Prostaglandin EP2 receptor

Jump to navigation Jump to search
VALUE_ERROR (nil)
Identifiers
Aliases
External IDsGeneCards: [1]
Orthologs
SpeciesHumanMouse
Entrez
Ensembl
UniProt
RefSeq (mRNA)

n/a

n/a

RefSeq (protein)

n/a

n/a

Location (UCSC)n/an/a
PubMed searchn/an/a
Wikidata
View/Edit Human

Prostaglandin E2 receptor 2 (53kDa), also known as EP2, is a prostaglandin receptor for prostaglandin E2 (PGE2) encoded by the human gene PTGER2: it is one of four identified EP receptors, the others being EP1, EP3, and EP4, which bind with and mediate cellular responses to PGE2 and also, but with lesser affinity and responsiveness, certain other prostanoids (see Prostaglandin receptors).[1] EP has been implicated in various physiological and pathological responses.[2]

Gene

The PTGER2 gene is located on human chromosome 14 at position p22.1 (i.e. 14q22.1), contains 2 introns and 3 exons, and codes for a G protein coupled receptor (GPCR) of the rhodopsin-like receptor family, Subfamily A14 (see rhodopsin-like receptors#Subfamily A14).[3]

Expression

EP2 is widely distributed in humans. Its protein is expressed in human small intestine, lung, media of arteries and arterioles of the kidney, thymus, uterus, brain cerebral cortex, brain striatum, brain hippocampus, corneal epithelium, corneal choriocapillaries, Myometriuml cells, eosinophiles, sclera of the eye, articular cartilage, the corpus cavernosum of the penis, and airway smooth muscle cells; its mRNA is expressed in gingival fibroblasts, monocyte-derived dendritic cells, aorta, corpus cavernosum of the penis, articular cartilage, airway smooth muscle, and airway epithelial cells. In rats, the receptor protein and/or mRNA has been found in lung, spleen, intestine, skin, kidney, liver, long bones, and rather extensively throughout the brain and other parts of the central nervous system.[4][5]

EP2 expression in fibroblasts from the lungs of mice with Bleomycin-induced pulmonary fibrosis and humans with Idiopathic pulmonary fibrosis is greatly reduced. In both instances, this reduced expression was associated with hypermethylation of CpG dinucleotide sites located in the first 420 base pairs upstream of the PTGER2 gene transcription start site of these fibroblasts. This suggests that EP2 expression is regulated by this methylation.[6]

Ligands

Activating ligands

The following standard prostaglandins have the following relative efficacies in binding to and activating EP2: PGE2>PGF2alpha>=PGI2>PGD2.[4] The receptor binding affinity Dissociation constant Kd (i.e. ligand concentration needed to bind with 50% of available EP1 receptors) is ~13 nM for PGE2 and ~10 nM for PGE1 with the human receptor and ~12 nM for PGE2 with the mouse receptor.[7][8] Because PGE2 activates multiple prostanoid receptors and has a short half-life in vivo due to its rapidly metabolism in cells by omega oxidation and beta oxidation, metabolically resistant EP2-selective activators are useful for the study of this receptor's function and could be clinically useful for the treatment of certain diseases. There are several such agonists including butaprost free acid and ONO-AE1-259-01 which have Ki inhibitory binding values (see Biochemis try#Receptor/ligand binding affinity) of 32 and 1.8 NM, respectively, and therefore are respectively ~2.5-fold less and 7-fold more potent than PGE2.[8]

Inhibiting ligands

PF-04418948 (Ki=16 nM), TG4-155 (Ki=9.9 nM), TG8-4, and TG6-129 are selective competitive antagonists for EP2 that have been used for studies in animal models of human diseases. Many of the earlier EP2 receptor antagonists used for such studies exhibited poor receptor selectivity, inhibiting, for example, other EP receptors.[8]

