Open access

Brain Tumor Metabolism — Unraveling Its Role in Finding New Therapeutic Targets

Written By

Vera Miranda Gonçalves, Fátima Baltazar and Rui Manuel Reis

Submitted: 28 April 2014 Published: 25 March 2015

DOI: 10.5772/59606

Chapter metrics overview

2,321 Chapter Downloads

View Full Metrics

1. Introduction

Primary tumors of brain account for approximately 2-3% of all cancers, with annual incidence approximately 15 patients per 100,000 people and the prevalence has been estimated in 69 patients per 100,000 people. Several brain tumor types evolve from glial or neuronal precursors, being the tumors of glial cells the most common and denominated gliomas [1, 2]. Gliomas are histologically classified according to the World Health Organization (WHO) classification into four malignancy grades[3, 4]. Pilocytic astrocytomas (WHO grade I) are benign tumors that can usually be cured after surgical resection. Diffuse astrocytomas (WHO grade II) exhibit a slow growth, but have an inevitable tendency to progress to higher grade lesions, such as anaplastic gliomas (WHO grade III) and glioblastomas (WHO grade IV). Anaplastic gliomas are rapidly growing malignant tumors that, in addition to surgery, require aggressive adjuvant therapy. Glioblastomas (GBMs) are the most malignant and frequent type of gliomas, which are preferentially manifested in aged adults with a peak of incidence between 50-60 years old [4]. Glioblastomas may evolve from lower grade tumors as described and are mentioned secondary glioblastomas, although most of GBMs arise rapidly without the evidence of less malignant lesion, and are denominated de novo or primary glioblastomas [2, 4].

The current standard therapy for GBM includes tumor resection followed by radiation and concomitant chemotherapy, with temozolomide being the only approved drug that shows some efficacy in this disease [5]. In the last decade, specific inhibitors of oncogenic signaling pathways such as EGFR, PI3K/Akt, and VEGF have made progress with some of them currently tested in clinical trials. Nowadays, bevacizumab (avastin®), a humanized monoclonal antibody against VEGF is approved as a second line of treatment for recurrent GBMs and is currently in phase III clinical trials for the treatment of initial GBMs [6]. Antiangiogenic therapy with avastin improved radiographic response and 6 month of progression free survival, however with modest or little effect on overall survival, when in combination with TMZ during and after radiotherapy [7, 8]. Besides, its role in promoting vascular normalization, the effect on tumor cell invasion is still controversial. Avastin treatment induces infiltration in U87 xenograft model and also was associated with diffusing invasive recurrence in some GBM patients [9, 10]. Additionally, it was observed that vasculature normalization with bevacizumab treatment leads to increased hypoxia and consequently acquisition of resistance [11]. Despite progress in new molecular-based therapies, the prognosis of glioblastomas patients is still very dismal [12, 13]. Thus, exploitation of new molecular targets becomes crucial in neuro-oncology.

In recent years, understanding the regulation of tumor metabolism has significantly improved. Accumulating evidence showed that tumor cells reprogram their metabolism to meet high energy demands, coordinate markedly elevated biosynthetic processes and energy production, which in turn promote rapid growth and division of tumor cells [14-17]. Thus, targeting metabolism has become a novel promising strategy for treating cancers, particularly glioblastomas.

Advertisement

2. Tumor metabolism

During cancer progression, molecular changes are associated to metabolic reprogramming [18, 19], which is nowadays defined as a new hallmark of cancer [20]. In mammalian cells, namely quiescent cells or differentiated tissues, glycolysis is reduced in the presence of oxygen and energy production arises from mitochondrial oxidative phosphorylation which oxidizes pyruvate to CO2 and H2O, known as “Pasteur effect” (Figure 1) [21]. However, in tumor cells, like proliferating tissues, there is high glycolytic activity even in the presence of oxygen, being glycolysis the major source of energy. This phenomenon is known as “Warburg effect”. As a result, tumor cells convert most of the incoming glucose into lactate (around 85 %) rather than metabolizing pyruvate in the mitochondria through oxidative phosphorylation (around 5%) (Figure 1) [16, 21, 22].

Figure 1.

Schematic representation of the metabolic differences between differentiated tissues and proliferating tissues. In the presence of oxygen, non-proliferating tissues metabolize glucose to pyruvate and oxidize it in mitochondria through oxidative phosphorylation. On the other hand, glucose is metabolized to lactate when in the absence of oxygen. In proliferative tissues, like tumor cells, glucose is metabolized to pyruvate and even in the presence of oxygen pryruvate is converted into lactate,a phenomenon denominated aerobic glycolysis or Warburg effect.

2.1. Glycolytic metabolism in brain tumors

As above mentioned, in tumor cells, even in the presence of oxygen, glucose is converted into lactate instead of being oxidized in mitochondria, being glycolysis the major source of energy [16]. It has been described that glioblastomas present metabolic remodeling, increasing glycolytic activity about 3-fold when compared to normal brain tissue [23, 24]. Thus, an increase in several glycolytic enzymes was observed, such as hexokinase II (HKII), pyruvate kinase (PKM), as well as the glucose transporters (GLUTs). Importantly, several studies reported these molecules as important mediators in glycolytic metabolism, constituting attractive molecular targets (Figure 2).

Figure 2.

Potential molecular targets in metabolic remodeling of glioblastomas. The green boxes represent the potential metabolic molecular targets in glioblastomas: enzymes involved in glycolytic metabolism, glutamine metabolism, lipid metabolism; different transporters and also the oncometabolite 2-hydroxyglutarate. Yellow boxes represent the different inhibitors of the specific molecular targets described. Abbreviations: GLUTs, glucose transporters; MCTs, monocarboxylate transporters; CAIX, carbonic anhydrase IX; HKII, hexokinase II, PKM2, pyruvate kinase M2; LDH-A, lactate dehydrogenase A; PDH, pyruvate dehydrogenase; PDK, pyruvate dehydrogenase kinase; IDH1/2, isocitrate dehydrogenase 1/2; IDH1/2 mut, isocitrate dehydrogenase ½ mutation; 2-HG, 2-hydroxyglutarate; GLS, glutaminase; ACC, acetyl-CoA carboxylase; FAS, fatty acid synthase; α-KG, α-ketoglutarate; VDAC, voltage-dependent anion channel; CHC, α-cyanohydroxycinnamic acid.

2.1.1. Glucose Transporters (GLUTs)

Glucose is the main source of energy in most tissues, including brain. GLUTs are transmembrane transporters that perform the uptake of glucose into the cell. The GLUT family is composed by 12 isoforms, however only GLUT1, GLUT3, and GLUT12 have been described as transporters of glucose [25]. GLUT1 is ubiquitously expressed and it is responsible for providing basal glucose to different tissues and cells. In brain, GLUT1 is expressed in astrocytes, whereas GLUT3 is observed in neurons [26].

