Oxidative Stress and Cancer-An Overview

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Ageing Research Reviews 12 (2013) 376–390

Contents lists available at SciVerse ScienceDirect

Ageing Research Reviews


journal homepage: www.elsevier.com/locate/arr

Review

Oxidative stress and cancer: An overview


Venus Sosa a , Teresa Moliné a , Rosa Somoza a , Rosanna Paciucci b , Hiroshi Kondoh c ,
Matilde E. LLeonart a,∗
a
Oncology and Molecular Pathology Group, Pathology Department, Institut de Recerca Hospital Vall d’Hebron, Passeig Vall d’Hebron 119-129, 08035 Barcelona, Spain
b
Unitat de Recerca Biomèdica, Institut de Recerca Hospital Vall d’Hebron, Passeig Vall d’Hebron 119-129, 08035 Barcelona, Spain
c
Geriatric Medicine Department, Graduate School of Medicine, Kyoto University, 54 Kawahara-cho, Schogoin, Sakyo-ku, Kyoto 606-8507, Japan

a r t i c l e i n f o a b s t r a c t

Article history: Reactive species, which mainly include reactive oxygen species (ROS), are products generated as a
Received 23 May 2012 consequence of metabolic reactions in the mitochondria of eukaryotic cells. In normal cells, low-level
Received in revised form 16 October 2012 concentrations of these compounds are required for signal transduction before their elimination. How-
Accepted 16 October 2012
ever, cancer cells, which exhibit an accelerated metabolism, demand high ROS concentrations to maintain
Available online 31 October 2012
their high proliferation rate. Different ways of developing ROS resistance include the execution of alter-
native pathways, which can avoid large amounts of ROS accumulation without compromising the energy
Keywords:
demand required by cancer cells. Examples of these processes include the guidance of the glycolytic path-
Cancer
Oxidative stress
way into the pentose phosphate pathway (PPP) and/or the generation of lactate instead of employing
Glycolysis aerobic respiration in the mitochondria. Importantly, ROS levels can be used as a thermostat to monitor
Cancer stem cells the damage that cells can bear. The implications for ROS regulation are highly significant for cancer ther-
Therapy apy because commonly used radio- and chemotherapeutic drugs influence tumor outcome through ROS
modulation. Moreover, the discovery of novel biomarkers that are able to predict the clinical response
to pro-oxidant therapies is a crucial challenge to overcome to allow for the personalization of cancer
therapies.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction pollutants, smoking and alcohol), and exercise. Reactive species


can be classified into four groups based on the main atom
Aerobic respiration generates energy in the mitochondria of involved: ROS, reactive nitrogen species (RNS), reactive sulfur
eukaryotic cells, and as a result of this oxidative metabolism, several species (RSS) and reactive chloride species (RCS) (Bannister, 2007).
compounds are produced. Most of these compounds are benefi- Of all the compounds derived from oxidative metabolism, ROS
cial; however, less than 5% of them can be toxic for the cell if are the most abundantly produced. Their half-lives range from
their concentration increases. These normally low-concentration a few nanoseconds to hours, depending on the stability of the
compounds that are derived from oxidative metabolism are nec- molecule. ROS include superoxide anion (O2 − ), hydrogen perox-
essary for certain subcellular events, including signal transduction, ide (H2 O2 ), hydroxyl radical (OH− ), singlet oxygen (1 O2 ) and ozone
enzyme activation, gene expression, disulfide bond formation dur- (O3 ) (Simic et al., 1989). ROS and RNS are produced during intracel-
ing the folding of new proteins in the endoplasmic reticulum, and lular metabolic processes, such as the electron transport chain. The
the control of the caspase activity that is activated during the apop- most abundant RNS is nitric oxide (NO− ), which is able to react with
totic mechanism. certain ROS, including the peroxynitrite anion and ONOO− , which
Sources of internal oxidative stress include peroxisomes and is produced by the interaction between the superoxide anion and
enzymes, particularly the detoxifying enzymes from the P450 com- nitric oxide; nitric oxide is later converted into peroxynitrous acid
plex, xanthine oxidase, and the nicotinamide adenine dinucleotide and ultimately into a hydroxyl radical and nitrite anion (NO2 − ).
(NADPH) oxidase complexes, which include the Nox family. Most The damage that these ROS can cause to the cell not only
of these enzymes act in the mitochondria, which is the main depends on their intracellular concentration but also on the equi-
source of oxidative stress. External sources of oxidative stress librium between the ROS and the endogenous antioxidant species.
include UV radiation, chemical compounds (e.g., environmental When the pro-oxidant/anti-oxidant equilibrium is lost, oxidative
stress is generated, altering and damaging many intracellular
molecules, including DNA, RNA, lipids and proteins (Veskoukis
∗ Corresponding author. Tel.: +34 93489 4169; fax: +34 93274 6708. et al., 2012). These reactive species cause nicks in the DNA and mal-
E-mail addresses: matilde.lleonart@vhir.org, melleona@ir.vhebron.net functions in the DNA repair mechanism. DNA oxidation by these
(M.E. LLeonart). reactive species generates 8-hydroxy-2 -deoxyguanosine, which is

1568-1637/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.arr.2012.10.004
V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390 377

a product that is able to generate mutations in DNA in a process that example, p38␣ acts as a key sensor of oxidative stress, and its redox-
enhances aging and carcinogenesis (Matsui et al., 2000). Moreover, sensing function is essential in the control of tumor development
the cell membrane is rich in polyunsaturated lipids that are suscep- (Luo et al., 2011). In contrast to other MAPKs, p38␣ suppresses
tible to oxidation by reactive species. Reactive species liberate lipid tumorigenesis by blocking proliferation or promoting apoptosis.
peroxidation reactions and consequently increase the permeability This review is focused on the relationship between oxidative
of the cell membrane, which could lead to cell death (Halliwell and stress and cancer from both the molecular and clinical point of
Chirico, 1993). Proteins are the most affected by a cellular environ- views.
ment with a high concentration of reactive species. Proteins suffer
from the generation and accumulation of carbonyl groups (i.e., alde-
hydes and ketones) and thiol groups (–SH) that may be converted 2. Metastasis-related processes that influence oxidative
into sulfur reactive radicals (Levine, 2002). Due to this oxidation- stress
induced modification, there is an alteration in the protein structure
and, consequently, changes or loss of protein function. 2.1. Epithelial-mesenchymal transition (EMT)
Natural antioxidants are the cell’s defense mechanisms that
scavenge reactive species, and they can be classified into differ- An important event that leads to metastasis is EMT, which is a
ent groups according to their properties: endogenous antioxidants, biological process by which epithelial cells undergo biological and
natural antioxidants and synthetic antioxidants. Endogenous chemical alterations that permit the development of a more aggres-
antioxidants include glutathione, alpha-lipoic acid, coenzyme Q, sive mesenchymal phenotype (Mani et al., 2008). These changes
ferritin, uric acid, bilirubin, metallothionein, l-carnitine, melatonin, are associated with an increase in ECM proteins, which provide
enzymatic superoxide dismutase (SOD), catalase (CAT), glutathione the cells with migratory properties and allow them to move to
peroxidases (GPXs), thioredoxins (TRX) and peroxiredoxins (PRXs). other regions of the body via the bloodstream. There are vari-
PRXs are a ubiquitous family of antioxidant enzymes (PRX I-VI) that ous ROS-associated signaling pathways that are implicated in the
also control cytokine-induced peroxide levels and mediate signal EMT process. Among these pathways, the proteins Smad (activated
transduction in mammalian cells. For example, PRX III scavenges by tumor growth factor ␤ (TGF-␤)), Snail, E-cadherin, ␤-catenin,
up to 90% of H2 O2 , and PRX V behaves more effectively as a scav- integrin, matrix metalloproteinases (MMPs), hepatocyte growth
enger of peroxynitrite. Natural antioxidants coexist in a delicate factor receptor (HGFR/c-Met), AP1 (activated by PKC activator),
balance with oxidative inputs. Other antioxidants can be obtained v-ets erythroblastosis virus E26 oncogene homolog 1 (Ets-1) and
from the diet, such as ascorbic acid (Vitamin C), tocopherol (Vita- transforming growth factor ␤-activated kinase 1 (TAK1) are mostly
min E), ␤-carotene (Vitamin A), lipoic acid, uric acid, glutathione activated in a ROS-dependent manner (Haorah et al., 2007; Ni et al.,
and polyphenol metabolites. Examples of synthetic antioxidants 2007; Omori et al., 2012). One of the first studies that established
include N-acetyl cysteine (NAC), tiron, pyruvate, selenium, buty- a direct connection between ROS and EMT is related to TGF-␤
lated hydroxytoluene, butylated hydroxyanisole, and propyl gallate signaling. TGF-␤ activation provokes an increase in intracellular
(Yoshida et al., 2003). ROS in a process that is associated with the phosphorylation of
Oxidative stress is important from a biomedical point of view Smad2, p38␣ and ERK1/2 (Rhyu et al., 2005). Moreover, intra-
because it is related to a wide variety of human diseases, such cellular ROS may regulate EMT through a mechanism involving
as neurodegenerative disease (e.g., Alzheimer’s, Parkinson’s, and NF-␬B in strict collaboration with hypoxia-inducible factor 1 (HIF-
amyotrophic lateral sclerosis), inflammatory disease (e.g., rheuma- 1␣) and cyclooxygenase-2 (COX-2). The role of ROS as a crucial
toid arthritis), cardiovascular disease (e.g., muscular dystrophy), EMT mediator is further corroborated by MMP-3 (matrix metal-
allergies, immune system dysfunctions, diabetes, aging and can- loproteinase 3), which is also known as Stromelysin-1. MMP-3 is
cer. For example, inflammatory cells release chemical mediators of an enzyme that is involved in the breakdown of the extracellular
inflammation, particularly ROS, in swollen tissue, which also affects matrix, which plays a key role in tumor metastases by degrading
normal cells. When this is a chronic process, the extremely high collagen, fibronectin, and laminin. MMP-3 has been reported to be
ROS levels saturate the cell defense mechanisms (i.e., antioxidants), upregulated in certain tumors, such as breast cancer, and is an EMT
and intracellular molecules become seriously damaged, affecting inducer in transgenic mice. In mice, MMP-3 secretion is associ-
surrounding neighboring cells. ated with Snail upregulation, E-cadherin loss and ␤-catenin nuclear
The mechanisms and pathways involved in oxidative stress are translocation events, which, in turn, are dependent on the small
conserved in mammalian cells. ROS can promote many aspects of GTP-binding protein Rac 1 (Rac1b) (Radisky et al., 2005; Rhyu et al.,
tumor development and progression, which can be classified into 2005). In addition to MMP-3, other metalloproteinases, including
the following biological processes: (a) cellular proliferation (e.g., MMP-2 and MMP-9, play important roles in Rac1b stimulation
extracellular-regulated kinase 1/2 (ERK1/2) activation and ligand- to influence ROS. Most of these proteins are oncogenic in mouse
independent RTK activation), (b) evasion of apoptosis or anoikis models, and their overexpression contributes to human tumorige-
(e.g., Src, NF-␬B and phosphatidylinositol-3 kinase (PI3K)/Akt acti- nesis.
vation), (c) tissue invasion and metastasis (e.g., metalloproteinase Other cellular pathways involved in metastasis are also targeted
(MMP) secretion into the extracellular matrix (ECM), Met overex- by ROS, such as Wnt/TCF (T cell factor), integrin-mediated MAPK
pression, and Rho-Rac interaction), and (d) angiogenesis (e.g., the signaling, protein kinase C (PKC), protein tyrosine phosphatases,
release of vascular endothelial growth factor (VEGF) and angiopoi- and p21-activated kinase 1 (PAK-1) signaling pathways (Almeida
etin). et al., 2007; Lee and Esselman, 2002). For example, ROS oxidize cys-
Regarding cellular proliferation, oxidative stress affects sev- teine residues in the PKC protein and inactivate protein tyrosine
eral biochemical pathways (from epidermal growth factor receptor phosphatases, which are important molecules in tumor cell inva-
(EGFR) to mTOR) that involve key signaling proteins, such as nuclear sion. ROS also regulate PAK-1, which is involved in Rac-associated
factor erythroid 2-related factor 2 (Nrf2), kelch-like protein 19 cytoskeleton remodeling, which is directly linked to metastasis and
(Keap1), Ras, Raf, mitogen activated protein kinases (MAPK) such as angiogenesis (Diebold et al., 2010). Proteins such as Snail, TGF-
ERK1/2, MEK, p38␣, c-Jun N-terminal kinase (JNK), c-myc, p53 and ␤, Twist, Wnt/␤-catenin and Zinc finger E-box-binding homeobox
PKC (Matsuzawa and Ichijo, 2008; Nguyen et al., 2009; Wiemer, 1 (Zeb), which induce EMT, also modulate the tumor microen-
2011). Among them, Nrf2 is considered to be the master regula- vironment. Rac1b enhances the expression of Zinc finger protein
tor of the antioxidant response, but others are also important. For SNAI1 (Slug), a transcriptional inhibitor of E-cadherin that can ease
378 V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390

tumorigenesis by activating the disheveled-3-mediated Wnt/␤- cells adapt to hypoxic conditions, which occur in the core of most
catenin signaling pathway (Esufali et al., 2007). Aberrant activation solid tumors, by upregulating HIF-1␣ and other pro-angiogenic
of the Wnt/␤-catenin pathway promotes tumorigenesis in several genes (e.g., VEGF and erythropoietin) and transcription factors that
cancer types and mainly has effects in breast and colorectal cancers are involved in cell survival, invasion, and glycolysis (Liao and
(Jeong et al., 2012). Johnson, 2007).
In particular, cancer stem cells (CSCs) are rare cells that have Lastly, glioblastoma CSCs express HIF-2␣, while HIF-1␣ is
many characteristics in common with stem cells (SCs), including expressed under more severe hypoxic conditions (Soeda et al.,
an indefinite potential for self-renewal and differentiation; thus, 2009). These two proteins have non-redundant biological roles
they have been proposed as the cells that drive tumorigenesis (Reya because they regulate unique target genes. Importantly, hypoxia
et al., 2001). Evidence for EMT has been described for CSCs (Chen induces the expression of genes related to SC functions, such as
et al., 2011). This link between SC/CSC markers and the EMT phe- Sox2 and Oct-4, which define a stemness signature (McCord et al.,
nomenon has implications in cancer therapy and cancer remission 2009). The relevance of this issue at the therapeutic level comes
as a whole. The relationship between CSCs and ROS will be dis- from the fact that CSCs possess a higher resistance to hypoxia com-
cussed elsewhere. pared to other cancer cells, which is fact that may explain their
enhanced radio- and chemoresistance.
2.2. Angiogenesis

