Modulation of Carbonic Anhydrase 9 (CA9) in Human ...

2 downloads 0 Views 1MB Size Report
common features of tumour cells (Warburg effect) which has at- tracted much ...... Warburg hypothesis and provide metabolic markers of cancer prognosis.
Current Pharmaceutical Design, 2010, 16, 000-000

1

Modulation of Carbonic Anhydrase 9 (CA9) in Human Brain Cancer Harun M. Said1, Claudiu T. Supuran2, Carsten Hageman3, Adrian Staab1,4, Buelent Polat1, Astrid Katzer1, Andrea Scozzafava2, Jelena Anacker5, Michael Flentje1 and Dirk Vordermark1,6 1

Dept of Radiation Oncology, Medical Faculty, University of Wuerzburg, Germany, 2Laboratorio di Chimica Bioinorganica, Dipartimento di Chimica, Università di Firenze188,Via Lastruccia 3, I-50019, Firenze, Italy, 3Dept of Neurosurgery, Medical Faculty, University of Wuerzburg, 4Paul Scherer Institute, Villingen, Switzerland, 5Gynaecology and Obstetrics, University of Wurzburg, 6 Dept of Radiation Oncology, Medical Faculty, University of Halle Wittenberg, Germany Abstract: Hypoxia is a crucial factor in tumour aggressiveness and its treatment resistance, particularly in human brain cancer. Tumour resistance against radiation- and chemo- therapy is facilitated by oxygenation reduction at tumour areas. HIF-1 regulated genes are mostly responsible for this type of resistance. Among these genes, carbonic anhydrase isoform 9 (CA9) is highly overexpressed in many types of cancer especially in high grade brain cancer like GBM. CA IX contributes to tumour environment acidification by catalyzing the carbon dioxide hydration to bicarbonate and protons, leading to the acquisition of metastasic phenotypes and chemoresistance to weakly basic anticancer drugs and therefore to inadequate application of radio-therapeutic or chemotherapeutic anti-cancer treatment strategies. Inhibition of this enzymatic activity by application of specific chemical CA9 inhibitors (sulphonamide derivative compounds) or indirect inhibitors like HIF-1 inhibitors (chetomin) or molecular inhibitors like CA9-siRNA leads to reversion of these processes, leading to the CA9 functional role inhibition during tumourigenesis. Hypoxia significantly influences the tumour microenvironment behaviour via activation of genes involved in the adaptation to the hypoxic stress. It also represents an important cancer prognosis indicator and is associated with aggressive growth, malignant progression, metastasis and poor treatment response. The main objective in malignant GBM therapy is either to eradicate the tumour or to convert it into a controlled, quiescent chronic disease. Sulfonamide derivative compounds with CA9 inhibitory characteristics represent one of the optimal treatment options beside other CA9 inhibitory agents or chemical inhibitory compounds against its main regulating transcription factor which is the hypoxia induced HIF-1 when applied against human cancers with hypoxic regions like GBM, bearing potential for an effective role in human brain tumour therapeutic strategies. Glycolytic inhibitors, when added in controlled doses under hypoxia, lead to a reduced accumulation of HIF-1 and can function as indirect hypoxia regulated genes inhibitors like CA9. These may be used as alternative or in conjunction with other direct inhibitors like the sulphonamide derivate compounds, chetomin or specific siRNAs, or other different chemical compounds possessing similar functionality making them as optimal tools for optimized therapy development in cancer treatment, especially against human brain cancer. Further experimental analysis towards the tumour stage specific inhibitory CA9 characteristics determination are necessary to find the optimal therapeutic solutions among the different available modalities; whether they are direct or indirect chemical, molecular or natural inhibitors to be able to set up successful treatment approaches against the different human tumour diseases.

Keywords: Brain Tumour, GBM, LGA, HIF-1, CA9, Tumour Therapy, Glycolysis, Sulfonamide derivatives, CA inhibitors (CAI). INTRODUCTION Hypoxia plays an important role within the solid tumour microenvironment and significantly influences the tumour cells behaviour via activation of genes encoding proteins involved in adaptation to hypoxic stress. Also, it plays an important role in tumour progression and it is selective for cells with enhanced glycolytic activity, causing production of large amounts of lactic acid, one of the most common features of tumour cells (Warburg effect) which has attracted much attention owing to its significant correlation with tumour progression, treatment results and the overall disease prognosis [1, 2]. Tumour tissue growth requires a sufficient O2 and nutrients supply. However, proliferating tumour cells quickly grow beyond the O2 diffusion distance from the nearest blood vessel (100150 μm), leading to a highly irregular and tortuous tumour vasculature containing arteriovenous shunts, blind ends and incomplete endothelial linings. This has an effect on blood flow, which is less efficient than in normal tissues [3, 4]. Tumour expansion is characterized by rapid growth of cancer cells when tumours establish themselves in organs host tissues. Rapid tumours growth is accompanied by cancer cell microenvironment alterations caused by the inability of local vasculature to supply enough O2 and nutrients to the rapidly dividing tumour cells [5], making hypoxia one common feature of solid tumours [6]. *Address correspondence to these authors at the Department of Radiation, Oncology, University of Wurzburg, Josef-Schneider Str. 11, D-97080 Wurzburg, Germany; Tel: ?????????; Fax: ?????????; E-mail: [email protected]

1381-6128/10 $55.00+.00

GBM are highly invasive brain tumours [7]. Transcription factor hypoxia-inducible factor-1 (HIF-1) is a key regulator of tumour cell adaptation and survival under hypoxic oxygenation conditions [8]. It regulates the expression of several genes related to O2 homeostasis in response to hypoxic stress [9]. Human GBM cells vary in their ability to survive under hypoxic conditions. Tumour hypoxia is a complex entity which is a function of O2 supply and demand. A shift towards anaerobic metabolism would decrease O 2 consumption rates leading to the improvement in tumour oxygenation [10, 11]. Under O2-limiting conditions, hypoxia-tolerant cells decrease their O2 consumption rate, whereas, hypoxia-sensitive cells continue to consume O2 at a relatively steady rate until the O 2 supply becomes exhausted [12, 13]. Tumour hypoxia is associated with adverse outcome in many malignancies. Tumour oxygenation is directly measured using microelectrodes or RuO2 microprobes, however, these techniques possess an invasive nature, restricting their use to accessible tumours [14]. Additional diagnostic tumour markers whose expression is induced by the tumour hypoxic microenvironment are necessary for definite advanced tumour characterisation according to its oxygenation status. Hypoxia-inducible factor-1 (HIF-1) is a transcription factor that consists of two subunits, HIF-1 and HIF-1 that are both bHLH proteins containing a PAS domain [15]. In normoxia, the von Hippel- Lindau tumour suppressor, which is the E3 ubiquitin ligase complex recognition component, targets HIF-1 [16, 17] leading to its ubiquitylation and consequent proteasomal degradation [16-21]. As a multi-subunit protein it regulates transcription at hypoxia response elements (HREs) containing a hypoxia binding sequence © 2010 Bentham Science Publishers Ltd.

2 Current Pharmaceutical Design, 2010, Vol. 16, No. 00

(HBS) and is composed of two basic helix-loop-helix proteins: a subunit, HIF-1, and the constitutively expressed HIF-1 subunits that is a non- O2 responsive nuclear protein playing various roles in transcription. Due to HRE positioning in the distal promoter region, HIF-1 also acts as a transcriptional enhancer. During hypoxia, the HIF-1/ heterodimer binds to the core pentanucleotide HBS (RCGTG) in the HREs of the target genes. [15, 22-26]. Among these genes, CA9 is one of the most strongly hypoxia-inducible [27]. CA IX protein is a transmembrane N-glycosylated isoenzyme localized at the cell surface in the form of dimers composed of monomeric subunits of 58/54 kDa [26, 28, and 29]. The -carbonic anhydrases (CAs, EC 4.2.1.1) are widespread metalloenzymes in higher vertebrates including humans [29]. Hypoxia, via the HIF-1 cascade, leads to strong overexpression of CA IX in many tumours with the overall consequence that the imbalance in pH in the tissue is increased. Indeed, most hypoxic tumours are acidic (pH6), in contrast to normal tissues (pH7.4). The role of CA IX in the acidification processes of hypoxic tumours has been demonstrated by the research work of the Pastorekova Group [30, 31]. Numerous potent (and sometimes selective) CA IX inhibitors have been developed in the past few years on the basis of the idea that tumour acidification processes inhibition and re-establishment of a more normal pH might lead to regression of the tumour, especially when used in combination with classical anticancer drugs. Results confirmed this hypothesis establishing CA IX as a novel drug target for the development of both diagnostic tools and therapeutic agents [32-34]. CA IX is frequently expressed in different tumour types and absent from their normal counterparts including bladder, kidney, breast, lung, head and neck, and cervix uteri tumours [29, 31] and slightly expressed in human brain [35-37]. Since HIF-1 subunits are highly inducible by different oxygenation conditions in human GBM cells, HIF-1 acts as a master regulator of numerous hypoxia inducible genes related to angiogenesis, cell proliferation/ survival, and glucose/iron metabolism including CA9. Hypoxia induced CA IX expression is usually concentrated in the perinecrotic tumour regions in GBM [38] and is associated with bad anti cancer therapeutic applications outcome especially chemotherapeutic and radiotherapeutic treatment modalities. HUMAN BRAIN TUMOUR CELLS METABOLISM Tumour cell proliferation requires rapid macromolecules synthesis including lipids, proteins, and nucleotides. Many tumour cells exhibit rapid glucose consumption, with most of the glucosederived carbon being secreted as lactate despite abundant O2 availability (the Warburg effect) [39]. Ordered pattern of metabolic and morphologic changes occurs during neoplastic cell transformation. Apparently, an increase in the flux of certain metabolic pathways such as the hexose-monophosphate shunt and glycolysis develops during transformation of many cell types. This metabolic aberration is conventionally explained as a consequence of a higher metabolic requirement [40]. Gliomas represent 50% of primary brain tumours, and their prognosis remains poor despite the advances in diagnosis and therapeutic strategies [41]. Positron emission tomography is used to study abnormalities in the glucose oxidative metabolism in human cerebral gliomas. Tumour regional cerebral glucose consumption was not depressed and regional glucose extraction ratios were similar for tumour and brain tissue, [42, and 43]. In some other experimental analysis, it was observed in human xenografted gliomas that Lactate / pyruvate ratios increased 3-4 fold and HK activity was of 2-4 fold lower than that of normal rat brain tissue, used as the control. The mitochondria-bound HK (mHK) fraction varied considerably and represented 9 to 69% of the total HK of that normal rat brain [44]. In another experimental series Lactate to pyruvate ratio was >1, suggesting that energy metabolism in LGG is glycolytic in nature, particularly in the tumours centre. Peripheral tumours samples showed increased glucose consumption and cytochrome c-oxidase activity [41]. HK binds to a mitochondrial porin involving peripheral benzodiazepine receptors. HK and peripheral

