activity in gene reporter assays.105
Another negative regulator of hTERT transcription is the
tumor suppressor protein p53, which induces cell cycle arrest or
apoptosis in response to DNA damage and other cellular
stresses to inhibit tumor formation. p53 has been shown to
inhibit telomerase activity through the repression of hTERT
transcription.106
Repression of telomerase activity was also
observed by overexpressing pRB in human squamous cell
carcinomas107
and E2F1 in head and neck squamous cell
carcinomas.108
pRB and E2F1 have the potential to transcrip-
tionally repress hTERT activity either independently or in
tandem with one another. Another tumor suppressor involved
in the transcriptional repression of hTERT is Wilm’s tumor
protein 1 (WT1). WT1 has also been shown to significantly
decrease hTERT mRNA and telomerase activity in HEK293
cells.109
The role of WT1 in hTERT repression is cell type
dependent, as the WT1 gene is expressed only in specific cells
such as the kidney, gonad, and spleen.
Other negative regulators of hTERT transcription include
the myeloid cell-specific zinc finger protein 2, interferon-α, and
TGF-β. Interferon-α- and -γ-induced sensitization to apoptosis
is associated with repressed transcriptional activity of the
hTERT promoter in multiple myeloma.110
Transforming
growth factor β, a potent tumor suppressor, inhibits telomerase
through the transcription factors SMAD3 and E2F.111,112
TGF-
β activated kinase 1 (TAK1) has been shown to repress
transcription of the human telomerase reverse transcriptase
gene. TAK1-induced repression was caused by the recruitment
of histone deacetylase to Sp1 at the hTERT promoter.113
T-cell
acute lymphoblastic leukemia protein 1 (TAL1) is also a
negative regulator of the hTERT promoter. Overexpression of
the proto-oncogenic protein TAL1 leads to a decrease in
hTERT mRNA levels and consequently reduced telomerase
activity.114
Histone deacetylase (HDAC) inhibitors, such as
trichostatin A, have been shown to induce apoptosis and down-
regulate telomerase activity via suppression of hTERT
expression in leukemic U937 cells.115−117
Dynamic assembly
of the E2F−pocket protein−HDAC complex plays a central
role in the regulation of hTERT. For example, a small molecule,
CGK 1026, inhibits the recruitment of HDAC into E2F-pocket
protein complexes assembled on the hTERT promoter.118
These data underscore the fact that several anticancer treatment
strategies are likely to result in telomerase inhibition indirectly.
Several reports demonstrate the indirect effects on
telomerase activity that are possible with small molecules.
Tea catechins such as (−)-epigallocatechin 3-gallate (15)
prevent growth of a variety of cancer cell lines and inhibit
telomerase activity in a concentration dependent manner by
decreasing hTERT mRNA levels.110,119
The changes in hTERT
mRNA levels can be attributed to chromatin remodeling due to
the hypoacetylation of histones at the hTERT gene promoter
and decreased methylation of the hTERT promoter at the E2F-
1 binding site.119
Consistent with the observed decrease activity
of the hTERT promoter, 15-treated U937 monoblastoid
leukemia and HT29 colon adenocarcinoma cells experienced
decreased proliferation potential commensurate with telomere
shortening. In addition, 15 was also shown to induce a severe
reduction in cell growth with enhanced membrane blebbing,
cell swelling, apoptosis, and necrosis in laryngeal squamous cell,
small-cell lung, and breast cancers.119−121
Further studies
showed that the tea catechin derivative MST-312 (N,N′-
bis(2,3-dihydroxybenzoyl)-1,2-phenylenediamine) appears to
inhibit telomere maintenance by directly targeting telomerase
(IC50 = 0.67 μM), resulting in telomere erosion at a 20-fold
lower dose than 15 in the monoblastic leukemia cell line
U937.122
GSK3 inhibition by 6-bromoindirubin-3′-oxime caused
suppression of hTERT expression, telomerase activity, and
telomere length in various cancer cell lines and decreased
hTERT expression in ovarian cancer xenografts.123
Other
repressors of hTERT transcription are retinoids, derivatives of
vitamin A, which have been shown to down-regulate telomerase
leading to telomere erosion.124
Retinoic acid receptor α and
retinoid-X receptor-specific agonists synergistically down-
regulated hTERT and induced tumor cell death.125
Posttranslational Regulators of hTERT. Perturbing
posttranslational modifications represents another mechanism
for targeting telomerase. Because telomerase activity and
cellular location is controlled by its phosphorylation state,
phosphatase and kinase inhibitors can have direct effects on
telomerase. For example, the phosphatase 2A inhibitor okadaic
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXG
acid prevents telomerase inhibition by phosphatase 2A in breast
cancer cells.126
Protein kinase C (PKC) has been shown to
enhance telomerase activity in certain cell types, and PKC
inhibitors have been shown to inactivate telomerase in
nasopharyngeal127,128
and cervical129
cancer cells among others.
Imatinib mesylate (Gleevec), a tyrosine kinase inhibitor, also
down-regulates telomerase activity and inhibits proliferation in
telomerase expressing cells.130
The protein kinase Akt has well-
established roles in telomerase activation. Akt phosphorylates
hTERT and activates telomerase in vitro, while the nonspecific
Akt inhibitor wortmannin causes loss of telomerase activity.127
The tyrosine kinase c-Abl is also capable of phosphorylating
hTERT. The SH3 domain of c-Abl interacts with and
phosphorylates hTERT resulting in a decrease in telomerase
activity.131
These observations demonstrate that telomerase
activity can be directly regulated by protein phosphorylation
independent of transcriptional regulation and highlight that
telomerase inhibition may contribute to the observed
anticancer effects of several drugs. In addition to regulation
by kinases, it has recently been determined that hTERT is a
substrate of caspases-6 and −7.132
Interestingly, the caspase
sites are unique and the cleavage products appear remarkably
stable.
Hsp90 Inhibitors. The telomerase ribonucleoprotein
complex requires the Hsp90 chaperone complex, including
p23, Hsp70, p60, and Hsp40/γdl for proper assemblage. Hsp90
associates directly with hTERT and is required to maintain
telomerase in an active conformation at least under some
conditions.133,134
Hsp90 inhibitors cause inactivation, destabi-
lization, and degradation of Hsp90 target proteins including
hTERT.46,133
Since several oncogenic proteins including c-Myc,
Akt, and mutated p53 require Hsp90 for their activity,
inhibition of this chaperone should block multiple oncogenic
pathways. Hsp90 inhibitors such as novobiocin, radicicol,
geldanamycin, and 17-AAG were also shown to inhibit
telomerase activity.134−137
Recently, curcumin has been
shown to inhibit nuclear localization of telomerase by
dissociating the Hsp90 co-chaperone p23 from hTERT.93
Curcumin treatment limited the association between p23 and
hTERT while not affecting Hsp90 interaction with telomerase.
■ G-QUADRUPLEX STABILIZERS
The most advanced area of small-molecule drug discovery
directed at the telomere is G-quadruplex-stabilizing ligands. G-
quadruplexes are unique structures formed by folding G-rich
nucleic acid sequences. These structures are heterogeneous
with a variety of structural topologies.138
Folding topologies are
influenced by sequence, buffer conditions including identity of
cations present, molecular crowding, and other parameters.
Significant effort toward determining the structures of human
telomeric DNA138,139
and the biological functions of G-
quadruplex DNA at the telomere and in other G-rich sequences
embedded in the chromosome140
has been reported. Because
telomerase must hybridize to its DNA primer to afford
telomeric DNA extension, it was proposed that G-quadruplex
formation would render a primer inert to telomerase.141
Seminal studies by Zahler et al. showing that telomerase
could not extend a folded G-quadruplex were consistent with
this hypothesis and initiated the field of G-quadruplex binding
ligands as a platform for telomerase inhibition.141
This report
was quickly followed by the cardinal work of the Hurley and
Neidle labs providing the first G-quadruplex ligands that
effectively inhibited telomerase.142,143
Since these reports, a
wide variety of pharmacophores have been explored as G-
quadruplex ligands, and most of these are polycationic aromatic
hydrocarbons (Figure 9).144
Since these initial studies
demonstrating a relationship between telomeres and G-
quadruplexes, these structures have become much more
complex as drug targets than originally assumed, as they are
predicted to be present in a wide variety of genomic locations
with increased density in promoter regions.145
Several of these
promoters, including those for hTERT,146
c-kit,147,148
and c-
myc,149,150
have verified abilities to fold into G-quadruplex
structures. Much of the work on G-quadruplex structure,
Figure 9. Representative G-quadruplex ligands that inhibit telomerase. (A) 16 (BRACO19), a 3,6,9 trisubstituted acridine derivative. (B) RHPS4, a
pentacyclic acridine derivative. (C) and (D) are models of 16 bound to a G-quadruplex formed by a dimer of TTAGGGTTAGGG. The structure
was rendered from PDB 3CE5. 16 is shown in red. dG residues are shown in cyan, and dA and dT residues are shown in gray.
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXH
biological relevance, and small-molecule targeting has been
recently and extensively reviewed, and readers are directed to
these.144,151,152
Here, we will simply highlight the major
concerns with targeting G-quadruplexes as a specifically
telomere-directed anticancer drug modality and highlight recent
approaches toward advancing telomere-binding ligands to the
clinic.
The first G-quadruplex ligands explored as telomerase
inhibitors were disubstituted anthraquinone derivatives.153
Since that time, several other polycyclic heteroaromatic
scaffolds and macrocycles have been explored allowing the
development of potent telomerase inhibitors with increasing
specificity for G-quadruplex DNA compared to duplex DNA,
and these ligands are available at the G-quadruplex ligands
database (http://www.g4ldb.org/ci2/index.php).154
Impor-
tantly, high resolution crystal and NMR structures of human
telomeric G-quadruplexes can inform rational design to
improve G-quadruplex affinity and specificity.152
From
structural studies, it is clear that critical features known to
enhance G-quadruplex binding include a delocalized π-electron
system capable of stacking onto G-tetrad a positive charge that
can rest in the center of the G-quartets to interact with the
carbonyl oxygens from guanine that form the core of the G-
quartets, and positively charged substituents that can interact
with the grooves of the G-quadruplex (Figure 9c and Figure
9d).138
From a purely biochemical view, the most significant
current issues in G-quadruplex ligand design are specificity for
G-quadruplex DNA compared to duplex DNA and specificity
for G-quadruplex type, i.e., telomeric versus other DNA and
RNA quadruplexes that may be present.155
From a
pharmaceutical standpoint, the major issue concerning trans-
lation of G-quadruplex ligands is the poor pharmacokinetic
profile associated with the typically hydrophilic and positively
charged molecules.156
One of the most significant recent advances toward selective
G-quadruplex targeting is the ability to rationally design
molecules using structure-based approaches. Structures of G-
quadruplexes formed by the human telomeric DNA sequence
by X-ray crystallography and NMR spectroscopy have provided
templates for rational design of G-quaduplex ligands.157,158
These advances have allowed for the development of rationally
designed molecules with increased.159,160
These promising
examples suggest that G-quadruplex specificity is achievable.
The more challenging hurdle that still remains is the poor
pharmacokinetic and pharmacodynamic properties of promis-
ing G-quadruplex compounds. These hurdles may not be
insurmountable, as evidenced by the G-quadruplex ligand and
fluoroquinolone derivative quarfloxin, which binds to nucleo-
lin/rDNA and inhibits Pol 1 transcription, thus inducing
apoptosis in adenocarcinoma cells, and which has had some
clinical success.161
■ DISRUPTION OF TELOMERE MAINTENANCE
USING ANTISENSE AND RELATED APPROACHES
Ribozymes. Hammerhead ribozymes are small catalytic
RNA molecules that have the capacity to specifically cleave
target RNA after NUH sequences, where N is any nucleotide,
U is uridine, and H is any nucleotide except G.162
Design of
hammerhead ribozymes is relatively simple, and specificity can
be achieved by incorporating sequences complementary to the
target. Because of this nature, ribozymes have been investigated
as a platform for telomerase inhibition via targeted RNA
cleavage of hTER or hTERT mRNA.
The template region of hTER is an ideal target for ribozyme
cleavage because it has the requisite target sequence for
Hammerhead or HDV ribozymes, is essential for telomerase
activity, and it may be solvent exposed even in the telomerase
holoenzyme. Studies using melanoma,163
hepatoma, colon
cancer,164
and endometrial cancer cells165
have correlated
ribozyme expression with a marked reduction in telomerase
activity resulting from diminished telomerase RNA levels.
While telomere lengths were greatly attenuated in some
carcinomas, other types did not display this phenotype.163,165
Treatment with template-targeting ribozymes also resulted in
significant growth retardation with proliferation rates reduced
compared to parent cells.166
Furthermore, treated cells
experienced telomere shortening in concert with altered or
dendritic morphologies, which are indicative of senescence,166
G0/G1 growth arrest, and increased apoptotic rate.164
Circularization and addition of a six nucleotide poly(A)
sequence to these ribozymes increased target RNA cleavage
while conferring protection from cellular exoribonuclease
degradation.167
TERT mRNA levels have been targeted by a Hammerhead
ribozyme targeting the initial nucleotides in hTERT mRNA,
which may be essential for the proper binding of the translation
initiation complex to the 5′ mRNA cap structure.168
Ribozyme
cleavage in the middle of hTERT mRNA has also demonstrated
decreased telomerase activity while also influencing rapid cell
death (4 days) even without diminished telomere length,
consistent with potentially nontelomeric roles of hTERT in
some cancer cells.169
Oligonucleotides and Template Antagonists. Antisense
(AS) oligonucleotides (ODNs) can inhibit telomerase by
several mechanisms including blocking hTERT translation,
reduction of hTERT mRNA and hTER levels, and sterically
blocking the association of telomerase with the telomere
(template antagonists).42
However, oligonucleotides present
many challenges toward achieving efficient intratumoral
telomerase inhibition. Nonetheless, studies have demonstrated
that unmodified antisense oligonucleotides can in fact potently
inhibit telomerase activity under in vitro conditions.170−172
One
of these was a genetic approach using a plasmid borne 125
nucleotide hTR-antisense RNA, which inhibited tumor growth
in mouse xenografts.170
To make up for the shortcomings of unmodified ODNs, a
spectrum of ODNs with chemically modified backbones have
been utilized to enhance the therapeutic potential of these
antisense strategies. Phosphorothioate (PS) ODNs composed
of telomere repeat sequences ranging from 6 to 24 bases and
targeting the template of hTER showed significant telomerase
inhibition in various cancer cell lines.173,174
While longer PS-
ODNs did not have any effect on tumor growth, a 6-mer PS-
ODN decreased cellular growth in Burkett’s lymphoma,
increased cell population doubling times, induced dramatic
levels of apoptosis, and reduced human xenograft tumors in
mice to undetectable sizes.174
2′-O-Methyl-PS chimeric ODNs
targeting the template sequence of hTER demonstrated potent
telomerase inhibition with IC50 values in the nanomolar
range.175,176
Similar ODNs blocked up to 95% of telomerase
activity after 1 day of treatment when delivered with a cationic
lipid delivery system and induced telomere erosion and
subsequent limited proliferation commensurate with marked
apoptosis in long-term studies with chronic treatment.177,178
Cellular delivery with chitosan-coated PLGA nanoparticles
afforded even higher cellular uptake and the similar phenotype
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXI
of progressive telomere shortening with an EC50 value of 10
nM.179
Direct addition of 2′-O-methoxyethyl-PS chimeric
ODNs decreased telomere length and proliferation and growth
rates in MCF-7 (breast)180
and DU145181
(prostate) cancer
cells in culture and xenograft tumor models.180,181
Synergistic
effects were seen in DU145 prostate cancer cells upon addition
of chemotherapeutic agents such as cisplatin or carboplatin.181
Peptide nucleic acids (PNAs) completely covering the
template region of hTER were 10−50 times more potent
than similar PS-ODNs with IC50 values in the nanomolar
range.182
The potent telomerase inhibition is due to high
affinity of the PNA to the single stranded RNA template and
suggests that affinity is not limited to base pairing but may also
include interactions between hTERT and the peptide back-
bone.182
Both cellular permeabilization and lipid-mediated
delivery demonstrated similar IC50 values in DU145 (prostate
cancer)83
and HME50-5 (immortal breast epithelial)182
cells.
Conjugation of short peptide sequences to PNAs targeting the
hTER template displayed increased telomerase inhibition with
IC50 values 10-fold lower than PNAs alone.183
Targeting of
hTER by 2-5A ODNs has led to a remarkable inhibition of
telomerase activity concomitant with rapid hTER degradation
and a greater than 50% decrease in cell viability in a time
dependent manner.41,184−186
2-5A ODNs targeting hTER can
also arrest cells in G2/M phase187
and induce apoptosis via
activation of the caspase signaling cascade.186
Direct admin-
istration of hTER-targeting 2-5A ODNs and combination
therapy with interferon-β has demonstrated antiproliferative
effects including a reduction in colony formation and a
significant retardation in tumor growth and volume over
time.184,185,188
In 2003, Asai et al. described a novel class of chemically
modified N3′→P5′-thiophosphoramidates (NPS), which con-
ferred high stability and sequence specific potency.42,189
The
efforts led to the development of 17 (GRN163), a 13-
nucleotide long NPS-ODN with sequence 5′-TAGGGTT-
AGACAA-3′ that targets the template region of hTER by
overlapping and extending four nucleotides beyond the 5′
template boundary (IC50 = 0.14 μM).189
Treatment of various
cancer cell lines with 17 significantly reduced telomerase
activity, promoted critical telomere loss, and instigated
apoptosis while concomitantly up-regulating caspase activity.189
These results combined with impressive data from long-term
experiments and its high potency and specificity indicated that
17 may be an effective anticancer agent.189
In order to improve its bioavailability, lipophilicity, and
cellular uptake, a 5′ terminal palmitoyl group was attached
through an amide bond to α-aminoglycerol of the thiophos-
phoramidate 17 to produce the clinical candidate 1 (Figure
10).190
Comparison studies with 17 proved that 1 was just as
potent at inhibiting telomerase in various cancers, could inhibit
telomerase activity effectively without the use of a transfection
agent, and demonstrated greater tumor uptake.190−192
Since
then, many studies have been published revealing the extensive
anticancer potential of this drug in a variety of cancers. A single
treatment of 1 μM 1 yielded complete inhibition of telomerase
activity in various cancers within hours of dosing and near
complete inhibition could be maintained for up to 6 days, while
chronic treatment allowed sustain inhibition with undetectable
telomerase activity.193−198
Continued treatment for 2 weeks
elicited morphological changes such as rounding and increased
extension of filliopodia-like structures,193,198
reduced colony
formation,194,196
and attenuated metastasis invasion.194
Longer
periods of treatment with 1 triggered decreased proliferation,196
decline in cell viability,197,199
disruption of focal adhesions and
actin filament organization,193
substantial telomere short-
ening,194−196
reduced capacity to form neurospheres and
mammospheres,191,199
down-regulation of cell cycle regula-
tors,192
G0/G1 cell cycle arrest,191,198
up-regulation of APG5
and DAPK2 expression (genes implicated in the induction of
apoptosis),197
disrupted membrane intergrity,191
and promo-
tion of apoptotic cell death.191,197
1 also promoted a higher
percentage of anaphase bridges, hallmarks of telomere
dysfunction, and increased the frequency of γ-H2AX and
53bp1 foci, known molecular markers that localize to sites of
DNA strand breaks and dysfunctional telomeres.200
In tumor
models, 1 prevented tumor formation and metastasis from
xenografts,194,195
significantly reduced growth and tumor
volume,191,200,201
and enhanced antitumor effects in combina-
tion therapy with doxorubicin, temozolomide, and irradia-
tion.191,200
The successful preclinical experiments demonstrated
the antitumor effectiveness of 1, and it remains the only
telomerase inhibitor to have been tested in clinical trials.
Lessons learned from clinical trials of 1 are essential for the
field, as this represents the first telomerase inhibitor to reach
the clinic. Currently, six phase I and phase I/II trials of 1 as a
single agent and three combination trials are ongoing. Major
objectives of these studies are safety, tolerability, and maximum
tolerated dose. Early reports indicate that 1 is well tolerated,
but nothing further has been reported.
RNAi. RNA interference or silencing RNA (RNAi or siRNA)
is double stranded RNAs that take advantage of a sequence-
specific and natural mechanism that mediates gene silenc-
ing.42,202
As an RNA that is complicit for telomerase activity,
hTER is an excellent target for siRNA therapies. Transfection
and intravenous injection of partially double-stranded PS-
ODNs targeting the template region of hTER reduced cell
proliferation in human cervical carcinoma and, more
importantly, strongly reduced the number of metastases in
murine malignant melanoma cells.203
siRNAs targeting
nucleotides 11−19 of hTER expressed in various cancer cell
lines including HCT116 colon cancer, LOX melanoma, HeLa
cervical cancer, T24 bladder cancer, MCF-7 and BT424 breast
cancers, and LNCaP prostate cancer, via lentiviral-vector
expression, caused rapid growth inhibition and apoptosis
while displaying negligible cellular cytotoxicity independent of
Figure 10. Clinical tested telomerase inhibitor. (A) 1 is a modified oligonucleotide that contains a 5′-palmotyl group. The sequence is
complementary to the hTER template domain. (B) The N3′→P5′-thiophosphoramidate linkage present in 1.
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXJ
telomere length.204
Cancer cells exhibited knockdown of
endogenous hTER as well as down-regulation of cell cycle
progression genes (cyclin G2 and Cdc27) and specific genes
that play pivotal roles in tumor growth, angiogenesis, and
metastasis (integrin αV and Met proto-oncogene), suggesting a
role for these genes in the hTER-mediated cell growth
inhibition.204
Expression of siRNAs targeting hTERT mRNA in various
cancers has caused decreased hTERT mRNA levels, reduced
hTERT expression, and notable telomerase inhibition within 24
h of dosing and lasting up to 4 days.120,176,205−209
Stable
transfection of these siRNAs led to complete suppression of
telomerase activity in 3 days concomitant with the absence of
hTERT mRNA expression.205,207
Cancers treated with siRNAs
that deplete hTERT experienced decreased cell viability,120
prolonged population doubling times,209
suppression of
anchorage-dependent proliferation,206
significant growth inhib-
ition,208
and morphological changes indicative of senescence.205
Esophageal and colon adenocarcinomas displayed extensive
telomere shortening,205,207
leading to chromosomes with
telomere free ends suggestive of complete telomere erosion.205
Genetic silencing of hTERT led to an up-regulation of genes
involved in DNA damage recognition and repair, p53 family
proteins and their regulators, cell cycle inhibitors, proapoptotic
proteins, and caspases.199,205,210
At the same time, hTERT
depletion significantly attenuated the expression of integrin αV
and β3 subunits and c-Met, proteins that are known to play
important roles in cell differentiation, adhesion, motility, and
tumor metastasis.206
In tumor models, hTERT siRNAs
disrupted cell adhesion, migration, and invasion, promoted
slower tumor growth,206
decreased cell viability and colony
formation, induced cell cycle arrest,199
reduced tumor sizes and
volumes in xenografts,206,208
and demonstrated additive effects
upon combination therapy with doxorubicin and intereferon-
γ.120,199
Interestingly, while expression of siRNAs targeting
both hTER and hTERT concomitantly did decrease mRNA,
RNA, and protein levels of telomerase components, cell
proliferation, and xenograft tumor growth and promote
pyknotic nuclear chromatin, the effect of silencing hTER
alone was more pronounced than silencing either hTERT or
both together.211
■ IMMUNOTHERAPY
The realization that hTERT is overexpressed in the vast
majority of cancers prompted the investigation of hTERT as a
tumor antigen. Soon after the discovery of widespread hTERT
expression in tumors, hTERT peptide fragments were identified
as tumor-associated antigens functioning through the major
histocompatibility complex class I and class II pathways. In
several cancer types, CD8+ cytotoxic T-cells specific for
hTERT peptides, primarily hTERT-I540 (ILAKFLHWL), are
detectable suggesting active hTERT-directed immune surveil-
lance for these cancers.212
Subsequently, over 20 hTERT-
derived peptides have been employed to boost antitumor
immunity.2
The majority of these elicit a CD8+ cytotoxic T-cell
response, while the others activate a CD4+ response. Several
clinical trials of hTERT-based cancer vaccines have demon-
strated the safety of this approach.213,214
Unfortunately, the
impact on disease progression reported to date is modest. Some
patients respond with significant tumor regression, but more
generally only a low percentage of patients respond and
responses tend to be transient. This is despite the fact that
vaccination with hTERT-derived peptides generally does result
in hTERT-specific T cell production, indicative of an
immunological response. Current efforts involve understanding
the differences between responsive and nonresponsive patients,
optimization of antigenic hTERT epitopes, and optimization of
specific immunotherapy platforms for telomerase-targeted
treatment strategies.
■ GENE DELIVERY
One of the main concerns with cancer gene therapy is the
selective expression of target genes in tumor cells while
avoiding expression in normal tissues. Because activation of
telomerase activity in cancer cells is largely driven by increased
expression of hTERT, it is reasonable to speculate that hTERT
promoter activity is commensurately increased. Taking
advantage of this increased promoter activity, several groups
have investigated the potential of the hTERT promoter for
cancer-selective gene therapy. The hTERT promoter is
extremely attractive for the selective expression of transgenes
for anticancer gene therapy because of the near universal
expression of hTERT in cancers and the much greater activity
of the promoter in cancer cells compared to normal cells, even
those that express telomerase. Most reports on the use of the
hTERT promoter for cancer gene therapy attempt to control
expression of anticancer genes in tumor cells. For example,
proapoptotic genes (FADD and TRAIL) and a suicide gene
(HSVtk) have been tested.215−217
In cell-based assays and in
mouse models, tumor-specific transgene expression has been
achieved and promising results in mouse studies highlight early
successes. Because these approaches utilize replication-deficient
virus in order to decrease harmful side effects, these approaches
may suffer from limited tumor-mass distribution. A different
and very promising approach that overcomes this limitation is
to utilize the hTERT promoter not for transgene expression
but instead viral replication, in essence establishing an
anticancer virus. One hTERT-promoter driven adenovirus,
telomelysin, has entered clinical trials in a small, 16 patient,
phase I trial in which the patients had a wide variety of cancer
diagnoses.35
Early data suggest that telomelysin was well
tolerated, but clinical response was limited, owing perhaps to
the single viral infection, which may have limited the viral titer.
Most promising was the response of one patient with a
significant, 56%, reduction in metastatic tumor size.35
■ PERSPECTIVE AND FUTURE DIRECTIONS
Much progress has been made toward translating telomere
research to the clinic since the seminal studies, demonstrating
the importance of telomeres, the identification of telomerase
and telomere-binding proteins, and the validation of telomere
length maintenance as a requirement of cancer cells. Early
clinical trials with telomerase inhibitors have shown that the
approach is well tolerated and does reduce telomerase activity.
However, the absence of reported clinical efficacy suggests
limited progress. Future trials examining which patients might
be best suited for telomerase inhibitor therapy and current trials
testing telomerase inhibitors in combination with other
traditional chemotherapeutics can be expected to advance the
field. We can also expect continued advances in gene therapy by
taking advantage of the robust activity of the hTERT promoter
in cancer cells as viral vectors are optimized. Likewise,
immunotherapy targeting hTERT antigens is expected to
continue its progress through clinical trials and arguably holds
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXK
the most promise currently for telomerase-targeting therapeu-
tics.
Surprisingly, small-molecule inhibitors of telomerase have
not fared well. This may change as advances in telomerase
structural biology; for example, the crystal structure of an insect
TERT,62
structures of several hTER domains,218,219
models of
hTERT based on the beetle crystal structure,220
