Sunteți pe pagina 1din 9

Cancer Letters 356 (2015) 156164

Contents lists available at ScienceDirect

Cancer Letters
journal homepage: www.elsevier.com/locate/canlet

Mini-review

The Warburg effect in tumor progression: Mitochondrial oxidative


metabolism as an anti-metastasis mechanism
Jianrong Lu a,, Ming Tan b, Qingsong Cai a
a
Department of Biochemistry and Molecular Biology, UF Health Cancer Center, University of Florida College of Medicine, Gainesville, FL 32610, United States
b
Mitchell Cancer Institute, University of South Alabama, Mobile, AL 36604, United States

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

Article history: Compared to normal cells, cancer cells strongly upregulate glucose uptake and glycolysis to give rise to
Received 24 January 2014 increased yield of intermediate glycolytic metabolites and the end product pyruvate. Moreover, glycolysis
Received in revised form 17 March 2014 is uncoupled from the mitochondrial tricarboxylic acid (TCA) cycle and oxidative phosphorylation
Accepted 5 April 2014
(OXPHOS) in cancer cells. Consequently, the majority of glycolysis-derived pyruvate is diverted to lactate
fermentation and kept away from mitochondrial oxidative metabolism. This metabolic phenotype is
known as the Warburg effect. While it has become widely accepted that the glycolytic intermediates pro-
Keywords:
vide essential anabolic support for cell proliferation and tumor growth, it remains largely elusive whether
Warburg effect
Glycolysis
and how the Warburg metabolic phenotype may play a role in tumor progression. We hereby review the
OXPHOS cause and consequence of the restrained oxidative metabolism, in particular in the context of tumor
ROS metastasis. Cells change or lose their extracellular matrix during the metastatic process. Inadequate/inap-
Anoikis propriate matrix attachment generates reactive oxygen species (ROS) and causes a specic type of cell
Metastasis death, termed anoikis, in normal cells. Although anoikis is a barrier to metastasis, cancer cells have often
acquired elevated threshold for anoikis and hence heightened metastatic potential. As ROS are inherent
byproducts of oxidative metabolism, forced stimulation of glucose oxidation in cancer cells raises oxida-
tive stress and restores cells sensitivity to anoikis. Therefore, by limiting the pyruvate ux into mitochon-
drial oxidative metabolism, the Warburg effect enables cancer cells to avoid excess ROS generation from
mitochondrial respiration and thus gain increased anoikis resistance and survival advantage for metasta-
sis. Consistent with this notion, pro-metastatic transcription factors HIF and Snail attenuate oxidative
metabolism, whereas tumor suppressor p53 and metastasis suppressor KISS1 promote mitochondrial
oxidation. Collectively, these ndings reveal mitochondrial oxidative metabolism as a critical suppressor
of metastasis and justify metabolic therapies for potential prevention/intervention of tumor metastasis.
2014 Elsevier Ireland Ltd. All rights reserved.

Introduction: the Warburg effect in cancer


Abbreviations: 2-DG, 2-deoxy-D-glucose; CoA, coenzyme A; COX, cytochrome c
oxidase; DCA, Dichloroacetate; EMT, epithelial-to-mesenchymal transition; ERR,
Altered metabolism is a universal property of most, if not all,
estrogen-related receptor; ETC, electron transport chain; F1, 6BP, fructose 1,6-
bisphosphate; F2, 6BP, fructose-2,6-bisphosphate; F6P, fructose-6-phosphate; FAO,
cancer cells [1,2]. One of the rst identied and most common bio-
fatty acid oxidation; FBP1, fructose-1,6-biphosphatase 1; FDG, 18-uorodeoxyglu- chemical characteristics of cancer cells is aberrant glucose metab-
cose; GAPDH, glyceraldehyde 3-phosphate dehydrogenase; GSH, reduced glutathi- olism. Glucose is a main source of energy and carbon for
one; HIF, hypoxia-inducible factor; IMS, mitochondrial intermembranous space; mammalian cells, providing not only energy (ATP) but also metab-
ISCU, iron-sulfur cluster assembly protein; LDH, lactate dehydrogenase; MnSOD,
olites for various anabolic pathways [3]. Glucose is taken up into
manganese superoxide dismutase; MOMP, mitochondrial outer membrane perme-
abilization; NAD+, nicotinamide adenine dinucleotide; NAMPT, nicotinamide the cell by glucose transporters and metabolized to pyruvate in
phosphoribosyltransferase; OXPHOS, oxidative phosphorylation; PDH, pyruvate the cytosol through a multi-step process known as glycolysis,
dehydrogenase; PDK, pyruvate dehydrogenase kinases; PDP, pyruvate dehydroge- which also yields a small amount of ATP. In normal (quiescent)
nase phosphatases; PET, positron emission tomography; PFK, phosphofructokinase; cells, the glycolysis-derived pyruvate is predominantly imported
PKM, pyruvate kinase; PPP, pentose phosphate pathway; ROS, reactive oxygen
species; TCA, tricarboxylic acid.
into the mitochondrial matrix where it is oxidized to acetyl coen-
Corresponding author. Tel.: +1 352 273 8200. zyme A (CoA) by the pyruvate dehydrogenase (PDH) complex.
E-mail address: jrlu@u.edu (J. Lu). Acetyl CoA is then fed into the tricarboxylic acid (TCA) cycle,

http://dx.doi.org/10.1016/j.canlet.2014.04.001
0304-3835/ 2014 Elsevier Ireland Ltd. All rights reserved.
J. Lu et al. / Cancer Letters 356 (2015) 156164 157

Fig. 1. Schematic illustration of glucose metabolism in normal and cancer cells under normoxia. (left) In normal (quiescent) cells, glucose is converted to pyruvate through
glycolysis, and most pyruvate enters mitochondrial oxidative metabolism for efcient energy generation (in the form of ATP). Glucose is predominantly used for energy
production. High levels of ATP attenuate glycolysis via feedback inhibition. (right) Cancer cells dramatically increase glucose uptake and glycolysis (indicated by bold arrows).
A signicant portion of glucose carbon is diverted to biosynthetic pathways to fuel cell proliferation. Pyruvate is preferentially shunted to lactate, resulting in increased
lactate production. Oxidative metabolism persists, but is uncoupled from increased glycolysis. The respiration byproducts ROS exhibit anti-metastasis activity, which may
explain why cancer cells keep glucose oxidation in check. The ux of glucose carbon is indicated by green arrows (The thickness of arrows reects the relative amount of the
ow). Major mitochondrial products are depicted in red; metastatic regulators are in purple; regulatory steps are in blue; and the metabolic effects on cancer are in pink. (For
interpretation of the references to colour in this gure legend, the reader is referred to the web version of this article.)

followed by oxidative phosphorylation (OXPHOS) for high-ef- regulatory mechanisms downregulating oxidative metabolism of
ciency ATP generation. The full oxidation of one molecule of glu- glucose in cancer cells (see below). Since a considerable portion
cose produces up to 38 ATP molecules (including 2 ATP of observed oxygen consumption in cancer cells may be attributed
generated by glycolysis). to oxidation of alternate fuels such as glutamine [8,9], the actual
By contrast, most cancer cells show conspicuous alterations in glucose oxidation in cancer cells is probably even lower.
glucose metabolism (Fig. 1): (i) Compared to normal cells, cancer The cause and consequence of the Warburg effect have been at
cells typically exhibit drastically increased glucose uptake and gly- the center of cancer metabolism study. Cancer is a disease arising
colytic rates. Increased glucose consumption generates more inter- from genetic and epigenetic alterations in oncogenes and tumor
mediate glycolytic metabolites and a signicant amount of ATP suppressors, many of which are also able to reprogram metabolism.
from glycolysis. (ii) Moreover, a substantial fraction of glucose car- The metabolic changes occurring in cancer were thus considered a
bon, in the form of assorted glycolytic intermediates, is shunted secondary effect to the transformation process. However, as rapid
into multiple biosynthetic pathways instead of giving rise to pyru- cell proliferation requires accelerated production of the basic cellu-
vate. (iii) Finally, following glycolysis, most pyruvate is converted lar building blocks for assembling new cells, it has now become
to lactate in the cytoplasm by the action of lactate dehydrogenase well recognized that alterations in cellular metabolism in turn fuel
(LDH) and secreted, rather than being oxidized through mitochon- tumor growth by maximally producing substrates for biosynthesis
drial metabolism. This occurs even in the presence of sufcient [3]. Glycolytic breakdown of glucose produces various intermediate
oxygen to support mitochondrial respiration. The metabolic phe- metabolites, which as precursors can be diverted into anabolic
nomenon was rst described by Otto Warburg and is referred to pathways including the pentose phosphate pathway (PPP), serine
as aerobic glycolysis or the Warburg effect [4]. Although human and triacylglycerol synthesis pathways for the de novo synthesis
cancers display a diverse range of metabolic proles [5], the War- of nucleotides, amino acids, and lipids [10]. While normal cells
burg metabolic phenotype is a widespread cancer-associated trait. metabolize glucose almost exclusively for maximal energy produc-
Indeed, enhanced glucose uptake by cancer cells has become the tion (though full energy extraction deprives cells from biosynthesis
basis for positron emission tomography (PET) with 18-uorode- of building blocks), cancer cells boost glucose consumption primar-
oxyglucose (FDG), which preferentially accumulates in tumor cells ily to provide a constant supply of glycolytic intermediates to sat-
as a result of their rapid uptake of glucose. Because of the preva- isfy the anabolic need of dividing cells. In this regard, glycolytic
lence of this phenotype, PET is an effective clinical imaging tech- intermediates seem to be more important than the nal product
nique to detect most cancers and monitor therapeutic responses. pyruvate. Cancer cells indeed use a variety of strategies to slow
It is noteworthy that mitochondrial function in most cancer down the last step of glycolysis that is catalyzed by pyruvate kinase
cells is intact. Warburg observed that the absolute rate of mito- (PKM) [11], allowing buildup of glycolytic intermediates for biosyn-
chondrial respiration in cancer cells remains comparable to that thesis. The altered glucose metabolism thus favors the conversion
of normal cells [4]. Oxidative metabolism indeed persists in the of glucose into biomass and sustains the highly proliferative nature
vast majority of tumors and remains a major source for ATP gener- of cancer [3]. Taken together, while alterations in oncogenes and
ation [6,7]. Nevertheless, while there is a profoundly elevated ux tumor suppressors drive inappropriate cell proliferation, they also
of glucose in cancer cells, it is predominantly directed to lactate concomitantly rewire and coordinate cellular metabolism to meet
fermentation, and the ow to oxidative metabolism does not the biosynthetic demands of continuous cell division.
increase proportionally. Simply put, increased glucose consump- Although the proliferative advantage offered by the Warburg
tion in cancer cells is devoted to lactate conversion and biosynthe- metabolic phenotype is well established, its signicance in
sis, and is uncoupled from oxidative metabolism. In fact, there exist metastatic progression has been much less clear. Moreover, if the
158 J. Lu et al. / Cancer Letters 356 (2015) 156164