Said et al.

benzodiazepine receptors inhibition by lonidamine and diazepam led to synergistic anti tumoural activity in xenografted gliomas. These two receptors co-inhibition will lead to a decrease in glycolysis, often elevated in these tumours, without modifying energetic normal cells metabolism [45]. POTENTIAL ROLE OF GLYCOLYTIC INHIBITORS IN HUMAN BRAIN CANCER TREATMENT Glycolysis is the transformation of glucose to pyruvate which is a key step of ATP acquisition in all mammalian cells, including cancer tissues. Glucose transporters are commonly overexpressed in human malignancies enhancing glucose influx in the proliferating cancer cells [46]. Glycolytic inhibitors are particularly effective against cancer cells with mitochondrial defects or under hypoxic conditions, which are frequently associated with cellular resistance to conventional anticancer drugs and radiation therapy [47]. The energy metabolism process is considered for brain cancer treatment through metabolic targeting in the normal orthotopic tissue. Glucose transporter, GLUT-1, is enriched in the brain capillary endothelial cells and mediates facilitated glucose diffusion through the blood brain barrier. Glucose is metabolized mostly, to pyruvate, which enters neurons and glia mitochondria and is converted to acetyl-CoA before entering the TCA cycle. In well-oxygenated normal cells, pyruvate enters the mitochondria and by the enzymic activity of pyruvate dehydrogenase, it is transformed to acetyl-Co A, the ATP production substrate through the Krebs cycle [48]. Under normal conditions, 13% of glycolytic pyruvate is converted to lactate [49]. Because intracellular acidosis triggers apoptosis blocking, increased glycolytic activity by down regulating HIF-1 may reduce apoptosis of the hypoxic cells [50]. Within this context, experimental results showed that iodide acetate minimized or inhibited both HIF-1 and CA IX protein and mRNA expression in GBM cells under long term in vitro hypoxia experimental cell oxygenation conditions [35-37] as well as in other tumour cells of different origin [51]. It has previously shown that CA IX may be a therapeutic target for cancer, since, inhibition of carbonic anhydrase isoenzymes with bacteriostatic or non-bacteriostatic sulfonamides e.g. acetazolamide results in either reduced tumour invasiveness or blocked tumour growth [34, 52]. Furthermore, CA isoenzyme antagonism has been observed to augment the cytotoxic effects of various chemotherapeutic agents, including platinum- based drugs. There is a clear glucose concentration involvement in the HIF1 and CA IX expression regulation, since glycolsis inhibitor iodide acetate application result in a minimized expression of both of them in different GBM cell lines. Accumulation of HIF-1 depends on Glucose concentration, therefore, glycolysis inhibitors, when added under hypoxia, lead to a reduced HIF-1 accumulation. It has also been shown that inhibition of mitochondrial respiration leads to the HIF-1 stabilization inhibition at low O2 concentrations [53]. Interference in the glycolysis path of GBM by IAA can represent a therapeutic alternative in CA IX involved therapeutic approaches and potentially other HIF-1  regulated hypoxia induced gene. On the other hand we can use CA IX status for diagnostic purposes to potentially aid in the selection of patients who might benefit from CAIX-targeted therapies [54]. CA9 can represent an optimal target for therapeutic applications in hypoxia related GBM tumours. In GBM inhibition of HIF-1 regulated hypoxia induced genes like CA9 is accomplished via the functional interference into the tumour cell glycolysis pathway via IAA and chetomin. Cancer cell energy metabolism deviates from that of normal tissues by maintaining high glycolytic rates which has also been associated with disease progression in several tumour entities [55-57]. We and others have shown that the hypoxic accumulation of HIF-1 occurs in a celltype-specific manner [35, 51] and is strongly dependent on glucose availability [58- 60]. Glucose metabolism modulation investigation using the glycolysis inhibitors iodoacetate (IAA) or 2-deoxyglucose (2-DG), affects the hypoxia influenced HIF-1 accumulation. This hypoxic HIF-1 accumulation depends strongly on the glucose

Modulation of Carbonic Anhydrase 9 (CA9) in Human Brain Cancer

availability [58, 60] beside the strong evidence shown that HIF-1 regulates glucose metabolism and maintenance of tumour growth [61, 62]. One explanation of this effect is the increased O2 availability for prolyl hydroxylation of HIF-1 when mitochondrial O2 consumption is reduced such that hypoxia is not recognized by prolyl hydroxylases [63]. Glycolysis inhibitors, when added under hypoxia, lead to a reduced accumulation of HIF-1. The regulation of HIF-1 by inhibition glycolysis is independent of the activation by prolyl hydroxylases in HT1080 cells. Furthermore, pyruvate was not able to increase hypoxic HIF-1 levels when glycolytic inhibitors were added to HT1080 cells, suggesting that the lack of pyruvate is not the reason for the reduced accumulation of HIF-1 under hypoxic conditions when glycolysis was inhibited. In contrast to previous reports, no evidence was found for a key role of pyruvate as a glycolytic metabolite promoting HIF-1 accumulation [64]. To further investigate the mechanism by which the glucose availability or glucose metabolites interact with HIF-1, we performed realtime RT-PCR to quantify the mRNA expression of the HIF-1 gene in cells under normoxia or hypoxia treated with IAA or 2-DG. We did not observe any significant changes in HIF-1 gene expression profiles after incubation of HT1080 cells with IAA or 2-DG. These findings corresponds to previous reports showing that hypoxia inhibits mRNA translation by suppressing multiple key regulators [65, 66] and limited nutrient availability can lead to drastically reduced protein synthesis [67]. We therefore presume that glucose levels affect the expression of HIF-1 on a translational level or by phosphorylation instead of transcriptional regulation. A likely explanation for the reduced hypoxic HIF-1 accumulation after glycolysis inhibition, as compared to full glucose availability, is the glucose dependence of mRNA translation. Interaction of glycolysis and the HIF-1 pathway may explain in part the effects of glycolysis inhibitors shown in preclinical and clinical studies where such agents have increased the efficacy in chemotherapy protocols and after radiation treatment [68, 69]. One can speculate that this effect depends on a reduced intra-tumoural accumulation of HIF-1 and thereby reduced expression of HIF-1- regulated genes. Previously, it was shown that glucose deprivation leads to an activation of multiple signal transduction pathways, changes in gene expression and induction of oxidative stress which mediate glucose-deprivationinduced cytotoxicity and metabolic oxidative stress in human cancer cells [70, 71]. Glycolysis modulation reduces the hypoxic accumulation of HIF-1 protein in human tumour cells by a translational or post-translational regulatory process. Therefore, tumour glucose levels manipulation represents a potential approach to therapeutically target HIF-1. A clear involvement of glucose availability in the hypoxic HIF-1a and CA IX expression in malignant glioma cells exist since application of the glycolysis inhibitor iodoacetate led to a sharply reduced expression of both proteins and CA9 mRNA in all cell lines tested. Most previous reports have focused on the effects of HIF-1 on glucose metabolism, rather than vice versa: Glycolytic enzymes are induced by hypoxia and lactate production by glycolysis is a major cause of the acidic extracellular pH of tumours [72]. Suboptimal O2 availability switches on cellular metabolism to anaerobic pathways for ATP production, which occurs through pyruvate transformation to lactic acid via the catalytic activity of lactate dehydrogenase 5 [73, 74]. Because intracellular acidosis triggers apoptosis blocking, increased glycolytic activity by down regulating HIF-1 may reduce apoptosis of hypoxic tumour cells. It was previously shown in non-brain tumour cell lines that the hypoxic accumulation of HIF-1 and expression of CA IX in vitro depend on the glucose concentration in the medium [60, 75]. As a consequence glycolysis inhibitors, when added under hypoxia, led to a reduced accumulation of HIF-1 via a translational or post-translational effect [51]. Interference with the glycolysis of malignant gliomas by IAA may therefore represent a therapeutic approach in targeting HIF-1 or CA IX. HIF-1 inhibition has been shown to slow tumour growth in vitro and in vivo tumour

