Sie sind auf Seite 1von 24

Hindawi Publishing Corporation

BioMed Research International


Volume 2014, Article ID 150845, 23 pages
http://dx.doi.org/10.1155/2014/150845

Review Article
Apoptosis and Molecular Targeting Therapy in Cancer

Mohamed Hassan,1,2 Hidemichi Watari,2 Ali AbuAlmaaty,1


Yusuke Ohba,3 and Noriaki Sakuragi2
1
Biotechnology Program, Department of Zoology, Port Said University, Faculty of Science, Port Said 42521, Egypt
2
Department of Obstetrics and Gynecology, Graduate School of Medicine, Hokkaido University, Sapporo 060-8638, Japan
3
Department of Cell Physiology, Graduate School of Medicine, Hokkaido University, Sapporo 060-8638, Japan

Correspondence should be addressed to Mohamed Hassan; mohamedkamel24@yahoo.com

Received 4 February 2014; Accepted 11 May 2014; Published 12 June 2014

Academic Editor: Elena Orlova

Copyright © 2014 Mohamed Hassan et al. This is an open access article distributed under the Creative Commons Attribution
License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly
cited.

Apoptosis is the programmed cell death which maintains the healthy survival/death balance in metazoan cells. Defect in apoptosis
can cause cancer or autoimmunity, while enhanced apoptosis may cause degenerative diseases. The apoptotic signals contribute into
safeguarding the genomic integrity while defective apoptosis may promote carcinogenesis. The apoptotic signals are complicated
and they are regulated at several levels. The signals of carcinogenesis modulate the central control points of the apoptotic
pathways, including inhibitor of apoptosis (IAP) proteins and FLICE-inhibitory protein (c-FLIP). The tumor cells may use some
of several molecular mechanisms to suppress apoptosis and acquire resistance to apoptotic agents, for example, by the expression
of antiapoptotic proteins such as Bcl-2 or by the downregulation or mutation of proapoptotic proteins such as BAX. In this review,
we provide the main regulatory molecules that govern the main basic mechanisms, extrinsic and intrinsic, of apoptosis in normal
cells. We discuss how carcinogenesis could be developed via defective apoptotic pathways or their convergence. We listed some
molecules which could be targeted to stimulate apoptosis in different cancers. Together, we briefly discuss the development of
some promising cancer treatment strategies which target apoptotic inhibitors including Bcl-2 family proteins, IAPs, and c-FLIP for
apoptosis induction.

1. Introduction the apoptotic signals help to safeguard the genomic integrity


[3, 6, 7] while dysregulation of the apoptotic pathways may
Apoptosis is a very tightly programmed cell death with dis- not only promote tumorigenesis but also render the cancer
tinct biochemical and genetic pathways that play a critical role cell resistant to treatment. Thus, the evasion of apoptosis is
in the development and homeostasis in normal tissues [1]. a prominent hallmark of cancer [8]. Cancer cells are, in fact,
It contributes to elimination of unnecessary and unwanted harboring alterations that result in impaired apoptotic signal-
cells to maintain the healthy balance between cell survival ing, which facilitates tumor development and metastasis [6–
and cell death in metazoan [2, 3]. It is critical to animals 8].
especially long-lived mammals that must integrate multiple Here, we provide an overview of mechanisms by which
physiological as well as pathological death signals. Evidence the main regulatory molecules govern apoptosis in normal
indicates that insufficient apoptosis can manifest as cancer or cells and describe models of apoptotic dysregulation based
autoimmunity, while accelerated cell death is evident in acute on alterations in their function that facilitate the evasion
and chronic degenerative diseases, immunodeficiency, and of apoptosis in cancer cells. We will also briefly discuss the
infertility. Under many stressful conditions like precancerous development of some promising cancer treatment strategies
lesions, activation of the DNA damage checkpoint pathway based on targeting the apoptotic inhibitor to stimulate the
can serve to remove potentially harmful DNA-damaged cells apoptotic signals. We will shed some light on this very
via apoptosis induction to block carcinogenesis [4, 5]. Thus, active field of endeavor that witnessed many breakthroughs
2 BioMed Research International

over the last few years. Notably, insights from proposed 2.1. Induction of Apoptosis. Apoptosis is caused by pro-
target molecules have included different portions of the death teases, known as “caspases,” which specifically target cysteine
pathway depending on the type of cancer. Applications of aspartyl [17, 18]. Upon receiving specific signals instructing
many of such targeted therapy in different cell types and the cells to undergo apoptosis a number of distinctive changes
signals studied in them have emphasized selected control occur in the cell. A family of proteins known as caspases
points for each targeted therapy. is typically activated in the early stages of apoptosis. These
proteins cleave key cellular components that are required
2. Apoptosis Defect and Cancer for normal cellular function including structural proteins in
the cytoskeleton and nuclear proteins such as DNA repair
Defects in programmed cell death (apoptosis) mechanisms enzymes. The caspases can also activate other degradative
play important roles in tumor pathogenesis, allowing neo- enzymes such as DNases, which begin to cleave the DNA in
plastic cells to survive over intended lifespans, subverting the the nucleus.
need for exogenous survival factors and providing protection During the apoptotic process, apoptotic cells display
from oxidative stress and hypoxia as the tumor mass expands. distinctive morphology. Typically, the cell begins to shrink
That gives time for accumulation of genetic alterations that following the cleavage of lamins and actin filaments in the
deregulate cell proliferation, interfere with differentiation, cytoskeleton. The apoptotic breakdown of chromatin in the
promote angiogenesis, and increase invasiveness during nucleus often leads to nuclear condensation and/or a “horse-
tumor progression [9]. Apoptosis defects are now considered shoe” like appearance. Cells continue to shrink, packaging
an important complement of protooncogene activation, as themselves into a form that allows for their removal by
many deregulated oncoproteins that drive cell division also macrophages. These phagocytic cells are responsible for
trigger apoptosis (e.g., Myc, E1a, and Cyclin-D1) [10]. On clearing the apoptotic cells from tissues in a clean and tidy
the other hand, the noncancerous cells have a DNA repair fashion that avoids many of the problems associated with
machinery. Defects in DNA repair and/or chromosome seg- necrotic cell death. In order to promote their phagocytosis
regation normally trigger cell suicide as a defense mechanism by macrophages, apoptotic cells often undergo plasma mem-
for eradicating genetically unstable cells and thus such suicide brane changes that trigger the macrophage response. One of
mechanism’s defects permit survival of genetically unstable such changes is the translocation of phosphatidylserine from
cells, providing opportunities for selection of progressively the inside of the cell to the outer surface. The end stages
aggressive clones and may promote tumorigenesis [11]. of apoptosis are often characterised by the appearance of
There are varieties of molecular mechanisms that tumor membrane blebs or blisters process and small vesicles called
cells use to suppress apoptosis. Tumor cells can acquire apoptotic bodies.
resistance to apoptosis by the expression of antiapoptotic The real workers in apoptosis mechanism are caspases.
proteins such as Bcl-2 or by the downregulation or mutation Caspases family is composed of intracellular cysteine pro-
of proapoptotic proteins such as BAX. Since the expression teases (𝑛 = 11 in humans), which collaborate in proteolytic
of both Bcl-2 and BAX is regulated by the p53 tumor cascades. These caspases activate themselves and each other.
suppressor gene [12], some forms of human B-cell lymphoma Within these proteolytic cascades, caspases can be posi-
have Bcl-2 overexpression. That example represents the first tioned as either upstream “initiators” or downstream “effec-
and strongest lines of evidence that failure of cell death tors” of apoptosis. Several pathways for activating caspases
contributes to cancer [13]. exist.
Apoptosis defects may allow epithelial cells to survive First, there are thirty members of the tumor necrosis
in a suspended state, without attachment to extracellular factor- (TNF-) family receptors; eight contain a so-called
matrix which facilitate metastasis [14]. They also promote death domain (DD) in their cytosolic tail [19]. Several
resistance to the immune system, including many weapons of these DD-containing TNF-family receptors use caspase
of cytolytic T cells (CTLs) and natural killer (NK) cells used activation as a signaling mechanism, including TNFR1/
for attacking tumors that depend on integrity of the apoptosis CD120a, Fas/APO1/CD95, DR3/Apo2/Weasle, DR4/TrailR1,
machinery [15]. Cancer-associated defects in apoptosis play DR5/TrailR2, and DR6. Ligation of these receptors at the
a role in treatment resistance with conventional therapies cell surface results in the recruitment of several intracellu-
like chemotherapy and radiotherapy, increasing the threshold lar proteins, including certain procaspases, to the cytoso-
for cell death and thereby requiring higher doses for tumor lic domains of these receptors, forming a “death-inducing
killing agents [16]. Thus, dysregulated or defective apoptosis signaling complex” (DISC) that triggers caspase activation,
regulation is a fundamental aspect of the tumor biology. constituting the so-called “extrinsic” pathway for apoptosis
Successful eradication of cancer cells by nonsurgical means is [20]. The caspase-8 and, in some cases, caspase-10 are the
ultimately approached via induction of apoptosis. Therefore, specific caspases summoned to the DISC. These caspases
all the cancer drug designers try either to activate the contain so-called death effector domains (DEDs) in their N-
inactivated apoptotic mechanism or rectify a defective one. terminal prodomains that bind to a corresponding DED in
Hence, all cytotoxic anticancer therapies currently in clinical the adaptor protein, FADD, thus linking them to the TNF-
use, when they work, induce apoptosis of malignant cells. family death receptor complexes.
Hence, deeper understanding of the molecular mechanisms Second is the intrinsic pathway, in which mitochondria
of apoptosis and its defective status opens the gate for a new induces apoptosis by releasing cytochrome-c (cyt-c) into the
class of targeted therapeutics. cytosol. The released cytochrome c assembles a multiprotein
BioMed Research International 3

caspase-activating complex, referred to as the “apoptosome” and Bfl-1/A1), which display sequence homology in all BH1-
[21]. The central component of the apoptosome is Apaf1, a BH4 domains, promote cell survival, whereas proapoptotic
caspase-activating protein that oligomerizes upon binding proteins mediate receptor-, mitochondria-, or endoplasmic
cyt-c and then binds procaspase-9 via interaction with its reticulum (ER) stress-dependent apoptosis. The latter group
caspase recruitment domain (CARD). The intrinsic pathway is subdivided into multidomain or BH3-only proteins. The
is activated by myriad stimuli, including growth factor former consists of Bax and Bak, which are essential for apop-
deprivation, oxidants, Ca2+ overload, oncogene activation, tosis [28]. The BH3-only proteins are further divided into two
DNA-damaging agents, and microtubule targeting drugs. subclasses: “activators” (e.g., Bim and tBid), which directly
In addition to cyt-c, mitochondria also releases endonucle- activate Bax/Bak to induce mitochondrial outer membrane
ase G, AIF (a death modulating flavor protein), and IAP permeabilization (MOMP), and “sensitizers/derepressors”
antagonists SMAC (DIABLO) and OMI (HtrA2). Some of (e.g., Bad, Bik, Bmf, Hrk, Noxa, and Puma), which do not
these molecules may promote caspase-independent (non- activate Bax/Bak directly but instead neutralize antiapoptotic
apoptotic) cell death [22, 23]. proteins [29, 30]. The central role that Bax/Bak play in
A third pathway for apoptosis induction is specific to apoptosis is supported by evidence that BH3-only proteins
CTL and NK cells, which spray apoptosis-inducing protease, fail to trigger apoptosis in Bax/Bak-deficient cells [31, 32].
granzyme B (GraB), onto target cells. GraB then piggybacks Antiapoptotic proteins block death signaling by antagonizing
into cells via mannose-6-phosphate receptors (IGFR2) and the actions of Bax/Bak through partially known mechanisms.
enters effective cellular compartments via perforin channels As recently summarized, antiapoptotic proteins prevent
[24]. GraB is a serine protease but, similar to the caspases, it Bax/Bak activation by sequestering/inhibiting “activator”
cleaves substrates at Asp residues, including several caspases BH3-only proteins and/or directly inhibiting Bax/Bak activa-
and some caspase substrates. The fourth pathway is a caspase tion [33]. “Sensitizer” BH3-only proteins displace “activator”
activation pathway. This pathway is proposed to be linked to BH3-only proteins from antiapoptotic proteins, leading to
endoplasmic reticulum (ER)/Golgi stress [17]; however, many Bax/Bak activation. Alternatively, BH3-only proteins may
mechanistic details are lacking. Finally, a nuclear pathway for directly neutralize/inhibit antiapoptotic proteins, releasing
apoptosis regulation was proposed. That pathway depends on their inhibition of Bax/Bak.
discrete nuclear organelles, called Pml oncogenic domains Overexpression of antiapoptotic Bcl-2 or Bcl-xL probably
(PODs) or nuclear bodies (NBs). Ablation of the pml gene occurs in more than half of all cancers [34], rendering tumor
in mice results in general resistance to apoptosis through cells resistant to myriad apoptotic stimuli, including most
unknown mechanisms. Several proteins that can promote cytotoxic anticancer drugs. For example, forced expression
apoptosis have been localized to PODs, including Daax, Zip of Bcl-2 protein from plasmid vectors, in contrast, abrogates
kinase, and Par4, and defects in assembly of these nuclear sensitivity to the apoptosis promoting effects of antiestro-
structures are documented in cancers [25]. How PODs are gens in breast cancer lines, while antisense Bcl-2 prevents
linked to caspase activation pathways is unknown. Several estrogen-mediated apoptosis suppression, thus establishing
endogenous antagonists of the caspase-activation pathways a direct functional connection between ER and Bcl-2 and
have been discovered and examples of dysregulation of their suppression of apoptosis [35].
expression or function in cancers have been obtained. These
apoptosis mediators help cells to decide successful apoptosis
2.4. Carcinogenesis via Extrinsic Pathway Signaling Defect.
or unsuccessful one. They sometimes act as targets for drug
The extrinsic pathway is activated in vivo by TNF family
discovery, with the idea that abrogating their cytoprotective
functions may restore apoptosis sensitivity to tumor cells. ligands that engage DD-containing receptors, resulting in
activation of DED-containing caspases. The “death ligands”
2.2. Resistance to Apoptosis in Cancer. Cancer is an exam- are expressed on CTLs, NK cells, and other types of immune-
ple where the normal mechanisms of cell cycle regulation relevant cells (activated monocytes/macrophages and den-
are dysfunctional, with either an over-proliferation of cells dritic cells) and are used as weapons for eradication of trans-
and/or decreased removal of cells [26]. In fact, suppression formed cells [19]. Mice-based studies on genetic alterations
of apoptosis during carcinogenesis is thought to play a central in genes encoding death ligands or their receptors, as well as
role in the development and progression of some cancers [27]. use of neutralizing antibodies and Fc-fusion proteins, have
There is a variety of molecular mechanisms that tumor cells provided evidence of important roles in tumor suppression by
use to suppress apoptosis. cellular immune mechanisms. Fas ligand (FasL) is important
for CTL-mediated killing of some tumor targets, and TRAIL
2.3. Carcinogenesis via Intrinsic Signaling Defects. Mitochon- (Apo2 ligand) is critical for NK-mediated tumor suppression.
dria-dependent apoptosis is one of the most important Some tumor cells resist the response of the death receptor
pathways for apoptosis induction. That is why disturbing pathway to FasL produced by T cells to evade immune
this pathway is an effective way to inhibit apoptosis. Bcl-2 destruction. Many tumor cell lines display intrinsic resistance
family proteins (𝑛 = 24 in humans) are central regulators to TRAIL even though they express the necessary cell surface
of the intrinsic pathway, which either suppress or promote receptors confirming that, during evolution of tumors in vivo,
changes in mitochondrial membrane permeability required selection occurs for malignant clones capable of withstanding
for release of cyt-c and other apoptogenic proteins [21, 24]. immune attack. This could occur in a variety of ways
Antiapoptotic proteins (e.g., Bcl-2, Bcl-xL, Bcl-W, Mcl-1, including downregulation of the Fas receptor, expression of
4 BioMed Research International

nonfunctioning Fas receptor, and secretion of high levels of a are required for binding and suppressing caspases-3 and -7,
soluble form of the Fas receptor. That will directly sequester while the third BIR domain binds caspase-9. Thus, different
the Fas ligand or expression of Fas ligand on the surface of domains in the multi-BIR containing IAPs are responsible
tumor cells (reviewed, [36]). Some tumor cells are capable of a for suppression of different caspases. IAP family member
Fas ligand-mediated “counter attack” that results in apoptotic Livin (ML-IAP) contains a single BIR and inhibits caspase-
depletion of activated tumor infiltrating lymphocytes [37]. 9 but not caspase-3 and -7. Survivin also contains a single BIR
Thus, successful biological therapy depends on restoring which associates with caspase-9.
competency of the extrinsic pathway. IAPs are overexpressed in many cancers, while more
Another molecule is the c-FLIP, cellular FADD-like detailed information is needed about the exact deregulation
interleukin-1beta-converting enzyme inhibitory proteins of the 8 members of this gene family in specific types of
long c-FLIP(L), which regulates caspase-8 activation. Despite cancer. Those IAPs elevations found in cancers are impor-
the identified dual functionality of c-FLIPL as a pro- or tant for maintaining tumor cell survival and resistance to
antiapoptotic factor in normal tissues, c-FLIPL has generally chemotherapy as confirmed by antisense-mediated reduc-
been shown to act as a key negative regulator of apoptosis in tions experiments. Targeting IAP-family members XIAP,
human cancer cells [38, 39]. Increased expression of c-FLIP cIAP1, Survivin, or Apollon can induce apoptosis in tumor
has been detected in many human malignancies, including cell lines in culture or sensitize cells to cytotoxic anticancer
melanoma, hepatocellular carcinoma [40], nonsmall cell drugs.
lung carcinoma [40], and endometrial [41], colon [42], Endogenous antagonists of IAPs keep these apoptosis
and prostate cancer [43–45]. While its overexpression suppressors in check, promoting apoptosis. Two of these
has been associated with cancer progression and/or poor naturally occurring IAP-antagonists, SMAC (Diablo) and
prognosis in BL, HCC, and ovarian, endometrial, colon, HtrA2 (Omi), are sequestered inside mitochondria, becom-
and prostate cancer [38, 39, 42, 46], elevated expression ing released into the cytosol during apoptosis [51]. SMAC and
of c-FLIP blocks caspase-8 and renders cells resistant to HtrA2 have N-terminal leader sequences that are truncated
cell receptor-mediated apoptosis [38]. Overexpression of by proteolysis upon import into mitochondria, exposing a
c-FLIPL has been reported to bind proteins involved in these novel tetrapeptide motif that binds the BIR domains of IAPs.
signaling pathways, including TRAF1, TRAF2, RIP, and These IAP antagonists compete with caspases for binding to
RAF1, and thereby promotes the activation of NF-𝜅𝛽 and IAPs, thus freeing caspases from the grip of the IAPs and
ERK as downstream molecules [39]. promote apoptosis. Thus, some modern synthetic peptides-
In TRAIL-resistant nonsmall cell lung cancer (NSCLC) based therapeutics that mimic SMAC and HtrA2 trigger
cells, c-FLIP and RIP, have been shown to be essential for apoptosis or sensitize tumor cell lines to apoptosis induced
TRAIL-induced formation of the DISC in nonraft domains by cytotoxic anticancer drugs or TRAIL in vitro and even in
of the plasma membrane and consequent activation of NF- some tumor xenograft models [52].
𝜅𝛽 and ERK cell survival signals. In contrast, knockdown
of c-FLIP has been found to redistribute the DISC to lipid 2.6. How the Cancer Cells Overcome the Therapeutic Agents-
rafts and switch DISC signaling to TRAIL-induced apoptosis Induced Apoptosis. The success of each therapeutic strategy
[47]. NF-𝜅𝛽 constitutive activation has been linked to the depends mainly on the ability of the therapeutic tool to
pathogenesis of many human cancers [48], and inhibition of induce apoptosis either by targeting the overexpressed anti-
NF-𝜅𝛽-induced transcription of c-FLIP has been shown to apoptotic proteins or by stimulating the expression of the
sensitize cells to death receptor-mediated apoptosis [49, 50]. proapoptotic molecules. However, many of the therapeutic
NF-𝜅𝛽-mediated upregulation of c-FLIP is implicated as an agents are still challenged by maneuvering(s) from the cancer
important factor in the evasion of cell death by cancer cells. cells to survive treatments. Alterations in the expression levels
or mutation of a chemotherapeutic drug’s target(s) may have
2.5. Carcinogenesis via Convergence Pathway Inhibition. The an impact on apoptosis by such drug. Here, we list some of
intrinsic and extrinsic pathways for caspase activation con- the mechanisms by which cancer cell can overcome three of
verge on downstream effector caspases. Mechanisms for the standard therapeutic agents.
suppressing apoptosis at this distal step have been revealed,
and their relevance to cancer is becoming progressively 2.6.1. 5-Fluorouracil (5-FU). The fluoropyrimidine 5-fluor-
clear. In this regard, the apoptosis-inhibiting proteins (IAPs) ouracil (5-FU) is widely used in cancer treatment includ-
represent a family of evolutionarily conserved apoptosis ing colorectal and breast cancer treatments [53, 54], but
suppressors (𝑛 = 8 in humans), many of which function resistance to that drug remains a major clinical obstacle; 5-
as endogenous inhibitors of caspases. All members of this is part of its antitumor activity and FU targets the tumor
family, by definition, contain at least one copy of a so-called suppressor p53 and subsequently triggers the cell cycle. 5-
BIR (baculovirus iap repeat) domain, a zinc binding fold, FU-induced apoptosis is p53-dependent; however, apoptosis
which is important for their antiapoptotic activity, present in can also occur in p53 mutant cell lines by a mechanism
1–3 copies. Caspases-3 and -7, as well as caspase-9 (intrinsic still unknown [55–57]. 5-FU is an analogue of uracil with a
pathway), are directly affected by the human IAP family fluorine atom at the C5 position of the pyrimidine ring. Inside
members, XIAP, cIAP1, and cIAP2. For XIAP, the second cells, 5-FU is converted into different active metabolites,
BIR domain and the linker region between BIR1 and BIR2 including fluorodeoxyuridine monophosphate (FdUMP),
BioMed Research International 5

