The Prokaryotic and Eukaryotic Nat Machinery: History
Please note this is an old version of this entry, which may differ significantly from the current revision.
Contributor: , ,

Prokaryotic and eukaryotic Nats belong to the general control non-repressible 5 (GCN5)-related N-acetyltransferases (GNAT) superfamily which counts thousands of members in all three domains of life. 

  • compartmentalization
  • co-translational modification
  • GNAT

1. Introduction: N-Terminal Acetylation—An Underestimated Protein Modification

Protein modifications are key modulators of protein fate and are often the first-aid tool for reprogramming cells in response to developmental or environmental cues. Together with phosphorylation and ubiquitination, acetylation is one of the most pervasive protein processing events [1]. Acetylation occurs at the α-amino group of protein N-termini (N-terminal acetylation, NTA) or at the ε-amino group of internal lysine residues (lysine acetylation, KA). Both NTA and KA are present throughout all kingdoms of life and are catalyzed by N-terminal acetyltransferases (Nats) or lysine acetyltransferases (Kats) which transfer acetyl moieties from acetyl coenzyme A (AcCoA) to their respective substrates. Prokaryotic and eukaryotic Nats belong to the general control non-repressible 5 (GCN5)-related N-acetyltransferases (GNAT) superfamily which counts thousands of members in all three domains of life [2][3][4]. Despite their low overall sequence homology (3–23%), the three-dimensional fold and catalytic domains of GNATs are well conserved (Figure 1A). The core GNAT fold consists of six to seven β-strands (β0–β6) and four α-helices (α1–α4). The loop connecting β4 and α3 harbors a highly conserved R/QxxGxA/G motif, which mediates AcCoA binding [2][5][6]. In higher eukaryotes, the bulk of cytosolic proteins (>80%) is co-translationally acetylated at their N-terminus, whereas KA affects selected proteins, most prominently histones [7][8]. While KA is widely recognized as transcriptional regulator, the overall biological significance of the more prevalent NTA remains unclear [9]. At the molecular level, NTA alters the electrostatic properties of proteins by neutralizing the positive charge at their N-terminus, which results in an increased overall hydrophobicity. In addition, NTA creates a new hydrogen bond acceptor and increases the nucleophilicity and basicity of the α-amine. Taken together, these changes have profound implications for the three-dimensional structure, activity, binding properties and lifetime of individual proteins [10]. Since up to date, no N-terminal deacetylases have been identified, these changes are considered irreversible [11][12]. Hence, NTA was for a long time perceived as a nonregulated, and consequently a static, co-translational process [13]. This dogma was challenged by the identification of regulatory mechanisms for Nats and a highly diversified family of post-translational Nats in higher eukaryotes [14][15][16][17][18][19]. Specifically, in plants, the characterization of plastid-localized GNAT proteins with dual Nat and Kat activity and the phytohormone-triggered regulation of the ribosome-tethered NatA contributed to this paradigm shift [20][21][22].
Figure 1. The typical GNAT fold is conserved throughout all domains of life. (A) The core GNAT fold consists of six to seven β-strands (β0–β6, light grey) and four α-helices (α1–α4, dark grey). The loop connecting β4 and α3 contains a conserved AcCoA binding motif (R/QxxGxA/G, red cross). Differences between GNAT structures are generally confined to the N-terminal β0 strand. (B) NTA frequency in different organisms as a percentage of the whole proteome. The bars represent the estimated upper limit reported for the individual organisms (1: [23], 2: [20], 3: [24], 4: [25], 5: [26], 6: [27], 7: [28], 8: [29] and 9: [30]).

2. The Prokaryotic Nat Machinery

While in humans and plants more than 80% of cytosolic proteins are N-terminally acetylated [20][23], the frequency of NTA declines in single-celled organisms (Figure 1B). In yeast for instance, only 60% of the proteome is N-terminally acetylated [15].
