You are on page 1of 15

Biomedicine & Pharmacotherapy 112 (2019) 108690

Contents lists available at ScienceDirect

Biomedicine & Pharmacotherapy


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

Cancer cells change their glucose metabolism to overcome increased ROS: T


One step from cancer cell to cancer stem cell?
Zahra Ghanbari Movaheda,1, Mohsen Rastegari-Pouyanib,1, Mohammad hossein Mohammadic,

Kamran Mansouria,d,
a
Medical Biology Research Center, Kermanshah University of Medical sciences, Kermanshah, Iran
b
Student Research Committee, Department of Immunology, School of Medicine, Shahid Beheshti University of Medical Sciences, Tehran, Iran
c
HSCT research center, Laboratory Hematology and blood Banking Department, School of Allied Medical Sciences, Shahid Beheshti University of Medical Science, Tehran,
Iran
d
Department of Molecular Medicine, School of Medicine, Kermanshah University of Medical Sciences, Kermanshah, Iran

A R T I C LE I N FO A B S T R A C T

Keywords: Cancer cells can adapt to low energy sources in the face of ATP depletion as well as to their high levels of ROS by
Cancer cell altering their metabolism and energy production networks which might also have a role in determining cell fate
Cancer stem cell and developing drug resistance. Cancer cells are generally characterized by increased glycolysis. This is while;
Glycolysis cancer stem cells (CSCs) exhibit an enhanced pentose phosphate pathway (PPP) metabolism. Based on the
ROS
current literature, we suggest that cancer cells when encountering ROS, first increase the glycolysis rate and then
Pentose phosphate pathway
following the continuation of oxidative stress, the metabolic balance is skewed from glycolysis to PPP. Therefore,
we hypothesize in this review that in cancer cells this metabolic deviation during persistent oxidative stress
might be a sign of cancer cells’ shift towards CSCs, an issue that might be pivotal in more effective targeting of
cancer cells and CSCs.

Abbreviations: ACOX1, acyl-CoA oxidase 1; ACOX2, acyl- CoA oxidase 2; ALD, aldolase; ALDH, aldehyde dehydrogenase; AML, acute myeloid leukemia; AMPK,
AMP-activated protein kinase; ARNT, aryl-hydrocarbon receptor nuclear translocator; ATM, ataxia telangiectasia mutated; BCKDH, Branched-chain alpha-keto acid
dehydrogenase; bHLH-PAS, basic helix-loop-helix/Per-Arnt-Sim; SBPase, sedoheptulose bisphosphatase; CSCs, cancer stem cells; DAO, D-amino acid oxidase; DDO, D-
aspartate oxidase; DHB, ethyl 3,4-dihydroxybenzoic acid; DHODH, dihydroorotate dehydrogenase; DMOG, dimethyloxallyl glycine; DUOX 2, dual oxidase 2; DUOX1,
dual oxidase 1; EMT, epithelial-mesenchymal transition; ENO, enolase; EPO, erythropoietin; ETC, electron transfer chain; ETFQOR, electron-transferring flavoprotein
ubiquinone oxidoreductase; F26-BP, fructose-2, 6-bisphosphate; F6P, fructose-6-phosphate; FA, fatty acid; FDH, folate dehydrogenases; FQR, ferredoxin quinone
reductase; G6P, glucose-6-phosphate; G6PD, glucose-6-phosphate dehydrogenase; GAPDH, Glyceraldehyde-3-phosphate dehydrogenase; GLUTs, glucose transpor-
ters; GPDH, glycerol-3-phosphate dehydrogenase; GPI, glucose-6-phosphate isomerase; GPX, glutathione peroxidase; GSH, glutathione; GSSG, glutathione disulfide;
GST, glutathione S-transferases; H2O2, hydrogen peroxide; HAO1, L-α-hydroxyacidoxidase 1; HAO2, L-α-hydroxyacidoxidase 2; HDAC, histone deacetylase; HIF-1,
hypoxia-inducible factor 1; HK, hexokinase; HMM, human malignant mesothelioma; iBCSCs, induced breast cancer stem cells; IDH, isocitrate dehydrogenases; iNOS,
inducible nitric oxide synthase; LDH, lactate dehydrogenase; LET, low linear energy transfer; MAOs, monoamine oxidases; ME, malic enzymes; mGPDH, mi-
tochondrial glycerol-3-phosphate dehydrogenase; MnSOD, manganese superoxide dismutase; mtDNA, mitochondrial DNA; MTHFD1, methylene tetrahydrofolate
dehydrogenase 1; MTHFD2L, MTHFD2-like; NGF, nerve growth factor; NOS, nitric oxide synthase; NOX1, NADPH oxidase 1; NOX2, NADPH oxidase 2; NOX4,
NADPH oxidase 4; O2%−, superoxide; OGDH, 2-oxoglutarate dehydrogenase; OGD, oxygen-glucose deprivation; OXPHOS, oxidative phosphorylation; PAOX, poly-
amine oxidase; PARP, poly(ADP-ribose) polymerase; PDH, pyruvate dehydrogenase; PDK-1, pyruvate dehydrogenase kinase 1; PEP, phosphoenolpyruvate; PFK1,
phosphofructokinase 1; PFKFB, 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatases; PG, phosphoglycerate; PGDH, phosphogluconate dehydrogenase; PGK,
phosphoglycerate kinase; PGL, phosphogluconolactonase; PHGDH, phosphoglycerate dehydrogenase; PHP, phosphopyruvate; PIKKs, phosphatidylinositol-3-kinase-
related protein kinases; PIPOX, L-pipecolic acid oxidase; PK, pyruvate kinase; PKL, liver-type pyruvate kinase; PKM2, pyruvate kinase M2; PKR, red blood cell
pyruvate kinase; PPP, pentose phosphate pathway; PSAT1, phosphoserine aminotransferase; P-Ser, phosphoserine; PSPH, phosphoserine phosphatase; RNS, reactive
nitrogen species; ROS, reactive oxygen species; RPE, ribulose 5-phosphate epimerase; RPI, ribose 5-phosphate isomerase; SAHA, suberoylanilide hydroxamic acid;
SHMT2, serine hydroxymethyltransferase

Corresponding author at: Kamran Mnasouri,Medical Biology Research Center, School of Medicine, Sorkheh Ligeh Blvd, P.O. Box 1568, Kermanshah, Iran.
E-mail address: kmansouri@kums.ac.ir (K. Mansouri).
1
The first two authors have equally contributed to this work.

https://doi.org/10.1016/j.biopha.2019.108690
Received 27 December 2018; Received in revised form 12 February 2019; Accepted 14 February 2019
0753-3322/ © 2019 The Authors. Published by Elsevier Masson SAS. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/BY-NC-ND/4.0/).
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

1. Introduction In this review, we first describe ROS and their sources as well as
control of intracellular ROS contents. In the next step we elaborate the
Many cancer cells exhibit persistently high reactive oxygen species role of glycolysis and PPP in coping with ROS and then proceed to the
(ROS) levels as a consequence of genetic, metabolic (higher metabolism possible molecular mechanisms involved in activating the Warburg
rate) and microenvironment-associated alterations [1]. effect and PPP in response to ROS. Finally, we discuss the possible
In cancer cells high ROS levels can originate from increased meta- molecular mechanisms of metabolic deviation from glycolysis to PPP, as
bolic activity, mitochondrial dysfunction, oncogene activity, peroxi- an appropriate target for more effective cancer treatment.
some activity, increased activity of oxidases, cyclooxygenases, lipox-
igenases, thymidine phosphorylase and increased cellular receptor 2. Reactive oxygen species (ROS) with an emphasis on hydrogen
signaling, or through crosstalk with infiltrating immune cells [2]. peroxide (H2O2)
As reported by Spitz and colleagues cancer cells produce greater
amounts of hydrogen peroxide (H2O2) and superoxide (O2%−) compared ROS is generally taken to include the initial species produced by
to normal cells [3,4]. oxygen reduction (hydrogen peroxide (H2O2) or superoxide (O2−%)) as
Cancer cell mitochondria may produce greater steady-state levels of well as their secondary reactive products. Also, reactive nitrogen spe-
H2O2 and O2%−, relative to normal cells [5]. cies (RNS) is in common usage to describe reactive species derived from
Moreover, glucose deprivation, as a tumor microenvironmental nitric oxide (NO) [30]. H2O2, as a key metabolite in oxidative stress, is
factor, induces oxidative stress in cancer cells [6]. believed to be a major player in ROS signaling owing to its physico-
Tumor suppressor and oncogenic pathways, which are frequently chemical properties, which include long half-life, a relatively low re-
mutated in cancer, commonly lead to increased accumulation of ROS activity and the ability to diffuse through membranes [31,32]. H2O2
[7–11]. Furthermore, conditions associated with tumorigenesis such as emerged as the main redox metabolite operative in redox sensing, sig-
hypoxia, inflammation, matrix detachment and mitochondrial dys- naling and redox regulation in physiological conditions (1–10 nM up to
function can all result in excess production of ROS [12–16]. approximately 100 nM). Supraphysiological concentrations of H2O2
In comparison to normal cells, which are characterized by lower (> 100 nM) result in damage of biomolecules, denoted as “oxidative
intracellular ROS levels, cancer cells have a higher antioxidant capacity distress” [33].
to counteract and scavenge ROS which represents a selective advantage Intracellular ROS produced under physiologic conditions serve as
and augments their survival under pro-oxidizing conditions, as high key signaling molecules regulating numerous cellular processes [34].
antioxidant capacity impairs cellular responses to anticancer therapy The abnormal accumulation of endogenous or exogenous ROS by the
and enhances cell survival [17,18]. mitochondrial electron transport chain (ETC) or in response to low
Studies have indicated that there is a close relationship between linear energy transfer (LET) ionizing radiation, like X-rays, may cause
cellular metabolism and redox homeostasis. Cancer cells use this con- protein denaturation, lipid peroxidation, or DNA mutation (oxidative
nection for cellular growth or to prevent oxidative damage. By other- distress) [35–37].
wise using substrates and metabolic intermediates in the biochemical ROS have both protumorigenic and antitumorigenic effects.
pathways which leads to the production of key antioxidant molecules, Escalated generation of ROS in tumor contributes to molecular and
malignant cells can directly support the mechanisms of ROS detox- biochemical alterations necessary for tumor initiation, promotion,
ification [14,19]. Studies have shown that acceleration of glycolysis; progression as well as chemo/radioresistance. ROS play play important
i.e., the Warburg effect; and pentose phosphate pathway (PPP) in roles in tumor cell signal transduction contributing to tumor cell sur-
cancer cells (human colon and breast cancer cell lines) could decrease vival, proliferation as well as tumor angiogenesis and metastasis. For
O2%− and H2O2 production [4] by less reliance on mitochondrial oxi- example, PI3K/AKT/mTOR survival signaling induced by ROS, results
dative phosphorylation (OXPHOS) and concomitantly augment the in the activation of mainly two signaling pathways: 1) Ras-MAPK
redox potential via an increase in NADPH [20,21]. leading to cell proliferation and 2) PI3K-Akt-eNOS which cause meta-
Warburg effect, which is a metabolic hallmark of most cancer cells, bolic modulation and cell survival. ROS also inactivates phosphatase
is characterized by an excessive conversion of glucose to lactate even in and tensin homolog (PTEN) pathway implicated in apoptosis initiation
the presence of ample oxygen [22]. On the other hand, intratumoral [38]. On the other hand, ROS can negatively regulate cellular FLICE-
heterogeneity, including those involving cancer stem cells (CSCs), is inhibitory protein (c-FLIP) by augmenting its proteasomal degradation
currently of interest in cancer research [23]. CSCs, characterized by the facilitating apoptosis. Also, H2O2 can iduce dimerization of apoptosis
ability to self-renew, are a small subset of cancer cells believed to be signal-regulating kinase 1 (ASK-1) resulting in its activation and sub-
responsible for tumor initiation, maintenance, growth, metastasis, re- sequent induction of apoptosis. ROS can also induce cytochrome-c (Cyt-
currence and drug resistance [24]. It is plausible that there exist dif- c) release and activate caspase-9 thereby inducing apoptosis [39].
ferences in metabolic state of heterogeneous tumors composed of dif- Tumor cells are characterized by an augmented metabolic activity
ferentiated cells and CSCs. CSCs have been reported to have an leading to changes of the cellular redox state that has to cope with the
enhanced PPP metabolism [25]. production of high levels of ROS [18]. In numerous cancer cells per-
Studies have shown that during oxidative stress, cancer cells will sistently upregulated H2O2-dependent signaling pathways are im-
shut down the glycolytic pathway and instead enhance glucose flux plicated in cell differentiation, growth and survival, yet high H2O2 le-
through the PPP in order to generate more NADPH for antioxidant vels can also induce cell cycle arrest or apoptosis in cells [40]. Human
defense [21]. tumor cells produce high levels of H2O2 [41]. The main antioxidant
PPP activity is increased in response to ionizing radiation, che- enzymes include superoxide dismutase (SOD), that converts O2−% into
motherapy or oxidative stress which increase ROS levels and trigger an H2O2, as well as catalase and glutathione peroxidase (GPX) which
adaptive response by augmenting the PPP [26]. Similarly, the popula- convert H2O2 into water [42]. It has been shown that SOD, via its
tion of CSCs increases in response to ionizing radiation, chemotherapy metabolic product H2O2, can induce migration, invasion, and epithelial-
or oxidative stress [27–29]. It has been documented that in multiple mesenchymal transition (EMT) (as a stem cell-like feature) of cancer
cancer cell lines, the acquisition of drug resistance, a feature of CSCs, is cells in pancreatic cancer cells through activation of the H2O2/ERK/NF-
accompanied by enhanced oxidative PPP activity [26]. κB axis [43].
Therefore, based on the conducted studies, we hypothesize that the Treatment of human malignant mesothelioma (HMM) cells with
metabolic deviation from glycolysis to PPP in the encounter of persis- H2O2 enhances EMT, as indicated by increased expression levels of vi-
tent oxidative stress, might represent alterations in cancer cells towards mentin, TWIST1 and SLUG and decreased E-cadherin expression. Also,
CSCs. expression of stemness genes such as oct4, sox2 and nanog significantly

2
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

Fig. 1. Sites for the generation of O2−% and H2O2 in mitochondria.

