1 s2.0 S030090841400337X Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Biochimie 114 (2015) 134e146

Contents lists available at ScienceDirect

Biochimie
journal homepage: www.elsevier.com/locate/biochi

Review

N-terminal protein modifications: Bringing back into play the


ribosome
Carmela Giglione a, b, *, Sonia Fieulaine a, b, Thierry Meinnel a, b, *
a
CNRS, Institut des Sciences du V eg ^t 23A, F-91198 Gif sur Yvette, France
etal, 1 Avenue de la Terrasse, Ba
b
Institute for Integrative Biology of the Cell (I2BC), CEA, CNRS, Universit
e Paris-Sud, 1 Avenue de la Terrasse, 91198 Gif-sur-Yvette cedex, France

a r t i c l e i n f o a b s t r a c t

Article history: N-terminal protein modifications correspond to the first modifications which in principle any protein
Received 23 September 2014 may undergo, before translation is completed by the ribosome. This class of essential modifications can
Accepted 10 November 2014 have different nature or function and be catalyzed by a variety of dedicated enzymes. Here, we review
Available online 18 November 2014
the current state of the major N-terminal co-translational modifications, with a particular emphasis to
their catalysts, which belong to metalloprotease and acyltransferase clans. The earliest of these modi-
Keywords:
fications corresponds to the N-terminal methionine excision, an ubiquitous and essential process leading
Co-translational modifications
to the removal of the first methionine. N-alpha acetylation occurs also in all Kingdoms although its
N-terminal
Peptide deformylase
extent appears to be significantly increased in higher eukaryotes. Finally, N-myristoylation is a crucial
Methionine aminopeptidase pathway existing only in eukaryotes. Recent studies dealing on how some of these co-translational
N-a-acetyltransferase modifiers might work in close vicinity of the ribosome is starting to provide new information on when
N-myristoyltransferase these modifications exactly take place on the elongating nascent chain and the interplay with other
ribosome biogenesis factors taking in charge the nascent chains. Here a comprehensive overview of the
recent advances in the field of N-terminal protein modifications is given.
© 2014 Elsevier B.V. and Socie te
 Française de Biochimie et Biologie Mole
culaire (SFBBM). All rights
reserved.

1. Introduction normal growth and function in most organisms and affect a vari-
able number of proteins.
It is recognized that the majority of proteins, to become func- NPMs usually take place co-translationally when the protein is
tional actives, needs to undergo several modifications which are still very short and yet bound to the ribosome. They are catalyzed by
usually catalyzed by specific enzymes. More than 400 different ubiquitous and essential enzymes being part of the so-called ribo-
protein modifications have been identified so far [1]. These modi- some-associated protein biogenesis factors (RPBs), which are non-
fications can be reversible or irreversible, thus escorting the protein ribosomal proteins acting as soon as a nascent polypeptide reaches
along its whole life cycle. Modifications affecting the protein N- the extremity of the ribosomal exit tunnel [2,3]. RPBs ensure a
termini, collectively known as N-terminal protein modifications number of co-translational events, including not only nascent pep-
(NPMs), are the earliest modifications which a protein undergoes. tide chain modifications but also protein folding and/or trans-
They involve the N-terminal methionine excision process (NME) location into various cellular compartments. Indeed, the ribosome
and acylations corresponding to N-a-acetylation (NTA) and N- functions as a central hub for a group of a dozen RPBs which act
myristoylation (MYR) (Fig. 1). These modifications can often be simultaneously or sequentially and ultimately define the functional
followed by further reversible modifications such as phosphoryla- biomolecules that populate the cell. The highly regulated mecha-
tion or further acylations such as S-palmitoylation (PAL). These nisms that govern the action of RPBs in space and time are still poorly
second-site modifications participate in the regulatory mecha- understood. Some models have recently appeared in the literature,
nisms associated to the NPM pathways. All NPMs are essential for highlighting coordination or competition events between RPBs.
The interest of the scientific community for NPMs has been
strongly revived during the past decade, probably because i) ge-
* Corresponding authors. 1 Avenue de la Terrasse, B^
at 23A, F-91198 Gif sur Yvette, nomics has provided further insights into the diversity and origin of
France.
the enzymes involved in these modifications, ii) proteomics issues
E-mail addresses: Carmela.Giglione@isv.cnrs-gif.fr (C. Giglione), Thierry.
Meinnel@isv.cnrs-gif.fr (T. Meinnel). are now often turning to protein modifications and iii) natural or

http://dx.doi.org/10.1016/j.biochi.2014.11.008
0300-9084/© 2014 Elsevier B.V. and Socie  te
 Française de Biochimie et Biologie Mole
culaire (SFBBM). All rights reserved.
C. Giglione et al. / Biochimie 114 (2015) 134e146 135

Fig. 1. The co-translational N-terminal protein modifications (NPMs). NPMs are catalyzed by several enzymes that act successively: peptide deformylase (PDF) and methionine
aminopeptidase (MetAP) which ensure the N-terminal methionine excision process (NME), followed by N-a-acetyltransferase (Nat) or N-myristoyltransferase (NMT) that acylate
proteins. Depending of the organism and cellular compartment, the sequence of NPMs is quite different, as well as the amount of modified proteins. The percentage displayed here
into brackets reflects the relative number of occurrences in a given proteome, knowing that in terms of protein accumulation the value can be significantly increased [see dis-
cussions in Ref. [67,80,173]]. For instance, NTA is known to reach 90% of the total protein mass in human cell extracts, while the value is lower in terms of diversity.

synthetic compounds targeting NPMs with antibiotic, antiparasitic, and in the organelles [6e9]. These rules are conserved even when a
antifungal or anti-cancer effects have been discovered. Despite that non-conventional initiation codon is used (i.e., different from the
the physiological function(s) of these NPMs remained for long time classic AUG start codon, including GUG, AUU, CUG…). However, the
largely unknown, remarkable efforts have been recently deployed first Met is often specifically removed from proteins accumulating
allowing a depth characterization of the co-translational modifi- at the steady state. Thus, the irreversible NME process corresponds
cations and important breakthroughs concerning the enzymes in bacteria and in the organelles of eukaryotes to the early removal
involved and the role of these NPMs in protein fate have been of the N-formyl group immediately followed by the removal (in
highlighted. The ways some of the enzymes involved in the process both bacteria and in eukaryotes) of the first Met. NME process is
of NPMs work and act in concert with other RPBs at the level of the ensured by two metalloenzymes known as peptide deformylases
ribosome have been refined in recent times. (PDFs) and methionine aminopeptidases (MetAPs) which sequen-
In this review, we provide an overview of the early N-terminal tially remove the formyl group, when it is present, and the N-ter-
protein modifications and we summarize and discuss the recent minal Met, respectively (Fig. 1).
data underlying the mechanistic details on how several of NPM
enzymes emerge to be involved in association with the ribosome. 2.1. Extent of the NME process

2. The NME process Between 30 and 60% of proteins are liberated of the N-terminal
Met, depending on the organism and cellular compartment [10]
Among NPMs, NME is the first occurring widespread protein (Fig. 1). More than 90% of characterized bacterial or chloroplast
modification. To understand what the NME process is and what it encoded proteins undergo deformylation [This laboratory, unpub-
induces on protein N-termini, it should be remembered that the lished results and Ref. [11]], whereas the percentage of
initiation process of ribosome-dependent protein synthesis mitochondria-encoded proteins undergoing deformylation is still
strongly differs depending on the type of organisms or cellular not clearly defined. Indeed, the number of mitochondrially-
compartments thereof (i.e., cytoplasm, mitochondrion or plastid) encoded proteins is variable depending on the organism [e.g., 13
[4,5]. In archaea and the cytoplasm of eukaryotes, protein synthesis in human [12] and 33 in the model plant Arabidopsis thaliana [13]]
always starts with an unblocked, so-called “free” methionine (Met), and because their high hydrophobicity the chemical characteriza-
whereas it starts with an N-formylated Met (Fo-Met) in Eubacteria tion has remained elusive for long time. By measuring the precise
136 C. Giglione et al. / Biochimie 114 (2015) 134e146

intact molecular masses of most of the 13 mitochondrial encoded protein half-life, NME was shown to fine tune the global gluta-
proteins from bovine heart it has been suggested that most likely thione redox homeostasis at least in plants, yeast and Archaea [31].
only the Cox III protein undergoes a complete NME process [12].
Nonetheless, it has been shown that the human mitochondrially
2.3. Peptide deformylases
PDF is able to efficiently deformylate, at least in vitro 7, formylated
peptides derived from the N-terminus of mitochondrial encoded
Peptide deformylase (PDF) is the first enzyme to modify the N-
proteins [14]. The sequencing of 8 of 32 mitochondrially encoded
terminal extremity of newly synthesized proteins. It is essential for
plant proteins clearly reveals that all the sequenced proteins with
eubacteria [32,33] and leads to serious dysfunction if altered in
one exception having their N-Formyl-Met being removed [15].
organelles [25,34e36]. PDFs are mostly monomeric hydrolases
belonging to the thermolysin-metzincin HEXXH motif containing
2.2. Control and physiological roles of NME protein family, shown to absolutely require a metal cation bound to
either of the His residues of the motif for hydrolysis to occur [37,38].
NME is conserved from bacteria to higher eukaryotes and is Although believed to exist only in bacteria, during the last decade
essential for normal growth and function in any organism investi- PDFs were found in almost all organisms, including not only pro-
gated [11]. Despite NME is still considered as a non-dynamic karyotes and most eukaryotes but also in viruses [39e42]. The low
imprinting modification of any proteome, many data highlighted sequence similarity observed for all identified PDFs contrasts with
that the enzymes ensuring the process are tightly regulated during the high evolutionary conservation of 3 short stretches of amino
development, tumorigenesis and in response to abiotic and biotic acids (motifs I, II and III) building the whole active site of these
stresses [16e23]. Nonetheless, a clear impact of these regulated enzymes (Fig. 2A). A Cys residue crucial as third ligand of the metal
processes on the modification state of the entire proteome is not cation features in motif II. Motif I creates series of hydrogen bonds
directly evidenced so far. Why up to two-third of proteins of any made with a Gln side chain and peptide bond residues for the
proteome undergo a very early NME process has remained elusive formyl-Met substrates to properly and specifically align the active
for long time. More than 25 years ago NME was suggested to be site, where the Glu of Motif III and the metal cation are crucial for
critical for protein half-life [24]. This is in line with the discovery hydrolysis to occur. Moreover, the 3D structures of many currently
that NME modulates the lifespan of a reduced number of chloro- available PDFs, coming from almost 30 different organisms, reveal
plast key proteins, suggesting that NME might directly contribute to that they all share identical structure, the active site fold is similar
protein stability, and that either removal or maintenance of the first and that the main structural difference concerns the C-terminal
Met can be perceived as stabilizing or destabilizing signals, part. Thus, based upon structure and sequence similarity, PDFs are
respectively [25]. The recent extraordinary accomplishments in this divided into three groups (1, 2 and 3), with Type 1 including a
field provided strong support for the “protein stability” hypothesis, subgroup 1A and a subgroup 1B. Type 1B PDF is considered as the
bridging the gap between NME and protein half-life. Particularly, in structural PDF model, with a globular core and a C-terminal a-helix
the context of the N-end rule, the best-characterized specific signal [43,44] (Fig. 2B). PDFs of Type 2 possess several insertions within
for protein degradation [for reviews see Refs. [26,27]], the ambig- the globular core, and harbor a totally different C-terminal ex-
uous role of the first Met as a destabilizing or stabilizing residue has tremity compared to Type 1B PDF [45] (Fig. 2B). Different to PDFs of
been recently elucidated and showed that the N-termini of proteins Type 1B that are found in bacteria and organelles, PDFs of Type 1A
starting with an N-alpha-acetylated Met (see next sections) or a are exclusively located in organelles [40,46e48]. They possess a
free Met may serve as the site of N-ubiquitination by regulation of long insertion that partially clogs the active site entrance,
their steric shielding [25,28e30]. Through its global control of contributing to a narrow ligand binding site, and harbors a C-