Mechanism of cell activation

EP2 is classified as a relaxant type of prostanoid receptor based on its ability, upon activation, to relax certain types of smooth muscle (see Prostaglandin receptors). When initially bound to PGE2 or any other of its agonists, it mobilizes G proteins containing the Gs alpha subunit (i.e. Gαs)-G beta-gamma complexes (i.e. Gβγ). The Gαs- Gβγ complexes dissociate into their Gαs and Gβγ subunits which in turn regulate cell signaling pathways. In particular, Gαs stimulates adenyl cyclase to raise cellular levels of cAMP thereby activating PKA; PKA activates various types of signaling molecules such as the transcription factor CREB which lead to different types of functional responses depending on cell type.[2][9] EP2 also activates the a) GSK-3 pathway which regulates cell migratory responses and innate immune responses including pro-inflammatory cytokine and interleukin production and b) Beta-catenin pathway which regulates not only cell–cell adhesion but also activates the Wnt signaling pathway which, in turn, stimulates the transcription of genes responsible for regulating cell migration and proliferation.[2] In many of these respects, EP2 actions resemble those of another type of relaxant prostanoid receptor, EP4 but differs from the contractile prostanoid receptors, EP1 and EP3 receptors which mobilize G proteins containing the q-Gβγ complex. EP2 also differs from all the other prostaglandin receptors in that it fails to undergo homologous desensitization. That is, following agonist-induced activation, the other prostaglandin (as well as most types of G protein coupled receptors) quickly become desensitized, often internalized, and whether or not internalized, incapable of activating their G protein targets. This effect limits the duration and extent to which agonists can stimulate cells. EP2, by failing to become desensitized, is able to function over prolong periods and later time points than other prostaglandin receptors and therefore potentially able to contribute to more delayed and chronic phases of cellular and tissue responses.[6]

Functions

Studies using animals genetically engineered to lack EP2 and supplemented by studies examining the actions of EP2 receptor antagonists and agonists in animals as well as animal and human tissues indicate that this receptor serves various functions.

Eye

When applied topically into the eyes of rodents, cats, rhesus monkeys, and humans PGE2 acts, apparently acting at least in part through EP2, decreases intraocular pressure by stimulating increases in the drainage of aqueous humor through the uveoskceral pathway, the principal aqueous humor outflow pathway in the eye.[10]

Reproduction

Female mice engineered to lack a functional Pgter2 gene show a modest reduction in ovulation and more severely impaired capacity for Fertilisation. Studies suggest that this impaired fertilization reflects the loss of EP2 functions in stimulating cumulus cells clusters which surround oocytes to: a) form the CCL7 chemokine which serves as a chemoattractant that guides sperm cells to oocytes and b) disassemble the extracellular matrix which in turn allows sperm cells to penetrate to the oocyte. These data allow that an EP2 receptor antagonist may be a suitable candidate as a contraceptive for women.[11]

Inflammation and allergy

Activation of EP2 contributes to regulating B cell immunoglobulin class switching, maturation of T lymphocyte CD4−CD8− cells to CD4+CD8+ cells, and the function of Antigen-presenting cells, particularly Dendritic cells. EP thereby contributes to the development of inflammation in rodent models of certain types of experimentally-induced joint and paw inflammation and the neurotoxic effects of endotoxin. However, EP2 activation also has anti-inflammatory actions on pro-inflammatory cells (e.g. neutrophils, monocytes, macrophages, dendritic cells, NK cells, TH1 cells, TH2 cells, and fibroblasts in various tissues and on microglia cells in the central nervous system,. These actions suppress certain forms of inflammation such NMDA receptor-related neurotoxicity and the rodent model of Bleomycin-induced pulmonary fibrosis.[2][12] EP2 activation also inhibits the phagocytosis and killing of pathogens by alveolar macrophages; these effects may serve an anti-inflammatory role but reduce host defense against these pathogens.[6]

Activation of EP2 also influences allergic inflammatory reactions. It dilates airways (bronchodilation) contracted by the allergic mediator, histamine; inhibits Immunoglobulin E-activated mast cells from releasing histamine and leukotrienes (viz., LTC4, LTD4, and LTE4), all of which have bronchoconstricting and otherwise pro-allergic actions; inhibits pro-allergic eosinophil apoptosis, chemotaxis, and release of pro-allergic granule contents; and reduces release of the pro-allergic cytokines Interleukin 5, Interleukin 4, and interleukin 13 from human blood mononuclear cells.[13][14]

Cardiovascular

EP2 receptor-deficient mice develop mild systolic and/or systemic hypertension which is worsened by high dietary intake of salt. These effects are thought to be due to the loss of EP2's vasodilation effects and/or ability to increase the urinary excretion of salt.[2][15][16]