Angiogenesis, which is the process of new blood vessel forma- 3. The relationship between oncogenes and tumor
tion from pre-existing vessels, is an important biological process suppressor genes and oxidative stress
for tumor growth and metastasis (Ushio-Fukai and Nakamura,
2008; Folkman, 1995). Tumors produce many pro-angiogenic fac- Genetic factors play key roles in the transforming events that
tors, such as vascular endothelial growth factor (VEGF) and its lead to cancer development. The high metabolism of cancer cells is
receptor (VEGFR), MMPs, angiopoietin-1, fibroblast growth factor generally associated with an increase in ROS; however, such levels
(FGF), interleukin-8 (IL-8), platelet-derived growth factor (PDGF) are less deleterious in cancer cells than they would be in normal
and TGF-␤. Specifically, VEGF, which is highly upregulated in most cells. For example, although the ROS level increases by a modest
human cancers, has recently emerged as the crucial rate-limiting degree, tumorigenic cells can induce a new redox balance, resulting
protein in the regulation of angiogenesis. Hypoxic conditions, ROS in cellular adaptation and proliferation. This is a striking feature of
production, growth factors, and cytokines all increase VEGF levels cancer cells that allows for the acquisition of resistance to oxidative
in many human tumors (Fei et al., 2009). Anti-angiogenic therapies stress inducers relative to normal cells. The resistance to oxidative
have especially focused on using antibodies and tyrosine kinase stress is one of the major adaptive advantages that permit cancer
inhibitors to block VEGF and VEGFR2 (Carmeliet and Jain, 2000; cells to increase their metabolic rate and proliferation and to escape
Folkman, 1995). free radical damage. However, this adaptive response to high doses
Other ROS-generating enzymes, such as NADPH oxidases (e.g., of ROS is not enough to explain the high metabolic rate of cancer
Nox: Nox1-5), activate redox signaling pathways that ultimately cells.
lead to angiogenesis (Arnold et al., 2001; Tojo et al., 2005). For Genetic alterations in genes that affect cancer cells are able to
example, Nox1 is overexpressed in colon and prostate cancers, and directly or indirectly modulate ROS. Apart from the protection pro-
it induces the production of H2 O2 , which augments VEGF, VEGFR vided by specific antioxidant enzymes, such as SOD, CAT, GPXs,
and MMP levels (Lim et al., 2005). Nox4 and Nox5 are increased TRXs and PRXs, the master regulator of the antioxidant response
in melanoma and prostate cancers, respectively, and act in a simi- is the transcription factor Nrf2. Nrf2 modulates the expression
lar manner as Nox1 (Govindarajan et al., 2007). ROS derived from of hundreds of genes, including not only the familiar antioxidant
these oxidases result in VEGFR2 autophosphorylation, which pro- enzymes but also a large number of genes that control several
motes the induction of transcription factors and genes involved processes, including immune and inflammatory responses, tissue
in angiogenesis. These reports underscore the important role for remodeling and fibrosis, carcinogenesis, and metastasis (Hybertson
ROS in the pathophysiological states involved in neovasculariza- et al., 2011). ROS levels are tightly controlled and predominantly
tion, tumor development and progression. regulated by Nrf2 and its repressor protein Keap1. Moreover, when
treated with oxidation-inducing drugs, mice that lack Nrf2 develop
2.3. Hypoxia more severe intestinal inflammation than controls, with increased
aberrant crypts, suggesting a function for Nrf2 in the prevention
Hypoxia is a pathological condition in which cells are deprived of of inflammation and carcinogenesis (Khor et al., 2006). Initially, it
an adequate oxygen supply, despite proper blood perfusion in a tis- was thought that Nrf2 was capable of regulating oxidative stress
sue. Cancer cells have high proliferation rates, which often generate levels by modulating the production of Antioxidant Response Ele-
a more hypoxic environment if new blood vessels are not rapidly ment (ARE)-regulated antioxidant enzymes. However, alternative
generated. To avoid hypoxic conditions, cancer cells must upregu- mechanisms for Nrf2 activation were further described, and they
late pro-angiogenic genes because the stimulation of angiogenesis are dependent upon kinase pathways, such as MAPK, PI3K, and
creates a new blood supply. Under normoxia, HIF-1␣ and HIF-2␣ other pathways (Yu et al., 1999; Kang et al., 2002). Importantly,
proteins are ubiquitinated through an oxygen-dependent interac- somatic mutations that disrupt the Nrf2-Keap1 interaction were
tion with the von Hippel-Lindau protein (pVLH) and degraded by recently identified in cancer patients (Singh et al., 2006; Kim et al.,
the 26S proteasome. However, under hypoxic conditions, HIF-1␣ 2010b). In breast cancer, the tumor suppressor gene breast can-
and HIF-2␣ proteins are not ubiquitinated, leading to their accu- cer gene 1 (BRCA1) is mutated in 40–50% of hereditary breast
mulation, translocation into the nucleus, dimerization with HIF-1␤ cancers and absent or expressed at low levels in 30–40% of spo-
and transactivation of target genes. Moreover, in cancer cells, Nox- radic breast cancers (Rosen et al., 2003). BRCA1 is a caretaker gene
derived ROS enzymes are involved in the induction of the HIF-1␣ that is responsible for repairing DNA, and it is able to upregulate
subunit by activating the PI3K/Akt/p706K and the MEK/ERK path- several genes involved in the antioxidant response by controlling
ways (Shi et al., 2005; Skinner et al., 2004). The MAPK signaling the activity of the transcription factors Nrf2 and Nf␬B (Bae et al.,
downstream of Ras has been shown to lead to HIF-1␣ phosphory- 2004; Benezra et al., 2003). In addition, Nrf2 induces cytoprotective
lation via a process that stimulates its translational activity. Cancer enzymes, such as GST, GPx and oxidoreductases, which help cells
V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390 379

to combat oxidative stress (Banning et al., 2005; Kohle and Bock, comes from findings that hTERT is localized in the mitochondria and
2007). Apart from the inhibitory action of BRCA1 on ROS genera- the ability of hTERT inhibitors to induce mitochondrial-dependent
tion, BRCA1 also reduces the levels of protein nitration due to RNS apoptosis in target cells (Indran et al., 2010).
accumulation in cells, and it enhances DNA repair processes that From a genetic point of view, many genes are implicated in the
ultimately help to cope with oxidative stress. In breast cancer cells, modulation of energy metabolism and cancer. For example, p53,
the redox factor 1/AP endonuclease 1 (Ref1/APE1) also participates one of the most frequently mutated genes in human cancers, mod-
in reducing the generation of ROS (Seo and Kinsella, 2009). ulates the balance between glycolysis and the utilization of the
One of the most important pathways related to oxidative stress respiratory chain in the mitochondria. A major mediator of this
and cancer is the Ras pathway (Yagoda et al., 2007). In humans, effect is cytochrome oxidase deficient homolog 2 (SCO2), whose
the Ras gene family consists of three small G proteins, Ha-, N- expression is critical for regulating the COX complex, the main
and Ki-ras, which participate in cell signaling (Malumbres and site of oxygen utilization in eukaryotic cells (Matoba et al., 2006).
Barbacid, 2003; Yagoda et al., 2007). Important for its relationship Reduced SCO2 synthesis may cause low respiration and high gly-
with ROS is the fact that the main mechanism of Ras activation colytic rates.
in tumors involves point mutations (i.e., at specific amino acids Sirtuins are a class of proteins that possess histone deacety-
that result in changes at positions 12, 13 and 61). Approximately lase activity and are implicated in aging, transcriptional regulation,
30% of human tumors contain activating mutations in the Ras fam- apoptosis, stress resistance, energy efficiency, and alertness during
ily of oncogenes, rendering a protein constitutively active (Bos, low-calorie situations. Sirtuins are NAD-dependent deacetylases
1989). Several studies have demonstrated that mutant Ras leads whose enzymatic activity is regulated by the ratio of NAD to NADH.
to an increase in ROS levels, which causes DNA damage and con- High NAD levels activate sirtuins, while high NADH levels inhibit
tributes to transformation (Maciag et al., 2004; Rai et al., 2011). them. Of the three mitochondrial sirtuins, Sirt3 is the most studied,
Mutant RasVal12 positively regulates the NADPH oxidase system and it has been linked to longevity in humans, acting as a tumor
Nox4-p22phox , which produces H2 O2 , in a process accompanied suppressor protein (Kim et al., 2010a). Kim and colleagues demon-
by ␥-H2A.X (foci of heterochromatin). Accordingly, knockdown strated that with the expression of a single oncogene, such as Myc
of NADPH oxidase decreases the RasVal12 -induced DNA damage or Ras in Sirt3 (−/−) murine embryonic fibroblasts (MEFs), the cells
response. Therefore, the NADPH oxidase Nox4 is a critical medi- exhibited increased glycolysis, decreased oxidative phosphoryla-
ator of oncogenic RasVal12 -induced DNA damage (Weyemi et al., tion, and increased ROS. Moreover, the absence of Sirt3 increases
2012). Cells that overexpress oncogenic Ras display increased mito- tumorigenesis in cancer cells in a ROS-dependent manner (Bell
chondrial mass and ROS accumulation. The ROS generated by the et al., 2011). This process is accompanied by HIF-1␣ target gene
respiratory chain in the mitochondria and by the Nox enzymes in activation under hypoxic conditions.
the cytoplasm are particularly important. In fact, Nox proteins are To identify new genes implicated in tumorigenesis, our group
now considered to be oncogenic proteins, and mitochondrial dys- has performed functional genome-wide screens using the retro-
function is associated with tumorigenesis (Graham et al., 2010). viral delivery of cDNA libraries. By these means, we found that
Mitochondrial dysfunction is also accompanied by DNA damage, a non-glycosylated membrane associated protein 17 (MAP17) was
drop in ATP levels, and protein kinase AMPK activation. The consti- associated with tumor malignancy due to its capacity to promote
tutive activation of mutant K-rasVal12 in non-transformed epithelial ROS production. In fact, we demonstrated that antioxidant treat-
cells leads to a highly significant increase in the levels of peroxides ment is useful in decreasing the tumorigenic properties that these
and an increase in the amount of DNA strand breakage compared cells acquired (Guijarro et al., 2007). Another gene identified was
with control cells. Peroxides may also be directly generated by COX- phosphoglycerate mutase 1 (PGM), which codifies for an enzyme
2 because COX-2 activity correlates with K-ras activity (Wang et al., involved in glycolysis and was independently found to be related to
2009). COX-2, one of the rate-limiting enzymes in the metabolism ROS (Kondoh et al., 2005; Guijarro et al., 2007). The action of PGM
of arachidonic acid, is reported to be involved in the pathogenesis was also found to decrease ROS levels by significantly stimulating
of many human tumors. Both peroxide generation and DNA single- the immortalization of murine cells, thereby increasing glycoly-
strand breaks are significantly reduced by pre-treatment with the sis (Kondoh et al., 2005). Furthermore, inherited defects in DNA
COX-2 specific inhibitor SC58125. Therefore, the dominant onco- repair genes correlate with increased oxidative stress and prema-
genicity of mutant K-rasVal12 can be influenced by several genes, ture aging (Barzilai et al., 2002). This is the case for ATM deficiency
such as COX-2 and/or HIF-1␣, which is a transcription factor that in patients with ataxia telangiectasia, which is a disease related to
is activated in response to low oxygen concentrations, and its ulti- premature aging and cancer. Unraveling the status of oncogenes
mate influence is on DNA damage. and tumor suppressor genes related to ROS modulation in cancer
Moreover, the overexpression of oncogenic proteins (i.e., Raf, patients promises to be beneficial in the development of successful
Mos, MEK, Myc, Cyclin E and Telomerase reverse transcriptase therapies.
(hTERT)) and the silencing of tumor suppressor genes (i.e., p53,
p21CIP1 and phosphatase and tensin homolog (Pten)) can induce
senescence by increasing ROS. Pten loss and Ras/MAPK activation 4. Environmental factors
cooperate to promote EMT and metastasis that are initiated from
prostate progenitor cells (Mulholland et al., 2012). Pten-deficient In the last few decades, our view of malignancy has broad-
glioblastoma cells, with high levels of Akt and ROS, undergo senes- ened considerably in relation to the contribution of environmental
cence, and such high ROS levels are indispensable for the observed factors and the tumor microenvironment to cancer development.
phenotype (Lee et al., 2011). While the conventional oncogenic role Hormones, such as estrogens, are important in the development
of hTERT has been linked to its ability to provide immortalization and progression of human breast cancer because they can be
in human cells by acting through the telomeres, evidence is accu- transformed into the carcinogenic metabolite 4-hydroxyestradiol
mulating that supports the contribution of non-canonical hTERT (4-OHE2 and 2-OHE2 ) by cytochrome P450 1B1 (Shao et al., 2008).
activity in cancer (Indran et al., 2010). To that end, hTERT has been Both metabolites can be further oxidized into a reactive quinone
implicated in redox-mediated events, and its expression has been that can react with DNA. The compounds 4-OHE2 and 2-OHE2
shown to impact the cellular redox status via the recruitment of and their quinone derivatives are able to generate oxidative stress
mitochondria, which are a critical intracellular source of ROS. Fur- (Cavalieri et al., 2006). Hormone-dependent tumors (e.g., estro-
ther evidence supporting the role of mitochondria in hTERT biology gen, androgen and progesterone) affect cell proliferation through
380 V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390

Fig. 1. The tumor microenvironment influences angiogenesis and tumor progression. ROS produced in cancer cells are released into the tumor microenvironment and
detected by adjacent fibroblasts (CAFs), initiating the onset of stromal oxidative stress, autophagy and mitophagy due to the activation of key transcription factors (i.e., HIF-
1␣) and signaling proteins (i.e., VEGF) that will ultimately contribute to angiogenesis. Moreover, CAFs also release MMPs and cytokines (i.e., IL-6, IL-10, TGF-␤, CCL2 and CCL5),
which stimulate tumor growth and block the natural immune response against cancer. Additionally, senescent cells can release cytokines into the tumor microenvironment,
which stimulate tumor growth. Immune surveillance can prevent and/or promote the growth and progression of cancer. For example, within the immune system, cytotoxic
CD8+ and CD4+ T cells, along with their characteristically produced cytokine IFN-␥, function as the major anti-tumor immune effector cells (−), whereas cancer-associated
macrophages (CAMs) and their derived cytokines IL-6, TNF, IL-1␤ and IL-23 are generally recognized as dominant tumor-promoting forces (+).