and the recent
three-dimensional model of human telomerase based on
electron microscopy images221
increase our understanding of
telomerase structure and provide templates for rational small-
molecule drug design. Rationally designed telomerase inhibitors
may allow increased specificity, bioavailability, and inhibition
mechanisms that have greater effects in a clinical setting than
those small molecules that are currently available. In addition,
the observation that G-quadruplex ligands can directly disrupt
telomere structure evidently by blocking binding of telomere-
binding proteins such as TRF2 and POT1 suggests that there
are more telomere-associated targets to be exploited, and we
can expect that near future studies will determine the
effectiveness of directly targeting the telomere as an anticancer
drug platform. Targeting telomere biology as a means to
improve human health remains an important objective that
drives discoveries in telomere biology and telomerase
biochemistry. The interdisciplinary approaches toward targeting
telomere biology have led to three discrete areas of clinical
trials: inhibition of telomerase, hTERT-directed immunother-
apy, and hTERT-promoter gene/viral therapy. It seems that the
future of pharmaceutical benefit from telomere-targeting agents
remains promising, and our increased understanding of
telomere biochemistry and biology is sure to inform more
effective approaches toward developing telomere biology
targeted agents.
■ AUTHOR INFORMATION
Corresponding Author
*Phone: 919-966-6422. E-mail: jarstfer@unc.edu.
Notes
The authors declare no competing financial interest.
Biographies
Vijay Sekaran is a postdoctoral fellow at the University of Colorado at
Boulder working with Drs. James Goodrich and Jennifer Kugel. He
received his B.S. in Biomedical Engineering from the Georgia Institute
of Technology in 2005 and his Ph.D. in Pharmaceutical Sciences from
the University of North Carolina at Chapel Hill under the guidance of
Michael Jarstfer in 2012. His dissertation focused on understanding
the structure and function of telomerase RNAs as well as
characterizing the effects of small molecules that bind to human
telomerase RNA and inhibit telomerase activity.
Joana Soares is a research scientist at Genezyme. She received her B.S.
in Chemistry from The College of William and Mary, VA, in 2004 and
her Ph.D. in Pharmaceutical Sciences from the University of North
Carolina at Chapel Hill in 2010 under the supervision of Michael
Jarstfer. Her dissertation focused on elucidating the mechanism of
telomerase activation or repression and its impact on cellular pathways
and elucidating the mechanism of action of new potential telomerase
inhibitors. Her research interrogated the relationship between
telomerase, aging, and cancer progression.
Michael B. Jarstfer is an Associate Professor of Chemical Biology and
Medicinal Chemistry in the Eshelman School of Pharmacy at the
University of North Carolina at Chapel Hill. He received his B.A. in
Biochemistry from Trinity University, San Antonio TX, and his Ph.D.
from the University of Utah under the guidance of Dale Poulter. He
conducted postdoctoral research in the laboratory of Thomas Cech at
the University of Colorado in Boulder. His research interests include
structure and function of telomerase, development of telomere
targeting agents, the roles of telomere biology in human health, and
the molecular basis for social motivation.
■ ABBREVIATIONS USED
ALT, alternative lengthening of telomeres; AS, antisense;
hTERT, human telomerase reverse transcriptase; HDAC,
histone deacetylase; hTER, human telomerase RNA; NPS,
N3′→P5′-thiophosphoramidates; ODN, oligonucleotide; PKC,
protein kinase C; PNA, peptide nucleic acid; POT1, Protection
of Telomeres 1; PS, phosphorothioate; QSAR, quantitative
structure−activity relationship; RNAi, RNA interference; SAR,
structure−activity relationship; siRNA, silencing RNA; TIF,
telomere dysfunction induced focus; TAK1, TGF-β activated
kinase 1; TAL1, T-cell acute lymphoblastic leukemia protein 1;
WT1, Wilm’s tumor protein 1
■ REFERENCES
(1) Shay, J. W.; Bacchetti, S. A survey of telomerase activity in human
cancer. Eur. J. Cancer 1997, 33, 787−791.
(2) Beatty, G. L.; Vonderheide, R. H. Telomerase as a universal
tumor antigen for cancer vaccines. Expert Rev. Vaccines 2008, 7, 881−
887.
(3) Fujiwara, T.; Urata, Y.; Tanaka, N. Telomerase-specific oncolytic
virotherapy for human cancer with the hTERT promoter. Curr. Cancer
Drug Targets 2007, 7, 191−201.
(4) Jakupciak, J. P. Real-time telomerase activity measurements for
detection of cancer. Expert Rev. Mol. Diagn. 2005, 5, 745−753.
(5) Deng, Y.; Chan, S. S.; Chang, S. Telomere dysfunction and
tumour suppression: the senescence connection. Nat. Rev. Cancer
2008, 8, 450−458.
(6) Palm, W.; de Lange, T. How shelterin protects mammalian
telomeres. Annu. Rev. Genet. 2008, 42, 301−334.
(7) Sperka, T.; Wang, J.; Rudolph, K. L. DNA damage checkpoints in
stem cells, ageing and cancer. Nat. Rev. Mol. Cell Biol. 2012, 13, 579−
590.
(8) Moon, I. K.; Jarstfer, M. B. The human telomere and its
relationship to human disease, therapy, and tissue engineering. Front.
Biosci. 2007, 12, 4595−4620.
(9) Blasco, M. A. Telomeres and human disease: ageing, cancer and
beyond. Nat. Rev. Genet. 2005, 6, 611−622.
(10) Collins, K. The biogenesis and regulation of telomerase
holoenzymes. Nat. Rev. Mol. Cell Biol. 2006, 7, 484−494.
(11) Sekaran, V. G.; Soares, J.; Jarstfer, M. B. Structures of telomerase
subunits provide functional insights. Biochim. Biophys. Acta 2010, 1804,
1190−1201.
(12) Feng, J.; Funk, W. D.; Wang, S. S.; Weinrich, S. L.; Avilion, A.
A.; Chiu, C. P.; Adams, R. R.; Chang, E.; Allsopp, R. C.; Yu, J.; Le, S.;
West, M. D.; Harley, C.; Andrews, W. H.; Greider, C. W.; Villeponteu,
B. The RNA component of human telomerase. Science 1995, 269,
1236−1241.
(13) Nakamura, T. M.; Morin, G. B.; Chapman, K. B.; Weinrich, S.
L.; Andrews, W. H.; Lingner, J.; Harley, C. B.; Cech, T. R. Telomerase
catalytic subunit homologs from fission yeast and human. Science 1997,
277, 955−959.
(14) Huang, F. W.; Hodis, E.; Xu, M. J.; Kryukov, G. V.; Chin, L.;
Garraway, L. A. Highly recurrent TERT promoter mutations in human
melanoma. Science 2013, 339, 957−959.
(15) Karlseder, J.; Smogorzewska, A.; de Lange, T. Senescence
induced by altered telomere state, not telomere loss. Science 2002, 295,
2446−2449.
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXL
(16) Griffith, J. D.; Comeau, L.; Rosenfield, S.; Stansel, R. M.;
Bianchi, A.; Moss, H.; de Lange, T. Mammalian telomeres end in a
large duplex loop. Cell 1999, 97, 503−514.
(17) Cao, Y.; Bryan, T. M.; Reddel, R. R. Increased copy number of
the TERT and TERC telomerase subunit genes in cancer cells. Cancer
Sci. 2008, 99, 1092−1099.
(18) Cesare, A. J.; Reddel, R. R. Telomere uncapping and alternative
lengthening of telomeres. Mech. Ageing Dev. 2008, 129, 99−108.
(19) Muntoni, A.; Reddel, R. R. The first molecular details of ALT in
human tumor cells. Hum. Mol. Genet. 2005, 14, R191−R196.
(20) Hahn, W. C.; Counter, C. M.; Lundbeg, A. S.; Beijersbergen, R.
L.; Brooks, M. W.; Weinberg, R. A. Creation of human tumour cells
with defined genetic elements. Nature 1999, 400, 464−468.
(21) Aubert, G.; Lansdorp, P. M. Telomeres and aging. Physiol. Rev.
2008, 88, 557−579.
(22) Bollmann, F. M. The many faces of telomerase: emerging
extratelomeric effects. BioEssays 2008, 30, 728−732.
(23) Artandi, S. E.; DePinho, R. A. Telomeres and telomerase in
cancer. Carcinogenesis 2010, 31, 9−18.
(24) Xu, L.; Li, S.; Stohr, B. A. The role of telomere biology in
cancer. Annu. Rev. Pathol. 2013, 8, 49−78.
(25) Ju, Z.; Rudolph, K. L. Telomeres and telomerase in cancer stem
cells. Eur. J. Cancer 2006, 42, 1197−1203.
(26) Armanios, M.; Greider, C. W. Telomerase and cancer stem cells.
Cold Spring Harbor Symp. Quant. Biol. 2005, 70, 205−208.
(27) 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.
(28) Svenson, U.; Roos, G. Telomere length as a biological marker in
malignancy. Biochim. Biophys. Acta 2009, 1792, 317−323.
(29) Hahn, W. C.; Stewart, S. A.; Brooks, M. W.; York, S. G.; Eaton,
E.; Kurachi, A.; Beijersbergen, R. L.; Knoll, J. H. M.; Maeyerson, M.;
Weinberg, R. A. Inhibition of telomerase limits the growth of human
cancer cells. Nat. Med. 1999, 5, 1164−1170.
(30) Zhang, X.; Mar, V.; Zhou, W.; Harrington, L.; Robinson, M. O.
Telomere shortening and apoptosis in telomerase-inhibited human
tumor cells. Genes Dev. 1999, 13, 2388−2399.
(31) Roth, A.; Harley, C. B.; Baerlocher, G. M. Imetelstat
(GRN163L)telomerase-based cancer therapy. Recent Results Cancer
Res. 2010, 184, 221−234.
(32) Wong, H. M.; Payet, L.; Huppert, J. L. Function and targeting of
G-quadruplexes. Curr. Opin. Mol. Ther. 2009, 11, 146−155.
(33) Monchaud, D.; Teulade-Fichou, M. P. A hitchhiker’s guide to G-
quadruplex ligands. Org. Biomol. Chem. 2008, 6, 627−636.
(34) 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.
(35) Nemunaitis, J.; Tong, A. W.; Nemunaitis, M.; Senzer, N.;
Phadke, A. P.; Bedell, C.; Adams, N.; Zhang, Y. A.; Maples, P. B.;
Chen, S.; Pappen, B.; Burke, J.; Ichimaru, D.; Urata, Y.; Fujiwara, T. A
phase I study of telomerase-specific replication competent oncolytic
adenovirus (telomelysin) for various solid tumors. Mol. Ther. 2010, 18,
429−434.
(36) Liu, J. P.; Chen, W.; Schwarer, A. P.; Li, H. Telomerase in
cancer immunotherapy. Biochim. Biophys. Acta, Rev. Cancer 2010, 1805,
35−42.
(37) Damm, K.; Hemmann, U.; Garin-Chesa, P.; Hauel, N.;
Kauffmann, I.; Priepke, H.; Niestroj, C.; Daiber, C.; Enenkel, B.;
Guilliard, B.; Lauritsch, I.; Muller, E.; Pascolo, E.; Sauter, G.; Pantic,
M.; Martens, U. M.; Wenz, C.; Lingner, J.; Kraut, N.; Rettig, W. J.;
Schnapp, A. A highly selective telomerase inhibitor limiting human
cancer cell proliferation. EMBO J. 2001, 20, 6958−6968.
(38) Huffman, K. E.; Levene, S. D.; Tesmer, V. M.; Shay, J. W.;
Wright, W. E. Telomere shortening is proportional to the size of the
G-rich telomeric 3′-overhang. J. Biol. Chem. 2000, 275, 19719−19722.
(39) Zou, Y.; Sfeir, A.; Gryaznov, S. M.; Shay, J. W.; Wright, W. E.
Does a sentinel or a subset of short telomeres determine replicative
senescence? Mol. Biol. Cell 2004, 15, 3709−3718.
(40) de Lange, T. Shelterin: the protein complex that shapes and
safeguards human telomeres. Genes Dev. 2005, 19, 2100−2110.
(41) Hu, J.; Hwang, S. S.; Liesa, M.; Gan, B.; Sahin, E.; Jaskelioff, M.;
Ding, Z.; Ying, H.; Boutin, A. T.; Zhang, H.; Johnson, S.; Ivanova, E.;
Kost-Alimova, M.; Protopopov, A.; Wang, Y. A.; Shirihai, O. S.; Chin,
L.; DePinho, R. A. Antitelomerase therapy provokes ALT and
mitochondrial adaptive mechanisms in cancer. Cell 2012, 148, 651−
663.
(42) Folini, M.; Zaffaroni, N. Targeting telomerase by antisense-
based approaches: perspectives for new anti-cancer therapies. Curr.
Pharm. Des. 2005, 11, 1105−1117.
(43) Cong, Y.; Shay, J. W. Actions of human telomerase beyond
telomeres. Cell Res. 2008, 18, 725−732.
(44) Park, J. I.; Venteicher, A. S.; Hong, J. Y.; Choi, J.; Jun, S.; Shkreli,
M.; Chang, W.; Meng, Z.; Cheung, P.; Ji, H.; McLaughlin, M.;
Veenstra, T. D.; Nusse, R.; McCrea, P. D.; Artandi, S. E. Telomerase
modulates Wnt signalling by association with target gene chromatin.
Nature 2009, 460, 66−72.
(45) Ghosh, A.; Saginc, G.; Leow, S. C.; Khattar, E.; Shin, E. M.; Yan,
T. D.; Wong, M.; Zhang, Z.; Li, G.; Sung, W. K.; Zhou, J.; Chng, W. J.;
Li, S.; Liu, E.; Tergaonkar, V. Telomerase directly regulates NF-
kappaB-dependent transcription. Nat. Cell Biol. 2012, 14, 1270−1281.
(46) De Cian, A.; Lacroix, L.; Douarre, C.; Temime-Smaali, N.;
Trentesaux, C.; Riou, J. F.; Mergny, J. L. Targeting telomeres and
telomerase. Biochimie 2008, 90, 131−155.
(47) Salvati, E.; Leonetti, C.; Rizzo, A.; Scarsella, M.; Mottolese, M.;
Galati, R.; Sperduti, I.; Stevens, M. F.; D’Incalci, M.; Blasco, M.;
Chiorino, G.; Bauwens, S.; Horard, B.; Gilson, E.; Stoppacciaro, A.;
Zupi, G.; Biroccio, A. Telomere damage induced by the G-quadruplex
ligand RHPS4 has an antitumor effect. J. Clin. Invest. 2007, 117, 3236−
3247.
(48) Miyazaki, T.; Pan, Y.; Joshi, K.; Purohit, D.; Hu, B.; Demir, H.;
Mazumder, S.; Okabe, S.; Yamori, T.; Viapiano, M.; Shin-ya, K.;
Seimiya, H.; Nakano, I. Telomestatin impairs glioma stem cell survival
and growth through the disruption of telomeric G-quadruplex and
inhibition of the proto-oncogene, c-Myb. Clin. Cancer Res. 2012, 18,
1268−1280.
(49) Gong, Y.; de Lange, T. A Shld1-controlled POT1a provides
support for repression of ATR signaling at telomeres through RPA
exclusion. Mol. Cell 2010, 40, 377−387.
(50) Melana, S. M.; Holland, J. F.; Pogo, B. G. Inhibition of cell
growth and telomerase activity of breast cancer cells in vitro by 3′-
azido-3′-deoxythymidine. Clin. Cancer Res. 1998, 4, 693−696.
(51) Strahl, C.; Blackburn, E. H. Effects of reverse transcriptase
inhibitors on telomere length and telomerase activity in two
immortalized human cell lines. Mol. Cell. Biol. 1996, 16, 53−65.
(52) Datta, A.; Bellon, M.; Sinha-Datta, U.; Bazarbachi, A.;
Lepelletier, Y.; Canioni, D.; Waldmann, T. A.; Hermine, O.; Nicot,
C. Persistent inhibition of telomerase reprograms adult T-cell leukemia
to p53-dependent senescence. Blood 2006, 108, 1021−1029.
(53) Johnston, J. S.; Johnson, A.; Gan, Y.; Wientjes, M. G.; Au, J. L.
Synergy between 3′-azido-3′-deoxythymidine and paclitaxel in human
pharynx FaDu cells. Pharm. Res. 2003, 20, 957−961.
(54) Yasuda, C.; Kato, M.; Kuroda, D.; Ohyanagi, H. Experimental
studies on potentiation of the antitumor activity of 5-fluorouracil with
3′-azido-3′-deoxythymidine for the gastric cancer cell line MKN28 in
vivo. Jpn. J. Cancer Res 1997, 88, 97−102.
(55) Chen, C.; Zhang, Y.; Wang, Y.; Huang, D.; Xi, Y.; Qi, Y.
Synergic effect of 3′-azido-3′-deoxythymidine and arsenic trioxide in
suppressing hepatoma cells. Anti-Cancer Drugs 2011, 22, 435−443.
(56) Hukezalie, K. R.; Thumati, N. R.; Cote, H. C.; Wong, J. M. In
vitro and ex vivo inhibition of human telomerase by anti-HIV
nucleoside reverse transcriptase inhibitors (NRTIs) but not by non-
NRTIs. PLoS One 2012, 7, e47505.
(57) Leeansyah, E.; Cameron, P. U.; Solomon, A.; Tennakoon, S.;
Velayudham, P.; Gouillou, M.; Spelman, T.; Hearps, A.; Fairley, C.;
Smit de, V.; Pierce, A. B.; Armishaw, J.; Crowe, S. M.; Cooper, D. A.;
Koelsch, K. K.; Liu, J. P.; Chuah, J.; Lewin, S. R. Inhibition of
telomerase activity by human immunodeficiency virus (HIV) nucleos-
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXM
(t)ide reverse transcriptase inhibitors: a potential factor contributing to
HIV-associated accelerated aging. J. Infect. Dis. 2013, 207, 1157−1165.
(58) Fletcher, T. M.; Cathers, B. E.; Ravikumar, K. S.; Mamiya, B. M.;
Kerwin, S. M. Inhibition of human telomerase by 7-deaza-2′-
deoxyguanosine nucleoside triphosphate analogs: potent inhibition
by 6-thio-7-deaza-2′-deoxyguanosine 5′-triphosphate. Bioorg. Chem.
2001, 29, 36−55.
(59) Hayakawa, N.; Nozawa, K.; Ogawa, A.; Kato, N.; Yoshida, K.;
Akamatsu, K.; Tsuchiya, M.; Nagasaka, A.; Yoshida, S. Isothiazolone
derivatives selectively inhibit telomerase from human and rat cancer
cells in vitro. Biochemistry 1999, 38, 11501−11507.
(60) Kim, J. H.; Lee, G. E.; Lee, J. E.; Chung, I. K. Potent inhibition
of human telomerase by nitrostyrene derivatives. Mol. Pharmacol.
2003, 63, 1117−1124.
(61) Kim, J. H.; Lee, G. E.; Kim, S. W.; Chung, I. K. Identification of
a quinoxaline derivative that is a potent telomerase inhibitor leading to
cellular senescence of human cancer cells. Biochem. J. 2003, 373, 523−
529.
(62) Gillis, A. J.; Schuller, A. P.; Skordalakes, E. Structure of the
Tribolium castaneum telomerase catalytic subunit TERT. Nature 2008,
455, 633−637.
(63) Liu, X. H.; Li, J.; Shi, J. B.; Song, B. A.; Qi, X. B. Design and
synthesis of novel 5-phenyl-N-piperidine ethanone containing 4,5-
dihydropyrazole derivatives as potential antitumor agents. Eur. J. Med.
Chem. 2012, 51, 294−299.
(64) Liu, X. H.; Ruan, B. F.; Liu, J. X.; Song, B. A.; Jing, L. H.; Li, J.;
Yang, Y.; Zhu, H. L.; Qi, X. B. Design and synthesis of N-phenylacetyl
(sulfonyl) 4,5-dihydropyrazole derivatives as potential antitumor
agents. Bioorg. Med. Chem. Lett. 2011, 21, 2916−2920.
(65) Luo, Y.; Zhang, S.; Qiu, K. M.; Liu, Z. J.; Yang, Y. S.; Fu, J.;
Zhong, W. Q.; Zhu, H. L. Synthesis, biological evaluation, 3D-QSAR
studies of novel aryl-2H-pyrazole derivatives as telomerase inhibitors.
Bioorg. Med. Chem. Lett. 2013, 23, 1091−1095.
(66) Liu, X. H.; Liu, H. F.; Chen, J.; Yang, Y.; Song, B. A.; Bai, L. S.;
Liu, J. X.; Zhu, H. L.; Qi, X. B. Synthesis and molecular docking study
of novel coumarin derivatives containing 4,5-dihydropyrazole moiety
as potential antitumor agents. Bioorg. Med. Chem. Lett. 2010, 20,
5705−5708.
(67) Zheng, Q. Z.; Zhang, X. M.; Xu, Y.; Cheng, K.; Jiao, Q. C.; Zhu,
H. L. Synthesis, biological evaluation, and molecular docking studies of
2-chloropyridine derivatives possessing 1,3,4-oxadiazole moiety as
potential antitumor agents. Bioorg. Med. Chem. 2010, 18, 7836−7841.
(68) Sun, J.; Zhu, H.; Yang, Z. M.; Zhu, H. L. Synthesis, molecular
modeling and biological evaluation of 2-aminomethyl-5-(quinolin-2-
yl)-1,3,4-oxadiazole-2(3H)-thione quinolone derivatives as novel
anticancer agent. Eur. J. Med. Chem. 2013, 60, 23−28.
(69) Huang, P. R.; Yeh, Y. M.; Wang, T. C. Potent inhibition of
human telomerase by helenalin. Cancer Lett. 2005, 227, 169−174.
(70) Soares, J.; Keppler, B. R.; Wang, X.; Lee, K. H.; Jarstfer, M. B.
Ortho-quinone tanshinones directly inhibit telomerase through an
oxidative mechanism mediated by hydrogen peroxide. Bioorg. Med.
Chem. Lett. 2011, 21, 7474−7478.
(71) Chen, Y. J.; Sheng, W. Y.; Huang, P. R.; Wang, T. C. Potent
inhibition of human telomerase by U-73122. J. Biomed. Sci. 2006, 13,
667−674.
(72) Ueno, T.; Takahashi, H.; Oda, M.; Mizunuma, M.; Yokoyama,
A.; Goto, Y.; Mizushina, Y.; Sakaguchi, K.; Hayashi, H. Inhibition of
human telomerase by rubromycins: implication of spiroketal system of
the compounds as an active moiety. Biochemistry 2000, 39, 5995−
6002.
(73) Mizushina, Y.; Ueno, T.; Oda, M.; Yamaguchi, T.; Saneyoshi,
M.; Sakaguchi, K. The biochemical mode of inhibition of DNA
polymerase beta by alpha-rubromycin. Biochim. Biophys. Acta 2000,
1523, 172−181.
(74) Nakai, R.; Ishida, H.; Asai, A.; Ogawa, H.; Yamamoto, Y.;
Kawasaki, H.; Akinaga, S.; Mizukami, T.; Yamashita, Y. Telomerase
inhibitors identified by a forward chemical genetics approach using a
yeast strain with shortened telomere length. Chem. Biol. 2006, 13,
183−190.
(75) Uchida, K.; Ogawa, T.; Yasuda, Y.; Mimura, H.; Fujimoto, T.;
Fukuyama, T.; Wakimoto, T.; Asakawa, T.; Hamashima, Y.; Kan, T.
Stereocontrolled total synthesis of (+)-UCS1025A. Angew. Chem.
2012, 51, 12850−12853.
(76) Lambert, T. H.; Danishefsky, S. J. Total synthesis of UCS1025A.
J. Am. Chem. Soc. 2006, 128, 426−427.
(77) Wong, L. H.; Unciti-Broceta, A.; Spitzer, M.; White, R.; Tyers,
M.; Harrington, L. A yeast chemical genetic screen identifies inhibitors
of human telomerase. Chem. Biol. 2013, 20, 333−340.
(78) Barma, D. K.; Elayadi, A.; Falck, J. R.; Corey, D. R. Inhibition of
telomerase by BIBR 1532 and related analogues. Bioorg. Med. Chem.
Lett. 2003, 13, 1333−1336.
(79) Pascolo, E.; Wenz, C.; Lingner, J.; Hauel, N.; Priepke, H.;
Kauffmann, I.; Garin-Chesa, P.; Rettig, W. J.; Damm, K.; Schnapp, A.
Mechanism of human telomerase inhibition by BIBR1532, a synthetic,
non-nucleosidic drug candidate. J. Biol. Chem. 2002, 277, 15566−
15572.
(80) El-Daly, H.; Kull, M.; Zimmermann, S.; Pantic, M.; Waller, C.
F.; Martens, U. M. Selective cytotoxicity and telomere damage in
leukemia cells using the telomerase inhibitor BIBR1532. Blood 2005,
105, 1742−1749.
(81) Ward, R. J.; Autexier, C. Pharmacological telomerase inhibition
can sensitize drug-resistant and drug-sensitive cells to chemo-
therapeutic treatment. Mol. Pharmacol. 2005, 68, 779−786.
(82) Pantic, M.; Zimmermann, S.; Waller, C. F.; Martens, U. M. The
level of telomere dysfunction determines the efficacy of telomerase-
based therapeutics in a lung cancer cell line. Int. J. Cancer 2005, 26,
1227−1232.
(83) Hamilton, S. E.; Simmons, C. G.; Kathiriya, I. S.; Corey, D. R.
Cellular delivery of peptide nucleic acids and inhibition of human
telomerase. Chem. Biol. 1999, 6, 343−351.
(84) Keppler, B. R.; Jarstfer, M. B. Inhibition of telomerase activity by
preventing proper assemblage. Biochemistry 2004, 43, 334−343.
(85) Pinto, I. G.; Guilbert, C.; Ulyanov, N. B.; Stearns, J.; James, T. L.
Discovery of ligands for a novel target, the human telomerase RNA,
based on flexible-target virtual screening and NMR. J. Med. Chem.
2008, 51, 7205−7215.
(86) Dominick, P. K.; Keppler, B. R.; Legassie, J. D.; Moon, I. K.;
Jarstfer, M. B. Nucleic acid-binding ligands identify new mechanisms
to inhibit telomerase. Bioorg. Med. Chem. Lett. 2004, 14, 3467−3471.
(87) Kaiser, M.; Sainlos, M.; Lehn, J. M.; Bombard, S.; Teulade-
Fichou, M. P. Aminoglycoside-quinacridine conjugates: towards
recognition of the P6.1 element of telomerase RNA. ChemBioChem
2006, 7, 321−329.
(88) Jain, N.; Francis, S.; Friedman, S. H. Inhibition of therapeutically
important polymerases with high affinity bis-intercalators. Bioorg. Med.
Chem. Lett. 2012, 22, 4844−4848.
(89) Rangarajan, S.; Friedman, S. H. Design, synthesis, and
evaluation of phenanthridine derivatives targeting the telomerase
RNA/DNA heteroduplex. Bioorg. Med. Chem. Lett. 2007, 17, 2267−
2273.
(90) Meyerson, M.; Counter, C. M.; Eaton, E. N.; Ellisen, L. W.;
Steiner, P.; Caddle, S. D.; Ziaugra, L.; Beijersbergen, R. L.; Davidoff,
M. J.; Liu, Q.; Bacchetti, S.; Haber, D. A.; Weinberg, R. A. hEST2, the
putative human telomerase catalytic subunit gene, is up-regulated in
tumor cells and during immortalization. Cell 1997, 90, 785−795.
(91) 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.
(92) Grandori, C.; Cowley, S. M.; James, L. P.; Eisenman, R. N. The
Myc/Max/Mad network and the transcriptional control of cell
behavior. Annu. Rev. Cell Dev. Biol. 2000, 16, 653−699.
(93) Moon, D. O.; Kim, M. O.; Lee, J. D.; Choi, Y. H.; Kim, G. Y.
Butein suppresses c-Myc-dependent transcription and Akt-dependent
phosphorylation of hTERT in human leukemia cells. Cancer Lett.
2009, 286, 172−179.
(94) Guo, Q. L.; Lin, S. S.; You, Q. D.; Gu, H. Y.; Yu, J.; Zhao, L.; Qi,
Q.; Liang, F.; Tan, Z.; Wang, X. Inhibition of human telomerase
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXN
reverse transcriptase gene expression by gambogic acid in human
hepatoma SMMC-7721 cells. Life Sci. 2006, 78, 1238−1245.
(95) Jagadeesh, S.; Kyo, S.; Banerjee, P. P. Genistein represses
telomerase activity via both transcriptional and posttranslational
mechanisms in human prostate cancer cells. Cancer Res. 2006, 66,
2107−2115.
(96) Shapira, S.; Granot, G.; Mor-Tzuntz, R.; Raanani, P.; Uziel, O.;
Lahav, M.; Shpilberg, O. Second-generation tyrosine kinase inhibitors
reduce telomerase activity in K562 cells. Cancer Lett. 2012, 323, 223−
231.
(97) Wooten, L. G.; Ogretmen, B. Sp1/Sp3-dependent regulation of
human telomerase reverse transcriptase promoter activity by the
bioactive sphingolipid ceramide. J. Biol. Chem. 2005, 280, 28867−
28876.
(98) Kyo, S.; Takakura, M.; Taira, T.; Kanaya, T.; Itoh, H.; Yutsudo,
M.; Ariga, H.; Inoue, M. Sp1 cooperates with c-Myc to activate
transcription of the human telomerase reverse transcriptase gene
(hTERT). Nucleic Acids Res. 2000, 28, 669−677.
(99) Misiti, S.; Nanni, S.; Fontemaggi, G.; Cong, Y. S.; Wen, J.; Hirte,
H. W.; Piaggio, G.; Sacchi, A.; Pontecorvi, A.; Bacchetti, S.; Farsetti, A.
Induction of hTERT expression and telomerase activity by estrogens
in human ovary epithelium cells. Mol. Cell. Biol. 2000, 20, 3764−3771.
(100) Kyo, S.; Takakura, M.; Kanaya, T.; Zhuo, W.; Fujimoto, K.;
Nishio, Y.; Orimo, A.; Inoue, M. Estrogen activates telomerase. Cancer
Res. 1999, 59, 5917−5921.
(101) Nanni, S.; Narducci, M.; Della Pietra, L.; Moretti, F.; Grasselli,
A.; De Carli, P.; Sacchi, A.; Pontecorvi, A.; Farsetti, A. Signaling
through estrogen receptors modulates telomerase activity in human
prostate cancer. J. Clin. Invest. 2002, 110, 219−227.
(102) Aldous, W. K.; Marean, A. J.; DeHart, M. J.; Matej, L. A.;
Moore, K. H. Effects of tamoxifen on telomerase activity in breast
carcinoma cell lines. Cancer 1999, 85, 1523−1529.
(103) Wang, Z.; Kyo, S.; Takakura, M.; Tanaka, M.; Yatabe, N.;
Maida, Y.; Fujiwara, M.; Hayakawa, J.; Ohmichi, M.; Koike, K.; Inoue,
M. Progesterone regulates human telomerase reverse transcriptase
gene expression via activation of mitogen-activated protein kinase
signaling pathway. Cancer Res. 2000, 60, 5376−5381.
(104) Xu, D.; Popov, N.; Hou, M.; Wang, Q.; Bjorkholm, M.;
Gruber, A.; Menkel, A. R.; Henriksson, M. Switch from Myc/Max to
Mad1/Max binding and decrease in histone acetylation at the
telomerase reverse transcriptase promoter during differentiation of
HL60 cells. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 3826−3831.
(105) Gunes, C.; Lichtsteiner, S.; Vasserot, A. P.; Englert, C.
Expression of the hTERT gene is regulated at the level of
transcriptional initiation and repressed by Mad1. Cancer Res. 2000,
60, 2116−2121.
(106) Kanaya, T.; Kyo, S.; Hamada, K.; Takakura, M.; Kitagawa, Y.;
Harada, H.; Inoue, M. Adenoviral expression of p53 represses
telomerase activity through down-regulation of human telomerase
reverse transcriptase transcription. Clin. Cancer Res. 2000, 6, 1239−
1247.
(107) Crowe, D. L.; Nguyen, D. C. Rb and E2F-1 regulate telomerase
activity in human cancer cells. Biochim. Biophys. Acta 2001, 1518, 1−6.
(108) Henderson, Y. C.; Breau, R. L.; Liu, T. J.; Clayman, G. L.
Telomerase activity in head and neck tumors after introduction of
wild-type p53, p21, p16, and E2F-1 genes by means of recombinant
adenovirus. Head Neck 2000, 22, 347−354.
(109) Oh, S.; Song, Y.; Yim, J.; Kim, T. K. The Wilms’ tumor 1
tumor suppressor gene represses transcription of the human
telomerase reverse transcriptase gene. J. Biol. Chem. 1999, 274,
37473−37478.
(110) Lindkvist, A.; Ivarsson, K.; Jernberg-Wiklund, H.; Paulsson-
Karlsson, Y. Interferon-induced sensitization to apoptosis is associated
with repressed transcriptional activity of the hTERT promoter in
multiple myeloma. Biochem. Biophys. Res. Commun. 2006, 341, 1141−
1148.
(111) Lacerte, A.; Korah, J.; Roy, M.; Yang, X. J.; Lemay, S.; Lebrun,
J. J. Transforming growth factor-beta inhibits telomerase through
SMAD3 and E2F transcription factors. Cell. Signalling 2008, 20, 50−
59.
(112) Li, H.; Xu, D.; Li, J.; Berndt, M. C.; Liu, J. P. Transforming
growth factor beta suppresses human telomerase reverse transcriptase
(hTERT) by Smad3 interactions with c-Myc and the hTERT gene. J.
Biol. Chem. 2006, 281, 25588−25600.
(113) Fujiki, T.; Miura, T.; Maura, M.; Shiraishi, H.; Nishimura, S.;
Imada, Y.; Uehara, N.; Tashiro, K.; Shirahata, S.; Katakura, Y. TAK1
represses transcription of the human telomerase reverse transcriptase
gene. Oncogene 2007, 26, 5258−5266.
(114) Terme, J. M.; Mocquet, V.; Kuhlmann, A. S.; Zane, L.;
Mortreux, F.; Wattel, E.; Duc Dodon, M.; Jalinot, P. Inhibition of the
hTERT promoter by the proto-oncogenic protein TAL1. Leukemia
2009, 23, 2081−2089.
(115) Wu, P.; Meng, L.; Wang, H.; Zhou, J.; Xu, G.; Wang, S.; Xi, L.;
Chen, G.; Wang, B.; Zhu, T.; Lu, Y.; Ma, D. Role of hTERT in
apoptosis of cervical cancer induced by histone deacetylase inhibitor.
Biochem. Biophys. Res. Commun. 2005, 335, 36−44.
(116) Woo, H. J.; Lee, S. J.; Choi, B. T.; Park, Y. M.; Choi, Y. H.
Induction of apoptosis and inhibition of telomerase activity by
trichostatin A, a histone deacetylase inhibitor, in human leukemic
U937 cells. Exp. Mol. Pathol. 2007, 82, 77−84.
(117) Choi, Y. H. Apoptosis of U937 human leukemic cells by
sodium butyrate is associated with inhibition of telomerase activity. Int.
J. Oncol. 2006, 29, 1207−1213.
(118) Won, J.; Chang, S.; Oh, S.; Kim, T. K. Small-molecule-based
identification of dynamic assembly of E2F-pocket protein-histone
deacetylase complex for telomerase regulation in human cells. Proc.
Natl. Acad. Sci. U.S.A. 2004, 101, 11328−11333.
(119) Berletch, J. B.; Liu, C.; Love, W. K.; Andrews, L. G.; Katiyar, S.
K.; Tollefsbol, T. O. Epigenetic and genetic mechanisms contribute to
telomerase inhibition by EGCG. J. Cell. Biochem. 2008, 103, 509−519.
(120) Wang, X.; Hao, M. W.; Dong, K.; Lin, F.; Ren, J. H.; Zhang, H.
Z. Apoptosis induction effects of EGCG in laryngeal squamous cell
carcinoma cells through telomerase repression. Arch. Pharmacal Res.
2009, 32, 1263−1269.
(121) Sadava, D.; Whitlock, E.; Kane, S. E. The green tea polyphenol,
epigallocatechin-3-gallate inhibits telomerase and induces apoptosis in
drug-resistant lung cancer cells. Biochem. Biophys. Res. Commun. 2007,
360, 233−237.
(122) Seimiya, H.; Oh-hara, T.; Suzuki, T.; Naasani, I.; Shimazaki, T.;
Tsuchiya, K.; Tsuruo, T. Telomere shortening and growth inhibition
of human cancer cells by novel synthetic telomerase inhibitors MST-
312, MST-295, and MST-1991. Mol. Cancer Ther. 2002, 1, 657−665.
(123) Bilsland, A. E.; Hoare, S.; Stevenson, K.; Plumb, J.; Gomez-
Roman, N.; Cairney, C.; Burns, S.; Lafferty-Whyte, K.; Roffey, J.;
Hammonds, T.; Keith, W. N. Dynamic telomerase gene suppression
via network effects of GSK3 inhibition. PLoS One 2009, 4, e6459.
(124) Pendino, F.; Flexor, M.; Delhommeau, F.; Buet, D.; Lanotte,
M.; Segal-Bendirdjian, E. Retinoids down-regulate telomerase and
telomere length in a pathway distinct from leukemia cell differ-
entiation. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 6662−6667.
(125) Pendino, F.; Dudognon, C.; Delhommeau, F.; Sahraoui, T.;
Flexor, M.; Bennaceur-Griscelli, A.; Lanotte, M.; Segal-Bendirdjian, E.
Retinoic acid receptor alpha and retinoid-X receptor-specific agonists
synergistically target telomerase expression and induce tumor cell
death. Oncogene 2003, 22, 9142−9150.
(126) Li, H.; Zhao, L. L.; Funder, J. W.; Liu, J. P. Protein phosphatase
2A inhibits nuclear telomerase activity in human breast cancer cells. J.
Biol. Chem. 1997, 272, 16729−16732.
(127) 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.
(128) Yu, C. C.; Lo, S. C.; Wang, T. C. Telomerase is regulated by
protein kinase C-zeta in human nasopharyngeal cancer cells. Biochem.
J. 2001, 355, 459−464.
Journal of Medicinal Chemistry Perspective
dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXO
Telomere maintenance as a target for drug discovery
Telomere maintenance as a target for drug discovery
Telomere maintenance as a target for drug discovery

