Gastro

Open journal

ISSN 2377-8369

The Role of Energy Metabolism in Driving Disease Progression in Inflammatory, Hypoxic and Angiogenic Microenvironments

James J. Phelan, C. O’Hanlon, John V. Reynolds and Jacintha O’Sullivan*

Jacintha O’Sullivan, PhD

Associate Professor, Department of Surgery, Institute of Molecular Medicine, Trinity College Dublin, St. James’s Hospital, Dublin 8, Ireland; E-mail: osullij4@tcd.ie

INTRODUCTION

Otto Warburg’s initial observation in 1956 demonstrated that tumours exhibit increased levels of aerobic glycolysis.1 This observation has since resulted in numerous studies investigating the role of mitochondrial energy metabolism in disease progression across many disease entities. As a reflection of its importance in the development of various cancers, the reprogramming of cellular energetics is now beginning to establish itself as one of the new hallmarks of cancer.2,3 In addition to significant quantities of adenosine triphosphate (ATP), metabolically demanding tumours require glucose for lipid and protein synthesis and de novo synthesis of nucleotides for rapid proliferation.4 More importantly, this altered metabolic phenotype allows tumours to maintain higher proliferative rates and resist apoptosis orchestrated by increased oxidative damage.5 Moreover, these metabolic phenotypes persist and are sometimes altered in distinct metabolically demanding microenvironments. Therefore, elucidating how diverse metabolic processes converse with distinct functional inflammatory, hypoxic and angiogenic pathways may infer significant insights into how several heterogeneous malignancies arise and subsequently advance beyond therapeutic intervention.

It has been widely documented that inflammation, hypoxia and angiogenesis all play independent roles in disease prevalence and in its subsequent stepwise progression. Some studies, however, have uncovered close associations between energy metabolism and these extensive processes. Therefore, this review focuses on the primary molecular mechanisms that link energy metabolism with inflammation, hypoxia and angiogenesis. In addition to investigating conventional mediators that link energy metabolism with inflammation and hypoxia, novel mediators that link energy metabolism to inflammation, hypoxia and angiogenesis will also be discussed. This review also explores the mechanisms linking energy metabolism with hypoxia by exploring the roles of glycolysis in rheumatoid arthritis and oxidative phosphorylation (OXPHOS) in circadian rhythms. This review subsequently focuses on the connection between energy metabolism and inflammation in greater detail by examining some of the reciprocal mechanisms linking both processes throughout the gastrointestinal tract. To conclude, this review explores contemporary metabolic-based treatments and multi-targeted therapies that target these key processes. 

MOLECULAR MEDIATORS LINKING ENERGY METABOLISM WITH INFLAMMATION, HYPOXIA AND ANGIOGENESIS

Conventional Mediators Linking Energy Metabolism with Inflammation and Hypoxia 

HIF1α 

Hypoxia Inducible Factor-1α (HIF1α) is an oxygen sensitive transcription factor subunit involved in various cellular processes including hypoxia, angiogenesis, cell survival, inflammation and energy metabolism.6 Interestingly, cells in hypoxic regions tend to be more resistant to the effects of radiotherapy and other conventional chemotherapeutic agents.7 As a result, these more resistant cells have been implicated in disease resistance and recurrence, and can lead to more aggressive phenotypes and contribute to subsequent metastasis.7,8 It is important to note, however, that hypoxia-induced alterations in energy metabolism are physiologically normal, for example, cardiomyocytes upregulate glycolytic ATP production under hypoxic stress.9 As Figure 1 shows, hypoxia affects metabolism by inducing the overexpression of various glycolytic enzymes, lactate dehydrogenase (LDH) and carbonic anhydrase in addition to inhibiting pyruvate dehydrogenase, a key enzyme that converts pyruvate into acetyl-CoA for subsequent oxidative metabolism.10 However, it has been shown that hypoxia, specifically HIF1α, plays a key role on T-cell function by modulating T-cell metabolism.

Upon activation, the metabolic demands of T-cells increase dramatically since activated T-cells are highly anabolic and exhibit marked increases in glycolytic metabolism.11,12 Interestingly, one study hypothesises that one possible mechanism responsible for T-cell anergy is failure to upregulate key metabolic machinery, since blocking energy metabolism mitigates T-cell activation and inhibition of these metabolic pathways during activation leads to anergy in Th1 cells.11 Hypoxia has differential effects on T-cell function, however, as lack of glucose in human CD4+ T lymphocytes results in increased dead cell numbers and increased reactive oxygen species under normoxia but not under hypoxic conditions.13 Hypoxia also stimulates increased levels of interleukin-1β (IL1β), IL10 and IL8 in these cells, but the lack of glucose reduces secretions of these cytokines, implying that CD4+ T cells are highly metabolically adaptable allowing for proper immune function under highly fluctuating bioenergetic microenvironments.13 HIF1α also regulates the balance between regulatory T cell and helper T cell differentiation.14 This differentiation has been shown to be regulated in both regulatory T cells and helper T cells via the glycolytic pathway in a HIF1α-dependent manner.15 In stimulated TH17 cells, glycolysis and various glycolytic enzymes were actively upregulated, although blocking glycolysis inhibited TH17 development while promoting TREG differentiation.15 HIF1α activity is key for mediating glycolytic activity and subsequently contributes to lineage choices between TH17 and TREG cells, whereas lack of HIF1α results in reduced TH17 development but enhances TREG differentiation.15 Some evidence suggests that HIF1α mediates this effect through mammalian target of Rapamycin (mTOR).15,16 These studies support the view that hypoxia mediates T-cell function and drives chronic inflammation through HIF1α by regulating T-cell metabolism.

 AMPK 

AMP-activated protein kinase, or AMPK, is a sensor of cellular energy metabolism and exhibits anti-Warburg effects by promoting fatty acid oxidation, mitochondrial biogenesis and the expression of genes necessary for oxidative metabolism (Figure 1).17,18,19,20 As aerobic glycolysis is a common entity in many cancer types, it is exciting to speculate that drugs that activate AMPK might therefore have therapeutic and clinical utility. Any metabolic imbalance that either inhibits the generation of ATP or accelerates ATP consumption results in increases in the ADP/ATP ratio resulting in AMPK activation due to the accumulation of ADP.17 As a result, activated AMPK acts to switch off ATP-consuming anabolic processes and restores energy imbalances by switching on alternative catabolic pathways that increase cellular ATP.17,21 One of these mechanisms involves down-regulating protein synthesis. For example, AMPK down-regulates protein synthesis of target of rapamycin complex 1 (TORC1), which is known to promote HIF1α translation, thereby reducing HIF1α expression and decreasing the expression of key glycolytic and glucose transporters required for aerobic glycolysis.22 Using a mouse model of Peutz-Jeghers syndrome, deficiency of either AMPK or Liver Kinase B1 (LKB1), the protein kinase responsible for induction of AMPK activation, led to the upregulation of HIF1α, hexokinase 2  and glucose transporter member 1 (GLUT1).23 

Activated immune cells tend to favour aerobic glycolysis whereas quiescent cells preferentially utilise oxidative metabolism.11,12 Therefore, agents that activate AMPK may have anti-inflammatory effects. LPS-induced activation of dendritic cells results in reduced activation of AMPK, whereas knockdown of AMPK leads to the maturation of dentritic cells that exhibit increased glucose uptake.24 Interestingly, AMPK downregulation in macrophages results in increased expression of various pro-inflammatory cytokines whereas expression of AMPK had the reverse effect.25 Therefore, AMPK promotes macrophage polarisation towards an anti-inflammatory M2 phenotype rather than the pro-inflammatory M1 phenotype. In addition, AMPK has been shown to monitor metabolic stress in cytotoxic T lymphocytes and control the differentiation switch from metabolically active cytotoxic T lymphocytes to metabolically quiescent CD8+ T cells, highlighting the important role of AMPK in various metabolic, immune and inflammatory processes.26

p53 

The transcription factor p53 regulates metabolism by lowering aerobic glycolysis and promoting oxidative phosphorylation through a variety of molecular mechanisms.27,28 p53 primarily supports oxidative phosphorylation by functioning as a mitochondrial checkpoint protein, regulating mitochondrial DNA copy number and mediating mitochondrial biogenesis.29,30 p53 promotes mitochondrial health via the p53-inducible protein Mieap that controls mitochondrial quality by repairing or eliminating unhealthy mitochondria.31 In intestinal metaplasia patients with progressive disease, oxidative-induced damage results in telomere shortening and mutations in the p53 gene abrogate p53’s role as the checkpoint of proliferation and apoptosis.32 Other studies have also shown that p53 plays a vital role in the synthesis of key components of the electron transfer chain.33,34,35

p53 mediates its central metabolic role through TP53-induced glycolysis and apoptosis regulator, or TIGAR.27,36 TIGAR acts as a phosphatase and degrades fructose-2,6-Bisphosphate (F26B) thereby decreasing the activity of phosphofructose kinase 1 (PFK1), a key enzyme of the glycolytic pathway.27 p53, via TIGAR, decreases glycolysis by diverting glycolytic intermediates into the pentose phosphate pathway (PPP).37 p53 also negatively regulates the expression of Pyruvate dehydrogenase kinase 2 (PDK-2) thereby inactivating the pyruvate dehydrogenase complex responsible for converting pyruvate to acetyl-CoA.38 Thus p53, activating the pyruvate dehydrogenase complex, favours oxidative phosphorylation through the production of acetyl-CoA.38 Furthermore, p53 directly downregulates the expression of GLUT1 and GLUT4.39 