their ligand-receptor interactions, and their expression is crucial contribute to tumor growth and affect cancer survival by (i) con-
for determining a specific therapy (Allred et al., 2012). The actions tributing to the vasculature by secreting VEGF and angiopoietin,
exerted by estrogen and its catabolites are essential processes in (ii) generating anti-apoptotic factors, (iii) promoting cell motil-
breast carcinogenesis. In its physiological concentration, estradiol ity and metastasis by secreting chemokines (e.g., CCL2, CCL5) and
is able to decrease the CAT enzyme and increase the GPX activities MMPs, and (iv) blocking the immune response through the secre-
in breast epithelial cells. Thioredoxin and thioredoxin reductase can tion of IL-6, IL-10 and TGF-␤ (Kayamori et al., 2010; Tsuyada et al.,
alter hydrogen peroxide levels and estrogen activity in estrogen- 2012) (Fig. 1). The importance of CAFs in mediating paracrine
responsive breast cancer cells (Hayashi et al., 1997). Oxidative signaling comes from the fact that conditioned media from stromal
stress has been shown to participate in the structural modifica- fibroblasts that are deficient for the Caveolin (Cav-1) gene, a nega-
tion of the estrogen and progesterone receptors, altering the clinical tive cytokine receptor signaling regulator, promote EMT, inducing
evolution of patients with endocrine-responsive breast cancer (Yau a mesenchymal appearance in human breast cancer-associated
and Benz, 2008). The cysteine residues of the zinc finger transcrip- fibroblasts (Sotgia et al., 2009). Other cells in the tumor microen-
tion factor Sp1 and the estrogen receptor (ESR1) are highly prone vironment can act independently or synergistically with CAFs,
to being oxidized by ROS, losing their structural and functional including CAMs and/or senescent fibroblasts. It is striking that pro-
properties, including their DNA-binding activity. For example, this teins produced by senescent cells can have opposite effects if they
failure has been observed in 30–35% of ER-positive breast cancers, act as autocrine or paracrine factors (Fig. 1). Moreover, senescent
and it was found to be related to a loss of the progesterone receptor. cells secrete pro-inflammatory cytokines and proteases that induce
Zinc finger transcription factors are necessary for the estrogenic cancer cells to promote tumor aggressiveness. In this context, can-
stimulation of certain genes, such as the progesterone receptor cer cells become “parasitic”: the ROS produced in cancer cells are
(PGR) and Bcl-2, and estrogen stimulation is inhibited by oxidative released into the tumor microenvironment and are detected by
stress. This fact would represent a problem for therapeutic inter- adjacent fibroblasts, initiating the onset of stromal oxidative stress,
vention because hormone-dependent tumors have the advantage autophagy and mitophagy due to the activation of key transcription
of being treated by an analog molecule that blocks ligand-receptor factors, such as HIF-1␣ and NF-␬B, which contribute to angiogen-
signaling, which downregulates the proliferative cascade. esis (Toullec et al., 2010). Furthermore, cancer cells may use the
Regarding the tumor microenvironment, it is important to nutrients from autophagic stromal fibroblasts to fuel the oxida-
consider that cancer cells are surrounded by other cell types tive mitochondrial metabolism to generate large amounts of ATP
that facilitate tumor growth. Tumor microenvironments consist of and protect themselves from apoptosis. At the same time, cancer
stromal cells (e.g., myofibroblasts and/or cancer-associated fibro- cells may establish a strong antioxidant defense by overexpressing
blasts (CAFs)), vascular cells (e.g., erythrocytes) and immune cells proteins, such as PRXs, or by undergoing TP-53-induced glycol-
(e.g., lymphocytes, natural killer cells, antigen-presenting cells and ysis (TIGAR). In this model, cancer cells released into the tumor
cancer-associated macrophages (CAMs)). CAFs have been shown to microenvironment induce the death of healthy cells and ultimately
V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390 381

Fig. 2. Glucose metabolism has different destinies. Under aerobic conditions, glucose will produce pyruvate, which is used by the mitochondria for respiration and ATP
production. Under anaerobic conditions, glucose will travel into the PPP and/or produce lactate, but its degradative products will not enter into the mitochondria. Glucose
metabolism through the PPP produces a reduced form of NADPH, which maintains glutathione in the reduced state. Under physiological conditions, the principal source
of ROS is mitochondria. During this process, superoxide is generated and immediately converted into H2 O2 by manganese superoxide dismutase (MnSOD), an important
antioxidant enzyme in mitochondria that protects against oxidative stress. H2 O2 must be rapidly eliminated by GPX, PRX or CAT. Glycolysis through the PPP pathway is the
standard response to a lack of oxygen in the cellular microenvironment.

benefit from the discarded metabolites of these cells, which are genetic background (oncogenes and tumor suppressor genes) and
captured by the cancer cells and used to maintain their own prolif- the tumor microenvironment, independent of O2 levels.
eration at a high rate. Dr. Otto Warburg discovered that cancer cells display a high gly-
colytic rate, even under high concentrations of O2 (20%), a finding
5. Glycolysis versus oxidative phosphorylation: Cancer later named the Warburg effect in his honor (Warburg, 1956; Kim
bioenergetics and Dang, 2006). The molecular mechanism behind this effect still
needs to be completely elucidated. Importantly, the Warburg effect
Glycolysis is the initial metabolic pathway of cellular respi- is exploited in the clinic to detect primary or metastatic tumors in
ration, and it consists of a series of reactions that result in the patients. The positron-emission tomography (PET) scanning tech-
conversion of glucose into pyruvic acid with the concomitant pro- nique is used to detect highly glycolytic areas in the body mass
duction of energy (Bar-Even et al., 2012). After glycolysis, there by measuring 2-[18 F] fluoro-2-deoxy-d-glucose, a glucose analog
are two branches of the pathway, depending on whether the that is concentrated in highly proliferative tumorigenic masses
cell chooses to undergo aerobic or anaerobic respiration. Under (Pritchard et al., 2012).
standard conditions, in which oxygen (O2 ) is abundant, and for This high glycolytic rate, or the Warburg effect, is interpreted
long-term maintenance, healthy cells choose aerobic respiration. as an accelerated glucose metabolism that is accompanied by an
In aerobic respiration, which takes place in the cytoplasm and increase in glycolytic enzymes, including PGM and glucose phos-
mitochondria, oxygen and glucose are used to generate 38 adeno- phate isomerase (GPI), events not observed in normal cells. A
sine tri-phosphate (ATP) molecules and carbon dioxide (CO2 ) is possible explanation of the Warburg effect lies in the fact that in
released and O2 is absorbed. In contrast, in anaerobic respiration, normal differentiated cells, the need for division is low. Glucose
which occurs only in the cytoplasm, 2 molecules of ATP are syn- metabolism is therefore optimized for efficient energy genera-
thesized in the electron transport chain, using inorganic molecules tion: glycolysis is linked to the TCA cycle in mitochondria. Feeding
other than O2 ; CO2 is released, but O2 is not required (Fig. 2). substrates to the TCA cycle is critical for the generation of ATP,
Although anaerobic glycolysis is energetically less efficient, the lipids and ROS. We and others suggest that the major role of
ATP generation in this process is faster than that in aerobic gly- glycolysis in cancer cells is to provide substrates to the PPP for
colysis associated with the tricarboxylic acid (TCA) cycle (or Krebs nucleotide synthesis, a great priority considering their high division
cycle). Anaerobic glycolysis generates lactate, and due to the low rate (Weinberg and Chandel, 2009; Kondoh et al., 2007a). In this
levels of energy produced (i.e., ATP), only low levels of ROS are way, the high glycolytic rate of tumor cells confers an advantage,
produced, resulting in cellular benefits. In contrast, while aerobic as they are hardly exposed to oxidative stress by avoiding ROS gen-
metabolism enables efficient energy production, high levels of ROS eration. Additionally, the production of NADPH by the PPP is used
are also generated, which damage key cellular components. In prin- by the cell in reductive biosynthesis reactions in the prevention
ciple, one could expect that, depending on the O2 concentration of oxidative stress (Kondoh et al., 2005; Vaughn and Deshmukh,
in the microenvironment, cells may choose one pathway or the 2008). This result is supported by specific findings, such as the
other. However, cancer cells in particular choose the most conve- enhancement of key enzymes of the PPP in cancer cells (Vizan
nient way to increase proliferation, which is determined by their et al., 2009). Thus, in primary wild-type cells, the Warburg effect not
382 V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390

only confers protection from oxidative damage, but this high gly- O2 consumption and superoxide production in a process that is
colytic flux is translated into a longer life span with the avoidance accompanied by massive cell death. During prolonged culture in
of ROS-induced senescence, whereas glycolysis inhibition triggers vitro, the cell growth rate gradually increased along with tumor
early senescence (Kondoh et al., 2005). The Warburg effect also forming potential, as observed in in vitro anchorage-independent
triggers local acidosis, requiring tumors to selectively develop acid growth assays and in vivo tumor formation assays in immuno-
resistance (Gatenby and Gillies, 2004). According to the “parasitic” deficient mice. Notably, the glucose-to-lactic acid flux increased
cancer cell hypothesis, this acidosis may dramatically affect CAFs, with the passage number, while the cellular oxygen consumption
but not cancer cells, because these cells are prepared to manage decreased. This conversion in metabolic properties is associated
acidosis and hypoxia. with a change in mitochondrial NAD+/NADH redox, which is indica-
Another example of the Warburg effect and oxidative damage tive of a decreased mitochondrial TCA cycle and OXPHOS activity.
are fibroblasts that are deficient for cytoplasmic polyadenyla- Moreover, the Ras oncogene increases OXPHOS activity in early
tion element binding protein (CPEB) protein, which have fewer transformed progenitor cells, such as mesenchymal SCs (de Groof
mitochondria compared with wild-type cells, but have a high gly- et al., 2009; Funes et al., 2007).
colytic rate (Burns and Richter, 2008). Moreover, in p53 null cells, We and others have shown that increased glucose consump-
aerobic respiration is reduced and glycolysis is elevated. Some tion and lactate production, also known as the Warburg effect, are
downstream genes whose expression is regulated by p53, includ- almost universal hallmarks of solid tumors that favor tumor growth
ing SCO2, TIGAR, or p53 inducible gene 3 (PIG3), may be responsible (Bensinger and Christofk, 2012; Kondoh et al., 2005). However, high
for the changes related to energy metabolism (Matoba et al., 2006; levels of glycolysis in the absence of adequate OXPHOS may not
Won et al., 2012; Kotsinas et al., 2012). be as beneficial for tumor growth in general. The knockdown of
Alternative explanations for the Warburg effect in tumors p32/hyaluronic acid binding protein 1 (HABP1), a mitochondrial
include an optimization of glucose usage, the presence of a scarce cell surface protein that is overexpressed in certain cancer cells,
metabolite in the tumor microenvironment, or an impairment shows a strong shift in cellular metabolism from OXPHOS to glycol-
in mitochondrial function. Overall, this high glycolytic rate is an ysis. This finding suggests that the function of proteins that regulate
advantage for the Darwinian selection of tumors in vivo; therefore, the balance between OXPHOS and glycolysis according to cellular
glycolytic inhibiting drugs may provide a benefit for cancer therapy needs, including HABP1, is important for cancer cells to guarantee
(Gatenby and Gillies, 2004). their immortalization and malignancy (Fogal et al., 2010).
In contrast, mitochondrial oxidative phosphorylation However, intense glycolysis is not present in all tumor types;
(OXPHOS), the major energy source in aerobic organisms, is a rapidly growing tumors depend more on this process than do
process by which glucose or other molecules (i.e., glutamine, slowly growing ones (Jose et al., 2011). Recent reports have found
glycine, alanine, glutamate, proline and alternatively, fatty acids, that the mitochondria in tumor cells are active at a low capacity
ketone bodies, short chain carboxylic acids, propionate, acetate and provide these cells with most of the needed ATP (Bustamante
and butyrate) are oxidized to CO2 and H2 O (Fig. 2). Electrons and Pedersen, 1977). Several cancer cells rely on the OXPHOS path-
from NADH or FADH2 , which are produced in glycolysis, fatty acid way to produce ATP rather than glycolysis (e.g., cervical carcinoma
oxidation, and the citric acid cycle, are transferred to O2 by electron and glioma cells). Overall, the fact that some cancer cells preferen-
carriers located in the mitochondrial inner membrane, leading tially use glycolysis as an energy source and are able to change from
to the pumping of protons out of the mitochondrial matrix. The aerobic glycolysis to OXPHOS in low-glucose conditions suggests
consequent unbalanced proton distribution creates a pH gradient that tumor cells have the capacity to reprogram their metabolism
and a transmembrane electrical potential that generates a proton- to rapidly adapt to genetic and/or microenvironmental changes
motive force. ATP synthase produces ATP when protons flow back (Beckner et al., 2005; Plecita-Hlavata et al., 2008; Rossignol et al.,
into the mitochondrial matrix (Berg, 2002). Mitochondrial energy 2004). For such cases, dual therapy, i.e., the inhibition of glycoly-
metabolism, which is necessary to produce ROS, is disrupted in sis in combination with mitochondrial respiration inhibiting drugs,
many cancer cells because mitochondrial DNA mutations exist in would be more efficient and promises to be more effective than
tumorigenic cells. Remarkably, mitochondrial DNA mutations pro- suppressing glycolysis alone.
mote tumorigenesis via alterations in ROS generation, ultimately
influencing the apoptotic response (Park et al., 2009).
A great variety of different cancer bioenergetic signatures have 6. ROS-regulated microRNAs involved in cancer
been described in tumorigenic cells. Some theories suggest that gly-
colysis and OXPHOS cooperate in cancer cells to provide these cells MicroRNAs are small (21–24 nucleotides), non-coding, evolu-
with the needed energy (Smolkova et al., 2010). Studies of cancer tionarily conserved RNA molecules that regulate gene expression
cells revealed that mitochondrial OXPHOS contributed to 79% of by base pairing with the 3 untranslated region (UTR) of tar-
the required energy; this percentage decreased to 30% in hypoxic get mRNAs, causing translational repression or mRNA cleavage
conditions (Guppy et al., 2002; Hervouet et al., 2008). A high (Krek et al., 2005; Miska et al., 2004). The human genome con-
glycolytic rate, which is associated with a rapid tumor growth pat- sists of more than 1000 microRNAs that are involved in a myriad
tern, leads to increased tumor cell malignancy, and OXPHOS is the of processes. MicroRNA expression is associated with cellular
alternative choice for limited glucose conditions. Several key onco- proliferation, death, and development, and microRNAs actively
genes/tumor suppressor genes that are involved in proliferation participate in different pathologies, including cancer. Ionizing
participate in the stimulation of glycolytic metabolism, including radiation, etoposide, and H2 O2 can induce alterations in global
HIF-1␣, Ras, c-myc, Src and p53. Changes in the tumor microen- microRNA expression patterns (Simone et al., 2009). Fibroblasts
vironment provoke some of these genes to enhance OXPHOS. exposed to such agents undergo changes in their microRNA expres-
The c-myc gene can provoke aerobic glycolysis and/or OXPHOS, sion, depending on the anticancer agent, the radiation dose,
depending on the tumor cell microenvironment, by increasing the and the time of exposure. A common signature in response to
expression of glycolytic genes or activating the mitochondrial oxi- exogenous genotoxic agents can be identified. It is noteworthy
dation of glutamine (Wise et al., 2008; Gao et al., 2009). OXPHOS that pre-treatment with the free radical scavenger NAC pre-
may also be required for the activation of certain tumor suppressor vented radiation-induced alterations in microRNA expression,
proteins, such as Bax and Bak, to control apoptosis (Tomiyama et al., suggesting that microRNAs actively respond to oxidative stress
2006). H-RasV12 /E1A transformation initially causes an increase in (Mathe et al., 2012).
V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390 383

Fig. 3. ROS levels dictate biological outcomes. Different ROS levels determine the cell response and are classified as follows: +(proliferation and differentiation), ++ (adaptative
response: senescence/quiescence), +++ (generalized DNA damage, also including damaged proteins and lipids), and ++++ (cell death). The ability of ROS to chemosensitize
cancer cells depends on the basal ROS levels in such cells. The ROS levels could represent a double-edged sword: ROS activation below a specific threshold promotes survival;
however, if certain limits are reached, such effects are incompatible with cell viability, and the cells will die. Examples of chemotherapeutic treatments with increased ROS
are paclitaxel, doxorubicin, and cisplatin. Radiotherapy also increases ROS. (N, Normal cells; S, Senescent cells; T, Tumor cells; SCs, Stem cells; CSCs, Cancer stem cells).