fluorodeoxyuridine triphosphate (FdUTP), and fluorouri- mechanisms of resistance to antimicrotubule agents remains
dine triphosphate (FUTP). These metabolites have been to be determined.
implicated in both global RNA metabolism due to the incor-
poration of the ribonucleotide FUMP into RNA and DNA 2.6.3. Cisplatin. The therapeutic activity of cisplatin and car-
metabolism due to thymidylate synthase (TS) inhibition or boplatin is mediated by an active species, formed by aqueous
direct incorporation of FdUMP into DNA, leading to a wide hydrolysis as the drug enters the cell. This active species
range of biological effects which trigger apoptotic cell death. interacts with DNA, RNA, and protein, but the cytotoxic
The inhibition of TS is believed to be the primary anticancer effect seems to be primarily mediated via the formation of
activity of 5-FU [58, 59]; however, 5-FU has been shown to DNA interstrand and intrastrand crosslinks. These platinum-
acutely induce TS expression in both cell lines and tumors DNA adducts are recognized by a number of proteins,
[60]. This induction of TS seems to be due to inhibition including those involved in nucleotide excision repair (NER),
of a negative feedback mechanism in which ligand-free TS mismatch repair (MMR), and high-mobility group proteins
binds to its own mRNA and inhibits its own translation (such as HMG1 and HMG2) [78]. Platinum-induced DNA
[61]. When stably bound by FdUMP, TS can no longer damage is normally repaired by the NER pathway, whereas
bind its own mRNA and suppress translation, resulting in the MMR pathway seems to trigger apoptosis [79, 80]. The
increased protein expression. This constitutes a potentially mechanism by which damage recognition results in apoptosis
important apoptosis resistance mechanism, as acute increases is unclear; however, in vitro data support a role for cisplatin-
in TS would facilitate recovery of enzyme activity. Thus, TS mediated activation of the c-ABL and JNK/SAPK pathway in
expression acted as a key determinant of 5-FU sensitivity apoptotic signalling in MMR-proficient cells [81]. In addition,
[62] while in vitro studies have validated the association the induction of p53, possibly by the ATM-CHK2 pathways,
between TS expression and apoptosis 5-FU [63, 64] while following DNA strand breaks leads to apoptosis via the
the improved response rates of 5-FU-based chemotherapy intrinsic pathway and also contributes to the cytotoxicity of
were observed in patients with low tumour TS expression cisplatin [82].
[65, 66]. Recently, genotyping studies have found that patients Bcl-2 members are localized in the mitochondria and
homozygous for a particular polymorphism in the TS pro- have either proapoptotic (Bax, Bak, Bid, and Bim) or anti-
moter (TSER3/TSER3) that increases TS expression are less apoptotic (Bcl-2, Bcl-xL, and Bcl-W) functions [8, 83, 84].
likely to respond to 5-FU-based chemotherapy than patients These members are involved in the downstream action by
who are heterozygous (TSER2/TSER3) or homozygous for cisplatin. These proteins form either homodimers (such as
the alternative polymorphism (TSER2/TSER2) [67]. Col- Bcl-2/Bcl-2) or heterodimers (e.g., Bcl-2/Bax) depending
lectively, many evidences indicate that high TS expression on the levels present of each component. Excess level of
correlates with increased 5-FU resistance. homodimers can either inhibit (e.g., Bcl-2/Bcl-2) or induce
(e.g., Bax/Bax) apoptosis. Although it is not confirmed
2.6.2. Antimicrotubule Agents (AMA). Microtubules are whether cisplatin can directly modulate levels of the anti-
highly dynamic cytoskeletal fibres composed of tubulin sub- apoptotic protein or not, there is evidence that cisplatin
units (most commonly 𝛼- and 𝛽-tubulin) that have crucial significantly transactivates the bax gene by wild-type p53.
roles in maintaining cell shape, in cell signalling, in cell divi- Thus, an increase in the Bax to Bcl-2 ratio by cisplatin-
sion and mitosis, and in the transport of vesicles and mito- induced p53 has been reported to activate the apoptotic
chondria and other cellular components throughout the cell process [85]. However, extrapolating experimental results
[68]. The rapid polymerization/depolymerization dynamics to the clinic results needs much caution, for instance, the
of microtubules are critical for proper spindle function and demonstration that experimental overexpression of bcl-2 in
accurate chromosome segregation during mitosis. tumors leads to the expected cisplatin resistance [86–88].
That was opposite to a clinical study, which reported that
Taxanes, such as paclitaxel and docetaxel, and vinca cisplatin surprisingly improved survival of ovarian cancer
alkaloids, such as vinblastine and vincristine, suppress micro- patients with increased bcl-2 gene expression [87]. Now, it is
tubule polymerization dynamics, which results in the slowing understood that proapoptotic homodimers affect cisplatin-
or blocking of mitosis at the metaphase-anaphase boundary induced apoptosis by first stimulating the mitochondria to
[69]. These compounds can block or slow mitosis and sub- release cytochrome c, which in turn activates a series of
sequently cells eventually die by apoptosis [70]. Microtubule proteases that includes caspase-1, -3, and -9 [84, 89–91] which
polymer levels and dynamics are regulated by many factors, are the final effectors of drug mediated apoptosis.
including expression of different tubulin isotypes [71]. Cells
can resist paclitaxel and vinca alkaloids by changing the
microtubules dynamics and levels of tubulin isotypes [72–74]. 3. Molecular Targeting
Sometimes, a reduction in total intracellular tubulin levels Therapies and Apoptosis
could be observed in paclitaxel-resistant cells. Therefore,
the development of resistance to antimicrotubule agents Expression of inhibitors of apoptosis as well as inactivation
may involve changes in the microtubule polymer mass and of apoptosis promoters is observed in human cancers [92].
expression of different tubulin isotypes. Furthermore, specific Moreover, functional defects in apoptotic signal transduction
𝛽-tubulin mutations that alter sensitivity to paclitaxel have may translate into drug resistance of cancer [86, 93]. Over
been reported in vitro [75–77]. The clinical relevance of these the past decade, significant advances have been made in
6 BioMed Research International

discovery and validation of several types of novel cancer interacting with retinoid receptors and other transcription
therapeutics. Such novel therapeutics tries to prime the factors. These inhibitors can modulate expression of Bcl-2
apoptotic machinery to act as promising apoptosis-inducing or Bcl-xL in some tumor lines [51]. These observations por-
agents, bearing high hopes for the management of cancers tend opportunities for exploiting endogenous transcriptional
resistant to conventional treatments. [94]. This therapeutics pathways for suppressing expression of antiapoptotic Bcl-2-
can be used either alone or in combination with some of the family genes in cancer, together with the idea of employing
conventional therapies. Several review articles have shed light them as chemo- or radiosensitizers, rather than relying on
to the great potential of apoptosis-targeted therapies [95, 96]. their antitumor activity as single agents. The genetic char-
Some of which focus on Bcl-2 family proteins [97–99], IAPs acteristics of cancer cells that dictate response or resistance
[100], or c-FLIP [101] and the approaches that have been to SRTF-ligands, as well as the complex pharmacological
used to modify their activity to reactivate apoptosis and thus interplay between this class of agents and conventional
eradicate cancer cells. cytotoxic drugs, need to be investigated more.
As below, we refer readers to these reviews for specific Because Bcl-2 antisense enhances sensitivity to cytotoxic
details regarding the benefits of some novel therapeutic anticancer drugs in vitro and in xenograft models, most
agents directed against some antiapoptotic targets, which clinical trials combine the antisense agent with conventional
have been shown to demonstrate enhanced apoptotic killing chemotherapy. More precisely, antisense oligonucleotides
and sensitize resistant cancer cells to antineoplastic agents. targeting the Bcl-2 mRNA have been introduced to Phase III
Nevertheless, we show below a few of the most promising clinical trials for melanoma, myeloma, CLL, and AML and
therapeutic strategies which target the induction of apoptosis. Phase II activity underway for a variety of solid tumors. [102].
Moreover, Phase II data suggest a benefit from adding Bcl-2
antisense to conventional therapy, but definitive proof awaits
3.1. Targeting Antiapoptotic Bcl-2 Family Members. The com- the Phase III results.
plex interplay between proapoptotic and antiapoptotic mem- Beside the antisense oligodeoxynucleutides (e.g., G3139
bers of the Bcl-2 family plays a crucial role in cellular [103]), various drugs molecules that target Bcl-2 and Mcl-
fate determination. Attempts to overcome the cytoprotective 1 have been tested for their abilities to induce apoptosis
effects of Bcl-2 and Bcl-xL in cancer include three strategies: in cancer cells [104]. Understanding how cells keep these
(1) shutting off gene transcription, (2) inducing mRNA antiapoptotic proteins controlled is the base on which the
degradation with antisense oligonucleotides, and (3) directly therapeutic potency for targeting the Bcl-2 and Bcl-xL pro-
attacking the proteins with small-molecule drugs. Herein, teins by these molecules is established. In this regard, a large
we will provide some of the most promising strategies for family of endogenous antagonists of Bcl-2 and Bcl-xL has
modulating the activity of apoptosis genes and their proteins been revealed (so-called “BH3-only proteins”), possessing a
for cancer therapy. conserved BH3 domain that binds a hydrophobic crevice
Some members of the steroid/retinoid superfamily of on the surface of Bcl-2 and Bcl-xL [100]. Synthetic BH3
ligand-activated transcription factors (SRTFs) represent peptides generated as small-molecule drugs that occupy
potentially “drugable” modulators of Bcl-2 and Bcl-xL gene the BH3 binding site on Bcl-2 or Bcl-xL, abrogating their
transcription although they are not components of the “core” cytoprotective functions, have been validated [105]. Many
apoptosis machinery. For example, expression of Bcl-2 is research groups have presented preclinical data regarding
estrogen-dependent in the mammary gland. Consequently, BH3-mimicking compounds. The optimal structure-activity
antiestrogens, such as tamoxifen, inhibit endogenous Bcl-2 relation (SAR) profile of these compounds had been deter-
expression in breast cancer cell lines, promoting sensitivity to mined. These molecules showed broad-spectrum activity
cytotoxic anticancer drugs such as doxorubicin. In addition against the various antiapoptotic members of the Bcl-2-
to antiestrogens in estrogen receptor- (ER-) positive breast family (Bcl-2, Bcl-xL, Mcl-1, Bcl-W, Bfl-1, and Bcl-B), in terms
cancers, expression of Bcl-2 or Bcl-xL can be downregulated of balancing antitumor efficacy. Of these molecules, which
in specific types of cancer and leukemia cells by small target Bcl-2, are the small molecule Bcl-2 inhibitors (e.g.,
molecule drugs that modulate the activity of retinoic acid HA14-1; [103]) and BH3 peptidomimetics [106]. The most
receptors (RAR), retinoid X receptors (RXR), PPAR, vitamin prominent molecules among them are the Bcl-2 antagonists
D receptors (VDR), and certain other members of the SRTF ABT-737 and ABT-263 [107, 108]. That novel Bcl-2/Bcl-
superfamily. RAR and RXR ligands are already approved for xL/Bcl-W inhibitor (ABT-737) has been developed. It showed
treatment of some types of leukemia and lymphoma and are enhanced activity of chemotherapeutic agents or ionizing
in advanced clinical testing for solid tumors. PPAR mod- radiation in the preclinical studies. That compound displays
ulators have demonstrated antitumor activity in xenograft impressive antitumor activity in human tumors including
models of breast and prostate cancer [10], sometimes dis- follicular lymphoma, chronic lymphocytic leukemia, and
playing synergy with retinoids, probably due, in part, to small cell lung cancer (SCLC) in vitro and in vivo [107] and
the fact that PPAR binds DNA as a heterodimer with RXR. in nonsmall lung cancer (NSCLC) [109].
Moreover, troglitazone, a potent PPAR agonist, can lower ABT specifically bind to Bcl-2, Bcl-xL, and Bcl-W flipping
serum PSA in men with advanced prostate cancer; however, them from Bak and Bax. ABT-737 mimics the BH3-only
its proapoptotic features are not confirmed in vivo yet. protein, Bad, by docking to the hydrophobic groove of
Compounds that inhibit histone deactylases (HDACs), antiapoptotic proteins, thereby disabling their capacity to
called HDAC inhibitors, act as transcriptional repressors by antagonize the actions of proapoptotic proteins. Therefore,
BioMed Research International 7

ABT-737 is highly potent (i.e., at nanomolar concentrations) transcription-independent role of p53 in apoptosis has been
in killing tumor cells which are dependent upon Bcl-2 for attributed to its ability to block the antiapoptotic function
survival [110]. Although ABT targets almost all the proapop- of Bcl-2 and Bcl-xL on the mitochondria [121, 122]. p53
totic family members of Bcl-2 except Mcl-1, it has been promotes, directly or indirectly, the apoptotic activities of
found that its combination with deoxyglucose has efficiently proapoptotic members of the Bcl-2 family, Bax and Bak
eliminated cancer cells in vitro and in vivo [111]. ABT-737 [121, 123]. Moreover, after targeting DNA by some damaging
was modified at three sites to create improved ABT-263 for agents, p53 contribute as a transcription factor. p53 transcrip-
oral bioavailability without the loss of affinity to the Bcl-2 tionally upregulates genes such as those encoding p21WAF-
family of proteins [108]. As a single agent, however, ABT- 1/CIP-1 and GADD45, which induce cell cycle arrest in
263/737 (ABT) induces apoptosis only in limited tumor types, response to DNA damage [124, 125]. However, p53 can trigger
such as lymphomas and some small cell lung carcinomas elimination of the damaged cells by promoting apoptosis
[107, 108, 112]. through the upregulation of proapoptotic genes such as Bax,
Other strategies for countering Bcl-2 and Bcl-xL in cancer NOXA, TRAIL-R2 (DR5), and Fas (CD95/Apo-1) [12, 55, 126,
include over-expression of opposing proapoptotic family 127]. However, the cell can choose between p53-dependent
members such as Bax with p53 adenovirus (in Phase III trials) cell cycle arrest or apoptosis. How cells choose between
or with Mda7 (IL-24) adenovirus gene therapy (completed this p53-mediated cell cycle arrest and apoptosis has three
Phase I trials), as well as Bax adenovirus gene therapy for possible models. (I) One model suggests that more profound
regional cancer control. Alternatively, it might be possible to DNA damage induces higher and prolonged activation of
activate the Bax and Bak proteins using drugs that mimic p53, which increases the chances of apoptosis over arrest
agonistic BH3 peptides [105]. [128]. (II) Another model suggests that different cell types
One further strategy to counteract Bcl-2 family members might keep different p53-regulated genes in regions of active
is the monoclonal antibodies which target the tetra-span chromatin, which determines the cassette of genes that
membrane protein CD20 (Rituximab) but can also down- are transcriptionally upregulated [129]. (III) A third model
regulate expression of Bcl-2 family members, Mcl-1 (intrin- proposes that the availability of transcriptional cofactors
sic pathway) and XIAP (convergence pathway) in certain determines the ability of p53 to activate different subsets of
leukemias, providing yet another example of unexpected genes [130, 131].
links to apoptosis pathways through receptor-mediated sig- DNA damage-inducing apoptotic factors induce the
naling [113]. activation of upstream kinases such as ATM (ataxia-
telangiectasia mutated), ATR (ATM and Rad-3 related),
3.2. Targeting Cell Cycle Control System for Apoptosis Induc- and DNA-PK (DNA-dependent protein kinase), which can
tion. The relationship between cell cycle arrest and apoptosis directly or indirectly activate p53. Phosphorylation of p53 by
is complex. Signaling pathways following apoptosis-inducing upstream kinases inhibits its negative regulation by MDM-
stimuli either arrest the cell cycle following damage signals to 2, which targets p53 for ubiquitin-mediated degradation
allow time for repair or switch to initiation of apoptosis after [132]. In fact, lack of functional p53 contributes to apoptosis
apoptogenic agents. These agents, such as ionizing radiation resistance due to the inability to undergo p53-mediated
or some chemotherapy, induce apoptosis in many cell types apoptosis sometimes due to p53 mutations. In vitro studies
[114]. Apoptosis is frequently associated with proliferating have reported that loss of p53 function reduces apoptosis
cells. This implies the existence of molecules in late G1 by 5-FU [57]. Other clinical studies have found that p53
and S phase, whose activities facilitate execution of the overexpression (a surrogate marker for p53 mutation) corre-
apoptotic process. Once the cells are committed to cell death, lates with resistance to apoptosis by 5-FU [133, 134]. Other
apoptogenic factors, including cytochrome c, are released studies have found no such correlation [135]. Such conflicting
from mitochondria to initiate a caspase cascade [115]. findings may be due, at least in part, to the fact that p53
An important cell cycle regulator plays a central role in overexpression does not actually reflect p53 mutation in as
the regulation of cell cycle arrest and cell death is the tumor many as 30–40% of cases [136]. A recent study has suggested
suppressor, p53 [116]. It has recently been discovered that the that certain p53 mutants may increase dUTPase expression,
p53 activates apoptosis through parallel pathways that may inhibiting apoptosis by 5-FU [137].
depend on transcription events or not [117]. Under many A number of in vitro studies have demonstrated decreased
types of stress, p53 induces cell cycle arrest and apoptosis to cisplatin-induced apoptosis in p53 mutant tumor cells [138–
maintain genomic integrity and prevent damaged DNA being 140]. Some studies have reported that disruption of p53 func-
passed on to daughter cells. Tumor cells have inactivated p53. tion actually enhances apoptosis by cisplatin [141, 142]. This
About 50% of this inactivation is achieved through mutations enhanced apoptosis was attributed to defects in p53-mediated
in its sequence and the rest by disabling key components cell cycle arrest, which reduced time for DNA repair. Some
that lie upstream or downstream of p53 in a common tumors with mutant p53 have a worse clinical outcome which
signaling pathway [118]. In p53 wild-type tumors, p53 may means failure to induce apoptosis by therapeutic drugs [143].
be compromised, by inactivation of positive regulators of p53 In vitro data from other laboratories and others suggest that
activity (such as p14ARF [119]) or over-activation of negative development of resistance to apoptosis by taxan is feasible
regulators of p53 activity (such as Akt [120]). in p53 null cells [144]. The clinical relevance of p53 status
Bax is one of the best known downstream p53 tran- for taxan-induced is yet to be determined. In vitro studies
scriptional targets that directly affect apoptosis. A further have found that p53 status does not affect paclitaxel-induced
8 BioMed Research International