principal goal of aerobic glycolysis is to produce glycolytic inter- F2,6BP is generated from F6P through phosphorylation by phos-
mediates for anabolic support for cell proliferation, it remains elu- phofructokinase-2 (PFK2). Increased glucose uptake and hexoki-
sive why after glycolysis, pyruvate is preferentially disposed of as nase activity leads to elevated levels of F6P, and consequently
secreted lactate rather than utilized through the TCA cycle in the yielding more F2,6BP. F2,6BP in turn allosterically stimulates
mitochondria [12]. This appears to be an inefcient use of carbon PFK1 and glycolysis. This represents a feedforward mechanism as
resources because acetyl-CoA is a major carbon provider for fatty glycolysis is enhanced when glucose is abundant. The steady-state
acid synthesis. TCA cycle reactions generate biosynthetic precur- concentrations of F2,6BP depend on the enzymatic activity of PFK2.
sors in a process termed cataplerosis [13]. When uncoupled from PFK2 is a bifunctional enzyme that catalyzes either the phosphor-
OXPHOS, intermediates of the TCA cycle (e.g. citrate) are diverted ylation of F6P to F2,6BP (kinase activity) or the dephosphorylation
into anabolic pathways and are essential for rapid tumor growth of F2,6BP to F6P (phosphatase activity). The enzyme exists in mul-
[1]. To compensate for the limited supply from glucose/pyruvate, tiple isozymic forms, encoded by different genes, which differ in
cancer cells increase consumption of glutamine to replenish the kinetics and regulatory properties. Among all PFK2 enzymes,
TCA cycle [8]. Therefore, it appears that cancer cells purposefully PFKFB3 has high kinase activity and almost no phosphatase activ-
restrain pyruvate from entry into mitochondrial oxidative ity, thereby exhibiting the highest kinase/phosphatase activity
metabolism. ratio, and preferentially driving the synthesis of F2,6BP. PFKFB3
is commonly overexpressed in human cancers and sustains high-
rate glycolysis [17]. PFK1 and glycolysis in such cancer cells are
Is glucose oxidation incompatible with high glycolysis in cancer expected to be largely refractory to negative feedback by the TCA
cells? cycle. There is also argument that citrate generated by mitochon-
drial TCA cycle is mainly used for lipid synthesis after exported into
One reason that cancer cells limit oxidative metabolism of glu- the cytosol, and never builds up to a sufcient concentration to
cose may be that it is not compatible with high rates of glycolysis. inhibit PFK1. As a result of these mechanisms, cancer cells can
The TCA cycle is a hub of metabolism, with central importance in potentially overcome the feedback inhibition by glucose oxidation
both energy production and biosynthesis. Cells must control the and maintain high levels of PFK1 activity and glycolysis.
TCA cycle to regulate energy balance and TCA metabolite levels
in the mitochondria. This regulation is fullled by feedback inhibi-
tion. As an important part of metabolism, metabolic ux through Is glucose oxidation limited because high glycolysis relies on
the glycolytic pathway is tightly regulated to respond to environ- NAD+ regenerated from increased lactate conversion?
ment and to meet the bioenergetic and anabolic needs.
A primary control point in glycolysis is the conversion of fruc- Glycolysis is a catabolic pathway that consumes nicotinamide
tose-6-phosphate (F6P) to fructose 1,6-bisphosphate (F1,6BP), a adenine dinucleotide (NAD+). In glycolysis, oxidation of glyceralde-
rate-limiting and irreversible early reaction in the pathway. This hyde 3-phosphate requires NAD+ as an electron acceptorit con-
step is catalyzed by phosphofructokinase-1 (PFK1), the prominent verts to NADH. This step is catalyzed by glyceraldehyde 3-
pacemaker of glycolysis. PFK1 is an allosteric enzyme and is sub- phosphate dehydrogenase (GAPDH) and NAD+ is a mandatory
ject to extensive regulation by various metabolites in the glucose coenzyme. A recent study shows that reduced availability of
metabolism pathways. PFK1 is controlled through allosteric activa- NAD+ attenuates glycolysis at the GAPDH step, resulting in the
tion or inhibition, and in this way, cell can increase or decrease the accumulation of glycolytic intermediates before this step and a
rate of glycolysis. A total of six ligand binding sites are found in decrease of glycolytic intermediates after the step [18]. Cells con-
PFK1: the substrate-binding sites for ATP and F6P, activator-bind- tain a limited supply of NAD+. Constant high-rate glycolysis will
ing sites for AMP and fructose-2,6-bisphosphate (F2,6BP), and use up the NAD+ pool. Glycolysis cannot sustain unless NAD+ is
inhibitor-binding sites for ATP and citrate. ATP is both a substrate regenerated. LDH-mediated reduction of pyruvate into lactate (fer-
and an allosteric inhibitor. Binding of ATP to the allosteric modula- mentation) is coupled to the oxidation of NADH into NAD+, thus
tion site, which is separated from the substrate binding site, facil- efciently replenish NAD+. Therefore, cancer cells may limit glu-
itates the formation of a conformation that lowers the afnity for cose oxidation in order to maximally increase lactate conversion
substrate F6P, thereby inhibiting the enzyme activity. Binding of and NAD+ recycling, allowing the continuation of high-rate
citrate enhances the inhibitory effect of ATP. During the evolution glycolysis.
of metazoans, eukaryotic PFK1 enzymes develop increased sensi- However, lactate fermentation alone is not sufcient to restore
tivity to citrate [14], suggesting the importance of citrate in down- all NAD+ consumed in glycolysis. Since a signicant portion of glu-
regulating PFK1 and hence the glycolytic ux. Complete oxidation cose carbon ows into anabolic pathways (e.g. serine synthesis),
of glucose through the TCA cycle produces high levels of ATP. Trun- the more diversion of glycolytic ux into biosynthesis, the less
cated TCA cycles result in intermediate metabolites, notably cit- NAD+ can be recovered by lactate fermentation. Lactate fermenta-
rate. When the TCA cycle is highly active, ATP and citrate tion cannot supply enough NAD+ to maintain glycolysis self-
accumulate, indicating an adequate supply of energy. Through neg- sufciency.
ative feedback, these metabolites allosterically suppress PFK1 Cellular NAD+ levels are maintained at stable levels via equilib-
activity and glycolysis. Therefore, active glucose oxidation is gener- rium between NAD+ consumption and NAD+ synthesis. There exist
ally not compatible with a high rate of glycolysis. multiple resources to produce NAD+. NAD+ can be synthesized
However, accumulating evidence suggests that cancer cells may from nicotinamide through the salvage pathway. Although there
have acquired ability to evade this feedback regulation. PFK1 is are alternate biosynthetic routes for NAD+ biosynthesis, virtually
often hyperactivated in cancer cells and becomes resistant to the all cells depend on nicotinamide to maintain adequate intracellular
feedback inhibition. PFK1 enzymes from cancer cells are less sensi- NAD+ levels. Biochemical and physiological studies have estab-
tive to allosteric inhibitors, and are more potently activated by lished nicotinamide as the most important and widely available
F2,6BP than those from normal cells [15,16]. F2,6BP is the most NAD+ precursor in mammals [19]. Nicotinamide phosphoribosyl-
potent allosteric activator of PFK1. The binding of F2,6BP to PFK1 transferase (NAMPT) is the rate-limiting enzyme catalyzing the
increases the afnity of the enzyme for F6P and overrides ATP- rst step in the salvage synthesis of NAD+ from nicotinamide. It
mediated inhibition. F2,6BP levels thus critically regulate the plays a central role in the regulation of NAD+ homeostasis. Inhibi-
glycolytic rate. tion of NAMPT reduces the total intracellular NAD+ levels, resulting
J. Lu et al. / Cancer Letters 356 (2015) 156164 159