Current Pharmaceutical Design, 2010, Vol. 16, No. 00

3

models [76, 77] and to act synergistically with other treatment modalities such as radiotherapy [78]. POTENTIAL ROLE OF CHEMICAL INHIBITORS IN HUMAN BRAIN CANCER TREATMENT Most anticancer drugs are transported by either active transport or passive diffusion into cells, where they frequently undergo further metabolism (79). It is well known that CA IX is highly over expressed in many types of cancer. Its expression, which is regulated by the HIF-1 [35, 80] transcription factor, is induced by hypoxia and correlates with a poor response to classical anti cancer therapeutic approaches like chemo- and radiotherapies (37). This chemo- and radio resistance occur due to the CA IX contribution to the tumour environment acidification by efficiently catalyzing the carbon dioxide hydration to bicarbonate and protons leading to metastatic phenotypes acquisition and chemoresistance to weakly basic anticancer drugs [81]. Among them are CA IX selective inhibitors, which can be inhibited via potent inhibitors derived from acetazolamide, benzenesulfonamides and ethoxzolamide which have been shown to inhibit the growth of several tumour cells invitro and in vivo [82, 83]. These drugs are pH sensitive, therefore, it is suggested that their cytotoxic activity depend on both intracellular pH (pHi) and pHe [84]. Targeting CAIX with such specific CA IX inhibitors [31], or also antibodies [85] should contribute, on one hand to the enhancing action of weakly basic drugs and on the other hand, to reduce the acquisition of metastatic phenotypes by pH imbalance controlling in the tumour cells. [86]. Selective CA IX inhibitors could prove useful for elucidating the CA IX role in hypoxic cancers, for pH imbalance in tumour cells controlling and for developing diagnostic or human tumour disease management therapeutic applications. CA9 specific enzymatic inhibition activity by specific inhibitors belonging to the group of sulphonamides like indisulam, reverts these processes, establishing a clear-cut role for CA IX in tumourigenesis. [86]. Practically, CA inhibitors have been previously shown to elicit synergistic effects when used in combination with other chemotherapeutics agents in animal models [87]. The CA inhibitors antiproliferative effect might also be due to their effect on other CA isoforms such as CA II or CA V, which provide the bicarbonate substrate for cell growth in carboxylation reactions involved in lipogenesis, nucleotide biosynthesis and gluconeogenesis, among others, thereby limiting the unrestrained tumour cells proliferation [33, 88, and 89]. Inhibition of this enzymatic activity by specific inhibitors, such as the sulfonamide indisulam reverses these processes, establishing a clear-cut role for CA IX in tumourigenesis. Thus, selective CA IX inhibitors could prove useful for elucidating the CA IX role in hypoxic cancers, for controlling the pH imbalance in tumour cells and for developing diagnostic or therapeutic applications for tumour management. CA IX inhibitors are undergoing continuous development in order to reach a high selectivity of these drugs and to avoid any side effects by other CA isozymes inhibition and avoiding them to play their physiological roles [32]. GLUCOSE METABOLISM IN HUMAN CANCER CELLS Glucose metabolism process may occur in nature either aerobically or anaerobically. In aerobic metabolism, glucose is converted to CO2 and H2O via the TCA cycle with the generation of about 36 moles of ATP per mole of glucose consumed. In anaerobic glycolysis, glucose is metabolized to lactic acid, producing 2 moles each of ATP and H+ ions per mole of glucose [90]. For fundamental thermodynamic reasons [91], the efficiency of aerobic metabolism is achieved at the cost of decreased maximum rate, and ATP production by the respiratory pathway rapidly saturates at high levels of glucose or limited O2 supply. In the lower-yield anaerobic pathway, more of the energy from glucose degradation is used to drive the reaction, allowing a greater maximum rate of metabolism. The net ATP production rate during the anaerobic pathway, in the presence of adequate glucose, can be similar to that of the aerobic route de-

4 Current Pharmaceutical Design, 2010, Vol. 16, No. 00

spite the relative inefficiency. Malignant brain tumours from either humans or animal models lack metabolic flexibility, in contrast to normal brain that oxidizes glucose as well as ketone bodies for energy. They are largely dependent on glucose for energy [45, 9297]. Enhanced glycolysis produces excess lactic acid that can return to the tumour as glucose through the Cori cycle [98]. Normal mammalian cells under physiological conditions utilize high-yield aerobic glucose metabolism, but can adapt to periods of hypoxia by elevating the anaerobic pathway, provided the transition to hypoxia is gradual and allows for induction of response mechanisms such as HIF. The energy cost of this transition is substantial, as the output of ATP per mole of glucose is reduced by over 90%. To compensate this decreased efficiency, glycolytic flux must increase severalfold. Warburg first demonstrated tumour glucose metabolism alteration [99]. Transformed cells in vivo and in vitro typically rely on anaerobic pathways to generate ATP from glucose even in the presence of abundant O2. A rough correlation between malignancy degree and glycolytic rate has long been noted [100]. The decreased anaerobic metabolism efficiency is compensated by increased glucose, flux, maintaining energy production sufficiently in excess of basal metabolic demands to allow for cellular proliferation. METABOLIC CONTROL ANALYSIS IN HUMAN BRAIN TUMOURS Metabolic control analysis evaluates the degree of flux in metabolic pathways and can be used to analyze and treat complex diseases [101, 102]. This approach is based on findings that compensatory genetic and biochemical pathways regulate the tumour cells phenotype and bioenergetic potential [101- 103]. As rate-controlling enzymatic steps in biochemical pathways are dependent on the physiological system metabolic environment, the management of disease phenotype depends more on the flux of the entire system than on the expression of any specific gene or enzyme alone [103105]. Complex disease phenotypes can be managed through selforganizing networks that display system wide dynamics involving glycolysis and respiration. Global manipulations of these metabolic networks can restore orderly adaptive behaviour to widely disordered states involving complex gene-environmental interactions [103-106]. HYPOXIA INDUCED GENE EXPRESSION DEPENDENCY ON GLUCOSE AVAILABILITY Glycolytic enzymes are induced by hypoxia and lactate production by glycolysis is a major cause of the acidic extracellular pH of tumours [72]. Suboptimal O2 availability switches on cellular metabolism to anaerobic pathways for ATP production which occurs through pyruvate transformation to lactic acid via the catalytic activity of lactate dehydrogenase 5 [73, 74]. Because intracellular acidosis triggers apoptosis blocking, increased glycolytic activity by down regulating HIF-1 may reduce apoptosis hypoxic tumour cells [50]. We could previously show in non-brain tumour cell lines that hypoxic HIF-1 accumulation and CA IX expression in-vitro depend on the glucose concentration in the medium [60, 75]. Glycolysis inhibitors, when added under hypoxia, led to a reduced HIF1 accumulation via a translational or post-translational effect [51]. Interference with the glycolysis of malignant GBM by IAA may therefore represent a therapeutic approach alternative. In GBM, there is a clear involvement of glucose availability in hypoxic HIF1 and CA IX expression in malignant GBM cells since application of the glycolysis inhibitor IAA led to a sharply reduced expression of both proteins and CA9 mRNA in all cell lines tested. Most previous reports have focused on the effects of HIF-1 on glucose metabolism, rather than vice versa. REGULATION VIA OTHER HYPOXIA GENE REGULATORS Early growth response factor 1 (Egr-1) is a transcription factor that triggers transcription of downstream genes within 15-30 min of

Said et al.

various stimulations [107]. In several publications [108, 109], it has been shown that the transcription factor Egr-1 is regulated via hypoxia, and hypothesized to be responsible for the hypoxia induced regulation of the NDRG1 gene in human tumours. The von HippelLindau tumour suppressor (pVHL), which is the recognition component of an E3 ubiquitin ligase complex, targets HIF-1 in normoxia [15, 16, and 17], leading to its ubiquitylation and consequent proteasomal degradation [17-19, 21]. Within this context, HIF-1 is responsible of several hypoxia induced genes regulation. These genes are expressed rapidly through specific promoter activation as addressed with its acting elements and regulative factors here previously. HIF-1 subunits are highly inducible and hypoxia responsive element (HRE) are bound by nuclear HIF-1 in human GBM cells in vitro under different oxygenation conditions, Also, the clear enhanced binding of nuclear extracts from GBM cell samples exposed to extreme hypoxic conditions confirms the HIF-1 western results [35-37, 110] and the NDRG1 regulation by HIF-1. Our findings [110] demonstrated that Egr-1 was not up-regulated in response to the extreme hypoxic (0.1 %O2) or even reoxygenative conditions after hypoxia in human GBM. As a consequence, HIF1 can still be considered as one of the promising targets for new therapeutic strategies in cancer research, especially therapeutic modulation of the adaptive hypoxic response. At the same time, we believe, based on the data that resulted from the research work of the different research groups worldwide, that Egr-1 could not be a new target for therapeutic modulation of the adaptive hypoxic response, at least in human GBM. SULPHONAMIDE DERIVATIVES AND THEIR ACTIVE AND POTENTIAL Role in the Treatment of Different Diseases Today, different chemical compounds exist either in nature or are synthetically developed possessing the functional ability to act as medical treatment agents that are of various natural appearance and chemical activity. Sulphonamide derivatives are among such compounds that play such an important role since more than half a century [88, 89]. The spectrum of diseases of different pathological origin or disease pathology that can be treated by them or can be the potentially treated by them is large. A group of such compounds function as anti-epileptic agents [111], anti-glaucoma agents [112], anti-Alzheimer’s disease agents [113], Diuretics [114], inhibition of the malaria parasite pathogenicity [115], migraine therapy [116], chronic parodontitis treatment [117], diabetic retinopathy prevention [118], and neuropathic pain inhibition [119] beside the treatment or potential treatment of many others diseases different other that need to be determined. Sulphonamide compounds beside other chemical compounds can function as diagnostic or medical treatment agents against cancer diseases in human especially as anti CA agents [120]. POTENTIAL ROLE OF CA9 INHIBITION BY CAI IN TUMOUR THERAPY CA9 inhibition is realized at various regulatory levels; while the indirect inhibitory occurs by the inhibition of the hypoxia induced regulatory transcription factor HIF-1, both on the protein and mRNA level [35-37, 51], the direct inhibition occurs after CA9 expression at CA IX protein or CA9-mRNA and level with the consequence that inhibitory power activity is in general different depending on the mode of inhibition, the level of inhibition-entry and the cell type where the functional inhibition take place. In general are two different alternative regulatory pathways how hypoxia induced CA9 regulative activity can be inhibited, direct or indirect manner. As shown in (Fig. 2). Indirect CA9 inhibition is mostly achieved via the hypoxia induced HIF-1 activity inhibition. This inhibition can be achieved either via the application of chemical substances like; CA inhibition via IAA, inhibition via Isoflavons [121], inhibition via Isothyocyanate [122], inhibition via Melotonin

