Abstract
Upregulated glycolysis, both in normoxic and hypoxic environments, is a nearly universal trait of cancer cells. The enormous difference in glucose metabolism offers a target for therapeutic intervention with a potentially low toxicity profile. The past decade has seen a steep rise in the development and clinical assessment of small molecules that target glycolysis. The enzymes in glycolysis have a highly heterogenous nature that allows for the different bioenergetic, biosynthetic, and signaling demands needed for various tissue functions. In cancers, these properties enable them to respond to the variable requirements of cell survival, proliferation and adaptation to nutrient availability. Heterogeneity in glycolysis occurs through the expression of different isoforms, post-translational modifications that affect the kinetic and regulatory properties of the enzyme. In this review we will explore this vast heterogeneity of glycolysis and discuss how this information might be exploited to better target glucose metabolism and offer possibilities for biomarker development.
Keywords: Glycolysis, isoenzymes, cancer, spliceforms, anti-glycolytic agents
1. Introduction
Increased glucose metabolism is one of the distinguishing features between normal cells and highly proliferating cells like cancer, stem and immune cells[1]. Recognized by Otto Warburg over 80 years ago, one of the main features is the observation that growing cells secrete lactate while consuming large amounts of glucose (the Warburg Effect)[2]. The Warburg Effect occurs both through the activity of oncogenes and tumor suppressor genes [3] and through adaptations to the tumor microenvironment[4]. Pre-clinical and clinical research into cancer metabolism has shown that small molecules that interfere with different metabolic pathways can indeed have marked effects on tumor proliferation, either alone[5-7] or in combination with longstanding chemotherapies[8-11]. In fact, some of the oldest classes of chemotherapy are anti-metabolites that interfere with one-carbon metabolism and nucleotide synthesis such as 5-fluorouracil (5-FU) and methotrexate[12]. Even though many of these drugs target one-carbon metabolism and nucleotide synthesis, the last decade has seen additional development of agents that target glycolysis. There are several rationales for targeting glycolysis. First, high glycolytic flux leads to new biology including the selective diversion of carbon into several anabolic pathways and thus provides precursors for nucleotide, protein and fatty acids synthesis and the maintenance of signal transduction processes that occur through changing the levels of metabolites[2]. Second, the rate at which glycolysis is upregulated in cancer cells compared to normal cells creates an opportunity to selectively target tumors[13]. Last, the high flux leading to lactate secretion has also been linked to promote favorable non cell autonomous conditions for uncontrolled proliferation such as evasion of the immune system[14] and induction of angiogenesis and metastasis[14] and these processes may be targeted by altering glycolysis. Nevertheless, there are still many unanswered questions on how to target glycolysis more effectively.
2. Heterogeneity in glycolysis
Heterogeneity in glycolysis is a typical feature of eukaryotic cells living either as interacting cell populations or as multi-cellular organisms[15]. This offers cells from various tissues or populations control to express the enzyme with the optimal kinetic and regulatory properties needed for the specific tissue or population function. Specifically, for multi-cellular organisms, this serves three main functions. Firstly, it allows for a coordinated control of glucose homeostasis within the body[16]. Second, it also allows an organisms cells to respond adequately to different forms of oxygen or glucose stress, such as hypoxia[15], ischemia[17] or changes in diet[16]. Lastly, glycolytic heterogeneity that allows for a high glycolytic rate, offers a subset of cells that need to sustain increased levels of proliferation during certain stages of physiological processes like angiogenesis[18], immune activation[19] and stem cell growth[20] (also see above). Research has shown that these isoenzymes also display a wide variety in expression according to tissue specific tumor development[21-23]. The functional basis for this tumor specific heterogeneity can to a certain degree be explained by the metabolic features of the tissue of origin[21]. There is also evidence that glycolytic isoenzymes are differentially regulated during the cell cycle[24]. Because proliferation rate and glycolysis also show a certain degree of correlation, this therefore explains another contributing factor to glycolytic heterogeneity in tumors with varying proliferation rates[21].
Many glycolytic isoenzymes are transcribed from different gene loci (see table 1 and figure 1). This offers cells the first level of heterogeneity. Additional heterogeneity is introduced because a specific gene locus can also give rise to different splice-forms. Some of these splice forms have only been predicted computationally or identified from preliminary large scale RNA-sequencing and still await further experimental validation[25]. Especially N-terminal splicing seems to be common for most glycolytic enzymes, with one of the most notable exceptions being the pyruvate kinase muscle isozyme (PKM) (see table 1). These isoenzymes have different enzyme kinetic parameters such as Michaelis constants (see table 1) and turnover rates. In addition many isoforms of glycolytic enzymes experience different allosteric regulation of small molecules. Furthermore, these enzymes have different regulation of their levels by promoters[26, 27], non-coding RNA[28, 29], expression of splice factors[30] etc. Further complexity could also be generated by the fact that glycolytic enzymes undergo a myriad of posttranslational modifications (PTMs) that are regulated by signal transduction events[31-40]. In some instances, modifications could allow for the exertion of ‘moonlighting’ functionalities such as modulation of gene-expression in the nucleus[41, 42], inhibition of apoptosis[43]. According to the phosphositeplus database[44], the most prevalent modifications are serine-, threonine- and tyrosine-phosphorylation, acetylation and ubiquitination (see table 1). Nevertheless despite these intriguing possibilities attributed to enzyme PTMs, it is not clear whether the stoichiometry of these modifications can ever reach sufficient levels to have a substantial impact on glycolytic flux[38, 45].
Table 1.
Various tissue specific, genetic, biochemical and kinetic properties of glycolytic enzymes.
| Posttranslational modifications | ||||||||||
|---|---|---|---|---|---|---|---|---|---|---|
|
| ||||||||||
| Enzyme (predominant complex formation) |
Isoforms with separate gene loci |
Gene locus | Splice forms (N- terminal |
Km substrate (mM) | Cancer associated isoform |
Organ specific expression | P | Ac | Ub | Other |
| HK (Monomer, dimer, tetramer) | HK1 | 10q22 | 4 (3) | 0.03 | Most tissues | 16 | 4 | 18 | 0 | |
| HK2 | 2p13 | / | 0.37 | x | Insulin-sensitive tissues | 10 | 0 | 15 | 0 | |
| HK3 | 5q35.2 | / | 0.003 | 12 | 0 | 2 | 0 | |||
| HK4 (Glucokinase) |
7p15.3-p15.1 | 3 (2) | 10 | Liver, pancreas | 1 | 1 | 11 | 6 | ||
| GPI (Dimer) | GPI | 19q13.1 | 2 | 0.4 | x | Most tissues | 20 | 7 | 20 | 1 |
| PFK1 (Homo- and hetero-tetramers) | PFKM | 12q13.3 | 3 | 0.6–2 (most sensitive to F26P) |
x | Muscle, heart, brain | 15 | 4 | 16 | 0 |
| PFKL | 21q22.3 | 2 | 0.35-0.55 (not in combination with L and M subunits) |
x | Liver, kidney | 9 | 3 | 5 | 0 | |
| PFKP | 10p15.3-p15.2 | 2 (1) | 1.4–4 (lowest affinity for F6P) |
x | Platelet | 15 | 7 | 22 | 1 | |
| PFK2 (Tetramers through dimerization of dimers) |
PFKFB1 | Xp11.21 | / | 0.015(F6P) / 0.0005 (F26P) |
Liver, adipose tissue, proliferating cells | 10 | 1 | 5 | 0 | |
| PFKFB2 | 1q31 | 2 | 0.03 (F6P) / 0.07(F26P) |
Heart, kidney | 15 | 0 | 0 | 0 | ||
| PFKFB3 | 10p15.1 | 4 (2) | 0.03 (F6P) / 0.13(F26P) |
xxx | All organs, most in muscle, most hypoxia responsive |
18 | 0 | 4 | 0 | |
| PFKFB4 | 3p22-p21 | 0.09 (F6P) / 0.02(F26P) |
x | Testis | 3 | 0 | 1 | 0 | ||
| ALDO (Homotetramer) | ALDOA | 16p11.2 | 2 (1) | 0.03 | x | Embryonic, most tissues, high in muscle and erythrocytes |
29 | 13 | 13 | 2 |
| ALDOB | 9q21.3-q22.2 | / | 0.003 | x | Liver, the kidney cortex | 12 | 10 | 15 | 0 | |
| ALDOC | 17cen-q12 | / | 0.01 | x | Brain, nervous tissue and smooth and heart muscle |
14 | 6 | 13 | 0 | |
| TPI1 (Homodimer) | TPI1 | 12p13 | 3 (2) | 1.6 (DHAP ) / 0.51 (G3P) |
xxx | Most tissues | 18 | 8 | 11 | 0 |
| GAPDH (Homotetramer) | GAPDH | 12p13 | 2 (1) | 0.19 | xxx | Most tissues | 41 | 17 | 25 | 1 |
| GAPDHS | 19q13.12 | / | Testis | 13 | 2 | 7 | 2 | |||
| PGK (Monomer) | PGK1 | Xq13.3 | / | 0.079 | xxx | Most tissues | 25 | 20 | 30 | 0 |
| PGK2 | 6p12.3 | / | Spermatogenesis | 3 | 6 | 3 | 0 | |||
| PGAM (Monomer) | PGAM1 | 1p31 | 2 (1) | 0.19 | xxx | Most tissues | 20 | 8 | 15 | 0 |
| PGAM2 | 4p14 | / | Muscle, embryonic | 8 | 1 | 4 | 0 | |||
| ENO (Homo- and heterodimers) | ENO1 (ENOα) | 1p36.2 | 2 (1) | 0.03 | xxx | Almost all human tissue | 35 | 31 | 31 | 0 |
| ENO2(ENOγ) | 12p13 | / | 0.03 | x | Neuron and neuroendocrine tissues | 14 | 1 | 3 | 0 | |
| ENO3(ENOß) | 17p13.2 | 3 | 0.03 | Muscle tissues | 15 | 1 | 14 | 0 | ||
| PK (Active tetrameric form and nearly inactive dimeric) |
PKM | 15q22 | 3 (1) | 0.03-0.5 (PKM2) / 0.033 PKM1) |
xxx | All somatic tissues (The M1 isoform mostly in muscle and brain, and the M2 isoform during embryogenesis, in adipose tissue and pancreas) |
42 | 14 | 29 | 1 |
| PKLR | 1q21 | 2 (1) | 0.3 (PKR) | Liver, kidney, small intestine and erythrocytes | 6 | 0 | 1 | 0 | ||
Isoform and spliceform information was collected from Uniprot (version 144)[151], chromosome locations from the NCBI website[152] and posttranslational modification sites from the phosphositeplus database, (release date May 5 2014) [44]. Kinetic information was gathered from: HK[81, 153, 154], GPI[155], PFK1[133], PFK2[156], ALDO[157, 158], TPI1[155], GAPDH[155], PGK[155],PGAM[155], ENO[159], PK[160, 161]. Tissue specificity: HK[81, 153, 154], GPI, PFK1[133], PFK2[91, 156, 162], ALDO[158], TPI1, GAPDH[163], PGK,PGAM[164], ENO[159, 165, 166], PK[114].