increases following treatment with H2O2 [44]. ROS is facilitated through antioxidant enzymes (i.e. SODs, peroxir-
Therefore, regarding the importance of H2O2 in cancer biology, in edoxins, catalases and glutathione system) specifically scavenging dif-
the current review we aimed to emphasize on this specific kind of ROS. ferent kinds of ROS or by non-enzymatic molecules (i.e. flavonoids,
glutathione, and vitamins A, C and E) [2]. As earlier mentioned
3. Sources of ROS [4,15,19–21], several biochemical pathways play a role in ROS detox-
ification, such as glycolysis and PPP. Here we proceed to discuss these
ROS can be found in the environment [45–48] such as in pollutants, two pathways.
iron salts, radiation, and tobacco smoke or can be generated inside
[48,49] or outside the cells [50–53] which can occur through multiple 4.1. Glycolysis and ROS detoxification
mechanisms.Generally, continuous production of O2−% and H2O2 takes
place in the mitochondria as the most important source of cellular ROS. 4.1.1. Glycolysis
This is the result of the electron transport chain (ETC) located in the Glycolysis occurs in all cells of the body. The enzymes involved in
mitochondrial membrane that is essential for the production of energy this pathway exist in the cytosomal fraction of the cell. Glycolysis can
within the cells (54,55). O2−% and H2O2 are produced in the mi- occur both in the absence of oxygen (anaerobic) and in the presence of
tochondria as by-products of oxidative phosphorylation for ATP oxygen (aerobic). Under anaerobic condition, lactate is the end product
synthesis and fatty acid (FA) metabolism [56,57]. of glycolysis. In the aerobic condition, pyruvate is generated, which is
In mitochondria, several proteins involved in glycolysis, β-oxida- then oxidized to H2O and CO2 [61].
tion, mitochondrial electron transport, and the TCA cycle are able to Two ATP molecules are synthesized in the anaerobic glycolysis
generate O2−%, H2O2 and/or other ROS (Fig. 1). while under aerobic conditions ˜seven ATPs are synthesized from glu-
Several oxidases localized in specialized stable (peroxisomes) or cose conversion to pyruvate [62,63].
transient (endosomes, phagosomes, multivesicular bodies, etc.) single-
membrane bound organelles are capable of generating substantial 4.1.2. Warburg effect
amounts of H2O2 [57]. A key change that cancer cells have to make in order to meet their
Mammalian peroxisomes play a pivotal role in various metabolic modified energy demands is to reprogram their cellular metabolism,
pathways such as fatty acid α- and β-oxidation, glyoxylate metabolism, one of aspects of which is to switch to an increased glycolysis pathway
ether-phospholipid biosynthesis, polyamine oxidation, amino acid cat- in spite of having access to enough oxygen [64].
abolism, and the oxidative part of PPP. Interestingly, many of the en- The metabolism of glucose is essential for supporting all mammalian
zymes involved in these pathways produce specific ROS or RNS as by- life. In mammals, the end product can be lactate or CO2 which is gen-
products of their normal catalytic function [47]. erated upon full oxidation of glucose via respiration in mitochondria. In
H2O2 is the main oxidant product of xanthine oxidase (XO) [58]. tumor cells and other developing or proliferating cells, glucose uptake
Reports regarding XO-derived ROS generation frequently address XO as rate is outstandingly augmented and even in the presence of oxygen and
the O2−%-producing form of xanthine dehydrogenase (XDH) and H2O2 fully functioning mitochondria, lactate is produced. This process is re-
is generated as a secondary byproduct of spontaneous or enzymatic cognized as the Warburg effect [65].
dismutation of O2−% [59]. In 1923, Otto Warburg for the first time reported an anomalous
ROS originating from cytosol could be delivered into extracellular characteristic of energy metabolism in cancer cells [66,67]. Warburg
space (free extracellular ROS) in different ways including: secretion as observed that cancer cells, even under fully oxygenated conditions,
typically occurs in the case of activated degranulating leukocytes, were characterized by accelerated glycolysis and excessive lactate for-
crossing the plasma membrane through some anion channels (O2−%) mation [68] Later in 1972, his discovery was named the ‘Warburg Ef-
and aquaporins (H2O2) as well as delivery via exosome release [60]. fect’ by Efraim Racker [69].
Respiration, generating some 32/30 mol of ATP per mole of glucose
4. Control of the intracellular concentrations of ROS [63,70], is downregulated in cancer cells in favor of fermentation which
only produces two moles of ATP per mole of glucose. Therefore, to
Under normal physiological conditions, intracellular ROS levels are supply sufficient ATP, the Warburg effect is associated with an in-
steadily maintained to prevent cells from damage. Detoxification from creased uptake of glucose (Fig. 2). Using this low ATP-yielding pathway

3
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

Fig. 2. Difference in glucose metabolic pathways in the presence and absence of oxygen in normal vs cancer cells.

by cancer cells seems amazing and several advantages of the Warburg thioredoxins and glutathione which in turn are used (directly or in-
effect for cancer cells have been proposed [71]. directly) by glutaredoxins, peroxiredoxins and glutathione peroxidases
Moreover, glycolysis upregulation is not only for ATP synthesis but to neutralize ROS. Therefore, for a large number of ROS-detoxifying
also for synthesis of biomass including ribonucleotides [72] and amino enzymes, NADPH acts as the ultimate donor of reductive power [79].
acids [70] for proliferation and growth in an ever-changing micro- Several metabolic pathways such as the reactions catalyzed by the
environment under several material limitations like shortages of oxygen malic enzymes (ME), folate dehydrogenases (FDH) and isocitrate de-
and nutrients. Upregulation of glycolysis could also be considered as a hydrogenases (IDH) can generate NADPH; but the main sources of
way to reduce ROS and therefore oxidative stress in cancer cells cellular NADPH are two enzymes of the oxidative branch of PPP: the
[73,74]. PPP rate-limiting enzyme G6PD and 6PGDH [80].

4.1.3. Warburg effect: coping with ROS 4.3. Mechanisms of activating the Warburg effect in response to ROS
In most mammalian cells, mitochondria are an important source of (DIRECT)
H2O2 and O2−% [53]. Upregulation of glycolysis rate in cancer cell
energy metabolism could decrease the production of H2O2 and O2−% by 4.3.1. GLUT1 and ROS
less reliance on mitochondrial OXPHOS (Fig. 3) [75] In order to involve in the Warburg effect, the cancer cells utilize
much more glucose compared to normal cells which derives most of
4.2. PPP and ROS detoxification their energy from aerobic metabolism [81]. Accordingly the ability of
such cells to bring glucose into their cytoplasm needs to be increased.
4.2.1. Pentose phosphate pathway (PPP) Various glucose transporters (GLUTs), primarily GLUTs 1, 3, and 4,
PPP is classified into two biochemical branches known as the oxi- have been shown to be upregulated in endometrial cancer, ovarian
dative and non-oxidative PPP. Non-oxidative PPP reactions, that occur cancer, gastric cancer, meningiomas, glioblastomas and squamous cell
virtually ubiquitously, play a central metabolic role by providing the carcinomas as well as in many of these cell lines [82]. GLUT-1, an
RNA backbone precursors erythrose 4-phosphate and ribose 5-phos- important member of GLUT family, is a basic high-affinity glucose
phate as precursors for aromatic amino acids. On the contrary, oxida- transporter normally expressed in erythrocytes, renal tubules, placenta
tive branch reactions of PPP are not universal and are absent in many and endothelial cells. GLUT-1 expression has been shown to be asso-
aerobic and thermophilic organisms [76–78]. However, the oxidative ciated with tumor aggressiveness and poor prognosis in many cancers
branch, being highly active in most of the eukaryotes, via the successive including lung, breast, stomach and kidney [83].
reactions of G6PD, 6-phosphogluconolactonase (6PGL) and 6-phos- Various signaling pathways are involved in glucose uptake stimu-
phogluconate dehydrogenase (6PGDH) converts the glycolytic meta- lation as an adaptive response during oxidative stress. ROS-induced
bolite glucose 6-phosphate (G6P) into ribulose 5-phosphate. This me- signals, by upregulating GLUT protein expression, induce glucose up-
tabolic sequence produces two NADPH molecules per each glucose 6- take. The rapid induction of glucose uptake can occur at the level of
phosphate metabolized [79]. GLUT translocation and regulation of GLUT intrinsic activity.
Interestingly, a number of the signaling proteins are ROS-sensitive
4.2.2. PPP: coping with ROS (such as AMP-activated protein kinase (AMPK) and thioredoxin-inter-
NADPH, in its reduced form, is the donor of reductive potential to acting protein (TXNIP)). This indicates that glucose uptake stimulation

4
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

Fig. 3. Warburg effect reduces the production of ROS by less reliance on mitochondrial OXPHOS. Pyruvate is a source for OXPHOS. Mitochondrial OXPHOS is a main
source of ROS production. By increasing the conversion of pyruvate to lactate, Warburg effect reduces ROS production by the mitochondrial OXPHOS.

induced by ROS is part of a/an (adaptive) mechanism triggered by genes with different functions have been identified as the target genes
oxidative and/or metabolic stress [54]. Translocation of GLUT1 to the activated by HIF-1 [90]. On the other hand, HIF-1α activity and
plasma membrane is regulated by ataxia telangiectasia mutated (ATM) accumulation are themselves found to be regulated by different
protein [84]. This protein is a member of the phosphatidylinositol-3- cellular levels of many other factors including Hsp90, TSGA10, etc.
kinase-related protein kinases (PIKKs) family playing an important role [91,92].
in the response to DNA damage after which it becomes phosphorylated.
However, ATM can also localize to the cytosol where following acti-
4.4.2. HIF-1 regulates glycolysis and the Warburg effect
vation by ROS, it is involved in cytosolic signaling [85]. ATM, following
GLUT1 and GLUT3 are needed for glucose shuttling across cell
thiol oxidation by ROS, forms an active dimer of two covalently linked
membranes. Due to the high affinity of GLUT1 and GLUT3 for glucose,
monomers. Activated ATM localizes near the mitochondria which are
expression of these two molecules guarantees efficient glucose uptake
necessary for ROS-induced ATM activation. This suggests that mi-
even in the conditions of glucose shortage. Both of the two isoforms
tochondrial ROS stimulate ATM dimerization and activation. There are
have a HIF-1-inducible gene expression coupling the glycolytic switch
evidence regarding the activation of ATM by ROS during treatment
to increased uptake of glucose in hypoxic cancer cells [93].
with the chemotherapeutic agent doxorubicin [54].
Hexokinase (HK) 2 that catalyses the first rate-limiting reaction of
glycolysis (glucose phosphorylation to G6P) involving phosphate
4.4. Mechanisms of activating the Warburg effect in response to ROS transfer from ATP, is a member of the HK family of enzymes. The ne-
(INDIRECT) gatively charged G6P is trapped inside the cell and can fuel both gly-
colysis and PPP. In comparison with other isoforms of HK, HK2 is a HIF-
4.4.1. Hypoxia-inducible factor 1 (HIF-1) and ROS 1 target gene product [94].
4.4.1.1. HIF-1. Hypoxia-inducible factor 1 (HIF-1) is a member of a Fructose-2,6-bisphosphate (F2,6-BP) is a key glycolysis regulator
family of heterodimeric transcription factors which contain a bHLH- which serves as an allosteric activator of phosphofructokinase 1 (PFK1),
PAS (basic helix-loop-helix/Per-Arnt-Sim) domain. HIF-1, with a one of the rate-controlling enzymes of glycolysis. F2,6-BP is generated
ubiquitous expression, is the key player in regulating hypoxic genes from fructose-6-phosphate (F6P) by the action of a family of homo-
and mediates the upregulation of hypoxia-inducible genes [86,87]. HIF- dimeric enzymes known as 6-phosphofructo-2-kinase/fructose-2,6-bi-
1 comprises an α-subunit isoform (HIF-1 α, HIF-2 α or HIF-3 α) and a sphosphatases (PFKFB). The family consists of four members among
common β-subunit (HIF-1 β) identical with aryl-hydrocarbon receptor which PFKFB1, PFKFB2, and PFKFB4 exhibit equal PFK2 (ATP-depen-
nuclear translocator (ARNT). Both of these subunits are constitutively dent phosphorylation of F6P to F2,6-BP) and FBPase (depho-
expressed in the cell with the α-subunit being oxygen-labile while the β- sphorylation of F2,6-BP to F6P) activities under basal conditions, while
subunit is not sensitive to changes in oxygenation [88]. Target genes of PFKFB3 has high PFK2 and almost no FBPase activity [94,95,96]. Hy-
HIF-1 are implicated in angiogenesis (VEGF or vascular endothelial poxia can induce the transcription of all four PFKFB genes but the main
growth factor), glucose metabolism and glucose transport (GLUT-1, induction is seen for the PFKFB3 gene which is a target of HIF-1
glycolytic enzymes), iron transport (transferrin, transferrin receptor) [97,98]. PFKFB3 that is highly expressed in several tumor types, sus-
and hematopoiesis (erythropoietin) and aim for the adaptation to tains high-rate glycolysis [93].
hypoxia on systemic, local and cellular levels [89]. More than 100 Pyruvate dehydrogenase (PDH), the enzyme which commits

5
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

pyruvate to enter into the Krebs cycle, is inhibited by phosphorylation mechanism by which aerobic glycolysis, i.e. the Warburg effect arises
via PDK-1 (pyruvate dehydrogenase kinase 1). Under hypoxic condi- and it is suggested that the Warburg effect is related to an inherently
tions, PDH inhibition follows two main objectives: 1) orienting pyr- high cellular ROS and HIF-1 levels [112].
uvate to the lactate dehydrogenase (LDH) 5 reaction to produce NAD+
and 2) preventing the excessive production of ROS by uncoupled mi- 4.5. Mechanisms of activating PPP in response to ROS
tochondria [99,100]. HIF-1α also hinders mitochondrial respiration
through inducing PDK-1 which is a PDH inhibitor [99]. High expression In response to ROS, cancer cells might initially upregulate glycolysis
of PDK-1 strongly correlates with poor outcome in head-and-neck and then following persistent ROS exposure the metabolic balance is
cancer [101]. skewed from glycolysis pathway to PPP to generate more NADPH for
LDHs are comprised of up to four copies of two different subunits: antioxidant defense.
subunit LDH-H, ubiquitously expressed in healthy tissues, is encoded by PK is a regulator for cellular anti-oxidative metabolism. PKM2 in-
the ldhb gene, while subunit LDH-M is encoded by the HIF-1-target gene hibition by ROS contributes to cellular antioxidant responses [73].
ldha and is hence induced by hypoxia. In comparison with LDH-H, LDH- However, this protein plays a unique regulatory role so that its de-
M has a higher Km for pyruvate and a higher Vmax for pyruvate re- creased catalytic activity is associated with tumor progression
duction [101,102]. As a result, LDH5/LDH-4 M preferentially catalyzes [113,114]. When the kinase activity of PK is low — as in cancer cells—
the reduction of pyruvate into lactate and plays pivotal roles in re- its substrate, phosphoenol pyruvate, accumulates which subsequently
sistance to apoptosis and in the maintenance of a high glycolytic flux. inhibits the glycolytic enzyme triose phosphate isomerase (TPI) re-
Augmented expression of LDH5 has an unfavorable prognostic sig- sulting in the activation of a pathway alternative to glycolysis i.e., PPP
nificance for many human tumors [103–105]. Conversely, silencing of [115,116]. Augmented PPP activity protects the cells against ROS in at
LDH1/LDH-4H is most commonly observed in glycolytic cancer cells, a least two ways. First, it provides NADPH, a reducing molecule required
process which involves hypermethylation of the ldhb gene promoter for antioxidant enzymes activity and for the recycling of the antioxidant
[106,107]. peptide glutathione (GSH). Also, NADPH compensates for the redox
Pyruvate kinase (PK) is one of the rate-limiting enzymes of the imbalance caused by increased synthesis of nucleotides and fatty-acids
glycolysis pathway which catalyzes the irreversible transpho- [117]. Second, PPP regulates gene expression in favor of adaptation to
sphorylation between phosphoenolpyruvate (PEP) and adenosine di- oxidative stress [118]. Several types of oxidative stresses, such as H2O2,
phosphate (ADP) producing pyruvate and ATP. This reaction is a hypoxia and diamide inactivate PKM2 in human cancer cells. Anasta-
committed step which results in glycolysis or oxidative phosphorylation siou D et al. showed that in A549 human lung cancer cells exposure to
of pyruvate. In mammals, the PK family comprises of four isoforms: red H2O2 (1 mM), diamide (250 μM), as a thiol-oxidizing compound, or
blood cell PK (PKR); liver-type PK (PKL); and PK muscle isozyme M1 hypoxia (1% O2) [which causes increased ROS generation by mi-
and M2 (PKM1 and PKM2, respectively). PKM1 and PKM2 are both tochondrial complex III [119]] was associated with increased in-
generated via alternative splicing of the primary RNA transcript of the tracellular concentrations of ROS leading to decreased PKM2 activity.
pkm gene. The constitutively active PKM1 is the non-allosteric isoform, In that study, acute increases of intracellular ROS levels in human lung
expressed in terminally differentiated tissues requiring a large ATP cancer cells caused inhibition of the glycolytic enzyme PKM2 through
supply. By contrast, PKM2, which is subject to complex allosteric reg- oxidation of Cys358. This inhibition of PKM2 is required to divert
ulation, is expressed in proliferating cells and cancer cells which have glucose flux into PPP thereby generating sufficient reducing potential
anabolic functions [108]. HIF-1 activates pyruvate kinase M2 (PKM2) for detoxification of ROS [73].
expression and on the other hand, PKM2 functions as a transcriptional Glyceraldehyde-3-phosphate dehydrogenase (GAPDH), which is an
coactivator for HIF-1α. HIF-1 activates PKM2 transcription leading to evolutionarily conserved enzyme, controls glucose flux through the
the production of a protein which interacts with HIF-1α to stimulate canonical Embden-Meyerhof glycolytic pathway. GAPDH, during pro-
coactivator recruitment, chromatin binding, and transcriptional acti- oxidant conditions, contributes to a glycolytic bottleneck favoring a
vation. This feed-forward loop augments expression of HIF-1 target metabolic switch towards PPP [120].
genes which code for proteins mediating the metabolic reprogramming GAPDH inhibition helps divert glycolytic flux into the oxidative
of cancer cells (e.g., LDH-M and PDK1) [109]. Sun et al showed that phase of PPP by allowing metabolites to accumulate upstream of the
PKM2 is a key glycolytic enzyme in the oncogenic mTOR-induced inhibition point consistent with the observed induction of PPP enzymes
Warbug effect, in which c-Myc-hnRNP and HIF-1α cascades act as the following H2O2 treatment [121,122].
transducers of mTOR regulation of PKM2 [108]. In mammalian cells (P388D1 cells), GAPDH can be inhibited within
minutes of exposure to H2O2 (100 μM) mainly via direct inactivation of
4.4.3. ROS regulates HIF-1 the enzyme and loss of NAD + cofactor likely through PARP (Poly
HIF-1 plays a key role in the cellular adaptive response to hypoxia, (ADP-ribose) polymerase) activation [123,124]. The active site cysteine
which is closely related to pathophysiological conditions, including of GAPDH is highly sensitive to inhibitory oxidative modifications of
cancer [110]. ROS induced by a direct application of H2O2 (0.5 mM) RNS and ROS [125].
and menadione (100 μM) to the human prostate carcinoma cell line, TIGAR, or TP53-induced glycolysis and apoptosis regulator, is
DU145, leads to HIF-1α protein accumulation through attenuating its identified as a P53 target gene induced by ionizing radiation. TIGAR
degradation and activating its transcriptional activity in an AMPK-de- has a single bisphosphatase (BPase) activity which degrades F-2,6-BP to
pendent manner. Inhibition of AMPK increases HIF-1α ubiquitination F-6-P. TIGAR-mediated decrease in F-2,6-BP levels, inhibits glycolytic
under oxidative stress conditions. Also, HIF-1 regulation by AMPK in flux downstream of PFK1. PFK1 inhibition results in the accumulation
response to H2O2 is under the control of Janus kinase 2 and c-Jun N- of G6P and F6P pools as their consumption is greatly diminished. The
terminal kinase pathways. Collectively, AMPK can be considered as a increased G6P can flow into the oxidative arm of the PPP to produce
pivotal determinant of HIF-1 functions in response to H2O2 and might NADPH. UV and ROS stress can induce P53-dependent TIGAR activa-
play a role in the sophisticated HIF-1 regulatory mechanisms (Fig. 4) tion which inhibits PFK1 [125].
[111].
Hepatoma cells have higher free radicals levels compared to im- 5. In response to increased ROS, which one do the cancer cells
mortalized normal hepatocyte cells. In hepatoma cells, increased level increase? Glycolysis or PPP
of ROS stress (SOD and XO gene transfection were used to produce
various cellular redox levels) can directly upregulate HIF-1 and induce Studies have also shown that increased glycolysis and PPP in cancer
glycolysis without requiring a hypoxic condition. This explains the cells could reduce the production of ROS [20,21].