Several ROS-related microRNAs have been described (Favaro consumption, which protects these cells from oxidative stress
et al., 2010). miR-210, which is overexpressed in breast and hepa- (Diehn et al., 2009; Funes et al., 2007; Iyer et al., 1998; Kondoh
tocellular carcinomas (HCC), is associated with invasion and poor et al., 2005).
clinical outcome (Rothe et al., 2011). miR-210 is strongly induced The importance of CSCs in molecular oncology results from
under hypoxic conditions and mediates the metastasis process the fact that they are responsible for tumor relapse, as they are
(Ying et al., 2011). Among the target mRNAs identified for miR-210 more resistant to chemo- and radiotherapeutic treatments than
are vacuole membrane protein 1 (VMP1) and iron sulfur scaffold cancer cells (i.e., non-CSCs). CSCs are present in most solid and non-
protein (ISCU), which is necessary for the formation of iron-sulfur solid tumors, including leukemia, brain, head and neck, breast and
clusters and cofactors for enzymes involved in the TCA cycle, elec- colon cancers, among others (Al-Hajj et al., 2003; Lapidot et al.,
tron transport, and iron metabolism. Remarkably, ISCU is also a 1994; O’Brien et al., 2007; Piccirillo et al., 2006; Prince et al., 2007;
target of HIF-1␣. Thus, miR-210 helps cancer cells adapt to hypoxic Ricci-Vitiani et al., 2007; Wulf et al., 2001). In contrast to dif-
conditions by regulating mitochondrial function and ROS genera- ferentiated cancer cells, CSCs generally maintain low ROS levels,
tion in situations in which enhanced cell survival and glycolysis exhibiting redox patterns similar to their corresponding wild-type
favor proliferation. SCs (Fig. 3). Several properties have been suggested in the mainte-
miR-128a is downregulated in medulloblastoma, the most nance of the low ROS levels found in CSCs, which appear to occur
malignant brain tumor in children, with a very poor progno- through a combination of mechanisms that may be unique to a
sis (Venkataraman et al., 2010). microRNA-128a re-expression given tumor. These properties include: (a) the expression of ROS-
is able to inhibit medulloblastoma proliferation by downregu- scavenging molecules, (b) more efficient DNA repair responses, (c)
lating the Bmi-1 oncogene, which, in turn, leads to an increase promotion of glycolysis, (d) promotion of autophagy and (e) FoxO
in p16INF4A expression and indirect modulation of E2F1 activity protein upregulation (Diehn et al., 2009; Piao et al., 2012). Some
(Jacobs et al., 1999; Molofsky et al., 2005). Bmi-1 downregulation of these mechanisms are derived from the transforming activity
by microRNA-128a alters the redox equilibrium by increasing ROS of certain oncogenes that are unexpectedly linked to a capacity to
in tumor cells and promoting cellular senescence. In turn, Bmi-1 maintain elevated intracellular levels of reduced glutathione (GSH),
re-expression reverses superoxide generation and the previously the main redox buffer. Because GSH maintenance largely relies
observed genetic and phenotypic effects. on glucose metabolism through the PPP cycle, this may partially
Two members of the microRNA-200 family, miR-141 and miR- explain the preference for CSCs in glycolytic metabolism and in a
200a, target p38␣, which is one of the main sensors of oxidative hypoxic microenvironment. For example, human gastrointestinal
stress (Mateescu et al., 2011). p38␣ inactivation is associated CSCs show an enhanced capacity for GSH synthesis and the defense
with ROS accumulation and the subsequent stimulation of antioxi- against ROS by a cystine-glutamate exchange transporter (Ishimoto
dant defenses. Primary high-grade human ovarian carcinomas, the et al., 2011). Alternatively, the low ROS levels of certain subsets of
most lethal gynecologic malignancies, overexpress the microRNA SC (i.e., non-CSC) populations may represent a reflection of their
miR-200a. Concomitant miR-200a overexpression and p38␣ down- state of quiescence in vivo.
regulation sensitizes ovarian cancer cells to chemotherapeutic ROS Interestingly, CD13, a HCC stem cell marker, has a protective
inducers. Therefore, the accumulation of microRNA-200 family role in regulating ROS levels to avoid DNA damage by controlling
members is associated with an encouraging therapeutic strategy, in angiogenesis, apoptosis and metastasis (Mishima et al., 2002; Kim
which tumors may be treated with ROS-inducing drugs and maxi- et al., 2012). The EMT that occurs as a result of the activation of
mal surgery (Hu et al., 2009). In contrast, ovarian tumors with lower TGF-␤ is associated with CD13+ CSC survival and ROS reduction in
amounts of miR-200a and no p38␣ effects should be treated by the liver. One possible explanation for the survival of liver CD13+
alternative methods that remain to be further elucidated. CSCs after chemotherapeutic treatment is the high expression of
aminopeptidase N, a ROS scavenger enzyme (Kim et al., 2012).
7. SCs/CSCs Immunohistochemical analysis revealed the presence of the con-
comitant expression of the CD13 and N-cadherin proteins in tumor
The immortalizing properties of SCs are a subject of debate. cells that survived chemotherapy. Accordingly, CD13 repression
We and others have demonstrated that SCs are more glycolytic generates apoptosis and inhibits the self-renewal ability of CSCs
than primary cells and possess reduced mitochondrial oxygen (Kim et al., 2012).
384 V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390

By upregulating the cells’ antioxidant capacity (in most cases, 2007). Moreover, the status of certain oncogenes/tumor suppres-
by increasing GSH or FoxO), CSCs couple the SC redox phenotype sor genes is also important in determining its effectiveness (Zhao
with radioresistance. Under these circumstances, it is expected that et al., 2006). Synthetic, metal-based antioxidants with biologically
they will respond to a specific CSC therapy in combination with active ligands are potential ROS scavenging molecules, and redox
pro-oxidant treatment. Importantly, pharmacological depletion of balance is restored in damaged cells. However, further research
ROS scavengers in CSCs markedly decreases their clonogenicity and is needed to establish an optimum design of these compounds.
results in radiosensitization (Diehn et al., 2009). Loss of the FoxO3 The combination of a metal ion and a redox-active organic ligand
gene causes myeloproliferative syndrome due to the increased appears to be a promising approach for creating an effective syn-
accumulation of ROS in hematopoietic progenitors, leading to the thetic antioxidant. The carefully controlled titration of the amount
activation of the Akt/mTOR signaling pathway (Yalcin et al., 2010). of catalytic antioxidants administered to treat some cancers could
Overall, the fact that SCs have high glycolysis levels and low facilitate ROS scavenging, restore the redox balance in tumor cells,
mitochondrial oxygen consumption suggests that these features and reduce their growth advantage.
are undoubtedly related to their immortalizing properties (Kondoh
et al., 2007b). This would explain, at least in part, CSCs’ higher 8.2. Pro-oxidant therapy
levels of resistance to ROS and radio- and chemoresistance. The
metabolic modulation-associated survival mechanisms employed Given the role of antioxidants in cancer prevention and treat-
by SCs are possibly mimicked by healthy cells to be immortalized ment, it somehow seems contradictory that a pro-oxidant therapy,
and transformed. which is based on the generation of oxygen radicals, could be used
for cancer annihilation. Although oxidative stress caused by ROS
accumulation promotes tumor growth, it can also increase the sen-
8. How therapy could target the oxidative stress
sitivity to treatment. This may be the reason for the limited success
mechanism
of antioxidant therapies in clinical trials (Appierto et al., 2009;
Mateescu et al., 2011). However, it remains a challenge to clas-
It is clear that oxidative stress plays a role in the develop-
sify the types of tumors that may benefit from pro-oxidant therapy
ment of age-associated pathologies, including cancer (Sohal and
based on their intrinsic properties. The first attempt to employ
Weindruch, 1996). Importantly, cancer therapy is highly associ-
in vivo pro-oxidant agents, such as antitumor agents, was reported
ated with ROS production because most cytotoxic drugs block
by Nathan and Cohn, who used glucose oxidase as an H2 O2 precur-
DNA replication, causing DNA damage. For example, the overex-
sor and obtained a significant decrease in tumor growth (Nathan
pression of tumor suppressor genes in combination with radiation,
and Cohn, 1981). Because cells can develop adaptive responses to
which is known to resume an apoptotic program, provokes massive
ROS by primarily increasing detoxification enzymes, free radical
cell killing, a process that is blocked by ROS scavenger molecules
research has focused on disease treatment with radical scavengers
(Yacoub et al., 2004).
that help in coping with the elevated ROS levels (Benhar et al., 2002;
Nathan et al., 1981). Currently, there is a major focus on cancer ther-
8.1. Antioxidants apies that increase ROS to reach a threshold that is incompatible
with cell viability, even for a highly tumorigenic cell (Figure 3). In
Antioxidants constitute the cellular mechanism of defense such a context, it will be necessary to first precisely define the ROS
against the consequences of high concentrations of reactive species levels in a particular tumor and to explore its ability to switch from
(mainly ROS) (Oh et al., 2008). There is evidence that these aerobic glycolysis to OXPHOS, considering that it would be possible
molecules prevent tumorigenesis and increase life span (Kovacic to manipulate it to obtain efficient cell killing in terms of therapy.
and Jacintho, 2001). Apart from their protective role as preven- Examples of drugs that induce ROS include arsenic trioxide, N-(4-
tive agents against cancer development, there is evidence that hydroxyphenyl) retinamide (HPR) and dithiophene (NSC656240).
antioxidant supplementation during chemotherapy has potential Moreover, certain commonly used chemotherapeutic agents
to reduce dose-limiting toxicities in patients with cancer (Block induce ROS. The anticancer drug 5-fluorouracil (5-FU) interferes
et al., 2008). with DNA replication and consequently with the inhibition of cell
Treatment with antioxidants (particularly polyphenols) is effec- division, which are effects that are highly associated with tumor
tive in combating tumor cells (Borrelli et al., 2011; Dhar et al., cells. Recent in vitro and in vivo data from animal studies and
2011). When the naturally occurring equilibrium between anti- and patient treatment have suggested that the cytotoxicity induced
pro-oxidant agents is lost, high ROS levels bypass the antioxidant by the chemotherapeutic drugs 5-FU and oxaliplatin is linked to
function and supplementary antioxidants may be necessary for increased ROS formation (Afzal et al., 2012).
therapy against ROS-induced cancer. However, it must be kept in One of the most exciting topics regarding pro-oxidant therapy
mind that the supplied synthetic antioxidant concentration is cru- is its effects on CSCs. For example, CD13+ cells reduce ROS-induced
cial for a correct cellular response; a high concentration could cause DNA damage after genotoxic chemo-radiation stress, a feature that
an opposing effect and act as a pro-oxidant. Initially, it was demon- protects these cells from apoptosis. In mouse xenograft models,
strated that the overexpression of genes that codify for antioxidant the combination of a CD13 inhibitor and 5-FU drastically reduced
enzymes was effective in cancer therapy. For example, overex- the tumor volume compared with either agent alone (Haraguchi
pression of SOD2 was found to suppress the malignant phenotype et al., 2010). Therefore, the combination of a CSC surface receptor
of human melanoma cells (Church et al., 1993). In addition, several inhibitor with ROS-inducing therapy (i.e., chemo- or radiotherapy)
Nrf2 activators constitute potential therapies for oxidative stress, may improve the treatment of aggressive tumors, such as liver can-
inflammation and chemoprevention. In a two-stage mouse skin cer, by precisely eradicating the cell type that permanently feeds
carcinogenesis model, a Protandim-supplemented diet (patented the tumor mass (i.e., CSCs). In this context, Venkataraman et al.
supplement consisting of several low-dose natural Nrf2 activators) (2010) envisioned a scenario in which miR-128a can be used to
was found to significantly reduce the incidence of skin tumors (Liu induce ROS in medulloblastoma SCs to make them more radiosen-
et al., 2009). The increased malignant cell behavior induced by sitive.
several oncogenes is associated with an increase in ROS produc- Paclitaxel, a mitotic inhibitory drug, stabilizes microtubules, and
tion, and the treatment of such malignant cells with antioxidants as a result, it interferes with the normal breakdown of microtubules
results in a reduction in the tumorigenic properties (Guijarro et al., during cell division. Paclitaxel also promotes ROS generation
V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390 385

