GET THE APP

The Role of Receptors for Advanced Glycation End Product in Pancr
Pancreatic Disorders & Therapy

Pancreatic Disorders & Therapy
Open Access

ISSN: 2165-7092

+44 1478 350008

Review Article - (2016) Volume 6, Issue 1

The Role of Receptors for Advanced Glycation End Product in Pancreatic Carcinogenesis

Ali Hafez El-Far*
Department of Biochemistry, Faculty of Veterinary Medicine, Damanhour University, Egypt
*Corresponding Author: Ali Hafez El-Far, Department of Biochemistry, Faculty of Veterinary Medicine, Damanhour University, Egypt, Tel: +20452167080 Email:

Abstract

Pancreatic cancer is the fourth-leading cause of cancer deaths worldwide that considered as the malignant tumor with the poorest prognosis and the lowest survival rate. This extremely violent disease is rarely diagnosed at an early level and difficult to treat due to its resistance to radiation therapy and chemotherapy. Here, we show that the receptor for advanced glycation end products (RAGE) and its ligands, advanced glycation end products (AGEs), high-mobility group box 1 (HMGB1) and S100 protein family are required for pancreatic cancer development, angiogenesis and metastasis through up-regulation of some anti-apoptotic molecules as matrix metalloproteinase -9 (MMP-9), kinase insert domain receptor (KDR), vascular endothelial growth factor (VEGF), platelet-derived growth factor-B (PDGF-B), hypoxia-inducible factor 1 (HIF1α), signal transducer and activator of transcription 3 (pSTAT3) and nuclear factor kappa B (NF-κB). In addition to the decrease in reactive oxygen species (ROS) those favouring the cancer cell growth and metastasis. The role of RAGE in pancreatic carcinogenesis needs further studies to investigate the relationship of RAGE with other anti-apoptotic and apoptotic molecules and examine the therapeutic potentials of anti-RAGE drugs for pancreatic cancer.

<

Keywords: Pancreatic cancer, RAGE, RAGE ligands

Introduction

Pancreatic cancer is a deadly disease, characterized by late diagnosis, early metastasis and chemotherapy resistant [1]. Pancreatic ductal adenocarcinoma (PDA) is a complex disease that arises from the genetic alterations of KRAS, BRCA1, SMAD4, CDKN2A/p16 and TP53 [2]. Activating KRAS mutations and p16 inactivation are genetic abnormalities, most frequently detected in PDA [3]. Oncogenic activation of the KRAS gene occurs in more than 90% of PDA and plays a critical role in PDA malignancy [4]. The mean survival is around six months, and the five years overall survival rate is less than 4% of the patients [5]. PDA is a highly mortal disease. It accounts for only 3% of cancer cases each year but is currently the fourth common cause of cancer mortality [6] due to advanced stage at diagnosis and poor response to current treatment [7]. By 2030, PDA is expected to be the 2nd leading cause of cancer death [8]. The best chance for survival is early detection when the tumor can be treated with surgical aid [9].

RAGE and its ligands

The cell cycle is strictly regulated and controlled by a complex network of cellular signalling pathways [10]. RAGE belongs to the immunoglobulin superfamily of receptors [11] with a molecular weight of about 55-kDa protein, has an extracellular part consisting of a variable (V) immunoglobulin-like domain followed by two constant domains, a single transmembrane domain and a cytosolic tail [12,13]. The N-terminus of the V domain is the ligand-binding site, and the cytosolic tail that necessary for RAGE-induced intracellular signalling [14]. Soluble RAGE (sRAGE) forms lack both the cytosolic and the transmembrane domains [14,15]. sRAGE can be distinguished in circulating blood [16] that can bind with RAGE ligands such as advanced glycation end product (AGE) [12].