6
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

Fig. 4. HIF-1 mediated effects of ROS (H2O2) on upregulation of


glycolysis. A number of ROS can activate HIF-1α. Activated HIF-
1α upregulates glucose transporters GLUT1 and GLUT3 to in-
crease glucose uptake into cancer cells. HIF-1α accelerates gly-
colysis by upregulating ENO, ALD, PGK1 and PKM2 and meta-
bolizing pyruvate. Pyruvate is not converted to acetyl CoA
because HIF-1α upregulates PDK-1, which inhibits PDH.

Activation of oxidative PPP, through increasing NADPH production phosphorylation resulting in antioxidant status (increased GSH and
and thus enhanced intracellular redox power of cancer cells, can help reduced ROS) of CSCs. Low PKM2 activity, by inhibiting the conversion
transformed cells escape oxidative stress [126]. The increased PPP in of pyruvate to lactate, promotes the flow of glycolytic intermediates
cancer cells produces high levels of NADPH to reduce ROS while con- into PPP. Also, cancer stem-like cells expressing high levels of CD44
comitantly producing high nucleotide levels for DNA synthesis and variant 8–10 (CD44v) do not show accumulation of intracellular ROS
repair. These activities of PPP might induce resistance to certain cancer and are protected from redox stress. CD44v stabilizes cysteine/gluta-
therapies which enhance oxidative stress or DNA damage [26]. mate antiporter at the cellular membrane which promotes the synthesis
As demonstrated by several studies, alterations of PPP considerably of GSH. On the other hand, cancer cells with high PKM2 level/activity
contribute to tumor growth and survival under certain stress condi- use a c-Myc-mediated strategy to preferentially splice the M2 isoform
tions. During oxidative stress, cancer cells are able to cope with oxi- over M1. It has been found that cells with high c-Myc activity also show
dative stress by redirecting the metabolic flux from glycolysis to PPP high PKM2/PKM1 ratios. A great deal of evidence indicates that c-Myc
[21,127]. stimulates glycolysis and is required to coordinate with HIF-1 for the
Therefore, cancer cells might initially respond to ROS by increasing regulation of cellular responses to hypoxia. Therefore, high levels of c-
glycolysis rate and then in response to increased ROS level/exposure, Myc activity are responsible for the increased PKM2/PKM1 ratios in
they increase PPP. cancer cells. C-Myc, in combination with HIFs, can promote glycolysis
by upregulating GLUT1, HK2 and PDK1. It has been shown that in
gastric cancer cells, CD44v and c-Myc exhibit an inversed expression
6. Glycolysis deviation to PPP: from cancer cell to cancer stem cell
manner. FBW7, a ubiquitin ligase component, can mediate ubiquitin-
proteasome-dependent degradation of c-Myc. Tumor tissue at the in-
Cancer stem cells (CSCs) are a small subset of cancer cells capable of
vasive front is regarded to be composed of heterogeneous cellular po-
self-renewing and initiating tumor growth. These cells are implicated in
pulations including proliferative cancer stem-like cells (CD44vhigh,
cancer initiation, maintenance, metastasis and recurrence (29).
FBW7low, c-Mychigh) and dormant cancer stem-like cells (CD44vhigh,
Traditional therapies of cancer such as chemo- and radiotherapy have
FBW7high, c-Myclow). Accordingly, CSCs with high expression of CD44v
several limitations that result in treatment failure and cancer recur-
and FBW7 and low expression of c-Myc tend to be quiescent (G0/G1
rence.
dormant phase) [134–136].
The Warburg effect is thought to be a common phenomenon in all
It has been determined that histone deacetylase (HDAC) inhibitors,
tumor cells. However, there seems to exist differences regarding me-
by reprogramming differentiated cancer cells into stem-like cells, pro-
tabolic state in heterogeneous tumors composed of differentiated cells
mote CSC expansion. In a study, Debeb et al. investigated metabolic
and CSCs. There are contradictory literature data regarding the meta-
differences between stem-like breast cancer cells and differentiated
bolic states of CSCs and their differentiated progeny. According to some
cancer cells. According to the results of their study, HDAC inhibitor-
studies comparing the metabolic state of leukemia, glioma, and breast
induced stem-like cancer cells exhibit an enhanced PPP metabolism.
CSCs with the differentiated cell population, CSCs are more dependent
PPP plays a substantial role in the production of cellular NADPH which
on oxidative phosphorylation than differentiated cells [128–130], while
is required for fatty acids synthesis and intracellular ROS detoxification.
some other studies point to more dependence of these cells on anae-
As expected, aldehyde dehydrogenase− (ALDH−) or ALDH + cells
robic glucose metabolism [131–133].
treated with valproic acid (VA) or suberoylanilide hydroxamic acid
CD44 expression by CSCs could be considered as one of the evidence
(SAHA) (as HDAC inhibitors), had significantly higher levels of NADPH
showing the metabolic deviation from glycolysis to PPP in such cells.
compared to the vehicle-treated cells [25]. Also in a study conducted by
CD44, as one of CSC markers of various kinds of solid tumors, is cate-
La Noce et al., HDAC2 depletion was shown to promote stemness in
gorized into standard (without variables exons, or CD44 s) and variant
osteosarcoma cell lines. They investigated the effects of valproic acid
(with variable exons) isoforms. Intracellular domain of CD44 s has been
(VPA), an HDAC inhibitor, and 5′azacytidine (DAC), a demethylating
shown to interact with and suppress PKM2 through increasing its

7
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

agent, on the stem phenotype of two osteosarcoma cell lines MG63 and increasing the production of ROS [144]. It has been shown that IR can
Saos2. They found an increased expression of the stemness markers: induce reprogramming of differentiated cancer cells into CSCs [150]. In
CD133, OCT4, SOX2 and NANOG as well as increased sarcospheres and patients with prostate cancer, radiotherapy elevates the CD44+ cells
colonies formation in soft agar (as a tumor stemness characteristic) population that show CSC properties [151]. Also, IR induces the re-
following treatment with VPA and DAC alone or their combination. expression of stem cell regulators like OCT4, SOX2, NANOG, and KLF4,
Also, they performed an analysis of histone modifications previously to promote stemness in cancer cells [144].
shown to be implicated in embryonic stemness and interestingly found Induced breast cancer stem cells (iBCSCs) have been found to be
that treatments with VPA and DAC alone and in combination, induced a generated by radiation. Dr. Pajonk and colleagues, after eliminating the
decrease of H3K9me3 (trimethylation of histone 3 lysin 9) and smaller pool of BCSCs and then irradiating the remaining breast cancer
H3K27me3 (as repressive histone markers) and an increase of cells, put these cells in mice. They found that the ability of these iBCSCs
H3K4me2 and H3K4me3 (as active histone markers) that were asso- to form tumors increased more than 30 folds compared to non-irra-
ciated with an increased acetylation of histone H3 and a significant diated breast cancer cells. Their findings demonstrate that threatening
reduction of global DNA methylation compared to untreated cells. differentiated breast cancer cells by exerting challenges on them (such
Furthermore. HDAC2-silenced osteosarcoma cell lines acquired a stem as radiation), induces the generation of iBCSCs which might, along with
phenotype and promoted in vivo tumorigenesis in mouse xenografts. surviving CSCs, produce more tumors [150].
Overall, they showed that VPA and DAC induce expansion of osteo- Normal stem cells have been shown to contain lower ROS levels
sarcoma CSCs and that HDAC2 is a key mediator in regulating both CSC compared to their more mature progeny, a characteristic which is cri-
phenotype and cancer growth in vivo [137]. In addition, PPP activity is tical for the maintenance of stem cell function. Low ROS levels are
enhanced in response to oxidative stress [138], ionizing radiation [139] required to maintain key properties of stem cells including quiescence
or chemotherapies [140], which induce high levels of ROS and trigger and self-renewal whereas high ROS levels result in stem cell differ-
an adaptive response by augmenting PPP. In multiple cancer cell lines, entiation, senescence and apoptosis. Therefore, ROS production in stem
acquisition of drug resistance has been documented to be associated cells is tightly regulated to ensure the ability to maintain tissue
with elevated oxidative PPP activities. Sustained high levels of GSH and homeostasis highlighting the presence of enhanced ROS scavenging
G6PD are hallmarks of enhanced oxidative PPP activity accounting for systems in such cells. Similar to normal tissue stem cells, CSCs exhibit
drug resistance [141,142]. Therefore, PPP inhibition seems to be a lower intracellular ROS contents and enhanced ROS defense (anti-
promising strategy for cancer targeting. Mele and coleagues in 2018 oxidant capabilities) in comparison with their non-tumorigenic progeny
showed anti-PPP effects of the natural molecule polydatin (3,4′,5-tri- which may contribute to tumor chemo/radio-resistance [151,5]. Nu-
hydroxystilbene-3-β-D-glucoside; transresveratrol 3-β-mono-D-gluco- clear factor erythroid2-related factor-2 (NRF2) is a central and pivotal
side; piceid) through targeting G6PD. In their study, blockade by transcription factor involved in cellular antioxidant defense system
polydatin of G6PD was shown to be associated with ROS accumulation (genes encoding G6PD and 6PGD enzymes of oxidative PPP as the most
and strong increase of endoplasmic reticulum (ER) stress leading to cell important providers of cellular NAPH responsible for antioxidant de-
cycle arrest at S phase, increased apoptosis (50%), and inhibition of in fense, are major targets of this transcription factor) and thus plays a key
vitro invasion (60%). Polydatin inhibited proliferation of head and neck role in cell homeostasis, proliferation and chemo/radioresistance. In
squamous cell carcinoma (HNSCC) cell lines and caused ER-induced cell various cancers, inhibition of NRF2-dependent antioxidant response
death (as a result of ROS accumulation) due to collaborating actions of using siRNA can lead to increased sensitivity to 5-FU, cisplatin, gem-
inositol-requiring enzyme 1 (IRE1; an enzyme activated during ER citabine and doxorubicin. Considered as a potential mediator of cancer
stress with intrinsic kinase and endoribonuclease activities) and protein cell resistance, NRF-2 has been shown to be associated with cancer cell
kinase RNA-like ER kinase (PERK) in inducing the transcription of stemness. For example, its silencing reduces the expression of aldehyde
CHOP (CCAAT-enhancer-binding protein homologous protein) which is dehydrogenase family, member A1 (ALDH1A1) and ALDH3A1 as bio-
involved in apoptosis activation. G6PD overexpression was shown to markers of pancreatic cancer stemness [152]. Taking these in mind and
counteract polydatin antitumor effects. In an experimental orthotopic also regarding the fact that the major part of cellular NADPH (re-
metastatic model of tongue cancer, polydatin reduced tumor size and sponsible for higher antioxidant capabilities and drug resistance) is
inhibited lymph node metastases. Also, polydatin increased che- produced in PPP, high PPP rate is deduced to play a key role in the
motherapeutic effects of cisplatin and afatinib [143]. function as well as survival of CSCs. Indeed, although augmented
Studies have also shown that exposure to stress associates with an aerobic glycolysis (Warburg effect) and function in parallel to enhanced
increase in stem cells populations, for example: Ionizing radiation (IR) PPP (which is necessary for providing nucleotides as building blocks for
can increase ROS production [144]. Radiolysis of water and early ac- proliferation as well as supports high antioxidant capabilities in cancer
tivation of nitric oxide synthases are main sources of ROS and RNS in cells) to support tumor growth [21], the key metabolic balance devia-
irradiated cells under ambient oxygen. Interestingly, the yield of these tion from glycolysis to PPP and more enhanced oxidative PPP rate
species is highly modulated by different radiation types. Increasing LET would occur in the encounter of increased ROS level/exposure and
of the irradiating particles, is associated with an increased yield of seem indispensable for development of CSC.
molecular products (such as H2O2) and a corresponding decreased yield
of radicals (such as %OH). In contrast, O2%− (or HO2%) is the most 7. Dual effects of ROS on HIF-1
abundant radical species generated by radiations with high LET char-
acter [145,146]. Increased levels of ROS in cancer cells can activate glycolysis, either
ROS have been demonstrated to have a significant role in mediating directly or indirectly. As previously mentioned, ROS can directly in-
the biological effects of IR. Although considered as a standard therapy fluence on glucose transporters and indirectly activate glycolysis
for a variety of malignant tumors, IR also paradoxically promotes tumor through affecting the HIF-1 axis.
metastasis and recurrence [144]. The epithelial mesenchymal transition O2−% and H2O2 are two species initiating cascades of redox chem-
(EMT) has been demonstrated to provide cancer cells with migratory istry which are central to redox biology. A network consisting of redox
and invasive characteristics, enabling the metastasis initiation active small molecules and enzymes maintains low steady-state levels of
[147–149]. It is known that IR induces EMT. EMT might be closely O2%− and H2O2. Different redox couples sense changes in O2−% and
linked to CSCs and the metabolic reprogramming of cancer cells. IR has H2O2 flux, each resulting in distinct patterns of gene expression. H2O2,
been demonstrated to induce CSC phenotype in multiple cancers, such as a signaling molecule, is a two-electron oxidant ideal for altering the
as: lung, breast and prostate cancers, as well as melanoma. Genotoxic redox poise of thiols, thereby activating signaling pathways that can
stress owing to chemotherapy or IR promotes a CSC-like phenotype by result in quiescence or differentiation, or at very high flux, apoptosis