(i.e., O2 − and H2 O2 ) by enhancing the activity of the NADPH We have previously demonstrated that treatment using the glu-
oxidase Nox and slowing tumor growth (Alexandre et al., 2007; cose analog 2-deoxyglucose in SCs is able to considerably decrease
Ramanathan et al., 2005). In breast cancer cells, this treatment their proliferative potential and immortalizing properties (Kondoh
caused the translocation of Rac1, a positive regulatory protein of et al., 2007b). Deoxyglucose and two other inhibitors of hexoki-
Nox. It has been established that the ROS generated by paclitaxel nase (HK), lonidamide and 3-Bromopiruvate (3-BrPa), have been
mainly accumulates outside of the cells, while the intracellular ROS successfully tested in pre-clinical trials of malignant tumors (Ko
remains unchanged. This accumulation outside of the cells pro- et al., 2001; Oudard et al., 1995). Importantly, 3-BrPa represents a
vokes lethal damage to bystander cancer cells that have not been novel therapy for liver cancer and other cancer types with poor sur-
exposed to paclitaxel (Alexandre et al., 2007). This unexpected vival rates (Geschwind et al., 2002). Along the same line of evidence,
cytotoxic bystander effect also occurs with other microtubule- targeting lactate efflux is another avenue to inhibit the glycolytic
targeted agents, such as vincristine and taxotere, but not with pathway and provoke cell death. This has been performed by uti-
5-FU or doxorubicin. This fact represents a novel basis on which lizing the small-molecule alpha cyano-4-hydroxy cinnamic acid
to improve the clinical use of chemotherapeutic agents that not (ACCA) (Colen et al., 2006). In this case, the targeted tumors show
only consider the tumor cells but also affect neighboring cells. In increased sensitivity if combined therapy is given, e.g., the addition
most cases, combined therapy appears to be effective for cancer of radiation treatment (Colen et al., 2006).
treatment. For example, it has been demonstrated that the anti- Bezielle (BZL101), an oral drug presently being tested in human
cancer effects of ionizing radiation combined with arsenic trioxide clinical trials for advanced breast cancer (Chen et al., 2012; Perez
increase the therapeutic efficacy compared to individual treat- et al., 2010), has proven to be efficient in inhibiting not only gly-
ments in human prostate cancer cells. Combined treatment exhibits colysis but also oxidative phosphorylation. Bezielle is derived from
enhanced ROS generation compared to each treatment alone, Scutellaria barbata, a plant with anti-cancer properties. This drug
provoking autophagy and apoptosis by inhibiting the Akt/mTOR has selective cytotoxic effects in vitro in cancer cells and in vivo
signaling pathways (Chiu et al., 2012). in renal, hepatocellular and prostate carcinomas (Cha et al., 2004;
In contrast, a gene of particular significance, melanoma differen- Marconett et al., 2010; Yin et al., 2004). Bezielle’s selectivity relies
tiation associated gene-7/interleukin-24 (mda-7/IL-24), selectively on generating large amounts of ROS in tumor cells, causing DNA
induces growth suppression and apoptosis in a broad spectrum damage and hyperactivation of PARP, which mainly targets the
of human cancers without eliciting apparent deleterious effects in mitochondria. This PARP hyperactivation provokes massive apop-
normal cells. One exception to the action of this gene is in pancreatic totic cell death (Zong et al., 2004). PARP hyperactivation in cancer
cancer, due to the lack of the capacity to convert mda-7/IL-24 mRNA cells greatly reduces the cellular supply of NAD+ and ATP, which are
into protein. Lebedeva et al. (Lebedeva et al., 2005) showed that needed for the PARylation of several proteins implicated in the DNA
pancreatic cells, regardless of the status of the K-ras gene, simulta- repair mechanism, the regulation of transcription, and cell cycle
neously express mda-7/IL-24 protein and undergo apoptosis when progression (Rouleau et al., 2010). A possible cause for the inhibi-
treated with ROS inducers, demonstrating that tumorigenesis was tion of glycolysis observed in tumors treated with Bezielle could
effectively suppressed in vivo and in vitro. The specificity of this be the limited cytosolic NAD+ concentration (Alano et al., 2010).
action is documented as the ability of ROS inhibitors to block this Importantly, the small amount of DNA damage that Bezielle causes
killing effect. This example represents a promising combinatorial in normal cells is restored by the DNA repair mechanism.
approach for pancreatic tumors, which are exceptionally aggressive Overexpression of the glycolytic enzyme pyruvate kinase M2
and have no long-term effective therapy. (PKM2) promotes anabolic processes and confers a selective advan-
Recently, a small molecule named piperlongumine, which tage to cancer cells, enabling them to sustain antioxidant responses
selectively kills cancer cells, has been described. Piperlongumine that lengthen cancer cell survival under intense oxidative stress
potently inhibits the growth of spontaneously formed malignant (Anastasiou et al., 2011). It has been reported that the high ROS
breast tumors and their associated metastases in mice by increas- levels that occur in human lung cancer cells repress PKM2 through
ing ROS levels and apoptosis. Importantly, piperlongumine has cysteine 358 oxidation, and consequently, glucose is directed to the
little effect on rapidly or slowly dividing primary normal cells (Raj PPP cycle. Under these circumstances, the cell is able to detoxify
et al., 2011), which represents a clear advantage of this compound ROS (Anastasiou et al., 2011). However, lung cancer cells in which
for therapeutic purposes. Fucoidan, an enzymatic, digested, low endogenous PKM2 is replaced with Cys358 to Ser358 oxidation-
molecular weight compound that is extracted from brown seaweed resistant mutants exhibit increased sensitivity to oxidative stress
and has anti-tumor properties, activates an independent caspase and impaired tumor formation in a xenograft model (Anastasiou
apoptotic pathway and is implicated in angiogenesis inhibition et al., 2011). A possible therapeutic approach is the use of small
(Liu et al., 2012). Such a pathway involves JNK, p38␣ and ERK1/2 PKM2 activator molecules that are able to interfere with tumor cell
phosphorylation in various cancer cells by activating ROS-mediated metabolism, as PKM2 inactivation ultimately increases ROS sensi-
MAP kinases and the regulation of bcl-2 protein family-mediated tivity. These results, which have been proven in mouse models, are
mitochondrial apoptosis (Zhang et al., 2011). Although this com- promising for the design of future therapies in patients with cancer.
pound has not been tested in patients, it appears to be efficient in Studies in cell lines that are defective for mitochondrial enzymes
leukemia, breast, colon, and lung cancer cells (Jin et al., 2010). have demonstrated that the oxidative phosphorylation activ-
Overall, a further understanding of the oncogenic/tumor sup- ity defines the rate of glucose utilization by aerobic glycolysis
pressor proteins that are respectively overexpressed or silenced (Sanchez-Arago et al., 2010). Importantly, abrogated mitochondrial
in human tumors and their relationship with ROS would permit respiration leads to a diminished ROS signaling potential in an
the classification of tumor types that are susceptible to efficient efficient response to 5-FU treatment. In contrast, an abrogated gly-
eradication by pro-oxidant therapy. colysis pathway in the same cells with intact mitochondria results
in a more efficient response to 5-FU.
8.3. Glycolysis inhibition Over the course of years, it has been discovered that glycolytic
inhibition per se has not been an effective chemotherapeutic strat-
A first choice for inhibiting glycolysis would be the use of non- egy, due in part to the systemic toxicity of antiglycolytic drugs.
hydrolyzable glucose analogs that prevent the initial glycolytic However, combined treatment that involves glycolysis inhibition
steps from advancing to the final cascade of metabolic degrada- and mitochondrial function may concomitantly help to over-
tion (e.g., past the hexokinase-mediated phosphorylation step). come this issue. Some examples include the mitochondria-targeted
386 V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390

drugs Mito-Q and Mito-CP, which, in combination with 2-deoxy-d- uptake due to enhanced glycolysis and increased lactate produc-
glucose, are more effective than inhibiting glycolysis alone (Cheng tion. Thus, in solid tumors, the Warburg effect could be the result
et al., 2012). of an adaptive mechanism that, in combination with angiogen-
Along the same lines of evidence, those drugs, which are equally esis, allows cells to overcome hypoxic stress. In addition, the
effective for glycolysis and OXPHOS inhibition, are more efficient Warburg effect confers protection from ROS for cancer cells in
than acting through a single process. For example, 3-BrPA inhibits comparison to normal cells. In highly glycolytic tumors, a poten-
cellular ATP production due to its effects on glycolysis and OXPHOS. tial therapeutic strategy would consist of inhibiting glycolysis to
It acts on several cellular enzymes that are involved in glycolysis, obtain a final effect on the re-activation of mitochondrial oxida-
such as GAPDH and PGK, and it inhibits mitochondrial respiration tive metabolism, and ultimately, increased ROS levels. However,
enzymes, including succinate dehydrogenase (SDH). The combined not all tumors exhibit augmented glycolysis. Some tumors prefer
results of 3-BrPA are responsible for a decrease in viability and to enhance OXPHOS, particularly under low glucose concentrations
the apoptotic death of HCC cells. These effects, which are linked in which glycolysis would be less efficient. In these cases, novel
to proton leaks, may reflect permeability in the transition pores energy restriction/mimetic agents promise to represent a better
of the mitochondrial membrane, which occurs prior to apoptotic therapeutic choice for reducing human tumor growth. Accumu-
processes. lating evidence shows that ROS levels in tumor cells are key for
Earlier studies have demonstrated that curcumin, betulinic acid determining a specific therapy. There is an urgent need for estab-
(BA), synthetic triterpenoid anticancerigenous drugs, and tolfe- lishing biomarkers (i.e., oncogenes, tumor suppressor genes, and
namic acid are able to reduce bcl2, survivin, hepatocyte growth microRNAs) that will predict the response of a pro-oxidant ther-
factor receptor (c-Met), epidermal growth factor receptor (EGFR), apy to define the best choice for cancer eradication for each cancer
cyclin D1, VEGF, and its receptors VEGFR1 and VEGFR2 in sev- patient. The association of CSCs and the resistance to therapy solic-
eral cancer cell types. BA is a naturally-occurring triterpenoid its a critical rethinking of the efficiency of pro-oxidant therapy on
and is a selective inhibitor of human melanoma, malignant brain CSCs. Because low ROS levels seem to be critical for maintaining SC
tumors, and human leukemia cells, but it is apparently inefficient in function, specific ROS induction in CSCs should render these cells
epithelial tumors. The BA mechanism mainly works by inhibiting sensitive to therapy.
topoisomerase, a process that is accompanied by caspase activa-
tion, mitochondrial membrane alterations and DNA fragmentation;
therefore, BA belongs to a class of mitochondria-targeted drugs (Li Acknowledgement
et al., 2010). Recently, BA has been suggested to act by repressing
the specificity protein (Sp) transcription factors Sp1, Sp3 and Sp4, Work in the authors’ laboratory is supported by the Healthy
which are overexpressed in cancer cells but underexpressed in nor- Ministry (FIS Project PI09/02193 to M.E.LL.).
mal tissues. In colon cancer cells, the mechanism of BA-induced
repression of the Sp protein is due to the induction of ROS, ROS-
mediated miR-27a repression, and the induction of the Sp repressor References
gene ZBTB10 (Chintharlapalli et al., 2011).
Another drug, ethyl 2-((2,3-bis(nitrooxy)propyl)disulfanyl) Afzal, S., Jensen, S.A., Sorensen, J.B., Henriksen, T., Weimann, A., Poulsen, H.E., 2012.
Oxidative damage to guanine nucleosides following combination chemotherapy
benzoate (GT-094), is a novel drug chimera that exhibits nonste- with 5-fluorouracil and oxaliplatin. Cancer Chemotherapy and Pharmacology
roidal anti-inflammatory properties, nitric oxide moieties and a 69, 301–307.
disulfide pharmacophore that possesses cancer chemopreventive Al-Hajj, M., Wicha, M.S., Benito-Hernandez, A., Morrison, S.J., Clarke, M.F., 2003.
Prospective identification of tumorigenic breast cancer cells. Proceedings of the
activity (Pathi et al., 2011). GT-094 was found to inhibit cell prolifer- National Academy of Sciences of the United States of America 100, 3983–3988.
ation and induce apoptosis (i.e., caspase-dependent PARP cleavage) Alano, C.C., Garnier, P., Ying, W., Higashi, Y., Kauppinen, T.M., Swanson, R.A., 2010.
in RKO and SW480 colon cancer cell lines, in addition to low- NAD+ depletion is necessary and sufficient for poly(ADP-ribose) polymerase-1-
mediated neuronal death. Journal of Neuroscience 30, 2967–2978.
ering the mitochondrial membrane potential and enhancing ROS
Alexandre, J., Hu, Y., Lu, W., Pelicano, H., Huang, P., 2007. Novel action of pacli-
production. These responses are reversed with antioxidant (glu- taxel against cancer cells: bystander effect mediated by reactive oxygen species.
tathione) treatment (Pathi et al., 2011). GT-094 has been proven to Cancer Research 67, 3512–3517.
Allred, D.C., Anderson, S.J., Paik, S., Wickerham, D.L., Nagtegaal, I.D., Swain, S.M.,
act by downregulating genes associated with cell growth (i.e., cyclin
Mamounas, E.P., Julian, T.B., Geyer Jr., C.E., Costantino, J.P., Land, S.R., Wolmark,
D1, c-Met, EGFR, bcl-2, and survivin) and angiogenesis, along with N., 2012. Adjuvant tamoxifen reduces subsequent breast cancer in women with
some regulators of these genes, including Sp1, Sp3, and Sp4, which estrogen receptor-positive ductal carcinoma in situ: a study based on NSABP
have been found to be overexpressed in different cancer cells. GT- protocol B-24. Journal of Clinical Oncology 30, 1268–1273.
Almeida, M., Han, L., Martin-Millan, M., O’Brien, C.A., Manolagas, S.C., 2007. Oxidative
094 is able to control the expression of these transcription factors stress antagonizes Wnt signaling in osteoblast precursors by diverting beta-
by repressing miR-27a and inducing ZBTB10 (Sp repressor). The catenin from T cell factor- to forkhead box O-mediated transcription. Journal of
repression of the above mentioned transcription factors is reversed Biological Chemistry 282, 27298–27305.
Anastasiou, D., Poulogiannis, G., Asara, J.M., Boxer, M.B., Jiang, J.K., Shen, M., Bellinger,
by a glutathione antioxidant, suggesting that the antitumorigenic G., Sasaki, A.T., Locasale, J.W., Auld, D.S., Thomas, C.J., Vander Heiden, M.G., Cant-
activity of GT-094 in colon cancer cells is partially due to the acti- ley, L.C., 2011. Inhibition of pyruvate kinase M2 by reactive oxygen species
vation of the ROS-miR-27a axis (Pathi et al., 2011). contributes to cellular antioxidant responses. Science 334, 1278–1283.
Appierto, V., Tiberio, P., Villani, M.G., Cavadini, E., Formelli, F., 2009. PLAB induction
in fenretinide-induced apoptosis of ovarian cancer cells occurs via a ROS-
dependent mechanism involving ER stress and JNK activation. Carcinogenesis
9. Conclusions 30, 824–831.
Arnold, R.S., Shi, J., Murad, E., Whalen, A.M., Sun, C.Q., Polavarapu, R., Parthasarathy,
S., Petros, J.A., Lambeth, J.D., 2001. Hydrogen peroxide mediates the cell growth
There is much evidence that normal cells are affected by sur- and transformation caused by the mitogenic oxidase Nox1. Proceedings of the
rounding cancer cells, helping the cancer cells to grow, invade National Academy of Sciences of the United States of America 98, 5550–5555.
Bae, I., Fan, S., Meng, Q., Rih, J.K., Kim, H.J., Kang, H.J., Xu, J., Goldberg, I.D., Jaiswal, A.K.,
and metastasize. Remarkably, cancer cells are able to reprogram
Rosen, E.M., 2004. BRCA1 induces antioxidant gene expression and resistance to
their metabolism by adapting it to microenvironmental changes. oxidative stress. Cancer Research 64, 7893–7909.
Research in cancer bioenergetics has acquired a better understand- Banning, A., Deubel, S., Kluth, D., Zhou, Z., Brigelius-Flohe, R., 2005. The GI-GPx gene
ing of the way in which energy metabolism is regulated during is a target for Nrf2. Molecular and Cellular Biology 25, 4914–4923.
Bannister, J., B.H.P.O., 2007. Free Radicals in Biology and Medicine 10, 250–266.
oncogenesis. Compared to normal cells, tumor cells exhibit impor- Bar-Even, A., Flamholz, A., Noor, E., Milo, R., 2012. Rethinking glycolysis: on the
tant metabolic changes. Tumor cells show an increase in glucose biochemical logic of metabolic pathways. Nature Chemical Biology 8, 509–517.
V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390 387