In the tumoral context, overexpression of specific isoforms of GLUTs has been reported [27, 28]. Most frequently, an increase in GLUT1 expression has been observed in several solid tumors compared with the corresponding normal tissue [27, 28]. However, it has been verified that their expression is tissue specific and some tumors overexpressed other isoforms, such as GLUT12 in prostate cancer [29]. Concerning brain tumors, few studies have evaluated GLUT expression, where it is described that glioblastomas have an increased expression of GLUT1 and GLUT3 when compared with low grade gliomas and normal brain [30, 31]. In fact, both the isoforms are downstream targets of hypoxia-inducible factor 1α (HIF-1α), a transcription factor that is frequently present in glioblastomas. GLUT1 expression is observed in vessels of the normal brain tissues and presents a focal expression in the perinecrotic regions of GBMs, suggesting that their expression is associated with hypoxic regions in glioblastomas (Miranda-Gonçalves V. et al, submitted). Several in vitro studies also reported overexpression of GLUT1 expression in GBMs cells when compared with normal astrocytes [32].

These findings raise the importance of GLUT inhibition in tumor therapy, however, at the moment, a glucose transporter inhibitor is not available at the clinical level. Nevertheless, in vitro studies have been using 2-deoxy-D-glucose has an inhibitor of glucose uptake promoting a decrease on tumor cell proliferation [33] (Figure 2). These studies reported the high dependence of brain tumors on glucose as source of energy and also for catabolic processes.

2.1.2. Hexokinase II

HK is one of the most important enzymes of the glycolytic pathway, which is responsible for the phosphorylation of glucose to glucose-6-phosphate (G6P), thereby preventing the efflux of glucose from the cell [34]. This enzyme has four isoforms (I-IV) identified in different mammalian tissues [35].

In most solid tumors, hexokinases type I and II are the most frequently upregulated [36]. In glioblastomas, HKII is highly expressed, whereas HKI is predominantly expressed in normal brain and low grade gliomas [37]. Additionally, HKII is expressed at low levels in neuronal tissue, but is highly expressed in mesenchymal subtype of glioblastomas [37]. As the first enzyme involved in the glycolytic pathway, HK controls glucose flux in glycolysis or the pentose phosphate pathway (PPP) [38]. HKII is a highly regulated form of hexokinase, being regulated by HIF-1α, glucose, p53, insulin, glucagon, cAMP, among others [36]. The four hexokinase types are normally expressed in the cytoplasm, however type I and II can bind to the outer membrane of the mitochondria via voltage-dependent anion channel (VDAC) (Figure 2) [39]. The translocation of HKII to mitochondria is regulated by growth factors and signaling pathways, such as EGFR and PI3K/AKT activation, which are known to be upregulated in glioblastomas [40]. Moreover, the association of HKII with mitochondria in gliomas, besides maintaining high glucose influx, also renders cells resistant to apoptosis, due to the prevention of cytochrome c release [41].

Several studies have described that the expression of HKII in gliomas promotes proliferation and increase in lactate production, being dependent on both mitochondrial localization and kinase activity [42]. Additionally, HKII overexpression in glioblastomas confers resistance to treatment with both temozolomide and radiation, being associated with poor overall survival [43]. Furthermore, silencing of HKII in glioma cells leads to decrease in glycolytic metabolism, observed by a decrease in lactate production and increase expression of OXPHOS proteins and oxygen consumption [43]. Finally, it was also demonstrated that reduction of HKII expression impaired tumor growth in vivo both on subcutaneous and intracranial xenograft models [37, 43].

Some drugs have been proposed for chemical inhibition of HKII (Figure 2). 3-bromopyruvate (3-BrPA), a pyruvate analogue, is an alkylating agent and also an inhibitor of glycolysis that decreases tumor growth, without apparent toxicity in subcutaneous hepatocellular carcinoma [44]. However, it is effective only at high concentrations (mM) and to the best our knowledge is not under clinical trials. Other known inhibitor is lonidamine, an inhibitor of HKII binding to the mitochondria, which is currently in clinical trials. In vitro studies showed a decrease in lactate production in high grade gliomas but not in low grade [33]. Despite that lonidamine treatment leads to a decrease in tumor growth in different solid tumors without adverse effects, the results of a phase I/II efficacy trial was disappointing in gliomas [45, 46]. Clotrimazole is another inhibitor of HKII localization that demonstrated promising results in vivo. In gliomas, clotrimazole increased the sensitivity to radiotherapy and also leads to decrease in tumor growth [47, 48].

2.1.3. Pyruvate Kinase (PK)

PK is an enzyme involved in the last irreversible step of the glycolytic pathway, converting phosphoenolpyruvate (PEP) to pyruvate [49, 50]. It is also regulated allosterically by the phosphotyrosine binding or phosphorylation and its expression is regulated by isoform selection [50]. Thus, PKM1 is mostly present in adult tissues, such as adult brain and muscle, whereas PKM2 is more frequent in proliferating tissues and embryonic tissues, namely in fetal brain and tumor cancer cells [49]. PKM1 and PKM2 presented different properties, which results in different activities. PKM1 is constitutively active, but PKM2 is regulated by fructose-1,6-biphosphate, presenting reduced activity, which allows the accumulation of glycolytic intermediates and promotes the entry of G6P into the oxidative metabolism of PPP for the production of energy and biosynthesis of proteins, lipids and nucleotides (macromolecules) [50-53]. In cancer cells, like glioblastomas, there is upregulation of PKM2 that favors aerobic glycolysis, increasing lactate production [51, 54, 55]. On the other hand, PKM2 favors the biosynthetic pathway, leading to increased biomass. This dual function potentiates tumor proliferation and aggressiveness. The dimeric form of PKM2 delays pyruvate formation and allows the accumulation of upstream glycolytic intermediates for biosynthetic pathways, whereas the tetrameric form favors aerobic glycolysis, increasing lactate production [56].

In lung cancer cell lines, replacing PKM2 by PKM1 decreases lactate production and increases oxygen consumption (reverse Warburg effect) and also decreases the proliferative capacity of cancer cells in nude mice [54]. It was demonstrated in glioblastomas that knockdown of PKM2 decreased cell proliferation and survival but this did not favor the switch from aerobic glycolysis to oxidative phosphorylation, unlike HKII knockdown [43]. Interestingly, PKM2 was identified as essential for survival of glioma stem cells [57].

Another important function of PKM2 has been associated to epigenetic regulation, being a regulator of histone phosphorylation and acetylation of EGFR-driven glioblastomas [58, 59]. Additionally, in glioblastomas, it was demonstrated that PKM2 is involved in the EGFR signaling pathway that induces its phosphorylation and translocation into the nucleus, which in turn promotes activation of the transcription factor c-Myc with consequent activation of downstream targets, namely genes involved in glycolytic metabolism [60]. Taken together, inhibition of PKM2, in order to deplete the dimer and tetramer formation, can be a new therapeutic strategy, since it can lead to inhibition of glycolysis, decreasing energy production and at the same time blocking the anabolic process in tumor cells (Figure 2).

