N-Terminal Methionine Excision: History
Please note this is an old version of this entry, which may differ significantly from the current revision.
Contributor:

In the cytosol of human cells, when a newly synthesized polypeptide emerges from the ribosomes, its fate can be determined by the enzymes that modify its N-terminal α-amino acid residue (Nα). These N-terminal modifications include excision of the initiator methionine (iMet), Nα-myristoylation, Nα-acetylation, Nα-methylation, and other less common modification events. Methionine aminopeptidases (MetAPs) are responsible for N-terminal iMet excision (NME).

  • methionine aminopeptidases
  • N-terminal iMet excision

1. Introduction

In the cytosol of human cells, when a newly synthesized polypeptide emerges from the ribosomes, its fate can be determined by the enzymes that modify its N-terminal α-amino acid residue (Nα). These N-terminal modifications include excision of the initiator methionine (iMet), Nα-myristoylation, Nα-acetylation, Nα-methylation, and other less common modification events such as Nα-propionylation, Nα-palmitoylation, Nα-arginylation, and Nα-ubiquitylation. Among these enzymes, methionine aminopeptidases (MetAPs) are responsible for N-terminal iMet excision (NME) [1][2]; N-terminal acetyltransferases (NATs) for Nα-acetylation [3]; N-terminal myristoyltransferase (NMTs) for Nα-myristoylation [4]; N-terminal methylation for Nα-methylation (NTMTs) [5]; N-terminal palmitoylacyltransferases (PATs) for Nα-palmitoylation [6]; and ubiquitin ligases for ubiquitylation of the N-terminal α-amino acid residue [7]. The NATs can also sometimes catalyze a much less understood modification: Nα-propionylation [8]. These modifications of proteins at their N-termini play critical roles in many important cellular processes, and dysregulation of these events could significantly impact the development and progression of certain human diseases [9][10]

2. N-Terminal Methionine Excision (NME)

Protein synthesis in the cytosol of a eukaryotic cell, in most cases, is initiated with methionine. When the second amino acid residue is a small and uncharged amino acid such as Ala, Cys, Gly, Pro, Ser, Thr, or Val, iMet is usually removed co-translationally by two types of methionine aminopeptidases (MetAPs) [11][12][13][14][15][16][17][18][19][20]. Although these two types do not share a high sequence identity, their catalytic domains belong to the same family of metalloproteases with a typical “pita-bread” protease fold [19]. The N-terminal domain of eukaryotic MetAP1s contains two zinc finger motifs; a RING-finger-like Cys2-Cys2 zinc finger and a Cys2-His2 zinc finger related to RNA-binding zinc fingers [21][22]. These two zinc fingers are essential for the regular functional association of MetAP1 with the ribosomes [21][22]. On the other hand, eukaryotic type 2 MetAPs (MetAP2s) contain an N-terminal domain with a positively charged Lys-rich region [16][17][18][19][20]. Deleting MetAP1 in yeast leads to a slow growth phenotype which can be rescued by overexpressing MetAP2, whereas knocking out both MetAPs is lethal, indicating that the NME process is vital for normal cell growth (Table 1) [17]. This finding strongly suggests that both enzymes act on similar groups of substrates in vivo. Structural studies of human MetAPs revealed a potential difference in the substrate specificity of their catalytic sites due to more steric restrictions in MetAP1 [20]. Proteomics analysis of the substrate specificity of human MetAPs indicates that MetAP2 prefers iMet-Val and iMet-Thr. However, substrate specificity significantly overlaps between human MetAP1 and MetAP2 [15].
Table 1. Impact of protein Nα-modifications on cellular functions.