Modulation of Carbonic Anhydrase 9 (CA9) in Human Brain Cancer

Current Pharmaceutical Design, 2010, Vol. 16, No. 00

5

Fig. (1). HIF-1 induced regulation of hypoxia induced CA9 expression in human tumour cells A) Under normoxic oxygenation conditions in the tumor cell microenvironment, HIF-1 is rapidly degraded via the von Hippel-Lindau tumour suppressor gene product (pVHL)-mediated ubiquitin proteasome pathway. B) When the tumor environment aeration conditions shifts from normoxic to hypoxic aeration conditions, HIF-1 subunit becomes stable, translocates into the cellular nucleus and interacts with co-activators of which its transcription machinery is consisted such as p300 / CBP to modulate the transcriptional activity of numerous hypoxia inducible genes, like CA9 in our case and about 61 other hypoxia induced genes [136, 137].

6 Current Pharmaceutical Design, 2010, Vol. 16, No. 00

Said et al.

Fig. (2). Overview of the potential CA9 inhibition alternatives in human tumour cells In general, there are two different possible alternatives in which hypoxia CA9 regulative activity can be inhibited in a direct or indirect manner. The indirect CA9 is mostly achieved via inhibition of the hypoxia induced HIF-1 activity. This inhibition can be achieved either via the application of chemical substances like; CA9-inhibition via IAA, inhibition via Isoflavons, Inhibition via Isothyocyanate, inhibition via Melotonin, inhibition via Epithidiketopiperazines, CA9inhibition via chetomin or direct functional molecular Inhibition via HIF-1-siRNA. Direct hypoxia induced CA9-inhibition can be achieved via the direct inhibitory activity of mainly chemical derivatives of sulphonamide compunds or via direct inhibition by the inhibitory activity of CA9-siRNA molecules.

[123], inhibition via Epithidiketopiperazines [124]. CA9 expression inhibition via chetomin is another inhibitory option, were HIF-1 targeting human fibrosarcoma cell line by chetomin (150 nM) leads to transcriptional response to hypoxia suppression and reduces hypoxic radioresistance in these tumours, in vitro, showing increased radiosensitivity of hypoxic cells in vitro in response to chetomin as a premier observation [125]. Experimental conditions chosen were based on previous observation that a near-maximal HIF-1 expression occurs after 12h of hypoxia at an O2 concentration of 0.1% O2, which represents a level of hypoxia that is frequently observed in solid tumours and radiobiologically relevant [60, 75]. At the dose level of 150nM, chetomin exhibited a maximum specific effect on HRE-regulated activity inhibition. Other available options are the direct functional molecular inhibition via HIF-1-siRNA [126-131]. Confirmed direct hypoxia induced CA9 inhibition can be achieved either via the direct inhibitory activity of mainly chemical derivatives of sulphonamide compounds [120, 132-134] or via direct inhibition by the inhibitory activity of CA9-siRNA molecules [121]. COMPERATIVE EXPERIMENTAL ANALYSIS OF CA9 EXPRESSION AND ITS Expression Inhibition in Human Glioblastoma Cells In different in vitro CA9 expression analysis series, [35-37], we could observe that different molecular mRNA analysis methods resulted in experimental results with similar CA9 expression tendency and repetitive regulative CA9 expression values. Whether

northern blotting or RT- PCR was used to examine the differences on hypoxia dependent CA9 mRNA expression levels between different GBM cells under different standardized aeration and time conditions (normoxic or hypoxic aeration conditions). Interesting differences on CA9 mRNA expression levels between these GBM cells depending on the tumour micromilieu aeration conditions existed and where displayed in a form of a clear O2 concentration dependent CA9-mRNA expression, while HIF-1 mRNA showed no expression regulation due to hypoxic development in cell lines examined. On the protein level [35] a weak CA IX protein expression pattern was shown under normoxic oxygenation conditions, in U251, U373 and GaMG, while, in U87-MG it reached the positive control expression level. Under different hypoxia conditions examined, an increase in CA IX till 24h and a relative stability over 48h after reoxygenation without differences in the O2 concentration occurred, with respect to the difference in the CA IX expression levels between the different GBM cell lines. Incubation under 24h of hypoxia, with 50M IAA (0.1% O2) minimized the CA IX expression 90 % when compared to the basic normoxic expression level. Expression of CA IX and HIF-1 were stable and comparable to the expression pattern of hypoxic and reoxygenation conditions after hypoxia. In a comparative in vivo CA9 expression analysis series from two groups of tumour specimens collected from brain tumour patients suffering from either LGA or GBM [37], this, semi quantitative RT-PCR results showed a tumour grade associated CA9 expression pattern. CA9-mRNA was found uniformly upregulated in GBM, compared to normal brain and to LGA. An increase

Modulation of Carbonic Anhydrase 9 (CA9) in Human Brain Cancer

of at least 2 fold in CA9 expression was shown in 3/15 patients (LGA) and 12/15 patients in (GBM). HIF-1 mRNA showed no regulation in vivo and in vitro due to hypoxic development in all cell lines examined. In parallel, expression analysis experiments on protein level in this two human patient groups analyzed showed that CA9 and HIF1 were expressed in GBM at a higher rate than in LGA. On the other hand, in another experiments series published here for the first time, we were able to determine the physiological behaviour of hypoxia induced CA9 in response to the regulative effect of sulphonamide derivative CAI in GBM (135). CA IX protein level expression under normoxic conditions was detectable in a cell-type specific manner. U87- MG exhibited a strong aerobic expression, U251 and U373 had moderate and GaMG very weak normoxic CA IX protein expression pattern as shown in Fig. 3. We could demonstrate that sulphonamide derivative CAI with a molecular weight of 157.20 g/mol [135] at functional concentrations between (3500 nM8000 nM) displayed inhibitory characteristics against hypoxia in-

Current Pharmaceutical Design, 2010, Vol. 16, No. 00

7

duced CA9 expression in the four GBM cell lines examined under 24h of hypoxia (0.1 % O2). Parallel experiments with the same tumour cell lines where CA9-siRNA under similar conditions applied confirmed these results. Also, we could demonstrate the ability of HIF-1 inhibitor chetomin in the hypoxia induced HIF-1 regulated CA9 expression in GBM (Fig. 4A, Fig. 4B), under similar oxygenation conditions at concentrations of chetomin between (150nM-500nM). These CA9 inhibition results especially via a functional CAI sulphonamide derivatives inhibition confirms the functional capability such compounds to regulate the expression of these important molecules towards the complete inhibition of these molecules, due to the cell proliferative inhibitory characteristics of these compound [120]. This parallel inhibition via other compounds like chetomin or alternative CA9-siRNA in this approach or the inhibition via IAA [35] under similar comparable conditions show that is possible to obtain such an hypoxia induced CA9 inhibitory function paired with a different level of inhibition with all the accompanied consequences. These sufonamide derivatives CAI repre-

Fig. (3). Comparative analysis of CA IX protein expression under in vitro hypoxia including hypoxia induced CA9 inhibition via chemical or non chemical alternatives-in vitro Early-passage U373, U251 and U87-MG human malignant GBM from the American Type Culture Collection (ATCC, Rockville, MD, USA) and GaMG, a cell line established from a patient with GBM (Gade Institut of the University Bergen, Norway) [138], were grown on glass Petri dishes in Dulbecco's modified Eagle's medium (DMEM) supplemented with 10% fetal bovine serum (FBS), non-essential amino acids, penicillin (100 IU/ml)/streptomycin (100 μg/ml) and 2 mM L-glutamine. Examined cells were treated with in vitro hypoxia for 1, 6, or 24 h at 5, 1 or 0.1% O2 as indicated in a Ruskinn Invivo2 hypoxic workstation (Cincinnatti, OH, USA), western blot analysis was performed as previously described [35] and detection was withn the M75 mouse monoclonal antibody against CA IX (Bayer Healthcare Co., diluted 1: 7200). The CA9-siRNA construct (Santa Cruz Biotechnology) were transfected into U373, U251, U87MG and GaMG GBM cells lines, these transient transfections were performed using Fugene6 transfection reagent (Roche Diagnostics GmbH, Mannheim, Germany) according to the manufacturer's instructions. Post-transfection cells were incubated for 8h under standard normoxic conditions (21% O2, 5% CO2) post transfection with further incubation under hypoxic conditions (0.1%) for 24 h and was able to completely inhibit the extreme hypoxia induced CA9. Densitometric evaluation of signal strengths in Western blotting or in semi-quantitative RT-PCR was performed with 1D Kodak Image Analysis Software. The proteins amount gave signals that were measured in Kodak light units (KLU) and divided by the loading control ß-actin corresponding signals. Also in parallel these cells were incubated under hypoxia with addition of (250nM-8000nM) of the CAI sulphonamide derivative with a molecular weight of 157.20 g/mol (135), or with chetomin (150-500nM). While in GaMG cells the CA9 inhibition was at comparatively low concentration of the CAI sulphonamide derivative compounds (4000nM) followed by U373 and U251, a complete inhibition was not possible in the PTEN mutated U87-MG. On the other hand the HIF-1 inhibitor chetomin was only able to down-regulate CA IX protein expression in GaMG to a basic expression level and did not change the expression substantially in the other three tumor cells analysed. Experiments were repeated for three times and the figure represents one representative experiment out of three. Treatment with 100 M DFO under aerobic conditions served as a positive control while cells incubated under aerobic conditions as negative control and -actin as a loading control.