2.2. Mitochondrial metabolism in brain tumors

In addition to glycolytic dependence, most tumors present abnormalities in the number and function of mitochondria, as the case of glioblastomas [61]. Otto Warburg hypothesized that the increase on glycolytic metabolism in cancer was due to mitochondrial dysfunction, however nowadays we know that most tumors maintain functional mitochondria [22, 62-64]. Moreover, increased glycolytic metabolism can be a consequence of mitochondrial metabolism impairment, due to abnormalities in components of the tricarboxylic acid (TCA) cycle, alterations in electron transport chain or deficiencies in oxidative phosphorylation [65, 66]. Concerning the selection theory in cancer cells, the dependence on glycolysis occurs gradually in order to compensate the respiratory failure. In contrast to normal brain cells, in which glycolysis and respiration are tightly coupled, tumor cells are defective in their ability to connect glycolysis and respiration [66].

Two mitochondrial enzymes are important in glioblastomas, such as pyruvate dehydrogenase kinase (PDK) and isocitrate dehydrogenases 1 and 2 (IDH1 and IDH2). The presence of mutations in IDH1 and IDH2 has been recently associated with gliomagenesis.

2.2.1. Pyruvate Dehydrogenase Kinase (PDK)

Pyruvate dehydrogenase (PDH) is a mitochondrial enzyme that controls the entry of pyruvate into mitochondria, promoting its oxidative decarboxylation into acetyl-CoA [67, 68]. The activity of PDH is inhibited by phosphorylation through PDK, resulting in its accumulation in the cytosol and consequent conversion into lactate [67, 68]. PDK is an important mitochondrial matrix protein comprising four isoforms (PDK1 to PDK4), being PDK2 highly expressed in glioblastomas compared to normal adjacent brain tissue [69].

Tumor cells present high levels of glycolysis as a consequence of increased hypoxic microenvironment, which leads to activation of HIF-1α and consequent upregulation of downstream target genes involved in glycolytic metabolism, such as PDK [67]. This enzyme is responsible for the uncoupling between glycolysis and mitochondrial oxidation of glucose, preventing the entry of pyruvate into the mitochondria with consequent increase in glycolytic rates, which confers resistance to apoptosis [67, 68]. Thus, PDK became an important target for glycolytic tumors (Figure 2). Dichloroacetate (DCA), a chemical PDK inhibitor, has been studied in several in vitro and in vivo models [68]. DCA promotes dephosphorylation of PDH, leading to entry of pyruvate into the mitochondria, decreasing glycolytic rates and lactate production [68, 70]. This leads to activation of oxidative phosphorylation, depolarization of mitochondria and consequent increase in production of reactive oxygen species (ROS), which promotes apoptosis of cancer cells, decreasing tumor cell proliferation [68, 70]. DCA treatment has demonstrated promising results in some tumors, particularly in non-small cell lung, breast and endometrial cancer, either experimentally using in vitro and in vivo models, as well as in clinical trials [70-73]. In glioblastomas, inhibition of PDK with DCA was evaluated pre-clinically. In C6 glioma cells, it was observed a decrease in lactate production, increase in ROS production, as well as depolarization of mitochondria, which results in a decrease on cell proliferation and induction of apoptosis [74]. In vivo it was verified that DCA decreased not only tumor growth but also the angiogenic capacity of glioma cells [74]. The effect of DCA was also tested in a series of glioblastoma patients with congenital acidosis, with a reduction in lactate levels, decrease in HKII localization in the mitochondria, as well as a decrease in mitochondrial polarization, which rendered tumor cells more sensitive to apoptosis [69]. Despite these encouraging results, no exact conclusions can yet be made regarding the efficacy and toxicity of DCA in glioblastoma patients. Thus, a large and randomized clinical study would be important to define the efficacy and toxicity of DCA. Additionally, whether DCA can sensitize GBM cells to temozolomide and radiotherapy remains undetermined.

2.2.2. Isocitrate Dehydrogenase (IDH)

IDH is an enzyme that catalyzes the oxidative decarboxylation of isocitrate to α-ketoglutarate, generating NADH in the mitochondria or NADPH in the cytoplasm [75]. It is composed by 5 genes, being the IDH1 and IDH2 the most explored and important in gliomas. IDH1 is presented in the cytoplasm and peroxisome, whereas IDH2 is presented in the mitochondria [76].

In 2008, recurrent somatic hotspot mutations of IDH1 and IDH2 were found in low grade gliomas and secondary glioblastomas [77, 78]. These mutations cause amino acid single change in one of the two alleles of the gene (arginine 132 for IDH1 and arginine 172 for IDH2), being classified has a dominant mutation [79]. It is described that the arginine mutation occurs in the binding site of the substrate isocitrate [42]. IDH1 mutations are reported in more than 80% secondary glioblastomas, but only 5% in primary glioblastomas [76, 80, 81]. Additionally, it occurs in 80% of diffuse astrocytomas (WHO grade II). These mutations are more frequent in younger patient secondary glioblastomas, associated with a proneural subtype and also with increased survival [82]. It was reported that the presence of IDH mutations leads to a decrease in α-ketoglutarate that is required for prolyl hydroxylase (PHD) activity that promote degradation of HIF-1α (ref). Thus, downregulation of intracellular α-ketoglutarate contributes to stabilization of HIF-1α, leading to pseudohypoxia [83]. Nevertheless, subsequent studies did not verify an alteration in α-ketoglutarate on mutant IDH1/2, instead, a gain of function activity it was found, which converts α-ketoglutarate to 2-hydroxyglutarate (Figure 2) [84, 85]. The latter has been recognized as an oncometabolite, which inhibits enzymes involved in the α-ketoglutarate pathway. Additionally, it was described that 2-hydroxyglutarate is involved in epigenetic regulation, promoting a hypermethylator phenotype in gliomas [86] and also keep cells in an undifferentiated status, or stem cell-like, which can be more permissive to transformation [17]. The presence of pseudohypoxia, due to the constitutive stabilization of HIF-1α, indicates that IDH1/2 mutations are involved in HIF-1α signaling pathway, which promotes glucose metabolism, angiogenesis and invasion [83]. These results suggest the paramount role of IDH mutations on the metabolic remodeling of glioblastomas, contributing to the “Warburg effect”. Therefore, study the involvement of IDH1/2 mutations in metabolic remodelling and in aerobic glycolysis opens a new window for investigation. Overall, IDH1 is an attractive target for therapy (Figure 2) since they are early events in the progression from low grade to high grade gliomas.

2.3. Glutamine metabolism and lipid synthesis in brain tumors

Like glucose, glutamine is a source of energy for tumor cells, functioning as a nitrogen donor [87, 88]. Glutamine metabolism has been reported to be upregulated in some tumors, being crucial for the biosynthetic processes, namely synthesis of cholesterol and fatty acids [14, 89, 90]. The shift to glutamine metabolism to produce the precursor acetyl-CoA for lipid biosynthesis is a mechanism of adaptation to glycolytic metabolism that prevents the entry of pyruvate into mitochondria, due to upregulation of PDK [91]. In fact, it has been observed an increased expression of glutaminase (GLS) enzyme in tumor cells. GLS is located in the mitochondria and catalyzes the conversion of glutamine to glutamate being transcriptionally regulated by the oncogenes c-Myc and NFκB. [92-94], An increased concentration of glutamine in glioblastomas compared to normal brain tissue was demonstrated by nuclear magnetic resonance (NMR) [95]. Additionally, there is a low expression of glutamine synthase that correlates with a better prognosis in glioblastomas [96].