Enzymes Cellular Functions References
MetAP1 Cell cycle progression, cell proliferation, enzyme function, protein stability, cellular localization [13][14][15][16][17][18][19][20][21][22][23]
MetAP2 Angiogenesis, B-cell differentiation, cell-specific Cytotoxicity [23][24][25][26][27][28][29][30][31][32][33][34]
NMTs Signal transduction, cellular transformation,
innate immune responses, adaptative immune response
[35][36][37][38]
NATs Actin cytoskeleton structure, cell cycle progression, cell proliferation
cell mobility
[39][40][41][42][43][44][45][46][47][48]
NTMTs Protein stability, protein-protein interaction, protein-DNA interaction, cellular localization, response to cellular stress, DNA repair, regulation of mitosis, chromatin interaction, tRNA transport, genome stability [5][49][50][51][52][53][54][55][56][57][58]
Since discovering that human MetAP2, not human MetAP1, is the molecular target of TNP-470, a potent anti-angiogenesis inhibitor, MetAP2 has become a drug target for treating cancer, obesity, Prader-Willi Syndrome (PWS), and autoimmunity (Table 2) [23][24][25][26]. Inhibitors for human MetAP2 are well tolerated in patients at therapeutically relevant doses and have been developed for a variety of pharmaceutical applications, including the treatment of cancer [27][28][29][30][31], diabetes, and obesity [32], as well as the modulation of autoimmunity [33][34]. Although none of these inhibitors have yet passed Phase III clinical trials, the interest of the drug development community remains high due to continued promising preclinical and clinical efficacy results for novel MetAP2 inhibitors. Unfortunately, most studies did not assess the impact of MetAP2 inhibition on cellular functions, making it harder to correlate the phenotypes to the inhibitors’ mode of action. Most of the time, more questions were raised than answered regarding the role of MetAP2 in these diseases after a new clinical study. For example, it is still being determined whether the molecular mechanisms driving each phenotype discovered during each clinical trial share the exact molecular mechanisms. The molecular mechanisms of MetAP2 inhibitor-induced weight loss or immune modulation remain to be established. Even the fundamental questions regarding the substrate specificity of MetAPs in different tissues still need to be better defined. Indeed, a better understanding of MetAP biology and the mode of action of MetAP2 inhibitors would undoubtedly improve the quality of biomarkers for patient screening, the identification of novel indications, and the development of evidence-based drug combinations in targeted disease treatment.
In the mitochondria of human cells, protein synthesis is initiated with formyl-methionine. A deformylase can remove the formyl group to expose an unmodified methionine, which becomes a substrate for MetAP. A search of the GenBank database with cytosolic MetAP1 and MetAP2 protein sequences led to the discovery of MetAP1D [59]. MetAP1D is a new member of the human MetAP family and belongs to the Type I MetAP subfamily. Phylogenetic analysis of human MetAP isoforms suggests that human MetAP1D pairs with mitochondrial MetAP orthologs previously identified in plants [90]. All three MetAP isoforms can remove Met from a Met-Ala-Ser peptide in vitro. However, the substrate specificity of MetAP1D has not been thoroughly investigated. MetAP1D is overexpressed in colon cancer cells and colon tumors. Reduced expression of MetAP1D by shRNA has been shown to decrease the ability of colon cancer cells to grow in soft agar, indicating that overexpression of MetAP1D may be necessary for tumorigenesis. Thus, MAP1D has been suggested to be a target for chemotherapy in colon carcinoma [59][60].
Recently, genomic analyses demonstrated that patients with intellectual disability (ID) harbor a novel homozygous nonsense mutation in the MetAP1 gene. ID is a common genetic and clinically heterogeneous disease, and underlying molecular pathogenesis can frequently be unidentified by whole-exome/genome testing. Improper neuronal function from losing essential proteins could lead to neurologic impairment and ID [91]. In addition, a mutation in the MetAP1D gene was identified as one candidate involved in the penetrance of Leber’s hereditary optic neuropathy (LHON) [92]. Though researchers are still very early in understanding how mutations in MetAPs could affect human health, NME excision processes provide a promising avenue in translational research.

This entry is adapted from the peer-reviewed paper 10.3390/life13071613

References

  1. Marino, G.; Eckhard, U.; Overall, C.M. Protein termini and their modifications revealed by positional proteomics. ACS Chem. Biol. 2015, 10, 1754–1764.
  2. Sherman, F.; Stewart, J.W.; Tsunasawa, S. Methionine or not methionine at the beginning of a protein. Bioessays 1985, 3, 27–31.
  3. Starheim, K.K.; Gevaert, K.; Arnesen, T. Protein N-terminal acetyltransferases: When the start matters. Trends Biochem. Sci. 2012, 37, 152–161.