The role of p53 is highlighted in hypoxic microenvironments. Through a hypoxia-induced HIF1α dependent mechanism, TIGAR has been shown to form a complex with hexokinase 2 at the mitochondria resulting in an increase in hexokinase 2 activity.36 This complex reduces glycolytic flux supporting pentose phosphate pathway activity, generates NADPH in the process and promotes antioxidant function thereby limiting reactive oxygen species-associated apoptosis and autophagy.36 p53 also represses the expression of monocarboxylate transporter 1 (MCT1) preventing the efflux of lactate under hypoxic conditions.40 It has been speculated that aberrant p53 expression may even promote tumour progression as some evidence suggests that p53 may enhance aerobic glycolysis rather than inhibit it.27,28,41 In addition, the mechanism by which p53 regulates the glycolytic pathway may be tissue and context specific which is thought to reflect different types of cellular stress, that is, metabolic, oxidative and hypoxic stress.27,42

NOVEL MEDIATORS LINKING ENERGY METABOLISM WITH INFLAMMATION, HYPOXIA AND ANGIOGENESIS 

NFκB 

Despite some early studies linking nuclear factor kappa B (NFκB) with energy metabolism, recent studies have increasingly shown NFκB to possess an equally important central role in various metabolic and pathological diseases.43 Inflammation is a key factor in the development of metabolic diseases such as atherosclerosis, insulin resistance, type 2 diabetes and obesity.43,44,45 The central role of NFκB in immunity, inflammation and carcinogenesis has been well documented.46,47,48

NFκB regulates cellular respiration in a p53-dependent manner (Figure 2).49 Translocation of the NFκB family member RelA to mitochondria is inhibited by p53, however, in the absence of p53, RelA is transported into mitochondria and recruited to the mitochondrial genome where it represses mitochondrial gene expression, oxygen consumption and cellular ATP levels, thereby contributing to the switch to glycolysis.49 Indeed, it was reported that the RelA subunit also upregulates transcription of GLUT3 resulting in increases in glucose uptake and glycolytic flux.50 The elevated glycolytic flux stimulates further IKK/NFκB pathway activity in a positive feedback loop that subsequently promotes H-Ras-induced oncogenic transformation in mouse embryonic fibroblasts.50 This was the first functional study to show that NFκB promotes cell growth and carcinogenesis by metabolic manipulation, but crucially, p53 was central to this pathway, as introduction of p53 disrupted the link between NFκB and glycolysis.50

Intriguingly, the role of NFκB is reversed in normal mouse embryonic fibroblasts upon glucose starvation, whereby NFκB inhibition causes cellular reprogramming to aerobic glycolysis.51 The role of NFκB in upregulating mitochondrial respiration in this circumstance involves  the p53-mediated upregulation of mitochondrial synthesis of cytochrome c oxidase 2 (SCO2), a key component of complex IV of the electron transport chain.51  Hence, NFκB can act as a focal checkpoint of metabolic homeostasis in conjunction with AMPK and p53 to regulate the response to low cellular ATP levels.51  Therefore, despite its prominent role in the Warburg effect, the metabolic plasticity of NFκB confers adaptivity in cells to adapt to fluctuating oxidative and hypoxic microenvironments. 

VEGF and PFKFB3 

In response to hypoxia-induced pro-angiogenic stimuli, endothelial cells rapidly switch from a metabolically inactive state of quiescence to an active migratory and proliferative state.52 Effective vascular sprouting relies on coordinated navigating tips cells and on proliferating stalk cells that elongate the sprout. Until recently, only genetic signals were known to play a role in this angiogenic switch. However, the angiogenic switch also requires a change in endothelial cell metabolism.52 Interestingly, endothelial cells are thought to be addicted to glycolysis as they rely minimally on oxidative phosphorylation for ATP generation.52,53 For instance, the glycolytic inhibitor 2-deoxy-D-glucose induces significant endothelial cell death.54

Endothelial cells increase their glycolytic rate by upregulating a range of glycolytic constituents including GLUT1, LDH and 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase-3 (PFKFB3).52 PFKFB3 has been shown to be critical for angiogenic sprouting, and its inactivation reduces endothelial cell proliferation and migration, and impairs motility and formation of endothelial cell lamellipodia and filopodia.54 Conversely, PFKFB3 overexpression stimulates the sprouting of mitotically-inactivated endothelial cells and promotes tip cell formation.54 This entire process of tip and stalk cell differentiation, however, is under tight control of vascular epithelial growth factor (VEGF) and Notch signalling (Figure 2).53

VEGF promotes tip cell induction and filopodia formation inducing the expression of the Notch ligand Delta-like 4 (DLL4).53 One of the main genetic signals of vessel sprouting is orchestrated through Notch.55 DLL4 subsequently activates Notch signalling in neighbouring cells and suppresses VEGF receptor 2 expression and tip cell behaviour.53 Therefore, the activation of VEGF receptor 2 in tips cells upregulates PFKFB3 levels and glycolysis but induces the expression of the Notch ligand DLL4 in neighbouring stalk cells activating Notch signalling, lowering VEGF receptor 2 expression resulting in lower PFKFB3 expression and glycolytic flux.53 Interestingly, overexpression of PFKFB3 overcomes the pro-stalk activity of Notch signalling thereby promoting tip cell behaviour, indicating that highly glycolytic endothelial cells can overcome inhibitory genetic signals.54

VEGF also controls angiogenesis through the glycolytic metabolite lactate.52 Once taken up by endothelial cells through MCT1, lactate competitively inhibits the oxygen-sensing propyl hydroxylase domain protein 2 (PHD2), resulting in activation of HIF1α and an increase in VEGF receptor 2 expression.56 Lactate signalling also induces VEGF expression.57 In addition to its angiogenic role, lactate also indirectly releases NFκB inducing IL8 expression, another promoter of angiogenesis.58  VEGF has also been shown to induce the production of IL8 in endothelial cells.59 In addition to promoting aerobic glycolysis, VEGF stimulates mitochondrial biogenesis through Akt-dependent signalling, and plays a significant role in fatty acid metabolism.60,61,62 These studies demonstrate a close relationship between VEGF-induced metabolism, hypoxia, angiogenesis and inflammation in endothelial cells and highlight how stressed endothelial cells adapt to an altering milieu that could potentially favour tumour progression.

mTOR, Succinate and STAT3 

mTOR is a serine/threonine kinase that controls cell proliferation and metabolism in response to a range of extracellular stimuli such as the availability of nutrients, growth factors and stress.63 mTOR plays an important role in the modulation of both innate and adaptive immune function (Figure 2).64 As discussed, activated T cells switch to an anabolic metabolism using aerobic glycolysis as a major supply of ATP to fuel the rapid synthesis of proteins , nucleotides and other biosynthetic products.63 TORC1, one of two currently recognised signalling forms of mTOR, has been shown to be heavily involved in the up-regulation of enzymes involved in glycolysis, glutaminolysis, the pentose phosphate pathway, surface expression of GLUT1 and expression of the glutamine transporter, SNAT-2.65,66,67 Similarly, inhibition of mTOR results in a metabolic bias towards oxidative phosphorylation and has been shown to produce a larger CD8 memory T cell pool.68 Ongoing clinical trials investigating the efficacy of mTOR inhibitors suggest that mTOR-mediated metabolism does play a central role in regulating biological outcomes within immune cells, however, a key question remaining is how mTOR-mediated metabolism is coupled to immune function.5

Increasing evidence also proposes that succinate, a citric acid cycle metabolite that accumulates due to succinate dehydrogenase mutations, transmits an oncogenic signal from the mitochondria to the cytosol, directly inhibiting PHD and resulting in HIF1α stabilization under normoxic conditions, with resultant increased expression of genes that facilitate angiogenesis, metastasis and glycolysis (Figure 2).69 By adding succinate to glioblastoma multiforme-derived cells cultured under hypoxic conditions, HIF1α stabilization is induced which increases stem cell fractions and preserves the tumour stem cell niche thereby promoting tumour survival.70 Recently, it has been reported that succinate as a metabolite in innate immune function enhances IL-1β production during inflammation through HIF1α thereby promoting disease progression.71

The signal transducer and activator of transcription factors (STATs) are a family of transcription factors that regulate cell growth, survival, differentiation and motility.72 One of the STAT members, STAT3, has long been recognised as a critical regulator of tumour cells.72 STAT3 has been recently found to act as one of the central mediators of aerobic glycolysis through both HIF1α and independent mechanisms (Figure 2).72,73 Upon translocation to the mitochondria, serine phosphorylation of STAT3 contributes to tumour cell transformation and tumourigenesis.74,75,76

EXPLORING THE MOLECULAR MECHANISMS THAT LINK ENERGY METABOLISM AND HYPOXIA 

Glycolysis, Hypoxia and Rheumatoid Arthritis               

It has been 35 years since the link between increased glycolytic activity and rheumatoid arthritis (RA) was first established.77 In normal synovial tissues, glycolysis is the primary pathway for mitochondrial substrate oxidation of pyruvate.78 Levels of two major glycolytic enzymes glyceraldehyde 3-phosphate dehydrogenase and LDH were found to be significantly increased in the synovial cells between fresh non-rheumatoid and rheumatoid synovial tissue.77 More recently, one study detected elevated lactate and reduced glucose levels in the synovial fluid in RA.79 Moreover, it is plausible that metabolic alterations that favour aerobic glycolysis are a result of hypoxia-induced mitochondrial mutagenesis and dysfunction.80 Despite studies lacking strong evidence of a direct relationship between inflammation and glycolysis in RA, it is interesting that some glycolytic components are characterised as being autoantigens, for example, glucose-6-phosphate isomerise, aldolase and enolase.78 However, studies need to be undertaken to examine the role of metabolic autoantigens in cancer initiation and progression.