Beyond the altered glycolytic and glutamine metabolism in tumor cells, the alteration in lipid metabolism is also recognized as a component of the metabolic reprogramming. It has been observed that tumor cells present reactivation of de novo fatty acid synthesis, important for the biogenesis of cellular membranes [97, 98]. Glioblastomas contain higher levels of unsaturated fatty acids compared to normal brain, indicating the presence of exacerbated lipogenesis, which is regulated by several key genes, such as SREBP-1 and its downstream-targeted genes acetyl-CoA carboxylase (ACC), fatty acid synthase (FAS) and low-density lipoprotein receptor (LDLR), which are upregulated in these tumors [99]. Importantly, the EGFR/PI3K/AkT signaling pathway regulates the metabolic reprogramming in glioblastomas [100]. Cholesterol is also an important component of cell membranes and cholesterol esters have been found to be abundantly present in high grade gliomas, but undetectable in normal tissues by NMR techniques [101, 102]. Recently, low density lipoprotein receptor (LDLR) has been described to be upregulated in GBM patients, xenografts and cell lines, and this upregulation was correlated with high levels of cholesterol esters in GBM cells [103]. Interestingly, LDLR is also be upregulated by EGFR/PI3K/Akt signaling, which was been shown to be mediated by SREBP-1 in GBMs [100]. However, little is known about the altered lipid metabolism in cancer cells, namely glioblastomas, and their role in the tumor context, being possible that lipogenesis in cancer cells could support the cell growth located within nutrient-limited areas, thereby contributing to symbiotic relationships within tumors. Once more, lipid metabolism, as well as glutamine metabolism, and their key enzymes are interesting targets in glioblastomas (Figure 2)

Advertisement

3. Lactate transport and pH regulation in brain tumors

A constitutive increase in the glycolytic phenotype of cancer cells leads to acute and chronic acidification of tumor microenvironment. Important proteins involved in acidification of the extracellular space are monocarboxylate transporters (MCTs) that co-transport H+and lactate, and carbonic anhydrases (CAIXs), which are activated by growth factors, oncogenic transformation, hypoxia, and low intracellular pH [21]. As it is known, tumor acidity is associated with cancer cell invasion behavior, i.e. increased migration, invasion and metastasis [104]. Further, tumor acidosis and lactate contributes to several features of tumor progression and malignancy, like immune escape, angiogenesis, and radioresistance, making lactate a key player in cancer aggressiveness. [105]. Still in line with a potential involvement of lactate in the invasion behavior, it has been shown that lactate up-regulates the expression of transforming growth factor (TGF-β2), which is associated with increased migration in glioblastomas [106].

3.1. Monocarboxylate transporters

The MCT family comprises 14 members with similar topology; however, only 4 isoforms (MCT1–MCT4) are proton-linked monocarboxylate transporters, performing the transmembrane transport of monocarboxylates, such as lactate, coupled with a proton, in an equimolar manner [107, 108].

In the last years, several studies reported up-regulation of MCTs in different human solid tumors, showing the importance of MCTs in cancer biology [109]. In brain tumors, the scare studies point to the importance of MCT expression, especially MCT1. Strong expression of MCT1 in the plasma membrane was found in high grade gliomas compared with low-grade lesions and normal adjacent tissues, which exhibited negative or weak MCT1 staining [110, 111],. A study in neuroblastomas showed, by mRNA quantification, that MCT1 was differently expressed and that its activity was highly associated with MYCN amplification, leading to the hypothesis that expression of MCT1 could be associated with higher malignancy [112]. Further, expression analysis revealed that SLC16A1 transcript, encoding MCT1, was elevated in 90% of the medulloblastomas analyzed [113]. It was also reported that inhibition of MCT activity, particularly MCT1, decreased the glycolytic phenotype (low glucose consumption and lactate production), cell proliferation and invasion, promoting increase in cell death [111, 114, 115]. This elucidates the importance of MCT1 activity in intracellular pH homeostasis and tumor aggressiveness of glioblastomas.

Although MCTs are not the major H+ transporters, the data available in the literature support the hypothesis of a major contribution of MCTs to the hyper-glycolytic and acid-resistant phenotype, as major adaptation to the hypoxic microenvironment [116]. Thus, MCT inhibition may be a useful therapeutic approach in brain tumors (Figure 2). Actually, it was demonstrated that in vitro MCT1 inhibition decreases intracellular pH, leads to cell death and, importantly, enhances cancer cell radiosensivity in gliomas [114, 115]. Importantly, promising results using in vivo models have also been reported, where treatment with the chemical inhibitor CHC retarded tumor growth, rendered tumor cells sensitive to radiation and decreased invasion [114, 117]. However, CHC is not a specific MCT inhibitor, having also other targets. Recently, novel MCT1 inhibitors have been designed and may constitute an effective strategy to block MCT1 activity in cancer [118]. A orally administered related compound, AZD3965 (AstraZeneca), is currently in Phase I/II clinical trials for advanced solid tumors [119].

3.2. Carbonic anhydrases

Carbonic anhydrase catalyzes the conversion of extracellular bicarbonate to CO2 and protons (H+), thereby contributing to extracellular acidification [120]. This family is composed by 15 isoforms described in mammals, which differ in cellular localization, catalytic activity and susceptibility to different class of inhibitors. Two carbonic anhydrases are overexpressed in many solid tumors, namely CAIX and CAXII, being associated with tumor progression and response to therapy [121]. It is verified that CAIX is mostly negative in normal tissues but increase in the corresponding tumor tissues, whereas CAXII present a diffuse distribution in healthy tissues [122, 123]. Glioblastomas present high levels of intratumoral hypoxia, with consequent HIF-1α activation which contributes to increased expression of glycolysis-related genes [124], including CAIX [125]. CAIX is overexpressed in these tumors with focal plasma membrane expression close to peri-necrotic regions (hypoxic) [126], being negative in normal adjacent tissues, making it a feasible treatment target [127]. Furthermore, it has been described that CAIX is associated to poor overall survival, because it confers resistance to chemotherapy, radiotherapy and anti-angiogenic therapy [128]. Increased expression of CAIX in advanced stages/grades of many tumor types also suggests its association with dedifferentiation [129].

In vitro and in vivo approaches have demonstrated the potential of CAIX inhibition (Figure 2). Knockdown of CAIX decreased tumor cell ATP levels under hypoxic and glycolytic conditions [126]. In addition, the susceptibility of U251 glioblastoma cells to chemotherapy and radiation treatment was strongly enhanced after CAIX downregulation, which is supported by a recent in vivo study [126]. Similarly, CAIX inhibition enhanced the effect of anti-angiogenic therapy with the anti-VEGF antibody bevacizumab [130]. Some inhibitors have been developed to inhibit CA activity, particularly CAIX and CAXII. Acetazolomide, enhances the apoptotic response of glioma cells to temozolomide [131] and an in vivo study using derivatives of acetazolamide showed retardation of mice carcinoma xenograft growth after 1 month of treatment [131]. Other studies have identified coumarins as CA inhibitors, however they were not tested yet in the cancer context [132, 133]. Furthermore, specific monoclonal antibodies for the mostly expressed isoforms in tumors, namely CAIX, have been developed, i.e, the M75 and WX-G250 for colorectal cancer and renal cell carcinoma, respectively [134, 135].