apoptosis, [145–147]. On the other hand, doxorubicin was and various feedback loops exist to activate Bax and thereby
more active against p53 wild-type tumor xenografts than to amplify the apoptotic cascade in hematopoietic and other
p53 mutant and null cells [148]. Furthermore, apoptosis cell types.
induction by doxorubicin in acute lymphoblastic leukaemia It is now clear that, through its dual role, cyclin E provides
cell lines was found to be dependent on the presence of a physiological balance between cell proliferation and death
wild-type p53 [149]. In the clinical setting, p53 mutations of hematopoietic cells [161]. Cyclin E has an initial limited
have been correlated with lack of response to doxorubicin activation of caspase-3 perhaps through a small amount
in patients with locally advanced breast cancer [150]. These of primed/activated mitochondrial Bax. Then, cyclin E is
data indicate that doxorubicin is more active against p53 wild- proteolytically cleaved by caspase-3 to generate p18-cyclin E,
type tumors. Fichtner et al. found no correlation between which interacts with Ku70 resulting in release of Bax from the
p53 status and sensitivity to the apoptogenic carboplatin, Ku70-Bax complex. This contributes to a robust activation of
in xenotransplanted colorectal tumors [151]. Similarly, Jacob Bax, leading to its mitochondrial targeting and amplification
et al. found that p53 status did not correlate with CPT- of the intrinsic pathway of apoptosis. This is consistent with
11-induced apoptosis in a panel of colorectal cancer cell the idea that limited caspase activation is not sufficient to
lines [152, 153]. Some studies have found that carboplatin release enough cytochrome c leading to significant apoptosis,
induces apoptosis to a similar extent in isogenic p53 wild- at least, in hematopoietic cells. Rather, an amplification loop
type and null cancer cells [118]. In general, all the above is required to trigger massive cytochrome c release from
mentioned studies suggest that p53 mutant tumors may be mitochondria resulting in abundant caspase-3 activation and
equally responsive to apoptosis inducers or possibly more cell death [115]. Caspase-3 and caspase-7 may participate in
sensitive than p53 wild-type tumors. That led to application a feedback amplification loop to promote the mitochondrial
of combinatorial drug therapy of paclitaxel and cisplatin for apoptotic pathway [162].
maximum apoptosis induction in p53 mutant ovarian tumors
than p53 wild-type tumors as introduced [154]. 3.2.2. Ku70, from Cyclin E to Bax. As we previously men-
tioned, the key regulatory step in the intrinsic cell death
3.2.1. Cyclines. Another important family, critically control- pathway is the activation of the proapoptotic molecule, Bax
ling cell cycle, is cyclines [155, 156]. Here, we will focus on [163]. Although the mechanism of Bax activation remains
an important member of this family, cyclin E, which has unclear, a recent study indicates that Bax is kept inactive
direct relation with apoptosis signals. Cyclin E has been in some cells, to counteract apoptosis, by interacting with
shown to have a Cdk2-independent function in regulating cytosolic Ku70 [163, 164]. Overexpression of Ku70 prevents
apoptosis of hematopoietic cells [157, 158]. The larger N- Bax-mediated apoptosis, whereas targeting Ku70 causes cells
terminal truncation of cyclin E that generates p18-cyclin to become more sensitive to a variety of apoptotic stim-
E prevents its binding to either Cdk2 or CKIs p21Waf1 uli. Importantly, we showed that some of these apoptotic
or p27Kip1. Therefore, unlike the other isoforms of cyclin stimuli which are not directly targeting DNA break like the
E, p18-cyclin E cannot bind Cdk2 and therefore interferes microtubule stabilizing agents, Taxane, and hyperthermia
with Cdk2-dependent (bound to cyclin E1, E2, or A) kinase stimulated caspase-dependent apoptosis after Ku70 inhibi-
activity. Cleavage of cyclin E results in loss of its Cdk2 tion (unpublished data). These results demonstrated that
associated kinase activity and consequently its function in Ku70 is a physiological inhibitor of Bax-induced apoptosis;
the cell cycle [157]. Recently, it has been reported that p18- a finding that expands on Ku70’s previously known role
cyclin E mediates p53 function-independent apoptosis by in DNA repair [163], particularly through nonhomologous
regulating the relationship between Bax and Ku70 [159]. That end joining (NHEJ). Moreover, cells from Ku70 knockout
study’s findings established a role for Ku70 in regulating Bax- mice are hypersensitive to agents that induce apoptosis in
mediated apoptosis in hematopoietic cells that generate p18- the absence of DNA damage, such as staurosporine [159].
cyclin E. By interacting with Ku70, p18-cyclin E releases Bax This suggests that Ku70 plays a role in suppressing apoptosis
from Ku70, thus allowing Bax to be activated during gentoxic
independently of its role in DNA repair [165]. Based on
stress and thereby leading to apoptosis [159]. Therefore,
these findings, nuclear Ku70 is therefore considered to be
p18-cyclin E, Bax, and Ku70 emerge as key regulators of
mainly responsible for repair of DNA damage, whereas
the apoptotic pathway at least in hematopoietic cells. The
sensitivity of these cells to genotoxic stress can be explained the cytosolic pool of Ku70 may be primarily a regulator
by their ability to generate p18-cyclin E. of Bax activation [162, 164]. Therefore, Ku70 plays a dual
The cleavage of cyclin E can be prevented by Bcl-2 role in NHEJ and regulation of Bax, which in turn may
overexpression, indicating that Bcl-2 functions upstream of determine the balance between cell survival and programmed
the mitochondrial apoptotic cascade [115, 158]. How Bax is cell death in hematopoietic cells. Ku70 may not be unique
released from its complex with Ku70 is still unknown. One in assuring Bax sequestration in the cytosol. Other Bax
mechanism may consist of a structural change in conforma- interactive proteins such as the peptide humanin [166], the
tion of the Bax binding-domain of Ku70 upon binding of p18- apoptosis repressor with caspase recruitment domain (ARC)
cyclin E at the amino terminus. Another possibility might be [167], or crystalline [168] may also function in various cellular
the recruitment of histone acetyl transferases, which would systems to sequester Bax. On the other hand, after apoptotic
acetylate several lysine residues at the carboxy terminus of stimulation other Bax-interacting proteins, such as p53 [169]
Ku70 [160]. However, it is quite possible that other pathways and Bif- [170] and apoptosis-associated speck-like protein
BioMed Research International 9

(ASC) [171], and MOAP-1 [172] promote Bax conformational result in hypersensitivity to cisplatin, and that restoration
change and mitochondrial translocation. of NER activity reduces sensitivity to more normal levels
[179]. NER is a complex process involving at least 17 different
3.2.3. Nonapoptotic Function of Caspases. Although caspases proteins; however, upregulation of only a few rate limiting
are classically known for their role in apoptosis, their components of the NER system is necessary to increase a cell’s
involvement in biological processes apart from cell death has capacity for NER [180]. One of these important rate-limiting
recently received increased attention. For example, they play factors is the excision repair cross-complementing 1 protein
an important role in the inflammatory signaling pathways (ERCC1), which plays a crucial role in response to apoptosis
[173, 174]. The nonapoptotic functions of these proteases inducer, cisplatin. Increased expression of ERCC1 associated
suggest that they may become activated independently of the with cisplatin resistance [181–183]. Clinically, high levels of
apoptotic cascade, thus leading to the cleavage of specific ERCC1 are correlated with poor response to platinum-based
substrates, such as kinases, cytokines, transcription factors, chemotherapy in ovarian, gastric, and nonsmall cell lung
and polymerases. cancers (NSCLC) [184–186]. In addition, high mRNA expres-
sion of ERCC1 and thymidin synthase (TS) has been acting
Functional analysis of conditional caspase-deficient mice
as predictors for poor response to combined oxaliplatin/5-
or derived cells confirmed the crucial roles of caspases in cell
FU treatment of advanced colorectal cancer [187]. Targeting
proliferation, differentiation, and inflammation. It has been
ERCC1 expression, using an antisense expression construct,
reported recently that caspase-3 is a negative regulator of B-
made ovarian cancer more sensitive to cisplatin than controls
cell cycling. This study shows that while the apoptotic path-
in cell line and xenograft models [188]. Another central
way is largely unaffected in caspase-3 knockout mice, splenic
component of the NER machinery, xeroderma pigmentosum
B-cell proliferation is enhanced in vivo, and in response to
group A (XPA), has been found to be overexpressed in
mitogenic stimulation in vitro, hyperproliferation is seen.
apoptosis resistant cancer cells under apoptosis inducer,
Genetic and biochemical data demonstrate that caspase-3
cisplatin, treatment [184]. Collectively, these studies suggest
is essential for the regulation of B-cell homeostasis [175].
an important role for NER in mediating resistance to DNA
In contrast to caspase-3, caspase-8 is a positive regulator of
damage and subsequent apoptosis by platinum.
cell cycling as B-cell proliferation since caspase-8 deficient
Another DNA damaging agent is 5-FU. 5-FU treatment
mice fail to proliferate [176, 177]. Recently, it has been
induces TS inhibition which subsequently causes nucleotide
suggested that caspase-3 and caspase-7 together play a role in
imbalances that severely disrupt DNA synthesis and repair
development, since double knockout mice exhibit defects in
[189]. TS inhibition results in the accumulation of deoxyuri-
heart development, leading to death immediately after birth.
dine triphosphate (dUTP), as the essential conversion process
In contrast, knockout of either caspase individually on the
of deoxyuridine monophosphate (dUMP) to deoxythymi-
same genetic background resulted in viable animals.
dine monophosphate (dTMP) is blocked [190]. Both dUTP
and another 5-FU metabolite, fluorodeoxyuridine triphos-
3.3. Targeting DNA Repair System for Apoptosis Induction. phate (FdUTP), can be misincorporated into DNA. Repair of
DNA repair can determine fate of a cell with or without uracil and 5-FU-containing DNA could be achieved by the
treatment stress which either resists over to survive or base excision repair enzyme uracil-DNA-glycosylase (UDG)
respond to toxic agents. When the repair mechanisms are [191]. However, this repair mechanism is futile in the presence
unsuccessful, it may cause cellular senescence (permanent of high (F)dUTP/dTTP ratios causing false nucleotide incor-
cell cycle arrest), oncogenesis, or apoptosis which plays an poration. These futile cycles of misincorporation, excision,
essential role in survival of the organisms by preventing the and repair eventually lead to DNA strand breaks. However,
multiplication of mutated chromosomes, as well as elimina- some cells show resistance to this mechanism by increased
tion of indisposed cells, normal embryonic development, and dUTPase. The increased dUTPase expression has been asso-
maintenance of cell homeostasis. [178]. In cancer treatments, ciated with resistance to TS inhibitors like 5-FU [192–194].
apoptosis-inducing chemotherapeutic drugs induce DNA Another mechanism is the DNA mismatch repair (MMR)
damage either directly (e.g., platinum drugs) or indirectly which is an important mechanism that controls the decision
(e.g., 5-FU and topoisomerase poisons). The response to from survival to apoptosis. The main function of the MMR
DNA damage is either repair or cell death. Only when system is to scan newly synthesized DNA and remove single
repair is incomplete, for example, when the DNA damage nucleotide mismatches that arise during replication. MMR
is too extensive, cells will undergo apoptosis. Herein, we deficiency is implicated in the development of apoptosis-
describe some major mechanisms through which cells can resistance against wide range of DNA-damaging agents,
overcome apoptosis by damaging agents and we introduce including platinum drugs [195]. It is possible that MMR
some example molecules which had successfully targeted in recognition of DNA damage may trigger an apoptotic path-
some types of cancer to stimulate or enhance apoptosis by way or that futile cycles of DNA damage/repair mediated
therapeutic agents. by the mismatch repair machinery generate lethal DNA
Nucleotide excision repair (NER) is the major pathway for strand breaks [196]. MMR has a role beyond that of repair
DNA repair after DNA stress. However, for example, cisplatin in response to DNA damage. Inherited defects in the DNA
induces platinum-DNA adduct which could be removed MMR genes, especially hMLH1 and hMSH2, are common
by NER as challenge mechanism by cancer cell to repair in certain familial forms of cancer such as hereditary non-
of platinum-induced DNA damage [178]. Defects in NER polyposis colon cancer (HNPCC). Such defects were also
10 BioMed Research International

observed in a variety of sporadic tumors, including colorectal, and hMSH2; thus, it plays a role in DNA repair [208]. BRCA1
breast, and ovarian cancers [197–200]. DNA MMR deficiency also plays a role in the activation of cell cycle checkpoints in
gives rise to the microsatellite instability (MSI) phenotype, response to different cellular stresses including DNA damage
which is detected as variations in lengths of DNA repeat and disruption of microtubule dynamics [209, 210]. Now,
sequences present in the genome [201]. Cisplatin-induced BRCA1 has been implicated in the regulation of apoptosis
apoptosis and acquisition of the drug resistance were found to [211]. It appears that, similar to p53, BRCA1 may function as
be associated with hypermethylation of the hMLH1 promoter a sensor of cell stress by relaying signals to either the cell-
[202]. One clinical study found that ovarian tumors that were cycle checkpoint machinery or cell death machinery. A recent
initially microsatellite stable before cisplatin-based stress study by Quinn et al. found that BRCA1 acts as a differential
exhibited MSI in the residual tumors after chemotherapy, and modulator of chemotherapy-induced apoptosis in breast can-
that was proposed to be due to loss of hMLH1 expression cer cells [212]. They found that BRCA1 enhanced sensitivity
[203]. Hence, preclinical studies strongly implicate loss of to apoptosis induced by antimicrotubule agents, such as
MMR, in particular loss of hMLH1, in overcoming apoptosis paclitaxel and vinorelbine, but inhibited apoptosis induced
by cisplatin. Reversal of hMLH1 promoter methylation by by DNA-damaging agents such as cisplatin and etoposide.
DNA methyl transferase inhibitor, 2-deoxy-5-azacytidine Moreover, inhibitors of poly (adenosine diphosphate- (ADP-)
(DAC), has been shown to resensitize cells to a range of ribose) polymerase (PARP) show the synthetically lethal
apoptotic agents, including platinum drugs [202]. Thus, effect in BRCA1-defective tumors because BRCA1 encodes
DAC has now entered clinical trials in combination with protein that is required for efficient homologous recombina-
carboplatin in some cancers. Finally, a clinical study found tion (HR) [213].
that MMR deficiency was a predictive factor for apoptosis
induction by carboplatin in advanced colorectal cancer [204]; 3.3.3. Ku Protein, from Gatekeeping to Caretaking. Double
however, MMR deficiency had no effect on sensitivity to the strand break (DSB) repair includes homologous recombina-
non-DNA-targeted apoptosis-inducing agents, paclitaxel and tion (HR) and nonhomologous end joining (NHEJ). DSB
docetaxel. repair through HR process is an error free pathway, since it
In addition to the above mentioned repair mechanisms requires an extensive region of sequence homology between
and other several repair machineries, such as mutagenic the damaged and template strands, whereas NHEJ is an
repair, which may become effective for the recovery of error prone alternate pathway for the repair of DSBs, which
genome against constant attack of numerous genotoxins, essentially joins broken chromosomal ends independent of
UV-radiation, ionizing radiation, and various chemicals are sequence homology. The NHEJ process is initiated by the
responsible for most of the mutagenesis due to a process binding of specific protein to the broken ends, which may
of translesion that inserts an incorrect nucleotide opposite act as end bridging factor [214]. The catalytic subunit of
to the lesion and then continues elongation [205]. Targeting DNA protein kinase (DNA-PKcs) is required in mammalian
any of such mechanisms could enhance apoptosis instead of NHEJ to bridge the DNA ends through their protein-protein
the repair and eradicate the defectively repaired cells under interactions [215].
apoptotic inducers. Cancer susceptibility genes have been categorized into
two classes, namely, caretakers and gatekeepers. Caretakers
are involved in DNA repair and whose inactivation can lead
3.3.1. DNA Double Strand Break Repair and Apoptosis.
to genomic instability while gatekeepers control cell death
Double-strand breaks (DSBs) repair mechanism is one of
and proliferation. Loss of functions in both caretaker and
the widespread mechanisms which efficiently repairs double-
gatekeeper genes results in an increase in cancer suscep-
strand breaks and single-strand gaps in damaged DNA
tibility. Ku86 and p53 are examples of such caretaker and
by a series of complex biochemical reactions, as a result
gatekeeper, respectfully. Their functions are closely linked
of ionizing radiation, UVR, ROS, and chemotherapeutic
[216]. Ku is a caretaker gene that maintains the integrity of
genotoxic chemicals [205]. The lethal effects of double strand
the genome by a mechanism that suppresses chromosomal
breaks (DSBs) can be conquered by the existence of two
rearrangements. Ku complex (a heterodimer of Ku70/Ku80
independent pathways, such as homologous recombination
[≈86]) is a major end binding factor in mammalian cells,
(HR) and nonhomologous end joining (NHEJ). Multiple
possessing end bridging activity [217, 218]. Ku exists as a
proteins are required for DSB repair by recombination, which
heterodimeric DNA binding complex in eukaryotes, con-
are conserved among all eukaryotes. Deficiencies in such
sisting of two protein subunits of about 70- and 80-kDa,
repair mechanism can develop cancer.
known as Ku70 and Ku80 (in S. cerevisiae) or Ku86 (in higher
eukaryotes). This complex was originally identified two
3.3.2. BRCA1. BRCA1 is a tumor suppressor gene that regu- decades ago as a major target of autoantibodies from Japanese
lates cellular responses to DNA stress by apoptotic inducers. patients with scleroderma-polymyositis overlap syndrome
Germline mutations of the BRCA1 gene account for about [219]. Functional Ku proteins are critical for the maintenance
10% of breast and ovarian cancer cases, and lower than nor- of genomic stability, DNA repair, and hence cellular and
mal BRCA1 expression may be an important factor that con- organismal viability [220]. Li and colleagues reported the
tributes into sporadic cancers [206, 207]. Recently, BRCA1 putative tumor suppressor function of Ku70. On the other
has been found to reside in a large DNA repair complex that hand, Ku was identified as a nuclear protein involved in both
includes various mismatch repair proteins including hMLH1 homologous recombination (HR) or nonhomologous end
BioMed Research International 11