In bacteria, NTA is an even rarer event. Unlike eukaryotes, bacteria initiate protein biosynthesis with formylated methionine (fMet). Before NTA can occur, the N-terminal formyl group has to be removed co-translationally by peptide deformylase (PDF). For the majority (60%) of proteins, deformylation is followed by the excision of the initiator methionine (iMet) by methionine aminopeptidase (MetAP). Acetylation marks were found on both N-termini with and without iMet and are added by one of the three known bacterial acetyltransferases “Ribosomal modification I” (RimI), RimJ, and RimL [30][31]. Of these three enzymes, RimJ seems to be the most promiscuous since the number of N-terminally acetylated proteins in E. coli drops significantly upon depletion of RimJ, but not RimI or RimL. RimJ predominantly targets N-termini starting with Ser and Thr, but also Ala [32]. Despite their role as ribosome-assembly factors, Rims are absent from mature ribosomes, suggesting that their catalytic activity is purely post-translational [33].
Initially, only five endogenous proteins were reported to be N-terminally acetylated in Escherichia coli, including the ribosomal proteins S5, L7/L12, and S18 as well as the elongation factor EF-Tu and the chaperone SecB [26][34][35][36][37][38]. Recent mass spectrometry-based proteome-studies expanded this originally short list of N-terminally-acetylated proteins in E. coli to over 100 entries, accounting for 10% of the E. coli proteins with experimentally assessed acetylation status [30][32]. In Pseudomonas aeruginosa PA14 and Mycobacterium tuberculosis for instance, between 18 and 29% of the proteome were found to be N-terminally-acetylated (Figure 1B) [28][29].
Acetylation levels are similar in archaea, where 13–29% of all proteins are affected by NTA [26][27][39]. Archaea express a single conserved Nat, which exhibits a broad substrate specificity. The active site of this Nat is a hybrid of known eukaryotic Nat active sites [40][41], suggesting that the cytosolic Nats in eukaryotes derived from this ancestral form [42]. The function of NTA in archaea has only been demonstrated for individual proteins. In the salt-loving archaea Haloferax volcanii for instance, the NTA of the α1 proteasome subunit mediates the efficiency of proteolysis by altering the conformation of the channel leading up to the proteasomal core [43]. On the organismal level, the importance of NTA in archaea remains to be elucidated.

3. The Eukaryotic Nat Machinery

So far, six evolutionary conserved Nats (NatA-F) have been identified in metazoans (Figure 2). The existence of five of those (NatA-C and NatE-F) has been experimentally confirmed in the model plant A. thaliana [20][44][45][46][47][48]. NatD has been proposed to exist in Arabidopsis based on the substantial homology to its human orthologue [7]. Unlike NatD and NatF, most cytosolic Nats are composed of one catalytic and one or more auxiliary subunits facilitating ribosome association and catalytic properties [49]. While NatA–E are thought to be ribosome-bound in humans and plants, NatF localizes to the plasma membrane in plants and the Golgi-membrane in humans [14][46]. In addition, a family of plastid-localized Nats (GNAT1-7 and GNAT10) with dual Kat/Nat activity was recently characterized in A. thaliana [21][22].
Figure 2. Phylogenetic tree of Nats from different domains of life based on protein sequence comparison. Homologous Nat sequences from the photosynthetic eukaryotes Arabidopsis thaliana (At) and Oryza sativa (Os), the non-photosynthetic eukaryotes Homo sapiens (Hs), Drosophila melanogaster (Ds) and Saccharomyces cerevisiae (Sc), as well as the bacterium Escherichia coli (Ec) and the archaeon Saccharolobus solfataricus (Ss) were aligned with ClustalW. For OsNAA50 and OsNAA60, only one protein could be identified by blasting the respective human orthologs against the rice proteome. The resulting phylogenetic tree was circularized with the iTOL tool (https://itol.embl.de, accessed on 20 October 2022).