Bone

EP2-deficient mice exhibit impaired generation of osteoclasts (cells that break down bone tissue) due to a loss in the capacity of osteoblastic cells to stimulate osteoclast formation. These mice have weakened bones compared with the wild type animals. When administered locally or systemically to animals, EP22-selective agonists stimulate the local or systemic formation of bone, augment bone mass, and accelerate the healing of fractures and other bone defects in animal models.[17]

Nervous system

EP2 deficient mice exhibit reduced Oxidative stress and beta amyloid formation. Activation of this receptor also has neuroprotective effects in models of Alzheimer's disease, Amyotrophic lateral sclerosis, multiple sclerosis, and stroke while its inhibition reduces Epileptic seizure. EP2 receptors that are located on either nerve or Neuroglia cells of the peripheral and central nervous system act to promote pain perception caused by inflammation, muscle stretch, temperature, and physical stimuli (see allodynia) in mice.[5][12]

Malignancy

The EP2 receptor can act as a tumor promoter. EP2 gene knockout mice have less lung, breast, skin, and colon cancers following exposure to carcinogens. Knockout of this gene in mice with the adenomatous polyposis coli mutation also causes a decrease in the size and number of pre-cancerous intestinal polyps that the animals develop. These effects are commonly ascribed to the loss of EP2-mediated: Vascular endothelial growth factor production and thereby of tumor vascularization; regulation of endothelial cell motility and survival; interference with transformng growth factor-β's anti-cell proliferation activity; and, more recently, regulation of host anti-tumor immune responses.[18]

Clinical significance

Therapeutics

Pre-clinical studies, as outlined above, indicate that EP2 may be a target for treating and/or preventing particular human disorders involving: allergic diseases such as asthma (particular aspirin and nonsteroidal inflammatory drug-induced asthma syndromes) and rhinitis;[13] glaucoma;[10] various diseases of the nervous system;[5] fractures, osteoporosis, and other bone abnormalities;[17] pulmonary fibrosis;[12] certain forms of malignant disease such as colon cancer including those that arise from Adenomatous polyposis coli mutations;[18] and salt-sensitive forms of hypertension;[16] This receptor has also been suggested to be a target for contraception.[11] To date, however, there has been little translational research to determine the possible beneficial effects of EP2 antagonists or agonists in humans. The following drugs that act on EP2 but also other prostaglandin receptors are in clinical use:

  • Iloprost activates EP2, EP3, and EP4 receptors to treat diseases involving pathological constriction of blood vessels such as pulmonary hypertension, Raynauds disease, and scleroderma. Presumably, it works by stimulating EP2, and EP4 receptors which have vasodilation actions.
  • Misoprostol, an EP3 and EP4 receptor agonist, to prevent ulcers; to induce labor in pregnancy, medical abortion, and late miscarriage; and to prevent and treat postpartum bleeding.

The following drugs are in development or proposed to be candidates for development as highly selective EP2 agonists for the indicated conditions:[8]

  • Butaprost for the treatment of pulmonary fibrosis and certain neurological diseases
  • CP533,536 for the stimulation of bone formation
  • Taprenepag isopropyl (PF-04217329) for the treatment of glaucoma and various neurological diseases (see above section on Nervous system)

Genomic studies

The Single-nucleotide polymorphism (SNP) variant rs17197[19] in the 3' untranslated region of PTGER2 has been associated with an increased incidence of essential hypertension in a population of Japanese men. SNP variant rs1254598[20] in a Spanish population; SNP variant uS5 located in a STAT-binding consensus sequence of the regulatory region of PTGER2 with reduced transcription activity in a Japanese population; and two PTGER2 SNP variants (-616C>G and -166G>A) in a Korean population have been associated with an increased incidence of Aspirin-induced asthma.[21]