RAGE is expressed in a variety of tissues as heart, lung, skeletal muscle and vessel wall. AGE-RAGE interactions inhibit the prostacyclin production by human endothelial cells, which prompt angiogenesis and thrombogenesis [17]. It is generally believed that RAGE-induced pathways are implicated in the development of various diseases [18,19]. RAGE plays a crucial role in numerous diseases including diabetes, inflammation, and cancer [20]. RAGE interacts with diverse ligands, including AGEs and β-amyloid fibrils, S100 protein family (S100B, S100P, S100A4, S100A6, S100A8, S100A9, S100A11, S100A12, and S100A13), high mobility group box-1 (HMGB1), and prions [21,22].

AGEs formed by a non-enzymatic reaction between the ketone and aldehyde groups of sugars and the amino groups of proteins, which called Maillard reaction that, have been concerned in aging and diabetes-related pathological complications [23]. These reactions can be triggered by glucose-6-phosphate, glyceraldehyde-3-phosphate, glyoxal (GO), methylglyoxal (MGO) and 3-deoxyglucosone (3DG) [24]. AGE and RAGE have been associated in cancer development [25]. AGEs binding to RAGE induces signalling pathways of mitogen-activated protein kinases (MAPKs) [26], cdc42/rac and Jak/STAT [22], which modulate the of some genes as the vascular endothelial growth factor (VEGF) [27]. AGE-RAGE activation increases transforming growth factor beta-1 (TGF-β1) levels, with enhanced activity of matrix metalloproteinase 2 (MMP-2); on the other hand, RAGE signalling promotes MMP-9 activity. MMP-2 and -9 induce modifications in collagen IV turnover [28]. AGEs have been shown to induce inflammation and intracellular reactive oxygen species (ROS), which leads to the expression of many atherosclerosis-related genes, including VEGF [29].

HMGB1 protein is a DNA-binding nuclear protein, released actively in response to cytokine stimulation, or passively during cell death [30], and it is present in almost all eukaryotic cells [31]. HMGB1 can activate a series of signalling components, including MAPKs and AKT, which play an important role in tumor growth through binding to RAGE and hastens cell-cycle progression [32]. In vitro studies with pancreatic cancer cells revealed that the targeted knockout or inhibition of HMGB1 and RAGE could increase apoptosis and defeat pancreatic cancer cell growth [33]. This phenomenon has been also noticed with other types of cancer cells [34].

S100 proteins are expressed in vertebrates, and exhibit somewhat cell-specific distribution [35]. S100P is a 95-amino acid member of the S100 family of protein, purified and characterized from placenta [36]. The term “S100” was coined to indicate a group of proteins soluble in a 100% saturated ammonium sulphate solution [37]. The designation “P” was coined to indicate that it was purified firstly from placenta [36]. S100P is a member of the large family of S100 calcium-binding proteins [38]. Expression of S100P was observed to be specific to pancreatic cancer cells [39]. The specificity for cancer cells was further confirmed in micro-dissected pancreatic cancer tissues and isolated primary cultures of cancer cells [40].

RAGE and Cancer

RAGE has been overexpressed in the brain, breast, colon, colorectal, lung, prostate, oral squamous cell, and ovarian cancers, in addition to lymphoma and melanoma [41] this clarifies the direct relationship between RAGE and cancer cell proliferation, survival, migration, and invasion of tumor cells [42]. Targeted knockdown of RAGE in the tumor cell, leads to increased apoptosis. In contrast, overexpression of RAGE is associated with enhanced autophagy and stop apoptosis [33]. The HMGB1-RAGE axis blockage suppressed tumor growth and explores the in vivo role of RAGE during cancer development [43]. RAGE and its ligand, S100P have been shown to mediate tumor growth, drug resistance, and metastasis [44]. RAGE activation has been considered to promote tumor vasculature, tumor growth and invasion [45] by induction of ROS, extracellular signal-regulated protein kinase (ERK1/2), p38MAPKs, phosphoinositol-J kinase, Janus kinase/signal transducer and activator of transcription pathway, nuclear factor kappa B (NF-κB), activator protein-1 (AP-1) and activator of transcription 3 (STAT-3) [22,46,47].