Nats are present in all plant organs. While NatA–E and the plastidic Nats are widely expressed in aereal organs except for the male reproductive parts, NatF is most strongly expressed in anther and pollen. Although the distribution of the plastidic Nats among different tissues is similar, there are differences between the transcription patterns of the individual enyzmes, indicating that they might fullfil different roles in specific organs. However, in specific organs, transcript levels of Nats barely change upon various biotic and abiotic stresses.
Furthermore, Nats may gain defined functions due to their specific subcellular compartments, which is summarized in Figure 3
Figure 3. Subcellular localization and substrate specificity of N-acetyltransferases in the model plant Arabidopsis thaliana. Catalytic subunits are schematically represented in red, whereas auxiliary subunits are depicted in orange. Subunits for which only predictions of subcellular localization are available are shown in lighter colors. From the plastid Nat family only NatG is shown for simplicity (1: [20][44][50][51][52]; 2: [47]; 3: [45]; 4: [46]; 5: [21][22], ?: debated in Arabidopsis). The pie chart shows the relative contribution of the individual acetyltransferases to the plant acetylome. Estimates are based on experimental data where acetyltransferases were assigned to acetylated N-termini based on their substrate specificity [20][53].
The substrate specificity of Nats is largely determined by the first two amino acids of their substrate proteins [11]. Consistent with the ability of Nats to acetylate distinct N-termini, the Nat catalytic sites differ in shape, size, and electrostatic properties (Figure 4). The catalytic mechanisms of AtNAA50 and AtNAA60 are very similar and rely on tyrosine and histidine residues that coordinate a catalytic water molecule [46][54]. Even though the catalytic mechanisms of AtNatA–NatC have not been uncovered yet, the residues required for catalysis in their human counterparts are conserved in plants [10].
Figure 4. Three-dimensional models of Arabidopsis thaliana Nats. The AcCoA-binding motives (A) of Arabidopsis Nats are strongly conserved (shown in red with conserved residues highlighted in ribbon mode). AcetylCoA is represented in grey. The Nat catalytic sites (B) have distinct surface characterizations in shape, size, and electrostatic properties, which is consistent with their ability to acetylate distinct substrate pools. Catalytically important residues were either reported in [1][2] for AtNatE and AtNatF or estimated based on their human and yeast counterparts for AtNatA–C [3] and are represented in stick mode. The crystal structures of AtNAA50 (6YZZ, green) and AtNAA60 (6TGX, cyan) were downloaded from the Protein Data Bank (https://www.rcsb.org, accessed on the 9 November 2022), whereas the three-dimensional structures of the other Nats were generated with SwissModel (https://swissmodel.expasy.org, accessed on the 9 November 2022) based on their human or yeast counterparts using the templates 6c9m.2.B (AtNAA10, blue), 7stx.1.A (AtNAA20, yellow) and 7l1k.1.A (AtNAA30, orange).
Interestingly, some proteins are not acetylated even though based on their primary sequence they fit the recognition potential of Nats. A search in the NterDB database (https://nterdb.i2bc.paris-saclay.fr/) reveals that of 1327 nuclear-encoded putative Arabidopsis NatA substrates 179 (14%) are not acetylated. Hence, substrate recognition might depend on so far unknown determinants. Those might include the three-dimensional properties of the nascent chain or competition of Nats with other ribosome-associated factors attracted by those nascent chains.

This entry is adapted from the peer-reviewed paper 10.3390/ijms232214492

References

  1. Mann, M.; Jensen, O.N. Proteomic analysis of post-translational modifications. Nat. Biotechnol. 2003, 21, 255–261.
  2. Favrot, L.; Blanchard, J.S.; Vergnolle, O. Bacterial GCN5-related N-acetyltransferases: From resistance to regulation. Biochem. Biophys. Res. Commun. 2016, 55, 989–1002.
  3. Krtenic, B.; Drazic, A.; Arnesen, T.; Reuter, N. Classification and phylogeny for the annotation of novel eukaryotic GNAT acetyltransferases. PLoS Comput. Biol. 2020, 16, e1007988.