See also

References

  1. https://www.ncbi.nlm.nih.gov/gene/5732
  2. 2.0 2.1 2.2 2.3 2.4 Woodward DF, Jones RL, Narumiya S (September 2011). "International Union of Basic and Clinical Pharmacology. LXXXIII: classification of prostanoid receptors, updating 15 years of progress". Pharmacological Reviews. 63 (3): 471–538. doi:10.1124/pr.110.003517. PMID 21752876.
  3. "PTGER2 prostaglandin E receptor 2 (human)". Entrez Gene. National Center for Biotechnology Information, U.S. National Library of Medicine.
  4. 4.0 4.1 "EP2 receptor". IUPHAR/BPS Guide to Pharmacology.
  5. 5.0 5.1 5.2 Yagami T, Koma H, Yamamoto Y (2016). "Pathophysiological Roles of Cyclooxygenases and Prostaglandins in the Central Nervous System". Molecular Neurobiology. 53 (7): 4754–71. doi:10.1007/s12035-015-9355-3. PMID 26328537.
  6. 6.0 6.1 6.2 Kalinski P (2012). "Regulation of immune responses by prostaglandin E2". Journal of Immunology. 188 (1): 21–8. doi:10.4049/jimmunol.1101029. PMC 3249979. PMID 22187483.
  7. Narumiya S, Sugimoto Y, Ushikubi F (October 1999). "Prostanoid receptors: structures, properties, and functions". Physiological Reviews. 79 (4): 1193–226. doi:10.1152/physrev.1999.79.4.1193. PMID 10508233.
  8. 8.0 8.1 8.2 8.3 Markovič T, Jakopin Ž, Dolenc MS, Mlinarič-Raščan I (January 2017). "Structural features of subtype-selective EP receptor modulators". Drug Discovery Today. 22 (1): 57–71. doi:10.1016/j.drudis.2016.08.003. PMID 27506873.
  9. Moreno JJ (2017). "Eicosanoid receptors: Targets for the treatment of disrupted intestinal epithelial homeostasis". European Journal of Pharmacology. 796: 7–19. doi:10.1016/j.ejphar.2016.12.004. PMID 27940058.
  10. 10.0 10.1 Doucette LP, Walter MA (2016). "Prostaglandins in the eye: Function, expression, and roles in glaucoma". Ophthalmic Genetics. 38: 1–9. doi:10.3109/13816810.2016.1164193. PMID 27070211.
  11. 11.0 11.1 Sugimoto Y, Inazumi T, Tsuchiya S (2015). "Roles of prostaglandin receptors in female reproduction". Journal of Biochemistry. 157 (2): 73–80. doi:10.1093/jb/mvu081. PMID 25480981.
  12. 12.0 12.1 12.2 Matsuoka T, Narumiya S (September 2007). "Prostaglandin receptor signaling in disease". TheScientificWorldJournal. 7: 1329–47. doi:10.1100/tsw.2007.182. PMID 17767353.
  13. 13.0 13.1 Machado-Carvalho L, Roca-Ferrer J, Picado C (August 2014). "Prostaglandin E2 receptors in asthma and in chronic rhinosinusitis/nasal polyps with and without aspirin hypersensitivity". Respiratory Research. 15: 100. doi:10.1186/s12931-014-0100-7. PMC 4243732. PMID 25155136.
  14. Torres R, Picado C, de Mora F (January 2015). "The PGE2-EP2-mast cell axis: an antiasthma mechanism". Molecular Immunology. 63 (1): 61–8. doi:10.1016/j.molimm.2014.03.007. PMID 24768319.
  15. Yang G, Chen L (2016). "An Update of Microsomal Prostaglandin E Synthase-1 and PGE2 Receptors in Cardiovascular Health and Diseases". Oxidative Medicine and Cellular Longevity. 2016: 5249086. doi:10.1155/2016/5249086. PMC 4993943. PMID 27594972.
  16. 16.0 16.1 Yang T, Du Y (2012). "Distinct roles of central and peripheral prostaglandin E2 and EP subtypes in blood pressure regulation". American Journal of Hypertension. 25 (10): 1042–9. doi:10.1038/ajh.2012.67. PMC 3578476. PMID 22695507.
  17. 17.0 17.1 Li M, Thompson DD, Paralkar VM (2007). "Prostaglandin E(2) receptors in bone formation". International Orthopaedics. 31 (6): 767–72. doi:10.1007/s00264-007-0406-x. PMC 2266676. PMID 17593365.
  18. 18.0 18.1 O'Callaghan G, Houston A (November 2015). "Prostaglandin E2 and the EP receptors in malignancy: possible therapeutic targets?". British Journal of Pharmacology. 172 (22): 5239–50. doi:10.1111/bph.13331. PMC 5341220. PMID 26377664.
  19. https://www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?rs=17197&pttaorrene=18Q7Q_47Q5bUsWk2qYtUploOjau55nf2rpc64rtWjoCMhn
  20. https://www.ncbi.nlm.nih.gov/SNP/snp_ref.cgi?rs=1254598&pt=1RUQQJU1TFGBKiod3TITUMQpUJ5_HXHPapAX-sGjOM_cchTks
  21. Cornejo-García JA, Perkins JR, Jurado-Escobar R, García-Martín E, Agúndez JA, Viguera E, Pérez-Sánchez N, Blanca-López N (2016). "Pharmacogenomics of Prostaglandin and Leukotriene Receptors". Frontiers in Pharmacology. 7: 316. doi:10.3389/fphar.2016.00316. PMC 5030812. PMID 27708579.

