Biology:CD36

From HandWiki
Short description: Mammalian protein found in Homo sapiens


A representation of the 3D structure of the protein myoglobin showing turquoise α-helices.
Generic protein structure example


CD36 (cluster of differentiation 36), also known as platelet glycoprotein 4, fatty acid translocase (FAT), scavenger receptor class B member 3 (SCARB3), and glycoproteins 88 (GP88), IIIb (GPIIIB), or IV (GPIV) is a protein that in humans is encoded by the CD36 gene. The CD36 antigen is an integral membrane protein found on the surface of many cell types in vertebrate animals. It imports fatty acids inside cells and is a member of the class B scavenger receptor family of cell surface proteins. CD36 binds many ligands including collagen,[1] thrombospondin,[2] erythrocytes parasitized with Plasmodium falciparum,[3] oxidized low density lipoprotein,[4][5] native lipoproteins,[6] oxidized phospholipids,[7] and long-chain fatty acids.[8]

Work in genetically modified rodents suggest a role for CD36 in fatty acid metabolism,[9][10][11] heart disease,[12] taste,[13][14][15] and dietary fat processing in the intestine.[16] It may be involved in glucose intolerance, atherosclerosis, arterial hypertension, diabetes, cardiomyopathy, Alzheimer's disease and various cancers, mostly of epithelial origin (breast, prostate, ovary, and colon) and also for hepatic carcinoma and gliomas.[17][18][19]

Structure

Primary

In humans, rats and mice, CD36 consists of 472 amino acids with a predicted molecular weight of approximately 53,000 Da. However, CD36 is extensively glycosylated and has an apparent molecular weight of 88,000 Da as determined by SDS polyacrylamide gel electrophoresis.[20]

Tertiary

Using Kyte-Doolittle analysis,[21] the amino acid sequence of CD36 predicts a hydrophobic region near each end of the protein large enough to span cellular membranes. Based on this notion and the observation that CD36 is found on the surface of cells, CD36 is thought to have a 'hairpin-like' structure with α-helices at the C- and N- termini projecting through the membrane and a larger extracellular loop (Fig. 1). This topology is supported by transfection experiments in cultured cells using deletion mutants of CD36.[22][23]

Based on the crystal structure of the homologous SCARB2, a model of the extracellular domain of CD36 has been produced.[24] Like SCARB2, CD36 is proposed to contain an antiparallel β-barrel core with many short α-helices adorning it. The structure is predicted to contain a hydrophobic transport tunnel. Disulfide linkages between 4 of the 6 cysteine residues in the extracellular loop are required for efficient intracellular processing and transport of CD36 to the plasma membrane.[25] It is not clear what role these linkages play on the function of the mature CD36 protein on the cell surface.

Posttranslational modification

Besides glycosylation, additional post-translational modifications have been reported for CD36. CD36 is modified with 4 palmitoyl chains, 2 on each of the two intracellular domains.[23] The function of these lipid modifications is currently unknown but they likely promote the association of CD36 with the membrane and possibly lipid rafts which appear to be important for some CD36 functions.[26][27] CD36 could be also phosphorylated at Y62, T92, T323,[28] ubiquitinated at K56, K469, K472 and acetylated at K52, K56, K166, K231, K394, K398, K403.[29][30][31]

Protein-protein interactions

In the absence of ligand, membrane bound CD36 exists primarily in a monomeric state. However exposure to the thrombospondin ligand causes CD36 to dimerize. This dimerization has been proposed to play an important role in CD36 signal transduction.[32]

Genetics

In humans, the gene is located on the long arm of chromosome 7 at band 11.2 (7q11.2[33]) and is encoded by 15 exons that extend over more than 32 kilobases. Both the 5' and the 3' untranslated regions contain introns: the 5' with two and the 3' one. Exons 1, 2 and first 89 nucleotides of exon 3 and as well as exon 15 are non-coding. Exon 3 contains encodes the N-terminal cytoplasmic and transmembrane domains. The C-terminal cytoplasmic and transmembrane regions is encoded by exon 14. The extracellular domain is encoded by the central 11 exons. Alternative splicing of the untranslated regions gives rise to at least two mRNA species.