Figure 1. Heterogeneity for glycolytic enzymes.
Heterogeneity on glycolytic enzymes can be imposed by differences in gene encoding, spliceforms and posttranslational modifications. This heterogeneity can alter kinetic and functional properties of the enzymes and also modify their response to various allosteric interactions and anti-glycolytic agents. Abbreviations: HK, hexokinase; BrPyr, 3-bromopyruvic acid; GAPDH, glyceraldehyde 3-phosphate dehydrogenase; G6P, glucose-6-phosphate; PFK, phosphofructokinase; PFKFB, 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase; 3PG, 3-phosphoglycerate; 2PG, 2-phosphoglycerate; PEP, phosphoenolpyruvate; PYR, pyruvate; PGAM, phosphoglycerate mutase; F16P, fructose-1,6-bisphosphate; F6P, fructose-6-phosphate, F26P, fructose-2,6-bisphosphate; PK, pyruvate kinase; ENO, Enolase; ALDO, Aldolase; PGI, Glucose-6-phosphate isomerase; GLC, glucose; TPI, triosephosphate isomerase; PGK, phosphoglycerate kinase; LDH, lactate dehydrogenase; LAC, lactate; DHAP, dihydroxyacetone phosphate; G3P, glyceraldehyde 3-phosphate
Together the extent of molecular diversity within glycolysis (see figure 1) offers the ability to rapidly tune enzyme kinetics to adapt to given environmental demands. In addition, this diversity also allows for a tremendously heterogeneous array of possibilities for flux through glycolysis and its control.
3. Therapeutic opportunities in glycolysis
Targeting glucose metabolism can occur both through altering systemic metabolism and through directly targeting enzymes in the diseased cell. For systemic metabolism examples include suppressing of hepatic gluconeogenesis by anti-diabetic agents like metformin[46-48], or the administration of the ketogenic diet [49-51] which enforces increased reliance on lipid oxidation as an energy source. Targeting glucose metabolism directly in the tumor is an alternative to this approach and aims to alter the biosynthetic, bioenergetic, and signaling processes that are differentially occurring as a result of the enhanced cell autonomous glucose metabolism[2, 52]. Both approaches are complementary and are actively being explored in the clinic[49, 53-55].
One approach to target glycolysis directly is to consider a single anti-glycolytic agent that targets a single enzyme with high specificity. This approach is promising in that it would offer potentially less toxicity but leads to difficulties in establishing whether there is sufficient efficacy, especially if tumors can develop resistance by expressing an alternate isoform of the targeted enzyme or establishing some other bypass mechanism. An additional approach is to consider anti-glycolytic agents that have multiple targets, both within the glycolysis pathway and in other pathways [56]. In response to this, some have argued for the development of “dirty drugs” or cocktails of specific molecules that are able to simultaneously target multiple nodes within the network and will be more effective[57, 58].
Often it is difficult to classify (anti-glycolytic) agents clearly into either class because specificity of small molecules for their proposed targets is usually dose dependent[59]. Moreover, identifying off targets effects or exact mechanism of action within a cell is also impeded by technical limitations[59, 60]. In the light of this uncertainty, it is interesting to describe the research into the mechanism of action of 3-Bromopyruvic acid (BrPyr). Originally identified as a an inhibitor of hexokinase, BrPyr is a small molecule and because of its structural homology has been described as a lactate and pyruvate homologue[61]. It is also an alkylating agent subject to nucleophilic attack and displacement of the alkyl bromide by a broad class of compounds[61]. Because of these properties BrPyr is highly reactive and therefore probably acts on multiple targets, although RNA/DNA alkylation, a common feature seen in most alkylating agent used in cancer therapy, has to our knowledge not been observed[62]. Three of the most frequently proposed targets of BrPyr include HK2 (hexokinase 2), glyceraldehyde 3-phosphate dehydrogenase (GAPDH) and succinate dehydrogenase complex A (SDHA)[62]. Pre-clinical studies have shown impressive results in eradicating tumors[61, 63, 64] and clinical trials are ongoing. Part of this efficacy might be explained that BrPyr alkylates predominantly the thiol group of cysteine that is often part of the catalytic mechanism of many metabolic enzymes but also mediates the stability of protein bridges through the formation of cysteine bridges. This highly reactive nature suggests that once localized in cells, BrPyr may function non selectively and independent of context. However, its transport into cells could be selectively mediated by the lactate transporter monocarboxylate transporter 1 (MCT1) as has been proposed. In this scenario, expression of MCT1 would serve as a biomarker for the predictive response of the anti-tumor effects of BrPyr[65]. Moreover, the overall low expression of MCT1 in normal cells could explain its apparent low level of toxicity in preclinical models. The generality of this model is unclear and the possibility nevertheless remains that BrPyr exerts cytotoxicity through its combinatorial targeting of defined enzymes in glycolysis and functions to selectively inhibit defined states of glucose metabolism that are observed in tumor cells. Of note, current chemotherapies used in the clinic disrupt a large number of biological processes that sustain cancer development and growth[66-71]. However, these therapies although efficacious, are also highly toxic[72-74]. Uptake of BrPyr by MCT1 could also seem in contradiction with the high glycolytic rate and lactate secretion associated with the Warburg effect, since MCT1 is a lactate importer under certain glucose/oxygen concentrations. Recent research into lactate transporters has shown that certain tumors are made up of symbiological populations where some fractions secrete lactate via the MCT4 exporter, while other fractions take up lactate as an energy source via the MCT1 importer[75, 76]. This process is thought to depend on oxygen and glucose concentration gradients within the tumor whereby MCT1 positive cells can develop the ability to migrate towards areas of increased glucose concentrations[75]. These subpopulations with different levels of MCT1/4 expression can therefore have implications not only for the efficacy of BrPyr treatment, but also whether drugs that are highly selective for either MCT1 or MCT4 or a pan MCT1/4 inhibitor will show the most efficacy[14].
Several outstanding reviews have described the use of anti-glycolytic agents focusing predominantly on targeting the glycolytic enzymes themselves[6, 7, 77]. Here we will focus more specifically on the heterogeneity of glycolysis and how different realizations of glycolysis might have targetable liabilities. In this review we will discuss this apparent heterogeneity for key glycolytic enzymes, its impact on enzymatic functionality and the use of some promising anti-glycolytic agents.
4. Targeting heterogeneous glycolytic enzymes
4.1. Hexokinases
For hexokinase (HK) both HK1 and HK2 isoenzymes are expressed in cancer cells, but HK2 appears more utilized in tumors[78-80]. All three HKs are sensitive to product inhibition by G6P. Interestingly, for HK2 and HK3, product inhibition is enhanced by elevated levels of inorganic phosphate(Pi), while for HK1, Pi has an antagonizing effect on product inhibition[81]. Patra et. al. showed that specific depletion of the HK2 isoform in tumor has been shown to reduce tumor growth in preclinical models[79]. The authors further demonstrated that a systemic depletion of HK2 was well tolerated, suggesting that HK2 specific inhibitors could target HK2 expressing tumors with potentially limited toxicity.
A number of small molecules have been proposed as inhibitors of hexokinases. Once considered a HK inhibitor, BrPyr has been shown to have many (metabolic) targets (see above), and research into its mechanisms of action are ongoing. Another well described hexokinase inhibitor is 2-deoxyglucose (2DG). 2DG is taken up through glucose transporters where it acts as a competitive inhibitor of hexokinases and gets metabolized into 2-DG-6-phosphate (2DG-6P). This inhibition results in decreases glycolytic flux, ATP-depletion and eventually cell death[82]. 2DG-6P, a phosphorylated molecule that does not diffuse through the plasma membrane, cannot be metabolized by glucose-6-phosphate isomerase (PGI) or through the pentose phosphate pathway. The subsequent buildup of 2DG-6P, a reactive phosphate, also has toxic effects on the cell. 2DG-6P buildup continues also in part since the Ki for product inhibition of 2DG-6P for hexokinase is much larger compared to G6P[83].
Currently to our knowledge no HK2 specific inhibitors have been developed but the differences in Km and protein sequences suggests this could be feasible. HK1 and HK2 also behave differently towards product inhibition in relation to concentrations of inorganic phosphate[79, 81].
4.2. Phosphofructokinases (PFK’s) and 6-phosphofructo-2-kinase/fructose-2,6-biphosphatases (PFKFB’s)
Phosphofructokinase catalyzes the phosphorylation of fructose-6-phosphate(F6P) to fructose-1,6-bisphosphate (F16P) and represents the second irreversible step in glycolysis. Phosphofructokinases are transcribed form three different genes and form a homo- or hetero tetrameric complexes called PFK1[84]. Phosphofructokinases display some of the most complex regulation within glycolysis. PFK1 is activated by fructose-2,6-bisphosphate (F26P), ADP/AMP and ammonium ions[85] and inhibited by ATP and citrate. F26P is synthesized by tetrameric complexes called PFK2 of which the components are encoded by PFKFB genes[86]. Especially the activation by F26P has caught widespread attention as inhibition of PFKFB enzymes can also down regulate glycolysis[87]. It’s interesting to note here that ammonia is an allosteric activator of PFK1[88]. Ammonia can be derived from deamination of glutamine, an amino acid that is taken up in large quantities by many tumor types and is one of the main components of cell culture media[89].