  4. Friedmann, D.R.; Marmorstein, R. Structure and mechanism of non-histone protein acetyltransferase enzymes. FEBS Lett. 2013, 280, 5570–5581.
  5. Vetting, M.W.; de Carvalho, L.P.S.; Yu, M.; Hegde, S.S.; Magnet, S.; Roderick, S.L.; Blanchard, J.S. Structure and functions of the GNAT superfamily of acetyltransferases. Arch. Biochem. Biophys. 2005, 433, 212–226.
  6. Rathore, O.S.; Faustino, A.; Prudêncio, P.; Van Damme, P.; Cox, C.J.; Martinho, R.G. Absence of N-terminal acetyltransferase diversification during evolution of eukaryotic organisms. Sci. Rep. 2016, 6, 21304.
  7. Bienvenut, W.V.; Sumpton, D.; Martinez, A.; Lilla, S.; Espagne, C.; Meinnel, T.; Giglione, C. Comparative large scale characterization of plant versus mammal proteins reveals similar and idiosyncratic Nα-acetylation features. Mol. Cell. Proteom. 2012, 11, M111.015131.
  8. Aksnes, H.; Drazic, A.; Marie, M.; Arnesen, T. First things frst: Vital protein marks by N-terminal acetyltransferases. Trends Biochem. Sci. 2016, 41, 746–760.
  9. Chen, Y.C.; Koutelou, E.; Dent, S.Y.R. Now open: Evolving insights to the roles of lysine acetylation in chromatin organization and function. Mol. Cell 2022, 82, 716–727.
  10. Deng, S.; Marmorstein, R. Protein N-terminal acetylation: Structural basis, mechanism, versatility, and regulation. Trends Biochem. Sci. 2021, 46, 15–27.
  11. Ree, R.; Varland, S.; Arnesen, T. Spotlight on protein N-terminal acetylation. Exp. Mol. Med. 2018, 50, 1–13.
  12. Drazic, A.; Myklebust, L.M.; Ree, R.; Arnesen, T. The world of protein acetylation. Biochim. Biophys. Acta Proteins Proteom. 2016, 1864, 1372–1401.
  13. Arnesen, T. Towards a functional understanding of protein N-terminal acetylation. PLoS Biol. 2011, 9, e1001074.
  14. Aksnes, H.; Van Damme, P.; Goris, M.; Starheim, K.K.; Marie, M.; Støve, S.I.; Hoel, C.; Kalvik, T.V.; Hole, K.; Glomnes, N.; et al. An Organellar Nα-Acetyltransferase, Naa60, Acetylates Cytosolic N Termini of Transmembrane Proteins and Maintains Golgi Integrity. Cell Rep. 2015, 10, 1362–1374.
  15. Van Damme, P.; Hole, K.; Pimenta-Marques, A.; Helsens, K.; Vandekerckhove, J.; Martinho, R.G.; Gevaert, K.; Arnesen, T. NatF contributes to an evolutionary shift in protein N-terminal acetylation and is important for normal chromosome segregation. PLoS Genet. 2011, 7, e1002169.
  16. Vo, T.T.L.; Park, J.H.; Lee, E.J.; Nguyen, Y.T.K.; Han, B.W.; Nguyen, H.T.T.; Mun, K.C.; Ha, E.; Kwon, T.K.; Kim, K.W.; et al. Characterization of lysine acetyltransferase activity of recombinant human ARD1/NAA10. Molecules 2020, 25, 588.
  17. Kang, J.; Chun, Y.S.; Huh, J.; Park, J.W. FIH permits NAA10 to catalyze the oxygen-dependent lysyl-acetylation of HIF-1α. Redox Biol. 2018, 19, 364–374.
  18. Drazic, A.; Aksnes, H.; Marie, M.; Boczkowska, M.; Varland, S.; Timmerman, E.; Foyn, H.; Glomnes, N.; Rebowski, G.; Impens, F.; et al. NAA80 is actin’s N-terminal acetyltransferase and regulates cytoskeleton assembly and cell motility. Proc. Natl. Acad. Sci. USA 2018, 115, 4399–4404.