8
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

and cell death. In contrast, a subset of iron-containing enzymes is the degradation and reduce HIF-1α levels in ischemic neurons and brains.
primary target of O2−%. Changes in the activity of these enzymes fol- Ethyl 3,4-dihydroxybenzoic acid (DHB) and dimethyloxallyl glycine
lowing reaction with O2−% result in changes in metabolism and gene (DMOG) are prolyl hydroxylase inhibitors and well-known activators of
expression patterns which promote cell division. Examples are activa- HIF-1. Recently, treatment with either DMOG or DHB during nerve
tion of the HIF system via prolyl hydroxylases and the inactivation of growth factor (NGF) deprivation has been observed to suppress the
aconitase. Superoxide dismutase, at the fulcrum of these two signaling accumulation of ROS to levels that are only slightly above the basal
pathways, contributes to the balance between 1-electron and 2-electron level measured in NGF-maintained neurons [164].
signaling. Therefore, the traditional view of antioxidant enzymes needs These results imply that inhibiting ROS might play a role in DHB
to be expanded. The antioxidant network of small molecules, proteins, and DMOG ability to increase HIF-1 expression. In explants of human
and enzymes, that senses changes in ROS/RNS flux, creates the ap- pial arteries, superoxide anion radical produced by xanthine/xanthine
propriate balance between 1-electron and 2-electron signaling. As a oxidase decrease HIF-1α levels [165]. Co-expression of the redox factor-
result, this apparent antioxidant network is best viewed as a set of in- 1 or the antioxidant pyrrolidine dithiocarbamate increase HIF-1α levels
terconnected redox networks establishing the basic biology of cells, in RacG12V-cells [166]. Glucose is an essential substrate required for
tissues, and organisms [153]. production of the main intracellular reductants GSH and NADPH by
Several reports suggest that mitochondria-derived ROS are required PPP and subsequently sustains a reduced cellular milieu in the brain. As
for induction of HIF-1 during hypoxia, the effect that is mediated by a consequence, glucose deprivation would lead to a condition of me-
H2O2. Exogenous H2O2 has also been observed to stabilize HIF-1 (in tabolic oxidative stress. Indeed, glucose deprivation has been reported
Hep3B cells treated with 25 or 40 μM H2O2) and catalase over- to increase ROS production in neurons under hypoxic conditions. A
expression prevents HIF-1 activity [154]. The PI3K and MAPK path- negative relationship has been observed between HIF-1α and H2O2
ways might mediate H2O2 -induced HIF-1 stabilization, as these path- levels. Moreover, N-acetyl cysteine (a ROS scavenger) elevates protein
ways are known to stabilize HIF-1 and H2O2 activates these pathways level of HIF-1α whereas L-buthionine sulfoximine, which induces oxi-
[155–157]. dative stress, inhibits HIF-1α expression in hypoxic neurons [161]. This
Since H2O2 and HIF-1 levels are thought to be increased in cancer finding suggests that ROS does not promote HIF-1α protein stabiliza-
cells and H2O2 is known to stabilize HIF-1, it seems conceivable that tion, but rather induces its degradation in hypoxic neurons.
increased HIF-1 levels in cancer cells might be due to elevated gen- Though, it is not obvious how ROS induces HIF-1α degradation in
eration of H2O2 by these cells [154]. ischemic brain, ROS might promote HIF-1α degradation through two
Activation of HIF not only elevates anaerobic glycolysis, but also mechanisms. First: activation of proteasomal degradation pathways.
attenuates mitochondrial respiration. The former occurs by upregula- The proteasomal proteolytic pathways are considered as the main me-
tion of the genes that encode GLUT, glucokinases, aldolase A and LDH- chanisms through which many intracellular proteins are degraded
M, the enzymes involved in the conversion of pyruvate to lactate. The [167]. These pathways play pivotal roles in the regulation of key cel-
latter occurs through the induction of PDK1, which inhibits PDH in- lular processes including apoptosis, differentiation, cytokine-induced
volved in the conversion of pyruvate to acetyl-CoA in the mitochondria. gene expression, and the stress response [168–170]. There are two
These two phenomena are known to hinder the entry of pyruvate to the forms of proteasome complexes: 26S and 20S. Protein substrate speci-
Krebs cycle and divert pyruvate toward lactate formation through ficities are usually different between the two complexes. The 26S pro-
glycolysis [158]. teasomal pathway is ubiquitin-dependent and is a key pathway for the
Studies show that MnSOD regulates HIF-1α expression through degradation of proteins in cells while the 20S proteasomal pathway is
modulating the steady-state level of O2−%. In MCF-7 cells an siRNA ubiquitin-independent and is primarily used for the degradation of
meadiated decrease of MnSOD was shown to be associated with an cellular oxidized proteins in oxidative stress conditions [167,171–174].
increase in the hypoxic accumulation of HIF-1α and removal of O2−% Under hypoxic conditions, ROS may activate the 26S pathway through
using O2−% scavenger Tempol or an SOD mimic (AEOL10113) or spin increasing prolyl hydroxylase activity [175–178]. ROS may also in-
traps (α-4-pyridyl-1-oxide-N-tert-butylnitrone and 5,5-dimethyl-1-pyr- crease 20S proteasomal activity [179]. HIF-1α, with its cysteine and
roline N-oxide) led to a decrease in HIF-1α protein, consistent with the methionine residues being vulnerable to oxidative damages, is sensitive
hypothesis that O2−% acts as an important molecular effecter in hypoxic to oxidative attacks [180,181]. In fact, according to a recent publica-
stabilization of HIF-1α. The evidence demonstrated that O2−% can tion, the 20S degradation pathway appears to be implicated in the
contribute to the stabilization of HIF-1α [159]. turnover of HIF-1α under hypoxic conditions [182]. Second: P53 has
Paradoxically, accumulated evidence demonstrates that cellular been reported to mediate proteasomal degradation of HIF-1α [183].
H2O2 are able to downregulate the stabilization of HIF-1α in cancer P53 is a redox sensitive protein that can be activated by oxidants
cells. For instance, exposure of Hela cells to H2O2 (1 mM) selectively [184,185]. ROS may induce the expression of P53 which in turn pro-
inhibits the accumulation of HIF-1α protein in hypoxia [87]. In HepG2 motes the degradation of HIF-1α. Furthermore, HIF-1α, under hypoxic
cells, inhibition of the Fenton reaction increases the levels of HIF-1α, conditions, may transcriptionally activate its own degradation, likely
HIF-1 transactivation, and also activates expression of the HIF-1 target mediated through acetylation [185].
genes [160]. According to other investigations, increased ROS levels in As shown by a study, low dose exogenous H2O2 or brief oxygen-
different cancer cells, also impairs the hypoxic signaling mechanism glucose deprivation (OGD) provides neuroprotection against severe
and HIF-1α expression [161]. In addition, in human breast carcinoma OGD 24 h later. H2O2 induces the expression of HIF-1 α in a dose-de-
MCF-7 cells, over-expression of thioredoxin, a thiol reducing protein, pendent manner with a maximal induction at 15 μM. On the other hand,
potentiates hypoxia-induced gene activation [162]. In normoxic cells, H2O2 at the dose of 15 μM, causes a time-dependent increase in the
the dramatic effect observed with the reducing thiol agent N-mercap- expression of HIF-1α with maximal levels at 32 h after preconditioning
topropionylglycine on erythropoietin (EPO) gene activation and HIF-1α [186].
formation provides a strong indication regarding the ability of ROS to Emerging evidence demonstrate that in addition to hypoxia (by
downregulate HIF-1α in Hep 3B and B-1 cells [163]. The mechanisms inhibiting prolyl hydroxylase-dependent degradation of HIF-1α), H2O2
through which ROS destabilizes HIF-1α may involve: 1- increasing HIF- is also a key factor in the activation of HIF-1 and regulation of its ac-
prolyl hydroxylase activity; 2- activating prolyl hydroxylase (via acting tivity. This regulation likely depends on the dose of H2O2 and also type
like oxygen); and 3- increasing oxidation of HIF-1α protein, as mildly of the cell, as pretreatment with high dose of H2O2 (1 mM) has been
oxidized proteins more likely can be degraded by ubiquitin-proteasome reported to inhibit hypoxia-induced HIF-1α protein accumulation in
systems [161]. human Hep3B cells and Hela cells. It has been shown that non-toxic
On the other hand, it has been shown that ROS may promote HIF-1α levels of H2O2 produced by the enzyme glucose oxidase have no effect

9
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

on HIF-1 α stabilization in HT1080 human fibrosarcoma cells metabolic switch towards PPP.
[88,181,187–193].
8.3. The role of TIGAR in glycolysis deviation to PPP
8. Possible mechanisms of glycolysis deviation to PPP
Another important molecule involved in the glycolysis deviation to
As previously suggested in this review, cancer cells, when en- PPP is TIGAR (TP53-induced glycolysis and apoptosis regulator). ROS
countering ROS, first increase the glycolysis rate and then following the and UV stress can trigger TIGAR activation that inhibits PFK1. PFK1
continuation of oxidative stress their metabolic balance is skewed from inhibition helps divert glycolytic flux into PPP [125].
glycolysis pathway to PPP. The possible molecular mechanisms are as In a study, TIGAR was shown to be activated with 720 μmol/L H2O2
follows: for 3 h in A549 human lung cancer cells [197] while as stated before, in
another study HIF-1 was shown to be activated with 15 μM H2O2 for
8.1. The role of PKM2 in glycolysis deviation to PPP 10 min [186]. Since activation of HIF-1 requires less ROS dose/ex-
posure time than that of TIGAR, HIF-1 is first activated. Therefore, first
PKM2 is an important molecule involved in the glycolysis deviation glycolysis is upregulated and then following the increase in ROS dose/
to PPP. Oxidation at Cys358 induced by acute increase in intracellular exposure time, HIF-1 is inactivated and TIGAR is activated. TIGAR
ROS, decreases PKM2 activity allowing accumulation of G6P which activation allows glycolytic flux to be diverted into PPP.
subsequently diverts glucose flux to PPP thereby generating sufficient
reducing potential (NADPH) for ROS detoxification [73,193]. There- 8.4. The role of P53 in glycolysis deviation to PPP
fore, PKM2 oxidation at Cys358 confers an advantage to cancer cells
(A549 human lung cancer cells), as it allows cells to tolerate ROS [73] P53 is also an important molecule involved in the glycolysis de-
and promotes cancer cell survival during conditions of oxidative stress viation to PPP. In response to ROS and UV stress, P53 activates TIGAR.
[108]. TIGAR degrades F2,6-BP thereby inhibiting PFK1. This allows glyco-
PKM2 is allosterically regulated allowing switching between high lytic flux to be diverted into PPP [125].
and low activity states. PKM2 exists in two tetrameric and dimeric In a study, it was shown that P53 is activated with 0.6 or 0.8 mM
states that are catalytically distinct. Tetrameric PKM2 possesses high H2O2 in 293 T cell line [198] Another study showed a slight increase in
catalytic activity associated with ATP synthesis and catabolic metabo- the expression of P53 in HeLa cells treated with 125 μM H2O2 in a time
lism in cells. Dimeric PKM2, which is the less active state of PKM2, has dependent manner (1 h–3 h), but this increase was not statistically
low catalytic activity and facilitates the production of glycolytic inter- significant [199]. In comparison with P53, HIF-1 activation occurs at
mediates to flux into glycolysisbranch pathways including PPP (in- lower H2O2 dose/exposure time [186]. As a consequence, in the face of
volved in NADPH and nucleotide synthesis) and glycerol synthesis H2O2, first glycolysis is upregulated.
pathway [108].
Multiple signals including oncogenes, growth factors and ROS have 9. In response to increased ROS, which one do the cancer cells
been shown to destabilize the tetrameric form via several mechanisms increase? PPP or other pathways of NADPH production
such as tyrosine phosphorylation, association with tyrosine phospho-
peptides or oxidation [194]. ROS can induce HIF-1 activation and Oxidative PPP which is a cytoplasmic branch of glycolysis providing
PKM2 inactivation which subsequently result in increased glycolysis the cell with both NADPH and ribose for nucleotide synthesis, is
and PPP rates, respectively. thought to be the main source of NADPH. However, there exist other
Studies show that ROS dose and exposure time are two important potential sources of NADPH such as the reactions involving different
factors mediating HIF-1 and PKM2 activation/inactivation. isoforms of isocitrate dehydrogenase (IDH) and malic enzyme (ME).
According to a study, in A549 human lung cancer cells PKM2 is Though less appreciated (until now), serine and glycine metabolism
inactivated in the presence of 250 μM diamide or 1 mM H2O2 for 15 min exploiting the tetrahydrofolate (THF) cycle, can also be regarded as
[73], while another study showed that HIF-1 is activated in 6–8 days in notable sources of NADPH [200].
vitro (DIV) primary cortical neurons from embryonic day 16 (E16) CD1 The cytosol possesses several sources of NADPH generation in-
mouse brains with 15 μM H2O2 for 10 min [186]. Since activation of cluding the oxidative PPP, IDH1, malic enzyme 1 (ME1), and one-
HIF-1 requires less H2O2 dose/exposure time than inactivation of carbon metabolism. In the mitochondria, NADPH generation is, in part,
PKM2, in the face of H2O2, first HIF-1 is activated. Therefore, in re- controlled by one-carbon metabolism and IDH2. Cytosolic sources of
sponse to ROS, it is glycolysis which is upregulated first. NADPH comprise the oxidative PPP, IDH1 and enzymes from one-
With the increase in ROS dose/exposure time, HIF-1 and PKM2 are carbon metabolism such as methylene tetrahydrofolate dehydrogenase
inactivated. Oxidation of PKM2 decreases enzyme activity and pro- 1 (MTHFD1). Mitochondrial sources of NADPH include IDH2, MTHFD2,
motes oxidative PPP flux. and MTHF2L [201].
There are two major NADPH-generating pathways in the cell. One is
8.2. The role of GAPDH in glycolysis deviation to PPP PPP which generates NADPH in the cytoplasm. The first step of PPP,
mediated by G6PD, involves the conversion of the glycolytic inter-
Another important molecule involved in the glycolysis deviation to mediate G6P and NADP + to 6-phosphogluconolactone and NADPH.
PPP is GAPDH. ROS inactivates GAPDH by directly targeting cysteine The other major pathway for NADPH generation is one-carbon meta-
residues [195]. GAPDH inhibition helps divert glycolytic flux into the bolism (1CM) which is also known as the folate cycle. Serine is the main
oxidative phase of PPP by allowing metabolites to accumulate upstream donor of one-carbon units to the folate cycle. The serine synthesis
of the inhibition point. HIF-1 activation by ROS increases glycolysis pathway (SSP) provides a mechanism for the carbons derived from
while ROS-mediated inactivation of GAPDH increases PPP. Studies have glucose to be deviated from glycolysis for de novo synthesis of serine
shown that a large amount of ROS is needed to inactivate GAPDH. For [202].
example, in a study, GAPDH was shown to be inactivated with 500 μM Serine is required for multiple biosynthetic and signaling pathways
H2O2 for 60 min or 2 mM H2O2 for 30 min in S. cerevisiae exponential- including synthesis of amino acids like cysteine and glycine both of
phase cells [196]. As stated before, HIF-1 is activated with 15 μM H2O2 which are required for glutathione synthesis. In the SSP, the glycolytic
for 10 min [186]. Therefore, in the face of H2O2, first glycolysis is ac- intermediate 3-phosphoglycerate (3-PG) is converted to 3-phospho-
tivated and then following the increase in H2O2 dose/exposure time, pyruvate (3-PHP) by the action of phosphoglycerate dehydrogenase
HIF-1 and GAPDH are inactivated. GAPDH inactivation leads to (PHGDH). In the next step, phosphoserine aminotransferase (PSAT1),

10
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

converts 3-PHP to phosphoserine (P-Ser) and in the final step, phos- consistent with the features like increased PPP activity. For example,
phoserine phosphatase (PSPH) converts P-Ser into serine. Then, serine increased PPP in cancer cells has been shown to associate with in-
and NADP + are utilized in 1CM pathway, either in the cytosol or mi- creased drug resistance. Based on the previous findings cancer cells are
tochondria, which produces glycine and NADPH. Mitochondrial 1CM is inclined to increase PPP rate following exposure to excess amounts of
catalyzed by the enzymes: serine hydroxymethyltransferase 2 (SHMT2), ROS. Since oxidative stress is a strong inducer of CSCs, the number of
MTHFD2, or MTHFD2-like (MTHFD2L), and MTHFD1L. In cytosolic these cells will increase following ROS exposure. Therefore, an increase
1CM, the same reactions are mediated by SHMT1 and MTHFD1 in ROS might indicate an increase in the population of CSCs with the
(MTHFD1 catalyzes the reactions performed by both MTHFD2/ feature of increased PPP. On such a basis, it may seem reasonable to
MTHFD2L and MTHFD1L) [202]. consider the increase in PPP, its metabolites and related proteins as
In hypoxia, HIFs activate the transcription of PHGDH, PSAT1, and markers for the detection of CSC overpopulation. Therefore, con-
PSPH to enhance conversion of glucose to serine; and SHMT2, sidering this issue could be of significant prognostic value and might
MTHFD2, and MTHFD1L to enhance generation of mitochondrial provide insight on more effective treatment of cancer.
NADPH (by mitochondrial 1CM), which is required to convert GSSG to ROS, have a double deck sword behaviour in cancer. On the one
GSH to protect against increased ROS generated by the ETC. Under hand, ROS can promote protumorigenic signaling which facilitates
hypoxic conditions, HIFs also activate the transcription of solute carrier proliferation, survival and metastasis of cancer cell. On the other hand,
family 7 member 11 (SLC7A11) and glutamate-cysteine ligase modifier ROS can promote antitumorigenic signaling and trigger oxidative
subunit (GCLM) to increase glutathione production [203]. stress–induced death in cancer cells. Production of ROS is considered as
In a study, IDH was treated with 1, 2, and 10 mM H2O2 for 12 min. a mechanism shared by non-surgical therapeutic approaches for can-
The results showed that incubation with 1, 2, and 10 mM H2O2 for cers, such as radiotherapy and chemotherapy. Therefore, due to the
12 min dose-dependently decreased IDH activity [204]. implication of such therapeutic approaches in triggering cell death, ROS
HIF-1 activates transcription of PHGDH, PSAT1, and PSPH to in- are also used to kill cancer cells. ROS are increasingly being appreciated
crease serine synthesis pathway; and SHMT2, MTHFD2, and MTHFD1L as potent signaling molecules implicated in the regulation of a number
to increase generation of mitochondrial NADPH. On one hand, HIF-1 is of different biological processes. Cancer cells exhibit higher ROS levels
inactivated at very high H2O2 levels, the conditions that cause the compared to their normal counterparts partly due to an enhanced
metabolic balance to be deviated from glycolysis to PPP. On the other metabolic rate as well as mitochondrial dysfunction. This increased
hand, under such high H2O2 concentrations IDH1 activity is decreased ROS in cancer cells contributes to molecular and biochemical altera-
and also due to HIF-1 inactivation transcription of PHGDH, PSAT1, tions necessary for tumor initiation, promotion, progression as well as
PSPH, SHMT2, MTHFD2, and MTHFD1L is hindered. Therefore it could chemoresistance (protumorigenic effects). However, in cancer cells
be concluded that under very high ROS levels that cause glycolysis further increase in ROS levels to a toxic level, through augmenting
deviation to PPP, other pathways of NADPH synthesis are down- chemosensitization and induction of various cell-death pathways, pro-
regulated. vides an avenue for cancer treatment (antitumorigenic effects).
Therefore, biological effects of ROS in cancer are complex, paradoxical
10. Conclusion and context-dependent. The dual aspects of ROS in cancer biology
provides us with a rational to develop two different contradicting ap-
Cancer cells, as compared to normal cells of the body, have higher proaches for cancer treatment. The first strategy involves lowering in-
levels of ROS. Alteration in glucose metabolism is one of the ways by tracellular ROS levels in cancer cells either by inhibiting ROS genera-
which cancer cells decrease the amount of ROS. Cancer cells can de- tion or by augmenting ROS scavenging activities using different
crease the amount of ROS by increasing glycolysis and PPP. In this antioxidants. For instance, metformin, as an anti-diabetic agent, has
review, by summarizing the current literature, we showed that cancer been shown to inhibit mitochondrial respiratory chain complex I within
cells, when encountering ROS, first activate the glycolysis pathway and cancer cells thereby reducing tumorigenesis. Also included in this
then following the increase in ROS levels/exposure time, the metabolic strategy, we have the use of antioxidants, either alone or in combina-
balance is skewed from glycolysis to PPP. What determines this meta- tion with other antitumor agents, as a promosing alternative to decrease
bolic deviation is probably differential responses of certain proteins to intracellular ROS-mediated tumorigenesis [39]. Based on the pivotal
different ROS exposure conditions (dose and time). For example, HIF-1 role ROS play as a key constituent in tumor biology [38] as well as
activation which upregulates glycolysis, occurs at low doses of ROS considering our hypothesis on the contribution of higher ROS levels to
(H2O2) while higher doses of ROS cause HIF-1 inactivation thereby CSC development, lowering ROS using antioxidants seems to be a
decreasing glycolysis. Inactivation of PKM2, which causes the deviation promising approach for cancer targeting. However, it is a challenging
towards PPP, requires higher doses of ROS and longer exposure times approach to adopt. Even though antioxidant therapy combats cancer
compared to what is required for HIF-1 activation. This confirms that through impeding ROS-induced carcinogenesis, it can also have pro-
HIF-1 is active in lower doses and promotes glycolysis, while following tumor effects by inhibiting cellular senescence or apoptosis induced by
the increase in ROS levels/exposure time HIF-1 and PKM2 proteins are higher ROS levels as a consequence of high genomic instability of
inactivated consequently decreasing glycolysis and promoting PPP. cancer cells [205].
Also, increased ROS causes GAPDH inactivation which promotes PPP. The second ROS-mediated strategy for cancer treatment is to in-
Inactivation of this protein occurs at higher ROS levels and longer ex- crease ROS levels in the cancer cell to a toxic level able to activate ROS-
posure times compared to HIF-1 activation. Activation of TIGAR and induced cell death pathways. This strategy can be achieved through the
P53, which cause the same metabolic deviation, also occur at higher use of agents that either augment ROS generation or inhibit antioxidant
ROS levels and longer exposure times compared to HIF-1 activation. systems, or a combination of both. Regarding the fact that in cancer cell
Therefore, exposure to higher doses of ROS (with longer exposure there is an already increased level of ROS compared to its normal
times) downregulates glycolysis and upregulates PPP. counterpart, therapeutic interventions further increasing the generation
In addition, in this review, we hypothesize that the metabolic de- of ROS can specifically induce cell death in cancer cell. For example, it
viation from glycolysis to PPP might represent alterations in cancer has been shown that beta-phenylethyl isothiocyanate and 2-methox-
cells towards CSCs. yoestradiol, by causing further increase in ROS, selectively induce
CSCs, unlike the differentiated cancer cells that exhibit increased death in human leukemia cells over normal lymphocytes. However, in
glycolysis, possess different metabolic characteristics including in- this respect eradicating cancer cells does not seem that easy. Cancer
creased PPP activity. CSCs are responsible for drug resistance, tumor cells also possess higher antioxidant capabilities to prevent excessive
initiation, metastasis and recurrence, the characteristics that are accumulation of ROS [39].

11
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

Based on the current review, we propose that treatments based on [14] Z.T. Schafer, A.R. Grassian, L. Song, Z. Jiang, Z. Gerhart-Hines, H.Y. Irie, S. Gao,
oxidative stress can not be considered as appropriate cancer treatment P. Puigserver, J.S. Brugge, Antioxidant and oncogene rescue of metabolic defects
caused by loss of matrix attachment, Nature 461 (2009) 109.
strategies, as these treatments could induce adaptive mechanisms in [15] K. Ishikawa, K. Takenaga, M. Akimoto, N. Koshikawa, A. Yamaguchi, H. Imanishi,
cancer cells, mechanisms that cause the deviation from glycolysis to K. Nakada, Y. Honma, J. Hayashi, ROS-generating mitochondrial DNA mutations
PPP which could be translated as a step from cancer cell to CSC. This can regulate tumor cell metastasis, Science 320 (2008) 661–664.
[16] F. Weinberg, R. Hamanaka, W.W. Wheaton, S. Weinberg, J. Joseph, M. Lopez,
point of view further complicates ROS-mediated therapeutic ap- B. Kalyanaraman, G.M. Mutlu, G.R. Budinger, N.S. Chandel, Mitochondrial me-
proaches in tumor context included in the second strategy and adds to tabolism and ROS generation are essential for Kras-mediated tumorigenicity, Proc.
the existing complexity and controversial aspects of ROS exploitation Natl. Acad. Sci. U. S. A. 107 (2010) 8788–8793.
[17] D. Trachootham, J. Alexandre, P. Huang, Targeting cancer cells by ROS-mediated
for cancer treatment. Considering this hypothesis, if we want to apply mechanisms: a radical therapeutic approach? Nat. Rev. Drug Discov. 8 (2009) 579.
the second strategy (activation of ROS-induced cell death pathways) to [18] A. Glasauer, N.S. Chandel, Targeting antioxidants for cancer therapy, Biochem.
target cancer cells, ROS-inducing agents are suggested to be used in Pharmacol. 92 (2014) 90–101.
[19] G.M. DeNicola, F.A. Karreth, T.J. Humpton, A. Gopinathan, C. Wei, K. Frese,
combination with antioxidant inhibitors in order to cope with CSCs
D. Mangal, K.H. Yu, C.J. Yeo, E.S. Calhoun, F. Scrimieri, J.M. Winter, R.H. Hruban,
(with high antioxidant capabilities) that are likely to develop following C. Iacobuzio-Donahue, S.E. Kern, I.A. Blair, D.A. Tuveson, Oncogene-induced Nrf2
our therapeutic intervention. Also combined with ROS-promoting transcription promotes ROS detoxification and tumorigenesis, Nature 475 (2011)
agents, we interestingly suggest the use of oxidative PPP inhibitors; for 106.
[20] S. Kamarajugadda, L. Stemboroski, Q. Cai, N.E. Simpson, S. Nayak, M. Tan, J. Lu,
example molecules that inhibit G6PD or 6PGD. It is also suggested to Glucose oxidation modulates anoikis and tumor metastasis, Mol. Cell. Biol. 32
highly focus on therapies specifically targeting important molecules (2012) 1893–1907.
mediating glycolysis deviation to PPP in combination with ROS-pro- [21] P. Jiang, W. Du, M. Wu, Regulation of the pentose phosphate pathway in cancer,
Protein Cell 5 (2014) 592–602.
moting agents. However, achieving a well-tailored ROS-mediated [22] J. Xie, H. Wu, C. Dai, Q. Pan, Z. Ding, D. Hu, B. Ji, Y. Luo, X. Hu, Beyond Warburg
therapeutic approach for cancer treatment involves a tremendous and effect–dual metabolic nature of cancer cells, Sci. Rep. 4 (2014) 4927.
detailed knowledge of dual aspects of ROS in tumor biology and calls [23] C.E. Meacham, S.J. Morrison, Tumour heterogeneity and cancer cell plasticity,
Nature 501 (2013) 328.
for further investigations in this amazing field of cancer research. [24] Q. Pan, Q. Li, S. Liu, N. Ning, X. Zhang, Y. Xu, A.E. Chang, M.S. Wicha, Concise
review: targeting cancer stem cells using immunologic approaches, Stem Cells 33
Conflict of interest (2015) 2085–2092.
[25] B.G. Debeb, L. Lacerda, R. Larson, A.R. Wolfe, S. Krishnamurthy, J.M. Reuben,
N.T. Ueno, M. Gilcrease, W.A. Woodward, Histone deacetylase inhibitor-induced
The authors declare that they have no conflict of interest. cancer stem cells exhibit high pentose phosphate pathway metabolism, Oncotarget
7 (2016) 28329.
[26] K.C. Patra, N. Hay, The phosphate pathway and cancer, Trends Biochem. Sci. 39
Acknowledgments
(2014) 347–354.
[27] T. Reya, S.J. Morrison, M.F. Clarke, I.L. Weissman, Stem cells, cancer, and cancer
This research was supported by Medical Biology Research Center, stem cells, Nature (2001) 105.
Kermanshah University of Medical Sciences, Kermanshah, Iran (96599). [28] Y.J. Kim, E.L. Siegler, N. Siriwon, P. Wang, Therapeutic strategies for targeting
cancer stem cells, J. Cancer Metastasis Treat. 2 (2016) 234.
We would like to specially thank to Prof. Reza Khodarahmi, Dr Bizhan [29] D.L. Dragu, L.G. Necula, C. Bleotu, C.C. Diaconu, M. Chivu-Economescu, Therapies
Solimani, Dr Hamid Rza Mohammadimotlagh and Mozhgan Jahani for targeting cancer stem cells: current trends and future challenges, World J. Stem
their valuable discussions and help with manuscript preparation. Cells 7 (2015) 1185.
[30] H.J. Forman, O. Augusto, R. Brigelius-Flohe, P.A. Dennery, B. Kalyanaraman,
H. Ischiropoulos, G.E. Mann, R. Radi, L.J. Roberts, J. Vina, K.J. Davies, Even free
References radicals should follow some rules: a guide to free radical research terminology and
methodology, Free Radic. Biol. Med. 78 (2015) 233–235.
[31] H.J. Forman, M. Maiorino, F. Ursini, Signaling functions of reactive oxygen spe-
[1] E. Panieri, M. Santoro, ROS homeostasis and metabolism: a dangerous liason in
cies, Biochemistry 49 (2010) 835–842.
cancer cells, Cell Death Dis. 7 (2016) e2253.
[32] C.C. Winterbourn, M.B. Hampton, Thiol chemistry and specificity in redox sig-
[2] G.Y. Liou, P. Storz, Reactive oxygen species in cancer, Free Radic. Res. 44 (2010)
naling, Free Radic. Biol. Med. 45 (2008) 549–561.
479–496.
[33] H. Sies, Hydrogen peroxide as a central redox signaling molecule in physiological
[3] I.M. Ahmad, N. Aykin-Burns, J.E. Sim, S.A. Walsh, R. Higashikubo, G.R. Buettner,
oxidative stress: oxidative eustress, Redox Biol. 11 (2017) 613–619.
S. Venkataraman, M.A. Mackey, S.W. Flanagan, L.W. Oberley, D.R. Spitz,
[34] E.I. Azzam, J.P. Jay-Gerin, D. Pain, Ionizing radiation-induced metabolic oxidative
Mitochondrial and H2O2 mediate glucose deprivation-induced stress in human
stress and prolonged cell injury, Cancer Lett. 327 (2012) 48–60.
cancer cells, J. Biol. Chem. 280 (2005) 4254–4263.
[35] W. Droge, Free radicals in the physiological control of cell function, Physiol. Rev.
[4] N. Aykin-Burns, I.M. Ahmad, Y. Zhu, L.W. Oberley, D.R. Spitz, Increased levels of
82 (2002) 47–95.
superoxide and H2O2 mediate the differential susceptibility of cancer cells versus
[36] W. Pendergrass, G. Zitnik, R. Tsai, N. Wolf, X-ray induced cataract is preceded by
normal cells to glucose deprivation, Biochem. J. 418 (2009) 29–37.
LEC loss, and coincident with accumulation of cortical DNA, and ROS; similarities
[5] D. Zhou, L. Shao, D.R. Spitz, Reactive oxygen species in normal and tumor stem
with age-related cataracts, Mol. Vis. 16 (2010) 1496.
cells, Adv. Cancer Res. 122 (2014) 1–67.
[37] P. Riley, Free radicals in biology: oxidative stress and the effects of ionizing ra-
[6] D.R. Spitz, J.E. Sim, L.A. Ridnour, S.S. Galoforo, Y.J. Lee, Glucose deprivation-
diation, Int. J. Radiat. Biol. 65 (1994) 27–33.
induced oxidative stress in human tumor cells: a fundamental defect in metabo-
[38] S. Kumari, A.K. Badana, M.M. G, S. G, R. Malla, Reactive Oxygen Species: A Key
lism? Ann. N. Y. Acad. Sci. 899 (2000) 349–362.
Constituent in Cancer Survival, Biomark. Insights 13 (2018) 1177271918755391..
[7] O. Vafa, M. Wade, S. Kern, M. Beeche, T.K. Pandita, G.M. Hampton, G.M. Wahl, c-
[39] S. Galadari, A. Rahman, S. Pallichankandy, F. Thayyullathil, Reactive oxygen
Myc can induce DNA damage, increase reactive oxygen species, and mitigate p53
species and cancer paradox: to promote or to suppress? Free Radic. Biol. Med. 104
function: a mechanism for oncogene-induced genetic instability, Mol. Cell. 9
(2017) 144–164.
(2002) 1031–1044.
[40] I. Ganan-Gomez, Y. Wei, H. Yang, M.C. Boyano-Adanez, G. Garcia-Manero,
[8] V. Nogueira, Y. Park, C.C. Chen, P.Z. Xu, M.-L. Chen, I. Tonic, T. Unterman,
Oncogenic functions of the transcription factor Nrf2, Free Radic. Biol. Med. 65
N. Hay, Akt determines replicative senescence and oxidative or oncogenic pre-
(2013) 750–764.
mature senescence and sensitizes cells to oxidative apoptosis, Cancer Cell 14
[41] T.P. Szatrowski, C.F. Nathan, Production of large amounts of hydrogen peroxide
(2008) 458–470.
by human tumor cells, Cancer Res. 51 (1991) 794–798.
[9] A.A. Sablina, A.V. Budanov, G.V. Ilyinskaya, L.S. Agapova, J.E. Kravchenko,
[42] A. Costa, A. Scholer-Dahirel, F. Mechta-Grigoriou, The role of reactive oxygen
P.M. Chumakov, The antioxidant function of the p53 tumor suppressor, Nat. Med.
species and metabolism on cancer cells and their microenvironment, Semin.
11 (2005) 1306.
Cancer Biol. 25 (2014) 23–32.
[10] K. Bensaad, E.C. Cheung, K.H. Vousden, Modulation of intracellular ROS levels by
[43] W. Li, L. Cao, L. Han, Q. Xu, Q. Ma, Superoxide dismutase promotes the epithelial-
TIGAR controls autophagy, EMBO J. 28 (2009) 3015–3026.
mesenchymal transition of pancreatic cancer cells via activation of the H2O2/
[11] W. Hu, C. Zhang, R. Wu, Y. Sun, A. Levine, Z. Feng, Glutaminase 2, a novel p53
ERK/NF-κB axis, Int. J. Oncol. 46 (2015) 2613–2620.
target gene regulating energy metabolism and antioxidant function, Proc. Natl.
[44] M.C. Kim, F.J. Cui, Y. Kim, Hydrogen peroxide promotes epithelial to mesench-
Acad. Sci. 107 (2010) 7455–7460.
ymal transition and stemness in human malignant mesothelioma cells, Asian Pac.
[12] S. Reuter, S.C. Gupta, M.M. Chaturvedi, B.B. Aggarwal, Oxidative stress, in-
J. Cancer Prev. 14 (2013) 3625–3630.
flammation, and cancer: how are they linked? Free Radic. Biol. Med. 49 (2010)
[45] V. Ribas, C. Garcia-Ruiz, J.C. Fernandez-Checa, Mitochondria, cholesterol and
1603–1616.
cancer cell metabolism, Clin. Transl. Med. 5 (2016) 22.
[13] B. Halliwell, Oxidative stress and cancer: have we moved forward? Biochem. J.
[46] A. Federico, F. Morgillo, C. Tuccillo, F. Ciardiello, C. Loguercio, Chronic in-
401 (2007) 1–11.
flammation and oxidative stress in human carcinogenesis, Int. J. Cancer 121