On the other hand, the link between hypoxia and inflammation has been well documented in-vivo.81,82,83,84 Significantly higher levels of synovial fluid tumour necrosis factor-α (TNFα), IL-1β, interferon-γ and macrophage inflammatory protein-3α in combination with low partial oxygen pressures of <20 mm Hg were found in patients with inflammatory arthritis.84 Interestingly, TNFα blocking therapy reverses joint inflammation and hypoxia.83 Another study also demonstrated that hypoxia-induced IL-17A expression is localised to neutrophils, mast cells and T cells within inflamed synovial tissue supporting the concept that IL-17A is a key mediator in inflammatory arthritis.81 

Numerous mechanistic processes within the inflammatory joint may alter energy metabolism profiles. RA is associated with increased levels of HIF1α and HIF2α.85,86 HIF1 also induces the expression of GLUT1 and GLUT3.87 Furthermore, HIF has been shown to regulate the levels of hexokinase II, glyceraldehyde 3-phosphate dehydrogenase, LDH and cytochrome oxidase in the inflammatory synovium.87,88,89,90,91 RA is also commonly associated with mutations in p53.92,93,94 As discussed, p53 can regulate glucose metabolism through NFκB, however, loss of p53 promotes the positive feedback cycle between the IKK-NFκB pathway and glycolysis thereby promoting oncogenic transformation.50,95 It is also plausible that angiogenic factors, such as VEGF, within the hypoxic inflammatory joint may induce alterations in energy metabolism profiles.96,97 Therefore, it may be enticing to speculate that metabolic perturbations within the inflammatory joint are a consequence of the combined contribution of many reciprocal mechanisms, for example, aberrantly expressed HIF1α, HIF2α, VEGF, NFκB and mutations in p53. 

AMPK, Hypoxia and Circadian Rhythms 

Significant time-of-day oscillations in glucose metabolism are observed in both humans and rodent models, at both the whole body and cellular level.98 It has been speculated that various mitochondrial functions may be regulated by the circadian clock thereby serving as a central coordinator between the clock and cellular energy metabolism.99 For example, cytochrome c oxidase activity is increased in the brains of 2 month old wistar rats during wakefulness compared to sleep to meet increased energy demands.100 AMPK is one of the main metabolic sensors responsible for transmitting energy dependent signals to the mammalian clock.101

A molecular oscillator exists whereby the transcription factors CLOCK and BMAL1 work together to drive the expression of many genes responsible for the mammalian molecular clock, including those encoding their own inhibitors, the period (PER1, PER2 and PER3) and cryptochrome (CRY 1 and CRY2) proteins.101 CRY1 and CRY2 are transcriptional repressors that are necessary for circadian clock function.102 The E3 ligase component F-box/LRR-repeat protein 3 (FBXL3) catalyzes the polyubiquitination of CYR1 and CRY2 and thus stimulates their proteosomal degradation.103 AMPK-mediated serine phosphorylation of CRY1 and CRY2 initiates the interaction between CRY1, CRY2 and FBXL3 and stimulates the degradation of both cryptochromes.104 Casein kinases, CKIε and CKIδ, are also important modulators of circadian rhythm in mammals.101 Genetic disruption or pharmacological inhibition of these casein kinases alters behavioural and cellular circadian rhythms in mice.105 Casein kinases phosphorylate serines in PER2, however, AMPK was reported to phosphorylate CKIε at serine 389 thereby increasing its enzymatic activity and indirectly leading to destabilisation of PER2 and alterations in circadian rhythm.106

 

AMPK has also been implicated in circadian rhythm entrainment in mice as pharmacological activation of AMPK by intraperitoneal injection of both 5-aminoimidazole-4-carboxyamide ribonucleaoside (AICAR) or metformin casues a phase shift of the liver clock.104,106 In addition, AICAR stimulation altered clock gene expression in wild type mice but not in mice lacking the AMPKγ3 subunit implying that AMPK activation may play a role in circadian entrainment.107 Furthermore, AMPK possesses a close relationship with silent mating type information regulation 2 homolog 1 (SIRT1), another fuel-sensing molecule key to nutritional status and circadian regulation. AMPK not only enhances SIRT1 activity by increasing NAD+ levels but activation of SIRT1 causes AMPK phosphorylation via LBK1 activation.108,109 AMPK is also associated with regulating other metabolic sensors known to have key roles in circadian regulation such as poly (ADP-ribose) polymerase 1 and nicotinamide phosphoribosyltransderase.110,111 These studies suggest that the effectiveness of widely prescribed drugs that regulate glucose homeostasis, such as metformin, may be ameliorated by altering the timing of treatment through AMPK-mediated control of circadian function by pharmacological intervention.

 

RECIPROCAL MECHANISMS LINKING ENERGY METABOLISM TO INFLAMMATION IN GASTROENTEROLOGICAL DISEASES

 

It is apparent thus far, that the combined effect of various molecular processes, can act in tandem to significantly alter the local microenvironment and attenuate disease progression. Little is known about how energy metabolism profiles cooperate with inflammatory processes to facilitate metaplastic progression in gastroenterological disease entities. However, some recent research has provided some insight on metabolic signatures in Barrett’s oesophagus, oesophageal adenocarcinoma, Inflammatory Bowel Disease (IBD), gastritis and gastric cancer.112-116

 

Recent research has demonstrated that both oxidative phosphorylation and glycolysis are reprogrammed early in the inflamed Barrett’s disease sequence and may act mutually to promote disease progression in Barrett’s oesophagus.117 Subsequent to screening 84 genes using a PCR microarray, validations utilising in-vitro and in-vivo models found that 3 genes associated with mitochondrial energy metabolism, ATP12A, COX4I2 and COX8C, were differentially expressed across the Barrett’s sequence.117 In addition, tissue microarrays demonstrated significant epithelial and stromal alterations using surrogate protein markers of oxidative phosphorylation, ATP synthase subunit 5 beta and heat shock protein 60, or ATP5B and HSP60 respectively.117 Moreover, significant alterations across the Barrett’s sequence were also demonstrated using surrogate protein markers of glycolysis, pyruvate kinase isozyme M2 and glyceraldehydes 3-phosphate dehydrogenase, or PKM2 and GAPDH respectively.117 Interestingly, ATP5B in sequential follow up surveillance biopsy material segregated Barrett’s non progressors and progressors to high grade dysplasia and adenocarcinoma thereby highlighting the prognostic advantage of metabolic profiles in these pre-neoplastic patients.117 Finally, utilising the in-vitro model, the authors present evidence that Barrett’s and adenocarcinoma cells exhibit significantly altered levels of various oxidative parameters, whereby the adenocarcinoma cell line maintains an equilibrium between both metabolic pathways while the Barrett’s cell line favours a more detrimental oxidative phenotype that may be selected for during early stages of disease progression .117

 

Other studies, although mostly indirectly, link inflammation to energy metabolism. IL-6, documented as being increased in myofibroblasts of Crohn’s disease patients, has also been shown to increase the expression of hexokinase 2 and PFKFB3 in murine embryonic fibroblasts.118,119 Moreover, increased secreted and immunological levels of IL-6 have been found in Barrett’s tissue compared to matched normal adjacent squamous epithelium.120 Aberrant expression of p53 is also associated with an increased risk of neoplastic progression in patients with Barrett’s oesophagus.121 Therefore, since mutated p53 enhances IL-6 promoter activity in renal cell carcinoma, it may be plausible that p53, known to modulate oxidative phosphorylation and glycolysis, simultaneously alters inflammatory and metabolic profiles in pre-neoplastic and neoplastic microenvironments of the oesophagus.50,122 Similarly, HIF1α, known to mediate hypoxia-induced alterations in glycolytic metabolism and to possess an intrinsic relationship with p53, has been shown to be differentially expressed across the Barrett’s sequence.6,36

 

Interestingly, despite increased oxidative phosphorylation in Barrett’s oesophagus, ulcerative colitis is associated with low levels of this metabolic pathway.114,123 Loss of oxidative phosphorylation precedes the development of dysplasia in ulcerative colitis and thus could potentially be utilised to predict cancer.114 Furthermore, following preneoplastic progression, cancer cells restore mitochondria indicative of an increase in energy demands for growth and proliferation.114 In addition, one study showed that increasing mucosal levels of ATP can protect mice from colitis and thus increasing ATP synthesis could be a plausible therapeutic approach for ulcerative colitis.123 Such reductions in oxidative phosphorylation, thought to be caused by defects in complex I of the electron transport chain, have also been reported in atrophic and active chronic gastritis.115,124 Gastric cancer is additionally associated with a complex I-induced defective electron transport chain.125 As well as decreased mitochondrial respiration, gastric cancer exhibits shifts to glycolysis.113,126 Decreased fructose-1,6-bisphosphatase (FBP), the enzyme which functions to antagonise glycolysis though NFκB, has been shown to be decreased in both gastric cell lines and gastric carcinomas thereby promoting glycolysis.127,128 PDK-1 and PKM2 are also overexpressed in gastric and colorectal tumour tissue and their expression is associated with poor survival.129-131 Moreover, knockdown of PKM2 has been shown to repress the proliferative and migratory capabilities of colorectal cancer cells.130

 

IBD patients are known to have high levels of HIF1α and HIF2α.132 IBD patients also exhibit increased colonic expression of various glycolytic enzymes and these alterations in metabolism are thought to be triggered by hypoxic stress.133 Such extensive regulation of various glycolytic intermediates could be mediated by central regulators of metabolism known to associate with HIF, for example, PFKFB. One study in gastric cancer cell lines and tissue found that both PFKFB3 and PFKFB4 significantly responded to hypoxia through HIF1α and this subsequently promoted the Warburg effect.134 Interestingly, alterations in the gut microbiome, as demonstrated by the fucosyltransferase 2 polymorphism in Crohn’s disease patients for example, could also affect the host mucosal state and thus increase disease susceptibility.135 Therefore, further studies directly linking inflammation with energy metabolism profiles through these distinct processes would enhance our understanding on the mechanisms involved in inflammatory-induced neoplastic progression in gasterological diseases.

