GET THE APP

Cancer Metabolism and Its Therapeutic Implications
Journal of Cell Science & Therapy

Journal of Cell Science & Therapy
Open Access

ISSN: 2157-7013

+44 1300 500008

Review Article - (2013) Volume 4, Issue 2

Cancer Metabolism and Its Therapeutic Implications

Weiqin Lu1*, Craig D Logsdon2 and James L Abbruzzese1*
1Department of GI Medical Oncology, The University of Texas MD Anderson Cancer Center, Houston, Texas, USA, E-mail: clogsdon@mdanderson.org
2Department of Cancer Biology, The University of Texas MD Anderson Cancer Center, Houston, Texas, USA, E-mail: clogsdon@mdanderson.org
*Corresponding Author(s): Weiqin Lu, Department of Gastrointestinal Medical Oncology, The University of Texas MD Anderson Cancer Center, Box 0426, 1515 Holcombe Boulevard Houston, TX 77030, USA, Tel: (713)745-2093 or (713)792-2828, Fax: (713)404-3333 Email:
James L Abbruzzese, Department of Gastrointestinal Medical Oncology, The University of Texas MD Anderson Cancer Center, Box 0426, 1515 Holcombe Boulevard Houston, TX 77030, USA, Tel: (713)745-2093 or (713)792-2828, Fax: (713)404-3333 Email:

Abstract

The goal of cancer therapy is to selectively kill cancer cells without significant adverse toxicity to normal cells. Tremendous progress has been made to understand the molecular mechanisms of cancer and we have realized that cancer cells have remarkable heterogeneity and adaptability, making them difficult to be targeted therapeutically. With the advanced developments in molecular biology and high throughput ‘omics’ technologies, metabolic alteration is revealed as the common feature of human cancer. The activation of oncogenes and/or inactivation of tumor suppressors are found to be important metabolic switch regulators. Some metabolic enzymes in cancer have gained non-enzymatic functions to facilitate tumor growth and adaptation. Metabolic reprogramming of cancer cells not only promotes survival and proliferation but also tolerates stress and induces drug resistance. Therefore, cancer metabolism might be a novel target for cancer therapy. This article reviews current understanding of cancer metabolism with major focus on glucose and glutamine metabolisms and their potential therapeutic implications.

Keywords: Pentose phosphate pathway; Hexokinase II; Reactive oxygen species

Introduction

Unlike normal cells which mainly metabolize glucose by using mitochondrial oxidative phosphorylation to generate ATP, cancer cells exhibit profound metabolic alterations by choosing aerobic glycolysis to convert glucose to lactate regardless of the presence of adequate oxygen [1]. Further studies have identified that cancer cells also have increased glutamine metabolism [2-4]. Such widespread metabolic alterations allow cancer cells to survive in adverse tumor microenvironment and under oxidative stress, to sustain higher proliferation rate through anabolic processes, to gain invasive and metastatic ability, and to resist drug treatment [5,6]. Oncogene activation and tumor suppressor inactivation are hallmarks in human cancer. Recent studies provide a plethora of evidence supporting the mechanistic link between oncogene activation, loss of tumor suppressor function, and the metabolic reprogramming of glucose and glutamine utilization. Metabolic enzymes, by expressing specific isoforms or mutations, regulate genetic and epigenetic events to favor cancer development and inhibit apoptosis. Since cancer cells have these metabolic dependencies that distinguish them from their normal counterparts, targeted inhibition of metabolic dependencies is considered a promising cancer therapeutic strategy.

Targeting Glucose Metabolism in Cancer

As a major fuel, glucose uptake in cancer cells is highly increased. This is due to increased expression of Na+ dependent Glucose Specific Transporters (GLUTs) and various glycolytic enzymes, including isoform specific Hexokinase II (HKII) and Pyruvate Kinase Muscle Isozyme 2 (PKM2). As illustrated in (Figure 1), glucose not only can be metabolized to pyruvate through the canonical glycolytic pathway, it also can branch out, generating ribose 5-phosphate for the synthesis of nucleotides and the reducing equivalent NADPH to combat Reactive Oxygen Species (ROS) induced stress and for the reductive biosynthesis of lipids and fats through the Pentose Phosphate Pathway (PPP) [7,8]. Glucose can also be metabolized to triose sugars, which are then converted to serine or glycerol for the synthesis of amino acids or lipids.

cell-science-therapy-glutamine-metabolism

Figure 1: Overview of glycolysis: the citric acid cycle: and glutamine metabolism in cancer
Cancer cells have profound metabolic alterations in glucose metabolism: the citric acid cycle: and glutamine metabolism. By over expressing glycolytic enzymes: such as isoform specific glycolytic enzymes GLUT1: HKII: PKM2: and LDHA: cancer cells have gained increased glucose uptake: increased lactate generation: increased anabolic processes: and increased ability to combat ROS stress. By expressing mutant citric acid cycle enzymes: such as mutant SDH: FH: and IDH1/2: cancer cells have gained the ability to reprogram metabolism through genetic and epigenetic alterations. By increasing canonical and non-canonical glutamine pathways: cancer cells have increased the utilization of glutamine for the anabolic processes and to adapt to ROS stress.

To accommodate rapid growth, cancer cells have diverse requirements. They need to replicate their genome, synthesize amino acids, fatty acids, membrane lipids, and package them into daughter cells. In addition, they need to deal with ROS stress and adverse microenvironments, such as hypoxia and acidic conditions generated by exported lactate from tumor cells. Glucose, as a key nutritional component, can fulfill these diverse cancer cell requirements through rearranging the established glycolytic pathways. There are many review articles published recently describing the pathway in great detail [6,9-12]. Here we highlight several key glycolytic enzymes and their potential roles in cancer therapy.

Hexokinase II (KII)

HKII is a key glycolytic enzyme that catalyzes the conversion of glucose to Glucose-6-Phosphate (G6P) and is transcriptionally enhanced in many cancers [13]. G6P generated from this reaction lies at the start point of glycolytic and the PPP pathways, thus having many possible fates. G6P may also be converted to glycogen for storage. Besides catalyzing glucose breakdown, HKII can also bind to mitochondria and interact with the mitochondrial permeability transition pore Voltage-Dependent Anion Channel (VDAC) to serve as a cell survival molecule by preventing BAX/BAK oligomerization on the mitochondrial membrane and thus suppressing apoptosis and preserve mitochondrial integrity [14,15]. HKII is also subject to epigenetic regulation. Adult normal brain with low HKII expression has methylation of CpG islands within intron 1 of HKII. In contrast, developing fetal brain and glioblastoma tissue expressing HKII have significantly lower level of methylation of CpG island [16]. By using conditional knockout of HKII in oncogenic Kras-driven lung cancer and ErbB2-driven breast cancer mouse models, HKII is found to be required for tumor initiation and maintenance despite continued HK1 expression. HKII deletion suppresses glucose-derived ribonucleotides and impaires glutamine-derived carbon utilization in lung cancer cells [17]. Systemic HKII deletion is therapeutic in mice bearing lung tumors without adverse physiological consequences, providing a rationale for the development of small-molecule inhibitors of HKII for cancer therapy. 2-deoxyglucose, the pyruvate analog 3-bromopyruvate, and lonidamine are all HK inhibitors, and have been explored for cancer therapy, either used as monotherapy or combined with other anticancer agents [18-21]. Metformin is recently reported to inhibit HK II through an allosteric modification of its molecular structure and blocking the synthesis of G6P in a human non-small cell lung cancer cell line. Metformin treatment causes release of HKII from the outer membrane of mitochondria, thus leading to the activation of apoptotic signals [22]. A study has shown that cerulenin, a fungal metabolite, disrupts the interaction between AIF and HKII, leading to the release of AIF from the mitochondria and eventual cell death [23].

Pyruvate Kinase Muscle Isozyme 2 (PKM2)

Pyruvate kinase, a key glycolytic rate-limiting enzyme, catalyzes the conversion of Phosphoenolpyruvate (PEP) to pyruvate. It has four isozymes (type L, R, M1, and M2). PKM1 and PKM2, resulting from mutually exclusive alternative splicing of the PKM pre-mRNA, have distinct functions in glucose metabolism [24]. By re-expressing the embryonic PKM2, proliferating cancer cells acquire the ability to use glucose for anabolic processes and this metabolic phenotype provides a selective growth advantage for tumor cells, whereas the adult isoform PKM1 promotes oxidative phosphorylation [24-26]. Replacing PKM2 with PKM1 leads to reduced lactate production, increased oxygen consumption, and reduced tumor formation in nude mouse xenografts [24]. Selective expression of PKM2 over PKM1 is due to oncogenic transcription factor c-Myc mediated transcriptional upregulation of splicing factors, such as a polypyrimidine tract binding protein (PTB, known as hnRNPI), hnRNPA1, and hnRNPA2 [27]. PKM2 may exist in a tetrameric or dimeric form. The tetrameric form of PKM2 has a high affinity to its substrate PEP and favors the formation of pyruvate and ATP. The dimeric form of PKM2 has less activity and drives glucose through the route of aerobic glycolysis and facilitates anabolic processes [28]. The dynamic conversion between the tetrameric form and dimeric form of PKM2 may provide an important layer of metabolic regulation and may be exploited for cancer therapy. PKM2, but not PKM1, can selectively bind to tyrosine-phosphorylated peptides and results in release of the allosteric activator fructose-1, 6-bisphosphate. This induces increased glycolysis and diverts glucose metabolites from energy production through the non-oxidative PPP pathway to anabolic processes [29]. TLN-232 (or CAP-232), an inhibitor of the glycolytic enzyme PKM2, is currently being explored in Phase II clinical trials for metastatic melanoma and renal cell carcinoma [30].