Interaction of full-length RAGE with its ligands, including AGEs, S100 protein family, and HMGB1, triggers the rapid activation of an array of key cell signalling pathways culminating in the activation of the NF-κB pathway [48]. NFκB is a transcription factor for a large group of genes which are involved in several different pathways. For instance, NFκB activates its own inhibitor (IκB) as well as groups of pro-apoptotic and anti-apoptotic genes [49]. Among the latter, NFκB activates transcription of a gene encoding for the inhibitor of apoptosis protein (IAP). This protein, in turn, contributes to downregulate the activity of the caspase cascade [50]. NF-κB is retained inert in the cytoplasm by the inhibitor protein, I-kappaB (IκB) [51]. Following stimulation of the cell by a variety of agents, IκB is degraded, allowing NF-κB to translocate to the nucleus and bind to the promoter regions of its multiple target genes to promote cell survival and proliferation [52].

RAGE and Pancreatic Cancer

RAGE has a crucial role in pancreatic cancer development [53,54]. PDA is a wasteful disease with low survival rates [55]. Loss of RAGE function inhibited the development of PDA in mouse models [3]. The immunohistochemical analysis confirmed the expression of RAGE and its ligands S100P, S100A4, and HMGB-1 in human PDA [56]. RAGE and S100 protein play important roles in the progression of PDA [55]. Indeed, S100 proteins interact with RAGE, which play a role in the degradation of the extracellular matrix facilitating the metastasis of pancreatic cancer [57]. AGE and RAGE are expressed in many tissues and cell types [58].

Induction of S phase

S100P expressed in more than 90% of pancreatic tumors, leading to tumor growth and invasion [59] as in Panc-1 and Mpanc96 cells [60] in which the increased S100P levels increased the growth of tumors in mice with subcutaneous-implanted Panc-1 cells. FACS analysis indicated an over 80% increase in the number of cells in S-phase cells in Panc-1 cells expressing S100P that interact with RAGE promoting Panc-1 cell migration and invasion [61]. This role of RAGE in pancreatic carcinogenesis was evidenced by incubation of wild-type BxPC3 cells with the anti-RAGE antibodies that inhibited the cell growth, migration, and invasion [43].

Induction of angiogenesis and metastasis

S100A4 synergize with VEGF, via the RAGE receptor, in promoting endothelial cell migration by increasing kinase insert domain receptor (KDR) expression and MMP-9 activity [62]. The expression of MMPs correlates with the extracellular matrix degradation and tumor metastasis [63]. The expression of MMP-9 is associated with metastasis of many human cancers because they play an important role in the degradation of type IV collagen, which is a major component of the basement membrane [64]. Therefore, MMP-9 may be involved in the process of cancer metastasis [65]. Meanwhile, VEGF is a signal protein produced by cells that stimulate vasculogenesis and angiogenesis that function as oxygen supply to tissues [66]. The role of angiogenesis in supporting tumor growth and metastasis [67].

AGE ligand–receptor interactions could play an active part in the progression of human pancreatic cancer cells (Mia PaCa-2) through the induction of autocrine platelet-derived growth factor-B (PDGF-B) [68]. Additionally, RAGE binds to oncogenic KRAS facilitates hypoxia-inducible factor-1 (HIF1α) activation and promotes pancreatic tumor growth [69]. HIF-1 potentiates the expression of proteins that promote angiogenesis and cell survival [70]. HIF1α is a transcription factor induced by low oxygen conditions and involved in the activation cancer cell angiogenesis and metastasis [71]. Increased HIF-1 activity promotes tumor progression, and inhibition of HIF-1 could represent a novel approach to cancer therapy [72].