  4. Martin, D.D.; Beauchamp, E.; Berthiaume, L.G. Posttranslational myristoylation: Fat matters in cellular life and death. Biochimie 2011, 93, 18–31.
  5. Stock, A.; Clarke, S.; Clarke, C.; Stock, J. N-Terminal Methylation of Proteins—Structure, Function and Specificity. FEBS Lett. 1987, 220, 8–14.
  6. Buglino, J.A.; Resh, M.D. Palmitoylation of Hedgehog proteins. Vitam. Horm. 2012, 88, 229–252.
  7. Ciechanover, A.; Ben-Saadon, R. N-terminal ubiquitination: More protein substrates join in. Trends Cell Biol. 2004, 14, 103–106.
  8. Foyn, H.; Van Damme, P.; Stove, S.I.; Glomnes, N.; Evjenth, R.; Gevaert, K.; Arnesen, T. Protein N-terminal acetyltransferases act as N-terminal propionyltransferases in vitro and in vivo. Mol. Cell. Proteom. 2013, 12, 42–54.
  9. Chen, L.; Kashina, A. Post-translational Modifications of the Protein Termini. Front. Cell. Dev. Biol. 2021, 9, 719590.
  10. Tooley, J.G.; Schaner Tooley, C.E. New roles for old modifications: Emerging roles of N-terminal post-translational modifications in development and disease. Protein. Sci. 2014, 23, 1641–1649.
  11. Boissel, J.P.; Kasper, T.J.; Shah, S.C.; Malone, J.I.; Bunn, H.F. Amino-terminal processing of proteins: Hemoglobin South Florida, a variant with retention of initiator methionine and Nalpha-acetylation. Proc. Natl. Acad. Sci. USA 1985, 82, 8448–8452.
  12. Tsunasawa, S.; Stewart, J.W.; Sherman, F. Amino-terminal processing of mutant forms of yeast iso-1-cytochrome c. The specificities of methionine aminopeptidase and acetyltransferase. J. Biol. Chem. 1985, 260, 5382–5391.
  13. Dummitt, B.; Micka, W.S.; Chang, Y.H. Yeast Glutamine-fructose-6-phosphate Amidotransferase (Gfa1) requires methionine aminopeptidase activity for proper function. J. Biol. Chem. 2005, 280, 14356–14360.
  14. Jonckheere, V.; Fijałkowska, D.; Van Damme, P. Omics Assisted N-terminal Proteoform and Protein Expression Profiling On Methionine Aminopeptidase 1 (MetAP1) Deletion. Mol. Cell Proteom. 2018, 17, 694–708.
  15. Xiao, Q.; Zhang, F.; Nacev, B.A.; Liu, J.O.; Pei, D. Protein N-terminal processing: Substrate specificity of Escherichia coli and human methionine aminopeptidases. Biochemistry 2010, 49, 5588–5599.
  16. Vetro, J.; Dummitt, B.; Chang, Y.H. Angiogenesis: Emerging Role of Methionine in Aminopeptidases, Aminopeptidases in Biology and Disease Series: Proteases in Biology and Disease; Hooper, N., Uwe, L., Eds.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2004; Volume 2.
  17. Li, X.; Chang, Y.H. Amino-terminal protein processing in Saccharomyces cerevisiae is an essential function that requires two distinct methionine aminopeptidases. Proc. Natl. Acad. Sci. USA 1995, 92, 12357–12361.
  18. Li, X.; Chang, Y.H. Evidence that the human homolog of a rat initiation factor- 2 associated protein (p67) is methionine aminopeptidase. Biochem. Biophys. Res. Commun. 1996, 227, 152–159.
  19. Arfin, S.M.; Kendall, R.L.; Hall, L.; Weaver, L.H.; Stewart, A.E.; Matthews, B.W.; Bradshaw, R.A. Eukaryotic methionyl aminopeptidases: Two classes of cobalt-dependent enzymes. Proc. Natl. Acad. Sci. USA 1995, 92, 7714–7718.
  20. Addlagatta, A.; Hu, X.; Liu, J.O.; Matthews, B.W. Structural basis for the functional differences between type I and type II human methionine aminopeptidases. Biochemistry 2005, 44, 14741–14749.
