Next Article in Journal
Cerebral Vasomotor Reactivity in COVID-19: A Narrative Review
Previous Article in Journal
Hemostatic Effects of Raloxifene in Ovariectomized Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Impact of Protein Nα-Modifications on Cellular Functions and Human Health

Edward A. Doisy Department of Biochemistry and Molecular Biology, Saint Louis University Medical School, Saint Louis, MO 63104, USA
Life 2023, 13(7), 1613; https://doi.org/10.3390/life13071613
Submission received: 21 June 2023 / Revised: 14 July 2023 / Accepted: 17 July 2023 / Published: 24 July 2023
(This article belongs to the Section Medical Research)

Abstract

:
Most human proteins are modified by enzymes that act on the α-amino group of a newly synthesized polypeptide. Methionine aminopeptidases can remove the initiator methionine and expose the second amino acid for further modification by enzymes responsible for myristoylation, acetylation, methylation, or other chemical reactions. Specific acetyltransferases can also modify the initiator methionine and sometimes the acetylated methionine can be removed, followed by further modifications. These modifications at the protein N-termini play critical roles in cellular protein localization, protein-protein interaction, protein-DNA interaction, and protein stability. Consequently, the dysregulation of these modifications could significantly change the development and progression status of certain human diseases. The focus of this review is to highlight recent progress in our understanding of the roles of these modifications in regulating protein functions and how these enzymes have been used as potential novel therapeutic targets for various human diseases.

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 (Figure 1). 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]. The focus of this review is to highlight recent progress in our understanding of the substrate specificity of these enzymes, their roles in the biological function of specific proteins, how they might be regulated, the crosstalk between different modifications, and how these enzymes are used as targets for potential novel therapeutical strategies.

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]. Eukaryotic type 2 MetAPs (MetAP2s), on the other hand, 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 is significantly overlapping for human MetAP1 and MetAP2 [15].
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 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 fully 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 we are still very early in understanding how mutations in MetAPs could affect human health, NME excision processes provide a promising avenue in translational research.

3. Nα-Myristoylation

N-myristoyltransferase (NMTs) are responsible for protein Nα-myristoylation. NMTs can transfer a C:14:0 acyl-CoA to the N-terminal glycine (Gly) of specific groups of proteins following the excision of iMet by MetAPs [93,94]. N-terminal Gly is absolutely required for NMT activity, with preferred substrates containing sequence: G2X3X4X5(C/S/T)6K7 [95,96,97,98] in which X3 favors a charged residue; X4 can be any residue; X5 favors Gly, Ala, Ser, Cys or Asn; X6 favors Cys, Ser, or Thr; whereas Trp, Phe, Tyr, and Pro are prohibited [97,98,99,100]; and K7 interacts with negatively charged residues in the binding pocket of MNTs [97,98,99,100]. The substrate specificity of MNTs might be regulated by their interacting partners in different cells and species [100]. There are variations in sequence preferences across species [96,98,100]. A tool predicting species-specific Nα-myristoylation before experimental data is available will help develop NMT-targeted therapy.
Most human tissues express 2 NMT isoenzymes (HsNMT1 and HsNMT2) [101,102,103,104,105]. These 2 isozymes share ~77% protein sequence identity. Although they share similar substrates, they are not considered functionally redundant [101,102,103,104,105,106,107,108,109]. At least 40 NMT substrates have been discovered in human cells. These proteins are usually inserted into the lipid rafts, plasma membrane, endoplasmic reticulum (ER), Golgi apparatus, nuclear membrane, or mitochondria in cells [109,110,111,112,113,114]. Thus, depending on the subcellular localization of the myristoylated protein, it can regulate diverse cellular functions [35,36,37,38,61], including signal transduction [36,37,38], cellular transformation [36,37,38], oncogenesis, both innate and adaptive immune responses, cancer, and human immunodeficiency virus (HIV) infection [61], as well as parasitic and fungal diseases [62,63]. Small-molecule NMT inhibitors have therapeutic potential in viral and parasitic infections and cancer. For example, IMP-366 (DDD85646), a bioactive pyrazole sulfonamide inhibitor of Trypanosoma brucei NMT with an apparent Ki value of 1.44 nM, is a widely used NMT inhibitor that can suppress picornavirus replication, as well as malaria and sleeping sickness parasites by inhibiting their NMTs [64,65,66]. IMP366 can also suppress breast and colon cancer cell growth [67]. PCLX-001 (DDD86481) is a potent, small molecule inhibitor of both human NMTs. Preclinical studies have shown that PCLX-001 markedly inhibits hematologic and lymphoma cell lines in tissue culture and achieves complete remissions in human cancers grown in immunodeficient mice [68] and tumor responses in solid cancers [68,69,70,115]. It can nullify Nα-myristoylation of Src family kinases and promote their degradation, leading to cancer cell death in vitro and xenograft models [68]. The molecule has been extensively investigated in non-clinical safety testing [70] and found suitable for formal drug development in humans. Recently, analysis of pharmacokinetic and pharmacodynamic endpoints revealed that PCLX-001 has pharmacokinetic properties suitable for continued development as an oral, once daily, cancer therapy. A more in-depth understanding of structural differences between human and pathogenic NMTs and human NMT1 and NMT2 will further advance the development of small molecules with increased selectivity and decreased toxicity.