8 Current Pharmaceutical Design, 2010, Vol. 16, No. 00

Said et al.

Fig. (4). Hypoxia induced CA9 gene mRNA expression analysis under in vitro hypoxia including the hypoxia induced CA9 inhibition via chemical or non chemical alternatives-in vitro Cell Incubation, oxygenation, treatment and transfection methods were the same as for the protein analysis and in a parallel approach. To compare the expression of the individual genes examined, , RNA was iolated, quantified and its quality was examined, RT-PCR was performed using primers designed as previously described [35, 37, 137], published information on -actin and CA9 mRNA sequences were obtained from GenBank. PCR was performed, and the PCR products were separated on agarose gels as previously described [37]. Densometric evaluation was similar as for the protein but off coarse the DNA signals were measured here and -actin served as house keeping gene. Also in parallel, these cells were incubated under hypoxia with addition of (250nM-5000nM) of the CAI sulphonamide derivative or with chetomin (150-500nM). While in GaMG cells the CA9 inhibition was at comparatively low concentrations of the sulphonamide derivative CAI (4000nM) followed by the other U373, U251 and U87-MG at 4500nM, on the other hand the HIF-1 inhibitor chetomin downregulated CAIX protein expression in GaMG at 400nM while in the other three tumor cells analysed it was possible at 300nM. Experiments were repeated for three times and the figure represents one representative experiment out of three. Treatment with 100 M DFO under aerobic conditions served as a positive control while cells incubated for 24 h under aerobic conditions as negative control and -actin as a loading control.

sent interesting candidates for the development of novel unconventional anticancer strategies targeting the hypoxic areas of tumours. We have to mention here that brain cell GBM is characterized by both, a very high CA IX expression and radiotherapy as well as chemotherapy unresponsiveness, rendering this CAI sulphonamide derivative be the leading candidate for such novel anti cancer therapeutic approaches. The CA9 inhibitory capacity of the sulphonamide derivative CAI, despite previous discussion about CA9 inhibition specificity achieved by them [120] showed from their inhibition results when they were compared to the parallel CA9 inhibition results of other modalities examined until now a similar inhibition tendency but with a different expression level inhibition and in a

tumour cell type specific manner as it is the case for the hypoxia induced expression of different cells [35-37, 51, 125], that we can say at least in GBM they are coming closer to the CA9 inhibition specificity point. Further experimental series are necessary to determine the CA9 inhibition specificity of each different CAI sufonamide derivative with taking in account the difference in both their functional activity and its specificity level depending on the tumour cell type examined. CONCLUSIONS Despite the ongoing research by the various research groups around the world, there is a further necessity towards identification

Modulation of Carbonic Anhydrase 9 (CA9) in Human Brain Cancer

and characterisation of substances that posses at least similar or as an optimal goal to be achieved better functional characteristics and that can function against a typical disease pattern of this nature where the patient treatment can not be positively affected by a such medical drugs. Interference in the GBM glycolysis pathway in addition to the other alternative approaches can represent a therapeutic alternative in CA9 involved therapeutic solutions, since CA9 represent a stable marker gene as well as one of the optimal target for therapeutic applications of human brain cancer, therefore, the expression modulation of CA9 and potentially other HIF-1 regulated genes beside CA9 cab be achieved by alternative therapeutic approaches of this nature. We were able here to provide the elementary information necessary for optimizing the human cancer treatment outcome and life quality of human brain cancer patients. Alternative optimizer of human brain cancer treatment modalities, especially further drug administration with further phase I, II and III series of clinical studies are necessary for determination of the minimal effective inhibitory concentration in brain and in other human organ tissues suffering from cancer diseases with no or non harmful minimal side effects on the treated human patient are required. Controlled CA9 expression with down-regulation to expression levels that do not affect the basic and essential cellular functions and that are able to neutralize the non beneficiary effect of the local CA9 expression in human brain tumours environment is necessary in order to improve human radio-therapeutic and / or chemotherapeutic treatment modalities to be applied in the human brain cancer patient treatment. In GBM therapy, the main aim, as with all cancer diseases is to either eradicate the tumour or convert it into a controlled, quiescent chronic disease. Angiogenesis and hypoxia induced, HIF-1 regulated genes inhibition remains the main parts of therapeutic approaches in human oncology. It is well known that cancer cell metabolism can be perturbed specifically at the level of glycolysis leading to interesting therapeutic activities in cancer that can be displayed. CA IX and also its main regulator HIF-1 (what is not the case for Egr-1, at least in GBM) represent interesting targets for anticancer drug development with considering the fact that the CAI design and effective CA9 targeting for anti-cancer purposes requires further knowledge about their molecular characteristics, functional regulation mechanisms and their distribution beside the different factors that might be unknown until now that are favouring this process.

Current Pharmaceutical Design, 2010, Vol. 16, No. 00

NDRG1 siRNA PAS

= = =

pVHL RuO2 TCA

= = =

N-myc Down-regulated Gene 1 Signal Inhibitory RNA Protein domain contained in many signalling proteins Von Hippel-Lindau tumour suppressor Ruthenium Oxide Microprobes Tricarboxylic acid

CONFLICTS OF INTERESTS STATEMENT The authors declare that they have no conflicts of interests related to the contents of this manuscript AKNOWLEDGEMENTS The authors would like to thank the University of Würzburg, Medical Faculty, Department of Radiation Oncology and the IZKF Würzburg (B25) to CH, GHV for financing this research and Bayer Healthcare Co. for provision of the M75 monoclonal antibody. Also, the authors would like to acknowledge the efforts & contributions made by the different research groups in the different aspects of the Carbonic Anhydrases research and tumour hypoxia signalling and regulation with their contributions, globally. REFERENCES [1] [2]

[3] [4] [5]

[6]

[7]

ABREVATIONS [2- D- G] = 2-Deoxy-D-glucose bHLH = Basic Helix-Loop-Helix Protein Transcription Factors CAI = Carbonic Anhydrase Inhibitors CA9 = Carbonic Anhydrase 9 DFO = Desferrioxamine Egr-1 = Early Growth Response gene 1 GBM = Gliobastoma Multiforme h = Incubation or exposition period in hours HBS = Hypoxia Binding Sequence HIF-1 = Hypoxia Induced Factor -1 (Hypoxia activated subunit) HIF-1 = Hypoxia Induced Factor -1 (constitutional subunit) HRE = Hypoxia Responsive Element of the gene promoter region HK = Hexokinases Hypoxia = O2 Concentration (0.1-5%) IAA = Iodide Acetate Normoxia = O2 Concentration (21%)

9

[8]

[9]

[10]

[11] [12]

[13] [14] [15]

Brown JM. Exploiting the hypoxic cancer cell: mechanisms and therapeutic strategies. Mol Med Today 2000; 6: 157-162. Kallio PJ, Pongratz I, Gradin K, McGuire J, Poellinger L. Activation of hypoxia-inducible factor 1alpha: posttranscriptional regulation and conformational change by recruitment of the Arnt transcription factor. PNAS 1997; 94(11): 5667-72. Pagé EL, Robitaille GA, Pouysségur J, Richard DE. Induction of hypoxia-inducible factor-1alpha by transcriptional and translational mechanisms. J Biol Chem 2002; 277(50): 48403-9. Guppy M. The hypoxic core: a possible answer to the cancer paradox. Biochem Biophys Res Commun 2002; 299(4): 676-80 Qu X, Xiao D, Weber HC. Human gastrin-releasing peptide receptor mediates sustained CREB phosphorylation and transactivation in HuTu 80 duodenal cancer cells. FEBS Lett 2002; 527(1-3): 10913. Angst E, Sibold S, Tiffon C, et al. Cellular differentiation determines the expression of the hypoxia-inducible protein NDRG1 in pancreatic cancer. Br J Cancer 2006; 95(3): 307-13. Elstner A, Holtkamp N, von Deimling A. Involvement of HIF-1 in desferrioxamine-induced invasion of glioblastoma cells. Clin Exp Metastasis. 2007; 24(1): 57-66. Hodges TW, Hossain CF, Kim YP, Zhou YD, Nagle DG. Molecular-targeted antitumor agents: the Saururus cernuus dineolignans manassantin B and 4-demethylmanassantin B are potent inhibitors of hypoxia-activated HIF-1. J Nat Prod 2004; 67(5): 767-71. Turcotte ML, Parliament M, Franko A, Allalunis-Turner J. Variation in mitochondrial function in hypoxia-sensitive and hypoxiatolerant human glioma cells. Br J Cancer 2002; 86(4): 619-24 Secomb TW, Hsu R, Braun RD, Ross JR, Gross JF, Dewhirst MW. Theoretical simulation of O2 transport to tumors by threedimensional networks of microvessels. Adv Exp Med Biol 1998; 454: 629-34. El-Kareh AW, Secomb TW. Theoretical models for drug delivery to solid tumors. Crit Rev Biomed Eng 1997; 25 (6): 503-71. Yoshida D, Kim K, Noha M, Teramoto A. Anti-apoptotic action by hypoxia inducible factor 1-alpha in human pituitary adenoma cell line, HP-75 in hypoxic condition. J Neurooncol. 2006; 78(3): 21725. Yoshida D, Kim K, Noha M, Teramoto A. Hypoxia inducible factor 1-alpha regulates of platelet derived growth factor-B in human glioblastoma cells. J Neurooncol 2006; 76(1): 13-21. Brown JM. Tumor microenvironment and the response to anticancer therapy. Cancer Biol Ther 2002; 1(5): 453-8. Wang GL, Jiang BH, Rue EA, Semenza GL. Hypoxia-inducible factor 1 is a basic-helix-loop-helix-PAS heterodimer regulated by cellular O2 tension. PNAS 1995; 92(12): 5510-4.

10 Current Pharmaceutical Design, 2010, Vol. 16, No. 00 [16]

[17]

[18] [19]

[20]

[21]

[22]

[23] [24]

[25]

[26] [27] [28]

[29]

[30] [31] [32]

[33]

[34] [35]

[36] [37]

[38]

Iwai K, Yamanaka K, Kamura T, et al. Identification of the von hippel lindau tumor-suppressor protein as part of an active E3 ubiquitin ligase complex. PNAS 1999; 96, 12436-12441. Lisztwan J, Imbert G, Wirbelauer C, Gstaiger M, Krek W. The von hippel-lindau tumor suppressor protein is a component of an E3 ubiquitin-protein ligase activity. Genes & Development 1999; 13, 1822-33. Cockman ME, Masson N, Mole DR, et al. Hypoxia inducible factor-alpha binding and ubiquitylation by the von hippel-lindau tumor suppressor protein. J Biological Chem 2000; 275, 25733-41. Kamura T, Sato S, Iwai K, Czyzyk-Krzeska M, Conaway RC, Conaway JW. Activation of HIF1alpha ubiquitination by a reconstituted von hippel-lindau (VHL) tumor suppressor complex. PNAS 2000; 97: 10430-5. Ohh M, Park CW, Ivan M, Hoffman MA, Kim TY, et al. Ubiquitination of hypoxia-inducible factor requires direct binding to the beta-domain of the von Hippel-Lindau protein. Nature Cell Biology 2000; 2: 423-7. Tanimoto K, Makino Y, Pereira T, Poellinger L. Mechanism of regulation of the hypoxia-inducible factor-1 alpha by the von hippel-lindau tumor suppressor protein. EMBO J 2000; 19(16): 4298309. Melillo G, Musso T, Sica A, Taylor LS, Cox GW, Varesio L. A hypoxia-responsive element mediates a novel pathway of activation of the inducible nitric oxide synthase promoter. J Exp Med 1995; 182(6): 1683-93. Dachs GU, Patterson AV, Firth JD, et al Targeting gene expression to hypoxic tumor cells. Nat Med 1997; 3(5): 515-20. Gu YZ, Hogenesch JB, Bradfield CA. The PAS superfamily: sensors of environmental and developmental signals. Annu Rev Pharm Toxicol 2000; 40: 519-61. Huang LE, Arany Z, Livingston DM, Bunn HF. Activation of hypoxia-inducible transcription factor depends primarily upon redoxsensitive stabilization of Its  subunit. J Biol Chem 1996; 271: 32253-9. Kuphal S, Winklmeier A, Warnecke C, Bosserhoff AK. Constitutive HIF-1 activity in malignant melanoma. Eur J Cancer 2010; 46(6): 1159-69. Lal A, Peters H, St Croix B, et al. Transcriptional response to hypoxia in human tumors. J Natl Cancer Inst 2001; 93(17): 1337-43. Pastorek J, Pastoreková S, Callebaut I, et al. Cloning and characterization of MN, a human tumor- associated protein with a domain homologous to carbonic anhydrase and a putative helix-loop-helix DNA binding segment 1994; 9: 2788-888. Závada J, Závadová Z, Pastoreková S, Ciampor F, Pastorek J, Zelník V. Expression of MaTu-MN protein in human tumor cultures and in clinical specimens. Int J Cancer 1993; 54(2): 268-74. Hynninen P, Hämäläinen JM, Pastorekova S, et al. Transmembrane carbonic anhydrase isozymes IX and XII in the female mouse reproductive organs. Reprod Biol Endocrinol 2004; 17(2): 73. Pastorekova S, Zavada J. Carbonic anhydrase IX (CA IX) as a potential target for cancer therapy. Cancer Ther 2004; 2: 245-62. Thiry A, Dogné JM, Masereel B, Supuran CT. Targeting tumorassociated carbonic anhydrase IX in cancer therapy. Trends Pharmacol Sci 2006; 27(11): 566-73. Supuran CT (2004) Carbonic anhydrases: catalytic and inhibition mechanisms, distribution and physiological roles. In: Carbonic Anhydrase. Its Inhibitors and Activators (Supuran, C.T. et al., Eds), pp. 1-23, CRC Press Svastová E, Hulíková A, Rafajová M, et al. Hypoxia activates the capacity of tumor-associated carbonic anhydrase IX to acidify extracellular pH. FEBS Lett 2004; 577(3): 439-45. Said HM, Staab A, Hagemann C, et al. Distinct patterns of hypoxic expression of carbonic anhydrase IX (CA IX) in human malignant glioma cell lines. J Neurooncol 2007; 81(1): 27-38. Said HM, Polat B, Staab A, et al. Rapid detection of the hypoxia regulated CA-IX and NDRG1 gene expression in different glioblatoma cells in vitro. Oncol Rep 2008; 20: 413-9. Said HM, Hagemann C, Staab A, et al. Expression patterns of the hypoxia-related genes osteopontin, CA9, erythropoietin, VEGF and HIF-1alpha in human glioma in vitro and in vivo. Radiother Oncol 2007; 83(3): 398-405. Vordermark D, Brown JM. Endogenous markers of tumor hypoxia predictors of clinical radiation resistance? Strahlenther Onkol 2003; 179(12): 801-11.

Said et al. [39]

[40] [41]

[42]

[43]

[44] [45]

[46]

[47] [48] [49]

[50]

[51] [52]

[53]

[54]

[55] [56]

[57]

[58]

[59] [60]

[61]

DeBerardinis RJ, Mancuso A, Daikhin E, et al. Beyond aerobic glycolysis: transformed cells can engage in glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis. PNAS 2007; 104(49): 19345-50. Bannasch P. Modulation of carbohydrate metabolism during carcinogenesis. Cancer Detect Prev 1986; 9(3-4): 243-9. Lamari F, La Schiazza R, Guillevin R, et al. Biochemical exploration of energetic metabolism and oxidative stress in low grade gliomas: central and peripheral tumor tissue analysis. Ann Biol Clin 2008; 66(2): 143-50. Rhodes CG, Wise RJ, Gibbs JM, et al. In vivo disturbance of the oxidative metabolism of glucose in human cerebral gliomas. Ann Neurol 1983; 14: 614-26. Miccoli L, Oudard S, Sureau F, Poirson F, Dutrillaux B, Poupon MF. Intracellular pH governs the subcellular distribution of hexokinase in a glioma cell line. Biochem J 1996; 313 (Pt3): 95762. Oudard S, Arvelo F, Miccoli L, et al. High glycolysis in gliomas despite low hexokinase transcription and activity correlated to chromosome 10 loss. Br J Cancer 1996; 74(6): 839-45. Oudard S, Boitier E, Miccoli L, Rousset S, Dutrillaux B, Poupon MF. Gliomas are driven by glycolysis putative roles of hexokinase, oxidative phosphorylation and mitochondrial ultrastructure. Anticancer Res 1997; 17(3C): 1903-11. Macheda ML, Rogers S, Best JD. Molecular and cellular regulation of glucose transporter (GLUT) proteins in cancer. J Cell Physiol 2005; 202(3): 654-62. Pelicano H, Martin DS, Xu RH, Huang P. Glycolysis inhibition for anticancer treatment. Oncogene 2006; 25(34): 4633-46. Harris RA, Bowker-Kinley MM, Huang B, Wu P. Regulation of the activity of the pyruvate dehydrogenase complex. Adv Enzyme Regul 2002; 42: 249-59. Clarke DD, Sokoloff L. (1999). Circulation and energy metabolism in the brain. In G. J. Siegel, B. W. Agranoff, R. W. Albers, S. K. Fisher, & M. D. Uhler (Eds.) Basic Neurochemistry (6th edition, pp. 637- 669). New York: Lippincott-Raven. Schmaltz C, Hardenbergh PH, Wells A, Fisher DE. Regulation of proliferation-survival decisions during tumor cell hypoxia. Mol Cell Biol 1998; 18(5): 2845-54. Staab A, Löffler J, Said HM, et al. Modulation of glucose metabolism inhibits hypoxic accumulation of hypoxia-inducible factor1alpha (HIF-1alpha). Strahlenther Onkol 2007; 183(7): 366-73 Nodin C, Nilsson M, Blomstrand F. Gap junction blockage limits intercellular spreading of astrocytic apoptosis induced by metabolic depression. J Neurochem 2005; 94(4): 1111-23. Mateo J, García-Lecea M, Cadenas S, Hernández C, Moncada S. Regulation of hypoxia-inducible factor-1alpha by nitric oxide through mitochondria-dependent and -independent pathways. Biochem J 2003; 376(Pt 2): 537-44. Bui MH, Seligson D, Han KR, et al. Carbonic anhydrase IX is an independent predictor of survival in advanced renal clear cell carcinoma: implications for prognosis and therapy. Clin Cancer Res 2003, 9(2): 802-11. Brizel DM, Schroeder T, Scher RL, et al. Elevated tumor lactate concentrations predict for an increased risk of metastases in headand-neck cancer. Int J Radiat Oncol Biol Phys 2001; 51(2): 349-53. Galarraga J, Loreck DJ, Graham JF, et al. Glucose metabolism in human gliomas: correspondence of in situ and in vitro metabolic rates and altered energy metabolism. Metab Brain Dis 1986; 1(4): 279-91. Isidoro A, Casado E, Redondo A, et al. Breast carcinomas fulfil the Warburg hypothesis and provide metabolic markers of cancer prognosis. Carcinogenesis 2005; 26(12): 2095-104 Kwon SJ, Lee YJ. Effect of low glutamine/glucose on hypoxiainduced elevation of hypoxia-inducible factor-1alpha in human pancreatic cancer MiaPaCa-2 and human prostatic cancer DU-145 cells. Clin Cancer Res 2005; 11(13): 4694-700. Sobhanifar S, Aquino-Parsons C, Stanbridge EJ, Olive P. Reduced expression of hypoxia-inducible factor-1alpha in perinecrotic regions of solid tumors. Cancer Res 2005; 65(16): 7259-66. Vordermark D, Kraft P, Katzer A, Bolling T, Willner J, Flentje M. Glucose requirement for hypoxic accumulation of hypoxiainducible factor-1 (HIF-1 ). Cancer Lett 2005; 230: 122-33. Griffiths JR, McSheehy PM, Robinson SP, et al. Metabolic changes detected by in vivo magnetic resonance studies of HEPA-1 wild-type tumors and tumors deficient in hypoxia-inducible factor-