  21. Xu, X.; Addlagatta, A.; Jun Lu, J.; Liu, J.O. Elucidation of the function of type 1 human methionine aminopeptidase during cell cycle progression. Proc. Natl. Acad. Sci. USA 2006, 103, 18148–18153.
  22. Vetro, J.; Chang, Y.H. Yeast MetAP1 is ribosome-associated and requires its N-terminal zinc finger domain for normal function in vivo. J. Cell. Biochem. 2002, 85, 678–688.
  23. Griffith, E.C.; Su, Z.; Niwayama, S.; Ramsay, C.A.; Chang, Y.H.; Liu, J.O. Molecular recognition of angiogenesis inhibitors fumagillin and ovarian by methionine aminopeptidase 2. Proc. Natl. Acad. Sci. USA 1998, 95, 15183–15188.
  24. Sin, N.; Meng, L.; Wang, M.Q.; Wen, J.J.; Bornmann, W.G.; Crews, C.M. The anti-angiogenic agent fumagillin covalently binds and inhibits the methionine aminopeptidase, MetAP-2. Proc. Natl. Acad. Sci. USA 1997, 94, 6099–6103.
  25. Turk, B.E.; Griffith, E.C.; Wolf, S.; Bieman, K.; Chang, Y.H.; Liu, J.O. Selective inhibition of N-terminal processing by TNP-470 and Ovalicin in endothelial cells. Chem. Biol. 1999, 6, 823–833.
  26. Grocin, A.G.; Kallemeijn, W.W.; Tate, E.W. Targeting methionine aminopeptidase 2 in cancer, obesity, and autoimmunity. Trends Pharmacol. Sci. 2021, 42, 870–882.
  27. Ingber, D.; Fujita, T.; Kishimoto, S.; Sudo, K.; Kanamaru, T.; Brem, H.; Folkman, J. Synthetic analogues of fumagillin that inhibit angiogenesis and suppress tumour growth. Nature 1990, 348, 555–557.
  28. Kruger, E.A.; Figg, W.D. TNP-470: An angiogenesis inhibitor in clinical development for cancer. Expert Opin. Investig. Drugs 2000, 9, 1383–1396.
  29. Wang, J.; Sheppard, G.S.; Lou, P.; Kawai, M.; BaMaung, N.; Erickson, S.A.; Tucker-Garcia, L.; Park, C.; Bouska, J.; Wang, Y.C.; et al. Tumor suppression by a rationally designed reversible inhibitor of methionine aminopeptidase-2. Cancer Res. 2003, 63, 7861–7869.
  30. Heinrich, T.; Seenisamy, J.; Blume, B.; Bomke, J.; Calderini, M.; Eckert, U.; Friese-Hamim, M.; Kohl, R.; Lehmann, M.; Leuthner, B.; et al. Discovery and structure-based optimization of next-generation reversible methionine aminopeptidase-2 (MetAP-2) inhibitors. J. Med. Chem. 2019, 62, 5025–5039.
  31. Heinrich, T.; Seenisamy, J.; Blume, B.; Bomke, J.; Dietz, M.; Eckert, U.; Friese-Hamim, M.; Gunera, J.; Hansen, K.; Leuthner, B.; et al. Identification of methionine aminopeptidase-2 (MetAP-2) inhibitor M8891, a clinical compound for the treatment of cancer. J. Med. Chem. 2019, 62, 11119–11134.
  32. Huang, H.-J.; Holub, C.; Rolzin, P.; Bilakovics, J.; Fanjul, A.; Satomi, Y.; Plonowski, A.; Larson, C.J.; Farrell, P.J. MetAP2 inhibition increases energy expenditure through direct action on brown adipocytes. J. Biol. Chem. 2019, 294, 9567–9575.
  33. Kanno, T.; Endo, H.; Takeuchi, K.; Morishita, Y.; Fukayama, M.; Mori, S. High expression of methionine aminopeptidase type 2 in germinal center B cells and their neoplastic counterparts. Lab. Investig. 2002, 82, 893–901.
  34. Priest, R.C.; Spaull, J.; Buckton, J.; Grimley, R.L.; Sims, M.; Binks, M.; Malhotra, R. Immunomodulatory activity of a methionine aminopeptidase-2 inhibitor on B cell differentiation. Clin. Exp. Immunol. 2009, 155, 514–522.