4. Nα-Acetylation

N-terminal acetylome analysis revealed that ~80–90% of the N-terminal α-amino group of soluble human proteins is acetylated. At least nine different N-terminal acetyltransferases (NATs) are involved in this modification event in which acetyl-CoA is used as a cofactor for the chemical reaction (Figure 1) [115,116]. This modification replaces the positive charge associated with the free α-amino group with a polar group and can block it for further changes. Each NAT has different substrate specificities [117,118], which can affect critical protein functions, including complex formation [39,40,41,42], protein localization [43,44,45], and protein degradation governed by the N-end rules (Table 1) [46,47,48]. NATs usually contain a unique catalytic subunit and one or two auxiliary subunits. The auxiliary subunits play various roles, including ribosomal anchoring [119,120]. In humans, 7 NATs have been identified: NatA (NAA10 and NAA15), NatB (NAA20 and NAA25), NatC (NAA 30, NAA 35, and NAA38), NatD (NAA40), NatE, NatF, and NatH (NAA80) [121,122,123]. NatA, NatB, and NatC are responsible for most protein Nα-acetylation [3]. NatA and NatB are heterodimeric complexes, each containing a specific catalytic subunit and a unique auxiliary subunit (NAA10 in NatA and NAA20 in NatB, as the catalytic subunit; NAA15 in NatA and NAA25 in NatB as the auxiliary subunit) [124]. NatC is a heterotrimeric complex.I It contains 1 catalytic subunit (NAA30) and 2 auxiliary subunits (NAA35 and NAA38) [125]. NatA, NatB, and NatC each have different substrate specificity. NatA acetylates small N-terminal residues after iMet excision [115], whereas NatB acetylates the iMet with sequences of MD- ME-, MN-, and MQ- [126,127]. NatC/E/F have overlapping substrates, acting on iMet when followed by residues that are not D, E, N, and Q [116,127,128,129,130,131]. When an additional catalytic subunit (NAA50) binds to NatA, NatE is formed as a dual enzyme complex with crosstalk between 2 catalytic subunits, NAA10 and NAA50 [71,121]. NatF binds to the Golgi membrane and acetylated transmembrane proteins [71,132,133]. NatD contains only a single catalytic unit, NAA40, with no auxiliary subunit. It acetylates the α-amino group of Ser of histones H4 and H2A after the exciton of iMet [134,135]. NatH targets all six mammalian actin isoforms in unique processing steps that differentiate muscle and non-muscle actins. Non-muscle actins contain a string of negatively charged residues following iMet, MDDD in β-actin, and MEEE in γ-actin. The iMet is acetylated by NatB and then removed by an unidentified actin N-acetyl-aminopeptidase (ANAP) [136]. Muscle actins, typically containing a Cys in the second position, are first processed by MetAPs to remove the iMet, followed by acetylation of the second Cys, likely by NatA, and the acetylated Cys will be further removed by ANAP to expose the acidic residue in the third position [136]. In all cases, this newly exposed acidic residue will be further acetylated post-translationally by NatH (Figure 2) [133]. Knockout of the NatH gene results in an increased filamentous to globular actin ratio, increased filopodia and lamellipodia formation, and accelerated cell motility [133].
Moreover, many studies have demonstrated that NatB and NatD have connections to various human diseases [71,72,73,74,75,76,77,78,79,80,81,137,138,139,140]. For example, NatB is upregulated in human hepatocellular carcinoma, and silencing NatB mRNA can block cell proliferation and tumor formation [139]. Thus, NatB could be a potential therapeutic target for certain cancers [139]. Nα-acetylation of α-Syn by NatB in human cells can increase its stability and lipid binding and reduce aggregation capacity [74,75,76,77,78]. Since α-Syn is a critical protein in Parkinson’s disease, NatB might play a role in PD pathogenesis (Table 2) [72,73]. A recent report on the Cro-EM structure of human NatB complexed with a CoA-α-Syn conjugate provided new insights into the mode of substrate selection of NAT enzymes, which will further facilitate the development of small molecule NatB probes [71]. NatD, on the other hand, plays essential roles in a diverse range of tumors, and its expression level correlates with poor survival of cancer patients [71,79]. However, unlike NatB, NatD is downregulated in hepatocellular carcinoma tissues, and ectopic NatD expression sensitizes hepatoma cancer cell lines to drug-induced apoptosis [54,71]. Moreover, a recent study has indicated that NatD is a critical regulator of cell invasion during lung cancer metastasis [139]. Interestingly, in colorectal cancer (CRC) cells, NatD plays a pro-survival role suggesting that it may stimulate cancer cell growth [71.81]. Despite the exciting discoveries of NatD’s essential roles in cancer development and metastasis, it remains elusive regarding its role in cancer chemotherapy response.
Figure 2. Nα-modifications of actins. (A) Muscle γ-actin and α-actins, typically containing a Cys in the second position followed by charged amino acid residues. After the iMet is removed by MetAPs, the second Cys, likely by NatA, is acetylated and the acetylated Cys will be further removed by ANAP to expose the acidic residue in the third position [136]. This newly exposed acidic residue will be further acetylated post-translationally by NatH. (B) Nonmuscle β-actin and γ-actin contain a string of negatively charged residues following iMet, MDDD in β-actin, and MEEE in γ-actin. The iMet is acetylated by NatB and then possibly removed by an unidentified actin N-acetyl-aminopeptidase (ANAP) [136]. Like its cytoplasmic partner γ-actin as discussed in Section 4, the iMet of β-actin acetylated by NatB, and then possibly removed by an unidentified ANAP is co-translationally. Next, the exposed second residue (Asp2 in β-actin, Glu2 in γ-actin) will be further acetylated by a dedicated N-acetyltransferase, NatH/NAA80 [141,142,143]. However, some β-actin Nα-termini will not be acetylated, instead they undergo further proteolytic processing, and the new Nα-termini (DD-) are then Nα-arginylated by ATE1. Nα-acetylation or Nα-arginylation of actins will change their N-terminal charge density and affect actin structure and function.
Figure 2. Nα-modifications of actins. (A) Muscle γ-actin and α-actins, typically containing a Cys in the second position followed by charged amino acid residues. After the iMet is removed by MetAPs, the second Cys, likely by NatA, is acetylated and the acetylated Cys will be further removed by ANAP to expose the acidic residue in the third position [136]. This newly exposed acidic residue will be further acetylated post-translationally by NatH. (B) Nonmuscle β-actin and γ-actin contain a string of negatively charged residues following iMet, MDDD in β-actin, and MEEE in γ-actin. The iMet is acetylated by NatB and then possibly removed by an unidentified actin N-acetyl-aminopeptidase (ANAP) [136]. Like its cytoplasmic partner γ-actin as discussed in Section 4, the iMet of β-actin acetylated by NatB, and then possibly removed by an unidentified ANAP is co-translationally. Next, the exposed second residue (Asp2 in β-actin, Glu2 in γ-actin) will be further acetylated by a dedicated N-acetyltransferase, NatH/NAA80 [141,142,143]. However, some β-actin Nα-termini will not be acetylated, instead they undergo further proteolytic processing, and the new Nα-termini (DD-) are then Nα-arginylated by ATE1. Nα-acetylation or Nα-arginylation of actins will change their N-terminal charge density and affect actin structure and function.
Life 13 01613 g002

5. Nα-Methylation

N-terminal methyltransferases catalyze protein N-terminal methylation (Nα-methylation). These enzymes are conserved between yeast (Tae1) and humans (NTMT1 and NTMT2). They catalyze the transfer of the methyl group from S-adenylmethionine (SAM) to the free α-amino group of the newly exposed X2 residue of a nascent polypeptide after iMet excision that contains a sequence motif, iMet-X2-P3-[K/R]4, in which X2 is A, S, G, or P (Figure 1). This motif is recognized as the canonical N-terminal motif for NTMTs [49,50]. The properties of Nα-methylated proteins differ according to the degree of methylation. Adding one methyl group to the α-amino group only slightly increases its basicity and introduces a minor steric hindrance that may slightly reduce its reactivity. However, adding two or three methyl groups can generate a permanent positive charge in the α-amino group. Until now, no eraser of this event has been identified, and protein Nα-methylation is considered irreversible. The two human NTMTs target various substrates associated with diverse biological pathways (Table 1) [50,51,52,53,54,55,56,57,58]. Nα-methylation regulates protein–protein and protein–DNA interactions [49,50,51]. For example, Nα-methylation (trimethylation) of CENP-A is critical for its formation of the centromere complex with two other partners, CENP-I and CENP-T, which is essential for cell cycle progress and cell survival [56,57]. On the other hand, the loss of Nα-trimethylation of CENP-B prevents it from binding to the centromeric DNA motif [56,57,58]. In HEK293T cells, Nα-methylation of DDB2 promotes its nuclear localization to UV light-induced cyclobutane pyrimidine dimer (CPD) foci and stimulates CPD repair, suggesting Nα-methylation’s protective role against UV damage [82,83]. Nα-methylation of MYL9, a transcriptional activator of intercellular adhesion molecule 1 (ICAM1), weakens its interaction with an actin-modulating protein, Cofilin-1, and promotes ICAM1 transcription [83]. Nα-methylation of RCC1 can regulate the RCC1-chromatin interaction by inhibiting the association of its core portion with histones H2A or H2B [84]. The binding of Ran to Nα-methylated RCC1 triggers the exposure of its histone-binding surface and promotes the interaction between the Nα-methylated RCC1 tail and negatively charged DNA [84,85,86]. This electrostatic interaction is Nα-methylation dependent. The loss of Nα-methylation reduces RCC1 binding to DNA and causes mitotic defects [85]. Interestingly, Nα-methylation of MRG15 was recently found to be modulated by m6A writers, leading to new regulation of Nα-methylation by the m6A-based epitranscriptome [86]. In summary, increasing evidence indicates that Nα-methylation is essential in regulating mitosis, chromatin interactions, DNA repair, tRNA transport, and maintaining genome stability (Table 1) [82,83,84,85,86,87,88,89]. Dysregulation of NTMTs has been implicated in the pathogenesis of various diseases, including breast, colorectal, pancreatic, and lung cancers (Table 2) [144,145].