  19. Deng, S.; McTiernan, N.; Wei, X.; Arnesen, T.; Marmorstein, R. Molecular basis for N-terminal acetylation by human NatE and its modulation by HYPK. Nat. Commun. 2020, 11, 818.
  20. Linster, E.; Stephan, I.; Bienvenut, W.V.; Maple-Grødem, J.; Myklebust, L.M.; Huber, M.; Reichelt, M.; Sticht, C.; Geir Møller, S.; Meinnel, T.; et al. Downregulation of N-terminal acetylation triggers ABA-mediated drought responses in Arabidopsis. Nat. Commun. 2015, 6, 7640.
  21. Bienvenut, W.V.; Brünje, A.; Boyer, J.-B.; Mühlenbeck, J.S.; Bernal, G.; Lassowskat, I.; Dian, C.; Linster, E.; Dinh, T.V.; Koskela, M.M.; et al. Dual lysine and N-terminal acetyltransferases reveal the complexity underpinning protein acetylation. Mol. Syst. Biol. 2020, 16, e9464.
  22. Dinh, T.V.; Bienvenut, W.V.; Linster, E.; Feldman-Salit, A.; Jung, V.A.; Meinnel, T.; Hell, R.; Giglione, C.; Wirtz, M. Molecular identification and functional characterization of the first Nα-acetyltransferase in plastids by global acetylome profiling. Proteomics 2015, 15, 2426–2435.
  23. Arnesen, T.; Van Damme, P.; Polevoda, B.; Helsens, K.; Evjenth, R.; Colaert, N.; Varhaug, J.E.; Vandekerckhove, J.; Lillehaug, J.R.; Sherman, F.; et al. Proteomics analyses reveal the evolutionary conservation and divergence of N-terminal acetyltransferases from yeast and humans. Proc. Natl. Acad. Sci. USA 2009, 106, 8157–8162.
  24. Goetze, S.; Qeli, E.; Mosimann, C.; Staes, A.; Gerrits, B.; Roschitzki, B.; Mohanty, S.; Niederer, E.M.; Laczko, E.; Timmerman, E.; et al. Identification and functional characterization of N-terminally acetylated proteins in Drosophila melanogaster. PLoS Biol. 2009, 7, e1000236.
  25. Van Damme, P.; Evjenth, R.; Foyn, H.; Demeyer, K.; De Bock, P.-J.; Lillehaug, J.R.; Vandekerckhove, J.; Arnesen, T.; Gevaert, K. Proteome-derived peptide libraries allow detailed analysis of the substrate specificities of Nα-acetyltransferases and point to hNaa10p as the post-translational actin Nα-acetyltransferase. Mol. Cell. Proteom. 2011, 10, M110.004580.
  26. Kirkland, P.A.; Humbard, M.A.; Daniels, C.J.; Maupin-Furlow, J.A. Shotgun proteomics of the haloarchaeon Haloferax volcanii. J. Proteome Res. 2008, 7, 5033–5039.
  27. Falb, M.; Aivaliotis, M.; Garcia-Rizo, C.; Bisle, B.; Tebbe, A.; Klein, C.; Konstantinidis, K.; Siedler, F.; Pfeiffer, F.; Oesterhelt, D. Archaeal N-terminal protein maturation commonly involves N-terminal acetylation: A large-scale proteomics survey. J. Mol. Biol. 2006, 362, 915–924.
  28. Kelkar, D.S.; Kumar, D.; Kumar, P.; Balakrishnan, L.; Muthusamy, B.; Yadav, A.K.; Shrivastava, P.; Marimuthu, A.; Anand, S.; Sundaram, H.; et al. Proteogenomic analysis of Mycobacterium tuberculosis by high resolution mass spectrometry. Mol. Cell. Proteom. 2011, 10, M111.011627.
