Loading

Metformin in Cancer: Chemical Pathways for Tumoral Control Independent of Amp-Dependent Kinase

Review Article | Open Access

  • 1. Endocrine and Metabolic Diseases Research Center. School of Medicine, University of Zulia, Venezuela
+ Show More - Show Less
Corresponding Authors
Joselyn Rojas, Endocrine and Metabolic Diseases Research Center, School of Medicine, University of Zulia, 20th Avenue, Maracaibo 4004, Venezuela, Tel: 582617597279; Fax: 582617597279.
Abstract

Dimethylbiguanide (Metformin) is an anti-hyperglycemic drug used in the management of insulin resistance-related diseases like type 2 diabetes mellitus for over 40 years. The molecular mechanisms of action related to its metabolic effects are linked to dependent and independent AMP-dependent kinase (AMPK) activation. Epidemiological evidence has suggested that metformin administration has been associated to lower cancer risk and cancer-related mortality in patients with type 2 diabetes mellitus. Anti-tumoral properties have been described via AMPK-dependent and independent pathways. Metformin is a cationic molecule capable of copper sequestration and subsequent mitochondrial electron transport inhibition, compromising oxidative phosphorylation. Moreover, metformin has also been confirmed to modulate pluripotency in cancer stem cells, blunting their survival and proliferation. Finally, the reduction of vitamin B12 availability has also been suggested to control one-carbon metabolism, DNA synthesis and cytostatic effects. The purpose of this review is to examine the AMPK-independent related mechanisms concerning tumor expansion control and survival.

Citation

Rojas J, Chávez-Castillo M, Torres W, Arraiz N, Cabrera M, et al. (2014) Metformin in Cancer: Chemical Pathways for Tumoral Control Independent of Amp-Dependent Kinase. J Endocrinol Diabetes Obes 2(2): 1036.

Keywords

•    Metformin
•    Biguanides
•    Cancer
•    AMPK
•    Copper sequestration
•    Mitochondrial toxicity
•    Vitamin B12

INTRODUCTION

Metformin is one of the most frequently prescribed drugs for insulin resistance-related conditions, such as Type 2 Diabetes Mellitus (T2DM) [1] and Polycystic Ovary Syndrome [2]. This medication has proven to enhance insulin sensitivity in liver [3] and skeletal muscle [4], improving metabolic control in these patients. The success of this outstanding drug relies on the activation of AMPK-dependent pathways [5] which result in greater glucose uptake [6], modulation of lactate production [7], inhibition of gluconeogenesis [8] and protein synthesis [9], activation of fatty acid β-oxidation [10], and central regulation of appetite [11]. Successful management of diabetic patients with metformin has been observed for over 20 years [12,13], being particularly preferred as first line of treatment as it allows inhibition of gluconeogenesis and improvement of impaired fasting glucose [14,15] while being less prone to inducing hypoglycemia, in comparison to other oral antidiabetic agents [16].

The relationship between diabetes and cancer is a two-way street. Johnson and Pollak have presented interesting insight explaining the intricate biological relationship between T2DM and inflammation-related neoplasia, as in colorectal, breast, and pancreatic cancer [17]. The mixture of a poor quality diet –low in antioxidants and rich in calories– associated with physical inactivity is related to hyperinsulinemia and insulin resistance. In turn, these factors have been related to development of obesity, inflammation and cancer [17,18]. Various meta-analyses have related diabetes with risk of several types of cancer after adjusting covariables, including bladder (RR=1.24; 95% CI: 1.08-1.42) [19], endometrial (RR=2.10, 95% CI: 1.75-2.53) [20], breast (RR=1.24; 95% CI:0.95-1.95) [21], hepatocellular (RR=2.5; 95% CI:1.9-3.2) [22] and even non-Hodgkin Lymphoma (RR=1.18; 95% CI:1.30- 2.47) [23], as well as higher mortality risk from colorectal cancer: RR=1.26, 95% CI: 1.05-1.50 [24]. In a population–based cohort study with over 65 thousand patients with no previous history of diabetes or cancer, Lin [25] reported that the greatest decrease in cancer risk was observed in those patients that had more than 1080 Defined Daily Doses, with a hazard ratio of 0.27 (95%CI 0.09-0.84), conferring a dosage-dependent effect to this drug. R

ecent evidence has suggested that T2DM patients who are long-term metformin users have lower risk for breast cancer. In a nested case-control study with T2DM patients over 50 years of age, Bosco [26] found long-term metformin users less likely to be diagnosed with breast cancer (OR=0.77; 95% IC: 0.61- 0.99). Likewise, Little [27] reported a borderline-significant negative association between metformin use and development of pancreatic cancer (OR=0.28; 95% CI: 0.06-1.22, p=0.09); which was absent in the group using sulphonylureas (OR=0.38; 95% CI: 0.08-1.70, p=0.20). These results were later supported by findings from Soranna [28]. Furthermore, a meta-analysis by Noto [29], ascertained a pooled risk ratio of 0.68 (95% CI: 0.53- 0.88) for all-cancer incidence and 0.66 (95% CI: 0.49-0.88) for cancer mortality, suggesting that metformin in T2DM patients lowered the risk for cancer development, including colorectal, liver and lung cancer.

Although current epidemiological evidence requires improved methodology in order to avoid time-related bias [30], this data have led to multiple in vitro and animal studies that appear to confirm the anti-tumoral activity of metformin, founded in various mechanisms dependent and independent of AMP kinase (AMPK) activity. The purpose of this review is to explore the anti-tumoral AMPK-independent properties of metformin, and correlate their efficacy against tumor cell metabolism.

PHARMACODYNAMICS AND PHARMACOKINETICS OF METFORMIN

Metformin, (N,N-Dimethylimidodicarbonimidic diamide or dimethylbiguanide), is a molecule belonging to biguanide family of drugs, very well-known for its hypoglycemic effect in animals since 1929, which was later confirmed in humans along with its sister drug, phenformin (phenylethyldiguanide) in 1957 [31]. In 1960, Kruger [32] pinpointed biguanides to exert their effects through modulation of carbohydrate metabolism and lowering oxygen uptake, inducing a pseudoanoxia state, which in turn induces anaerobic glycolysis, lactate production and lower glycemia. Additionally, these effects are observed solely in individuals with altered glucose metabolism –and not in healthy subjects– and they don’t appear to involve stimulation of insulin secretion [33].

Dimethylbiguanide is a hydrophilic molecule with positive charge at physiologic pH, which is not metabolized during its passage through the liver and circulatory system [34]. It has a molecular weight of 129,164 g/mol, with 50±16% bioavailability a half-life of ~5 hours and a renal clearance rate of 510±130 mL/min in subjects with appropriate kidney function [35]. Therapeutic levels of the drug range between 0.5-1.0 mg/L [34], with a threshold of 2.5 mg/L for lactic acidosis [35].

Absorption and distribution of metformin rely on a series of bidirectional transporters from 3 protein families in the intestine, liver and kidney [36,37]; Figure 1. Translocation to the enterocyte depends on two transporters: PMAT (Plasma Membrane Monoamine Transporter, SLC29A4; OMIM 609149) and OCT3 (Organic Cation Transporters, SLC22A3; OMIM 604842). PMAT is a 530-aminoacid protein which contains 11 transmembrane segments, whereas OCT3 codifies a 556-aminoacid protein with 12 transmembrane segments. Once inside the enterocyte, metformin enters interstitial fluid and subsequently reaches the liver via the OCT1 (SLC22A1; OMIM 602607), OCT2 (SLC22A2; OMIM 602608) OCT3 transporters, which are located in the basolateral cell membranes [38]. In the liver, the Multidrug and Toxicity Extrusion protein, MATE1 (SLC47A1; OMIM 609832), allows transport of metformin through canaliculi [39]. Once in the circulatory system, metformin enters the kidney via OCT2 which uptakes the drug from blood; and then excretes it through tubular ducts, via OCT1, MATE1 and MATE2 (SLC47A2; OMIM 609833) [40-43]. Despite this wide array of transporters implicated in the distribution and excretion of metformin, only polymorphisms of OCT1 appear to effectively impact the potential pharmaceutical effect of the drug [44,45].

Once inside target cells, metformin induces the activation of anti-tumoral effects through both AMPK-independent and dependent pathways. The former mechanisms include inhibition of the mitochondrial respiratory chain [46], modulation of stemness profile [47], and copper sequestration [48]. The latter property depends on metformin’s chemical properties: As a cationic protein, it is capable of forming a pseudoaromatic square planar complex between 2 dimethylbiguanide and 1 copper ion via delocalization of their π-electron [48,49], a quality which is mandatory for effective anti-hyperglycemic biguanides Figure 2 [50]. This modification allows the conformation of a pseudoaromatic structure, relying on the van der Waals isosurfaces which make the Y-aromaticity possible [51] and the assembly of Shift-base ligands (Salen-like complexes) [52], interfering with copper mitochondrial metabolism [53]. And even more interestingly, the activation of AMPK is blunted by the absence of the delocated electron, rather than by the phosphorylation of its downstream target, protein S6 [48]. Metformin’s effects on the respiratory chain and stemness profile modulation will be further discussed in the Mitochondrial Toxicity section.