6. Other Nα-Modifications

6.1. Nα-Palmitoylation

Unlike Nα-myristoylation, much less is known about Nα-palmitoylation. Protein palmitoylation usually occurs at an internal Cys residue [146], but researchers have recently identified several Nα-palmitoylated proteins. For example, a palmitoyl group is found to be attached to the α-amino group of the N-terminal Gly residue of the α-subunit of the heterotrimeric G protein that is responsible for the activation of adenylyl cyclase [146,147]. In addition, the secreted vertebrate signaling proteins Hedgehog (Hh) and Sonic Hedgehog (Shh) are Nα-palmitoylated at the Cys residue after the cleavage of the N-terminal signal sequence [148,149]. Hedgehog protein acyltransferase (Hhat) is suggested to be responsible for palmitoylating Shh [150,151,152,153]. This modification regulates Shh signaling strength [150,151,152,153,154,155]. That belongs to the family of transmembrane proteins termed MBOAT (membrane-bound O-acyltransferase) [154] which acylates Shh during its passage through the secretory pathway [150].

6.2. Nα-Ubiquitylation

Protein ubiquitylation usually refers to the addition of ubiquitin to the ε-amino group of an internal Lys residue via a combined activity of ubiquitin-activating (E1), conjugating (E2), and ligating (E3) enzymes. However, protein Nα-ubiquitylation refers to adding ubiquitin to the newly exposed α-amino group of a protein. In both cases, the ubiquityl group may serve as a target for polyubiquitylation, a well-known degradation signal recognized by the proteasome [156,157]. Nα-ubiquitylation was first discovered by Ciechanover’s lab [158]. However, the first direct evidence revealed by MS analysis was the Nα-ubiquitylation of HPV-58 oncoprotein E7 [7,159]. As HPV-58 E7 contains no lysine residues, its degradation is likely solely dependent upon Nα-ubiquitylation. Recently, α-synuclein and a tau tetra repeat domain were found to be Nα-ubiquitylated in vitro. Nα-ubiquitylation affected its aggregation properties and was proposed to enable targeting of the modified α-synuclein, and a tau tetra repeat domain to the proteasome for degradation, suggesting a role of Nα-ubiquitylation in removing amyloidogenic proteins [160]. A total of 2 enzymes, E3 HUWE1 and E2 Ube2w, catalyze Nα-ubiquitylation [161,162,163]. HUWE1 was shown to ubiquitylate the N-terminus of a MyoD mutant that contains no Lysine residues [163]. Ube2w, on the other hand, can successfully ubiquitylate the N-terminus of a lysine-free version of Ataxin-3 and Tau [161]. There are some distinctive differences when comparing the active site of Ube2w to that of classical E2s. The unique structure features of Ube2w make its novel active site more suitable to accommodate an α-amino group rather than a Lys side chain [161,162]. Interestingly, a recent report found that Pro in positions 2 to 4 of an unstructured peptide backbone has an inhibitory effect on Ube2w activity [164]. We are still very early in studying the biochemistry and biology of protein Nα-ubiquitylation.

6.3. Nα-Arginylation

Protein arginylation was discovered in 1963 [165,166]. The enzyme arginyltransferase (ATE1) responsible for this modification was first cloned and characterized in yeast [167]. ATE1 catalyzes the transfer of Arg from aminoacyl-tRNA to target proteins post-translationally [167]. The ATE1 gene exists in nearly all eukaryotes. The ATE1 gene encodes four isoforms in humans and mice, generated by alternative splicing [168,169]. ATE1 preferentially targets the unacetylated acidic N-terminal residues, including Asp and Glu [170]. It has also been found to target oxidized Cys [171,172] at a far lesser frequency and targeting oxidized Cys was found to differ between different ATE1 isoforms [173,174]. Recent high throughput analysis of arginylation suggested the existence of a consensus motif that may potentially be used to predict arginylation sites in vivo [173,174]. ATE1-mediated Nα-arginylation has been initially characterized as part of the N-degron (N-end rule) pathway that regulates the protein’s half-life [175,176]. N-terminally arginylated proteins can be recognized by specific E3 ligases of the ubiquitin–proteasome system (UPS) to ubiquitinate a nearby Lys for the follow-up degradation. However, if there is no accessible Lys for the E3 ligases, the N-terminally arginylated proteins will remain metabolically stable. Calreticulin [177] is one of ATE1 target. Nα-arginylation of calreticulin is less susceptible to proteasomal degradation than the non-arginylated form [178], and the modification triggers its translocation from the ER into the cytosol, increasing apoptotic response [179]. Actin is another known target of ATE1. Nα-arginylation or Nα-acetylation of cytoplasmic β-actin is emerging as a first-line mechanism to regulate cell migration [180,181,182]. Like its cytoplasmic partner γ-actin as discussed in Section 4, the iMet of β-actin acetylated by NatB, and then removed by an unidentified ANAP is co-translationally. Next, the exposed second residue (Asp2 in β-actin, Glu2 in γ-actin) will be further acetylated by a dedicated N-acetyltransferase, NatH/NAA80 [141,142,143]. The acetylated Asp2 or Glu2 can be removed, and then the Asp3 or Glu 3 can be arginylated by the nonspecific arginyltransferase Ate1 (Figure 2) [173,183]. Arginylation of γ-actin with slower translation leads to its immediate proteasomal degradation [184]. However, arginylated β-actin (R-actin) has been shown to specifically relocate to the leading edge upon induction of cell migration (Figure 2) [181]. NatH knockout (KO) cells show an increase in R-actin level by seven-fold [173,184,185], which supports the hypothesis that acetylation and arginylation of β-actin are mutually exclusive and that the increased level of R-actin could be an essential factor for the enhanced motility of NatH KO cells. However, it remains unclear whether the impact of Nα-modification on actin directly changes its interactions with the associated proteins or indirectly affects its interaction with the associated proteins via altered intrinsic interactions between actin monomers within the actin filament [185]. Recently, Nguyen KT et al. found that ubiquitin is a target of Ate1 by identifying the arginylated ubiquitin (RE-Ub) in yeast [186]. Ubiquitin (Ub) starts with Met-Gln-Ile (MQ) N-terminus. The authors proposed that N-terminal modifications of mammalian ubiquitin based on their new findings involve NME by unknown MetAP, N-terminal deamination by NTAQ1 N-terminal Gln amidase, and N-terminal arginylation by ATE1 arginyltransferase [187,188,189] ubiquitin might be processed by the NME-provoked cascade reactions of N-terminal deamination and N-terminal arginylation to yield RE-Ub (Figure 3) [186]. However, according to the specificity of NatB, iMet is likely to be acetylated. If so, the acetylated iMet will be removed by an unknown deacetylase. Then, the newly exposed Gln is converted to Glu by NTAQ1 N-terminal Gln amidase, followed by Nα-arginylation by ATE1 (Figure 3). Further studies of this exciting N-terminal modification of selected targets will eventually unravel the full complexity of the N-terminal arginylome and the biological significance of this event.