Barzilai, A., Rotman, G., Shiloh, Y., 2002. ATM deficiency and oxidative stress: a new Esufali, S., Charames, G.S., Pethe, V.V., Buongiorno, P., Bapat, B., 2007. Activation
dimension of defective response to DNA damage. DNA Repair (Amsterdam) 1, of tumor-specific splice variant Rac1b by dishevelled promotes canonical Wnt
3–25. signaling and decreased adhesion of colorectal cancer cells. Cancer Research 67,
Beckner, M.E., Gobbel, G.T., Abounader, R., Burovic, F., Agostino, N.R., Laterra, J., Pol- 2469–2479.
lack, I.F., 2005. Glycolytic glioma cells with active glycogen synthase are sensitive Favaro, E., Ramachandran, A., McCormick, R., Gee, H., Blancher, C., Crosby, M., Devlin,
to PTEN and inhibitors of PI3K and gluconeogenesis. Laboratory Investigation 85, C., Blick, C., Buffa, F., Li, J.L., Vojnovic, B., Pires das, N.R., Glazer, P., Iborra, F., Ivan,
1457–1470. M., Ragoussis, J., Harris, A.L., 2010. MicroRNA-210 regulates mitochondrial free
Bell, E.L., Emerling, B.M., Ricoult, S.J., Guarente, L., 2011. SirT3 suppresses hypoxia radical response to hypoxia and krebs cycle in cancer cells by targeting iron
inducible factor 1alpha and tumor growth by inhibiting mitochondrial ROS pro- sulfur cluster protein ISCU. PLoS One 5, e10345.
duction. Oncogene 30, 2986–2996. Fei, J., Hong, A., Dobbins, T.A., Jones, D., Lee, C.S., Loo, C., Al-Ghamdi, M., Harnett, G.B.,
Benezra, M., Chevallier, N., Morrison, D.J., MacLachlan, T.K., El-Deiry, W.S., Licht, Clark, J., O’Brien, C.J., Rose, B., 2009. Prognostic significance of vascular endothe-
J.D., 2003. BRCA1 augments transcription by the NF-kappaB transcription fac- lial growth factor in squamous cell carcinomas of the tonsil in relation to human
tor by binding to the Rel domain of the p65/RelA subunit. Journal of Biological papillomavirus status and epidermal growth factor receptor. Annals of Surgical
Chemistry 278, 26333–26341. Oncology 16, 2908–2917.
Benhar, M., Engelberg, D., Levitzki, A., 2002. ROS, stress-activated kinases and stress Fogal, V., Richardson, A.D., Karmali, P.P., Scheffler, I.E., Smith, J.W., Ruoslahti, E.,
signaling in cancer. EMBO Reports 3, 420–425. 2010. Mitochondrial p32 protein is a critical regulator of tumor metabolism
Bensinger, S.J., Christofk, H.R., 2012. New aspects of the Warburg effect in cancer via maintenance of oxidative phosphorylation. Molecular and Cellular Biology
cell biology. Seminars in Cell & Developmental Biology 23, 352–361. 30, 1303–1318.
Berg, J.M., Tymoczko, J.L., Stryer, L., 2002. Biochemistry, 5th edition. W H Freeman, Folkman, J., 1995. Angiogenesis in cancer, vascular, rheumatoid and other disease.
New York. Nature Medicine 1, 27–31.
Block, K.I., Koch, A.C., Mead, M.N., Tothy, P.K., Newman, R.A., Gyllenhaal, C., 2008. Funes, J.M., Quintero, M., Henderson, S., Martinez, D., Qureshi, U., Westwood, C.,
Impact of antioxidant supplementation on chemotherapeutic toxicity: a sys- Clements, M.O., Bourboulia, D., Pedley, R.B., Moncada, S., Boshoff, C., 2007.
tematic review of the evidence from randomized controlled trials. International Transformation of human mesenchymal stem cells increases their depend-
Journal of Cancer 123, 1227–1239. ency on oxidative phosphorylation for energy production. Proceedings of
Borrelli, A., Schiattarella, A., Mancini, R., Morelli, F., Capasso, C., De, L.V., Gori, E., the National Academy of Sciences of the United States of America 104,
Mancini, A., 2011. The leader peptide of a human rec. MnSOD as molecular car- 6223–6228.
rier which delivers high amounts of Cisplatin into tumor cells inducing a fast Gao, P., Tchernyshyov, I., Chang, T.C., Lee, Y.S., Kita, K., Ochi, T., Zeller, K.I., De
apoptosis in vitro. International Journal of Cancer 128, 453–459. Marzo, A.M., Van Eyk, J.E., Mendell, J.T., Dang, C.V., 2009. c-Myc suppression
Bos, J.L., 1989. ras oncogenes in human cancer: a review. Cancer Research 49, of miR-23a/b enhances mitochondrial glutaminase expression and glutamine
4682–4689. metabolism. Nature 458, 762–765.
Burns, D.M., Richter, J.D., 2008. CPEB regulation of human cellular senescence, energy Gatenby, R.A., Gillies, R.J., 2004. Why do cancers have high aerobic glycolysis. Nature
metabolism, and p53 mRNA translation. Genes & Development 22, 3449–3460. Reviews Cancer 4, 891–899.
Bustamante, E., Pedersen, P.L., 1977. High aerobic glycolysis of rat hepatoma cells in Geschwind, J.F., Ko, Y.H., Torbenson, M.S., Magee, C., Pedersen, P.L., 2002. Novel ther-
culture: role of mitochondrial hexokinase. Proceedings of the National Academy apy for liver cancer: direct intraarterial injection of a potent inhibitor of ATP
of Sciences of the United States of America 74, 3735–3739. production. Cancer Research 62, 3909–3913.
Carmeliet, P., Jain, R.K., 2000. Angiogenesis in cancer and other diseases. Nature 407, Govindarajan, B., Sligh, J.E., Vincent, B.J., Li, M., Canter, J.A., Nickoloff, B.J., Rodenburg,
249–257. R.J., Smeitink, J.A., Oberley, L., Zhang, Y., Slingerland, J., Arnold, R.S., Lambeth, J.D.,
Cavalieri, E., Chakravarti, D., Guttenplan, J., Hart, E., Ingle, J., Jankowiak, R., Muti, Cohen, C., Hilenski, L., Griendling, K., Martinez-Diez, M., Cuezva, J.M., Arbiser,
P., Rogan, E., Russo, J., Santen, R., Sutter, T., 2006. Catechol estrogen quinones J.L., 2007. Overexpression of Akt converts radial growth melanoma to vertical
as initiators of breast and other human cancers: implications for biomarkers growth melanoma. Journal of Clinical Investigation 117, 719–729.
of susceptibility and cancer prevention. Biochimica et Biophysica Acta 1766, Graham, K.A., Kulawiec, M., Owens, K.M., Li, X., Desouki, M.M., Chandra, D., Singh,
63–78. K.K., 2010. NADPH oxidase 4 is an oncoprotein localized to mitochondria. Cancer
Cha, Y.Y., Lee, E.O., Lee, H.J., Park, Y.D., Ko, S.G., Kim, D.H., Kim, H.M., Kang, I.C., Kim, Biology and Therapy 10, 223–231.
S.H., 2004. Methylene chloride fraction of Scutellaria barbata induces apoptosis Guijarro, M.V., Leal, J.F., Blanco-Aparicio, C., Alonso, S., Fominaya, J., Lleonart, M.,
in human U937 leukemia cells via the mitochondrial signaling pathway. Clinica Castellvi, J., Cajal, S., Carnero, A., 2007. MAP17 enhances the malignant behavior
Chimica Acta 348, 41–48. of tumor cells through ROS increase. Carcinogenesis 28, 2096–2104.
Chen, C., Wei, Y., Hummel, M., Hoffmann, T.K., Gross, M., Kaufmann, A.M., Albers, Guppy, M., Leedman, P., Zu, X., Russell, V., 2002. Contribution by different fuels
A.E., 2011. Evidence for epithelial-mesenchymal transition in cancer stem cells and metabolic pathways to the total ATP turnover of proliferating MCF-7 breast
of head and neck squamous cell carcinoma. PLoS One 6, e16466. cancer cells. Biochemical Journal 364, 309–315.
Chen, V., Staub, R.E., Fong, S., Tagliaferri, M., Cohen, I., Shtivelman, E., 2012. Bezielle Halliwell, B., Chirico, S., 1993. Lipid peroxidation: its mechanism, measurement, and
selectively targets mitochondria of cancer cells to inhibit glycolysis and OXPHOS. significance. American Journal of Clinical Nutrition 57, 715S–724S.
PLoS One 7, e30300. Haorah, J., Ramirez, S.H., Schall, K., Smith, D., Pandya, R., Persidsky, Y., 2007. Oxida-
Cheng, G., Zielonka, J., Dranka, B.P., McAllister, D., Mackinnon Jr., A.C., Joseph, tive stress activates protein tyrosine kinase and matrix metalloproteinases
J., Kalyanaraman, B., 2012. Mitochondria-targeted drugs synergize with 2- leading to blood-brain barrier dysfunction. Journal of Neurochemistry 101,
deoxyglucose to trigger breast cancer cell death. Cancer Research 72, 566–576.
2634–2644. Haraguchi, N., Ishii, H., Mimori, K., Tanaka, F., Ohkuma, M., Kim, H.M., Akita, H., Tak-
Chintharlapalli, S., Papineni, S., Lei, P., Pathi, S., Safe, S., 2011. Betulinic acid inhibits iuchi, D., Hatano, H., Nagano, H., Barnard, G.F., Doki, Y., Mori, M., 2010. CD13 is
colon cancer cell and tumor growth and induces proteasome-dependent and a therapeutic target in human liver cancer stem cells. Journal of Clinical Inves-
-independent downregulation of specificity proteins (Sp) transcription factors. tigation 120, 3326–3339.
BMC Cancer 11, 371. Hayashi, S., Hajiro-Nakanishi, K., Makino, Y., Eguchi, H., Yodoi, J., Tanaka, H., 1997.
Chiu, H.W., Chen, Y.A., Ho, S.Y., Wang, Y.J., 2012. Arsenic trioxide enhances the Functional modulation of estrogen receptor by redox state with reference to
radiation sensitivity of androgen-dependent and -independent human prostate thioredoxin as a mediator. Nucleic Acids Research 25, 4035–4040.
cancer cells. PLoS One 7, e31579. Hervouet, E., Cizkova, A., Demont, J., Vojtiskova, A., Pecina, P., Franssen-van Hal,
Church, S.L., Grant, J.W., Ridnour, L.A., Oberley, L.W., Swanson, P.E., Meltzer, P.S., N.L., Keijer, J., Simonnet, H., Ivanek, R., Kmoch, S., Godinot, C., Houstek, J., 2008.
Trent, J.M., 1993. Increased manganese superoxide dismutase expression sup- HIF and reactive oxygen species regulate oxidative phosphorylation in cancer.
presses the malignant phenotype of human melanoma cells. Proceedings of the Carcinogenesis 29, 1528–1537.
National Academy of Sciences of the United States of America 90, 3113–3117. Hu, X., Macdonald, D.M., Huettner, P.C., Feng, Z., El Naqa, I.M., Schwarz, J.K., Mutch,
Colen, C.B., Seraji-Bozorgzad, N., Marples, B., Galloway, M.P., Sloan, A.E., Mathupala, D.G., Grigsby, P.W., Powell, S.N., Wang, X., 2009. A miR-200 microRNA cluster
S.P., 2006. Metabolic remodeling of malignant gliomas for enhanced sensitiza- as prognostic marker in advanced ovarian cancer. Gynecologic Oncology 114,
tion during radiotherapy: an in vitro study. Neurosurgery 59, 1313–1323. 457–464.
de Groof, A.J., te Lindert, M.M., van Dommelen, M.M., Wu, M., Willemse, M., Smift, Hybertson, B.M., Gao, B., Bose, S.K., McCord, J.M., 2011. Oxidative stress in health
A.L., Winer, M., Oerlemans, F., Pluk, H., Fransen, J.A., Increased, Wieringa B., 2009. and disease: the therapeutic potential of Nrf2 activation. Molecular Aspects of
OXPHOS activity precedes rise in glycolytic rate in H-RasV12/E1A transformed Medicine 32, 234–246.
fibroblasts that develop a Warburg phenotype. Molecular Cancer 8, 54. Indran, I.R., Hande, M.P., Pervaiz, S., 2010. Tumor cell redox state and mitochondria
Dhar, S.K., Tangpong, J., Chaiswing, L., Oberley, T.D., St Clair, D.K., 2011. Manganese at the center of the non-canonical activity of telomerase reverse transcriptase.
superoxide dismutase is a p53-regulated gene that switches cancers between Molecular Aspects of Medicine 31, 21–28.
early and advanced stages. Cancer Research 71, 6684–6695. Ishimoto, T., Nagano, O., Yae, T., Tamada, M., Motohara, T., Oshima, H., Oshima, M.,
Diebold, I., Petry, A., Djordjevic, T., Belaiba, R.S., Fineman, J., Black, S., Schreiber, Ikeda, T., Asaba, R., Yagi, H., Masuko, T., Shimizu, T., Ishikawa, T., Kai, K., Taka-
C., Fratz, S., Hess, J., Kietzmann, T., Gorlach, A., 2010. Reciprocal regulation of hashi, E., Imamura, Y., Baba, Y., Ohmura, M., Suematsu, M., Baba, H., Saya, H.,
Rac1 and PAK-1 by HIF-1alpha: a positive-feedback loop promoting pulmonary 2011. CD44 variant regulates redox status in cancer cells by stabilizing the xCT
vascular remodeling. Antioxidants & Redox Signaling 13, 399–412. subunit of system xc(−) and thereby promotes tumor growth. Cancer Cell 19,
Diehn, M., Cho, R.W., Lobo, N.A., Kalisky, T., Dorie, M.J., Kulp, A.N., Qian, D., Lam, 387–400.
J.S., Ailles, L.E., Wong, M., Joshua, B., Kaplan, M.J., Wapnir, I., Dirbas, F.M., Somlo, Iyer, N.V., Kotch, L.E., Agani, F., Leung, S.W., Laughner, E., Wenger, R.H., Gassmann,
G., Garberoglio, C., Paz, B., Shen, J., Lau, S.K., Quake, S.R., Brown, J.M., Weiss- M., Gearhart, J.D., Lawler, A.M., Yu, A.Y., Semenza, G.L., 1998. Cellular and devel-
man, I.L., Clarke, M.F., 2009. Association of reactive oxygen species levels and opmental control of O2 homeostasis by hypoxia-inducible factor 1 alpha. Genes
radioresistance in cancer stem cells. Nature 458, 780–783. & Development 12, 149–162.
388 V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390