The transcription initiation site of the CD36 gene has been mapped to 289 nucleotides upstream from the translational start codon and a TATA box and several putative cis regulatory regions lie further 5'. A binding site for PEBP2/CBF factors has been identified between -158 and -90 and disruption of this site reduces expression. The gene is the transcriptional control of the nuclear receptor PPAR/RXR heterodimer (Peroxisome proliferator-activated receptorRetinoid X receptor) and gene expression can be up regulated using synthetic and natural ligands for PPAR and RXR, including the thiazolidinedione class of anti-diabetic drugs and the vitamin A metabolite 9-cis-retinoic acid respectively.

Tissue distribution

CD36 is found on platelets, erythrocytes, monocytes, differentiated adipocytes, skeletal muscle, mammary epithelial cells, spleen cells and some skin microdermal endothelial cells.

Function

The protein itself belongs to the class B scavenger receptor family which includes receptors for selective cholesteryl ester uptake, scavenger receptor class B type I (SR-BI) and lysosomal integral membrane protein II (LIMP-II).

CD36 interacts with a number of ligands, including collagen types I and IV, thrombospondin, erythrocytes parasitized with Plasmodium falciparum, platelet-agglutinating protein p37, oxidized low density lipoprotein and long-chain fatty acids.[34]

On macrophages CD36 forms part of a non-opsonic receptor (the scavenger receptor CD36/alpha-v beta-3 complex) and is involved in phagocytosis.[35]

CD36 has also been implicated in hemostasis, thrombosis, malaria, inflammation, lipid metabolism and atherogenesis.[36]

On binding a ligand the protein and ligand are internalized. This internalization is independent of macropinocytosis and occurs by an actin dependent mechanism requiring the activation Src-family kinases, JNK and Rho-family GTPases.[37] Unlike macropinocytosis this process is not affected by inhibitors of phosphatidylinositol 3-kinase or Na+/H+ exchange.

CD36 ligands have also been shown to promote sterile inflammation through assembly of a Toll-like receptor 4 and 6 heterodimer.[38]

Recently, CD36 was linked to store-operated calcium flux, phospholipase A2 activation, and production of prostaglandin E2[39]

CD36 function in long-chain fatty acid uptake and signaling can be irreversibly inhibited by sulfo-N-succinimidyl oleate (SSO), which binds lysine 164 within a hydrophobic pocket shared by several CD36 ligands, e.g. fatty acid and oxLDL.[30] Recent research concluded that CD36 is involved in the fat taste transduction (oleogustus).

Clinical significance

Malaria

Infections with the human malaria parasite Plasmodium falciparum are characterized by sequestration of erythrocytes infected with mature forms of the parasite and CD36 has been shown to be a major sequestration receptor on microvascular endothelial cells. Parasitised erythrocytes adhere to endothelium at the trophozoite/schizonts stage simultaneous with the appearance of the var gene product (erythrocyte membrane protein 1) on the erythrocyte surface. The appearance of Plasmodium falciparum erythrocyte membrane protein 1 (PfEMP1) on the erythrocyte surface is a temperature dependent phenomenon which is due to increased protein trafficking to the erythrocyte surface at the raised temperature. PfEMP1 can bind other endothelial receptors - thrombospondin (TSP) and intercellular adhesion molecule 1 (ICAM-1) – in addition to CD36 - and genes other than PfEMP1 also bind to CD36: cytoadherence linked protein (clag) and sequestrin. The PfEMP1 binding site on CD36 is known to be located on exon 5.

CD36 on the surface of the platelets has been shown to be involved in adherence but direct adherence to the endothelium by the infected erythrocytes also occurs. Autoaggregation of infected erythrocytes by platelets has been shown to correlate with severe malaria and cerebral malaria in particular and antiplatelet antibodies may offer some protection.

Several lines of evidence suggest that mutations in CD36 are protective against malaria: mutations in the promoters and within introns and in exon 5 reduce the risk of severe malaria. Gene diversity studies suggest there has been positive selection on this gene presumably due to malarial selection pressure. Dissenting reports are also known suggesting that CD36 is not the sole determinant of severe malaria. In addition a role for CD36 has been found in the clearance of gametocytes (stages I and II).

CD36 has been shown to have a role in the innate immune response to malaria in mouse models.[40] Compared with wild type mice CD36 (-/-) mice the cytokine induction response and parasite clearance were impaired. Earlier peak parasitemias, higher parasite densities and higher mortality were noted. It is thought that CD36 is involved in the Plasmodium falciparum glycophosphatidylinositol (PfGPI) induced MAPK activation and proinflammatory cytokine secretion. When macrophages were exposed to PfGPI the proteins ERK1/2, JNK, p38, and c-Jun became phosphorylated. All these proteins are involved as secondary messengers in the immune response. These responses were blunted in the CD36 (-/-) mice. Also in the CD36 (-/-) macrophages secreted significantly less TNF-alpha on exposure to PfGPI. Work is ongoing to determine how these exactly how these responses provide protection against malaria.