in blockade of glycolysis at the GAPDH step [18]. Conversely, extracellular matrix that elicits anti-apoptotic and pro-survival sig-
increased dosage of NAMPT increases cellular NAD+ levels. NAMPT nals. Matrix detachment in anoikis-competent cells activates both
is indeed overexpressed in a number of cancers [20,21], suggesting extrinsic and intrinsic apoptotic pathways. Cell detachment leads
elevated salvage synthesis of NAD+ is the primary source for NAD+ to upregulation and activation of several BH3-only pro-apoptotic
generation in cancer cells to support glycolysis. proteins (e.g., BMF, BIM, and BID) [30]. These proteins then activate
Collectively, these ndings suggest cancer cells may be able to the pro-apoptotic members of the BCL-2 family of proteins (BAK
maintain high-rate glycolysis irrespective to the fate of pyruvate. and BAX), resulting in mitochondrial outer membrane permeabili-
Theoretically, cancer cells may simultaneously keep both glycoly- zation (MOMP) that facilitates the release of IMS apoptogenic fac-
sis and glucose oxidation at high levels. This conicts with the pre- tors, and consequently downstream caspase activation and
valent Warburg effect. apoptosis [31].
Furthermore, matrix detachment markedly stimulates the gen-
eration of reactive oxygen species (ROS) specically in the mito-
Control of glucose oxidation by PDH and PDKs
chondria of detached cells, such as endothelial cells and
mammary epithelial cells [32,33]. The mitochondrial origin of
At the center of glucose oxidative metabolism lies the multisub-
increased ROS was based on measurement by the mitochondrial
unit PDH complex. PDH complex commits pyruvate into the TCA
indicator MitoSOX. ROS have dual functions. Low levels of ROS
cycle by catalyzing the rate-limiting oxidative decarboxylation of
can activate various signaling pathways to stimulate cell prolifera-
pyruvate into acetyl-CoA. It interconnects glycolysis and the TCA
tion and survival, whereas excess ROS irreversibly damage cellular
cycle, thus representing a key regulatory step in glucose metabo-
macromolecular components (proteins, lipids, nucleic acids) and
lism. The activity of PDH is tightly regulated by a variety of alloste-
cause cell death (including apoptosis). ROS may activate intrinsic
ric effectors and by reversible phosphorylation. PDH complex E1a
mitochondrial apoptotic pathway in part through lipid peroxida-
subunit can be phosphorylated by PDH kinases (PDKs) and dephos-
tion. Cytochrome c is normally retained to the mitochondrial inner
phorylated by PDH phosphatases (PDPs) [22]. Phosphorylation of
membrane due to its association with cardiolipin, a mitochondria-
PDH inhibits the enzyme and is the primary determinant of its
specic anionic phospholipid. ROS peroxidate cardiolipin, resulting
activity. By phosphorylation and inactivation of PDH, PDKs are a
in the liberation of cytochrome c [34], which may translocate
major gatekeeper of pyruvate entry into the TCA cycle.
through damaged mitochondrial outer membrane (as a conse-
There are four different PDKs (PDK1-4) in human that are differ-
quence of MOMP) into cytosol, and activate the caspase cascade
entially expressed in tissues [22]. Different PDKs respond to vari-
and apoptosis. ROS play an essential role in anoikis induction.
ous hormonal and nutritional stimuli, including hypoxia and
Treatment of detached cells with antioxidants suppresses anoikis
nutrient levels, to critically control glucose metabolism [23,24].
[32,35,36].
Among the 4 PDKs, PDK4 is a principal isoenzyme responsible for
ROS are constantly generated as byproducts of normal mito-
the modulation of PDH activity and fuel selection between glucose
chondrial oxidative metabolism under physiological conditions
and fatty acid for energy utilization [25]. PDK4 is highly expressed
[37]. Mitochondrial respiration is indeed the major source of intra-
in metabolically active tissues including heart, skeletal muscle,
cellular ROS. The electron transport chain (ETC) respiratory com-
liver, pancreas, and kidney. Selective fuel utilization depending
plexes produce superoxide anion (the rst member in a plethora
on the fed-fast cycle is a crucial metabolic control in all mammals.
of ROS) when single electrons are transferred to O2 during electron
In the fed state, increased plasma glucose stimulates PDH activity,
transport. Approximately 15% of the total oxygen consumed dur-
glucose oxidation, and fatty acid synthesis. In the starved state,
ing normal respiration is converted to superoxide radicals. To
PDH activity is downregulated to limit the use of glucose. Instead,
counter the deleterious effect of ROS, cells express anti-oxidants
free fatty acids released from adipose tissue are used for oxidative
including anti-oxidant enzymes to detoxify ROS and prevent them
ATP generation. Perturbation of the selection of glucose or fatty
from accumulating at high concentrations [38]. The mitochondria-
acids as energy source is a key part of diabetes and metabolic syn-
located manganese superoxide dismutase (MnSOD, or SOD2) ef-
drome. PDK4 expression is selectively induced in most tissues and
ciently converts superoxide to the less reactive hydrogen peroxide
organs in response to metabolic stress conditions [25]. Starvation,
that breaks down further into water and dioxygen by other enzy-
fasting, and glucose deprivation markedly upregulate PDK4 levels.
matic and non-enzymatic antioxidants. The balance between the
The abundance of PDK4 is also stimulated by a variety of dietary
production and elimination of ROS leads to redox homeostasis.
lipids and hormones, and is negatively regulated by insulin. Tran-
The metabolic root of ROS raised the possibility that metabo-
scription of PDK4 is activated by transcription factors FOXOs, ERRs,
lism is implicated in anoikis regulation. In fact, cell detachment
PPARs, and coactivator PGC1, all of which are well-established key
causes dramatic global metabolic changes characterized by
global metabolic regulators [26]. Increased PDK4 expression
reduced uptake of nutrients (including glucose) and consequently
reduces PDH activity, thereby conserving glucose and preventing
decreased glycolysis and the TCA cycle [35,39,36]. In particular,
hypoglycemia. The importance of PDK4 in glucose homeostasis
when detached from the matrix, untransformed mammary epithe-
has indeed been validated in PDK4 knockout mice [27]. The PDK4
lial cells robustly upregulate PDK4, thereby inhibiting PDH and dis-
mutant mice exhibit elevated PDH activity and increased glucose
proportionally decreasing the ux of glucose carbon into the TCA
utilization. Therefore, PDK4 is a pivotal regulator in nutrient sens-
cycle [39,36]. Depletion of PDK4 or forced activation of PDH
ing and control of PDH activity.
increases mitochondrial respiration in suspended cells, further
exacerbating induction of ROS and anoikis [36]. Conversely, over-
Metabolic control of anoikis in normal cells expression of PDKs antagonizes anoikis by prolonging survival of
cells in suspension [36]. Moreover, cell detachment also strongly
Metabolism is intrinsically linked to cell death, as mitochondria activates expression of MnSOD, the principal mitochondrial antiox-
play a central role in both energy metabolism and apoptosis [28]. idant enzyme, to detoxify mitochondrial ROS resulting from
Mitochondrial intermembranous space (IMS) contains key pro- detachment [33]. Cells depleted of MnSOD are hypersensitive to
apoptotic factors, such as cytochrome c, which trigger apoptosis matrix detachment. Taken together, although normal cells gener-
if released into the cytosol. A specic type of apoptosis, termed ate ROS and undergo anoikis when detached from matrix, detached
anoikis, is induced by loss of cell-matrix interaction [29]. Survival cells manage to keep ROS under control by (i) shutting down glu-
of normal cells relies on integrin-mediated attachment to the cose oxidation to evade excess ROS production and (ii) increasing
160 J. Lu et al. / Cancer Letters 356 (2015) 156164