Weitere ähnliche Inhalte

Was ist angesagt? (20)

Epigenetics diet and cancer
Epigenetics diet and cancerEpigenetics diet and cancer
Epigenetics diet and cancer
 
Real ROCH
Real ROCHReal ROCH
Real ROCH
 
Aging genetics and aging
Aging genetics and agingAging genetics and aging
Aging genetics and aging
 
ZG_SS14_pageformat
ZG_SS14_pageformatZG_SS14_pageformat
ZG_SS14_pageformat
 
Peptidimimetics in the Treatment of Cancer
Peptidimimetics in the Treatment of CancerPeptidimimetics in the Treatment of Cancer
Peptidimimetics in the Treatment of Cancer
 
Proteasome inhibitors in cancer therapy
Proteasome inhibitors in cancer therapyProteasome inhibitors in cancer therapy
Proteasome inhibitors in cancer therapy
 
Regulation of deoxynucleotide metabolism
Regulation of deoxynucleotide metabolismRegulation of deoxynucleotide metabolism
Regulation of deoxynucleotide metabolism
 
Okoye_NMDAR Inhibition Poster
Okoye_NMDAR Inhibition PosterOkoye_NMDAR Inhibition Poster
Okoye_NMDAR Inhibition Poster
 
Canine oncoprotein targets for Melanoma, Breast Cancer, Osteosarcoma
Canine oncoprotein targets for Melanoma, Breast Cancer, OsteosarcomaCanine oncoprotein targets for Melanoma, Breast Cancer, Osteosarcoma
Canine oncoprotein targets for Melanoma, Breast Cancer, Osteosarcoma
 