CD36 deficiency and alloimmune thrombocytopenia

CD36 is also known as glycoprotein IV (gpIV) or glycoprotein IIIb (gpIIIb) in platelets and gives rise to the Naka antigen. The Naka null phenotype is found in 0.3% of Caucasians and appears to be asymptomatic. The null phenotype is more common in African (2.5%), Japan ese, and other Asian populations (5-11%).

Mutations in the human CD36 gene were first identified in a patient who, despite multiple platelet transfusions, continued to exhibit low platelet levels.[41][42] This condition is known as refractoriness to platelet transfusion. Subsequent studies have shown that CD36 found on the surface of platelets. This antigen is recognized by the monoclonal antibodies (MAbs) OKM5 and OKM8. It is bound by the Plasmodium falciparum protein sequestrin.[43]

Depending on the nature of the mutation in codon 90 CD36 may be absent either on both platelets and monocytes (type 1) or platelets alone (type 2). Type 2 has been divided into two subtypes - a and b. Deficiency restricted to the platelets alone is known as type 2a; if CD36 is also absent from the erythroblasts the phenotype is classified as type 2b.[44] The molecular basis is known for some cases: T1264G in both Kenyans and Gambians; C478T (50%), 539 deletion of AC and 1159 insertion of an A, 1438-1449 deletion and a combined 839-841 deletion GAG and insertion of AAAAC in Japanese.

In a study of 827 apparently healthy Japanese volunteers, type I and II deficiencies were found in 8 (1.0%) and 48 (5.8%) respectively.[45] In 1127 healthy French blood donors (almost all of whom were white Europeans) no CD36 deficiency was found.[46] In a second group only 1 of 301 white test subjects was found to be CD36 deficient. 16 of the 206 sub-Saharan black Africans and 1 of 148 black Caribbeans were found to be CD36 -ve. Three of 13 CD36 -ve persons examined had anti CD36 antibodies. In a group of 250 black American blood donors 6 (2.4%) were found to be Naka antigen negative.[47]

CD36 deficiency may be a cause of post transfusion purpura.[48]

Blood pressure

Below normal levels of CD36 expression in the kidneys has been implicated as a genetic risk factor for hypertension (high blood pressure).[49]

Fatty acid uptake

An association with myocardial fatty acid uptake in humans has been noted.[50] The data suggest a link between hypertrophic cardiomyopathy and CD36 but this needs to be confirmed.

Tuberculosis

RNAi screening in a Drosophila model has revealed that a member of the CD36 family is required for phagocytosis of Mycobacterium tuberculosis into macrophage phagosomes.[51]

Toxoplasmosis

Avirulent strains of Toxoplasma gondii bind to CD36 but virulent parasites fail to engage CD36. In mice, CD36 is required for disease tolerance but not for the development of immunity or resistance.[52]

Obesity

CD36's association with the ability to taste fats has made it a target for various studies regarding obesity and alteration of lipid tasting. CD36 mRNA expression was found to be reduced in taste bud cells (TBC) of obese sand rats (P. obesus) compared to lean controls, implicating an association between CD36 and obesity.[53] Although actual levels of CD36 protein were not different between the obese and control rat cells, Abdoul-Azize et al. hypothesize that the physical distribution of CD36 could differ in obese rat cells.[53] Changes in calcium mediation have been associated with CD36 and obesity as well. Taste bud cells (more specifically, cells from the circumvallate papillae) containing CD36 that were isolated from obese mice exhibited a significantly smaller increase in calcium after fatty acid stimulation when compared to control mice:[54] CD36 associated calcium regulation is impaired when mice are made to be obese (but not in normal weight mice), and this could be a mechanism contributing to behavior changes in the obese mice, such as decreased lipid taste sensitivity and decreased attraction to fats.[54]

There has been some investigation into human CD36 as well. A study examined oral detection of fat in obese subjects with genetic bases for high, medium, and low expression of the CD36 receptor. Those subjects with high CD36 expression were eight times more sensitive to certain fats (oleic acid and triolein) than the subjects with low CD36 expression.[14] Those subjects with an intermediate amount of CD36 expression were sensitive to fat at a level between the high and low groups.[14] This study demonstrates that there is a significant relationship between oral fat sensitivity and the amount of CD36 receptor expression, but further investigation into CD36 could be useful for learning more about lipid tasting in the context of obesity, as CD36 may be a target for therapies in the future.