Fig. 2. Sequence and structure of peptide deformylases (PDFs). PDFs are classified into three groups (types 1, 2 and 3) and two type 1 subgroups (1A and 1B). (A) Sequence
alignment around the three conserved motifs, with two representative sequences of each PDF type. Escherichia coli and Pseudomonas aeruginosa for type 1B; cyanophage S-SSM7
and viral 1502.1906 [39,41] for C-terminal truncated type 1B; Bacillus stearothermophilus and Staphylococcus aureus for type 2; Homo sapiens and Arabidopsis thaliana for type 1A;
Trypanosoma brucei and Leishmania major for type 3. The 38 and 90 last residues of Type 3 PDFs has been truncated (noted X38 and X90). Identical residues are boxed in red when
strictly conserved or red-colored when conserved in all sequences except for one or two sequences [174]. (B) Comparison of archetype PDF structures for types 1B, 2 and 1A.
Structures are directed towards the substrate binding cleft of the catalytic site, with the three characteristic motifs of PDF enzymes highlighted in magenta. The variable C-terminal
extremity is shown in green and specific insertions are highlighted in red or blue. The comparison of C-terminal extremity folding is shown through superimposition of the three
types of PDF structures with the globular core of Escherichia coli PDF represented as molecular surface (right).
C. Giglione et al. / Biochimie 114 (2015) 134e146 137

terminal extremity different to that found in Type 1B and 2 [14,47]


(Fig. 2B). This C-terminal part is totally absent or shorter in the
recent discovered viral PDFs, all belonging to the Type 1B subgroup
(Fig. 2A) [39,41,42,49]. Despite these differences at the C-terminus,
for one of these viral PDFs it has been recently shown to display a
classical PDF activity in vitro [41]. Type 3 PDFs, for which no 3D
structure is available to date, are also characterized by a much
longer C-terminal extension [50] (Fig. 2A), which is likely suggested
to fold as an a-helical structure (data not shown). Type 3 PDFs are
suggested to be inactive because they display substitutions in
crucial amino acids of motif I (Fig. 2A). To date, the exact role of
these inactive PDFs is not known, but it has been shown that
frequently the organisms exhibiting these inactive isoforms have
concomitantly active PDFs paralogs [51,52].
PDFs deformylate newly synthesized proteins in a co-
translational way. It has been shown that E. coli subtype 1B PDFs
interact with the ribosome in the exit tunnel region, with domain I
of 23S ribosomal RNA, uL22 and probably bL17 and bL32 proteins.
The interaction is driven by the C-terminal helix extension of the
PDF, which therefore acts as a master regulator of PDF-ribosome
binding [53]. The significance of this positively charged C-termi-
nal a-helix for binding to the ribosome and influencing cell viability
and growth rate raises the question of how the viral PDFs engender
their interaction with ribosomes. In a more general context, it as
well remains unclear how other PDF enzymes that do not belong to
the same class as the E. coli PDF interact with the ribosome. Indeed,
the C-terminal extensions of subtype 1A or 2 PDFs adopt different
folding compared to the C-terminal a-helix of subtype 1B [49]
(Fig. 2B). It should be also pointed out that PDFs of subtype 1B
and 1A or 2 can be present in the same bacteria or organelle
[45,54,55], triggering the question whether different types of PDFs
can simultaneously bind the same ribosome and more importantly
what are the physiological consequences for an organism to express
slightly different PDFs [49].

2.4. Methionine aminopeptidases

Methionine aminopeptidases (MetAPs) remove the initiator


methionine of newly synthesized proteins, which as a prerequisite
has to be deformylated by PDF. The MetAPs are essential in all
kingdoms of life, from bacteria to higher eukaryotes including yeast
and plants [11,56,57]. MetAPs are dinuclear metalloenzymes
evolutionary closed to creatinase, prolidase (DPP) and aminopep-
tidase P (APP) [11]. They all belong to the so called “pita-bread”
family or “clan MG” M24 protease family featuring a pseudo two-
fold axis of symmetry with the catalytic site and metal ion bind-
ing located at the interfaces between the domains (see legend of
Fig. 3 for more details) [56,58]. Metal requirement studies indicate
that MetAPs can be activated by various divalent metal cations but
unlike PDFs the physiological metals of MetAPs are still a matter of
debate [56,59]. As for PDFs, a low level of MetAPs sequence simi-
larity is compensated by similar 3D structures of the active site and
a conserved 5-residue metal binding site.
From sequence similarity and structural characterization two
Fig. 3. Domain architecture and structure of methionine aminopeptidases (MetAPs). types of MetAPs, MetAP1 and MetAP2, can be distinguished. Pro-
MetAPs are classified into two groups (types 1 and 2) and several subgroups (1a-d and karyotes possess homologs of either MetAP1s (bacteria) or
2a-b), distinguished by specific additional domains. (A) All MetAPs possess a common
catalytic domain (pink), in which MetAPs of type 2 harbors two insertions (designed as
N- and C-terminal insertions, shown in green and blue respectively). MetAP1a' from all
Streptococci and several Lactobacilli differs from the classical MetAP1a enzymes furiosus). MetAP1a' from Streptococcus pneumoniae and human Ebp1 are also shown.
through the presence of two insertions within the conserved catalytic domain. Structures are directed towards the substrate binding cleft of the catalytic site, with
Eukaryotic MetAPs display N-terminal extensions of variable length and sequences catalytic residues of EcMetAP1a highlighted in magenta. The insertions within the
(red). Ebp1 is a human protein related to type 2 MetAP group, i.e., with insertions in the catalytic domain of type 2 MetAPs are highlighted in green and blue, and various N-
MetAP related catalytic domain. It also possesses a C-terminal domain, composed of terminal extensions in red (the N-terminal polycharged domain of human MetAP2b is
basic and acidic stretches. (B) Comparison of archetype MetAP structures for types 1a, partially visible in the X-ray structure). The binary complex formed by fungal Arx1 and
1b, 1c, 2a and 2b (E. coli, Homo sapiens, Mycobacterium tuberculosis, Pyrococcus 60S ribosomal subunit is shown, with the same color code for Arx1 compared to
MetAPs (bottom right).
138 C. Giglione et al. / Biochimie 114 (2015) 134e146

MetAP2s (archaea), while eukaryotes have both classes. The major association of Ebp1 with mature ribosome has been also demon-
difference between MetAP1s and MetAP2s resides in the presence strated [74] and the crystal structure of the yeast Ebp1 homolog
of two additional compact sub-domains of unknown function Arx1 with the 60S ribosomal subunit has been recently solved
within the C-terminal catalytic domain of MetAP2s (Fig. 3A). The [75,76]. The binding of Arx1 has been mapped at the exit tunnel by
absence or the presence of specific N-terminal extensions of both directly contacting (Fig. 3) several rRNA elements and ribosomal
prokaryotic and eukaryotic MetAPs further differentiates these proteins uL23, uL29 and uL24 [75e77]. Although the 3D map of
enzymes. Thus, MetAP1s can be classified into four subclasses Arx1 bound to the 60S ribosomal subunit is not detailed, a never-
(Fig. 3). For example, E. coli possesses a MetAP1a displaying only the theless well-defined major contribution to the ribosome binding is
catalytic domain [60] (Fig. 3), whereas other bacteria such as acti- carried by the C-terminal catalytic domain insertion (Fig. 3B).
nobacteria display an additional MetAP1c characterized by an extra Hence, it was suggested (but not demonstrated) that MetAP2b
N-terminal stretch of 40 residues bearing an SH3 binding P-X-X-P might bind to a ribosome in a similar manner and position as Arx1
motif that precedes the catalytic domain and is suggested to be [75]. An interaction between uL23 and bL17 of the ribosome and
involved in binding to the ribosome [61] (Fig. 3). Since several ri- bacterial E. coli MetAP1a has been also shown [78] and involves a
bosomal proteins have SH3 domains and as uL24 is located adjacent positively charged loop as a binding site for ribosomes (the putative
to the ribosomal exit tunnel, it has been suggested that MetAP1c ribosome binding loop, Fig. 3B). Although no data is yet available
and the human type 1 MetAP might bind to this ribosomal protein about how many distinct MetAP family members interact in a
[61]. Recent findings have revealed that the N-terminal extension of productive manner with ribosomes and the nascent polypeptide
MtMetAP1c is indispensable for its activity likely by regulating chains, the two above cited studies together with other evidence
active site residues through a distance mechanism [62]. In the seem to decline a common strategy that might be exploited by the
cytoplasm of eukaryotes, on the other hand, MetAP1b contains two different MetAP family members.
domains composed by the catalytic domain followed by a 50-
residue linker and an N-terminal zinc finger domain proposed to 3. Protein N-a-Acetylation
be responsible of interaction with the ribosome [63] (Fig. 3A). The
fourth subclass of MetAP1s is represented by the organelle- After NME is completed, N-a-Acetylation (NTA) is probably the
targeted MetAP1d showing an N-terminal extension containing a second most common protein modification as it affects more than
transient N-terminal signal peptide removed upon translocation of 50% and 80% of yeast and human cytosolic proteins, respectively
the protein into the organelles [40] (Fig. 3). A genetic variant of [79e81]. Although protein acetylation is frequently associated with
MetAP1s (MetAP1a'), present predominantly in Streptococci and histone and Lys modifications, protein NTA is distinct and co-
Lactobacilli and displaying a new insert in the catalytic domain (i.e., translationally targets the N-terminus amino group of the pro-
Streptococcus pneumoniae MetAP1a', Fig. 3B), has been recently teins exclusively. This modification occurs also in archaea but with
described [64]. This insertion is likely to undergo post-translational a reduced frequency [82]. Available data indicate that bacterial NTA
modifications. Finally, MetAP2s are arranged in two subclasses also occurs on some proteins such as elongation factor Tu and some
(MetAP2a and MetAP2b) with MetAP2b containing an extra N- ribosomal proteins but more rarely than in other organisms (Fig. 1).
terminal domain composed by polybasic and polyacidic residue NTA was the subject of recent large scale proteomic studies that
blocks folding as an a-helix [65] (Fig. 3). One or two additional highlights the high frequency and the conservation of this process
insertions are present also in the catalytic domain of type 2 MetAPs even in organelles of eukaryotes [83]. Indeed, the presence of such
(Fig. 3). modification has been shown on the N-terminus of chloroplastic
The role(s) of each of the additional domains of the different proteins with an unexpected high frequency (more than 30% of the
MetAP enzymes is still largely unknown. It is nevertheless clear plastid proteome). Since NTA occurs on nuclear-encoded chloro-
that they are not required for aminopeptidase activity [66] and, plastic proteins after the cleavage of a transit peptide, this N-ter-
considering their location with respect to the active site, they are minal acetylation appears to be post-translational despite it could
unlikely to explain the slight differences of substrate specificity of also be co-translational for the few chloroplast-encoded proteins as
MetAP types [67]. Interestingly, the additional domains protrude on well (Fig. 1). Less clear is the impact of NTA on mitochondrial
the very same side of the 3D structure (Fig. 3B). It is therefore likely proteins to date, although the few data available on the mature N-
that these domains mediate interactions with various partners, and terminus of the nuclear encoded mitochondrial proteins seems to
contribute to regulate the NME process. Accordingly, it has been suggest that this modification should be rare.
shown that the first N-terminal polycharged Lys-rich block of Although it has been proven that this modification is essential
MetAP2b, previously called p67 [68], stores a POEP (protection of for cell viability [84e89] and survival [90,91], very little is known
eIF2a phosphorylation) activity [58], preventing the phosphoryla- about the physiological reasons associated with this crucial role.
tion of the alpha subunit of eukaryotic initiation factor 2 (eIF2a) by Since most accumulating cytoplasmic eukaryotic proteins are
interacting with it [69e71]. When unphosphorylated, eIF2a can terminally acetylated, NTA was considered to protect proteins from
bind to the ribosome and initiate translation, while its phosphor- degradation. Nonetheless, recent studies on specific protein targets
ylated form cannot bind, leading to inhibition of translation initi- have revealed that one biological function of NTA is to create spe-
ation [58]. The whole process contributes to regulate the overall cific degrons inducing degradation of the protein by the protea-
rate of protein synthesis. It is still unclear how the two independent some [28e30]. Involvement of NTA has been also shown to
activities, POEP and NME, are balanced taking into account the influence membrane targeting, proteineprotein interactions
extremely low level of the protein in the cell. Nonetheless, the [92e94] and changes in proteostasis [95].
generation by autoproteolysis or calpain action of shorter MetAP2
molecules of about 50e57 kDa and 26 kDa, preserving an NME and 3.1. The N-a-acetyltransferase enzyme
POEP activity respectively, might imply a separation of these two
activities at least in particular cellular conditions [72]. The ErbB-3 NTA is catalyzed by N-a-acetyltransferase (Nat) complexes
receptor binding protein (Ebp1), showing a conserved pita-bread generally composed in eukaryotes by a minimum of two sub-
MetAP fold devoid of NME activity, appears to also have a POEP units: a catalytic subunit accepting an acetyl-CoA as donor for
activity situated at its C-terminal region (Fig. 3), allowing the in- the transfer of the acetate moiety to the N-a-amino group
hibition of eIF2a phosphorylation in the cell [73,74]. Interestingly, substrate and an auxiliary subunit (Table 1). In eukaryotes, 6
C. Giglione et al. / Biochimie 114 (2015) 134e146 139

Table 1
Nats composition and their protein specificity. N-a-acetyltransferases (Nats) enzymes have been identified in all kingdoms of life. They contain at least one catalytic subunit,
which can be associated with one or two auxiliary subunit(s); names in parenthesis are for yeast proteins. No auxiliary subunits have been described for bacterial and archaeal
Nat. Naa10 and Naa15 are common to NatA and NatE, for which they play distinct roles. The substrate specificity is indicated for each class of Nats.