Advertisement

4. Future perspectives and conclusions

Metabolic transformation plays a major role in gliomas development, tumor progression and adaptation to tumor microenvironment. The interplay between tumor angiogenesis, hypoxia, pH regulation and energy metabolism, glycolysis related enzymes and transporters, as well as pH regulator transporters, may provide promising molecular targets for drug development. In addition to glycolysis, glutaminolysis and fatty acid synthesis represent key metabolic events with potentially interesting drug targets. Furthermore, mutations in IDH1/2, detected in a genome wide screen on GBMs, point to new specific transforming events in gliomas. These metabolic pathways are tightly linked and also controlled by signaling events often deregulated in gliomas, underlying the flexibility of glioma cells to develop adaptive mechanisms when exposed to oxygen or nutrient deprivation. This highlights the need of targeting several pathways simultaneously and linking the metabolic targets to the genetic makeup of GBM tumors.

References

  1. 1. Ohgaki, H. and P. Kleihues, Epidemiology and etiology of gliomas. Acta Neuropathol, 2005. 109(1): p. 93-108.
  2. 2. Huse, J.T. and E.C. Holland, Targeting brain cancer: advances in the molecular pathology of malignant glioma and medulloblastoma. Nat Rev Cancer, 2010. 10(5): p. 319-31.
  3. 3. Louis, D.N., et al., The 2007 WHO classification of tumours of the central nervous system. Acta Neuropathol, 2007. 114(2): p. 97-109.
  4. 4. Riemenschneider, M.J. and G. Reifenberger, Molecular neuropathology of gliomas. Int J Mol Sci, 2009. 10(1): p. 184-212.
  5. 5. Stupp, R., et al., Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma. N Engl J Med, 2005. 352(10): p. 987-96.
  6. 6. Lai, A., et al., Phase II study of bevacizumab plus temozolomide during and after radiation therapy for patients with newly diagnosed glioblastoma multiforme. J Clin Oncol, 2011. 29(2): p. 142-8.
  7. 7. Chinot, O.L., et al., Bevacizumab plus radiotherapy-temozolomide for newly diagnosed glioblastoma. N Engl J Med, 2014. 370(8): p. 709-22.
  8. 8. Weathers, S.P. and M.R. Gilbert, Advances in treating glioblastoma. F1000Prime Rep, 2014. 6: p. 46.
  9. 9. de Groot, J.F., et al., Tumor invasion after treatment of glioblastoma with bevacizumab: radiographic and pathologic correlation in humans and mice. Neuro Oncol, 2010. 12(3): p. 233-42.
  10. 10. Narayana, A., et al., Bevacizumab in recurrent high-grade pediatric gliomas. Neuro Oncol, 2010. 12(9): p. 985-90.
  11. 11. Mesti, T., et al., Metabolic impact of anti-angiogenic agents on U87 glioma cells. PLoS One, 2014. 9(6): p. e99198.
  12. 12. Gaspar, N., et al., MGMT-independent temozolomide resistance in pediatric glioblastoma cells associated with a PI3-kinase-mediated HOX/stem cell gene signature. Cancer Res, 2010. 70(22): p. 9243-52.
  13. 13. Costa, B.M., et al., Prognostic value of MGMT promoter methylation in glioblastoma patients treated with temozolomide-based chemoradiation: a Portuguese multicentre study. Oncol Rep, 2010. 23(6): p. 1655-62.
  14. 14. DeBerardinis, R.J., et al., The biology of cancer: metabolic reprogramming fuels cell growth and proliferation. Cell Metab, 2008. 7(1): p. 11-20.
  15. 15. DeBerardinis, R.J. and C.B. Thompson, Cellular metabolism and disease: what do metabolic outliers teach us? Cell, 2012. 148(6): p. 1132-44.
  16. 16. Vander Heiden, M.G., L.C. Cantley, and C.B. Thompson, Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science, 2009. 324(5930): p. 1029-33.
  17. 17. Schulze, A. and A.L. Harris, How cancer metabolism is tuned for proliferation and vulnerable to disruption. Nature, 2012. 491(7424): p. 364-73.
  18. 18. Dang, C.V. and G.L. Semenza, Oncogenic alterations of metabolism. Trends Biochem Sci, 1999. 24(2): p. 68-72.
  19. 19. Gatenby, R.A. and R.J. Gillies, A microenvironmental model of carcinogenesis. Nat Rev Cancer, 2008. 8(1): p. 56-61.
  20. 20. Hanahan, D. and R.A. Weinberg, Hallmarks of cancer: the next generation. Cell, 2011. 144(5): p. 646-74.
  21. 21. Gatenby, R.A. and R.J. Gillies, Why do cancers have high aerobic glycolysis? Nat Rev Cancer, 2004. 4(11): p. 891-9.
  22. 22. Warburg, O., On respiratory impairment in cancer cells. Science, 1956. 124(3215): p. 269-70.
  23. 23. Oudard, S., et al., High glycolysis in gliomas despite low hexokinase transcription and activity correlated to chromosome 10 loss. Br J Cancer, 1996. 74(6): p. 839-45.
  24. 24. Tabatabaei, P., et al., Glucose metabolites, glutamate and glycerol in malignant glioma tumours during radiotherapy. J Neurooncol, 2008. 90(1): p. 35-9.
  25. 25. Zhao, F.Q. and A.F. Keating, Functional properties and genomics of glucose transporters. Curr Genomics, 2007. 8(2): p. 113-28.
  26. 26. Leybaert, L., Neurobarrier coupling in the brain: a partner of neurovascular and neurometabolic coupling? J Cereb Blood Flow Metab, 2005. 25(1): p. 2-16.
  27. 27. Medina, R.A. and G.I. Owen, Glucose transporters: expression, regulation and cancer. Biol Res, 2002. 35(1): p. 9-26.
  28. 28. Macheda, M.L., S. Rogers, and J.D. Best, Molecular and cellular regulation of glucose transporter (GLUT) proteins in cancer. J Cell Physiol, 2005. 202(3): p. 654-62.
  29. 29. Chandler, J.D., et al., Expression and localization of GLUT1 and GLUT12 in prostate carcinoma. Cancer, 2003. 97(8): p. 2035-42.
  30. 30. Boado, R.J., K.L. Black, and W.M. Pardridge, Gene expression of GLUT3 and GLUT1 glucose transporters in human brain tumors. Brain Res Mol Brain Res, 1994. 27(1): p. 51-7.
  31. 31. Flynn, J.R., et al., Hypoxia-regulated protein expression, patient characteristics, and preoperative imaging as predictors of survival in adults with glioblastoma multiforme. Cancer, 2008. 113(5): p. 1032-42.