  29. Ouidir, T.; Jarnier, F.; Cosette, P.; Jouenne, T.; Hardouin, J. Characterization of N-terminal protein modifications in Pseudomonas aeruginosa PA14. Proteomics 2015, 114, 214–225.
  30. Bienvenut, W.V.; Giglione, C.; Meinnel, T. Proteome-wide analysis of the amino terminal status of Escherichia coli proteins at the steady-state and upon deformylation inhibition. Proteomics 2015, 15, 2503–2518.
  31. Solbiati, J.; Chapman-Smith, A.; Miller, J.L.; Miller, C.G.; Cronan, J.E. Processing of the N-termini of nascent polypeptide chains requires deformylation prior to methionine removal. J. Mol. Biol. 1999, 290, 607–614.
  32. Schmidt, A.; Kochanowski, K.; Vedelaar, S.; Ahrné, E.; Volkmer, B.; Callipo, L.; Knoops, K.; Bauer, M.; Aebersold, R.; Heinemann, M. The quantitative and condition-dependent Escherichia coli proteome. Nat. Biotechnol. 2016, 34, 104–110.
  33. Roy-Chaudhuri, B.; Kirthi, N.; Kelley, T.; Culver, G.M. Suppression of a cold-sensitive mutation in ribosomal protein S5 reveals a role for RimJ in ribosome biogenesis. Mol. Microbiol. 2008, 68, 1547–1559.
  34. Charbaut, E.; Redeker, V.; Rossier, J.; Sobel, A. N-terminal acetylation of ectopic recombinant proteins in Escherichia coli. FEBS Lett. 2002, 529, 341–345.
  35. Arai, K.; Clark, B.F.; Duffy, L.; Jones, M.D.; Kaziro, Y.; Laursen, R.A.; L’Italien, J.; Miller, D.L.; Nagarkatti, S.; Nakamura, S.; et al. Primary structure of elongation factor Tu from Escherichia coli. Proc. Natl. Acad. Sci. USA 1980, 77, 1326–1330.
  36. Yoshikawa, A.; Isono, S.; Sheback, A.; Isono, K. Cloning and nucleotide sequencing of the genes rimI and rimJ which encode enzymes acetylating ribosomal proteins S18 and S5 of Escherichia coli K12. Mol. Gen. Genet. 1987, 209, 481–488.
  37. Smith, V.F.; Schwartz, B.L.; Randall, L.L.; Smith, R.D. Electrospray mass spectrometric investigation of the chaperone SecB. Proc. Natl. Acad. Sci. USA 1996, 5, 488–494.
  38. Tanka, S.; Matsushita, Y.; Yoshikawa, A.; Isono, K. Cloning and molecular characterization of the gene RimL which encodes an enzyme acetylating ribosomal protein L12 of Escherichia coli K12. Mol. Gen. Genet. 1989, 217, 289–293.
  39. Aivaliotis, M.; Gevaert, K.; Falb, M.; Tebbe, A.; Konstantinidis, K.; Bisle, B.; Klein, C.; Martens, L.; Staes, A.; Timmerman, E.; et al. Large-scale identification of N-terminal peptides in the halophilic archaea Halobacterium salinarum and Natronomonas pharaonis. J. Proteome Res. 2007, 6, 2195–2204.
  40. Mackay, D.T.; Botting, C.H.; Taylor, G.L.; White, M.F. An acetylase with relaxed specificity catalyses protein N-terminal acetylation in Sulfolobus solfataricus. Mol. Microbiol. 2007, 64, 1540–1548.
  41. Liszczak, G.; Marmorstein, R. Implications for the evolution of eukaryotic amino-terminal acetyltransferase (NAT) enzymes from the structure of an archaeal ortholog. Proc. Natl. Acad. Sci. USA 2013, 110, 14652–14657.
  42. Hug, L.A.; Baker, B.J.; Anantharaman, K.; Brown, C.T.; Probst, A.J.; Castelle, C.J.; Butterfield, C.N.; Hernsdorf, A.W.; Amano, Y.; Ise, K.; et al. A new view of the tree of life. Nat. Microbiol. 2016, 1, 16048.