Establishment of cellular senescence

Upregulation of CD36 could contribute to membrane remodeling during senescence.[55] In response to various senescence‐inducing stimuli, CD36 stimulate NF-κB‐dependent inflammatory cytokine and chemokine production, a phenomenon known as the senescence‐associated secretory phenotype (SASP).[56] This secretory molecule production leads to the onset of a comprehensive senescent cell fate. Reducing the burden of senescent cells, or reducing their inflammatory secretome through CD36 neutralization, accelerates regeneration in young and old mice.[57]

Cancer

CD36 plays a role in the regulation of angiogenesis, which may be a therapeutic strategy for controlling the spread of cancer.[58] Some data from in vitro and animal studies suggested that fatty acid uptake through CD36 may promote cancer cell migration and proliferation in hepatocellular carcinoma, glioblastoma, and potentially other cancers; there was limited data from observational studies in people that low CD36 may correlate with a slightly better outcome in glioblastoma.[59]

Interactions

CD36 has been shown to interact with FYN.[60][61]

Related proteins

CD36 family
4f7b 1.png
Structure of Limp-II. PDB entry 4f7b
Identifiers
SymbolCD36
PfamPF01130
InterProIPR002159

Other human scavenger receptors related to CD36 are SCARB1 and SCARB2 proteins.