Eukaryotes Bacteria Archae


*
NatA NatB NatC NatD NatE NatF Rim Nat

Catalytic subunit Naa10 (Ard1) or Naa20 (Nat3) Naa30 (Mak3) Naa40 (Nat4) Naa50 (Nat5/San) Naa60 RimI/J/L NAT
Naa11a (Ard2) YhY
Auxilliary subunit(s) Naa15 (Nat1) or Naa25 (Mdm20) Naa35 (Mak10) d Naa10 d d d
Naa16a (Nat2) Naa38 (Mak31) Naa15
Substrates Ser- Met-Leu- Ser- Ser- / Ala-
Ala- Met-Asp- Met-Leu- Met-Leu- Met-Phe- Ala- Thr- / Gly-
Thr- Met-Glu- Met-Phe- Ser- Met-Phe- Met-Ile- Thr- Val- / Cys-
Gly- Met-Asn- Met-Ile- (H2A and Met-Ile- Met-Trp- Met-Leu-
Val- Met-Gln- Met-Trp- H4 histones) Met-Trp- Met-Lys- Met-Phe-
Cys- Met-Ile-
Met-Trp-
a
Naa11 and Naa16 are orthologues of Naa11 and Naa15 respectively, and are specific for higher eukaryotic NatAs. Each catalytic subunit of NatA (Naa10 and Naa11) can
therefore be associated to either one of the auxiliary subunits, leading to four different complexes. NatF has been identified only in higher eukaryotes.

distinct complexes have been identified so far (NatA, NatB, NatC, structures are generally positioned at the N and C-termini [109]
NatD, NatE, NatF), each of them able to modify a specific cate- (Fig. 4A). Moreover, the 3D structure of the active subunit of hu-
gory of N-terminal substrates with some overlap for various Nats man NatE (Naa50) (Table 1) reveals crucial differences at the level
differing also in subunit composition (Table 1) [for reviews see of the substrate-binding groove showing a more constricted
Refs. [81,96,97]]. NatA, NatB and NatC are thought to be protein-substrate binding site allowing to the enzyme to accom-
responsible for the majority of NTA events. The NatA complex, modate only an a-amino substrate versus side chain Lys substrates
which is composed of the catalytic subunits Naa10 or Naa11 and [106]. The structure of the S. pombe NatA complex shows that the
the auxiliary subunits Naa15 or Naa16, acetylates N-termini auxiliary Naa15 subunit is composed of 37 a-helices, among which
exposed after removal of the first Met by NME (Fig. 1), mainly 13 have conserved bundle tetratricopeptide repeat (TPR) motifs,
Ser, Ala, Gly, Thr and Val (Table 1). Monomeric Naa10 has been and wraps completely the Naa10 catalytic subunit. A number of
shown to also acetylate post-translationally mature actins these TPR-helices are involved in Naa10 interaction and it has
exposing Asp or Glu N-termini but not traditional NatA sub- been also suggested to favor interaction with other NatA binding
strates [98]. Moreover Naa10 seems to have also a lysine-specific partners (Naa50 in NatE, see Table 1) or the ribosome [105].
acetyltransferase KAT activity, transferring the acetyl group form Interestingly, the comparison of the 3D structures of fungal Naa10
the Ac-CoA to the ε-amino groups of internal lysines of proteins, shows a conformational change of the protein when it interacts
and even an AT-independent activity, controlling the activity or with Naa15, especially in the loop connecting a-helices 1 and 2
binding with other partners [for review see Ref. [97]]. NatB and [105]. This suggests that the auxiliary subunit Naa15 induces an
NatC are potentially NME-independent and acetylate proteins allosteric conformational change of this Naa10 loop, leading to an
retaining the initial Met when the second residue is acidic active conformation of the NatA catalytic subunit active site
(NatB) or hydrophobic (NatC) (Table 1). In eubacteria, RimI, RimJ, required for the catalytic mechanism of the NatA complex
RimL and YhY are believed to be homologs of eukaryotic Nats, employing its traditional substrates [105]. Finally, the superim-
whereas only a single Nat complex with a broader substrate position of the Naa10 and Naa50 active sites revealed the impor-
specificity than eukaryotic NatA complexes seems to occur in tance of Naa10 Glu24 for excluding Glu and Met as amino-
archaea [82,99,100] (Table 1). Although large proteomic analyses terminal substrates and highlights the significance of Val29 in
of several plant species have shown substrate similarity to hu- human Naa50 for N-terminal Met recognition and acetylation
man Nats, limited information is available on plant NTA ma- [105,106] (Fig. 4B). Interestingly, the 3D structure of the archaeal
chinery [83,101e104]. ortholog, able to acetylate NatA and NatE substrates, presents the
To date, several Nats have been structurally characterized. This acetyl-CoA binding core region and the flanking N and C-terminal
includes 2 eukaryotic Nats, the Schizosaccharomyces pombe NatA parts highly similar to the active form of Naa10 and Naa50.
complex consisting of the Naa10 and Naa15 subunits, the human However, it displays a non-conserved feature in the catalytic core
NatE (the catalytic subunit Naa50), the unique archaeal Nat from of the enzyme composed by an extended b3-b4 loop important to
Sulfolobus solfataricus and several bacterial Nats [105e108]. The 3D stabilize the position of key residues such as His137 involved
structures of all reported Nats place these enzymes in the Gcn5- particularly in Met amino terminal substrate acetylation [100].
related N-acetyltransferase (GNAT) family, a large superfamily of Besides, there are several GNAT enzymes that take advantage of
enzymes that use acyl-CoAs to acylate their substrates [109]. the b3-b4 loop variability for functional gain [100]. It was there-
Members of the GNAT family are indeed characterized by a nearly fore suggested that the archaeal Nat, by showing a larger substrate
universally structurally conserved fold including an N-terminal specificity than eukaryotic orthologs, represents an ancestral Nat
strand followed by two helices, three antiparallel b strands, a variant.
central helix signature followed by a fourth a helix and a final b
strand [109]. The GNAT members use the same structural element 3.2. Co-translational N-a-acetylation and its control
to serve as their oxyanion hole, i.e., the backbone amides in a
conserved “b bulge” located on strand b4 to perform acylation via Several lines of evidence indicate that NTA can occur co- or post-
a variety of different chemical stategies involving sometimes translationally [83,102,110e112]. Studies using synthetic substrates
dramatic conformational changes. Differences between GNAT and purified Nats have concluded that the protein substrate must
140 C. Giglione et al. / Biochimie 114 (2015) 134e146

Fig. 4. Similarities and differences between N-a-acetyltransferases (Nats) and N-myristoyltransferases (NMTs). (A) Both Nats and NMTs belong to the GNAT superfamily, which
members use acyl-CoAs to acylate their substrates. All members of GNAT superfamily display a common GNAT fold, consisting of an acyl-CoA binding site and a peptide binding site.
The GNAT fold is highly conserved around these ligand binding sites. Measurement of evolutionary conservation of amino acid positions in GNAT family members was achieved
with ConSurf 2010 [175,176]. The search for homologous sequences of Naa10 in its active form (PDB code 4KVX) was performed using PSI-BLAST against UniRef90 sequences
databank, the other parameters being the default ones. The continuous conservation scores are divided into a discrete scale of nine grades for visualization, from the most variable
positions (grade 1) colored turquoise, through intermediately conserved positions (grade 5) colored white, to the most conserved positions (grade 9) colored brown. These con-
servation scores are projected onto Naa10 structure represented as molecular surface. (B) Nats display the classical GNAT fold, with acetyl-CoA and peptide binding sites. Except for
NatD, the Nat substrate specificity lies within the first two to five residues, leading to a largely open peptide binding site (top inset). Residues contributing to the substrate specificity
are shown for Naa10 and Naa50 (middle and bottom insets). (C) The NMT fold is composed of two adjacent GNAT domains (highlighted in yellow and magenta), with the myristoyl-
CoA binding site carried by the N-terminal GNAT domain. The peptide binding site, recognizing the first eight amino acids, is shared by both the two GNAT domains (top and middle
insets). Residues involved in peptide binding are shown for human NMT1 (bottom inset).

contain a minimum of 10 and a maximum of 25 residues to be 4. N-myristoylation


acetylated [113,114]. Alternatively, in vivo investigations of yeast
and mouse NatA defined that association of the complex with a N-terminal Myristoylation (MYR) is one of the most critical lipid
ribosome is mediated by the auxiliary subunit Naa15 and required a protein-modifications and consists in the addition of a saturated
length of the nascent polypeptide emerging from the ribosome C:14 fatty acid to a free N-terminal Gly residue of eukaryotic and
(40e70 amino acids) longer than nascent chain necessary for viral proteins exhibiting a favorable context at the level of the first 8
interaction with the nascent polypeptide-associated complex NAC, residues [122e125]. In rare cases, proteins from lower eukaryotic
an eukaryotic chaperone dimer, or the Ssb1p/Ssb2p proteins parasites and bacteria can also undergo MYR by the eukaryotic host
forming a chaperone triad with the ribosome-associated complex cell [126]. An N-terminal Gly is exposed mostly as a consequence of
RAC [115e119]. This ribosome association seems not to be depen- NME very early during protein biosynthesis (Fig. 1), although recent
dent on the nature of the amino acid sequence of the nascent studies indicate that such a modification can occur also later on a
polypeptide [116]. NatB, NatC and NatE have also been shown to newly exposed N-terminal Gly after a proteolytic cleavage medi-
colocalize with ribosome and interaction between Naa15 and uL23 ated by caspases during apoptosis or by removal of a leader peptide
and uL29 has been highlighted [111]. from bacterial proteins imported into eukaryotic cells [10,127].
Until recently, NTA was mostly considered as an irreversible Thus, MYR as NTA can occur in a co- or a post-translational manner.
modification although some recent results [31,79,102,120] have MYR is believed to cover a significant part of any eukaryotic
uncovered the occurrence of both the acetylated and non- proteome, reaching 1e4% of all proteins in a given proteome [80]
acetylated forms in the same sample of eukaryote species and (Fig. 1). Nonetheless, the exact composition of such a proteome
involving about 9% of all experimentally identified N-termini in remains still problematic to identify because i) it is technically
human cells [79]. Moreover, E. coli ribosomal protein L12/L7 is also difficult to evidence the lipidated proteins from homogenous bio-
known to exist in two different forms i.e., N-acetylated and the logical samples, ii) the available experimental approaches aimed at
unacetylated counterpart. The total amount of both forms remains validating MYR from complex in vivo samples are limited as the
constant but their ratio is dependent on the growth cycle [121]. No corresponding protein usually belongs to the low abundance
N-alpha deacetylase is known so far but the characterization of membrane proteome in any organism and iii) it is tricky to obtain
both modified and unmodified products change radically the idea reliable MYR prediction with the bioinformatics tools available so
of a completeness of this modification. The origin of the unacety- far [128]. Hence, only a very small subset of putative myristoylated
lated counterpart of a protein is yet to be ascertained. Several non- proteins has been experimentally checked through individual
exclusive hypotheses can be formulated such as: i) recurrent weak studies [for review see discussion in Refs. [129]], while bioinfor-
identification of the N-terminal substrates by Nats (see above), ii) matics prediction tools developed by various teams worldwide
regulation of Nat activity, e.g., availability of AcCoA has been shown indicate that the majority of myristoylated protein targets remain
to control the degree of NTA of specific proteins [91] or/and iii) to be confirmed [80,130e133].
regulation of ribosome Nat-binding.
NTA and the enzymes involved in this modification are currently 4.1. N-myristoyltransferases
among the most challenging topics of molecular cell biology with
many open questions [96] which are expected to find an answer in MYR is catalyzed by an acylase called myristoyl CoA:protein N-
the near future. myristoyltransferase (NMT), which transfers the myristoyl group
C. Giglione et al. / Biochimie 114 (2015) 134e146 141