The PFK1 subunits are transcribed from three different genes called PFKM(Muscle), PFKL(liver) and PFKP(platelet). While the PFKM is the only isoform to be tissue specific, the other three isoforms have been found to show distinct expression in other tissues[90]. The three isoforms share a relatively low sequence similarity between 66 and 70%. Although substrate affinities are relatively similar (see table 1), the isoforms have marked difference in response to the allosteric activator by F26P. Changes in PFK1 isoform composition have been associated with cancer progression[22, 23].
For PFKFB’s, the PFKFB4 and especially the PFKFB3 isoforms seem to be the most relevant for cancer tissues[91] and are also induced by hypoxia[92]. One of the most well studied PFKFB3 inhibitors is 3-PO[93, 94], which has been shown to lower F-2,6-BP and the glucose uptake in cancer cells. 3-PO was developed using computational methods in combination with a PFKFB3 enzymatic assay and appears to be specific for PFKFB3, but the extent of this specificity warrants further investigation[91, 94]. Recent research into PFKFB3 provides us with an example of how this isoenzyme is not only relevant for cancer cells but is also involved in angiogenesis along the VEGF gradient[95, 96]. Quiescent endothelial cells (ECs) already have a high baseline level of glycolysis likely functioning to avoid diversion of oxygen away from tissues that require oxygenation[97]. During vessel sprouting however, activated endothelial cells upregulate glycolysis to meet the demands cell proliferation and migration. Blockade of PFKFB3 by 3-PO was also able to reduce vessel sprouting. Intriguingly, this blockade was also able to abrogate the residual hyper-branching of proliferating ECs after loss of directed growth along the VEGF gradient due to Notch and VEGFR2 inhibition[96]. Such a synergism holds promise for more effective angiogenic therapy, especially taking into account the relative low toxicity profile of 3-PO.
PFK’s also undergo acetylations and phosphorylation’s[44]. PFKFB3 activity can be enhanced through a Ser461 phosphorylation by activated AMPK. AMPK activation and its role in cancer has an expansive literature described in detail elsewhere[98, 99]. PFKFB3 is also phosphorylated by protein kinase A (PKA) at Ser461, enhancing its kinase activity and increasing the activity of PFK1[100]. Acetylations have been shown to destabilize or inactivate a number of glycolytic enzymes. This could explain the strong synergistic effect found between the glycolytic inhibitor 2-DG and Lysine-(K)-deacetylase (KDAC) inhibitors[101]. Protein kinase C (PKC) phosphorylation of PFKFB3 at Ser461 (the same site as AMPK and PKA phosphorylation) also induces up-regulation of its kinase activity and activation of PFK1[100]. PFK1 can also be activated through Akt kinase phosphorylation of PGK2[91, 102]. Akt also phosphorylates and activates the glycolytic enzymes HK and glucose transporter 1 (GLUT1)[103, 104]. Phosphorylation of PFK1’s could interfere with allosteric inhibition by lactate and has been proposed as a driver for the Warburg effect, as it abrogates one the main negative feedback loops for fermentative glycolysis[90]. Yi et. al. recently described a O-linked β-N-acetylglucosamine (O-GlcNAc) modification on a serine residues (Ser529) for PFKM which is induced under hypoxia or glucose deprivation[105]. They further demonstrated that this modification is dynamically controlled by its substrate, uridine diphospho-N-acetylglucosamine (UDP-GlcNAc), a key integrator of several biosynthetic and bioenergetic pathways. This modification was shown to inhibit PFK1 activity and redirected glycolytic flux into the biosynthetic pentose phosphate pathway. This finding is especially intriguing, since PFKM (85-kDa) has been shown to be posttranslationally cleaved, resulting in a constitutively active 47-kDa N-terminal fragment[106]. The cleavage of the C-terminal domain not only contains the sites for allosteric inhibition with citrate and ATP, but also the site for inhibition by O-GlcNAc modification on Ser529[107]. PFKM also gets phosphorylated during mitosis on Ser667, which is also located on the cleaved off C-terminal, and would thus also remove any regulation imposed on PFKM during mitosis[44]. For O-GlcNAc modifications only one enzyme has been discovered so far, namely O-linked N-acetylglucosamine transferase (OGT)[108]. OGT inhibitors could therefore provide a suitable synergistic drug for anti-glycolysis treatment. Both glucose and glutamine participate in UDP-GlcNAc synthesis. How the increased uptake of both these nutrients relates to O-GlcNAc modification is currently not known, but it has been postulated that decreased activity of several glycolytic enzymes by O-GlcNAc modification diverts part of the glycolytic flux into anabolic pathways[109].
4.3. Phosphoglycerate mutases
Phosphoglycerate mutases (PGAM’s) catalyze the reversible reaction of 3-phosphoglycerate (3PG) to 2-phosphoglycerate (2-PG) but has also been implicated in the conversion of phosphoenolpyruvate (PEP) to pyruvate (PYR)[110]. This transfer of PEP to a histidine residue (His11) in the catalytic site actually increases the mutase function of PGAM1. This alternative conversion is thought to occur more efficiently at higher PEP concentrations generated by the less active downstream PKM2 spliceforms the PKM enzyme, the main converter of PEP to pyruvate in cells. Such a mechanism maintains glycolytic flux, but ensures a buildup of glycolytic intermediates needed for anabolic pathways.
PGAM is acetylated and enhances it enzymatic activity[111]. Its substrate, 3PG can inhibit the pentose phosphate pathway [112, 113]. Thus increased priming by high PEP levels might ensure a more efficient catalytic turnover that balances the anabolic fluxes into the pentose phosphate pathway and serine synthesis. This regulatory behavior therefore makes PGAM1 a suitable target to inhibit cancer metabolism. A recently developed small molecule inhibitor, PGMI-004A resulted in increased 3PG and decreased 2PG levels in cancer cells, which resulted in significantly decreased glycolysis, pentose phosphate pathway flux and biosynthesis, as well as attenuated cell proliferation and tumor growth[113].
4.4. Pyruvate Kinases
Pyruvate kinases (PK) perform the last major step in glycolysis whereby PEP is transformed into pyruvate together with the generation of one ATP molecule from ADP. PK are transcribed from two different gene locus’s that are designated by the gene names PKLR and PKM. PKLR gives rise to the liver and erythrocyte specific enzymes, whereby the differential expression among the two cell types is accomplished by N-terminal splicing[114]. Whether PKLR isoforms contribute to tumor glycolysis is difficult to assess[114], but the similar Km of PKLR and PKM2 (see table 1) suggest this could be possible. The PKM enzyme is transcribed from a single gene locus located on chromosome 15 but has two well described isoforms that are of major importance to cancer metabolic specific adaptations. The isoform occurring in most normal tissues is PKM1 and has the lowest Km for PEP. PKM1 form highly constitutively active tetramers protein complexes. PKM2 (in its dimeric form) on the other hand has a lower affinity for PEP and is almost universally expressed in cancer cell alongside PKM1. Besides cancer specific expression, PKM2 has several other properties that set it aside from PKM1 and could make it an interesting target to interfere with cancer metabolism. PKM2 has allosteric binding sites that are regulated by the glycolytic intermediate as F16BP[115]. More recently, PKM2 activation has been linked to serine biosynthesis, when it was shown that PKM2 can be activated by serine[116] and a small molecule intermediate from the purine synthesis pathway called succinyl-5-aminoimidazole-4-carboxamide-1-ribose-5′-phosphate (SAICAR)[117].
The requirements of pyruvate kinase activity and expression of any given isoform is thus a complex subject. Strikingly, tumors were shown to grow in complete absence of activity[118] and deletion of PKM2 could in some instances accelerate tumor growth[119]. In addition, activators of PKM2, like TEP-46[120], have been shown to reduce tumor growth. Together, these findings indicate that decreased flux through the pyruvate kinase step in glycolysis could have selective advantages for tumor growth. Because PKM2 has lower catalytic activity, it is thought to build up glycolytic intermediates that can be used for anabolic pathways. Activation of PKM thus depletes these intermediates and results in reduced cell proliferation. Buildup of anabolic intermediates has recently been identified a consequence of a newly identified PTM that seems especially prevalent on glycolytic enzymes. What is especially intriguing is that this PTM is established by 1,3-biphosphoglycerate (13BPG), the product of GAPDH[121]. Accumulation of highly reactive 1,3-BPG results in modification of lysine to lysine-phosphoglycerate (Kpg). This modification not only results in auto-modification of GAPDH itself, but also in modification of the glycolytic enzymes PGAM1, enolase 1 (ENO1) and PKM2. This modification can then lead to significant alterations in enzyme activity and excessive buildup of glycolytic intermediates. How Kpg modification alters PKM2 activity is currently not know but since exon10 is thought to be involved in the formation of dimer and tetramers, there could be an effect on enzyme kinetics.
5. Conclusions and future perspectives
The large heterogeneity in glycolytic enzymes offered by different genes, spliceforms and posttranslational modifications gives cancer cells an array of tools to react differently upon treatment with anti-glycolytic agents. This heterogeneity also makes it possible to give marked different responses to allosteric regulators that originate within glycolysis and other metabolic pathways and have consequences for drugs that target these allosteric sites. Lastly, this heterogeneity also provides cancer cells with rapid switches between the bioenergetics, signaling and biosynthetic needs. Current technical and computational advances in the assessment of glycolytic heterogeneity hold promise to better target this heterogeneity and thereby improve therapeutic outcomes.
5.1. Assessment of the heterogeneity in glycolytic enzymes and its dynamic properties
Several reports on differential enzyme expression between different tumor types have already surfaced indicating their relevance in cancer metabolism. Most of these studies have used transcription profiling [21] or protein expression assays[122, 123]. Although we have some knowledge of how these isoenzymes are expressed in different normal tissues (see table 1), our understanding of how these isoenzymes are expressed between different tumors or within a tumor is often not well defined. Also the large number of (predicted) spliceforms has seen relatively little functional characterization. With the advent of next-generation sequencing[124] and (targeted) proteomics[122, 125], a better understanding of the quantitative nature of isoenzyme expression is to be expected and could serve as basis for biomarker discovery for anti-glycolytic agents. However, the relevant phenotypic consequence of the alteration in mRNA and protein expression is a change in metabolism that is characterized by changes in metabolite concentrations and fluxes[126, 127]. Several in vivo imaging strategies are now commonplace in assessing changes in flux through glycolysis by using positron emission tomography with radioactive glucose (FDG-PET)[128] or by advanced technologies in that exploit the properties of heavy labeled, hyperpolarized nutrients such as choline, acetate, and glutamine[129, 130]. High resolution mass spectrometry offers additional possibilities for characterizing the state of glycolysis in a tumor through either profiling with metabolomics the state of a biopsy or a biofluid such as plasma or sera[131].