joining (NHEJ). Moreover, Ku proteins have been implicated resulting in the release of Bax from Ku70. Subsequently, acti-
in numerous other cellular processes, including telomere vated Bax translocates to the mitochondria. In our laboratory
maintenance, antigen receptor gene arrangements (e.g., vari- we identified Ku70, particularly as a required molecule for the
able (diversity) joining {V(D)J} recombination), regulation cytostatic arrest under hyperthermia treatment in nonsmall
of specific gene transcription and apoptosis, regulation of lung cancer cells. In that study, small interference RNA
heat shock-induced responses, and a newly identified role in against Ku70 could inhibit the cytostatic arrest and switch
regulation of the G2 and M phases of the cell cycle [221]. cells to cytotoxic arrest and caspase-dependent apoptosis
Moreover, an association of the DNA-PK-Ku70-Ku86v with mediated by Bax activation [232].
a number of cell cycle regulatory proteins such as p53, CDK4,
and E2F-4 had been reported, suggesting a pivotal role of Ku 3.4. Targeting Death Receptors (DR). Many tumors refract to
in cell cycle regulation and malignant transformation [222]. the classic apoptosis-inducing cancer therapies which pro-
Recently, Ku70 has been shown to bind to the proapop- voke intrinsic pathway-based apoptosis due to development
totic protein BAX and inhibit BAX-mediated apoptosis in of defects in that pathway under treatment stress [233]. To
vitro by preventing its relocalization to the mitochondria enhance cancer cell apoptosis and thus overcome treatment
[223]. This antiapoptotic function is mediated by a domain failure, many therapeutic strategies targeting other molecules
in the carboxy-terminal of Ku70 and does not require the implicated in apoptosis resistance have been developed [95,
cooperative effects of Ku80. Although the exact function of 96]. Hence, alternative therapies which provoke the extrin-
Ku in these processes is presently still not clear, what is sic pathway-based apoptosis were developed. For example,
certain is that this protein is of fundamental importance in attempts to apply TNF as a cancer treatment were stymied
cell viability especially under treatment stress [223]. Hence, by the proinflammatory effects of this cytokine, because it
the idea of targeting Ku70 had emerged to enhance apoptosis induces both caspase activation pathways and NFKappaB.
by different therapies. That was because tumor cells depend Fortunately, the TNF family cytokines, FasL and TRAIL,
on DNA repair mechanisms to a greater extent than nor- trigger caspase activation without concomitant induction of
mal cells; hence, many cancer therapies target such DNA NFKappaB giving chances for successful cancer therapy via
repair machinery [224]. As DNA-PK plays a crucial role in apoptosis, where TNF failed due to toxicity. The antibodies
the DSBR, the kinase activity of the DNA-PK complex is that trigger the receptor Fas (CD95) unfortunately showed
regulated through the DNA end-binding (DEB) activity of high toxicity to liver [224], while TRAIL and agonistic anti-
its regulatory subunit Ku. Therefore, the regulation of Ku bodies that bind TRAIL receptors appear to be well tolerated
expression and function also plays a vital role in cancer cells in vivo. As a fact, Phase I trial in humans was recently com-
resistance to anticancer therapy. Thus, targeting Ku protein pleted using an agonistic antibody directed against TRAIL-
led to changes in susceptibility to anticancer drug-induced receptor-1 (TRAIL-R1; DR4). In mouse xenograft models
apoptosis [225, 226]. Most exciting are the new clinical bearing human tumor cells lines, TRAIL and its agonis-
studies that target Ku as their primary strategy. Reports have tic antibodies directed against TRAIL receptors have been
shown that the status of Ku or NHEJ in tumors can enhance demonstrated to possess potent antitumor activity [234],
apoptosis by radiotherapy or chemotherapy, suggesting that supporting the idea of using these biological agents as a novel
Ku may be a potential target to overcome resistance during approach to cancer treatment and thereby mimicking some of
cancer treatment. Recently, novel therapies such as the engi- the effector mechanisms normally employed by the immune
neering of selective small molecule, DNA-PK inhibitors, that system in its defense against transformed cells and potentially
enhance radiation-induced tumor control in a mouse-human bypass the defective intrinsic (mitochondrial) pathway to
xenograft model of cancer have been developed and have apoptosis.
shown promising results [224]. Other examples include the Synthetic triterpenoids, such as CDDO and CDDOm,
use of adenovirus-mediated, heat activated antisense Ku70 sensitize solid tumor cell lines to TRAIL and induce apop-
expression that radiosensitizes human tumors [227] and the tosis as single agents in leukemia cells, through a caspase-
use of a Ku antibody to block MM cell adhesion and abrogates 8-dependent mechanism that remains operative even in
the protective effect of membrane Ku [228]. chemorefractory cells [235–237]. Moreover, in leukemia,
Further understanding of the possible link of the main the protein kinase C (PKC) modulator byrostatin induces
functions of Ku in different cell systems may also help in myeloid leukemia cell lines to produce TNF, resulting
answering the intriguing question of how autoantibodies in autocrine engagement of TNF-receptors and apoptosis
against Ku are generated in various autoimmune diseases. induction through a mechanism that is suppressible by
This provides an exciting insight into the future of anticancer TNFR-Fc fusion protein and caspase-8 dominant-negative
development. It has shown that Ku70 siRNA sensitized [238]. Similarly, all-trans-retinoic acid (ATRA) induces
mammalian cells to radiation and topoisomerase II inhibitor, TRAIL production and triggers autocrine induction of apop-
etoposide [229]. Recently, small molecule, HDAC inhibitors, tosis in acute promyelomonocytic leukemia (APML) cells
which targets Ku70 acetylation enhanced DNA damage by that harbor the RAR-PML fusion protein resulting from
DNA-damaging agents in prostate cancer and they enhanced t(15; 17) chromosomal translocations [239].
radiation effect in melanoma cells [230, 231]. Moreover, Sub- TRAIL, a candidate for targeted therapy, which can pref-
ramanian used HDAC inhibitors, which target ku70 acetyla- erentially kill cancer cells, is frequently hampered by various
tion to induce apoptosis in neuroblastoma cells [163]. These mechanisms of cancer cell resistance to apoptosis [240, 241].
molecules increase the acetylation of Ku70 in the cytoplasm, As a cytokine, TRAIL induces apoptosis by binding one of
12 BioMed Research International

its cognate death receptors and plays an important role in or early clinical development. Bivalent SMAC mimetics,
immune surveillance against tumors [49, 240, 241]. Neverthe- which bind to both the BIR2 and BIR3 domains within a
less, in many types of cancer cells, effective antitumor activity single molecule of XIAP as well as the BIR3 domains of two
of TRAIL requires the suppression of aberrantly expressed cIAP molecules, induce their dimerization, and subsequently
negative regulators of apoptosis [240, 241]. In general, as trigger their proteasome mediated degradation, are more
multiple cell signaling mechanisms may control apoptosis, significant potent than their monovalent counterparts [257,
cancer cells can employ a number of different strategies to 258]. Despite evidence indicating that SMAC mimetics show
suppress a protective apoptotic response [6–8]. From other great promise for cancer therapy [257], they can induce
side, survival signaling pathways are linked to the apoptotic loss of cIAP1 and cIAP2, resulting in NIK stabilization and
machinery. These pathways can modulate components of the consequent NF-kB activation [48, 257].
apoptotic machinery itself or key regulatory molecules within Targeting c-FLIP appears to be a highly promising thera-
the core apoptotic pathways [242, 243]. Notably, various peutic strategy to reduce the apoptotic threshold because its
upstream signaling pathways often impinge on the same few overexpression can protect cancer cells from both TRAIL and
central apoptosis control points, which involve some crucial chemotherapy-induced apoptosis, while small interfering
members of the Bcl-2 family, inhibitor of apoptosis (IAP) RNA- (siRNA-) mediated downregulation of c-FLIP is able to
proteins, and FLICE-inhibitory protein (c-FLIP) [6, 7, 38, 40]. sensitize cancer cells to FASL, TRAIL, and chemotherapeutic
agents [39, 101]. Targeting FLIP by siRNAs in a variety
of cancer cell types showed their potential as therapeutic
3.5. NF-𝜅B as a Target for Apoptosis Induction. NF-𝜅B is a
agents to induce apoptosis [101]. Nevertheless, the widespread
transcription factor that was discovered in the nucleus of B
use of such strategy in clinical settings depends on the
cells and binds to the enhancer of the kappa light chain of
safety delivery of siRNA [261]. Thus, many research efforts
immunoglobulin. It is expressed widely in the cytoplasm of all
have been devoted to improve RNA interference (RNAi)
types of cells. NF-𝜅B is an important molecular link between
therapeutics, which lead to important advances in the field
chronic inflammation, cell cycle, cancer development, and
[262, 263]. In May 2008, the first Phase I clinical trial for
cell death [6, 244–247]. The transactive NF-𝜅B suppresses
the treatment of solid tumors involving systemic delivery
apoptosis by inducing the expression of some apoptosis-
of siRNA via targeted nanoparticles was initiated (Clini-
inhibitory genes, including IAPs, cFLIP, TNF receptor asso-
calTrials.gov identifier: NCT00689065) [262]. Importantly,
ciated factor 1 (TRAF1), and TRAF2 [248]. Two typical
data from this trial demonstrated, for the first time, that
prosurvival NF-𝜅B targets are Bcl-xL and XIAP, which can
systemically delivered siRNA can reduce protein levels of a
block apoptosis at multiple steps [249, 250]. Importantly,
specific gene through an RNAi mechanism in humans [264].
Chuang et al. demonstrated that a wide range of cytotoxic
This finding opens the gate for future application of siRNA
drugs (5-FU, doxorubicin, paclitaxel, and cisplatin) activated
as a gene-specific tool for cancer treatment, clinically. These
NF-𝜅B in a panel of cancer cell lines [251]. This suggests that
applications facilitate targeting the antiapoptotic genes’ prod-
NF-𝜅B activation is a general feature of cancer cell response
ucts to kill cancer cell either alone or in combination with
to chemotherapy. For instance, increased NF-𝜅B activity
classic therapies. Recently, c-FLIP and other antiapoptotic
in patients with oesophageal cancer has been correlated
proteins are promising targets for the development of RNAi
with reduced response to neoadjuvant chemotherapy and
therapeutics [101]. As well as RNAi-mediated downregulation
radiotherapy [252]. Now it is believed that NF-𝜅B appears
of c-FLIP, RNAi-based reduction of Bcl-2, Bcl-xL, XIAP, and
to be a critical determinant of drug resistance; its activation
survivin has also been shown to sensitize cancers cells to a
reduces apoptosis by chemotherapy [247]. A number of in
range of cancer therapies, including chemotherapeutic drugs
vitro studies demonstrated that inhibition of NF-𝜅B sensitizes
and TRAIL [265].
cancer cells to chemotherapy-induced apoptosis [253–255].
Furthermore, NF-𝜅B is believed to be a major target for
proteasome inhibitors, as proteasome inhibition prevents 3.7. Targeting Heat Shock Chaperons
degradation of I𝜅B, blocking NF-𝜅B nuclear translocation
3.7.1. HSP90. Heat shock proteins 90 (Hsp90s) are abun-
[255, 256]. Clinical trials with proteasome inhibitors such as
dantly and ubiquitously expressed proteins required for the
bortezomib are underway to inhibit NF-𝜅B signalling and
energy-driven stabilisation, conformation, and function of
enhance drug-induced apoptosis in cancers.
a large number of cellular proteins, named clinets [266,
267]. Several key Hsp90 proteins are involved in the pro-
3.6. Targeting Multiple Pathways Regulators. Like antiapop- cesses characteristic to the malignant phenotype, such as
totic Bcl-2 proteins overexpression, IAPs overexpression invasion, angiogenesis metastasis, and treatment resistance
is associated with poor prognosis and chemoresistance in [268–270]. They also contribute to leading mechanism of
many cancers [257–259]. It acted as a therapeutic target mitogen-activated protein kinases (MAPK) and nuclear fac-
for apoptosis-inducing strategies [260]. Antisense oligonu- tor kappa B (NF-𝜅B) induction [271–273]. Moreover, Hsp90
cleotides against XIAP and survivin have been developed stabilises Raf-1, Akt, and ErbB2 proteins [274–276], which
[249]. In addition small-molecule inhibitors that bind to the are known to be associated with protection against radiation-
BIR2 or BIR3 domain of XIAP, cIAP1, and cIAP2 enhance induced cell death [277–279]. Molecular studies suggested
apoptosis [249]. Most of small-molecule IAP antagonists are that inhibiting HSP90 by targeted therapy could provide
SMAC mimetics, several of which are in advanced preclinical a promising strategy to enhance apoptosis by conventional
BioMed Research International 13

therapies [280–286]. The inhibitor of Hsp90, geldanamycin, in vivo experiments. That opened the gate for inventing the
and its derivatives successfully enhance the radiosensitivity first generation ASO in which the phosphoryl oxygen of
of tumor cell lines derived from different origins, including DNA was replaced with a sulphate to create a phosphothiate
glioma, prostate, pancreas, and cervix [276, 280–284, 287]. backbone, ASOs, intended for cancer therapy. Recently, more
However, geldanamycins have several limitations, includ- modifications were made to the ribose, the 20-position,
ing poor solubility, formulation difficulties, hepatotoxicity, to improve the pharmacokinetic characteristics of second-
and extensive metabolism by polymorphic enzymes, along generation phosphothioate ASOs, for example, OGX-011
with drug efflux by P-glycoprotein; therefore, efforts to [313].
design small synthetic inhibitors of Hsp90 with improved In advanced breast cancer, targeting CLU by OGX-011
bioavailability and lower toxicity are devoted [288–291]. Now, enhanced the apoptotic effect by Trastuzumab, an HER-
pyrazole resorcinol compounds act as stronger inhibitors 2-targeted monoclonal antibody used in the clinical man-
of Hsp90 than geldanamycin derivatives. Currently, the agement of advanced breast cancer patients. Importantly,
isoxazole resorcinol NVP-AUY922 shows the highest affin- only the combination of OGX-011 and Tratuzumab leads to
ity for the NH2-terminal nucleotide-binding site of Hsp90 activation of apoptosis that was not observed with either
[290, 291], whereas NVP-BEP800 represents a novel fully agent alone. Phase I and phase II studies have suggested a
synthetic, orally available 2-aminothieno [2,3-d] pyrimidine combined therapy using ASOs against both of CLU and Bcl-2,
class Hsp90 inhibitor [292]. Both compounds have good along with Trastuzumab for advanced breast cancer patients
pharmaceutical and pharmacological properties. They also [314, 315]. Targeting CLU by ASO enhanced apoptosis by
exhibit strong antiproliferative activity against various tumor paclitaxel or radiation in PC-3, prostate cancer cells [314].
cell lines and primary tumors in vitro and in vivo at well- Phases I and II trials in prostate cancer have shown
tolerated doses [292]. The novel inhibitors of Hsp90 can that OGX-011 significantly enhanced the apoptotic response
sensitize various tumor cells to apoptosis by destabilisation and efficacy of chemotherapy, radiation, and androgen with-
and depletion of several Hsp90 client proteins, thus causing drawal by inhibiting the expression of CLU. In other prostate
the depletion of S phase and G2/M arrest by heat shock, cancer phase I trials, OGX-011 was given prior to radical
increased DNA damage, and repair protraction and finally prostoctomy, while some “well tolerated” phase II trials were
apoptosis [293]. established based on the biologic effectiveness of OGX-011
[315].
3.7.2. Clusterin. Clusterin (CLU) is a cytoprotective heat In our laboratory we reported for the first time the asso-
chaperone protein, involved in numerous physiological pro- ciation between upregulation of the sCLU and persistence
cesses important for carcinogenesis and tumor growth, against apoptosis by paclitaxel in ovarian cancer cells and
including apoptotic cell death, cell cycle regulation, DNA human tumors. Importantly, when we targeted CLU either
repair, cell adhesion, and theromprotection [294]. It is now by OGX-011 or RNAi we successfully enhanced caspase-
accepted that the primary function of the 67–80 kDa cyto- dependent apoptosis and the cells restored their sensitivity
plasmic CLU (sCLU) protein form is cytoprotective as a stress to paclitaxel [144]. Similarly, CLU ASO enhanced cisplatin-
response gene. sCLU expression is consistently reported to induced apoptosis, human bladder cancer cells, and human
be increased with many apoptosis-inducers, chemotherapy or lung adenocarcinoma both in vitro and in vivo [304].
radiotherapy, treatment as an adaptive cell survival molecule
in vivo and in vitro to block apoptosis [295]. That was Conflict of Interests
also shown in human tumor tissues from prostate [296,
297], kidney [298], breast [299], ovarian [300], colon [301], The authors declare that there is no conflict of interests
lung [302], pancreas [303], cervix [304], glioma [305], and regarding the publication of this paper.
anaplastic large cell lymphoma [306].
It has recently been shown that sCLU knockdown in References
human cancer cells, using siRNA-mediated CLU gene silenc-
ing, induces significant reduction of cellular growth and [1] R. A. Lockshin and C. M. Williams, “Programmed cell death-I.
higher rates of spontaneous apoptosis [294]. Thus, we con- Cytology of degeneration in the intersegmental muscles of the
sider the fact that clusterin, as a chaperon molecule, has been Pernyi silkmoth,” Journal of Insect Physiology, vol. 11, no. 2, pp.
123–133, 1965.
found to bind with the active form of Bax under apoptotic
[2] T. G. Cotter, “Apoptosis and cancer: the genesis of a research
stimuli inhibiting its dimerization in the mitochondria and
field,” Nature Reviews Cancer, vol. 9, no. 7, pp. 501–507, 2009.
block apoptosis. Consequently, it has been hypothesized
[3] J. F. Kerr, A. H. Wyllie, and A. R. Currie, “Apoptosis: a
that sCLU gene silencing using siRNA or other techniques basic biological phenomenon with wide-ranging implications
may ultimately develop into attractive antitumor apoptotic in tissue kinetics,” The British Journal of Cancer, vol. 26, no. 4,
induction [294]. Thus, many researchers targeted sCLU pp. 239–257, 1972.
expression silencing using antisense oligonucleotides (ASO) [4] T. D. Halazonetis, V. G. Gorgoulis, and J. Bartek, “An oncogene-
or RNAi, as promising tools for cancer therapy [307–309]. induced DNA damage model for cancer development,” Science,
However, that was more effective when given in combination vol. 319, no. 5868, pp. 1352–1355, 2008.
with convention therapies to synergize the effect of such [5] S. Negrini, V. G. Gorgoulis, and T. D. Halazonetis, “Genomic
combination therapy [310–312]. The ASO was challenged by instability an evolving hallmark of cancer,” Nature Reviews
the rapid intracellular degeneration of oligonucleotides in Molecular Cell Biology, vol. 11, no. 3, pp. 220–228, 2010.
14 BioMed Research International