7. Regulation and Crosstalk

The substrate specificity and catalytic efficiency of MetAP2, but not MetAP1, are regulated by an allosteric disulfide bond, Cys228-Cys448, located at the rim of the active site. Oxidized and reduced isoforms of MetAP2 have different catalytic activities on their peptide substrates [190]. When solid tumor cells adapt to a limiting blood supply, they experience different degrees of stress, such as hypoxia, glucose deprivation, and growth factor. This cellular stress can lead to increased production of reactive oxygen species that may lead to an increased level of oxidized MetAP2 with altered substrate specificity [190]. Since excision of iMet is a prerequisite for specific Nα-modifications such as Nα-myristoylation, Nα-methylation, and Nα-acetylation events via NatA, altered MetAP activity could indirectly affect the downstream Nα-modifications. Nα-modifications, except the abovementioned possible indirect regulation by MetAP activity, are generally considered static because no specific modification eraser(s) have been discovered yet. Many recent reports indicate that dynamic crosstalk occurs between Nα-methylation and other modifications, including Nα-acetylation, internal methylation on lysine or arginine, phosphorylation, and m6A modifications in RNA [79]. Knockdown of NatD reduced metastasis and invasion of lung cancer cells. The proposed mechanism was that Nα-acetylation of histone H4 antagonizes the CK2α-mediated phosphorylation on the same serine residue (H4S1ph) to regulate Slug expression and metastasis [79]. The differential impact of Nα-methylation and Nα-acetylation on the subcellular localization of MYL9 is the first report on the interplay between methylation and acetylation at the same site [89,191,192]. Both Nα-methylation and Ser phosphorylation on the N-terminal tail of RCC1 were concurrent during mitosis [134,193]. In asynchronous HeLa cells, S1 phosphorylation decreased by about 25% in the RCC1 Nα methylation-defective mutant compared with wild-type RCC1, suggesting that Nα-methylation has a positive effect on the phosphorylation of S1 [134]. In mitotic cells, no significant change was observed in the total phosphorylation levels of two Ser residues (S1 and S10) regardless of Nα-methylation. However, the phosphorylation level of S2 increased by 10% in the absence of Nα-methylation [134]. Furthermore, a recent study demonstrated that NTMT1 protein expression might be regulated by readers, writers, and erasers involved in the N6-methyladenosine (m6A) modification of mRNA, providing a critical first piece of evidence for the regulation of Nα-methylation [134,193], which introduces additional dimensions that govern the interplay among different modifications at the Nα-position.

8. Conclusions

Our understanding of protein Nα-modifications has been significantly advanced in the past decades. Protein Nα-modifications are critical in controlling protein/protein interaction, protein-DNA interaction, cellular protein localization, and protein stability. The dysfunction or dysregulation of the enzymes involved in Nα-modification has been connected to various human diseases, including cancer, neurodegenerative diseases, and infectious diseases. Therefore, many of these enzymes have been used as targets for developing novel therapeutic approaches for disease treatment, including cancer, diabetes, and obesity. We can anticipate that scientists using cutting-edge technologies such as cryo-EM and multi-omics approaches will make significant advances in this exciting field in the future.

Funding

This work was supported by the National Institute of Health (Grant number: HHSN272201300021I) and the National Science Foundation (Grant number: MCB9512655).

Data Availability Statement

No new data were created.

Conflicts of Interest

The author declares no conflict of interest.