The PKM2 isoform is selected through alternative splicing over PKM1 by cancer cells not only to facilitate shunting of glucose to the anabolic processes as mentioned above, but also to cause chromatin remodeling through its translocation to the nucleus where it serves as a protein kinase. EGFR activation in glioblastomas results in PKM2 phosphorylation and translocation to nucleus, where it phosphorylates Histone 3 at Threonine 11 (H3-T11) [31]. This histone modification results in the dissociation of Histone Deacetylase 3 (HDAC3) from the promoters of cyclin D1 and c-Myc. Nuclear PKM2 then forms a complex with c-Src phosphorylated ß-catenin at the promoter regions of cyclin D1 and c-Myc, thereby leading to acetylation of histone H3 at lysine 9 and promoting the expression of CCND1 (gene encoding cyclin D1) and Myc [31,32]. This is an example of how a cancer specific metabolic enzyme serves to regulate oncogene expression and support tumorigenesis through its non-enzymatic function. Nuclear PKM2 is also reported to phosphorylate STAT3, which then activates MEK5 transcription [28].

Lactate dehydrogenase A (LDHA)

LDHA, catalyzing the conversion of pyruvate and NADH to lactate and electron acceptor NAD+, is important to maintain continued glycolysis. Cancer cells produce more lactate and have higher LDHA activity than normal cells [33]. Lactate generated from LDHA can exit cancer cells via the Monocarboxylate Transporters (MCTs) and cause acidification of the tumor microenvironment, thus facilitating cancer cell invasion [34]. Lactate is reported to enhance HIF1α accumulation regardless of hypoxia and lactate induced acidosis often precedes angiogenesis [35]. LDHA knockdown in tumor cells stimulates mitochondrial oxidative phosphorylation, decreases cell proliferation under hypoxia, and suppresses tumorigenicity [36]. LDHA is reported to be acetylated at lysine-5 and this modified LDHA can be recognized by the HSC70 chaperone, leading to the lysosome mediated LDHA degradation [37]. Importantly, lysine-5 acetylation of LDHA is reduced in human pancreatic cancer, which provides a mechanism of LDHA upregulation and a therapeutic window for pancreatic cancer [38]. The LDHA inhibitor FX11 is found to reduce ATP levels, induce ROS mediated cell death, and inhibit lymphoma and pancreatic tumor progression [39]. By using Magnetic Resonance Spectroscopy (MRS) to real-time monitor metabolic fluxes in murine xenografts of human lymphoma, FX11 inhibition of LDHA suppresses the conversion of hyperpolarized 13C-pyruvate to lactate, providing a promising technique for clinical development of imaging biomarkers to monitor metabolic pharmacodynamics [40].

Pentose phosphate pathway (PPP)

The PPP, one of the major branches of the glycolytic pathway, is essential for cancer cell survival because this pathway can generate pentose-5-phosphate for the synthesis of nucleotides and the reducing equivalent NADPH needed for cell proliferation and scavenging of ROS. Both the oxidative branch, including G6PD and the 6-phosphogluconate dehydrogenase enzymes, and the non-oxidative branch, including transketolase-like-1 and transaldolases, of the PPP have been reported to be activated during carcinogenesis [41-43]. TAp73 is a structural homologue of tumor suppressor p53, but is frequently overexpressed in human tumors. By stimulating the expression of G6PD, TAp73 increases PPP flux and directs glucose to the production of ribose and NADPH, for the synthesis of macromolecules and detoxification of ROS. The growth defect of TAp73-deficient cells can be rescued by either G6PD over-expression or the presence of nucleosides plus an ROS scavenger, suggesting the critical role of TAp73 in regulating the PPP metabolism to support oncogenic cell growth [44]. Targeted inhibition of both branches of the PPP is effective in preventing tumor metastasis [45]. Oxythiamine is a thiamine antagonist and therefore inhibits transketolase and pyruvate dehydrogenase that use Thiamine Pyrophosphate (TPP) as cofactor for their enzymatic activities. Oxythiamine has demonstrated anticancer activity both in vitro and in vivo [46,47]. 6-Aminonicotinamide (6-AN), another PPP inhibitor targeting G6PD, exhibits anticancer activity, causes oxidative stress, and sensitizes cells to anticancer agents [48,49].

Other strategies

A large body of evidence suggests that tumor cells have enhanced transcription of genes encoding glycolytic enzymes when compared with non-tumorigenic cells. Compounds have been developed that can directly inhibit almost every enzyme of the glycolytic pathway (Table 1). In addition to the enzymes and their inhibitors discussed above, an inhibitor of Phosphoglycerate Mutase 1 (PGAM1), called MJE3, is identified as the most potent anti-breast cancer drug from chemical screening results [50]. The Monocarboxylate Transporters (MCTs) function to export lactate out of the cells. These transporters are upregulated in various human tumor cells and can be exploited for targeted inhibition as well [51]. Disruption of glucose transporters could be another strategy to inhibit glucose metabolism. Small metabolic inhibitors such as Fasentin, Phloretin, and WZB117 have been developed to inhibit GLUT1 [52-54]. Dichloroacetate (DCA), a small-molecule inhibitor of mitochondrial pyruvate dehydrogenase kinase involved in diverting pyruvate to lactate production, is being tested in patients with metastatic solid tumors and glioma and is shown to decrease lactate levels, activate pyruvate dehydrogenase activity, and induce apoptosis in a variety of human cancer cells [55,56]. DCA also enhances the effectiveness of hypoxia-specific cytotoxic chemotherapy in solid tumors [57]. Diclofenac, a nonsteroidal antiinflammatory drug, significantly decreases GLUT1, LDHA, and MCT1 gene expression resulting in a decrease in glucose uptake and lactate secretion, holding potential as a clinically applicable glycolysis inhibitor for cancer therapies [58,59].

Targeted  metabolism Metabolic targets  Agents References
Glycolysis GLUT1
HKII
PGAM
PKM2
LDHA
Transketolase     G6PD 
MCT1
Fasentin:  WZB117: Diclofenac
2-deoxyglucose: 3-bromopyruvate: lonidamine: metformin: cerulenin
MJE3
TLN-232 (or CAP-232)
FX11: Diclofenac
Oxythiamine
6-Aminonicotinamide (6-AN)
Diclofenac
[52-54]
[18-23]
[50]
[30]
[39-40: 58-59]
[46-47]       
[48-49]
[58-59]
Glutamine Glutaminase Glutamine ASCT2 (SLC1A5) SLC7A5/SLC3A2 Compand 968
Phenylacetate
L-γ-glutamyl-p-nitroanilide (GPNA)
2-aminobicyclo-(2:2:1) heptanecarboxylic acid(BCH)
[3]
[65-66]
[82]
[83]
Citric acid Master regulators PDK Mutant IDH HIF1α AKT Dichloroacetate
Antibody MsMab-1: AGI-5198:
KC7F2: RX-0047
MK-2206                                                               
[55-57]
[145-146]
[76-77] [86]
GLUT1: Glucose Transporter 1; HKII: Hexokinase II; PGAM1: Phosphoglycerate Mutase 1; PKM2: Pyruvate Kinase Muscle Isozyme 2; LDHA: Lactate Dehydrogenase A; G6PD: Glucose-6-Phosphate Dehydrogenase; MCT1: Monocarboxylate Transporter 1; ASCT2: Solute Carrier Family 1 (Neutral Amino Acid Transporter): Member 5 (also called SLC1A5); SLC7A5: Solute Carrier Family 7 (Amino Acid Transporter Light Chain: l system): Member 5; SLC3A2: Solute Carrier Family 3 (Amino Acid Transporter Heavy Chain) member 2; PDK: Pyruvate Dehydrogenase Kinase; IDH: Isocitrate Dehydrogenase; HIF1: Hypoxia Inducible Factor 1; Alpha Subunit (basic helix-loop-helix transcription factor); AKT: v-akt Murine Thymoma Viral Oncogene Homolog

Table 1: Targeting glucose/glutamine metabolism for cancer therapy.