Cancerous tissues were reported to produce increased amounts


of ROS [44]. However, cancer cells also develop heightened antiox-
idant systems to cope with the high oxidative stress [45]. While
MnSOD is induced following matrix detachment in normal cells
[33], many antioxidants, including MnSOD, are constitutively over-
expressed in cancer. Elevated expression of MnSOD is associated
with adverse clinical outcomes [46,47,33] (Fig. 3). Consistent with
these ndings, increased glucose consumption by cancer cells
diverts more glucose carbon into the oxidative branch of PPP, a
principal mechanism to generate NADPH [45]. NADPH is an essen-
tial cofactor for replenishing reduced glutathione (GSH)cells
most critical antioxidant. Cancer cells can further enhance this
antioxidant generation pathway by inhibiting PKM when oxidative
stress rises [48]. Enhanced antioxidant capacity allows cancer cells
to better survive detachment-induced oxidative stress and metas-
tasize. Consistently, in a lung cancer mouse model, treatment with
antioxidants reduced oxidative stress and accelerated lung cancer
progression (as evidenced by development of more tumors and
Fig. 2. Normal cells reprogram metabolism and redox regulation to counter at more advanced stages) [49]. In our hands, normal cells (MCF10A
oxidative stress induced by matrix detachment. Although they undergo anoikis, human mammary epithelial cells) apparently accumulate more
detached normal cells shut down glucose oxidation and increase antioxidant
ROS than highly glycolytic and metastatic cancer cells (MDA-MB-
capacity by upregulating of PDK4 and MnSOD, respectively, to contain ROS and
delay anoikis. 231 human breast cancer cells) following matrix detachment
(not shown). This might be attributed partly to higher antioxidants
in cancer cells.
antioxidant capacity to alleviate oxidative stress (Fig. 2). Such met- Taken together, the Warburg effect brings multiple benets to
abolic reprogramming renders cells survive longer in the absence cancer cells: increased glucose consumption provides anabolic
of anchorage and mitigate anoikis. Without these metabolic support for proliferation and enhances antioxidant capacity,
endeavors, detached cells would die much faster. These ndings whereas the shift of pyruvate away from mitochondrial oxidation
highlight the metabolic control of anoikis and also reinforce the allows cells to avoid excess ROS generation. Therefore, cancer cells
vital role of ROS in anoikis. are better equipped to tolerate oxidative stress generated by
matrix detachment, and thus acquire anoikis resistance and meta-
The Warburg effect contributes to anoikis resistance and static potential. In essence, the Warburg effect promotes tumor
metastasis growth and importantly, metastasis.

The development of metastasis is a complex process including


Metastasis regulators impact oxidative metabolism
detachment of tumor cells from the primary site, local invasion
and migration, intravasation, survival in the circulation, extravasa-
As discussed in many excellent reviews [1,3,50], many well-
tion, and colonization of the secondary sites. During the process,
established oncogenes and tumor suppressors, such as hypoxia-
cells are displaced from their natural matrix niche or completely
inducible factor (HIF), Akt, Myc, and p53, exert direct impact on
deprived of matrix support (during circulation). Therefore, resis-
metabolism, most notably on glucose uptake and glycolysis. How-
tance to anoikis is a prerequisite for tumor metastasis.
ever, increased glycolysis does not evenly increase downstream
In contrast to normal cells that are sensitive to matrix detach-
metabolic pathways (Fig. 1). In particular, pyruvate is preferen-
ment, cancer cells have attained increased resistance to anoikis.
tially diverted to lactate fermentation, with very little to the mito-
This is in part because apoptotic pathway is often compromised
chondrial TCA cycle. The regulatory mechanisms for oxidative
in cancer cells. In addition, given the relationship between metab-
metabolism of glucose are distinct from those for glycolysis. For
olism and anoikis, one may wonder whether cells metabolic pro-
instance, Akt stimulates glucose uptake and glycolysis without
le may impact its anoikis sensitivity and metastatic potential.
affecting the rate of oxidative phosphorylation [51]. Because oxida-
While normal cells attenuate glucose oxidation upon detachment,
tive metabolism antagonizes metastasis, here we discuss several
many cancer cells already limit glucose oxidation prior to detach-
metastasis regulators, namely HIF, Snail, p53, and KISS1, for their
ment thanks to the Warburg effect, and may hence inherently pos-
effects on mitochondrial metabolism.
sess a survival advantage when detached. Normal cells activate
PDK4 to inhibit PDH activity in response to detachment to repro-
gram glucose metabolism, cancer cells express high levels of PDKs HIF promotes metastasis and suppresses oxidative metabolism
(such as PDK1) even under attached conditions [36]. Depletion of
PDK or activation of PDH in cancer cells stimulates glucose oxida- Development of intratumoral hypoxia is a nearly universal fea-
tion and ROS production, and restores their susceptibility to anoi- ture of rapidly growing solid tumors. Hypoxia occurs when cancer
kis, leading to decreased metastatic potential [36]. Therefore, the cells outgrow the blood vessels or as a result of unstable tumor
Warburg effect allows cancer cells to evade excess ROS that would vasculature. Hypoxia is a negative prognostic factor because it
be generated by glucose oxidation. It has previously been reported enhances invasiveness, stemness, metabolic shift, angiogenesis,
that glycolysis indeed diminishes cellular oxidative stress [41]. and metastatic potential of tumor cells [52]. Cells pretreated with
Thus, reducing ROS levels and protecting against ROS-mediated hypoxia gain increased metastatic ability than their normoxic
anoikis, which promotes metastasis, may represent an advantage counterparts [53]. The adaptive changes are primarily mediated
conferred by a Warburg metabolic phenotype. Consistently, by the transcription factor HIF, a master regulator to balance oxy-
expression of PDK1 and PDK3 in human cancer positively gen supply and demand [54]. HIF is stabilized and activated under
correlates with tumor histological grades and negatively with hypoxic environment, leading to activation of its target genes. HIF
disease-free survival [42,43] (Fig. 3). is also activated by loss of tumor suppressor (e.g., VHL), activation
J. Lu et al. / Cancer Letters 356 (2015) 156164 161

Fig. 3. Expression of PDKs and MnSOD in breast cancer correlates with tumor grades and unfavorable clinical outcome. (A) Correlation between expression of PDK1, PDK3,
and MnSOD, and tumor histological grades in breast cancer based on the dataset GSE3494 (n = 251). (B) Higher levels of PDK1, PDK3, and MnSOD predict worse relapse-free
survival (RFS) in a large cohort of breast cancer patients (n = 3455). Analysis was performed using the webtool developed by [40].

of oncogenic signaling, and increased ROS levels [52,55]. Therefore, There are multiple ROS production sites in the ETC, which consists
many tumors may have constitutively activated the HIF-dependent of four complexes. The ubiquinone reduction site of Complex I and
program even under normoxia. Higher levels of HIF are associated the outer quinone-binding site of the Q cycle in Complex III possess
with higher cancer mortality [56]. the greatest maximum capacity of ROS production [65]. Hypoxia
HIF is a prominent determinant of the metabolic shift from glu- upregulates microRNA-210 (miR-210), which targets iron-sulfur
cose oxidation to aerobic glycolysis in cancer. Besides stimulating cluster assembly proteins (ISCU1/2) [66]. These proteins facilitate
glucose uptake and glycolysis, HIF transcriptionally activates the assembly of iron-sulfur clusters (including those in Complex I
PDK1 and PDK3 [23,24,57]. These PDKs phosphorylate and inacti- and Complex III). Iron-sulfur clusters are critical for electron trans-
vate PDH, prevent pyruvate from fueling the mitochondrial TCA port and mitochondrial redox reactions. Consequently, miR-210
cycle, and hence reduce mitochondrial oxygen consumption and disrupts the integrity of iron-sulfur clusters and represses mito-
prevent the excessive production of ROS. chondrial respiration [66]. Furthermore, Complex I (NADH:ubiqui-
HIF also enhances LDH-mediated pyruvate-to-lactate conver- none oxidoreductase) catalyzes the rst step in the ETC. NADH
sion. LDH is a tetrameric enzyme consisting of LDH-H subunit dehydrogenase (ubiquinone) 1 subcomplex subunit 4-like 2 (NDU-
(encoded by ubiquitously expressed LDHB gene) and/or LDH-M FA4L2), considered as an inhibitory component of Complex I, was
subunit (encoded by LDHA). Generally, LDH-catalyzed reaction is identied as a direct transcriptional target of HIF [67]. HIF induces
reversible. However, LDH enzyme comprising pure LDH-M sub- expression of NDUFA4L2, which in turn inhibits ETC Complex I
units preferentially converts pyruvate into lactate. LDHA is a direct activity, thereby attenuating oxygen consumption and mitochon-
transcriptional target of HIF and highly inducible by hypoxia. drial ROS production [67].
Therefore, HIF promotes the formation of LDH enzyme made of Together, as a master commander, HIF stimulates glycolysis,
only LDH-M and more efciently turns pyruvate into lactate. This switches glucose metabolism from oxidative phosphorylation to
may indirectly decrease the pyruvate ux into mitochondria by lactate fermentation, and attenuates overall mitochondrial respira-
competition. tion. These metabolic effects protect cells against oxidative stress,
HIF cooperates with other transcriptional regulators, such as and may impart cancer cells with a selective advantage during
Myc. Myc is an oncogenic transcription factor involved in the reg- metastasis.
ulation of cell metabolism and cell proliferation, and is commonly
deregulated in cancer [58]. Acting alone (under normoxia) or in EMT inducer Snail inhibits mitochondrial respiration
conjunction with HIF, Myc activates PDK1 and LDHA [59]. Estro-
gen-related receptors (ERRs), normally controlling the fuel selec- Epithelial cells are able to acquire mesenchymal properties
tion between glucose and fatty acids by suppressing glucose through epithelial-to-mesenchymal transition (EMT). EMT is a
oxidation via upregulation of PDK4, are associated with tumor pro- characteristic of embryonic development, but also contributes to
gression [60,61]. Our previous studies suggest that ERRs enhance tumor progression by bestowing cancer cells with increased meta-
HIF and Myc activity and contribute to the Warburg effect static potential, cancer stem cell (CSC)-like traits, and therapeutic
[62,63]. Collectively, HIF-dependent transcriptional program resistance [68,69]. A dening feature of EMT is resistance to anoi-
decreases PDH activity (via upregulation of PDKs) and increases kis, which contributes to metastasis [70]. The mechanistic coupling
LDH-mediated pyruvate-to-lactate conversion, thereby efciently between EMT and resistance to anoikis remains poorly understood
diverting pyruvate away from mitochondrial oxidation. [70]. The transcriptional repressor Snail is a central driver of EMT
PDKs only control the ow of glucose carbons, and do not and its expression correlates with metastasis and poor clinical
directly regulate metabolites derived from fatty acid oxidation prognosis [71,72].
and glutaminolysis, which may also fuel the TCA cycle. Recently, As part of ETC Complex IV, cytochrome c oxidase (COX) is the
it was shown that HIF shifts glutamine metabolism from oxidation last enzyme in the respiratory chain that transfers the electrons
to reductive carboxylation as well [64]. In fact, HIF is able to to oxygen during mitochondrial respiration. COX is an oligomeric
broadly attenuate mitochondrial respiration and ROS generation. protein complex that is composed of 13 different subunits encoded
162 J. Lu et al. / Cancer Letters 356 (2015) 156164