[6] S. Fulda, “Evasion of apoptosis as a cellular stress response in [26] K. L. King and J. A. Cidlowski, “Cell cycle regulation and
cancer,” International Journal of Cell Biology, vol. 2010, Article apoptosis,” Annual Review of Physiology, vol. 60, pp. 601–617,
ID 370835, 6 pages, 2010. 1998.
[7] J. Plati, O. Bucur, and R. Khosravi-Far, “Dysregulation of [27] J. F. Kerr and J. Searle, “A mode of cell loss in malignant
apoptotic signaling in cancer: molecular mechanisms and neoplasms,” Journal of Pathology, vol. 106, no. 1, 1972.
therapeutic opportunities,” Journal of Cellular Biochemistry, vol. [28] J. C. Reed, “Proapoptotic multidomain Bcl-2/Bax-family pro-
104, no. 4, pp. 1124–1149, 2008. teins: mechanisms, physiological roles, and therapeutic oppor-
[8] D. Hanahan and R. A. Weinberg, “The hallmarks of cancer,” Cell, tunities,” Cell Death and Differentiation, vol. 13, no. 8, pp. 1378–
vol. 100, no. 1, pp. 57–70, 2000. 1386, 2006.
[9] J. C. Reed, J. M. Jurgensmeier, and S. Matsuyama, “Bcl-2 family [29] A. Letai, “BCL-2: found bound and drugged!,” Trends in
proteins and mitochondria,” Biochimica et Biophysica Acta, vol. Molecular Medicine, vol. 11, no. 10, pp. 442–444, 2005.
1366, no. 1-2, pp. 127–137, 1998. [30] H. Kim, M. Rafiuddin-Shah, H.-C. Tu et al., “Hierarchical
[10] D. R. Green and G. I. Evan, “A matter of life and death,” Cancer regulation of mitochondrion-dependent apoptosis by BCL-2
Cell, vol. 1, no. 1, pp. 19–30, 2002. subfamilies,” Nature Cell Biology, vol. 8, no. 12, pp. 1348–1358,
[11] Y. Ionov, H. Yamamoto, S. Krajewski, J. C. Reed, and M. Peru- 2006.
cho, “Mutational inactivation of the proapoptotic gene BAX [31] M. C. Wei, W.-X. Zong, E. H.-Y. Cheng et al., “Proapoptotic
confers selective advantage during tumor clonal evolution,” BAX and BAK: a requisite gateway to mitochondrial dysfunc-
Proceedings of the National Academy of Sciences of the United tion and death,” Science, vol. 292, no. 5517, pp. 727–730, 2001.
States of America, vol. 97, no. 20, pp. 10872–10877, 2000. [32] W.-X. Zong, T. Lindsten, A. J. Ross, G. R. MacGregor, and C.
[12] T. Miyashita, S. Krajewski, M. Krajewska et al., “Tumor suppres- B. Thompson, “BH3-only proteins that bind pro-survival Bcl-2
sor p53 is a regulator of bcl-2 and bax gene expression in vitro family members fail to induce apoptosis in the absence of Bax
and in vivo,” Oncogene, vol. 9, no. 6, pp. 1799–1805, 1994. and Bak,” Genes and Development, vol. 15, no. 12, pp. 1481–1486,
[13] D. L. Vaux, “Immunopathology of apoptosis—introduction and 2001.
overview,” Springer Seminars in Immunopathology, vol. 19, no. 3, [33] D. R. Green, “At the gates of death,” Cancer Cell, vol. 9, no. 5, pp.
pp. 271–278, 1998. 328–330, 2006.
[14] S. M. Frisch and R. A. Screaton, “Anoikis mechanisms,” Current [34] S. A. Amundson, T. G. Myers, D. Scudiero, S. Kitada, J. C.
Opinion in Cell Biology, vol. 13, no. 5, pp. 555–562, 2001. Reed, and J. Fornace A.J., “An informatics approach identifying
[15] J. Tschopp, F. Martinon, and K. Hofmann, “Apoptosis: silencing markers of chemosensitivity in human cancer cell lines,” Cancer
the death receptors,” Current Biology, vol. 9, no. 10, pp. R381– Research, vol. 60, no. 21, pp. 6101–6110, 2000.
R384, 1999. [35] C. Teixeira, J. C. Reed, and M. A. C. Pratt, “Estrogen promotes
[16] G. Makin and J. A. Hickman, “Apoptosis and cancer chemother- chemotherapeutic drug resistance by a mechanism involving
apy,” Cell and Tissue Research, vol. 301, no. 1, pp. 143–152, 2000. Bcl-2 proto-oncogene expression in human breast cancer cells,”
Cancer Research, vol. 55, no. 17, pp. 3902–3907, 1995.
[17] V. Cryns and J. Yuan, “Proteases to die for,” Genes and Develop-
ment, vol. 12, no. 11, pp. 1551–1570, 1998. [36] S. Elmore, “Apoptosis: a review of programmed cell death,”
Toxicologic Pathology, vol. 35, no. 4, pp. 495–516, 2007.
[18] N. A. Thornberry and Y. Lazebnik, “Caspases: enemies within,”
Science, vol. 281, no. 5381, pp. 1312–1316, 1998. [37] S. Koyama, N. Koike, and S. Adachi, “Fas receptor counterattack
against tumor-infiltrating lymphocytes in vivo as a mechanism
[19] R. M. Locksley, N. Killeen, and M. J. Lenardo, “The TNF and of immune escape in gastric carcinoma,” Journal of Cancer
TNF receptor superfamilies: integrating mammalian biology,” Research and Clinical Oncology, vol. 127, no. 1, pp. 20–26, 2001.
Cell, vol. 104, no. 4, pp. 487–501, 2001.
[38] M. Bagnoli, S. Canevari, and D. Mezzanzanica, “Cellular
[20] D. Wallach, E. E. Varfolomeev, N. L. Malinin, Y. V. Goltsev, A. V. FLICE-inhibitory protein (c-FLIP) signalling: a key regulator
Kovalenko, and M. P. Boldin, “Tumor necrosis factor receptor of receptor-mediated apoptosis in physiologic context and in
and Fas signaling mechanisms,” Annual Review of Immunology, cancer,” International Journal of Biochemistry and Cell Biology,
vol. 17, pp. 331–367, 1999. vol. 42, no. 2, pp. 210–213, 2010.
[21] D. R. Green and G. Kroemer, “The pathophysiology of mito- [39] A. R. Safa, T. W. Day, and C.-H. Wu, “Cellular FLICE-like
chondrial cell death,” Science, vol. 305, no. 5684, pp. 626–629, inhibitory protein (C-FLIP): a novel target for cancer therapy,”
2004. Current Cancer Drug Targets, vol. 8, no. 1, pp. 37–46, 2008.
[22] D. Chauhan, T. Hideshima, S. Rosen, J. C. Reed, S. Kharbanda, [40] T. R. Wilson, K. M. Redmond, K. M. McLaughlin et al.,
and K. C. Anderson, “Apaf-1/Cytochrome c-independent and “Procaspase 8 overexpression in non-small-cell lung cancer
Smac-dependent Induction of Apoptosis in Multiple Myeloma promotes apoptosis induced by FLIP silencing,” Cell Death and
(MM) Cells,” Journal of Biological Chemistry, vol. 276, no. 27, pp. Differentiation, vol. 16, no. 10, pp. 1352–1361, 2009.
24453–24456, 2001.
[41] L.-Y. Chen, T.-H. Chen, P.-Y. Wen et al., “Differential expression
[23] G. Kroemer and J. C. Reed, “Mitochondrial control of cell of NUDT9 at different phases of the menstrual cycle and
death,” Nature Medicine, vol. 6, no. 5, pp. 513–519, 2000. in different components of normal and neoplastic human
[24] B. Motyka, G. Korbutt, M. J. Pinkoski et al., “Mannose 6- endometrium,” Taiwanese Journal of Obstetrics and Gynecology,
phosphate/insulin-like growth factor II receptor is a death vol. 48, no. 2, pp. 96–107, 2009.
receptor for granzyme B during cytotoxic T cell-induced apop- [42] P. Korkolopoulou, A. A. Saetta, G. Levidou et al., “c-FLIP
tosis,” Cell, vol. 103, no. 3, pp. 491–500, 2000. expression in colorectal carcinomas: association with Fas/FasL
[25] P. Salomoni and P. P. Pandolfi, “The role of PML in tumor expression and prognostic implications,” Histopathology, vol. 51,
suppression,” Cell, vol. 108, no. 2, pp. 165–170, 2002. no. 2, pp. 150–156, 2007.
BioMed Research International 15

[43] S. Gao, H. Wang, P. Lee et al., “Androgen receptor and prostate [59] R. M. Schultz, V. F. Patel, J. F. Worzalla, and C. Shih, “Role of
apoptosis response factor-4 target the c-FLIP gene to determine thymidylate synthase in the antitumor activity of the multitar-
survival and apoptosis in the prostate gland,” Journal of Molec- geted antifolate, LY231514,” Anticancer Research, vol. 19, no. 1 A,
ular Endocrinology, vol. 36, no. 3, pp. 463–483, 2006. pp. 437–443, 1999.
[44] X. Zhang, T.-G. Jin, H. Yang, W. C. Dewolf, R. Khosravi-Far, and [60] E. Chu, D. M. Koeller, P. G. Johnston, S. Zinn, and C. J. Allegra,
A. F. Olumi, “Persistent c-FLIP(L) expression is necessary and “Regulation of thymidylate synthase in human colon cancer
sufficient to maintain resistance to tumor necrosis factor-related cells treated with 5-fluorouracil and interferon-𝛾,” Molecular
apoptosis-inducing ligand-mediated apoptosis in prostate can- Pharmacology, vol. 43, no. 4, pp. 527–533, 1993.
cer,” Cancer Research, vol. 64, no. 19, pp. 7086–7091, 2004. [61] E. Chu, D. M. Voeller, K. L. Jones et al., “Identification of
[45] X. Zhang, L. Zhang, H. Yang et al., “c-Fos as a proapoptotic agent a thymidylate synthase ribonucleoprotein complex in human
in TRAIL-induced apoptosis in prostate cancer cells,” Cancer colon cancer cells,” Molecular and Cellular Biology, vol. 14, no. 1,
Research, vol. 67, no. 19, pp. 9425–9434, 2007. pp. 207–213, 1994.
[46] X. Du, G. Bao, X. He et al., “Expression and biological signifi- [62] D. B. Longley, D. P. Harkin, and P. G. Johnston, “5-Fluorouracil:
cance of c-FLIP in human hepatocellular carcinomas,” Journal mechanisms of action and clinical strategies,” Nature Reviews
of Experimental and Clinical Cancer Research, vol. 28, no. 1, Cancer, vol. 3, no. 5, pp. 330–338, 2003.
article 24, 2009. [63] D. B. Longley, P. R. Ferguson, J. Boyer et al., “Characterization
of a thymidylate synthase (TS)-inducible cell line: a model
[47] J. H. Song, M. C. L. Tse, A. Bellail et al., “Lipid rafts and non-
system for studying sensitivity to TS- and non-TS-targeted
rafts mediate tumor necrosis factor-related apoptosis-inducing
chemotherapies,” Clinical Cancer Research, vol. 7, no. 11, pp.
ligand-induced apoptotic and nonapoptotic signals in non-
3533–3539, 2001.
small cell lung carcinoma cells,” Cancer Research, vol. 67, no. 14,
pp. 6946–6955, 2007. [64] D. B. Longley, J. Boyer, W. L. Allen et al., “The role of
thymidylate synthase induction in modulating p53-regulated
[48] V. Baud and M. Karin, “Is NF-𝜅B a good target for cancer gene expression in response to 5-fluorouracil and antifolates,”
therapy? Hopes and pitfalls,” Nature Reviews Drug Discovery, Cancer Research, vol. 62, no. 9, pp. 2644–2649, 2002.
vol. 8, no. 1, pp. 33–40, 2009.
[65] P. G. Johnston, H.-J. Lenz, C. G. Leichman et al., “Thymidylate
[49] M. E. Guicciardi and G. J. Gores, “Life and death by death synthase gene and protein expression correlate and are associ-
receptors,” The FASEB Journal, vol. 23, no. 6, pp. 1625–1637, 2009. ated with response to 5-fluorouracil in human colorectal and
[50] T. S. Jani, J. DeVecchio, T. Mazumdar, A. Agyeman, and J. gastric tumors,” Cancer Research, vol. 55, no. 7, pp. 1407–1412,
A. Houghton, “Inhibition of NF-𝜅B signaling by quinacrine is 1995.
cytotoxic to human colon carcinoma cell lines and is synergistic [66] H.-J. Lenz, K. Hayashi, D. Salonga et al., “p53 point mutations
in combination with tumor necrosis factor-related apoptosis- and thymidylate synthase messenger RNA levels in dissemi-
inducing ligand (TRAIL) or Oxaliplatin,” Journal of Biological nated colorectal cancer: an analysis of response and survival,”
Chemistry, vol. 285, no. 25, pp. 19162–19172, 2010. Clinical Cancer Research, vol. 4, no. 5, pp. 1243–1250, 1998.
[51] X. Saelens, N. Festjens, L. Vande Walle, M. Van Gurp, G. [67] S. Marsh and H. L. McLeod, “Thymidylate synthase pharmaco-
Van Loo, and P. Vandenabeele, “Toxic proteins released from genetics in colorectal cancer,” Clinical Colorectal Cancer, vol. 1,
mitochondria in cell death,” Oncogene, vol. 23, no. 16, pp. 2861– no. 3, pp. 175–179, 2001.
2874, 2004. [68] J. F. Dillman III, L. P. Dabney, and K. K. Pfister, “Cytoplasmic
[52] S. Fulda, I. Jeremias, and K.-M. Debatin, “Cooperation of dynein is associated with slow axonal transport,” Proceedings of
betulinic acid and TRAIL to induce apoptosis in tumor cells,” the National Academy of Sciences of the United States of America,
Oncogene, vol. 23, no. 46, pp. 7611–7620, 2004. vol. 93, no. 1, pp. 141–144, 1996.
[53] J. Ajani, “Review of capecitabine as oral treatment of gastric, [69] M. A. Jordan, R. J. Toso, D. Thrower, and L. Wilson, “Mechanism
gastroesophageal, and esophageal cancers,” Cancer, vol. 107, no. of mitotic block and inhibition of cell proliferation by taxol
2, pp. 221–231, 2006. at low concentrations,” Proceedings of the National Academy of
Sciences of the United States of America, vol. 90, no. 20, pp. 9552–
[54] W. B. Ershler, “Capecitabine use in geriatric oncology: an
9556, 1993.
analysis of current safety, efficacy, and quality of life data,”
Critical Reviews in Oncology/Hematology, vol. 58, no. 1, pp. 68– [70] M. A. Jordan, K. Wendell, S. Gardiner, W. B. Derry, H. Copp,
78, 2006. and L. Wilson, “Mitotic block induced in HeLa cells by low
concentrations of paclitaxel (taxol) results in abnormal mitotic
[55] I. Petak, D. M. Tillman, and J. A. Houghton, “p53 Dependence of exit and apoptotic cell death,” Cancer Research, vol. 56, no. 4, pp.
Fas induction and acute apoptosis in response to 5-fluorouracil- 816–825, 1996.
leucovorin in human colon carcinoma cell lines,” Clinical
[71] D. Panda, H. P. Miller, A. Banerjee, R. F. Ludueña, and L. Wilson,
Cancer Research, vol. 6, no. 11, pp. 4432–4441, 2000.
“Microtubule dynamics in vitro are regulated by the tubulin
[56] H. H. J. Backus, D. Wouters, C. G. Ferreira et al., “Thymidylate isotype composition,” Proceedings of the National Academy of
synthase inhibition triggers apoptosis via caspases-8 and -9 Sciences of the United States of America, vol. 91, no. 24, pp. 11358–
in both wild-type and mutant p53 colon cancer cell lines,” 11362, 1994.
European Journal of Cancer, vol. 39, no. 9, pp. 1310–1317, 2003. [72] S. Ranganathan, C. A. Benetatos, P. J. Colarusso, D. W. Dexter,
[57] F. Bunz, P. M. Hwang, C. Torrance et al., “Disruption of p53 in and G. R. Hudes, “Altered 𝛽-tubulin isotype expression in
human cancer cells alters the responses to therapeutic agents,” paclitaxel-resistant human prostate carcinoma cells,” The British
Journal of Clinical Investigation, vol. 104, no. 3, pp. 263–269, Journal of Cancer, vol. 77, no. 4, pp. 562–566, 1998.
1999. [73] M. Haber, C. A. Burkhart, D. L. Regl, J. Madafiglio, M. D. Norris,
[58] Y. M. Rustum, “Thymidylate synthase: a critical target in cancer and S. B. Horwitz, “Altered expression of M𝛽2, the class II 𝛽-
therapy?” Frontiers in Bioscience, vol. 9, pp. 2467–2473, 2004. tubulin isotype, in a murine J774.2 cell line with a high level of
16 BioMed Research International

taxol resistance,” Journal of Biological Chemistry, vol. 270, no. 52, [90] J. J. Turchi, K. M. Henkels, I. L. Hermanson, and S. M. Patrick,
pp. 31269–31275, 1995. “Interactions of mammalian proteins with cisplatin-damaged
[74] K. Kamath, L. Wilson, F. Cabral, and M. A. Jordan, “𝛽III- DNA,” Journal of Inorganic Biochemistry, vol. 77, no. 1-2, pp. 83–
tubulin induces paclitaxel resistance in association with reduced 87, 1999.
effects on microtubule dynamic instability,” Journal of Biological [91] G. Gebauer, A. T. Peter, D. Onesime, and N. Dhanasekaran,
Chemistry, vol. 280, no. 13, pp. 12902–12907, 2005. “Apoptosis of ovarian granulosa cells: correlation with the
[75] F. Cabral, I. Abraham, and M. M. Gottesman, “Isolation of a reduced activity of ERK-signaling module,” Journal of Cellular
taxol-resistant Chinese hamster ovary cell mutant that has an Biochemistry, vol. 75, no. 4, pp. 547–554, 1999.
alteration in 𝛼-tubulin,” Proceedings of the National Academy of [92] S. H. Kaufmann and D. L. Vaux, “Alterations in the apoptotic
Sciences of the United States of America, vol. 78, no. 7 I, pp. 4388– machinery and their potential role in anticancer drug resis-
4391, 1981. tance,” Oncogene, vol. 22, no. 47, pp. 7414–7430, 2003.
[76] C. H. A. Cheung, S.-Y. Wu, T.-R. Lee et al., “Cancer cells [93] C. A. Schmitt, C. T. Rosenthal, and S. W. Lowe, “Genetic analysis
acquire mitotic drug resistance properties through beta i- of chemoresistance in primary murine lymphomas,” Nature
tubulin mutations and alterations in the expression of beta- Medicine, vol. 6, no. 9, pp. 1029–1035, 2000.
tubulin isotypes,” PLoS ONE, vol. 5, no. 9, Article ID e12564, [94] J. Meiler and M. Schuler, “Therapeutic targeting of apoptotic
pp. 1–11, 2010. pathways in cancer,” Current Drug Targets, vol. 7, no. 10, pp.
[77] S. Yin, R. Bhattacharya, and F. Cabral, “Human mutations that 1361–1369, 2006.
confer paclitaxel resistance,” Molecular Cancer Therapeutics, vol. [95] P. Giménez-Bonafé, A. Tortosa, and R. Pérez-Tomás, “Over-
9, no. 2, pp. 327–335, 2010.
coming drug resistance by enhancing apoptosis of tumor cells,”
[78] M. Kartalou and J. M. Essigmann, “Recognition of cisplatin Current Cancer Drug Targets, vol. 9, no. 3, pp. 320–340, 2009.
adducts by cellular proteins,” Mutation Research—Fundamental
[96] T. R. Wilson, P. G. Johnston, and D. B. Longley, “Anti-apoptotic
and Molecular Mechanisms of Mutagenesis, vol. 478, no. 1-2, pp.
mechanisms of drug resistance in cancer,” Current Cancer Drug
1–21, 2001.
Targets, vol. 9, no. 3, pp. 307–319, 2009.
[79] F. J. Dijt, A. M. J. Fichtinger-Schepman, F. Berends, and J.
[97] M. H. Kang and C. P. Reynolds, “BcI-2 Inhibitors: targeting
Reedijk, “Formation and repair of cisplatin-induced adducts to
mitochondrial apoptotic pathways in cancer therapy,” Clinical
DNA in cultured normal and repair-deficient human fibrob-
Cancer Research, vol. 15, no. 4, pp. 1126–1132, 2009.
lasts,” Cancer Research, vol. 48, no. 21, pp. 6058–6062, 1988.
[98] M. R. Patel, A. Masood, P. S. Patel, and A. A. Chanan-Khan,
[80] R. P. Perez, “Cellular and molecular determinants of cisplatin
“Targeting the Bcl-2,” Current Opinion in Oncology, vol. 21, no.
resistance,” European Journal of Cancer, vol. 34, no. 10, pp. 1535–
6, pp. 516–523, 2009.
1542, 1998.
[81] A. Nehmé, R. Baskaran, S. Aebi et al., “Differential induction of [99] B. Leibowitz and J. Yu, “Mitochondrial signaling in cell death
c-Jun NH2-terminal kinase and c-Abl kinase in DNA mismatch via the Bcl-2 family,” Cancer Biology and Therapy, vol. 9, no. 6,
repair-proficient and -deficient cells exposed to cisplatin,” Can- pp. 417–422, 2010.
cer Research, vol. 57, no. 15, pp. 3253–3257, 1997. [100] M. C. De Almagro and D. Vucic, “The inhibitor of apoptosis
[82] H. Niedner, R. Christen, X. Lin, A. Kondo, and S. B. Howell, (IAP) proteins are critical regulators of signaling pathways and
“Identification of genes that mediate sensitivity to cisplatin,” targets for anti-cancer therapy,” Experimental Oncology, vol. 34,
Molecular Pharmacology, vol. 60, no. 6, pp. 1153–1160, 2001. no. 3, pp. 200–211, 2012.
[83] S. N. Farrow and R. Brown, “New members of the Bcl-2 family [101] T. W. Day and A. R. Safa, “RNA interference in cancer: targeting
and their protein partners,” Current Opinion in Genetics and the anti-apoptotic protein c-FLIP for drug discovery,” Mini-
Development, vol. 6, no. 1, pp. 45–49, 1996. Reviews in Medicinal Chemistry, vol. 9, no. 6, pp. 741–748, 2009.
[84] R. Agarwal and S. B. Kaye, “Ovarian cancer: strategies for over- [102] K. N. Chi, M. E. Gleave, R. Klasa et al., “A phase I dose-
coming resistance to chemotherapy,” Nature Reviews Cancer, finding study of combined treatment with an antisense Bcl-2
vol. 3, no. 7, pp. 502–516, 2003. oligonucleotide (Genasense) and mitoxantrone in patients with
metastatic hormone-refractory prostate cancer,” Clinical Cancer
[85] A. G. Eliopoulos, D. J. Kerr, J. Herod et al., “The control of
Research, vol. 7, no. 12, pp. 3920–3927, 2001.
apoptosis and drug resistance in ovarian cancer: influence of
p53 and Bcl-2,” Oncogene, vol. 11, no. 7, pp. 1217–1228, 1995. [103] J. C. Reed and M. Pellecchia, “Apoptosis-based therapies for
[86] A. Strasser, A. W. Harris, T. Jacks, and S. Cory, “DNA damage hematologic malignancies,” Blood, vol. 106, no. 2, pp. 408–418,
can induce apoptosis in proliferating lymphoid cells via p53- 2005.
independent mechanisms inhibitable by Bcl-2,” Cell, vol. 79, no. [104] D. M. Hockenbery, “Targeting mitochondria for cancer ther-
2, pp. 329–339, 1994. apy,” Environmental and Molecular Mutagenesis, vol. 51, no. 5,
[87] J. J. O. Herod, A. G. Eliopoulos, J. Warwick, G. Niedobitek, L. S. pp. 476–489, 2010.
Young, and D. J. Kerr, “The prognostic significance of Bcl-2 and [105] A. Letai, M. C. Bassik, L. D. Walensky, M. D. Sorcinelli, S. Weiler,
p53 expression in ovarian carcinoma,” Cancer Research, vol. 56, and S. J. Korsmeyer, “Distinct BH3 domains either sensitize or
no. 9, pp. 2178–2184, 1996. activate mitochondrial apoptosis, serving as prototype cancer
[88] H. Miyake, I. Hara, K. Yamanaka, S. Arakawa, and S. Kamidono, therapeutics,” Cancer Cell, vol. 2, no. 3, pp. 183–192, 2002.
“Synergistic enhancement of resistance to cisplatin in human [106] L. D. Walensky, A. L. Kung, I. Escher et al., “Activation of
bladder cancer cells by overexpression of mutant-type p53 and apoptosis in vivo by a hydrocarbon-stapled BH3 helix,” Science,
Bcl-2,” Journal of Urology, vol. 162, no. 6, pp. 2176–2181, 1999. vol. 305, no. 5689, pp. 1466–1470, 2004.
[89] S. Kondo, “Apoptosis by antitumor agents and other factors in [107] T. Oltersdorf, S. W. Elmore, A. R. Shoemaker et al., “An inhibitor
relation to cell cycle checkpoints,” Journal of radiation research, of Bcl-2 family proteins induces regression of solid tumours,”
vol. 36, no. 1, pp. 56–62, 1995. Nature, vol. 435, no. 7042, pp. 677–681, 2005.
BioMed Research International 17