Rectal cancer
Rectal cancerRectal cancer
Rectal cancer
 
Proteome
ProteomeProteome
Proteome
 
Omics era
Omics eraOmics era
Omics era
 
Hallmarks of cancer
Hallmarks of cancerHallmarks of cancer
Hallmarks of cancer
 
New stratigies for_combating_cancer
New stratigies for_combating_cancerNew stratigies for_combating_cancer
New stratigies for_combating_cancer
 
Chapter 2.1 mi rna
Chapter 2.1 mi rnaChapter 2.1 mi rna
Chapter 2.1 mi rna
 
Bmc research note
Bmc research noteBmc research note
Bmc research note
 
Thesis-Final Draft
Thesis-Final DraftThesis-Final Draft
Thesis-Final Draft
 
ComparisonMitoMorphology
ComparisonMitoMorphologyComparisonMitoMorphology
ComparisonMitoMorphology
 
Canine Mast Cell Tumor
Canine Mast Cell TumorCanine Mast Cell Tumor
Canine Mast Cell Tumor
 
Cancer Cells
Cancer CellsCancer Cells
Cancer Cells
 

Ähnlich wie Telomere maintenance as a target for drug discovery

Telomere and telomerase inbitors
Telomere and telomerase inbitorsTelomere and telomerase inbitors
Telomere and telomerase inbitorsJeet Mazumder
 
Telomerase Inhibition as Novel Cancer Therapeutic Method
Telomerase Inhibition as Novel Cancer Therapeutic MethodTelomerase Inhibition as Novel Cancer Therapeutic Method
Telomerase Inhibition as Novel Cancer Therapeutic MethodVincensanicko
 
The clinical value of telomere testing – dr. Mark Tager A4M May 2016
The clinical value of telomere testing – dr. Mark Tager A4M May 2016The clinical value of telomere testing – dr. Mark Tager A4M May 2016
The clinical value of telomere testing – dr. Mark Tager A4M May 2016Life Length
 
Life at the End of Chromosomes
Life at the End of ChromosomesLife at the End of Chromosomes
Life at the End of Chromosomestelomerescience
 
Telomere and telomerase
Telomere and telomeraseTelomere and telomerase
Telomere and telomerasefuad76
 
TELOMERE PROTIENS.pptx
TELOMERE PROTIENS.pptxTELOMERE PROTIENS.pptx
TELOMERE PROTIENS.pptxAzkaJavaid6
 
Telomerase inhibition abolishes the tumorigenicity of pediatric ependymoma tu...
Telomerase inhibition abolishes the tumorigenicity of pediatric ependymoma tu...Telomerase inhibition abolishes the tumorigenicity of pediatric ependymoma tu...
Telomerase inhibition abolishes the tumorigenicity of pediatric ependymoma tu...Joshua Mangerel
 
Telomerase its role in aging and cancer
Telomerase its role in aging and cancerTelomerase its role in aging and cancer
Telomerase its role in aging and cancerHimadri Nath
 
Telomere, Functions & Role in Aging & Cancer
Telomere, Functions & Role in Aging & CancerTelomere, Functions & Role in Aging & Cancer
Telomere, Functions & Role in Aging & CancerZohaib HUSSAIN
 
Senescence and immortalization
Senescence and immortalizationSenescence and immortalization
Senescence and immortalizationbharathichellam
 
Rohan's presentation
Rohan's presentationRohan's presentation
Rohan's presentationrohanparikh
 
Aging and the telomere connection dr. Jerry w. Shay - april 2012
Aging and the telomere connection   dr. Jerry w. Shay - april 2012Aging and the telomere connection   dr. Jerry w. Shay - april 2012
Aging and the telomere connection dr. Jerry w. Shay - april 2012Life Length
 

Ähnlich wie Telomere maintenance as a target for drug discovery (20)

Telomeres kiran br
Telomeres kiran brTelomeres kiran br
Telomeres kiran br
 
Telomere and telomerase inbitors
Telomere and telomerase inbitorsTelomere and telomerase inbitors
Telomere and telomerase inbitors
 
A TALE OF LOBSTERS
A TALE OF LOBSTERSA TALE OF LOBSTERS
A TALE OF LOBSTERS
 
N-Terminal Domains
N-Terminal DomainsN-Terminal Domains
N-Terminal Domains
 
Telomerase Inhibition as Novel Cancer Therapeutic Method
Telomerase Inhibition as Novel Cancer Therapeutic MethodTelomerase Inhibition as Novel Cancer Therapeutic Method
Telomerase Inhibition as Novel Cancer Therapeutic Method
 
The clinical value of telomere testing – dr. Mark Tager A4M May 2016
The clinical value of telomere testing – dr. Mark Tager A4M May 2016The clinical value of telomere testing – dr. Mark Tager A4M May 2016
The clinical value of telomere testing – dr. Mark Tager A4M May 2016
 
Life at the End of Chromosomes
Life at the End of ChromosomesLife at the End of Chromosomes
Life at the End of Chromosomes
 
Telomere and telomerase
Telomere and telomeraseTelomere and telomerase
Telomere and telomerase
 
Treating cancer
Treating cancer Treating cancer
Treating cancer
 
TELOMERE PROTIENS.pptx
TELOMERE PROTIENS.pptxTELOMERE PROTIENS.pptx
TELOMERE PROTIENS.pptx
 
Telomerase inhibition abolishes the tumorigenicity of pediatric ependymoma tu...
Telomerase inhibition abolishes the tumorigenicity of pediatric ependymoma tu...Telomerase inhibition abolishes the tumorigenicity of pediatric ependymoma tu...
Telomerase inhibition abolishes the tumorigenicity of pediatric ependymoma tu...
 
Telomerase its role in aging and cancer
Telomerase its role in aging and cancerTelomerase its role in aging and cancer
Telomerase its role in aging and cancer
 
Telomere, Functions & Role in Aging & Cancer
Telomere, Functions & Role in Aging & CancerTelomere, Functions & Role in Aging & Cancer
Telomere, Functions & Role in Aging & Cancer
 
Cancer ppt
Cancer pptCancer ppt
Cancer ppt
 
Telomers and Telomerases
Telomers and TelomerasesTelomers and Telomerases
Telomers and Telomerases
 
Senescence and immortalization
Senescence and immortalizationSenescence and immortalization
Senescence and immortalization
 
Rohan's presentation
Rohan's presentationRohan's presentation
Rohan's presentation
 
Aging and the telomere connection dr. Jerry w. Shay - april 2012
Aging and the telomere connection   dr. Jerry w. Shay - april 2012Aging and the telomere connection   dr. Jerry w. Shay - april 2012
Aging and the telomere connection dr. Jerry w. Shay - april 2012
 
Telomere
TelomereTelomere
Telomere
 
Telomere ppt
Telomere pptTelomere ppt
Telomere ppt
 

Kürzlich hochgeladen

Book Call Girls in Kasavanahalli - 7001305949 with real photos and phone numbers
Book Call Girls in Kasavanahalli - 7001305949 with real photos and phone numbersBook Call Girls in Kasavanahalli - 7001305949 with real photos and phone numbers
Book Call Girls in Kasavanahalli - 7001305949 with real photos and phone numbersnarwatsonia7
 
Call Girls Viman Nagar 7001305949 All Area Service COD available Any Time
Call Girls Viman Nagar 7001305949 All Area Service COD available Any TimeCall Girls Viman Nagar 7001305949 All Area Service COD available Any Time
Call Girls Viman Nagar 7001305949 All Area Service COD available Any Timevijaych2041
 
Call Girls Service Noida Maya 9711199012 Independent Escort Service Noida
Call Girls Service Noida Maya 9711199012 Independent Escort Service NoidaCall Girls Service Noida Maya 9711199012 Independent Escort Service Noida
Call Girls Service Noida Maya 9711199012 Independent Escort Service NoidaPooja Gupta
 
Call Girls Frazer Town Just Call 7001305949 Top Class Call Girl Service Avail...
Call Girls Frazer Town Just Call 7001305949 Top Class Call Girl Service Avail...Call Girls Frazer Town Just Call 7001305949 Top Class Call Girl Service Avail...
Call Girls Frazer Town Just Call 7001305949 Top Class Call Girl Service Avail...narwatsonia7
 
97111 47426 Call Girls In Delhi MUNIRKAA
97111 47426 Call Girls In Delhi MUNIRKAA97111 47426 Call Girls In Delhi MUNIRKAA
97111 47426 Call Girls In Delhi MUNIRKAAjennyeacort
 
Call Girls Service Nandiambakkam | 7001305949 At Low Cost Cash Payment Booking
Call Girls Service Nandiambakkam | 7001305949 At Low Cost Cash Payment BookingCall Girls Service Nandiambakkam | 7001305949 At Low Cost Cash Payment Booking
Call Girls Service Nandiambakkam | 7001305949 At Low Cost Cash Payment BookingNehru place Escorts
 
Housewife Call Girls Hsr Layout - Call 7001305949 Rs-3500 with A/C Room Cash ...
Housewife Call Girls Hsr Layout - Call 7001305949 Rs-3500 with A/C Room Cash ...Housewife Call Girls Hsr Layout - Call 7001305949 Rs-3500 with A/C Room Cash ...
Housewife Call Girls Hsr Layout - Call 7001305949 Rs-3500 with A/C Room Cash ...narwatsonia7
 
Glomerular Filtration rate and its determinants.pptx
Glomerular Filtration rate and its determinants.pptxGlomerular Filtration rate and its determinants.pptx
Glomerular Filtration rate and its determinants.pptxDr.Nusrat Tariq
 
VIP Call Girls Mumbai Arpita 9910780858 Independent Escort Service Mumbai
VIP Call Girls Mumbai Arpita 9910780858 Independent Escort Service MumbaiVIP Call Girls Mumbai Arpita 9910780858 Independent Escort Service Mumbai
VIP Call Girls Mumbai Arpita 9910780858 Independent Escort Service Mumbaisonalikaur4
 
call girls in munirka DELHI 🔝 >༒9540349809 🔝 genuine Escort Service 🔝✔️✔️
call girls in munirka  DELHI 🔝 >༒9540349809 🔝 genuine Escort Service 🔝✔️✔️call girls in munirka  DELHI 🔝 >༒9540349809 🔝 genuine Escort Service 🔝✔️✔️
call girls in munirka DELHI 🔝 >༒9540349809 🔝 genuine Escort Service 🔝✔️✔️saminamagar
 
Housewife Call Girls Bangalore - Call 7001305949 Rs-3500 with A/C Room Cash o...
Housewife Call Girls Bangalore - Call 7001305949 Rs-3500 with A/C Room Cash o...Housewife Call Girls Bangalore - Call 7001305949 Rs-3500 with A/C Room Cash o...
Housewife Call Girls Bangalore - Call 7001305949 Rs-3500 with A/C Room Cash o...narwatsonia7
 
Russian Call Girls Chickpet - 7001305949 Booking and charges genuine rate for...
Russian Call Girls Chickpet - 7001305949 Booking and charges genuine rate for...Russian Call Girls Chickpet - 7001305949 Booking and charges genuine rate for...
Russian Call Girls Chickpet - 7001305949 Booking and charges genuine rate for...narwatsonia7
 
Call Girl Koramangala | 7001305949 At Low Cost Cash Payment Booking
Call Girl Koramangala | 7001305949 At Low Cost Cash Payment BookingCall Girl Koramangala | 7001305949 At Low Cost Cash Payment Booking
Call Girl Koramangala | 7001305949 At Low Cost Cash Payment Bookingnarwatsonia7
 
Call Girls Hebbal Just Call 7001305949 Top Class Call Girl Service Available
Call Girls Hebbal Just Call 7001305949 Top Class Call Girl Service AvailableCall Girls Hebbal Just Call 7001305949 Top Class Call Girl Service Available
Call Girls Hebbal Just Call 7001305949 Top Class Call Girl Service Availablenarwatsonia7
 
Low Rate Call Girls Mumbai Suman 9910780858 Independent Escort Service Mumbai
Low Rate Call Girls Mumbai Suman 9910780858 Independent Escort Service MumbaiLow Rate Call Girls Mumbai Suman 9910780858 Independent Escort Service Mumbai
Low Rate Call Girls Mumbai Suman 9910780858 Independent Escort Service Mumbaisonalikaur4
 
High Profile Call Girls Kodigehalli - 7001305949 Escorts Service with Real Ph...
High Profile Call Girls Kodigehalli - 7001305949 Escorts Service with Real Ph...High Profile Call Girls Kodigehalli - 7001305949 Escorts Service with Real Ph...
High Profile Call Girls Kodigehalli - 7001305949 Escorts Service with Real Ph...narwatsonia7
 
Call Girls Service in Bommanahalli - 7001305949 with real photos and phone nu...
Call Girls Service in Bommanahalli - 7001305949 with real photos and phone nu...Call Girls Service in Bommanahalli - 7001305949 with real photos and phone nu...
Call Girls Service in Bommanahalli - 7001305949 with real photos and phone nu...narwatsonia7
 
Call Girls Jayanagar Just Call 7001305949 Top Class Call Girl Service Available
Call Girls Jayanagar Just Call 7001305949 Top Class Call Girl Service AvailableCall Girls Jayanagar Just Call 7001305949 Top Class Call Girl Service Available
Call Girls Jayanagar Just Call 7001305949 Top Class Call Girl Service Availablenarwatsonia7
 
Pharmaceutical Marketting: Unit-5, Pricing
Pharmaceutical Marketting: Unit-5, PricingPharmaceutical Marketting: Unit-5, Pricing
Pharmaceutical Marketting: Unit-5, PricingArunagarwal328757
 
Air-Hostess Call Girls Madambakkam - Phone No 7001305949 For Ultimate Sexual ...
Air-Hostess Call Girls Madambakkam - Phone No 7001305949 For Ultimate Sexual ...Air-Hostess Call Girls Madambakkam - Phone No 7001305949 For Ultimate Sexual ...
Air-Hostess Call Girls Madambakkam - Phone No 7001305949 For Ultimate Sexual ...Ahmedabad Escorts
 

Kürzlich hochgeladen (20)

Book Call Girls in Kasavanahalli - 7001305949 with real photos and phone numbers
Book Call Girls in Kasavanahalli - 7001305949 with real photos and phone numbersBook Call Girls in Kasavanahalli - 7001305949 with real photos and phone numbers
Book Call Girls in Kasavanahalli - 7001305949 with real photos and phone numbers
 
Call Girls Viman Nagar 7001305949 All Area Service COD available Any Time
Call Girls Viman Nagar 7001305949 All Area Service COD available Any TimeCall Girls Viman Nagar 7001305949 All Area Service COD available Any Time
Call Girls Viman Nagar 7001305949 All Area Service COD available Any Time
 
Call Girls Service Noida Maya 9711199012 Independent Escort Service Noida
Call Girls Service Noida Maya 9711199012 Independent Escort Service NoidaCall Girls Service Noida Maya 9711199012 Independent Escort Service Noida
Call Girls Service Noida Maya 9711199012 Independent Escort Service Noida
 
Call Girls Frazer Town Just Call 7001305949 Top Class Call Girl Service Avail...
Call Girls Frazer Town Just Call 7001305949 Top Class Call Girl Service Avail...Call Girls Frazer Town Just Call 7001305949 Top Class Call Girl Service Avail...
Call Girls Frazer Town Just Call 7001305949 Top Class Call Girl Service Avail...
 
97111 47426 Call Girls In Delhi MUNIRKAA
97111 47426 Call Girls In Delhi MUNIRKAA97111 47426 Call Girls In Delhi MUNIRKAA
97111 47426 Call Girls In Delhi MUNIRKAA
 
Call Girls Service Nandiambakkam | 7001305949 At Low Cost Cash Payment Booking
Call Girls Service Nandiambakkam | 7001305949 At Low Cost Cash Payment BookingCall Girls Service Nandiambakkam | 7001305949 At Low Cost Cash Payment Booking
Call Girls Service Nandiambakkam | 7001305949 At Low Cost Cash Payment Booking
 
Housewife Call Girls Hsr Layout - Call 7001305949 Rs-3500 with A/C Room Cash ...
Housewife Call Girls Hsr Layout - Call 7001305949 Rs-3500 with A/C Room Cash ...Housewife Call Girls Hsr Layout - Call 7001305949 Rs-3500 with A/C Room Cash ...
Housewife Call Girls Hsr Layout - Call 7001305949 Rs-3500 with A/C Room Cash ...
 
Glomerular Filtration rate and its determinants.pptx
Glomerular Filtration rate and its determinants.pptxGlomerular Filtration rate and its determinants.pptx
Glomerular Filtration rate and its determinants.pptx
 
VIP Call Girls Mumbai Arpita 9910780858 Independent Escort Service Mumbai
VIP Call Girls Mumbai Arpita 9910780858 Independent Escort Service MumbaiVIP Call Girls Mumbai Arpita 9910780858 Independent Escort Service Mumbai
VIP Call Girls Mumbai Arpita 9910780858 Independent Escort Service Mumbai
 
call girls in munirka DELHI 🔝 >༒9540349809 🔝 genuine Escort Service 🔝✔️✔️
call girls in munirka  DELHI 🔝 >༒9540349809 🔝 genuine Escort Service 🔝✔️✔️call girls in munirka  DELHI 🔝 >༒9540349809 🔝 genuine Escort Service 🔝✔️✔️
call girls in munirka DELHI 🔝 >༒9540349809 🔝 genuine Escort Service 🔝✔️✔️
 
Housewife Call Girls Bangalore - Call 7001305949 Rs-3500 with A/C Room Cash o...
Housewife Call Girls Bangalore - Call 7001305949 Rs-3500 with A/C Room Cash o...Housewife Call Girls Bangalore - Call 7001305949 Rs-3500 with A/C Room Cash o...
Housewife Call Girls Bangalore - Call 7001305949 Rs-3500 with A/C Room Cash o...
 
Russian Call Girls Chickpet - 7001305949 Booking and charges genuine rate for...
Russian Call Girls Chickpet - 7001305949 Booking and charges genuine rate for...Russian Call Girls Chickpet - 7001305949 Booking and charges genuine rate for...
Russian Call Girls Chickpet - 7001305949 Booking and charges genuine rate for...
 
Call Girl Koramangala | 7001305949 At Low Cost Cash Payment Booking
Call Girl Koramangala | 7001305949 At Low Cost Cash Payment BookingCall Girl Koramangala | 7001305949 At Low Cost Cash Payment Booking
Call Girl Koramangala | 7001305949 At Low Cost Cash Payment Booking
 
Call Girls Hebbal Just Call 7001305949 Top Class Call Girl Service Available
Call Girls Hebbal Just Call 7001305949 Top Class Call Girl Service AvailableCall Girls Hebbal Just Call 7001305949 Top Class Call Girl Service Available
Call Girls Hebbal Just Call 7001305949 Top Class Call Girl Service Available
 
Low Rate Call Girls Mumbai Suman 9910780858 Independent Escort Service Mumbai
Low Rate Call Girls Mumbai Suman 9910780858 Independent Escort Service MumbaiLow Rate Call Girls Mumbai Suman 9910780858 Independent Escort Service Mumbai
Low Rate Call Girls Mumbai Suman 9910780858 Independent Escort Service Mumbai
 
High Profile Call Girls Kodigehalli - 7001305949 Escorts Service with Real Ph...
High Profile Call Girls Kodigehalli - 7001305949 Escorts Service with Real Ph...High Profile Call Girls Kodigehalli - 7001305949 Escorts Service with Real Ph...
High Profile Call Girls Kodigehalli - 7001305949 Escorts Service with Real Ph...
 
Call Girls Service in Bommanahalli - 7001305949 with real photos and phone nu...
Call Girls Service in Bommanahalli - 7001305949 with real photos and phone nu...Call Girls Service in Bommanahalli - 7001305949 with real photos and phone nu...
Call Girls Service in Bommanahalli - 7001305949 with real photos and phone nu...
 
Call Girls Jayanagar Just Call 7001305949 Top Class Call Girl Service Available
Call Girls Jayanagar Just Call 7001305949 Top Class Call Girl Service AvailableCall Girls Jayanagar Just Call 7001305949 Top Class Call Girl Service Available
Call Girls Jayanagar Just Call 7001305949 Top Class Call Girl Service Available
 
Pharmaceutical Marketting: Unit-5, Pricing
Pharmaceutical Marketting: Unit-5, PricingPharmaceutical Marketting: Unit-5, Pricing
Pharmaceutical Marketting: Unit-5, Pricing
 
Air-Hostess Call Girls Madambakkam - Phone No 7001305949 For Ultimate Sexual ...
Air-Hostess Call Girls Madambakkam - Phone No 7001305949 For Ultimate Sexual ...Air-Hostess Call Girls Madambakkam - Phone No 7001305949 For Ultimate Sexual ...
Air-Hostess Call Girls Madambakkam - Phone No 7001305949 For Ultimate Sexual ...
 