  35. Bhargava, M.; Cajas, J.M.; Wainberg, M.A.; Klein, M.B.; Pant Pai, N. Do HIV-1 non-B subtypes differentially impact resistance mutations and clinical disease progression in treated populations? Evidence from a systematic review. J. Int. AIDS Soc. 2014, 17, 18944.
  36. Boutin, J.A. Myristoylation. Cell Signal 1997, 9, 15–35.
  37. Taniguchi, H. Protein myristoylation in protein-lipid and protein-protein interactions. Biophys. Chem. 1999, 82, 129–137.
  38. Towler, D.A.; Gordon, J.I.; Adams, S.P.; Glaser, L. The biology and enzymology of eukaryotic protein acylation. Annu. Rev. Biochem. 1988, 57, 69–99.
  39. Scott, D.C.; Monda, J.K.; Bennett, E.J.; Harper, J.W.; Schulman, B.A. N-terminal acetylation acts as an avidity enhancer within an interconnected multiprotein complex. Science 2011, 334, 674–678.
  40. Monda, J.K.; Scott, D.C.; Miller, D.J.; Lydeard, J.; King, D.; Harper, J.W.; Bennett, E.J.; Schulman, B.A. Structural conservation of distinctive N-terminal acetylation-dependent interactions across a family of mammalian NEDD8 ligation enzymes. Structure 2013, 21, 42–53.
  41. Yang, H.; Ni, H.M.; Ding, W.X. The double-edged sword of MTOR in autophagy deficiency induced-liver injury and tumorigenesis. Autophagy 2019, 15, 1671–1673.
  42. Arnaudo, N.; Fernández, I.S.; McLaughlin, S.H.; Peak-Chew, S.Y.; Rhodes, D.; Martino, F. The N-terminal acetylation of Sir3 stabilizes its binding to the nucleosome core particle. Nat. Struct. Mol. Biol. 2013, 20, 1119–1121.
  43. Behnia, R.; Panic, B.; Whyte, J.R.; Munro, S. Targeting of the Arf-like GTPase Arl3p to the Golgi requires N-terminal acetylation and the membrane protein Sys1p. Nat. Cell Biol. 2004, 6, 405–413.
  44. Setty, S.R.; Strochlic, T.I.; Tong, A.H.; Boone, C.; Burd, C.G. Golgi targeting of ARF-like GTPase Arl3p requires its Nalpha-acetylation and the integral membrane protein Sys1p. Nat. Cell Biol. 2004, 6, 414–419.
  45. Dikic, I.; Elazar, Z. Mechanism and medical implications of mammalian autophagy. Nat. Rev. Mol. Cell Biol. 2018, 19, 349–364.
  46. Hwang, C.S.; Shemorry, A.; Varshavsky, A. N-terminal acetylation of cellular proteins creates specific degradation signals. Science 2010, 327, 973–977.
  47. Shemorry, A.; Hwang, C.S.; Varshavsky, A. Control of protein quality and stoichiometries by N-terminal acetylation and the N-end rule pathway. Mol. Cell 2013, 50, 540–551.
  48. Park, S.E.; Kim, J.M.; Seok, O.H.; Cho, H.; Wadas, B.; Kim, S.Y.; Varshavsky, A.; Hwang, C.S. Control of mammalian G protein signaling by N-terminal acetylation and the N-end rule pathway. Science 2015, 347, 1249–1252.
  49. Dong, C.; Mao, Y.; Tempel, W.; Qin, S.; Li, L.; Loppnau, P.; Huang, R.; Min, J. Structural Basis for Substrate Recognition by the Human N-Terminal Methyltransferase 1. Genes Dev. 2015, 29, 2343–2348.
  50. Huang, R. Chemical Biology of Protein N-Terminal Methyltransferases. ChemBioChem 2019, 20, 976–984.
  51. Chen, P.; Sobreira, T.J.P.; Hall, M.C.; Hazbun, T.R. Discovering the N-Terminal Methylome by Repurposing of Proteomic Datasets. J. Proteome Res. 2021, 20, 4231–4247.