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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Sherman, F.; Stewart, J.W.; Tsunasawa, S. Methionine or not methionine at the beginning of a protein. Bioessays 1985, 3, 27–31. [Google Scholar] [CrossRef] [PubMed]
  3. Starheim, K.K.; Gevaert, K.; Arnesen, T. Protein N-terminal acetyltransferases: When the start matters. Trends Biochem. Sci. 2012, 37, 152–161. [Google Scholar] [CrossRef] [PubMed]
  4. Martin, D.D.; Beauchamp, E.; Berthiaume, L.G. Posttranslational myristoylation: Fat matters in cellular life and death. Biochimie 2011, 93, 18–31. [Google Scholar] [CrossRef] [PubMed]
  5. Stock, A.; Clarke, S.; Clarke, C.; Stock, J. N-Terminal Methylation of Proteins—Structure, Function and Specificity. FEBS Lett. 1987, 220, 8–14. [Google Scholar] [CrossRef] [Green Version]
  6. Buglino, J.A.; Resh, M.D. Palmitoylation of Hedgehog proteins. Vitam. Horm. 2012, 88, 229–252. [Google Scholar]
  7. Ciechanover, A.; Ben-Saadon, R. N-terminal ubiquitination: More protein substrates join in. Trends Cell Biol. 2004, 14, 103–106. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [Green Version]
  9. Chen, L.; Kashina, A. Post-translational Modifications of the Protein Termini. Front. Cell. Dev. Biol. 2021, 9, 719590. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [PubMed]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  36. Boutin, J.A. Myristoylation. Cell Signal 1997, 9, 15–35. [Google Scholar] [CrossRef] [PubMed]
  37. Taniguchi, H. Protein myristoylation in protein-lipid and protein-protein interactions. Biophys. Chem. 1999, 82, 129–137. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef]
  45. Dikic, I.; Elazar, Z. Mechanism and medical implications of mammalian autophagy. Nat. Rev. Mol. Cell Biol. 2018, 19, 349–364. [Google Scholar] [CrossRef]
  46. Hwang, C.S.; Shemorry, A.; Varshavsky, A. N-terminal acetylation of cellular proteins creates specific degradation signals. Science 2010, 327, 973–977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  50. Huang, R. Chemical Biology of Protein N-Terminal Methyltransferases. ChemBioChem 2019, 20, 976–984. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  71. Deng, S.; Marmorstein, R. Protein N-Terminal acetylation: Structural basis, mechanism, versatility, and regulation. Trends Biochem. Sci. 2020, 8, 005. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Watson, M.D.; Lee, J.C. N-Terminal acetylation affects a-Synuclein fibril polymorphism. Biochemistry 2019, 58, 3630–3633. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  77. Fernández, R.D.; Lucas, H.R. Mass spectrometry data confirming tetrameric a-synuclein N-terminal acetylation. Data Brief 2018, 20, 1686–1691. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  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. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  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. [Google Scholar] [CrossRef]
  90. Giglione, C.; Vallon, O.; Meinnel, T. Control of protein life-span by N-terminal methionine excision. EMBO J. 2003, 22, 13–23. [Google Scholar] [CrossRef] [Green Version]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  93. Bhatnagar, R.S.; Ashrafi, K.; Futterer, K.; Waksman, G.; Gordon, J.I. Biology and Enzymology of Protein N-Myristoylation. In Protein Lipidation; Tamanoi, F., Sigman, D.S., Eds.; Academic Press: Cambridge, MA, USA, 2001; Volume XXI, pp. 241–286. [Google Scholar]
  94. Thinon, E.; Serwa, R.A.; Broncel, M.; Brannigan, J.A.; Brassat, U.; Wright, M.H.; Heal, W.P.; Wilkinson, A.J.; Mann, D.J.; Tate, E.W. Global profiling of co- and post-translationally N-myristoylated proteomes in human cells. Nat. Commun. 2014, 5, 4919. [Google Scholar] [CrossRef] [Green Version]
  95. Jiang, H.; Zhang, X.; Chen, X.; Aramsangtienchai, P.; Tong, Z.; Lin, H. Protein Lipidation: Occurrence, Mechanisms, Biological Functions, and Enabling Technologies. Chem. Rev. 2018, 118, 919–988. [Google Scholar] [CrossRef]
  96. Castrec, B.; Dian, C.; Ciccone, S.; Ebert, C.L.; Bienvenut, W.V.; Le Caer, J.P.; Steyaert, J.M.; Giglione, C.; Meinnel, T. Structural and genomic decoding of human and plant myristoylomes reveals a definitive recognition pattern. Nat. Chem. Biol. 2018, 14, 671–679. [Google Scholar] [CrossRef] [PubMed]
  97. Towler, D.A.; Adams, S.P.; Eubanks, S.R.; Towery, D.S.; Jackson-Machelski, E.; Glaser, L.; Gordon, J.I. Purification and characterization of yeast myristoyl CoA:protein N-myristoyltransferase. Proc. Natl. Acad. Sci. USA 1987, 84, 2708–2712. [Google Scholar] [CrossRef] [PubMed]
  98. Towler, D.A.; Adams, S.P.; Eubanks, S.R.; Towery, D.S.; Jackson-Machelski, E.; Glaser, L.; Gordon, J.I. Myristoyl CoA:protein N-myristoyltransferase activities from rat liver and yeast possess overlapping yet distinct peptide substrate specificities. J. Biol. Chem. 1988, 263, 1784–1790. [Google Scholar] [CrossRef] [PubMed]
  99. Kosciuk, T.; Lin, H. NMT as a glycine and lysine myristoyltransferase in cancer, immunity, and infections. ACS Chem. Biol. 2020, 15, 1747–1758. [Google Scholar] [CrossRef]
  100. Traverso, J.A.; Giglione, C.; Meinnel, T. High-throughput profiling of N-myristoylation substrate specificity across species including pathogens. Proteomics 2013, 13, 25–36. [Google Scholar] [CrossRef]
  101. Glover, C.J.; Hartman, K.D.; Felsted, R.L. Human N-myristoyltransferase amino-terminal domain involved in targeting the enzyme to the ribosomal subcellular fraction. J. Biol. Chem. 1997, 272, 28680–28689. [Google Scholar] [CrossRef] [Green Version]
  102. Duronio, R.J.; Reed, S.I.; Gordon, J.I. Mutations of human myristoyl-CoA:protein N-myristoyltransferase cause temperature-sensitive myristic acid auxotrophy in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA 1992, 89, 4129–4133. [Google Scholar] [CrossRef]
  103. Giang, D.K.; Cravatt, B.F. A second mammalian N-myristoyltransferase. J. Biol. Chem. 1998, 273, 6595–6598. [Google Scholar] [CrossRef] [Green Version]
  104. Weston, S.A.; Camble, R.; Colls, J.; Rosenbrock, G.; Taylor, I.; Egerton, E.; Tucker, A.D.; Tunnicliffe, A.; Mistry, A.; Mancia, F.; et al. Crystal structure of the anti-fungal target N-myristoyl transferase. Nat. Struct. Biol. 1998, 5, 213–221. [Google Scholar] [CrossRef] [PubMed]
  105. Farazi, T.A.; Manchester, J.K.; Gordon, J.I. Transient-state kinetic analysis of Saccharomyces cerevisiae myristoyl CoA:protein N-myristoyltransferase reveals that a step after chemical transformation is rate limiting. Biochemistry 2000, 39, 15807–15816. [Google Scholar] [CrossRef] [PubMed]
  106. Ducker, C.E.; Upson, J.J.; French, K.J.; Smith, C.D. Two N-myristoyltransferase isozymes play unique roles in protein myristoylation, proliferation, and apoptosis. Mol. Cancer Res. 2005, 3, 463–476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Yang, S.H.; Shrivastav, A.; Kosinski, C.; Sharma, R.K.; Chen, M.-H.; Berthiaume, L.G.; Peters, L.L.; Chuang, P.T.; Young, S.G.; Bergo, M.O. N-myristoyltransferase 1 is essential in early mouse development. J. Biol. Chem. 2005, 280, 18990–18995. [Google Scholar] [CrossRef] [Green Version]
  108. Rioux, V.; Beauchamp, E.; Pedrono, F.; Daval, S.; Molle, D.; Catheline, D.; Legrand, P. Identification and characterization of recombinant and native rat myristoyl-CoA: Protein N-myristoyltransferases. Mol. Cell. Biochem. 2006, 286, 161–170. [Google Scholar] [CrossRef]
  109. Podell, S.; Gribskov, M. Predicting N-terminal myristoylation sites in plant proteins. BMC Genom. 2004, 5, 37. [Google Scholar] [CrossRef] [Green Version]
  110. Sugii, M.; Okada, R.; Matsuno, H.; Miyano, S. Performance improvement in protein N-myristoyl classification by BONSAI with insignificant indexing symbol. Genome Inform. Int. Conf. Genome Inform. 2007, 18, 277–286. [Google Scholar]