See also

References

  1. "Identification of glycoprotein IV (CD36) as a primary receptor for platelet-collagen adhesion". The Journal of Biological Chemistry 264 (13): 7576–83. May 1989. doi:10.1016/S0021-9258(18)83273-2. PMID 2468670. 
  2. "Sense and antisense cDNA transfection of CD36 (glycoprotein IV) in melanoma cells. Role of CD36 as a thrombospondin receptor". The Journal of Biological Chemistry 267 (23): 16607–12. August 1992. doi:10.1016/S0021-9258(18)42046-7. PMID 1379600. 
  3. "CD36 directly mediates cytoadherence of Plasmodium falciparum parasitized erythrocytes". Cell 58 (1): 95–101. July 1989. doi:10.1016/0092-8674(89)90406-6. PMID 2473841. 
  4. "CD36 is a receptor for oxidized low density lipoprotein". The Journal of Biological Chemistry 268 (16): 11811–6. June 1993. doi:10.1016/S0021-9258(19)50272-1. PMID 7685021. 
  5. "Oxidized LDL binds to CD36 on human monocyte-derived macrophages and transfected cell lines. Evidence implicating the lipid moiety of the lipoprotein as the binding site". Arteriosclerosis, Thrombosis, and Vascular Biology 15 (2): 269–75. February 1995. doi:10.1161/01.ATV.15.2.269. PMID 7538425. 
  6. "Human CD36 is a high affinity receptor for the native lipoproteins HDL, LDL, and VLDL". Journal of Lipid Research 39 (4): 777–88. April 1998. doi:10.1016/S0022-2275(20)32566-9. PMID 9555943. http://www.jlr.org/cgi/pmidlookup?view=long&pmid=9555943. [yes|permanent dead link|dead link}}]
  7. "Identification of a novel family of oxidized phospholipids that serve as ligands for the macrophage scavenger receptor CD36". The Journal of Biological Chemistry 277 (41): 38503–16. October 2002. doi:10.1074/jbc.M203318200. PMID 12105195. 
  8. "Reversible binding of long-chain fatty acids to purified FAT, the adipose CD36 homolog". The Journal of Membrane Biology 153 (1): 75–81. September 1996. doi:10.1007/s002329900111. PMID 8694909. 
  9. "Defective fatty acid uptake modulates insulin responsiveness and metabolic responses to diet in CD36-null mice". The Journal of Clinical Investigation 109 (10): 1381–9. May 2002. doi:10.1172/JCI14596. PMID 12021254. 
  10. "Transgenic expression of CD36 in the spontaneously hypertensive rat is associated with amelioration of metabolic disturbances but has no effect on hypertension". Physiological Research 52 (6): 681–8. 2003. doi:10.33549/physiolres.930380. PMID 14640889. http://www.biomed.cas.cz/physiolres/pdf/52/52_681.pdf. 
  11. "Free fatty-acid transport via CD36 drives β-oxidation-mediated hematopoietic stem cell response to infection". Nature Communications 12 (1): 7130. December 2021. doi:10.1038/s41467-021-27460-9. PMID 34880245. Bibcode2021NatCo..12.7130M. 
  12. "Targeted disruption of the class B scavenger receptor CD36 protects against atherosclerotic lesion development in mice". The Journal of Clinical Investigation 105 (8): 1049–56. April 2000. doi:10.1172/JCI9259. PMID 10772649. 
  13. "CD36 involvement in orosensory detection of dietary lipids, spontaneous fat preference, and digestive secretions". The Journal of Clinical Investigation 115 (11): 3177–84. November 2005. doi:10.1172/JCI25299. PMID 16276419. 
  14. 14.0 14.1 14.2 "The fatty acid translocase gene CD36 and lingual lipase influence oral sensitivity to fat in obese subjects". Journal of Lipid Research 53 (3): 561–6. March 2012. doi:10.1194/jlr.M021873. PMID 22210925. 
  15. "Is fat taste ready for primetime?". Physiology & Behavior 136: 145–54. September 2014. doi:10.1016/j.physbeh.2014.03.002. PMID 24631296. 
  16. "CD36 deficiency impairs intestinal lipid secretion and clearance of chylomicrons from the blood". The Journal of Clinical Investigation 115 (5): 1290–7. May 2005. doi:10.1172/JCI21514. PMID 15841205. 
  17. "Molecular basis of human CD36 gene mutations". Molecular Medicine 13 (5–6): 288–96. 2007. doi:10.2119/2006-00088.Rac. PMID 17673938. 
  18. Ana-Maria Enciu; Eugen Radu; Ionela Daniela Popescu; Mihail Eugen Hinescu; Laura Cristina Ceafalan (2018). "Targeting CD36 as Biomarker for Metastasis Prognostic: How Far from Translation into Clinical Practice?". BioMed Research International 2018: 1–12. doi:10.1155/2018/7801202. PMID 30069479. 
  19. Jingchun Wang; Yongsheng Li (2019). "CD36 tango in cancer: signaling pathways and functions". Theranostics 9 (17): 4893–490. doi:10.7150/thno.36037. PMID 31410189. 
  20. "PAS IV, an integral membrane protein of mammary epithelial cells, is related to platelet and endothelial cell CD36 (GP IV)". Biochemistry 29 (30): 7054–9. July 1990. doi:10.1021/bi00482a015. PMID 1699598. 