Jacobs, J.J., Kieboom, K., Marino, S., DePinho, R.A., van, L.M., 1999. The oncogene and Lim, S.D., Sun, C., Lambeth, J.D., Marshall, F., Amin, M., Chung, L., Petros, J.A., Increased,
Polycomb-group gene bmi-1 regulates cell proliferation and senescence through Arnold R.S., 2005. Nox1 and hydrogen peroxide in prostate cancer. Prostate 62,
the ink4a locus. Nature 397, 164–168. 200–207.
Jeong, W.J., Yoon, J., Park, J.C., Lee, S.H., Lee, S.H., Kaduwal, S., Kim, H., Yoon, J.B., Choi, Liu, F., Wang, J., Chang, A.K., Liu, B., Yang, L., Li, Q., Wang, P., Zou, X., 2012. Fucoidan
K.Y., 2012. Ras stabilization through aberrant activation of Wnt/beta-catenin extract derived from Undaria pinnatifida inhibits angiogenesis by human umbil-
signaling promotes intestinal tumorigenesis. Science Signaling 5, ra30. ical vein endothelial cells. Phytomedicine 19, 797–803.
Jin, J.O., Song, M.G., Kim, Y.N., Park, J.I., Kwak, J.Y., 2010. The mechanism of fucoidan- Liu, J., Gu, X., Robbins, D., Li, G., Shi, R., McCord, J.M., Zhao, Y., 2009. Protandim
induced apoptosis in leukemic cells: involvement of ERK1/2, JNK, glutathione, a fundamentally new antioxidant approach in chemoprevention using mouse
and nitric oxide. Molecular Carcinogenesis 49, 771–782. two-stage skin carcinogenesis as a model. PLoS One 4, e5284.
Jose, C., Bellance, N., Rossignol, R., 2011. Choosing between glycolysis and oxida- Luo, Y., Zou, P., Zou, J., Wang, J., Zhou, D., Liu, L., 2011. Autophagy regulates ROS-
tive phosphorylation: a tumor’s dilemma. Biochimica et Biophysica Acta 1807, induced cellular senescence via p21 in a p38 MAPKalpha dependent manner.
552–561. Experimental Gerontology 46, 860–867.
Kang, K.W., Lee, S.J., Park, J.W., Kim, S.G., 2002. Phosphatidylinositol 3-kinase Maciag, A., Sithanandam, G., Anderson, L.M., 2004. Mutant K-rasV12 increases COX-
regulates nuclear translocation of NF-E2-related factor 2 through actin 2, peroxides and DNA damage in lung cells. Carcinogenesis 25, 2231–2237.
rearrangement in response to oxidative stress. Molecular Pharmacology 62, Malumbres, M., Barbacid, M., 2003. RAS oncogenes: the first 30 years. Nature
1001–1010. Reviews Cancer 3, 459–465.
Kayamori, K., Sakamoto, K., Nakashima, T., Takayanagi, H., Morita, K., Omura, K., Mani, S.A., Guo, W., Liao, M.J., Eaton, E.N., Ayyanan, A., Zhou, A.Y., Brooks, M., Rein-
Nguyen, S.T., Miki, Y., Iimura, T., Himeno, A., Akashi, T., Yamada-Okabe, H., Ogata, hard, F., Zhang, C.C., Shipitsin, M., Campbell, L.L., Polyak, K., Brisken, C., Yang,
E., Yamaguchi, A., 2010. Roles of interleukin-6 and parathyroid hormone-related J., Weinberg, R.A., 2008. The epithelial-mesenchymal transition generates cells
peptide in osteoclast formation associated with oral cancers: significance of with properties of stem cells. Cell 133, 704–715.
interleukin-6 synthesized by stromal cells in response to cancer cells. American Marconett, C.N., Morgenstern, T.J., San Roman, A.K., Sundar, S.N., Singhal, A.K., Fire-
Journal of Pathology 176, 968–980. stone, G.L., 2010. BZL101, a phytochemical extract from the Scutellaria barbata
Khor, T.O., Huang, M.T., Kwon, K.H., Chan, J.Y., Reddy, B.S., Kong, A.N., 2006. plant, disrupts proliferation of human breast and prostate cancer cells through
Nrf2-deficient mice have an increased susceptibility to dextran sulfate sodium- distinct mechanisms dependent on the cancer cell phenotype. Cancer Biology &
induced colitis. Cancer Research 66, 11580–11584. Therapy 10, 397–405.
Kim, H.M., Haraguchi, N., Ishii, H., Ohkuma, M., Okano, M., Mimori, K., Eguchi, H., Mateescu, B., Batista, L., Cardon, M., Gruosso, T., de, F.Y., Mariani, O., Nicolas, A.,
Yamamoto, H., Nagano, H., Sekimoto, M., Doki, Y., Mori, M., 2012. Increased CD13 Meyniel, J.P., Cottu, P., Sastre-Garau, X., Mechta-Grigoriou, F., 2011. miR-141 and
expression reduces reactive oxygen species, promoting survival of liver cancer miR-200a act on ovarian tumorigenesis by controlling oxidative stress response.
stem cells via an epithelial-mesenchymal transition-like phenomenon. Annals Nature Medicine 17, 1627–1635.
of Surgical Oncology 19, S539–S548. Mathe, E., Nguyen, G.H., Funamizu, N., He, P., Moake, M., Croce, C.M., Hussain, S.P.,
Kim, H.S., Patel, K., Muldoon-Jacobs, K., Bisht, K.S., Aykin-Burns, N., Pennington, 2012. Inflammation regulates microRNA expression in cooperation with p53
J.D., van der Meer, R., Nguyen, P., Savage, J., Owens, K.M., Vassilopoulos, A., and nitric oxide. International Journal of Cancer 131, 760–765.
Ozden, O., Park, S.H., Singh, K.K., Abdulkadir, S.A., Spitz, D.R., Deng, C.X., Gius, D., Matoba, S., Kang, J.G., Patino, W.D., Wragg, A., Boehm, M., Gavrilova, O., Hurley, P.J.,
2010a. SIRT3 is a mitochondria-localized tumor suppressor required for mainte- Bunz, F., Hwang, P.M., 2006. p53 regulates mitochondrial respiration. Science
nance of mitochondrial integrity and metabolism during stress. Cancer Cell 17, 312, 1650–1653.
41–52. Matsui, A., Ikeda, T., Enomoto, K., Hosoda, K., Nakashima, H., Omae, K., Watanabe,
Kim, Y.R., Oh, J.E., Kim, M.S., Kang, M.R., Park, S.W., Han, J.Y., Eom, H.S., Yoo, N.J., M., Hibi, T., Kitajima, M., 2000. Increased formation of oxidative DNA damage, 8-
Lee, S.H., 2010b. Oncogenic NRF2 mutations in squamous cell carcinomas of hydroxy-2 -deoxyguanosine, in human breast cancer tissue and its relationship
oesophagus and skin. Journal of Pathology 220, 446–451. to GSTP1 and COMT genotypes. Cancer Letters 151, 87–95.
Kim, J.W., Dang, C.V., 2006. Cancer’s molecular sweet tooth and the Warburg effect. Matsuzawa, A., Ichijo, H., 2008. Redox control of cell fate by MAP kinase: physi-
Cancer Research 66, 8927–8930. ological roles of ASK1-MAP kinase pathway in stress signaling. Biochimica et
Ko, Y.H., Pedersen, P.L., Geschwind, J.F., 2001. Glucose catabolism in the rabbit VX2 Biophysica Acta 1780, 1325–1336.
tumor model for liver cancer: characterization and targeting hexokinase. Cancer McCord, A.M., Jamal, M., Shankavaram, U.T., Lang, F.F., Camphausen, K., Tofilon,
Letters 173, 83–91. P.J., 2009. Physiologic oxygen concentration enhances the stem-like proper-
Kohle, C., Bock, K.W., 2007. Coordinate regulation of Phase I and II xenobiotic ties of CD133+ human glioblastoma cells in vitro. Molecular Cancer Research
metabolisms by the Ah receptor and Nrf2. Biochemical Pharmacology 73, 7, 489–497.
1853–1862. Mishima, Y., Matsumoto-Mishima, Y., Terui, Y., Katsuyama, M., Yamada, M., Mori,
Kondoh, H., Lleonart, M.E., Gil, J., Wang, J., Degan, P., Peters, G., Martinez, D., Carnero, M., Ishizaka, Y., Ikeda, K., Watanabe, J., Mizunuma, N., Hayasawa, H., Hatake, K.,
A., Beach, D., 2005. Glycolytic enzymes can modulate cellular life span. Cancer 2002. Leukemic cell-surface CD13/aminopeptidase N and resistance to apopto-
Research 65, 177–185. sis mediated by endothelial cells. Journal of the National Cancer Institute 94,
Kondoh, H., Lleonart, M.E., Bernard, D., Gil, J., 2007a. Protection from oxidative stress 1020–1028.
by enhanced glycolysis; a possible mechanism of cellular immortalization. His- Miska, E.A., Alvarez-Saavedra, E., Townsend, M., Yoshii, A., Sestan, N., Rakic, P.,
tology and Histopathology 22, 85–90. Constantine-Paton, M., Horvitz, H.R., 2004. Microarray analysis of microRNA
Kondoh, H., Lleonart, M.E., Nakashima, Y., Yokode, M., Tanaka, M., Bernard, D., expression in the developing mammalian brain. Genome Biology 5, R68.
Gil, J., Beach, D., 2007b. A high glycolytic flux supports the proliferative Molofsky, A.V., He, S., Bydon, M., Morrison, S.J., Pardal, R., 2005. Bmi-1 promotes neu-
potential of murine embryonic stem cells. Antioxidants & Redox Signaling 9, ral stem cell self-renewal and neural development but not mouse growth and
293–299. survival by repressing the p16Ink4a and p19Arf senescence pathways. Genes &
Kotsinas, A., Aggarwal, V., Tan, E.J., Levy, B., Gorgoulis, V.G., 2012. PIG3: A novel link Development 19, 1432–1437.
between oxidative stress and DNA damage response in cancer. Cancer Letters Mulholland, D.J., Kobayashi, N., Ruscetti, M., Zhi, A., Tran, L.M., Huang, J., Gleave, M.,
327, 97–102. Wu, H., 2012. Pten loss and RAS/MAPK activation cooperate to promote EMT and
Kovacic, P., Jacintho, J.D., 2001. Mechanisms of carcinogenesis: focus on oxidative metastasis initiated from prostate cancer stem/progenitor cells. Cancer Research
stress and electron transfer. Current Medicinal Chemistry 8, 773–796. 72, 1878–1889.
Krek, A., Grun, D., Poy, M.N., Wolf, R., Rosenberg, L., Epstein, E.J., MacMenamin, P., Nathan, C.F., Arrick, B.A., Murray, H.W., DeSantis, N.M., Cohn, Z.A., 1981.
da, P.I., Gunsalus, K.C., Stoffel, M., Rajewsky, N., 2005. Combinatorial microRNA Tumor cell anti-oxidant defenses. Inhibition of the glutathione redox cycle
target predictions. Nature Genetics 37, 495–500. enhances macrophage-mediated cytolysis. Journal of Experimental Medicine
Lapidot, T., Sirard, C., Vormoor, J., Murdoch, B., Hoang, T., Caceres-Cortes, J., Min- 153, 766–782.
den, M., Paterson, B., Caligiuri, M.A., Dick, J.E., 1994. A cell initiating human Nathan, C.F., Cohn, Z.A., 1981. Antitumor effects of hydrogen peroxide in vivo. Journal
acute myeloid leukaemia after transplantation into SCID mice. Nature 367, of Experimental Medicine 154, 1539–1553.
645–648. Nguyen, T., Nioi, P., Pickett, C.B., 2009. The Nrf2-antioxidant response element
Lebedeva, I.V., Su, Z.Z., Sarkar, D., Gopalkrishnan, R.V., Waxman, S., Yacoub, A., Dent, signaling pathway and its activation by oxidative stress. Journal of Biological
P., Fisher, P.B., 2005. Induction of reactive oxygen species renders mutant and Chemistry 284, 13291–13295.
wild-type K-ras pancreatic carcinoma cells susceptible to Ad.mda-7-induced Ni, W., Zhan, Y., He, H., Maynard, E., Balschi, J.A., Oettgen, P., 2007. Ets-1 is a
apoptosis. Oncogene 24, 585–596. critical transcriptional regulator of reactive oxygen species and p47(phox)
Lee, J.J., Kim, B.C., Park, M.J., Lee, Y.S., Kim, Y.N., Lee, B.L., Lee, J.S., 2011. PTEN gene expression in response to angiotensin II. Circulation Research 101,
status switches cell fate between premature senescence and apoptosis in 985–994.
glioma exposed to ionizing radiation. Cell Death and Differentiation 18, O’Brien, C.A., Pollett, A., Gallinger, S., Dick, J.E., 2007. A human colon cancer cell capa-
666–677. ble of initiating tumour growth in immunodeficient mice. Nature 445, 106–110.
Lee, K., Esselman, W.J., 2002. Inhibition of PTPs by H(2)O(2) regulates the activation Oh, J.Y., Giles, N., Landar, A., Darley-Usmar, V., 2008. Accumulation of 15-deoxy-
of distinct MAPK pathways. Free Radical Biology and Medicine 33, 1121–1132. delta(12,14)-prostaglandin J2 adduct formation with Keap1 over time: effects
Levine, R.L., 2002. Carbonyl modified proteins in cellular regulation, aging, and dis- on potency for intracellular antioxidant defence induction. Biochemical Journal
ease. Free Radical Biology and Medicine 32, 790–796. 411, 297–306.
Li, Y., He, K., Huang, Y., Zheng, D., Gao, C., Cui, L., Jin, Y.H., 2010. Betulin induces Omori, E., Inagaki, M., Mishina, Y., Matsumoto, K., Ninomiya-Tsuji, J., 2012. Epithe-
mitochondrial cytochrome c release associated apoptosis in human cancer cells. lial transforming growth factor beta-activated kinase 1 (TAK1) is activated
Molecular Carcinogenesis 49, 630–640. through two independent mechanisms and regulates reactive oxygen species.
Liao, D., Johnson, R.S., 2007. Hypoxia: a key regulator of angiogenesis in cancer. Proceedings of the National Academy of Sciences of the United States of America
Cancer and Metastasis Reviews 26, 281–290. 109, 3365–3370.
V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390 389