 

INNOVATIVE METABOLIC-BASED TREATMENTS AND MULTI-TARGETED THERAPIES

 

In order to replicate and divide, tumour cells need to possess the ability to acquire large quantities of proteins, lipids and nucleotides.5 As these processes are highly metabolically demanding, cells additionally require vast quantities of ATP. Consequently, targeting glucose metabolism and nucleotide biosynthesis could have significant advantages on combating metabolic transformation. In addition, altering the metabolism of susceptible or predisposed pre-neoplastic or neoplastic tissue may prevent subsequent disease progression.

 

Table 1 highlights some of the diverse therapeutic strategies currently being employed to target various aspects of energy metabolism. Significant research has begun to focus on targeting upstream regulators of metabolic pathways such as HIF, phosphoinositide 3-kinase (PI3K), Akt, mTOR and AMPK.5 For example, PI3K inhibitors such as BEZ235 have been shown to target metabolism leading to cancer regression in Kras-mutant murine lung adenocarcinomas.136 The AMPK activator metformin, primarily used to treat patients with type 2 diabetes, has been shown to be protective as those treated with metformin were found to be cancer free over 8 years versus those on alternative treatment regimes.137 Additional AMPK activators are also being investigated for their potential therapeutic use.138 Interestingly, methotrexate, a chemotherapeutic known to target nucleotide biosynthesis, enhances the antianabolic and antiproliferative effects of AICAR, an alternative AMPK agonist.139 Moreover, targeting nucleotide biosynthesis may be more favourable as nucleotide building blocks necessary for proliferating tumour cells can be synthesised by endogenous glucose and glutamine due to poor vascularisation.5 Therefore, blocking ribose-5-phosphate synthesis, with 5-fluorouracil (5-FU) for example, could provide a better therapeutic window. Dichloroacetate, an inhibitor of PDK-1, has also been shown to re-sensitise gastric cancer cells with hypoxia-induced resistance to 5-FU through the alteration of glycolysis.140

 

Table 1: Metabolic-based compounds.
Drug Target Mechanism Reference(s)
2-deoxyglucose Glycolytic Pathway Inhibits the production of glucose-6-phosphate (5,142)
3-bromopyruvate Glycolytic Pathway Inhibits GAPDH (5)
3-PO Glycolytic Pathway Inhibits PFKFB3 (145)
5-FU Nucleotide Biosynthetic Pathway Inhibits cell proliferation (5)
BEZ235 PI3K/mTOR Pathways Inhibits PI3K signalling & mTORC1/mTORC2 (136)
Dichloroacetate Glycolytic Pathway Inhibits PDK-1 (140)
Lonidamine Glycolytic Pathway Inhibits hexokinase and mitochondrial respiration (5)
Metformin AMPK agonist Activates AMPK (137)
Methotrexate AMPK / Nucleotide Biosynthetic Pathway Activates AMPK (5,139)
Phloretin Glucose Transport Inhibits sodium-glucose transporters 1 & 2 (143,144)
Phloridzin Glucose Transport Inhibits sodium-glucose transporters 1 & 2 (143,144)
PX-478 HIF1α Inhibits HIF signalling (5)
Salicylate AMPK agonist Activates AMPK (138)
TLN-232 Glycolytic Pathway Inhibits PKM2 (5)

 

 

Therapeutic agents that target the glycolytic pathway such as 2-deoxyglucose, lonidamine, 3-bromopyruvate and TLN-232 have also shown significant promise.5 Despite not showing substantial effects on tumour growth as monotherapeutic drugs, their use in conjunction with other chemotherapeutic reagents seems to sensitive tumours by reducing ATP levels and perhaps by indirectly limiting the availability of macromolecules synthesised through anapleurotic interactions.141 In addition to being combined with radiotherapy, some glycolytic inhibitors are currently being used in phase I, II and III clinical trial.5,142 Inhibitors that target glucose transport across the plasma membrane, such as phloridzin and phloretin, have also shown efficacy in inhibiting in-vitro, xenograft and in-vivo tumour growth.143,144 Inhibition of PFKFB3 with 3-(3-pyridinyl)-1-(4-pyridinyl)-2-propen-1-one, or 3-PO, known to be a selective agent against neoplastic cells, has been shown to reduce cellular lactate, ATP, NAD+ and other cellular metabolites within several human malignant hematopoietic and adenocarcinoma cell lines.145

 

                        Additional therapeutic modalities that can be exploited include targeting HIF1α, lactate transporters, amino acid metabolism and lipid metabolism.5 Altering diet is also a unique and beneficial therapeutic approach. For example, a ketogenic diet relies on food that does not increase plasma glucose but produces ketone bodies that can be used as a carbon source thereby bypassing glycolysis.146 Even though the ketogenic diet has been shown to have mixed results, further studies may reveal that it is a cancer specific therapy.146 More recently, micro RNAs have shown promise at targeting cancer metabolic pathways. A recent study demonstrated that mir-122 targets PKM2 and affects metabolism in hepatocellular carcinoma.147 Despite the encouraging evolution of metabolic-based treatments and multi-targeted therapies, more work is required to understand which pathways are activated in distinct tumour types thereby allowing the identification of pharmacological targets that can avert disease progression and alleviate tumour burden.

 

CONCLUSION

 

Cellular energy metabolism plays a crucial role in inflammatory, hypoxic and angiogenic microenvironments by supporting malignant progression in a range of disease entities. Pre-neoplastic and neoplastic tissue must use a diverse range of molecular components to alter their metabolism to adapt to fluctuating oxidative, hypoxic and metabolic stresses. This involves exploiting various molecular elements such as HIF1α, AMPK or p53 that have the potential to function rapidly to acute onsets of stress. It is evident from ongoing research, however, that tumour cells can survive these stresses by adjusting their metabolism through a range of alternative pathways and novel mediators such as NFκB, VEGF and mTOR. Substantiating the reciprocal relationship between energy metabolism in inflammatory and hypoxic diseases is evident in RA and circadian rhythms. In addition, it is clear that the inflammatory microenvironment of the gastrointestinal tract presents clear indication of this mutual association. Therefore, understanding the underlying mechanisms that permit premalignant cells to transform, survive, thrive and subsequently adapt in response to a range of metabolic-based therapies will aid considerably in the development of effective and specific multi-targeted therapies.

 

ACKNOWLEDGEMENTS

 

The authors would kindly wish to acknowledge funding from Science Foundation Ireland (SFI 11/RFP.1/CAN/3137).

 

CONFLICTS OF INTEREST

 

The authors declare no conflicts of interest.

1. Warburg O. On the origin of cancer cells. Science. 1956; 123(3191): 309-314. doi: 10.1126/science.123.3191.309

2. Hanahan D, Weinberg RA. The hallmarks of cancer. Cell. 2000; 100(1): 57-70. doi: 10.1016/S0092-8674(00)81683-9

3. Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011; 144(5): 646-674. doi: 10.1016/j.cell.2011.02.013

4. Weinberg F, Chandel NS. Mitochondrial metabolism and cancer. Annals of the New York Academy of Sciences. 2009; 1177: 66-73. doi: 10.1111/j.1749-6632.2009.05039.x

5. Tennant DA, Duran RV, Gottlieb E. Targeting metabolic transformation for cancer therapy. Nature reviews Cancer. 2010; 10(4): 267-277. doi: 10.1038/nrc2817

6. Ling FC, Khochfar J, Baldus SE, et al. HIF-1alpha protein expression is associated with the environmental inflammatory reaction in Barrett’s metaplasia. Diseases of the esophagus : official journal of the International Society for Diseases of the Esophagus / ISDE. 2009; 22(8): 694-699. doi: 10.1111/j.1442-2050.2009.00957.x

7. Schmaltz C, Hardenbergh PH, Wells A, Fisher DE. Regulation of proliferation-survival decisions during tumor cell hypoxia. Molecular and cellular biology. 1998; 18(5): 2845-2854. doi: 10.1128%2Fmcb.18.5.2845

8. Brizel DM, Scully SP, Harrelson JM, et al. Tumor oxygenation predicts for the likelihood of distant metastases in human soft tissue sarcoma. Cancer research. 1996; 56(5): 941-943.