  21. "A simple method for displaying the hydropathic character of a protein". Journal of Molecular Biology 157 (1): 105–32. May 1982. doi:10.1016/0022-2836(82)90515-0. PMID 7108955. 
  22. "CD36 is a ditopic glycoprotein with the N-terminal domain implicated in intracellular transport". Biochemical and Biophysical Research Communications 275 (2): 446–54. August 2000. doi:10.1006/bbrc.2000.3333. PMID 10964685. 
  23. 23.0 23.1 "CD36 is palmitoylated on both N- and C-terminal cytoplasmic tails". The Journal of Biological Chemistry 271 (37): 22315–20. September 1996. doi:10.1074/jbc.271.37.22315. PMID 8798390. 
  24. "Structure of LIMP-2 provides functional insights with implications for SR-BI and CD36". Nature 504 (7478): 172–6. December 2013. doi:10.1038/nature12684. PMID 24162852. Bibcode2013Natur.504..172N. 
  25. "Formation of one or more intrachain disulphide bonds is required for the intracellular processing and transport of CD36". The Biochemical Journal 328 (2): 635–42. December 1997. doi:10.1042/bj3280635. PMID 9371725. 
  26. "Endocytosis of oxidized low density lipoprotein through scavenger receptor CD36 utilizes a lipid raft pathway that does not require caveolin-1". The Journal of Biological Chemistry 278 (46): 45931–6. November 2003. doi:10.1074/jbc.M307722200. PMID 12947091. 
  27. "FAT/CD36-mediated long-chain fatty acid uptake in adipocytes requires plasma membrane rafts". Molecular Biology of the Cell 16 (1): 24–31. January 2005. doi:10.1091/mbc.E04-07-0616. PMID 15496455. 
  28. "CD36 (human) protein page". PhosphoSitePlus. Cell Signaling Technology, Inc.. http://www.phosphosite.org/proteinAction.do?id=18224&showAllSites=true. 
  29. "Opposite regulation of CD36 ubiquitination by fatty acids and insulin: effects on fatty acid uptake". The Journal of Biological Chemistry 283 (20): 13578–85. May 2008. doi:10.1074/jbc.M800008200. PMID 18353783. 
  30. 30.0 30.1 "Sulfo-N-succinimidyl oleate (SSO) inhibits fatty acid uptake and signaling for intracellular calcium via binding CD36 lysine 164: SSO also inhibits oxidized low density lipoprotein uptake by macrophages". The Journal of Biological Chemistry 288 (22): 15547–55. May 2013. doi:10.1074/jbc.M113.473298. PMID 23603908. 
  31. "Proteomic analysis of lysine acetylation sites in rat tissues reveals organ specificity and subcellular patterns". Cell Reports 2 (2): 419–31. August 2012. doi:10.1016/j.celrep.2012.07.006. PMID 22902405. 
  32. "Thrombospondin induces dimerization of membrane-bound, but not soluble CD36". Thrombosis and Haemostasis 78 (2): 897–901. August 1997. doi:10.1055/s-0038-1657649. PMID 9268192. 
  33. "Gene encoding the collagen type I and thrombospondin receptor CD36 is located on chromosome 7q11.2". Genomics 17 (3): 759–61. September 1993. doi:10.1006/geno.1993.1401. PMID 7503937. 
  34. "Structural organization of the gene for human CD36 glycoprotein". The Journal of Biological Chemistry 269 (29): 18985–91. July 1994. doi:10.1016/S0021-9258(17)32263-9. PMID 7518447. 
  35. "CD36 and TLR interactions in inflammation and phagocytosis: implications for malaria". Journal of Immunology 183 (10): 6452–9. November 2009. doi:10.4049/jimmunol.0901374. PMID 19864601. 
  36. "Vascular biology of CD36: roles of this new adhesion molecule family in different disease states". Thrombosis and Haemostasis 78 (1): 65–9. July 1997. doi:10.1055/s-0038-1657502. PMID 9198129. 
  37. "Uptake of oxidized low density lipoprotein by CD36 occurs by an actin-dependent pathway distinct from macropinocytosis". The Journal of Biological Chemistry 284 (44): 30288–97. October 2009. doi:10.1074/jbc.M109.045104. PMID 19740737. 
  38. "CD36 ligands promote sterile inflammation through assembly of a Toll-like receptor 4 and 6 heterodimer". Nature Immunology 11 (2): 155–61. February 2010. doi:10.1038/ni.1836. PMID 20037584. 
  39. "CD36 protein is involved in store-operated calcium flux, phospholipase A2 activation, and production of prostaglandin E2". The Journal of Biological Chemistry 286 (20): 17785–95. May 2011. doi:10.1074/jbc.M111.232975. PMID 21454644. 
  40. "Disruption of CD36 impairs cytokine response to Plasmodium falciparum glycosylphosphatidylinositol and confers susceptibility to severe and fatal malaria in vivo". Journal of Immunology 178 (6): 3954–61. March 2007. doi:10.4049/jimmunol.178.6.3954. PMID 17339496. 
  41. "A new platelet-specific antigen, Naka, involved in the refractoriness of HLA-matched platelet transfusion". Vox Sanguinis 57 (3): 213–7. 1989. doi:10.1111/j.1423-0410.1989.tb00826.x. PMID 2617957. 