Modulation of Carbonic Anhydrase 9 (CA9) in Human Brain Cancer

[62]

[63] [64]

[65] [66] [67]

[68]

[69]

[70]

[71] [72]

[73] [74]

[75]

[76] [77]

[78]

[79] [80]

[81] [82]

[83]

[84]

1 (HIF-1 ): evidence of an anabolic role for the HIF-1 pathway. Cancer Res 2002; 62: 688-95. Yasuda S, Arii S, Mori A, et al. Hexokinase II and VEGF expression in liver tumors: correlation with hypoxia-inducible factor-1 and its significance. J Hepatol 2004; 40: 117-23. Hagen T, Taylor CT, Lam F. Moncada S. Redistribution of intracellular O2 in hypoxia by nitric oxide: effect on HIF1alpha. Science 2003; 302: 1975-8. Lu H, Forbes RA, Verma A. Hypoxia-inducible factor 1 activation by aerobic glycolysis implicates the Warburg effect in carcinogenesis. J Biol Chem 2002; 277: 23111-5. Liu L, Simon C. Regulation of transcription and translation by hypoxia. Cancer Biol Ther 2004; 3: 492-7. Liu L, Cash TP, Jones RG, Keith B, Thompson CB, Simon MC. Hypoxia-induced energy stress regulates mRNA translation and cell growth. Mol Cell 2006; 21: 521-31. Faulhammer F, Konrad G, Brankatschk B, Tahirovic S, Knödler A, Mayinger P. Cell growth-dependent coordination of lipid signaling and glycosylation is mediated byinteractions between Sac1p and D pm1p. J Cell Biol 2005; 168: 185-91. Maher JC, Krishan A, Lampidis TJ. Greater cell cycle inhibition and cytotoxicity induced by 2 deoxy-D-glucose in tumor cells treated under hypoxic vs aerobic conditions. Cancer Chemother Pharmacol 2004; 53: 116-22. Varshney R, Dwarakanath B, Jain V. Radiosensitization by 6aminonicotinamide and 2-deoxy D glucose in human cancer cells. Int J Radiat Biol 2005; 81: 397-408. Yun H, Lee M, Kim SS, Ha J. Glucose deprivation increases mRNA stability of vascular endothelial growth factor through activation of AMP-activated protein kinase in DU145 prostate carcinoma. J Biol Chem 2005; 280: 9963-72. Ahmad IM, Aykin-Burns N, Sim JE, et al. Mitochondrial O2 and H2O2 mediate glucose deprivation-induced cytotoxicity and oxidative stress in human cancer cells. J Biol Chem 2005; 280: 4254-63. Semenza GL, Roth PH, Fang HM, Wang LW. Transcriptional regulation of genes encoding glycolytic enzymes by hypoxiainducible factor 1. J Biol Chem 1994; 269: 23757-63. Harris RA, Bowker-Kinley MM, Hyang B, Wu P. Regulation of the activity of the pyruvate dehydrogenase complex. Adv Enzyme Regul, 2002; 42: 249-59. Holbrook JJ, Gutfreund H. Approaches to the study of enzyme mechanisms lactate dehydrogenase. FEBS Lett 1973; 15; 31(2): 157-69. Vordermark D, Kaffer A, Riedl S, Katzer A, Flentje, M. Characterization of carbonic anhydrase IX (CA IX) as an endogenous marker of chronic hypoxia in live human tumor cells. Int J Radiat Oncol Biol Phys 2005; 61: 1197-207. Kung AL, Zabludoff SD, France DS, et al. Small molecule blockade of transcriptional coactivation of the hypoxia-inducible factor pathway. Cancer Cell 2004; 6: 33-43 Welsh S, Williams R, Kirkpatrick L, Paine-Murrieta G, Powis G. Antitumor activity and pharmacodynamic properties of PX-478, an inhibitor of hypoxia-inducible factor-1alpha. Mol Cancer Ther 2004; 3: 233-44. Moeller BJ, Cao Y, Li CY, Dewhirst MW. Radiation activates HIF1 to regulate vascular radiosensitivity in tumors: role of reoxygenation, free radicals, and stress granules. Cancer Cell 2004; 5: 429-41 Stubbs M, McSheehy PM, Griffiths JR, Bashford CL. Causes and consequences of tumour acidity and implications for treatment. Mol Med Today 2000; 6(1): 15-9. Wykoff CC, Beasley N, Watson P, et al. Hypoxia-inducible regulation of tumor-associated carbonic anhydrases. Cancer Res 2000; 60: 7075-83. Supuran CT, Scozzafava A. Carbonic anhydrase inhibitors: aromatic sulfonamides and disulfonamides act as efficient tumor growth inhibitors. J Enzyme Inhib 2000; 15(6): 597-610. De Simone G, Vitale RM, Di Fiore A, et al. Carbonic anhydrase inhibitors: Hypoxia-activatable sulfonamides incorporating disulfide bonds that target the tumor-associated isoform IX. J Med Chem 2006; 49(18): 5544-51. Hilvo M, Supuran CT, Parkkila S. Characterization and inhibition of the recently discovered carbonic anhydrase isoforms CA XIII, XIV and XV. Curr Top Med Chem 2007; 7(9): 893-9. Supuran C. Carbonic Anhydrases-An Overview. Curr Pharm Des 2008; 14(7): 603-14.

Current Pharmaceutical Design, 2010, Vol. 16, No. 00 [85]

[86] [87]

[88] [89] [90] [91]

[92] [93] [94]

[95]

[96] [97]

[98] [99] [100]

[101]

[102] [103] [104] [105] [106]

[107] [108]

[109]

[110]