  111. Maurer-Stroh, S.; Eisenhaber, B.; Eisenhaber, F. N-terminal N-myristoylation of proteins: Refinement of the sequence motif and its taxon-specific differences. J. Mol. Biol. 2002, 317, 523–540. [Google Scholar] [CrossRef]
  112. Giglione, C.; Fieulaine, S.; Meinnel, T. N-terminal protein modifications: Bringing back into play the ribosome. Biochimie 2015, 114, 134–146. [Google Scholar] [CrossRef]
  113. Maurer-Stroh, S.; Eisenhaber, B.; Eisenhaber, F. N-terminal N-myristoylation of proteins: Prediction of substrate proteins from amino acid sequence. J. Mol. Biol. 2002, 317, 541–557. [Google Scholar] [CrossRef]
  114. Boisson, B.; Giglione, C.; Meinnel, T. Unexpected protein families including cell defense components feature in the N-myristoylome of a higher eukaryote. J. Biol. Chem. 2003, 278, 43418–43429. [Google Scholar] [CrossRef] [Green Version]
  115. Arnesen, T.; Van Damme, P.; Polevoda, B.; Helsens, K.; Evjenth, R.; Colaert, K.; 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. [Google Scholar] [CrossRef] [PubMed]
  116. 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. [Google Scholar] [CrossRef] [Green Version]
  117. Neuwald, A.F.; Landsman, D. GCN5-related histone Nacetyltransferases belong to a diverse superfamily that includes the yeast SPT10 protein. Trends Biochem. Sci. 1997, 22, 154–155. [Google Scholar] [CrossRef] [PubMed]
  118. Polevoda, B.; Norbeck, J.; Takakura, H.; Blomberg, A.; Sherman, F. Identification and specificities of N-terminal acetyltransferases from Saccharomyces cerevisiae. EMBO J. 1999, 18, 6155–6168. [Google Scholar] [CrossRef] [PubMed]
  119. Gautschi, M.; Just, S.; Mun, A.; Ross, S.; Rücknagel, P.; Dubaquié, Y.; Ehrenhofer-Murray, A.; Rospert, S. The yeast N(alpha)-acetyltransferase NatA is quantitatively anchored to the ribosome and interacts with nascent polypeptides. Mol. Cell. Biol. 2003, 23, 7403–7414. [Google Scholar] [CrossRef] [Green Version]
  120. Polevoda, B.; Brown, S.; Cardillo, T.S.; Rigby, S.; Sherman, F. Yeast N(alpha)-terminal acetyltransferases are associated with ribosomes. J. Cell. Biochem. 2008, 103, 492–508. [Google Scholar] [CrossRef]
  121. Aksnes, H.; Drazic, A.; Marie, M.; Arnesen, T. First Things First: Vital Protein Marks by N-Terminal Acetyltransferases. Trends Biochem. Sci. 2016, 41, 746–760. [Google Scholar] [CrossRef] [Green Version]
  122. Aksnes, H.; Marie, M.; Arnesen, T.; Drazic, A. Actin polymerization and cell motility are affected by NAA80-mediated posttranslational N-terminal acetylation of actin. Commun. Integr. Biol. 2018, 11, e1526572. [Google Scholar] [CrossRef] [Green Version]
  123. Aksnes, H.; Ree, R.; Arnesen, T. Co-translational, Post-translational, and Non-catalytic Roles of N-Terminal Acetyltransferases. Mol. Cell 2019, 73, 1097–1114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Arnesen, T.; Anderson, D.; Baldersheim, C.; Lanotte, M.; Varhaug, J.E.; Lillehaug, J.R. Identification and characterization of the human ARD1-NATH protein acetyltransferase complex. Biochem. J. 2015, 386, 433–443. [Google Scholar] [CrossRef] [Green Version]
  125. Stalheims, K.K.; Gromyko, D.; Evjenth, R.; Ryningen, A.; Varhaug, J.E.; Lillehaug, J.R.; Arnesen, T. Knockdown of human N alpha-terminal acetyltransferase complex C leads to p53-dependent apoptosis and aberrant human Arl8b localization. Mol. Cell Biol. 2009, 29, 3569–3581. [Google Scholar]
  126. Van Damme, P.; Lasa, M.; Polevoda, B.; Gazquez, C.; Elosegui-Artola, A.; Kim, D.S.; De Juan-Pardo, E.; Demeyer, K.; Hole, K.; Larrea, E.; et al. N-terminal acetylome analyses and functional insights of the N-terminal acetyltransferase NatB. Proc. Natl. Acad. Sci. USA 2012, 109, 12449–12454. [Google Scholar] [CrossRef] [PubMed]
  127. Van Damme, P.; Kalvik, T.V.; Starheim, K.K.; Jonckheere, V.; Myklebust, L.M.; Menschaert, G.; Varhaug, J.E.; Gevaert, K.; Arnesen, T. A Role for Human N-alpha Acetyltransferase 30 (Naa30) in Maintaining Mitochondrial Integrity. Mol. Cell. Proteom. 2016, 15, 3361–3372. [Google Scholar] [CrossRef] [Green Version]
  128. Tercero, J.C.; Dinman, J.D.; Wickner, R.B. Yeast MAK3 N-acetyltransferase recognizes the N-terminal four amino acids of the major coat protein (gag) of the L-A double-stranded RNA virus. J. Bacteriol. 1993, 175, 3192–3194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Evjenth, R.; Hole, K.; Karlsen, O.A.; Ziegler, M.; Arnesen, T.; Lillehaug, J.R. Human Naa50p (Nat5/San) displays both protein N alpha- and N epsilon-acetyltransferase activity. J. Biol. Chem. 2009, 284, 31122–31129. [Google Scholar] [CrossRef] [Green Version]
  130. Støve, S.I.; Magin, R.S.; Foyn, H.; Haug, B.E.; Marmorstein, R.; Arnesen, T. Crystal structure of the Golgi-Associated human Na-Acetyltransferase 60 reveals the molecular determinants for Substrate-Specific acetylation. Structure 2016, 24, 1044–1056. [Google Scholar] [CrossRef] [Green Version]
  131. Liszczak, G.; Arnesen, T.; Marmorstein, R. Structure of a ternary Naa50p (NAT5/SAN) N-terminal acetyltransferase complex reveals the molecular basis for substrate-specific acetylation. J. Biol. Chem. 2011, 286, 37002–37010. [Google Scholar] [CrossRef] [Green Version]
  132. 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. [Google Scholar] [CrossRef] [Green Version]
  133. 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Song, O.K.; Wang, X.R.; Waterborg, J.H.; Sternglanz, R. An N-alpha-acetyltransferase responsible for acetylation of the N-terminal residues of histones H4 and H2A. J. Biol. Chem. 2003, 278, 38109–38112. [Google Scholar] [CrossRef] [Green Version]
  135. Hole, K.; Van Damme, P.; Dalva, M.; Aksnes, H.; Glomnes, N.; Varhaug, J.E.; Lillehaug, J.R.; Gevaert, K.; Arnesen, T. The human N-alpha-acetyltransferase 40 (hNaa40p/hNatD) is conserved from yeast and N-terminally acetylates histones H2A and H4. PLoS ONE 2011, 6, e24713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Sheff, D.R.; Rubenstein, P. Isolation and characterization of the rat liver actin N-acetylaminopeptidase. J. Biol. Chem. 1992, 267, 20217–20224. [Google Scholar] [CrossRef] [PubMed]
  137. Starheim, K.K.; Arnesen, T.; Gromyko, D.; Ryningen, A.; Varhaug, J.E.; Lillehaug, J.R. Identification of the human N (alpha)-acetyltransferase complex B (hNatB): A complex important for cell-cycle progression. Biochem. J. 2008, 415, 325–331. [Google Scholar] [CrossRef] [Green Version]
  138. Polevoda, B.; Cardillo, T.S.; Doyle, T.C.; Bedi, G.S.; Sherman, F. Nat3p and Mdm20p are required for function of yeast NatB Nalpha-terminal acetyltransferase and of actin and tropomyosin. J. Biol. Chem. 2003, 278, 30686–30697. [Google Scholar] [CrossRef] [Green Version]
  139. Ametzazurra, A.; Larrea, E.; Civeira, M.P.; Prieto, J.; Aldabe, R. Implication of human N-a-acetyltransferase 5 in cellular proliferation and carcinogenesis. Oncogene 2008, 27, 7296–7306. [Google Scholar] [CrossRef] [Green Version]
  140. Neri, L.; Lasa, M.; Elosegui-Artola, A.; D’Avola, D.; Carte, B.; Gazquez, C.; Alve, S.; Roca-Cusachs, P.; Iñarrairaegui, M.; Herrero, J.; et al. NatB-mediated protein N-a-terminal acetylation is a potential therapeutic target in hepatocellular carcinoma. Oncotarget 2017, 8, 40967–40981. [Google Scholar] [CrossRef] [Green Version]
  141. Arnesen, T.; Marmorstein, R.; Dominguez, R. Actin’s N-terminal acetyltransferase uncovered. Cytoskeleton 2018, 75, 318–322. [Google Scholar] [CrossRef]
  142. Goris, M.; Magin, R.S.; Foyn, H.; Myklebust, L.M.; Varland, S.; Ree, R.; Varland, S.; Ree, R.; Drazic, A.; Bhambra, P.; et al. Structural determinants and cellular environment define processed actin as the sole substrate of the N-terminal acetyltransferase NAA80. Proc. Natl. Acad. Sci. USA 2018, 115, 4405–4410. [Google Scholar] [CrossRef] [Green Version]