  43. Humbard, M.A.; Zhou, G.; Maupin-Furlow, J.A. The N-terminal penultimate residue of 20S proteasome α1 influences its Nα-acetylation and protein levels as well as growth rate and stress responses of Haloferax volcanii. J. Bacteriol. 2009, 191, 3794–3803.
  44. Armbruster, L.; Linster, E.; Boyer, J.-B.; Brünje, A.; Eirich, J.; Stephan, I.; Bienvenut, W.V.; Weidenhausen, J.; Meinnel, T.; Hell, R.; et al. NAA50 is an enzymatically active Nα-acetyltransferase that is crucial for development and regulation of stress responses. Plant Physiol. 2020, 183, 1502–1516.
  45. Pesaresi, P.; Gardner, N.A.; Masiero, S.; Dietzmann, A.; Eichacker, L.; Wickner, R.; Salamini, F.; Leister, D. Cytoplasmic N-terminal protein acetylation is required for efficient photosynthesis in Arabidopsis. Plant Cell 2003, 15, 1817–1832.
  46. Linster, E.; Layer, D.; Bienvenut, W.V.; Dinh, T.V.; Weyer, F.A.; Leemhuis, W.; Brünje, A.; Hoffrichter, M.; Miklankova, P.; Kopp, J.; et al. The Arabidopsis Nα-acetyltransferase NAA60 locates to the plasma membrane and is vital for the high salt stress response. New Phytol. 2020, 228, 554–569.
  47. Ferrández-Ayela, A.; Micol-Ponce, R.; Sánchez-García, A.B.; Alonso-Peral, M.M.; Micol, J.L.; Ponce, M.R. Mutation of an Arabidopsis NatB N-alpha-terminal acetylation complex component causes pleiotropic developmental defects. PLoS ONE 2013, 8, e80697.
  48. Huber, M.; Bienvenut, W.V.; Linster, E.; Stephan, I.; Armbruster, L.; Sticht, C.; Layer, D.; Lapouge, K.; Meinnel, T.; Sinning, I.; et al. NatB-mediated N-terminal acetylation affects growth and biotic stress responses. Plant Physiol. 2020, 182, 792–806.
  49. Aksnes, H.; Ree, R.; Arnesen, T. Co-translational, post-translational, and non-catalytic roles of N-terminal acetyltransferases. Mol. Cell 2019, 73, 1097–1114.
  50. Feng, J.; Li, R.; Yu, J.; Ma, S.; Wu, C.; Li, Y.; Cao, Y.; Ma, L. Protein N-terminal acetylation is required for embryogenesis in Arabidopsis. J. Exp. Bot. 2016, 67, 4779–4789.
  51. Chen, H.; Li, S.; Li, L.; Wu, W.; Ke, X.; Zou, W.; Zhao, J. Nα-acetyltransferases 10 and 15 are required for the correct initiation of endosperm cellularization in Arabidopsis. Plant Cell Physiol. 2018, 59, 2113–2128.
  52. Gong, X.; Huang, Y.; Liang, Y.; Yuan, Y.; Liu, Y.; Han, T.; Li, S.; Gao, H.; Lv, B.; Huang, X.; et al. OsHYPK-mediated protein N-terminal acetylation coordinates plant development and abiotic stress responses in rice. Mol. Plant 2022, 15, 740–754.
  53. Linster, E.; Wirtz, M. N-terminal acetylation: An essential protein modification emerges as an important regulator of stress responses. J. Exp. Bot. 2018, 69, 4555–4568.
  54. Weidenhausen, J.; Kopp, J.; Armbruster, L.; Wirtz, M.; Lapouge, K.; Sinning, I. Structural and functional characterization of the N-terminal acetyltransferase NAA50. Structure 2021, 29, 413–425.e415.
More
This entry is offline, you can click here to edit this entry!
ScholarVision Creations