from a myristoyl-CoA onto the N-terminal Gly of the targeted response functions in animals and plants [126]. Together with
protein (Fig. 1). Higher eukaryotes share very similar N-myr- other proximal signals, the main role of this lipidation is to target
istoylation enzyme systems, involving two NMTs (NMT1 and the protein to specific membrane compartments where the protein
NMT2) among which NMT1 is the most abundant enzyme will be an essential regulatory component of signal transduction
[123,125] and most efficient in vitro [129]. Lower eukaryotes display pathways involving many kinases or small G proteins. Due to its
only one NMT. NMT1 and NMT2 display a high level of sequence function in cellular communication networks, it should be expected
identity with the major differences residing at the N-terminus of that the myristoylated proteome is intrinsically dynamic to ensure
the enzymes. As Nats, NMTs are monomeric enzymes belonging to a proper cell communication. However, MYR has been considered
the GNAT superfamily [109]. Interestingly, the first 3D structures of for long time an irreversible modification being necessary but not
fungal NMTs revealed the presence of both an N-terminal and a C- sufficient to allow a protein to membrane anchoring [148,149]
terminal GNAT domains, with low sequence homology to each suggesting that the dynamics of the proteome was mainly
other, connected by a long loop [134,135] (Fig. 4C). Myristoyl-CoA is ensured by the reversibility of the second signal. For several evi-
bound to the N-terminal GNAT domain while the peptide is bound denced myristoylated proteins it has been shown that the second
to a substrate binding site shared by both GNAT domains (Fig. 4C). signal can be provided by attachment of another lipid a palmitate,
Particularly, the N-terminal Gly of a substrate-peptide becomes or a polybasic motif or a domain that will interact with another
positioned towards the N-terminal GNAT domain, whereas the rest membrane-bound protein. Surprisingly, the dynamic scenario of
of the recognized-peptide lies on a long groove situated on the C- the myristoylated proteome has been revised by recent break-
terminal GNAT domain (Fig. 4C). From structural and biochemical through studies revealing for the first time the enzymatic removal
investigation of the yeast and human NMTs it seems that these of the N-terminal moiety on specific proteins by the Shigella viru-
enzymes follow an ordered biebi kinetic mechanism proceeding lence factor IpaJ [150] and that MYR alone localizes the proteins to
through a myristoyl-CoA binding to the free enzyme, followed by the endomembrane system with a partitioning between this
the peptide substrate, and the released of CoA before the myr- membrane compartment and the cytosol correlated with the cat-
istoylated protein [136e138]. The conserved oxyanion hole alytic efficiency of the NMT [151,152]. In the latter case, it was
involved during the catalytic mechanism, formed by the main chain suggested that NMT and Nat complexes could compete at the level
amide backbones of both F170 and L171, polarizes the carbonyl of of the ribosome for binding to specific nascent chains. This was
the thioester of the myristoyl-CoA and stabilizes the tetrahedral shown for two protein mutants, R3G-TNF and R3M-TNF, found to
intermediate generated after the binding of the peptide Glycine- become both N-acetylated and myristoylated and in an in vitro
residue followed by nucleophilic attack by the amine and elimi- translational system showing that the same subset of proteins
nation of the thiol of CoA. displays both modifications [153,154] Indeed, most available pre-
To date, it is assumed that all NMTs have common characteris- diction tools of these modifications, even if quite accurate, show
tics but they are not completely redundant, suggesting distinctive weaknesses in the prediction of NTA and MYR on N-Gly residues
peptide substrate specificities [123,124,139e141]. Together with [83], indicating that the two patterns might overlap and thus
the failure in finding a unique consensus sequence for the myr- compete for the corresponding modification enzymes when they
istoylable site, this implies that NMTs might adopt different cata- emerge from the exit tunnel of the ribosome.
lytic strategies, as already observed for other GNAT members
including Nats (see above) to accomplish substrate-specific MYR. 5. Nascent polypeptide meeting ribosome-associated NPMs
This would enable the enzyme to recognize several consensus and interplay with other RPBs
motifs with different catalytic efficiencies [80,129,142].
Translation requires intricate communication among multiple
4.2. N-myristoyltransferase and translation components to achieve speed, accuracy, and regulation. Commu-
nication between ribosomes, translation factors, tRNA, mRNA, and
The N-terminus of NMTs is not required for catalysis [123,125] other cellular pathways is the mechanistic basis for the multiple
and it was suggested that it might be involved in targeting the and tight levels of control that exist in translation. Recent findings
enzyme to various subcellular environments such as the trans- have revealed the functional importance of non-ribosomal proteins
lational machinery [143]. Indeed, interaction of NMT with poly- that permanently or transiently associate with the ribosome.
ribosomes has been evidenced in the past [144,145]. Since it was Among non-ribosomal proteins, ribosome-associated protein
observed that during protein biosynthesis the myristate was more biogenesis factors (RPBs) are molecules acting as soon as a nascent
efficiently attached to a model myristoylated protein in the first polypeptide reaches the exit from the ribosomal tunnel. Indeed, the
10 min of translation and considering an average rate of in vitro ribosome is a platform for co-translational events affecting the
polypeptide chain elongation for this protein of 10.5 residues per nascent polypeptide, including protein folding, targeting to various
min, it was suggested for all predicted myristoylated proteins that cellular compartments for integration into membrane or trans-
this modification might occur co-translationally when the length of location, protein quality control and protein modifications. How the
the nascent chain is less than 100 amino acids [146]. Recently, it RPBs involved in these processes act upon the nascent polypeptide
was shown for both human NMT1 and 2 that the N-terminal basic chain and how they dialog at the level of the exit tunnel turn out to
amino acid region called K box is sufficient and essential for binding be one of the most exciting and challenging questions of the last
to the ribosome [147]. However, this K box is not conserved in other years. As described in this review and others [see for instance
NMTs, leaving open the question of which part of these NMTs is [3,155]], the region around the exit tunnel seems to constitute a
involved in ribosome binding or even which region on the ribo- general docking site for almost all investigated RPBs.
some is concerned.
5.1. Ribosome as NPMs partner
4.3. The physiological functions of myristoylation
Since early seventies, it was accepted that PDFs and MetAPs
Despite the poor knowledge on the whole pool of proteins un- were the first RPBs to interact with the nascent chain. Even before
dergoing to MYR, it is already known that this modification is cloning the first PDF gene, it was noticed that in bacteria significant
linked with fundamental cancer-related [125] or pathogen- amount of deformylation always occurred within the first 5e8 s, a
142 C. Giglione et al. / Biochimie 114 (2015) 134e146

time that can increase to 10e20 s during starvation of one or more these enzymes to scan ribosomes in a very short time compared to
amino acids [156]. Considering a constant elongation rate of 5 translation speed [37,49]. Although no similar kinetic constants
residues/sec in vivo, it was concluded that deformylation was taking are available for MetAPs, same results were extrapolated for these
place when the length of the nascent chain was between 25 and processing enzymes. For the moment, it remains a puzzle what is
100 amino acids. In parallel, studies on specific bacterial nascent the proportional weight of the respective independent contribu-
chains have demonstrated that removal of the first formyl-Met was tion of PDF/MetAP-ribosome binding and the rapid PDF/MetAP
occurring when the nascent chains where about 40e50 amino kinetics for an efficient and rapid NME process on a whole pro-
acids long confirming previous results [78,157e159] (Fig. 5). Many teome. Even more diffuse is our knowledge on the interplay be-
genes encoding PDFs and MetAPs with classical enzymatic activity tween the NME enzymes and either bacterial or organellar
but displaying specific features have been discovered in most or- chaperone protein trigger factors (TF), for which the bacterial form
ganisms and cellular compartments in the last fifteen years (see is known to interact with the bacterial ribosomal proteins uL23,
above) and several of them have been shown to be present and act uL24 and uL29 [162e164]. After an initial hypothesis that PDF and
simultaneously. Although most of these PDFs and MetAPs have TF can act in a concerted fashion [53], it was later shown that
been extensively characterized enzymatically and structurally, E. coli TF and PDF1B compete for ribosome binding in vivo [165],
nothing is known about their behavior in respect to their rate of i.e., TF overexpression slightly reduced cell growth, a phenotype
deformylation and Met removal from the nascent chain in vivo, i.e., particularly evident when cells expressing endogenous E. coli PDF
the length of the nascent chain when these processes are ensured were treated with the specific PDF inhibitor actinonin. This inhi-
by each of them. No impact of the nature of the sequence has been bition by TF overexpression was even exacerbated when actinonin
shown in vitro for several PDFs [160,161], whereas no data are inhibiting cells expressed the C-terminally truncated version of
available to date on PDF nascent chain association properties E. coli PDF instead of the WT form [78]. Since it was previously
in vivo. This is true also for the different MetAP enzymes for which a shown that the 21 last residues folding the C-terminal helical
predominant dependency on the nature of the second residue and extension of E. coli PDF is not necessary for ensuring the essential
of the 4e5 following residues has been shown in vitro [67]. None- role of PDF in vivo [161,166], it is uneasy to fully explain these data
theless, using ribosome-nascent chain complexes comprising 40 N- to date, although interaction with the ribosome of the latter
terminal residues of enolase fused to the SecM arrest sequence it extension was observed [53]. To even further complicate the sce-
was shown that this nascent chain was not able to increase the nario several in vitro binding data show that TF and PDF or MetAP
ribosome affinity for E. coli PDF1B and MetAP1b [78]. can bind simultaneously to ribosomes or RNCs [78,167]. Particu-
As discussed in the previous section, some Nat auxiliary sub- larly, it has been proposed that TF remained bound when PDF
units have been shown to interact with the ribosomal docking interacts with specific RNCs or non-translating ribosomes (only
protein uL23 and uL29 (Fig. 5) but no information to our knowledge partial or no changes of Kd values for binding of TF to ribosome
is available on NMT ribosomal binding proteins. Since the two were observed in the presence of different PDF concentrations),
classes of enzymes display the same GNAT fold it is tempting to but in a modified arrangement [167]. The same results were re-
speculate that NMTs will also lie on the ribosome in proximity of ported for E. coli MetAP. Recent in vitro studies have also consid-
the exit tunnel. Nonetheless, knowledge on how exactly NMTs and ered the interplay at the level of the exit tunnel of the ribosome of
Nats interact with the ribosome and other RPBs will shed light not another RPB factor, the signal recognition particle (SRP) involved
only in the mechanistic aspects of their interaction with ribosome in translocation of specific membrane proteins, with PDF or MetAP
but also on their essential but not fully comprehended function(s). and have revealed that the binding of PDF or MetAP to RNCs
doesn't influence the binding of SRP as already observed with TF
[167].
5.2. Cooperation and competition between RPBs

In vitro biochemical studies have shown that E. coli PDF1B and 6. Conclusions-future perspective
MetAP1b interact with ribosome-binding sites very close to each
other (around uL22, bL17 and bL32 for the first enzyme and bL17 The ribosome was proposed early on as a partner of NPM en-
and uL23 for the second one, Fig. 5) [53,78], raising the specula- zymes [see for instance as a pioneering work Ref. [168]]. However,
tion that in vivo they might influence each other's interaction on the inability to isolate these enzymes stably associated with ribo-
the ribosome. Although not complete, an in vitro competition some preparations but being able to fully reproduce their activity
between E. coli PDF1B and MetAP1b for binding to non-translating and specificity in vitro in the absence of ribosomes have led the
or translating ribosomes has been evidenced [78]. However, it was scientific community to neglecting the “ribosome contribution” on
suggested that the extremely rapid kinetics of PDFs should allow these NPMs. During the recent years, several data have clearly

Fig. 5. Currently known ribosomal proteins involved in binding with NPM enzymes. This picture summarizes the whole set of different N-terminal modifications, the related
enzymatic factors (so-called RPB factors), the suggested nascent chain length when these modifications are expected to occur as well as the known ribosomal proteins interacting
with each RPB factors. The location of these ribosomal proteins is highlighted in bacterial and eukaryotic cytoplasmic ribosomes (right).
C. Giglione et al. / Biochimie 114 (2015) 134e146 143

established that the ribosome is indeed an additional player at least [7] T. Meinnel, Y. Mechulam, F. Dardel, J.M. Schmitter, C. Hountondji, S. Brunie,
P. Dessen, G. Fayat, S. Blanquet, Methionyl-tRNA synthetase from E. coliea
for some NPMs. Several alternative routes for involving an NPM are
review, Biochimie 72 (1990) 625e632.
nevertheless feasible, as NPMs appear to be added also post- [8] E. Schmitt, J.M. Guillon, T. Meinnel, Y. Mechulam, F. Dardel, S. Blanquet,
translationally without the ribosome as a required partner of the Molecular recognition governing the initiation of translation in Escherichia
respective enzymes involved. Thus, the exact role and the very coli. A review, Biochimie 78 (1996) 543e554.
[9] E. Schmitt, T. Meinnel, M. Panvert, Y. Mechulam, S. Blanquet, Two acidic
requirement of ribosome association need to be cautiously residues of Escherichia coli methionyl-tRNA synthetase act as negative dis-
addressed. Unambiguously however, physical interaction of NPMs criminants towards the binding of non-cognate tRNA anticodons, J. Mol. Biol.
with ribosomes brings closer the nascent chain and the protein 233 (1993) 615e628.
[10] T. Meinnel, C. Giglione, Tools for analyzing and predicting N-terminal protein
modifiers. Such proximity is probably required to ensure full modifications, Proteomics 8 (2008) 626e649.
modification of a protein without any possibility for it to escape the [11] C. Giglione, A. Boularot, T. Meinnel, Protein N-terminal methionine excision,
modification prior to being completed. Cell. Mol. Life Sci. 61 (2004) 1455e1474.
[12] J.E. Walker, J. Carroll, M.C. Altman, I.M. Fearnley, Chapter 6 Mass spectro-
It is important to mention that despite the organization of the metric characterization of the thirteen subunits of bovine respiratory com-
proteins around the exit tunnel is quite conserved in all studied ri- plexes that are encoded in mitochondrial DNA, Methods Enzymol. 456
bosomes, a much greater number of proteins is involved in the (2009) 111e131.
[13] M. Unseld, J.R. Marienfeld, A. Brennicke, The mitochondrial genome of Ara-
eukaryotic ribosomal exit tunnel core itself (Fig. 5) and that the bidopsis thaliana contains 57 genes in 366,924 nucleotides, Nat. Genet. 15
eukaryotic “welcome committee” of proteins more loosely associ- (1997) 57e61.
ated to the ribosome (including NPMs) is larger compared to its [14] S. Escobar-Alvarez, Y. Goldgur, G. Yang, O. Ouerfelli, Y. Li, D.A. Scheinberg,
Structure and activity of human mitochondrial peptide deformylase, a novel
prokaryotic counterpart. Nonetheless, the recent high resolution
cancer target, J. Mol. Biol. 387 (2009) 1211e1228.
structure of the large subunits of both yeast and mammal mito- [15] C. Giglione, T. Meinnel, Organellar peptide deformylases: universality of the
ribosome revealed several unique features including the region of N-terminal methionine cleavage mechanism, Trends Plant Sci. 6 (2001)
the exit tunnel making this complex divergent from other types of 566e572.
[16] H. Randhawa, S. Chikara, D. Gehring, T. Yildirim, J. Menon, K.M. Reindl,
ribosomes [169e171]. Moreover, differences have been noticed even Overexpression of peptide deformylase in breast, colon, and lung cancers,
between the yeast and mammal mito-ribosomes and involving also BMC Cancer 13 (2013) 321.
the previous identified exit tunnels (e.g., the conventional poly- [17] X. Bao, N.D. Pachikara, C.B. Oey, A. Balakrishnan, L.F. Westblade, M. Tan,
T. Chase Jr., B.E. Nickels, H. Fan, Non-coding nucleotides and amino acids near
peptide exit site, PES, and the polypeptide accessible site, PAS). This the active site regulate peptide deformylase expression and inhibitor sus-
led to the proposal that they both are putative alternative interaction ceptibility in Chlamydia trachomatis, Microbiology 157 (2011) 2569e2581.
sites for different classes of enzymes involved in the co-translational [18] J. Chiu, J.W. Wong, P.J. Hogg, Redox regulation of methionine aminopepti-
dase 2 activity, J. Biol. Chem. 289 (2014) 15035e15043.
events. However, the presence of a long mitochondria-specific [19] B.D. Dill, S. Dessus-Babus, J.E. Raulston, Identification of iron-responsive
extension of uL23 in the PES of yeast suggested that PAS is the proteins expressed by Chlamydia trachomatis reticulate bodies during
only active protein channel in yeast mito-ribosomes. This seems not intracellular growth, Microbiology 155 (2009) 210e219.
[20] M.D. Cleary, U. Singh, I.J. Blader, J.L. Brewer, J.C. Boothroyd, Toxoplasma gondii
to be the case for the mammal mito-ribosome leaving still open the asexual development: identification of developmentally regulated genes and
issue of the functionality of both PES and/or PAS [169,172]. Protein distinct patterns of gene expression, Eukaryot. Cell. 1 (2002) 329e340.
expansion segments have been found on most of the ribosomal [21] B. Dasgupta, Y. Yi, B. Hegedus, J.D. Weber, D.H. Gutmann, Cerebrospinal fluid
proteomic analysis reveals dysregulation of methionine aminopeptidase-2
proteins from cytosolic, chloroplastic or mitochondrial ribosome,
expression in human and mouse neurofibromatosis 1-associated glioma,
suggesting that this may be a prerequisite for keeping or even Cancer Res. 65 (2005) 9843e9850.
improving efficient, timely and specific interactions with the [22] P. Selvakumar, A. Lakshmikuttyamma, J.R. Dimmock, R.K. Sharma, Methio-
increased number of co-translational NPM enzymes. nine aminopeptidase 2 and cancer, Biochim. Biophys. Acta 1765 (2006)
148e154.
[23] P. Selvakumar, A. Lakshmikuttyamma, Z. Lawman, K. Bonham, J.R. Dimmock,
R.K. Sharma, Expression of methionine aminopeptidase 2, N-myristoyl-
Conflict of interest transferase, and N-myristoyltransferase inhibitor protein 71 in HT29, Bio-
chem. Biophys. Res. Commun. 322 (2004) 1012e1017.
The authors declare no conflict of interest. [24] S.M. Arfin, R.A. Bradshaw, Cotranslational processing and protein turnover in
eukaryotic cells, Biochemistry 27 (1988) 7979e7984.
[25] C. Giglione, O. Vallon, T. Meinnel, Control of protein life-span by N-terminal
methionine excision, EMBO J. 22 (2003) 13e23.
Acknowledgments [26] A. Varshavsky, The N-end rule pathway and regulation by proteolysis, Pro-
tein Sci. 20 (2011) 1298e1345.
This work was supported by ANR projects Ribo-Dyn (ANR-2010- [27] D.J. Gibbs, J. Bacardit, A. Bachmair, M.J. Holdsworth, The eukaryotic N-end
rule pathway: conserved mechanisms and diverse functions, Trends Cell Biol.
BLAN-1510) and PalMyProt (ANR-2010-BLAN-1611) and grant ARC 24 (2014) 603e611.
SFI20111203841. The authors thank Dr Bruno Gronenborn for crit- [28] C.S. Hwang, A. Shemorry, A. Varshavsky, N-terminal acetylation of cellular
ical reading of the manuscript and Dr Adina Breiman (Tel Aviv proteins creates specific degradation signals, Science 327 (2010) 973e977.
[29] A. Shemorry, C.S. Hwang, A. Varshavsky, Control of protein quality and
University) for strong support. stoichiometries by N-terminal acetylation and the N-end rule pathway, Mol.
Cell 50 (2013) 540e551.
[30] H.K. Kim, R.R. Kim, J.H. Oh, H. Cho, A. Varshavsky, C.S. Hwang, The N-terminal
References methionine of cellular proteins as a degradation signal, Cell 156 (2014)
158e169.
[1] R.G. Krishna, F. Wold, Post-translational modification of proteins, Adv. [31] F. Frottin, C. Espagne, J.A. Traverso, C. Mauve, B. Valot, C. Lelarge-Trouverie,
Enzymol. Relat. Areas Mol. Biol. 67 (1993) 265e298. M. Zivy, G. Noctor, T. Meinnel, C. Giglione, Cotranslational proteolysis
[2] U. Raue, S. Oellerer, S. Rospert, Association of protein biogenesis factors at dominates glutathione homeostasis to support proper growth and devel-
the yeast ribosomal tunnel exit is affected by the translational status and opment, Plant Cell. 21 (2009) 3296e3314.
nascent polypeptide sequence, J. Biol. Chem. 282 (2007) 7809e7816. [32] D. Mazel, S. Pochet, P. Marliere, Genetic characterization of polypeptide
[3] G. Kramer, D. Boehringer, N. Ban, B. Bukau, The ribosome as a platform for deformylase, a distinctive enzyme of eubacterial translation, EMBO J. 13
co-translational processing, folding and targeting of newly synthesized (1994) 914e923.
proteins, Nat. Struct. Mol. Biol. 16 (2009) 589e597. [33] T. Meinnel, S. Blanquet, Characterization of the Thermus thermophilus locus
[4] M. Kozak, Comparison of initiation of protein synthesis in procaryotes, eu- encoding peptide deformylase and methionyl-tRNA(fMet) formyltransferase,
caryotes, and organelles, Microbiol. Rev. 47 (1983) 1e45. J. Bacteriol. 176 (1994) 7387e7390.
[5] N. Malys, J.E. McCarthy, Translation initiation: variations in the mechanism [34] L.M. Dirk, M.A. Williams, R.L. Houtz, Eukaryotic peptide deformylases.
can be anticipated, Cell. Mol. Life Sci. 68 (2011) 991e1003. Nuclear-encoded and chloroplast- targeted enzymes in Arabidopsis, Plant
[6] J.M. Guillon, Y. Mechulam, J.M. Schmitter, S. Blanquet, G. Fayat, Disruption of Physiol. 127 (2001) 97e107.
the gene for Met-tRNA(fMet) formyltransferase severely impairs growth of [35] A. Serero, C. Giglione, T. Meinnel, Distinctive features of the two classes of
Escherichia coli, J. Bacteriol. 174 (1992) 4294e4301. eukaryotic peptide deformylases, J. Mol. Biol. 314 (2001) 695e708.
144 C. Giglione et al. / Biochimie 114 (2015) 134e146

[36] N. Bouzaidi-Tiali, C. Giglione, Y. Bulliard, M. Pusnik, T. Meinnel, A. Schneider, [64] T. Arya, C. Kishor, V. Saddanapu, R. Reddi, A. Addlagatta, Discovery of a new
Type 3 peptide deformylases are required for oxidative phosphorylation in genetic variant of methionine aminopeptidase from Streptococci with
Trypanosoma brucei, Mol. Microbiol. 65 (2007) 1218e1228. possible post-translational modifications: biochemical and structural char-
[37] S. Ragusa, S. Blanquet, T. Meinnel, Control of peptide deformylase activity by acterization, PLoS One 8 (2013) e75207.
metal cations, J. Mol. Biol. 280 (1998) 515e523. [65] S. Liu, J. Widom, C.W. Kemp, J. Clardy, Structure of human methionine
[38] C. Giglione, M. Pierre, T. Meinnel, Peptide deformylase as a target for new aminopeptidase-2 complexed with fumagillin, Science 282 (1998)
generation, broad spectrum antimicrobial agents, Mol. Microbiol. 36 (2000) 1324e1327.
1197e1205. [66] G. Yang, R.B. Kirkpatrick, T. Ho, G.F. Zhang, P.H. Liang, K.O. Johanson,
[39] I. Sharon, N. Battchikova, E.M. Aro, C. Giglione, T. Meinnel, F. Glaser, R.Y. Pinter, D.J. Casper, M.L. Doyle, J.P. Marino Jr., S.K. Thompson, W. Chen, D.G. Tew,
M. Breitbart, F. Rohwer, O. Beja, Comparative metagenomics of microbial traits T.D. Meek, Steady-state kinetic characterization of substrates and metal-ion
within oceanic viral communities, ISME J. 5 (2011) 1178e1190. specificities of the full-length and N-terminally truncated recombinant hu-
[40] C. Giglione, A. Serero, M. Pierre, B. Boisson, T. Meinnel, Identification of man methionine aminopeptidases (type 2), Biochemistry 40 (2001)
eukaryotic peptide deformylases reveals universality of N-terminal protein 10645e10654.
processing mechanisms, EMBO J. 19 (2000) 5916e5929. [67] F. Frottin, A. Martinez, P. Peynot, S. Mitra, R.C. Holz, C. Giglione, T. Meinnel,
[41] J.A. Frank, D. Lorimer, M. Youle, P. Witte, T. Craig, J. Abendroth, F. Rohwer, The proteomics of N-terminal methionine cleavage, Mol. Cell. Proteomics 5
R.A. Edwards, A.M. Segall, A.B. Burgin Jr., Structure and function of a (2006) 2336e2349.
cyanophage-encoded peptide deformylase, ISME J. 7 (2013) 1150e1160. [68] X. Li, Y.H. Chang, Evidence that the human homologue of a rat initiation
[42] V. Seguritan, I.W. Feng, F. Rohwer, M. Swift, A.M. Segall, Genome sequences factor-2 associated protein (p67) is a methionine aminopeptidase, Biochem..
of two closely related Vibrio parahaemolyticus phages, VP16T and VP16C, Biophys. Res. Commun. 227 (1996) 152e159.
J. Bacteriol. 185 (2003) 6434e6447. [69] R. Datta, P. Choudhury, M. Bhattacharya, F. Soto Leon, Y. Zhou, B. Datta,
[43] M.K. Chan, W. Gong, P.T. Rajagopalan, B. Hao, C.M. Tsai, D. Pei, Crystal Protection of translation initiation factor eIF2 phosphorylation correlates
structure of the Escherichia coli peptide deformylase, Biochemistry 36 (1997) with eIF2-associated glycoprotein p67 levels and requires the lysine-rich
13904e13909. domain I of p67, Biochimie 83 (2001) 919e931.
[44] F. Dardel, S. Ragusa, C. Lazennec, S. Blanquet, T. Meinnel, Solution structure of [70] R. Datta, R. Tammali, B. Datta, Negative regulation of the protection of
nickel-peptide deformylase, J. Mol. Biol. 280 (1998) 501e513. eIF2alpha phosphorylation activity by a unique acidic domain present at the
[45] J.P. Guilloteau, M. Mathieu, C. Giglione, V. Blanc, A. Dupuy, M. Chevrier, P. Gil, N-terminus of p67, Exp. Cell. Res. 283 (2003) 237e246.
A. Famechon, T. Meinnel, V. Mikol, The crystal structures of four peptide [71] B. Datta, R. Datta, A. Ghosh, A. Majumdar, The binding between p67 and
deformylases bound to the antibiotic actinonin reveal two distinct types: a eukaryotic initiation factor 2 plays important roles in the protection of
platform for the structure-based design of antibacterial agents, J. Mol. Biol. eIF2alpha from phosphorylation by kinases, Arch. Biochem. Biophys. 452
320 (2002) 951e962. (2006) 138e148.
[46] A. Serero, C. Giglione, A. Sardini, J. Martinez Sanz, T. Meinnel, An unusual [72] T. Clinkinbeard, S. Ghoshal, S. Craddock, L. Creed Pettigrew, R.P. Guttmann,
peptide deformylase features in the human mitochondrial N-terminal Calpain cleaves methionine aminopeptidase-2 in a rat model of ischemia/
methionine excision pathway, J. Biol. Chem. 278 (2003) 52953e52963. reperfusion, Brain Res. 1499 (2013) 129e135.
[47] S. Fieulaine, C. Juillan-Binard, A. Serero, F. Dardel, C. Giglione, T. Meinnel, [73] E. Kowalinski, G. Bange, B. Bradatsch, E. Hurt, K. Wild, I. Sinning, The crystal
J.L. Ferrer, The crystal structure of mitochondrial (Type 1A) peptide defor- structure of Ebp1 reveals a methionine aminopeptidase fold as binding
mylase provides clear guidelines for the design of inhibitors specific for the platform for multiple interactions, FEBS Lett. 581 (2007) 4450e4454.
bacterial forms, J. Biol. Chem. 280 (2005) 42315e42324. [74] M. Squatrito, M. Mancino, L. Sala, G.F. Draetta, Ebp1 is a dsRNA-binding
[48] S. Fieulaine, M. Desmadril, T. Meinnel, C. Giglione, Understanding the highly protein associated with ribosomes that modulates eIF2alpha phosphoryla-
efficient catalysis of prokaryotic peptide deformylases by shedding light on tion, Biochem. Biophys. Res. Commun. 344 (2006) 859e868.
the determinants specifying the low activity of the human counterpart, Acta [75] B.J. Greber, D. Boehringer, C. Montellese, N. Ban, Cryo-EM structures of Arx1
Crystallogr. D Biol. Crystallogr. 70 (2014) 242e252. and maturation factors Rei1 and Jjj1 bound to the 60S ribosomal subunit,
[49] C. Giglione, S. Fieulaine, T. Meinnel, Cotranslational processing mechanisms: Nat. Struct. Mol. Biol. 19 (2012) 1228e1233.
towards a dynamic 3D model, Trends Biochem. Sci. 34 (2009) 417e426. [76] B. Bradatsch, C. Leidig, S. Granneman, M. Gnadig, D. Tollervey, B. Bottcher,
[50] T. Meinnel, Peptide deformylase of eukaryotic protists: a target for new R. Beckmann, E. Hurt, Structure of the pre-60S ribosomal subunit with nu-
antiparasitic agents? Parasitol. Today 16 (2000) 165e168. clear export factor Arx1 bound at the exit tunnel, Nat. Struct. Mol. Biol. 19
[51] P.S. Margolis, C.J. Hackbarth, D.C. Young, W. Wang, D. Chen, Z. Yuan, (2012) 1234e1241.
R. White, J. Trias, Peptide deformylase in Staphylococcus aureus: resistance to [77] N.J. Hung, A.W. Johnson, Nuclear recycling of the pre-60S ribosomal subunit-
inhibition is mediated by mutations in the formyltransferase gene, Anti- associated factor Arx1 depends on Rei1 in Saccharomyces cerevisiae, Mol.
microb. Agents Chemother. 44 (2000) 1825e1831. Cell. Biol. 26 (2006) 3718e3727.
[52] P. Margolis, C. Hackbarth, S. Lopez, M. Maniar, W. Wang, Z. Yuan, R. White, [78] A. Sandikci, F. Gloge, M. Martinez, M.P. Mayer, R. Wade, B. Bukau,
J. Trias, Resistance of Streptococcus pneumoniae to deformylase inhibitors is G. Kramer, Dynamic enzyme docking to the ribosome coordinates N-ter-
due to mutations in defB, Antimicrob. Agents Chemother. 45 (2001) minal processing with polypeptide folding, Nat. Struct. Mol. Biol. 20 (2013)
2432e2435. 843e850.
[53] R. Bingel-Erlenmeyer, R. Kohler, G. Kramer, A. Sandikci, S. Antolic, T. Maier, [79] T. Arnesen, P. Van Damme, B. Polevoda, K. Helsens, R. Evjenth, N. Colaert,
C. Schaffitzel, B. Wiedmann, B. Bukau, N. Ban, A peptide deformylase- J.E. Varhaug, J. Vandekerckhove, J.R. Lillehaug, F. Sherman, K. Gevaert, Pro-
ribosome complex reveals mechanism of nascent chain processing, Nature teomics analyses reveal the evolutionary conservation and divergence of N-
452 (2008) 108e111. terminal acetyltransferases from yeast and humans, Proc. Natl. Acad. Sci. U. S.
[54] M. Haas, D. Beyer, R. Gahlmann, C. Freiberg, YkrB is the main peptide A. 106 (2009) 8157e8162.
deformylase in Bacillus subtilis, a eubacterium containing two functional [80] A. Martinez, J.A. Traverso, B. Valot, M. Ferro, C. Espagne, G. Ephritikhine,
peptide deformylases, Microbiology 147 (2001) 1783e1791. M. Zivy, C. Giglione, T. Meinnel, Extent of N-terminal modifications in
[55] J.K. Park, K.H. Kim, J.H. Moon, E.E. Kim, Characterization of peptide defor- cytosolic proteins from eukaryotes, Proteomics 8 (2008) 2809e2831.
mylase2 from B. cereus, J. Biochem. Mol. Biol. 40 (2007) 1050e1057. [81] P. Van Damme, T. Arnesen, K. Gevaert, Protein alpha-N-acetylation studied
[56] W.T. Lowther, B.W. Matthews, Metalloaminopeptidases: common functional by N-terminomics, FEBS J. 278 (2011) 3822e3834.
themes in disparate structural surroundings, Chem. Rev. 102 (2002) 4581e4608. [82] M. Falb, M. Aivaliotis, C. Garcia-Rizo, B. Bisle, A. Tebbe, C. Klein,
[57] S. Ross, C. Giglione, M. Pierre, C. Espagne, T. Meinnel, Functional and K. Konstantinidis, F. Siedler, F. Pfeiffer, D. Oesterhelt, Archaeal N-terminal
developmental impact of cytosolic protein N-terminal methionine excision protein maturation commonly involves N-terminal acetylation: a large-scale
in Arabidopsis, Plant Physiol. 137 (2005) 623e637. proteomics survey, J. Mol. Biol. 362 (2006) 915e924.
[58] B. Datta, MAPs and POEP of the roads from prokaryotic to eukaryotic king- [83] W.V. Bienvenut, D. Sumpton, A. Martinez, S. Lilla, C. Espagne, T. Meinnel,
doms, Biochimie 82 (2000) 95e107. C. Giglione, Comparative large scale characterization of plant versus mammal
[59] O.A. Olaleye, W.R. Bishai, J.O. Liu, Targeting the role of N-terminal methio- proteins reveals similar and idiosyncratic N-alpha-acetylation features, Mol.
nine processing enzymes in Mycobacterium tuberculosis, Tuberculosis 89 Cell. Proteomics 11 (2012). M111 015131.
(Suppl. 1) (2009) S55eS59. [84] T. Arnesen, D. Gromyko, F. Pendino, A. Ryningen, J.E. Varhaug, J.R. Lillehaug,
[60] S.L. Roderick, B.W. Matthews, Structure of the cobalt-dependent methionine Induction of apoptosis in human cells by RNAi-mediated knockdown of
aminopeptidase from Escherichia coli: a new type of proteolytic enzyme, hARD1 and NATH, components of the protein N-alpha-acetyltransferase
Biochemistry 32 (1993) 3907e3912. complex, Oncogene 25 (2006) 4350e4360.
[61] A. Addlagatta, M.L. Quillin, O. Omotoso, J.O. Liu, B.W. Matthews, Identifica- [85] K.K. Starheim, D. Gromyko, R. Evjenth, A. Ryningen, J.E. Varhaug,
tion of an SH3-binding motif in a new class of methionine aminopeptidases J.R. Lillehaug, T. Arnesen, Knockdown of human N alpha-terminal acetyl-
from Mycobacterium tuberculosis suggests a mode of interaction with the transferase complex C leads to p53-dependent apoptosis and aberrant hu-
ribosome, Biochemistry 44 (2005) 7166e7174. man Arl8b localization, Mol. Cell. Biol. 29 (2009) 3569e3581.
[62] P. Kanudia, M. Mittal, S. Kumaran, P.K. Chakraborti, Amino-terminal exten- [86] D. Gromyko, T. Arnesen, A. Ryningen, J.E. Varhaug, J.R. Lillehaug, Depletion of
sion present in the methionine aminopeptidase type 1c of Mycobacterium the human Nalpha-terminal acetyltransferase A induces p53-dependent
tuberculosis is indispensible for its activity, BMC Biochem. 12 (2011) 35. apoptosis and p53-independent growth inhibition, Int. J. Cancer 127
[63] J.A. Vetro, Y.H. Chang, Yeast methionine aminopeptidase type 1 is ribosome- (2010) 2777e2789.
associated and requires its N-terminal zinc finger domain for normal func- [87] C.F. Lee, D.S. Ou, S.B. Lee, L.H. Chang, R.K. Lin, Y.S. Li, A.K. Upadhyay, X. Cheng,
tion in vivo, J. Cell. Biochem. 85 (2002) 678e688. Y.C. Wang, H.S. Hsu, M. Hsiao, C.W. Wu, L.J. Juan, hNaa10p contributes to
C. Giglione et al. / Biochimie 114 (2015) 134e146 145