  143. Rebowski, G.; Boczkowska, M.; Drazic, A.; Ree, R.; Goris, M.; Arnesen, T.; Goris, M.; Arnesen, T.; Dominguez, R. Mechanism of actin N-terminal acetylation. Sci. Adv. 2020, 6, eaay8793. [Google Scholar] [CrossRef] [Green Version]
  144. Jia, K.; Huang, G.; Wu, W.; Shrestha, R.; Wu, B.; Xiong, Y.; Li, P. In Vivo Methylation of OLA1 Revealed by Activity-Based Target Profiling of NTMT1. Chem. Sci. 2019, 10, 8094–8099. [Google Scholar] [CrossRef] [Green Version]
  145. Liu, S.; Hausmann, S.; Carlson, S.M.; Fuentes, M.E.; Francis, J.W.; Pillai, R.; Lofgren, S.M.; Hulea, L.; Tandoc, K.; Lu, J.; et al. METTL13 Methylation of EEF1A Increases Translational Output to Promote Tumorigenesis. Cell 2019, 176, 491–504. [Google Scholar] [CrossRef] [Green Version]
  146. Linder, M.E.; Middleton, P.; Hepler, J.R.; Taussig, R.; Gilman, A.G.; Mumby, S.M. Lipid modifications of G proteins: Alpha subunits are palmitoylated. Proc. Natl. Acad. Sci. USA 1993, 90, 3675–3679. [Google Scholar] [CrossRef] [PubMed]
  147. Kleuss, C.; Krause, E. Galpha(s) is palmitoylated at the N-terminal glycine. EMBO J. 2003, 22, 826–832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Resh, M.D. Palmitoylation of ligands, receptors, and intracellular signaling molecules. Sci. STKE 2006, 2006, 14. [Google Scholar] [CrossRef] [PubMed]
  149. Pepinsky, R.B.; Zeng, C.; Wen, D.; Rayhorn, P.; Baker, D.P.; Williams, K.P.; Bixler, S.A.; Ambrose, C.M.; Garber, E.A.; Miatkowski, K.; et al. Identification of a palmitic acid-modified form of human Sonic hedgehog. J. Biol. Chem. 1998, 273, 14037–14045. [Google Scholar] [CrossRef] [Green Version]
  150. Buglino, J.A.; Resh, M.D. Hhat is a palmitoylacyltransferase with specificity for N-palmitoylation of Sonic Hedgehog. J. Biol. Chem. 2008, 283, 22076–22208. [Google Scholar] [CrossRef] [Green Version]
  151. Chamoun, Z.; Mann, R.K.; Nellen, D.; von Kessler, D.P.; Bellotto, M.; Beachy, P.A.; Basler, K. Skinny hedgehog, an acyltransferase required for palmitoylation and activity of the hedgehog signal. Science 2001, 293, 2080–2084. [Google Scholar] [CrossRef]
  152. Lee, J.D.; Treisman, J.E. Sightless has homology to transmembrane acyltransferases and is required to generate active Hedgehog protein. Curr. Biol. 2001, 11, 1147–1152. [Google Scholar] [CrossRef] [Green Version]
  153. Micchelli, C.A.; The, I.; Selva, E.; Mogila, V.; Perrimon, N. Rasp, a putative transmembrane acyltransferase, isrequired for Hedgehog signaling. Development 2002, 129, 843–851. [Google Scholar] [CrossRef] [PubMed]
  154. Hofmann, K. A superfamily of membrane-bound Oacyltransferases with implications forwnt signaling. Trends Biochem. Sci. 2000, 25, 111–112. [Google Scholar] [CrossRef]
  155. Taylor, F.R.; Wen, D.; Garber, E.A.; Carmillo, A.N.; Baker, D.P.; Arduini, R.M.; Williams, K.P.; Weinreb, P.H.; Rayhorn, P.; Hronowski, X.; et al. Enhanced potency of human Sonic hedgehog by hydrophobic modification. Biochemistry 2001, 40, 4359–4371. [Google Scholar] [CrossRef]
  156. Ciechanover, A.; Stanhill, A. The complexity of recognition of ubiquitinated substrates by the 26S proteasome. Biochim. Biophys. Acta 2014, 1843, 86–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Varshavsky, A. The ubiquitin system, an immense realm. Annu. Rev. Biochem. 2012, 81, 167–176. [Google Scholar] [CrossRef]
  158. Breitschopf, K.; Bengal, E.; Ziv, T.; Admon, A.; Ciechanover, A. A novel site for ubiquitination: The N-terminal residue, and not internal lysines of MyoD, is essential for conjugation and degradation of the protein. EMBO J. 1998, 17, 5964–5973. [Google Scholar] [CrossRef] [Green Version]
  159. Ben-Saadon, R.; Fajerman, I.; Ziv, T.; Hellman, U.; Schwartz, A.L.; Ciechanover, A. The tumor suppressor protein p16(INK4a) and the human papillomavirus oncoprotein-58 E7 are naturally occurring lysineless proteins that are degraded by the ubiquitin system. Direct evidence for ubiquitination at the N-terminal residue. J. Biol. Chem. 2004, 279, 41414–41421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Ye, Y.; Klenerman, D.; DanielFinley, D. N-Terminal Ubiquitination of Amyloidogenic Proteins Triggers Removal of Their Oligomers by the Proteasome Holoenzyme. J. Mol. Biol. 2020, 432, 585–596. [Google Scholar] [CrossRef]
  161. Scaglione, K.M.; Basrur, V.; Ashraf, N.S.; Konen, J.R.; Elenitoba-Johnson, K.S.J.; Todi, S.V.; Paulson, H.L. The ubiquitin-conjugating enzyme (E2) Ube2w ubiquitinates the N terminus of substrates. J. Biol. Chem. 2013, 288, 18784–18788. [Google Scholar] [CrossRef] [Green Version]
  162. Tatham, M.H.; Plechanovova, A.; Jaffray, E.G.; Salmen, H.; Hay, R.T. Ube2W conjugates ubiquitin to alpha-amino groups of protein N-termini. Biochem. J. 2013, 453, 137–145. [Google Scholar] [CrossRef]
  163. Noy, T.; Suad, O.; Taglicht, D.; Ciechanover, A. HUWE1ubiquitinates MyoD and targets it for proteasomal degradation. Biochem. Biophys. Res. Commun. 2012, 418, 408–413. [Google Scholar] [CrossRef] [PubMed]
  164. Vittal, V.; Shi, L.; Wenzel, D.M.; Scaglione, K.M.; Duncan, E.D.; Basrur, V.; Elenitoba-Johnson, K.S.J.; Baker, D.; Paulson, H.L.; Brzovic, P.S.; et al. Intrinsic disorder drives N-terminal ubiquitination by Ube2w. Nat. Chem. Biol. 2015, 11, 83–89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Kaji, A.; Kaji, H.; Novelli, G.D. A soluble amino acid incorporating system. Biochem. Biophys. Res. Commun. 1963, 10, 406–409. [Google Scholar] [CrossRef] [PubMed]
  166. Kaji, H.; Novelli, G.D.; Kaji, A. A soluble amino acid-incorporatingsystem from rat liver. Biochim. Biophys. Acta 1963, 76, 474–477. [Google Scholar] [CrossRef] [PubMed]
  167. Balzi, E.; Choder, M.; Chen, W.; Varshavsky, A.; Goffeau, A. Cloning and functional analysis of the arginyl-tRNA-protein transferase gene ATE1 of Saccharomyces cerevisiae. J. Biol. Chem. 1990, 265, 7464–7471. [Google Scholar] [CrossRef]
  168. Rai, R.; Kashina, A. Identification of mammalian arginyltransferasesthat modify a specific subset of protein substrates. Proc. Natl. Acad. Sci. USA 2005, 102, 10123–10128. [Google Scholar] [CrossRef]
  169. Rai, R.; Mushegian, A.; Makarova, K.; Kashina, A. Molecular dissection of arginyltransferases guided by similarity to bacterial peptidoglycan synthases. EMBO J. 2006, 7, 800–805. [Google Scholar] [CrossRef] [Green Version]
  170. Wang, J.; Pejaver, V.R.; Dann, G.P.; Wolf, M.Y.; Kellis, M.; Huang, Y.; Kellis, M.; Huang, Y.; Garcia, B.A.; Radivojac, P.; et al. Target site specificity and in vivo complexity of the mammalian arginylome. Sci. Rep. 2018, 8, 16177. [Google Scholar] [CrossRef] [Green Version]
  171. Hu, R.G.; Brower, C.S.; Wang, H.; Davydov, I.V.; Sheng, J.; Zhou, J.; Sheng, J.; Zhou, J.; Kwon, Y.T.; Varshavsky, A. Arginyltransferase, its specificity, putative substrates, bidirectional promoter, and splicing-derived isoforms. J. Biol. Chem. 2006, 281, 32559–32573. [Google Scholar] [CrossRef] [Green Version]
  172. Sriram, S.M.; Kim, B.Y.; Kwon, Y.T. The N-end rule pathway: Emerging functions and molecular principles of substrate recognition. Nat. Rev. Mol. Cell Biol. 2011, 12, 735–747. [Google Scholar] [CrossRef]
  173. Wong, C.C.; Xu, T.; Rai, R.; Bailey, A.O.; Yates, J.R., III; Wolf, Y.I.; Zebroski, H.; Kashina, A. Global analysis of posttranslational protein arginylation. PLoS Biol. 2007, 5, e258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Wang, J.; Yates, J.R., III; Kashina, A. Biochemical analysis of protein arginylation. Methods Enzymol. 2019, 626, 89–113. [Google Scholar] [PubMed]
  175. Bachmair, A.; Finley, D.; Varshavsky, A. In vivo half-life of a protein is a function of its amino-terminal residue. Science 1986, 234, 179–186. [Google Scholar] [CrossRef] [PubMed]