Better kinetic characterization of glycolytic enzymes especially when isoforms are expressed in parallel can be helpful to elucidate the dynamic behavior of glycolysis in various (tumor) cell types though mathematical modeling using techniques such as Metabolic Control Analysis (MCA)[132, 133] and can find use in in identifying better drug targets[134, 135].
A better understanding of the influence and functionality of isoenzyme expression can also be attained by flux studies using isotope tracers[136]. Furthermore understanding concentration patterns of metabolites in glycolysis in the context of basal rates of glycolysis and their changes upon partial inhibition by therapeutics will serve as a helpful guideline for therapeutic efficacy[137]. In addition excretory fluxes that may be altered as a result of manipulating fluxes in glycolysis could result in the production of metabolites that could be measured in tumors or in biofluids. The measurement of these molecules could allow for amplification of signals that are derived from the internal state of glycolysis. Such measurements may be useful for identifying biomarkers that so far have been difficult to identify. Together such studies could also determine drug efficacy[138, 139], highlight rewiring in metabolic pathways in sensitive and resistant tumor cells and provide crucial metabolic biomarkers of drug efficacy[140].
We will also need to make use of cell culture and in vivo models that are more representative of the spatial and temporal differences in pH, pO2 and nutrient and hormonal concentrations that exist in tumors and different tissues. Recent studies have shown that under these conditions tumor (subpopulations) exist with marked heterogeneity in isoenzyme expression and flux[76, 141, 142].
5.2. Opportunities for targeting glycolytic heterogeneity
The techniques described above to resolve the quantitative and dynamic properties of glycolytic heterogeneity holds promise to improve therapeutic intervention. By providing insight into the proliferation strategies of cells within the tumor environment opportunities exist to highlight points of liability that current anti-glycolytic agents and/or systemic strategies could exploit[143]. Currently it is not well established what region of a solid tumor should be targeted by anti-glycolysis agents. One option is to target the well oxygenated/high glucose vascularized component of the tumor[75, 144], but additional focus on targeting cells in nutrient deprived areas such as the hypoxic core of the tumor has also been considered[145]. Differences in pH and nutrient availability are essential parameters for controlling flux through glycolysis and these properties affect both the pharmacological properties of the compound[146] and the control properties of glycolysis[141]. Other variables including genetic status and availability of growth factors are also likely key mediators of the regional response to targeting glycolysis.
Additional insight will also be gained by a better understanding of glycolytic heterogeneity in tumor associated physiological process like angiogenesis[95, 96] and inhibition of immune activation[147, 148]. Moreover, by the use of integrative approaches that combine experimental and computational techniques, a better understanding of tumor and organism interactions is to be expected[149, 150]
Currently almost every enzyme or transporter involved in glycolysis is subject of intensive research as an anti-cancer target by small molecules[77]. It is now becoming clear that the expression of different isoenzymes and regulation of their PTMs could have profound influences on the efficacy of these drugs. Only a minority of this glycolytic heterogeneity has been functionally characterized and provides opportunities for more optimal targeting of cancer metabolism and biomarker development for current therapeutic strategies.
Acknowledgements
None of the authors have any financial conflict of interest. JWL is supported by awards R00CA168997 and 401 R01AI110613 from the National Institutes of Health and a Future Leader award from the 402 International Life Sciences Institute.
Abbreviations
- UDP-GlcNAc
uridine diphospho-N-acetylglucosamine
- HK
hexokinase
- BrPyr
3-bromopyruvic acid
- EC
endothelial cells
- MCT
monocarboxylate transporter
- 2DG
2-deoxyglucose
- F26P
fructose-2,6-bisphosphate
- PFKFB
6-phosphofructo-2-kinase/fructose-2,6-biphosphatase
- O-GlcNAc
O-linked β-N-acetylglucosamine
- OGT
O-linked N-acetylglucosamine (GlcNAc) transferase
- AMPK
AMP-activated protein kinase
- PKA
protein kinase A
- PKC
protein kinase C
- GLUT1
glucose transporter 1
- SER
serine
- PPP
pentose phosphate pathway
- SAICAR
succinyl-5-aminoimidazole-4-carboxamide-1-ribose-5′-phosphate
- PTM
posttranslational modification
- Kpg
lysine-phosphoglycerate
- 13BGP
1,3-biphosphoglycerate
- MCA
metabolic control analysis
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. 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.
Chemical Compounds studied in this article:
Bromopyruvate (PubChem CID: 70684); 2-deoxyglucose (PubChem CID: 9062); 3-PO (PubChem CID: 5720233); PGMI-004A (PubChem CID: 66521681); TEP-46 (Chem CID: 44246499)
References
- [1].Lunt SY, Vander Heiden MG. Aerobic glycolysis: meeting the metabolic requirements of cell proliferation. Annual review of cell and developmental biology. 2011;27:441–64. doi: 10.1146/annurev-cellbio-092910-154237. [DOI] [PubMed] [Google Scholar]
- [2].Locasale JW. The consequences of enhanced cell-autonomous glucose metabolism. Trends in endocrinology and metabolism: TEM. 2012;23:545–51. doi: 10.1016/j.tem.2012.07.005. [DOI] [PubMed] [Google Scholar]
- [3].Levine AJ, Puzio-Kuter AM. The control of the metabolic switch in cancers by oncogenes and tumor suppressor genes. Science. 2010;330:1340–4. doi: 10.1126/science.1193494. [DOI] [PubMed] [Google Scholar]
- [4].Bailey KM, Wojtkowiak JW, Hashim AI, Gillies RJ. Targeting the metabolic microenvironment of tumors. Advances in pharmacology. 2012;65:63–107. doi: 10.1016/B978-0-12-397927-8.00004-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [5].Michelakis ED, Sutendra G, Dromparis P, Webster L, Haromy A, Niven E, et al. Metabolic modulation of glioblastoma with dichloroacetate. Science translational medicine. 2010;2:31ra4. doi: 10.1126/scitranslmed.3000677. [DOI] [PubMed] [Google Scholar]
- [6].Pelicano H, Martin DS, Xu RH, Huang P. Glycolysis inhibition for anticancer treatment. Oncogene. 2006;25:4633–46. doi: 10.1038/sj.onc.1209597. [DOI] [PubMed] [Google Scholar]
- [7].Galluzzi L, Kepp O, Vander Heiden MG, Kroemer G. Metabolic targets for cancer therapy. Nature reviews Drug discovery. 2013;12:829–46. doi: 10.1038/nrd4145. [DOI] [PubMed] [Google Scholar]
- [8].Dwarakanath B, Jain V. Targeting glucose metabolism with 2-deoxy-D-glucose for improving cancer therapy. Future oncology. 2009;5:581–5. doi: 10.2217/fon.09.44. [DOI] [PubMed] [Google Scholar]
- [9].Maschek G, Savaraj N, Priebe W, Braunschweiger P, Hamilton K, Tidmarsh GF, et al. 2-deoxy-D-glucose increases the efficacy of adriamycin and paclitaxel in human osteosarcoma and non-small cell lung cancers in vivo. Cancer research. 2004;64:31–4. doi: 10.1158/0008-5472.can-03-3294. [DOI] [PubMed] [Google Scholar]
- [10].Warmoes M, Jaspers JE, Xu G, Sampadi BK, Pham TV, Knol JC, et al. Proteomics of genetically engineered mouse mammary tumors identifies fatty acid metabolism members as potential predictive markers for cisplatin resistance. Molecular & cellular proteomics : MCP. 2013;12:1319–34. doi: 10.1074/mcp.M112.024182. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [11].Xie J, Wang BS, Yu DH, Lu Q, Ma J, Qi H, et al. Dichloroacetate shifts the metabolism from glycolysis to glucose oxidation and exhibits synergistic growth inhibition with cisplatin in HeLa cells. International journal of oncology. 2011;38:409–17. doi: 10.3892/ijo.2010.851. [DOI] [PubMed] [Google Scholar]
- [12].Locasale JW. Serine, glycine and one-carbon units: cancer metabolism in full circle. Nature reviews Cancer. 2013;13:572–83. doi: 10.1038/nrc3557. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [13].Vander Heiden MG. Targeting cancer metabolism: a therapeutic window opens. Nature reviews Drug discovery. 2011;10:671–84. doi: 10.1038/nrd3504. [DOI] [PubMed] [Google Scholar]
- [14].Doherty JR, Cleveland JL. Targeting lactate metabolism for cancer therapeutics. The Journal of clinical investigation. 2013;123:3685–92. doi: 10.1172/JCI69741. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [15].Webster KA. Evolution of the coordinate regulation of glycolytic enzyme genes by hypoxia. The Journal of experimental biology. 2003;206:2911–22. doi: 10.1242/jeb.00516. [DOI] [PubMed] [Google Scholar]
- [16].Herman MA, Kahn BB. Glucose transport and sensing in the maintenance of glucose homeostasis and metabolic harmony. The Journal of clinical investigation. 2006;116:1767–75. doi: 10.1172/JCI29027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [17].Hue L, Beauloye C, Marsin AS, Bertrand L, Horman S, Rider MH. Insulin and ischemia stimulate glycolysis by acting on the same targets through different and opposing signaling pathways. Journal of molecular and cellular cardiology. 2002;34:1091–7. doi: 10.1006/jmcc.2002.2063. [DOI] [PubMed] [Google Scholar]
- [18].Eelen G, Cruys B, Welti J, De Bock K, Carmeliet P. Control of vessel sprouting by genetic and metabolic determinants. Trends in endocrinology and metabolism: TEM. 2013;24:589–96. doi: 10.1016/j.tem.2013.08.006. [DOI] [PubMed] [Google Scholar]