OVERVIEW OF CANCER CELL METABOLISM

The key to understanding metformin as a potential anti-tumoral drug lies in the unique metabolism of cancer cells, which is characterized by a high proliferation rate, increased glucose oxidation, and the capacity to thrive during hypoxia [54]. Two hypotheses –the Warburg Effect and the Crabtree Effect– attempt to explain the acquisition of multiplicative capacity in the ever-increasing hypoxic environment typical of these cells [55,56].

The Warburg effect, named after the iconic publication by Dr. Otto Warburg in 1956 [57], describes the activation of an aerobic glycolytic phenotype, with induction of a “fermentative” metabolism and lactate production (Figure 3). This scenario is characterized by increased expression of glucose transporters [58], LDH5 isoform [59] and pyruvate kinase M2 (PK-M2) isoform [60], alongside induction of glutaminolysis and acceleration of the Krebs cycle [61]. Because cancer cells survive on a pseudoanaerobic metabolism, glucose uptake must increase parallel to the their needs; therefore, glucose transporters such as SGLT-1 (SLC5A1; OMIM 182380), Glut1 (SLC2A1; OMIM 138140), Glut2 (SLC2A2; OMIM 138160) and Glut3 (SLC2A3; OMIM 138170) are overexpressed in cancers such as colorectal, head and neck, lung, pancreas, endometrial, breast, ovary, and liver [62], while Glut4 (SLC2A4; OMIM 138190) has been observed in breast and gastric cancer [62]. The expression of LDH5 favors the metabolism from pyruvate to lactate due to its higher affinity for pyruvate [63], granting the fermentative phenotype. Finally, the PK-M2 isoform has been considered fundamental to maintain tumor growth and expansion due to its lower affinity for phosphoenolpyruvate (PEP) and lower enzymatic activity [64,65]. However, this lower activity of PK-M2 has not been associated with impaired conversion of PEP to pyruvate; instead, Vander Heiden [66,67] have observed that Phosphoglycerate Mutase-1 (PGM1) can convert PEP by transference of the phosphate to the catalytic pocket of the mutase, concomitantly increasing serine and glycine biosynthesis [66,67].

Aminoacid metabolism in cancer cells is complex, fundamentally encompassing 3 amino acids: glutamine, glycine and serine. Glutaminolysis is key for providing carbon skeletons to fuel the Krebs cycle via production of α-keto-glutarate, increasing the production of ATP [68,69]. Serine de novo synthesis is associated with the increased offer of one-carbons for DNA synthesis and proliferation [70] and acquisition of selective advantage [71], mechanism especially observed in triple negative breast cancer [72]. Interestingly, serine is a known allosteric activator of PK-M2, fulfilling a full circle on the relationship between PK-M2, PGM1, and pyruvate diversion towards mitochondria [66,67,73]. Finally, glycine de novo biosynthesis from threonine is associated with nucleotide synthesis, tumor cell multiplication and poor prognosis [74].

To summarize, it has been confirmed that cancer cells survive on glucose-dependent pathways, whose metabolites are deviated towards the production of lactic acid and ATP. Despite anaplerotic pathways contributing to the Krebs cycle (e.g., glutaminolysis), this process is weakened because it is also used as a source for macromolecular synthesis during cell proliferation [75]. Glucose-induced inhibition of mitochondrial respiration is known as the Crabtree effect [76,77]; Figure 4. This phenomenon has been related to the competition between glycolysis and respiratory chain for ADP as a main substrate and the effects of calcium inside the mitochondria [77,78]. This “custodial fight” stops mitochondrial functionality, and ATP production becomes exclusively cytosolic [77].

One cell model to explain the intricacy of these effects was published by Suchorolski [79], using Barrett´s esophagus (BE) cell lines. These investigators chose this cell line due to their energetic characteristics during oncogenesis, where esophageal adenocarcinomas show the Warburg effect [80]. Consequently, BE cells are at a crossroads towards choosing a glycolytic pathway or adaptation towards dysfunctional mitochondria. The study [79] demonstrated that in the early stage of BE the cells used oxidative phosphorylation as the main source of ATP, conveying functional mitochondria. Nonetheless, as transformation continues, late stages of BE are mainly Crabtree-dependent yet retain active mitochondria. Finally, when adenocarcinoma fully develops, mitochondrial metabolism is uncoupled and the Warburg effect ensues. In metaplastic cells, the Crabtree effect provides a survival advantage and allows endurance of the tumor cells even during variations of glucose and oxygen availability. These biochemical features are also observed in cancerous cells, such as leukemia cells [81] which demonstrate various metabolic profiles relating to aerobic glycolysis and mitochondrial production of ATP.

METFORMIN AND MITOCHONDRIAL TOXICITY IN CANCER CELLS

As can be observed from the previous sections, viable and active mitochondria are necessary; albeit progressively losing their efficacy and coupling properties during tumor cell expansion and intensification of hypoxia. Mitochondrial toxicity is one of the most effective mechanisms at killing highly replicating cells, such as cancerous cells. Metformin has been proven to exert 2 specific mitochondriotoxic effects: inhibition of respiratory complex I and inhibition of respiratory complex IV and other copper-dependent proteins.

The mitochondrial respiratory chain is a supercomplex ensemble, conformed by Complexes I through IV [82,83]; Figure 5. Complex I encompasses an L-shaped protein NADH:ubiquinone oxidoreductase, coupled with Flavin Mononucleotide and 7 Ferrous-Sulfur (FeS) clusters [82-84]. This first complex oxidizes intermediary metabolism-derived NADH to NAD+ , pumping protons towards the intermembrane space and passing electrons towards Complex II. Succinate:ubiquinone oxidoreductase is the enzyme component of Complex II, along with 3 FeS clusters [82,83,85]. This second complex is not a proton pump, yet it serves to mobilize reducing equivalents from FAD2 towards the next complex. The next in line is Complex III, the Cytochrome c oxidoreductase, which oxidizes ubiquinone moving electrons via cytochrome b/c1 towards Complex IV, while pumping protons outside mitochondrial matrix [82,83,86]. Finally, the Cytochrome c oxidase is the last piece of this machinery, ending the reduction of 1 oxygen molecule with 4 protons, rendering 2 H2 O molecules [82,83]. The model of activation and coordinated functioning for respiratory complexes are encircled in the super complexes models [82], where the assembly of all the complexes nearby the ATP Synthase (Complex V) [87] ensures electron flux, controlled reactive oxygen species (ROS) production, and maintains a thermodynamically working environment. The uncoupling of these supercomplexes, especially the separation of Complex I from the rest of the components, is associated with higher generation of ROS [88].

For over 10 years, it has been known that metformin can exert anti-hyperglycemic effects through the inhibition of mitochondrial energy metabolism. Owen [46] designed an elegant in vitro model with hepatoma cells for the evaluation of the inhibitory effect of metformin on mitochondrial membrane potential, reporting that the drug was capable of inducing a time-and dose-dependent blockage of the Krebs cycle, increasing the production of lactate and lactate shuttles. Metformin accumulation on mitochondria depends on a steady membrane potential of ~180 mV, with an average mitochondrial quantum 1000-fold that of plasma concentration (~100 mM); yet the liver it’s the first-passage organ and the levels that circulate in this organ are more than enough to induce this inhibitory effect on the respiratory chain, blocking 50-80% of glycolysis and paralleled induction of glucose uptake via Glut1/4. These results have also been observed in other cells lines such as pancreatic ductal adenocarcinoma [89,90], colorectal cancer [91], and other hepatoma/liver mitochondria experiments [92,93].

Inhibition of Complex I am associated with enhanced production of ROS and induction of apoptosis, observed not only in physiologic mitochondrial investigation [88], but also during experiments with Delocalized Lipophilic Cations (DLC). These cationic proteins are highly hydrophobic positively charged components, intensely drawn towards the mitochondrial matrix, which is negatively charged. It has been recognized that DLC´s inhibit oxidative phosphorylation and cell death, making them perfect weapons against tumor cells [94,95].One of these DLC is [Cu(isaepy)2], also known as Isatin-Schiff base Copper(II), which is a known apoptosis inducer due to respiratory Complex I inhibition in neuroblastoma cells [96]. Induction of cell death was blunted when a copper chelator was incubated with the cells, proving the key role of this ion in this kind of DLC [93]. The chemical structure of [Cu(isaepy)2] is quite similar to that of the pseudoaromatic square planar complex between cupper and metformin [48-53], and they share physical characteristics, which makes it very plausible that metformin works as a DLC molecule when sequestering copper inside the mitochondria, inducing oxidative stress [48,53].

Indeed, copper mitochondrial metabolism is a highly conserved mechanism which relies on copper chaperones and their copper-mobilizing capacity from cytosol to mitochondrial matrix [97,98]. In tumor cells, copper is essential for growth, regulation of oxidative phosphorylation and modulation of glucose expenditure for energy production (99); therefore, copper insufficiency is considered cytotoxic for such cells [99,100]. This copper chelating property of metformin is likely responsible for the inhibition of Complex IV, due to the disassembly of copper binding sites CuA and CuB [101,102]. The Cytochrome c oxidase is fundamental to maintain membrane voltage during high-input periods, thanks to the gating property of Glu242 which is close to Heme α3 and CuB [101].

Overall, metformin seems to be a bona fide copper chelator and DLC, which disrupts activity of Complexes I and IV; and probably interferes with copper-dependent proteins such as Superoxide Dismutase.

THE ROLE OF METFORMIN IN CELL CYCLE AND PLURIPOTENCY MODULATION

As a well-known activator of AMPK [6], the analysis of metformin’s AMPK-independent functions can be quite challenging, especially concerning its effects on the cell cycle. Metformin-activated AMPK is known to consequently activate TSC2 (Tuberous Sclerosis Complex), which inhibits RagC/mTOR1 in several cancer lines from breast cancer [103], pancreatic intreaepithelial neoplasia [104], head/neck [105] and skin [106] squamous cell carcinoma and nasopharyngeal carcinoma [107]. The blocking on mTOR1 modulates processes further down such as: protein synthesis, lipogenesis, pentose phosphate pathway and the very important inhibition of autophagy [108]. However, some properties have been attributed to mechanisms still unclear, yet seemingly AMPK-dependent, such as the control of pluripotency [47] and activation of p53 [109]; Figure 6.

Vazquez-Martin [110] published their results about the influence of AMPK in the modulation of stemness in induced Pluripotent Stem Cells (iPSC), centering in the expression of reprogramming factors such as Oct4, Klf4, Sox2 and c-Myc. This pioneer work showed that metformin activated-AMPK prevented the expression of Oct4, putting a stop in the “immortalization” process of iPSCs by blocking bioenergetic glycolytic metabotype. Moreover, the same team has reported that metformin is able to modulate Embrionic-Mesenchymal Phenotype (EMP) via restriction of its genetic controllers, twist-zeb-snail and upregulation of microRNA let-7 [111], forcing a mesenchymal phenotype which is a “mature”-like cell, in an AMPK-independent manner. Such results have been also reported in a pancreatic cell lines by Li [112], who also found the increased expression of microRNA-26a and microRNA-192. Likewise, metformin blocks TGF-β activation of the EMP genetic program and with it, loss of expression of E-cadherin in breast cancer cells [113], limiting their migration and metastasis [114]. And finally, this biguanide also intervenes in inflammation modulation, another link between diabetes and cancer [18,19]. Hirsch [115] published their results on the effect of metformin on cancer stem cells, showing that the biguanide inhibits nuclear transcription factor NF-κB via phosphorylation and activation of IKB. The blockage of NF-κB lowers the expression of IL-6 and LIN28 (natural repressor of let-7), and Src-dependent cancerous transformation [115,116].

On the other hand, metformin appears to activate p53, the guardian of the genome [109]. This tumor suppressor-related protein is responsible for one of the most complex intracellular signaling cascades, which not only controls DNA damage response and tumor suppression [117], but is also a glucose metabolism controller [118,119]. Activation of p53 is associated with downregulation of Glut-1/3/4, phosphoglycerate mutase, while upregulation of TIGAR (TP53-inducible glycolysis and apoptosis regulator) and glutamine synthase [115,116]. Using mouse embryo fibroblasts, He [109], determined that AMPK could phosphorylate and inactivate MDMX, a natural repressor of p53, allowing its stabilization and activation. Such property has been related to induction of senescence in endothelial cells [120] and non-small cell lung cancer [121]; and with increased radiation response in melanoma [122] and colon cancer cells [123].

A novel molecular axis has been proposed for metformin, the p53/REDD1 pathway. The Regulated in Development and DNA Damage responses-1 (REDD1) is a hypoxia-related target gene which is associated with stressed cell survival [124], whose main effect is the inhibition of mTOR1 via TSC2 activation [125]. Ben Sahra [126] published new insights on this new pathway, proving that metformin inhibited mTOR1 by inducing p53 and REDD1 in prostate cancer cells, results which have confirmed by Mohammed [104]. As it seems, metformin is able to modulate cancer cell nutrition and fuel-regulated survival [90,104,108], which could impaired proliferation and migration phenotypes.

THE ROLE OF VITAMIN B12 IN METFORMIN´S ANTITUMORAL ACTIVITY

Perhaps the most unusual anti-tumoral effect that has been linked to metformin is its role in the metabolism of monocarbons derived from serine, glycine and threonine. DNA synthesis requires the methyl donation from S-adenosylmethionine (SAM), a unique molecular species generate only on the one-carbon metabolism [127] (Figure 7). This process requires a numbers of enzymes and it has the property of overlapping the folate, methionine and glutathione cycles [128], requiring the presence of 4 vitamins: folate, vitamine B6, B2 and B12 [129]. The absence of vitamin B12 (cyanocobalamin) compromises the progression of the cycles in a phenomenon called “folate bottleneck”, which is characterized by the blunted production of SAM and stalled DNA synthesis and methylation [128,129].

As previously stated, serine, glycine and glutamine metabolisms are associated with tumor cells proliferation and survival advantage [68-72,74,75]. In fact, inhibition of the one-carbon metabolism using methotrexate, a widely used antifolate chemotherapeutic [130], slows cancer cell proliferation due to halting of the one-carbon metabolism and ATP shortage [131]. Interruption of folate metabolism and stalling of purine synthesis has been proven to induce AMPK and senescence in in vitro models of prostate [132] and breast [133] cancer. Moreover, Vazquez [134] have uncovered novel ATP generation pathways that involve one-carbon metabolism, glycine cleavage and serine biosynthesis, highlighting the importance of this biochemical pathway in cancer cell proliferation, survival, and even, act as a methotrexate-sensitive marker [135].

Cyanocobalamin deficiency has been demonstrated in T2DM patients being treated with merformin [136-138], sometimes associated with megalobastic anemia in this group [139,140]. Bauman [141] suggested that these low levels of B12 were related to decrease absorption due to a calcium-dependent ileal membrane antagonism, and such effect seems amplified with the concomitant use of proton pumps inhibitors [142]. Taking into account that Vitamin B12 deficiency induces a folate bottleneck, a cancer cell which is living in a pseudo-folate deficiency state is even more susceptible to the cytostatic effects of metformin [143]. Therefore, it has been suggested that metformin and methotrexate should be considered as partners in treatment of diseases such as cancer [143] and psoriasis [144].

PERSPECTIVE

Metronomic chemotherapy is a modality of cancer treatment characterized by continuous low-dose administration of conventional chemotherapy drugs, without extended drug-free periods of time [145]. Such therapy has been known to offer better angiogenic control and tumor regression than conventional high-dose schemes [146-148]. This kind of therapy has been known to offer higher survival rates, complete positive responses and fewer cases of grade 3/4 adverse effects [149]. Metformin has been proposed as a novel metronomic chemotherapy drug due to its AMPK-dependent properties [149], especially after the results from Obajimi [132] which evaluated the anti-proliferative effects of AMPK agonists. Nevertheless, as explained in this review, the impact of AMPK-independent effects may be enough to strongly consider the use of metformin as a metronomic agent, especially due to its effects in energy production. However, prospective studies are needed to properly ascertain adequate dosage and time-windows of opportunity; as well as in vitro models in order to fully describe the intracellular effects of metformin.

ACKNOWLEDGMENT

This work was supported by Research Grant no. CC-0437-10- 21-09-10 from Consejo de Desarrollo Científico, Humanístico y Tecnológico (CONDES), University of Zulia, and Research Grant No. FZ-0058-2007 from Fundacite-Zulia.

REFERENCES

1. Correia S, Carvalho C, Santos MS, Seiça R, Oliveira CR, Moreira PI. Mechanisms of action of metformin in type 2 diabetes and associated complications: an overview. Mini Rev Med Chem. 2008; 8: 1343-1354.

2. Al-Nozha O, Habib F, Mojaddidi M, El-Bab MF. Body weight reduction and metformin: Roles in polycystic ovary syndrome. Pathophysiology. 2013; 20: 131-137.