Counteract the ROS-induced oxidative stress

Activation of RAGE by S100P stimulates several cellular signalling pathways, including MAP kinase and NFκB [44] inhibiting S100P-RAGE interactions significantly reduce basal levels of NFκB activity in PDA [59]. Excessive ROS production can lead to oxidation of macromolecules and has been implicated in mitochondrial DNA (mtDNA) mutations, aging, and cell death. Mitochondrial generated ROS play an important role in the release of proapoptotic proteins, which can trigger caspase activation and apoptosis [73]. Exposure of pancreatic tumor cells to H2O2 provoked a nuclear factor kappa B (NF-κB)-dependent increase in RAGE expression [74] that decreases ROS-induced oxidative injury [20].

Mitochondrial ATP production

Tumor cells have increased energy requirements. ATP production occurs through glycolysis and oxidative phosphorylation within mitochondria. The STAT3 has a role in mitochondrial function, regulating complex I activation and ATP production [75], promoting the oncogenic property of Ras-mediated cellular transformation [76]. RAGE expression appears to enhance the cancer cell survival that depends on autophagy, autocrine IL-6 production, and IL-6–promoted mitochondrial STAT3 phosphorylation and localization, which in turn promotes enhanced ATP production in pancreatic cancer cells [1].

RAGE and HMGB1 coordinately enhanced tumor cell mitochondrial complex I activity, ATP production that promote tumor cell proliferation and migration. Lack of RAGE or inhibition of HMGB1 diminished ATP production and slowed tumor growth in vitro and in vivo [69]. The role of HMGB1 in pancreatic tumor survival might involve alterations in tumor bioenergetics [69]. RAGE-mediated autophagy is required for IL-6-induced mitochondrial translocation of STAT3 and subsequently, IL-6/STAT3-mediated ATP production [1]. There are at least two different mechanisms involved in RAGE-mediated ATP production: mitRAGE dependent and RAGE-mediated autophagy dependent [69].

Down-regulation of apoptotic molecules

RAGE knockdown was associated with increased apoptosis that reversed in part by treatment with pan-caspase inhibitors [33]. In cells treated with a pharmacologic inhibitor of p53, pifithrin α, and p53 knock out tumor cell lines we observed abrogation of the increased cell death observed with RAGE knockdown [77].

Conclusion

From this review, we can conclude that RAGE and its ligands have the crucial role in pancreatic carcinogenesis through up-regulation of some anti-apoptotic molecules. Besides, more studies have been in need to understand the mechanisms by which RAGE induces the pancreatic cancer development and survival by further studies to know the relationship between RAGE and other anti-apoptotic and apoptotic molecules by molecular and proteomic tool through:

Study the relationship between RAGE and apoptotic molecules such as p53, p21, caspases and BAX.

Study the relationship between RAGE and anti-apoptotic molecules like Akt, BCL2, PARP and inhibitors of apoptosis proteins.

Drug discovery of new novel anti-RAGE agents.

References

  1. Kang R, Loux T, Tang D, Schapiro NE, Vernon P, et al. (2012) The expression of the receptor for advanced glycationendproducts (RAGE) is permissive for early pancreatic neoplasia. ProcNatlAcadSci USA 109: 7031-7036.
  2. Kang R, Tang D, Lotze MT, Zeh HJ 3rd (2012) AGER/RAGE-mediated autophagy promotes pancreatic tumorigenesis and bioenergetics through the IL6-pSTAT3 pathway. Autophagy 8: 989-991.
  3. DiNorcia J, Lee MK, Moroziewicz DN, Winner M, Suman P, et al. (2012) RAGE gene deletion inhibits the development and progression of ductal neoplasia and prolongs survival in a murine model of pancreatic cancer. J GastrointestSurg 16: 104-112.
  4. Wang P, Sun YC, Lu WH, Huang P, et al. (2015) Selective killing of K-ras-transformed pancreatic cancer cells by targeting NAD(P)H oxidase. Chin J Cancer 34: 166-176.
  5. Jemal A, Siegel R, Ward E, Murray T, Xu J, et al. (2006) Cancer statistics, 2006. CA Cancer J Clin 56: 106-130.
  6. Poruk KE, Firpo MA, Adler DG, Mulvihill SJ (2013) Screening for pancreatic cancer: why, how, and who? Ann Surg257: 17-26.
  7. Kang R, Tang D (2012) Autophagy in pancreatic cancer pathogenesis and treatment. Am J Cancer Res 2: 383-396.
  8. Rahib L, Smith BD, Aizenberg R, Rosenzweig AB, Fleshman JM, et al. (2014) Projecting cancer incidence and deaths to 2030: the unexpected burden of thyroid, liver, and pancreas cancers in the United States. Cancer Res 74: 2913-2921.
  9. Siegel R, Ward E, Brawley O, Jemal A (2011) Cancer statistics, 2011: the impact of eliminating socioeconomic and racial disparities on premature cancer deaths. CA Cancer J Clin 61: 212-236.
  10. Franovic A, Holterman CE, Payette J, Lee S (2009) Human cancers converge at the HIF-2alpha oncogenic axis. ProcNatlAcadSci U S A 106: 21306-21311.
  11. Ramasamy R, Vannucci SJ, Yan SS, Herold K, Yan SF, et al. (2005) Advanced glycation end products and RAGE: a common thread in aging, diabetes, neurodegeneration, and inflammation. Glycobiology 15: 16R-28R.
  12. Raucci A, Cugusi S, Antonelli A, Barabino SM, Monti L, et al. (2008) A soluble form of the receptor for advanced glycationendproducts (RAGE) is produced by proteolytic cleavage of the membrane-bound form by the sheddase a disintegrin and metalloprotease 10 (ADAM10). FASEB J 22: 3716-3727.
  13. Sorci G, Riuzzi F, Giambanco I, Donato R (2013) RAGE in tissue homeostasis, repair and regeneration. BiochimBiophysActa 1833: 101-109.
  14. Kalea AZ, Schmidt AM, Hudson BI (2011) Alternative splicing of RAGE: roles in biology and disease. Front Biosci (Landmark Ed) 16: 2756-2770.
  15. Diamanti-Kandarakis E, Piperi C, Kalofoutis A, Creatsas G (2005) Increased levels of serum advanced glycation end-products in women with polycystic ovary syndrome. ClinEndocrinol (Oxf) 62: 37-43.
  16. Diamanti-Kandarakis E (2006) Insulin resistance in PCOS. Endocrine 30: 13-17.
  17. Yamagishi S, Yamamoto Y, Harada S, Hsu CC, and Yamamoto H (1996) Advanced glycosylation end products stimulate the growth but inhibit the prostacyclin-producing ability of endothelial cells through interactions with their receptors. FEBS Lett. 384: 103-6.
  18. Riehl A, Németh J, Angel P, Hess J (2009) The receptor RAGE: Bridging inflammation and cancer. Cell Commun Signal 7: 12.
  19. Riuzzi F, Sorci G, Sagheddu R, Donato R (2012) HMGB1-RAGE regulates muscle satellite cell homeostasis through p38-MAPK- and myogenin-dependent repression of Pax7 transcription. J Cell Sci 125: 1440-1454.
  20. Kang R, Tang D, Lotze MT, Zeh HJ (2011) RAGE regulates autophagy and apoptosis following oxidative injury. Autophagy 7: 442-444.
  21. Schmidt AM, Yan SD, Yan SF, Stern DM (2001) The multiligand receptor RAGE as a progression factor amplifying immune and inflammatory responses. J Clin Invest 108: 949-955.
  22. Bierhaus A, Humpert PM, Morcos M, Wendt T, Chavakis T, et al. (2005) Understanding RAGE, the receptor for advanced glycation end products. J Mol Med (Berl) 83: 876-886.
  23. Al-Abed Y, Kapurniotu A, Bucala R (1999) Advanced glycation end products: detection and reversal. Methods Enzymol 309: 152-172.
  24. Thornalley PJ (2005) Dicarbonyl intermediates in the maillard reaction. Ann N Y AcadSci 1043: 111-117.
  25. Grote VA, Nieters A, Kaaks R, Tjønneland A, Roswall N, et al. (2012) The associations of advanced glycation end products and its soluble receptor with pancreatic cancer risk: a case-control study within the prospective EPIC Cohort. Cancer Epidemiol Biomarkers Prev 21: 619-628.
  26. Schmidt AM, Yan SD, Wautier JL, Stern D (1999) Activation of receptor for advanced glycation end products: a mechanism for chronic vascular dysfunction in diabetic vasculopathy and atherosclerosis. Circ Res 84: 489-497.
  27. Goldin A, Beckman JA, Schmidt AM, Creager MA (2006) Advanced glycation end products: sparking the development of diabetic vascular injury. Circulation 114: 597-605.
  28. Serban AI, Stanca L, Geicu OI, Munteanu MC, Costache M (2015) Extracellular matrix is modulated in advanced glycation end products milieu via a RAGE receptor dependent pathway boosted by transforming growth factor-beta1 RAGE. J Diabetes 7: 114-24.
  29. Yamagishi S (2011) Role of advanced glycation end products (AGEs) and receptor for AGEs (RAGE) in vascular damage in diabetes. ExpGerontol 46: 217-224.
  30. Scaffidi P, Misteli T, Bianchi ME (2002) Release of chromatin protein HMGB1 by necrotic cells triggers inflammation. Nature 418: 191-195.
  31. Bonaldi T, Talamo F, Scaffidi P, Ferrera D, Porto A, et al. (2003) Monocytic cells hyperacetylate chromatin protein HMGB1 to redirect it towards secretion. EMBO J 22: 5551-5560.
  32. Sparvero LJ, Asafu-Adjei D, Kang R, Tang D, Amin N, et al. (2009) RAGE (Receptor for Advanced GlycationEndproducts), RAGE ligands, and their role in cancer and inflammation. J Transl Med 7: 17.
  33. Kang R, Tang D, Schapiro NE, Livesey KM, Farkas Aet al. (2010) The receptor for advanced glycation end products (RAGE) sustains autophagy and limits apoptosis, promoting pancreatic tumor cell survival. Cell Death Differ 17: 666-76.
  34. Ellerman JE, Brown CK, Vera M, Zeh HJ, Billiar T, et al. (2007) Masquerader: high mobility group box-1 and cancer. Clin Cancer Res 13: 2836-2848.
  35. Donato R (2001) S100: a multigenic family of calcium-modulated proteins of the EF-hand type with intracellular and extracellular functional roles. Int J Biochem Cell Biol 33: 637-668.
  36. Becker T, Gerke V, Kube E, Weber K (1992) S100P, a novel Ca(2+)-binding protein from human placenta. cDNA cloning, recombinant protein expression and Ca2+ binding properties. Eur J Biochem 207: 541-547.
  37. Moore BW (1965) A soluble protein characteristic of the nervous system. BiochemBiophys Res Commun 19: 739-744.
  38. Zhang H, Wang G, Ding Y, Wang Z, Barraclough R, et al. (2003) The crystal structure at 2A resolution of the Ca2+ -binding protein S100P. J MolBiol 325: 785-794.
  39. Logsdon CD, Simeone DM, Binkley C, Arumugam T, Greenson JKet al. (2003) Molecular profiling of pancreatic adenocarcinoma and chronic pancreatitis identifies multiple genes differentially regulated in pancreatic cancer. Cancer Res. 63: 2649-57.
  40. Ohuchida K, Mizumoto K, Egami T, Yamaguchi H, Fujii Ket al. (2006) S100P is an early developmental marker of pancreatic carcinogenesis. Clin Cancer Res. 12: 5411-6.
  41. Logsdon CD, Fuentes MK, Huang EH, and Arumugam T (2007) RAGE and RAGE ligands in cancer. CurrMol Med. 7: 777-89.
  42. Abe R, Yamagishi S (2008) AGE-RAGE system and carcinogenesis. Curr Pharm Des 14: 940-945.
  43. Taguchi A, Blood DC, del Toro G, Canet A, Lee DC et al. (2000) Blockade of RAGE-amphoterin signalling suppresses tumour growth and metastases. Nature 405: 354-60.
  44. Arumugam T and Logsdon CD (2011) S100P: a novel therapeutic target for cancer. Amino Acids. 41: 893-9.
  45. Rojas A, Figueroa H, Morales E (2010) Fueling inflammation at tumor microenvironment: the role of multiligand/RAGE axis. Carcinogenesis 31: 334-341.
  46. Buckley ST, Ehrhardt C (2010) The receptor for advanced glycation end products (RAGE) and the lung. J Biomed Biotechnol 2010: 917108.
  47. Gebhardt C, Riehl A, Durchdewald M, Németh J, Fürstenberger G, et al. (2008) RAGE signaling sustains inflammation and promotes tumor development. J Exp Med 205: 275-285.
  48. Duan Z, Chen G, Chen L, Stolzenberg-Solomon R, Weinstein SJet al. (2014) Determinants of concentrations of N(epsilon)-carboxymethyl-lysine and soluble receptor for advanced glycation end products and their associations with risk of pancreatic cancer. Int J MolEpidemiol Genet 5: 152-63.
  49. Perkins ND (2007) Integrating cell-signalling pathways with NF-kappaB and IKK function. Nat Rev Mol Cell Biol 8: 49-62.
  50. Danial NN, Korsmeyer SJ (2004) Cell death: critical control points. Cell 116: 205-219.
  51. Thanos D, Maniatis T (1995) NF-kappa B: a lesson in family values. Cell 80: 529-532.
  52. Pahl HL (1999) Activators and target genes of Rel/NF-kappaB transcription factors. Oncogene 18: 6853-6866.
  53. Vernon PJ, Zeh Iii HJ, Lotze MT (2013) The myeloid response to pancreatic carcinogenesis is regulated by the receptor for advanced glycation end-products. Oncoimmunology 2: e24184.
  54. Tesarova P, Cabinakova M, Mikulova V, Zima T, Kalousova M (2015) RAGE and its ligands in cancer - culprits, biomarkers, or therapeutic targets? Neoplasma 62: 353-364.
  55. Leclerc E, Vetter SW (2015) The role of S100 proteins and their receptor RAGE in pancreatic cancer. BiochimBiophysActa 1852: 2706-2711.
  56. Arumugam T, Ramachandran V, Gomez SB, Schmidt AM, Logsdon CD (2012) S100P-derived RAGE antagonistic peptide reduces tumor growth and metastasis. Clin Cancer Res 18: 4356-4364.
  57. Ji YF, Huang H, Jiang F, Ni RZ, Xiao MB (2014) S100 family signaling network and related proteins in pancreatic cancer (Review). Int J Mol Med 33: 769-776.
  58. Singh R, Barden A, Mori T, Beilin L (2001) Advanced glycation end-products: a review. Diabetologia 44: 129-146.
  59. Arumugam T, Ramachandran V, Logsdon CD (2006) Effect of cromolyn on S100P interactions with RAGE and pancreatic cancer growth and invasion in mouse models. J Natl Cancer Inst 98: 1806-1818.