by 3 mitochondrial genes and 10 nuclear genes. Three COX sub- KISS1 promotes oxidative metabolism
units, COX6c, COX7a, and COX7c, were identied as direct targets
of Snail [73]. Snail binds to their promoters and inhibits their KISS1 is a member of metastasis suppressor genes, which are
expression. Increased Snail expression inhibits COX expression dened by their ability to block metastasis without affecting pri-
and activity, and consequently represses oxygen consumption mary tumor development [82]. KISS1 encodes for Kisspeptin,
and mitochondrial respiration [73]. which is a peptide ligand for G protein-coupled receptor KISS1R.
Snail also regulates glucose metabolism. Fructose-1,6-biphos- KISS1 exhibits anti-metastatic effects in numerous human cancers,
phatase 1 (FBP1) is a rate-limiting enzyme in the pathway of gluco- including melanoma, thyroid, ovarian, bladder, gastric, esophageal,
neogenesis. FBP1 catalyzes the hydrolysis of F1,6BP to F6P, which is pancreatic, lung, and pituitary cancers [83,84]. Clinically, reduced
opposite to the rate-limiting glycolytic reaction catalyzed by PFK1. KISS1 expression is associated with poor prognosis in cancer
Therefore, FBP1 functions to antagonize glycolysis. In cancer cells, patients. It is proposed that KISS1 suppresses metastasis through
expression of FBP1 is often downregulated due to promoter hyper- inhibition of cancer cell migration and invasion [84]. A recent
methylation [74]. Restoration of FBP1 expression reduces glucose study links KISS1-dependent metastatic suppression to metabo-
uptake, glycolysis, and lactate generation, meanwhile increases lism [85]. Expression of KISS1 shifts metabolism from aerobic gly-
mitochondrial oxygen consumption as well as ROS production colysis to oxidative phosphorylation, apparently via pathways that
[74,75]. These metabolic effects suppress anchorage-independent boost mitochondrial respiration and/or biogenesis through stabil-
growth and tumorigenecity [74]. FBP1 was recently discovered as ization of PGC1a. PGC1a is a transcriptional co-activator for most
a direct target of Snail [75]. Snail binds to the FBP1 promoter and genes in the TCA cycle and ETC, and thus is a master regulator of
represses FBP1 expression, thereby enhancing glycolysis. Together mitochondrial respiration and oxidative metabolism [86]. Impor-
with its ability to suppress mitochondrial respiration [73], Snail tantly, PGC1a seems to be essential for KISS1-mediated metabolic
facilitates the metabolic shift towards aerobic glycolysis. changes and suppression of metastasis [85], suggesting the pro-
Snail not only induces EMT but also suppresses mitochondrial oxidative metabolic effect of KISS1 is important for its metasta-
oxidative metabolism. As EMT confers resistance to anoikis, it is sis-suppressing activity.
conceivable that Snail-mediated metabolic changes may contrib-
ute to anoikis resistance and metastasis.
Metabolic modulation for anti-metastasis therapy
p53 suppresses the Warburg effect
The Warburg effect not only fuels tumor growth but also facil-
As one of the most frequently mutated genes in cancer, muta- itates metastatic progression. Current studies position the War-
tions in the p53 tumor suppressor occur in the majority of human burg effect as a central player in malignancy. Metastasis is the
cancers. p53 suppresses tumorigenesis primarily by orchestrating primary cause of cancer mortality. Discovery of its root in altered
cell cycle arrest and apoptosis in response to cellular stress [76]. metabolism may expose its vulnerability, making metabolic mod-
Accumulating evidence suggests that p53 functions at multiple ulation a viable therapeutic anti-metastasis approach. An impor-
stages of cancer, including metastasis. Tumor-derived mutant tant step towards metabolic therapy is to identify and
p53 can drive tumor survival, invasion, and metastasis [77]. p53 manipulate critical biochemical nodes that are deregulated in can-
has also been discovered to be an important regulator of metabolic cer metabolism. The Warburg effect is a metabolic hallmark of can-
pathways. p53 has been shown to control glycolysis, oxidative cer cells and is generally reversible, therefore may represent an
phosphorylation, glutaminolysis, fatty acid oxidation, and redox attractive target for therapy. As part of the Warburg phenotype,
[50,78]. p53-decient cells exhibit higher rates of glycolysis, glucose oxidation is restrained in cancer cells. Forced activation
increased lactate production, and decreased mitochondrial respira- of oxidative metabolism may curtail cancer cells anoikis resistance
tion compared with wild-type cells [79], suggesting p53 sup- and metastatic potential.
presses the Warburg effect. PDH is a key determinant of pyruvate entry into the mitochon-
Mechanistically, p53 regulates genes involved in the balance dria. PDH activity is diminished in cancer cells partly due to
between the utilization of respiratory and glycolytic pathways. increased PDK expression. The small molecule Dichloroacetate
Basal levels of p53 can promote oxidative metabolism through (DCA) is a pyruvate mimetic and can compete with pyruvate for
transcriptional activation of SCO2 [79]. SCO2 is critical for regulat- binding to PDKs [87]. DCA inhibits PDKs and was initially used in
ing the assembly of the COX complex in the ETC, the major site of clinical treatment for lactic acidosis [88]. Inhibition of PDK by
oxygen consumption in the mitochondria. Cells with mutant p53 DCA enhances pyruvate entry into the TCA cycle and shifts metab-
have compromised oxidative phosphorylation. Disruption of the olism from glycolysis to glucose oxidation. This metabolic normal-
SCO2 gene in cells with wild-type p53 recapitulates the aerobic ization increases mitochondrial respiration and ROS generation,
glycolysis phenotype that is exhibited by p53-decient cells. In and induces apoptosis in certain cancer cells even under attached
addition, p53 downregulates PDK2 to promote the entry of pyru- conditions [89,90]. In our study of breast cancer cells, ectopic acti-
vate into mitochondria for oxidative metabolism [80]. Moreover, vation of PDH increases oxidative stress, restores anoikis sensitiv-
p53 activates transcription of glutaminase 2 (GLS2) to promote ity, and reduces metastasis, but has little effect on the viability of
glutaminolysis to fuel the TCA cycle and also facilitates fatty acid attached cells [36]. The discrepancy may originate from the differ-
oxidation (FAO) as an alternative energy source [78]. ence in the antioxidant capacity of different cancer cells. In the
In coordination with activation of oxidative metabolism, p53 mouse brain tumor model, DCA is able to cross brain blood barrier
restricts glycolytic ux through a number of mechanisms [78], and induce regression of metastatic tumor mass [90]. DCA has been
which include transcriptional activation of TIGAR. TIGAR acts as a on several clinical trials, including for glioma and refractory meta-
fructose-2,6-bisphosphatase that lowers F2,6BP levels and thereby static breast cancer (ClinicalTrials.gov Identier: NCT01111097,
reduces the activity of PFK1. Hence, p53, via TIGAR, can decrease NCT01386632, and NCT01029925) [91,92].
the rate of glycolysis [81]. Collectively, p53 counteracts the War- Several other regulators of glycolysis are also amenable to anti-
burg effect by favoring oxidative phosphorylation and minimizing cancer therapy. Modulation of such metabolic targets may indi-
the glycolytic phenotype. p53 suppresses tumor development and rectly enhance oxidative metabolism as well [91,93]. Inhibition
progression perhaps in part through its function at the level of of LDH re-directs pyruvate into mitochondrial oxidation [94]. Gly-
metabolism. colysis inhibition by 2-deoxy-D-glucose (2-DG) increases oxygen
J. Lu et al. / Cancer Letters 356 (2015) 156164 163