11

Chrastina A, Závada J, Parkkila S, et al. Biodistribution and pharmacokinetics of 125I-labeled monoclonal antibody M75 specific for carbonic anhydrase IX, an intrinsic marker of hypoxia, in nude mice xenografted with human colorectal carcinoma. Int J Cancer 2003; 105(6): 873-81. Thiry A, Supuran CT, Masereel B, Dogné JM. Recent developments of carbonic anhydrase inhibitors as potential anticancer drugs. J Med Chem 2008; 51(11): 3051-6. Teicher BA, Liu SD, Liu JT, Holden SA, Herman TS. Carbonic Anhydrase inhibitor as a potential modulator of cancer therapies. Anticancer Res 1993; 13: 1549-56 Supuran CT. Carbonic anhydrases: novel therapeutic applications for inhibitors and activators. Nature Reviews Drug Discovery 2008; 7(2): 168-81. Supuran CT, Scozzafava A, Casini A. Carbonic anhydrase inhibitors. Med Res Rev 2003; 23: 146-89. Stryer, L. (1988). Biochemistry (3rd edition, pp. 420 -1). New York: Freeman and Company. Pfeiffer T, Schuster S, Bonhoeffer S. Cooperation and competition in the evolution of ATP-producing pathways. Science 2001; 292, 504-7. Seyfried TN, Sanderson TM, El-Abbadi MM, McGowan R, Mukherjee P. Role of glucose and ketone bodies in the metabolic control of experimental brain cancer. Br J Cancer 2003; 89: 1375-82. Mangiardi JR, Yodice P. Metabolism of the malignant astrocytoma. Neurosurgery 1990; 26: 1-19. Nagamatsu S., Nakamichi Y., Inoue N., Inoue M., Nishino H, Sawa H. Rat C6 glioma cell growth is related to glucose transport and metabolism. Biochem J 1996; 319 (Pt 2), 477-82. Roslin M, Henriksson R, Bergstrom P, Ungerstedt U, Bergenheim AT. Baseline levels of glucose metabolites, glutamate and glycerol in malignant glioma assessed by stereotactic microdialysis. J Neurooncol 2003; 61: 151-60. Floridi A, Paggi MG, Fanciulli M. Modulation of glycolysis in neuroepithelial tumors. J Neurosurg Sci 1989; 33: 55-64. Mies G, Paschen W, Ebhardt G, Hossmann KA. Relationship between of blood flow, glucose metabolism, protein synthesis, glucose and ATP content in experimentally-induced glioma (RG1 2.2) of rat brain. J Neurooncol 1990; 9: 17-28. Tisdale MJ. Biology of cachexia. J Natl Cancer Inst 1997; 89: 1763-73. Vazquez A, Liu J, Zhou Y, Oltvai ZN. Catabolic efficiency of aerobic glycolysis: The Warburg effect revisited. BMC Syst Biol 2010; 6: 4(1): 58. Burk D, Woods M, Hunter J. On the significance of glycolysis for cancer growth, with special reference to Morris rat hepatomas. J Natl Cancer Inst 1967; 38: 839-63. Veech RL. The therapeutic implications of ketone bodies: the effects of ketone bodies in pathological conditions: ketosis, ketogenic diet, redox states, insulin resistance, and mitochondrial metabolism. Prostaglandins Leukot Essent Fatty Acids 2004; 70: 309-19 Greene AE, Todorova MT, Seyfried TN. Perspectives on the metabolic management of epilepsy through dietary reduction ofglucose and elevation of ketone bodies. J Neurochem 2003; 86: 529-37. Strohman R. Maneuvering in the complex path from genotype to phenotype. Science 2002; 296: 701-3. Kacser H, Burns JA. The molecular basis of dominance. Genetics 1981; 97: 639-66. Greenspan RJ. The flexible genome. Nat Rev Genet 2001; 2: 3837. Seyfried TN, Mukherjee P, Adams E, Mulroony T, Abate LE. Metabolic control of brain cancer: Role of glucose and ketone bodies. Proc Amer Assoc Cancer Res 2005; 46: 1147. Akutagawa O, Nishi H, Kyo S, et al. Early growth response-1 mediates downregulation of telomerase in cervical cancer. Cancer Sci 2008; 99(7): 1401-6. Ellen T, Ke Q, Zhang P, Costa M. NDRG1, a growth and cancer related gene: regulation of gene expression and function in normal and disease states. Carcinogenesis 2008; 29(1): 2-8. Zhang P, Tchou-Wong KM, Costa M. Egr-1 mediates hypoxiainducible transcription of the NDRG1 gene through an overlapping Egr-1/Sp1 binding site in the promoter. Cancer Res 2007; 67(19): 9125-33. Said HM, Polat B, Hagemann C, Vince GH, Anacker J, Kämmerer U, et al. Egr-1 is not up-regulated in response to Hypoxic and oxy-

12 Current Pharmaceutical Design, 2010, Vol. 16, No. 00

[111]

[112] [113]

[114] [115]

[116] [117]

[118] [119]

[120] [121]

[122] [123]

[124]

[125]

genation conditions in Human Glioblastoma-in vitro, Molecular Medicine Reports 2009; 2: 757-63. Hilvo M, Innocenti A, Monti SM, De Simone G, Supuran CT, Parkkila S. Recent advances in research on the most novel carbonic anhydrases, CA XIII and XV. Curr Pharm Des. 2008; 14(7): 672-8. De Simone G, Di Fiore A, Supuran CT. Are carbonic anhydrase inhibitors suitable for obtaining antiobesity drugs?. Curr Pharm Des 2008; 14(7): 655-60. Temperini C, Scozzafava A, Supuran CT. Carbonic anhydrase activation and the drug design. Curr Pharm Des 2008; 14(7): 70815. Supuran CT. Diuretics: from classical carbonic anhydrase inhibitors to novel applications of the sulfonamides. Curr Pharm Des 2008; 14(7): 641-8 Krungkrai J, Supuran CT. The alpha-carbonic anhydrase from the malaria parasite and its inhibition. Curr Pharm Des 2008; 14(7): 631-40 Newman LC, Broner SW, Lay CL. Reversible anorgasmia with topiramate therapy for migraine. Neurology 2005; 65(8): 1333-4 Kurti B, Tüter G, Serdar M, Pinar S, Demirel I, Toyman U. Gingival crevicular fluid prostaglandin E(2) and thiobarbituric acid reactive substance levels in smokers and non-smokers with chronic periodontitis following phase I periodontal therapy and adjunctive use of flurbiprofen. J Periodontol 2007; 78(1): 104-11. Einarsdóttir AB, Stefánsson E. Prevention of diabetic retinopathy. Lancet 2009; 373(9672): 1316-8. Asiedu M, Ossipov MH, Kaila K, Price TJ. Acetazolamide and midazolam act synergistically to inhibit neuropathic pain. Pain 2010; 148(2): 302-8. Pastorekova S, Zatovicova M, Pastorek J. Cancer-associated carbonic anhydrases and their inhibition. Curr Pharm Des 2008; 14(7): 685-98. Singh-Gupta V, Zhang H, Banerjee S, et al. Radiation-induced HIF-1alpha cell survival pathway is inhibited by soy isoflavones in prostate cancer cells. Int J Cancer 2009; 124(7): 1675-84. Wang XH, Cavell BE, Syed Alwi SS, Packham G. Inhibition of hypoxia inducible factor by phenethyl isothiocyanate. Biochem Pharmacol 2009; 78(3): 261-72 Wang XH, Chen SF, Jin HM, Hu RM. Differential analyses of angiogenesis and expression of growth factors in micro- and macrovascular endothelial cells of type 2 diabetic rats. Life Sci 2009; 84(7-8): 240-9. Cook KM, Hilton ST, Mecinovic J, Motherwell WB, Figg WD, Schofield CJ. Epidithiodiketopiperazines block the interaction between hypoxia-inducible factor-1alpha (HIF-1) and p300 by a zinc ejection mechanism. J Biol Chem 2009; 284(39): 26831-8 Staab A, Loeffler J, Said HM, et al. Effects of HIF-1 inhibition by chetomin on hypoxia-related transcription and radiosensitivity in HT 1080 human fibrosarcoma cells. BMC Cancer 2007; 7: 213.

Received: July 5, 2010

Accepted: July 29, 2010

Said et al. [126]

[127]

[128]

[129]

[130]

[131]

[132] [133]

[134]

[135]

[136] [137] [138]

Bryant CS, Munkarah AR, Kumar S, et al. Reduction of hypoxiainduced angiogenesis in ovarian cancer cells by inhibition of HIF-1 alpha gene expression. Arch Gynecol Obstet 2010, In Press Bartholomeusz G, Cherukuri P, Kingston J, et al. In vivo therapeutic silencing of hypoxia-inducible factor 1 (HIF-1) using singlewalled carbon nanotubes noncovalently coated with siRNA. Nano Res 2009; 2(4): 279-91. Kamlah F, Eul BG, Li S, et al. Intravenous injection of siRNA directed against hypoxia-inducible factors prolongs survival in a Lewis lung carcinoma cancer model. Cancer Gene Ther 2009; 16(3): 195-205. Takahashi Y, Nishikawa M, Takakura Y. Inhibition of tumor cell growth in the liver by RNA interference-mediated suppression of HIF-1 expression in tumor cells and hepatocytes. Gene Ther 2008; 15(8): 572-82. Jiang M, Wang B, Wang C, et al. Inhibition of hypoxia-inducible factor-1alpha and endothelial progenitor cell differentiation by adenoviral transfer of small interfering RNA in vitro. J Vasc Res 2006; 43(6): 511-21. Mizuno T, Nagao M, Yamada Y, et al. Small interfering RNA expression vector targeting hypoxia-inducible factor 1 inhibits tumor growth in hepatobiliary and pancreatic cancers. Cancer Gene Ther 2006; 13(2): 131-40. Viñas-Castells R, Beltran M, Valls G, et al. The hypoxia-controlled FBXL14 ubiquitin ligase targets SNAIL1 for proteasome degradation. J Biol Chem 2010; 285(6): 3794-805. Berchner-Pfannschmidt U, Petrat F, Doege K, et al. Chelation of cellular calcium modulates hypoxia-inducible gene expression through activation of hypoxia-inducible factor-1alpha. J Biol Chem 2004; 279(43): 44976-86. Berchner-Pfannschmidt U, Wotzlaw C, Merten E, Acker H, Fandrey J. Visualization of the three-dimensional organization of hypoxia-inducible factor-1 alpha and interacting cofactors in subnuclear structures. Biol Chem 2004; 385(3-4): 231-7. Garaj V, Puccetti L, Fasolis G, et al. Carbonic anhydrase inhibitors: synthesis and inhibition of cytosolic/tumor-associated carbonic anhydrase isozymes I, II, and IX with sulfonamides incorporating 1,2,4-triazine moieties. Bioorg Med Chem Lett 2004; 14(21): 542733. Semenza GL.Targeting HIF-1 for cancer therapy.Nat Rev Cancer 2003; 3(10): 721-732 Said HM, Stein S, Hagemann C, et al. Oxygen-dependent regulation of NDRG1 in human glioblastoma cells in vitro and in vivo. Oncol Rep 2009; 21(1): 237-246. Akslen LA, Andersen KJ, Bjerkvig R. Characteristics of human and rat glioma cells grown in a defined medium. Anticancer Res 1988; 8: 797-803