12
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

(2007) 2381–23896. Microbiol. Mol. Biol. Rev. 78 (2014) 89–175.


[47] M. Fransen, M. Nordgren, B. Wang, O. Apanasets, Role of peroxisomes in ROS/ [79] A. Stincone, A. Prigione, T. Cramer, M. Wamelink, K. Campbell, E. Cheung,
RNS-metabolism: implications for human disease, Biochim. Biophys. Acta. Mol. V. Olin-Sandoval, N.M. Gruning, A. Kruger, M. Tauqeer Alam, M.A. Keller,
Basis Dis. 1822 (2012) 1363–1373. M. Breitenbach, K.M. Brindle, J.D. Rabinowitz, M. Ralser, The return of metabo-
[48] V.D. Antonenkov, S. Grunau, S. Ohlmeier, J.K. Hiltunen, Peroxisomes are oxida- lism: biochemistry and physiology of the pentose phosphate pathway, Biol Rev. 90
tive organelles, Antioxid. Redox Signal. 13 (2010) 525–537. (2015) 927–963.
[49] M. Valko, C. Rhodes, J. Moncol, M. Izakovic, M. Mazur, Free radicals, metals and [80] P.J. Fernandez-Marcos, S. Nobrega-Pereira, NADPH: new oxygen for the ROS
antioxidants in oxidative stress-induced cancer, Chem.-Biol. Interact. 160 (2006) theory of aging, Oncotarget 7 (2016) 50814.
1–40. [81] R.B. Hamanaka, N.S. Chandel, Targeting glucose metabolism for cancer therapy, J.
[50] M. Inoue, E.F. Sato, M. Nishikawa, A.M. Park, Y. Kira, I. Imada, Utsumi K, Exp. Med. 209 (2012) 211–215.
Mitochondrial generation of reactive oxygen species and its role in aerobic life, [82] C.M. Labak, P.Y. Wang, R. Arora, M.R. Guda, S. Asuthkar, A.J. Tsung, K.V. Kiran,
Curr. Med. Chem. 10 (2003) 2495–2505. Glucose transport: meeting the metabolic demands of cancer, and applications in
[51] D.E. Handy, J. Loscalzo, Redox regulation of mitochondrial function, Antioxid. glioblastoma treatment, Am. J. Cancer Res. 6 (2016) 1599.
Redox Signal. 16 (2012) 1323–1367. [83] T.B. Zhang, Y. Zhao, Z.X. Tong, Y.F. Guan, Inhibition of glucose-transporter 1
[52] M.M. Briehl, Oxygen in human health from life to death–an approach to teaching (GLUT-1) expression reversed Warburg effect in gastric cancer cell MKN45, Int. J.
redox biology and signaling to graduate and medical students, Redox Boil 5 (2015) Clin. Exp. Med. 8 (2015) 2423.
124–139. [84] S. Andrisse, G.D. Patel, J.E. Chen, A.M. Webber, L.D. Spears, R.M. Koehler,
[53] M.P. Murphy, How mitochondria produce reactive oxygen species, Biochem. J. Ching Robinson-Hill, R.M, J.K, I. Jeong, J.S, ATM and GLUT1-S490 phosphor-
417 (2009) 1–13. ylation regulate GLUT1 mediated transport in skeletal muscle, PLoS One 8 (2013)
[54] D.C. Liemburg-Apers, P.H. Willems, W.J. Koopman, S. Grefte, Interactions be- e66027.
tween mitochondrial reactive oxygen species and cellular glucose metabolism, [85] A. Alexander, C.L. Walker, Differential localization of ATM is correlated with ac-
Arch. Toxicol. 89 (2015) 1209–1226. tivation of distinct downstream signaling pathways, Cell Cycle 9 (2010)
[55] F. Cosentino, P. Francia, G.G. Camici, P.G. Pelicci, M. Volpe, T.F. Luscher, Final 3709–3710.
common molecular pathways of aging and cardiovascular disease: role of the [86] G.L. Semenza, Hypoxia-inducible factor 1: master regulator of O 2 homeostasis,
p66Shc protein, Arterioscler. Thromb. Vasc. Biol. 28 (2008) 622–628. Curr. Opin. Genet. Dev. 8 (1998) 588–594.
[56] K.M. Holmstrom, Finkel T, Cellular mechanisms and physiological consequences [87] G.L. Wang, G.L. Semenza, Purification and characterization of hypoxia-inducible
of redox-dependent signalling, Nat. Rev. Mol. Cell Biol. 15 (2014) 411. factor 1, J. Biol. Chem. 270 (1995) 1230–1237.
[57] N.Y. Spencer, J.F. Engelhardt, The basic biology of redoxosomes in cytokine- [88] L.E. Huang, Z. Arany, D.M. Livingston, H.F. Bunn, Activation of hypoxia-inducible
mediated signal transduction and implications for disease-specific therapies, transcription factor depends primarily upon redox-sensitive stabilization of its α
Biochemistry 53 (2014) 1551–1564. subunit, J. Biol. Chem. 271 (1996) 32253–32259.
[58] E.E. Kelley, N.K. Khoo, N.J. Hundley, U.Z. Malik, B.A. Freeman, M.M. Tarpey, [89] R.S. Bel Aiba, E.Y. Dimova, A. Gorlach, T. Kietzmann, The role of hypoxia in-
Hydrogen peroxide is the major oxidant product of xanthine oxidase, Free Radical. ducible factor-1 in cell metabolism–a possible target in cancer therapy, Expert
Bio. Med. 48 (2010) 493–498. Opin. Ther. Targets 10 (2006) 583–599.
[59] I. Fridovich, Quantitative aspects of the production of superoxide anion radical by [90] M. Jahani, K. Mansouri, Role of hypoxia and hypoxia inducible factor in physio-
milk xanthine oxidase, J. Biol. Chem. 245 (1970) 4053–4057. logical and pathological conditions, J. Kermanshah Univ. Med Sci. 21 (2017)
[60] M. Tafani, L. Sansone, F. Limana, T. Arcangeli, E. De Santis, M. Polese, F. Massimo, 126–132.
A.R. Matteo, The interplay of reactive oxygen species, hypoxia, inflammation, and [91] G.N. Masoud, W. Li, HIF-1α pathway: role, regulation and intervention for cancer
sirtuins in cancer initiation and progression, Oxid. Med. Cell. Longev. (2016). therapy, Acta Pharm. Sin. B 5 (2015) 378–389.
[61] M. Allard, B. Schonekess, S. Henning, D. English, G.D. Lopaschuk, Contribution of [92] K. Mansouri, A. Mostafie, D. Rezazadeh, M. Shahlaei, M.H. Modarressi, New
oxidative metabolism and glycolysis to ATP production in hypertrophied hearts, function of TSGA10 gene in angiogenesis and tumor metastasis: a response to a
Am. J. Physiol. Heart Circ. Physiol. 267 (1994) 742–750. challengeable paradox, Hum. Mol. Genet. 25 (2015) 233–244.
[62] M.G. Vander Heiden, L.C. Cantley, C.B. Thompson, Understanding the Warburg [93] P.E. Porporato, R.K. Dadhich, S. Dhup, T. Copetti, P. Sonveaux, Anticancer targets
effect: the metabolic requirements of cell proliferation, Science 324 (2009) in the glycolytic metabolism of tumors: a comprehensive review, Front.
1029–1033. Pharmacol. 2 (2011) 49.
[63] D. Nelson, Lehninger Principles of Biochemistry, 5nd edn, Macmillan, 2008. [94] S.P. Mathupala, A. Rempel, P.L. Pedersen, Glucose catabolism in cancer cells
[64] M. Jahani, F. Noroznezhad, K. Mansouri, Arginine: challenges and opportunities of identification and characterization of a marked activation response of the type II
this two-faced molecule in cancer therapy, Biomed. Pharmacother. 102 (2018) hexokinase gene to hypoxic conditions, J. Biol. Chem. 276 (2001) 43407–43412.
594–601. [95] D.A. Okar, A.J. Lange, Fructose‐2, 6‐bisphosphate and control of carbohydrate
[65] O. Warburg, On the origin of cancer cells, Science 123 (1956) 309–314. metabolism in eukaryotes, Biofactors 10 (1999) 1–14.
[66] S.W. House, O. Warburg, D. Burk, A.L. Schade, On respiratory impairment in [96] D.A. Okar, A.J. Lange, A. Manzano, A. Navarro-Sabate, Ls. Riera, R. Bartrons, PFK-
cancer cells, Science 124 (1956) 267–272. 2/FBPase-2: maker and breaker of the essential biofactor fructose-2, 6-bispho-
[67] Warburg O, Minami S, Versuche an uberlebendem carcinom-gewebe, J. Mol. Med. sphate, Trends Biochem. Sci. 26 (2001) 30–35.
2 (1923) 776–777. [97] A. Minchenko, I. Leshchinsky, I. Opentanova, N. Sang, V. Srinivas, V. Armstead,
[68] O. Warburg, Uber den Stoffwechsel der Carcinomzelle, Klin. Wochenschr. 4 (1925) J. Caro, Hypoxia-inducible Factor-1-mediated Expression of the 6-Phosphofructo-
534–536. 2-kinase/fructose-2, 6-bisphosphatase-3 (PFKFB3) Gene its possible role in the
[69] E. Racker, Bioenergetics and the problem of tumor growth: an understanding of warburg effect, J. Biol. Chem. 277 (2002) 6183–6187.
the mechanism of the generation and control of biological energy may shed light [98] O. Minchenko, I. Opentanova, J. Caro, Hypoxic regulation of the 6‐phospho-
on the problem of tumor growth, Am. Sci. 60 (1972) 56–63. fructo‐2‐kinase/fructose‐2, 6‐bisphosphatase gene family (PFKFB‐1–4) expression
[70] J.C. Kotz, P.M. Treichel, J. Townsend, Chemistry and Chemical Reactivity, 8 ed., in vivo, FEBS Lett. 554 (2003) 264–270.
Cengage Learning, 2012. [99] Jw. Kim, I. Tchernyshyov, G.L. Semenza, C.V. Dang, HIF-1-mediated expression of
[71] X. Tong, F. Zhao, C.B. Thompson, The molecular determinants of de novo nu- pyruvate dehydrogenase kinase: a metabolic switch required for cellular adapta-
cleotide biosynthesis in cancer cells, Curr. Opin. Genet. Dev. 19 (2009) 32–37. tion to hypoxia, Cell Metab. 3 (2006) 177–185.
[72] J.W. Locasale, A.R. Grassian, T. Melman, C.A. Lyssiotis, K.R. Mattaini, A.J. Bass, [100] I. Papandreou, R.A. Cairns, L. Fontana, A.L. Lim, N.C. Denko, HIF-1 mediates
G. Heffron, C.M. Metallo, T. Muranen, H. Sharfi, A.T. Sasaki, D. Anastasiou, adaptation to hypoxia by actively downregulating mitochondrial oxygen con-
E. Mullarky, N.I. Vokes, M. Sasaki, R. Beroukhim, G. Stephanopoulos, A.H. Ligon, sumption, Cell Metab. 3 (2006) 187–197.
M. Meyerson, A.L. Richardson, L. Chin, G. Wagner, J.M. Asara, J.S. Brugge, [101] S. Wigfield, S. Winter, A. Giatromanolaki, J. Taylor, M. Koukourakis, A. Harris,
L.C. Cantley, M.G. Vander Heiden, Phosphoglycerate dehydrogenase diverts gly- PDK-1 regulates lactate production in hypoxia and is associated with poor prog-
colytic flux and contributes to oncogenesis, Nat. Genet. 43 (2011) 869. nosis in head and neck squamous cancer, Br. J. Cancer 98 (2008) 1975.
[73] D. Anastasiou, G. Poulogiannis, J.M. Asara, M.B. Boxer, Jk. Jiang, M. Shen, [102] C.L. Markert, J.B. Shaklee, G.S. Whitt, Evolution of a gene, Science 189 (1975)
G. Bellinger, A.T. Sasaki, J.W. Locasale, D.S. Auld, C.J. Thomas, M.G. Vander 102–114.
Heiden, L.C. Cantley, Inhibition of pyruvate kinase M2 by reactive oxygen species [103] M. Koukourakis, A. Giatromanolaki, E. Sivridis, G. Bougioukas, V. Didilis,
contributes to cellular antioxidant responses, Science 334 (2011) 1278–1283. K. Gatter, A.L. Harris, Lactate dehydrogenase-5 (LDH-5) overexpression in non-
[74] R.B. Hamanaka, N.S. Chandel, Warburg effect and redox balance, Science 334 small-cell lung cancer tissues is linked to tumour hypoxia, angiogenic factor
(2011) 1219–1220. production and poor prognosis, Br. J. Cancer 89 (2003) 877.
[75] M. Lee, J.H. Yoon, Metabolic interplay between glycolysis and mitochondrial [104] M.I. Koukourakis, A. Giatromanolaki, E. Sivridis, K.C. Gatter, T. Trarbach,
oxidation: the reverse Warburg effect and its therapeutic implication, World J. G. Folprecht, M.M. Shi, D. Lebwohl, T. Jalava, D. Laurent, G. Meinhardt,
Biol. Chem. 6 (2015) 148. A.L. Harris, Prognostic and predictive role of lactate dehydrogenase 5 expression
[76] L.L. Grochowski, H. Xu, R.H. White, Ribose-5-phosphate biosynthesis in in colorectal cancer patients treated with PTK787/ZK 222584 (vatalanib) anti-
Methanocaldococcus jannaschii occurs in the absence of a pentose-phosphate angiogenic therapy, Clin. Cancer Res. 17 (2011) 4892–4900.
pathway, J. Bacteriol. 187 (2005) 7382–7389. [105] M.I. Koukourakis, A. Giatromanolaki, S. Winter, R. Leek, E. Sivridis, A.L. Harris,
[77] T. Nunoura, Y. Takaki, J. Kakuta, S. Nishi, J. Sugahara, H. Kazama, G.J. Chee, Lactate dehydrogenase 5 expression in squamous cell head and neck cancer relates
M. Hattori, A. Kanai, H. Atomi, K. Takai, H. Takami, Insights into the evolution of to prognosis following radical or postoperative radiotherapy, Oncology 77 (2009)
Archaea and eukaryotic protein modifier systems revealed by the genome of a 285–292.
novel archaeal group, Nucleic Acids Res. 39 (2010) 3204–3223. [106] A. Leiblich, S. Cross, J. Catto, J. Phillips, H. Leung, F. Hamdy, A.L. Harris, Lactate
[78] C. Brasen, D. Esser, B. Rauch, B. Siebers, Carbohydrate metabolism in Archaea: dehydrogenase-B is silenced by promoter hypermethylation in human prostate
current insights into unusual enzymes and pathways and their regulation, cancer, Oncogene 25 (2006) 2953.

13
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

[107] M. Thangaraju, K.N. Carswell, P.D. Prasad, V. Ganapathy, Colon cancer cells initiating cells, Stem Cells 32 (2014) 1734–1745.
maintain low levels of pyruvate to avoid cell death caused by inhibition of [134] G.J. Yoshida, H. Saya, Inversed relationship between CD44 variant and c-Myc due
HDAC1/HDAC3, Biochem. J. 417 (2009) 379–389. to oxidative stress-induced canonical Wnt activation, Biochem. Biophys. Res.
[108] G. Dong, Q. Mao, W. Xia, Y. Xu, J. Wang, L. Xu, F. Jiang, PKM2 and cancer: the Commun. 443 (2014) 622–627.
function of PKM2 beyond glycolysis, Oncol. Lett. 11 (2016) 1980–1986. [135] G.J. Yoshida, Emerging roles of Myc in stem cell biology and novel tumor thera-
[109] G.L. Semenza, HIF-1 mediates metabolic responses to intratumoral hypoxia and pies, J. Exp. Clin. Cancer Res. 37 (2018) 173.
oncogenic mutations, J. Clin. Invest. 123 (2013) 3664–3671. [136] N. Wong, J. De Melo, D. Tang, PKM2, a central point of regulation in cancer
[110] G.N. Masoud, W. Li, HIF-1α pathway: role, regulation and intervention for cancer metabolism, Int. J. Biochem. Cell Biol. (2013).
therapy, Acta Pharm. Sin. B 5 (2015) 378–389. [137] La Noce M, F. Paino, L. Mele, G. Papaccio, T. Regad, A. Lombardi, F. Papaccio,
[111] S.N. Jung, W.K. Yang, J. Kim, H.S. Kim, E.J. Kim, H. Yun, H. Park, S.S. Kim, V. Desiderio, V. Tirino, HDAC2 depletion promotes osteosarcoma’s stemness both
W. Choe, I. Kang, J. Ha, Reactive oxygen species stabilize hypoxia-inducible in vitro and in vivo: a study on a putative new target for CSCs directed therapy, J.
factor-1 alpha protein and stimulate transcriptional activity via AMP-activated Exp. Clin. Cancer Res. 37 (2018) 296.
protein kinase in DU145 human prostate cancer cells, Carcinogenesis 29 (2008) [138] E. Przybytkowski, D.A. Averill-Bates, Correlation between glutathione and sti-
713–721. mulation of the pentose phosphate cycle in situ in chinese Hamster ovary cells
[112] Dy. Shi, Fz. Xie, C. Zhai, J.S. Stern, Y. Liu, Sl. Liu, The role of cellular oxidative exposed to hydrogen peroxide, Arch. Biochem. Biophys. 325 (1996) 91–98.
stress in regulating glycolysis energy metabolism in hepatoma cells, Mol. Cancer 8 [139] S. Tuttle, T. Stamato, M.L. Perez, J. Biaglow, Glucose-6-phosphate dehydrogenase
(2009) 32. and the oxidative pentose phosphate cycle protect cells against apoptosis induced
[113] T. Hitosugi, S. Kang, M.G. Vander Heiden, T.W. Chung, S. Elf, K. Lythgoe, S. Dong, by low doses of ionizing radiation, Radiat. Res. 153 (2000) 781–787.
S. Lonial, X. Wang, G.Z. Chen, J. Xie, T.L. Gu, R.D. Polakiewicz, J.L. Roesel, [140] G.C. Yeh, S.J. Occhipinti, K.H. Cowan, B.A. Chabner, C.E. Myers, Adriamycin re-
T.J. Boggon, F.R. Khuri, D.G. Gilliland, L.C. Cantley, J. Kaufman, J. Chen, Tyrosine sistance in human tumor cells associated with marked alterations in the regulation
phosphorylation inhibits PKM2 to promote the Warburg effect and tumor growth, of the hexose monophosphate shunt and its response to oxidant stress, Cancer Res.
Sci. Signal. 2 (2009) ra73. 47 (1987) 5994–5999.
[114] H.R. Christofk, M.G. Vander Heiden, M.H. Harris, A. Ramanathan, R.E. Gerszten, [141] C. Friesen, Y. Kiess, K. Debatin, A critical role of glutathione in determining
R. Wei, M.D. Fleming, S.L. Schreiber, L.C. Cantley, The M2 splice isoform of apoptosis sensitivity and resistance in leukemia cells, Cell Death Differ. 11
pyruvate kinase is important for cancer metabolism and tumour growth, Nature (2004) 73.
452 (2008) 230. [142] T. Gessner, L.A. Vaughan, B.C. Beehler, C.J. Bartels, R.M. Baker, Elevated pentose
[115] M.G. Vander Heiden, J.W. Locasale, K.D. Swanson, H. Sharfi, G.J. Heffron, cycle and glucuronyltransferase in daunorubicin-resistant P388 cells, Cancer Res.
D. Amador-Noguez, H.R. Christofk, G. Wagner, J.D. Rabinowitz, J.M. Asara, 50 (1990) 3921–3927.
L.C. Cantley, Evidence for an alternative glycolytic pathway in rapidly pro- [143] L. Mele, F. Paino, F. Papaccio, T. Regad, D. Boocock, P. Stiuso, A. Lombardi,
liferating cells, Science 329 (2010) 1492–1499. D. Liccardo, G. Aquino, A. Barbieri, C. Arra, C. Coveney, M. La Noce, G. Papaccio,
[116] N.M. Gruning, M. Rinnerthaler, K. Bluemlein, M. Mulleder, M.M. Wamelink, M. Caraglia, V. Tirino, V. Desiderio, A new inhibitor of glucose-6-phosphate de-
H. Lehrach, C. Jakobs, M. Breitenbach, M. Ralser, Pyruvate kinase triggers a me- hydrogenase blocks pentose phosphate pathway and suppresses malignant pro-
tabolic feedback loop that controls redox metabolism in respiring cells, Cell Metab. liferation and metastasis in vivo, Cell Death Dis. 9 (2018) 572.
14 (2011) 415–427. [144] S.Y. Lee, E.K. Jeong, M.K. Ju, H.M. Jeon, M.Y. Kim, C.H. Kim, H.G. Park, S.I. Han,
[117] R.A. Cairns, I.S. Harris, T.W. Mak, Regulation of cancer cell metabolism, Nat. Rev. H.S. Kang, Induction of metastasis, cancer stem cell phenotype, and oncogenic
Cancer 11 (2011) 85. metabolism in cancer cells by ionizing radiation, Mol. Cancer 16 (2017) 10.
[118] A. Kruger, N.M. Gruning, M.M. Wamelink, M. Kerick, A. Kirpy, D. Parkhomchuk, [145] A. Mozumder, Y. Hatano, Charged Particle and Photon Interactions With Matter:
K. Bluemlein, M.R. Schweiger, A. Soldatov, H. Lehrach, C. Jakobs, M. Ralser, The Chemical, Physicochemical, and Biological Consequences With Applications, 1 ed.,
pentose phosphate pathway is a metabolic redox sensor and regulates transcrip- CRC Press, 2003.
tion during the antioxidant response, Antioxid. Redox Signal. 15 (2011) 311–324. [146] J. Meesungnoen, J. Jay-Gerin, Radiation Chemistry of Liquid Water With Heavy
[119] J.K. Brunelle, E.L. Bell, N.M. Quesada, K. Vercauteren, V. Tiranti, M. Zeviani, Ions: Monte Carlo Simulation Studies, CRC Press, Taylor & Francis, 2011, pp.
R.C. Scarpulla, N.S. Chandel, Oxygen sensing requires mitochondrial ROS but not 355–400.
oxidative phosphorylation, Cell Metab. 1 (2005) 409–414. [147] J.H. Tsai, J. Yang, Epithelial–mesenchymal plasticity in carcinoma metastasis,
[120] M. Ralser, M.M. Wamelink, A. Kowald, B. Gerisch, G. Heeren, E.A. Struys, E. Klipp, Genes Dev. 27 (2013) 2192–2206.
C. Jakobs, M. Breitenbach, H. Lehrach, S. Krobitsch, Dynamic rerouting of the [148] B. De Craene, G. Berx, Regulatory networks defining EMT during cancer initiation
carbohydrate flux is key to counteracting oxidative stress, J. Biol. 6 (2007) 10. and progression, Nat. Rev. Cancer 13 (2013) 97.
[121] D. Shenton, C.M. Grant, Protein S-thiolation targets glycolysis and protein [149] S. Lamouille, J. Xu, R. Derynck, Molecular mechanisms of epithelial–mesenchymal
synthesis in response to oxidative stress in the yeast Saccharomyces cerevisiae, transition, Nat. Rev. Mol. Cell Biol. 15 (2014) 178.
Biochem. J. 374 (2003) 513–519. [150] C. Lagadec, E. Vlashi, L. Della Donna, C. Dekmezian, F. Pajonk, Radiation‐induced
[122] C. Godon, G. Lagniel, J. Lee, J.M. Buhler, S. Kieffer, M. Perrot, H. Boucherie, reprogramming of breast cancer cells, Stem Cells 30 (2012) 833–844.
M.B. Toledano, J. Labarre, The H2O2 stimulon in Saccharomyces cerevisiae, J. [151] M. Diehn, R.W. Cho, N.A. Lobo, T. Kalisky, M.J. Dorie, A.N. Kulp, D. Qian,
Biol. Chem. 273 (1998) 22480–22489. J.S. Lam, L.E. Ailles, M. Wong, B. Joshua, M.J. Kaplan, I. Wapnir, F.M. Dirbas,
[123] C.G. Cochrane, Cellular injury by oxidants, Am. J. Med. 91 (1991) 23–30. G. Somlo, C. Garberoglio, B. Paz, J. Shen, S.K. Lau, S.R. Quake, J.M. Brown,
[124] P. Hyslop, D. Hinshaw, Jr.W. Halsey, I. Schraufstatter, R. Sauerheber, R. Spragg, I.L. Weissman, M.F. Clarke, Association of reactive oxygen species levels and
J.H. Jackson, C.G. Cochrane, The glycolytic and mitochondrial pathways of ADP radioresistance in cancer stem cells, Nature 458 (2009) 780–783.
phosphorylation are major intracellular targets inactivated by hydrogen peroxide, [152] H.Q. Duong, K.S. You, S. Oh, S.J. Kwak, Y.S. Seong, Silencing of NRF2 reduces the
J. Biol. Chem. 263 (1988) 1665–1675. expression of ALDH1A1 and ALDH3A1 and sensitizes to 5-FU in pancreatic cancer
[125] E. Mullarky, L.C. Cantley, Diverting glycolysis to combat oxidative stress, cells, Antioxidants 6 (2017) 52.
InnoMed. 2015 (2015) 3–23. [153] G.R. Buettner, One-electron and two-electron signaling in redox biology, Free
[126] M.A. Kowalik, A. Columbano, A. Perra, Emerging role of the pentose phosphate Radic. Biol. Med. 49 (2010) S7.
pathway in hepatocellular carcinoma, Front. Oncol. 7 (2017) 87. [154] M. Lopez-Lazaro, HIF-1: hypoxia-inducible factor or dysoxia-inducible factor?
[127] N. Verma, M. Pink, S. Boland, A.W. Rettenmeier, S. Schmitz-Spanke, Benzo [a] FASEB J. 20 (2006) 828–832.
pyrene-induced metabolic shift from glycolysis to pentose phosphate pathway in [155] R.D. Guzy, B. Hoyos, E. Robin, H. Chen, L. Liu, K.D. Mansfield, M.C. Simon,
the human bladder cancer cell line RT4, Sci. Rep. 7 (2017) 9773. U. Hammerling, P.T. Schumacker, Mitochondrial complex III is required for hy-
[128] E. Vlashi, C. Lagadec, L. Vergnes, T. Matsutani, K. Masui, M. Poulou, R. Popescu, poxia-induced ROS production and cellular oxygen sensing, Cell Metab. 1 (2005)
L. Della Donna, P. Evers, C. Dekmezian, K. Reue, H. Christofk, P.S. Mischel, 401–408.
F. Pajonk, Metabolic state of glioma stem cells and nontumorigenic cells, Proc. [156] K.D. Mansfield, R.D. Guzy, Y. Pan, R.M. Young, T.P. Cash, P.T. Schumacker,
Natl. Acad. Sci. U. S. A. 108 (2011) 16062–16067. M.C. Simon, Mitochondrial dysfunction resulting from loss of cytochrome c im-
[129] E.D. Lagadinou, A. Sach, K. Callahan, R.M. Rossi, S.J. Neering, M. Minhajuddin, pairs cellular oxygen sensing and hypoxic HIF-α activation, Cell Metab. 1 (2005)
J.M. Ashton, S. Pei, V. Grose, K.M. O’Dwyer, J.L. Liesveld, P.S. Brookes, 393–399.
M.W. Becker, C.T. Jordan, BCL-2 inhibition targets oxidative phosphorylation and [157] Q. Cao, K.M. Mak, C. Ren, C.S. Lieber, Leptin stimulates tissue inhibitor of me-
selectively eradicates quiescent human leukemia stem cells, Cell Stem Cell 12 talloproteinase-1 in human hepatic stellate cells respective roles of the JAK/STAT
(2013) 329–341. and JAK-Mediated H2O2-Dependent MAPK pathways, J. Biol. Chem. 279 (2004)
[130] E. Vlashi, C. Lagadec, L. Vergnes, K. Reue, P. Frohnen, M. Chan, Y. Alhiyari, 4292–4304.
M.B. Dratver, F. Pajonk, Metabolic differences in breast cancer stem cells and [158] D. Trachootham, W. Lu, M.A. Ogasawara, N.R.D. Valle, P. Huang, Redox regula-
differentiated progeny, Breast Cancer Res. Treat. 146 (2014) 525–534. tion of cell survival, Antioxid. Redox Signal. 10 (2008) 1343–1374.
[131] D. Ciavardelli, C. Rossi, D. Barcaroli, S. Volpe, A. Consalvo, M. Zucchelli, A. De [159] S. Kaewpila, S. Venkataraman, G.R. Buettner, L.W. Oberley, Manganese super-
Cola, E. Scavo, R. Carollo, D. DAgostino, F. ForlI, S. DAguanno, M. Todaro, oxide dismutase modulates hypoxia-inducible factor-1α induction via superoxide,
G. Stassi, C. Di Ilio, V. De Laurenzi, A. Urbani, Breast cancer stem cells rely on Cancer Res. 68 (2008) 2781–2788.
fermentative glycolysis and are sensitive to 2-deoxyglucose treatment, Cell Death [160] Q. Liu, U. Berchner-Pfannschmidt, U. Moller, M. Brecht, C. Wotzlaw, H. Acker,
Dis. 5 (2014) e1336. K. Jungermann, T. Kietzmann, A Fenton reaction at the endoplasmic reticulum is
[132] Y.A. Shen, C.Y. Wang, Y.T. Hsieh, Y.J. Chen, Y.H. Wei, Metabolic reprogramming involved in the redox control of hypoxia-inducible gene expression, Proc. Natl.
orchestrates cancer stem cell properties in nasopharyngeal carcinoma, Cell Cycle Acad. Sci. U. S. A. 101 (2004) 4302–4307.
14 (2015) 86–98. [161] H. Shi, Hypoxia inducible factor 1 as a therapeutic target in ischemic stroke, Curr.
[133] W. Feng, A. Gentles, R.V. Nair, M. Huang, Y. Lin, C.Y. Lee, S. Cai, F.A. Scheeren, Med. Chem. 16 (2009) 4593–4600.
A.H. Kuo, M. Diehn, Targeting unique metabolic properties of breast tumor [162] S.J. Welsh, W.T. Bellamy, M.M. Briehl, G. Powis, The redox protein thioredoxin-1

14
Z. Ghanbari Movahed, et al. Biomedicine & Pharmacotherapy 112 (2019) 108690

(Trx-1) increases hypoxia-inducible factor 1α protein expression: Trx-1 over- downregulation of the hypoxia marker carbonic anhydrase IX, Mol. Cell. Biol. 24
expression results in increased vascular endothelial growth factor production and (2004) 5757–5766.
enhanced tumor angiogenesis, Cancer Res. 62 (2002) 5089–5095. [184] M. Jahani, M. Azadbakht, F. Norooznezhad, K. Mansouri, L-arginine alters the
[163] S. Salceda, J. Caro, Hypoxia-inducible factor 1α (HIF-1α) protein is rapidly de- effect of 5-fluorouracil on breast cancer cells in favor of apoptosis, Biomed.
graded by the ubiquitin-proteasome system under normoxic conditions its stabi- Pharmacother. 88 (2017) 114–123.
lization by hypoxia depends on redox-induced changes, J. Biol. Chem. 272 (1997) [185] Z.N. Demidenko, A. Rapisarda, M. Garayoa, P. Giannakakou, G. Melillo,
22642–22647. M.V. Blagosklonny, Accumulation of hypoxia-inducible factor-1α is limited by
[164] D.J. Lomb, L.A. Desouza, J.L. Franklin, R.S. Freeman, Prolyl hydroxylase inhibitors transcription-dependent depletion, Oncogene 24 (2005) 4829.
depend on extracellular glucose and hypoxia-inducible factor (HIF)-2α to inhibit [186] S. Chang, X. Jiang, C. Zhao, C. Lee, D.M. Ferriero, Exogenous low dose hydrogen
cell death caused by nerve growth factor (NGF) deprivation: evidence that HIF-2α peroxide increases hypoxia-inducible factor-1alpha protein expression and induces
has a role in NGF-promoted survival of sympathetic neurons, Mol. Pharmacol. 75 preconditioning protection against ischemia in primary cortical neurons, Neurosci.
(2009) 1198–1209. Lett. 441 (2008) 134–138.
[165] T.L. Wellman, J. Jenkins, P.L. Penar, B. Tranmer, R. Zahr, K.M. Lounsbury, Nitric [187] E.L. Bell, T.A. Klimova, J. Eisenbart, C.T. Moraes, M.P. Murphy, G.S. Budinger,
oxide and reactive oxygen species exert opposing effects on the stability of hy- N.S. Chandel, The Qo site of the mitochondrial complex III is required for the
poxia-inducible factor-1α (HIF-1α) in explants of human pial arteries, FASEB J. 18 transduction of hypoxic signaling via reactive oxygen species production, J. Cell
(2004) 379–381. Biol. 177 (2007) 1029–1036.
[166] A. Gorlach, U. Berchner-Pfannschmidt, C. Wotzlaw, R.H. Cool, J. Fandrey, [188] E.L. Bell, T.A. Klimova, J. Eisenbart, P.T. Schumacker, N.S. Chandel,
H. Acker, K. Jungermann, T. Kietzmann, Reactive oxygen species modulate HIF-1 Mitochondrial reactive oxygen species trigger hypoxia-inducible factor-dependent
mediated PAI-1 expression: involvement of the GTPase Rac1, Thromb. Haemost. extension of the replicative life span during hypoxia, Mol. Cell. Biol. 27 (2007)
89 (2003) 926–935. 5737–5745.
[167] Q. Ding, E. Dimayuga, J.N. Keller, Proteasome regulation of oxidative stress in [189] N.S. Chandel, D.S. McClintock, C.E. Feliciano, T.M. Wood, J.A. Melendez,
aging and age-related diseases of the CNS, Antioxid. Redox Signal. 8 (2006) A.M. Rodriguez, P.T. Schumacker, Reactive oxygen species generated at mi-
163–172. tochondrial complex III stabilize hypoxia-inducible factor-1α during hypoxia a
[168] A. Ciechanover, The ubiquitin proteolytic system from a vague idea, through basic mechanism of O2 sensing, J. Biol. Chem. 275 (2000) 25130–25138.
mechanisms, and onto human diseases and drug targeting, Neurology 66 (2006) [190] R.D. Guzy, B. Sharma, E. Bell, N.S. Chandel, P.T. Schumacker, Loss of the SdhB,
7–19. but not the SdhA, subunit of complex II triggers reactive oxygen species-dependent
[169] K.D. Wilkinson, Roles of ubiquitinylation in proteolysis and cellular regulation, hypoxia-inducible factor activation and tumorigenesis, Mol. Cell. Biol. 28 (2008)
Annu. Rev. Nutr. 15 (1995) 161–189. 718–731.
[170] C. Brahimi-Horn, J. Pouyssegur, When hypoxia signalling meets the ubiquitin- [191] X. Lin, C.A. David, J.B. Donnelly, M. Michaelides, N.S. Chandel, X. Huang,
proteasomal pathway, new targets for cancer therapy, Crit. Rev. Oncol. Hematol. U. Warrior, F. Weinberg, K.V. Tormos, S.W. Fesik, A. Shen, Chemical genomics
53 (2005) 115–123. screen highlights the essential role of mitochondria in HIF-1 regulation, Proc. Natl.
[171] R. Shringarpure, T. Grune, J. Mehlhase, K.J. Davies, Ubiquitin conjugation is not Acad. Sci. U. S. A. 105 (2008) 174–179.
required for the degradation of oxidized proteins by proteasome, J. Biol. Chem. [192] E.L. Page, D.A. Chan, A.J. Giaccia, M. Levine, D.E. Richard, Hypoxia-inducible
278 (2003) 311–318. factor-1α stabilization in nonhypoxic conditions: role of oxidation and in-
[172] T. Grune, T. Reinheckel, K.J. Davies, Degradation of oxidized proteins in K562 tracellular ascorbate depletion, Mol. Biol. Cell 19 (2008) 86–94.
human hematopoietic cells by proteasome, J. Biol. Chem. 271 (1996) [193] V. Srinivas, I. Leshchinsky, N. Sang, M.P. King, A. Minchenko, J. Caro, Oxygen
15504–15509. sensing and HIF-1 activation does not require an active mitochondrial respiratory
[173] T. Grune, T. Reinheckel, K. Davies, Degradation of oxidized proteins in mamma- chain electron-transfer pathway, J. Biol. Chem. 276 (2001) 21995–21998.
lian cells, FASEB J. 11 (1997) 526–534. [194] M. Demaria, V. Poli, PKM2, STAT3 and HIF-1α: The Warburg’s vicious circle, Jak-
[174] T. Grune, T. Reinheckel, M. Joshi, K.J. Davies, Proteolysis in cultured liver epi- Stat. 1 (2012) 194–196.
thelial cells during oxidative stress Role of the multicatalytic proteinase complex, [195] I. Schuppe‐Koistinen, P. Moldeus, T. Bergman, I.A. Cotgreave, S‐Thiolation of
proteasome, J. Biol. Chem. 270 (1995) 2344–2351. human endothelial cell glyceraldehyde‐3‐phosphate dehydrogenase after hy-
[175] J.J. Haddad, Antioxidant and prooxidant mechanisms in the regulation of redox drogen peroxide treatment, FEBS J. 221 (1994) 1033–1037.
(y)-sensitive transcription factors, Cell. Signal. 14 (2002) 879–897. [196] L. Cyrne, F. Antunes, A. Sousa-Lopes, J. Diaz-Berrio, H.S. Marinho,
[176] M. Callapina, J. Zhou, T. Schmid, R. Kohl, B. Brune, NO restores HIF-1α hydro- Glyceraldehyde-3-phosphate dehydrogenase is largely unresponsive to low reg-
xylation during hypoxia: role of reactive oxygen species, Free Radic. Biol. Med. 39 ulatory levels of hydrogen peroxide in Saccharomyces cerevisiae, BMC Biochem.
(2005) 925–936. 11 (2010) 49.
[177] H. Elkon, E. Melamed, D. Offen, Oxidative stress, induced by 6-hydroxydopamine, [197] L. Cao, J. Chen, M. Li, Y.Y. Qin, M. Sun, R. Sheng, F. Han, G. Wang, Z.H. Qin,
reduces proteasome activities in PC12 cells, J. Mol. Neurosci. 24 (2004) 387–400. Endogenous level of TIGAR in brain is associated with vulnerability of neurons to
[178] T. Reinheckel, N. Sitte, O. Ullrich, U. Kuckelkorn, Kj. Davies, T. Grune, ischemic injury, Neurosci. Bull. 31 (2015) 527–540.
Comparative resistance of the 20S and 26S proteasome to oxidative stress, [198] J. Xiang, C. Wan, R. Guo, D. Guo, Is hydrogen peroxide a suitable apoptosis in-
Biochem. J. 335 (1998) 637–642. ducer for all cell types? Biomed Res. Int. 2016 (2016) 1–6.
[179] O. Ullrich, T. Reinheckel, N. Sitte, R. Hass, T. Grune, K.J. Davies, Poly-ADP ribose [199] M. Singh, H. Sharma, N. Singh, Hydrogen peroxide induces apoptosis in HeLa cells
polymerase activates nuclear proteasome to degrade oxidatively damaged his- through mitochondrial pathway, Mitochondrion 7 (2007) 367–373.
tones, Proc. Natl. Acad. Sci. U. S. A. 6 (1999) 6223–6228. [200] O.D. Maddocks, C.F. Labuschagne, K.H. Vousden, Localization of NADPH pro-
[180] G.L. Wang, B.H. Jiang, G.L. Semenza, Effect of altered redox states on expression duction: a wheel within a wheel, Mol. Cell 55 (2014) 158–160.
and DNA-binding activity of hypoxia-inducible factor 1, Biochem. Biophys. Res. [201] R.J. DeBerardinis, N.S. Chandel, Fundamentals of cancer metabolism, Sci. Adv. 2
Commun. 212 (1995) 550–556. (2016) e1600200.
[181] B.M. Emerling, L.C. Platanias, E. Black, A.R. Nebreda, R.J. Davis, N.S. Chandel, [202] D. Samanta, G.L. Semenza, Serine synthesis helps hypoxic cancer stem cells reg-
Mitochondrial reactive oxygen species activation of p38 mitogen-activated protein ulate redox, Cancer Res. 76 (2016) 6458–6462.
kinase is required for hypoxia signaling, Mol. Cell. Biol. 25 (2005) 4853–4862. [203] D. Samanta, G.L. Semenza, Maintenance of redox homeostasis by hypoxia-in-
[182] X. Kong, B. Alvarez-Castelao, Z. Lin, J.G. Castano, J. Caro, Constitutive/hypoxic ducible factors, Redox Biol. 13 (2017) 331–335.
degradation of HIF-α proteins by the proteasome is independent of von Hippel [204] S.M. Lee, T.L. Huh, J.W. Park, Inactivation of NADP+-dependent isocitrate de-
Lindau protein ubiquitylation and the transactivation activity of the protein, J. hydrogenase by reactive oxygen species, Biochimie 83 (2001) 1057–1065.
Biol. Chem. 282 (2007) 15498–15505. [205] K.M. Holmström, T. Finkel, Cellular mechanisms and physiological consequences
[183] M. Kaluzova, S. Kaluz, M.I. Lerman, E.J. Stanbridge, DNA damage is a prerequisite of redox-dependent signalling, Nat. Rev. Mol. Cell Biol. 15 (2014) 411–421.
for p53-mediated proteasomal degradation of HIF-1α in hypoxic cells and

15

You might also like