Targeting Glutamine Metabolism in Cancer

To meet the high proliferation rates, cancer cells require large quantities of nucleotides, proteins, and lipids. However, for the synthesis of key cellular elements, glucose can only provide a carbon source. Since it carries two nitrogens, glutamine is an effective source of nitrogen for the biosynthesis of nucleic acids and some amino acids, such as proline, arginine and asparagine. Glutamine is the most abundant amino acid in the serum [60,61]. Glutamine is also a good source of carbon as it can be deaminated to glutamate through a phosphate-dependent glutaminase in the mitochondrion and further transaminated to α-ketoglutarate that enters the citric acid cycle. In addition, glutamate generated from deamination of glutamine is important for glutathione synthesis allowing cancer cells to survive ROS stress. High level of Glutaminase 1 (GLS1) is a common feature of cancer and rapidly dividing cells. Studies find that deficiency in glutamine but not glucose induces Myc-dependent apoptosis in human cells [62]. Further study identifies that c-Myc enhances mitochondrial GLS1 expression and glutamine metabolism through suppression of miR-23a/b [2]. These studies provide a molecular link between oncogene activation and glutamine metabolism. Another oncogene, Rho GTPase, is identified to promote glutamine metabolism through NF-κB mediated upregulation of GLS1 activity, contributing to malignant transformation [3]. Interestingly, the tumor suppressor p53 is found to regulate glutamine metabolism by transcriptionally activating liver-type mitochondrial phosphate-activated Glutaminase 2 (GLS2) to combat oxidative stress and suppress tumor growth [63].

Targeted inhibition of GLS1 by a tetrahydrobenzo derivative (small compound 968) decreases malignant transformation and tumor cell growth [3]. Inhibition of glutaminase expression by antisense mRNA decreases growth and tumorigenicity of tumor cells [64]. Phenylacetate, a drug that conjugates with glutamine to form renally excretable phenylacetylglutamine to reduce plasma glutamine levels, was in Phase II clinical testing in patient with glioma, and medulloblastoma [65,66]. Due to its relatively nontoxic features, phenylacetate is clinically appealing and early results have shown that phenylacetate can suppress tumor growth and induce differentiation of some solid tumors [65,66].

Oncogenes, Tumor Suppressors, and Cancer Metabolism

We now know that oncogene activation and loss of tumor suppressor functions have mechanistic links to cancer metabolism. One mechanism linking genetic alterations with cancer cell metabolism is through altered expression in nuclear transcription factors. For example, HIF1 transcriptionally upregulates the expression of glycolytic related genes, such as GLUT1, HKII, Phosphofructokinase (PFK), aldolase A, Phosphoglycerate Kinase (PGK), enolase 1, PKM2, LDHA, and MCT4 [67,68]. The increased glycolytic gene expression due to HIF1α stabilization leads to enhanced glycolysis [69]. HIF1α transcriptionally upregulates Pyruvate Dehydrogenase Kinase (PDK) to inhibit pyruvate entry into the citric acid cycle through phosphorylation and inactivation of pyruvate dehydrogenase [70,71]. In addition, metabolic enzyme mutations can also lead to HIF1α stabilization. Patients with inherited mutations in Succinate Dehydrogenase (SDH) and Fumarate Hydratase (FH) develop hereditary paraganglioma or skin leiomyomata and papillary renal cell cancer, respectively [72,73]. This is due to the accumulation of succinate or fumarate, resulting in the inhibition of HIF1α prolyl hydroxylases in the cytosol, leading to stabilization and activation of HIF1α [74,75]. HIF1α inhibitors are being developed for therapeutic potential. KC7F2, a lead compound with a central structure of cystamine, has been identified as a novel HIF1 inhibitor. It markedly inhibits HIF-mediated transcription in cancer cells, and exhibits enhanced cytotoxicity under hypoxia [76]. Another small molecule inhibitor, called RX-0047, inhibits the formation of human lung metastasis in xenograft mouse models and reduces tumor size in flank models [77].

Deregulated expression of the c-Myc gene is a common feature of human malignancies [78]. c-Myc expression enhances aerobic glycolysis by directly upregulating the expression of glycolytic genes such as HK II, PFK, GAPDH, enolase A, LDHA, and Glucose Transporter 1 (GLUT1) independent of hypoxia [79]. Substantial experimental data support that the pivotal role of c-Myc in glutamine metabolism. c-Myc upregulates glutamine metabolism through repression expression of miR-23a/b, which both target the 3’-UTR sequence to down-regulate glutaminase translation [2,80]. ASCT2 (also called SLC1A5) and SLC7A1, both involved in glutamine transport, are direct c-Myc target genes [80,81]. L-γ-glutamyl-p-nitroanilide (GPNA), an ASCT2 inhibitor, can inhibit glutamine uptake and inhibit glutamine-dependent mTOR activation in vitro [82]. 2-aminobicyclo-(2,2,1) heptanecarboxylic acid (BCH), an inhibitor of SLC7A5/SLC3A2, blocks the glutamine-dependent activation of mTORC1 and induces autophagy [83].

AKT activation can increase GLUT localization to the cell membrane to facilitate glucose uptake. It can also phosphorylate and activate PFK2 to generate fructose-2,6-bisphosphate for the activation of PFK1, and increase HKII association to the mitochondrial membrane promoting glycolysis [84]. A recent study has further confirmed that Akt phosphorylates HKII at Thr-473 and increases HKII association to mitochondria [85]. MK-2206, a novel allosteric Akt inhibitor, is reported to synergistically inhibit cancer cell proliferation and tumor growth in vitro and in vivo in combination with several anticancer agents [86].

Mutations of K-Ras are found in over 90% of pancreatic ductal adenocarcinomas and are critical in pancreatic cancer initiation. Oncogenic transformation induced by K-Ras attenuates mitochondrial function, increases NADPH oxidase enzyme activity to support aerobic glycolysis, and increases ROS stress [87,88]. Studies exploring inducible expression of oncogenic K-Ras also indicate that oncogenic K-Ras serves a vital role in controlling tumor metabolism through stimulation of glucose uptake and channeling of glucose intermediates into the hexosamine biosynthesis and PPP pathways [89].

Through the transcriptional regulation of GLUT1, HKII, Phosphoglycerate Mutase (PGAM), TP53-induced glycolysis and apoptosis regulator (TIGAR), and synthesis of cytochrome c oxidase (SCO2), tumor suppressor p53 prevents a metabolic switch from aerobic respiration to glycolysis [90-94]. Mutations in p53 cause decreased mitochondrial respiration and increased glycolysis, which provide a molecular link between genetic alterations and the Warburg effect.

ROS and Cancer Metabolism

The major by-products of cellular metabolism are collectively known as ROS. Chemically, these species mainly include superoxide (O2 -), hydrogen peroxide (H2O2), and the hydroxyl radical (OH-) [95]. ROS are mainly produced from the mitochondrial electron transport chain or enzymes such as NADPH oxidases [87,96]. With unpaired electrons, ROS are highly reactive, producing chemical modifications that damage proteins, lipids, nucleotides, and carbohydrates. Moderate elevation of ROS regulates cell signaling and promotes cell proliferation [97], while highly increased level of ROS can lead to cell death [98]. ROS stress is a common feature of human cancer. In order to survive oxidative stress, cancer cells need to rearrange their metabolic flow to promote the synthesis of the major reducing equivalent NADPH and glutathione to maintain ROS at a moderate level. This promotes cancer cell proliferation and continued genomic instability.

The discovery of how ROS modifies the key glycolytic enzyme PKM2 illustrates how ROS regulates cell metabolism. This study demonstrates that nuclear translocation of PKM2 can be triggered by ROS [99]. An acute increase in intracellular ROS level is reported to decrease PKM2 activity through oxidation of PKM2 at cysteine 358, resulting in shunting glucose to the PPP, generating the reducing equivalent NADPH for the detoxification of ROS [100].

HIF1α and c-Myc are two important metabolic regulators as discussed previously. Both are subjected to ROS regulation. By oxidizing the Prolyl Hydroxylase (PHD) cofactors ferrous iron and ascorbate, ROS inhibits PHD activity under normoxic condition and stabilizes HIF1α and therefore alters metabolism into aerobic glycolysis which favors NADPH production through the PPP and allows cancer cells to adapt to ROS stress [101]. H2O2 is reported to induce c-Myc gene expression through ROS-mediated enhancement of IκB degradation, thus allowing NF-κB to translocate into the nucleus, where NF-κB serves as a transcription factor directing c-Myc gene expression [102].

Research conducted recently suggests that metabolic reprogramming might also help cancer cells adapt to ROS stress. 6-phosphofructo-2-kinase/fructose-2,6-biphosphases 4 (PFKBF4) is a bifunctional enzyme responsible for maintaining the cellular level of fructose-2,6-bisphosphate, which is important for the dynamic regulation of glycolytic flux by allosterically activating PFK1. By screening 222 different enzymes and other proteins involved in regulating energy production, PFKBF4 was found to be essential for balancing glycolytic activity and antioxidant production in order to maintain cellular redox balance – making sure that the cell’s energy needs are met without allowing ROS to accumulate and trigger cell death [103]. Blocking PFKBF4 stalls cancer cell growth and triggers a catastrophic build-up of ROS, suggesting that it could be a suitable drug target [103,104].

Oncogenic K-Ras, the signature genetic alteration in PDAC, is known to regulate a non-canonical pathway of glutamine metabolism to favor the production of NADPH and to maintain the cellular redox status [8]. Through upregulation of aspartate transaminase (GOT1), oncogenic Kras promotes the conversion of glutamine-derived aspartate into oxaloacetate, which is converted into malate and then into pyruvate, increasing the NADPH to NADP+ ratio [8]. Targeted inhibition of the enzymes in the pathway leads to an increase in ROS and a profound suppression of PDAC growth [8]. The metabolic vulnerability of this pathway in PDAC may provide novel therapeutic strategies to treat this lethal cancer.

Mitochondria and Cancer Metabolism

Mitochondria play central roles in energy metabolism, maintenance of ion homeostasis, redox balance, and regulation of apoptosis. As the power house of the cell, mitochondrion is an important metabolic organelle intimately linked to glucose, glutamine, amino acids, and lipid metabolisms. Mitochondrial dysfunction, either due to mitochondrial DNA mutation induced defective mitochondrial electron transport chain, excessive HKII association to the mitochondrial membrane, or citric acid cycle enzyme such as FH, SDH, or Isocitrate Dehydrogenase 2 (IDH2) gene mutations, has been observed to alter cell metabolism in a wide spectrum of human cancers. A recent study has discovered that tumors harboring mutations in FH or Electron Transport Chain (ETC) with disabled oxidative mitochondrial function use glutaminedependent reductive carboxylation rather than oxidative metabolism as the major pathway of citrate formation. This pathway uses mitochondrial and cytosolic isoforms of NADP+/NADPH-dependent isocitrate dehydrogenase, IDH1 or IDH2, to metabolize glutaminederived α-ketoglutarate to citrate to provide the acetyl-coenzyme A for lipid synthesis [105]. Moreover, human tumors endure profound hypoxia and the stabilized HIF1α due to hypoxia inhibits the entry of pyruvate into the citric acid cycle and prevents the synthesis of citrate for the production of cytoplasmic acetyl-coenzyme A, a major source of acetyl groups for protein acetylation and lipid synthesis. Under this mitochondrial inhibition condition, glutamine-dependent reductive carboxylation by IDH1 and IDH2 is also found as the major pathway to generate citrate for lipid synthesis to support cell growth and viability [106,107].

The normal mitochondrial oxygen consumption by respiration represents a fundamental important function of mitochondria, which is to maintain redox homeostasis. Mitochondrial dysfunction causes electron leakage, which can be captured by molecular oxygen to generate superoxide, mainly by complex I and III. Superoxide can be converted to other forms of free radicals and released into the intermembrane space for the generation of cytoplasmic ROS. As mentioned previous, ROS levels can directly affect the metabolic activity of cancer cells. A cell model employing a tetracycline controlled expression of a domain-negative form of DNA polymerase gamma has been established to allow real-time monitoring of the metabolic switch from mitochondrial oxidative phosphorylation to aerobic glycolysis [87]. Increases in NADPH oxidase activity were observed to support the metabolic switch with moderate ROS generation. Further study has identified that in order to survive the mitochondrial dysfunction, transcription factor ZNF143 was activated and mediated upregulation of the major antioxidant enzyme Glutathione Peroxidase 1 (GPX1) to keep ROS in the moderate range [108]. Compelling evidence suggests that mitochondrial dysfunction, either due to mitochondrial genetic alterations, disrupted redox balancing, specific metabolic enzyme mutations, or a hypoxic tumor environment alters mitochondrial metabolic flow by providing precursors to build macromolecules and combat ROS stress so as to promote cancer cell survival and drug resistance.

The Interplay of microRNA, Oncogenes/Tumor Suppressors, and Metabolic Enzymes

MicroRNAs (miRNAs) are small, noncoding RNAs with about 22 nucleotides in length. They regulate gene expression by interacting with the 3’-untranslated region (3’-UTR) of their target mRNAs to inhibit translation and increase mRNA degradation. More than 900 miRNAs have been annotated in the human genome and about 30% of human genes are regulated by miRNAs. miRNAs directly or indirectly regulate several key metabolic enzymes and master regulators that cause alterations in metabolic pathways, resulting in tumor progression and metastasis [109,110]. Evidence also indicates that metabolic alterations, such as ROS tress, hypoxia, glucose content, critically orchestrate miRNA activity in various pathologies [111,112].

miRNAs Directly Regulate Metabolic Enzymes

Several key glycolytic enzymes are demonstrated to be regulated by miRNAs. GLUT1 is a key glucose transporter upregulated in various human cancers. miR-130b, miR-301a, and miR-19a/b are reported to increase GLUT1 expression in renal cell carcinoma [113]. Consistent with this discovery, miR-130b and miR-301a are highly expressed in pancreatic cancer, suggesting the role of these miRNAs in regulating glucose metabolism through GLUT1 [114]. In contrast, miR-138, miR- 150, miR-199a, and miR532-5p down-regulate GLUT1 expression [113]. miR-143, downregulated by Mammalian Target of Rapamycin (mTOR) activation, reduces glucose metabolism and inhibits cancer cell proliferation and tumor formation through targeting HKII [115]. miR-122 and the miR-15a/16-1 cluster target AldoA [116,117]. In glioma cells, PKM2 is found to be a target of the tumor suppressive miR-326 [118]. miR-133a and miR-133b target PKM2 in squamous cell carcinoma of tongue [119]. LDHA, the key enzyme responsible for lactate generation, is targeted by miR-34a [120]. Researchers have also demonstrated that miRNAs, such as miR-29a, miR-29b, and miR-124, modulate lactate flow of the cancer cells by regulating MCTs [121,122].

miRNAs Interact With Metabolic Master Regulators

HIF and c-Myc are master regulators of cancer metabolism. Recently, miRNAs have been identified that interact with HIF and c-Myc in tumors. miR-199a targets HIF1α in a feedback manner through Twist-1 [123]. Interestingly, HIF1 is reported as a direct target of the miR-17-92 cluster, which is under the control of c-Myc [124,125]. c-Myc is aberrantly expressed and contributes to the genesis of many human cancers [126]. Myc can not only activate miRNAs, such as miR-17-92, it also represses miRNAs, including miR-let-7, miR-23, and miR-34a [126]. Studied have also demonstrated that miRNAs can regulate c-Myc. miR-let-7a is reported to regulate c-Myc in human colon cancer [127]. miR-let-7a stably transfected cells reduces c-Myc levels and inhibits tumor growth in xenograft models, confirming the tumor suppressor function of miR-let-7a [128]. In addition, tumor suppressor miR-34c regulates c-Myc levels by directly binding the 3’- UTR region of c-Myc mRNA [129]. Halting Myc on the path to cancer by targeting Myc-induced miRNA expression may provide novel opportunities for cancer therapies.

 

RAS mutations are a common feature for tumorigenesis in cancers with multiple origins. Kras is a direct target of let-7 miRNA in lung cancer [130]. In addition, both miR-96 and miR-217 target K-ras to inhibit tumor growth in pancreatic adenocarcinoma [131,132].

P53, an important metabolic regulator, regulates miRNAs and is also regulated by miRNAs. P53-inducible miR-34a is found to regulate several glycolytic enzymes, such as HKI, HKII, and LDHA to inhibit glycolysis [133]. miR-34a also directly targets PDK1 to regulate mitochondrial oxidative phosphorylation. Both miR-125a and miR-125b are reported to target p53 and they are highly expressed in pancreatic cancer for tumorigenesis [134].

miRNAs have emerged as important regulators of cancer metabolism. The potential of miRNA mediated therapy might be a novel direction for cancer treatment. Besides miRNAs, long noncoding RNAs (lncRNAs) are reported to interact with oncogenes and tumor suppressors, such as c-Myc and p53 [135]. The role of lncRNAs in cancer metabolism awaits investigation.

Metabolic Enzymes and Chromatin Remodeling in Cancer

Unique functional alterations of some metabolic enzymes are recently reported to play an important role in the reprogramming of cancer cell metabolism. PKM2 nuclear localization and IDH1/2 mutation are reported to have functions related to chromatin remodeling. PKM2 and chromatin remodeling in cancer has been discussed previously. Here, we only discuss IDH1/2 in chromatin remodeling.

Isocitrate Dehydrogenase Mutation and Gain of Function in Cancer

IDHs are a family of enzymes with three isoforms that catalyzes oxidative decarboxylation of isocitrate to α-KG. IHD1 and IHD2 are NADP+ dependent homodimers, mainly participating in reductive glutamine metabolism in the cytosol/peroxisomes and the mitochondrial compartment, respectively. IDH3 is a NAD+ dependent heterotetramer and is the main functional form of IDH in the citric acid cycle under physiological conditions. Genome-wide mutation analyses and high-throughput deep sequencing have identified that 70% of grade II-III gliomas and secondary glioblastomas as well as 10% of acute myeloid leukemia have mutations in IDH1 or IDH2 [136-138]. The mutations in IDH1 and IDH2 confer a gain of function alteration to the enzymes resulting in the production of oncometabolite 2-Hydroxyglutarate (2HG) from α-KG [139]. As α-KG competitor, elevated 2HG effectively inhibits dioxygenase enzymes, such as the 5’-methylcytosine hydroxylase TET2, resulting in increased histone and DNA methylation [140-143]. This area of study demonstrates that the aberrant metabolite 2HG that results from IDH1/2 mutations can promote cancer by changing the epigenetic landscape of the cells to alter gene expression [144]. 2HG is currently being assessed as a biomarker for cancer detection and a small molecular inhibitor called AGI-5198 and a monoclonal antibody called MsMab-1 specific to mutant IDH1/2 are developed for their therapeutic potentials [145-147]. Using MRS, 2-HG can be detected with high sensitivity and specificity in gliomas harboring IDH mutations ex vivo and in vivo, suggesting that 2-HG can be a clinical valuable biomarker [148].

Chromatin remodeling by alterations of metabolic enzymes in cancer is a novel area of research. Continued study of the dynamic interplay between metabolism and epigenetics in cancer should lead to improved understanding of disease development and may provide novel targets for cancer therapy.

Conclusions

Metabolic reprogramming is a hallmark of cancer. Through oncogene activation, loss of tumor suppressor functions, and epigenetic alterations, cancer cells alter the metabolism of glucose and glutamine for the synthesis of nucleotides, lipids, fatty acids, and proteins to meet the proliferative needs of cancer cells and to maintain a survival advantage in the adverse tumor microenvironment. Beyond catalyzing glucose and glutamine breakdown and providing substrates for the anabolic reactions, cancer metabolic enzymes, such as HKII, PKM2, and mutant IDH, also contribute to tumor adaptation through alternative non-enzymatic functions of these enzymes. Sustained aerobic glycolysis and glutamine metabolism may also protect cancer cells from ROS induced cell death [100]. As a result of the altered metabolism of glucose, the citric acid cycle, and glutamine, targeted inhibition of these metabolic differences in cancer versus normal cells may have beneficial clinical effects. However, the full magnitude of cancer cell alterations in glycolysis, the citric acid cycle, mitochondrial respiration, glutamine utilization, ROS stress, and their interconnection with genetic and epigenetic alterations in cancer remains to be fully elucidated. In addition, the magnitude of metabolic heterogeneity occurring in cancers of different cellular origins as well as within an individual’s cancer is unknown. It is hoped that by targeting altered metabolism of cancer cells, novel therapeutic strategies with high selectivity can be developed.

Acknowledgements

We thank the reviewers for comments. This work was supported by a grant from The University of Texas MD Anderson Cancer Center Sheikh Ahmed Bin Zayed Al Nahyan Center for Pancreatic Cancer Research. We apologize for omissions of important references due to space limitations.

References

  1. Warburg O (1956) On respiratory impairment in cancer cells. Science 124: 269-270.
  2. Gao P, Tchernyshyov I, Chang TC, Lee YS, Kita K, et al. (2009) c-Myc suppression of miR-23a/b enhances mitochondrial glutaminase expression and glutamine metabolism. Nature 458: 762-765.
  3. Wang JB, Erickson JW, Fuji R, Ramachandran S, Gao P, et al. (2010) Targeting mitochondrial glutaminase activity inhibits oncogenic transformation. Cancer Cell 18: 207-219.
  4. Jiang H, Zhao D, Feng T, Liao S, Chen Y (2013) Construction of classifier based on MPCA and QSA and its application on classification of pancreatic diseases. Comput Math Methods Med 2013: 713174.
  5. Cairns RA, Harris IS, Mak TW (2011) Regulation of cancer cell metabolism. Nat Rev Cancer 11: 85-95.
  6. DeBerardinis RJ, Thompson CB (2012) Cellular metabolism and disease: what do metabolic outliers teach us? Cell 148: 1132-1144.
  7. Lunt SY, Vander Heiden MG (2011) Aerobic glycolysis: meeting the metabolic requirements of cell proliferation. Annu Rev Cell Dev Biol 27: 441-464.
  8. Schulze A, Harris AL (2012) How cancer metabolism is tuned for proliferation and vulnerable to disruption. Nature 491: 364-373.
  9. Chen Z, Lu W, Garcia-Prieto C, Huang P (2007) The Warburg effect and its cancer therapeutic implications. J Bioenerg Biomembr 39: 267-274.
  10. Lu W, Huang P (2010) Glycolytic Pathway as a Target for Tumor Inhibition. The Tumor Microenvironment Book 91-118.
  11. Levine AJ, Puzio-Kuter AM (2010) The control of the metabolic switch in cancers by oncogenes and tumor suppressor genes. Science 330: 1340-1344.
  12. Dang CV (2012) Links between metabolism and cancer. Genes Dev 26: 877-890.
  13. Kim JW, Dang CV (2006) Cancer's molecular sweet tooth and the Warburg effect. Cancer Res 66: 8927-8930.
  14. Mathupala SP, Ko YH, Pedersen PL (2006) Hexokinase II: cancer's double-edged sword acting as both facilitator and gatekeeper of malignancy when bound to mitochondria. Oncogene 25: 4777-4786.
  15. Golshani-Hebroni SG, Bessman SP (1997) Hexokinase binding to mitochondria: a basis for proliferative energy metabolism. J Bioenerg Biomembr 29: 331-338.
  16. Wolf A, Agnihotri S, Munoz D, Guha A (2011) Developmental profile and regulation of the glycolytic enzyme hexokinase 2 in normal brain and glioblastoma multiforme. Neurobiol Dis 44: 84-91.
  17. Patra KC, Wang Q, Bhaskar PT, Miller L, Wang Z, et al. (2013) Hexokinase 2 is required for tumor initiation and maintenance and its systemic deletion is therapeutic in mouse models of cancer. Cancer Cell 24: 213-228.
  18. Geschwind JF, Georgiades CS, Ko YH, Pedersen PL (2004) Recently elucidated energy catabolism pathways provide opportunities for novel treatments in hepatocellular carcinoma. Expert Rev Anticancer Ther 4: 449-457.
  19. Xu RH, Pelicano H, Zhou Y, Carew JS, Feng L, et al. (2005) Inhibition of glycolysis in cancer cells: a novel strategy to overcome drug resistance associated with mitochondrial respiratory defect and hypoxia. Cancer Res 65: 613-621.
  20. Ko YH, Pedersen PL, Geschwind JF (2001) Glucose catabolism in the rabbit VX2 tumor model for liver cancer: characterization and targeting hexokinase. Cancer Lett 173: 83-91.
  21. Chen Z, Zhang H, Lu W, Huang P (2009) Role of mitochondria-associated hexokinase II in cancer cell death induced by 3-bromopyruvate. Biochim Biophys Acta 1787: 553-560.
  22. Salani B, Marini C, Rio AD, Ravera S, Massollo M, et al. (2013) Metformin impairs glucose consumption and survival in Calu-1 cells by direct inhibition of hexokinase-II. Sci Rep 3: 2070.
  23. Jeong NY, Yoo YH (2012) Cerulenin-induced apoptosis is mediated by disrupting the interaction between AIF and hexokinase II. Int J Oncol 40: 1949-1956.
  24. Christofk HR, Vander Heiden MG, Harris MH, Ramanathan A, Gerszten RE, et al. (2008) The M2 splice isoform of pyruvate kinase is important for cancer metabolism and tumour growth. Nature 452: 230-233.
  25. Vander Heiden MG, Christofk HR, Schuman E, Subtelny AO, Sharfi H, et al. (2010) Identification of small molecule inhibitors of pyruvate kinase M2. Biochem Pharmacol 79: 1118-1124.
  26. Altenberg B, Greulich KO (2004) Genes of glycolysis are ubiquitously overexpressed in 24 cancer classes. Genomics 84: 1014-1020.
  27. David CJ, Chen M, Assanah M, Canoll P, Manley JL (2010) HnRNP proteins controlled by c-Myc deregulate pyruvate kinase mRNA splicing in cancer. Nature 463: 364-368.
  28. Gao X, Wang H, Yang JJ, Liu X, Liu ZR (2012) Pyruvate kinase M2 regulates gene transcription by acting as a protein kinase. Mol Cell 45: 598-609.
  29. Christofk HR, Vander Heiden MG, Wu N, Asara JM, Cantley LC (2008) Pyruvate kinase M2 is a phosphotyrosine-binding protein. Nature 452: 181-186.
  30. Tennant DA, Durán RV, Gottlieb E (2010) Targeting metabolic transformation for cancer therapy. Nat Rev Cancer 10: 267-277.
  31. Yang W, Xia Y, Hawke D, Li X, Liang J, et al. (2012) PKM2 phosphorylates histone H3 and promotes gene transcription and tumorigenesis. Cell 150: 685-696.
  32. Yang W, Xia Y, Ji H, Zheng Y, Liang J, et al. (2011) Nuclear PKM2 regulates β-catenin transactivation upon EGFR activation. Nature 480: 118-122.
  33. Hove H, Rye Clausen M, Brøbech Mortensen P (1993) Lactate and pH in faeces from patients with colonic adenomas or cancer. Gut 34: 625-629.
  34. Gatenby RA, Gillies RJ (2004) Why do cancers have high aerobic glycolysis? Nat Rev Cancer 4: 891-899.
  35. Lu H, Forbes RA, Verma A (2002) Hypoxia-inducible factor 1 activation by aerobic glycolysis implicates the Warburg effect in carcinogenesis. J Biol Chem 277: 23111-23115.
  36. Fantin VR, St-Pierre J, Leder P (2006) Attenuation of LDH-A expression uncovers a link between glycolysis, mitochondrial physiology, and tumor maintenance. Cancer Cell 9: 425-434.
  37. Zhao D, Zou SW, Liu Y, Zhou X, Mo Y, et al. (2013) Lysine-5 acetylation negatively regulates lactate dehydrogenase A and is decreased in pancreatic cancer. Cancer Cell 23: 464-476.
  38. Zhao D, Xiong Y, Lei QY, Guan KL (2013) LDH-A Acetylation: implication in Pancreatic Cancer Initiation and Diagnosis. Oncotarget .
  39. Le A, Cooper CR, Gouw AM, Dinavahi R, Maitra A, et al. (2010) Inhibition of lactate dehydrogenase A induces oxidative stress and inhibits tumor progression. Proc Natl Acad Sci U S A 107: 2037-2042.
  40. Dutta P, Le A, Vander Jagt DL, Tsukamoto T, Martinez GV, et al. (2013) Evaluation of LDH-A and Glutaminase Inhibition In Vivo by Hyperpolarized 13C-Pyruvate Magnetic Resonance Spectroscopy of Tumors. Cancer Res 73: 4190-4195.
  41. Heinrich PC, Morris HP, Weber G (1976) Behavior of transaldolase (EC 2.2.1.2) and transketolase (EC 2.2.1.1) Activities in normal, neoplastic, differentiating, and regenerating liver. Cancer Res 36(9 pt.
  42. Coy JF, Dressler D, Wilde J, Schubert P (2005) Mutations in the transketolase-like gene TKTL1: clinical implications for neurodegenerative diseases, diabetes and cancer. Clin Lab 51: 257-273.
  43. Langbein S, Zerilli M, Zur Hausen A, Staiger W, Rensch-Boschert K, et al. (2006) Expression of transketolase TKTL1 predicts colon and urothelial cancer patient survival: Warburg effect reinterpreted. Br J Cancer 94: 578-585.
  44. Du W, Jiang P, Mancuso A, Stonestrom A, Brewer MD, et al. (2013) TAp73 enhances the pentose phosphate pathway and supports cell proliferation. Nat Cell Biol 15: 991-1000.
  45. Langbein S, Frederiks WM, zur Hausen A, Popa J, Lehmann J, et al. (2008) Metastasis is promoted by a bioenergetic switch: new targets for progressive renal cell cancer. Int J Cancer 122: 2422-2428.
  46. Raïs B, Comin B, Puigjaner J, Brandes JL, Creppy E, et al. (1999) Oxythiamine and dehydroepiandrosterone induce a G1 phase cycle arrest in Ehrlich's tumor cells through inhibition of the pentose cycle. FEBS Lett 456: 113-118.
  47. Comín-Anduix B, Boren J, Martinez S, Moro C, Centelles JJ, et al. (2001) The effect of thiamine supplementation on tumour proliferation. A metabolic control analysis study. Eur J Biochem 268: 4177-4182.
  48. Budihardjo II, Walker DL, Svingen PA, Buckwalter CA, Desnoyers S, et al. (1998) 6-Aminonicotinamide sensitizes human tumor cell lines to cisplatin. Clin Cancer Res 4: 117-130.
  49. Varshney R, Adhikari JS, Dwarakanath BS (2003) Contribution of oxidative stress to radiosensitization by a combination of 2-DG and 6-AN in human cancer cell line. Indian J Exp Biol 41: 1384-1391.
  50. Evans MJ, Saghatelian A, Sorensen EJ, Cravatt BF (2005) Target discovery in small-molecule cell-based screens by in situ proteome reactivity profiling. Nat Biotechnol 23: 1303-1307.
  51. Sonveaux P, Végran F, Schroeder T, Wergin MC, Verrax J, et al. (2008) Targeting lactate-fueled respiration selectively kills hypoxic tumor cells in mice. J Clin Invest 118: 3930-3942.
  52. Liu Y, Cao Y, Zhang W, Bergmeier S, Qian Y, et al. (2012) 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: 1672-1682.
  53. Wood TE, Dalili S, Simpson CD, Hurren R, Mao X, et al. (2008) A novel inhibitor of glucose uptake sensitizes cells to FAS-induced cell death. Mol Cancer Ther 7: 3546-3555.
  54. Cao X, Fang L, Gibbs S, Huang Y, Dai Z, et al. (2007) Glucose uptake inhibitor sensitizes cancer cells to daunorubicin and overcomes drug resistance in hypoxia. Cancer Chemother Pharmacol 59: 495-505.
  55. Bonnet S, Archer SL, Allalunis-Turner J, Haromy A, Beaulieu C, et al. (2007) A mitochondria-K+ channel axis is suppressed in cancer and its normalization promotes apoptosis and inhibits cancer growth. Cancer Cell 11: 37-51.
  56. Cao W, Yacoub S, Shiverick KT, Namiki K, Sakai Y, et al. (2008) Dichloroacetate (DCA) sensitizes both wild-type and over expressing Bcl-2 prostate cancer cells in vitro to radiation. Prostate 68: 1223-1231.
  57. Cairns RA, Papandreou I, Sutphin PD, Denko NC (2007) Metabolic targeting of hypoxia and HIF1 in solid tumors can enhance cytotoxic chemotherapy. Proc Natl Acad Sci U S A 104: 9445-9450.
  58. Gottfried E, Lang SA, Renner K, Bosserhoff A, Gronwald W, et al. (2013) New Aspects of an Old Drug - Diclofenac Targets MYC and Glucose Metabolism in Tumor Cells. PLoS One 8: e66987.
  59. Chirasani SR, Leukel P, Gottfried E, Hochrein J, Stadler K, et al. (2013) Diclofenac inhibits lactate formation and efficiently counteracts local immune suppression in a murine glioma model. Int J Cancer 132: 843-853.
  60. Curthoys NP, Watford M (1995) Regulation of glutaminase activity and glutamine metabolism. Annu Rev Nutr 15: 133-159.
  61. Lu W, Pelicano H, Huang P (2010) Cancer metabolism: is glutamine sweeter than glucose? Cancer Cell 18: 199-200.
  62. Yuneva M, Zamboni N, Oefner P, Sachidanandam R, Lazebnik Y (2007) Deficiency in glutamine but not glucose induces MYC-dependent apoptosis in human cells. J Cell Biol 178: 93-105.
  63. Suzuki S, Tanaka T, Poyurovsky MV, Nagano H, Mayama T, et al. Phosphate-activated glutaminase (GLS2), a p53-inducible regulator of glutamine metabolism and reactive oxygen species. Proc Natl Acad Sci U S A 107: 7461-7466.
  64. Lobo C, Ruiz-Bellido MA, Aledo JC, Márquez J, Núñez De Castro I, et al. (2000) Inhibition of glutaminase expression by antisense mRNA decreases growth and tumourigenicity of tumour cells. Biochem J 348 Pt 2: 257-261.
  65. Chang SM, Kuhn JG, Robins HI, Schold SC, Spence AM, et al. (1999) Phase II study of phenylacetate in patients with recurrent malignant glioma: a North American Brain Tumor Consortium report. J Clin Oncol 17: 984-990.
  66. Li XN, Parikh S, Shu Q, Jung HL, Chow CW, et al. (2004) Phenylbutyrate and phenylacetate induce differentiation and inhibit proliferation of human medulloblastoma cells. Clin Cancer Res 10: 1150-1159.
  67. Dang CV, Semenza GL (1999) Oncogenic alterations of metabolism. Trends Biochem Sci 24: 68-72.
  68. Denko NC (2008) Hypoxia, HIF1 and glucose metabolism in the solid tumour. Nat Rev Cancer 8: 705-713.
  69. Safran M, Kaelin WG Jr (2003) HIF hydroxylation and the mammalian oxygen-sensing pathway. J Clin Invest 111: 779-783.
  70. Dang CV, Kim JW, Gao P, Yustein J (2008) The interplay between MYC and HIF in cancer. Nat Rev Cancer 8: 51-56.
  71. Kim JW, Tchernyshyov I, Semenza GL, Dang CV (2006) HIF-1-mediated expression of pyruvate dehydrogenase kinase: a metabolic switch required for cellular adaptation to hypoxia. Cell Metab 3: 177-185.
  72. Baysal BE, Ferrell RE, Willett-Brozick JE, Lawrence EC, Myssiorek D, et al. (2000) Mutations in SDHD, a mitochondrial complex II gene, in hereditary paraganglioma. Science 287: 848-851.
  73. Tomlinson IP, Alam NA, Rowan AJ, Barclay E, Jaeger EE, et al. (2002) Germline mutations in FH predispose to dominantly inherited uterine fibroids, skin leiomyomata and papillary renal cell cancer. Nat Genet 30: 406-410.
  74. Pollard PJ, Brière JJ, Alam NA, Barwell J, Barclay E, et al. (2005) Accumulation of Krebs cycle intermediates and over-expression of HIF1alpha in tumours which result from germline FH and SDH mutations. Hum Mol Genet 14: 2231-2239.
  75. Koivunen P, Hirsilä M, Remes AM, Hassinen IE, Kivirikko KI, et al. Inhibition of hypoxia-inducible factor (HIF) hydroxylases by citric acid cycle intermediates: possible links between cell metabolism and stabilization of HIF. J Biol Chem 282: 4524-4532.
  76. Narita T, Yin S, Gelin CF, Moreno CS, Yepes M, et al. (2009) Identification of a novel small molecule HIF-1alpha translation inhibitor. Clin Cancer Res 15: 6128-6136.
  77. Dikmen ZG, Gellert GC, Dogan P, Yoon H, Lee YB, et al. (2008) In vivo and in vitro effects of a HIF-1alpha inhibitor, RX-0047. J Cell Biochem 104: 985-994.
  78. Little CD, Nau MM, Carney DN, Gazdar AF, Minna JD (1983) Amplification and expression of the c-myc oncogene in human lung cancer cell lines. Nature 306: 194-196.
  79. Dang CV, O'Donnell KA, Zeller KI, Nguyen T, Osthus RC, et al. (2006) The c-Myc target gene network. Semin Cancer Biol 16: 253-264.
  80. Wise DR, DeBerardinis RJ, Mancuso A, Sayed N, Zhang XY, et al. (2008) Myc regulates a transcriptional program that stimulates mitochondrial glutaminolysis and leads to glutamine addiction. Proc Natl Acad Sci U S A 105: 18782-18787.
  81. Dang CV (2010) Rethinking the Warburg effect with Myc micromanaging glutamine metabolism. Cancer Res 70: 859-862.
  82. Esslinger CS, Cybulski KA, Rhoderick JF (2005) Ngamma-aryl glutamine analogues as probes of the ASCT2 neutral amino acid transporter binding site. Bioorg Med Chem 13: 1111-1118.
  83. Nicklin P, Bergman P, Zhang B, Triantafellow E, Wang H, et al. (2009) Bidirectional transport of amino acids regulates mTOR and autophagy. Cell 136: 521-534.
  84. Robey RB, Hay N (2009) Is Akt the "Warburg kinase"?-Akt-energy metabolism interactions and oncogenesis. Semin Cancer Biol 19: 25-31.
  85. Roberts DJ, Tan-Sah VP, Smith JM, Miyamoto S (2013) Akt Phosphorylates HK-II at Thr-473 and Increases Mitochondrial HK-II Association to Protect Cardiomyocytes. J Biol Chem 288: 23798-23806.
  86. Hirai H, Sootome H, Nakatsuru Y, Miyama K, Taguchi S, et al. (2010) MK-2206, an allosteric Akt inhibitor, enhances antitumor efficacy by standard chemotherapeutic agents or molecular targeted drugs in vitro and in vivo. Mol Cancer Ther 9: 1956-1967.
  87. Lu W, Hu Y, Chen G, Chen Z, Zhang H, et al. (2012) Novel role of NOX in supporting aerobic glycolysis in cancer cells with mitochondrial dysfunction and as a potential target for cancer therapy. PLoS Biol 10: e1001326.
  88. Hu Y, Lu W, Chen G, Wang P, Chen Z, et al. (2012) K-ras(G12V) transformation leads to mitochondrial dysfunction and a metabolic switch from oxidative phosphorylation to glycolysis. Cell Res 22: 399-412.
  89. Ying H, Kimmelman AC, Lyssiotis CA, Hua S, Chu GC, et al. (2012) Oncogenic Kras maintains pancreatic tumors through regulation of anabolic glucose metabolism. Cell 149: 656-670
  90. Shaw RJ (2006) Glucose metabolism and cancer. Curr Opin Cell Biol 18: 598-608.
  91. Green DR, Chipuk JE (2006) p53 and metabolism: Inside the TIGAR. Cell 126: 30-32.
  92. Bensaad K, Tsuruta A, Selak MA, Vidal MN, Nakano K, et al. (2006) TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell 126: 107-120.
  93. Matoba S, Kang JG, Patino WD, Wragg A, Boehm M, et al. (2006) p53 regulates mitochondrial respiration. Science 312: 1650-1653.
  94. Feng Z, Levine AJ (2010) The regulation of energy metabolism and the IGF-1/mTOR pathways by the p53 protein. Trends Cell Biol 20: 427-434.
  95. Lu W, Ogasawara MA, Huang P (2007) Models of reactive oxygen species in cancer. Drug Discov Today Dis Models 4: 67-73.
  96. Dröge W (2002) Free radicals in the physiological control of cell function. Physiol Rev 82: 47-95.
  97. Burdon RH (1995) Superoxide and hydrogen peroxide in relation to mammalian cell proliferation. Free Radic Biol Med 18: 775-794.
  98. Jabs T (1999) Reactive oxygen intermediates as mediators of programmed cell death in plants and animals. Biochem Pharmacol 57: 231-245.
  99. Steták A, Veress R, Ovádi J, Csermely P, Kéri G, et al. (2007) Nuclear translocation of the tumor marker pyruvate kinase M2 induces programmed cell death. Cancer Res 67: 1602-1608.
  100. Anastasiou D, Poulogiannis G, Asara JM, Boxer MB, Jiang JK, et al. (2011) Inhibition of pyruvate kinase M2 by reactive oxygen species contributes to cellular antioxidant responses. Science 334: 1278-1283.
  101. Gerald D, Berra E, Frapart YM, Chan DA, Giaccia AJ, et al. (2004) JunD reduces tumor angiogenesis by protecting cells from oxidative stress. Cell 118: 781-794.