[108] C. Tse, A. R. Shoemaker, J. Adickes et al., “ABT-263: a potent [124] G. P. Dotto, “p21(WAF1/Cip1): more than a break to the cell
and orally bioavailable Bcl-2 family inhibitor,” Cancer Research, cycle?” Biochimica et Biophysica Acta—Reviews on Cancer, vol.
vol. 68, no. 9, pp. 3421–3428, 2008. 1471, no. 1, pp. M43–M56, 2000.
[109] E. Wesarg, S. Hoffarth, R. Wiewrodt et al., “Targeting BCL-2 [125] Q. Zhan, I. T. Chen, M. J. Antinore, and A. J. Fornace Jr., “Tumor
family proteins to overcome drug resistance in non-small cell suppressor p53 can participate in transcriptional induction of
lung cancer,” International Journal of Cancer, vol. 121, no. 11, pp. the GADD45 promoter in the absence of direct DNA binding,”
2387–2394, 2007. Molecular and Cellular Biology, vol. 18, pp. 2768–2778, 1998.
[110] M. Certo, V. D. G. Moore, M. Nishino et al., “Mitochondria [126] M. Schuler and D. R. Green, “Mechanisms of p53-dependent
primed by death signals determine cellular addiction to anti- apoptosis,” Biochemical Society Transactions, vol. 29, no. 6, pp.
apoptotic BCL-2 family members,” Cancer Cell, vol. 9, no. 5, pp. 684–688, 2001.
351–365, 2006. [127] J. Yu, L. Zhang, P. M. Hwang, C. Rago, K. W. Kinzler, and B.
[111] R. Yamaguchi, E. Janssen, G. Perkins, M. Ellisman, S. Kitada, Vogelstein, “Identification and classification of p53-regulated
and J. C. Reed, “Efficient elimination of cancer cells by genes,” Proceedings of the National Academy of Sciences of the
deoxyglucose-ABT-263/737 combination therapy,” PLoS ONE, United States of America, vol. 96, no. 25, pp. 14517–14522, 1999.
vol. 6, no. 9, Article ID e24102, 2011. [128] K. H. Vousden, “p53: death star,” Cell, vol. 103, no. 5, pp. 691–694,
[112] C. L. Hann, V. C. Daniel, E. A. Sugar et al., “Therapeutic efficacy 2000.
of ABT-737, a selective inhibitor of BCL-2, in small cell lung [129] S. Vossio, E. Palescandolo, N. Pediconi et al., “DN-p73 is
cancer,” Cancer Research, vol. 68, no. 7, pp. 2321–2328, 2008. activated after DNA damage in a p53-dependent manner to
[113] J. C. Byrd, S. Kitada, I. W. Flinn et al., “The mechanism of regulate p53-induced cell cycle arrest,” Oncogene, vol. 21, no. 23,
tumor cell clearance by rituximab in vivo in patients with B-cell pp. 3796–3803, 2002.
chronic lymphocytic leukemia: evidence of caspase activation [130] K. Oda, H. Arakawa, T. Tanaka et al., “p53AIP1, a potential
and apoptosis induction,” Blood, vol. 99, no. 3, pp. 1038–1043, mediator of p53-dependent apoptosis, and its regulation by ser-
2002. 46-phosphorylated p53,” Cell, vol. 102, no. 6, pp. 849–862, 2000.
[114] B. Gong, Q. Chen, B. Endlich, S. Mazumder, and A. Almasan, [131] Y. Samuels-Lev, D. J. O'Connor, D. Bergamaschi et al., “ASPP
“Ionizing radiation-induced, bax-mediated cell death is depen- proteins specifically stimulate the apoptotic function of p53,”
dent on activation of cysteine and serine proteases,” Cell Growth Molecular Cell, vol. 8, no. 4, pp. 781–794, 2001.
and Differentiation, vol. 10, no. 7, pp. 491–502, 1999. [132] M. Ljungman, “Dial 9-1-1 for p53: mechanisms of p53 activation
[115] Q. Chen, B. Gong, and A. Almasan, “Distinct stages of by cellular stress,” Neoplasia, vol. 2, no. 3, pp. 208–225, 2000.
cytochrome c release from mitochondria: evidence for a feed- [133] J.-T. Liang, K.-C. Huang, Y.-M. Cheng et al., “P53 Overexpres-
back amplification loop linking caspase activation to mitochon- sion predicts poor chemosensitivity to high-dose 5-fluorouracil
drial dysfunction in genotoxic stress induced apoptosis,” Cell plus leucovorin chemotherapy for stage IV colorectal cancers
Death and Differentiation, vol. 7, no. 2, pp. 227–233, 2000. after palliative bowel resection,” International Journal of Cancer,
[116] B. Vogelstein, D. Lane, and A. J. Levine, “Surfing the p53 vol. 97, no. 4, pp. 451–457, 2002.
network,” Nature, vol. 408, no. 6810, pp. 307–310, 2000. [134] D. J. Ahnen, P. Feigl, G. Quan et al., “Ki-ras mutation and
[117] J. J. Fuster, S. M. Sanz-González, U. M. Moll, and V. Andrés, p53 overexpression predict the clinical behavior of colorectal
“Classic and novel roles of p53: prospects for anticancer ther- cancer: a Southwest Oncology Group study,” Cancer Research,
apy,” Trends in Molecular Medicine, vol. 13, no. 5, pp. 192–199, vol. 58, no. 6, pp. 1149–1158, 1998.
2007. [135] A. Paradiso, G. Simone, S. Petroni et al., “Thymidilate synthase
[118] J. Boyer, E. G. McLean, S. Aroori et al., “Characterization of p53 and p53 primary tumour expression as predictive factors for
wild-type and null isogenic colorectal cancer cell lines resistant advanced colorectal cancer patients,” The British Journal of
to 5-fluorouracil, oxaliplatin and irinotecan,” Clinical Cancer Cancer, vol. 82, no. 3, pp. 560–567, 2000.
Research, vol. 10, no. 6, pp. 2158–2167, 2004. [136] S. Sjögren, M. Inganäs, T. Norberg et al., “The p53 gene in breast
[119] T. Kamijo, J. D. Weber, G. Zambetti, F. Zindy, M. F. Roussel, cancer: prognostic value of complementary DNA sequencing
and C. J. Sherr, “Functional and physical interactions of the versus immunohistochemistry,” Journal of the National Cancer
ARF tumor suppressor with p53 and Mdm2,” Proceedings of the Institute, vol. 88, no. 3-4, pp. 173–182, 1996.
National Academy of Sciences of the United States of America, [137] E. N. Pugacheva, A. V. Ivanov, J. E. Kravchenko, B. P. Kopnin, A.
vol. 95, no. 14, pp. 8292–8297, 1998. J. Levine, and P. M. Chumakov, “Novel gain of function activity
[120] J. Feng, R. Tamaskovic, Z. Yang et al., “Stabilization of Mdm2 via of p53 mutants: activation of the dUTPase gene expression
decreased ubiquitination is mediated by protein kinase B/Akt- leading to resistance to 5-fluorouracil,” Oncogene, vol. 21, no. 30,
dependent phosphorylation,” Journal of Biological Chemistry, pp. 4595–4600, 2002.
vol. 279, no. 34, pp. 35510–35517, 2004. [138] S. Fan, W. S. El-Deiry, I. Bae et al., “p53 Gene mutations are
[121] J. I.-J. Leu, P. Dumont, M. Hafey, M. E. Murphy, and D. L. associated with decreased sensitivity of human lymphoma cells
George, “Mitochondrial p53 activates Bak and causes disruption to DNA damaging agents,” Cancer Research, vol. 54, no. 22, pp.
of a Bak-Mcl1 complex,” Nature Cell Biology, vol. 6, no. 5, pp. 5824–5830, 1994.
443–450, 2004. [139] P. Perego, M. Giarola, S. C. Righetti et al., “Association between
[122] M. Mihara, S. Erster, A. Zaika et al., “p53 has a direct apopto- cisplatin resistance and mutation of p53 gene and reduced bax
genic role at the mitochondria,” Molecular Cell, vol. 11, no. 3, pp. expression in ovarian carcinoma cell systems,” Cancer Research,
577–590, 2003. vol. 56, no. 3, pp. 556–562, 1996.
[123] J. E. Chipuk, T. Kuwana, L. Bouchier-Hayes et al., “Direct [140] W. M. Gallagher, M. Cairney, B. Schott, I. B. Roninson, and
activation of Bax by p53 mediates mitochondrial membrane R. Brown, “Identification of p53 genetic suppressor elements
permeabilization and apoptosis,” Science, vol. 303, no. 5660, pp. which confer resistance to cisplatin,” Oncogene, vol. 14, no. 2,
1010–1014, 2004. pp. 185–193, 1997.
18 BioMed Research International

[141] S. Fan, M. L. Smith, D. J. Rivet II et al., “Disruption of p53 [156] D. Resnitzky and S. I. Reed, “Different roles for cyclins D1 and E
function sensitizes breast cancer MCF-7 cells to cisplatin and in regulation of the G1-to-S transition,” Molecular and Cellular
pentoxifylline,” Cancer Research, vol. 55, no. 8, pp. 1649–1654, Biology, vol. 15, no. 7, pp. 3463–3469, 1995.
1995. [157] S. Mazumder, B. Gong, Q. Chen, J. A. Drazba, J. C. Buchsbaum,
[142] D. S. Hawkins, G. W. Demers, and D. A. Galloway, “Inactiva- and A. Almasan, “Proteolytic cleavage of cyclin E leads to
tion of p53 enhances sensitivity to multiple chemotherapeutic inactivation of associated kinase activity and amplification
agents,” Cancer Research, vol. 56, no. 4, pp. 892–898, 1996. of apoptosis in hematopoietic cells,” Molecular and Cellular
[143] B. A. Goff, J. A. Ries, L. P. Els, M. D. Coltrera, and A. M. Biology, vol. 22, no. 7, pp. 2398–2409, 2002.
Gown, “Immunophenotype of ovarian cancer as predictor of [158] S. Mazumder, E. L. DuPree, and A. Almasan, “A dual role of
clinical outcome: evaluation at primary surgery and second- cyclin E in cell proliferation and apoptosis may provide a target
look procedure,” Gynecologic Oncology, vol. 70, no. 3, pp. 378– for cancer therapy,” Current Cancer Drug Targets, vol. 4, no. 1,
385, 1998. pp. 65–75, 2004.
[144] M. K. Hassan, H. Watari, Y. Han et al., “Clusterin is a poten- [159] S. Mazumder, D. Plesca, M. Kinter, and A. Almasan, “Interac-
tial molecular predictor for ovarian cancer patient's survival: tion of a cyclin E fragment with Ku70 regulates Bax-mediated
targeting Clusterin improves response to paclitaxel,” Journal of apoptosis,” Molecular and Cellular Biology, vol. 27, no. 9, pp.
Experimental and Clinical Cancer Research, vol. 30, no. 1, article 3511–3520, 2007.
113, 2011. [160] C. Subramanian, A. W. Opipari Jr., X. Bian, V. P. Castle, and
[145] S. C. Righetti, G. Della Torre, S. Pilotti et al., “A compara- R. P. S. Kwok, “Ku70 acetylation mediates neuroblastoma cell
tive study of p53 gene mutations, protein accumulation, and death induced by histone deacetylase inhibitors,” Proceedings of
response to cisplatin-based chemotherapy in advanced ovarian the National Academy of Sciences of the United States of America,
carcinoma,” Cancer Research, vol. 56, no. 4, pp. 689–693, 1996. vol. 102, no. 13, pp. 4842–4847, 2005.
[161] J. M. Adams and S. Cory, “The Bcl-2 apoptotic switch in cancer
[146] D. Marx, H. Meden, T. Ziemek, T. Lenthe, W. Kuhn, and A.
development and therapy,” Oncogene, vol. 26, no. 9, pp. 1324–
Schauer, “Expression of the p53 tumour suppressor gene as a
1337, 2007.
prognostic marker in platinum-treated patients with ovarian
cancer,” European Journal of Cancer, vol. 34, no. 6, pp. 845–850, [162] S. A. Lakhani, A. Masud, K. Kuida et al., “Caspases 3 and 7: key
1998. mediators of mitochondrial events of apoptosis,” Science, vol.
311, no. 5762, pp. 847–851, 2006.
[147] D. Mayr, U. Pannekamp, G. B. Baretton et al., “Immunohis-
tochemical analysis of drug resistance-associated proteins in [163] C. Subramanian, A. W. Opipari Jr., V. P. Castle, and R. P. S.
ovarian carcinomas,” Pathology Research and Practice, vol. 196, Kwok, “Histone deacetylase inhibition induces apoptosis in
no. 7, pp. 469–475, 2000. neuroblastoma,” Cell Cycle, vol. 4, no. 12, pp. 1741–1743, 2005.
[148] D. A. Dart, S. M. Picksley, P. A. Cooper, J. A. Double, and M. [164] H. Y. Cohen, S. Lavu, K. J. Bitterman et al., “Acetylation of the
C. Bibby, “The role of p53 in the chemotherapeutic responses C terminus of Ku70 by CBP and PCAF controls Bax-mediated
to cisplatin, doxorubicin and 5-fluorouracil treatment,” Interna- apoptosis,” Molecular Cell, vol. 13, no. 5, pp. 627–638, 2004.
tional Journal of Oncology, vol. 24, no. 1, pp. 115–125, 2004. [165] Y. Ma, H. Lu, K. Schwarz, and M. R. Lieber, “Repair of double-
strand DNA breaks by the human non-homologous DNA end
[149] V. Lam, J. P. McPherson, L. Salmena et al., “p53 gene status and
joining pathway: the iterative processing model,” Cell Cycle, vol.
chemosensitivity of childhood acute lymphoblastic leukemia
4, no. 9, pp. 1193–1200, 2005.
cells to adriamycin,” Leukemia Research, vol. 23, no. 10, pp. 871–
880, 1999. [166] D. Zhai, F. Luciano, X. Zhu, B. Guo, A. C. Satterthwait, and J. C.
Reed, “Humanin binds and nullifies bid activity by blocking its
[150] S. Geisler, P. E. Lønning, T. Aas et al., “Influence of TP53
activation of Bax and Bak,” Journal of Biological Chemistry, vol.
gene alterations and c-erbB-2 expression on the response to
280, no. 16, pp. 15815–15824, 2005.
treatment with doxorubicin in locally advanced breast cancer,”
Cancer Research, vol. 61, no. 6, pp. 2505–2512, 2001. [167] Å. B. Gustafsson, J. G. Tsai, S. E. Logue, M. T. Crow, and R. A.
Gottlieb, “Apoptosis repressor with caspase recruitment domain
[151] I. Fichtner, W. Slisow, J. Gill et al., “Anticancer drug response and protects against cell death by interfering with Bax activation,”
expression of molecular markers in early-passage xenotrans- Journal of Biological Chemistry, vol. 279, no. 20, pp. 21233–21238,
planted colon carcinomas,” European Journal of Cancer, vol. 40, 2004.
no. 2, pp. 298–307, 2004.
[168] Y.-W. Mao, J.-P. Liu, H. Xiang, and D. W.-C. Li, “Human
[152] B. R. Jacob, “Surviving and thriving,” Rehab Management, vol. 𝛼A- and 𝛼B-crystallins bind to Bax and Bcl-Xs to sequester
15, no. 1, pp. 50–51, 2002. their translocation during staurosporine-induced apoptosis,”
[153] V. Pavillard, V. Charasson, A. Laroche-Clary, I. Soubeyran, Cell Death and Differentiation, vol. 11, no. 5, pp. 512–526, 2004.
and J. Robert, “Cellular parameters predictive of the clinical [169] Y. Deng and X. Wu, “Peg3/Pw1 promotes p53-mediated apop-
response of colorectal cancers to irinotecan. A Preliminary tosis by inducing Bax translocation from cytosol to mitochon-
study,” Anticancer Research, vol. 24, no. 2, pp. 579–585, 2004. dria,” Proceedings of the National Academy of Sciences of the
[154] C. Lavarino, S. Pilotti, M. Oggionni et al., “p53 Gene status United States of America, vol. 97, no. 22, pp. 12050–12055, 2000.
and response to platinum/paclitaxel-based chemotherapy in [170] Y. Takahashi, M. Karbowski, H. Yamaguchi et al., “Loss of Bif-1
advanced ovarian carcinoma,” Journal of Clinical Oncology, vol. suppresses Bax/Bak conformational change and mitochondrial
18, no. 23, pp. 3936–3945, 2000. apoptosis,” Molecular and Cellular Biology, vol. 25, no. 21, pp.
[155] M. Ohtsubo, A. M. Theodoras, J. Schumacher, J. M. Roberts, and 9369–9382, 2005.
M. Pagano, “Human cyclin E, a nuclear protein essential for the [171] T. Ohtsuka, H. Ryu, Y. A. Minamishima et al., “ASC is a Bax
G1-to-S phase transition,” Molecular and Cellular Biology, vol. adaptor and regulates the p53-Bax mitochondrial apoptosis
15, no. 5, pp. 2612–2624, 1995. pathway,” Nature Cell Biology, vol. 6, no. 2, pp. 121–128, 2004.
BioMed Research International 19