  52. Webb, K.J.; Lipson, R.S.; Al-Hadid, Q.; Whitelegge, J.P.; Clarke, S.G. Identification of Protein N-Terminal Methyltransferases in Yeast and Humans. Biochemistry 2010, 49, 5225–5235.
  53. Kimura, Y.; Kurata, Y.; Ishikawa, A.; Okayama, A.; Kamita, M.; Hirano, H. N-Terminal Methylation of Proteasome Subunit Rpt1 in Yeast. Proteomics 2013, 13, 3167–3174.
  54. Hamey, J.J.; Winter, D.L.; Yagoub, D.; Overall, C.M.; Hart-Smith, G.; Wilkins, M.R. Novel N-Terminal and Lysine Methyltransferases That Target Translation Elongation Factor 1A in Yeast and Human. Mol. Cell. Proteom. 2016, 15, 164–176.
  55. Schaner Tooley, C.E.; Petkowski, J.J.; Muratore-Schroeder, T.L.; Balsbaugh, J.L.; Shabanowitz, J.; Sabat, M.; Minor, W.; Hunt, D.F.; Macara, I.G. NRMT Is an α-NMethyltransferase That Methylates RCC1 and Retinoblastoma Protein. Nature 2010, 466, 1125–1128.
  56. Bailey, A.O.; Panchenko, T.; Sathyan, K.M.; Petkowski, J.J.; Pai, P.-J.; Bai, D.L.; Russell, D.H.; Macara, I.G.; Shabanowitz, J.; Hunt, D.F.; et al. Posttranslational Modification of CENP-A Influences the Conformation of Centromeric Chromatin. Proc. Natl. Acad. Sci. USA 2013, 110, 11827–11832.
  57. Sathyan, K.M.; Fachinetti, D.; Foltz, D.R. α-Amino Trimethylation of CENP-A by NRMT Is Required for Full Recruitment of the Centromere. Nat. Commun. 2017, 8, 14678.
  58. Dai, X.; Otake, K.; You, C.; Cai, Q.; Wang, Z.; Masumoto, H.; Wang, Y. Identification of Novel α-n-Methylation of CENP-B That Regulates Its Binding to the Centromeric DNA. J. Proteome Res. 2013, 12, 4167–4175.
  59. Leszczyniecka, M.; Bhatia, U.; Cueto, M.; Nirmala, N.R.; Towbin, H.; Vattay, A.; Wang, B.; Zabludoff, S.; Phillips, P.E. MAP1D, a novel methionine aminopeptidase family member is overexpressed in colon cancer. Oncogene 2006, 25, 3471–3478.
  60. Randhawa, H.; Chikara, S.; Gehring, D.; Yildirim, T.; Menon, J.; Reindl, K.M. Overexpression of peptide deformylase in breast, colon, and lung cancers. BMC Cancer 2013, 13, 321.
  61. Udenwobele, D.I.; Su, R.-C.; Good, S.V.; Ball, T.B.; Shrivastav, S.V.; Shrivastav, A. Myristoylation: An Important Protein Modification in the Immune Response. Front. Immunol. 2017, 8, 751.
  62. Chan, X.W.; Wrenger, C.; Stahl, K.; Bergmann, B.; Winterberg, M.; Muller, I.B.; Saliba, K.J. Chemical and genetic validation of thiamine utilization as an antimalarial drug target. Nat. Commun. 2013, 4, 2060.
  63. Tan, Y.W.; Hong, W.J.; Chu, J.J. Inhibition of enterovirus VP4 myristoylation is a potential antiviral strategy for hand, foot and mouth disease. Antivir. Res. 2016, 133, 191–195.
  64. Ramljak, I.; Stanger, J.; Real-Hohn, A.; Dreier, D.; Wimmer, L.; Redlberger-Fritz, M.; Fischl, W.; Klingel, K.; Mihovilovic, M.D.; Blaas, D.; et al. Cellular N-myristoyltransferases play a crucial picornavirus genus-specific role in viral assembly, virion maturation, and infectivity. PLoS Pathog. 2018, 14, e1007203.