Telomere maintenance as a target for drug discovery

  • 1. Telomere Maintenance as a Target for Drug Discovery Vijay Sekaran, Joana Soares, and Michael B. Jarstfer* Division of Chemical Biology and Medicinal Chemistry, Eshelman School of Pharmacy, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina 27599, United States ABSTRACT: The observation that the enzyme telomerase is up-regulated in 80−90% of cancer cells isolated from primary human tumors but is absent in neighboring cells of healthy tissue has resulted in significant efforts to validate telomerase as an anticancer drug target and to develop effective approaches toward its inhibition. In addition to inhibitors that target the enzymatic function of telomerase, efforts toward immunotherapy using peptides derived from its catalytic subunit hTERT and hTERT-promoter driven gene therapy have made significant advances. The increased level of telomerase in cancer cells also provides a potential platform for cancer diagnostics. Telomerase inhibition leads to disruption of a cell’s ability to maintain the very ends of the chromosomes, which are called telomeres. Thus, the telomere itself has also attracted attention as an anticancer drug target. In this Perspective, interdisciplinary efforts to realize the therapeutic potential of targeting telomere maintenance with a focus on telomerase are discussed. ■ INTRODUCTION Pathways exhibiting differential activity in cancer cells when compared to healthy cells are attractive targets for therapeutic intervention. Accordingly, the observation that the enzyme telomerase is up-regulated in 80−90% of all cancer cells isolated from primary human tumors but is absent in neighboring cells of healthy tissue has resulted in significant efforts to validate telomerase as an anticancer drug target and to develop effective approaches toward its inhibition.1 In addition to inhibitors that target the enzymatic function of telomerase, efforts toward immunotherapy using peptides derived from the catalytic telomerase subunit hTERT (human telomerase reverse tran- scriptase)2 and hTERT-promoter driven gene therapy3 have made significant advances. The increased level of telomerase activity in cancer cells also provides a potential platform for cancer diagnostics.4 Telomerase inhibition leads to disruption of a cell’s ability to maintain the very ends of the chromosomes, which are called telomeres. Thus, the telomere itself has also attracted attention as an anticancer drug target. Here, we discuss the interdisciplinary efforts to realize the therapeutic potential of targeting telomere maintenance with a specific focus on telomerase. ■ TELOMERE MAINTENANCE AND TELOMERASE IN HUMAN CANCER CELLS Telomeres are the physical ends of linear chromosomes, and their importance to human health is underscored by the Nobel Prize in Physiology or Medicine awarded to Elizabeth Blackburn, Jack Szostak, and Carol Grieder in 2009 for their work on telomere biology and biochemistry. Telomeres are so important because proper telomere maintenance is an absolute requirement for continued cellular proliferation. Functional telomeres require a specific and repetitive DNA sequence and several telomere-specific proteins.5,6 These telomere-associated proteins function in two primary capacities: maintaining the structural and functional integrity of the telomere and ensuring complete replication of telomeric DNA. In the absence of functional telomeres, chromosome ends are recognized as DNA damage, and the ensuing DNA damage response generally results in either senescence or apoptosis, depending on the cell type.7 Mammalian telomeric DNA contains several thousand base pairs of repetitive DNA with the sequence 5′- TTAGGG/3′-AATCCC and a single stranded 3′ overhang of the G-rich sequence.8,9 The canonical DNA polymerases responsible for DNA replication cannot efficiently replicate the very ends of linear DNA templates. As a result, proliferative cells must employ additional pathways to overcome this end- replication problem. In human cells, this additional activity resides primarily in telomerase. Telomerase is a ribonucleo- protein complex that contains several protein subunits and one RNA subunit.10,11 Human telomerase RNA (hTER) is a highly structured 451-nucleotide long RNA that contains several functional domains in addition to a templating sequence used to code for the G-rich telomeric DNA strand.12 The protein subunit hTERT functions as a reverse transcriptase by using the templating portion of hTER to synthesize telomeric DNA.13 In the majority of cancer cells, hTERT is up-regulated, increasing the telomerase activity required for continuous proliferation.8,14 In contrast, telomerase activity is down-regulated in most normal human diploid cells, with the exception of germline and stem cells, primarily owing to inhibition of hTERT expression during cellular differentiation. Because telomerase activity is absent in most differentiated human cells, telomeres erode during the cellular aging process. This erosion allows telomere length to function like a mitotic hourglass that regulates cellular replication and arguably functions as a tumor suppressor mechanism.5 Cells lacking Received: April 11, 2013 Perspective pubs.acs.org/jmc © XXXX American Chemical Society A dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXX
  • 2. telomerase activity experience erosion of their telomeric DNA when they divide. At a critical length, the shortened telomeres are recognized as DNA damage and form a telomere dysfunction induced focus (TIF). Typically, cells respond to TIFs through senescence or apoptosis. Importantly, the telomere-length dependence of proliferation appears to be a result of the structure and not the length of the telomeric DNA per se. This is evidenced in part by experiments in the de Lange lab showing that the negative consequences of short telomeres can be overcome by the overexpression of the telomere-binding protein TRF2.15 It is likely that as telomeres erode in cells lacking an active telomere maintenance pathway, the telomere structure, predicted to be a lasso-like fold back structure termed a t-loop,16 becomes increasingly difficult to sustain eventually leading to TIFs and senescence or apoptosis if telomere function is not reestablished.6 Telomere maintenance directly impacts cancer in two specific ways. The most obvious is the absolute requirement for telomeric DNA maintenance to offset telomere loss from the end-replication problem, processing, and degradation. Most cancer cells up-regulate expression of hTERT, in order to increase telomerase activity.17 In 10−15% of cancers, a recombination based telomere lengthening pathway called alternative lengthening of telomeres (ALT) supplants the need for telomerase.18 The ALT pathway is particularly prevalent in astrocytic brain tumors and osteosarcomas.19 A second connection between telomere biology and neoplasias is the effect of overall length of telomeric DNA and the actions of telomere-associated proteins on genetic stability. Telomere erosion has been argued to serve as a tumor suppression mechanism because it can activate the senescence pathway, which suppresses neoplastic transformation.5 However, even in the presence of telomere erosion, senescence is not activated in the absence of the Rb/p16 or p53 pathways, leading to genetic instability and increased potential for transformation. These pivotal roles of telomere maintenance in tumorigenicity are evidenced by the ability to inhibit tumor growth by inhibiting telomerase in cancer cells and by the ability to convert primary human cells to cancer cells by combining constitutive telomerase expression with oncogene expression and/or loss of a tumor repressor gene. For example, the simian virus 40 large-T oncoprotein together with an oncogenic allele of H-ras and ectopic expression of hTERT has been demonstrated to transform a variety of cell types.20 For more comprehensive discussions on the roles of telomerase and telomere maintenance on tumor biology, several recent reviews are recommended.21−24 ■ TARGETING TELOMERE MAINTENANCE FOR CANCER THERAPY When considering anticancer drug targets, several advantages for targeting telomerase and telomere maintenance exist when compared to other targets and pathways. Most important is the apparently universal requirement for telomere maintenance in cancer cells including putative cancer stem cells.25,26 In addition, normal human cells including stem cells express lower telomerase activity and generally maintain telomeres at longer lengths when compared to cancer cells.27,28 These features provide a level of specificity with telomerase inhibition. Telomerase activity itself was validated as an anticancer drug target in 1999 when two independent studies reported that overexpressing dominant-negative-hTERT mutants resulted in telomerase inhibition, telomere shortening, morphological changes, growth arrest, and apoptosis in a variety of tumor cell lines.29,30 Since that time, several strategies for targeting telomere maintenance and telomerase-positive cancer cells have been explored (Figure 1). The initial approach explored was the direct inhibition of telomerase’s enzymatic activity. As validation of the approach, the telomerase inhibitor imetelstat (1, GRN163L) has been explored in anticancer therapy in several clinical trials.31 Direct targeting of the telomere itself has also been accomplished with specific DNA binding agents that bind to and stabilize the G-quadruplex structures formed by telomeric DNA.32,33 The high activity of the hTERT promoter in cancer cells has resulted in the exploration of hTERT- promoter driven expression of cancer killing genes and viruses,34 and one clinical trial for hTERT-driven replication of an oncolytic virus is underway.35 Immunotherapy targeting hTERT-derived peptides has also been explored leading to several clinical trials.36 Each of these platforms will be discussed further below. ■ CONSEQUENCE OF TELOMERASE INHIBITION AND TELOMERE DISRUPTION If telomere maintenance is to be considered a viable anticancer drug discovery platform, it is paramount to understand the consequences of disrupting telomere maintenance for both healthy and cancerous cells. The simplest and best explored telomere-targeted anticancer drug platform is the inhibition of telomerase enzymatic activity. When telomerase activity is inhibited, DNA replication attending cellular proliferation results in telomeric DNA erosion.29,37 Erosion of telomeric DNA eventually results in dysfunctional telomeres that initiate TIF formation.5,6 Generally, proliferating human cells lose 50− 200 base pairs of telomeric DNA during each cell cycle in the absence of telomerase activity.38 Therefore, a lag between the phenotypic consequences of telomerase inhibition and the initiation of telomerase inhibition therapy is likely to occur. The duration of this phenotypic lag depends on both the initial length of telomeric DNA, which ranges from 5000 to 15 000 base pairs in human cancer cells, and the rate of erosion. Importantly, it is suspected that the length of a few telomeres, Figure 1. Many ways to target the overexpression of telomerase in cancer cells have been developed. Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXB
  • 3. not the average length of all telomeres, dictates the cellular response to telomerase inhibition.39 This suggests that some tumor types may be especially sensitive to telomerase inhibition. Once telomeres erode to a critically short length, they cannot function properly,40 perhaps because they cannot be folded into the proper protected state, presumably a t- loop.16 In most cancer cells, the loss of telomere function results in apoptosis, whereas in normal cells, telomere loss results in cell cycle arrest. Notably, telomerase inhibition has also been observed to promote the ALT phenotype in tumors concomitant with up-regulation of the PGC-1β, a master regulator of mitochondrial biogenesis.41 This report offers critical insight into the mechanisms of adaptation that may be in play when tumors are subjected to telomerase inhibition. Inhibition of telomerase by some methods, for example, antisense oligonucleotides targeting the hTERT mRNA,42 can result in immediate effects that are not preceded by the phenotypic lag characteristic of telomerase activity inhibitors and are not coupled to erosion of telomeric DNA. While a precise explanation remains missing, one intriguing possibility is that hTERT has several noncanonical properties that are separate from its enzymatic activity, and inhibition of these may have immediate effects.43 Proposed additional activities of telomerase include telomere capping, inhibition of apoptosis, activation of the Wnt pathway, and activation of the NFκB pathway.44,45 Thus, down-regulation of hTERT can potentially lead to immediate antiproliferative effects. Understanding why some cells are immediately impacted by the loss of hTERT and others are not remains to be determined. Perhaps some cancer cells become oncogenically addicted to hTERT and require hTERT to maintain elevated growth signaling. Currently, one of the most studied small-molecule telomerase inhibitors are G-quadruplex stabilizers.46 This class of molecules can impact telomere functions in at least two primary mechanisms. One is by inhibiting the enzymatic activity of telomerase by preventing telomerase from binding to its primer or by interacting with the nascent DNA telomerase product. The second is to dissociate or block telomere-binding proteins from the telomere, disrupting the formation of the protective cap. As expected, the former mechanism results in a delayed but highly telomerase-expression-specific effect while the second mechanism is more immediate but may suffer from reduced specificity because formation of proper telomere structure is required in all cells including normal somatic and stem cells. Early evidence suggests that the therapeutic index for telomere disrupters is surprisingly high at least in a mouse model.47,48 One potential reason that telomere disruption may be well tolerated in normal cells is that the DNA damage response at the telomere may be transient and reversible. Evidence for this hypothesis comes from experiments using a ligand-stabilized version of the telomere binding protein Protection of Telomeres 1 (POT1). POT1 is essential for telomere protection, and the ligand-stabilized POT1 allowed POT1 to be inhibited transiently and reversibly. In cancer cells, POT1 inhibition resulted in cell death. But in normal cells, POT1 inhibition resulted in arrest that was reversed by reactivating POT1.49 ■ SMALL-MOLECULE TELOMERASE INHIBITORS Once the importance of telomerase activity for cancer cell growth was recognized, several academic and pharmaceutical groups initiated programs to discover small molecule telomerase inhibitors. Approaches toward the identification of effective telomerase inhibitors have included repurposing of viral reverse transcriptase inhibitors, screening of natural products, high throughput screens of diverse compound libraries, and more recently, the beginnings of structure-based discovery and optimization. Despite these efforts, small molecule telomerase inhibitors have yet to reach the clinic. In this section, the varieties of approaches toward direct telomerase activity inhibition with small molecules will be discussed. ■ NUCLEOSIDE ANALOGUES Nucleoside analogues developed as viral reverse transcriptase inhibitors have been investigated as telomerase inhibitors (Figure 2). Nucleoside analogues can hypothetically affect the telomere through several mechanisms, including telomerase inhibition and chain termination that can lead to telomere shortening and perturbation of telomere sequence that can lead to disruption of telomere function. The chain terminator 3′- azido-2′,3′-dideoxythymidine (2, AZT), the first nucleoside reverse transcriptase inhibitor approved for the treatment of HIV-1, inhibits telomerase in vitro and in vivo, giving rise to telomere shortening.50,51 In cultured cells, treatment with 2 resulted in telomere shortening, cell-cycle perturbation, and increased p14ARF expression,52 and 2 exhibited synergistic effects when combined with other chemotherapeutic drugs.53−55 Several nucleoside HIV reverse transcriptase inhibitors including stavudine, tenofovir, didanosine, and abacavir also inhibit telomerase activity and lead to telomere erosion in cultured cells, but non-nucleoside HIV RT inhibitors did not inhibit telomerase in these studies.56 HIV patients have been taking these and other nucleoside reverse transcriptase inhibitors for decades, leading to possible effects from long- term telomerase inhibition. This potential side effect has been investigated in a small patient population, and it was found that telomere length was inversely correlated with total time patients were exposed to nucleoside reverse transcriptase inhibitors.57 Although other nucleoside triphosphate analogues, including arabinofuranylguanosine and 6-thio-7-deaza-2′-deox- yguanosine 5′-triphosphate, ddGTP, ddATP, and ddTTP, produced substantial inhibition of telomerase and telomere shortening, they have not been further tested in cell cultures or in animal models and are likely to be nonspecific, thus limiting Figure 2. Representative nucleoside analogues identified as telomerase inhibitors. Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXC
  • 4. their clinical utility.51,58 As telomerase structural biology increases, it may be possible to rationally design nucleoside analogues with the requisite telomerase specificity for this platform to be viable. ■ NON-NUCLEOSIDE ANALOGUES Non-nucleoside, small-molecule telomerase inhibitors have been identified by a variety of approaches. High throughput screening of chemical libraries has been used to identify telomerase inhibitors (Figure 3). The isothiazolone derivative 2-[3-(trifluoromethyl)phenyl]isothiazolin-3-one (3, IC50 = 1 μM) was identified from a screen of a 16 000-member, chemical library using rat telomerase (Figure 3).59 The reducing agent dithiothreitol was found to protect telomerase from inhibition by 3, suggesting that inhibition resulted from interaction between the isothiazolone moiety of 3 with the sulfhydryl group of key cysteine residue(s).59 The nitrostyrene derivative 3-(3,5-dichlorophenoxy)nitrostyrene (4, IC50 = 0.4 μM)60 and the quinoxaline derivative 2,3,7-trichloro-5-nitroquinoxaline (5),61 two noncompetitive telomerase inhibitors, were also identified by screening chemical libraries. At low concentrations (1 μM), 4 was not acutely cytotoxic to HeLa cells, but chronic exposure led to progressive telomere erosion followed by the induction of senescence consistent with telomerase inhibition.60 Long-term treatment of the breast cancer cell line MCF7 with 5 (1 μM) also resulted in telomere erosion, chromosome abnormalities, and senescence.61 The crystal structure of TERT from the beetle Tribolium castaneum62 has been used in limited structure based design of telomerase inhibitors. Because 2H-pyrazole and dihydropyr- azole derivatives have been shown to have strong anticancer, antmalarial, antiviral, and anti-inflammatory activity, they were selected as a starting scaffold for telomerase inhibition.63−65 Aryl-2H-pyrazoles with 1,3-benzodioxoles and electron with- drawing aryls groups and sulfonyl-4,5-dihydropyrazoles dis- played telomerase inhibition with IC50 values ranging from 0.8 to 10 μM in a variety of cancer cells (Figure 4).63−65 Since previous studies have shown that tetracyclic coumarins have notable anti-HIV activities and that the HIV and telomerase reverse transcriptase proteins display structural homology, dihydropyrazolepiperidine derivatives containing coumarin moieties were also synthesized.63,66 These molecules showed anticancer activity against gastric cancer and the most potent inhibited telomerase in crude cell extracts with IC50 values in the 1−2 μM range.63,66 Results from docking studies were used to suggest that these molecules bind the hTERT active site through key intermolecular hydrogen bonding and π−cation interactions with invariant residues Lys189, Asp254, and Gln308 and by projecting aryl and dihyrdopyrazole rings into the hydrophobic dNTP-binding pocket.63,65,66 The π−cation interaction is consistent with the structure−activity relationship (SAR) trend in which electron withdrawing groups on the aromatic ring attenuate inhibition, whereas electron donating groups increased inhibition. 3D quantitative SAR (QSAR) was used to rationalize the proposed binding mechanism, and a correlation coefficient of r2 = 0.966 was found for predicted and observed IC50 for a small, chemically similar test set.65 Other telomerase inhibitors proposed to bind telomerase using docking studies include 2-chloropyridines and 2-aminomethyl- 5-(quinolin-2-yl)-1,3,4-oxadiazole-2(3H)-thionequinoline de- rivatives, although these studies did not offer strong validation of the binding models.67,68 Several natural products have been identified as telomerase inhibitors as well (Figure 5). Helenalin (6), a natural sesquiterpene lactone, was found to inhibit telomerase apparently by reaction with a cysteine residue similar to 3.69 o-Quinone containing compounds, for example, tanshinone 2A (7), a plant diterpene known to exhibit diverse pharmacological activities, also inhibit telomerase (IC50 = 3 μM). The quinone was fully responsible for inhibition that was time- and DTT- dependent. It was also found that 7 undergoes redox cycling to produce hydrogen peroxide. The data suggest that quinone redox cycling inhibits telomerase, perhaps by oxidation of conserved cysteine residues.70 Another apparently reversible inhibitor is 8 (U-73122), commonly used as an inhibitor of Figure 3. Examples of small-molecule telomerase inhibitors identified by screening small molecule libraries. Figure 4. Telomerase inhibitors reported to bind the telomerase active site based on docking studies. Figure 5. Natural product and natural product derivatives identified as telomerase inhibitors. Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXD
  • 5. phospholipase C, which was reported to be a potent (IC50 = 0.2 μM) telomerase inhibitor.71 Inhibition was shown to be dependent on the pyrrole-2,5-dione functional group, suggest- ing potential alkylation of telomerase. β-Rubromycin (9), a quinone antibiotic, inhibited telomerase with an IC50 of 3 μM.72 Mechanistically, 9 appears to be a competitive inhibitor versus the telomerase primer, it also inhibits retroviral reverse transcriptases, mammalian DNA polymerases, and terminal deoxynucleotidyl transferase,73 indicating that rubromycins are not telomerase specific. The search for more biologically effective telomerase inhibitors has encouraged innovative yeast-based screens. Natural products chrolactomycin (10) and UCS1025A (11) were identified as direct telomerase inhibitors using a forward chemical genetics screen in a yeast system (Figure 6).74 A yeast strain with artificially shortened telomeres was used to screen a microbial metabolite library. Since telomeres were shortened, the strain was more sensitive to telomerase inhibition than wild type. Therefore, compounds with selective activity in the telomere shortened strain were more likely to be telomerase inhibitors. With this assay, two new telomerase inhibitors 10 (IC50 = 0.5 μM) and 11 (IC50 = 1.3 μM) were identified. In addition, the Hsp90 inhibitor radicicol was also identified as a telomerase maintenance inhibitor in screen.74 The synthesis of 11 has been accomplished, but structure−activity relationships of analogues have not been reported.75,76 More recently, an alternative yeast based assay was developed using recombinant human telomerase. Expression of human telomerase activity in yeast results in a growth delay; therefore, potential inhibitors of human telomerase could rescue the transfected cells. By use of this screen, several potential telomerase inhibitors with moderate potency (IC50 values of ∼1−10 μM) were identified to confirm the screening platform.77 The most potent compound identified in the screen was the dichlorothiophene 12 (CD11359). One promising class of small molecule telomerase inhibitors is the potent enoylaminobenzoic acids represented by 13 (BIBR1532, IC50 = 0.093 μM) and 14 (BIBR1591, IC50 = 0.47 μM) that function as selective mixed-type, noncompetitive telomerase inhibitors and were developed from systematic structure−activity correlations (Figure 7).37 Limited structure− activity relationship data are publically available and show that the α−β unsaturated amide is essential and modifications to the naphthyl group are not well tolerated.78 The mechanism of inhibition by 13 is distinct from those by other non-nucleoside compounds or antisense oligonucleotides, as it does not cause chain termination events but rather inhibits the formation of long reaction products. Compound 13 seems to act allosteri- cally to impair the unique ability of telomerase to catalyze addition of multiple telomeric repeats onto a DNA primer.79 Acute treatment with high concentrations of 13 caused direct cytotoxic effects on malignant cells of the hematopoietic system, resulting in end to end fusions, phosphorylated p53, and decreased expression of TRF2, consistent with a telomere- associated effect.80 Compound 13 is also capable of sensitizing drug-resistant cells to chemotherapeutic agents.81 Chronic Figure 6. Small molecules identified as telomerase inhibitors utilizing yeast-based assays. Figure 7. Treatment with 14 reversibly inhibits telomere length maintenance. (A) Structures of 13 and 14. (B) Chronic treatment of NCI-H460 cells with 14 results in growth arrest that can be reversed after ending the treatment. Open circles are untreated cells. Filled triangles are cells treated with 14. Treatment was halted at day 220; note the return to proliferative growth. (C) Southern blot analysis of average telomere length of telomerase inhibitor treated cells and cells released from telomerase inhibition. Treatment with 14 resulted in loss of telomere length and exhibited ∼1.5 kb of telomeric DNA (lanes 1 and 2). Removal of 14 resulted in telomere elongation to ∼3 kb of telomeric DNA (lane 3). Untreated cells are shown in lane 4 for comparison and have 3 kb of telomeric DNA. Parts B and C are reprinted by permission from Macmillan Publishers Ltd.: The EMBO Journal (http://www.nature.com/emboj/index.html) (Damm, K.; Hemmann, U.; Garin-Chesa, P.; Hauel, N.; Kauffmann, I.; Priepke, H.; Niestroj, C.; Daiber, C.; Enenkel, B.; Guilliard, B.; Lauritsch, I.; Muller, E.; Pascolo, E.; Sauter, G.; Pantic, M.; Martens, U. M.; Wenz, C.; Lingner, J.; Kraut, N.; Rettig, W. J.; Schnapp, A. EMBO J. 2001, 20, 6958−6968),37 copyright 2001. Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXE
  • 6. exposure to lower doses of 14 resulted in progressive erosion of telomeric DNA leading to senescence.37 However, the lag phase between telomerase inhibition and the impact on proliferative capacity under the chronic treatment condition is a major concern,37 as is the apparently high protein binding evidenced by the significantly decreased potency in crude cell extracts compared to purified telomerase.78 More than 120 days, approximately 135 population doublings, was necessary for treatment with 13 to inhibit cell proliferation of NCI-H460 lung cancer.82 Interestingly, washout of 14 in telomerase positive cells allowed recovery of telomere length (Figure 7).37 ■ SMALL MOLECULES THAT TARGET hTER AND THE hTER-TELOMERIC DNA HETERODUPLEX In addition to targeting the enzymatic activity of telomerase, exploratory efforts in identifying ligands that target hTER suggest that RNA-binding ligands may be a productive platform for targeting telomerase. In two separate reports, oligonucleo- tides were used to identify regions of hTER that can be specifically targeted to prevent proper telomerase assem- blage.83,84 Specifically, the CR4-CR5 and pseudoknot domains of hTER are susceptible. Detailed analysis demonstrated that discrete domains of hTER are blocked without losing the ability of hTERT to associate with hTER.84 This suggests that the CR4-CR5 and pseudoknot domain targeting oligonucleotides effectively cause formation of a dysfunctional telomerase complex. Subsequently, small molecules that bind hTER have been identified,85−87 and several of these including amino- glycosides and the bisbenzimide Heochst 33258 have been shown to inhibit telomerase,86 presumably by binding hTER and perturbing proper assemblage of the holoenzyme complex (Figure 8). Efforts to target the telomerase template RNA− DNA primer/nascent product heteroduplex have also been reported.88,89 In these studies, several intercalators were identified that inhibit telomerase and bind preferentially to the RNA−DNA heteroduplex as opposed to G-quadruplex DNA. Advances in this area include second generation intercalators and bis-intercalators that are fairly potent telomerase inhibitors in cell free assays. A small library of bis- ethidium derivatives with varying linkers was screened. Surprisingly, the library exhibited a narrow range of moderately potent telomerase inhibition without significant improvement over the parent monointercalator.88 Data for cell based assays using small molecule targeting of hTER or the hTER/DNA heteroduplex have not yet been reported, so it is still too early to tell if this approach will be clinical viable. ■ TELOMERASE INHIBITION BY INDIRECT MEANS hTERT Transcription Regulators. Cellular telomerase activity is regulated at various levels including transcription, mRNA splicing, post-transcriptional and post-translational modifications of hTER and hTERT, transport and subcellular localization, assembly of active telomerase ribonucleoprotein, and telomeric DNA accessibility. Transcriptional regulation of hTERT is thought to be the major mechanism of telomerase regulation.90 The hTERT promoter contains binding sites for several transcription factors, suggesting that the regulation of hTERT expression may be subject to multiple levels of control by various factors and dependent on cellular contexts.91 A primary activator of hTERT is the c-Myc oncogene, a known regulator of cell proliferation, growth, and apoptosis.92 Targets of the Myc family are involved in various aspects of cellular functions, including cell cycle, growth, differentiation, and life span. c-Myc expression correlates with hTERT, and both are Figure 8. Ligands that inhibit telomerase by binding to the RNA subunit hTER (Hoechst 33258 and neomycin) and the telomerase DNA/RNA heteroduplex (bis-ethidium compounds). Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXF
  • 7. up-regulated in highly proliferative and immortal cells but down-regulated during terminal differentiation and in non- proliferative cells. Therefore, it is not surprising that inhibitors of c-Myc, butein,93 gamboic acid,94 and genistein,95 also inhibit telomerase. Additionally, genistein treatment resulted in down- regulation of the protein kinase Akt and thus hTERT phosphorylation, indicating that genistein can inhibit telomer- ase by at least two mechansisms: decreased hTERT tran- scription and decreased posttranslational activation of hTERT.95 Second generation tyrosine kinase inhibitors and the lipid- derived signaling molecule ceramide have been shown to inhibit hTERT promoter activity by altering the functions of Sp1 and Sp3.96,97 Sp1 has been shown to collaborate with c-Myc in order to initiate hTERT transcription.98 Imatinib, dasatinib, and nilotininb decrease telomerase activity and hTERT expression by decreasing Sp1 nuclear levels and activation and reducing the nuclear translocation of hTERT via suppression of Akt phosphorylation.96 Agents that act on hormonal pathways have also been linked to telomerase activity regulation. For example, estrogen activates telomerase in hormone-sensitive tissues such as mammary and ovary epithelial cells,99,100 while the selective estrogen receptor modulator tamoxifen reduces telomerase activity in cells where it functions as an estrogen antago- nist.101,102 Progesterone also has an antagonistic effect on estrogen-induced hTERT expression.103 These observations highlight the fact that the polypharmacological effects of many anticancer drugs are anticipated to include inhibition of telomere maintenance as a mechanism of their anticancer effects. Negative Regulators of hTERT Transcription. The lack of telomerase activity in somatic cells appears to be a consequence of transcriptional repression of the hTERT expression. The loss of transcriptional repression results in the up-regulation of hTERT expression and telomerase activity as seen during carcinogenesis and cellular immortalization. Several transcriptional repressors of hTERT have been identified. One such negative regulator of hTERT transcription is Mad1. Mad1, c-Myc, and Max are part of a family of transcription factors that bind to E-box-containing promoters to either down-regulate or up-regulate gene expression. Mad1 is up-regulated in telomerase negative somatic cells, c-Myc is up- regulated in cancer cell lines, and Max is expressed ubiquitously.104 Consistent with its role in regulating telomerase, overexpression of Mad1 resulted in decreased hTERT promoter activity in gene reporter assays.105 Another negative regulator of hTERT transcription is the tumor suppressor protein p53, which induces cell cycle arrest or apoptosis in response to DNA damage and other cellular stresses to inhibit tumor formation. p53 has been shown to inhibit telomerase activity through the repression of hTERT transcription.106 Repression of telomerase activity was also observed by overexpressing pRB in human squamous cell carcinomas107 and E2F1 in head and neck squamous cell carcinomas.108 pRB and E2F1 have the potential to transcrip- tionally repress hTERT activity either independently or in tandem with one another. Another tumor suppressor involved in the transcriptional repression of hTERT is Wilm’s tumor protein 1 (WT1). WT1 has also been shown to significantly decrease hTERT mRNA and telomerase activity in HEK293 cells.109 The role of WT1 in hTERT repression is cell type dependent, as the WT1 gene is expressed only in specific cells such as the kidney, gonad, and spleen. Other negative regulators of hTERT transcription include the myeloid cell-specific zinc finger protein 2, interferon-α, and TGF-β. Interferon-α- and -γ-induced sensitization to apoptosis is associated with repressed transcriptional activity of the hTERT promoter in multiple myeloma.110 Transforming growth factor β, a potent tumor suppressor, inhibits telomerase through the transcription factors SMAD3 and E2F.111,112 TGF- β activated kinase 1 (TAK1) has been shown to repress transcription of the human telomerase reverse transcriptase gene. TAK1-induced repression was caused by the recruitment of histone deacetylase to Sp1 at the hTERT promoter.113 T-cell acute lymphoblastic leukemia protein 1 (TAL1) is also a negative regulator of the hTERT promoter. Overexpression of the proto-oncogenic protein TAL1 leads to a decrease in hTERT mRNA levels and consequently reduced telomerase activity.114 Histone deacetylase (HDAC) inhibitors, such as trichostatin A, have been shown to induce apoptosis and down- regulate telomerase activity via suppression of hTERT expression in leukemic U937 cells.115−117 Dynamic assembly of the E2F−pocket protein−HDAC complex plays a central role in the regulation of hTERT. For example, a small molecule, CGK 1026, inhibits the recruitment of HDAC into E2F-pocket protein complexes assembled on the hTERT promoter.118 These data underscore the fact that several anticancer treatment strategies are likely to result in telomerase inhibition indirectly. Several reports demonstrate the indirect effects on telomerase activity that are possible with small molecules. Tea catechins such as (−)-epigallocatechin 3-gallate (15) prevent growth of a variety of cancer cell lines and inhibit telomerase activity in a concentration dependent manner by decreasing hTERT mRNA levels.110,119 The changes in hTERT mRNA levels can be attributed to chromatin remodeling due to the hypoacetylation of histones at the hTERT gene promoter and decreased methylation of the hTERT promoter at the E2F- 1 binding site.119 Consistent with the observed decrease activity of the hTERT promoter, 15-treated U937 monoblastoid leukemia and HT29 colon adenocarcinoma cells experienced decreased proliferation potential commensurate with telomere shortening. In addition, 15 was also shown to induce a severe reduction in cell growth with enhanced membrane blebbing, cell swelling, apoptosis, and necrosis in laryngeal squamous cell, small-cell lung, and breast cancers.119−121 Further studies showed that the tea catechin derivative MST-312 (N,N′- bis(2,3-dihydroxybenzoyl)-1,2-phenylenediamine) appears to inhibit telomere maintenance by directly targeting telomerase (IC50 = 0.67 μM), resulting in telomere erosion at a 20-fold lower dose than 15 in the monoblastic leukemia cell line U937.122 GSK3 inhibition by 6-bromoindirubin-3′-oxime caused suppression of hTERT expression, telomerase activity, and telomere length in various cancer cell lines and decreased hTERT expression in ovarian cancer xenografts.123 Other repressors of hTERT transcription are retinoids, derivatives of vitamin A, which have been shown to down-regulate telomerase leading to telomere erosion.124 Retinoic acid receptor α and retinoid-X receptor-specific agonists synergistically down- regulated hTERT and induced tumor cell death.125 Posttranslational Regulators of hTERT. Perturbing posttranslational modifications represents another mechanism for targeting telomerase. Because telomerase activity and cellular location is controlled by its phosphorylation state, phosphatase and kinase inhibitors can have direct effects on telomerase. For example, the phosphatase 2A inhibitor okadaic Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXG
  • 8. acid prevents telomerase inhibition by phosphatase 2A in breast cancer cells.126 Protein kinase C (PKC) has been shown to enhance telomerase activity in certain cell types, and PKC inhibitors have been shown to inactivate telomerase in nasopharyngeal127,128 and cervical129 cancer cells among others. Imatinib mesylate (Gleevec), a tyrosine kinase inhibitor, also down-regulates telomerase activity and inhibits proliferation in telomerase expressing cells.130 The protein kinase Akt has well- established roles in telomerase activation. Akt phosphorylates hTERT and activates telomerase in vitro, while the nonspecific Akt inhibitor wortmannin causes loss of telomerase activity.127 The tyrosine kinase c-Abl is also capable of phosphorylating hTERT. The SH3 domain of c-Abl interacts with and phosphorylates hTERT resulting in a decrease in telomerase activity.131 These observations demonstrate that telomerase activity can be directly regulated by protein phosphorylation independent of transcriptional regulation and highlight that telomerase inhibition may contribute to the observed anticancer effects of several drugs. In addition to regulation by kinases, it has recently been determined that hTERT is a substrate of caspases-6 and −7.132 Interestingly, the caspase sites are unique and the cleavage products appear remarkably stable. Hsp90 Inhibitors. The telomerase ribonucleoprotein complex requires the Hsp90 chaperone complex, including p23, Hsp70, p60, and Hsp40/γdl for proper assemblage. Hsp90 associates directly with hTERT and is required to maintain telomerase in an active conformation at least under some conditions.133,134 Hsp90 inhibitors cause inactivation, destabi- lization, and degradation of Hsp90 target proteins including hTERT.46,133 Since several oncogenic proteins including c-Myc, Akt, and mutated p53 require Hsp90 for their activity, inhibition of this chaperone should block multiple oncogenic pathways. Hsp90 inhibitors such as novobiocin, radicicol, geldanamycin, and 17-AAG were also shown to inhibit telomerase activity.134−137 Recently, curcumin has been shown to inhibit nuclear localization of telomerase by dissociating the Hsp90 co-chaperone p23 from hTERT.93 Curcumin treatment limited the association between p23 and hTERT while not affecting Hsp90 interaction with telomerase. ■ G-QUADRUPLEX STABILIZERS The most advanced area of small-molecule drug discovery directed at the telomere is G-quadruplex-stabilizing ligands. G- quadruplexes are unique structures formed by folding G-rich nucleic acid sequences. These structures are heterogeneous with a variety of structural topologies.138 Folding topologies are influenced by sequence, buffer conditions including identity of cations present, molecular crowding, and other parameters. Significant effort toward determining the structures of human telomeric DNA138,139 and the biological functions of G- quadruplex DNA at the telomere and in other G-rich sequences embedded in the chromosome140 has been reported. Because telomerase must hybridize to its DNA primer to afford telomeric DNA extension, it was proposed that G-quadruplex formation would render a primer inert to telomerase.141 Seminal studies by Zahler et al. showing that telomerase could not extend a folded G-quadruplex were consistent with this hypothesis and initiated the field of G-quadruplex binding ligands as a platform for telomerase inhibition.141 This report was quickly followed by the cardinal work of the Hurley and Neidle labs providing the first G-quadruplex ligands that effectively inhibited telomerase.142,143 Since these reports, a wide variety of pharmacophores have been explored as G- quadruplex ligands, and most of these are polycationic aromatic hydrocarbons (Figure 9).144 Since these initial studies demonstrating a relationship between telomeres and G- quadruplexes, these structures have become much more complex as drug targets than originally assumed, as they are predicted to be present in a wide variety of genomic locations with increased density in promoter regions.145 Several of these promoters, including those for hTERT,146 c-kit,147,148 and c- myc,149,150 have verified abilities to fold into G-quadruplex structures. Much of the work on G-quadruplex structure, Figure 9. Representative G-quadruplex ligands that inhibit telomerase. (A) 16 (BRACO19), a 3,6,9 trisubstituted acridine derivative. (B) RHPS4, a pentacyclic acridine derivative. (C) and (D) are models of 16 bound to a G-quadruplex formed by a dimer of TTAGGGTTAGGG. The structure was rendered from PDB 3CE5. 16 is shown in red. dG residues are shown in cyan, and dA and dT residues are shown in gray. Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXH
  • 9. biological relevance, and small-molecule targeting has been recently and extensively reviewed, and readers are directed to these.144,151,152 Here, we will simply highlight the major concerns with targeting G-quadruplexes as a specifically telomere-directed anticancer drug modality and highlight recent approaches toward advancing telomere-binding ligands to the clinic. The first G-quadruplex ligands explored as telomerase inhibitors were disubstituted anthraquinone derivatives.153 Since that time, several other polycyclic heteroaromatic scaffolds and macrocycles have been explored allowing the development of potent telomerase inhibitors with increasing specificity for G-quadruplex DNA compared to duplex DNA, and these ligands are available at the G-quadruplex ligands database (http://www.g4ldb.org/ci2/index.php).154 Impor- tantly, high resolution crystal and NMR structures of human telomeric G-quadruplexes can inform rational design to improve G-quadruplex affinity and specificity.152 From structural studies, it is clear that critical features known to enhance G-quadruplex binding include a delocalized π-electron system capable of stacking onto G-tetrad a positive charge that can rest in the center of the G-quartets to interact with the carbonyl oxygens from guanine that form the core of the G- quartets, and positively charged substituents that can interact with the grooves of the G-quadruplex (Figure 9c and Figure 9d).138 From a purely biochemical view, the most significant current issues in G-quadruplex ligand design are specificity for G-quadruplex DNA compared to duplex DNA and specificity for G-quadruplex type, i.e., telomeric versus other DNA and RNA quadruplexes that may be present.155 From a pharmaceutical standpoint, the major issue concerning trans- lation of G-quadruplex ligands is the poor pharmacokinetic profile associated with the typically hydrophilic and positively charged molecules.156 One of the most significant recent advances toward selective G-quadruplex targeting is the ability to rationally design molecules using structure-based approaches. Structures of G- quadruplexes formed by the human telomeric DNA sequence by X-ray crystallography and NMR spectroscopy have provided templates for rational design of G-quaduplex ligands.157,158 These advances have allowed for the development of rationally designed molecules with increased.159,160 These promising examples suggest that G-quadruplex specificity is achievable. The more challenging hurdle that still remains is the poor pharmacokinetic and pharmacodynamic properties of promis- ing G-quadruplex compounds. These hurdles may not be insurmountable, as evidenced by the G-quadruplex ligand and fluoroquinolone derivative quarfloxin, which binds to nucleo- lin/rDNA and inhibits Pol 1 transcription, thus inducing apoptosis in adenocarcinoma cells, and which has had some clinical success.161 ■ DISRUPTION OF TELOMERE MAINTENANCE USING ANTISENSE AND RELATED APPROACHES Ribozymes. Hammerhead ribozymes are small catalytic RNA molecules that have the capacity to specifically cleave target RNA after NUH sequences, where N is any nucleotide, U is uridine, and H is any nucleotide except G.162 Design of hammerhead ribozymes is relatively simple, and specificity can be achieved by incorporating sequences complementary to the target. Because of this nature, ribozymes have been investigated as a platform for telomerase inhibition via targeted RNA cleavage of hTER or hTERT mRNA. The template region of hTER is an ideal target for ribozyme cleavage because it has the requisite target sequence for Hammerhead or HDV ribozymes, is essential for telomerase activity, and it may be solvent exposed even in the telomerase holoenzyme. Studies using melanoma,163 hepatoma, colon cancer,164 and endometrial cancer cells165 have correlated ribozyme expression with a marked reduction in telomerase activity resulting from diminished telomerase RNA levels. While telomere lengths were greatly attenuated in some carcinomas, other types did not display this phenotype.163,165 Treatment with template-targeting ribozymes also resulted in significant growth retardation with proliferation rates reduced compared to parent cells.166 Furthermore, treated cells experienced telomere shortening in concert with altered or dendritic morphologies, which are indicative of senescence,166 G0/G1 growth arrest, and increased apoptotic rate.164 Circularization and addition of a six nucleotide poly(A) sequence to these ribozymes increased target RNA cleavage while conferring protection from cellular exoribonuclease degradation.167 TERT mRNA levels have been targeted by a Hammerhead ribozyme targeting the initial nucleotides in hTERT mRNA, which may be essential for the proper binding of the translation initiation complex to the 5′ mRNA cap structure.168 Ribozyme cleavage in the middle of hTERT mRNA has also demonstrated decreased telomerase activity while also influencing rapid cell death (4 days) even without diminished telomere length, consistent with potentially nontelomeric roles of hTERT in some cancer cells.169 Oligonucleotides and Template Antagonists. Antisense (AS) oligonucleotides (ODNs) can inhibit telomerase by several mechanisms including blocking hTERT translation, reduction of hTERT mRNA and hTER levels, and sterically blocking the association of telomerase with the telomere (template antagonists).42 However, oligonucleotides present many challenges toward achieving efficient intratumoral telomerase inhibition. Nonetheless, studies have demonstrated that unmodified antisense oligonucleotides can in fact potently inhibit telomerase activity under in vitro conditions.170−172 One of these was a genetic approach using a plasmid borne 125 nucleotide hTR-antisense RNA, which inhibited tumor growth in mouse xenografts.170 To make up for the shortcomings of unmodified ODNs, a spectrum of ODNs with chemically modified backbones have been utilized to enhance the therapeutic potential of these antisense strategies. Phosphorothioate (PS) ODNs composed of telomere repeat sequences ranging from 6 to 24 bases and targeting the template of hTER showed significant telomerase inhibition in various cancer cell lines.173,174 While longer PS- ODNs did not have any effect on tumor growth, a 6-mer PS- ODN decreased cellular growth in Burkett’s lymphoma, increased cell population doubling times, induced dramatic levels of apoptosis, and reduced human xenograft tumors in mice to undetectable sizes.174 2′-O-Methyl-PS chimeric ODNs targeting the template sequence of hTER demonstrated potent telomerase inhibition with IC50 values in the nanomolar range.175,176 Similar ODNs blocked up to 95% of telomerase activity after 1 day of treatment when delivered with a cationic lipid delivery system and induced telomere erosion and subsequent limited proliferation commensurate with marked apoptosis in long-term studies with chronic treatment.177,178 Cellular delivery with chitosan-coated PLGA nanoparticles afforded even higher cellular uptake and the similar phenotype Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXI
  • 10. of progressive telomere shortening with an EC50 value of 10 nM.179 Direct addition of 2′-O-methoxyethyl-PS chimeric ODNs decreased telomere length and proliferation and growth rates in MCF-7 (breast)180 and DU145181 (prostate) cancer cells in culture and xenograft tumor models.180,181 Synergistic effects were seen in DU145 prostate cancer cells upon addition of chemotherapeutic agents such as cisplatin or carboplatin.181 Peptide nucleic acids (PNAs) completely covering the template region of hTER were 10−50 times more potent than similar PS-ODNs with IC50 values in the nanomolar range.182 The potent telomerase inhibition is due to high affinity of the PNA to the single stranded RNA template and suggests that affinity is not limited to base pairing but may also include interactions between hTERT and the peptide back- bone.182 Both cellular permeabilization and lipid-mediated delivery demonstrated similar IC50 values in DU145 (prostate cancer)83 and HME50-5 (immortal breast epithelial)182 cells. Conjugation of short peptide sequences to PNAs targeting the hTER template displayed increased telomerase inhibition with IC50 values 10-fold lower than PNAs alone.183 Targeting of hTER by 2-5A ODNs has led to a remarkable inhibition of telomerase activity concomitant with rapid hTER degradation and a greater than 50% decrease in cell viability in a time dependent manner.41,184−186 2-5A ODNs targeting hTER can also arrest cells in G2/M phase187 and induce apoptosis via activation of the caspase signaling cascade.186 Direct admin- istration of hTER-targeting 2-5A ODNs and combination therapy with interferon-β has demonstrated antiproliferative effects including a reduction in colony formation and a significant retardation in tumor growth and volume over time.184,185,188 In 2003, Asai et al. described a novel class of chemically modified N3′→P5′-thiophosphoramidates (NPS), which con- ferred high stability and sequence specific potency.42,189 The efforts led to the development of 17 (GRN163), a 13- nucleotide long NPS-ODN with sequence 5′-TAGGGTT- AGACAA-3′ that targets the template region of hTER by overlapping and extending four nucleotides beyond the 5′ template boundary (IC50 = 0.14 μM).189 Treatment of various cancer cell lines with 17 significantly reduced telomerase activity, promoted critical telomere loss, and instigated apoptosis while concomitantly up-regulating caspase activity.189 These results combined with impressive data from long-term experiments and its high potency and specificity indicated that 17 may be an effective anticancer agent.189 In order to improve its bioavailability, lipophilicity, and cellular uptake, a 5′ terminal palmitoyl group was attached through an amide bond to α-aminoglycerol of the thiophos- phoramidate 17 to produce the clinical candidate 1 (Figure 10).190 Comparison studies with 17 proved that 1 was just as potent at inhibiting telomerase in various cancers, could inhibit telomerase activity effectively without the use of a transfection agent, and demonstrated greater tumor uptake.190−192 Since then, many studies have been published revealing the extensive anticancer potential of this drug in a variety of cancers. A single treatment of 1 μM 1 yielded complete inhibition of telomerase activity in various cancers within hours of dosing and near complete inhibition could be maintained for up to 6 days, while chronic treatment allowed sustain inhibition with undetectable telomerase activity.193−198 Continued treatment for 2 weeks elicited morphological changes such as rounding and increased extension of filliopodia-like structures,193,198 reduced colony formation,194,196 and attenuated metastasis invasion.194 Longer periods of treatment with 1 triggered decreased proliferation,196 decline in cell viability,197,199 disruption of focal adhesions and actin filament organization,193 substantial telomere short- ening,194−196 reduced capacity to form neurospheres and mammospheres,191,199 down-regulation of cell cycle regula- tors,192 G0/G1 cell cycle arrest,191,198 up-regulation of APG5 and DAPK2 expression (genes implicated in the induction of apoptosis),197 disrupted membrane intergrity,191 and promo- tion of apoptotic cell death.191,197 1 also promoted a higher percentage of anaphase bridges, hallmarks of telomere dysfunction, and increased the frequency of γ-H2AX and 53bp1 foci, known molecular markers that localize to sites of DNA strand breaks and dysfunctional telomeres.200 In tumor models, 1 prevented tumor formation and metastasis from xenografts,194,195 significantly reduced growth and tumor volume,191,200,201 and enhanced antitumor effects in combina- tion therapy with doxorubicin, temozolomide, and irradia- tion.191,200 The successful preclinical experiments demonstrated the antitumor effectiveness of 1, and it remains the only telomerase inhibitor to have been tested in clinical trials. Lessons learned from clinical trials of 1 are essential for the field, as this represents the first telomerase inhibitor to reach the clinic. Currently, six phase I and phase I/II trials of 1 as a single agent and three combination trials are ongoing. Major objectives of these studies are safety, tolerability, and maximum tolerated dose. Early reports indicate that 1 is well tolerated, but nothing further has been reported. RNAi. RNA interference or silencing RNA (RNAi or siRNA) is double stranded RNAs that take advantage of a sequence- specific and natural mechanism that mediates gene silenc- ing.42,202 As an RNA that is complicit for telomerase activity, hTER is an excellent target for siRNA therapies. Transfection and intravenous injection of partially double-stranded PS- ODNs targeting the template region of hTER reduced cell proliferation in human cervical carcinoma and, more importantly, strongly reduced the number of metastases in murine malignant melanoma cells.203 siRNAs targeting nucleotides 11−19 of hTER expressed in various cancer cell lines including HCT116 colon cancer, LOX melanoma, HeLa cervical cancer, T24 bladder cancer, MCF-7 and BT424 breast cancers, and LNCaP prostate cancer, via lentiviral-vector expression, caused rapid growth inhibition and apoptosis while displaying negligible cellular cytotoxicity independent of Figure 10. Clinical tested telomerase inhibitor. (A) 1 is a modified oligonucleotide that contains a 5′-palmotyl group. The sequence is complementary to the hTER template domain. (B) The N3′→P5′-thiophosphoramidate linkage present in 1. Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXJ
  • 11. telomere length.204 Cancer cells exhibited knockdown of endogenous hTER as well as down-regulation of cell cycle progression genes (cyclin G2 and Cdc27) and specific genes that play pivotal roles in tumor growth, angiogenesis, and metastasis (integrin αV and Met proto-oncogene), suggesting a role for these genes in the hTER-mediated cell growth inhibition.204 Expression of siRNAs targeting hTERT mRNA in various cancers has caused decreased hTERT mRNA levels, reduced hTERT expression, and notable telomerase inhibition within 24 h of dosing and lasting up to 4 days.120,176,205−209 Stable transfection of these siRNAs led to complete suppression of telomerase activity in 3 days concomitant with the absence of hTERT mRNA expression.205,207 Cancers treated with siRNAs that deplete hTERT experienced decreased cell viability,120 prolonged population doubling times,209 suppression of anchorage-dependent proliferation,206 significant growth inhib- ition,208 and morphological changes indicative of senescence.205 Esophageal and colon adenocarcinomas displayed extensive telomere shortening,205,207 leading to chromosomes with telomere free ends suggestive of complete telomere erosion.205 Genetic silencing of hTERT led to an up-regulation of genes involved in DNA damage recognition and repair, p53 family proteins and their regulators, cell cycle inhibitors, proapoptotic proteins, and caspases.199,205,210 At the same time, hTERT depletion significantly attenuated the expression of integrin αV and β3 subunits and c-Met, proteins that are known to play important roles in cell differentiation, adhesion, motility, and tumor metastasis.206 In tumor models, hTERT siRNAs disrupted cell adhesion, migration, and invasion, promoted slower tumor growth,206 decreased cell viability and colony formation, induced cell cycle arrest,199 reduced tumor sizes and volumes in xenografts,206,208 and demonstrated additive effects upon combination therapy with doxorubicin and intereferon- γ.120,199 Interestingly, while expression of siRNAs targeting both hTER and hTERT concomitantly did decrease mRNA, RNA, and protein levels of telomerase components, cell proliferation, and xenograft tumor growth and promote pyknotic nuclear chromatin, the effect of silencing hTER alone was more pronounced than silencing either hTERT or both together.211 ■ IMMUNOTHERAPY The realization that hTERT is overexpressed in the vast majority of cancers prompted the investigation of hTERT as a tumor antigen. Soon after the discovery of widespread hTERT expression in tumors, hTERT peptide fragments were identified as tumor-associated antigens functioning through the major histocompatibility complex class I and class II pathways. In several cancer types, CD8+ cytotoxic T-cells specific for hTERT peptides, primarily hTERT-I540 (ILAKFLHWL), are detectable suggesting active hTERT-directed immune surveil- lance for these cancers.212 Subsequently, over 20 hTERT- derived peptides have been employed to boost antitumor immunity.2 The majority of these elicit a CD8+ cytotoxic T-cell response, while the others activate a CD4+ response. Several clinical trials of hTERT-based cancer vaccines have demon- strated the safety of this approach.213,214 Unfortunately, the impact on disease progression reported to date is modest. Some patients respond with significant tumor regression, but more generally only a low percentage of patients respond and responses tend to be transient. This is despite the fact that vaccination with hTERT-derived peptides generally does result in hTERT-specific T cell production, indicative of an immunological response. Current efforts involve understanding the differences between responsive and nonresponsive patients, optimization of antigenic hTERT epitopes, and optimization of specific immunotherapy platforms for telomerase-targeted treatment strategies. ■ GENE DELIVERY One of the main concerns with cancer gene therapy is the selective expression of target genes in tumor cells while avoiding expression in normal tissues. Because activation of telomerase activity in cancer cells is largely driven by increased expression of hTERT, it is reasonable to speculate that hTERT promoter activity is commensurately increased. Taking advantage of this increased promoter activity, several groups have investigated the potential of the hTERT promoter for cancer-selective gene therapy. The hTERT promoter is extremely attractive for the selective expression of transgenes for anticancer gene therapy because of the near universal expression of hTERT in cancers and the much greater activity of the promoter in cancer cells compared to normal cells, even those that express telomerase. Most reports on the use of the hTERT promoter for cancer gene therapy attempt to control expression of anticancer genes in tumor cells. For example, proapoptotic genes (FADD and TRAIL) and a suicide gene (HSVtk) have been tested.215−217 In cell-based assays and in mouse models, tumor-specific transgene expression has been achieved and promising results in mouse studies highlight early successes. Because these approaches utilize replication-deficient virus in order to decrease harmful side effects, these approaches may suffer from limited tumor-mass distribution. A different and very promising approach that overcomes this limitation is to utilize the hTERT promoter not for transgene expression but instead viral replication, in essence establishing an anticancer virus. One hTERT-promoter driven adenovirus, telomelysin, has entered clinical trials in a small, 16 patient, phase I trial in which the patients had a wide variety of cancer diagnoses.35 Early data suggest that telomelysin was well tolerated, but clinical response was limited, owing perhaps to the single viral infection, which may have limited the viral titer. Most promising was the response of one patient with a significant, 56%, reduction in metastatic tumor size.35 ■ PERSPECTIVE AND FUTURE DIRECTIONS Much progress has been made toward translating telomere research to the clinic since the seminal studies, demonstrating the importance of telomeres, the identification of telomerase and telomere-binding proteins, and the validation of telomere length maintenance as a requirement of cancer cells. Early clinical trials with telomerase inhibitors have shown that the approach is well tolerated and does reduce telomerase activity. However, the absence of reported clinical efficacy suggests limited progress. Future trials examining which patients might be best suited for telomerase inhibitor therapy and current trials testing telomerase inhibitors in combination with other traditional chemotherapeutics can be expected to advance the field. We can also expect continued advances in gene therapy by taking advantage of the robust activity of the hTERT promoter in cancer cells as viral vectors are optimized. Likewise, immunotherapy targeting hTERT antigens is expected to continue its progress through clinical trials and arguably holds Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXK
  • 12. the most promise currently for telomerase-targeting therapeu- tics. Surprisingly, small-molecule inhibitors of telomerase have not fared well. This may change as advances in telomerase structural biology; for example, the crystal structure of an insect TERT,62 structures of several hTER domains,218,219 models of hTERT based on the beetle crystal structure,220 and the recent three-dimensional model of human telomerase based on electron microscopy images221 increase our understanding of telomerase structure and provide templates for rational small- molecule drug design. Rationally designed telomerase inhibitors may allow increased specificity, bioavailability, and inhibition mechanisms that have greater effects in a clinical setting than those small molecules that are currently available. In addition, the observation that G-quadruplex ligands can directly disrupt telomere structure evidently by blocking binding of telomere- binding proteins such as TRF2 and POT1 suggests that there are more telomere-associated targets to be exploited, and we can expect that near future studies will determine the effectiveness of directly targeting the telomere as an anticancer drug platform. Targeting telomere biology as a means to improve human health remains an important objective that drives discoveries in telomere biology and telomerase biochemistry. The interdisciplinary approaches toward targeting telomere biology have led to three discrete areas of clinical trials: inhibition of telomerase, hTERT-directed immunother- apy, and hTERT-promoter gene/viral therapy. It seems that the future of pharmaceutical benefit from telomere-targeting agents remains promising, and our increased understanding of telomere biochemistry and biology is sure to inform more effective approaches toward developing telomere biology targeted agents. ■ AUTHOR INFORMATION Corresponding Author *Phone: 919-966-6422. E-mail: jarstfer@unc.edu. Notes The authors declare no competing financial interest. Biographies Vijay Sekaran is a postdoctoral fellow at the University of Colorado at Boulder working with Drs. James Goodrich and Jennifer Kugel. He received his B.S. in Biomedical Engineering from the Georgia Institute of Technology in 2005 and his Ph.D. in Pharmaceutical Sciences from the University of North Carolina at Chapel Hill under the guidance of Michael Jarstfer in 2012. His dissertation focused on understanding the structure and function of telomerase RNAs as well as characterizing the effects of small molecules that bind to human telomerase RNA and inhibit telomerase activity. Joana Soares is a research scientist at Genezyme. She received her B.S. in Chemistry from The College of William and Mary, VA, in 2004 and her Ph.D. in Pharmaceutical Sciences from the University of North Carolina at Chapel Hill in 2010 under the supervision of Michael Jarstfer. Her dissertation focused on elucidating the mechanism of telomerase activation or repression and its impact on cellular pathways and elucidating the mechanism of action of new potential telomerase inhibitors. Her research interrogated the relationship between telomerase, aging, and cancer progression. Michael B. Jarstfer is an Associate Professor of Chemical Biology and Medicinal Chemistry in the Eshelman School of Pharmacy at the University of North Carolina at Chapel Hill. He received his B.A. in Biochemistry from Trinity University, San Antonio TX, and his Ph.D. from the University of Utah under the guidance of Dale Poulter. He conducted postdoctoral research in the laboratory of Thomas Cech at the University of Colorado in Boulder. His research interests include structure and function of telomerase, development of telomere targeting agents, the roles of telomere biology in human health, and the molecular basis for social motivation. ■ ABBREVIATIONS USED ALT, alternative lengthening of telomeres; AS, antisense; hTERT, human telomerase reverse transcriptase; HDAC, histone deacetylase; hTER, human telomerase RNA; NPS, N3′→P5′-thiophosphoramidates; ODN, oligonucleotide; PKC, protein kinase C; PNA, peptide nucleic acid; POT1, Protection of Telomeres 1; PS, phosphorothioate; QSAR, quantitative structure−activity relationship; RNAi, RNA interference; SAR, structure−activity relationship; siRNA, silencing RNA; TIF, telomere dysfunction induced focus; TAK1, TGF-β activated kinase 1; TAL1, T-cell acute lymphoblastic leukemia protein 1; WT1, Wilm’s tumor protein 1 ■ REFERENCES (1) Shay, J. W.; Bacchetti, S. A survey of telomerase activity in human cancer. Eur. J. Cancer 1997, 33, 787−791. (2) Beatty, G. L.; Vonderheide, R. H. Telomerase as a universal tumor antigen for cancer vaccines. Expert Rev. Vaccines 2008, 7, 881− 887. (3) Fujiwara, T.; Urata, Y.; Tanaka, N. Telomerase-specific oncolytic virotherapy for human cancer with the hTERT promoter. Curr. Cancer Drug Targets 2007, 7, 191−201. (4) Jakupciak, J. P. Real-time telomerase activity measurements for detection of cancer. Expert Rev. Mol. Diagn. 2005, 5, 745−753. (5) Deng, Y.; Chan, S. S.; Chang, S. Telomere dysfunction and tumour suppression: the senescence connection. Nat. Rev. Cancer 2008, 8, 450−458. (6) Palm, W.; de Lange, T. How shelterin protects mammalian telomeres. Annu. Rev. Genet. 2008, 42, 301−334. (7) Sperka, T.; Wang, J.; Rudolph, K. L. DNA damage checkpoints in stem cells, ageing and cancer. Nat. Rev. Mol. Cell Biol. 2012, 13, 579− 590. (8) Moon, I. K.; Jarstfer, M. B. The human telomere and its relationship to human disease, therapy, and tissue engineering. Front. Biosci. 2007, 12, 4595−4620. (9) Blasco, M. A. Telomeres and human disease: ageing, cancer and beyond. Nat. Rev. Genet. 2005, 6, 611−622. (10) Collins, K. The biogenesis and regulation of telomerase holoenzymes. Nat. Rev. Mol. Cell Biol. 2006, 7, 484−494. (11) Sekaran, V. G.; Soares, J.; Jarstfer, M. B. Structures of telomerase subunits provide functional insights. Biochim. Biophys. Acta 2010, 1804, 1190−1201. (12) Feng, J.; Funk, W. D.; Wang, S. S.; Weinrich, S. L.; Avilion, A. A.; Chiu, C. P.; Adams, R. R.; Chang, E.; Allsopp, R. C.; Yu, J.; Le, S.; West, M. D.; Harley, C.; Andrews, W. H.; Greider, C. W.; Villeponteu, B. The RNA component of human telomerase. Science 1995, 269, 1236−1241. (13) Nakamura, T. M.; Morin, G. B.; Chapman, K. B.; Weinrich, S. L.; Andrews, W. H.; Lingner, J.; Harley, C. B.; Cech, T. R. Telomerase catalytic subunit homologs from fission yeast and human. Science 1997, 277, 955−959. (14) Huang, F. W.; Hodis, E.; Xu, M. J.; Kryukov, G. V.; Chin, L.; Garraway, L. A. Highly recurrent TERT promoter mutations in human melanoma. Science 2013, 339, 957−959. (15) Karlseder, J.; Smogorzewska, A.; de Lange, T. Senescence induced by altered telomere state, not telomere loss. Science 2002, 295, 2446−2449. Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXL
  • 13. (16) Griffith, J. D.; Comeau, L.; Rosenfield, S.; Stansel, R. M.; Bianchi, A.; Moss, H.; de Lange, T. Mammalian telomeres end in a large duplex loop. Cell 1999, 97, 503−514. (17) Cao, Y.; Bryan, T. M.; Reddel, R. R. Increased copy number of the TERT and TERC telomerase subunit genes in cancer cells. Cancer Sci. 2008, 99, 1092−1099. (18) Cesare, A. J.; Reddel, R. R. Telomere uncapping and alternative lengthening of telomeres. Mech. Ageing Dev. 2008, 129, 99−108. (19) Muntoni, A.; Reddel, R. R. The first molecular details of ALT in human tumor cells. Hum. Mol. Genet. 2005, 14, R191−R196. (20) Hahn, W. C.; Counter, C. M.; Lundbeg, A. S.; Beijersbergen, R. L.; Brooks, M. W.; Weinberg, R. A. Creation of human tumour cells with defined genetic elements. Nature 1999, 400, 464−468. (21) Aubert, G.; Lansdorp, P. M. Telomeres and aging. Physiol. Rev. 2008, 88, 557−579. (22) Bollmann, F. M. The many faces of telomerase: emerging extratelomeric effects. BioEssays 2008, 30, 728−732. (23) Artandi, S. E.; DePinho, R. A. Telomeres and telomerase in cancer. Carcinogenesis 2010, 31, 9−18. (24) Xu, L.; Li, S.; Stohr, B. A. The role of telomere biology in cancer. Annu. Rev. Pathol. 2013, 8, 49−78. (25) Ju, Z.; Rudolph, K. L. Telomeres and telomerase in cancer stem cells. Eur. J. Cancer 2006, 42, 1197−1203. (26) Armanios, M.; Greider, C. W. Telomerase and cancer stem cells. Cold Spring Harbor Symp. Quant. Biol. 2005, 70, 205−208. (27) 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. (28) Svenson, U.; Roos, G. Telomere length as a biological marker in malignancy. Biochim. Biophys. Acta 2009, 1792, 317−323. (29) Hahn, W. C.; Stewart, S. A.; Brooks, M. W.; York, S. G.; Eaton, E.; Kurachi, A.; Beijersbergen, R. L.; Knoll, J. H. M.; Maeyerson, M.; Weinberg, R. A. Inhibition of telomerase limits the growth of human cancer cells. Nat. Med. 1999, 5, 1164−1170. (30) Zhang, X.; Mar, V.; Zhou, W.; Harrington, L.; Robinson, M. O. Telomere shortening and apoptosis in telomerase-inhibited human tumor cells. Genes Dev. 1999, 13, 2388−2399. (31) Roth, A.; Harley, C. B.; Baerlocher, G. M. Imetelstat (GRN163L)telomerase-based cancer therapy. Recent Results Cancer Res. 2010, 184, 221−234. (32) Wong, H. M.; Payet, L.; Huppert, J. L. Function and targeting of G-quadruplexes. Curr. Opin. Mol. Ther. 2009, 11, 146−155. (33) Monchaud, D.; Teulade-Fichou, M. P. A hitchhiker’s guide to G- quadruplex ligands. Org. Biomol. Chem. 2008, 6, 627−636. (34) 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. (35) Nemunaitis, J.; Tong, A. W.; Nemunaitis, M.; Senzer, N.; Phadke, A. P.; Bedell, C.; Adams, N.; Zhang, Y. A.; Maples, P. B.; Chen, S.; Pappen, B.; Burke, J.; Ichimaru, D.; Urata, Y.; Fujiwara, T. A phase I study of telomerase-specific replication competent oncolytic adenovirus (telomelysin) for various solid tumors. Mol. Ther. 2010, 18, 429−434. (36) Liu, J. P.; Chen, W.; Schwarer, A. P.; Li, H. Telomerase in cancer immunotherapy. Biochim. Biophys. Acta, Rev. Cancer 2010, 1805, 35−42. (37) Damm, K.; Hemmann, U.; Garin-Chesa, P.; Hauel, N.; Kauffmann, I.; Priepke, H.; Niestroj, C.; Daiber, C.; Enenkel, B.; Guilliard, B.; Lauritsch, I.; Muller, E.; Pascolo, E.; Sauter, G.; Pantic, M.; Martens, U. M.; Wenz, C.; Lingner, J.; Kraut, N.; Rettig, W. J.; Schnapp, A. A highly selective telomerase inhibitor limiting human cancer cell proliferation. EMBO J. 2001, 20, 6958−6968. (38) Huffman, K. E.; Levene, S. D.; Tesmer, V. M.; Shay, J. W.; Wright, W. E. Telomere shortening is proportional to the size of the G-rich telomeric 3′-overhang. J. Biol. Chem. 2000, 275, 19719−19722. (39) Zou, Y.; Sfeir, A.; Gryaznov, S. M.; Shay, J. W.; Wright, W. E. Does a sentinel or a subset of short telomeres determine replicative senescence? Mol. Biol. Cell 2004, 15, 3709−3718. (40) de Lange, T. Shelterin: the protein complex that shapes and safeguards human telomeres. Genes Dev. 2005, 19, 2100−2110. (41) Hu, J.; Hwang, S. S.; Liesa, M.; Gan, B.; Sahin, E.; Jaskelioff, M.; Ding, Z.; Ying, H.; Boutin, A. T.; Zhang, H.; Johnson, S.; Ivanova, E.; Kost-Alimova, M.; Protopopov, A.; Wang, Y. A.; Shirihai, O. S.; Chin, L.; DePinho, R. A. Antitelomerase therapy provokes ALT and mitochondrial adaptive mechanisms in cancer. Cell 2012, 148, 651− 663. (42) Folini, M.; Zaffaroni, N. Targeting telomerase by antisense- based approaches: perspectives for new anti-cancer therapies. Curr. Pharm. Des. 2005, 11, 1105−1117. (43) Cong, Y.; Shay, J. W. Actions of human telomerase beyond telomeres. Cell Res. 2008, 18, 725−732. (44) Park, J. I.; Venteicher, A. S.; Hong, J. Y.; Choi, J.; Jun, S.; Shkreli, M.; Chang, W.; Meng, Z.; Cheung, P.; Ji, H.; McLaughlin, M.; Veenstra, T. D.; Nusse, R.; McCrea, P. D.; Artandi, S. E. Telomerase modulates Wnt signalling by association with target gene chromatin. Nature 2009, 460, 66−72. (45) Ghosh, A.; Saginc, G.; Leow, S. C.; Khattar, E.; Shin, E. M.; Yan, T. D.; Wong, M.; Zhang, Z.; Li, G.; Sung, W. K.; Zhou, J.; Chng, W. J.; Li, S.; Liu, E.; Tergaonkar, V. Telomerase directly regulates NF- kappaB-dependent transcription. Nat. Cell Biol. 2012, 14, 1270−1281. (46) De Cian, A.; Lacroix, L.; Douarre, C.; Temime-Smaali, N.; Trentesaux, C.; Riou, J. F.; Mergny, J. L. Targeting telomeres and telomerase. Biochimie 2008, 90, 131−155. (47) Salvati, E.; Leonetti, C.; Rizzo, A.; Scarsella, M.; Mottolese, M.; Galati, R.; Sperduti, I.; Stevens, M. F.; D’Incalci, M.; Blasco, M.; Chiorino, G.; Bauwens, S.; Horard, B.; Gilson, E.; Stoppacciaro, A.; Zupi, G.; Biroccio, A. Telomere damage induced by the G-quadruplex ligand RHPS4 has an antitumor effect. J. Clin. Invest. 2007, 117, 3236− 3247. (48) Miyazaki, T.; Pan, Y.; Joshi, K.; Purohit, D.; Hu, B.; Demir, H.; Mazumder, S.; Okabe, S.; Yamori, T.; Viapiano, M.; Shin-ya, K.; Seimiya, H.; Nakano, I. Telomestatin impairs glioma stem cell survival and growth through the disruption of telomeric G-quadruplex and inhibition of the proto-oncogene, c-Myb. Clin. Cancer Res. 2012, 18, 1268−1280. (49) Gong, Y.; de Lange, T. A Shld1-controlled POT1a provides support for repression of ATR signaling at telomeres through RPA exclusion. Mol. Cell 2010, 40, 377−387. (50) Melana, S. M.; Holland, J. F.; Pogo, B. G. Inhibition of cell growth and telomerase activity of breast cancer cells in vitro by 3′- azido-3′-deoxythymidine. Clin. Cancer Res. 1998, 4, 693−696. (51) Strahl, C.; Blackburn, E. H. Effects of reverse transcriptase inhibitors on telomere length and telomerase activity in two immortalized human cell lines. Mol. Cell. Biol. 1996, 16, 53−65. (52) Datta, A.; Bellon, M.; Sinha-Datta, U.; Bazarbachi, A.; Lepelletier, Y.; Canioni, D.; Waldmann, T. A.; Hermine, O.; Nicot, C. Persistent inhibition of telomerase reprograms adult T-cell leukemia to p53-dependent senescence. Blood 2006, 108, 1021−1029. (53) Johnston, J. S.; Johnson, A.; Gan, Y.; Wientjes, M. G.; Au, J. L. Synergy between 3′-azido-3′-deoxythymidine and paclitaxel in human pharynx FaDu cells. Pharm. Res. 2003, 20, 957−961. (54) Yasuda, C.; Kato, M.; Kuroda, D.; Ohyanagi, H. Experimental studies on potentiation of the antitumor activity of 5-fluorouracil with 3′-azido-3′-deoxythymidine for the gastric cancer cell line MKN28 in vivo. Jpn. J. Cancer Res 1997, 88, 97−102. (55) Chen, C.; Zhang, Y.; Wang, Y.; Huang, D.; Xi, Y.; Qi, Y. Synergic effect of 3′-azido-3′-deoxythymidine and arsenic trioxide in suppressing hepatoma cells. Anti-Cancer Drugs 2011, 22, 435−443. (56) Hukezalie, K. R.; Thumati, N. R.; Cote, H. C.; Wong, J. M. In vitro and ex vivo inhibition of human telomerase by anti-HIV nucleoside reverse transcriptase inhibitors (NRTIs) but not by non- NRTIs. PLoS One 2012, 7, e47505. (57) Leeansyah, E.; Cameron, P. U.; Solomon, A.; Tennakoon, S.; Velayudham, P.; Gouillou, M.; Spelman, T.; Hearps, A.; Fairley, C.; Smit de, V.; Pierce, A. B.; Armishaw, J.; Crowe, S. M.; Cooper, D. A.; Koelsch, K. K.; Liu, J. P.; Chuah, J.; Lewin, S. R. Inhibition of telomerase activity by human immunodeficiency virus (HIV) nucleos- Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXM
  • 14. (t)ide reverse transcriptase inhibitors: a potential factor contributing to HIV-associated accelerated aging. J. Infect. Dis. 2013, 207, 1157−1165. (58) Fletcher, T. M.; Cathers, B. E.; Ravikumar, K. S.; Mamiya, B. M.; Kerwin, S. M. Inhibition of human telomerase by 7-deaza-2′- deoxyguanosine nucleoside triphosphate analogs: potent inhibition by 6-thio-7-deaza-2′-deoxyguanosine 5′-triphosphate. Bioorg. Chem. 2001, 29, 36−55. (59) Hayakawa, N.; Nozawa, K.; Ogawa, A.; Kato, N.; Yoshida, K.; Akamatsu, K.; Tsuchiya, M.; Nagasaka, A.; Yoshida, S. Isothiazolone derivatives selectively inhibit telomerase from human and rat cancer cells in vitro. Biochemistry 1999, 38, 11501−11507. (60) Kim, J. H.; Lee, G. E.; Lee, J. E.; Chung, I. K. Potent inhibition of human telomerase by nitrostyrene derivatives. Mol. Pharmacol. 2003, 63, 1117−1124. (61) Kim, J. H.; Lee, G. E.; Kim, S. W.; Chung, I. K. Identification of a quinoxaline derivative that is a potent telomerase inhibitor leading to cellular senescence of human cancer cells. Biochem. J. 2003, 373, 523− 529. (62) Gillis, A. J.; Schuller, A. P.; Skordalakes, E. Structure of the Tribolium castaneum telomerase catalytic subunit TERT. Nature 2008, 455, 633−637. (63) Liu, X. H.; Li, J.; Shi, J. B.; Song, B. A.; Qi, X. B. Design and synthesis of novel 5-phenyl-N-piperidine ethanone containing 4,5- dihydropyrazole derivatives as potential antitumor agents. Eur. J. Med. Chem. 2012, 51, 294−299. (64) Liu, X. H.; Ruan, B. F.; Liu, J. X.; Song, B. A.; Jing, L. H.; Li, J.; Yang, Y.; Zhu, H. L.; Qi, X. B. Design and synthesis of N-phenylacetyl (sulfonyl) 4,5-dihydropyrazole derivatives as potential antitumor agents. Bioorg. Med. Chem. Lett. 2011, 21, 2916−2920. (65) Luo, Y.; Zhang, S.; Qiu, K. M.; Liu, Z. J.; Yang, Y. S.; Fu, J.; Zhong, W. Q.; Zhu, H. L. Synthesis, biological evaluation, 3D-QSAR studies of novel aryl-2H-pyrazole derivatives as telomerase inhibitors. Bioorg. Med. Chem. Lett. 2013, 23, 1091−1095. (66) Liu, X. H.; Liu, H. F.; Chen, J.; Yang, Y.; Song, B. A.; Bai, L. S.; Liu, J. X.; Zhu, H. L.; Qi, X. B. Synthesis and molecular docking study of novel coumarin derivatives containing 4,5-dihydropyrazole moiety as potential antitumor agents. Bioorg. Med. Chem. Lett. 2010, 20, 5705−5708. (67) Zheng, Q. Z.; Zhang, X. M.; Xu, Y.; Cheng, K.; Jiao, Q. C.; Zhu, H. L. Synthesis, biological evaluation, and molecular docking studies of 2-chloropyridine derivatives possessing 1,3,4-oxadiazole moiety as potential antitumor agents. Bioorg. Med. Chem. 2010, 18, 7836−7841. (68) Sun, J.; Zhu, H.; Yang, Z. M.; Zhu, H. L. Synthesis, molecular modeling and biological evaluation of 2-aminomethyl-5-(quinolin-2- yl)-1,3,4-oxadiazole-2(3H)-thione quinolone derivatives as novel anticancer agent. Eur. J. Med. Chem. 2013, 60, 23−28. (69) Huang, P. R.; Yeh, Y. M.; Wang, T. C. Potent inhibition of human telomerase by helenalin. Cancer Lett. 2005, 227, 169−174. (70) Soares, J.; Keppler, B. R.; Wang, X.; Lee, K. H.; Jarstfer, M. B. Ortho-quinone tanshinones directly inhibit telomerase through an oxidative mechanism mediated by hydrogen peroxide. Bioorg. Med. Chem. Lett. 2011, 21, 7474−7478. (71) Chen, Y. J.; Sheng, W. Y.; Huang, P. R.; Wang, T. C. Potent inhibition of human telomerase by U-73122. J. Biomed. Sci. 2006, 13, 667−674. (72) Ueno, T.; Takahashi, H.; Oda, M.; Mizunuma, M.; Yokoyama, A.; Goto, Y.; Mizushina, Y.; Sakaguchi, K.; Hayashi, H. Inhibition of human telomerase by rubromycins: implication of spiroketal system of the compounds as an active moiety. Biochemistry 2000, 39, 5995− 6002. (73) Mizushina, Y.; Ueno, T.; Oda, M.; Yamaguchi, T.; Saneyoshi, M.; Sakaguchi, K. The biochemical mode of inhibition of DNA polymerase beta by alpha-rubromycin. Biochim. Biophys. Acta 2000, 1523, 172−181. (74) Nakai, R.; Ishida, H.; Asai, A.; Ogawa, H.; Yamamoto, Y.; Kawasaki, H.; Akinaga, S.; Mizukami, T.; Yamashita, Y. Telomerase inhibitors identified by a forward chemical genetics approach using a yeast strain with shortened telomere length. Chem. Biol. 2006, 13, 183−190. (75) Uchida, K.; Ogawa, T.; Yasuda, Y.; Mimura, H.; Fujimoto, T.; Fukuyama, T.; Wakimoto, T.; Asakawa, T.; Hamashima, Y.; Kan, T. Stereocontrolled total synthesis of (+)-UCS1025A. Angew. Chem. 2012, 51, 12850−12853. (76) Lambert, T. H.; Danishefsky, S. J. Total synthesis of UCS1025A. J. Am. Chem. Soc. 2006, 128, 426−427. (77) Wong, L. H.; Unciti-Broceta, A.; Spitzer, M.; White, R.; Tyers, M.; Harrington, L. A yeast chemical genetic screen identifies inhibitors of human telomerase. Chem. Biol. 2013, 20, 333−340. (78) Barma, D. K.; Elayadi, A.; Falck, J. R.; Corey, D. R. Inhibition of telomerase by BIBR 1532 and related analogues. Bioorg. Med. Chem. Lett. 2003, 13, 1333−1336. (79) Pascolo, E.; Wenz, C.; Lingner, J.; Hauel, N.; Priepke, H.; Kauffmann, I.; Garin-Chesa, P.; Rettig, W. J.; Damm, K.; Schnapp, A. Mechanism of human telomerase inhibition by BIBR1532, a synthetic, non-nucleosidic drug candidate. J. Biol. Chem. 2002, 277, 15566− 15572. (80) El-Daly, H.; Kull, M.; Zimmermann, S.; Pantic, M.; Waller, C. F.; Martens, U. M. Selective cytotoxicity and telomere damage in leukemia cells using the telomerase inhibitor BIBR1532. Blood 2005, 105, 1742−1749. (81) Ward, R. J.; Autexier, C. Pharmacological telomerase inhibition can sensitize drug-resistant and drug-sensitive cells to chemo- therapeutic treatment. Mol. Pharmacol. 2005, 68, 779−786. (82) Pantic, M.; Zimmermann, S.; Waller, C. F.; Martens, U. M. The level of telomere dysfunction determines the efficacy of telomerase- based therapeutics in a lung cancer cell line. Int. J. Cancer 2005, 26, 1227−1232. (83) Hamilton, S. E.; Simmons, C. G.; Kathiriya, I. S.; Corey, D. R. Cellular delivery of peptide nucleic acids and inhibition of human telomerase. Chem. Biol. 1999, 6, 343−351. (84) Keppler, B. R.; Jarstfer, M. B. Inhibition of telomerase activity by preventing proper assemblage. Biochemistry 2004, 43, 334−343. (85) Pinto, I. G.; Guilbert, C.; Ulyanov, N. B.; Stearns, J.; James, T. L. Discovery of ligands for a novel target, the human telomerase RNA, based on flexible-target virtual screening and NMR. J. Med. Chem. 2008, 51, 7205−7215. (86) Dominick, P. K.; Keppler, B. R.; Legassie, J. D.; Moon, I. K.; Jarstfer, M. B. Nucleic acid-binding ligands identify new mechanisms to inhibit telomerase. Bioorg. Med. Chem. Lett. 2004, 14, 3467−3471. (87) Kaiser, M.; Sainlos, M.; Lehn, J. M.; Bombard, S.; Teulade- Fichou, M. P. Aminoglycoside-quinacridine conjugates: towards recognition of the P6.1 element of telomerase RNA. ChemBioChem 2006, 7, 321−329. (88) Jain, N.; Francis, S.; Friedman, S. H. Inhibition of therapeutically important polymerases with high affinity bis-intercalators. Bioorg. Med. Chem. Lett. 2012, 22, 4844−4848. (89) Rangarajan, S.; Friedman, S. H. Design, synthesis, and evaluation of phenanthridine derivatives targeting the telomerase RNA/DNA heteroduplex. Bioorg. Med. Chem. Lett. 2007, 17, 2267− 2273. (90) Meyerson, M.; Counter, C. M.; Eaton, E. N.; Ellisen, L. W.; Steiner, P.; Caddle, S. D.; Ziaugra, L.; Beijersbergen, R. L.; Davidoff, M. J.; Liu, Q.; Bacchetti, S.; Haber, D. A.; Weinberg, R. A. hEST2, the putative human telomerase catalytic subunit gene, is up-regulated in tumor cells and during immortalization. Cell 1997, 90, 785−795. (91) 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. (92) Grandori, C.; Cowley, S. M.; James, L. P.; Eisenman, R. N. The Myc/Max/Mad network and the transcriptional control of cell behavior. Annu. Rev. Cell Dev. Biol. 2000, 16, 653−699. (93) Moon, D. O.; Kim, M. O.; Lee, J. D.; Choi, Y. H.; Kim, G. Y. Butein suppresses c-Myc-dependent transcription and Akt-dependent phosphorylation of hTERT in human leukemia cells. Cancer Lett. 2009, 286, 172−179. (94) Guo, Q. L.; Lin, S. S.; You, Q. D.; Gu, H. Y.; Yu, J.; Zhao, L.; Qi, Q.; Liang, F.; Tan, Z.; Wang, X. Inhibition of human telomerase Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXN
  • 15. reverse transcriptase gene expression by gambogic acid in human hepatoma SMMC-7721 cells. Life Sci. 2006, 78, 1238−1245. (95) Jagadeesh, S.; Kyo, S.; Banerjee, P. P. Genistein represses telomerase activity via both transcriptional and posttranslational mechanisms in human prostate cancer cells. Cancer Res. 2006, 66, 2107−2115. (96) Shapira, S.; Granot, G.; Mor-Tzuntz, R.; Raanani, P.; Uziel, O.; Lahav, M.; Shpilberg, O. Second-generation tyrosine kinase inhibitors reduce telomerase activity in K562 cells. Cancer Lett. 2012, 323, 223− 231. (97) Wooten, L. G.; Ogretmen, B. Sp1/Sp3-dependent regulation of human telomerase reverse transcriptase promoter activity by the bioactive sphingolipid ceramide. J. Biol. Chem. 2005, 280, 28867− 28876. (98) Kyo, S.; Takakura, M.; Taira, T.; Kanaya, T.; Itoh, H.; Yutsudo, M.; Ariga, H.; Inoue, M. Sp1 cooperates with c-Myc to activate transcription of the human telomerase reverse transcriptase gene (hTERT). Nucleic Acids Res. 2000, 28, 669−677. (99) Misiti, S.; Nanni, S.; Fontemaggi, G.; Cong, Y. S.; Wen, J.; Hirte, H. W.; Piaggio, G.; Sacchi, A.; Pontecorvi, A.; Bacchetti, S.; Farsetti, A. Induction of hTERT expression and telomerase activity by estrogens in human ovary epithelium cells. Mol. Cell. Biol. 2000, 20, 3764−3771. (100) Kyo, S.; Takakura, M.; Kanaya, T.; Zhuo, W.; Fujimoto, K.; Nishio, Y.; Orimo, A.; Inoue, M. Estrogen activates telomerase. Cancer Res. 1999, 59, 5917−5921. (101) Nanni, S.; Narducci, M.; Della Pietra, L.; Moretti, F.; Grasselli, A.; De Carli, P.; Sacchi, A.; Pontecorvi, A.; Farsetti, A. Signaling through estrogen receptors modulates telomerase activity in human prostate cancer. J. Clin. Invest. 2002, 110, 219−227. (102) Aldous, W. K.; Marean, A. J.; DeHart, M. J.; Matej, L. A.; Moore, K. H. Effects of tamoxifen on telomerase activity in breast carcinoma cell lines. Cancer 1999, 85, 1523−1529. (103) Wang, Z.; Kyo, S.; Takakura, M.; Tanaka, M.; Yatabe, N.; Maida, Y.; Fujiwara, M.; Hayakawa, J.; Ohmichi, M.; Koike, K.; Inoue, M. Progesterone regulates human telomerase reverse transcriptase gene expression via activation of mitogen-activated protein kinase signaling pathway. Cancer Res. 2000, 60, 5376−5381. (104) Xu, D.; Popov, N.; Hou, M.; Wang, Q.; Bjorkholm, M.; Gruber, A.; Menkel, A. R.; Henriksson, M. Switch from Myc/Max to Mad1/Max binding and decrease in histone acetylation at the telomerase reverse transcriptase promoter during differentiation of HL60 cells. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 3826−3831. (105) Gunes, C.; Lichtsteiner, S.; Vasserot, A. P.; Englert, C. Expression of the hTERT gene is regulated at the level of transcriptional initiation and repressed by Mad1. Cancer Res. 2000, 60, 2116−2121. (106) Kanaya, T.; Kyo, S.; Hamada, K.; Takakura, M.; Kitagawa, Y.; Harada, H.; Inoue, M. Adenoviral expression of p53 represses telomerase activity through down-regulation of human telomerase reverse transcriptase transcription. Clin. Cancer Res. 2000, 6, 1239− 1247. (107) Crowe, D. L.; Nguyen, D. C. Rb and E2F-1 regulate telomerase activity in human cancer cells. Biochim. Biophys. Acta 2001, 1518, 1−6. (108) Henderson, Y. C.; Breau, R. L.; Liu, T. J.; Clayman, G. L. Telomerase activity in head and neck tumors after introduction of wild-type p53, p21, p16, and E2F-1 genes by means of recombinant adenovirus. Head Neck 2000, 22, 347−354. (109) Oh, S.; Song, Y.; Yim, J.; Kim, T. K. The Wilms’ tumor 1 tumor suppressor gene represses transcription of the human telomerase reverse transcriptase gene. J. Biol. Chem. 1999, 274, 37473−37478. (110) Lindkvist, A.; Ivarsson, K.; Jernberg-Wiklund, H.; Paulsson- Karlsson, Y. Interferon-induced sensitization to apoptosis is associated with repressed transcriptional activity of the hTERT promoter in multiple myeloma. Biochem. Biophys. Res. Commun. 2006, 341, 1141− 1148. (111) Lacerte, A.; Korah, J.; Roy, M.; Yang, X. J.; Lemay, S.; Lebrun, J. J. Transforming growth factor-beta inhibits telomerase through SMAD3 and E2F transcription factors. Cell. Signalling 2008, 20, 50− 59. (112) Li, H.; Xu, D.; Li, J.; Berndt, M. C.; Liu, J. P. Transforming growth factor beta suppresses human telomerase reverse transcriptase (hTERT) by Smad3 interactions with c-Myc and the hTERT gene. J. Biol. Chem. 2006, 281, 25588−25600. (113) Fujiki, T.; Miura, T.; Maura, M.; Shiraishi, H.; Nishimura, S.; Imada, Y.; Uehara, N.; Tashiro, K.; Shirahata, S.; Katakura, Y. TAK1 represses transcription of the human telomerase reverse transcriptase gene. Oncogene 2007, 26, 5258−5266. (114) Terme, J. M.; Mocquet, V.; Kuhlmann, A. S.; Zane, L.; Mortreux, F.; Wattel, E.; Duc Dodon, M.; Jalinot, P. Inhibition of the hTERT promoter by the proto-oncogenic protein TAL1. Leukemia 2009, 23, 2081−2089. (115) Wu, P.; Meng, L.; Wang, H.; Zhou, J.; Xu, G.; Wang, S.; Xi, L.; Chen, G.; Wang, B.; Zhu, T.; Lu, Y.; Ma, D. Role of hTERT in apoptosis of cervical cancer induced by histone deacetylase inhibitor. Biochem. Biophys. Res. Commun. 2005, 335, 36−44. (116) Woo, H. J.; Lee, S. J.; Choi, B. T.; Park, Y. M.; Choi, Y. H. Induction of apoptosis and inhibition of telomerase activity by trichostatin A, a histone deacetylase inhibitor, in human leukemic U937 cells. Exp. Mol. Pathol. 2007, 82, 77−84. (117) Choi, Y. H. Apoptosis of U937 human leukemic cells by sodium butyrate is associated with inhibition of telomerase activity. Int. J. Oncol. 2006, 29, 1207−1213. (118) Won, J.; Chang, S.; Oh, S.; Kim, T. K. Small-molecule-based identification of dynamic assembly of E2F-pocket protein-histone deacetylase complex for telomerase regulation in human cells. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 11328−11333. (119) Berletch, J. B.; Liu, C.; Love, W. K.; Andrews, L. G.; Katiyar, S. K.; Tollefsbol, T. O. Epigenetic and genetic mechanisms contribute to telomerase inhibition by EGCG. J. Cell. Biochem. 2008, 103, 509−519. (120) Wang, X.; Hao, M. W.; Dong, K.; Lin, F.; Ren, J. H.; Zhang, H. Z. Apoptosis induction effects of EGCG in laryngeal squamous cell carcinoma cells through telomerase repression. Arch. Pharmacal Res. 2009, 32, 1263−1269. (121) Sadava, D.; Whitlock, E.; Kane, S. E. The green tea polyphenol, epigallocatechin-3-gallate inhibits telomerase and induces apoptosis in drug-resistant lung cancer cells. Biochem. Biophys. Res. Commun. 2007, 360, 233−237. (122) Seimiya, H.; Oh-hara, T.; Suzuki, T.; Naasani, I.; Shimazaki, T.; Tsuchiya, K.; Tsuruo, T. Telomere shortening and growth inhibition of human cancer cells by novel synthetic telomerase inhibitors MST- 312, MST-295, and MST-1991. Mol. Cancer Ther. 2002, 1, 657−665. (123) Bilsland, A. E.; Hoare, S.; Stevenson, K.; Plumb, J.; Gomez- Roman, N.; Cairney, C.; Burns, S.; Lafferty-Whyte, K.; Roffey, J.; Hammonds, T.; Keith, W. N. Dynamic telomerase gene suppression via network effects of GSK3 inhibition. PLoS One 2009, 4, e6459. (124) Pendino, F.; Flexor, M.; Delhommeau, F.; Buet, D.; Lanotte, M.; Segal-Bendirdjian, E. Retinoids down-regulate telomerase and telomere length in a pathway distinct from leukemia cell differ- entiation. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 6662−6667. (125) Pendino, F.; Dudognon, C.; Delhommeau, F.; Sahraoui, T.; Flexor, M.; Bennaceur-Griscelli, A.; Lanotte, M.; Segal-Bendirdjian, E. Retinoic acid receptor alpha and retinoid-X receptor-specific agonists synergistically target telomerase expression and induce tumor cell death. Oncogene 2003, 22, 9142−9150. (126) Li, H.; Zhao, L. L.; Funder, J. W.; Liu, J. P. Protein phosphatase 2A inhibits nuclear telomerase activity in human breast cancer cells. J. Biol. Chem. 1997, 272, 16729−16732. (127) 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. (128) Yu, C. C.; Lo, S. C.; Wang, T. C. Telomerase is regulated by protein kinase C-zeta in human nasopharyngeal cancer cells. Biochem. J. 2001, 355, 459−464. Journal of Medicinal Chemistry Perspective dx.doi.org/10.1021/jm400528t | J. Med. Chem. XXXX, XXX, XXX−XXXO