Further reading

  • Duncan AM, Anderson LL, Funk CD, Abramovitz M, Adam M (February 1995). "Chromosomal localization of the human prostanoid receptor gene family". Genomics. 25 (3): 740–2. doi:10.1016/0888-7543(95)80022-E. PMID 7759114.
  • Wu H, Wu T, Hua W, Dong X, Gao Y, Zhao X, Chen W, Cao W, Yang Q, Qi J, Zhou J, Wang J (March 2015). "PGE2 receptor agonist misoprostol protects brain against intracerebral hemorrhage in mice". Neurobiology of Aging. 36 (3): 1439–50. doi:10.1016/j.neurobiolaging.2014.12.029. PMC 4417504. PMID 25623334.
  • Regan JW, Bailey TJ, Pepperl DJ, Pierce KL, Bogardus AM, Donello JE, Fairbairn CE, Kedzie KM, Woodward DF, Gil DW (August 1994). "Cloning of a novel human prostaglandin receptor with characteristics of the pharmacologically defined EP2 subtype". Molecular Pharmacology. 46 (2): 213–20. PMID 8078484.
  • Wu H, Wu T, Han X, Wan J, Jiang C, Chen W, Lu H, Yang Q, Wang J (January 2017). "Cerebroprotection by the neuronal PGE2 receptor EP2 after intracerebral hemorrhage in middle-aged mice". Journal of Cerebral Blood Flow and Metabolism. 37 (1): 39–51. doi:10.1177/0271678X15625351. PMID 26746866.
  • Bastien L, Sawyer N, Grygorczyk R, Metters KM, Adam M (April 1994). "Cloning, functional expression, and characterization of the human prostaglandin E2 receptor EP2 subtype". The Journal of Biological Chemistry. 269 (16): 11873–7. PMID 8163486.
  • An S, Yang J, Xia M, Goetzl EJ (November 1993). "Cloning and expression of the EP2 subtype of human receptors for prostaglandin E2". Biochemical and Biophysical Research Communications. 197 (1): 263–70. doi:10.1006/bbrc.1993.2470. PMID 8250933.
  • Stillman BA, Breyer MD, Breyer RM (September 1999). "Importance of the extracellular domain for prostaglandin EP(2) receptor function". Molecular Pharmacology. 56 (3): 545–51. PMID 10462542.
  • Smock SL, Pan LC, Castleberry TA, Lu B, Mather RJ, Owen TA (September 1999). "Cloning, structural characterization, and chromosomal localization of the gene encoding the human prostaglandin E(2) receptor EP2 subtype". Gene. 237 (2): 393–402. doi:10.1016/S0378-1119(99)00323-6. PMID 10521663.
  • Desai S, April H, Nwaneshiudu C, Ashby B (December 2000). "Comparison of agonist-induced internalization of the human EP2 and EP4 prostaglandin receptors: role of the carboxyl terminus in EP4 receptor sequestration". Molecular Pharmacology. 58 (6): 1279–86. PMID 11093764.
  • Duckworth N, Marshall K, Clayton JK (February 2002). "An investigation of the effect of the prostaglandin EP2 receptor agonist, butaprost, on the human isolated myometrium from pregnant and non-pregnant women" (PDF). The Journal of Endocrinology. 172 (2): 263–9. doi:10.1677/joe.0.1720263. PMID 11834444.
  • Kyveris A, Maruscak E, Senchyna M (March 2002). "Optimization of RNA isolation from human ocular tissues and analysis of prostanoid receptor mRNA expression using RT-PCR". Molecular Vision. 8: 51–8. PMID 11951086.
  • Takafuji VA, Evans A, Lynch KR, Roche JK (January 2002). "PGE(2) receptors and synthesis in human gastric mucosa: perturbation in cancer". Prostaglandins, Leukotrienes, and Essential Fatty Acids. 66 (1): 71–81. doi:10.1054/plef.2001.0299. PMID 12051958.