  65. Frearson, J.A.; Brand, S.; McElroy, S.P.; Cleghorn, L.A.; Smid, O.; Stojanovski, L.; Price, H.P.; Guther, M.L.; Torrie, L.S.; Robinson, D.A.; et al. N-myristoyltransferase inhibitors as new leads to treat sleeping sickness. Nature 2021, 464, 728–732.
  66. Wright, M.H.; Paape, D.; Price, H.P.; Smith, D.F.; Tate, E.W. Global Profiling and Inhibition of Protein Lipidation in Vector and Host Stages of the Sleeping Sickness Parasite Trypanosoma brucei. ACS Infect. Dis. 2016, 2, 427–441.
  67. Thinon, E.; Morales-Sanfrutos, J.; Mann, D.J.; Tate, E.W. N-Myristoyltransferase Inhibition Induces ER-Stress, Cell Cycle Arrest, and Apoptosis in Cancer Cells. ACS Chem. Biol. 2016, 11, 2165–2176.
  68. Beauchamp, E.; Yap, M.C.; Iyer, A.; Perinpanayagam, M.A.; Gamma, J.M.; Vincent, K.M.; Lakshmanan, M.; Raju, A.; Tergaonkar, V.; Tan, S.Y.; et al. Targeting N-myristoylation for therapy of B-cell lymphomas. Nat. Commun. 2020, 11, 5348.
  69. Mackey, J.R.; Lai, J.; Chauhan, U.; Beauchamp, E.; Dong, W.F.; Glubrecht, D.; Sim, Y.W.; Ghosh, S.; Bigras, G.; Lai, R.; et al. N-myristoyltransferase proteins in breast cancer: Prognostic relevance and validation as a new drug target. Breast Cancer Res. Treat. 2021, 186, 79–87.
  70. Weickert, M.; Dillberger, J.; Mackey, J.R.; Wyatt, P.; Gray, D.; Read, K.; Li, C.; Parenteau, A.; Berthiaume, L.G. Initial Characterization and Toxicology of an Nmt Inhibitor in Development for Hematologic Malignancies. Blood 2019, 134, 3362.
  71. Deng, S.; Marmorstein, R. Protein N-Terminal acetylation: Structural basis, mechanism, versatility, and regulation. Trends Biochem. Sci. 2020, 8, 005.
  72. Halliday, G.M.; Holton, J.L.; Revesz, T.; Dickson, D.W. Neuropathology underlying clinical variability in patients with synucleinopathies. Acta Neuropathol. 2011, 122, 187–204.
  73. Spillantini, M.G.; Crowther, R.A.; Jakes, R.; Hasegawa, M.; Goedert, M. alpha-Synuclein in filamentous inclusions of lewy bodies from Parkinson’s disease and dementia with lewy bodies. Proc. Natl. Acad. Sci. USA 1998, 95, 6469–6473.
  74. Dikiy, I.; Eliezer, D. N-terminal acetylation stabilizes N-terminal helicity in lipid- and micelle-bound a-synuclein and increases its affinity for physiological membranes. J. Biol. Chem. 2014, 289, 3652–3665.
  75. Watson, M.D.; Lee, J.C. N-Terminal acetylation affects a-Synuclein fibril polymorphism. Biochemistry 2019, 58, 3630–3633.
  76. Mason, R.J.; Paskins, A.R.; Dalton, C.F.; Smith, D.P. Copper binding and subsequent aggregation of a-Synuclein are modulated by N-Terminal acetylation and ablated by the H50Q missense mutation. Biochemistry 2016, 55, 4737–4741.
  77. Fernández, R.D.; Lucas, H.R. Mass spectrometry data confirming tetrameric a-synuclein N-terminal acetylation. Data Brief 2018, 20, 1686–1691.
  78. Iyer, A.; Roeters, S.J.; Schilderink, N.; Hommersom, B.; Heeren, R.M.; Woutersen, S.; Claessens, M.M.; Subramaniam, V. The impact of N-terminal acetylation of a-Synuclein on phospholipid membrane binding and fibril structure. J. Biol. Chem. 2016, 291, 21110–21122.
  79. Ju, J.; Chen, A.; Deng, Y.; Liu, M.; Wang, Y.; Wang, Y.; Nie, M.; Wang, C.; Ding, H.; Yao, B.; et al. NatD promotes lung cancer progression by preventing histone H4 serine phosphorylation to activate Slug expression. Nat. Commun. 2017, 8, 928.