  102. de Souza GA, Godoy LM, Teixeira VR, Otake AH, Sabino A, et al. (2006) Proteomic and SAGE profiling of murine melanoma progression indicates the reduction of proteins responsible for ROS degradation. Proteomics 6: 1460-1470.
  103. Ros S, Santos CR, Moco S, Baenke F, Kelly G, et al. (2012) Functional metabolic screen identifies 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 4 as an important regulator of prostate cancer cell survival. Cancer Discov 2: 328-343.
  104. Dang CV (2012) Cancer cell metabolism: there is no ROS for the weary. Cancer Discov 2: 304-307.
  105. Mullen AR, Wheaton WW, Jin ES, Chen PH, Sullivan LB, et al. (2011) Reductive carboxylation supports growth in tumour cells with defective mitochondria. Nature 481: 385-388.
  106. Wise DR, Ward PS, Shay JE, Cross JR, Gruber JJ, et al. (2011) Hypoxia promotes isocitrate dehydrogenase-dependent carboxylation of alpha-ketoglutarate to citrate to support cell growth and viability. Proc Natl Acad Sci U S A 108: 19611-19616.
  107. Metallo CM, Gameiro PA, Bell EL, Mattaini KR, Yang J, et al. (2011) Reductive glutamine metabolism by IDH1 mediates lipogenesis under hypoxia. Nature 481: 380-384.
  108. Lu W, Chen Z, Zhang H, Wang Y, Luo Y, et al. (2012) ZNF143 transcription factor mediates cell survival through upregulation of the GPX1 activity in the mitochondrial respiratory dysfunction. Cell Death Dis 3: e422.
  109. Singh PK, Brand RE, Mehla K (2012) MicroRNAs in pancreatic cancer metabolism. Nat Rev Gastroenterol Hepatol 9: 334-344.
  110. Singh PK, Mehla K, Hollingsworth MA, Johnson KR (2011) Regulation of Aerobic Glycolysis by microRNAs in Cancer. Mol Cell Pharmacol 3: 125-134.
  111. Wang Z, Liu Y, Han N, Chen X, Yu W, et al. (2010) Profiles of oxidative stress-related microRNA and mRNA expression in auditory cells. Brain Res 1346: 14-25.
  112. Kulshreshtha R, Davuluri RV, Calin GA, Ivan M (2008) A microRNA component of the hypoxic response. Cell Death Differ 15: 667-671.
  113. Chow TF, Mankaruos M, Scorilas A, Youssef Y, Girgis A, et al. (2010) The miR-17-92 cluster is over expressed in and has an oncogenic effect on renal cell carcinoma. J Urol 183: 743-751.
  114. Szafranska AE, Davison TS, John J, Cannon T, Sipos B, et al. (2007) MicroRNA expression alterations are linked to tumorigenesis and non-neoplastic processes in pancreatic ductal adenocarcinoma. Oncogene 26: 4442-4452.
  115. Fang R, Xiao T, Fang Z, Sun Y, Li F, et al. (2012) MicroRNA-143 (miR-143) regulates cancer glycolysis via targeting hexokinase 2 gene. J Biol Chem 287: 23227-23235.
  116. Esau C, Davis S, Murray SF, Yu XX, Pandey SK, et al. (2006) miR-122 regulation of lipid metabolism revealed by in vivo antisense targeting. Cell Metab 3: 87-98.
  117. Calin GA, Cimmino A, Fabbri M, Ferracin M, Wojcik SE, et al. (2008) MiR-15a and miR-16-1 cluster functions in human leukemia. Proc Natl Acad Sci U S A 105: 5166-5171.
  118. Kefas B, Comeau L, Erdle N, Montgomery E, Amos S, et al. (2010) Pyruvate kinase M2 is a target of the tumor-suppressive microRNA-326 and regulates the survival of glioma cells. Neuro Oncol 12: 1102-1112.
  119. Wong TS, Liu XB, Chung-Wai Ho A, Po-Wing Yuen A, Wai-Man Ng R, et al. (2008) Identification of pyruvate kinase type M2 as potential oncoprotein in squamous cell carcinoma of tongue through microRNA profiling. Int J Cancer 123: 251-257.
  120. Kaller M, Liffers ST, Oeljeklaus S, Kuhlmann K, Röh S, et al. (2011) Genome-wide characterization of miR-34a induced changes in protein and mRNA expression by a combined pulsed SILAC and microarray analysis. Mol Cell Proteomics 10: M111.
  121. Pullen TJ, da Silva Xavier G, Kelsey G, Rutter GA (2011) miR-29a and miR-29b contribute to pancreatic beta-cell-specific silencing of monocarboxylate transporter 1 (Mct1). Mol Cell Biol 31: 3182-3194.
  122. Li KK, Pang JC, Ching AK, Wong CK, Kong X, et al. (2009) miR-124 is frequently down-regulated in medulloblastoma and is a negative regulator of SLC16A1. Hum Pathol 40: 1234-1243.
  123. Rane S, He M, Sayed D, Vashistha H, Malhotra A, et al. (2009) Downregulation of miR-199a derepresses hypoxia-inducible factor-1alpha and Sirtuin 1 and recapitulates hypoxia preconditioning in cardiac myocytes. Circ Res 104: 879-886.
  124. Taguchi A, Yanagisawa K, Tanaka M, Cao K, Matsuyama Y, et al. (2008) Identification of hypoxia-inducible factor-1 alpha as a novel target for miR-17-92 microRNA cluster. Cancer Res 68: 5540-5545.
  125. O'Donnell KA, Wentzel EA, Zeller KI, Dang CV, Mendell JT (2005) c-Myc-regulated microRNAs modulate E2F1 expression. Nature 435: 839-843.
  126. Akao Y, Nakagawa Y, Naoe T (2006) let-7 microRNA functions as a potential growth suppressor in human colon cancer cells. Biol Pharm Bull 29: 903-906.
  127. He XY, Chen JX, Zhang Z, Li CL, Peng QL, et al. (2010) The let-7a microRNA protects from growth of lung carcinoma by suppression of k-Ras and c-Myc in nude mice. J Cancer Res Clin Oncol 136: 1023-1028.
  128. Cannell IG, Bushell M (2010) Regulation of Myc by miR-34c: A mechanism to prevent genomic instability? Cell Cycle 9: 2726-2730.
  129. Johnson SM, Grosshans H, Shingara J, Byrom M, Jarvis R, et al. (2005) RAS is regulated by the let-7 microRNA family. Cell 120: 635-647.
  130. Yu S, Lu Z, Liu C, Meng Y, Ma Y, et al. (2010) miRNA-96 suppresses KRAS and functions as a tumor suppressor gene in pancreatic cancer. Cancer Res 70: 6015-6025.
  131. Zhao WG, Yu SN, Lu ZH, Ma YH, Gu YM, et al. (2010) The miR-217 microRNA functions as a potential tumor suppressor in pancreatic ductal adenocarcinoma by targeting KRAS. Carcinogenesis 31: 1726-1733.
  132. Kim HR, Roe JS, Lee JE, Cho EJ, Youn HD (2013) p53 regulates glucose metabolism by miR-34a. Biochem Biophys Res Commun 437: 225-231.
  133. Le MT, Teh C, Shyh-Chang N, Xie H, Zhou B, et al. (2009) MicroRNA-125b is a novel negative regulator of p53. Genes Dev 23: 862-876.
  134. Hauptman N, Glavač D (2013) Long non-coding RNA in cancer. Int J Mol Sci 14: 4655-4669.
  135. Balss J, Meyer J, Mueller W, Korshunov A, Hartmann C, et al. (2008) Analysis of the IDH1 codon 132 mutation in brain tumors. Acta Neuropathol 116: 597-602.
  136. Hartmann C, Meyer J, Balss J, Capper D, Mueller W, et al. (2009) Type and frequency of IDH1 and IDH2 mutations are related to astrocytic and oligodendroglial differentiation and age: a study of 1,010 diffuse gliomas. Acta Neuropathol 118: 469-474.
  137. Paschka P, Schlenk RF, Gaidzik VI, Habdank M, Krönke J, et al. (2010) IDH1 and IDH2 mutations are frequent genetic alterations in acute myeloid leukemia and confer adverse prognosis in cytogenetically normal acute myeloid leukemia with NPM1 mutation without FLT3 internal tandem duplication. J Clin Oncol 28: 3636-3643.
  138. Dang L, White DW, Gross S, Bennett BD, Bittinger MA, et al. (2009) Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 462: 739-744.
  139. Figueroa ME, Abdel-Wahab O, Lu C, Ward PS, Patel J, et al. (2010) Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell 18: 553-567.
  140. Sasaki M, Knobbe CB, Munger JC, Lind EF, Brenner D, et al. (2012) IDH1(R132H) mutation increases murine haematopoietic progenitors and alters epigenetics. Nature 488: 656-659.
  141. Lu C, Ward PS, Kapoor GS, Rohle D, Turcan S, et al. (2012) IDH mutation impairs histone demethylation and results in a block to cell differentiation. Nature 483: 474-478.
  142. Turcan S, Rohle D, Goenka A, Walsh LA, Fang F, et al. (2012) IDH1 mutation is sufficient to establish the glioma hypermethylator phenotype. Nature 483: 479-483.
  143. Losman JA, Looper RE, Koivunen P, Lee S, Schneider RK, et al. (2013) (R)-2-hydroxyglutarate is sufficient to promote leukemogenesis and its effects are reversible. Science 339: 1621-1625.
  144. Kato Kaneko M, Ogasawara S, Kato Y (2013) Establishment of a multi-specific monoclonal antibody MsMab-1 recognizing both IDH1 and IDH2 mutations. Tohoku J Exp Med 230: 103-109.
  145. Rohle D, Popovici-Muller J, Palaskas N, Turcan S, Grommes C, et al. (2013) An inhibitor of mutant IDH1 delays growth and promotes differentiation of glioma cells. Science 340: 626-630.
  146. Dunn GP, Andronesi OC, Cahill DP (2013) From genomics to the clinic: biological and translational insights of mutant IDH1/2 in glioma. Neurosurg Focus 34: E2.
  147. Esmaeili M, Vettukattil R, Bathen TF (2013) 2-hydroxyglutarate as a magnetic resonance biomarker for glioma subtyping. Transl Oncol 6: 92-98.
Citation: Lu W, Logsdon CD, Abbruzzese JL (2013) Cancer Metabolism and Its Therapeutic Implications. J Cell Sci Ther 4:143.

Copyright: © 2013 Lu W, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top