tumorigenesis by facilitating DNMT1-mediated tumor suppressor gene [112] K. Helsens, P. Van Damme, S. Degroeve, L. Martens, T. Arnesen,
silencing, J. Clin. Invest. 120 (2010) 2920e2930. J. Vandekerckhove, K. Gevaert, Bioinformatics analysis of a Saccharomyces
[88] J.H. Lim, J.W. Park, Y.S. Chun, Human arrest defective 1 acetylates and acti- cerevisiae N-terminal proteome provides evidence of alternative translation
vates beta-catenin, promoting lung cancer cell proliferation, Cancer Res. 66 initiation and post-translational N-terminal acetylation, J. Proteome Res. 10
(2006) 10677e10682. (2011) 3578e3589.
[89] M. Yu, M. Ma, C. Huang, H. Yang, J. Lai, S. Yan, L. Li, M. Xiang, D. Tan, Cor- [113] R. Yamada, R.A. Bradshaw, Rat liver polysome N alpha-acetyltransferase:
relation of expression of human arrest-defective-1 (hARD1) protein with substrate specificity, Biochemistry 30 (1991) 1017e1021.
breast cancer, Cancer Invest. 27 (2009) 978e983. [114] K. Kamitani, K. Narita, F. Sakiyama, Purification and characterization of hen
[90] A.F. Rope, K. Wang, R. Evjenth, J. Xing, J.J. Johnston, J.J. Swensen, oviduct N alpha-acetyltransferase, J. Biol. Chem. 264 (1989) 13188e13193.
W.E. Johnson, B. Moore, C.D. Huff, L.M. Bird, J.C. Carey, J.M. Opitz, [115] T. Arnesen, D. Anderson, C. Baldersheim, M. Lanotte, J.E. Varhaug,
C.A. Stevens, T. Jiang, C. Schank, H.D. Fain, R. Robison, B. Dalley, S. Chin, J.R. Lillehaug, Identification and characterization of the human ARD1-NATH
S.T. South, T.J. Pysher, L.B. Jorde, H. Hakonarson, J.R. Lillehaug, L.G. Biesecker, protein acetyltransferase complex, Biochem. J. 386 (2005) 433e443.
M. Yandell, T. Arnesen, G.J. Lyon, Using VAAST to identify an X-linked dis- [116] M. Gautschi, S. Just, A. Mun, S. Ross, P. Rucknagel, Y. Dubaquie,
order resulting in lethality in male infants due to N-terminal acetyl- A. Ehrenhofer-Murray, S. Rospert, The yeast N(alpha)-acetyltransferase NatA
transferase deficiency, Am. J. Hum. Genet. 89 (2011) 28e43. is quantitatively anchored to the ribosome and interacts with nascent
[91] C.H. Yi, H. Pan, J. Seebacher, I.H. Jang, S.G. Hyberts, G.J. Heffron, M.G. Vander polypeptides, Mol. Cell. Biol. 23 (2003) 7403e7414.
Heiden, R. Yang, F. Li, J.W. Locasale, H. Sharfi, B. Zhai, R. Rodriguez-Mias, [117] A. Pestana, H.C. Pitot, Acetylation of nascent polypeptide chains on rat liver
H. Luithardt, L.C. Cantley, G.Q. Daley, J.M. Asara, S.P. Gygi, G. Wagner, C.F. Liu, polyribosomes in vivo and in vitro, Biochemistry 14 (1975) 1404e1412.
J. Yuan, Metabolic regulation of protein N-alpha-acetylation by Bcl-xL Pro- [118] R. Jackson, T. Hunter, Role of methionine in the initiation of haemoglobin
motes cell survival, Cell 146 (2011) 607e620. synthesis, Nature 227 (1970) 672e676.
[92] D.C. Scott, J.K. Monda, E.J. Bennett, J.W. Harper, B.A. Schulman, N-terminal [119] R.D. Palmiter, J. Gagnon, K.A. Walsh, Ovalbumin: a secreted protein without a
acetylation acts as an avidity enhancer within an interconnected multi- transient hydrophobic leader sequence, Proc. Natl. Acad. Sci. U. S. A. 75
protein complex, Science 334 (2011) 674e678. (1978) 94e98.
[93] G.M. Forte, M.R. Pool, C.J. Stirling, N-terminal acetylation inhibits protein [120] S. Goetze, E. Qeli, C. Mosimann, A. Staes, B. Gerrits, B. Roschitzki, S. Mohanty,
targeting to the endoplasmic reticulum, PLoS Biol. 9 (2011) e1001073. E.M. Niederer, E. Laczko, E. Timmerman, V. Lange, E. Hafen, R. Aebersold,
[94] S. Bischof, K. Baerenfaller, T. Wildhaber, R. Troesch, P.A. Vidi, B. Roschitzki, J. Vandekerckhove, K. Basler, C.H. Ahrens, K. Gevaert, E. Brunner, Identifi-
M. Hirsch-Hoffmann, L. Hennig, F. Kessler, W. Gruissem, S. Baginsky, Plastid cation and functional characterization of N-terminally acetylated proteins in
proteome assembly without Toc159: photosynthetic protein import and Drosophila melanogaster, PLoS Biol. 7 (2009) e1000236.
accumulation of N-acetylated plastid precursor proteins, Plant Cell. 23 [121] S. Ramagopal, A.R. Subramanian, Alteration in the acetylation level of ribo-
(2011) 3911e3928. somal protein L12 during growth cycle of Escherichia coli, Proc. Natl. Acad.
[95] W.M. Holmes, B.K. Mannakee, R.N. Gutenkunst, T.R. Serio, Loss of amino- Sci. U. S. A. 71 (1974) 2136e2140.
terminal acetylation suppresses a prion phenotype by modulating global [122] J.A. Boutin, Myristoylation, Cell. Signal 9 (1997) 15e35.
protein folding, Nat. Commun. 5 (2014) 4383. [123] R.V. Rajala, R.S. Datla, T.N. Moyana, R. Kakkar, S.A. Carlsen, R.K. Sharma, N-
[96] K.K. Starheim, K. Gevaert, T. Arnesen, Protein N-terminal acetyltransferases: myristoyltransferase, Mol. Cell. Biochem. 204 (2000) 135e155.
when the start matters, Trends Biochem. Sci. 37 (2012) 152e161. [124] T.A. Farazi, G. Waksman, J.I. Gordon, The biology and enzymology of protein
[97] T.V. Kalvik, T. Arnesen, Protein N-terminal acetyltransferases in cancer, N-myristoylation, J. Biol. Chem. 276 (2001) 39501e39504.
Oncogene 32 (2013) 269e276. [125] P. Selvakumar, A. Lakshmikuttyamma, A. Shrivastav, S.B. Das, J.R. Dimmock,
[98] P. Van Damme, R. Evjenth, H. Foyn, K. Demeyer, P.J. De Bock, J.R. Lillehaug, R.K. Sharma, Potential role of N-myristoyltransferase in cancer, Prog. Lipid
J. Vandekerckhove, T. Arnesen, K. Gevaert, Proteome-derived peptide li- Res. 46 (2007) 1e36.
braries allow detailed analysis of the substrate specificities of N(alpha)- [126] S. Maurer-Stroh, F. Eisenhaber, Myristoylation of viral and bacterial proteins,
acetyltransferases and point to hNaa10p as the post-translational actin Trends Microbiol. 12 (2004) 178e185.
N(alpha)-acetyltransferase, Mol. Cell. Proteomics 10 (2011). M110 004580. [127] D.D. Martin, E. Beauchamp, L.G. Berthiaume, Post-translational myr-
[99] D.T. Mackay, C.H. Botting, G.L. Taylor, M.F. White, An acetylase with relaxed istoylation: fat matters in cellular life and death, Biochimie 93 (2011) 18e31.
specificity catalyses protein N-terminal acetylation in Sulfolobus solfataricus, [128] T. Meinnel, C. Giglione, Protein lipidation meets proteomics, Front. Biosci. 13
Mol. Microbiol. 64 (2007) 1540e1548. (2008) 6326e6340.
[100] G. Liszczak, R. Marmorstein, Implications for the evolution of eukaryotic [129] J.A. Traverso, C. Giglione, T. Meinnel, High-throughput profiling of N-myr-
amino-terminal acetyltransferase (NAT) enzymes from the structure of an istoylation substrate specificity across species including pathogens, Prote-
archaeal ortholog, Proc. Natl. Acad. Sci. U. S. A. 110 (2013) 14652e14657. omics 13 (2013) 25e36.
[101] P. Pesaresi, N.A. Gardner, S. Masiero, A. Dietzmann, L. Eichacker, R. Wickner, [130] S. Maurer-Stroh, B. Eisenhaber, F. Eisenhaber, N-terminal N-myristoylation
F. Salamini, D. Leister, Cytoplasmic N-terminal protein acetylation is required of proteins: refinement of the sequence motif and its taxon-specific differ-
for efficient photosynthesis in Arabidopsis, Plant Cell. 15 (2003) 1817e1832. ences, J. Mol. Biol. 317 (2002) 523e540.
[102] W.V. Bienvenut, C. Espagne, A. Martinez, W. Majeran, B. Valot, M. Zivy, [131] S. Maurer-Stroh, M. Gouda, M. Novatchkova, A. Schleiffer, G. Schneider,
O. Vallon, Z. Adam, T. Meinnel, C. Giglione, Dynamics of post-translational F.L. Sirota, M. Wildpaner, N. Hayashi, F. Eisenhaber, MYRbase: analysis of
modifications and protein stability in the stroma of Chlamydomonas rein- genome-wide glycine myristoylation enlarges the functional spectrum of
hardtii chloroplasts, Proteomics 11 (2011) 1734e1750. eukaryotic myristoylated proteins, Genome Biol. 5 (2004) R21.
[103] A. Ferrandez-Ayela, R. Micol-Ponce, A.B. Sanchez-Garcia, M.M. Alonso-Peral, [132] G. Bologna, C. Yvon, S. Duvaud, A.L. Veuthey, N-Terminal myristoylation
J.L. Micol, M.R. Ponce, Mutation of an Arabidopsis NatB N-alpha-terminal predictions by ensembles of neural networks, Proteomics 4 (2004)
acetylation complex component causes pleiotropic developmental defects, 1626e1632.
PLoS One 8 (2013) e80697. [133] S. Podell, M. Gribskov, Predicting N-terminal myristoylation sites in plant
[104] S. Hoshiyasu, K. Kohzuma, K. Yoshida, M. Fujiwara, Y. Fukao, A. Yokota, proteins, BMC Genomics 5 (2004) 37.
K. Akashi, Potential involvement of N-terminal acetylation in the quantita- [134] S.A. Weston, R. Camble, J. Colls, G. Rosenbrock, I. Taylor, M. Egerton,
tive regulation of the epsilon subunit of chloroplast ATP synthase under A.D. Tucker, A. Tunnicliffe, A. Mistry, F. Mancia, E. de la Fortelle, J. Irwin,
drought stress, Biosci. Biotechnol. Biochem. 77 (2013) 998e1007. G. Bricogne, R.A. Pauptit, Crystal structure of the anti-fungal target N-myr-
[105] G. Liszczak, J.M. Goldberg, H. Foyn, E.J. Petersson, T. Arnesen, R. Marmorstein, istoyl transferase, Nat. Struct. Biol. 5 (1998) 213e221.
Molecular basis for N-terminal acetylation by the heterodimeric NatA [135] R.S. Bhatnagar, K. Futterer, T.A. Farazi, S. Korolev, C.L. Murray, E. Jackson-
complex, Nat. Struct. Mol. Biol. 20 (2013) 1098e1105. Machelski, G.W. Gokel, J.I. Gordon, G. Waksman, Structure of N-myristoyl-
[106] G. Liszczak, T. Arnesen, R. Marmorstein, Structure of a Ternary Naa50p (NAT5/ transferase with bound myristoylCoA and peptide substrate analogs, Nat.
SAN) N-terminal acetyltransferase complex reveals the molecular basis for Struct. Biol. 5 (1998) 1091e1097.
substrate-specific acetylation, J. Biol. Chem. 286 (2011) 37002e37010. [136] D.A. Rudnick, C.A. McWherter, G.W. Gokel, J.I. Gordon, MyristoylCoA:protein
[107] M.W. Vetting, D.C. Bareich, M. Yu, J.S. Blanchard, Crystal structure of RimI N-myristoyltransferase, Adv. Enzymol. Relat. Areas Mol. Biol. 67 (1993)
from Salmonella typhimurium LT2, the GNAT responsible for N(alpha)- 375e430.
acetylation of ribosomal protein S18, Protein Sci. 17 (2008) 1781e1790. [137] D.A. Rudnick, C.A. McWherter, W.J. Rocque, P.J. Lennon, D.P. Getman,
[108] M.W. Vetting, L.P. de Carvalho, S.L. Roderick, J.S. Blanchard, A novel dimeric J.I. Gordon, Kinetic and structural evidence for a sequential ordered Bi Bi
structure of the RimL Nalpha-acetyltransferase from Salmonella typhimu- mechanism of catalysis by Saccharomyces cerevisiae myristoyl-CoA:protein
rium, J. Biol. Chem. 280 (2005) 22108e22114. N-myristoyltransferase, J. Biol. Chem. 266 (1991) 9732e9739.
[109] M.W. Vetting, S.d.C. LP, M. Yu, S.S. Hegde, S. Magnet, S.L. Roderick, [138] T.A. Farazi, J.K. Manchester, G. Waksman, J.I. Gordon, Pre-steady-state kinetic
J.S. Blanchard, Structure and functions of the GNAT superfamily of acetyl- studies of Saccharomyces cerevisiae myristoylCoA:protein N-myristoyl-
transferases, Arch. Biochem. Biophys. 433 (2005) 212e226. transferase mutants identify residues involved in catalysis, Biochemistry 40
[110] J.A. Traugh, S.B. Sharp, Protein modification enzymes associated with the (2001) 9177e9186.
protein-synthesizing complex from rabbit reticulocytes. Protein kinase, [139] R.O. Heuckeroth, D.A. Towler, S.P. Adams, L. Glaser, J.I. Gordon, 11-(Ethylthio)
phosphoprotein phosphatase, and acetyltransferase, J. Biol. Chem. 252 undecanoic acid. A myristic acid analogue of altered hydrophobicity which is
(1977) 3738e3744. functional for peptide N-myristoylation with wheat germ and yeast acyl-
[111] B. Polevoda, S. Brown, T.S. Cardillo, S. Rigby, F. Sherman, Yeast N(alpha)- transferase, J. Biol. Chem. 263 (1988) 2127e2133.
terminal acetyltransferases are associated with ribosomes, J. Cell. Biochem. [140] D.A. Towler, J.I. Gordon, S.P. Adams, L. Glaser, The biology and enzymology of
103 (2008) 492e508. eukaryotic protein acylation, Annu. Rev. Biochem. 57 (1988) 69e99.
146 C. Giglione et al. / Biochimie 114 (2015) 134e146