  80. Liu, Z.; Liu, Y.; Wang, H.; Ge, X.; Jin, Q.; Ding, G.; Hu, Y.; Zhou, B.; Chen, Z.; Ge, X.; et al. Patt1, a novel protein acetyltransferase that is highly expressed in liver and downregulated in hepatocellular carcinoma, enhances apoptosis of hepatoma cells. Int. J. Biochem. Cell. Biol. 2009, 41, 2528–2537.
  81. Demetriadou, C.; Pavlou, D.; Mpekris, F.; Achilleos, C.; Stylianopoulos, T.; Zaravinos, A.; Papageorgis, P.; Kirmizis, A. NAA40 contributes to colorectal cancer growth by controlling PRMT5 expression. Cell Death Dis. 2019, 10, 236.
  82. Dai, X.; Wang, Y.; Fu, L.; Wang, Z.; Gan, N.; Cai, Q. α- N -Methylation of Damaged DNA-binding Protein 2 (DDB2) and Its Function in Nucleotide Excision Repair. J. Biol. Chem. 2014, 289, 16046–16056.
  83. Moser, J.; Volker, M.; Kool, H.; Alekseev, S.; Vrieling, H.; Yasui, A.; Van Zeeland, A.A.; Mullenders, L.H.F. The UV-damaged DNA binding protein mediates efficient targeting of the nucleotide excision repair complex to UV-induced photo lesions. DNA Repair 2005, 4, 571–582.
  84. Chen, T.; Muratore, T.L.; Schaner-Tooley, C.E.; Shabanowitz, J.; Hunt, D.F.; Macara, I.F. N-terminal α-methylation of RCC1 is necessary for stable chromatin association and normal mitosis. Nat. Cell Biol. 2007, 9, 596–603.
  85. Hitakomate, E.; Hood, F.E.; Sanderson, H.S.; Clarke, P.R. The methylated N-terminal tail of RCC1 is required for stabilisation of its interaction with chromatin by Ran in live cells. BMC Cell Biol. 2010, 11, 43–52.
  86. Hao, Y.; Macara, I.G. Regulation of chromatin binding by a conformational switch in the tail of the Ran exchange factor RCC1. J. Cell Biol. 2008, 182, 827–836.
  87. Bade, D.; Cai, Q.; Li, L.; Yu, K.; Dai, X.; Miao, W.; Wang, Y. Modulation of N-Terminal Methyltransferase 1 by an N6-Methyladenosine-Based Epitranscriptomic Mechanism. Biochem. Biophys. Res. Commun. 2021, 546, 54–58.
  88. Dai, X.; Rulten, S.L.; You, C.; Caldecott, K.W.; Wang, Y. Identification and Functional Characterizations of N-Terminal α-NMethylation and Phosphorylation of Serine 461 in Human Poly(ADPRibose) Polymerase 3. J. Proteome Res. 2015, 14, 2575–2582.
  89. Nevitt, C.; Tooley, J.G.; Schaner Tooley, C.E. N-Terminal Acetylation and Methylation Differentially Affect the Function of MYL9. Biochem. J. 2018, 475, 3201–3219.
  90. Giglione, C.; Vallon, O.; Meinnel, T. Control of protein life-span by N-terminal methionine excision. EMBO J. 2003, 22, 13–23.
  91. Caglayan, A.O.; Aktar, F.; Bilguvar, K.; Baranoski, J.F.; Akgumus, G.T.; Harmanci, A.S.; Erson-Omay, E.Z.; Yasuno, K.; Caksen, H.; Gunel, M. MetAP1 mutation is a novel candidate for autosomal recessive intellectual disability. J. Hum. Genet. 2021, 66, 215–218.
  92. Cheng, H.-C.; Chi, S.-C.; Liang, C.-Y.; Yu, J.-Y.; Wang, A.-G. Candidate Modifier Genes for the Penetrance of Leber’s Hereditary Optic Neuropathy. Int. J. Mol. Sci. 2022, 23, 11891.
More
This entry is offline, you can click here to edit this entry!
Video Production Service