[172] M. D. Vos, A. Dallol, K. Eckfeld et al., “The RASSF1A tumor [187] Y. Shirota, J. Stoehlmacher, J. Brabender et al., “ERCC1 and
suppressor activates bax via MOAP-1,” Journal of Biological thymidylate synthase mRNA levels predict survival for col-
Chemistry, vol. 281, no. 8, pp. 4557–4563, 2006. orectal cancer patients receiving combination oxaliplatin and
[173] H. H. Park, E. Logette, S. Raunser et al., “Death domain fluorouracil chemotherapy,” Journal of Clinical Oncology, vol. 19,
assem¬bly mechanism revealed by crystal structure of the no. 23, pp. 4298–4304, 2001.
oligomeric PIDDosome core complex,” Cell, vol. 128, no. 3, pp. [188] M. Selvakumaran, D. A. Pisarcik, R. Bao, A. T. Yeung, and T.
533–546, 2007. C. Hamilton, “Enhanced cisplatin cytotoxicity by disturbing the
[174] A. Sekiyama, H. Ueda, S.-I. Kashiwamura et al., “A stress- nucleotide excision repair pathway in ovarian cancer cell lines,”
induced, superoxide-mediated caspase-1 activation pathway Cancer Research, vol. 63, no. 6, pp. 1311–1316, 2003.
causes plasma IL-18 upregulation,” Immunity, vol. 22, no. 6, pp. [189] J. A. Houghton, D. M. Tillman, and F. G. Harwood, “Ratio of 2󸀠 -
669–677, 2005. Deoxyadenosine-5󸀠 -triphosphate/thymidine-5󸀠 -triphosphate
[175] M. Woo, R. Hakem, C. Furlonger et al., “Caspase-3 regulates cell influences the commitment of human colon carcinoma cells to
cycle in B cells: a consequence of substrate specificity,” Nature thymineless death,” Clinical Cancer Research, vol. 1, no. 7, pp.
Immunology, vol. 4, no. 10, pp. 1016–1022, 2003. 723–730, 1995.
[176] D. R. Beisner, I. L. Ch'en, R. V. Kolla, A. Hoffmann, and S. M. [190] G. W. Aherne, A. Hardcastle, F. Raynaud, and A. L. Jack-
Hedrick, “Cutting edge: innate immunity conferred by B cells is man, “Immunoreactive dUMP and TTP pools as an index of
regulated by caspase-8,” Journal of Immunology, vol. 175, no. 6, thymidylate synthase inhibition; effect of tomudex (ZD1694)
pp. 3469–3473, 2005. and a nonpolyglutamated quinazoline antifolate (CB30900) in
L1210 mouse leukaemia cells,” Biochemical Pharmacology, vol.
[177] R. P. Rastogi and R. P. Sinha, “Apoptosis: molecular mechanisms
51, no. 10, pp. 1293–1301, 1996.
and pathogenicity,” EXCLI Journal, vol. 8, pp. 155–181, 2009.
[191] T. Lindahl, “An N glycosidase from Escherichia coli that
[178] J. T. Reardon, A. Vaisman, S. G. Chaney, and A. Sancar,
releases free uracil from DNA containing deaminated cytosine
“Efficient nucleotide excision repair of cisplatin, oxaliplatin, and
residues,” Proceedings of the National Academy of Sciences of the
bis-acetoammine-dichloro-cyclohexylamine-platinum(IV)
United States of America, vol. 71, no. 9, pp. 3649–3653, 1974.
(JM216) platinum intrastrand DNA diadducts,” Cancer
Research, vol. 59, no. 16, pp. 3968–3971, 1999. [192] R. D. Ladner, “The role of dUTPase and uracil-DNA repair in
cancer chemotherapy,” Current Protein and Peptide Science, vol.
[179] T. Furuta, T. Ueda, G. Aune, A. Sarasin, K. H. Kraemer, and
2, no. 4, pp. 361–370, 2001.
Y. Pommier, “Transcription-coupled nucleotide excision repair
as a determinant of cisplatin sensitivity of human cells,” Cancer [193] S. D. Webley, A. Hardcastle, R. D. Ladner, A. L. Jackman,
Research, vol. 62, no. 17, pp. 4899–4902, 2002. and G. W. Aherne, “Deoxyuridine triphosphatase (dUTPase)
expression and sensitivity to the thymidylate synthase (TS)
[180] E. Reed, “Platinum-DNA adduct, nucleotide excision repair and
inhibitor ZD9331,” The British Journal of Cancer, vol. 83, no. 6,
platinum based anti-cancer chemotherapy,” Cancer Treatment
pp. 792–799, 2000.
Reviews, vol. 24, no. 5, pp. 331–344, 1998.
[194] S. D. Webley, S. J. Welsh, A. L. Jackman, and G. W. Aherne, “The
[181] K. B. Lee, R. J. Parker, V. Bohr, T. Cornelison, and E. Reed,
ability to accumulate deoxyuridine triphosphate and cellular
“Cisplatin sensitivity/resistance in UV repair-deficient Chinese
response to thymidylate synthase (TS) inhibition,” The British
hamster ovary cells of complementation groups 1 and 3,”
Journal of Cancer, vol. 85, no. 3, pp. 446–452, 2001.
Carcinogenesis, vol. 14, no. 10, pp. 2177–2180, 1993.
[195] D. Fink, S. Aebi, and S. B. Howell, “The role of DNA mismatch
[182] D. W. Melton, A.-M. Ketchen, F. Núñez et al., “Cells from
repair in drug resistance,” Clinical Cancer Research, vol. 4, no. 1,
ERCC1-deficient mice show increased genome instability and
pp. 1–6, 1998.
a reduced frequency of S-phase-dependent illegitimate chro-
mosome exchange but a normal frequency of homologous [196] K. A. D. Narine, A. M. Keuling, R. Gombos, V. A. Tron, S.
recombination,” Journal of Cell Science, vol. 111, no. 3, pp. 395– E. Andrew, and L. C. Young, “Defining the DNA mismatch
404, 1998. repair-dependent apoptotic pathway in primary cells: evidence
for p53-independence and involvement of centrosomal caspase
[183] C. K. Youn, M. H. Kim, H. J. Cho et al., “Oncogenic H-Ras up-
2,” DNA Repair, vol. 9, no. 2, pp. 161–168, 2010.
regulates expression of ERCC1 to protect cells from platinum-
based anticancer agents,” Cancer Research, vol. 64, pp. 4849– [197] J.-P. Issa, “The epigenetics of colorectal cancer,” Annals of the
4857, 2004. New York Academy of Sciences, vol. 910, pp. 140–155, 2000.
[184] M. Dabholkar, J. Vionnet, F. Bostick-Bruton, J. J. Yu, and [198] B. L. King, M.-L. Carcangiu, D. Carter et al., “Microsatellite
E. Reed, “Messenger RNA levels of XPAC and ERCC1 in instability in ovarian neoplasms,” The British Journal of Cancer,
ovarian cancer tissue correlate with response to platinum-based vol. 72, no. 2, pp. 376–382, 1995.
chemotherapy,” Journal of Clinical Investigation, vol. 94, no. 2, [199] T. G. Paulson, F. A. Wright, B. A. Parker, V. Russack, and
pp. 703–708, 1994. G. M. Wahl, “Microsatellite instability correlates with reduced
[185] R. Metzger, C. G. Leichman, K. D. Danenberg et al., “ERCC1 survival and poor disease prognosis in breast cancer,” Cancer
mRNA levels complement thymidylate synthase mRNA levels Research, vol. 56, no. 17, pp. 4021–4026, 1996.
in predicting response and survival for gastric cancer patients [200] K. K. Herfarth, T. P. Brent, R. P. Danam et al., “A specific CpG
receiving combination cisplatin and fluorouracil chemother- methylation pattern of the MGMT promoter region associated
apy,” Journal of Clinical Oncology, vol. 16, no. 1, pp. 309–316, with reduced MGMT expression in primary colorectal cancers,”
1998. Molecular Carcinogenesis, vol. 24, pp. 90–98, 1999.
[186] R. V. N. Lord, J. Brabender, D. Gandara et al., “Low ERRC1 [201] R. Fishel and R. D. Kolodner, “Identification of mismatch repair
expression correlates with prolonged survival after cisplatin genes and their role in the development of cancer,” Current
plus gemcitabine chemotherapy in non-small cell lung cancer,” Opinion in Genetics and Development, vol. 5, no. 3, pp. 382–395,
Clinical Cancer Research, vol. 8, no. 7, pp. 2286–2291, 2002. 1995.
20 BioMed Research International

[202] J. A. Plumb, G. Strathdee, J. Sludden, S. B. Kaye, and R. Brown, [219] T. Mimori, M. Akizuki, H. Yamagata, S. Inada, S. Yoshida,
“Reversal of drug resistance in human tumor xenografts by and M. Homma, “Characterization of a high molecular weight
2󸀠 -deoxy-5-azacytidine-induced demethylation of the hMLH1 acidic nuclear protein recognized by autoantibodies in sera
gene promoter,” Cancer Research, vol. 60, no. 21, pp. 6039–6044, from patients with polymyositis-scleroderma overlap,” Journal
2000. of Clinical Investigation, vol. 68, no. 3, pp. 611–620, 1981.
[203] Y. Watanabe, M. Koi, H. Hemmi, H. Hoshai, and K. Noda, “A [220] G. C. Li, H. Ouyang, X. Li et al., “Ku70: a candidate tumor
change in microsatellite instability caused by cisplatin-based suppressor gene for murine T cell lymphoma,” Molecular Cell,
chemotherapy of ovarian cancer,” The British Journal of Cancer, vol. 2, no. 1, pp. 1–8, 1998.
vol. 85, no. 7, pp. 1064–1069, 2001. [221] R. Tuteja and N. Tuteja, “Ku autoantigen: a multifunctional
[204] D. Fallik, F. Borrini, V. Boige et al., “Microsatellite instability is a DNA-binding protein,” Critical Reviews in Biochemistry and
predictive factor of the tumor response to irinotecan in patients Molecular Biology, vol. 35, no. 1, pp. 1–33, 2000.
with advanced colorectal cancer,” Cancer Research, vol. 63, no. [222] C. Gullo, M. Au, G. Feng, and G. Teoh, “The biology of Ku and
18, pp. 5738–5744, 2003. its potential oncogenic role in cancer,” Biochimica et Biophysica
[205] E. C. Friedberg, G. C. Walker, W. Siede, R. D. Wood, R. A. Acta—Reviews on Cancer, vol. 1765, no. 2, pp. 223–234, 2006.
Schultz, and T. Ellenberger, DNA Repair and Mutagenesis, ASM [223] M. Sawada, W. Sun, P. Hayes, K. Leskov, D. A. Boothman, and
Press, Washington, DC, USA, 2006. S. Matsuyama, “Ku70 suppresses the apoptotic translocation of
[206] E. C. Friedberg, “DNA damage and repair,” Nature, vol. 421, no. bax to mitochondria,” Nature Cell Biology, vol. 5, no. 4, pp. 320–
6921, pp. 436–440, 2003. 329, 2003.
[207] R. D. Kennedy, J. E. Quinn, P. B. Mullan, P. G. Johnston, and [224] A. Kashishian, H. Douangpanya, D. Clark et al., “DNA-
D. P. Harkin, “The role of BRCA1 in the cellular response to dependent protein kinase inhibitors as drug candidates for the
chemotherapy,” Journal of the National Cancer Institute, vol. 96, treatment of cancer,” Molecular Cancer Therapeutics, vol. 2, no.
no. 22, pp. 1659–1668, 2004. 12, pp. 1257–1264, 2003.
[208] Y. Wang, D. Cortez, P. Yazdi, N. Neff, S. J. Elledge, and J. Qin, [225] V. L. Gabai, A. B. Meriin, D. D. Mosser et al., “Hsp70 prevents
“BASC, a super complex of BRCA1-associated proteins involved activation of stress kinases: a novel pathway of cellular thermo-
in the recognition and repair of aberrant DNA structures,” tolerance,” Journal of Biological Chemistry, vol. 272, no. 29, pp.
Genes and Development, vol. 14, no. 8, pp. 927–939, 2000. 18033–18037, 1997.
[209] R. I. Yarden, S. Pardo-Reoyo, M. Sgagias, K. H. Cowan, and L.
[226] A. Samali and T. G. Cotter, “Heat shock proteins increase
C. Brody, “BRCA1 regulates the G2/M checkpoint by activating
resistance to apoptosis,” Experimental Cell Research, vol. 223, no.
Chk1 kinase upon DNA damage,” Nature Genetics, vol. 30, no.
1, pp. 163–170, 1996.
3, pp. 285–289, 2002.
[227] G. C. Li, F. He, X. Shao et al., “Adenovirus-mediated heat-
[210] P. B. Mullan, J. E. Quinn, P. M. Gilmore et al., “BRCA1 and
activated antisense Ku70 expression radiosensitizes tumor cells
GADD45 mediated G2/M cell cycle arrest in response to
in vitro and in vivo,” Cancer Research, vol. 63, no. 12, pp. 3268–
antimicrotubule agents,” Oncogene, vol. 20, no. 43, pp. 6123–
3274, 2003.
6131, 2001.
[228] Y.-T. Tai, K. Podar, S.-K. Kraeft et al., “Translocation of
[211] D. P. Harkin, “Uncovering functionally relevant signaling path-
Ku86/Ku70 to the multiple myeloma cell membrane: functional
ways using microarray-based expression profiling,” Oncologist,
implications,” Experimental Hematology, vol. 30, no. 3, pp. 212–
vol. 5, no. 6, pp. 501–507, 2000.
220, 2002.
[212] J. E. Quinn, R. D. Kennedy, P. B. Mullan et al., “BRCA1
functions as a differential modulator of chemotherapy-induced [229] I. S. Ayene, L. P. Ford, and C. J. Koch, “Ku protein targeting
apoptosis,” Cancer Research, vol. 63, no. 19, pp. 6221–6228, 2003. by Ku70 small interfering RNA enhances human cancer cell
response to topoisomerase II inhibitor and 𝛾 radiation,” Molec-
[213] G. Guo, F. Zhang, R. Gao, R. Delsite, Z. Feng, and S. N. Powell,
ular Cancer Therapeutics, vol. 4, no. 4, pp. 529–536, 2005.
“DNA repair and synthetic lethality,” International Journal of
Oral Science, vol. 3, no. 4, pp. 176–179, 2011. [230] C.-S. Chen, Y.-C. Wang, H.-C. Yang et al., “Histone deacetylase
inhibitors sensitize prostate cancer cells to agents that produce
[214] S. Thode, A. Schafer, P. Pfeiffer, and W. Vielmetter, “A novel
DNA double-strand breaks by targeting Ku70 acetylation,”
pathway of DNA end-to-end joining,” Cell, vol. 60, no. 6, pp.
Cancer Research, vol. 67, no. 11, pp. 5318–5327, 2007.
921–928, 1990.
[215] P. Baumann and S. C. West, “DNA end-joining catalyzed by [231] A. Munshi, J. F. Kurland, T. Nishikawa et al., “Histone deacety-
human cell-free extracts,” Proceedings of the National Academy lase inhibitors radiosensitize human melanoma cells by sup-
of Sciences of the United States of America, vol. 95, no. 24, pp. pressing DNA repair activity,” Clinical Cancer Research, vol. 11,
14066–14070, 1998. no. 13, pp. 4912–4922, 2005.
[216] M. J. Difilippantonio, J. Zhu, H. T. Chen et al., “DNA repair pro- [232] M. K. Hassan, H. Watari, A. E. Salah-Eldin et al., “Histone
tein Ku80 suppresses chromosomal aberrations and malignant deacetylase inhibitors sensitize lung cancer cells to hyperther-
transformation,” Nature, vol. 404, no. 6777, pp. 510–514, 2000. mia: involvement of Ku70/SirT-1 in thermo-protection,” PLoS
ONE, vol. 9, no. 4, Article ID e94213, 2014.
[217] R. B. Cary, S. R. Peterson, J. Wang, D. G. Bear, E. M. Bradbury,
and D. J. Chen, “DNA looping by Ku and the DNA-dependent [233] D. R. Green and J. C. Reed, “Mitochondria and apoptosis,”
protein kinase,” Proceedings of the National Academy of Sciences Science, vol. 281, no. 5381, pp. 1309–1312, 1998.
of the United States of America, vol. 94, no. 9, pp. 4267–4272, [234] A. Ashkenazi, “Targeting the extrinsic apoptosis pathway in
1997. cancer,” Cytokine and Growth Factor Reviews, vol. 19, no. 3-4,
[218] D. A. Ramsden and M. Geliert, “Ku protein stimulates DNA pp. 325–331, 2008.
end joining by mammalian DNA ligases: a direct role for Ku [235] Y. Ito, P. Pandey, A. Place et al., “The novel triterpenoid 2-
in repair of DNA double-strand breaks,” EMBO Journal, vol. 17, cyano-3,12-dioxoolean-1,9-dien-28-oic acid induces apoptosis
no. 2, pp. 609–614, 1998. of human myeloid leukemia cells by a caspase-8-dependent
BioMed Research International 21