utilization and decreases lactate production, shifting metabolism bisphosphatase (iPFK-2; PFKFB3) in human cancers, Cancer Res. 62 (20) (2002)
58815887.
toward OXPHOS and reverting the metastatic phenotype in vitro
[18] B. Tan, D.A. Young, Z.-H. Lu, T. Wang, T.I. Meier, R.L. Shepard, K. Roth, Y. Zhai, K.
and in vivo [95]. Overall, these approaches reverse the Warburg Huss, M.-S. Kuo, J. Gillig, S. Parthasarathy, T.P. Burkholder, M.C. Smith, S.
glycolytic phenotype, boost oxidative metabolism, heighten oxida- Geeganage, G. Zhao, Pharmacological inhibition of nicotinamide
tive stress, and may restore anoikis sensitivity and suppress metas- phosphoribosyltransferase (NAMPT), an enzyme essential for NAD+
biosynthesis, in human cancer cells: metabolic basis and potential clinical
tasis. Furthermore, because commonly used radiation and implications, J. Biol. Chem. 288 (5) (2013) 35003511.
chemotherapy depend on the induction of free radicals to kill can- [19] M. Gall, F. Van Gool, A. Rongvaux, F. Andris, O. Leo, The nicotinamide
cer cells, the pro-oxidative metabolic modulation may enhance phosphoribosyltransferase: a molecular link between metabolism,
inammation, and cancer, Cancer Res. 70 (2010) 811.
their therapeutic efcacy. However, the currently available meta- [20] T. Bi, X. Che, Nampt/PBEF/visfatin and cancer, Cancer Biol. Ther. 10 (2) (2010)
bolic modulators are very limited. The potency and specicity of 119125.
DCA are quite low. All these justify the quest for new reagents tar- [21] A. Garten, S. Petzold, A. Krner, S.-I. Imai, W. Kiess, Nampt: linking NAD
biology, metabolism and cancer, Trends Endocrinol. Metab. 20 (3) (2009) 130
geting aerobic glycolysis for anti-metastasis treatment. 138.
[22] M.S. Patel, L.G. Korotchkina, Regulation of the pyruvate dehydrogenase
complex, Biochem. Soc. Trans. 34 (Pt 2) (2006) 217222.
Conict of Interest [23] J. Kim, I. Tchernyshyov, G.L. Semenza, C.V. Dang, HIF-1-mediated expression of
pyruvate dehydrogenase kinase: a metabolic switch required for cellular
None. adaptation to hypoxia, Cell Metab. 3 (2006) 177185.
[24] I. Papandreou, R.A. Cairns, L. Fontana, A.L. Lim, N.C. Denko, HIF-1 mediates
adaptation to hypoxia by actively downregulating mitochondrial oxygen
Acknowledgement consumption, Cell Metab. 3 (2006) 187197.
[25] H.-S. Kwon, R.A. Harris, Mechanisms responsible for regulation of pyruvate
dehydrogenase kinase 4 gene expression, Adv. Enzyme Regul. 44 (2004) 109
This work was supported by NIH Grants R01CA137021 (to J.L.) 121.
and R01CA149646 (to M.T.). [26] J.Y. Jeong, N.H. Jeoung, K.-G. Park, I.-K. Lee, Transcriptional regulation of
pyruvate dehydrogenase kinase, Diabetes Metab. J. 36 (5) (2012) 328335.
[27] N.H. Jeoung, P. Wu, M.A. Joshi, J. Jaskiewicz, C.B. Bock, A.A. Depaoli-Roach, R.A.
References Harris, Role of pyruvate dehydrogenase kinase isoenzyme 4 (PDHK4) in
glucose homoeostasis during starvation, Biochem. J. 397 (3) (2006) 417425.
[28] J.L. Andersen, S. Kornbluth, The tangled circuitry of metabolism and apoptosis,
[1] G. Kroemer, J. Pouyssegur, Tumor cell metabolism: cancers Achilles heel,
Mol. Cell 49 (3) (2013) 399410.
Cancer Cell 13 (6) (2008) 472482.
[29] S.M. Frisch, E. Ruoslahti, Integrins and anoikis, Curr. Opin. Cell Biol. 9 (5)
[2] P.P. Hsu, D.M. Sabatini, Cancer cell metabolism: Warburg and beyond, Cell 134
(1997) 701706.
(5) (2008) 703707.
[30] A.P. Gilmore, Anoikis, Cell Death Differ. 12 (Suppl 2) (2005) 14731477.
[3] M.G. Vander Heiden, L.C. Cantley, C.B. Thompson, Understanding the Warburg
[31] F. Llambi, T. Moldoveanu, S.W.G. Tait, L. Bouchier-Hayes, J. Temirov, L.L.
effect: the metabolic requirements of cell proliferation, Science 324 (5930)
McCormick, C.P. Dillon, D.R. Green, A unied model of mammalian BCL-2
(2009) 10291033.
protein family interactions at the mitochondria, Mol. Cell 44 (4) (2011) 517
[4] W.H. Koppenol, P.L. Bounds, C.V. Dang, Otto Warburgs contributions to current
531.
concepts of cancer metabolism, Nat. Rev. Cancer 11 (5) (2011) 325337.
[32] A.E. Li, H. Ito, I.I. Rovira, K.S. Kim, K. Takeda, Z.Y. Yu, V.J. Ferrans, T. Finkel, A role
[5] C. Jose, N. Bellance, R. Rossignol, Choosing between glycolysis and oxidative
for reactive oxygen species in endothelial cell anoikis, Circ. Res. 85 (4) (1999)
phosphorylation: a tumors dilemma?, Biochim Biophys. Acta 1807 (6) (2011)
304310.
552561.
[33] S. Kamarajugadda, Q. Cai, H. Chen, S. Nayak, J. Zhu, M. He, Y. Jin, Y. Zhang, L. Ai,
[6] I. Marin-Valencia, C. Yang, T. Mashimo, S. Cho, H. Baek, X.-L. Yang, K.N.
S.S. Martin, M. Tan, J. Lu, Manganese superoxide dismutase promotes anoikis
Rajagopalan, M. Maddie, V. Vemireddy, Z. Zhao, L. Cai, L. Good, B.P. Tu, K.J.
resistance and tumor metastasis, Cell Death Dis. 4 (2013) e504.
Hatanpaa, B.E. Mickey, J.M. Mats, J.M. Pascual, E.A. Maher, C.R. Malloy, R.J.
[34] S. Orrenius, V. Gogvadze, B. Zhivotovsky, Mitochondrial oxidative stress:
Deberardinis, R.M. Bachoo, Analysis of tumor metabolism reveals
implications for cell death, Annu. Rev. Pharmacol. Toxicol. 47 (2007) 143