Oudard, S., Poirson, F., Miccoli, L., Bourgeois, Y., Vassault, A., Poisson, M., Magdelenat, cell line under normoxic conditions: involvement of PI-3K/Akt and MEK1/ERK
H., Dutrillaux, B., Poupon, M.F., 1995. Mitochondria-bound hexokinase as target pathways. Journal of Pathology 205, 530–536.
for therapy of malignant gliomas. International Journal of Cancer 62, 216–222. Simic, M.G., Bergtold, D.S., Karam, L.R., 1989. Generation of oxy radicals in biosys-
Park, J.S., Sharma, L.K., Li, H., Xiang, R., Holstein, D., Wu, J., Lechleiter, J., Naylor, S.L., tems. Mutation Research 214, 3–12.
Deng, J.J., Lu, J., Bai, Y., 2009. A heteroplasmic, not homoplasmic, mitochondrial Simone, N.L., Soule, B.P., Ly, D., Saleh, A.D., Savage, J.E., Degraff, W., Cook, J., Harris,
DNA mutation promotes tumorigenesis via alteration in reactive oxygen species C.C., Gius, D., Mitchell, J.B., 2009. Ionizing radiation-induced oxidative stress
generation and apoptosis. Human Molecular Genetics 18, 1578–1589. alters miRNA expression. PLoS One 4, e6377.
Pathi, S.S., Jutooru, I., Chadalapaka, G., Sreevalsan, S., Anand, S., Thatcher, G.R., Safe, Singh, A., Misra, V., Thimmulappa, R.K., Lee, H., Ames, S., Hoque, M.O., Herman, J.G.,
S., 2011. GT-094, a NO-NSAID, inhibits colon cancer cell growth by activation of Baylin, S.B., Sidransky, D., Gabrielson, E., Brock, M.V., Biswal, S., 2006. Dysfunc-
a reactive oxygen species-microRNA-27a: ZBTB10-specificity protein pathway. tional KEAP1-NRF2 interaction in non-small-cell lung cancer. PLoS Medicine 3,
Molecular Cancer Research 9, 195–202. e420.
Perez, A.T., Arun, B., Tripathy, D., Tagliaferri, M.A., Shaw, H.S., Kimmick, G.G., Cohen, Skinner, H.D., Zheng, J.Z., Fang, J., Agani, F., Jiang, B.H., 2004. Vascular endothelial
I., Shtivelman, E., Caygill, K.A., Grady, D., Schactman, M., Shapiro, C.L., 2010. growth factor transcriptional activation is mediated by hypoxia-inducible factor
A phase 1B dose escalation trial of Scutellaria barbata (BZL101) for patients 1alpha, HDM2, and p70S6K1 in response to phosphatidylinositol 3-kinase/AKT
with metastatic breast cancer. Breast Cancer Research and Treatment 120, signaling. Journal of Biological Chemistry 279, 45643–45651.
111–118. Smolkova, K., Bellance, N., Scandurra, F., Genot, E., Gnaiger, E., Plecita-Hlavata, L.,
Piao, L.S., Hur, W., Kim, T.K., Hong, S.W., Kim, S.W., Choi, J.E., Sung, P.S., Song, M.J., Jezek, P., Rossignol, R., 2010. Mitochondrial bioenergetic adaptations of breast
Lee, B.C., Hwang, D., Yoon, S.K., 2012. CD133+ liver cancer stem cells modu- cancer cells to aglycemia and hypoxia. Journal of Bioenergetics and Biomem-
late radioresistance in human hepatocellular carcinoma. Cancer Letters 315, branes 42, 55–67.
129–137. Soeda, A., Park, M., Lee, D., Mintz, A., Androutsellis-Theotokis, A., McKay, R.D., Engh,
Piccirillo, S.G., Reynolds, B.A., Zanetti, N., Lamorte, G., Binda, E., Broggi, G., Brem, H., J., Iwama, T., Kunisada, T., Kassam, A.B., Pollack, I.F., Park, D.M., 2009. Hypoxia
Olivi, A., Dimeco, F., Vescovi, A.L., 2006. Bone morphogenetic proteins inhibit promotes expansion of the CD133-positive glioma stem cells through activation
the tumorigenic potential of human brain tumour-initiating cells. Nature 444, of HIF-1alpha. Oncogene 28, 3949–3959.
761–765. Sohal, R.S., Weindruch, R., 1996. Oxidative stress, caloric restriction, and aging. Sci-
Plecita-Hlavata, L., Lessard, M., Santorova, J., Bewersdorf, J., Jezek, P., 2008. Mito- ence 273, 59–63.
chondrial oxidative phosphorylation and energetic status are reflected by Sotgia, F., Del, G.F., Casimiro, M.C., Bonuccelli, G., Mercier, I., Whitaker-Menezes,
morphology of mitochondrial network in INS-1E and HEP-G2 cells viewed by D., Daumer, K.M., Zhou, J., Wang, C., Katiyar, S., Xu, H., Bosco, E., Quong, A.A.,
4Pi microscopy. Biochimica et Biophysica Acta 1777, 834–846. Aronow, B., Witkiewicz, A.K., Minetti, C., Frank, P.G., Jimenez, S.A., Knudsen, E.S.,
Prince, M.E., Sivanandan, R., Kaczorowski, A., Wolf, G.T., Kaplan, M.J., Dalerba, P., Pestell, R.G., Lisanti, M.P., 2009. Caveolin-1−/− null mammary stromal fibro-
Weissman, I.L., Clarke, M.F., Ailles, L.E., 2007. Identification of a subpopulation of blasts share characteristics with human breast cancer-associated fibroblasts.
cells with cancer stem cell properties in head and neck squamous cell carcinoma. American Journal of Pathology 174, 746–761.
Proceedings of the National Academy of Sciences of the United States of America Tojo, T., Ushio-Fukai, M., Yamaoka-Tojo, M., Ikeda, S., Patrushev, N., Alexander, R.W.,
104, 973–978. 2005. Role of gp91phox (Nox2)-containing NAD(P)H oxidase in angiogenesis in
Pritchard, K.I., Julian, J.A., Holloway, C.M., McCready, D., Gulenchyn, K.Y., George, response to hindlimb ischemia. Circulation 111, 2347–2355.
R., Hodgson, N., Lovrics, P., Perera, F., Elavathil, L., O’Malley, F.P., Down, Tomiyama, A., Serizawa, S., Tachibana, K., Sakurada, K., Samejima, H., Kuchino, Y.,
N., Bodurtha, A., Shelley, W., Levine, M.N., 2012. Prospective study of 2- Kitanaka, C., 2006. Critical role for mitochondrial oxidative phosphorylation in
[18F]fluorodeoxyglucose positron emission tomography in the assessment the activation of tumor suppressors Bax and Bak. Journal of the National Cancer
of regional nodal spread of disease in patients with breast cancer: an Institute 98, 1462–1473.
Ontario clinical oncology group study. Journal of Clinical Oncology 30, Toullec, A., Gerald, D., Despouy, G., Bourachot, B., Cardon, M., Lefort, S., Richardson,
1274–1279. M., Rigaill, G., Parrini, M.C., Lucchesi, C., Bellanger, D., Stern, M.H., Dubois, T.,
Radisky, D.C., Levy, D.D., Littlepage, L.E., Liu, H., Nelson, C.M., Fata, J.E., Leake, D., Sastre-Garau, X., Delattre, O., Vincent-Salomon, A., Mechta-Grigoriou, F., 2010.
Godden, E.L., Albertson, D.G., Nieto, M.A., Werb, Z., Bissell, M.J., 2005. Rac1b and Oxidative stress promotes myofibroblast differentiation and tumour spreading.
reactive oxygen species mediate MMP-3-induced EMT and genomic instability. EMBO Molecular Medicine 2, 211–230.
Nature 436, 123–127. Tsuyada, A., Chow, A., Wu, J., Somlo, G., Chu, P., Loera, S., Luu, T., Li, X., Wu, X., Ye,
Rai, P., Young, J.J., Burton, D.G., Giribaldi, M.G., Onder, T.T., Weinberg, R.A., W., Chen, S., Zhou, W., Yu, Y., Wang, Y.Z., Ren, X., Li, H., Scherle, P., Kuroki,
2011. Enhanced elimination of oxidized guanine nucleotides inhibits onco- Y., Wang, S.E., 2012. CCL2 mediates crosstalk between cancer cells and stro-
genic RAS-induced DNA damage and premature senescence. Oncogene 30, mal fibroblasts that regulates breast cancer stem cells. Cancer Research 72,
1489–1496. 2768–2779.
Raj, L., Ide, T., Gurkar, A.U., Foley, M., Schenone, M., Li, X., Tolliday, N.J., Golub, T.R., Ushio-Fukai, M., Nakamura, Y., 2008. Reactive oxygen species and angiogenesis:
Carr, S.A., Shamji, A.F., Stern, A.M., Mandinova, A., Schreiber, S.L., Lee, S.W., 2011. NADPH oxidase as target for cancer therapy. Cancer Letters 266, 37–52.
Selective killing of cancer cells by a small molecule targeting the stress response Vaughn, A.E., Deshmukh, M., 2008. Glucose metabolism inhibits apoptosis in neu-
to ROS. Nature 475, 231–234. rons and cancer cells by redox inactivation of cytochrome c. Nature Cell Biology
Ramanathan, B., Jan, K.Y., Chen, C.H., Hour, T.C., Yu, H.J., Pu, Y.S., 2005. Resistance to 10, 1477–1483.
paclitaxel is proportional to cellular total antioxidant capacity. Cancer Research Venkataraman, S., Alimova, I., Fan, R., Harris, P., Foreman, N., Vibhakar, R., 2010.
65, 8455–8460. MicroRNA 128a increases intracellular ROS level by targeting Bmi-1 and inhibits
Reya, T., Morrison, S.J., Clarke, M.F., Weissman, I.L., 2001. Stem cells, cancer, and medulloblastoma cancer cell growth by promoting senescence. PLoS One 5,
cancer stem cells. Nature 414, 105–111. e10748.
Rhyu, D.Y., Yang, Y., Ha, H., Lee, G.T., Song, J.S., Uh, S.T., Lee, H.B., 2005. Role of reactive Veskoukis, A.S., Tsatsakis, A.M., Kouretas, D., 2012. Dietary oxidative stress and
oxygen species in TGF-beta1-induced mitogen-activated protein kinase acti- antioxidant defense with an emphasis on plant extract administration. Cell
vation and epithelial-mesenchymal transition in renal tubular epithelial cells. Stress and Chaperones 17, 11–21.
Journal of Americam Society of Nephrology 16, 667–675. Vizan, P., Alcarraz-Vizan, G., Diaz-Moralli, S., Solovjeva, O.N., Frederiks, W.M., Cas-
Ricci-Vitiani, L., Lombardi, D.G., Pilozzi, E., Biffoni, M., Todaro, M., Peschle, C., De, cante, M., 2009. Modulation of pentose phosphate pathway during cell cycle
M.R., 2007. Identification and expansion of human colon-cancer-initiating cells. progression in human colon adenocarcinoma cell line HT29. International Jour-
Nature 445, 111–115. nal of Cancer 124, 2789–2796.
Rosen, E.M., Fan, S., Pestell, R.G., Goldberg, I.D., 2003. BRCA1 gene in breast cancer. Wang, X.Q., Li, H., Van, P.V., Winn, R.A., Heasley, L.E., Nemenoff, R.A., 2009. Oncogenic
Journal of Cellular Physiology 196, 19–41. K-Ras regulates proliferation and cell junctions in lung epithelial cells through
Rossignol, R., Gilkerson, R., Aggeler, R., Yamagata, K., Remington, S.J., Capaldi, R.A., induction of cyclooxygenase-2 and activation of metalloproteinase-9. Molecular
2004. Energy substrate modulates mitochondrial structure and oxidative capac- Biology of the Cell 20, 791–800.
ity in cancer cells. Cancer Research 64, 985–993. Warburg, O., 1956. On the origin of cancer cells. Science 123, 309–314.
Rothe, F., Ignatiadis, M., Chaboteaux, C., Haibe-Kains, B., Kheddoumi, N., Majjaj, S., Weinberg, F., Chandel, N.S., 2009. Mitochondrial metabolism and cancer. Annals of
Badran, B., Fayyad-Kazan, H., Desmedt, C., Harris, A.L., Piccart, M., Sotiriou, C., the New York Academy of Sciences 1177, 66–73.
2011. Global microRNA expression profiling identifies MiR-210 associated with Weyemi, U., Lagente-Chevallier, O., Boufraqech, M., Prenois, F., Courtin, F., Caillou,
tumor proliferation, invasion and poor clinical outcome in breast cancer. PLoS B., Talbot, M., Dardalhon, M., Al, G.A., Bidart, J.M., Schlumberger, M., Dupuy,
One 6, e20980. C., 2012. ROS-generating NADPH oxidase NOX4 is a critical mediator in onco-
Rouleau, M., Patel, A., Hendzel, M.J., Kaufmann, S.H., Poirier, G.G., 2010. PARP inhi- genic H-Ras-induced DNA damage and subsequent senescence. Oncogene 31,
bition: PARP1 and beyond. Nature Reviews Cancer 10, 293–301. 1117–1129.
Sanchez-Arago, M., Chamorro, M., Cuezva, J.M., 2010. Selection of cancer cells with Wiemer, E.A., 2011. Stressed tumor cell, chemosensitized cancer. Nature Medicine
repressed mitochondria triggers colon cancer progression. Carcinogenesis 31, 17, 1552–1554.
567–576. Wise, D.R., DeBerardinis, R.J., Mancuso, A., Sayed, N., Zhang, X.Y., Pfeiffer, H.K., Nissim,
Seo, Y., Kinsella, T.J., 2009. Essential role of DNA base excision repair on survival in I., Daikhin, E., Yudkoff, M., McMahon, S.B., Thompson, C.B., 2008. Myc regulates a
an acidic tumor microenvironment. Cancer Research 69, 7285–7293. transcriptional program that stimulates mitochondrial glutaminolysis and leads
Shao, C., Folkard, M., Held, K.D., Prise, K.M., 2008. Estrogen enhanced cell-cell to glutamine addiction. Proceedings of the National Academy of Sciences of the
signalling in breast cancer cells exposed to targeted irradiation. BMC Cancer United States of America 105, 18782–18787.
8, 184. Won, K.Y., Lim, S.J., Kim, G.Y., Kim, Y.W., Han, S.A., Song, J.Y., Lee, D.K., 2012. Reg-
Shi, Y.H., Wang, Y.X., Bingle, L., Gong, L.H., Heng, W.J., Li, Y., Fang, W.G., 2005. In vitro ulatory role of p53 in cancer metabolism via SCO2 and TIGAR in human breast
study of HIF-1 activation and VEGF release by bFGF in the T47D breast cancer cancer. Human Pathology 43, 221–228.
390 V. Sosa et al. / Ageing Research Reviews 12 (2013) 376–390

Wulf, G.G., Wang, R.Y., Kuehnle, I., Weidner, D., Marini, F., Brenner, M.K., Andreeff, Ying, Q., Liang, L., Guo, W., Zha, R., Tian, Q., Huang, S., Yao, J., Ding, J., Bao, M., Ge,
M., Goodell, M.A., 2001. A leukemic stem cell with intrinsic drug efflux capacity C., Yao, M., Li, J., He, X., 2011. Hypoxia-inducible microRNA-210 augments the
in acute myeloid leukemia. Blood 98, 1166–1173. metastatic potential of tumor cells by targeting vacuole membrane protein 1 in
Yacoub, A., Mitchell, C., Hong, Y., Gopalkrishnan, R.V., Su, Z.Z., Gupta, P., Sauane, hepatocellular carcinoma. Hepatology 54, 2064–2075.
M., Lebedeva, I.V., Curiel, D.T., Mahasreshti, P.J., Rosenfeld, M.R., Broaddus, W.C., Yoshida, T., Oka, S., Masutani, H., Nakamura, H., Yodoi, J., 2003. The role of thioredoxin
James, C.D., Grant, S., Fisher, P.B., Dent, P., 2004. MDA-7 regulates cell growth in the aging process: involvement of oxidative stress. Antioxidants & Redox
and radiosensitivity in vitro of primary (non-established) human glioma cells. Signaling 5, 563–570.
Cancer Biology and Therapy 3, 739–751. Yu, R., Lei, W., Mandlekar, S., Weber, M.J., Der, C.J., Wu, J., Kong, A.N., 1999.
Yagoda, N., von, R.M., Zaganjor, E., Bauer, A.J., Yang, W.S., Fridman, D.J., Wolpaw, A.J., Role of a mitogen-activated protein kinase pathway in the induction of phase
Smukste, I., Peltier, J.M., Boniface, J.J., Smith, R., Lessnick, S.L., Sahasrabudhe, S., II detoxifying enzymes by chemicals. Journal of Biological Chemistry 274,
Stockwell, B.R., 2007. RAS-RAF-MEK-dependent oxidative cell death involving 27545–27552.
voltage-dependent anion channels. Nature 447, 864–868. Zhang, Z., Teruya, K., Eto, H., Shirahata, S., 2011. Fucoidan extract induces apoptosis
Yalcin, S., Marinkovic, D., Mungamuri, S.K., Zhang, X., Tong, W., Sellers, R., Ghaffari, in MCF-7 cells via a mechanism involving the ROS-dependent JNK activation and
S., 2010. ROS-mediated amplification of AKT/mTOR signalling pathway leads to mitochondria-mediated pathways. PLoS One 6, e27441.
myeloproliferative syndrome in Foxo3(−/−) mice. EMBO Journal 29, 4118–4131. Zhao, R., Xiang, N., Domann, F.E., Zhong, W., 2006. Expression of p53 enhances
Yau, C., Benz, C.C., 2008. Genes responsive to both oxidant stress and loss of estrogen selenite-induced superoxide production and apoptosis in human prostate can-
receptor function identify a poor prognosis group of estrogen receptor positive cer cells. Cancer Research 66, 2296–2304.
primary breast cancers. Breast Cancer Research 10, R61. Zong, W.X., Ditsworth, D., Bauer, D.E., Wang, Z.Q., Thompson, C.B., 2004. Alkyl-
Yin, X., Zhou, J., Jie, C., Xing, D., Zhang, Y., 2004. Anticancer activity and mechanism ating DNA damage stimulates a regulated form of necrotic cell death. Genes
of Scutellaria barbata extract on human lung cancer cell line A549. Life Sciences & Development 18, 1272–1282.
75, 2233–2244.

You might also like