3. Miller RA, Chu Q, Xie J, Foretz M, Viollet B, Birnbaum MJ. Biguanides suppress hepatic glucagon signalling by decreasing production of cyclic AMP. Nature. 2013; 494: 256-260.

4. Joya-Galeana J, Fernandez M, Cervera A, Reyna S, Ghosh S, Triplitt C, et al. Effects of insulin and oral anti-diabetic agents on glucose metabolism, vascular dysfunction and skeletal muscle inflammation in type 2 diabetic subjects. Diabetes Metab Rev Res. 2011; 27: 373- 382.

5. Rojas J, Arraiz N, Aguirre M, Velasco M, Bermúdez V. AMPK as Target for Intervention in Childhood and Adolescent Obesity. J Obes. 2011; 2011: 252817.

6. Mackenzie RW, Elliott BT. Akt/PKB activation and insulin signaling: a novel insulin signaling pathway in the treatment of type 2 diabetes. Diabetes Metab Syndr Obes. 2014; 7: 55-64.

7. Richardson MC, Ingamells S, Simonis CD, Cameron IT, Sreekumar R, Vijendren A, et al. Stimulation of lactate production in human granulosa cells by metformin and potential involvement of adenosine 5’ monophosphate-activated protein kinase. J Clin Endocrinol Metab. 2009; 94: 670-677.

8. Song S, Andrikopoulos S, Filippis C, Thorburn AW, Khan D, Proietto J. Mechanism of fat-induced hepatic gluconeogenesis: effect of metformin. Am J Physiol Endocrinol Metab. 2001; 281: E275-282.

9. Sinnett-Smith J, Kisfalvi K, Kui R, Rozengurt E. Metformin inhibition of mTORC1 activation, DNA synthesis and proliferation in pancreatic cancer cells: dependence on glucose concentration and role of AMPK. Biochem Biophys Res Commun. 2013; 430: 352-357.

10. González-Barroso MM, Anedda A, Gallardo-Vara E, Redondo-Horcajo M, Rodríguez-Sánchez L, Rial E. Fatty acids revert the inhibition of respiration caused by the antidiabetic drug metformin to facilitate their mitochondrial ß-oxidation. Biochim Biophys Acta. 2012; 1817: 1768-1775.

11. Kim HJ, Park EY, Oh MJ, Park SS, Shin KH, Choi SH, et al. Central administration of metformin into the third ventricle of C57BL/6 mice decreases meal size and number and activates hypothalamic S6 kinase. Am J Physiol Regul Integr Comp Physiol. 2013; 305: R499-505.

12. Ravina A, Minuchin O. Bedtime administration of metformin may reduce insulin requirements. Harefuah. 1990; 119: 200-203.

13. Hermann LS, Scherstén B, Melander A. Antihyperglycaemic efficacy, response prediction and dose-response relations of treatment with metformin and sulphonylurea, alone and in primary combination. Diabet Med. 1994; 11: 953-960.

14. Sajan MP, Ivey RA 3rd, Farese RV. Metformin action in human hepatocytes: coactivation of atypical protein kinase C alters 5’-AMP-activated protein kinase effects on lipogenic and gluconeogenic enzyme expression. Diabetologia. 2013; 56: 2507-2516.

15. He L, Sabet A, Djedjos S, Miller R, Sun X, Hussain MA, et al. Metformin and insulin suppress hepatic gluconeogenesis through phosphorylation of CREB binding protein. Cell. 2009; 137: 635-646.

16. Barnett AH. Avoiding hypoglycaemia while achieving good glycaemic control in type 2 diabetes through optimal use of oral agent therapy. Curr Med Res Opin. 2010; 26: 1333-1342.

17. Johnson JA, Pollak M. Insulin, glucose and the increased risk of cancer in patients with type 2 diabetes. Diabetologia. 2010; 53: 2086-2088.

18. Garg SK, Maurer H, Reed K, Selagamsetty R. Diabetes and cancer: two diseases with obesity as a common risk factor. Diabetes Obes Metab. 2013;.

19. Larsson SC, Orsini N, Brismar K, Wolk A. Diabetes mellitus and risk of bladder cancer: a meta-analysis. Diabetologia. 2006; 49: 2819-2823.

20. Friberg E, Orsini N, Mantzoros CS, Wolk A. Diabetes mellitus and risk of endometrial cancer: a meta-analysis. Diabetologia. 2007; 50: 1365- 1374.

21. Larsson SC, Mantzoros CS, Wolk A. Diabetes mellitus and risk of breast cancer: a meta-analysis. Int J Cancer. 2007; 121: 856-862.

22. El-Serag HB, Hampel H, Javadi F. The association between diabetes and hepatocellular carcinoma: a systematic review of epidemiologic evidence. Clin Gastroenterol Hepatol. 2006; 4: 369-380.

23. Chao C, Page JH. Type 2 diabetes mellitus and risk of non-Hodgkin lymphoma: a systematic review and meta-analysis. Am J Epidemiol. 2008; 168: 471-480.

24. Larsson SC, Orsini N, Wolk A. Diabetes mellitus and risk of colorectal cancer: a meta-analysis. J Natl Cancer Inst. 2005; 97: 1679-1687.

25. Lin HC, Kachingwe BH, Lin HL, Cheng HW, Uang YS, Wang LH. Effects of metformin dose on cancer risk reduction in patients with type 2 diabetes mellitus: a 6-year follow-up study. Pharmacotherapy. 2014; 34: 36-45.

26. Bosco JL, Antonsen S, Sørensen HT, Pedersen L, Lash TL. Metformin and incident breast cancer among diabetic women: a population-based case-control study in Denmark. Cancer Epidemiol Biomarkers Prev. 2011; 20: 101-111.

27. Little MW, Pugh TF, Carey FJ, Ndokera R, Ing H, Robinson RJ, et al. The potential protective effect of metformin against pancreatic cancer: preliminary results from a case-control study in two UK centres. Gut 2011; 60: A78-A79.

28. Soranna D, Scotti L, Zambon A, Bosetti C, Grassi G, Catapano A, et al. Cancer risk associated with use of metformin and sulfonylurea in type 2 diabetes: a meta-analysis. Oncologist. 2012; 17: 813-822.

29. Noto H, Goto A, Tsujimoto T, Noda M. Cancer risk in diabetic patients treated with metformin: a systematic review and meta-analysis. PLoS ONE. 2012; 7: e33411.

30. Suissa S, Azoulay L. Metformin and the risk of cancer: time-related biases in observational studies. Diabetes Care. 2012; 35: 2665-2673.

31. Gottlieb B, Auld WH. Metformin in treatment of diabetes mellitus. Br Med J. 1962; 1: 680-682.

32. Kruger FA, Skillman TG, Hamwi GJ, Grubbs RC, Danforth N. The mechanism of action of hypoglycemic guanidine derivatives. Diabetes. 1960; 9: 170-173.

33. Cusi K, Consoli A, DeFronzo RA. Metabolic effects of metformin on glucose and lactate metabolism in noninsulin-dependent diabetes mellitus. J Clin Endocrinol Metab. 1996; 81: 4059-4067

34. Scheen AJ. Clinical pharmacokinetics of metformin. Clin Pharmacokinet. 1996; 30: 359-371.

35. Graham GG, Punt J, Arora M, Day RO, Doogue MP, Duong JK, et al. Clinical pharmacokinetics of metformin. Clin Pharmacokinet. 2011; 50: 81-98.

36. Marin JJ. Plasma membrane transporters in modern liver pharmacology. Scientifica (Cairo). 2012; 2012: 428139.

37. Gong L, Goswami S, Giacomini KM, Altman RB, Klein TE. Metformin pathways: pharmacokinetics and pharmacodynamics. Pharmacogenet Genomics. 2012; 22: 820-827.

38. Jonker JW, Schinkel AH. Pharmacological and physiological functions of the polyspecific organic cation transporters: OCT, 2, and 3 (SLC22A1-3). J Pharmacol Exp Ther. 2004; 308: 2-9.

39. Damme K, Nies AT, Schaeffeler E, Schwab M. Mammalian MATE (SLC47A) transport proteins: impact on efflux of endogenous substrates and xenobiotics. Drug Metab Rev. 2011; 43: 499-523.

40. Takane H, Shikata E, Otsubo K, Higuchi S, Ieiri I. Polymorphism in human organic cation transporters and metformin action. Pharmacogenomics. 2008; 9: 415-422.

41. Sato T, Masuda S, Yonezawa A, Tanihara Y, Katsura T, Inui K. Transcellular transport of organic cations in double-transfected MDCK cells expressing human organic cation transporters hOCT1/hMATE1 and hOCT2/hMATE1. Biochem Pharmacol. 2008; 76: 894-903.

42. Tsuda M, Terada T, Ueba M, Sato T, Masuda S, Katsura T, Inui K. Involvement of human multidrug and toxin extrusion 1 in the drug interaction between cimetidine and metformin in renal epithelial cells. J Pharmacol Exp Ther. 2009; 329: 185-191.