- [19].Pearce EL, Poffenberger MC, Chang CH, Jones RG. Fueling immunity: insights into metabolism and lymphocyte function. Science. 2013;342:1242454. doi: 10.1126/science.1242454. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [20].Ito K, Suda T. Metabolic requirements for the maintenance of self-renewing stem cells. Nature reviews Molecular cell biology. 2014;15:243–56. doi: 10.1038/nrm3772. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [21].Hu J, Locasale JW, Bielas JH, O’Sullivan J, Sheahan K, Cantley LC, et al. Heterogeneity of tumor-induced gene expression changes in the human metabolic network. Nature biotechnology. 2013;31:522–9. doi: 10.1038/nbt.2530. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [22].Wang G, Xu Z, Wang C, Yao F, Li J, Chen C, et al. Differential phosphofructokinase-1 isoenzyme patterns associated with glycolytic efficiency in human breast cancer and paracancer tissues. Oncology letters. 2013;6:1701–6. doi: 10.3892/ol.2013.1599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [23].Vora S, Halper JP, Knowles DM. Alterations in the activity and isozymic profile of human phosphofructokinase during malignant transformation in vivo and in vitro: transformation- and progression-linked discriminants of malignancy. Cancer research. 1985;45:2993–3001. [PubMed] [Google Scholar]
- [24].Marjanovic S, Skog S, Heiden T, Tribukait B, Nelson BD. Expression of glycolytic isoenzymes in activated human peripheral lymphocytes: cell cycle analysis using flow cytometry. Experimental cell research. 1991;193:425–31. doi: 10.1016/0014-4827(91)90116-c. [DOI] [PubMed] [Google Scholar]
- [25].Richard H, Schulz MH, Sultan M, Nurnberger A, Schrinner S, Balzereit D, et al. Prediction of alternative isoforms from exon expression levels in RNA-Seq experiments. Nucleic acids research. 2010;38:e112. doi: 10.1093/nar/gkq041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [26].Johnson C, Warmoes MO, Shen X, Locasale JW. Epigenetics and cancer metabolism. Cancer letters. 2013 doi: 10.1016/j.canlet.2013.09.043. In Press, Corrected Proof. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [27].Prigione A, Rohwer N, Hoffmann S, Mlody B, Drews K, Bukowiecki R, et al. HIF1alpha modulates cell fate reprogramming through early glycolytic shift and upregulation of PDK1-3 and PKM2. Stem cells. 2014;32:364–76. doi: 10.1002/stem.1552. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [28].Singh PK, Mehla K, Hollingsworth MA, Johnson KR. Regulation of Aerobic Glycolysis by microRNAs in Cancer. Molecular and cellular pharmacology. 2011;3:125–34. [PMC free article] [PubMed] [Google Scholar]
- [29].Kornfeld JW, Bruning JC. Regulation of metabolism by long, non-coding RNAs. Frontiers in genetics. 2014;5:57. doi: 10.3389/fgene.2014.00057. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [30].Chen M, Zhang J, Manley JL. Turning on a fuel switch of cancer: hnRNP proteins regulate alternative splicing of pyruvate kinase mRNA. Cancer research. 2010;70:8977–80. doi: 10.1158/0008-5472.CAN-10-2513. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [31].Yamamoto T, Takano N, Ishiwata K, Ohmura M, Nagahata Y, Matsuura T, et al. Reduced methylation of PFKFB3 in cancer cells shunts glucose towards the pentose phosphate pathway. Nature communications. 2014;5:3480. doi: 10.1038/ncomms4480. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [32].Doulias PT, Tenopoulou M, Greene JL, Raju K, Ischiropoulos H. Nitric oxide regulates mitochondrial fatty acid metabolism through reversible protein S-nitrosylation. Science signaling. 2013;6:rs1. doi: 10.1126/scisignal.2003252. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [33].Yang M, Soga T, Pollard PJ. Oncometabolites: linking altered metabolism with cancer. The Journal of clinical investigation. 2013;123:3652–8. doi: 10.1172/JCI67228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [34].Peng C, Lu Z, Xie Z, Cheng Z, Chen Y, Tan M, et al. The first identification of lysine malonylation substrates and its regulatory enzyme. Molecular & cellular proteomics : MCP. 2011;10:M111–012658. doi: 10.1074/mcp.M111.012658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [35].Wong BW, Kuchnio A, Bruning U, Carmeliet P. Emerging novel functions of the oxygen-sensing prolyl hydroxylase domain enzymes. Trends in biochemical sciences. 2013;38:3–11. doi: 10.1016/j.tibs.2012.10.004. [DOI] [PubMed] [Google Scholar]
- [36].Seo J, Jeong J, Kim YM, Hwang N, Paek E, Lee KJ. Strategy for comprehensive identification of post-translational modifications in cellular proteins, including low abundant modifications: application to glyceraldehyde-3-phosphate dehydrogenase. Journal of proteome research. 2008;7:587–602. doi: 10.1021/pr700657y. [DOI] [PubMed] [Google Scholar]
- [37].Mustafa AK, Gadalla MM, Sen N, Kim S, Mu W, Gazi SK, et al. H2S signals through protein S-sulfhydration. Science signaling. 2009;2:ra72. doi: 10.1126/scisignal.2000464. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [38].Locasale JW, Cantley LC. Metabolic flux and the regulation of mammalian cell growth. Cell metabolism. 2011;14:443–51. doi: 10.1016/j.cmet.2011.07.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [39].Slawson C, Hart GW. O-GlcNAc signalling: implications for cancer cell biology. Nature reviews Cancer. 2011;11:678–84. doi: 10.1038/nrc3114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [40].Robey RB, Hay N. Is Akt the “Warburg kinase”?-Akt-energy metabolism interactions and oncogenesis. Seminars in cancer biology. 2009;19:25–31. doi: 10.1016/j.semcancer.2008.11.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [41].Yang W, Xia Y, Ji H, Zheng Y, Liang J, Huang W, et al. Nuclear PKM2 regulates beta-catenin transactivation upon EGFR activation. Nature. 2011;480:118–22. doi: 10.1038/nature10598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [42].Tristan C, Shahani N, Sedlak TW, Sawa A. The diverse functions of GAPDH: views from different subcellular compartments. Cellular signalling. 2011;23:317–23. doi: 10.1016/j.cellsig.2010.08.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [43].Tarze A, Deniaud A, Le Bras M, Maillier E, Molle D, Larochette N, et al. GAPDH, a novel regulator of the pro-apoptotic mitochondrial membrane permeabilization. Oncogene. 2007;26:2606–20. doi: 10.1038/sj.onc.1210074. [DOI] [PubMed] [Google Scholar]
- [44].Hornbeck PV, Kornhauser JM, Tkachev S, Zhang B, Skrzypek E, Murray B, et al. PhosphoSitePlus: a comprehensive resource for investigating the structure and function of experimentally determined post-translational modifications in man and mouse. Nucleic acids research. 2012;40:D261–70. doi: 10.1093/nar/gkr1122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [45].Oliveira AP, Sauer U. The importance of post-translational modifications in regulating Saccharomyces cerevisiae metabolism. FEMS yeast research. 2012;12:104–17. doi: 10.1111/j.1567-1364.2011.00765.x. [DOI] [PubMed] [Google Scholar]
- [46].Quinn BJ, Kitagawa H, Memmott RM, Gills JJ, Dennis PA. Repositioning metformin for cancer prevention and treatment. Trends in endocrinology and metabolism: TEM. 2013;24:469–80. doi: 10.1016/j.tem.2013.05.004. [DOI] [PubMed] [Google Scholar]
- [47].Niraula S, Dowling RJ, Ennis M, Chang MC, Done SJ, Hood N, et al. Metformin in early breast cancer: a prospective window of opportunity neoadjuvant study. Breast cancer research and treatment. 2012;135:821–30. doi: 10.1007/s10549-012-2223-1. [DOI] [PubMed] [Google Scholar]
- [48].Del Barco S, Vazquez-Martin A, Cufi S, Oliveras-Ferraros C, Bosch-Barrera J, Joven J, et al. Metformin: multi-faceted protection against cancer. Oncotarget. 2011;2:896–917. doi: 10.18632/oncotarget.387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [49].Schmidt M, Pfetzer N, Schwab M, Strauss I, Kammerer U. Effects of a ketogenic diet on the quality of life in 16 patients with advanced cancer: A pilot trial. Nutrition & metabolism. 2011;8:54. doi: 10.1186/1743-7075-8-54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [50].Allen BG, Bhatia SK, Buatti JM, Brandt KE, Lindholm KE, Button AM, et al. Ketogenic diets enhance oxidative stress and radio-chemo-therapy responses in lung cancer xenografts. Clinical cancer research : an official journal of the American Association for Cancer Research. 2013;19:3905–13. doi: 10.1158/1078-0432.CCR-12-0287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [51].Ho VW, Leung K, Hsu A, Luk B, Lai J, Shen SY, et al. A low carbohydrate, high protein diet slows tumor growth and prevents cancer initiation. Cancer research. 2011;71:4484–93. doi: 10.1158/0008-5472.CAN-10-3973. [DOI] [PubMed] [Google Scholar]
- [52].Dang CV. Links between metabolism and cancer. Genes & development. 2012;26:877–90. doi: 10.1101/gad.189365.112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [53].Thompson AM. Molecular pathways: preclinical models and clinical trials with metformin in breast cancer. Clinical cancer research : an official journal of the American Association for Cancer Research. 2014;20:2508–15. doi: 10.1158/1078-0432.CCR-13-0354. [DOI] [PubMed] [Google Scholar]
- [54].Woolf EC, Scheck AC. The Ketogenic Diet for the Treatment of Malignant Glioma. Journal of lipid research. 2014 doi: 10.1194/jlr.R046797. Epub ahead of print. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [55].Franciosi M, Lucisano G, Lapice E, Strippoli GF, Pellegrini F, Nicolucci A. Metformin therapy and risk of cancer in patients with type 2 diabetes: systematic review. PloS one. 2013;8:e71583. doi: 10.1371/journal.pone.0071583. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [56].Zhao Y, Butler EB, Tan M. Targeting cellular metabolism to improve cancer therapeutics. Cell death & disease. 2013;4:e532. doi: 10.1038/cddis.2013.60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [57].Moreno-Sanchez R, Saavedra E, Rodriguez-Enriquez S, Gallardo-Perez JC, Quezada H, Westerhoff HV. Metabolic control analysis indicates a change of strategy in the treatment of cancer. Mitochondrion. 2010;10:626–39. doi: 10.1016/j.mito.2010.06.002. [DOI] [PubMed] [Google Scholar]
- [58].Lu JJ, Pan W, Hu YJ, Wang YT. Multi-target drugs: the trend of drug research and development. PloS one. 2012;7:e40262. doi: 10.1371/journal.pone.0040262. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [59].Kamb A, Wee S, Lengauer C. Why is cancer drug discovery so difficult? Nature reviews Drug discovery. 2007;6:115–20. doi: 10.1038/nrd2155. [DOI] [PubMed] [Google Scholar]
- [60].Xie L, Xie L, Bourne PE. Structure-based systems biology for analyzing off-target binding. Current opinion in structural biology. 2011;21:189–99. doi: 10.1016/j.sbi.2011.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [61].Ko YH, Smith BL, Wang Y, Pomper MG, Rini DA, Torbenson MS, et al. Advanced cancers: eradication in all cases using 3-bromopyruvate therapy to deplete ATP. Biochemical and biophysical research communications. 2004;324:269–75. doi: 10.1016/j.bbrc.2004.09.047. [DOI] [PubMed] [Google Scholar]
- [62].Shoshan MC. 3-Bromopyruvate: targets and outcomes. Journal of bioenergetics and biomembranes. 2012;44:7–15. doi: 10.1007/s10863-012-9419-2. [DOI] [PubMed] [Google Scholar]
- [63].Gong L, Wei Y, Yu X, Peng J, Leng X. 3-Bromopyruvic acid, a hexokinase II inhibitor, is an effective antitumor agent on the hepatoma cells : in vitro and in vivo findings. Anti-cancer agents in medicinal chemistry. 2014;14:771–6. doi: 10.2174/1871520614666140416105309. [DOI] [PubMed] [Google Scholar]
- [64].Yun J, Rago C, Cheong I, Pagliarini R, Angenendt P, Rajagopalan H, et al. Glucose deprivation contributes to the development of KRAS pathway mutations in tumor cells. Science. 2009;325:1555–9. doi: 10.1126/science.1174229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [65].Birsoy K, Wang T, Possemato R, Yilmaz OH, Koch CE, Chen WW, et al. MCT1-mediated transport of a toxic molecule is an effective strategy for targeting glycolytic tumors. Nature genetics. 2013;45:104–8. doi: 10.1038/ng.2471. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [66].Ricci MS, Zong WX. Chemotherapeutic approaches for targeting cell death pathways. The oncologist. 2006;11:342–57. doi: 10.1634/theoncologist.11-4-342. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [67].Longley DB, Harkin DP, Johnston PG. 5-fluorouracil: mechanisms of action and clinical strategies. Nature reviews Cancer. 2003;3:330–8. doi: 10.1038/nrc1074. [DOI] [PubMed] [Google Scholar]
- [68].Abal M, Andreu JM, Barasoain I. Taxanes: microtubule and centrosome targets, and cell cycle dependent mechanisms of action. Current cancer drug targets. 2003;3:193–203. doi: 10.2174/1568009033481967. [DOI] [PubMed] [Google Scholar]
- [69].Fu D, Calvo JA, Samson LD. Balancing repair and tolerance of DNA damage caused by alkylating agents. Nature reviews Cancer. 2012;12:104–20. doi: 10.1038/nrc3185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [70].Nitiss JL. Targeting DNA topoisomerase II in cancer chemotherapy. Nature reviews Cancer. 2009;9:338–50. doi: 10.1038/nrc2607. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [71].Chabner BA, Roberts TG., Jr Timeline: Chemotherapy and the war on cancer. Nature reviews Cancer. 2005;5:65–72. doi: 10.1038/nrc1529. [DOI] [PubMed] [Google Scholar]
- [72].Sorrentino MF, Kim J, Foderaro AE, Truesdell AG. 5-fluorouracil induced cardiotoxicity: review of the literature. Cardiology journal. 2012;19:453–8. doi: 10.5603/cj.2012.0084. [DOI] [PubMed] [Google Scholar]
- [73].Hennenfent KL, Govindan R. Novel formulations of taxanes: a review. Old wine in a new bottle? Annals of oncology : official journal of the European Society for Medical Oncology / ESMO. 2006;17:735–49. doi: 10.1093/annonc/mdj100. [DOI] [PubMed] [Google Scholar]
- [74].Curigliano G, Cardinale D, Suter T, Plataniotis G, de Azambuja E, Sandri MT, et al. Cardiovascular toxicity induced by chemotherapy, targeted agents and radiotherapy: ESMO Clinical Practice Guidelines. Annals of oncology : official journal of the European Society for Medical Oncology / ESMO. 2012;23(Suppl 7):vii155–66. doi: 10.1093/annonc/mds293. [DOI] [PubMed] [Google Scholar]
- [75].De Saedeleer CJ, Porporato PE, Copetti T, Perez-Escuredo J, Payen VL, Brisson L, et al. Glucose deprivation increases monocarboxylate transporter 1 (MCT1) expression and MCT1-dependent tumor cell migration. Oncogene. 2013 doi: 10.1038/onc.2013.454. Epub ahead of print. [DOI] [PubMed] [Google Scholar]
- [76].Sonveaux P, Vegran F, Schroeder T, Wergin MC, Verrax J, Rabbani ZN, et al. Targeting lactate-fueled respiration selectively kills hypoxic tumor cells in mice. The Journal of clinical investigation. 2008;118:3930–42. doi: 10.1172/JCI36843. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [77].Granchi C, Minutolo F. Anticancer agents that counteract tumor glycolysis. ChemMedChem. 2012;7:1318–50. doi: 10.1002/cmdc.201200176. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [78].Patra KC, Hay N. Hexokinase 2 as oncotarget. Oncotarget. 2013;4:1862–3. doi: 10.18632/oncotarget.1563. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [79].Patra KC, Wang Q, Bhaskar PT, Miller L, Wang Z, Wheaton W, et al. Hexokinase 2 is required for tumor initiation and maintenance and its systemic deletion is therapeutic in mouse models of cancer. Cancer cell. 2013;24:213–28. doi: 10.1016/j.ccr.2013.06.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [80].Mathupala SP, Rempel A, Pedersen PL. Glucose catabolism in cancer cells: identification and characterization of a marked activation response of the type II hexokinase gene to hypoxic conditions. The Journal of biological chemistry. 2001;276:43407–12. doi: 10.1074/jbc.M108181200. [DOI] [PubMed] [Google Scholar]
- [81].Wilson JE. Isozymes of mammalian hexokinase: structure, subcellular localization and metabolic function. The Journal of experimental biology. 2003;206:2049–57. doi: 10.1242/jeb.00241. [DOI] [PubMed] [Google Scholar]
- [82].Kurtoglu M, Maher JC, Lampidis TJ. Differential toxic mechanisms of 2-deoxy-D-glucose versus 2-fluorodeoxy-D-glucose in hypoxic and normoxic tumor cells. Antioxidants & redox signaling. 2007;9:1383–90. doi: 10.1089/ars.2007.1714. [DOI] [PubMed] [Google Scholar]
- [83].Crane RK, Sols A. The non-competitive inhibition of brain hexokinase by glucose-6-phosphate and related compounds. The Journal of biological chemistry. 1954;210:597–606. [PubMed] [Google Scholar]
- [84].Staal GE, Kalff A, Heesbeen EC, van Veelen CW, Rijksen G. Subunit composition, regulatory properties, and phosphorylation of phosphofructokinase from human gliomas. Cancer research. 1987;47:5047–51. [PubMed] [Google Scholar]
- [85].Dunaway GA. A review of animal phosphofructokinase isozymes with an emphasis on their physiological role. Molecular and cellular biochemistry. 1983;52:75–91. doi: 10.1007/BF00230589. [DOI] [PubMed] [Google Scholar]
- [86].Mor I, Cheung EC, Vousden KH. Control of glycolysis through regulation of PFK1: old friends and recent additions. Cold Spring Harbor symposia on quantitative biology. 2011;76:211–6. doi: 10.1101/sqb.2011.76.010868. [DOI] [PubMed] [Google Scholar]
- [87].Yalcin A, Telang S, Clem B, Chesney J. Regulation of glucose metabolism by 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatases in cancer. Experimental and molecular pathology. 2009;86:174–9. doi: 10.1016/j.yexmp.2009.01.003. [DOI] [PubMed] [Google Scholar]
- [88].Abrahams SL, Younathan ES. Modulation of the kinetic properties of phosphofructokinase by ammonium ions. The Journal of biological chemistry. 1971;246:2464–7. [PubMed] [Google Scholar]
- [89].Eagle H. Nutrition needs of mammalian cells in tissue culture. Science. 1955;122:501–14. doi: 10.1126/science.122.3168.501. [DOI] [PubMed] [Google Scholar]
- [90].Sola-Penna M, Da Silva D, Coelho WS, Marinho-Carvalho MM, Zancan P. Regulation of mammalian muscle type 6-phosphofructo-1-kinase and its implication for the control of the metabolism. IUBMB life. 2010;62:791–6. doi: 10.1002/iub.393. [DOI] [PubMed] [Google Scholar]
- [91].Ros S, Schulze A. Balancing glycolytic flux: the role of 6-phosphofructo-2-kinase/fructose 2,6-bisphosphatases in cancer metabolism. Cancer & metabolism. 2013;1:8. doi: 10.1186/2049-3002-1-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [92].Minchenko O, Opentanova I, Caro J. Hypoxic regulation of the 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase gene family (PFKFB-1-4) expression in vivo. FEBS letters. 2003;554:264–70. doi: 10.1016/s0014-5793(03)01179-7. [DOI] [PubMed] [Google Scholar]
- [93].Clem BF, O’Neal J, Tapolsky G, Clem AL, Imbert-Fernandez Y, Kerr DA, 2nd, et al. Targeting 6-phosphofructo-2-kinase (PFKFB3) as a therapeutic strategy against cancer. Molecular cancer therapeutics. 2013;12:1461–70. doi: 10.1158/1535-7163.MCT-13-0097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [94].Clem B, Telang S, Clem A, Yalcin A, Meier J, Simmons A, et al. Small-molecule inhibition of 6-phosphofructo-2-kinase activity suppresses glycolytic flux and tumor growth. Molecular cancer therapeutics. 2008;7:110–20. doi: 10.1158/1535-7163.MCT-07-0482. [DOI] [PubMed] [Google Scholar]