  • Scandella E, Men Y, Gillessen S, Förster R, Groettrup M (August 2002). "Prostaglandin E2 is a key factor for CCR7 surface expression and migration of monocyte-derived dendritic cells". Blood. 100 (4): 1354–61. doi:10.1182/blood-2001-11-0017. PMID 12149218.
  • Okuyama T, Ishihara S, Sato H, Rumi MA, Kawashima K, Miyaoka Y, Suetsugu H, Kazumori H, Cava CF, Kadowaki Y, Fukuda R, Kinoshita Y (August 2002). "Activation of prostaglandin E2-receptor EP2 and EP4 pathways induces growth inhibition in human gastric carcinoma cell lines". The Journal of Laboratory and Clinical Medicine. 140 (2): 92–102. doi:10.1016/s0022-2143(02)00023-9. PMID 12228765.
  • Konger RL, Scott GA, Landt Y, Ladenson JH, Pentland AP (December 2002). "Loss of the EP2 prostaglandin E2 receptor in immortalized human keratinocytes results in increased invasiveness and decreased paxillin expression". The American Journal of Pathology. 161 (6): 2065–78. doi:10.1016/S0002-9440(10)64485-9. PMC 1850902. PMID 12466123.
  • Abulencia JP, Gaspard R, Healy ZR, Gaarde WA, Quackenbush J, Konstantopoulos K (August 2003). "Shear-induced cyclooxygenase-2 via a JNK2/c-Jun-dependent pathway regulates prostaglandin receptor expression in chondrocytic cells". The Journal of Biological Chemistry. 278 (31): 28388–94. doi:10.1074/jbc.M301378200. PMID 12743126.
  • Richards JA, Brueggemeier RW (June 2003). "Prostaglandin E2 regulates aromatase activity and expression in human adipose stromal cells via two distinct receptor subtypes". The Journal of Clinical Endocrinology and Metabolism. 88 (6): 2810–6. doi:10.1210/jc.2002-021475. PMID 12788892.
  • Sun HS, Hsiao KY, Hsu CC, Wu MH, Tsai SJ (September 2003). "Transactivation of steroidogenic acute regulatory protein in human endometriotic stromalcells is mediated by the prostaglandin EP2 receptor". Endocrinology. 144 (9): 3934–42. doi:10.1210/en.2003-0289. PMID 12933667.
  • Bradbury DA, Newton R, Zhu YM, El-Haroun H, Corbett L, Knox AJ (December 2003). "Cyclooxygenase-2 induction by bradykinin in human pulmonary artery smooth muscle cells is mediated by the cyclic AMP response element through a novel autocrine loop involving endogenous prostaglandin E2, E-prostanoid 2 (EP2), and EP4 receptors". The Journal of Biological Chemistry. 278 (50): 49954–64. doi:10.1074/jbc.M307964200. PMID 14517215.
  • Moreland RB, Kim N, Nehra A, Goldstein I, Traish A (October 2003). "Functional prostaglandin E (EP) receptors in human penile corpus cavernosum". International Journal of Impotence Research. 15 (5): 362–8. doi:10.1038/sj.ijir.3901042. PMID 14562138.
  • Sugimoto Y, Nakato T, Kita A, Takahashi Y, Hatae N, Tabata H, Tanaka S, Ichikawa A (March 2004). "A cluster of aromatic amino acids in the i2 loop plays a key role for Gs coupling in prostaglandin EP2 and EP3 receptors". The Journal of Biological Chemistry. 279 (12): 11016–26. doi:10.1074/jbc.M307404200. PMID 14699136.

External links

  • "Prostanoid Receptor: EP2". IUPHAR Database of Receptors and Ion Channels. International Union of Basic and Clinical Pharmacology.

This article incorporates text from the United States National Library of Medicine, which is in the public domain.