43. Tzvetkov MV, Vormfelde SV, Balen D, Meineke I, Schmidt T, Sehrt D, et al. The effects of genetic polymorphisms in the organic cation transporters OCT, OCT2, and OCT3 on the renal clearance of metformin. Clin Pharmacol Ther. 2009; 86: 299-306.

44. Christensen MM, Brasch-Andersen C, Green H, Nielsen F, Damkier P, Beck-Nielsen H, et al. The pharmacogenetics of metformin and its impact on plasma metformin steady-state levels and glycosylated hemoglobin A1c. Pharmacogenet Genomics. 2011; 21: 837-850.

45. Shu Y, Sheardown SA, Brown C, Owen RP, Zhang S, Castro RA, et al. Effect of genetic variation in the organic cation transporter 1 (OCT1) on metformin action. J Clin Invest. 2007; 117: 1422-1431.

46. Owen MR, Doran E, Halestrap AP. Evidence that metformin exerts its anti-diabetic effects through inhibition of complex 1 of the mitochondrial respiratory chain. Biochem J. 2000; 348 Pt 3: 607-614.

47. Vazquez-Martin A, López-Bonetc E, Cufí S, Oliveras-Ferraros C, Del Barco S, Martin-Castillo B, et al. Repositioning chloroquine and metformin to eliminate cancer stem cell traits in pre-malignant lesions. Drug Resist Updat. 2011; 14: 212-223.

48. Logie L, Harthill J, Patel K, Bacon S, Hamilton DL, Macrae K, et al. Cellular responses to the metal-binding properties of metformin. Diabetes. 2012; 61: 1423-1433.

49. Zhu M, Lu L, Yang P, Jin X. Bis (1,1-dimethylbiguanido)copper(II) octahydrate. Acta Cryst. 2002; E58: m217-m219.

50. Fanshawe WJ, Bauer VJ, Ullman EF, Safir SR. Synthesis of unsymmetrically substituted malonamides. J Org Chem. 1964; 29: 308–311.

51. Rozas I, Sánchez-Sanz G, Alkorta I, Elguero J. Solvent effects on guanidinium-anion interactions and the problem of guanidinium Y-aromaticity. J Phys Organic Chem. 2013; 26: 378-385.

52. Cozzi PG. Metal-Salen Schiff base complexes in catalysis: practical aspects. Chem Soc Rev. 2004; 33: 410-421.

53. Repiscak P, Erhardt S, Rena G, Paterson MJ. Biomolecular mode of action of metformin in relation to its copper binding properties. Biochemistry. 2014; 53: 787-795.

54. Cantor JR, Sabatini DM. Cancer cell metabolism: one hallmark, many faces. Cancer Discov. 2012; 2: 881-898.

55. Horsman MR. Measurement of tumor oxygenation. Int J Radiat Oncol Biol Phys. 1998; 42: 701-704.

56. Airley R, Loncaster J, Davidson S, Bromley M, Roberts S, Patterson A, et al. Glucose transporter glut-1 expression correlates with tumor hypoxia and predicts metastasis-free survival in advanced carcinoma of the cervix. Clin Cancer Res. 2001; 7: 928-934.

57. Warburg O. On the origin of cancer cells. Science. 1956; 123: 309-314.

58. Yamamoto T, Seino Y, Fukumoto H, Koh G, Yano H, Inagaki N, Yamada Y. Over-expression of facilitative glucose transporter genes in human cancer. Biochem Biophys Res Commun. 1990; 170: 223-230.

59. Koukourakis MI, Giatromanolaki A, Simopoulos C, Polychronidis A, Sivridis E. Lactate dehydrogenase 5 (LDH5) relates to up-regulated hypoxia inducible factor pathway and metastasis in colorectal cancer. Clin Exp Metastasis. 2005; 22: 25-30.

60. Atsumi T, Chesney J, Metz C, Leng L, Donnelly S, Makita Z, et al. High expression of inducible 6-phosphofructo-2-kinase/fructose-2,6- bisphosphatase (iPFK-2; PFKFB3) in human cancers. Cancer Res. 2002; 62: 5881-5887.

61. Seyfried TN, Shelton LM. Cancer as a metabolic disease. Nutr Metab (Lond). 2010; 7: 7.

62. Szablewski L. Expression of glucose transporters in cancers. Biochim Biophys Acta. 2013; 1835: 164-169.

63. Dawson DM, Goodfriend TL, Kaplan NO. Lactic Dehydrogenases: Functions of the two types rates of Synthesis of the two major forms can be correlated with metabolic differentiation. Science. 1964; 143: 929-933.

64. Cortés-Cros M, Hemmerlin C, Ferretti S, Zhang J, Gounarides JS, Yin H, Muller A. M2 isoform of pyruvate kinase is dispensable for tumor maintenance and growth. Proc Natl Acad Sci U S A. 2013; 110: 489- 494.

65. Christofk HR, Vander Heiden MG, Harris MH, Ramanathan A, Gerszten RE, Wei R, et al. The M2 splice isoform of pyruvate kinase is important for cancer metabolism and tumour growth. Nature. 2008; 452: 230- 233.

66. Vander Heiden MG, Locasale JW, Swanson KD, Sharfi H, Heffron GJ, Amador-Noguez D, et al. Evidence for an alternative glycolytic pathway in rapidly proliferating cells. Science. 2010; 329: 1492-1499.

67. Vander Heiden MG, Lunt SY, Dayton TL, Fiske BP, Israelsen WJ, Mattaini KR, et al. Metabolic pathway alterations that support cell proliferation. Cold Spring Harb Symp Quant Biol. 2011; 76: 325-334.

68. Icard P, Poulain L, Lincet H. Understanding the central role of citrate in the metabolism of cancer cells. Biochim Biophys Acta. 2012; 1825: 111-116.

69. Dang CV. Glutaminolysis: supplying carbon or nitrogen or both for cancer cells? Cell Cycle. 2010; 9: 3884-3886.

70. Locasale JW. Serine, glycine and one-carbon units: cancer metabolism in full circle. Nat Rev Cancer. 2013; 13: 572-583.

71. Locasale JW, Cantley LC. Genetic selection for enhanced serine metabolism in cancer development. Cell Cycle. 2011; 10: 3812-3813.

72. Possemato R, Marks KM, Shaul YD, Pacold ME, Kim D, Birsoy K, et al. Functional genomics reveal that the serine synthesis pathway is essential in breast cancer. Nature. 2011; 476: 346-350.

73. Chaneton B, Hillmann P, Zheng L, Martin AC, Maddocks OD, Chokkathukalam A, et al. Serine is a natural ligand and allosteric activator of pyruvate kinase M2. Nature. 2012; 491: 458-462.

74. Jain M, Nilsson R, Sharma S, Madhusudhan N, Kitami T, Souza AL, et al. Metabolite profiling identifies a key role for glycine in rapid cancer cell proliferation. Science. 2012; 336: 1040-1044.

75. Dell’ Antone P. Energy metabolism in cancer cells: how to explain the Warburg and Crabtree effects? Med Hypotheses. 2012; 79: 388-392.

76. Crabtree HG. Observations on the carbohydrate metabolism of tumours. Biochem J. 1929; 23: 536-545.

77. Wojtczak L. The Crabtree effect: a new look at the old problem. Acta Biochim Pol. 1996; 43: 361-368.

78. Diaz-Ruiz R, Rigoulet M, Devin A. The Warburg and Crabtree effects: On the origin of cancer cell energy metabolism and of yeast glucose repression. Biochim Biophys Acta. 2011; 1807: 568-576.

79. Suchorolski MT, Paulson TG, Sanchez CA, Hockenbery D, Reid BJ. Warburg and Crabtree effects in premalignant Barrett’s esophagus cell lines with active mitochondria. PLoS One. 2013; 8: e56884.

80. Taylor MD, Smith PW, Brix WK, Wick MR, Theodosakis N, Swenson BR, et al. Correlations between selected tumor markers and fluorodeoxyglucose maximal standardized uptake values in esophageal cancer. Eur J Cardiothorac Surg. 2009; 35: 699–705.

81. Suganuma K, Miwa H, Imai N, Shikami M, Gotou M, Goto M, et al. Energy metabolism of leukemia cells: glycolysis versus oxidative phosphorylation. Leuk Lymphoma. 2010; 51: 2112-2119.

82. Acin-Perez R, Enriquez JA. The function of the respiratory supercomplexes: the plasticity model. Biochim Biophys Acta. 2014; 1837: 444-450.

83. Chaban Y, Boekema EJ, Dudkina NV. Structures of mitochondrial oxidative phosphorylation supercomplexes and mechanisms for their stabilisation. Biochim Biophys Acta. 2014; 1837: 418-426.

84. Efremov RG, Baradaran R, Sazanov LA. The architecture of respiratory complex I. Nature. 2010; 465: 441-445.

85. Sun F, Huo X, Zhai Y, Wang A, Xu J, Su D, et al. Crystal structure of mitochondrial respiratory membrane protein complex II. Cell. 2005; 121: 1043-1057.

86. Meunier B, Fisher N, Ransac S, Mazat JP, Brasseur G. Respiratory complex III dysfunction in humans and the use of yeast as a model organism to study mitochondrial myopathy and associated diseases. Biochim Biophys Acta. 2013; 1827: 1346-1361.

87. Wittig I, Schägger H. Structural organization of mitochondrial ATP synthase. Biochim Biophys Acta. 2008; 1777: 592-598.

88. Maranzana E, Barbero G, Falasca AI, Lenaz G, Genova ML. Mitochondrial respiratory supercomplex association limits production of reactive oxygen species from complex I. Antioxid Redox Signal. 2013; 19: 1469-1480.

89. Lonardo E, Cioffi M, Sancho P, Sanchez-Ripoll Y, Trabulo SM, Dorado J, et al. Metformin targets the metabolic achilles heel of human pancreatic cancer stem cells. PLoS One. 2013; 8: e76518.

90. Cantoria MJ, Boros LG, Meuillet EJ. Contextual inhibition of fatty acid synthesis by metformin involves glucose-derived acetyl-CoA and cholesterol in pancreatic tumor cells. Metabolomics. 2014; 10: 91-104.

91. Habibollahi P, van den Berg NS, Kuruppu D, Loda M, Mahmood U. Metformin--an adjunct antineoplastic therapy--divergently modulates tumor metabolism and proliferation, interfering with early response prediction by 18F-FDG PET imaging. J Nucl Med. 2013; 54: 252-258.

92. Ota S, Horigome K, Ishii T, Nakai M, Hayashi K, Kawamura T, et al. Metformin suppresses glucose-6-phosphatase expression by a complex I inhibition and AMPK activation-independent mechanism. Biochem Biophys Res Commun. 2009; 388: 311-316.

93. El-Mir MY, Nogueira V, Fontaine E, Avéret N, Rigoulet M, Leverve X. Dimethylbiguanide inhibits cell respiration via an indirect effect targeted on the respiratory chain complex I. J Biol Chem. 2000; 275: 223-228.

94. Modica-Napolitano JS, Aprille JR. Delocalized lipophilic cations selectively target the mitochondria of carcinoma cells. Adv Drug Deliv Rev. 2001; 49: 63-70.

95. Kurtoglu M, Lampidis TJ. From delocalized lipophilic cations to hypoxia: blocking tumor cell mitochondrial function leads to therapeutic gain with glycolytic inhibitors. Mol Nutr Food Res. 2009; 53: 68-75.

96. Filomeni G, Cerchiaro G, Da Costa Ferreira AM, De Martino A, Pedersen JZ, Rotilio G, et al. Pro-apoptotic activity of novel Isatin-Schiff base copper(II) complexes depends on oxidative stress induction and organelle-selective damage. J Biol Chem. 2007; 282: 12010-12021.

97. Cobine PA, Pierrel F, Winge DR. Copper trafficking to the mitochondrion and assembly of copper metalloenzymes. Biochim Biophys Acta. 2006; 1763: 759-772.

98. Leary SC, Winge DR, Cobine PA. “Pulling the plug” on cellular copper: the role of mitochondria in copper export. Biochim Biophys Acta. 2009; 1793: 146-153.

99. Ishida S, Andreux P, Poitry-Yamate C, Auwerx J, Hanahan D. Bioavailable copper modulates oxidative phosphorylation and growth of tumors. Proc Natl Acad Sci U S A. 2013; 110: 19507-19512.

100. Lombardo MF, Ciriolo MR, Rotilio G, Rossi L. Prolonged copper depletion induces expression of antioxidants and triggers apoptosis in SH-SY5Y neuroblastoma cells. Cell Mol Life Sci. 2003; 60: 1733- 1743.

101. Kim YC, Hummer G. Proton-pumping mechanism of cytochrome c oxidase: a kinetic master-equation approach. Biochim Biophys Acta. 2012; 1817: 526-536.

102. Whitaker-Menezes D, Martinez-Outschoorn UE, Flomenberg N, Birbe RC, Witkiewicz AK, Howell A, et al. Hyperactivation of oxidative mitochondrial metabolism in epithelial cancer cells in situ: visualizing the therapeutic effects of metformin in tumor tissue. Cell Cycle 2011; 10: 4047-4064.

103. Larsson O, Morita M, Topisirovic I, Alain T, Blouin MJ, Pollak M, et al. Distinct perturbation of the translatome by the antidiabetic drug metformin. Proc Natl Acad Sci U S A. 2012; 109: 8977-8982.

104. Mohammed A, Janakiram NB, Brewer M, Ritchie RL, Marya A, Lightfoot S, et al. Antidiabetic Drug Metformin Prevents Progression of Pancreatic Cancer by Targeting in Part Cancer Stem Cells and mTOR Signaling. Transl Oncol. 2013; 6: 649-659.

105. Vitale-Cross L, Molinolo AA, Martin D, Younis RH, Maruyama T, Patel V, et al. Metformin prevents the development of oral squamous cell carcinomas from carcinogen-induced premalignant lesions. Cancer Prev Res (Phila). 2012; 5: 562-573.

106. Wu CL, Qiang L, Han W, Ming M, Viollet B, He YY. Role of AMPK in UVB-induced DNA damage repair and growth control. Oncogene. 2013; 32: 2682-2689.

107. Zhao L, Wen ZH, Jia CH, Li M, Luo SQ, Bai XC. Metformin induces G1 cell cycle arrest and inhibits cell proliferation in nasopharyngeal carcinoma cells. Anat Rec (Hoboken). 2011; 294: 1337-1343. 

108. Dibble CC, Manning BD. Signal integration by mTORC1 coordinates nutrient input with biosynthetic output. Nat Cell Biol. 2013; 15: 555- 564.

109. He G, Zhang YW, Lee JH, Zeng SX, Wang YV, Luo Z, et al. AMP-activated protein kinase induces p53 by phosphorylating MDMX and inhibiting its activity. Mol Cell Biol. 2014; 34: 148-157.

110. Vazquez-Martín A, Vellon L, Quirós PM, Ruiz de GalarretaE, Oliveras-Ferraros C, Martin AG, et al. Activation of AMP-activated protein kinase (AMPK) provides a metabolic barrier to reprogramming somatic cells into stem cells. Cell Cycle. 2012; 11: 974-989.

111. Vazquez-Martin A, López-Bonetc E, Cufí S, Oliveras-Ferraros C, Del Barco S, Martin-Castillo B, et al. Repositioning chloroquine and metformin to eliminate cancer stem cell traits in pre-malignant lesions. Drug Resist Updat. 2011; 14: 212-223.

112. Li W, Yuan Y, Huang L, Qiao M, Zhang Y. Metformin alters the expression profiles of microRNAs in human pancreatic cancer cells. Diabetes Res Clin Pract. 2012; 96: 187-195.

113. Cufí S, Vazquez-Martin A, Oliveras-Ferraros C, Martin-Castillo B, Joven J, Menendez JA. Metformin against TGFβ- induced epithelial-to-mesenchymal transition (EMT): from cancer stem cells to aging-associated fibrosis. Cell Cycle. 2010; 9: 4461-4468.

114. Zheng H, Kang Y. Multilayer control of the EMT master regulators. Oncogene 2013.

115. Hirsch HA, Iliopoulos D, Struhl K. Metformin inhibits the inflammatory response associated with cellular transformation and cancer stem cell growth. Proc Natl Acad Sci U S A. 2013; 110: 972- 977.

116. Anastasiou D. Metformin: a case of divide and conquer. Breast Cancer Res. 2013; 15: 306.

117. Efeyan A, Serrano M. p53: guardian of the genome and policeman of the oncogenes. Cell Cycle. 2007; 6: 1006-1010.

118. Soga T. Cancer metabolism: key players in metabolic reprogramming. Cancer Sci. 2013; 104: 275-281.

119. Vousden KH, Ryan KM. p53 and metabolism. Nat Rev Cancer. 2009; 9: 691-700.

120. Arunachalam G, Samuel SM, Marei I, Ding H, Triggle CR. Metformin modulates hyperglycaemia-induced endothelial senescence and apoptosis through SIRT1. Br J Pharmacol. 2014; 171: 523-535.

121. Storozhuk Y, Hopmans SN, Sanli T, Barron C, Tsiani E, Cutz JC, et al. Metformin inhibits growth and enhances radiation response of non-small cell lung cancer (NSCLC) through ATM and AMPK. Br J Cancer. 2013; 108: 2021-2032.

122. Cerezo M, Tichet M, Abbe P, Ohanna M, Lehraiki A, Rouaud F, et al. Metformin blocks melanoma invasion and metastasis development in AMPK/p53-dependent manner. Mol Cancer Ther. 2013; 12: 1605- 1615.

123. Muaddi H, Chowdhury S, Vellanki R, Zamiara P, Koritzinsky M. Contributions of AMPK and p53 dependent signaling to radiation response in the presence of metformin. Radiother Oncol. 2013; 108: 446-450.

124. Ellisen LW, Ramsayer KD, Johannessen CM, Yang A, Beppu H, Minda K, et al. REDD, a developmentally regulated transcriptional target of p63 and p53, links p63 to regulation of reactive oxygen species. Mol Cell. 2002; 10: 995-1005.

125. Brugarolas J, Lei K, Hurley RL, Manning BD, Reiling JH, Hafen E, et al. Regulation of mTOR function in response to hypoxia by REDD1 and the TSC1/TSC2 tumor suppressor complex. Genes Dev. 2004; 18: 2893-2904.

126. Ben Sahra I, Regazzetti C, Robert G, Laurent K, Le Marchand-Brustel Y, Auberger P, et al. Metformin, independent of AMPK, induces mTOR inhibition and cell-cycle arrest through REDD1. Cancer Res. 2011; 71: 4366-4372.

127. Selhub J. Homocysteine metabolism. Annu Rev Nutr. 1999; 19: 217- 246.

128. Locasale JW. Serine, glycine and one-carbon units: cancer metabolism in full circle. Nat Rev Cancer. 2013; 13: 572-583.

129. Anderson OS, Sant KE, Dolinoy DC. Nutrition and epigenetics: an interplay of dietary methyl donors, one-carbon metabolism and DNA methylation. J Nutr Biochem. 2012; 23: 853-859.

130. Jolivet J, Cowan KH, Curt GA, Clendeninn NJ, Chabner BA. The pharmacology and clinical use of methotrexate. N Engl J Med. 1983; 309: 1094-1104.

131. Tedeschi PM, Markert EK, Gounder M, Lin H, Dvorzhinski D, Dolfi SC, et al. Contribution of serine, folate and glycine metabolism to the ATP, NADPH and purine requirements of cancer cells. Cell Death Dis. 2013; 4: e877.

132. Obajimi O, Keen JC, Melera PW. Inhibition of de novo purine synthesis in human prostate cells results in ATP depletion, AMPK activation and induces senescence. Prostate. 2009; 69: 1206-1221.

133. Allegra CJ, Hoang K, Yeh GC, Drake JC, Baram J. Evidence for direct inhibition of de novo purine synthesis in human MCF-7 breast cells as a principal mode of metabolic inhibition by methotrexate. J Biol Chem. 1987; 262: 13520-13526.

134. Vazquez A, Markert EK, Oltvai ZN. Serine biosynthesis with one carbon catabolism and the glycine cleavage system represents a novel pathway for ATP generation. PLoS One. 2011; 6: e25881.

135. Vazquez A, Tedeschi PM, Bertino JR. Overexpression of the mitochondrial folate and glycine-serine pathway: a new determinant of methotrexate selectivity in tumors. Cancer Res. 2013; 73: 478- 482.

136. Nervo M, Lubini A, Raimundo FV, Faulhaber GA, Leite C, Fischer LM, et al. Vitamin B12 in metformin-treated diabetic patients: a cross-sectional study in Brazil. Rev Assoc Med Bras. 2011; 57: 46-49.

137. Calvo Romero JM, Ramiro Lozano JM. Vitamin B(12) in type 2 diabetic patients treated with metformin. Endocrinol Nutr. 2012; 59: 487-490.

138. de Groot-Kamphuis DM, van Dijk PR, Groenier KH, Houweling ST, Bilo HJ, Kleefstra N. Vitamin B12 deficiency and the lack of its consequences in type 2 diabetes patients using metformin. Neth J Med. 2013; 71: 386-390.

139. Pierce SA, Chung AH, Black KK. Evaluation of vitamin B12 monitoring in a veteran population on long-term, high-dose metformin therapy. Ann Pharmacother. 2012; 46: 1470-1476.

140. Filioussi K, Bonovas S, Katsaros T. Should we screen diabetic patients using biguanides for megaloblastic anaemia? Aust Fam Physician. 2003; 32: 383-384.

141. Bauman WA, Shaw S, Jayatilleke E, Spungen AM, Herbert V. Increased intake of calcium reverses vitamin B12 malabsorption induced by metformin. Diabetes Care. 2000; 23: 1227-1231.

142. Long AN, Atwell CL, Yoo W, Solomon SS. Vitamin B(12) deficiency associated with concomitant metformin and proton pump inhibitor use. Diabetes Care. 2012; 35: e84.

143. Corominas-Faja B, Quirantes-Piné R, Oliveras-Ferraros C, Vazquez- Martin A, Cufí S, Martin-Castillo B, et al. Metabolomic fingerprint reveals that metformin impairs one-carbon metabolism in a manner similar to the antifolate class of chemotherapy drugs. Aging 2012; 4: 480-498

144. Glossmann H, Reider N. A marriage of two “Methusalem” drugs for the treatment of psoriasis?: Arguments for a pilot trial with metformin as add-on for methotrexate. Dermatoendocrinol. 2013; 5: 252-263.

145. Hanahan D, Bergers G, Bergsland E. Less is more, regularly: metronomic dosing of cytotoxic drugs can target tumor angiogenesis in mice. J Clin Invest. 2000; 105: 1045-1047.

146. Kerbel RS, Kamen BA. The anti-angiogenic basis of metronomic chemotherapy. Nat Rev Cancer. 2004; 4: 423-436.

147. Browder T, Butterfield CE, Kräling BM, Shi B, Marshall B, O’Reilly MS, et al. Antiangiogenic scheduling of chemotherapy improves efficacy against experimental drug-resistant cancer. Cancer Res. 2000; 60: 1878-1886.

148. Klement G, Baruchel S, Rak J, Man S, Clark K, Hicklin DJ, et al. Continuous low-dose therapy with vinblastine and VEGF receptor-2 antibody induces sustained tumor regression without overt toxicity. J Clin Invest. 2000; 105: R15-24.

149. Lien K, Georgsdottir S, Sivanathan L, Chan K, Emmenegger U. Low-dose metronomic chemotherapy: a systematic literature analysis. Eur J Cancer. 2013; 49: 3387-3395.

Rojas J, Chávez-Castillo M, Torres W, Arraiz N, Cabrera M, et al. (2014) Metformin in Cancer: Chemical Pathways for Tumoral Control Independent of Amp-Dependent Kinase. J Endocrinol Diabetes Obes 2(2): 1036.

Received : 05 Apr 2014
Accepted : 26 May 2014
Published : 28 May 2014
Journals
Annals of Otolaryngology and Rhinology
ISSN : 2379-948X
Launched : 2014
JSM Schizophrenia
Launched : 2016
Journal of Nausea
Launched : 2020
JSM Internal Medicine
Launched : 2016
JSM Hepatitis
Launched : 2016
JSM Oro Facial Surgeries
ISSN : 2578-3211
Launched : 2016
Journal of Human Nutrition and Food Science
ISSN : 2333-6706
Launched : 2013
JSM Regenerative Medicine and Bioengineering
ISSN : 2379-0490
Launched : 2013
JSM Spine
ISSN : 2578-3181
Launched : 2016
Archives of Palliative Care
ISSN : 2573-1165
Launched : 2016
JSM Nutritional Disorders
ISSN : 2578-3203
Launched : 2017
Annals of Neurodegenerative Disorders
ISSN : 2476-2032
Launched : 2016
Journal of Fever
ISSN : 2641-7782
Launched : 2017
JSM Bone Marrow Research
ISSN : 2578-3351
Launched : 2016
JSM Mathematics and Statistics
ISSN : 2578-3173
Launched : 2014
Journal of Autoimmunity and Research
ISSN : 2573-1173
Launched : 2014
JSM Arthritis
ISSN : 2475-9155
Launched : 2016
JSM Head and Neck Cancer-Cases and Reviews
ISSN : 2573-1610
Launched : 2016
JSM General Surgery Cases and Images
ISSN : 2573-1564
Launched : 2016
JSM Anatomy and Physiology
ISSN : 2573-1262
Launched : 2016
JSM Dental Surgery
ISSN : 2573-1548
Launched : 2016
Annals of Emergency Surgery
ISSN : 2573-1017
Launched : 2016
Annals of Mens Health and Wellness
ISSN : 2641-7707
Launched : 2017
Journal of Preventive Medicine and Health Care
ISSN : 2576-0084
Launched : 2018
Journal of Chronic Diseases and Management
ISSN : 2573-1300
Launched : 2016
Annals of Vaccines and Immunization
ISSN : 2378-9379
Launched : 2014
JSM Heart Surgery Cases and Images
ISSN : 2578-3157
Launched : 2016
Annals of Reproductive Medicine and Treatment
ISSN : 2573-1092
Launched : 2016
JSM Brain Science
ISSN : 2573-1289
Launched : 2016
JSM Biomarkers
ISSN : 2578-3815
Launched : 2014
JSM Biology
ISSN : 2475-9392
Launched : 2016
Archives of Stem Cell and Research
ISSN : 2578-3580
Launched : 2014
Annals of Clinical and Medical Microbiology
ISSN : 2578-3629
Launched : 2014
JSM Pediatric Surgery
ISSN : 2578-3149
Launched : 2017
Journal of Memory Disorder and Rehabilitation
ISSN : 2578-319X
Launched : 2016
JSM Tropical Medicine and Research
ISSN : 2578-3165
Launched : 2016
JSM Head and Face Medicine
ISSN : 2578-3793
Launched : 2016
JSM Cardiothoracic Surgery
ISSN : 2573-1297
Launched : 2016
JSM Bone and Joint Diseases
ISSN : 2578-3351
Launched : 2017
JSM Bioavailability and Bioequivalence
ISSN : 2641-7812
Launched : 2017
JSM Atherosclerosis
ISSN : 2573-1270
Launched : 2016
Journal of Genitourinary Disorders
ISSN : 2641-7790
Launched : 2017
Journal of Fractures and Sprains
ISSN : 2578-3831
Launched : 2016
Journal of Autism and Epilepsy
ISSN : 2641-7774
Launched : 2016
Annals of Marine Biology and Research
ISSN : 2573-105X
Launched : 2014
JSM Health Education & Primary Health Care
ISSN : 2578-3777
Launched : 2016
JSM Communication Disorders
ISSN : 2578-3807
Launched : 2016
Annals of Musculoskeletal Disorders
ISSN : 2578-3599
Launched : 2016
Annals of Virology and Research
ISSN : 2573-1122
Launched : 2014
JSM Renal Medicine
ISSN : 2573-1637
Launched : 2016
Journal of Muscle Health
ISSN : 2578-3823
Launched : 2016
JSM Genetics and Genomics
ISSN : 2334-1823
Launched : 2013
JSM Anxiety and Depression
ISSN : 2475-9139
Launched : 2016
Clinical Journal of Heart Diseases
ISSN : 2641-7766
Launched : 2016
Annals of Medicinal Chemistry and Research
ISSN : 2378-9336
Launched : 2014
JSM Pain and Management
ISSN : 2578-3378
Launched : 2016
JSM Women's Health
ISSN : 2578-3696
Launched : 2016
Clinical Research in HIV or AIDS
ISSN : 2374-0094
Launched : 2013
Journal of Endocrinology, Diabetes and Obesity
ISSN : 2333-6692
Launched : 2013
Journal of Substance Abuse and Alcoholism
ISSN : 2373-9363
Launched : 2013
JSM Neurosurgery and Spine
ISSN : 2373-9479
Launched : 2013
Journal of Liver and Clinical Research
ISSN : 2379-0830
Launched : 2014
Journal of Drug Design and Research
ISSN : 2379-089X
Launched : 2014
JSM Clinical Oncology and Research
ISSN : 2373-938X
Launched : 2013
JSM Bioinformatics, Genomics and Proteomics
ISSN : 2576-1102
Launched : 2014
JSM Chemistry
ISSN : 2334-1831
Launched : 2013
Journal of Trauma and Care
ISSN : 2573-1246
Launched : 2014
JSM Surgical Oncology and Research
ISSN : 2578-3688
Launched : 2016
Annals of Food Processing and Preservation
ISSN : 2573-1033
Launched : 2016
Journal of Radiology and Radiation Therapy
ISSN : 2333-7095
Launched : 2013
JSM Physical Medicine and Rehabilitation
ISSN : 2578-3572
Launched : 2016
Annals of Clinical Pathology
ISSN : 2373-9282
Launched : 2013
Annals of Cardiovascular Diseases
ISSN : 2641-7731
Launched : 2016
Journal of Behavior
ISSN : 2576-0076
Launched : 2016
Annals of Clinical and Experimental Metabolism
ISSN : 2572-2492
Launched : 2016
Clinical Research in Infectious Diseases
ISSN : 2379-0636
Launched : 2013
JSM Microbiology
ISSN : 2333-6455
Launched : 2013
Journal of Urology and Research
ISSN : 2379-951X
Launched : 2014
Journal of Family Medicine and Community Health
ISSN : 2379-0547
Launched : 2013
Annals of Pregnancy and Care
ISSN : 2578-336X
Launched : 2017
JSM Cell and Developmental Biology
ISSN : 2379-061X
Launched : 2013
Annals of Aquaculture and Research
ISSN : 2379-0881
Launched : 2014
Clinical Research in Pulmonology
ISSN : 2333-6625
Launched : 2013
Journal of Immunology and Clinical Research
ISSN : 2333-6714
Launched : 2013
Annals of Forensic Research and Analysis
ISSN : 2378-9476
Launched : 2014
JSM Biochemistry and Molecular Biology
ISSN : 2333-7109
Launched : 2013
Annals of Breast Cancer Research
ISSN : 2641-7685
Launched : 2016
Annals of Gerontology and Geriatric Research
ISSN : 2378-9409
Launched : 2014
Journal of Sleep Medicine and Disorders
ISSN : 2379-0822
Launched : 2014
JSM Burns and Trauma
ISSN : 2475-9406
Launched : 2016
Chemical Engineering and Process Techniques
ISSN : 2333-6633
Launched : 2013
Annals of Clinical Cytology and Pathology
ISSN : 2475-9430
Launched : 2014
JSM Allergy and Asthma
ISSN : 2573-1254
Launched : 2016
Journal of Neurological Disorders and Stroke
ISSN : 2334-2307
Launched : 2013
Annals of Sports Medicine and Research
ISSN : 2379-0571
Launched : 2014
JSM Sexual Medicine
ISSN : 2578-3718
Launched : 2016
Annals of Vascular Medicine and Research
ISSN : 2378-9344
Launched : 2014
JSM Biotechnology and Biomedical Engineering
ISSN : 2333-7117
Launched : 2013
Journal of Hematology and Transfusion
ISSN : 2333-6684
Launched : 2013
JSM Environmental Science and Ecology
ISSN : 2333-7141
Launched : 2013
Journal of Cardiology and Clinical Research
ISSN : 2333-6676
Launched : 2013
JSM Nanotechnology and Nanomedicine
ISSN : 2334-1815
Launched : 2013
Journal of Ear, Nose and Throat Disorders
ISSN : 2475-9473
Launched : 2016
JSM Ophthalmology
ISSN : 2333-6447
Launched : 2013
Journal of Pharmacology and Clinical Toxicology
ISSN : 2333-7079
Launched : 2013
Annals of Psychiatry and Mental Health
ISSN : 2374-0124
Launched : 2013
Medical Journal of Obstetrics and Gynecology
ISSN : 2333-6439
Launched : 2013
Annals of Pediatrics and Child Health
ISSN : 2373-9312
Launched : 2013
JSM Clinical Pharmaceutics
ISSN : 2379-9498
Launched : 2014
JSM Foot and Ankle
ISSN : 2475-9112
Launched : 2016
JSM Alzheimer's Disease and Related Dementia
ISSN : 2378-9565
Launched : 2014
Journal of Addiction Medicine and Therapy
ISSN : 2333-665X
Launched : 2013
Journal of Veterinary Medicine and Research
ISSN : 2378-931X
Launched : 2013
Annals of Public Health and Research
ISSN : 2378-9328
Launched : 2014
Annals of Orthopedics and Rheumatology
ISSN : 2373-9290
Launched : 2013
Journal of Clinical Nephrology and Research
ISSN : 2379-0652
Launched : 2014
Annals of Community Medicine and Practice
ISSN : 2475-9465
Launched : 2014
Annals of Biometrics and Biostatistics
ISSN : 2374-0116
Launched : 2013
JSM Clinical Case Reports
ISSN : 2373-9819
Launched : 2013
Journal of Cancer Biology and Research
ISSN : 2373-9436
Launched : 2013
Journal of Surgery and Transplantation Science
ISSN : 2379-0911
Launched : 2013
Journal of Dermatology and Clinical Research
ISSN : 2373-9371
Launched : 2013
JSM Gastroenterology and Hepatology
ISSN : 2373-9487
Launched : 2013
Annals of Nursing and Practice
ISSN : 2379-9501
Launched : 2014
JSM Dentistry
ISSN : 2333-7133
Launched : 2013
Author Information X