mitochondrial glucose oxidation in genetically diverse human glioblastomas
183.
in the mouse brain in vivo, Cell Metab. 15 (6) (2012) 827837.
[35] Z.T. Schafer, A.R. Grassian, L. Song, Z. Jiang, Z. Gerhart-Hines, H.Y. Irie, S. Gao, P.
[7] M. Guppy, P. Leedman, X. Zu, V. Russell, Contribution by different fuels and
Puigserver, J.S. Brugge, Antioxidant and oncogene rescue of metabolic defects
metabolic pathways to the total ATP turnover of proliferating MCF-7 breast
caused by loss of matrix attachment, Nature 461 (7260) (2009) 109113.
cancer cells, Biochem. J. 364 (Pt 1) (2002) 309315.
[36] S. Kamarajugadda, L. Stemboroski, Q. Cai, N.E. Simpson, S. Nayak, M. Tan, J. Lu,
[8] R.J. DeBerardinis, A. Mancuso, E. Daikhin, I. Nissim, M. Yudkoff, S. Wehrli, C.B.
Glucose oxidation modulates anoikis and tumor metastasis, Mol. Cell. Biol. 32
Thompson, Beyond aerobic glycolysis: transformed cells can engage in
(10) (2012) 18931907.
glutamine metabolism that exceeds the requirement for protein and
[37] D.F. Stowe, A.K.S. Camara, Mitochondrial reactive oxygen species production in
nucleotide synthesis, Proc. Natl. Acad. Sci. USA 104 (49) (2007) 1934519350.
excitable cells: modulators of mitochondrial and cell function, Antioxid. Redox
[9] D.R. Wise, R.J. DeBerardinis, A. Mancuso, N. Sayed, X.-Y. Zhang, H.K. Pfeiffer, I.
Signal. 11 (6) (2009) 13731414.
Nissim, E. Daikhin, M. Yudkoff, S.B. McMahon, C.B. Thompson, Myc regulates a
[38] D. Trachootham, J. Alexandre, P. Huang, Targeting cancer cells by ROS-
transcriptional program that stimulates mitochondrial glutaminolysis and
mediated mechanisms: a radical therapeutic approach?, Nat Rev. Drug Discov.
leads to glutamine addiction, Proc. Natl. Acad. Sci. USA 105 (48) (2008) 18782
8 (7) (2009) 579591.
18787.
[39] A.R. Grassian, C.M. Metallo, J.L. Coloff, G. Stephanopoulos, J.S. Brugge, Erk
[10] S.Y. Lunt, M.G. Vander Heiden, Aerobic glycolysis: meeting the metabolic
regulation of pyruvate dehydrogenase ux through PDK4 modulates cell
requirements of cell proliferation, Annu. Rev. Cell Dev. Biol. 27 (Jan. 2011)
proliferation, Genes Dev. 25 (16) (2011) 17161733.
441464.
[40] B. Gyrffy, A. Lanczky, A.C. Eklund, C. Denkert, J. Budczies, Q. Li, Z. Szallasi, An
[11] M.G. Vander Heiden, S.Y. Lunt, T.L. Dayton, B.P. Fiske, W.J. Israelsen, K.R.
online survival analysis tool to rapidly assess the effect of 22,277 genes on
Mattaini, N.I. Vokes, G. Stephanopoulos, L.C. Cantley, C.M. Metallo, J.W.
breast cancer prognosis using microarray data of 1,809 patients, Breast Cancer
Locasale, Metabolic pathway alterations that support cell proliferation, Cold
Res. Treat. 123 (3) (2010) 725731.
Spring Harb. Symp. Quant. Biol. 76 (2011) 325334.
[41] K.A. Brand, U. Hermsse, Aerobic glycolysis by proliferating cells: a protective
[12] S.L. McKnight, On getting there from here, Science 330 (6009) (2010) 1338
strategy against reactive oxygen species, FASEB J. 11 (5) (1997) 388395.
1339.
[42] S.M. Wigeld, S.C. Winter, A. Giatromanolaki, J. Taylor, M.L. Koukourakis, A.L.
[13] R.J. DeBerardinis, J.J. Lum, G. Hatzivassiliou, C.B. Thompson, The biology of
Harris, PDK-1 regulates lactate production in hypoxia and is associated with
cancer: metabolic reprogramming fuels cell growth and proliferation, Cell
poor prognosis in head and neck squamous cancer, Br. J. Cancer 98 (12) (2008)
Metab. 7 (1) (2008) 1120.
19751984.
[14] A. Usenik, M. Legia, Evolution of allosteric citrate binding sites on 6-
[43] C.-W. Lu, S.-C. Lin, C.-W. Chien, S.-C. Lin, C.-T. Lee, B.-W. Lin, J.-C. Lee, S.-J. Tsai,
phosphofructo-1-kinase, PLoS One 5 (11) (2010) e15447.
Overexpression of pyruvate dehydrogenase kinase 3 increases drug resistance
[15] S. Vora, J.P. Halper, D.M. Knowles, Alterations in the activity and isozymic
and early recurrence in colon cancer, Am. J. Pathol. 179 (3) (2011) 14051414.
prole of human phosphofructokinase during malignant transformation
[44] G. Haklar, E. Sayin-Ozveri, M. Yksel, A.O. Aktan, A.S. Yalin, Different kinds of
in vivo and in vitro: transformation- and progression-linked discriminants of
reactive oxygen and nitrogen species were detected in colon and breast
malignancy, Cancer Res. 45 (7) (1985) 29933001.
tumors, Cancer Lett. 165 (2) (2001) 219224.
[16] G.E. Staal, A. Kalff, E.C. Heesbeen, C.W. van Veelen, G. Rijksen, Subunit
[45] R.A. Cairns, I.S. Harris, T.W. Mak, Regulation of cancer cell metabolism, Nat.
composition, regulatory properties, and phosphorylation of
Rev. Cancer 11 (2) (2011) 8595.
phosphofructokinase from human gliomas, Cancer Res. 47 (19) (1987) 5047
[46] G. Pani, R. Colavitti, B. Bedogni, S. Fusco, D. Ferraro, S. Borrello, T. Galeotti,
5051.
Mitochondrial superoxide dismutase: a promising target for new anticancer
[17] T. Atsumi, J. Chesney, C. Metz, L. Leng, S. Donnelly, Z. Makita, R. Mitchell, R.
therapies, Curr. Med. Chem. 11 (10) (2004) 12991308.
Bucala, High expression of inducible 6-phosphofructo-2-kinase/fructose-2,6-
164 J. Lu et al. / Cancer Letters 356 (2015) 156164