9. Casey TM, Pakay JL, Guppy M, Arthur PG. Hypoxia causes downregulation of protein and RNA synthesis in noncontracting Mammalian cardiomyocytes. Circulation research. 2002; 90(7): 777-783. doi: 10.1161/01.RES.0000015592.95986.0

10. Semenza GL. Regulation of metabolism by hypoxia-inducible factor 1. Cold Spring Harbor symposia on quantitative biology. 2011; 76: 347-353. doi: 10.1101/sqb.2011.76.010678

11. Zheng Y, Delgoffe GM, Meyer CF, Chan W, Powell JD. Anergic T cells are metabolically anergic. Journal of immunology. 2009; 183(10): 6095-6101. doi: 10.4049/jimmunol.0803510

12. Fox CJ, Hammerman PS, Thompson CB. Fuel feeds function: energy metabolism and the T-cell response. Nature reviews Immunology. 2005; 5(11): 844-852. doi: 10.1038/nri1710

13. Dziurla R, Gaber T, Fangradt M, et al. Effects of hypoxia and/or lack of glucose on cellular energy metabolism and cytokine production in stimulated human CD4+ T lymphocytes. Immunology letters. 2010; 131(1): 97-105. doi: 10.1016/j.imlet.2010.02.008

14. Dang EV, Barbi J, Yang HY, et al. Control of T(H)17/T(reg) balance by hypoxia-inducible factor 1. Cell. 2011; 146(5): 772-784. doi: 10.1016/j.cell.2011.07.033

15. Shi LZ, Wang R, Huang G, et al. HIF1alpha-dependent glycolytic pathway orchestrates a metabolic checkpoint for the differentiation of TH17 and Treg cells. The Journal of experimental medicine. 2011; 208(7): 1367-1376. doi: 10.1084/jem.20110278

16. Finlay DK, Rosenzweig E, Sinclair LV, et al. PDK1 regulation of mTOR and hypoxia-inducible factor 1 integrate metabolism and migration of CD8+ T cells. The Journal of experimental medicine. 2012; 209(13): 2441-2453. doi: 10.1084/jem.20112607

17. Dandapani M, Hardie DG. AMPK: opposing the metabolic changes in both tumour cells and inflammatory cells? Biochemical Society Transactions. 2013; 41(2): 687-693. doi: 10.1042/BST20120351

18. Merrill GF, Kurth EJ, Hardie DG, Winder WW. AICA riboside increases AMP-activated protein kinase, fatty acid oxidation, and glucose uptake in rat muscle. The American Journal of Physiology. 1997; 273(6 Pt 1): E1107-E1112. doi: 10.1152/ajpendo.1997.273.6.e1107

19. Zong H, Ren JM, Young LH, et al. AMP kinase is required for mitochondrial biogenesis in skeletal muscle in response to chronic energy deprivation. Proceedings of the National Academy of Sciences of the United States of America. 2002; 99(25): 15983-15987. doi: 10.1073/pnas.252625599

20. Winder WW, Holmes BF, Rubink DS, Jensen EB, Chen M, Holloszy JO. Activation of AMP-activated protein kinase increases mitochondrial enzymes in skeletal muscle. Journal of applied physiology (Bethesda, Md : 1985). 2000; 88(6): 2219-2226. doi: 10.1152/jappl.2000.88.6.2219

21. Hardie DG, Ross FA, Hawley SA. AMPK: a nutrient and energy sensor that maintains energy homeostasis. Nature reviews Molecular cell biology. 2012; 13(4): 251-262. doi: 10.1038/nrm3311

22. Hoppe S, Bierhoff H, Cado I, et al. AMP-activated protein kinase adapts rRNA synthesis to cellular energy supply. Proceedings of the National Academy of Sciences of the United States of America. 2009; 106(42): 17781-17786. doi: 10.1073/pnas.0909873106

23. Shackelford DB, Vasquez DS, Corbeil J, et al. mTOR and HIF-1alpha-mediated tumor metabolism in an LKB1 mouse model of Peutz-Jeghers syndrome. Proceedings of the National Academy of Sciences of the United States of America. 2009; 106(27): 11137-11142. doi: 10.1073/pnas.0900465106

24. Krawczyk CM, Holowka T, Sun J, et al. Toll-like receptor-induced changes in glycolytic metabolism regulate dendritic cell activation. Blood. 2010; 115(23): 4742-4749. doi: 10.1182/blood-2009-10-249540

25. Sag D, Carling D, Stout RD, Suttles J. Adenosine 5’-monophosphate-activated protein kinase promotes macrophage polarization to an anti-inflammatory functional phenotype. Journal of immunology (Baltimore, Md : 1950). 2008; 181(12): 8633-8641. doi: 10.4049/​jimmunol.181.12.8633

26. Rolf J, Zarrouk M, Finlay DK, Foretz M, Viollet B, Cantrell DA. AMPKalpha1: a glucose sensor that controls CD8 T-cell memory. European journal of immunology. 2013; 43(4): 889-896. doi: 10.1002/eji.201243008

27. Berkers CR, Maddocks OD, Cheung EC, Mor I, Vousden KH. Metabolic regulation by p53 family members. Cell Metabolism. 2013; 18(5): 617-633. doi: 10.1016/j.cmet.2013.06.019

28. Vousden KH, Ryan KM. p53 and metabolism. Nature reviews Cancer. 2009; 9(10): 691-700. doi: 10.1038/nrc2715

29. Kulawiec M, Ayyasamy V, Singh KK. p53 regulates mtDNA copy number and mitocheckpoint pathway. Journal of carcinogenesis. 2009; 8: 8. doi: 10.4103/1477-3163.50893

30. Sahin E, Colla S, Liesa M, et al. Telomere dysfunction induces metabolic and mitochondrial compromise. Nature. 2011; 470(7334): 359-365. doi: 10.1038/nature09787

31. Kitamura N, Nakamura Y, Miyamoto Y, et al. Mieap, a p53-inducible protein, controls mitochondrial quality by repairing or eliminating unhealthy mitochondria. PLoS One. 2011; 6(1): e16060. doi: 10.1371/journal.pone.0016060

32. Cardin R, Piciocchi M, Tieppo C, et al. Oxidative DNA damage in Barrett mucosa: correlation with telomeric dysfunction and p53 mutation. Annals of surgical oncology. 2013; 20 Suppl 3: S583-S589. doi: 10.1245/s10434-013-3043-1

33. Matoba S, Kang JG, Patino WD, et al. p53 regulates mitochondrial respiration. Science. 2006; 312(5780): 1650-1653. doi: 10.1126/science.1126863

34. Okamura S, Ng CC, Koyama K, et al. Identification of seven genes regulated by wild-type p53 in a colon cancer cell line carrying a well-controlled wild-type p53 expression system. Oncology research. 1999; 11(6): 281-285.

35. Stambolsky P, Weisz L, Shats I, et al. Regulation of AIF expression by p53. Cell death and differentiation. 2006; 13(12): 2140-2149. doi: 10.1038/sj.cdd.4401965

36. Cheung EC, Ludwig RL, Vousden KH. Mitochondrial localization of TIGAR under hypoxia stimulates HK2 and lowers ROS and cell death. Proceedings of the National Academy of Sciences of the United States of America. 2012; 109(50): 20491-20496. doi: 10.1073/pnas.1206530109

37. Li H, Jogl G. Structural and biochemical studies of TIGAR (TP53-induced glycolysis and apoptosis regulator). The Journal of biological chemistry. 2009; 284(3): 1748-1754. doi: 10.1074/jbc.M807821200

38. Contractor T, Harris CR. p53 negatively regulates transcription of the pyruvate dehydrogenase kinase Pdk2. Cancer research. 2012; 72(2): 560-567. doi: 10.1158/0008-5472.CAN-11-1215

39. Schwartzenberg-Bar-Yoseph F, Armoni M, Karnieli E. The tumor suppressor p53 down-regulates glucose transporters GLUT1 and GLUT4 gene expression. Cancer research. 2004; 64(7): 2627-2633. doi: 10.1158/0008-5472.CAN-03-0846

40. Boidot R, Vegran F, Meulle A, et al. Regulation of monocarboxylate transporter MCT1 expression by p53 mediates inward and outward lactate fluxes in tumors. Cancer research. 2012; 72(4): 939-948. doi: 10.1158/0008-5472.CAN-11-2474

41. Mathupala SP, Heese C, Pedersen PL. Glucose catabolism in cancer cells. The type II hexokinase promoter contains functionally active response elements for the tumor suppressor p53. The Journal of biological chemistry. 1997; 272(36): 22776-22780. doi: 10.1074/jbc.272.36.22776

42. Maddocks OD, Vousden KH. Metabolic regulation by p53. Journal of molecular medicine (Berlin, Germany). 2011; 89(3): 237-245. doi: 10.1007/s00109-011-0735-5

43. Hotamisligil GS, Erbay E. Nutrient sensing and inflammation in metabolic diseases. Nature reviews Immunology. 2008; 8(12): 923-934. doi: 10.1038/nri2449

44. Solinas G, Karin M. JNK1 and IKKbeta: molecular links between obesity and metabolic dysfunction. FASEB journal : official publication of the Federation of American Societies for Experimental Biology. 2010; 24(8): 2596-2611. doi: 10.1096/fj.09-151340

45. Tornatore L, Thotakura AK, Bennett J, Moretti M, Franzoso G. The nuclear factor kappa B signaling pathway: integrating metabolism with inflammation. Trends in cell biology. 2012; 22(11): 557-566. doi: 10.1016/j.tcb.2012.08.001

46. Marusawa H, Jenkins BJ. Inflammation and gastrointestinal cancer: an overview. Cancer Letters. 2014; 345(2): 153-156. doi: 10.1016/j.canlet.2013.08.025

47. DiDonato JA, Mercurio F, Karin M. NF-kappaB and the link between inflammation and cancer. Immunological reviews. 2012; 246(1): 379-400. doi: 10.1111/j.1600-065X.2012.01099.x

48. Grivennikov SI, Greten FR, Karin M. Immunity, inflammation, and cancer. Cell. 2010; 140(6): 883-899. doi: 10.1016/j.cell.2010.01.025

49. Johnson RF, Witzel, II, Perkins ND. p53-dependent regulation of mitochondrial energy production by the RelA subunit of NF-kappaB. Cancer research. 2011; 71(16): 5588-5597. doi: 10.1158/0008-5472.CAN-10-4252

50. Kawauchi K, Araki K, Tobiume K, Tanaka N. p53 regulates glucose metabolism through an IKK-NF-kappaB pathway and inhibits cell transformation. Nature cell biology. 2008; 10(5): 611-618. doi: 10.1038/ncb1724

51. Mauro C, Leow SC, Anso E, et al. NF-kappaB controls energy homeostasis and metabolic adaptation by upregulating mitochondrial respiration. Nature cell biology. 2011; 13(10): 1272-1279. doi: 10.1038/ncb2324

52. Eelen G, Cruys B, Welti J, De Bock K, Carmeliet P. Control of vessel sprouting by genetic and metabolic determinants. Trends in endocrinology and metabolism: TEM. 2013; 24(12): 589-596. doi: 10.1016/j.tem.2013.08.006

53. De Bock K, Georgiadou M, Carmeliet P. Role of endothelial cell metabolism in vessel sprouting. Cell metabolism. 2013; 18(5): 634-647. doi: 10.1016/j.cmet.2013.08.001

54. De Bock K, Georgiadou M, Schoors S, et al. Role of PFKFB3-driven glycolysis in vessel sprouting. Cell. 2013; 154(3): 651-663. doi: 10.1016/j.cell.2013.06.037

55. Potente M, Gerhardt H, Carmeliet P. Basic and therapeutic aspects of angiogenesis. Cell. 2011; 146(6): 873-887. doi: 10.1016/j.cell.2011.08.039

56. Sonveaux P, Copetti T, De Saedeleer CJ, et al. Targeting the lactate transporter MCT1 in endothelial cells inhibits lactate-induced HIF-1 activation and tumor angiogenesis. PLoS One. 2012; 7(3): e33418g. doi: 10.1371/journal.pone.0033418

57. Hunt TK, Aslam RS, Beckert S, et al. Aerobically derived lactate stimulates revascularization and tissue repair via redox mechanisms. Antioxidants & redox signaling. 2007; 9(8): 1115-1124. doi: 10.1089/ars.2007.1674

58. Vegran F, Boidot R, Michiels C, Sonveaux P, Feron O. Lactate influx through the endothelial cell monocarboxylate transporter MCT1 supports an NF-kappaB/IL-8 pathway that drives tumor angiogenesis. Cancer Research. 2011; 71(7): 2550-2560. doi: 10.1158/0008-5472.CAN-10-2828

59. Hao Q, Wang L, Tang H. Vascular endothelial growth factor induces protein kinase D-dependent production of proinflammatory cytokines in endothelial cells. American journal of physiology Cell physiology. 2009; 296(4): C821-C827. doi: 10.1152/ajpcell.00504.2008

60. Wright GL, Maroulakou IG, Eldridge J, et al. VEGF stimulation of mitochondrial biogenesis: requirement of AKT3 kinase. FASEB journal : official publication of the Federation of American Societies for Experimental Biology. 2008; 22(9): 3264-3275. doi: 10.1096/fj.08-106468

61. Elmasri H, Karaaslan C, Teper Y, et al. Fatty acid binding protein 4 is a target of VEGF and a regulator of cell proliferation in endothelial cells. FASEB journal : official publication of the Federation of American Societies for Experimental Biology. 2009; 23(11): 3865-3873. doi: 10.1096/fj.09-134882

62. Hagberg CE, Mehlem A, Falkevall A, et al. Targeting VEGF-B as a novel treatment for insulin resistance and type 2 diabetes. Nature. 2012; 490(7420): 426-430. doi: 10.1038/nature11464

63. Xu X, Ye L, Araki K, Ahmed R. mTOR, linking metabolism and immunity. Seminars in immunology. 2012; 24(6): 429-435. doi: 10.1016/j.smim.2012.12.005

64. Thomson AW, Turnquist HR, Raimondi G. Immunoregulatory functions of mTOR inhibition. Nature reviews Immunology. 2009; 9(5): 324-337. doi: 10.1038/nri2546

65. Buller CL, Loberg RD, Fan MH, et al. A GSK-3/TSC2/mTOR pathway regulates glucose uptake and GLUT1 glucose transporter expression. American journal of physiology Cell physiology. 2008; 295(3): C836-C843. doi: 10.1152/ajpcell.00554.2007

66. Duvel K, Yecies JL, Menon S, et al. Activation of a metabolic gene regulatory network downstream of mTOR complex 1. Molecular cell. 2010; 39(2): 171-183. doi: 10.1016/j.molcel.2010.06.022

67. Wang R, Dillon CP, Shi LZ, et al. The transcription factor Myc controls metabolic reprogramming upon T lymphocyte activation. Immunity. 2011; 35(6): 871-882. doi: 10.1016/j.immuni.2011.09.021

68. Araki K, Turner AP, Shaffer VO, et al. mTOR regulates memory CD8 T-cell differentiation. Nature. 2009; 460(7251): 108-112. doi: 10.1038/nature08155

69. Selak MA, Armour SM, MacKenzie ED, et al. Succinate links TCA cycle dysfunction to oncogenesis by inhibiting HIF-alpha prolyl hydroxylase. Cancer Cell. 2005; 7(1): 77-85. doi: 10.1016/j.ccr.2004.11.022

70. Pistollato F, Abbadi S, Rampazzo E, et al. Hypoxia and succinate antagonize 2-deoxyglucose effects on glioblastoma. Biochemical pharmacology. 2010; 80(10): 1517-1527. doi: 10.1016/j.bcp.2010.08.003

71. Tannahill GM, Curtis AM, Adamik J, et al. Succinate is an inflammatory signal that induces IL-1beta through HIF-1alpha. Nature. 2013; 496(7444): 238-242. doi: 10.1038/nature11986

72. Qi QR, Yang ZM. Regulation and function of signal transducer and activator of transcription 3. World journal of biological chemistry. 2014; 5(2): 231-239. doi: 10.4331/wjbc.v5.i2.231

73. Haigis MC, Deng CX, Finley LW, Kim HS, Gius D. SIRT3 is a mitochondrial tumor suppressor: a scientific tale that connects aberrant cellular ROS, the Warburg effect, and carcinogenesis. Cancer research. 2012; 72(10): 2468-2472. doi: 10.1158/0008-5472.CAN-11-3633

74. Tateno T, Asa SL, Zheng L, Mayr T, Ullrich A, Ezzat S. The FGFR4-G388R polymorphism promotes mitochondrial STAT3 serine phosphorylation to facilitate pituitary growth hormone cell tumorigenesis. PLoS genetics. 2011; 7(12): e1002400. doi: 10.1371/journal.pgen.1002400

75. Gough DJ, Corlett A, Schlessinger K, Wegrzyn J, Larner AC, Levy DE. Mitochondrial STAT3 supports Ras-dependent oncogenic transformation. Science. 2009; 324(5935): 1713-1716. doi: 10.1126/science.1171721

76. Gough DJ, Marie IJ, Lobry C, Aifantis I, Levy DE. STAT3 supports experimental K-RasG12D-induced murine myeloproliferative neoplasms dependent on serine phosphorylation. Blood. 2014; 22. doi: 10.1182/blood-2013-02-484196

77. Henderson B, Bitensky L, Chayen J. Glycolytic activity in human synovial lining cells in rheumatoid arthritis. Annals of the rheumatic diseases. 1979; 38(1): 63-67. doi: 10.1136/ard.38.1.63

78. Chang X, Wei C. Glycolysis and rheumatoid arthritis. International journal of rheumatic diseases. 2011; 14(3): 217-222. doi: 10.1111/j.1756-185X.2011.01598.x

79. Ciurtin C, Cojocaru VM, Miron IM, et al. Correlation between different components of synovial fluid and pathogenesis of rheumatic diseases. Romanian journal of internal medicine = Revue roumaine de medecine interne. 2006; 44(2): 171-181.

80. Biniecka M, Fox E, Gao W, et al. Hypoxia induces mitochondrial mutagenesis and dysfunction in inflammatory arthritis. Arthritis and rheumatism. 2011; 63(8): 2172-2182. doi: 10.1002/art.30395

81. Moran EM, Heydrich R, Ng CT, et al. IL-17A expression is localised to both mononuclear and polymorphonuclear synovial cell infiltrates. PLoS One. 2011; 6(8): e24048. doi: 10.1371/journal.pone.0024048

82. Biniecka M, Kennedy A, Ng CT, et al. Successful tumour necrosis factor (TNF) blocking therapy suppresses oxidative stress and hypoxia-induced mitochondrial mutagenesis in inflammatory arthritis. Arthritis research & therapy. 2011; 13(4): R121. doi: 10.1186/ar3424

83. Kennedy A, Ng CT, Chang TC, et al. Tumor necrosis factor blocking therapy alters joint inflammation and hypoxia. Arthritis and rheumatism. 2011; 63(4): 923-932. doi: 10.1002/art.30221

84. Ng CT, Biniecka M, Kennedy A, et al. Synovial tissue hypoxia and inflammation in vivo. Annals of the rheumatic diseases. 2010; 69(7): 1389-1395. doi: 10.1136/ard.2009.119776

85. Hollander AP, Corke KP, Freemont AJ, Lewis CE. Expression of hypoxia-inducible factor 1alpha by macrophages in the rheumatoid synovium: implications for targeting of therapeutic genes to the inflamed joint. Arthritis and rheumatism. 2001; 44(7): 1540-1544. doi: 10.1002/1529-0131(200107)44:7<1540::AID-ART277>3.0.CO;2-7

86. Giatromanolaki A, Sivridis E, Maltezos E, et al. Upregulated hypoxia inducible factor-1alpha and -2alpha pathway in rheumatoid arthritis and osteoarthritis. Arthritis research & therapy. 2003; 5(4): R193-R201. doi: 10.1186/ar756

87. Mobasheri A, Richardson S, Mobasheri R, Shakibaei M, Hoyland JA. Hypoxia inducible factor-1 and facilitative glucose transporters GLUT1 and GLUT3: putative molecular components of the oxygen and glucose sensing apparatus in articular chondrocytes. Histology and histopathology. 2005; 20(4): 1327-1338. doi: 10.14670/hh-20.1327

88. Gaber T, Dziurla R, Tripmacher R, Burmester GR, Buttgereit F. Hypoxia inducible factor (HIF) in rheumatology: low O2! See what HIF can do! Annals of the rheumatic diseases. 2005; 64(7): 971-980. doi: 10.1136/ard.2004.031641

89. Bodamyali T, Stevens CR, Billingham ME, Ohta S, Blake DR. Influence of hypoxia in inflammatory synovitis. Annals of the rheumatic diseases. 1998; 57(12): 703-710. doi: 10.1136/ard.57.12.703

90. Distler JH, Wenger RH, Gassmann M, et al. Physiologic responses to hypoxia and implications for hypoxia-inducible factors in the pathogenesis of rheumatoid arthritis. Arthritis and rheumatism. 2004; 50(1): 10-23. doi: 10.1002/art.11425

91. Taylor PC, Sivakumar B. Hypoxia and angiogenesis in rheumatoid arthritis. Current opinion in rheumatology. 2005; 17(3): 293-298. doi: 10.1097/01.bor.0000155361.83990.5b

92. Firestein GS, Echeverri F, Yeo M, Zvaifler NJ, Green DR. Somatic mutations in the p53 tumor suppressor gene in rheumatoid arthritis synovium. Proceedings of the National Academy of Sciences of the United States of America. 1997; 94(20): 10895-10900. doi: 10.1073/pnas.94.20.10895

93. Reme T, Travaglio A, Gueydon E, Adla L, Jorgensen C, Sany J. Mutations of the p53 tumour suppressor gene in erosive rheumatoid synovial tissue. Clinical and experimental immunology. 1998; 111(2): 353-358. doi: 10.1046/j.1365-2249.1998.00508.x

94. Kullmann F, Judex M, Neudecker I, et al. Analysis of the p53 tumor suppressor gene in rheumatoid arthritis synovial fibroblasts. Arthritis and rheumatism. 1999; 42(8): 1594-1600. doi: 10.1002/1529-0131(199908)42:8<1594::AID-ANR5>3.0.CO;2-#

95. Kawauchi K, Araki K, Tobiume K, Tanaka N. Loss of p53 enhances catalytic activity of IKKbeta through O-linked beta-N-acetyl glucosamine modification. Proceedings of the National Academy of Sciences of the United States of America. 2009; 106(9): 3431-3436. doi: 10.1073/pnas.0813210106

96. Biniecka M, Connolly M, Gao W, et al. Redox mediated angiogenesis in the hypoxic joint of inflammatory arthritis. Arthritis & rheumatology (Hoboken, NJ). 2014; 66(12): 3300-3310. doi: 10.1002/art.38822

97. Yang M, Guo M, Hu Y, Jiang Y. Scube regulates synovial angiogenesis-related signaling. Medical hypotheses. 2013; 81(5): 948-953. doi: 10.1016/j.mehy.2013.09.001

98. Bailey SM, Udoh US, Young ME. Circadian regulation of metabolism. The Journal of endocrinology. 2014; 222(2): R75-R96. doi: 10.1530/JOE-14-0200

99. Langmesser S, Albrecht U. Life time-circadian clocks, mitochondria and metabolism. Chronobiology international. 2006; 23(1-2): 151-157. doi: 10.1080/07420520500464437

100. Nikonova EV, Vijayasarathy C, Zhang L, et al. Differences in activity of cytochrome C oxidase in brain between sleep and wakefulness. Sleep. 2005; 28(1): 21-27. doi: 10.1093/sleep/28.1.21

101. Jordan SD, Lamia KA. AMPK at the crossroads of circadian clocks and metabolism. Molecular and cellular endocrinology. 2013; 366(2): 163-169. doi: 10.1016/j.mce.2012.06.017

102. Sancar A. Regulation of the mammalian circadian clock by cryptochrome. The Journal of biological chemistry. 2004; 279(33): 34079-34082. doi: 10.1074/jbc.R400016200

103. Busino L, Bassermann F, Maiolica A, et al. SCFFbxl3 controls the oscillation of the circadian clock by directing the degradation of cryptochrome proteins. Science. 2007; 316(5826): 900-904. doi: 10.1126/science.1141194

104. Lamia KA, Sachdeva UM, DiTacchio L, et al. AMPK regulates the circadian clock by cryptochrome phosphorylation and degradation. Science. 2009; 326(5951): 437-440. doi: 10.1126/science.1172156

105. Etchegaray JP, Machida KK, Noton E, et al. Casein kinase 1 delta regulates the pace of the mammalian circadian clock. Molecular and cellular biology. 2009; 29(14): 3853-3866. doi: 10.1128/MCB.00338-09

106. Um JH, Yang S, Yamazaki S, et al. Activation of 5’-AMP-activated kinase with diabetes drug metformin induces casein kinase Iepsilon (CKIepsilon)-dependent degradation of clock protein mPer2. J Biol Chem. 2007; 282(29): 20794-20798. doi: 10.1074/jbc.C700070200

107. Vieira E, Nilsson EC, Nerstedt A, et al. Relationship between AMPK and the transcriptional balance of clock-related genes in skeletal muscle. American journal of physiology Endocrinology and metabolism. 2008; 295(5): E1032-E1037. doi: 10.1152/ajpendo.90510.2008

108. Canto C, Gerhart-Hines Z, Feige JN, et al. AMPK regulates energy expenditure by modulating NAD+ metabolism and SIRT1 activity. Nature. 2009; 458(7241): 1056-1060. doi: 10.1038/nature07813

109. Lan F, Cacicedo JM, Ruderman N, Ido Y. SIRT1 modulation of the acetylation status, cytosolic localization, and activity of LKB1. Possible role in AMP-activated protein kinase activation. The Journal of biological chemistry. 2008; 283(41): 27628-27635. doi: 10.1074/jbc.M805711200

110. Walker JW, Jijon HB, Madsen KL. AMP-activated protein kinase is a positive regulator of poly(ADP-ribose) polymerase. Biochemical and biophysical research communications. 2006; 342(1): 336-341. doi: 10.1016/j.bbrc.2006.01.145

111. Fulco M, Cen Y, Zhao P, et al. Glucose restriction inhibits skeletal myoblast differentiation by activating SIRT1 through AMPK-mediated regulation of Nampt. Developmental cell. 2008; 14(5): 661-673. doi: 10.1016/j.devcel.2008.02.004

112. Zhang X, Xu L, Shen J, et al. Metabolic signatures of esophageal cancer: NMR-based metabolomics and UHPLC-based focused metabolomics of blood serum. Biochimica et biophysica acta. 2013; 1832(8): 1207-1216. doi: 10.1016/j.bbadis.2013.03.009

113. Abbassi-Ghadi N, Kumar S, Huang J, Goldin R, Takats Z, Hanna GB. Metabolomic profiling of oesophago-gastric cancer: a systematic review. European journal of cancer. 2013; 49(17): 3625-3637. doi: 10.1016/j.ejca.2013.07.004

114. Ussakli CH, Ebaee A, Binkley J, et al. Mitochondria and tumor progression in ulcerative colitis. Journal of the National Cancer Institute. 2013; 105(16): 1239-1248. doi: 10.1093/jnci/djt167

115. Gruno M, Peet N, Tein A, et al. Atrophic gastritis: deficient complex I of the respiratory chain in the mitochondria of corpus mucosal cells. Journal of gastroenterology. 2008; 43(10): 780-788. doi: 10.1007/s00535-008-2231-4

116. O’Sullivan KE, Phelan JJ, O’Hanlon C, Lysaght J, O’Sullivan JN, Reynolds JV. The role of inflammation in cancer of the esophagus. Expert review of gastroenterology & hepatology. 2014; 8(7): 749-760. doi: 10.1586/17474124.2014.913478

117. Phelan JJ, MacCarthy F, Feighery R, et al. Differential expression of mitochondrial energy metabolism profiles across the metaplasia-dysplasia-adenocarcinoma disease sequence in Barrett’s oesophagus. Cancer Letters. 2014; 354(1): 122-131. doi: 10.1016/j.canlet.2014.07.035

118. Ando M, Uehara I, Kogure K, et al. Interleukin 6 enhances glycolysis through expression of the glycolytic enzymes hexokinase 2 and 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase-3. Journal of Nippon Medical School = Nippon Ika Daigaku zasshi. 2010; 77(2): 97-105. doi: 10.1272/jnms.77.97

119. Catarzi S, Favilli F, Romagnoli C, et al. Oxidative state and IL-6 production in intestinal myofibroblasts of Crohn’s disease patients. Inflammatory bowel diseases. 2011; 17(8): 1674-1684. doi: 10.1002/ibd.21552

120. Dvorakova K, Payne CM, Ramsey L, et al. Increased expression and secretion of interleukin-6 in patients with Barrett’s esophagus. Clinical cancer research : an official journal of the American Association for Cancer Research. 2004; 10(6): 2020-2028. doi: 10.1158/1078-0432.CCR-0437-03

121. Kastelein F, Biermann K, Steyerberg EW, et al. Aberrant p53 protein expression is associated with an increased risk of neoplastic progression in patients with Barrett’s oesophagus. Gut. 2013; 62(12): 1676-1683. doi: 10.1136/gutjnl-2012-30359

122. Angelo LS, Talpaz M, Kurzrock R. Autocrine interleukin-6 production in renal cell carcinoma: evidence for the involvement of p53. Cancer research. 2002; 62(3): 932-940.

123. Bar F, Bochmann W, Widok A, et al. Mitochondrial gene polymorphisms that protect mice from colitis. Gastroenterology. 2013; 145(5): 1055e3-1063e3. doi: 10.1053/j.gastro.2013.07.015

124. Gruno M, Peet N, Seppet E, et al. Oxidative phosphorylation and its coupling to mitochondrial creatine and adenylate kinases in human gastric mucosa. American journal of physiology Regulatory, integrative and comparative physiology. 2006; 291(4): R936-R946. doi: 10.1152/ajpregu.00162.2006

125. Puurand M, Peet N, Piirsoo A, et al. Deficiency of the complex I of the mitochondrial respiratory chain but improved adenylate control over succinate-dependent respiration are human gastric cancer-specific phenomena. Molecular and cellular biochemistry. 2012; 370(1-2): 69-78. doi: 10.1007/s11010-012-1399-3

126. Chan AW, Gill RS, Schiller D, Sawyer MB. Potential role of metabolomics in diagnosis and surveillance of gastric cancer. World journal of gastroenterology : WJG. 2014; 20(36): 12874-12882. doi: 10.3748/wjg.v20.i36.12874

127. Li H, Wang J, Xu H, et al. Decreased fructose-1,6-bisphosphatase-2 expression promotes glycolysis and growth in gastric cancer cells. Molecular cancer. 2013; 12(1): 110. doi: 10.1186/1476-4598-12-110

128. Liu X, Wang X, Zhang J, et al. Warburg effect revisited: an epigenetic link between glycolysis and gastric carcinogenesis. Oncogene. 2010; 29(3): 442-450. doi: 10.1038/onc.2009.332

129. Kwon OH, Kang TW, Kim JH, et al. Pyruvate kinase M2 promotes the growth of gastric cancer cells via regulation of Bcl-xL expression at transcriptional level. Biochemical and biophysical research communications. 2012; 423(1): 38-44. doi: 10.1016/j.bbrc.2012.05.063

130. Zhou CF, Li XB, Sun H, et al. Pyruvate kinase type M2 is upregulated in colorectal cancer and promotes proliferation and migration of colon cancer cells. IUBMB life. 2012; 64(9): 775-782. doi: 10.1002/iub.1066

131. Hur H, Xuan Y, Kim YB, et al. Expression of pyruvate dehydrogenase kinase-1 in gastric cancer as a potential therapeutic target. International journal of oncology. 2013; 42(1): 44-54. doi: 10.3892/ijo.2012.1687

132. Giatromanolaki A, Sivridis E, Maltezos E, et al. Hypoxia inducible factor 1alpha and 2alpha overexpression in inflammatory bowel disease. Journal of clinical pathology. 2003; 56(3): 209-213. doi: 10.1136/jcp.56.3.209

133. Vermeulen N, Vermeire S, Arijs I, et al. Seroreactivity against glycolytic enzymes in inflammatory bowel disease. Inflammatory bowel diseases. 2011; 17(2): 557-564. doi: 10.1002/ibd.21388

134. Bobarykina AY, Minchenko DO, Opentanova IL, et al. Hypoxic regulation of PFKFB-3 and PFKFB-4 gene expression in gastric and pancreatic cancer cell lines and expression of PFKFB genes in gastric cancers. Acta biochimica Polonica. 2006; 53(4): 789-799.

135. Tong M, McHardy I, Ruegger P, et al. Reprograming of gut microbiome energy metabolism by the FUT2 Crohn’s disease risk polymorphism. The ISME journal. 2014; 8(11): 2193-2206. doi: 10.1038/ismej.2014.64

136. Engelman JA, Chen L, Tan X, et al. Effective use of PI3K and MEK inhibitors to treat mutant Kras G12D and PIK3CA H1047R murine lung cancers. Nature medicine. 2008; 14(12): 1351-1356. doi: 10.1038/nm.1890

137. Libby G, Donnelly LA, Donnan PT, Alessi DR, Morris AD, Evans JM. New users of metformin are at low risk of incident cancer: a cohort study among people with type 2 diabetes. Diabetes Care. 2009; 32(9): 1620-1625. doi: 10.2337/dc08-2175

138. Hawley SA, Fullerton MD, Ross FA, et al. The ancient drug salicylate directly activates AMP-activated protein kinase. Science. 2012; 336(6083): 918-922. doi: 10.1126/science.1215327

139. Beckers A, Organe S, Timmermans L, et al. Methotrexate enhances the antianabolic and antiproliferative effects of 5-aminoimidazole-4-carboxamide riboside. Molecular cancer therapeutics. 2006; 5(9): 2211-2217. doi: 10.1158/1535-7163.MCT-06-0001

140. Xuan Y, Hur H, Ham IH, et al. Dichloroacetate attenuates hypoxia-induced resistance to 5-fluorouracil in gastric cancer through the regulation of glucose metabolism. Experimental cell research. 2014; 321(2): 219-230. doi: 10.1016/j.yexcr.2013.12.009

141. Maschek G, Savaraj N, Priebe W, et al. 2-deoxy-D-glucose increases the efficacy of adriamycin and paclitaxel in human osteosarcoma and non-small cell lung cancers in vivo. Cancer research. 2004; 64(1): 31-34. doi: 10.1158/0008-5472.CAN-03-3294

142. Singh D, Banerji AK, Dwarakanath BS, et al. Optimizing cancer radiotherapy with 2-deoxy-d-glucose dose escalation studies in patients with glioblastoma multiforme. Strahlentherapie und Onkologie : Organ der Deutschen Rontgengesellschaft [et al]. 2005; 181(8): 507-514. doi: 10.1007/s00066-005-1320-z

143. Nelson JA, Falk RE. Phloridzin and phloretin inhibition of 2-deoxy-D-glucose uptake by tumor cells in vitro and in vivo. Anticancer research. 1993; 13(6A): 2293-2299.

144. Nelson JA, Falk RE. The efficacy of phloridzin and phloretin on tumor cell growth. Anticancer research. 1993; 13(6A): 2287-2292.

145. Clem B, Telang S, Clem A, et al. Small-molecule inhibition of 6-phosphofructo-2-kinase activity suppresses glycolytic flux and tumor growth. Molecular cancer therapeutics. 2008; 7(1): 110-120. doi: 10.1158/1535-7163.MCT-07-0482

146. Nebeling LC, Miraldi F, Shurin SB, Lerner E. Effects of a ketogenic diet on tumor metabolism and nutritional status in pediatric oncology patients: Two case reports. J Am Coll Nutr. 1995; 14(2): 202-208. doi: 10.1080/07315724.1995.10718495

147. Liu AM, Xu Z, Shek FH, et al. miR-122 targets pyruvate kinase M2 and affects metabolism of hepatocellular carcinoma. PLoS One. 2014; 9(1): e86872. doi: 10.1371/journal.pone.0086872

LATEST ARTICLES

Unraveling the Mysteries of Type-A Aortic Dissection Using POCUS/Echocardiography

Syeda Rukh*, Sathyanarayana Machani and Milind Awale

doi.

Blood Sample from the Patient

Hypertriglyceridemia-Induced Pancreatitis: A Case Report and Literature Review

Maarten Bulterys, Melvin Willems* and Agnes Meersman

doi.

From Neck Pain to a Life-Threatening Condition: A Case Report

Floris Vandewoude* and Sören Verstraete

doi.

LATEST ARTICLES