  60. Arumugam T, Simeone DM, Van Golen K, Logsdon CD (2005) S100P promotes pancreatic cancer growth, survival, and invasion. Clin Cancer Res 11: 5356-5364.
  61. Arumugam T, Ramachandran V, Sun D, Peng Z, Pal A, et al. (2013) Designing and developing S100P inhibitor 5-methyl cromolyn for pancreatic cancer therapy. Mol Cancer Ther 12: 654-662.
  62. Hernández JL, Padilla L, Dakhel S, Coll T, Hervas R, et al. (2013) Therapeutic targeting of tumor growth and angiogenesis with a novel anti-S100A4 monoclonal antibody. PLoS One 8: e72480.
  63. Chambers AF, Matrisian LM (1997) Changing views of the role of matrix metalloproteinases in metastasis. J Natl Cancer Inst 89: 1260-1270.
  64. Waas ET, Lomme RM, DeGroot J, Wobbes T, Hendriks T (2002) Tissue levels of active matrix metalloproteinase-2 and -9 in colorectal cancer. Br J Cancer 86: 1876-1883.
  65. Mendes O, Kim HT, Stoica G (2005) Expression of MMP2, MMP9 and MMP3 in breast cancer brain metastasis in a rat model. ClinExp Metastasis 22: 237-246.
  66. Cooper ME, Vranes D, Youssef S, Stacker SA, Cox AJ et al. (1999) Increased renal expression of vascular endothelial growth factor (VEGF) and its receptor VEGFR-2 in experimental diabetes. Diabetes 48: 2229-39.
  67. Folkman J (1995) Angiogenesis in cancer, vascular, rheumatoid and other disease. Nat Med 1: 27-31.
  68. Yamamoto Y, Yamagishi S, Hsu CC, and Yamamoto H (1996) Advanced glycationendproducts-receptor interactions stimulate the growth of human pancreatic cancer cells through the induction of platelet-derived growth factor-B. BiochemBiophys Res Commun222: 700-5.
  69. Kang R, Tang D, Schapiro NE, Loux T, Livesey KMet al. (2014) The HMGB1/RAGE inflammatory pathway promotes pancreatic tumor growth by regulating mitochondrial bioenergetics. Oncogene 33: 567-77.
  70. Koh MY, Spivak-Kroizman TR, Powis G (2010) HIF-1alpha and cancer therapy. Recent Results Cancer Res 180: 15-34.
  71. Marin-Hernandez A, Gallardo-Perez JC, Ralph SJ, Rodriguez-Enriquez S, and Moreno-Sanchez R (2009) HIF-1alpha modulates energy metabolism in cancer cells by inducing over-expression of specific glycolytic isoforms. Mini Rev Med Chem. 9: 1084-101.
  72. Semenza GL (2002) HIF-1 and tumor progression: pathophysiology and therapeutics. Trends Mol Med 8: S62-67.
  73. Kiffin R, Bandyopadhyay U, Cuervo AM (2006) Oxidative stress and autophagy. Antioxid Redox Signal 8: 152-162.
  74. Kang R, Tang D, Livesey KM, Schapiro NE, LotzeMT (2011) The Receptor for Advanced Glycation End-products (RAGE) protects pancreatic tumor cells against oxidative injury. Antioxid Redox Signal 15: 2175-84.
  75. Wegrzyn J, Potla R, Chwae YJ, Sepuri NB, Zhang Q, et al. (2009) Function of mitochondrial Stat3 in cellular respiration. Science 323: 793-797.
  76. Gough DJ, Corlett A, Schlessinger K, Wegrzyn J, Larner AC, et al. (2009) Mitochondrial STAT3 supports Ras-dependent oncogenic transformation. Science 324: 1713-1716.
  77. Chipuk JE, Kuwana T, Bouchier-Hayes L, Droin NM, Newmeyer DDet al. (2004) Direct activation of Bax by p53 mediates mitochondrial membrane permeabilization and apoptosis. Science 303: 1010-4.
Citation: El-Far AH (2016) The Role of Receptors for Advanced Glycation End Product in Pancreatic Carcinogenesis. Pancreat Disord Ther 6:166.

Copyright: © 2016 El-Far AH. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top