[47] M. Landriscina, F. Maddalena, G. Laudiero, F. Esposito, Adaptation to oxidative [73] S.Y. Lee, H.M. Jeon, M.K. Ju, C.H. Kim, G. Yoon, S.I. Han, H.G. Park, H.S. Kang,
stress, chemoresistance, and cell survival, Antioxid. Redox Signal. 11 (11) Wnt/Snail signaling regulates cytochrome C oxidase and glucose metabolism,
(2009) 27012716. Cancer Res. 72 (2012) 36073617.
[48] D. Anastasiou, G. Poulogiannis, J.M. Asara, M.B. Boxer, J. Jiang, M. Shen, G. [74] X. Liu, X. Wang, J. Zhang, E.K.Y. Lam, V.Y. Shin, A.S.L. Cheng, J. Yu, F.K.L. Chan,
Bellinger, A.T. Sasaki, J.W. Locasale, D.S. Auld, C.J. Thomas, M.G. Vander Heiden, J.J.Y. Sung, H.C. Jin, Warburg effect revisited: an epigenetic link between
L.C. Cantley, Inhibition of pyruvate kinase M2 by reactive oxygen species glycolysis and gastric carcinogenesis, Oncogene 29 (2010) 442450.
contributes to cellular antioxidant responses, Science 334 (6060) (2011) [75] C. Dong, T. Yuan, Y. Wu, Y. Wang, T.W.M. Fan, S. Miriyala, Y. Lin, J. Yao, J. Shi, T.
12781283. Kang, P. Lorkiewicz, D. St Clair, M.-C. Hung, B.M. Evers, B.P. Zhou, Loss of FBP1
[49] V.I. Sayin, M.X. Ibrahim, E. Larsson, J.A. Nilsson, P. Lindahl, M.O. Bergo, by Snail-mediated repression provides metabolic advantages in basal-like
Antioxidants accelerate lung cancer progression in mice, Sci. Transl. Med. 6 breast cancer, Cancer Cell 23 (2013) 316331.
(221) (2014) 221ra15. [76] K.H. Vousden, C. Prives, Blinded by the light: the growing complexity of p53,
[50] A.J. Levine, A.M. Puzio-Kuter, The control of the metabolic switch in cancers by Cell 137 (3) (May 2009) 413431.
oncogenes and tumor suppressor genes, Science 330 (6009) (2010) 1340 [77] P.A.J. Muller, K.H. Vousden, P53 mutations in cancer, Nat. Cell Biol. 15 (1)
1344. (2013) 28.
[51] R.L. Elstrom, D.E. Bauer, M. Buzzai, R. Karnauskas, M.H. Harris, D.R. Plas, H. [78] C.R. Berkers, O.D.K. Maddocks, E.C. Cheung, I. Mor, K.H. Vousden,
Zhuang, R.M. Cinalli, A. Alavi, C.M. Rudin, C.B. Thompson, Akt stimulates Metabolic regulation by p53 family members, Cell Metab. 18 (5) (2013)
aerobic glycolysis in cancer cells, Cancer Res. 64 (11) (2004) 38923899. 617633.
[52] G.L. Semenza, HIF-1 mediates metabolic responses to intratumoral hypoxia [79] S. Matoba, J.-G. Kang, W.D. Patino, A. Wragg, M. Boehm, O. Gavrilova, P.J.
and oncogenic mutations, J. Clin. Invest. 123 (2013) 36643671. Hurley, F. Bunz, P.M. Hwang, P53 regulates mitochondrial respiration, Science
[53] E.K. Rofstad, T. Danielsen, Hypoxia-induced metastasis of human melanoma 312 (5780) (2006) 16501653.
cells: involvement of vascular endothelial growth factor-mediated [80] T. Contractor, C.R. Harris, P53 negatively regulates transcription of the
angiogenesis, Br. J. Cancer 80 (11) (1999) 16971707. pyruvate dehydrogenase kinase Pdk2, Cancer Res. 72 (2) (2012) 560
[54] W.G. Kaelin, P.J. Ratcliffe, Oxygen sensing by metazoans: the central role of the 567.
HIF hydroxylase pathway, Mol. Cell 30 (4) (2008) 393402. [81] K. Bensaad, A. Tsuruta, M.A. Selak, M.N.C. Vidal, K. Nakano, R. Bartrons, E.
[55] A.J. Majmundar, W.J. Wong, M.C. Simon, Hypoxia-inducible factors and the Gottlieb, K.H. Vousden, TIGAR, a p53-inducible regulator of glycolysis and
response to hypoxic stress, Mol. Cell 40 (2010) 294309. apoptosis, Cell 126 (1) (2006) 107120.
[56] G.L. Semenza, Dening the role of hypoxia-inducible factor 1 in cancer biology [82] S.C. Smith, D. Theodorescu, Learning therapeutic lessons from metastasis
and therapeutics, Oncogene 29 (5) (2010) 625634. suppressor proteins, Nat. Rev. Cancer 9 (2009) 253264.
[57] C.-W. Lu, S.-C. Lin, K.-F. Chen, Y.-Y. Lai, S.-J. Tsai, Induction of pyruvate [83] B.H. Beck, D.R. Welch, The KISS1 metastasis suppressor: a good night kiss for
dehydrogenase kinase-3 by hypoxia-inducible factor-1 promotes metabolic disseminated cancer cells, Eur. J. Cancer 46 (2010) 12831289.
switch and drug resistance, J. Biol. Chem. 283 (2008) 2810628114. [84] D. Cvetkovic, A.V. Babwah, M. Bhattacharya, Kisspeptin/KISS1R system in
[58] C.V. Dang, MYC on the path to cancer, Cell 149 (2012) 2235. breast cancer, J. Cancer 4 (8) (2013) 653661.
[59] C.V. Dang, The interplay between MYC and HIF in the Warburg effect, Ernst [85] W. Liu, B.H. Beck, K.S. Vaidya, K.T. Nash, K.P. Feeley, S.W. Ballinger, K.M.
Schering Found. Symp. Proc., 2007, pp. 3553. Pounds, W.L. Denning, A.R. Diers, A. Landar, A. Dhar, T. Iwakuma, D.R. Welch,
[60] G. Deblois, V. Gigure, Oestrogen-related receptors in breast cancer: control of Metastasis suppressor KISS1 appears to reverse the Warburg effect by
cellular metabolism and beyond, Nat. Rev. Cancer 13 (1) (2013) 2736. enhancing mitochondrial biogenesis, Cancer Res. (2013).
[61] C. Chang, D.P. McDonnell, Molecular pathways: the metabolic regulator [86] C. Handschin, B.M. Spiegelman, Peroxisome proliferator-activated receptor
estrogen-related receptor a as a therapeutic target in cancer, Clin. Cancer gamma coactivator 1 coactivators, energy homeostasis, and metabolism,
Res. 18 (22) (2012) 60896095. Endocr. Rev. 27 (7) (2006) 728735.
[62] A. Ao, H. Wang, S. Kamarajugadda, J. Lu, Involvement of estrogen-related [87] T.R. Knoechel, A.D. Tucker, C.M. Robinson, C. Phillips, W. Taylor, P.J. Bungay,
receptors in transcriptional response to hypoxia and growth of solid tumors, S.A. Kasten, T.E. Roche, D.G. Brown, Regulatory roles of the N-terminal domain
Proc. Natl. Acad. Sci. USA 105 (2008) 78217826. based on crystal structures of human pyruvate dehydrogenase kinase 2
[63] Q. Cai, T. Lin, S. Kamarajugadda, J. Lu, Regulation of glycolysis and the Warburg containing physiological and synthetic ligands, Biochemistry 45 (2) (2006)
effect by estrogen-related receptors, Oncogene 32 (16) (2013) 20792086. 402415.
[64] R.C. Sun, N.C. Denko, Hypoxic regulation of glutamine metabolism through [88] P.W. Stacpoole, A.C. Lorenz, R.G. Thomas, E.M. Harman, Dichloroacetate in the
HIF1 and SIAH2 supports lipid synthesis that is necessary for tumor growth, treatment of lactic acidosis, Ann. Int. Med. 108 (1) (1988) 5863.
Cell Metab. 19 (2) (2014) 285292. [89] S. Bonnet, S.L. Archer, J. Allalunis-Turner, A. Haromy, C. Beaulieu, R. Thompson,
[65] M.P. Murphy, How mitochondria produce reactive oxygen species, Biochem. J. C.T. Lee, G.D. Lopaschuk, L. Puttagunta, S. Bonnet, G. Harry, K. Hashimoto, C.J.
417 (2009) 113. Porter, M.A. Andrade, B. Thebaud, E.D. Michelakis, A mitochondria-K+ channel
[66] S.Y. Chan, Y.-Y. Zhang, C. Hemann, C.E. Mahoney, J.L. Zweier, J. Loscalzo, axis is suppressed in cancer and its normalization promotes apoptosis and
MicroRNA-210 controls mitochondrial metabolism during hypoxia by inhibits cancer growth, Cancer Cell 11 (1) (2007) 3751.
repressing the iron-sulfur cluster assembly proteins ISCU1/2, Cell Metab. 10 [90] E.D. Michelakis, G. Sutendra, P. Dromparis, L. Webster, A. Haromy, E. Niven, C.
(2009) 273284. Maguire, T.-L. Gammer, J.R. Mackey, D. Fulton, B. Abdulkarim, M.S. McMurtry,
[67] D. Tello, E. Balsa, B. Acosta-Iborra, E. Fuertes-Yebra, A. Elorza, . Ordez, M. K.C. Petruk, Metabolic modulation of glioblastoma with dichloroacetate, Sci.
Corral-Escariz, I. Soro, E. Lpez-Bernardo, E. Perales-Clemente, A. Martnez- Transl. Med. 2 (31) (2010) 31ra34.
Ruiz, J.A. Enrquez, J. Aragons, S. Cadenas, M.O. Landzuri, Induction of the [91] P.E. Porporato, S. Dhup, R.K. Dadhich, T. Copetti, P. Sonveaux, Anticancer
mitochondrial NDUFA4L2 Protein by HIF-1a decreases oxygen consumption targets in the glycolytic metabolism of tumors: a comprehensive review,
by inhibiting Complex I activity, Cell Metab. 14 (2011) 768779. Front. Pharmacol. 2 (2011) 49.
[68] K. Polyak, R.A. Weinberg, Transitions between epithelial and mesenchymal [92] I. Papandreou, T. Goliasova, N.C. Denko, Anticancer drugs that target
states: acquisition of malignant and stem cell traits, Nat. Rev. Cancer 9 (4) metabolism: Is dichloroacetate the new paradigm?, Int J. Cancer 128 (5)
(2009) 265273. (2011) 10011008.
[69] J.P. Thiery, H. Acloque, R.Y.J. Huang, M.A. Nieto, Epithelial-mesenchymal [93] E.B. Butler, Y. Zhao, C. Muoz-Pinedo, J. Lu, M. Tan, Stalling the engine of
transitions in development and disease, Cell 139 (2009) 871890. resistance: targeting cancer metabolism to overcome therapeutic resistance,
[70] S.M. Frisch, M. Schaller, B. Cieply, Mechanisms that link the oncogenic Cancer Res. 73 (2013) 27092717.
epithelial-mesenchymal transition to suppression of anoikis, J. Cell Sci. 126 [94] A. Le, C.R. Cooper, A.M. Gouw, R. Dinavahi, A. Maitra, L.M. Deck, R.E. Royer, D.L.
(2013) 2129. Vander Jagt, G.L. Semenza, C.V. Dang, Inhibition of lactate dehydrogenase A
[71] H. Peinado, D. Olmeda, A. Cano, Snail, Zeb and bHLH factors in tumour induces oxidative stress and inhibits tumor progression, Proc. Natl. Acad. Sci.
progression: an alliance against the epithelial phenotype?, Nat Rev. Cancer 7 USA 107 (2010) 20372042.
(6) (2007) 415428. [95] J.L. Sottnik, J.C. Lori, B.J. Rose, D.H. Thamm, Glycolysis inhibition by 2-deoxy-D-
[72] B. De Craene, G. Berx, Regulatory networks dening EMT during cancer glucose reverts the metastatic phenotype in vitro and in vivo, Clin. Exp.
initiation and progression, Nat. Rev. Cancer 13 (2) (2013) 97110. Metastasis 28 (8) (2011) 865875.

S-ar putea să vă placă și