- [95].Schoors S, Cantelmo AR, Georgiadou M, Stapor P, Wang X, Quaegebeur A, et al. Incomplete and transitory decrease of glycolysis: a new paradigm for anti-angiogenic therapy? Cell cycle. 2014;13:16–22. doi: 10.4161/cc.27519. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [96].Schoors S, De Bock K, Cantelmo AR, Georgiadou M, Ghesquiere B, Cauwenberghs S, et al. Partial and transient reduction of glycolysis by PFKFB3 blockade reduces pathological angiogenesis. Cell metabolism. 2014;19:37–48. doi: 10.1016/j.cmet.2013.11.008. [DOI] [PubMed] [Google Scholar]
- [97].De Bock K, Georgiadou M, Carmeliet P. Role of endothelial cell metabolism in vessel sprouting. Cell metabolism. 2013;18:634–47. doi: 10.1016/j.cmet.2013.08.001. [DOI] [PubMed] [Google Scholar]
- [98].Rehman G, Shehzad A, Khan AL, Hamayun M. Role of AMP-Activated Protein Kinase in Cancer Therapy. Archiv der Pharmazie. 2014;347:457–68. doi: 10.1002/ardp.201300402. [DOI] [PubMed] [Google Scholar]
- [99].Luo Z, Zang M, Guo W. AMPK as a metabolic tumor suppressor: control of metabolism and cell growth. Future oncology. 2010;6:457–70. doi: 10.2217/fon.09.174. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [100].Okamura N, Sakakibara R. A common phosphorylation site for cyclic AMP-dependent protein kinase and protein kinase C in human placental 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase. Bioscience, biotechnology, and biochemistry. 1998;62:2039–42. doi: 10.1271/bbb.62.2039. [DOI] [PubMed] [Google Scholar]
- [101].Egler V, Korur S, Failly M, Boulay JL, Imber R, Lino MM, et al. Histone deacetylase inhibition and blockade of the glycolytic pathway synergistically induce glioblastoma cell death. Clinical cancer research : an official journal of the American Association for Cancer Research. 2008;14:3132–40. doi: 10.1158/1078-0432.CCR-07-4182. [DOI] [PubMed] [Google Scholar]
- [102].Manes NP, El-Maghrabi MR. The kinase activity of human brain 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase is regulated via inhibition by phosphoenolpyruvate. Archives of biochemistry and biophysics. 2005;438:125–36. doi: 10.1016/j.abb.2005.04.011. [DOI] [PubMed] [Google Scholar]
- [103].Majewski N, Nogueira V, Bhaskar P, Coy PE, Skeen JE, Gottlob K, et al. Hexokinase-mitochondria interaction mediated by Akt is required to inhibit apoptosis in the presence or absence of Bax and Bak. Molecular cell. 2004;16:819–30. doi: 10.1016/j.molcel.2004.11.014. [DOI] [PubMed] [Google Scholar]
- [104].Bellacosa A, Kumar CC, Di Cristofano A, Testa JR. Activation of AKT kinases in cancer: implications for therapeutic targeting. Advances in cancer research. 2005;94:29–86. doi: 10.1016/S0065-230X(05)94002-5. [DOI] [PubMed] [Google Scholar]
- [105].Yi W, Clark PM, Mason DE, Keenan MC, Hill C, Goddard WA, 3rd, et al. Phosphofructokinase 1 glycosylation regulates cell growth and metabolism. Science. 2012;337:975–80. doi: 10.1126/science.1222278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [106].Smerc A, Sodja E, Legisa M. Posttranslational modification of 6-phosphofructo-1-kinase as an important feature of cancer metabolism. PloS one. 2011;6:e19645. doi: 10.1371/journal.pone.0019645. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [107].Olsen JV, Vermeulen M, Santamaria A, Kumar C, Miller ML, Jensen LJ, et al. Quantitative phosphoproteomics reveals widespread full phosphorylation site occupancy during mitosis. Science signaling. 2010;3:ra3. doi: 10.1126/scisignal.2000475. [DOI] [PubMed] [Google Scholar]
- [108].Hart GW, Slawson C, Ramirez-Correa G, Lagerlof O. Cross talk between O-GlcNAcylation and phosphorylation: roles in signaling, transcription, and chronic disease. Annual review of biochemistry. 2011;80:825–58. doi: 10.1146/annurev-biochem-060608-102511. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [109].Ferrer CM, Lynch TP, Sodi VL, Falcone JN, Schwab LP, Peacock DL, et al. O-GlcNAcylation Regulates Cancer Metabolism and Survival Stress Signaling via Regulation of the HIF-1 Pathway. Molecular cell. 2014 doi: 10.1016/j.molcel.2014.04.026. Epub. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [110].Vander Heiden MG, Locasale JW, Swanson KD, Sharfi H, Heffron GJ, Amador-Noguez D, et al. Evidence for an alternative glycolytic pathway in rapidly proliferating cells. Science. 2010;329:1492–9. doi: 10.1126/science.1188015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [111].Hallows WC, Yu W, Denu JM. Regulation of glycolytic enzyme phosphoglycerate mutase-1 by Sirt1 protein-mediated deacetylation. The Journal of biological chemistry. 2012;287:3850–8. doi: 10.1074/jbc.M111.317404. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [112].Hitosugi T, Zhou L, Fan J, Elf S, Zhang L, Xie J, et al. Tyr26 phosphorylation of PGAM1 provides a metabolic advantage to tumours by stabilizing the active conformation. Nature communications. 2013;4:1790. doi: 10.1038/ncomms2759. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [113].Hitosugi T, Zhou L, Elf S, Fan J, Kang HB, Seo JH, et al. Phosphoglycerate mutase 1 coordinates glycolysis and biosynthesis to promote tumor growth. Cancer cell. 2012;22:585–600. doi: 10.1016/j.ccr.2012.09.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [114].Chaneton B, Gottlieb E. Rocking cell metabolism: revised functions of the key glycolytic regulator PKM2 in cancer. Trends in biochemical sciences. 2012;37:309–16. doi: 10.1016/j.tibs.2012.04.003. [DOI] [PubMed] [Google Scholar]
- [115].Mulukutla BC, Khan S, Lange A, Hu WS. Glucose metabolism in mammalian cell culture: new insights for tweaking vintage pathways. Trends in biotechnology. 2010;28:476–84. doi: 10.1016/j.tibtech.2010.06.005. [DOI] [PubMed] [Google Scholar]
- [116].Chaneton B, Hillmann P, Zheng L, Martin AC, Maddocks OD, Chokkathukalam A, et al. Serine is a natural ligand and allosteric activator of pyruvate kinase M2. Nature. 2012;491:458–62. doi: 10.1038/nature11540. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [117].Keller KE, Doctor ZM, Dwyer ZW, Lee YS. SAICAR induces protein kinase activity of PKM2 that is necessary for sustained proliferative signaling of cancer cells. Molecular cell. 2014;53:700–9. doi: 10.1016/j.molcel.2014.02.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [118].Cortes-Cros M, Hemmerlin C, Ferretti S, Zhang J, Gounarides JS, Yin H, et al. M2 isoform of pyruvate kinase is dispensable for tumor maintenance and growth. Proceedings of the National Academy of Sciences of the United States of America. 2013;110:489–94. doi: 10.1073/pnas.1212780110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [119].Israelsen WJ, Dayton TL, Davidson SM, Fiske BP, Hosios AM, Bellinger G, et al. PKM2 isoform-specific deletion reveals a differential requirement for pyruvate kinase in tumor cells. Cell. 2013;155:397–409. doi: 10.1016/j.cell.2013.09.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [120].Anastasiou D, Yu Y, Israelsen WJ, Jiang JK, Boxer MB, Hong BS, et al. Pyruvate kinase M2 activators promote tetramer formation and suppress tumorigenesis. Nature chemical biology. 2012;8:839–47. doi: 10.1038/nchembio.1060. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [121].Moellering RE, Cravatt BF. Functional lysine modification by an intrinsically reactive primary glycolytic metabolite. Science. 2013;341:549–53. doi: 10.1126/science.1238327. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [122].Aebersold R, Burlingame AL, Bradshaw RA. Western blots versus selected reaction monitoring assays: time to turn the tables? Molecular & cellular proteomics : MCP. 2013;12:2381–2. doi: 10.1074/mcp.E113.031658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [123].Panasyuk G, Espeillac C, Chauvin C, Pradelli LA, Horie Y, Suzuki A, et al. PPARgamma contributes to PKM2 and HK2 expression in fatty liver. Nature communications. 2012;3:672. doi: 10.1038/ncomms1667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [124].Martin JA, Wang Z. Next-generation transcriptome assembly. Nature reviews Genetics. 2011;12:671–82. doi: 10.1038/nrg3068. [DOI] [PubMed] [Google Scholar]
- [125].Zhou W, Liotta LA, Petricoin EF. Cancer metabolism: what we can learn from proteomic analysis by mass spectrometry. Cancer genomics & proteomics. 2012;9:373–81. [PMC free article] [PubMed] [Google Scholar]
- [126].ter Kuile BH, Westerhoff HV. Transcriptome meets metabolome: hierarchical and metabolic regulation of the glycolytic pathway. FEBS letters. 2001;500:169–71. doi: 10.1016/s0014-5793(01)02613-8. [DOI] [PubMed] [Google Scholar]
- [127].Griffin JL, Shockcor JP. Metabolic profiles of cancer cells. Nature reviews Cancer. 2004;4:551–61. doi: 10.1038/nrc1390. [DOI] [PubMed] [Google Scholar]
- [128].de Geus-Oei LF, Vriens D, van Laarhoven HW, van der Graaf WT, Oyen WJ. Monitoring and predicting response to therapy with 18F-FDG PET in colorectal cancer: a systematic review. Journal of nuclear medicine : official publication, Society of Nuclear Medicine. 2009;50(Suppl 1):43S–54S. doi: 10.2967/jnumed.108.057224. [DOI] [PubMed] [Google Scholar]
- [129].Plathow C, Weber WA. Tumor cell metabolism imaging. Journal of nuclear medicine : official publication, Society of Nuclear Medicine. 2008;49(Suppl 2):43S–63S. doi: 10.2967/jnumed.107.045930. [DOI] [PubMed] [Google Scholar]