[141] B. Boisson, C. Giglione, T. Meinnel, Unexpected protein families including cell [159] D. Housman, D. Gillespie, H.F. Lodish, Removal of formyl-methionine residue
defense components feature in the N-myristoylome of a higher eukaryote, from nascent bacteriophage f2 protein, J. Mol. Biol. 65 (1972) 163e166.
J. Biol. Chem. 278 (2003) 43418e43429. [160] T. Meinnel, L. Patiny, S. Ragusa, S. Blanquet, Design and synthesis of substrate
[142] D.A. Towler, S.R. Eubanks, D.S. Towery, S.P. Adams, L. Glaser, Amino-terminal analogue inhibitors of peptide deformylase, Biochemistry 38 (1999)
processing of proteins by N-myristoylation. Substrate specificity of N-myr- 4287e4295.
istoyl transferase, J. Biol. Chem. 262 (1987) 1030e1036. [161] S. Ragusa, P. Mouchet, C. Lazennec, V. Dive, T. Meinnel, Substrate recognition
[143] M.H. Wright, W.P. Heal, D.J. Mann, E.W. Tate, Protein myristoylation in and selectivity of peptide deformylase. Similarities and differences with
health and disease, J. Chem. Biol. 3 (2010) 19e35. metzincins and thermolysin, J. Mol. Biol. 289 (1999) 1445e1457.
[144] C. Wilcox, J.S. Hu, E.N. Olson, Acylation of proteins with myristic acid occurs [162] G. Kramer, T. Rauch, W. Rist, S. Vorderwulbecke, H. Patzelt, A. Schulze-
cotranslationally, Science 238 (1987) 1275e1278. Specking, N. Ban, E. Deuerling, B. Bukau, L23 protein functions as a chap-
[145] C.J. Glover, K.D. Hartman, R.L. Felsted, Human N-myristoyltransferase amino- erone docking site on the ribosome, Nature 419 (2002) 171e174.
terminal domain involved in targeting the enzyme to the ribosomal sub- [163] L. Ferbitz, T. Maier, H. Patzelt, B. Bukau, E. Deuerling, N. Ban, Trigger factor in
cellular fraction, published erratum appears in J Biol Chem 1998 Mar complex with the ribosome forms a molecular cradle for nascent proteins,
6;273(10):5988, J. Biol. Chem. 272 (1997) 28680e28689. Nature 431 (2004) 590e596.
[146] I. Deichaite, L.P. Casson, H.P. Ling, M.D. Resh, In vitro synthesis of pp60v-src: [164] F. Schlunzen, D.N. Wilson, P. Tian, J.M. Harms, S.J. McInnes, H.A. Hansen,
myristylation in a cell-free system, Mol. Cell. Biol. 8 (1988) 4295e4301. R. Albrecht, J. Buerger, S.M. Wilbanks, P. Fucini, The binding mode of the
[147] N. Takamune, T. Kuroe, N. Tanada, S. Shoji, S. Misumi, Suppression of human trigger factor on the ribosome: implications for protein folding and SRP
immunodeficiency virus type-1 production by coexpression of catalytic- interaction, Structure 13 (2005) 1685e1694.
region-deleted N-myristoyltransferase mutants, Biological Pharm. Bull. 33 [165] E. Oh, A.H. Becker, A. Sandikci, D. Huber, R. Chaba, F. Gloge, R.J. Nichols,
(2010) 2018e2023. A. Typas, C.A. Gross, G. Kramer, J.S. Weissman, B. Bukau, Selective ribosome
[148] R.M. Peitzsch, S. McLaughlin, Binding of acylated peptides and fatty acids to profiling reveals the cotranslational chaperone action of trigger factor
phospholipid vesicles: pertinence to myristoylated proteins, Biochemistry 32 in vivo, Cell 147 (2011) 1295e1308.
(1993) 10436e10443. [166] T. Meinnel, C. Lazennec, F. Dardel, J.M. Schmitter, S. Blanquet, The C-terminal
[149] C.T. Pool, T.E. Thompson, Chain length and temperature dependence of the domain of peptide deformylase is disordered and dispensable for activity,
reversible association of model acylated proteins with lipid bilayers, FEBS Lett. 385 (1996) 91e95.
Biochemistry 37 (1998) 10246e10255. [167] T. Bornemann, W. Holtkamp, W. Wintermeyer, Interplay between trigger
[150] N. Burnaevskiy, T.G. Fox, D.A. Plymire, J.M. Ertelt, B.A. Weigele, A.S. Selyunin, factor and other protein biogenesis factors on the ribosome, Nat. Commun. 5
S.S. Way, S.M. Patrie, N.M. Alto, Proteolytic elimination of N-myristoyl (2014) 4180.
modifications by the Shigella virulence factor IpaJ, Nature 496 (2013) [168] M. Takeda, R.E. Webster, Protein chain initiation and deformylation in
106e109. B. subtilis homogenates, Proc. Natl. Acad. Sci. U. S. A. 60 (1968) 1487e1494.
[151] J.A. Traverso, C. Micalella, A. Martinez, S.C. Brown, B. Satiat-Jeunemaitre, [169] B.J. Greber, D. Boehringer, A. Leitner, P. Bieri, F. Voigts-Hoffmann,
T. Meinnel, C. Giglione, Roles of N-Terminal fatty acid acylations in mem- J.P. Erzberger, M. Leibundgut, R. Aebersold, N. Ban, Architecture of the large
brane compartment partitioning: arabidopsis h-type thioredoxins as a case subunit of the mammalian mitochondrial ribosome, Nature 505 (2014)
study, Plant Cell. 25 (2013) 1056e1077. 515e519.
[152] L. Renna, G. Stefano, W. Majeran, C. Micalella, T. Meinnel, C. Giglione, [170] A. Amunts, A. Brown, X.C. Bai, J.L. Llacer, T. Hussain, P. Emsley, F. Long,
F. Brandizzi, Golgi traffic and integrity depend on N-myristoyl transferase-1 G. Murshudov, S.H. Scheres, V. Ramakrishnan, Structure of the yeast mito-
in arabidopsis, Plant Cell. 25 (2013) 1756e1773. chondrial large ribosomal subunit, Science 343 (2014) 1485e1489.
[153] T. Utsumi, K. Nakano, T. Funakoshi, Y. Kayano, S. Nakao, N. Sakurai, H. Iwata, [171] B.J. Greber, D. Boehringer, M. Leibundgut, P. Bieri, A. Leitner, N. Schmitz,
R. Ishisaka, Vertical-scanning mutagenesis of amino acids in a model N- R. Aebersold, N. Ban, Nature. 515 (7526) (2014 Nov 13) 283e286, http://
myristoylation motif reveals the major amino-terminal sequence re- dx.doi.org/10.1038/nature13895. Epub 2014 Sep 1.
quirements for protein N-myristoylation, Eur. J. Biochem. 271 (2004) [172] A. Brown, A. Amunts, X.C. Bai, Y. Sugimoto, P.C. Edwards, G. Murshudov,
863e874. S.H. Scheres, V. Ramakrishnan, Structure of the large ribosomal subunit from
[154] T. Utsumi, M. Sato, K. Nakano, D. Takemura, H. Iwata, R. Ishisaka, Amino acid human mitochondria, Science 346 (2014) 718e722.
residue penultimate to the amino-terminal gly residue strongly affects two [173] T. Meinnel, P. Peynot, C. Giglione, Processed N-termini of mature proteins in
cotranslational protein modifications, N-myristoylation and N-acetylation, higher eukaryotes and their major contribution to dynamic proteomics,
J. Biol. Chem. 276 (2001) 10505e10513. Biochimie 87 (2005) 701e712.
[155] F. Gloge, A.H. Becker, G. Kramer, B. Bukau, Co-translational mechanisms of [174] X. Robert, P. Gouet, Deciphering key features in protein structures with the
protein maturation, Curr. Opin. Struct. Biol. 24C (2014) 24e33. new ENDscript server, Nucleic Acids Res. 42 (2014) W320eW324.
[156] M.J. Pine, Kinetics of maturation of the amino termini of the cell proteins of [175] M. Landau, I. Mayrose, Y. Rosenberg, F. Glaser, E. Martz, T. Pupko, N. Ben-Tal,
Escherichia coli, Biochim. Biophys. Acta 174 (1969) 359e372. ConSurf 2005: the projection of evolutionary conservation scores of residues
[157] D.B. Wilson, H. Dintzis, Initiation of the alpha chain of rabbit hemoglobin, on protein structures, Nucleic Acids Res. 33 (2005) W299eW302.
Cold Spring Harbor Symp. Quant. Biol. 34 (1969) 313e317. [176] H. Ashkenazy, E. Erez, E. Martz, T. Pupko, N. Ben-Tal, ConSurf 2010: calcu-
[158] A. Rich, E.F. Eikenberry, L.I. Malkin, Experiments on hemoglobin polypeptide lating evolutionary conservation in sequence and structure of proteins and
chain initiation and on the shielding actionof the ribosome, Cold Spring nucleic acids, Nucleic Acids Res. 38 (2010) W529eW533.
Harbor Symp. Quant. Biol. 31 (1966) 303e310.

You might also like