  32. 32. Jensen, R.L., Brain tumor hypoxia: tumorigenesis, angiogenesis, imaging, pseudoprogression, and as a therapeutic target. J Neurooncol, 2009. 92(3): p. 317-35.
  33. 33. Stieber, D., S.A. Abdul Rahim, and S.P. Niclou, Novel ways to target brain tumour metabolism. Expert Opin Ther Targets, 2011. 15(10): p. 1227-39.
  34. 34. Smith, T.A., Mammalian hexokinases and their abnormal expression in cancer. Br J Biomed Sci, 2000. 57(2): p. 170-8.
  35. 35. Wilson, J.E., Isozymes of mammalian hexokinase: structure, subcellular localization and metabolic function. J Exp Biol, 2003. 206(Pt 12): p. 2049-57.
  36. 36. Pedersen, P.L., et al., Mitochondrial bound type II hexokinase: a key player in the growth and survival of many cancers and an ideal prospect for therapeutic intervention. Biochim Biophys Acta, 2002. 1555(1-3): p. 14-20.
  37. 37. Wolf, A., et al., Hexokinase 2 is a key mediator of aerobic glycolysis and promotes tumor growth in human glioblastoma multiforme. J Exp Med, 2011. 208(2): p. 313-26.
  38. 38. Agnihotri, S., et al., A GATA4-regulated tumor suppressor network represses formation of malignant human astrocytomas. J Exp Med, 2011. 208(4): p. 689-702.
  39. 39. Mathupala, S.P., Y.H. Ko, and P.L. Pedersen, Hexokinase-2 bound to mitochondria: cancer's stygian link to the "Warburg Effect" and a pivotal target for effective therapy. Semin Cancer Biol, 2009. 19(1): p. 17-24.
  40. 40. Mellinghoff, I.K., et al., Molecular determinants of the response of glioblastomas to EGFR kinase inhibitors. N Engl J Med, 2005. 353(19): p. 2012-24.
  41. 41. Pastorino, J.G., N. Shulga, and J.B. Hoek, Mitochondrial binding of hexokinase II inhibits Bax-induced cytochrome c release and apoptosis. J Biol Chem, 2002. 277(9): p. 7610-8.
  42. 42. Wolf, A., S. Agnihotri, and A. Guha, Targeting metabolic remodeling in glioblastoma multiforme. Oncotarget, 2010. 1(7): p. 552-62.
  43. 43. Wolf, A., et al., Developmental profile and regulation of the glycolytic enzyme hexokinase 2 in normal brain and glioblastoma multiforme. Neurobiol Dis, 2011. 44(1): p. 84-91.
  44. 44. Ko, Y.H., et al., Advanced cancers: eradication in all cases using 3-bromopyruvate therapy to deplete ATP. Biochem Biophys Res Commun, 2004. 324(1): p. 269-75.
  45. 45. Oudard, S., et al., Phase II study of lonidamine and diazepam in the treatment of recurrent glioblastoma multiforme. J Neurooncol, 2003. 63(1): p. 81-6.
  46. 46. Carapella, C.M., et al., The potential role of lonidamine (LND) in the treatment of malignant glioma. Phase II study. J Neurooncol, 1989. 7(1): p. 103-8.
  47. 47. Liu, H., Y. Li, and K.P. Raisch, Clotrimazole induces a late G1 cell cycle arrest and sensitizes glioblastoma cells to radiation in vitro. Anticancer Drugs, 2010. 21(9): p. 841-9.
  48. 48. Khalid, M.H., et al., Inhibition of tumor growth and prolonged survival of rats with intracranial gliomas following administration of clotrimazole. J Neurosurg, 2005. 103(1): p. 79-86.
  49. 49. Altenberg, B. and K.O. Greulich, Genes of glycolysis are ubiquitously overexpressed in 24 cancer classes. Genomics, 2004. 84(6): p. 1014-20.
  50. 50. Mazurek, S., et al., Pyruvate kinase type M2 and its role in tumor growth and spreading. Semin Cancer Biol, 2005. 15(4): p. 300-8.
  51. 51. Mazurek, S., et al., Pyruvate kinase type M2: a crossroad in the tumor metabolome. Br J Nutr, 2002. 87 Suppl 1: p. S23-9.
  52. 52. Christofk, H.R., et al., Pyruvate kinase M2 is a phosphotyrosine-binding protein. Nature, 2008. 452(7184): p. 181-6.
  53. 53. Eigenbrodt, E., et al., Double role for pyruvate kinase type M2 in the expansion of phosphometabolite pools found in tumor cells. Crit Rev Oncog, 1992. 3(1-2): p. 91-115.
  54. 54. Christofk, H.R., et al., The M2 splice isoform of pyruvate kinase is important for cancer metabolism and tumour growth. Nature, 2008. 452(7184): p. 230-3.
  55. 55. Kefas, B., et al., Pyruvate kinase M2 is a target of the tumor-suppressive microRNA-326 and regulates the survival of glioma cells. Neuro Oncol, 2010. 12(11): p. 1102-12.
  56. 56. Soga, T., Cancer metabolism: key players in metabolic reprogramming. Cancer Sci, 2013. 104(3): p. 275-81.
  57. 57. Goidts, V., et al., RNAi screening in glioma stem-like cells identifies PFKFB4 as a key molecule important for cancer cell survival. Oncogene, 2012. 31(27): p. 3235-43.
  58. 58. Yang, W., et al., PKM2 phosphorylates histone H3 and promotes gene transcription and tumorigenesis. Cell, 2012. 150(4): p. 685-96.
  59. 59. Yang, W., et al., EGFR-induced and PKCepsilon monoubiquitylation-dependent NF-kappaB activation upregulates PKM2 expression and promotes tumorigenesis. Mol Cell, 2012. 48(5): p. 771-84.
  60. 60. Yang, W., et al., ERK1/2-dependent phosphorylation and nuclear translocation of PKM2 promotes the Warburg effect. Nat Cell Biol, 2012. 14(12): p. 1295-304.
  61. 61. Katsetos, C.D., H. Anni, and P. Draber, Mitochondrial dysfunction in gliomas. Semin Pediatr Neurol, 2013. 20(3): p. 216-27.
  62. 62. Warburg, O., On the origin of cancer cells. Science, 1956. 123(3191): p. 309-14.
  63. 63. Pedersen, P.L., Tumor mitochondria and the bioenergetics of cancer cells. Prog Exp Tumor Res, 1978. 22: p. 190-274.
  64. 64. Meixensberger, J., et al., Metabolic patterns in malignant gliomas. J Neurooncol, 1995. 24(2): p. 153-61.
  65. 65. Seyfried, T.N. and P. Mukherjee, Targeting energy metabolism in brain cancer: review and hypothesis. Nutr Metab (Lond), 2005. 2: p. 30.
  66. 66. Seyfried, T.N., et al., Metabolic management of brain cancer. Biochim Biophys Acta, 2011. 1807(6): p. 577-94.
  67. 67. Kim, J.W., et al., HIF-1-mediated expression of pyruvate dehydrogenase kinase: a metabolic switch required for cellular adaptation to hypoxia. Cell Metab, 2006. 3(3): p. 177-85.
  68. 68. Michelakis, E.D., L. Webster, and J.R. Mackey, Dichloroacetate (DCA) as a potential metabolic-targeting therapy for cancer. Br J Cancer, 2008. 99(7): p. 989-94.
  69. 69. Michelakis, E.D., et al., Metabolic modulation of glioblastoma with dichloroacetate. Sci Transl Med, 2010. 2(31): p. 31ra34.
  70. 70. Bonnet, S., et al., A mitochondria-K+channel axis is suppressed in cancer and its normalization promotes apoptosis and inhibits cancer growth. Cancer Cell, 2007. 11(1): p. 37-51.
  71. 71. Cairns, R.A., et al., Metabolic targeting of hypoxia and HIF1 in solid tumors can enhance cytotoxic chemotherapy. Proc Natl Acad Sci U S A, 2007. 104(22): p. 9445-50.
  72. 72. Cao, W., et al., Dichloroacetate (DCA) sensitizes both wild-type and over expressing Bcl-2 prostate cancer cells in vitro to radiation. Prostate, 2008. 68(11): p. 1223-31.
  73. 73. Wong, J.Y., et al., Dichloroacetate induces apoptosis in endometrial cancer cells. Gynecol Oncol, 2008. 109(3): p. 394-402.
  74. 74. Duan, Y., et al., Antitumor activity of dichloroacetate on C6 glioma cell: in vitro and in vivo evaluation. Onco Targets Ther, 2013. 6: p. 189-98.
  75. 75. Reitman, Z.J. and H. Yan, Isocitrate dehydrogenase 1 and 2 mutations in cancer: alterations at a crossroads of cellular metabolism. J Natl Cancer Inst, 2010. 102(13): p. 932-41.
  76. 76. Kim, W. and L.M. Liau, IDH mutations in human glioma. Neurosurg Clin N Am, 2012. 23(3): p. 471-80.
  77. 77. Parsons, D.W., et al., An integrated genomic analysis of human glioblastoma multiforme. Science, 2008. 321(5897): p. 1807-12.
  78. 78. Yan, H., et al., IDH1 and IDH2 mutations in gliomas. N Engl J Med, 2009. 360(8): p. 765-73.
  79. 79. Fu, Y., et al., Glioma-derived mutations in IDH: from mechanism to potential therapy. Biochem Biophys Res Commun, 2010. 397(2): p. 127-30.
  80. 80. DeAngelis, L.M. and I.K. Mellinghoff, Virchow 2011 or how to ID(H) human glioblastoma. J Clin Oncol, 2011. 29(34): p. 4473-4.
  81. 81. Watanabe, T., et al., IDH1 mutations are early events in the development of astrocytomas and oligodendrogliomas. Am J Pathol, 2009. 174(4): p. 1149-53.
  82. 82. Verhaak, R.G., et al., Integrated genomic analysis identifies clinically relevant subtypes of glioblastoma characterized by abnormalities in PDGFRA, IDH1, EGFR, and NF1. Cancer Cell, 2010. 17(1): p. 98-110.
  83. 83. Zhao, S., et al., Glioma-derived mutations in IDH1 dominantly inhibit IDH1 catalytic activity and induce HIF-1alpha. Science, 2009. 324(5924): p. 261-5.
  84. 84. Ward, P.S., et al., The common feature of leukemia-associated IDH1 and IDH2 mutations is a neomorphic enzyme activity converting alpha-ketoglutarate to 2-hydroxyglutarate. Cancer Cell, 2010. 17(3): p. 225-34.
  85. 85. Dang, L., et al., Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature, 2009. 462(7274): p. 739-44.
  86. 86. Lu, C. and C.B. Thompson, Metabolic regulation of epigenetics. Cell Metab, 2012. 16(1): p. 9-17.
  87. 87. Zielke, H.R., C.L. Zielke, and P.T. Ozand, Glutamine: a major energy source for cultured mammalian cells. Fed Proc, 1984. 43(1): p. 121-5.
  88. 88. Reitzer, L.J., B.M. Wice, and D. Kennell, Evidence that glutamine, not sugar, is the major energy source for cultured HeLa cells. J Biol Chem, 1979. 254(8): p. 2669-76.
  89. 89. DeBerardinis, R.J., et al., Beyond aerobic glycolysis: transformed cells can engage in glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis. Proc Natl Acad Sci U S A, 2007. 104(49): p. 19345-50.
  90. 90. Rajagopalan, K.N. and R.J. DeBerardinis, Role of glutamine in cancer: therapeutic and imaging implications. J Nucl Med, 2011. 52(7): p. 1005-8.
  91. 91. Daye, D. and K.E. Wellen, Metabolic reprogramming in cancer: unraveling the role of glutamine in tumorigenesis. Semin Cell Dev Biol, 2012. 23(4): p. 362-9.
  92. 92. Dang, C.V., A. Le, and P. Gao, MYC-induced cancer cell energy metabolism and therapeutic opportunities. Clin Cancer Res, 2009. 15(21): p. 6479-83.
  93. 93. Rathore, M.G., et al., The NF-kappaB member p65 controls glutamine metabolism through miR-23a. Int J Biochem Cell Biol, 2012. 44(9): p. 1448-56.
  94. 94. Wang, J.B., et al., Targeting mitochondrial glutaminase activity inhibits oncogenic transformation. Cancer Cell, 2010. 18(3): p. 207-19.
  95. 95. Kallenberg, K., et al., Untreated glioblastoma multiforme: increased myo-inositol and glutamine levels in the contralateral cerebral hemisphere at proton MR spectroscopy. Radiology, 2009. 253(3): p. 805-12.
  96. 96. Rosati, A., et al., Epilepsy in glioblastoma multiforme: correlation with glutamine synthetase levels. J Neurooncol, 2009. 93(3): p. 319-24.
  97. 97. Menendez, J.A. and R. Lupu, Fatty acid synthase and the lipogenic phenotype in cancer pathogenesis. Nat Rev Cancer, 2007. 7(10): p. 763-77.
  98. 98. Abramson, H.N., The lipogenesis pathway as a cancer target. J Med Chem, 2011. 54(16): p. 5615-38.
  99. 99. Gopal, K., et al., Lipid Composition of Human Intracranial Tumors: A Biochemical Study. Acta Neurochir (Wien), 1963. 11: p. 333-47.
  100. 100. Guo, D., et al., EGFR signaling through an Akt-SREBP-1-dependent, rapamycin-resistant pathway sensitizes glioblastomas to antilipogenic therapy. Sci Signal, 2009. 2(101): p. ra82.
  101. 101. Tugnoli, V., et al., Characterization of lipids from human brain tissues by multinuclear magnetic resonance spectroscopy. Biopolymers, 2001. 62(6): p. 297-306.
  102. 102. Yates, A.J., et al., Lipid composition of human neural tumors. J Lipid Res, 1979. 20(4): p. 428-36.
  103. 103. Rudling, M.J., et al., Low density lipoprotein receptor activity in human intracranial tumors and its relation to the cholesterol requirement. Cancer Res, 1990. 50(3): p. 483-7.
  104. 104. Dhup, S., et al., Multiple biological activities of lactic acid in cancer: influences on tumor growth, angiogenesis and metastasis. Curr Pharm Des, 2012. 18(10): p. 1319-30.
  105. 105. Hirschhaeuser, F., U.G. Sattler, and W. Mueller-Klieser, Lactate: a metabolic key player in cancer. Cancer Res, 2011. 71(22): p. 6921-5.
  106. 106. Baumann, F., et al., Lactate promotes glioma migration by TGF-beta2-dependent regulation of matrix metalloproteinase-2. Neuro Oncol, 2009. 11(4): p. 368-80.
  107. 107. Halestrap, A.P. and N.T. Price, The proton-linked monocarboxylate transporter (MCT) family: structure, function and regulation. Biochem J, 1999. 343 Pt 2: p. 281-99.
  108. 108. Enerson, B.E. and L.R. Drewes, Molecular features, regulation, and function of monocarboxylate transporters: implications for drug delivery. J Pharm Sci, 2003. 92(8): p. 1531-44.
  109. 109. Pinheiro, C., et al., Role of monocarboxylate transporters in human cancers: state of the art. J Bioenerg Biomembr, 2012. 44(1): p. 127-39.
  110. 110. Froberg, M.K., et al., Expression of monocarboxylate transporter MCT1 in normal and neoplastic human CNS tissues. Neuroreport, 2001. 12(4): p. 761-5.
  111. 111. Miranda-Goncalves, V., et al., Monocarboxylate transporters (MCTs) in gliomas: expression and exploitation as therapeutic targets. Neuro Oncol, 2013. 15(2): p. 172-88.
  112. 112. Fang, J., et al., The H+-linked monocarboxylate transporter (MCT1/SLC16A1): a potential therapeutic target for high-risk neuroblastoma. Mol Pharmacol, 2006. 70(6): p. 2108-15.
  113. 113. Li, K.K., et al., miR-124 is frequently down-regulated in medulloblastoma and is a negative regulator of SLC16A1. Hum Pathol, 2009. 40(9): p. 1234-43.
  114. 114. Colen, C.B., et al., Metabolic targeting of lactate efflux by malignant glioma inhibits invasiveness and induces necrosis: an in vivo study. Neoplasia, 2011. 13(7): p. 620-32.
  115. 115. Colen, C.B., et al., Metabolic remodeling of malignant gliomas for enhanced sensitization during radiotherapy: an in vitro study. Neurosurgery, 2006. 59(6): p. 1313-23; discussion 1323-4.
  116. 116. Pouyssegur, J., F. Dayan, and N.M. Mazure, Hypoxia signalling in cancer and approaches to enforce tumour regression. Nature, 2006. 441(7092): p. 437-43.
  117. 117. Sonveaux, P., et al., Targeting lactate-fueled respiration selectively kills hypoxic tumor cells in mice. J Clin Invest, 2008. 118(12): p. 3930-42.
  118. 118. Bueno, V., et al., The specific monocarboxylate transporter (MCT1) inhibitor, AR-C117977, a novel immunosuppressant, prolongs allograft survival in the mouse. Transplantation, 2007. 84(9): p. 1204-7.
  119. 119. Porporato, P.E., et al., Anticancer targets in the glycolytic metabolism of tumors: a comprehensive review. Front Pharmacol, 2011. 2: p. 49.
  120. 120. Damaghi, M., J.W. Wojtkowiak, and R.J. Gillies, pH sensing and regulation in cancer. Front Physiol, 2013. 4: p. 370.
  121. 121. Supuran, C.T., Carbonic anhydrases: novel therapeutic applications for inhibitors and activators. Nat Rev Drug Discov, 2008. 7(2): p. 168-81.
  122. 122. Tureci, O., et al., Human carbonic anhydrase XII: cDNA cloning, expression, and chromosomal localization of a carbonic anhydrase gene that is overexpressed in some renal cell cancers. Proc Natl Acad Sci U S A, 1998. 95(13): p. 7608-13.
  123. 123. Hilvo, M., et al., Biochemical characterization of CA IX, one of the most active carbonic anhydrase isozymes. J Biol Chem, 2008. 283(41): p. 27799-809.
  124. 124. Brahimi-Horn, M.C., G. Bellot, and J. Pouyssegur, Hypoxia and energetic tumour metabolism. Curr Opin Genet Dev, 2011. 21(1): p. 67-72.
  125. 125. Chiche, J., M.C. Brahimi-Horn, and J. Pouyssegur, Tumour hypoxia induces a metabolic shift causing acidosis: a common feature in cancer. J Cell Mol Med, 2010. 14(4): p. 771-94.
  126. 126. Proescholdt, M.A., et al., Function of carbonic anhydrase IX in glioblastoma multiforme. Neuro Oncol, 2012. 14(11): p. 1357-66.
  127. 127. Zatovicova, M., et al., Carbonic anhydrase IX as an anticancer therapy target: preclinical evaluation of internalizing monoclonal antibody directed to catalytic domain. Curr Pharm Des, 2010. 16(29): p. 3255-63.
  128. 128. Sedlakova, O., et al., Carbonic anhydrase IX, a hypoxia-induced catalytic component of the pH regulating machinery in tumors. Front Physiol, 2014. 4: p. 400.
  129. 129. Currie, M.J., et al., Immunohistochemical analysis of cancer stem cell markers in invasive breast carcinoma and associated ductal carcinoma in situ: relationships with markers of tumor hypoxia and microvascularity. Hum Pathol, 2013. 44(3): p. 402-11.
  130. 130. McIntyre, A., et al., Carbonic anhydrase IX promotes tumor growth and necrosis in vivo and inhibition enhances anti-VEGF therapy. Clin Cancer Res, 2012. 18(11): p. 3100-11.
  131. 131. De Simone, G., et al., Carbonic anhydrase inhibitors: Hypoxia-activatable sulfonamides incorporating disulfide bonds that target the tumor-associated isoform IX. J Med Chem, 2006. 49(18): p. 5544-51.
  132. 132. Maresca, A., et al., Non-zinc mediated inhibition of carbonic anhydrases: coumarins are a new class of suicide inhibitors. J Am Chem Soc, 2009. 131(8): p. 3057-62.
  133. 133. Maresca, A., et al., Deciphering the mechanism of carbonic anhydrase inhibition with coumarins and thiocoumarins. J Med Chem, 2010. 53(1): p. 335-44.
  134. 134. Chrastina, A., S. Pastorekova, and J. Pastorek, Immunotargeting of human cervical carcinoma xenograft expressing CA IX tumor-associated antigen by 125I-labeled M75 monoclonal antibody. Neoplasma, 2003. 50(1): p. 13-21.
  135. 135. Siebels, M., et al., A clinical phase I/II trial with the monoclonal antibody cG250 (RENCAREX(R)) and interferon-alpha-2a in metastatic renal cell carcinoma patients. World J Urol, 2011. 29(1): p. 121-6.

Written By

Vera Miranda Gonçalves, Fátima Baltazar and Rui Manuel Reis

Submitted: 28 April 2014 Published: 25 March 2015