mechanism,” Cell Growth and Differentiation, vol. 11, no. 5, pp. Barrett's metaplasia, survival, and response to neoadjuvant
261–267, 2000. chemoradiotherapy,” Annals of Surgery, vol. 239, no. 4, pp. 491–
[236] K. B. Kim, R. Lotan, P. Yue et al., “Identification of a novel 500, 2004.
synthetic triterpenoid, methyl-2-cyano-3,12-dioxooleana-1,9- [253] A. Arlt, A. Gehrz, S. Müerköster et al., “Role of NF-𝜅B and
dien-28-oate, that potently induces caspase-mediated apoptosis Akt/PI3K in the resistance of pancreatic carcinoma cell lines
in human lung cancer cells,” Molecular Cancer Therapeutics, vol. against gemcitabine-induced cell death,” Oncogene, vol. 22, no.
1, no. 3, pp. 177–184, 2002. 21, pp. 3243–3251, 2003.
[237] I. M. Pedersen, S. Kitada, A. Schimmer et al., “The triterpenoid [254] T. Kato, D. C. Duffey, F. G. Ondrey et al., “Cisplatin and radiation
CDDO induces apoptosis in refractory CLL B cells,” Blood, vol. sensitivity in human head and neck squamous carcinomas
100, no. 8, pp. 2965–2972, 2002. are independently modulated by glutathione and transcription
[238] L. Cartee, R. Smith, Y. Dai et al., “Synergistic induction of factor NF-kappaB,” Head & Neck, vol. 22, pp. 748–759, 2000.
apoptosis in human myeloid leukemia cells by phorbol 12- [255] J. C. Cusack Jr., “Rationale for the treatment of solid tumors
myristate 13-acetate and flavopiridol proceeds via activation of with the proteasome inhibitor bortezomib,” Cancer Treatment
both the intrinsic and tumor necrosis factor-mediated extrinsic Reviews, vol. 29, no. 1, pp. 21–31, 2003.
cell death pathways,” Molecular Pharmacology, vol. 61, no. 6, pp. [256] P. Richardson, “Clinical update: proteasome inhibitors in
1313–1321, 2002. hematologic malignancies,” Cancer Treatment Reviews, vol. 29,
[239] L. Altucci and H. Gronemeyer, “Nuclear receptors in cell life and supplement 1, pp. 33–39, 2003.
death,” Trends in Endocrinology and Metabolism, vol. 12, no. 10, [257] E. C. LaCasse, D. J. Mahoney, H. H. Cheung, S. Plenchette, S.
pp. 460–468, 2001. Baird, and R. G. Korneluk, “IAP-targeted therapies for cancer,”
[240] F. A. E. Kruyt, “TRAIL and cancer therapy,” Cancer Letters, vol. Oncogene, vol. 27, no. 48, pp. 6252–6275, 2008.
263, no. 1, pp. 14–25, 2008. [258] S. Fulda, “Targeting inhibitor of apoptosis proteins (IAPs) for
[241] D. Mahalingam, E. Szegezdi, M. Keane, S. D. Jong, and A. cancer therapy,” Anti-Cancer Agents in Medicinal Chemistry, vol.
Samali, “TRAIL receptor signalling and modulation: are we on 8, no. 5, pp. 533–539, 2008.
the right TRAIL?” Cancer Treatment Reviews, vol. 35, no. 3, pp. [259] S. Fulda, “Tumor resistance to apoptosis,” International Journal
280–288, 2009. of Cancer, vol. 124, no. 3, pp. 511–515, 2009.
[242] G. Brumatti, M. Salmanidis, and P. G. Ekert, “Crossing paths: [260] R. Mannhold, S. Fulda, and E. Carosati, “IAP antagonists:
interactions between the cell death machinery and growth promising candidates for cancer therapy,” Drug Discovery
factor survival signals,” Cellular and Molecular Life Sciences, vol. Today, vol. 15, no. 5-6, pp. 210–219, 2010.
67, no. 10, pp. 1619–1630, 2010. [261] K. A. Whitehead, R. Langer, and D. G. Anderson, “Knocking
[243] V. Duronio, “The life of a cell: apoptosis regulation by the down barriers: advances in siRNA delivery,” Nature Reviews
PI3K/PKB pathway,” Biochemical Journal, vol. 415, no. 3, pp. Drug Discovery, vol. 8, no. 2, pp. 129–138, 2009.
333–344, 2008. [262] M. E. Davis, “The first targeted delivery of siRNA in humans via
[244] T. Gallenne, F. Gautier, L. Oliver et al., “Bax activation by a self-assembling, cyclodextrin polymer-based nanoparticle:
the BH3-only protein Puma promotes cell dependence on from concept to clinic,” Molecular Pharmaceutics, vol. 6, no. 3,
antiapoptotic Bcl-2 family members,” Journal of Cell Biology, pp. 659–668, 2009.
vol. 185, no. 2, pp. 279–290, 2009. [263] M. Baker, “RNA interference: homing in on delivery,” Nature,
[245] A. Letai, “Puma strikes Bax,” Journal of Cell Biology, vol. 185, no. vol. 464, no. 7292, pp. 1225–1228, 2010.
2, pp. 189–191, 2009. [264] M. E. Davis, J. E. Zuckerman, C. H. J. Choi et al., “Evidence
[246] M. Karin, Y. Cao, F. R. Greten, and Z.-W. Li, “NF-𝜅B in cancer: of RNAi in humans from systemically administered siRNA via
from innocent bystander to major culprit,” Nature Reviews targeted nanoparticles,” Nature, vol. 464, no. 7291, pp. 1067–
Cancer, vol. 2, no. 4, pp. 301–310, 2002. 1070, 2010.
[247] A. S. Baldwin, “Control of oncogenesis and cancer therapy [265] B. M. Ryan, N. O’Donovan, and M. J. Duffy, “Survivin: a
resistance by the transcription factor NF-𝜅B,” Journal of Clinical newtarget for anti-cancer therapy,” Cancer Treatment Reviews,
Investigation, vol. 107, no. 3, pp. 241–246, 2001. vol. 35, pp. 553–562, 2009.
[248] B. Leber, J. Lin, and D. W. Andrews, “Embedded together: the [266] D. Picard, “Heat-shock protein 90, a chaperone for folding and
life and death consequences of interaction of the Bcl-2 family regulation,” Cellular and Molecular Life Sciences, vol. 59, no. 10,
with membranes,” Apoptosis, vol. 12, no. 5, pp. 897–911, 2007. pp. 1640–1648, 2002.
[249] S. Fulda, “Inhibitor of apoptosis proteins in hematological [267] L. Whitesell and S. L. Lindquist, “HSP90 and the chaperoning of
malignancies,” Leukemia, vol. 23, no. 3, pp. 467–476, 2009. cancer,” Nature Reviews Cancer, vol. 5, no. 10, pp. 761–772, 2005.
[250] M. P. Boldin, I. L. Mett, E. E. Varfolomeev et al., “Self- [268] L. Neckers and K. Neckers, “Heat-shock protein 90 inhibitors
association of the 'death domains'of the p55 tumor necrosis as novel cancer chemotherapeutic agents,” Expert Opinion on
factor (TNF) receptor and Fas/APO1 prompts signaling for TNF Emerging Drugs, vol. 7, no. 2, pp. 277–288, 2002.
and Fas/APO1 effects,” Journal of Biological Chemistry, vol. 270, [269] L. Neckers, “Heat shock protein 90 is a rational molecular target
no. 1, pp. 387–391, 1995. in breast cancer,” Breast Disease, vol. 15, pp. 53–60, 2002.
[251] S.-E. Chuang, P.-Y. Yeh, Y.-S. Lu et al., “Basal levels and patterns [270] L. Neckers, “Hsp90 inhibitors as novel cancer chemotherapeutic
of anticancer drug-induced activation of nuclear factor-𝜅B (NF- agents,” Trends in Molecular Medicine, vol. 8, no. 4, supplement,
𝜅B), and its attenuation by tamoxifen, dexamethasone, and pp. S55–S61, 2002.
curcumin in carcinoma cells,” Biochemical Pharmacology, vol. [271] N. Sato, T. Yamamoto, Y. Sekine et al., “Involvement of heat-
63, no. 9, pp. 1709–1716, 2002. shock protein 90 in the interleukin-6-mediated signaling path-
[252] M. M. M. Abdel-Latif, J. O'Riordan, H. J. Windle et al., “NF- way through STAT3,” Biochemical and Biophysical Research
𝜅B activation in esophageal adenocarcinoma: relationship to Communications, vol. 300, no. 4, pp. 847–852, 2003.
22 BioMed Research International

[272] C. S. Mitsiades, N. S. Mitsiades, C. J. McMullan et al., radioresistance,” Radiation Research, vol. 165, no. 4, pp. 410–423,
“Antimyeloma activity of heat shock protein-90 inhibition,” 2006.
Blood, vol. 107, no. 3, pp. 1092–1100, 2006. [286] Y.-C. Wu, W.-Y. Yen, T.-C. Lee, and L.-H. Yih, “Heat shock
[273] M. Chatterjee, S. Jain, T. Stühmer et al., “STAT3 and MAPK protein inhibitors, 17-DMAG and KNK437, enhance arsenic
signaling maintain overexpression of heat shock proteins 90𝛼 trioxide-induced mitotic apoptosis,” Toxicology and Applied
and 𝛽 in multiple myeloma cells, which critically contribute to Pharmacology, vol. 236, no. 2, pp. 231–238, 2009.
tumor-cell survival,” Blood, vol. 109, no. 2, pp. 720–728, 2007. [287] R. Enmon, W.-H. Yang, Å. M. Ballangrud et al., “Combination
[274] T. W. Schulte, M. V. Blagosklonny, C. Ingui, and L. Neckers, treatment with 17-N-allylamino-17-demethoxy geldanamycin
“Disruption of the Raf-1-Hsp90 molecular complex results and acute irradiation produces supra-additive growth suppres-
in destabilization of Raf-1 and loss of Raf-1-Ras association,” sion in human prostate carcinoma spheroids,” Cancer Research,
Journal of Biological Chemistry, vol. 270, no. 41, pp. 24585– vol. 63, no. 23, pp. 8393–8399, 2003.
24588, 1995. [288] L. R. Kelland, “Small molecule anticancer drugs,” IDrugs, vol. 2,
[275] S. Sato, N. Fujita, and T. Tsuruo, “Modulation of Akt kinase no. 6, pp. 550–552, 1999.
activity by binding to Hsp90,” Proceedings of the National [289] J. L. Eiseman, J. Lan, T. F. Lagattuta et al., “Pharmacokinetics
Academy of Sciences of the United States of America, vol. 97, no. and pharmacodynamics of 17-demethoxy 17-[[(2-
20, pp. 10832–10837, 2000. dimethylamino)ethyl]amino]geldanamycin (17DMAG,
[276] E. E. A. Bull, H. Dote, K. J. Brady et al., “Enhanced NSC 707545) in C.B-17 SCID mice bearing MDA-MB-231
tumor cell radiosensitivity and abrogation of G2 and S phase human breast cancer xenografts,” Cancer Chemotherapy and
arrest by the Hsp90 inhibitor 17-(dimethylaminoethylamino)- Pharmacology, vol. 55, no. 1, pp. 21–32, 2005.
17-demethoxygeldanamycin,” Clinical Cancer Research, vol. 10, [290] P. A. Brough, W. Aherne, X. Barril et al., “4,5-Diarylisoxazole
no. 23, pp. 8077–8084, 2004. Hsp90 chaperone inhibitors: potential therapeutic agents for the
treatment of cancer,” Journal of Medicinal Chemistry, vol. 51, no.
[277] K. F. Pirollo, Z. Hao, A. Rait, C. W. Ho, and E. H. Chang,
2, pp. 196–218, 2008.
“Evidence supporting a signal transduction pathway leading
to the radiation-resistant phenotype in human tumor cells,” [291] S. A. Eccles, A. Massey, F. I. Raynaud et al., “NVP-AUY922: a
Biochemical and Biophysical Research Communications, vol. 230, novel heat shock protein 90 inhibitor active against xenograft
no. 1, pp. 196–201, 1997. tumor growth, angiogenesis, and metastasis,” Cancer Research,
vol. 68, no. 8, pp. 2850–2860, 2008.
[278] A. K. Gupta, V. J. Bakanauskas, G. J. Cerniglia et al., “The ras
radiation resistance pathway,” Cancer Research, vol. 61, no. 10, [292] P. A. Brough, X. Barril, J. Borgognoni et al., “Combin-
pp. 4278–4282, 2001. ing hit identification strategies: fragment-based and in silico
approaches to orally active 2-aminothieno[2,3-d]pyrimidine
[279] S. Tanno, N. Yanagawa, A. Habiro et al., “Serine/threonine inhibitors of the Hsp90 molecular chaperone,” Journal of Medic-
kinase AKT is frequently activated in human bile duct can- inal Chemistry, vol. 52, no. 15, pp. 4794–4809, 2009.
cer and is associated with increased radioresistance,” Cancer
[293] L. Stingl, T. Stühmer, M. Chatterjee, M. R. Jensen, M. Flentje,
Research, vol. 64, no. 10, pp. 3486–3490, 2004.
and C. S. Djuzenova, “Novel HSP90 inhibitors, NVP-AUY922
[280] K. S. Bisht, C. M. Bradbury, D. Mattson et al., “Geldanamycin and NVP-BEP800, radiosensitise tumour cells through cell-
and 17-allylamino-17-demethoxygeldanamycin potentiate the in cycle impairment, increased DNA damage and repair protrac-
vitro and in vivo radiation response of cervical tumor cells via tion,” The British Journal of Cancer, vol. 102, no. 11, pp. 1578–1591,
the heat shock protein 90-mediated intracellular signaling and 2010.
cytotoxicity,” Cancer Research, vol. 63, no. 24, pp. 8984–8995,
[294] I. P. Trougakos, A. So, B. Jansen, M. E. Gleave, and E. S.
2003.
Gonos, “Silencing expression of the clusterin/apolipoprotein
[281] H. Machida, Y. Matsumoto, M. Shirai, and N. Kubota, “Gel- j gene in human cancer cells using small interfering RNA
danamycin, an inhibitor of Hsp90, sensitizes human tumour induces spontaneous apoptosis, reduced growth ability, and cell
cells to radiation,” International Journal of Radiation Biology, sensitization to genotoxic and oxidative stress,” Cancer Research,
vol. 79, no. 12, pp. 973–980, 2003. vol. 64, no. 5, pp. 1834–1842, 2004.
[282] J. S. Russell, W. Burgan, K. A. Oswald, K. Camphausen, and [295] B. Shannan, M. Seifert, K. Leskov et al., “Challenge and promise:
P. J. Tofilon, “Enhanced cell killing induced by the combina- roles for clusterin in pathogenesis, progression and therapy of
tion of radiation and the heat shock protein 90 inhibitor 17- cancer,” Cell Death and Differentiation, vol. 13, no. 1, pp. 12–19,
allylamino-17-demethoxygeldanamycin: a multitarget approach 2006.
to radiosensitization,” Clinical Cancer Research, vol. 9, no. 10 I, [296] H. Miyake, M. Muramaki, T. Kurahashi et al., “Expression of
pp. 3749–3755, 2003. clusterin in prostate cancer correlates with gleason score but not
[283] K. Harashima, T. Akimoto, T. Nonaka, K. Tsuzuki, N. Mit- with prognosis in patients undergoing radical prostatectomy
suhashi, and T. Nakano, “Heat shock protein 90 (Hsp90) without neoadjuvant hormonal therapy,” Urology, vol. 68, no.
chaperone complex inhibitor, Radicicol, potentiated radiation- 3, pp. 609–614, 2006.
induced cell killing in a hormone-sensitive prostate cancer cell [297] J. Steinberg, R. Oyasu, S. Lang et al., “Intracellular levels of SGP-
line through degradation of the androgen receptor,” Interna- 2 (Clusterin) correlate with tumor grade in prostate cancer,”
tional Journal of Radiation Biology, vol. 81, no. 1, pp. 63–76, 2005. Clinical Cancer Research, vol. 3, no. 10, pp. 1707–1711, 1997.
[284] H. Dote, W. E. Burgan, K. Camphausen, and P. J. Tofilon, [298] K. Parczyk, C. Pilarsky, U. Rachel, and C. Koch-Brandt, “Gp80
“Inhibition of Hsp90 compromises the DNA damage response (clusterin; TRPM-2) mRNA level is enhanced in human renal
to radiation,” Cancer Research, vol. 66, no. 18, pp. 9211–9220, clear cell carcinomas,” Journal of Cancer Research and Clinical
2006. Oncology, vol. 120, no. 3, pp. 186–188, 1994.
[285] A. E. Kabakov, Y. V. Malyutina, and D. S. Latchman, “Hsf1- [299] M. Redondo, E. Villar, J. Torres-Munoz, T. Tellez, M. Morell,
mediated stress response can transiently enhance cellular and C. K. Petito, “Overexpression of clusterin in human breast
BioMed Research International 23

carcinoma,” The American Journal of Pathology, vol. 157, no. 2, Trastuzumab,” Journal of Cellular Physiology, vol. 204, no. 2, pp.
pp. 393–399, 2000. 463–469, 2005.
[300] D. Xie, H. L. Sze, J. S. T. Sham et al., “Up-regulated expression [315] M. John, A. Hinke, M. Stauch et al., “Weekly paclitaxel plus
of cytoplasmic clusterin in human ovarian carcinoma,” Cancer, trastuzumab in metastatic breast cancer pretreated with anth-
vol. 103, no. 2, pp. 277–283, 2005. racyclines—a phase II multipractice study,” BMC Cancer, vol.
[301] S. Pucci, E. Bonanno, F. Pichiorri, C. Angeloni, and L. G. 12, article 165, 2012.
Spagnoli, “Modulation of different clusterin isoforms in human
colon tumorigenesis,” Oncogene, vol. 23, no. 13, pp. 2298–2304,
2004.
[302] L. V. July, E. Beraldi, A. So et al., “Nucleotide-based ther-
apies targeting clusterin chemosensitize human lung adeno-
carcinoma cells both in vitro and in vivo,” Molecular Cancer
Therapeutics, vol. 3, no. 3, pp. 223–232, 2004.
[303] N. Mourra, A. Couvelard, E. Tiret, S. Olschwang, and
J.-F. Flejou, “Clusterin is highly expressed in pancreatic
endocrine tumours but not in solid pseudopapillary tumours,”
Histopathology, vol. 50, no. 3, pp. 331–337, 2007.
[304] H. Watari, Y. Ohta, M. K. Hassan, Y. Xiong, S. Tanaka, and
N. Sakuragi, “Clusterin expression predicts survival of invasive
cervical cancer patients treated with radical hysterectomy and
systematic lymphadenectomy,” Gynecologic Oncology, vol. 108,
pp. 527–532, 2008.
[305] M. Danik, J.-G. Chabot, C. Mercier et al., “Human gliomas
and epileptic foci express high levels of a mRNA related to rat
testicular sulfated glycoprotein 2, a purported marker of cell
death,” Proceedings of the National Academy of Sciences of the
United States of America, vol. 88, no. 19, pp. 8577–8581, 1991.
[306] A. Wellmann, C. Thieblemont, S. Pittaluga et al., “Detection
of differentially expressed genes in lymphomas using cDNA
arrays: identification of clusterin as a new diagnostic marker for
anaplastic large-cell lymphomas,” Blood, vol. 96, no. 2, pp. 398–
404, 2000.
[307] M. Gleave and H. Miyake, “Use of antisense oligonucleotides
targeting the cytoprotective gene, clusterin, to enhance
androgen- and chemo-sensitivity in prostate cancer,” World
Journal of Urology, vol. 23, no. 1, pp. 38–46, 2005.
[308] M. E. Gleave and B. P. Monia, “Antisense therapy for cancer,”
Nature Reviews Cancer, vol. 5, no. 6, pp. 468–479, 2005.
[309] H. Miyake, I. Hara, S. Kamidono, and M. E. Gleave, “Syn-
ergistic chemsensitization and inhibition of tumor growth
and metastasis by the antisense oligodeoxynucleotide targeting
clusterin gene in a human bladder cancer model,” Clinical
Cancer Research, vol. 7, no. 12, pp. 4245–4252, 2001.
[310] I. P. Trougakos and E. S. Gonos, “Clusterin/Apolipoprotein J in
human aging and cancer,” International Journal of Biochemistry
and Cell Biology, vol. 34, no. 11, pp. 1430–1448, 2002.
[311] M. Gleave and B. Jansen, “Clusterin and IGFBPs as antisense
targets in prostate cancer,” Annals of the New York Academy of
Sciences, vol. 1002, pp. 95–104, 2003.
[312] T. Criswell, M. Beman, S. Araki et al., “Delayed activation of
insulin-like growth factor-1 receptor/Src/ MAPK/Egr-1 signal-
ing regulates clusterin expression, a pro-survival factor,” Journal
of Biological Chemistry, vol. 280, no. 14, pp. 14212–14221, 2005.
[313] T. Zellweger, H. Miyake, S. Cooper et al., “Antitumor activity
of antisense clusterin oligonucleotides is improved in vitro and
in vivo by incorporation of 2󸀠-O-(2-methoxy) ethyl chemistry,”
Journal of Pharmacology and Experimental Therapeutics, vol.
298, no. 3, pp. 934–940, 2001.
[314] A. Biroccio, C. D'Angelo, B. Jansen, M. E. Gleave, and G.
Zupi, “Antisense clusterin oligodeoxynucleotides increase the
response of HER-2 gene amplified breast cancer cells to
International Journal of

Peptides

Advances in
BioMed
Research International
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014
Stem Cells
International
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014
Virolog y
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014
International Journal of
Genomics
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014

Journal of
Nucleic Acids

Zoology
 International Journal of

Hindawi Publishing Corporation Hindawi Publishing Corporation


http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

Submit your manuscripts at


http://www.hindawi.com

Journal of The Scientific


Signal Transduction
Hindawi Publishing Corporation
World Journal
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

Genetics Anatomy International Journal of Biochemistry Advances in


Research International
Hindawi Publishing Corporation
Research International
Hindawi Publishing Corporation
Microbiology
Hindawi Publishing Corporation
Research International
Hindawi Publishing Corporation
Bioinformatics
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

Enzyme International Journal of Molecular Biology Journal of


Archaea
Hindawi Publishing Corporation
Research
Hindawi Publishing Corporation
Evolutionary Biology
Hindawi Publishing Corporation
International
Hindawi Publishing Corporation
Marine Biology
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014 http://www.hindawi.com Volume 2014

Das könnte Ihnen auch gefallen