- [130].Kurhanewicz J, Vigneron DB, Brindle K, Chekmenev EY, Comment A, Cunningham CH, et al. Analysis of cancer metabolism by imaging hyperpolarized nuclei: prospects for translation to clinical research. Neoplasia. 2011;13:81–97. doi: 10.1593/neo.101102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [131].Armitage EG, Barbas C. Metabolomics in cancer biomarker discovery: current trends and future perspectives. Journal of pharmaceutical and biomedical analysis. 2014;87:1–11. doi: 10.1016/j.jpba.2013.08.041. [DOI] [PubMed] [Google Scholar]
- [132].Mulukutla BC, Yongky A, Daoutidis P, Hu WS. Bistability in glycolysis pathway as a physiological switch in energy metabolism. PloS one. 2014;9:e98756. doi: 10.1371/journal.pone.0098756. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [133].Marin-Hernandez A, Gallardo-Perez JC, Ralph SJ, Rodriguez-Enriquez S, Moreno-Sanchez R. HIF-1alpha modulates energy metabolism in cancer cells by inducing over-expression of specific glycolytic isoforms. Mini reviews in medicinal chemistry. 2009;9:1084–101. doi: 10.2174/138955709788922610. [DOI] [PubMed] [Google Scholar]
- [134].Alberghina L, Westerhoff HV. Systems biology: definitions and perspectives. 2005;13:1–408. [Google Scholar]
- [135].Link H, Christodoulou D, Sauer U. Advancing metabolic models with kinetic information. Current opinion in biotechnology. 2014;29C:8–14. doi: 10.1016/j.copbio.2014.01.015. [DOI] [PubMed] [Google Scholar]
- [136].Duckwall CS, Murphy TA, Young JD. Mapping cancer cell metabolism with(13)C flux analysis: Recent progress and future challenges. Journal of carcinogenesis. 2013;12:13. doi: 10.4103/1477-3163.115422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [137].Beger RD. A review of applications of metabolomics in cancer. Metabolites. 2013;3:552–74. doi: 10.3390/metabo3030552. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [138].Wei S, Liu L, Zhang J, Bowers J, Gowda GA, Seeger H, et al. Metabolomics approach for predicting response to neoadjuvant chemotherapy for breast cancer. Molecular oncology. 2013;7:297–307. doi: 10.1016/j.molonc.2012.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [139].Hou Y, Yin M, Sun F, Zhang T, Zhou X, Li H, et al. A metabolomics approach for predicting the response to neoadjuvant chemotherapy in cervical cancer patients. Molecular bioSystems. 2014;10:2126–33. doi: 10.1039/c4mb00054d. [DOI] [PubMed] [Google Scholar]
- [140].Spratlin JL, Serkova NJ, Eckhardt SG. Clinical applications of metabolomics in oncology: a review. Clinical cancer research : an official journal of the American Association for Cancer Research. 2009;15:431–40. doi: 10.1158/1078-0432.CCR-08-1059. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [141].Xie J, Wu H, Dai C, Pan Q, Ding Z, Hu D, et al. Beyond Warburg effect - dual metabolic nature of cancer cells. Scientific reports. 2014;4:4927. doi: 10.1038/srep04927. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [142].Iqbal MA, Siddiqui FA, Gupta V, Chattopadhyay S, Gopinath P, Kumar B, et al. Insulin enhances metabolic capacities of cancer cells by dual regulation of glycolytic enzyme pyruvate kinase M2. Molecular cancer. 2013;12:72. doi: 10.1186/1476-4598-12-72. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [143].Gacche RN, Meshram RJ. Targeting tumor micro-environment for design and development of novel anti-angiogenic agents arresting tumor growth. Progress in biophysics and molecular biology. 2013;113:333–54. doi: 10.1016/j.pbiomolbio.2013.10.001. [DOI] [PubMed] [Google Scholar]
- [144].De Bock K, Mazzone M, Carmeliet P. Antiangiogenic therapy, hypoxia, and metastasis: risky liaisons, or not? Nature reviews Clinical oncology. 2011;8:393–404. doi: 10.1038/nrclinonc.2011.83. [DOI] [PubMed] [Google Scholar]
- [145].Rapisarda A, Melillo G. Overcoming disappointing results with antiangiogenic therapy by targeting hypoxia. Nature reviews Clinical oncology. 2012;9:378–90. doi: 10.1038/nrclinonc.2012.64. [DOI] [PubMed] [Google Scholar]
- [146].Wilson WR, Hay MP. Targeting hypoxia in cancer therapy. Nature reviews Cancer. 2011;11:393–410. doi: 10.1038/nrc3064. [DOI] [PubMed] [Google Scholar]
- [147].Chang CH, Curtis JD, Maggi LB, Jr., Faubert B, Villarino AV, O’Sullivan D, et al. Posttranscriptional control of T cell effector function by aerobic glycolysis. Cell. 2013;153:1239–51. doi: 10.1016/j.cell.2013.05.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [148].Dong H, Bullock TN. Metabolic influences that regulate dendritic cell function in tumors. Frontiers in immunology. 2014;5:24. doi: 10.3389/fimmu.2014.00024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [149].Carmona-Fontaine C, Bucci V, Akkari L, Deforet M, Joyce JA, Xavier JB. Emergence of spatial structure in the tumor microenvironment due to the Warburg effect. Proceedings of the National Academy of Sciences of the United States of America. 2013;110:19402–7. doi: 10.1073/pnas.1311939110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [150].Werner HM, Mills GB, Ram PT. Cancer Systems Biology: a peek into the future of patient care? Nature reviews Clinical oncology. 2014;11:167–76. doi: 10.1038/nrclinonc.2014.6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [151].UniProt C. Activities at the Universal Protein Resource (UniProt) Nucleic acids research. 2014;42:D191–8. doi: 10.1093/nar/gkt1140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [152].Sayers EW, Barrett T, Benson DA, Bryant SH, Canese K, Chetvernin V, et al. Database resources of the National Center for Biotechnology Information. Nucleic acids research. 2009;37:D5–15. doi: 10.1093/nar/gkn741. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [153].Massa ML, Gagliardino JJ, Francini F. Liver glucokinase: An overview on the regulatory mechanisms of its activity. IUBMB life. 2011;63:1–6. doi: 10.1002/iub.411. [DOI] [PubMed] [Google Scholar]
- [154].Ahn KJ, Kim J, Yun M, Park JH, Lee JD. Enzymatic properties of the N- and C-terminal halves of human hexokinase II. BMB reports. 2009;42:350–5. doi: 10.5483/bmbrep.2009.42.6.350. [DOI] [PubMed] [Google Scholar]
- [155].Marin-Hernandez A, Gallardo-Perez JC, Rodriguez-Enriquez S, Encalada R, Moreno-Sanchez R, Saavedra E. Modeling cancer glycolysis. Biochimica et biophysica acta. 2011;1807:755–67. doi: 10.1016/j.bbabio.2010.11.006. [DOI] [PubMed] [Google Scholar]
- [156].Sakakibara R, Kato M, Okamura N, Nakagawa T, Komada Y, Tominaga N, et al. Characterization of a human placental fructose-6-phosphate, 2-kinase/fructose-2,6-bisphosphatase. Journal of biochemistry. 1997;122:122–8. doi: 10.1093/oxfordjournals.jbchem.a021719. [DOI] [PubMed] [Google Scholar]
- [157].Arakaki TL, Pezza JA, Cronin MA, Hopkins CE, Zimmer DB, Tolan DR, et al. Structure of human brain fructose 1,6-(bis)phosphate aldolase: linking isozyme structure with function. Protein science : a publication of the Protein Society. 2004;13:3077–84. doi: 10.1110/ps.04915904. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [158].Eagles PA, Iqbal M. A comparative study of aldolase from human muscle and liver. The Biochemical journal. 1973;133:429–39. doi: 10.1042/bj1330429. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [159].Shimizu A, Suzuki F, Kato K. Characterization of alpha alpha, beta beta, gamma gamma and alpha gamma human enolase isozymes, and preparation of hybrid enolases (alpha gamma, beta gamma and alpha beta) from homodimeric forms. Biochimica et biophysica acta. 1983;748:278–84. doi: 10.1016/0167-4838(83)90305-9. [DOI] [PubMed] [Google Scholar]
- [160].Rodriguez-Horche P, Luque J, Perez-Artes E, Pineda M, Pinilla M. Comparative kinetic behaviour and regulation by fructose-1,6-bisphosphate and ATP of pyruvate kinase from erythrocytes, reticulocytes and bone marrow cells. Comparative biochemistry and physiology B, Comparative biochemistry. 1987;87:553–7. doi: 10.1016/0305-0491(87)90051-4. [DOI] [PubMed] [Google Scholar]
- [161].Mazurek S. Pyruvate kinase type M2: a key regulator of the metabolic budget system in tumor cells. The international journal of biochemistry & cell biology. 2011;43:969–80. doi: 10.1016/j.biocel.2010.02.005. [DOI] [PubMed] [Google Scholar]
- [162].Rider MH, Bertrand L, Vertommen D, Michels PA, Rousseau GG, Hue L. 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase: head-to-head with a bifunctional enzyme that controls glycolysis. The Biochemical journal. 2004;381:561–79. doi: 10.1042/BJ20040752. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [163].Shchutskaya YY, Elkina YL, Kuravsky ML, Bragina EE, Schmalhausen EV. Investigation of glyceraldehyde-3-phosphate dehydrogenase from human sperms. Biochemistry Biokhimiia. 2008;73:185–91. doi: 10.1134/s0006297908020107. [DOI] [PubMed] [Google Scholar]
- [164].Qiu H, Zhao S, Xu X, Yerle M, Liu B. Assignment and expression patterns of porcine muscle-specific isoform of phosphoglycerate mutase gene. Journal of genetics and genomics. 2008;35:257–60. doi: 10.1016/S1673-8527(08)60036-3. [DOI] [PubMed] [Google Scholar]
- [165].Capello M, Ferri-Borgogno S, Cappello P, Novelli F. alpha-Enolase: a promising therapeutic and diagnostic tumor target. The FEBS journal. 2011;278:1064–74. doi: 10.1111/j.1742-4658.2011.08025.x. [DOI] [PubMed] [Google Scholar]
- [166].Soh MA, Garrett SH, Somji S, Dunlevy JR, Zhou XD, Sens MA, et al. Arsenic, cadmium and neuron specific enolase (ENO2, gamma-enolase) expression in breast cancer. Cancer cell international. 2011;11:41. doi: 10.1186/1475-2867-11-41. [DOI] [PMC free article] [PubMed] [Google Scholar]