  42. "A platelet membrane glycoprotein (GP) deficiency in healthy blood donors: Naka- platelets lack detectable GPIV (CD36)". Blood 76 (9): 1698–703. November 1990. doi:10.1182/blood.V76.9.1698.1698. PMID 1699620. http://www.bloodjournal.org/cgi/pmidlookup?view=long&pmid=1699620. 
  43. "Sequestrin, a CD36 recognition protein on Plasmodium falciparum malaria-infected erythrocytes identified by anti-idiotype antibodies". Proceedings of the National Academy of Sciences of the United States of America 88 (8): 3175–9. April 1991. doi:10.1073/pnas.88.8.3175. PMID 1707534. Bibcode1991PNAS...88.3175O. 
  44. "Erythroid involvement in CD36 deficiency". Experimental Hematology 29 (10): 1194–200. October 2001. doi:10.1016/S0301-472X(01)00691-9. PMID 11602321. 
  45. "Phenotype-genotype correlation in CD36 deficiency types I and II". Thrombosis and Haemostasis 84 (3): 436–41. September 2000. doi:10.1055/s-0037-1614041. PMID 11019968. 
  46. "CD36 deficiency is frequent and can cause platelet immunization in Africans". Transfusion 39 (8): 873–9. August 1999. doi:10.1046/j.1537-2995.1999.39080873.x. PMID 10504124. 
  47. "Incidence of the Nak(a)-negative platelet phenotype in African Americans is similar to that of Asians". Transfusion 36 (4): 331–4. April 1996. doi:10.1046/j.1537-2995.1996.36496226147.x. PMID 8623134. 
  48. "Posttransfusion purpura-like syndrome associated with CD36 (Naka) isoimmunization". Transfusion 35 (9): 777–82. September 1995. doi:10.1046/j.1537-2995.1995.35996029165.x. PMID 7570941. 
  49. "Identification of renal Cd36 as a determinant of blood pressure and risk for hypertension". Nature Genetics 40 (8): 952–4. August 2008. doi:10.1038/ng.164. PMID 18587397. 
  50. "CD36 abnormality and impaired myocardial long-chain fatty acid uptake in patients with hypertrophic cardiomyopathy". Japanese Circulation Journal 62 (7): 499–504. July 1998. doi:10.1253/jcj.62.499. PMID 9707006. 
  51. "Drosophila RNAi screen reveals CD36 family member required for mycobacterial infection". Science 309 (5738): 1251–3. August 2005. doi:10.1126/science.1116006. PMID 16020694. Bibcode2005Sci...309.1251P. 
  52. "CD36 mediates phagocyte tropism and avirulence of Toxoplasma gondii". Journal of Immunology 207 (6): 1507–1512. August 2021. doi:10.4049/jimmunol.2100605. ISSN 0022-1767. PMID 34400524. 
  53. 53.0 53.1 "Oro-gustatory perception of dietary lipids and calcium signaling in taste bud cells are altered in nutritionally obesity-prone Psammomys obesus". PLOS ONE 8 (8): e68532. 2013. doi:10.1371/journal.pone.0068532. PMID 23936306. Bibcode2013PLoSO...868532A. 
  54. 54.0 54.1 "Obesity alters the gustatory perception of lipids in the mouse: plausible involvement of lingual CD36". Journal of Lipid Research 54 (9): 2485–94. September 2013. doi:10.1194/jlr.M039446. PMID 23840049. 
  55. Saitou, Marie; Lizardo, Darleny Y.; Taskent, Recep Ozgur; Millner, Alec; Gokcumen, Omer; Atilla-Gokcumen, Gunes Ekin (2018). "An evolutionary transcriptomics approach links CD36 to membrane remodeling in replicative senescence". Molecular Omics (Royal Society of Chemistry (RSC)) 14 (4): 237–246. doi:10.1039/c8mo00099a. ISSN 2515-4184. PMID 29974107. https://www.biorxiv.org/content/biorxiv/early/2018/04/03/294512.full.pdf. 
  56. "CD36 initiates the secretory phenotype during the establishment of cellular senescence". EMBO Reports 19 (6). 2018. doi:10.15252/embr.201745274. PMID 29777051. 
  57. "Senescence atlas reveals an aged-like inflamed niche that blunts muscle regeneration". Nature 612 (7941): 169–178. 2022. doi:10.1038/s41586-022-05535-x. PMID 36544018. 
  58. "CD36: a multiligand molecule". Laboratory Hematology 11 (1): 31–7. 2005. doi:10.1532/LH96.04056. PMID 15790550. 
  59. "Attacking the supply wagons to starve cancer cells to death". FEBS Letters 590 (7): 885–907. April 2016. doi:10.1002/1873-3468.12121. PMID 26938658. 
  60. "Membrane glycoprotein IV (CD36) is physically associated with the Fyn, Lyn, and Yes protein-tyrosine kinases in human platelets". Proceedings of the National Academy of Sciences of the United States of America 88 (17): 7844–8. September 1991. doi:10.1073/pnas.88.17.7844. PMID 1715582. Bibcode1991PNAS...88.7844H. 
  61. "Src-related protein tyrosine kinases are physically associated with the surface antigen CD36 in human dermal microvascular endothelial cells". FEBS Letters 351 (1): 41–4. August 1994. doi:10.1016/0014-5793(94)00814-0. PMID 7521304. 

Further reading

External links