Download as pdf or txt
Download as pdf or txt
You are on page 1of 59

Journal Pre-proof

Targeting glucose metabolism to suppress cancer progression:


prospective of anti-glycolytic cancer therapy

Ali F. Abdel-Wahab, Waheed Mahmoud, Randa M. Al-Harizy

PII: S1043-6618(19)31673-1
DOI: https://doi.org/10.1016/j.phrs.2019.104511
Reference: YPHRS 104511

To appear in: Pharmacological Research

Received Date: 11 August 2019


Revised Date: 19 October 2019
Accepted Date: 23 October 2019

Please cite this article as: Abdel-Wahab AF, Mahmoud W, Al-Harizy RM, Targeting glucose
metabolism to suppress cancer progression: prospective of anti-glycolytic cancer therapy,
Pharmacological Research (2019), doi: https://doi.org/10.1016/j.phrs.2019.104511

This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.

© 2019 Published by Elsevier.


Targeting glucose metabolism to suppress cancer progression:
prospective of anti-glycolytic cancer therapy

Ali F. Abdel-Wahaba,b,⁎, Waheed Mahmoudc, Randa M. Al-Harizyd,e

a
Departments of Clinical Pharmacology, Faculty of Medicine, Cairo University, Egypt
b
Departments of Pharmacology and Toxicology, Faculty of Medicine, Umm Al-Qura University, Saudi Arabia
c
Departments of Medical Biochemistry, Faculty of Medicine, Cairo University, Egypt

of
d
Departments of Internal Medicine, Faculty of Medicine, Cairo University, Egypt
e
Departments of Internal Medicine, Ibn Sina National College for Medical Sciences, Saudi Arabia

ro
-p
Corresponding author:

Ali F. Abdel-Wahab, MD
re
Department of Pharmacology, Faculty of Medicine, Cairo University, Cairo, Egypt; and
Department of Pharmacology, Faculty of Medicine, Umm Al-Qura University, Makkah, Saudi
Arabia
lP

E-mail: alifatehi@hotmail.com
afabdulwahab@uqu.edu.sa

Tel: 00966548085598
na

00201024035540
0020225329227
Fax: 0096625352834
ur
Jo

Graphical Abstract:

1
of
Highlights:

ro
 Cancer is the second major cause of mortality worldwide despite current chemotherapy
 Targeting aerobic glycolysis has become a research focus for developing anticancer drugs

-p
 These antiglycolytic agents might also sensitize tumor cells to other cytotoxic therapies
 Although preclinical studies are promising, this antiglycolytic therapy has not yet been
re
successfully translated into clinical practice
 This review presents the current state of this emerging glycolytic inhibitors discussing the
lP

challenges and the opportunities for this antiglycolytic cancer therapy


na

Abstract
Most solid tumor cells adapt to their heterogeneous microenvironment by depending largely on
aerobic glycolysis for energy production, a phenomenon called the Warburg effect, which is a
ur

hallmark of cancer. The altered energy metabolism not only provides cancer cell with ATP for
cellular energy, but also generate essential metabolic intermediates that play a pivotal role in the
Jo

biosynthesis of macromolecules, to support cell proliferation, invasiveness, and chemoresistance.


The cellular metabolic reprogramming in cancer is regulated by several oncogenic proteins and
tumor suppressors such as hypoxia-inducible factor (HIF-1), Myc, p53, and PI3K/Akt/mTOR
pathway. A better understanding of the mechanisms involved in the regulation of aerobic
glycolysis can help in developing glycolytic inhibitors as anticancer agents. These metabolic
antiglycolytic agents could be more effective if used in drug combinations to combat cancer.

2
Several preclinical and early clinical studies have shown the effectiveness of targeting the
glycolytic pathway as a therapeutic approach to suppress cancer progression. This review aimed
to present the most recent data on the emerging drug candidate targeting enzymes and
intermediates involved in glucose metabolism to provide therapeutic opportunities and challenges
for antiglycolytic cancer therapy.

Chemical compounds:

of
Chemical compounds studied in this article;
3-Bromopyruvate (PubChem CID: 70684); Dichloroacetate (PubChem CID: 517326); Fasentin
(PubChem CID: 879520); Koningic acid (PubChem CID: 124361); Lonidamine (PubChem CID:

ro
39562); Metformin (PubChem CID: 4091); Omeprazole (PubChem CID:4549); Oxamate
(PubChem CID: 974); Phloretin (PubChem CID: 4788); Rapamycin (PubChem CID: 5284616).

Abbreviations:
-p
re
3-BrPA, 3-bromopyruvate; 2-DG, 2-deoxy-D-glucose; 3PO, 3-(3-pyridinyl)-1-(4-pyridinyl)-2-
propen-1-one; ALDH, aldehyde dehydrogenase; AMP, adenosine monophosphate; AMPK, AMP-
lP

activated protein kinase; ATP, adenosine triphosphate; BET, bromodomain and extra-terminal
motif; β-CD, β-cyclodextrin; ß-CD–3BrPA, micro-encapsulated formulation of 3BrPA; CAFs,
cancer associated fibroblasts; DCA, dichloroacetate; DNA, deoxyribonucleic acid; EGF,
na

endothelial growth factor; EGFR, EGF receptor; EMT, epithelial-mesenchymal transition; ENO1,
enolase 1; ERRα: estrogen-related receptor alpha; ETC, electron transport chain; GBM,
18
glioblastoma multiforme; Glo1, glyoxalase1; FDG, fluoro-2-deoxy-D-glucose; F-2,6-BP,
ur

fructose-2,6-bisphosphate; GAPDH, glycerladehyde-3-phosphate dehydrogenase; G-6-P, glucose-


6-phosphate; GLUTs, glucose transporters; GPx, glutathione peroxidase; GSH, glutathione;
Jo

H2O2, hydrogen peroxide; HCC; hepatocellular carcinoma; HF, halofuginone; HK2, hexokinase
2; HIF-1 α, hypoxia-inducible factor-1 alpha; LDHA, lactate dehydrogenase A; LN, lonidamine;
MCT, monocarboxylate transporters; mtDNA, mitochondrial DNA; mTOR, mammalian target of
rapamycin; mTORC1, mTOR kinase1; NAD, nicotinamide adenine dinucleotide; NADP,
nicotinamide adenine dinucleotide phosphate; OxPhos, oxidative phosphorylation; OA, oleanolic
acid; PDAC, pancreatic ductal adenocarcinoma; PDH, pyruvate dehydrogenase, PDK1, pyruvate

3
dehydrogenase kinase1; PEL, primary effusion lymphoma; PEP, phosphoenolpyruvate; PET,
positron emission tomography; PFK1, phosphofructokinase1; PFKFB, 6-phosphofructo-2-
kinase/fructose-2,6-bisphosphatases; PFK15, 1-(4-pyridinyl)-3-(2-quinolinyl)-2-propen-1-one;
PGK, phosphoglycerate kinase; PGM, phosphoglycerate mutase; PHD, prolylhydroxylase;
PHGDH, phosphoglycerate dehydrogenase; PI3K, phosphoinositide 3-kinase; PK, pyruvate
kinase; PK-M2, pyruvate kinase M2; PPP, pentose phosphate pathway; RCC, renal cell carcinoma;
ROS, reactive oxygen species; SGLT1, sodium-glucose linked transporter 1, siRNA, small
interfering RNA; TCA, tricarboxylic acid; TKTL1, transketolase1; TNBC, triple negative breast

of
cancer, VEGF, vascular endothelial growth factor.

Keywords: Aerobic glycolysis; antiglycolytic agents; cancer therapy; Warburg effect

ro
1- Introduction

-p
Cancer incidence is rapidly growing with an expected increase of 70% over the next two decades.
Cancer is ranked as a leading cause of death, accounting for 13% of all mortality worldwide
re
(GLOBOCAN 2018 database) [1]. Metabolism of cancer cells is different from that of normal
cells, which allows them to sustain a high rate of proliferation and resist signals of apoptosis [2].
lP

Normally, cells utilize multiple metabolic pathways to produce energy depending on the
availability of metabolites and biosynthetic requirements for cellular function. Cells typically use
glycolysis to convert glucose into pyruvate in the cytosol. The pyruvate is further metabolized in
na

the mitochondria by oxidative phosphorylation (OxPhos) through the tricarboxylic acid (TCA)
cycle and electron transport chain (ETC), to produce the energy-storing adenosine triphosphate
(ATP). Under hypoxic conditions, cells can utilize anaerobic glycolysis, and converts pyruvate
ur

into lactate, producing much less amount of ATP, but at a faster rate [3]. Mitochondrial oxidation
of one glucose molecule yields 36 molecules of ATP, while its metabolism to lactate by glycolysis
Jo

produces only 2 ATP molecules. Under aerobic conditions, cells can also utilize fatty-acid
oxidation (called beta-oxidation) or glutamine oxidation, if these metabolites are available [3].
Tumor cells, unlike normal cells, depend largely on glycolysis for producing energy even in the
presence of adequate levels of oxygen, a process termed aerobic glycolysis [4]. Thus, inhibition of
glycolytic pathways has the potential to provide an effective approach to cancer research aiming
to develop new targeted anticancer agents. This approach has been proven effective in suppressing

4
tumor progression, and several of these glycolytic inhibitors are currently under investigation in
preclinical and clinical studies with promising results [5-7]. This review will present the most
recent data on the emerging candidate agents targeting glycolytic enzymes and intermediates to be
useful in cancer therapy.

2- Tumor glucose metabolism and Warburg effect


Cancer progression involves an inappropriate proliferation of cells, which have enhanced abilities
for energy production to resist metabolic stresses [8]. Tumor cell metabolism is reprogrammed in

of
favor of aerobic glycolysis despite the presence of plentiful oxygen [9]. This observation was first
reported many decades ago by the German scientist Otto Warburg and is thus referred to as the
“Warburg effect” [10]. This metabolic alteration to a high glycolysis rate has been observed in a

ro
variety of malignant tumors using positron emission tomography (PET) [11]. Warburg first
hypothesized that the high rate of glycolysis is a result of mitochondrial injury and this can convert

-p
differentiated cells into proliferating malignant cells [10]. However, this primary defect in
mitochondrial function is not supported by later studies in cancer [12], and many observations
re
have suggested that mitochondrial OxPhos is the main source of ATP in most cancer tissues [13].
The switch in the metabolism of some cancer cells appears to happen because of the altered
lP

conversion of phosphoenolpyruvate to pyruvate, which is catalyzed by the enzyme pyruvate kinase


M2 overexpressed in cancer cells [14]. The generation of pyruvate through this unique enzymatic
mechanism in cancer cells, is uncoupled with ATP production but is converted primarily into lactic
na

acid, rather than acetyl-CoA to enter the TCA cycle [7].


Cancer cells adapt to the low energy yield of glycolysis by increasing uptake of glucose to support
a higher glycolytic rate. This increased glucose uptake has been exploited clinically in diagnosis
ur

and follows up of cancer via the use of 18Fluoro-2-deoxy-D-glucose (FDG), a radiolabeled glucose
analog, in PET [11,15]. It is likely that, the high rate of glycolysis benefit cancer cells by providing
Jo

a high rate of ATP production, in addition to providing many intermediates, that are used in
subsidiary metabolic pathways for de novo synthesis of nucleotides, amino acids, lipids and
NADPH, that are required for rapid cell proliferation [16]. Besides, recent evidence indicated a
critical role for mitochondria in cancer cell metabolism [17]. Inhibition of mitochondrial OxPhos
has been shown to inhibit tumor invasion in hepatocellular and breast cancer [18], and to reduce
the multidrug resistance of melanoma cells [19]. Indeed, it has been observed that mitochondrial

5
OxPhos and glutaminolysis contributes to cancer progression and metastasis [20]. However, the
presence of aerobic glycolysis under the normoxic condition and functionally efficient
mitochondria is a very interesting fingerprint of cancer cells. Thus, the use of tumor glycolysis as
a potential target for cancer therapy is the most intriguing.

3- Dual metabolic phenotype and hybrid state


The glycolytic phenotype is known to be expressed by many cancers [21], but the dependence of
cancer cells on such phenotype remains unclear. It has been demonstrated cancer cell death can be

of
induced by reversing the glycolytic state to OxPhos [22]. The glycolytic cancer cells can exhibit a
non-glycolytic phenotype under acidic conditions by intracellular lactic acid accumulation. Lactic
acidosis is a common consequence of the Warburg effect in most solid tumors [23]. Tumor cells

ro
cultured with sufficient glucose are initially glycolytic and then lactate generation leads to
acidification of the medium. The lowered cellular pH decreases glycolytic flux and inhibits the

-p
activities of glycolytic enzymes, leading to the suppression of glycolysis. The limited supply of
glucose can lead to the rapid death of tumor cells, but under lactate acidosis, the cells switch from
re
Warburg effect to a non-glycolytic oxidative phenotype and metabolize glucose at a slower rate to
support cell survival [24]. During hypoxia or mitochondrial dysfunction, the cells can switch
lP

energy metabolism from mitochondrial respiration to glycolysis, thus maintaining cancer growth
[25]. It has been demonstrated that breast cancer cells can resist radiation injury by shifting from
glycolytic phenotype to OxPhos, thus producing more ATP to enhance survival [26].
na

In addition to the concept of a central role of the Warburg effect in cancer metabolism, some
studies suggested the concept of dual metabolic nature or hybrid state (Warburg effect and
oxidative phenotype) in cancer cells under the stressful microenvironment [27]. It is proposed that
ur

normal cells can exhibit either a glycolytic or oxidative state, but tumor cells can have a hybrid
metabolism, where both states coexist, enhanced by increased reactive oxygen species (ROS)
Jo

and/or activation of oncogenes such as RAS, MYC, and c-SRC [28,29]. The hybrid metabolic
state, with coexisting oxidative respiration and glycolysis, has been demonstrated in some
aggressive tumor cell lines, such as SiHa and HeLa, due to strong activation of hypoxia-inducible
factor-1 (HIF-1) from lactate accumulation [30]. The hybrid state is also verified by in vitro
experiments with aggressive triple-negative breast cancer (TNBC) cells [31]. These data support
the hypothesis of a hybrid (glycolytic and oxidative) model and its critical role in tumorigenesis.

6
The hybrid state model could explain the phenomenon of oxygen shock [32] when glycolytic cells
reach the blood vessel and exposed to a large amount of oxygen. These cells could switch to the
hybrid metabolic phenotype, which further induces metastasis [33]. Moreover, the hybrid
phenotype enhances the metabolic plasticity and enables malignant cells to switch their phenotypes
between glycolysis and OxPhos. This metabolic plasticity facilitates cancer cell adaptation to the
various tumor microenvironment, such as hypoxia and acidic conditions, enhancing invasion and
metastasis [34]. Therefore, targeting the hybrid state to eliminate metabolic plasticity could be a
new therapeutic strategy against cancer progression.

of
Recent evidence has demonstrated that tumor cells can acquire a hybrid metabolic state, thereby
using both glycolysis and OxPhos for the production of energy and formation of the required
macromolecules [35]. The hybrid glycolysis/OxPhos phenotype enhances the metabolic plasticity

ro
of tumor cells supporting cancer invasion, metastasis, and chemoresistance [20]. Also, it is noted
that the metabolic phenotype is heterogeneous in different tumors, even in the same type of tumor

-p
[36]. Detection of the metabolic plasticity of cancer could be helpful in targeting specific metabolic
pathways to prevent the hybrid state and suppress tumor growth. Moreover, metastasis of cancer
re
cells is facilitated by epithelial-mesenchymal transition (EMT), where epithelial tumor cells lose
adhesion and acquire the mesenchymal characteristics of migration and invasiveness [37]. Rapid
lP

growth and expansion of solid tumors usually outpace angiogenesis leading to a progressive
hypoxic condition, inducing HIF-1α, which induces both EMT and aerobic glycolysis [38]. This
hybrid glycolysis/OxPhos metabolic state indicates that dual blockade of both metabolic processes,
na

glycolysis and OxPhos, could be more effective in suppressing tumor and preventing metabolic
plasticity [39]. This combination treatment could drive tumor cells away from the hybrid
phenotype leading to more effective anticancer response [36].
ur

4- Reverse Warburg effect or metabolic coupling


Jo

Another type of tumor metabolism has been recently observed in certain types of cancers called
the “reverse Warburg effect” or “metabolic coupling” which have high mitochondrial respiration
and low glycolysis rate [40]. In this type, the tumor cells and adjacent stromal fibroblasts form a
two-compartment model of cancer metabolism, where aerobic glycolysis occurs in fibroblasts, and
the generated metabolites are transferred to malignant cells, to fuel the TCA cycle and maintain
ATP generation [41]. This metabolic coupling reveals a parasite-host relationship between tumor

7
cells and the cancer-associated fibroblasts, as observed in some forms of breast cancer [42]. The
two-compartment model of metabolism may contribute to certain forms of drug resistance and
therapeutic failure in some types of cancers [43]. It has been demonstrated that tamoxifen-sensitive
MCF7 cancer cells become tamoxifen-resistant when co-cultured with fibroblasts. The drug
combination of tamoxifen and the tyrosine kinase inhibitor dasatinib has been shown to decrease
this fibroblast-induced tamoxifen-resistance and reprogram tumor cells to the glycolytic
phenotype, leading to the death of MCF7 cancer cells [44]. However, although recent advances
have shed more light on the biological significance of tumor metabolism, the underlying

of
mechanistic details of regulation and consequences of such metabolic types remain unclear.

5- Warburg effect and tumor acidosis

ro
The microenvironment of solid tumors is markedly heterogenous, so tumor cells tend to increase
their uptake of glucose to maintain energy production [13]. Increased glycolysis and decreased

-p
mitochondrial oxidation lead to increased formation of lactic acid, increased glutaminolysis,
increased beta-oxidation of fatty acids and activation of the pentose phosphate pathway [45]. To
re
maintain the intracellular pH (pHi), tumor cells promptly export the excess intracellular acid load
to the extracellular compartment. The rapid turnover of cancer cells, glucose fermentation, and
lP

hypoxia, all can result in the generation and release of excessive amounts of protons into the
extracellular space [46]. In contrast to normal cells, cancer cells show a reversed pH gradient with
increased pHi and decreased extracellular pH (pHe) [47]. To survive in this hostile
na

microenvironment, cancer cells develop prompt ATP-requiring mechanisms to extrude protons.


These mechanisms include the vacuolar H+-ATPase, Na+/H+ exchanger, Na-bicarbonate
cotransport via carbonic anhydrase (CA) 9 or 12, and monocarboxylate transporters (MCTs) [46].
ur

The alkaline pHi and acidic pHe, enhance metabolic adaptation, proliferation, invasiveness and
metastatic behavior of tumor cells [48]. Moreover, many chemotherapeutic drugs are weak bases
Jo

that are trapped and neutralized in the acidic microenvironment of tumors contributing to
chemoresistance. It has been demonstrated that restoration of mitochondrial function could
suppress tumor growth and restore chemosensitivity [49].

6- Tumor glycolysis and its clinical relevance to cancer progression

8
The glycolysis process takes place in the cytoplasm by converting glucose into pyruvate through
nine reaction steps, involving several glycolytic enzymes. First, glucose is transported into tumor
cells at a high rate by glucose transporters (GLUTs), GLUT1 and sodium-glucose linked
transporter 1 (SGLT1) which are overexpressed in most cancers [50]. Glucose is phosphorylated
into glucose-6-phosphate by hexokinase (HK), a rate-limiting step that provides direct feedback
inhibition to preserve energy. The HK is bound to the mitochondrial membrane and has a high
affinity for glucose, facilitating initiation of glycolysis with low glucose levels [51]. There are four
HK isoforms (HK 1-4), HK1 is ubiquitously expressed, whereas HK2 is expressed in insulin-

of
sensitive adipose tissue and muscles [52] and is overexpressed in many cancer cells [53]. Glucose-
6-phosphate isomerase then converts glucose-6-phosphate into fructose-6-phosphate, which is
further phosphorylated to form fructose-1,6-biphosphate and fructose-2,6-biphosphate, under the

ro
effect of phosphofructokinase-1 (PFK1) and PFK2, respectively, and consuming one ATP
molecule. The PFK1 is a crucial driver of glycolysis and is inhibited by high ATP levels. The

-p
PFK2 is overexpressed in tumor cells, generating excess fructose-2,6-bisphosphate, which
activates PFK1, leading to maintenance of high glycolytic rate irrespective of ATP level [54].
re
Next, fructose-1,6-bisphosphate is transformed by aldolase enzyme into glyceraldehyde-3-
phosphate and dihydroxyacetone phosphate. Glyceraldehyde-3-phosphate is transformed by
lP

glyceraldehyde-3-phosphate dehydrogenase (GAPDH) into glycerate-1,3-diphosphate, which is


further transformed by phosphoglycerate kinase (PGK) into 3-phosphoglycerate and producing
two ATP molecules. The phosphoglycerate mutase (PGM) isomerizes 3-phosphoglycerate into 2-
na

phosphoglycerate, followed by the formation of phosphoenolpyruvate (PEP). Lastly, PEP is


converted into pyruvate by pyruvate kinase (PK), with the generation of one ATP molecule. The
PK catalyzes a rate-limiting step of glycolysis and its activity is affected by the cellular pH and
ur

ATP/AMP ratio [55]. The PK exists in four isoforms; PKL, PKR, PKM1, and PKM2 [56]. The
PKM2 is overexpressed in tumor cells and tumor-associated fibroblasts [57]. Also, PKM2 can be
Jo

translocated into the nucleus, acting as a protein kinase to regulate gene transcription [58]. Unlike
PKM2, PKM1 seems to have little effect on cell proliferation [59]. The ratio of PKM1/PKM2 has
been shown to decrease during malignant progression [60] and switching PKM2 to PKM1 can
have therapeutic implications. Replacing PKM2 with PKM1 has been demonstrated to inhibit
glycolysis in lung cancer cells and suppress tumor xenograft formation in nude mice [14,61]. The
pyruvate resulting from glycolysis can be converted into acetyl-CoA to enter the TCA cycle for

9
OxPhos, but during hypoxia, pyruvate is transformed into lactic acid by lactate dehydrogenase
(LDH). The LDH is overexpressed in tumor cells thus maintaining high glycolytic flux [62].
Although glycolysis yields less amount of ATP (18 times lower) compared to mitochondrial
oxidation, it can provide many benefits for cancer cells competing for shared energy sources [63].
The accelerated rate of glycolysis leads to faster and greater ATP production, which reaches 100
times faster than oxidative phosphorylation [64]. Besides ATP generation, glycolysis provides
tumor cells with metabolic intermediates and precursors that fuel pathways for biosynthesis of
macromolecules required for cell proliferation and tumor progression [65]. The accumulated

of
glycolytic intermediates enhance the pentose phosphate pathway (PPP) with the formation of
NADPH and ribose-5-phosphate, which are needed for the synthesis of phospholipids and nucleic
acids. Also, NADPH enables cancer cells to maintain supplies of the antioxidant glutathione

ro
(GSH), required for maintaining the intracellular redox status and protecting cancer cells against
the damaging effects of chemotherapeutic agents [66]. The glycolytic pathway also produces

-p
NADH and serine, which can be used in the formation of signaling molecules and important amino
acids such as glycine and cysteine [67] (figure 1).
re
Several lines of evidence have established a link between cancer progression and overexpression
of GLUTs and glycolytic enzymes including HK2, GAPDH, LDH, and PFK-2 [68,69]. A
lP

biochemical link has also been demonstrated between cancer glycolysis and resistance to
chemotherapy and radiotherapy [70,71]. Moreover, the transketolase1 (TKTL1) enzyme, involved
in PPP, plays an important role in cell survival under starvation and stresses [72]. The TKTL1 has
na

been shown to affect cancer cell sensitivity to drugs such as imatinib [73] and cetuximab [74].
Also, the pyruvate dehydrogenase kinase (PDK) isoforms, PDK1 and PDK3, have been found to
confer chemotherapeutic resistance in the cervical cancer cell line, HeLa [75]. Similarly, in colon
ur

carcinoma cell line, LoVo, drug resistance correlated with increased lactate production by aerobic
glycolysis [76]. This key signature of cancer metabolism, providing energy and supporting
Jo

uninterrupted growth, could be an attractive target for cancer therapy and sensitization to
chemotherapeutics [77].
A high lactate level, indicating the prevalence of glycolytic phenotype, correlated significantly
with tumor growth, spread, and recurrence [78]. Lactate was originally thought to be an acidifying
molecule that can cause a dangerous lowering of intracellular pH if not exported from tumor cells.
However, different roles of lactate efflux and influx have been proposed that contribute to cancer

10
cell survival through some type of “metabolic symbiosis” between normoxic and hypoxic tumor
cells co-existing in the heterogeneous microenvironment of most solid tumors [79]. The lactate is
produced and exported by hypoxic glycolytic tumor cells, to be imported and utilized by normoxic
tumor cells for energy production by mitochondrial OxPhos [80]. This process of lactate transfer
is achieved through the monocarboxylate transporters (MCTs), mainly MCT4 for lactate release
and MCT1 for lactate uptake, which is overexpressed in most tumors [81]. The MCT1 is abundant
in normoxic non-glycolytic cancer cells of vascularized area whereas MCT4 is expressed in
hypoxic glycolytic cancer cells. A significant correlation has recently been demonstrated, in breast

of
cancer cell lines (MCF-7 and MDA-MB-231), between the distribution of MCT isoforms (1 and
4) and the expression profile of LDH isoforms (A and B) [82]. In the MDA-MB-231cell line,
LDHA is abundant to convert pyruvate into lactate and MCT4 is overexpressed to release lactate.

ro
While in the MCF-7 cell line, MCT1 is overexpressed to uptake lactate and LDHB is abundant to
convert lactate back into pyruvate to fuel the TCA cycle [83]. Thus, cancer cells organize their

-p
glycolytic phenotype to achieve efficient energy supply.
re
7- Non-glycolytic functions of glycolytic enzymes and metabolic intermediates
Many glycolytic enzymes have also important roles in several non-glycolytic processes involved
lP

in cellular functions that support cancer cell survival and growth [84]. For instance, the
mitochondrial membrane-bound HK2 can antagonize the proapoptotic pathway in cancer cells
[85]. Also, HK2 acting as a nuclear enzyme is involved in transcriptional regulation of some
na

nuclear proteins [84]. Similarly, GAPDH has a critical role in maintaining the cellular redox
balance by catalyzing the production of NADH, and protection against free radical-induced injury
[86]. Also, GAPDH has some nuclear functions contributing to the pro-apoptosis and the
ur

oncogenic process by affecting nucleic acid binding properties of hepatitis viruses [87]. Several
non-glycolytic functions have been demonstrated for PKM2, including phosphorylation of histone
Jo

H3 to favor tumorigenesis [88], transactivation of β-catenin [89], and binding phosphotyrosine to


interact with other proteins [90].
Results revealed that GAPDH can bind directly to telomeric DNA preventing its degradation by
chemotherapy [91]. In prostate cancer, GAPDH has been shown to enhance the transcriptional
activity of androgen receptors [92]. The LDH has been found to cooperate with the transcriptional
factor Oct-4, and LDH gene silencing leads to the downregulation of Oct-4 and suppression of

11
gastric tumorigenesis. Also, the nuclear translocation of LDH can affect the functions of DNA
polymerases [93]. Moreover, the influx and efflux of lactate achieved through the MCTs, are
involved in the regulation of the CD147, a matrix metalloproteinase inducer, that increases cancer
cell invasion and metastasis [94]. Also, fructose-1,6 biphosphate has an anti-apoptotic effect in
tumor cells by reducing cytochrome C [95]. Besides, pyruvate has been involved in resistance to
chemotherapy by over-expression of p-glycoprotein [96]. Therefore, evidence indicates that many
glycolytic enzymes and intermediates participate in non-glycolytic processes at various subcellular
locations such as mitochondria, nucleus, and cytosol, to support cancer progression.

of
8- Molecular regulation of tumor glycolysis
It is increasingly evident that coordinated networks of signaling pathways regulate reprogramming

ro
of cancer cells to balance their metabolic state, supporting tumor growth and stress resistance.
Several studies have demonstrated the affection of cancer cell metabolism by many regulators

-p
including protooncogenes (e.g. Myc), transcription factors (e.g. HIF-1), signaling pathways (e.g.
PI3K/Akt/mTOR), and tumor suppressors (e.g. p53) [97]. The c-Myc is a transcription factor,
re
encoded by Myc oncogene, that control cellular growth, and metabolism, and its expression is
upregulated in many cancers, such as breast, colon, prostate, and bladder cancers [98].
lP

Experimental studies have demonstrated that overexpression of c-Myc in the liver of transgenic
mice can increase the activity of the glycolytic enzymes with the overproduction of lactic acid
[99]. Activation of the c-Myc has been found to upregulate many genes of the glycolytic enzyme,
na

such as GLUTs, HK2, PFK, enolase1 (ENO1), LDHA, and pyruvate dehydrogenase kinase1
(PDK1) [62,100]. Also, c-Myc is linked to increased mitochondrial ROS, leading to mitochondrial
dysfunction and switching cancer cells to glycolysis for energy production.
ur

The PI3K/Akt pathway contributes to several cellular processes including inflammation,


autophagy, and tumorigenesis. The Akt oncogene has been shown to stimulate the metabolism of
Jo

glucose and production of lactate without increasing oxygen consumption, in glioblastoma and
hematopoietic cancer cells [101]. Activation of Akt, the serine/threonine kinase, has been found
to mobilize GLUTs and activate HK2, thus enhancing glycolysis and promoting cancer growth
[102]. The mammalian Target of Rapamycin (mTOR) kinase exists in two forms, mTORC1 and
mTORC2, both are involved in the regulation of metabolism and cell proliferation. The mTORC1,
as a downstream effector of PI3K/Akt signaling, can enhance protein translation, lipogenesis, and

12
glycolysis [103]. Activation of mTORC1 induces GLUT1 and HK2, leading to increased glucose
uptake and a high rate of glycolysis [104]. The mTORC2 acts primarily through phosphorylating
Akt on serine 473, to enhance the expression of GLUT1, and activate HK2 and PFK-1, increasing
glycolysis rate [105]. Also, mTORC2 was shown to upregulate intracellular c-Myc and enhance
glycolysis in glioblastoma [106].
Hypoxia-inducible factor-1 (HIF-1) consists of oxygen-labile α subunit and constitutive β subunit,
and control gene transcriptions, regulating many processes such as angiogenesis, erythropoiesis,
inflammation and energy production [107]. The level of HIF-1α is sensitive to oxygen and is

of
regulated by HIF prolyl hydroxylase (PHD), which enhances its degradation in normoxic
conditions [108]. During hypoxia, the HIF-1α protein increases due to lower degradation by PHD
[109]. Thus, under hypoxic condition, cancer cell accumulates the HIF-1α, leading to upregulation

ro
of GLUTs and glycolytic enzymes, thereby increasing the rate of glycolysis. Studies in human
glioblastoma multiforme (GBM) cells demonstrated overexpression of HIF-1α, PDK1 and EGF

-p
receptor (EGFR) under hypoxic conditions [110]. Also, HIF-1 activates PDK1 to inhibit the PDH
activity and suppress the conversion of pyruvate to acetyl-CoA, thereby impairing mitochondrial
re
function [111]. Furthermore, activation of EGFR was found to enhance nuclear translocation of
PKM2 in cells of glioblastoma, breast and prostate cancers, leading to increased expression of
lP

cyclin D1 and glycolytic enzymes, thus promoting aerobic glycolysis and cellular proliferation
[88,89].
The tumor suppressor p53 can negatively regulate cell growth by inhibiting mTOR via inducing
na

the transcription of several genes. The p53 can inhibit glycolysis by decreasing GLUTs and
increasing fructose-2,6-bisphosphatase activity [112]. Also, p53 can enhance mitochondrial
oxidation through activation of the SCO2 gene of the respiratory chain [113]. Also, AMP-activated
ur

protein kinase (AMPK), the main sensor of cellular energy, is often associated with p53 mutation
in human cancers, leading to enhanced aerobic glycolysis [43].
Jo

Phloretin, fasentin,
Glucose
WZB117, ritonavir
GLUTs

2-DG, LN, 3-BrPA,


Glucose genistein-27, benserazide
HK
NADPH
ribose-5-phosphate Nucleotide
Glucose-6-P
13 synthesis
G-6-PI

Fructose-6-P
ATP PFK [PFKFB3] 3PO, PFK15

Fructose-1,6-bis-P

Aldolase

Dihydroxyacetone Lipid
Glyceraldehyde-3-P
phosphate synthesis

of
NADH GAPDH
3-BrPA, KA
1,3-bisphosphoglycerate

ro
2ATP PGK
Amino acid
3-phosphoglycerate Serine synthesis
PGM

Phosphoenolpyruvate -p
re
ATP PDK
DCA

TCA
lP

Pyruvate Acetyl-CoA cycle


LDH
Oxamate, FX11
Lactate
na

MCTs

Cinnamate, AZD3965
Lactate (H+)
ur

Figure 1: The glycolysis process with catalytic enzymes, metabolic targets (red) and glycolytic
inhibitors (green)
Jo

9- Targeting tumor metabolism and Glycolytic inhibitors


Recent cancer research focuses on selective inhibition of metabolic pathways to deprive cancer
cells of essential metabolic needs and interfere with tumor growth. The improved understanding
of aerobic glycolysis as a hallmark of cancer and underlying mechanisms may pave the way for
the development of targeted metabolic agents for antiglycolytic cancer therapy [114]. There are
several approaches to disrupt energy production and prevent glucose utilization by cancer cells.

14
Indeed, carbohydrate-restricted diets have been reported to have therapeutic benefits in cancer
patients [115]. Targeting glycolytic pathways to achieve cancer treatment seems appealing since
the involved enzymes are attractive molecular targets. However, to be useful targets, these
enzymes must have a significant difference in the activity or expression between cancer cells and
normally proliferating cells.
Currently, several agents inhibiting glycolysis are under intensive investigation in preclinical and
clinical studies exploiting the glycolytic activity of tumor cells [116]. These antiglycolytic agents
can target glucose transporters (e.g. phloretin, WZB117) or glycolytic enzymes such as HK2 (e.g.

of
2-deoxy-D-glucose, lonidamine), GAPDH (e.g. 3-bromopyruvate, koningic acid), LDH-A (e.g.
oxamate) or PDK (e.g. dichloroacetate) (Table 1). Although preliminary results of tumor growth
inhibition are promising, there are concerns of significant toxicity related to the wide expression

ro
of these target enzymes in normally proliferating cells [117]. This review will present the available
preclinical and clinical results with recent drug candidates to provide a future perspective of

-p
therapeutic opportunities for antiglycolytic cancer therapy.
re
9.1- Targeting glucose uptake (GLUTs inhibitors)
A direct antiglycolytic approach would be to block the uptake of glucose in malignant cells via
lP

GLUTs, leading to a total disruption of energy production pathways. Several small molecules can
selectively inhibit GLUTs including phloretin, fasentin, genistein, ritonavir, STF-31, WZB117,
and cytochalasin B. These agents have demonstrated anticancer effects in preclinical models by
na

inhibiting glucose uptake, thus leading to cell death through glucose and energy deprivation [118].
The polyphenol phloretin was recently shown to inhibit GLUT2 in TNBC leading to suppression
of tumor growth and metastasis [119]. Also, phloretin can inhibit GLUT1 that is overexpressed in
ur

the hypoxic area of resistant colon cancer cell lines and induce apoptosis by activating p53-
mediated signaling, leading to suppression of growth in resistant cancer cells [120]. Fasentin and
Jo

its analogs have shown to inhibit glucose uptake and decrease the resistance of caspase activation,
which is involved in chemoresistance of tumor cells [121]. Other studies have demonstrated that
STF-31 is a selective GLUT1 inhibiting agent that can kill the VHL-deficient renal cell carcinoma
(RCC), suppressing tumor growth with lower toxicity to normal cells [122]. The WZB117 is a bis-
hydroxybenzoate compound that produces fast selective irreversible blocking of glucose uptake
by GLUT1, leading to inhibition of tumor glycolysis and reduction of cellular ATP levels with the

15
arrest of cell-cycle [123]. This drug also produced a synergistic effect when combined with other
anticancer drugs, such as paclitaxel or cisplatin, against lung and breast cancer cell lines in vitro
and in vivo [118]. Moreover, the HIV-protease inhibitor, ritonavir was shown to inhibit GLUT4
leading to suppression of multiple myeloma cells, with synergistic effects with other drugs as
metformin [124]. However, since GLUTs are expressed ubiquitously in all cells, selective
inhibition of cancer glucose uptake is an important challenge. For instance, several glucose
transport inhibitors, tested in phase I clinical trials for hepatocellular and prostate cancer, are
associated with significant side effects [125].

of
9.2- Targeting glucose phosphorylation (HK inhibitors)
Another antiglycolytic approach is to target the HK enzymes responsible for the first, rate-limiting

ro
step of glycolysis. The HK2 is overexpressed in many cancer cells, and its inhibition produced
effective anticancer activity in preclinical studies [53]. Several HK inhibitors have been exploited

-p
for the anticancer effect such as 2-deoxy-D-glucose (2-DG), lonidamine, 3-Bromopyruvate (3-
BrPA), genistein-27 (GEN-27) and benserazide. Many natural products such as resveratrol,
re
astragalin, chrysin, have also been shown to inhibit HK2, suppressing growth and inducing
apoptosis in hepatocellular carcinoma cells [126-128]. The genistein derivative, GEN-27, is a
lP

synthetic flavonoid that was shown to suppress breast cancer cells via inhibition of HK2 and
induction of cell apoptosis [129]. Also, benserazide was found to specifically inhibit HK2 and
induce apoptosis in colorectal cancer cells, suppressing tumor growth in xenograft cancer model
na

[130].
9.2.1- 2-deoxy-D-glucose (2-DG)
The glucose analog 2-deoxy-D-glucose (2-DG) is one of the well-studied antiglycolytic agents,
ur

with promising anticancer effects in preclinical models [131]. The 2-DG enters the cell via GLUTs
competitively inhibiting glucose uptake. Inside the cell, 2-DG competes with glucose-6-phosphate
Jo

interfering with HK activity. 2-DG is phosphorylated to 2-DG-P which is an end substance that
cannot be metabolized and accumulate intracellularly, thereby blocking the glycolysis process
[132]. Subsequently, therapeutic doses of 2-DG caused ATP depletion and oxidative stress, and
eventually facilitated cell death [133]. Although 2-DG has been shown to inhibit the proliferation
of several cancer cell lines in vitro, animal studies revealed heterogeneous results with 2-DG [134-
136]. So, contrary to its expected anti-cancer effects, 2-DG may enhance some pro-survival

16
pathways in tumor cells [137]. Also, hypoxic tumor cells may resist the therapeutic effect of 2-DG
by overexpression of HIF-1 leading to the upregulation of GLUT and glycolytic enzymes [138].
These disappointing results challenge the use of 2-DG as a single agent for antiglycolytic cancer
therapy. Experimental studies demonstrated that 2-DG combined with the PI3K/mTOR inhibitor,
PF-04691502, can convert aerobic glycolysis into mitochondrial OxPhos in the cells of primary
effusion lymphoma (PEL). This drug combination was cytotoxic to PEL cells with less toxicity to
normal lymphocytes [139]. Also, it is suggested that 2-DG can be used with radiotherapy to
prevent the recovery of tumor cells from radiation damage, acting as a radiosensitizer [140].

of
Enhanced killing of malignant cells has also been observed in several preclinical studies using 2-
DG in combination with irradiation [141,142].
Despite the promising results of 2-DG in preclinical studies, results of clinical trials are

ro
inconsistent. Recently, 2-DG activity has been tested in phase II clinical trial in patients with
pancreatic cancer, but the trial was terminated due to the slow accrual [143]. The anticancer

-p
activity of 2-DG is tested in two other clinical trials, but there are real concerns about significant
inhibition of the glycolytic metabolism in the heart and brain [135,144]. Daily administrations of
re
2-DG at therapeutic doses were associated with hypoglycemia-like symptoms that can limit its use
as a single-agent therapy in vivo [145,146]. The safety of 2-DG is evaluated in phase I dose-
lP

escalation trial in prostatic cancer and identified 45 mg/kg as the recommended oral dose, before
stoppage due to insignificant effects on tumor growth [147]. Overall, the preliminary results from
clinical trials of 2-DG as a monotherapy are inconclusive and ambiguous.
na

Currently, 2-DG is reintroduced for use in combination approach, as reported in more recent
preclinical and clinical studies, using 2-DG at lower doses to produce synergistic anticancer effects
with other chemotherapeutic agents [136,148]. These synergistic effects of 2-DG have been
ur

observed in combination with cytotoxic agents in the in vitro studies and in phase I clinical trial
with docetaxel [149,150]. Also, a combination of 2-DG with cisplatin has been shown enhanced
Jo

cytotoxicity in head and neck cancer by increasing the ROS levels [136]. A recent phase I clinical
trials have shown that the use of 2-DG is well tolerated and has low toxicity [151,152]. An early
clinical trial examined the safety and effectiveness of orally administered 2-DG and showed
favorable results in patients with cerebral glioma [153]. Another clinical trial in glioblastoma using
2-DG 250 mg/kg, in combination with radiotherapy, showing a favorable toxicity profile, but
significant side effects occurred at higher doses in two of six patients [154]. Histological

17
examination of tumors in patients treated with the recommended dosage schedule of 2-DG
revealed extensive necrosis of tumor tissue with preserved surrounding tissue [155]. Similar good
tumor control, in addition to improved quality of life, was achieved in patients with brain tumors
using 2-DG combination regimens [156]. Moreover, combining 2-DG with metformin produced a
synergistic inhibitory effect on energy metabolism with depleting ATP levels in the cancer cell
lines [157]. This combination therapy was more effective in the resistant poorly glycolytic
pancreatic cancer cells, indicating that metformin sensitizes cancer cells to 2-DG [158]. Thus,
combination treatments using 2-DG may have encouraging outcomes providing a new opportunity

of
for cancer combination therapy.
9.2.2- Lonidamine (LN)
Lonidamine (LN) is an indole derivative, was used as an anti-spermatogenic agent, but now

ro
recognized with anti-tumor and pro-apoptotic activity [159]. The mechanism of action is not clear,
but studies indicated that LN acts as an inhibitor of HK-2 suppressing glycolysis and/or

-p
mitochondrial respiration. LN is an ANT ligand inducing mitochondrial channel formation and
causing inhibition of complex II and complex I [160]. This leads to decreased mitochondrial uptake
re
of pyruvate and decreased formation of fumarate and malate in treated cells, together with the
accumulation of succinate [161]. It is also suggested that LN can inhibit the MCTs responsible for
lP

lactate efflux, leading to decreased extracellular lactate [160].


LN is an emerging anticancer agent alone and in combination with other anticancer therapies. This
agent has completed pre-clinical studies and entered phase II clinical trials for cancer treatment
na

[162]. Inhibition of glycolytic enzyme HK2, with LN, has also been tested in several types of
cancers, such as lung, breast and ovarian cancer [163-166], but elevated toxicity recorded with no
significant survival benefit for patients. Also, early clinical testing of this drug showed no benefits
ur

for patients and the trial was terminated [167]. Results of combination therapy studies showed that
LN combined with other chemotherapeutics, such as doxorubicin, produced better anticancer
Jo

effects for the treatment of breast, prostate and ovarian tumors [168,169]. LN was found to enhance
the apoptotic response to many anticancer agents and γ irradiation both in vivo and in vitro [162].
However, the clinical success of LN has been impaired by significant pancreatic and hepatic
toxicities [170]. Therefore, current research focused on developing alternative dosage forms or
local targeted delivery of LN to reduce its toxicity.

18
9.3- Targeting fructose phosphorylation by PFK (PFKFB3 inhibitors)
Another glycolytic enzyme, PFK, that catalyzes the rate-limiting step of fructose phosphorylation,
is a potential target for anticancer agents. PFK is allosterically activated by AMP and fructose 2,6-
bisphosphate (F-2,6-BP). The PFK activity is liable to feed-back inhibition by excess ATP, but
this inhibition can be overcome by the elevated F-2,6-BP level to maintain uninterrupted glycolytic
flux in tumor cells [171]. The PFK is also controlled by a family of bi-functional enzymes
including PFKFBs, which is overexpressed in many cancers, providing an interesting target for
cancer therapy. The PFKFB3 is specifically inhibited by 3-(3-pyridinyl)-1-(4-pyridinyl)-2-propen-

of
1-one (3PO), and its derivative, 1-(4-pyridinyl)-3-(2-quinolinyl)-2-propen-1-one (PFK15).
Administration of 3PO was shown to produce a rapid reduction of glucose uptake, lactate
production and ATP generation in the cells of Jurkat T-cell leukemia [172]. Also, PFK15 has been

ro
reported to exhibit significant anti-tumor activity by reducing 18FDG uptake and F-2,6-BP level in
xenografted tumors. Moreover, PFK15 exhibits a pro-apoptotic effect in the transformed tumor

-p
cells in vivo and in vitro [173]. Thus, preliminary studies with PFKFB3 inhibitors revealed
potentially useful anticancer effects.
re
9.4- Targeting glucotrioses metabolism (GAPDH inhibitors)
lP

One of the most promising therapeutic approaches in antiglycolytic cancer therapy has been
targeting the enzyme GAPDH, which catalyzes the unique reaction for producing energy in the
form of NADH with the conversion of 3-phosphoglycerate into serine. The NADH molecule plays
na

a critical role in the regulation of intracellular ROS levels, and the biosynthesis of macromolecules.
Recent studies demonstrated that the growth of some subsets of melanoma and breast cancers are
dependent on the expression of high levels of PHGDH enzyme to enhance tumorigenesis and
ur

chemoresistance [174,175]. Therefore, inhibition of this multifunctional enzyme, GAPDH, would


result in multipronged effects within cancer cells, including reduced NADH and ATP production
Jo

with disturbed cellular redox balance, leading eventually to cell death [176]. Metabolism of the
glucotrioses, glyceraldehyde-3-phosphate, and dihydroxyacetone phosphate is inhibited, and their
partial degradation leads to the formation of methylglyoxal. This cytotoxic metabolite is usually
detoxified by the glyoxalase system (Glo 1 and 2). In presence of excess ROS and depletion of
GSH, the glyoxalase (Glo1) activity diminishes leading to the accumulation of the toxic

19
methylglyoxal [177]. Thus, inhibiting GAPDH represents an attractive therapeutic strategy, not
only affecting glycolysis but also exploiting other cytotoxic mechanisms in cancer cells.
A reversed pH gradient, manifested by extracellular acidosis and intracellular alkalization, with
metabolic reprogramming is a hallmark of cancer metabolism. By integrating the pH-dependent
enzyme activities, Persi et al., [178], developed a computational approach to explore how pHi can
modulate metabolic adaptation. They showed that, alkaline pHi enhances tumor cell proliferation
associated with increased glycolysis (more Warburgness). On the other hand, acidic pHi inhibits
tumor growth and prevents metabolic adaptation of cancer cells to be more oxidative (less

of
Warburgness). Their analysis explored GAPDH as a main metabolic modulator of cancer cells and
its inhibition at acidic pHi increases the anti-Warburg effect of low pHi on tumor growth [178].
Several GAPDH inhibitors, such as 3-BrPA, arsenate, iodoacetate, together with the natural

ro
compound Koningic acid, have been studied in cell cultures and animal models, with evident
anticancer efficacy [179, 180].
9.4.1- Koningic acid (KA)

-p
Koningic acid (KA), also called heptelidic acid, is a sesquiterpene metabolite isolated from a
re
fungus, that inhibits GAPDH by covalent binding to its active site [181]. The KA is cytotoxic to
highly glycolytic tumor cells due to ATP depletion associated with the progression of glucose
lP

phosphorylation. Thus, KA may be effective in disrupting metabolic pathways in aggressively


glycolytic tumors with little effect on healthy cells [182]. Furthermore, KA produced a marked
toxic effect on tumor cells with minimal systemic toxicity after its intraperitoneal administration
na

[183]. These data reinvigorate the hope to exploit this selective targeting of GAPDH by KA in the
treatment of cancer.
9.4.2- 3-Bromopyruvate (3-BrPA)
ur

The pyruvate analog, 3-Bromopyruvate (3-BrPA), is one of the most effective antiglycolytic drugs
currently under evaluation. This small molecule analog of pyruvate has demonstrated an
Jo

interesting ability to inhibit GAPDH and HK enzymes by alkylation of the active site.
Administration of 3-BrPA causes profound inhibition of tumor glycolysis and mitochondrial
OxPhos, leading to profound depletion of intracellular ATP, depriving cancer cells of energy
[184,185]. This alkylating agent can react by covalent binding with thiol groups in cysteine of
many targets involved in the glycolytic pathway including mt-HNK, SERCA-1, MCT1, GAPDG,
and others [186]. However, the predominant mechanism of action of 3-BrPA is selective inhibition

20
of GAPDH, which has been demonstrated in vitro in many cancer cell lines and, also in vivo in
animal studies [187,188]. Inhibition of GAPDH disrupts the metabolism of glucotriosis leading to
marked energy depletion, redox imbalance and eventually cancer cell death [189,190]. The
anticancer effects of 3-BrPA included suppression of tumor growth, invasion, metastasis,
angiogenesis, together with increased oxidative stress, and regulation of apoptotic pathways [191].
Moreover, 3-BrPA has been shown to enhance the cytotoxic effect and decrease resistance to other
anticancer drugs by inhibiting the ATP-dependent MDR transporter, providing a promising
candidate in combination therapy [192-194].

of
Interestingly, 3-BrPA is more stable in the acidic microenvironment of tumors with a longer half-
life, thus producing more marked cytotoxic effects in cancer cells without serious organ toxicity
[191]. Also, the cellular uptake of 3-BrPA is mediated through MCTs that is upregulated in most

ro
cancers, thereby providing a preferential uptake of 3-BrPA in tumor cells. The selective uptake of
this potentially toxic agent by cancer cells has recently been demonstrated [195]. However,

-p
concern remains about the rather non-specific alkylating properties of 3-BrPA that may be
associated with significant toxicity limiting the use of this potent anticancer agent. Thus, many
re
studies have focused on the local regional catheter-based delivery of 3-BrPA for the treatment of
solid tumors, such as hepatic VX2 tumor model in rabbits [195,196], murine orthotopic pancreatic
lP

and breast cancer models [197,198]. In the liver VX2 tumor model, histopathological assessment
of liver samples 4 days after treatment, revealed complete necrosis of tumor cells with no evident
damage to surrounding tissue or other organs [199]. Subsequent dose-escalation study in the VX2
na

tumor-bearing rabbits defined 1.75 mM 3-BrPA as a therapeutic concentration, to be injected intra-


arterially or given as continuous intra-arterial infusion for 1 h duration is equally effective. Higher
concentrations of 2.5 mM 3-BrPA caused peripheral liver necrosis [200]. Treatment with 3-BrPA
ur

was found to achieve survival benefits in treated cancer animal models [201-204].
The systemic delivery of 3-BrPA has been investigated in an orthotopic xenograft mouse model
Jo

of pancreatic cancer, after micro-encapsulation of 3-BrPA using β-cyclodextrin (β-CD), as a


molecular carrier, to achieve stability of 3-BrPA molecule for systemic administration and reduce
its nontarget toxicity. This micro-encapsulated formulation (ß-CD–3BrPA) showed similar
anticancer efficacy as that observed with the free drug in vitro [184]. A recent in vivo experiment
using daily intraperitoneal injections of free or micro-encapsulated 3-BrPA (up to 5 mg/kg) in
tumor-bearing mice showed similar efficacy with tumor eradication, confirmed by

21
histopathological analysis, under both formulations of systemic 3-BrPA therapy. However, the
micro-encapsulated formulation was associated with lower toxicity as compared with the free 3-
BrPA [205].
Early clinical trials showed promising results for efficacy and safety of 3-BrPA in the local
regional delivery, but drug toxicity is considered the main obstacle for its systemic delivery. In a
clinical trial, two patients diagnosed with end-stage liver cancer (HCC and cholangiocarcinoma,
respectively) were treated with intra-arterial 3-BrPA, showing favorable results with no significant
toxicity [5]. Also, 3-BrPA has been accepted by the US FDA as an investigational new drug

of
application for use in Phase I clinical trial in patients with liver cancer [206]. However, most
scheduled trials regarding 3-BrPA are suspended after the reports of three deaths, in 2016. All
these patients died within a few days after receiving 3-BrPA as alternative medicine by a

ro
nonmedical practitioner in Germany and the condition is under investigation by the prosecutor
about the possible improper use of 3-BrPA [207]. Currently, no approved clinical trials on 3-BrPA

-p
are available in the database website of approved clinical studies; https: //www.clinicaltrials.gov/.
There are only two reported case studies on using 3-BrPA for the treatment of patients with
re
resistant cancers. The first is reported by Ko et al., [208], on voluntary treating a boy with advanced
fibrolamellar hepatocellular carcinoma using transcatheter arterial administration of formulated 3-
lP

BrPA over several months. The boy survived with a higher quality of life for a much longer period
than expected without major toxicity. Unfortunately, the patient died after two years of diagnosis,
due to overload liver function [208]. The second is reported by El Sayed et al., [209], using an
na

intravenous infusion of 3-BrPA (1–2.2 mg/kg) to treat a man with stage IV metastatic melanoma
and response was monitored by serum LDH level. Repeated 6 infusion doses of 3-BrPA over 10
days led to a partial reduction of serum LDH level. The patient was also given paracetamol as a
ur

GSH scavenger, to avoid tumor resistance by the high GSH level. This is followed by a sharp drop
of LDH level, and no significant cytotoxicity observed except burning sensation at the infusion
Jo

site. However, the man died due to respiratory distress and hypoxemia [209]. Currently, it is
suggested that unformulated 3-BrPA should not be used and only formulated forms can be used in
clinical oncology studies [191].

9.5- Targeting pyruvate formation (PKM2 inhibitors)

22
Pyruvate is regarded as a "hub" metabolite, regulating metabolic reprogramming of cancer cells,
and linking glycolysis in the cytosol to mitochondrial oxidation. The enzyme PK catalyzes the
formation of pyruvate from phosphoenolpyruvate (PEP) with the generation of ATP. This essential
enzyme for aerobic glycolysis is liable to regulation by allosteric effectors as fructose-1,6-
bisphosphate and by phosphorylation [210]. The isoform PKM2 is overexpressed in tumors
enhancing cellular proliferation, and thus could be an attractive target for anticancer therapy
[61,211]. Interestingly, data are suggesting that either inhibition or activation of PKM2 could
diminish the growth of cancer cells. The PKM2 inhibitors could allow glycolytic intermediates to

of
accumulate and feed biosynthetic pathways, promoting tumor growth. For instance, Anastasiou et
al. [212], have demonstrated that inhibition of PKM2 by oxidation on cysteine 358, leads to
diversion of glycolytic intermediates into the PPP, with NADPH production and promotion of

ro
redox balance, supporting tumor growth. Thus, the anticancer effect can be achieved by the
expression of a non-oxidizable PKM2 mutant or by PKM2 activators that suppress the PPP and

-p
increase oxidative stress. On the contrary, Goldberg and Sharp [213], found that using small
interfering RNA (siRNA) to inhibit PKM2, could inhibit tumor growth and increase cell apoptosis
re
in vitro. Also in vivo delivery of siRNA to inhibit PKM2 in mouse xenograft model can induce
tumor regression.
lP

These contradictory results could be explained by variable responses to various degrees of


hypoxia. Under moderate hypoxia hydrogen peroxide (H2O2) formation is increased, activating
signaling pathways involved in the cellular response to hypoxia. Thus, PKM2 activity is
na

suppressed by its oxidation leading to increased flux through the PPP and enhanced redox balance,
preventing oxidative cellular damage [212]. On the other hand, in severe hypoxia mitochondrial
production of ATP and generation of H2O2 stops. Thus, cells dependent on PK activity for energy
ur

production and can be affected by PKM2 inhibitors. This implies that the anticancer effects of
PKM2 inhibitors or activators could be tumor-dependent according to tumor size and
Jo

vascularization. So, PKM2 activators can enhance oxidative damage in moderately hypoxic cells,
while PKM2 inhibitors can prevent energy production in severely hypoxic cells. Also, PKM2
inhibition can induce caspase-dependent cell death, increasing apoptosis [213]. Recent studies
revealed that PKM2 can increase the transcriptional activity of HIF-1, which mediates cellular
adaptation to hypoxic stress via transcription of diverse targets such as GLUTs, HK2 and LDH-A

23
[214]. Thus, it would be also important to clarify the effects of PKM2 inhibition or activation on
the non-glycolytic functions involved in tumorigenesis.
Several compounds such as TT-232, VK3, VK5 and Compound 3, can inhibit PKM2 leading to
suppression of glycolysis in tumor cells [211]. The natural compound oleanolic acid (OA), has
been recently found to exert its anticancer effect by shifting PKM2 to PKM1, thus inhibiting
aerobic glycolysis [215]. However, it has been recently observed that knockdown of PKM2 in the
breast cancer model could enhance tumor formation, indicating that PKM2 inhibition may not be
efficient alone in the treatment of cancer [216]. These results highlighted the importance of

of
understanding the tumor metabolic needs to target metabolic pathways for cancer therapy.

9.6- Targeting pyruvate oxidation (PDK inhibitor)

ro
Pyruvate in the cytosol is converted into mitochondrial acetyl-CoA to enter the Krebs cycle by
pyruvate dehydrogenase (PDH) enzymes. The PDH is negatively regulated by the enzyme

-p
pyruvate dehydrogenase kinase (PDK), leading to a shift of glucose from oxidative
phosphorylation to glycolytic metabolism [217]. Thus, targeting PDK may be an attractive
re
approach to inhibit cellular proliferation and cancer growth by inducing tumor cell reprogramming
from the glycolytic to oxidative phenotypes. This reversing of the Warburg effect has been
lP

demonstrated in glioblastoma multiforme (GBM) cells using the PDK inhibitor, dichloroacetate
(DCA), that lowered PDK1-EGFR activation and shifted glycolytic phenotype to oxidative
metabolism [110].
na

9.6.1- Dichloroacetate (DCA)


Dichloroacetate (DCA) is a small molecule recently introduced for antiglycolytic therapy. This
drug has been used for more than 25 years in the treatment of congenital mitochondrial
ur

dysfunctions [218]. The mechanism of DCA action involves its ability to activate mitochondrial
PDH by inhibiting its regulator PDK. Thus, DCA enhances the reprogramming of energy
Jo

production from the glycolytic state toward mitochondrial OxPhos with the consumption of
pyruvate and decreasing lactate accumulation [219]. So, cancer cells become more susceptible to
apoptotic signals and their proliferation is suppressed [220]. During hypoxia, HIF-1-alpha
activates gene transcription of some glycolytic regulators such as GLUTs, PDK and LDHA. This
leads to redirection of cells from oxidative to glycolytic state with accumulation of pyruvate [111].

24
The compelling concept of the ability to reverse ‘Warburg effect' in tumor cells prompted several
oncologic studies to investigate the potential use of DCA in antiglycolytic cancer therapy. A study
conducted in rats having subcutaneous cancer cells showed that orally administered DCA can
induce apoptosis and decrease cell proliferation [221]. Similar promising anticancer effects were
demonstrated for DCA administration in tumor xenograft studies in animals and humans with
glioblastoma [222]. However, it has been recently observed that DCA treatment could not suppress
or even enhance cell proliferation in cancer cell lines of human breast cancer, and cell lines of
human and murine neuroblastoma [223]. On the contrary, it has been reported that treatment with

of
DCA can overcome the resistance to sorafenib in a mouse model of sorafenib-resistant xenograft
HCC cells [224]. Also, combining oral DCA with adriamycin leads to enhanced cytotoxic effects
in treated hepatoma cells in vitro and in mice with subcutaneous xenografts in vivo [225]. Thus,

ro
preclinical data of DCA use in cancer therapy are not conclusive and need further confirmation.
Several phase I/ II clinical trials are ongoing for testing DCA activity and safety as an anticancer

-p
agent. A single-arm prospective study examined the safety of oral DCA in patients with brain
tumors, revealed no acute dose-limiting toxicities after at least one 4-week cycle [222]. Some of
re
the patients treated with DCA for about 75 days remained with stable disease, but two patients
reported grade 0–1 paresthesia. Another dose-escalation trial included 24 patients with solid
lP

malignancies to receive oral treatment with DCA at a dose of 6.25 mg/kg. Dose-limiting fatigue,
vomiting, and diarrhea are reported by three patients at a dose of 12.5 mg/kg. Tumor response
assessed by FDG-PET revealed stable disease in eight patients but no response in others [226].
na

Overall, the current clinical results of DCA use in antiglycolytic therapy are preliminary,
supporting a favorable toxicity profile rather than prominent anticancer efficacy.
ur

9.7- Targeting lactate dehydrogenase (LDHA inhibitors)


The LDH is a major glycolytic enzyme, that has five isozymes catalyzing the interconversion of
Jo

pyruvate and lactate coupled with NAD+ recycling. Accumulation of lactate can drastically lower
pH inside the cells. Thus, it is necessary for cells to actively efflux lactate into the outside, through
specific transporters (MCTs), as the final port of the lactate shuttle. The MCT1 and MCT4
isoforms play a crucial role in cancer metabolism through the efflux and influx of lactate between
adjacent cells [81]. Experimental studies demonstrated the overexpression of LDHA in several
tumor cells and its inhibition can markedly delay tumor progression [227]. LDHA is directly

25
regulated by the c-myc oncogene, and so the genetic or pharmacologic inhibition can diminish
myc-dependent tumors [62].
Inhibitors of LDHA, such as FX11 and oxamate, have demonstrated encouraging results in
preclinical studies and remain to show similar promising results in clinical trials. FXII, a catechol-
containing small compound that inhibits LDHA, was shown to inhibit tumor growth in xenografts
[228]. Oxamate, another LDHA inhibitor, can sensitize resistant tumor cells to the cytotoxic effect
of chemotherapeutic agents [229]. The LDHA inhibitors probably act by a mechanism similar to
that of DCA, involving increased delivery of pyruvate into the mitochondria to undergo

of
decarboxylation to acetyl-CoA, with enhanced mitochondrial activity, thereby increasing
oxidative stress leading to cancer cell death [228].

ro
9.8- Targeting tumor acidosis:
Targeting tumor acidosis and decreasing the intracellular alkalosis can be a useful addition to

-p
metabolic anticancer therapy [230]. Tumor acidity can be targeted by inhibition of proton release
using proton pump inhibitors (PPIs) such as omeprazole, either alone or in combination with
re
chemotherapeutic agents [47]. It has been demonstrated that PPIs can inhibit vacuolar-type
ATPase to increase the lysosomal pH and increase lysosomal-endosomal turnover in tumor cells
lP

[231]. In a case-control study, it was found that breast cancer patients were 25% less likely to have
prior use of PPIs compared with control women [232]. Also, a strong association was detected
between PPIs use and improved survival in patients with head and neck cancer [233]. Several other
na

reports revealed survival benefits from the combined use of PPIs with conventional
chemotherapeutics in many types of cancers including osteosarcoma [234], breast cancer [235] or
gut cancers [236]. The safety profile of PPIs on long-term use makes these drugs ideal candidates
ur

for more wide clinical anticancer studies.


Targeting the lactate shuttle through MCTs, as important regulators of pH status, could also be
Jo

exploited for anti-tumor effects. The specific MCT1 inhibitor, AZD3965, has shown promising
anticancer effects in preclinical studies [237]. The mechanism of the anticancer effect involves
intracellular trapping of lactate, decreasing pHi, with the death of tumor cells. AZD3965 is
currently under testing in phase I clinical trials in patients with B-cell lymphoma and prostate
cancer [238,239]. Inhibition of MCTs using α-cyano-4-hydroxycinnamic acid has also, been
demonstrated to produce a significant effect on tumor growth of malignant glioma [240].

26
Cinnamate is another small molecule inhibitor of MCTs, that was shown to decrease pHi and
inhibit tumor growth [241]. Further investigations are required to confirm the target specificity
and effectiveness of these candidate drugs before translation into anticancer clinical studies.
Moreover, Carbonic anhydrase 9 (CA9) enzyme is active at low pH and is involved in tumor
acidosis. CA9 is an important prognostic indicator in many tumors, associated with invasion and
metastasis [242]. Inhibition of CA9 was shown to suppress growth and metastasis in breast cancer
[243]. The CA9/CA12 inhibitor SLC0111 is investigated in phase I clinical trial in solid tumors
showing favorable results (NCT02215850).

of
9.9- Targeting other regulatory aspects of energy metabolism
The molecular regulators of glycolysis could also be targets for potential anticancer activity. For

ro
instance, KRAS mutation is found in most cases (>90%) of pancreatic cancer [244], where it
facilitates glycolysis and drive the PPP for the synthesis of nucleic acids [245], to enhance

-p
tumorigenesis [246]. Inhibition of KRAS or its downstream pathways revealed promising
anticancer effects in preclinical studies [247,248]. However, clinical trials in pancreatic cancer
re
patients reported no effect on overall survival from these KRAS inhibitors [249-251]. The c-MYC
oncogene is also a driver of glycolysis and tumorigenesis in many cancers [252]. Agents targeting
lP

MYC showed encouraging anticancer effects, and many of these agents are currently under
investigation in clinical trials. For example, inhibition of BET proteins downregulates MYC
expression and inhibits tumor growth [253,254]. Moreover, MYC inhibition can prevent the
na

development of resistance to mitochondrial inhibitors [255], making combination therapy with


these agents a potentially useful anticancer strategy.
The components of PI3K/Akt/mTOR signaling pathways are extensively investigated for potential
ur

anticancer effects. Many inhibitor agents for the PI3K/Akt pathway are under testing in preclinical
and early clinical studies as targeted anticancer therapies, including Afuresertib (GSK2110183),
Jo

Uprosertib (GSK2141795), Ipatasertib (GDC-0068), MK-2206, GDC‑ 0068, TCN and TCN-P
[256]. Most of these clinical trials with Akt inhibitors showed modest activity as single anticancer
agents, being most effective in cancers with PIK3 mutations and PTEN deficiency [257]. They
produce more promising effects when combined with other chemotherapeutic agents or with
radiotherapy [258].

27
Besides, inhibition of mTOR kinase was proposed as an effective approach for tumor suppression.
Recently, mTOR inhibitors have demonstrated significant activity in the treatment of several types
of tumors such as neuroendocrine, endometrial and breast cancers, medulloblastoma,
glioblastoma, osteosarcoma, leukemia and lymphoma [259-265]. The mTOR inhibitors have been
classified into three generations; including rapamycin and other rapalogs, temsirolimus,
everolimus, and ridaforolimus, as first-generation; the ATP-competitive inhibitor of mTOR kinase
(inhibit both mTORC1 and mTORC2), MLN0128, AZD2014, AZD8055 and CC223, as second-
generation inhibitors; and the dual PI3K/mTOR inhibitors, PP242, MLN0128, KU-0063794 and

of
BEZ235, as third-generation inhibitors [266]. Rapalogs, are mTORC1 inhibitors, showed limited
activity in clinical trials, as single anticancer agents, probably related to the various cross-talks of
the complicated mTOR pathway with other signaling pathways [267]. The newer generations of

ro
dual mTOR kinase inhibitors are less liable to induce tumor resistance than the rapalogs. These
agents are tested in preclinical studies and recently entered some clinical trials [268]. Also,

-p
halofuginone (HF) treatment have demonstrated potent anticancer activity in human colorectal
cancer cells. Halofuginone can reduce the Akt/mTORC1 signaling pathway, decreasing HK2 and
re
GlU1, leading to glycolysis inhibition [269]. Combination therapy targeting multiple pathways
may produce a better therapeutic outcome and overcome tumor resistance [266].
lP

Currently, the concept of reverse Warburg effect provides a new approach to anticancer strategies
by targeting glycolysis and oxidative phosphorylation [27]. Thus, it could be relevant for cancer
therapy to suppress glycolysis by targeting autophagy in the stroma decreasing the fuel supplied
na

to mitochondria of cancer cells, thereby inhibiting the two-compartment tumor metabolism.


Several such drugs have been approved including N-acetylcysteine, metformin,
hydroxychloroquine, and rapamycin [43]. Metformin is a biguanide oral antidiabetic drug, that
ur

activates AMPK, inhibits mitochondrial ETC Complex-1 and inhibits mTOR, which further
inhibits HIF1. It can reduce the activity of mitochondrial complex I, to inhibit OxPhos and reduce
Jo

ATP production, suppressing cancer progression [270]. Recent epidemiological studies revealed
that metformin-treated diabetic patients have lower cancer risk. Later the drug has been proven to
exhibit significant anti-proliferative activity in many clinical trials in cancer patients with or
without diabetes [271]. This antitumor activity involved several mechanisms including inhibition
of the respiratory complex I, and reduction of mTOR activity, in addition to the reduction of blood
glucose and insulin levels, decreasing its mitogen effect [272]. However, despite the promising

28
results in preclinical studies, the results of some clinical trials are inconclusive [273]. Inhibition of
both the main metabolic pathway and its main escape mechanism can completely suppress
malignant cells. Thus, the use of dual metabolic inhibitors such as metformin (inhibits
mitochondrial ETC complex 1) and 2-DG, (inhibits of glycolysis), was shown to produce profound
metabolic inhibition with more effective anticancer activity in preclinical studies [39]. The
synergetic effect has also been demonstrated for metformin combination with either the
bromodomain and extra-terminal motif (BET) inhibitor JQ-1 in pancreatic cancer [255] or PI3K
inhibition for ovarian cancer [274], which produces simultaneous inhibition of OxPhos and

of
glycolysis. Similarly, glycolysis inhibitor can be combined with other DNA damaging
chemotherapeutic agents could provide an effective approach to overcome anticancer resistance
through increasing cytotoxicity and decreasing repair capacity.

ro
The hypoxic condition prevailing in rapidly growing tumors can switch the balance between the
angiogenic activators and inhibitors, in favor of the angiogenesis activators such as vascular

-p
endothelial growth factor (VEGF), transforming growth factor-beta, fibroblast growth factor-2,
platelet-derived growth factor, interleukin-8, and angiopoietins [275]. The newly formed blood
re
vessels in tumors are structurally and functionally abnormal with tortuosity, leakiness and poor
covering by vascular supportive cells. These poorly functioning vessels lead to chaotic blood flow
lP

with the incorporation of tumor cells into the endothelial wall, enhancing blood dissemination of
tumors [276]. Agents inhibiting tumor angiogenesis have been tested for anticancer therapy,
including agents targeting the angiogenic factor VEGF by the monoclonal antibody, bevacizumab
na

[277], or by the fusion protein, aflibercept [278]. Several other agents are tyrosine kinase inhibitors
(sorafenib, sunitinib, regorafenib) that target the angiogenic receptors [279]. These agents have
shown promising results in the treatment of various types of cancers but, with little survival benefit
ur

[280]. However, complete vascular suppression can lead to marked hypoxia with decreased
perfusion and decreased the supply of tumors with oxygen, nutrients and also drugs. This may
Jo

result in a more aggressive tumor with a worse prognosis [281]. Moreover, anti-VEGF drugs can
decrease the vascular density in normal tissues, with reduced function of healthy organs [282].
Thus, further studies are required to support the beneficial role of antiangiogenic agents in cancer
therapy.

Table 1: Glycolytic targets, and main antiglycolytic agents in preclinical and clinical
development for anticancer therapy

29
Target metabolic Glycolytic inhibitors State of References
pathway development
Glucose transporters Phloretin Preclinical [119,120]
(GLUTs) Fasentin Preclinical [121]
STF-31 Preclinical [122]
WZB117 Preclinical [118,123]
Ritonavir Preclinical [124]
Silybin Phase I [125]
Hexokinase2 (HK2) 2-Deoxy-D-glucose (2-DG) Phase II [142,143,147,150,153]
lonidamine (LN) Phase II [163-168]
Genistein-27 Preclinical [129]
Benserazide Preclinical [130]

of
Resveratrol Preclinical [126]
Astragalin Preclinical [127]
Chrysin Preclinical [128]
Glyceraldehyde-3- 3-Bromopyruvate (3-BrPA) Phase I [204-209]

ro
phosphate dehydrogenase Koningic acid Preclinical [181-182]
(GAPDH) Iodoacetate Preclinical [179]
Pyruvate dehydrogenase Dichloroacetate (DCA) Phase I [222-226]

-p
kinase (PDK)
Lactate dehydrogenase Oxamate Preclinical [289]
(LDHA) FX11 Preclinical [288]
Monocarboxylate transportersα-Cyano-4-hydroxycinnamic acid Preclinical [240]
re
(MCTs) Cinnamate Preclinical [241]
AZD3965 Phase I [238,239]
Carbonic anhydrase (CA) SLC0111 Phase I [243]
lP

9/12
Proton pump Omeprazole, Esomeprazole, Phase I [232-236]
(H+/K+-ATPase) Lansoprazole, Pantoprazole
Mitochondrial AMPK Metformin Phase III [270-273]
Akt/mTORC1 Halofuginone Phase I [269]
na

mTORC1 Rapamycin, temsirolimus, Phase I [259-265]


everolimus, ridaforolimus
mTORC1/mTORC2 MLN0128, OSI-027 Phase I [267]
PI3K/mTOR PP242, NVP-BEZ235 Phase I [268]
ur

PI3K/Akt Afuresertib Phase I [256-258]


Uprosertib Phase I
Ipatasertib Phase I
KRAS mutation Sorafenib Phase 1 [249]
Jo

Trametinib Phase I [250]


Selumetinib Phase 1 [251]
Angiogenesis factors (e.g. Bevacizumab Phase II [277]
VEGF) and receptors Aflibercept Phase II [278]
Regorafenib Phase II [279, 280]

10- Conclusion and future perspective for metabolic cancer therapy

30
Although aerobic glycolysis is a known signature of cancer cells, targeting this pathway for therapy
has not yet been successfully translated into clinical practice. Recently, this hallmark of cancer
metabolism has become a focus of cancer research and drug discovery, aiming for the introduction
of effective metabolic agents as a promising strategy to combat cancer. These candidate drugs
might also sensitize tumor cells to other more effective cytotoxic therapies. Many emerging drugs
have been tested in preclinical studies with promising results and some are under investigation in
clinical trials with mixed favorable and disappointing results. For example, 2-DG is recently
resurrected in clinical trials in combination therapy with other chemotherapeutics or with

of
irradiation, after disappointing results as monotherapy. Other new antiglycolytic agents such as
DCA and 3-BrPA have been tested for monotherapy in solid glycolytic tumors such as HCC with
encouraging results and improved survival in animal models.

ro
However, a great concern arises about the significant toxicities of inhibiting metabolic enzymes in
normal cells especially in the immune system and stem cells that also display aerobic glycolysis.

-p
Novel solutions to circumvent this problem of systemic toxicity include selective targeted delivery
using recent imaging technology to deliver the drug directly into the tumor or intraarterially near
re
the tumor. Other approaches included micro-encapsulation and developing antiglycolytic agents
with high specificity for the molecular target, such as 3-BrPA that enters cells through MCTs
lP

which is overexpressed in cancer.


Moreover, cancer cells display metabolic plasticity and can overcome the inhibition of a specific
metabolic pathway via the expression or up-regulation of alternative pathways. Also, adjacent cells
na

such as fibroblasts and adipocytes can offer metabolic intermediates for the synthetic needs of
cancer cells. The great metabolic heterogeneity and cellular plasticity observed in solid tumors
made metabolic inhibitors unlikely to become effective as monotherapy for cancer. For instance,
ur

combination treatment with two or more antimetabolic gents inhibiting different metabolic
pathways simultaneously would decrease resistance and prevent relapse. Also, tumor glycolysis
Jo

has an important role in cancer resistance to chemotherapeutic agents, therefore antiglycolytic


agents could be helpful in sensitization of tumor cells and improving the therapeutic outcome for
treatment of resistant cancers as shown in preclinical models. Therefore, it is worth to explore the
outcome of combining conventional drugs with antiglycolytic agents to provide novel therapeutic
or preventive strategies for fighting against cancer.

31
Conflict of interest
The authors declare that there is no conflict of interests.

Funding
This research did not receive any specific grant from funding agencies in the public, commercial,
or not-for-profit sectors.

Acknowledgments

of
We thank Dr. Abdul-Salam Noor Waly, Dean Faculty of Medicine, Umm Al-Qura University,
Makkah, for providing access to the Saudi Digital Library.

ro
References
[1] F. Bray, J. Ferlay, I. Soerjomataram, R.L. Siegel, L.A. Torre, A. Jemal, Global cancer statistics

-p
2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185
countries. CA-Cancer J. Clin. 68 (2018) 394–424.
re
[2] H. Liu, Y.P. Hu, N. Savaraj, W. Priebe, T.J. Lampidis, Hypersensitization of tumor cells to
glycolytic inhibitors, Biochemistry 40(18) (2001) 5542–5547.
lP

[3] H. Pelicano, D.S. Martin, R.H. Xu, P. Huang, Glycolysis inhibition for anticancer treatment,
Oncogene 25 (34) (2006) 4633–4646.
[4] S. Ganapathy-Kanniappan, J-F.H. Geschwind, Tumor glycolysis as a target for cancer therapy:
na

progress and prospects, Mol. Cancer 12 (2013) 152.


[5] L.J. Savic, J. Chapiro, G. Duwe, J.F. Geschwind, Targeting glucose metabolism in cancer: new
class of agents for loco-regional and systemic therapy of liver cancer and beyond? Hepat. Oncol.
ur

3(1) (2016) 19–28. doi: 10.2217/hep.15.36.


[6] X.S. Chen, L. Li, Y. Guan, J.M. Yang, Y. Cheng, Anticancer strategies based on the metabolic
Jo

profile of tumor cells: therapeutic targeting of the Warburg effect, Acta Pharmacologica Sinica
37 (2016) 1013–1019.
[7] N.S. Akins, T.C. Nielson, H.V. Le, Inhibition of Glycolysis and Glutaminolysis: An Emerging
Drug Discovery Approach to Combat Cancer, Curr. Top. Med. Chem. 18(6) (2018) 494–504.
doi: 10.2174/1568026618666180523111351.

32
[8] M.G. Vander Heiden, L.C. Cantley, C.B. Thompson, Understanding the Warburg effect: the
metabolic requirements of cell proliferation, Science 324 (2009) 1029–1033.
[9] O. Warburg, On the origin of cancer cells, Science 123(3191) (1956) 309–314.
doi:10.1126/science.123.3191.309. 0036–8075.
[10] O. Warburg, K. Posener, E. Negelein, Über den Stoffwechsel der Carcinomzelle, Biochem.
Zeitschr. 152 (1924) 309–344.
[11] J.B. Bomanji, D.C. Costa, P.J. Ell, Clinical role of positron emission tomography in
oncology, Lancet Oncol 2(3) (2001) 157–164. doi: 10.1016/S1470-2045(00)00257-6.

of
[12] C. Frezza, E. Gottlieb, Mitochondria in cancer: not just innocent bystanders, Semin. Cancer
Biol. 19 (2009) 4–11. doi: 10.1016/j.semcancer.2008.11.008.
[13] X.L. Zu, M. Guppy, Cancer metabolism: facts, fantasy, and fiction, Biochem. Biophys. Res.

ro
Commun. 313 (2004) 459–465. doi: 10.1016/j.bbrc.2003.11.136.
[14] H.R. Christofk, M.G. Vander Heiden, M.H. Harris, A. Ramanathan, R.E. Gerszten, R. Wei,

-p
M.D. Fleming, S.L. Schreiber, L.C. Cantley, The M2 splice isoform of pyruvate kinase is
important for cancer metabolism and tumour growth, Nature 452 (2008) 230-233.
re
[15] M. Weiler-Sagie, O. Bushelev, R. Epelbaum, E.J. Dann, N. Haim, I. Avivi, et al., 18F-FDG
avidity in lymphoma readdressed: a study of 766 patients, J. Nucl. Med. 51 (2010) 25–30.
lP

[16] S.Y. Lunt, M.G. Vander Heiden, Aerobic glycolysis: meeting the metabolic requirements of
cell proliferation, Annu. Rev. Cell Dev. Biol. 27 (2011) 441–464 10.1146/annurev-cellbio-
092910-154237.
na

[17] S.P. Mathupala, Y.H. Ko, P.L. Pedersen, The pivotal roles of mitochondria in cancer:
Warburg and beyond and encouraging prospects for effective therapies, Biochimica et Biophysica
Acta (BBA) Bioenergetics 1797 (2010) 1225–1230.
ur

[18] H. Zhou, B. Zhang, J. Zheng, M. Yu, T. Zhou, K. Zhao, et al., The inhibition of migration and
invasion of cancer cells by graphene via the impairment of mitochondrial respiration, Biomaterials
Jo

35 (2014) 1597–607.
[19] A. Roesch, A. Vultur, I. Bogeski, H. Wang, K.M. Zimmermann, D. Speicher, et al.,
Overcoming intrinsic multidrug resistance in melanoma by blocking the mitochondrial respiratory
chain of slow-cycling JARID1B high cells, Cancer Cell 23 (2013) 811–825.
[20] M.C. Maiuri, G. Kroemer, Essential Role for Oxidative Phosphorylation in Cancer
Progression, Cell Metabolism 21 (2015) 11–12.

33
[21] Y. Zhang, J.M. Yang, Altered energy metabolism in cancer: a unique opportunity for
therapeutic intervention, Cancer Biol. Ther. 14(2) (2013) 81–89. doi: 10.4161/cbt.22958.
[22] S. Bonnet, S.L. Archer, J. Allalunis-Turner, A. Haromy, C. Beaulieu, R. Thompson, C.T. Lee,
G.D. Lopaschuk, L. Puttagunta, S. Bonnet, G. Harry, K. Hashimoto, C.J. Porter, M.A. Andrade,
B. Thebaud, E.D. Michelakis, A mitochondria-K + channel axis is suppressed in cancer and its
normalization promotes apoptosis and inhibits cancer growth, Cancer Cell 11(1) (2007) 37–51.
doi: 10.1016/j.ccr.2006.10.020.
[23] R.A. Gatenby, R.J. Gillies, Why do cancers have high aerobic glycolysis? Nat. Rev. Cancer 4

of
(2004) 891–9.
[24] H. Wu, Z. Ding, D. Hu, F. Sun, C. Dai, J. Xie, et al., Central role of lactic acidosis in cancer
cell resistance to glucose deprivation-induced cell death, J. Pathol. 227 (2012) 189–99.

ro
[25] W. Lu, Y. Hu, G. Chen, Z. Chen, H. Zhang, F. Wang, L. Feng, H. Pelicano, H. Wang, M.J.
Keating, J. Liu, W. McKeehan, H. Wang, Y. Luo, P. Huang, Novel role of NOX in supporting

-p
aerobic glycolysis in cancer cells with mitochondrial dysfunction and as a potential target for
cancer therapy, PLoS Biol. 10(5) (2012) e1001326. doi: 10.1371/journal.pbio.1001326.
re
[26] C.L. Lu, L. Qin, H.C. Liu, D. Candas, M. Fan, J.J. Li, Tumor Cells Switch to Mitochondrial
Oxidative Phosphorylation under Radiation via mTOR-Mediated Hexokinase II Inhibition-A
lP

Warburg-Reversing Effect, PloS One 10 (2015) 10.


[27] J. Xie, H. Wu, C. Dai, Q. Pan, Z. Ding, D. Hu, et al., Beyond Warburg effect–dual metabolic
nature of cancer cells, Sci. Rep. 4 (2014) 4927.
na

[28] C.V. Dang, Rethinking the Warburg effect with Myc micromanaging glutamine
metabolism, Cancer Research 70 (2010) 859–862.
[29] M.C. Caino, J.C. Ghosh, Y.C. Chae, V. Vaira, D.B. Rivadeneira, A. Faversani, et al., PI3K
ur

therapy reprograms mitochondrial trafficking to fuel tumor cell invasion, Proc. Natl. Acad. Sci.
112 (2015) 8638–8643.
Jo

[30] C.J. De Saedeleer, T. Copetti, P.E. Porporato, J. Verrax, O. Feron, P. Sonveaux, Lactate
activates HIF-1 in oxidative but not in Warburg-phenotype human tumor cells, PloS One 7 (2012)
e46571.
[31] D. Jia, M. Lu, K.H. Jung, J.H. Park, L. Yu, J.N. Onuchic, B.A. Kaipparettu, H. Levine,
Elucidating cancer metabolic plasticity by coupling gene regulation with metabolic pathways,
Proc. Natl. Acad. Sci. USA 116(9) (2019) 3909-3918. doi: 10.1073/pnas.1816391116.

34
[32] C.R. Mantel, H.A. O’Leary, B.R. Chitteti, X. Huang, S. Cooper, G. Hangoc, et al., Enhancing
hematopoietic stem cell transplantation efficacy by mitigating oxygen shock, Cell 161 (2015)
1553-1565.
[33] V.S. LeBleu, J.T. O’Connell, K.N.G. Herrera, H. Wikman, K. Pantel, M.C. Haigis, et al.,
PGC-1α mediates mitochondrial biogenesis and oxidative phosphorylation in cancer cells to
promote metastasis, Nature Cell Biology 16 (2014) 992–1003.
[34] P.L. Pedersen, Mitochondria in relation to cancer metastasis: introduction to a minireview
series, J. Bioenergy Biomembr. 44(6) (2012) 615–617.

of
[35] D. Jia, J.H. Park, K.H. Jung, H. Levine, B.A. Kaipparettu, Elucidating the Metabolic Plasticity
of Cancer: Mitochondrial Reprogramming and Hybrid Metabolic States, Cells 7 (2018) 21.
doi:10.3390/cells7030021.

ro
[36] L. Yu, M. Lu, D. Jia, J. Ma, E. Ben-Jacob, H. Levine, B.A. Kaipparettu, J.N. Onuchic,
Modeling the Genetic Regulation of Cancer Metabolism: Interplay between Glycolysis and

-p
Oxidative Phosphorylation. Cancer Res. 77 (2017) 1564–1574.
[37] R. Kalluri, R.A. Weinberg, The basics of epithelial-mesenchymal transition. J. Clin. Investing.
re
119 (2009) 1420–1428.
[38] L. Zhang, G. Huang, X. Li, Y. Zhang, Y. Jiang, J. Shen, J. Liu, Q. Wang, et al., Hypoxia
lP

induces epithelial-mesenchymal transition via activation of SNAI1 by hypoxia-inducible factor -


1α in hepatocellular carcinoma. BMC Cancer 13 (2013) 108.
[39] J.H. Cheong, E.S. Park, J. Liang, J.B. Dennison, D. Tsavachidou, C. Nguyen-Charles, et al.,
na

Dual inhibition of tumor energy pathway by 2-deoxyglucose and metformin is effective against a
broad spectrum of preclinical cancer models. Mol. Cancer Ther. 10 (2011) 2350–2362.
[40] S. Pavlides, D. Whitaker-Menezes, R. Castello-Cros, N. Flomenberg, A.K. Witkiewicz, P.G.
ur

Frank, et al., The reverse Warburg effect: aerobic glycolysis in cancer associated fibroblasts and
the tumor stroma, Cell Cycle 8 (2009) 3984–4001.
Jo

[41] F. Sotgia, U.E. Martinez-Outschoorn, S. Pavlides, A. Howell, R.G. Pestell, Understanding the
Warburg effect and the prognostic value of stromal caveolin-1 as a marker of a lethal tumor
microenvironment, Breast Cancer Res. 13 (2011) 213.
[42] K. Sun, S. Tang, Y. Hou, L. Xi, Y. Chen, J. Yin, M. Peng, M. Zhao, X. Cui, M. Liu, Oxidized
ATM-mediated glycolysis enhancement in breast cancer-associated fibroblasts contributes to
tumor invasion through lactate as metabolic coupling, E. Bio. Medicine 41 (2019) 370 –383.

35
[43] C.D. Gonzalez, S. Alvarez, A. Ropolo, C. Rosenzvit, M.F. Bagnes, M.I. Vaccaro, Autophagy,
Warburg, and Warburg reverse effects in human cancer, Biomed. Res. Int. 2014 (2014) 926729.
[44] U.E. Martinez-Outschoorn, Z. Lin, Y.H. Ko, A.F. Goldberg, N. Flomenberg, C. Wang, et
al., Understanding the metabolic basis of drug resistance: therapeutic induction of the Warburg
effect kills cancer cells, Cell Cycle 10 (2011) 2521–8.
[45] T.N. Seyfried, L.M. Shelton, Cancer as a metabolic disease, Nutr. Metab. 7 (2010) 7.
doi:10.1186/1743-7075-7-7.
[46] E.P. Spugnini, P. Sonveaux, C. Stock, M. Perez-Sayans, A. De Milito, S. Avnet, A.G. Garcìa,

of
S. Harguindey, S. Fais, Proton channels and exchangers in cancer., Biochim. Biophys. Acta BBA
Biomembr. 1848 (2014) 2715–2726. doi:10.1016/j.bbamem.2014.10.015.
[47] S. Taylor, E.P. Spugnini, Y.G. Assaraf, T. Azzarito, C. Rauch, S. Fais, Microenvironment

ro
acidity as a major determinant of tumor chemoresistance: Proton pump inhibitors (PPIs) as a novel
therapeutic approach, Drug Resist. Updat. Rev. Comment. Antimicrob. Anticancer Chemother. 23
(2015) 69–78. doi:10.1016/j.drup.2015.08.004.

-p
[48] V. Huber, A. DeMilito, S. Harguindey, S.J. Reshkin, M.L. Wahl, C. Rauch, A. Chiesi, J.
re
Pouysségur, R.A. Gatenby, L. Rivoltini, S. Fais, Proton dynamics in cancer., J. Transl. Med. 8
(2010) 57. doi:10.1186/1479-5876-8-57.
lP

[49] B.A. Kaipparettu, Y. Ma, J.H. Park, T.L. Lee, Y. Zhang, P. Yotnda, C.J. Creighton, W.Y.
Chan, L.J.C. Wong, Crosstalk from non-cancerous mitochondria can inhibit tumor properties of
metastatic cells by suppressing oncogenic pathways, PloS One. 8 (2013) e61747.
na

doi:10.1371/journal.pone.0061747.
[50] V.F. Casneuf, P. Fonteyne, N. Van Damme, P. Demetter, P. Pauwels, et al., Expression of
SGLT1, Bcl-2 and p53 in primary pancreatic cancer related to survival, Cancer Invest. 26 (2008)
ur

852–859.
[51] S.P. Mathupala, Y.H. Ko, P.L. Pedersen, Hexokinase-2 bound to mitochondria: cancer’s
Jo

stygian link to the “Warburg Effect” and a pivotal target for effective therapy, Semin. Cancer
Biol. 19(1) (2009) 17–24. doi: 10.1016/j.semcancer.2008.11.006.
[52] R.B. Robey, N. Hay, Mitochondrial hexokinases, novel mediators of the antiapoptotic effects
of growth factors and Akt. Oncogene, 25 (2006) 4683–4696. doi: 10.1038/sj.onc.1209595.
[53] H.J. Jae, J.W. Chung, H.S. Park, M.J. Lee, K.C. Lee, H.C. Kim, J.H. Yoon, H. Chung, J.H.
Park, The antitumor effect and hepatotoxicity of a hexokinase II inhibitor 3-bromopyruvate: in

36
vivo investigation of intraarterial administration in a rabbit VX2 hepatoma model, Korean J.
Radiol. 10 (2009) 596–603. doi: 10.3348/kjr.2009.10.6.596.
[54] S.L. Colombo, M. Palacios-Callender, N. Frakich, J. De Leon, C.A. Schmitt, L. Boorn, N.
Davis, S. Moncada, Anaphase-promoting complex/cyclosome-Cdh1 coordinates glycolysis and
glutaminolysis with transition to S phase in human T lymphocytes, Proc. Natl. Acad. Sci. USA
107 (2010) 18868–18873. doi: 10.1073/pnas.1012362107.
[55] J. Kinderlerer, S. Ainsworth, C.N. Morris, N. Rhodes, The regulatory properties of yeast
pyruvate kinase: Effect of Ph, Biochem. J. 234 (1986) 699–703.

of
[56] K. Imamura, T. Tanaka, Multimolecular forms of pyruvate kinase from rat and other
mammalian tissues I: Electrophoretic studies, J. Biochem. 71 (1972) 1043–51.
[57] W. Luo, G.L. Semenza, Emerging roles of PKM2 in cell metabolism and cancer

ro
progression, Trends Endocrinol. Metab. 23 (2012) 560–566.
[58] W. Yang, Z. Lu, Nuclear PKM2 regulates the Warburg effect, Cell Cycle 12 (2013) 3154–
3158.

-p
[59] J.F. Barger, D.R. Plas, Balancing biosynthesis and bioenergetics: metabolic programs in
re
oncogenesis, Endocr. Relat. Cancer 17 (2010) R287–304.
[60] H.R. Christofk, M.G. Vander Heiden, N. Wu, J.M. Asara, L.C. Cantley, Pyruvate kinase M2
lP

is a phosphotyrosine binding protein, Nature 452 (2008) 181–186.


[61] K. Bluemlein, N.M. Grüning, R.G. Feichtinger, H. Lehrach, B. Kofler, M. Ralser, No
evidence for a shift in pyruvate kinase PKM1 to PKM2 expression during
na

tumorigenesis, Oncotarget 2 (2011) 393-400.


[62] H. Shim, C. Dolde, B.C. Lewis, C.S. Wu, G. Dang, R.A. Jungmann, R. Dalla-Favera, C.V.
Dang, c-Myc transactivation of LDH-A: implications for tumor metabolism and growth, Proc.
ur

Natl. Acad. Sci. USA 94 (1997) 6658–6663. doi: 10.1073/pnas.94.13.6658.


[63] T. Pfeiffer, S. Schuster, S. Bonhoeffer, Cooperation and competition in the evolution of ATP-
Jo

producing pathways, Science 292(5516) (2001) 504–507. doi: 10.1126/science.1058079.


[64] J.W. Locasale, L.C. Cantley, Altered metabolism in cancer, BMC Biol 8; (2010): 88.
[65] R.J. Deberardinis, N. Sayed, D. Ditsworth, C.B. Thompson, Brick by brick: metabolism and
tumor cell growth, Curr. Opin. Genet. Dev. 18(1) (2008) 54–61. doi: 10.1016/j.gde.2008.02.003.
[66] D.S. Backos, C.C. Franklin, P. Reigan, The role of glutathione in brain tumor drug
resistance, Biochem. Pharmacol. 83(8) (2012) 1005–1012. doi: 10.1016/j.bcp.2011.11.016.

37
[67] R. Possemato, K.M. Marks, Y.D. Shaul, et al., Functional genomics reveal that the serine
synthesis pathway is essential in breast cancer, Nature 476(7360) (2011) 346–350.
[68] K. Mikuriya, Y. Kuramitsu, S. Ryozawa, M. Fujimoto, S. Mori, M. Oka, K. Hamano, K.
Okita, I. Sakaida, K. Nakamura, Expression of glycolytic enzymes is increased in pancreatic
cancerous tissues as evidenced by proteomic profiling by two-dimensional electrophoresis and
liquid chromatography-mass spectrometry/mass spectrometry, Int. J. Oncol. 30(4) (2007) 849–
855.
[69] C.S. Yeh, J.Y. Wang, F.Y. Chung, S.C. Lee, M.Y. Huang, C.W. Kuo, M.J. Yang, S.R. Lin,

of
Significance of the glycolytic pathway and glycolysis related-genes in tumorigenesis of human
colorectal cancers, Oncol. Rep. 19(1) (2008) 81–91.
[70] Y. Zhou, F. Tozzi, J. Chen, F. Fan, L. Xia, J. Wang, G. Gao, A. Zhang, X. Xia, H. Brasher,

ro
W. Widger, L.M. Ellis, Z. Weihua, Intracellular ATP levels are a pivotal determinant of
chemoresistance in colon cancer cells, Cancer Res. 72(1) (2012) 304–314. doi: 10.1158/0008-
5472.CAN-11-1674.

-p
[71] S.P. Pitroda, B.T. Wakim, R.F. Sood, M.G. Beveridge, M.A. Beckett, D.M. MacDermed,
re
R.R. Weichselbaum, N.N. Khodarev, STAT1-dependent expression of energy metabolic pathways
links tumour growth and radioresistance to the Warburg effect, BMC Med. 7 (2009) 68. doi:
lP

10.1186/1741-7015-7-68.
[72] X. Xu, A. Zur Hausen, J.F. Coy, M. Lochelt, Transketolase-like protein 1 (TKTL1) is required
for rapid cell growth and full viability of human tumor cells, Int. J. Cancer 124(6) (2009) 1330-
na

1337. doi: 10.1002/ijc.


[73] F. Zhao, A. Mancuso, T.V. Bui, X. Tong, J.J.Gruber, C.R. Swider, P.V. Sanchez, J.J. Lum,
N. Sayed, J.V. Melo, A.E. Perl, M. Carroll, S.W. Tuttle, C.B. Thompson, Imatinib resistance
ur

associated with BCR-ABL upregulation is dependent on HIF-1alpha-induced metabolic


reprograming, Oncogene 29(20) (2010) 2962–2972. doi: 10.1038/onc.2010.67.
Jo

[74] F. Monteleone, R. Rosa, M. Vitale, C. D'Ambrosio, M. Succoio, L. Formisano, L. Nappi,


M.F. Romano, A. Scaloni, G. Tortora, R. Bianco, N. Zambrano, Increased anaerobic metabolism
is a distinctive signature in a colorectal cancer cellular model of resistance to antiepidermal growth
factor receptor antibody, Proteomics 13(5) (2013) 866–877. doi: 10.1002/pmic.201200303.

38
[75] C.W. Lu, S.C. Lin, K.F. Chen, Y.Y. Lai, S.J. Tsai, Induction of pyruvate dehydrogenase
kinase-3 by hypoxia-inducible factor-1 promotes metabolic switch and drug resistance, J. Biol.
Chem. 283(42) (2008) 28106–28114. doi: 10.1074/jbc.M803508200.
[76] M. Fanciulli, T. Bruno, A. Giovannelli, F.P. Gentile, M. Di Padova, O. Rubiu, A. Floridi,
Energy metabolism of human LoVo colon carcinoma cells: correlation to drug resistance and
influence of lonidamine, Clin. Cancer Res. 6(4) (2000) 1590–1597.
[77] K. Birsoy, D.M. Sabatini, R. Possemato, Untuning the tumor metabolic machine: targeting
cancer metabolism: a bedside lesson, Nat. Med. 18(7) (2012) 1022–1023. doi: 10.1038/nm.2870.

of
[78] S. Walenta, M. Wetterling, M. Lehrke, G. Schwickert, K. Sundfor, E.K. Rofstad, W. Mueller-
Klieser, High lactate levels predict likelihood of metastases, tumor recurrence, and restricted
patient survival in human cervical cancers, Cancer Res. 60(4) (2000) 916–921.

ro
[79] P. Sonveaux, F. Vegran, T. Schroeder, M.C. Wergin, J. Verrax, Z.N. Rabbani, C.J. De
Saedeleer, K.M. Kennedy, C. Diepart, B.F. Jordan, M.J. Kelley, B. Gallez, M.L. Wahl, O. Feron,

-p
M.W. Dewhirst, Targeting lactate-fueled respiration selectively kills hypoxic tumor cells in
mice, J. Clin. Invest. 118(12) (2008) 3930–3942.
re
[80] N. Draoui, O. Feron, Lactate shuttles at a glance: from physiological paradigms to anti-cancer
treatments, Dis. Model Mech. 4(6) (2011) 727–732. doi: 10.1242/dmm.007724.
lP

[81] C. Pinheiro, A. Longatto-Filho, J. Azevedo-Silva, M. Casal, F.C. Schmitt, F. Baltazar, Role


of monocarboxylate transporters in human cancers: state of the art, J. Bioenerg. Biomembr. 44(1)
(2012) 127–139. doi: 10.1007/s10863-012-9428-1.
na

[82] R. Hussien, G.A. Brooks, Mitochondrial and plasma membrane lactate transporter and lactate
dehydrogenase isoform expression in breast cancer cell lines, Physiol. Genomics 43(5) (2011)
255–264.
ur

[83] O. Feron, Pyruvate into lactate and back: from the Warburg effect to symbiotic energy fuel
exchange in cancer cells, Radiother. Oncol. 92(3) (2009) 329–333. doi:
Jo

10.1016/j.radonc.2009.06.025.
[84] J.W. Kim, C.V. Dang, Multifaceted roles of glycolytic enzymes, Trends Biochem. Sci. 30(3)
(2005) 142-150.
[85] N. Majewski, V. Nogueira, P. Bhaskar, P.E. Coy, J.E. Skeen, K. Gottlob, N.S. Chandel, C.B.
Thompson, R.B. Robey, N. Hay, Hexokinase-mitochondria interaction mediated by Akt is required
to inhibit apoptosis in the presence or absence of Bax and Bak, Mol. Cell 16(5) (2004) 819–830.

39
[86] N.W. Seidler, GAPDH and intermediary metabolism, Adv. Exp. Med. Biol. 985 (2013) 37–
59.
[87] S. Ganapathy-Kanniappan, R. Kunjithapatham, J.F. Geschwind, Glyceraldehyde-3-phosphate
dehydrogenase: a promising target for molecular therapy in hepatocellular carcinoma, Oncotarget
3(9) (2012) 940–953.
[88] W. Yang, Y. Xia, D. Hawke, X. Li, J. Liang, D. Xing, K. Aldape, T. Hunter, W.K. Alfred
Yung, Z. Lu, PKM2 phosphorylates histone H3 and promotes gene transcription and
tumorigenesis, Cell 150(4) (2012) 685–696. doi: 10.1016/j.cell.2012.07.018.

of
[89] W. Yang, Y. Xia, H. Ji, Y. Zheng, J. Liang, W. Huang, X. Gao, K. Aldape, Z. Lu, Nuclear
PKM2 regulates beta-catenin transactivation upon EGFR activation, Nature 480(7375) (2011)
118–122.

ro
[90] F.V. Filipp, Cancer metabolism meets systems biology: Pyruvate kinase isoform PKM2 is a
metabolic master regulator, J. Carcinog. 12 (2013) 14. doi: 10.4103/1477-3163.115423.

-p
[91] N.A. Demarse, S. Ponnusamy, E.K. Spicer, E. Apohan, J.E. Baatz, B. Ogretmen, C. Davies,
Direct binding of glyceraldehyde 3-phosphate dehydrogenase to telomeric DNA protects
re
telomeres against chemotherapy-induced rapid degradation, J. Mol. Biol. 394(4) (2009) 789-803.
doi: 10.1016/j.jmb.2009.
lP

[92] N. Harada, R. Yasunaga, Y. Higashimura, R. Yamaji, K. Fujimoto, J. Moss, H. Inui, Y.


Nakano, Glyceraldehyde-3-phosphate dehydrogenase enhances transcriptional activity of
androgen receptor in prostate cancer cells, J. Biol. Chem. 282(31) (2007) 22651–22661. doi:
na

10.1074/jbc.M610724200.
[93] O. Popanda, G. Fox, H.W. Thielmann, Modulation of DNA polymerases alpha, delta and
epsilon by lactate dehydrogenase and 3-phosphoglycerate kinase, Biochim. Biophys. Acta 1397(1)
ur

(1998) 102-117.
[94] N. Pertega-Gomes, JR. Vizcaino, V. Miranda-Goncalves, C. Pinheiro, J. Silva, H. Pereira, P.
Jo

Monteiro, R.M. Henrique, R.M. Reis, C. Lopes, F. Baltazar, Monocarboxylate transporter 4


(MCT4) and CD147 overexpression is associated with poor prognosis in prostate cancer. BMC
Cancer 11 (2011) 312.
[95] R. Diaz-Ruiz, N. Averet, D. Araiza, B. Pinson, S. Uribe-Carvajal, A. Devin, M. Rigoulet,
Mitochondrial oxidative phosphorylation is regulated by fructose 1,6-bisphosphate. A possible

40
role in Crabtree effect induction? J. Biol. Chem. 283(40) (2008) 26948–26955. doi:
10.1074/jbc.M800408200.
[96] M. Wartenberg, M. Richter, A. Datchev, S. Gunther, N. Milosevic, M.M. Bekhite, H.R.
Figulla, J.M. Aran, J. Petriz, H. Sauer, Glycolytic pyruvate regulates P-Glycoprotein expression
in multicellular tumor spheroids via modulation of the intracellular redox state, J. Cell Biochem.
109(2) (2010) 434–446.
[97] H. Pelicano, D.S. Martin, R.H. Xu, P. Huang, Glycolysis inhibition for anticancer
treatment, Oncogene 25(34) (2006) 4633–4646. doi: 10.1038/sj.onc.1209597.

of
[98] C.V. Dang, A. Le, P. Gao, MYC-induced cancer cell energy metabolism and therapeutic
opportunities, Clin. Cancer Res. 15 (2009) 6479–83.
[99] A. Valera, A. Pujol, X. Gregori, E. Riu, J. Visa, F. Bosch, Evidence from transgenic mice that

ro
myc regulates hepatic glycolysis, Faseb. J. 9 (1995) 1067–78.
[100] J.W. Kim, P. Gao, Y.C. Liu, G.L. Semenza, C.V. Dang, Hypoxia-inducible factor 1 and

-p
dysregulated c-Myc cooperatively induce vascular endothelial growth factor and metabolic
switches hexokinase 2 and pyruvate dehydrogenase kinase 1, Mol. Cell Biol. 27 (2007) 7381–
re
7393.
[101] R.L. Elstrom, D.E. Bauer, M. Buzzai, R. Karnauskas, M.H. Harris, D.R. Plas, et al., Akt
lP

stimulates aerobic glycolysis in cancer cells, Cancer Res 64 (2004) 3892–3899.


[102] K. Gottlob, N. Majewski, S. Kennedy, E. Kandel, R.B. Robey, N. Hay, Inhibition of early
apoptotic events by Akt/PKB is dependent on the first committed step of glycolysis and
na

mitochondrial hexokinase, Genes Dev. 15 (2001) 1406–1418.


[103] I. Ben-Sahra, J.J. Howell, J.M. Asara, B.D. Manning, Stimulation of de novo pyrimidine
synthesis by growth signaling through mTOR and S6K1, Science 339 (2013) 1323–1328.
ur

[104] K. Maiese, Z.Z. Chong, Y.C. Shang, S. Wang, mTOR: ontarget for novel therapeutic
strategies in the nervous system, Trends Mol. Med. 19 (2013) 51–60.
Jo

[105] D.D. Sarbassov, D.A. Guertin, S.M. Ali, D.M. Sabatini, Phosphorylation and regulation of
Akt/PKB by the rictor-mTOR complex, Science 307 (2005) 1098–1101.
[106] K. Masui, K. Tanaka, D. Akhavan, I. Babic, B. Gini, T. Matsutani, et al., mTOR complex 2
controls glycolytic metabolism in glioblastoma through FoxO acetylation and upregulation of c-
Myc, Cell Metab. 18 (2013) 726-39

41
[107] S.H. Liao, X.Y. Zhao, Y.H. Han, J. Zhang, L.S. Wang, L. Xia, et al., Proteomics-based
identification of two novel direct targets of hypoxia-inducible factor-1 and their potential roles in
migration/invasion of cancer cells, Proteomics 9 (2009) 3901–3912.
[108] P.H. Maxwell, M.S. Wiesener, G.W. Chang, S.C. Clifford, E.C. Vaux, M.E. Cockman, et
al., The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent
proteolysis, Nature 399 (1999) 271–275.
[109] W.J. Kaelin, P.J. Ratcliffe, Oxygen sensing by metazoans: the central role of the HIF
hydroxylase pathway, Mol. Cell 30 (2008) 393–402.

of
[110] K.K. Velpula, A. Bhasin, S. Asuthkar, A.J. Tsung, Combined targeting of PDK1 and EGFR
triggers regression of glioblastoma by reversing the Warburg effect, Cancer Res. 73 (2013) 7277–
7289.

ro
[111] J.W. Kim, I. Tchernyshyov, G.L. Semenza, C.V. Dang, HIF-1-mediated expression of
pyruvate dehydrogenase kinase: a metabolic switch required for cellular adaptation to
hypoxia, Cell Metab. 3 (2006) 177–185.

-p
[112] Z. Feng, A.J. Levine, The regulation of energy metabolism and the IGF-1/mTOR pathways
re
by the p53 protein, Trends Cell Biol. 20 (2010) 427–434.
[113] S.J. Bensinger, C. HR, New aspects of the Warburg effect in cancer cell biology, Semin.
lP

Cell Dev. Biol. 23 (2012) 352–361.


[114] R.B. Hamanaka, N.S. Chandel, Targeting glucose metabolism for cancer therapy, J. Exp.
Med. 209(2) (2012) 211–215.
na

[115] R.J. Klement, U. Kammerer, Is there a role for carbohydrate restriction in the treatment and
prevention of cancer? Nutr. Metab. (Lond) 8 (2011) 75. doi: 10.1186/1743-7075-8-75.
[116] P. Jagust, B. de Luxán-Delgado, B. Parejo-Alonso, P. Sancho, Metabolism-Based
ur

Therapeutic Strategies Targeting Cancer Stem Cells, Front. Pharmacol. 10 (2019) 203. doi:
10.3389/fphar.2019.00203
Jo

[117] M. Jang, S.S. Kim, J. Lee, Cancer cell metabolism: implications for therapeutic targets, Exp.
Mol. Med. 45 (2013) e45. doi: 10.1038/emm.2013.85.
[118] Y. Liu, Y. Cao, W. Zhang, S. Bergmeier, Y. Qian, H. Akbar, R. Colvin, J. Ding, L. Tong, S.
Wu, J. Hines, X. Chen, A small-molecule inhibitor of glucose transporter 1 downregulates
glycolysis, induces cell-cycle arrest, and inhibits cancer cell growth in vitro and in vivo, Mol.
Cancer Ther. 11(8) (2012) 1672–1682. doi: 10.1158/1535-7163.MCT-12-0131.

42
[119] K.H. Wu, C.T. Ho, Z.F. Chen, L.C. Chen, J. Whang-Peng, T.N. Lin, Y.S. Ho, The apple
polyphenol phloretin inhibits breast cancer cell migration and proliferation via inhibition of signals
by type 2 glucose transporter, J. Food Drug Anal. 26 (2018) 221-231.
[120] S.T. Lin, S.H. Tu, P.S. Yang, S.P. Hsu, W.H. Lee, C.T. Ho, C.H. Wu, Y.H. Lai, M.Y. Chen,
L.C. Chen, Apple polyphenol phloretin inhibits colorectal cancer cell growth via inhibition of the
type 2 glucose transporter and activation of p53mediated signaling, J. Agric. Food Chem. 64
(2016) 6826-6837.
[121] T.E. Wood, S. Dalili, C. Simpson, R. Hurren, X. Mao, F.S. Saiz, M. Gronda, et al., A novel

of
inhibitor of glucose uptake sensitizes cells to fas-induced cell death, Mol. Cancer Ther. 7 (2008)
3546-3555.
[122] D.A. Chan, P.D. Sutphin, P. Nguyen, S. Turcotte, E.W. Lai, A. Banh, G.E. Reynolds, J.T.

ro
Chi, J. Wu, D.E. Solow-Cordero, et al., Targeting GLUT1 and the Warburg effect in renal cell
carcinoma by chemical synthetic lethality, Sci. Transl. Med. 3 (2011) 94ra70.

-p
[123] Y. Zhao, E.B. Butler, M. Tan, Targeting cellular metabolism to improve cancer therapeutics,
Cell Death Dis. 4 (2013) e532.
re
[124] S. Dalva-Aydemir, R. Bajpai, M. Martinez, K.U. Adekola, I. Kandela, C. Wei, S. Singhal,
et al., Targeting the metabolic plasticity of multiple myeloma with fda-approved ritonavir and
lP

metformin, Clin. Cancer Res. 21 (2015) 1161-1171.


[125] T.W. Flaig, D.L. Gustafson, L.J. Su, J.A. Zirrolli, F. Crighton, G.S. Harrison, et al., A phase
I and pharmacokinetic study of silybin-phytosome in prostate cancer patients, Invest. New Drugs
na

25 (2006) 139–146. doi: 10.1007/s10637006-9019-2.


[126] W. Dai, F. Wang, J. Lu, Y. Xia, L. He, K. Chen, J. Li, S. Li, T. Liu, Y. Zheng, et al., By
reducing hexokinase 2, resveratrol induces apoptosis in HCC cells addicted to aerobic glycolysis
ur

and inhibits tumor growth in mice, Oncotarget, 6 (2015) 13703-13717.


[127] W. Li, J. Hao, L. Zhang, Z. Cheng, X. Deng, G. Shu, Astragalin reduces hexokinase 2
Jo

through increasing miR-125b to inhibit the proliferation of hepatocellular carcinoma cells in Vitro
and in Vivo, J. Agric. Food Chem. 65 (2017) 5961-5972.
[128] D. Xu, J. Jin, H. Yu, Z. Zhao, D. Ma, C. Zhang, H. Jiang, Chrysin inhibited tumor glycolysis
and induced apoptosis in hepatocellular carcinoma by targeting hexokinase-2, J. Exp. Clin. Cancer
Res. 36 (2017) 44.

43
[129] L. Tao, L. Wei, Y. Liu, Y. Ding, X. Liu, X. Zhang, X. Wang, Y. Yao, J. Lu, Q. Wang, R.
Hu, Gen-27, a newly synthesized flavonoid, inhibits glycolysis and induces cell apoptosis via
suppression of hexokinase II in human breast cancer cells, Biochem. Pharmacol.125 (2017)12-25.
[130] W. Li, M. Zheng, S. Wu, S. Gao, M. Yang, Z. Li, Q. Min, W. Sun, L. Chen, G. Xiang, H.
Li, Benserazide, a dopadecarboxylase inhibitor, suppresses tumor growth by targeting hexokinase
2, J. Exp. Clin. Cancer Res. 36 (2017) 58.
[131] J.C. Maher, A. Krishan, T.J. Lampidis, Greater cell cycle inhibition and cytotoxicity induced
by 2-deoxy-D-glucose in tumor cells treated under hypoxic vs aerobic conditions, Cancer

of
Chemother. Pharmacol. 53(2) (2004) 116–122. doi: 10.1007/s00280-003-0724-7.
[132] B. Dwarakanath, V. Jain, Targeting glucose metabolism with 2-deoxy-D-glucose for
improving cancer therapy, Future Oncol. 5(5) (2009) 581–585.

ro
[133] D. Zhang, J. Li, F. Wang, J. Hu, S. Wang, Y. Sun, 2-Deoxy-D-glucose targeting of glucose
metabolism in cancer cells as a potential therapy, Cancer Lett. 355(2) (2014) 176–183.

-p
[134] Z. Wang, L. Zhang, D. Zhang, R. Sun, Q. Wang, X. Liu, Glycolysis inhibitor 2-deoxy-D-
glucose suppresses carcinogen-induced rat hepatocarcinogenesis by restricting cancer cell
re
metabolism, Mol. Med. Rep. 11(3) (2015) 1917–1924.
[135] R. Vijayaraghavan, D. Kumar, S.N. Dube, et al., Acute toxicity and cardio-respiratory effects
lP

of 2-deoxy-D-glucose: a promising radiosensitizer, Biomed. Environ. Sci. 19(2) (2006) 96–103.


[136] A.L. Simons, I.M. Ahmad, D.M. Mattson, K.J. Dornfeld, D.R. Spitz, 2-Deoxy-D-glucose
combined with cisplatin enhances cytotoxicity via metabolic oxidative stress in human head and
na

neck cancer cells, Cancer Res. 67 (2007) 3364–3370.


[137] D. Zhong, L. Xiong, T. Liu, X. Liu, J. Chen, S.Y. Sun, F.R. Khuri, Y. Zong, Q. Zhou, W.
Zhou, The glycolytic inhibitor 2-deoxyglucose activates multiple prosurvival pathways through
ur

IGF1R, J. Biol. Chem. 284(35) (2009) 23225–23233. doi: 10.1074/jbc.M109.005280.


[138] J.C. Maher, M. Wangpaichitr, N. Savaraj, M. Kurtoglu, T.J. Lampidis, Hypoxia-inducible
Jo

factor-1 confers resistance to the glycolytic inhibitor 2-deoxy-D-glucose, Mol. Cancer Ther. 6(2)
(2007) 732–741. doi: 10.1158/1535-7163.MCT-06-0407.
[139] L. Mediani, F. Gibellini, J. Bertacchini, C. Frasson, R. Bosco, B. Accordi, et al., Reversal of
the glycolytic phenotype of primary effusion lymphoma cells by combined targeting of cellular
metabolism and PI3K/Akt/ mTOR signaling, Oncotarget 7 (2015) 5521–5537.

44
[140] F. Aghaee, J. Pirayesh Islamian, B. Baradaran, Enhanced radiosensitivity and
chemosensitivity of breast cancer cells by 2-deoxy-D-glucose in combination therapy, J. Breast
Cancer 15(2) (2012) 141–147.
[141] B.S. Dwarakanath, D. Singh, A.K. Banerji, et al., Clinical studies for improving radiotherapy
with 2-deoxy-D-glucose: present status and future prospects, J. Cancer Res. Ther. 5(Suppl. 1)
(2009) S21–S26.
[142] B.K. Mohanti, G.K. Rath, N. Anantha, et al., Improving cancer radiotherapy with 2-deoxy-
D-glucose: Phase I/II clinical trials on human cerebral gliomas, Int. J. Radiat. Oncol. Biol. Phys.

of
35(1) (1996) 103–111.
[143] M.C. Coleman, C.R. Asbury, D. Daniels, J. Du, N. Aykin-Burns, B.J. Smith, et al., 2-Deoxy-
d-glucose causes cytotoxicity, oxidative stress, and radiosensitization in pancreatic cancer, Free

ro
Radic. Biol. Med. 44 (2008) 322–331. doi: 10.1016/j.freeradbiomed.2007.08.032
[144] M. Kurtoglu, N. Gao, J. Shang, J.C. Maher, M.A. Lehrman, M. Wangpaichitr, N. Savaraj,

-p
A.N. Lane, T.J. Lampidis, Under normoxia, 2-deoxy-D-glucose elicits cell death in select tumor
types not by inhibition of glycolysis but by interfering with N-linked glycosylation, Mol. Cancer
re
Ther. 6(11) (2007) 3049–3058. doi: 10.1158/1535-7163.MCT-07-0310.
[145] S. Fulda, L. Galluzzi, G. Kroemer, Targeting mitochondria for cancer therapy, Nat. Rev.
lP

Drug Discov. 9 (2010) 447–464.


[146] P. Muley, A. Olinger, H. Tummala, 2-Deoxyglucose induces cell cycle arrest and apoptosis
in colorectal cancer cells independent of its glycolysis inhibition, Nutr. Cancer 67(3) (2015) 514–
na

522.
[147] M. Stein, H. Lin, C. Jeyamohan, et al., Targeting tumor metabolism with 2-deoxyglucose in
patients with castrate-resistant prostate cancer and advanced malignancies, Prostate 70(13) (2010)
ur

1388–1394.
[148] G. Maschek, N. Savaraj, W. Priebe, P. Braunschweiger, K. Hamilton, G.F. Tidmarsh, L.R.
Jo

De Young, T.J. Lampidis, 2-deoxy-D-glucose increases the efficacy of adriamycin and paclitaxel
in human osteosarcoma and non-small cell lung cancers in vivo, Cancer Res. 64(1) (2004) 31–34.
doi: 10.1158/0008-5472.CAN-03-3294.
[149] A. Takemura, X.F. Che, T. Tabuchi, S. Moriya, K. Miyazawa, A. Tomoda, Enhancement of
cytotoxic and pro-apoptotic effects of 2-aminophenoxazine-3-one on the rat hepatocellular
carcinoma cell line dRLh-84, the human hepatocellular carcinoma cell line HepG2, and the rat

45
normal hepatocellular cell line RLN-10 in combination with 2-deoxy-Dglucose, Oncol. Rep. 27(2)
(2012) 347–355.
[150] L.E. Raez, K. Papadopoulos, A.D. Ricart, et al., A phase I dose-escalation trial of 2-deoxy-
Dglucose alone or combined with docetaxel in patients with advanced solid tumors, Cancer
Chemother. Pharmacol. 71(2) (2013) 523–530.
[151] V.K. Kalia, S. Prabhakara, V. Narayanan, Modulation of cellular radiation responses by 2-
deoxy-D-glucose and other glycolytic inhibitors: implications for cancer therapy, J. Cancer Res.
Ther. 5(Suppl. 1) (2009) S57–S60.

of
[152] B. Dwarakanath, V. Jain, Targeting glucose metabolism with 2-deoxy-D-glucose for
improving cancer therapy, Future Oncol. 5(5) (2009) 581–585. doi: 10.2217/fon.09.44.
[153] D. Singh, A.K. Banerji, B.S. Dwarakanath, et al., Optimizing cancer radiotherapy with 2-

ro
deoxy-D-glucose dose escalation studies in patients with glioblastoma multiforme. Strahlenther.
Onkol. 181(8) (2005) 507–514.

-p
[154] V.K. Prasanna, N.K. Venkataramana, B.S. Dwarakanath, V. Santhosh, Differential
responses of tumors and normal brain to the combined treatment of 2-DG and radiation in
re
glioablastoma. J. Cancer Res. Ther. 5(Suppl. 1) (2009) S44–S47.
[155] N.K. Venkataramana, P.K. Venkatesh, B.S. Dwarakanath, S. Vani, Protective effect on
lP

normal brain tissue during a combinational therapy of 2-deoxy-D-glucose and hypofractionated


irradiation in malignant gliomas. Asian. J. Neurosurg. 8(1) (2013) 9–14.
[156] C.R. Kennedy, S.B. Tilkens, H. Guan, J.A. Garner, P.M. Or, A.M. Chan, Diff erential
na

sensitivities of glioblastoma cell lines towards metabolic and signaling pathway inhibitions.
Cancer Lett. 336 (2013) 299–306. doi: 10.1016/j.canlet. 2013.03.020.
[157] G. Cheng, J. Zielonka, B.P. Dranka, et al., Mitochondria-targeted drugs synergize with 2-
ur

deoxyglucose to trigger breast cancer cell death, Cancer Res. 72(10) (2012) 2634–2644.
[158] G. Cheng, J. Zielonka, D. McAllister, S. Tsai, M.B. Dwinell, B. Kalyanaraman, Profiling
Jo

and targeting of cellular bioenergetics: inhibition of pancreatic cancer cell proliferation, Br. J.
Cancer 111(1) (2014) 85–93.
[159] A.S. Belzacq, C. El Hamel, H.L. Vieira, et al., Adenine nucleotide translocator mediates the
mitochondrial membrane permeabilization induced by lonidamine, arsenite and CD437, Oncogene
20 (2001) 7579–7587.

46
[160] K. Nath, L. Guo, B. Nancolas, et al., Mechanism of antineoplastic activity of lonidamine,
Biochim. Biophys. Acta 1866(2) (2016) 151–162.
[161] Guo L, Shestov AA, A.J. Worth, et al., Inhibition of mitochondrial complex II by the
anticancer agent lonidamine, J. Biol. Chem. 291 (2016) 42–57.
[162] S. Di Cosimo, G. Ferretti, P. Papaldo, P. Carlini, A. Fabi, F. Cognetti, Lonidamine: efficacy
and safety in clinical trials for the treatment of solid tumors, Drugs Today 39 (2003) 157–174.
[163] F. DeMarinis, M. Rinaldi, A. Ardizzoni, P. Bruzzi, M.C. Pennucci, L. Portalone, et al., The
role of vindesine and lonidamine in the treatment of elderly patients with advanced non-small cell

of
lung cancer: a phase III randomized fonicap trial, Tumori. 85 (1999) 177–182. doi:
10.1177/03008916990850 0306.
[164] A. Berruti, R. Bitossi, G. Gorzegno, A. Bottini, P. Alquati, A. De Matteis, et al., Time to

ro
progression in metastatic breast cancer patients treated with epirubicin is not improved by the
addition of either cisplatin or lonidamine: final results of a phase III study with a factorial design,

-p
J. Clin. Oncol. 20 (2002) 4150–4159. doi: 10.1200/JCO.2002.08.012.
[165] A. Gadducci, I. Brunetti, M.P. Muttini, A. Fanucchi, F. Dargenio, P.G. Giannessi, et al.,
re
Epidoxorubicin and lonidamine in refractory or recurrent epithelial ovarian cancer, Eur. J. Cancer
30 (1994) 1432–1435. doi: 10.1016/09598049(94)00231-S.
lP

[166] M. De Lena, V. Lorusso, C. Bottalico, M. Brandi, A. De Mitrio, A. Catino, et al., Revertant


and potentiating activity of lonidamine in patients with ovarian cancer previously treated with
platinum, J. Clin. Oncol. 15 (1997) 3208–3213. doi:10.1200/JCO.1997.15.10.3208.
na

[167] S. Oudard, A. Carpentier, E. Banu, et al., Phase II study of lonidamine and diazepam in the
treatment of recurrent glioblastoma multiforme, J. Neurooncol. 63 (2003) 81–86.
[168] D. Amadori, G.L. Frassineti, A. De Matteis, G. Mustacchi, A. Santoro, S. Cariello, et
ur

al., Modulating effect of lonidamine on response to doxorubicin in metastatic breast cancer


patients: results from a multicenter prospective randomized trial, Breast Cancer Res. Treat. 49
Jo

(1998) 209–217.
[169] K. Nath, D.S. Nelson, D.F. Heitjan, D.B. Leeper, R. Zhou, J.D. Glickson, Lonidamine
induces intracellular tumor acidification and ATP depletion in breast, prostate and ovarian cancer
xenografts and potentiates response to doxorubicin, Nmr. Biomed. 28 (2015) 281–290.
[170] G.S. Price, R.L. Page, J.E. Riviere, J.M. Cline, D.E. Thrall, Pharmacokinetics and toxicity
of oral and intravenous lonidamine in dogs, Cancer Chemother. Pharmacol. 38(2) (1996) 129–135.

47
[171] J. Chesney, 6-Phosphofructo-2-Kinase/fructose-2,6-Bisphosphatase and Tumor Cell
Glycolysis, Curr. Opin. Clin. Nutr. Metab. Care 9(5) (2006) 535–539. doi:
10.1097/01.mco.0000241661.15514.fb.
[172] B. Clem, S. Telang, A. Clem, A. Yalcin, J. Meier, A. Simmons, M.A. Rasku, S. Arumugam,
W.L. Dean, J. Eaton, A. Lane, J.O. Trent, J. Chesney, Small-molecule inhibition of 6-
phosphofructo-2-kinase activity suppresses glycolytic flux and tumor growth, Mol. Cancer
Ther. 7(1) (2008) 110–120. doi: 10.1158/1535-7163.MCT-07-0482.
[173] B.F. Clem, J. O'Neal, G. Tapolsky, A.L. Clem, Y. Imbert-Fernandez, D.N. Kerr, et

of
al., Targeting 6-phosphofructo-2-kinase (PFKFB3) as a therapeutic strategy against cancer, Mol.
Cancer Ther. 12 (2013) 1461–1470.
[174] J.W. Locasale, A.R. Grassian, T. Melman, C.A. Lyssiotis, K.R. Mattaini, A.J. Bass, G.

ro
Heffron, C.M. Metallo, T. Muranen, H. Sharfi, et al., Phosphoglycerate dehydrogenase diverts
glycolytic flux and contributes to oncogenesis, Nat. Genet. 43 (2011) 869–874.

-p
[175] R. Possemato, K.M. Marks, Y.D. Shaul, M.E. Pacold, D. Kim, K. Birsoy, S. Sethumadhavan,
H.K. Woo, H.G. Jang, A.K. Jha, et al., Functional genomics reveal that the serine synthesis
re
pathway is essential in breast cancer, Nature 476 (2011) 346–350. doi: 10.1038/nature10350.
[176] Y. Zhou, X. Yi, J.B. Stoffer, et al., The multifunctional protein glyceraldehyde–3– phosphate
lP

dehydrogenase is both regulated and controls colony-stimulating factor-1 messenger RNA stability
in ovarian cancer, Mol. Cancer Res. 6(8) (2008) 1375–1384.
[177] P.J. Thornalley, N. Rabbani, Glyoxalase in tumourigenesis and multidrug resistance, Semin.
na

Cell Dev. Biol. 22(3) (2011) 318–325. doi: 10.1016/j.semcdb.2011.02.006.


[178] E. Persi, M. Duran-Frigola, M. Damaghi, W.R. Roush, P. Aloy, J.L. Cleveland, R.J. Gillies,
E. Ruppin, Systems analysis of intracellular pH vulnerabilities for cancer therapy, Nature
ur

Communications 9 (2018) 2997. doi: 10.1038/s41467-018-05261-x


[179] S. Ganapathy-Kanniappan, Evolution of GAPDH as a druggable target of tumor glycolysis?
Jo

Expert Opinion on Therapeutic Targets, 22(4) (2018) 295–298.


https://doi.org/10.1080/14728222.2018.1449834
[180] M. Wintzell, L. Löfstedt, J. Johansson, A.B. Pedersen, J. Fuxe, M. Shoshan, Repeated
cisplatin treatment can lead to a multiresistant tumor cell population with stem cell features and
sensitivity to 3-Bromopyruvate, Cancer Biol. Ther. 13(14) (2012) 1454–1462.

48
[181] N.J. Rahier, N. Molinier, C. Long, S.K. Deshmukh, A.S. Kate, P. Ranadive, S.A. Verekar,
et al., Anticancer activity of koningic acid and semisynthetic derivatives, Bioorganic & Medicinal
Chemistry 23 (2015) 3712 –3721. https://doi.org/10.1016/j.bmc.2015.04.004.
[182] S. Kumagai, R. Narasaki, K. Hasumi, Glucose dependent active ATP depletion by koningic
acid kills high glycolytic cells, Biochemical and Biophysical Research Communications 365
(2008) 362–368. https://doi.org/10.1016/j.bbrc. 2007.10.199.
[183] M.V. Liberti, Z. Dai, S.E. Wardell, et al., A predictive model for selective targeting of the
Warburg effect through GAPDH inhibition with a natural product, Cell Metab. 3, 26(4) (2017)

of
648659.e8.
[184] Y.H. Ko, B.L. Smith, Y. Wang, M.G. Pomper, D.A. Rini, M.S. Torbenson, J. Hullihen, P.L.
Pedersen, Advanced cancers: eradication in all cases using 3-bromopyruvate therapy to deplete

ro
ATP, Biochem. Biophys. Res. Commun. 324(1) (2004) 269-275. doi: 10.1016/j.bbrc.2004.09.047.
[185] A.P. Pereira da Silva, T. El-Bacha, N. Kyaw, R.S. dos Santos, W.S. da-Silva, F.C. Almeida,

-p
A.T. Da Poian, A. Galina, Inhibition of energy-producing pathways of HepG2 cells by 3-
bromopyruvate, Biochem. J. 417(3) (2009) 717–726. doi: 10.1042/BJ20080805.
re
[186] S. Ganapathy-Kanniappan, J.F. Geschwind, R. Kunjithapatham, et al., Glyceraldehyde-
3phosphate dehydrogenase (GAPDH) is pyruvylated during 3-bromopyruvate mediated cancer
lP

cell death, Anticancer Res. 29(12) (2009) 4909–4918.


[187] S. Ganapathy-Kanniappan, R. Kunjithapatham, J.F. Geschwind. Glyceraldehyde-3-
phosphate dehydrogenase: a promising target for molecular therapy in hepatocellular carcinoma,
na

Oncotarget 3(9) (2012) 940–953.


[188] S. Ganapathy-Kanniappan, R. Kunjithapatham, J.F. Geschwind, Anticancer efficacy of the
metabolic blocker 3-bromopyruvate: specific molecular targeting, Anticancer Res. 33(1) (2013)
ur

13–20.
[189] S. Ganapathy-Kanniappan, J.F. Geschwind, 3-Bromopyruvate induces endoplasmic
Jo

reticulum stress, overcomes autophagy and causes apoptosis in human HCC cell lines, Anticancer
Res. 30 (2010) 923–936.
[190] J.S. Kim, K.J. Ahn, J.A. Kim, et al., Role of reactive oxygen species-mediated mitochondrial
dysregulation in 3-bromopyruvate induced cell death in hepatoma cells: ROS-mediated cell death
by 3-BrPA, J. Bioenerg. Biomembr. 40(6) (2008) 607–618.

49
[191] T. Fan, G. Sun, X. Sun, L. Zhao, R. Zhong, Y. Peng, Tumor energy metabolism and potential
of 3-bromopyruvate as an inhibitor of aerobic glycolysis: implications in tumor treatment, Cancers
11 (2019) 317. doi:10.3390/cancers11030317.
[192] A. Nakano, D. Tsuji, H. Miki, et al., Glycolysis inhibition inactivates ABC transporters to
restore drug sensitivity in malignant cells, PLoS ONE 6(11) (2011) e27222.
[193] L.S. Ihrlund, E. Hernlund, O. Khan, M.C. Shoshan, 3-Bromopyruvate as inhibitor of tumor
cell energy metabolism and chemopotentiator of platinum drugs, Mol. Oncol. 2(1) (2008) 94–101.
[194] X. Cao, G. Jia, T. Zhang, et al., Non-invasive MRI tumor imaging and synergistic anticancer

of
effect of HSP90 inhibitor and glycolysis inhibitor in RIP1-Tag2 transgenic pancreatic tumor
model, Cancer Chemother. Pharmacol. 62(6) (2008) 985–994.
[195] Y.H. Ko, P.L Pedersen, J.F. Geschwind, Glucose catabolism in the rabbit VX2 tumor model

ro
for liver cancer: characterization and targeting hexokinase, Cancer Lett. 173(1) (2001) 83–91.
[196] E. Liapi, J.F. Geschwind, Interventional oncology: new options for interstitial treatments

-p
and intravascular approaches: targeting tumor metabolism via a loco-regional approach: a new
therapy against liver cancer, J. Hepatobiliary Pancreat. Sci. 17(4) (2010) 405–406.
re
[197] S. Ota, J.F. Geschwind, M. Buijs, J.W. Wijlemans, B.K. Kwak, S. Ganapathy-Kanniappan,
Ultrasound-guided direct delivery of 3-Bromopyruvate blocks tumor progression in an orthotopic
lP

mouse model of human pancreatic cancer, Target. Oncol. 8(2) (2013) 145–151.
[198] M. Buijs, J.W. Wijlemans, B.K. Kwak, S. Ota, J.F. Geschwind, Antiglycolytic Therapy
combined with an image-guided minimally invasive delivery strategy for the treatment of breast
na

cancer, J. Vasc. Interv. Radiol. 24(5) (2013) 737–743.


[199] J.F. Geschwind, Y.H. Ko, M.S. Torbenson, C. Magee, P.L. Pedersen, Novel therapy for liver
cancer: direct intra-arterial injection of a potent inhibitor of ATP production, Cancer Res. 62(14)
ur

(2002) 3909–3913.
[200] M. Vali, E. Liapi, J. Kowalski, et al., Intra-arterial therapy with a new potent inhibitor of
Jo

tumor metabolism (3-bromopyruvate): identification of therapeutic dose and method of injection


in an animal model of liver cancer, J. Vasc. Interv. Radiol. 18(1Pt1) (2007) 95–101.
[201] M. Vali, J.A. Vossen, M. Buijs, et al., Targeting of VX2 rabbit liver tumor by selective
delivery of 3-bromopyruvate: a biodistribution and survival study, J. Pharmacol. Exp. Ther. 327(1)
(2008) 32–37.

50
[202] S. Ganapathy-Kanniappan, R. Kunjithapatham, M.S. Torbenson, et al., Human
hepatocellular carcinoma in a mouse model: assessment of tumor response to percutaneous
ablation by using glyceraldehyde-3-phosphate dehydrogenase antagonists, Radiology 262(3)
(2012) 834–845.
[203] J.A. Vossen, M. Buijs, L. Syed, et al., Development of a new orthotopic animal model of
metastatic liver cancer in the rabbit VX2 model: effect on metastases after partial hepatectomy,
intra-arterial treatment with 3-bromopyruvate and chemoembolization, Clin. Exp. Metastasis 25(7)
(2008) 811–817.

of
[204] E. Liapi, J.F. Geschwind, M. Vali, et al., Assessment of tumoricidal efficacy and response
to treatment with 18F-FDG PET/CT after intra-arterial infusion with the antiglycolytic agent 3-
bromopyruvate in the VX2 model of liver tumor, J. Nucl. Med. 52(2) (2011) 225–230.

ro
[205] J. Chapiro, S. Sur, L.J. Savic, S. Ganapathy-Kanniappan, J. Reyes, R. Duran, et al., Systemic
delivery of microencapsulated 3-bromopyruvate for the therapy of pancreatic cancer, Clin. Cancer

-p
Res. 20(24) (2014) 6406–6417. doi: 10.1158/1078-0432.CCR-14-1271.
[206] L. Prescience Labs, PreScience Labs announced that the FDA accepts IND application for
re
novel oncology drug, (2013). http:// presciencelabs.com/prescience-labs-investors/ press.php
[207] H. Feldwisch-Drentrup, Candidate cancer drug suspected after death of three patients at an
lP

alternative medicine clinic, Available online: http://dx.doi.org/10.1126/science.aah7192 (accessed


on September 2019).
[208] Y.H. Ko, H.A. Verhoeven, M.J. Lee, D.J. Corbin, T.J. Vogl, P.L Pedersen, A translational
na

study “case report” on the small molecule “energy blocker” 3-bromopyruvate (3BP) as a potent
anticancer agent: From bench side to bedside, J. Bioenerg. Biomembr. 44 (2012) 163–170.
[209] S.M. El Sayed, W.G. Mohamed, M.A.H. Seddik, A.S.A. Ahmed, A.G. Mahmoudi, W.H.
ur

Amer, et al., Safety and outcome of treatment of metastatic melanoma using 3-bromopyruvate: A
concise literature review and case study, Chin. J. Cancer 33 (2014) 356–364.
Jo

[210] G.A. Spoden, S. Mazurek, D. Morandell, N. Bacher, M.J. Ausserlechner, P. Jansen-Durr, E.


Eigenbrodt, W. Zwerschke, Isotype-specific inhibitors of the glycolytic key regulator pyruvate
kinase subtype M2 moderately decelerate tumor cell proliferation, Int. J. Cancer 123(2) (2008)
312–321.

51
[211] M.G. Vander Heiden, H.R. Christofk, E. Schuman, A.O. Subtelny, H. Sharfi, E.E. Harlow,
J. Xian, L.C. Cantley, Identification of small molecule inhibitors of pyruvate kinase M2, Biochem.
Pharmacol. 79 (2010) 1118–1124. doi: 10.1016/j.bcp.2009.12.003.
[212] D. Anastasiou, G. Poulogiannis, J.M. Asara, M.B. Boxer, J.K. Jiang, M. Shen, G. Bellinger,
A.T. Sasaki, J.W. Locasale, D.S. Auld, et al., Inhibition of pyruvate kinase M2 by reactive oxygen
species contributes to cellular antioxidant responses, Science 334 (2011) 1278–1283. doi:
10.1126/science.1211485.
[213] M.S. Goldberg, P.A. Sharp, Pyruvate kinase M2-specific siRNA induces apoptosis and

of
tumor regression, J. Exp. Med. 209(2) (2012) 217–224. doi: 10.1084/jem.20111487.
[214] W. Luo, H. Hu, R. Chang, J. Zhong, M. Knabel, R. O’Meally, R.N. Cole, A. Pandey, G.L.
Semenza, Pyruvate kinase M2 is a PHD3-stimulated coactivator for hypoxia-inducible

ro
factor1, Cell 145 (2011) 732–744. doi: 10.1016/j.cell.2011.03.054.
[215] J. Liu, N. Wu, L. Ma, M. Liu, G. Liu, Y. Zhang, et al., Oleanolic acid suppresses aerobic

-p
glycolysis in cancer cells by switching pyruvate kinase type M isoforms, PLoS One 9 (2014)
e91606.
re
[216] W.J. Israelsen, T.L. Dayton, S.M. Davidson, B.P. Fiske, A.M. Hosios, G. Bellinger, J. Li,
Y. Yu, M. Sasaki, J.W. Horner, et al., PKM2 isoform-specific deletion reveals a differential
lP

requirement for pyruvate kinase in tumor cells, Cell 155 (2013) 397–409.
[217] M.J. Holness, M.C. Sugden, Regulation of pyruvate dehydrogenase complex activity by
reversible phosphorylation, Biochem. Soc. Trans. 31 (Pt 6) (2003) 1143–1151.
na

[218] S. Kankotia, P.W. Stacpoole. Dichloroacetate and cancer: new home for an orphan drug?
Biochim. Biophys. Acta 1846(2) (2014) 617–629.
[219] K. Birsoy, T. Wang, R. Possemato, et al., MCT1-mediated transport of a toxic molecule is
ur

an effective strategy for targeting glycolytic tumors, Nat. Genet. 45(1) (2013) 104–108.
[220] S. Bonnet, S.L. Archer, J. Allalunis-Turner, et al., A mitochondria-K+ channel axis is
Jo

suppressed in cancer and its normalization promotes apoptosis and inhibits cancer growth, Cancer
Cell 11(1) (2007) 37–51.
[221] S. Vella, M. Conti, R. Tasso, R. Cancedda, A. Pagano, Dichloroacetate inhibits
neuroblastoma growth by specifically acting against malignant undifferentiated cells, Int. J. Cancer
130(7) (2012) 1484–1493.

52
[222] E.M. Dunbar, B.S. Coats, A.L. Shroads, et al., Phase 1 trial of dichloroacetate (DCA) in
adults with recurrent malignant brain tumors, Invest. New Drugs 32 (2014) 452–464.
[223] B. Feuerecker, C. Seidl, S. Pirsig, G. Bruchelt, R. Senekowitsch-Schmidtke, DCA promotes
progression of neuroblastoma tumors in nude mice, Am. J. Cancer Res. 5(2) (2015) 812–820.
[224] Y.C. Shen, D.L. Ou, C. Hsu, et al., Activating oxidative phosphorylation by a pyruvate
dehydrogenase kinase inhibitor overcomes sorafenib resistance of hepatocellular carcinoma, Br. J.
Cancer 108(1) (2013) 72–81.
[225] Y. Dai, X. Xiong, G. Huang, et al., Dichloroacetate enhances adriamycin-induced hepatoma

of
cell toxicity in vitro and in vivo by increasing reactive oxygen species levels, PLoS One 9(4)
(2014) e92962.
[226] Q.S. Chu, R. Sangha, J. Spratlin, et al., A phase I open-labeled, single-arm, dose-escalation,

ro
study of dichloroacetate (DCA) in patients with advanced solid tumors, Invest. New Drugs 33(3)
(2015) 603–610.

-p
[227] P. Miao, S. Sheng, X. Sun, J. Liu, G. Huang, Lactate dehydrogenase A in cancer: a promising
target for diagnosis and therapy, IUBMB Life 65 (11) (2013) 904–910.
re
[228] A. Le, C.R. Cooper, A.M. Gouw, R. Dinavahi, A. Maitra, L.M. Deck, R.E. Royer, D.L.
Vander Jagt, G.L. Semenza, C.V. Dang, Inhibition of lactate dehydrogenase A induces oxidative
lP

stress and inhibits tumor progression, Proc. Natl. Acad. Sci. USA 107(5) (2010) 2037–2042. doi:
10.1073/pnas.0914433107.
[229] M. Zhou, Y. Zhao, Y. Ding, H. Liu, Z. Liu, O. Fodstad, A.I. Riker, S. Kamarajugadda, J.
na

Lu, L.B. Owen, S.P. Ledoux, M. Tan, Warburg effect in chemosensitivity: targeting lactate
dehydrogenase-A re-sensitizes taxol-resistant cancer cells to taxol, Mol. Cancer 9 (2010) 33.
[230] M. Stubbs, P.M. McSheehy, J.R. Griffiths, C.L. Bashford, Causes and consequences of
ur

tumour acidity and implications for treatment, Molecular Medicine Today 6 (2000) 15-19.
[231] A. DeMilito, M.L. Marino, S. Fais, A rationale for the use of proton pump inhibitors as
Jo

antineoplastic agents, Current Pharmaceutical Design 18 (2012) 1395 –1406.


[232] C.H. Chen, C.Z. Lee, Y.C. Lin, L.T. Kao, H.C. Lin, Negative association of proton pump
inhibitors with subsequent development of breast cancer: a nationwide population-based study,
Journal of Clinical Pharmacology 59 (2018) 350 –355. https:// doi.org/10.1002/jcph.1329.
[233] S. Papagerakis, E. Bellile, L.A. Peterson, M. Pliakas, K. Balaskas, S. Selman, et al., Proton
pump inhibitors and histamine 2 blockers are associated with improved overall survival in patients

53
with head and neck squamous carcinoma, Cancer Prevention Research (Philadelphia, Pa.) 7 (2014)
1258-1269. https://doi.org/10.1158/19406207.CAPR-14-0002.
[234] S. Ferrari, F. Perut, F. Fagioli, A.B. Del Prever, C. Meazza, A. Parafioriti, et al., Proton
pump inhibitor chemosensitization in human osteosarcoma: from the bench to the patients’ bed, J.
Transl. Med. 11 (2013) 268. doi:10.1186/1479-5876-11-268.
[235] B.Y. Wang, J. Zhang, J.L. Wang, S. Sun, Z.H. Wang, L.P. Wang, Q.L. Zhang, et al.,
Intermittent high dose proton pump inhibitor enhances the antitumor effects of chemotherapy in
metastatic breast cancer, J. Exp. Clin. Cancer Res. CR. 34 (2015) 85. doi:10.1186/s13046-015-

of
0194-x.
[236] R. Falcone, M. Roberto, C. D’Antonio, A. Romiti, A. Milano, C.E. Onesti, P. Marchetti, S.
Fais, High-doses of proton pumps inhibitors in refractory gastro-intestinal cancer: A case series

ro
and the state of art, Dig. Liver Dis. Off. J. Ital. Soc. Gastroenterol. Ital. Assoc. Study Liver (2016).
doi:10.1016/j.dld.2016.08.126.

-p
[237] I. Marchiq, R. LeFloch, D. Roux, M.P. Simon, J. Pouyssegur, Genetic disruption of
lactate/H+ symporters (MCTs) and their subunit CD147/BASIGIN sensitizes glycolytic tumor
re
cells to phenformin, Cancer Research 75 (2015) 171-180. https://doi. org/10.1158/0008-
5472.CAN-14-2260.
lP

[238] U.E. Martinez-Outschoorn, M. Peiris-Pages, R.G. Pestell, F. Sotgia, M.P. Lisanti, Cancer
metabolism: a therapeutic perspective, Nature Reviews. Clinical Oncology 14 (2017) 11–31.
https://doi.org/10.1038/nrclinonc.2016.60.
na

[239] I. Marchiq, J. Pouyssegur, Hypoxia, cancer metabolism and the therapeutic benefit of
targeting lactate/H(+) symporters, J. Mol. Med. (Berl) 94 (2016) 155–171.
https://doi.org/10.1007/s00109015-1307-x.
ur

[240] C.B. Colen, Y. Shen, F. Ghoddoussi, P. Yu, T.B. Francis, B.J. Koch, M.D. Monterey, M.P.
Galloway, A.E. Sloan, S.P. Mathupala, Metabolic targeting of lactate efflux by malignant glioma
Jo

inhibits invasiveness and induces necrosis: an in vivo study, Neoplasia 13(7) (2011) 620–632.
[241] A.P. Halestrap, The monocarboxylate transporter family– structure and functional
characterization, IUBMB Life 64 (2012)1 –9. https://doi.org/10.1002/iub.573.
[242] C.T. Supuran, J.Y. Winum, Carbonic anhydrase IX inhibitors in cancer therapy: an update,
Future Medicinal Chemistry 7 (2015) 1407 –1414. https://doi.org/10.4155/fmc.15.71.

54
[243] Y. Lou, P.C. McDonald, A. Oloumi, S. Chia, C. Ostlund, A. Ahmadi, A. Kyle, et al.,
Targeting tumor hypoxia: suppression of breast tumor growth and metastasis by novel carbonic
anhydrase IX inhibitors, Cancer Research 71 (2011) 3364 –3376. https://doi.org/ 10.1158/0008-
5472.CAN-10-4261.
[244] P. Bailey, D.K. Chang, K. Nones, A.L. Johns, A.M. Patch, M.C. Gingras, et al., Genomic
analyses identify molecular subtypes of pancreatic cancer, Nature 531 (2016) 47–52. doi:
10.1038/nature16965.
[245] H. Ying, A.C. Kimmelman, C.A. Lyssiotis, S. Hua, G.C. Chu, E. Fletcher-Sananikone, et

of
al., Oncogenic kras maintains pancreatic tumors through regulation of anabolic glucose
metabolism, Cell 149 (2012) 656–670. doi: 10.1016/j.cell.2012.01.058.
[246] G.Y. Liou, H. Döppler, K.E. DelGiorno, L. Zhang, M. Leitges, H.C. Crawford, et al., Mutant

ro
KRas-induced mitochondrial oxidative stress in acinar cells upregulates EGFR signaling to drive
formation of pancreatic precancerous lesions, Cell Rep. 14 (2016) 2325–2336. doi:
10.1016/j.celrep.2016.02.029.

-p
[247] C. Xie, Y. Li, L. Li, X. Fan, Y.W. Wang, C.L. Wei, et al., Identification of a new potent
re
inhibitor targeting KRAS in non-small cell lung cancer cells, Front. Pharmacol. 8 (2017) 823. doi:
10.3389/fphar.2017. 00823.
lP

[248] M. Zeng, J. Lu, J. Li, F. Feru, C. Quan, T.W. Gero, et al., Potent and selective covalent
quinazoline inhibitors of KRAS G12C, Cell Chem. Biol. 24 (2017) 1005–1016.e3. doi:
10.1016/j.chembiol.2017.06.017.
na

[249] H.L. Kindler, K. Wroblewski, J.A. Wallace, M.J. Hall, G. Locker, S. Nattam, et al.,
Gemcitabine plus sorafenib in patients with advanced pancreatic cancer: a phase II trial of the
University of Chicago Phase II Consortium, Investig. New Drugs 30 (2012) 382–386.
ur

doi:10.1007/s10637-010-9526-z.
[250] J.R. Infante, B.G. Somer, J.O. Park, C.P. Li, M.E. Scheulen, S.M. Kasubhai, et al., A
Jo

randomised, double-blind, placebo-controlled trial of trametinib, an oral MEK inhibitor, in


combination with gemcitabine for patients with untreated metastatic adenocarcinoma of the
pancreas, Eur. J. Cancer 50 (2014) 2072–2081. doi: 10.1016/j.ejca.2014.04.024.
[251] V. Chung, S. McDonough, P.A. Philip, D. Cardin, A. Wang-Gillam, L. Hui, et al., Eff ect of
selumetinib and MK-2206 vsoxaliplatin and fluorouracil in patients with metastatic pancreatic
cancer after prior therapy, JAMA Oncol. 3 (2017) 516–522. doi:10.1001/jamaoncol.2016.5383.

55
[252] T.L. He, Y.J. Zhang, H. Jiang, X.H. Li, H. Zhu, K.L. Zheng, The c-Myc–LDHA axis
positively regulates aerobic glycolysis and promotes tumor progression in pancreatic cancer, Med.
Oncol. 32 (2015) 187. doi:10.1007/s12032-0150633-8.
[253] P.K. Mazur, A. Herner, S.S. Mello, M. Wirth, S. Hausmann, F.J. Sánchez-Rivera, et al.,
Combined inhibition of BET family proteins and histone deacetylases as a potential epigenetics-
based therapy for pancreatic ductal adenocarcinoma, Nat. Med. 21 (2015) 1163–1171.
doi:10.1038/nm.3952.
[254] P.L. Garcia, A.L. Miller, K.M. Kreitzburg, L.N. Council, T.L. Gamblin, J.D. Christein, et

of
al., The BET bromodomain inhibitor JQ1 suppresses growth of pancreatic ductal adenocarcinoma
in patient-derived xenograft models, Oncogene 35 (2016) 833–845. doi:10.1038/onc.2015.126.
[255] P. Sancho, E. Burgos-Ramos, A. Tavera, T.B. Kheir, P. Jagust, M. Schoenhals, et al.,

ro
MYC/PGC-1α balance determines the metabolic phenotype and plasticity of pancreatic cancer
stem cells, Cell Metab. 22 (2015) 590–605. doi:10.1016/ j.cmet.2015.08.015.

-p
[256] G.M. Nitulescu, D. Margina, P. Juzenas, Q. Peng, O.T. Olaru, E. Saloustros, C. Fenga, D.Α.
Spandidos, M. Libra, A.M. Tsatsakis, Akt inhibitors in cancer treatment: The long journey from
re
drug discovery to clinical use (Review), International Journal of Oncology 48 (2016) 869-885.
[257] R. Dienstmann, J. Rodon, V. Serra, J. Tabernero, Picking the point of inhibition: A
lP

comparative review of PI3K/AKT/mTOR pathway inhibitors, Mol. Cancer Ther. 13 (2014)


1021‑ 1031.
[258] A. Carnero, The PKB/AKT pathway in cancer, Curr. Pharm. Des. 16 (2010) 34‑ 44.
na

[259] J. Chan, M. Kulke, Targeting the mTOR signaling pathway in neuroendocrine tumors, Curr.
Treat. Opt. Oncol. 15 (2014) 365–379.
[260] A.C. deMelo, E. Paulino, A.H. Garces, A review of mTOR pathway inhibitors in
ur

gynecologic cancer, Oxid. Med. Cell Longev. 2017 (2017) 4809751.


[261] V. Dimitrova, A. Arcaro, Targeting the PI3K/AKT/mTOR signaling pathway in
Jo

medulloblastoma, Curr. Mol. Med. 15 (2015) 82–93.


[262] X. Li, C. Wu, N. Chen, H. Gu, A. Yen, L. Cao, E. Wang, L. Wang, PI3K/Akt/mTOR
signaling pathway and targeted therapy for glioblastoma, Oncotarget 7 (2016) 33440–33450.
[263] K. Hu, H.B. Dai, Z.L. Qiu, mTOR signaling in osteosarcoma: Oncogenesis and therapeutic
aspects (Review), Oncol. Rep. 36 (2016) 1219–1225.

56
[264] S. Dinner, L.C. Platanias, Targeting the mTOR Pathway in Leukemia, J. Cell Biochem. 117
(2016) 1745–1752.
[265] J.S. Blachly, R.A. Baiocchi, Targeting PI3-kinase (PI3K), AKT and mTOR axis in
lymphoma, Br. J. Haematol. 167 (2014) 19–32.
[266] T. Tian, X. Li, J. Zhang, mTOR signaling in cancer and mTOR inhibitors in solid tumor
targeting therapy, Int. J. Mol. Sci. 20 (2019) 755. doi:10.3390/ijms20030755.
[267] M.C. Mendoza, E.E. Er, J. Blenis, The Ras-ERK and PI3K-mTOR pathways: Cross-talk and
compensation, Trends Biochem. Sci. 36 (2011) 320–328.

of
[268] Y.J. Zhang, Y. Duan, X.F. Zheng, Targeting the mTOR kinase domain: The second
generation of mTOR inhibitors, Drug Discov. Today 16 (2011) 325–331.
[269] G.Q. Chen, C.F. Tang, X.K. Shi, C.Y. Lin, S. Fatima, X.H. Pan, et al., Halofuginone inhibits

ro
colorectal cancer growth through suppression of Akt/mTORC1 signaling and glucose metabolism,
Oncotarget 6 (2015) 24148-24162.

-p
[270] S.A. Hawley, A.E. Gadalla, G.S. Olsen, D.G. Hardie, The antidiabetic drug metformin
activates the AMP-activated protein kinase cascade via an adenine nucleotide-independent
re
mechanism, Diabetes 51 (2002) 2420–2425.
[271] J. Kasznicki, A. Sliwinska, J. Drzewoski, Metformin in cancer prevention and therapy, Ann.
lP

Transl. Med. 2 (2014) 57.


[272] H.R. Bridges, A.J.Y. Jones, M.N. Pollak, J. Hirst, Effects of metformin and other biguanides
on oxidative phosphorylation in mitochondria, Biochem. J. 462 (2014) 475–487.
na

[273] J.H. Kim, K.J. Lee, Y. Seo, J.H. Kwon, J.P. Yoon, J.Y. Kang, et al., Eff ects of metformin
on colorectal cancer stem cells depend on alterations in glutamine metabolism, Sci. Rep. 8 (2018)
409. doi:10.1038/s41598-017-18762-4.
ur

[274] C. Li, V.W. Liu, D.W. Chan, K.M. Yao, H.Y. Ngan, LY294002 and metformin
cooperatively enhance the inhibition of growth and the induction of apoptosis of ovarian cancer
Jo

cells, Int. J. Gynecol. Cancer 22 (2012) 15–22. doi: 10.1097/IGC.0b013e3182322834.


[275] G. Bergers, D. Hanahan, Modes of resistance to anti-angiogenic therapy, Nat. Rev. Cancer
8 (2008) 592–603.
[276] K. Hosaka, Y. Yang, T. Seki, M. Nakamura, P. Andersson, P. Rouhi, et al., Tumour PDGF-
BB expression levels determine dual effects of anti-PDGF drugs on vascular remodelling and
metastasis, Nat. Commun. 4 (2013) 2129.

57
[277] C.G. Willett, Y. Boucher, E. diTomaso, D.G. Duda, L.L. Munn, R.T. Tong, et al., Direct
evidence that the VEGF-specific antibody bevacizumab has antivascular effects in human rectal
cancer, Nat. Med. 10 (2004) 145–147.
[278] A.C. Lockhart, M.L. Rothenberg, J. Dupont, W. Cooper, P. Chevalier, L. Sternas, et al.,
Phase I study of intravenous vascular endothelial growth factor trap, aflibercept, in patients with
advanced solid tumors, J. Clin. Oncol. 28 (2010) 207-14.
[279] K.J. Gotink, H.M. Verheul, Anti-angiogenic tyrosine kinase inhibitors: what is their
mechanism of action? Angiogenesis 13 (2010) 1–14.

of
[280] A. Grothey, E. VanCutsem, A. Sobrero, S. Siena, A. Falcone, M. Ychou, et al., Regorafenib
monotherapy for previously treated metastatic colorectal cancer (CORRECT): an international,
multicentre, randomised, placebo-controlled, phase 3 trial, Lancet 381 (2013) 303–312.

ro
[281] D. Zhang, E.M. Hedlund, S. Lim, F. Chen, Y. Zhang, B. Sun, et al., Antiangiogenic agents
significantly improve survival in tumor-bearing mice by increasing tolerance to chemotherapy-

-p
induced toxicity, Proc. Natl. Acad. Sci. USA 108 (2011) 4117–4122.
[282] Y. Yang, Y. Zhang, Z. Cao, H. Ji, X. Yang, H. Iwamoto, et al., Anti-VEGF- and anti-VEGF
re
receptor-induced vascular alteration in mouse healthy tissues, Proc. Natl. Acad. Sci. USA 110
(2013) 12018–23.
lP
na
ur
Jo

58

You might also like