  176. Varshavsky, A. N-degron and C-degron pathways of protein degradation. Proc. Natl. Acad. Sci. USA 2019, 116, 358–366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Carpio, M.A.; Sambrooks, C.; Durand, E.S.; Hallak, M.E. The arginylation-dependent association of calreticulin with stress granules is regulated by calcium. Biochem. J. 2010, 429, 63–72. [Google Scholar] [CrossRef] [PubMed]
  178. Goitea, V.E.; Hallak, M.E. Calreticulin and arginylated calreticulin have different susceptibilities to proteasomal degradation. J. Biol. Chem. 2015, 290, 16403–16414. [Google Scholar] [CrossRef] [Green Version]
  179. Comba, A.; Bonnet, L.V.; Goitea, V.E.; Hallak, M.E.; Galiano, M.R. Arginylated calreticulin increases apoptotic response induced by Bortezomib in Glioma Cells. Mol. Neurobiol. 2019, 56, 1653–1664. [Google Scholar] [CrossRef]
  180. Varland, S.; Vandekerckhove, J.; Drazic, A. Actin posttranslational modifications: The cinderella of cytoskeletal control. Trends Biochem. Sci. 2019, 44, 502–516. [Google Scholar] [CrossRef] [Green Version]
  181. Pavlyk, I.; Leu, N.A.; Vedula, P.; Kurosaka, S.; Kashina, A. Rapid and dynamic arginylation of the leading edge β-actin is required for cell migration. Traffic 2018, 19, 263–272. [Google Scholar] [CrossRef] [Green Version]
  182. Vedula, P.; Kurosaka, S.; MacTaggart, B.; Ni, Q.; Papoian, G.; Jiang, Y.; Papoian, G.; Jiang, Y.; Dong, D.W.; Kashina, A. Different translation dynamics of β- and γ-actin regulates cell migration. eLife 2021, 10, e68712. [Google Scholar] [CrossRef]
  183. Karakozova, M.; Kozak, M.; Wong, C.C.L.; Bailey, A.O.; Yates, J.R., 3rd; Mogilner, A.; Zebroski, H.; Kashina, A. Arginylation of ß-actin regulates actin cytoskeleton and cell motility. Science 2006, 313, 192–196. [Google Scholar] [CrossRef]
  184. Zhang, F.; Saha, S.; Shabalina, S.A.; Kashina, A. Differential arginylation of actin isoforms is regulated by coding sequence–dependent degradation. Science 2010, 329, 1534–1537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Drazic, A.; Timmerman, E.; Kajan, U.; Marie, M.; Varland, S.; Impens, F.; Gevaert, K.; Arnesen, T. The final maturation state of β-actin Involves N-terminal acetylation by NAA80, not N-terminal arginylation by ATE1. J. Mol. Biol. 2022, 434, 167397. [Google Scholar] [CrossRef]
  186. Nguyen, K.T.; Ju, S.; Kim, S.-Y.; Lee, C.-S.; Lee, C.; Hwang, C.-S. N-Terminal Modifications of Ubiquitin via Methionine Excision, Deamination, and Arginylation Expand the Ubiquitin Code. Mol. Cells 2022, 45, 158–167. [Google Scholar] [CrossRef]
  187. Dittmar, G.; Selbach, M. Deciphering the ubiquitin code. Mol. Cell 2017, 65, 779–780. [Google Scholar] [CrossRef] [Green Version]
  188. Mattiroli, F.; Penengo, L. Histone ubiquitination: An integrative signaling platform in genome stability. Trends Genet. 2021, 37, 566–581. [Google Scholar] [CrossRef] [PubMed]
  189. Swatek, K.N.; Komander, D. Ubiquitin modifications. Cell Res. 2016, 26, 399–422. [Google Scholar] [CrossRef] [Green Version]
  190. Chiu, J.; Wong, J.W.H.; Hogg, P.J. Redox Regulation of Methionine Aminopeptidase 2 Activity. J. Biol. Chem. 2014, 289, 15035–15043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Demetriadou, C.; Koufaris, C.; Kirmizis, A. Histone N-alpha terminal modifications: Genome regulation at the tip of the tail. Epigenetics Chromatin 2020, 13, 29. [Google Scholar] [CrossRef]
  192. Lawrence, M.; Daujat, S.; Schneider, R. Lateral thinking: How histone modifications regulate gene expression. Trends Genet. 2016, 32, 42–56. [Google Scholar] [CrossRef] [Green Version]
  193. Bao, X.; Liu, Z.; Zhang, W.; Gladysz, K.; Fung, Y.M.E.; Tian, G.; Xiong, Y.; Wong, J.W.H.; Yuen, K.W.Y.; Li, X.D. Glutarylation of histone H4 lysine 91 regulates chromatin dynamics. Mol Cell. 2019, 76, 660–675. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Summary of Nα-modifications of Cytosolic Human Proteins. 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]; and N-terminal methylation for Nα-methylation (NTMTs) [5].
Figure 1. Summary of Nα-modifications of Cytosolic Human Proteins. 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]; and N-terminal methylation for Nα-methylation (NTMTs) [5].
Life 13 01613 g001
Figure 3. Nα-modifications of ubiquitin. (A) Ubiquitin (Ub) starts with Met-Gln-Ile (MQ) N-terminus. Nα-modifications of ubiquitin involve NME possibly by an unknown MetAP, N-terminal deamination by NTAQ1 N-terminal Gln amidase, and N-terminal arginylation by ATE1 arginyltransferase [187,188,189] ubiquitin might be processed by the NME-provoked cascade reactions of N-terminal deamination and N-terminal arginylation to yield RE-Ub (Figure 3) [186]. (B) However, according to the specificity of NatB, iMet is likely to be acetylated by NatB. If so, the acetylated iMet will be removed by an unknown deacetylase. Then, the newly exposed Gln is converted to Glu by NTAQ1 N-terminal Gln amidase, followed by Nα-arginylation by ATE1.
Figure 3. Nα-modifications of ubiquitin. (A) Ubiquitin (Ub) starts with Met-Gln-Ile (MQ) N-terminus. Nα-modifications of ubiquitin involve NME possibly by an unknown MetAP, N-terminal deamination by NTAQ1 N-terminal Gln amidase, and N-terminal arginylation by ATE1 arginyltransferase [187,188,189] ubiquitin might be processed by the NME-provoked cascade reactions of N-terminal deamination and N-terminal arginylation to yield RE-Ub (Figure 3) [186]. (B) However, according to the specificity of NatB, iMet is likely to be acetylated by NatB. If so, the acetylated iMet will be removed by an unknown deacetylase. Then, the newly exposed Gln is converted to Glu by NTAQ1 N-terminal Gln amidase, followed by Nα-arginylation by ATE1.
Life 13 01613 g003
Table 1. Impact of protein Nα-modifications on cellular functions.
Table 1. Impact of protein Nα-modifications on cellular functions.
EnzymesCellular FunctionsReferences
MetAP1Cell cycle progression, cell proliferation, enzyme function, protein stability, cellular localization[13,14,15,16,17,18,19,20,21,22,23]
MetAP2Angiogenesis, B-cell differentiation, cell specific Cytotoxicity[23,24,25,26,27,28,29,30,31,32,33,34]
NMTsSignal transduction, cellular transformation,
innate immune responses, adaptative immune response
[35,36,37,38]
NATsActin cytoskeleton structure, cell cycle progression, cell proliferation
cell mobility
[39,40,41,42,43,44,45,46,47,48]
NTMTsProtein 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]
Table 2. Potential targets for developing a novel treatment for human diseases.
Table 2. Potential targets for developing a novel treatment for human diseases.
EnzymesTargeted Human DiseasesReferences
MetAP1Antibiotics, cancer[20]
MetAP1DColon cancer[59,60]
MetAP2Cancer, obesity, diabetes, Prader-Willi syndrome, autoimmunity[23,24,25,26,27,28,29,30,31,32,33,34]
NMTsCancer; HIV, fungal, and parasitic infection[61,62,63,64,65,66,67,68,69,70]
NATsCancer (lung, liver, colon), Parkinson’s disease[71,72,73,74,75,76,77,78,79,80,81]
NTMTsCancer (breast, colon, pancreatic, lung)[82,83,84,85,86,87,88,89]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chang, Y.-H. Impact of Protein Nα-Modifications on Cellular Functions and Human Health. Life 2023, 13, 1613. https://doi.org/10.3390/life13071613

AMA Style

Chang Y-H. Impact of Protein Nα-Modifications on Cellular Functions and Human Health. Life. 2023; 13(7):1613. https://doi.org/10.3390/life13071613

Chicago/Turabian Style

Chang, Yie-Hwa. 2023. "Impact of Protein Nα-Modifications on Cellular Functions and Human Health" Life 13, no. 7: 1613. https://doi.org/10.3390/life13071613

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop