ISSN 0006-2979, Biochemistry (Moscow), 2025, Vol. 90, No. 11, pp. 1536-1552 © Pleiades Publishing, Ltd., 2025.
Russian Text © The Author(s), 2025, published in Biokhimiya, 2025, Vol. 90, No. 11, pp. 1638-1656.
1536
REVIEW
New Aspects of Protein Biosynthesis Inhibition
by Proline-Rich Antimicrobial Peptides
Olga V. Shulenina
1
, Eugene A. Tolstyko
1
, Andrey L. Konevega
1,2,3,a
*,
and Alena Paleskava
1,2,b
*
1
Petersburg Nuclear Physics Institute named by B.P. Konstantinov
of National Research Center “Kurchatov Institute”, 188300 Gatchina, Russia
2
Peter the Great St. Petersburg Polytechnic University, 195251 St. Petersburg, Russia
3
National Research Center “Kurchatov Institute”, 123098 Moscow, Russia
a
e-mail: konevega_al@pnpi.nrcki.ru 
b
e-mail: polesskova_ev@pnpi.nrcki.ru
Received July 31, 2025
Revised October 30, 2025
Accepted November 3, 2025
AbstractProline-rich antimicrobial peptides (PrAMPs) are promising compounds for overcoming antibiotic
resistance, one of the global health threats, and stand out from other types of AMPs by their high safety
profile. The main cellular target of PrAMPs, like most modern antibiotics, is the conservative cellular struc-
ture – the ribosome. PrAMPs bind in the ribosomal tunnel, forming multiple interactions with nucleotides of
23SrRNA, and are divided into two classes depending on their mechanism of action: inhibition of elongation
or termination. The N-terminal part of the peptides, which is important for the activity of class I peptides,
extends into the A-site pocket, preventing the binding of aminoacyl-tRNA. A new family of PrAMPs, rumicidins,
was discovered using genomic search methods. Its representatives have the longest N-terminal part, as well
as a unique pair of amino acids Trp23 and Phe24 at the C-terminus. The Trp-Phe dyad forms a spacer at the
constriction site of the ribosomal tunnel, stabilizing the binding and leading to increased antibacterial activity.
New structural studies of the class I peptide Bac5 have demonstrated its ability to disrupt the correct position-
ing of the CCA-end of the P-site tRNA in the peptidyltransferase center of the ribosome, which can affect the
assembly of functional initiation complexes. Class II PrAMPs, according to new data, have additional binding
sites on the ribosome and have a complex effect on the bacterial cell: they disrupt the termination of pro-
tein synthesis, block the cellular ribosome release system, prevent the correct assembly of the 50S ribosomal
subunits, and, possibly, affect the first stage of translocation. Recent studies expand our understanding of the
antimicrobial activity of PrAMPs and contribute to the creation of future therapeutic drugs based on AMPs.
DOI: 10.1134/S0006297925602394
Keywords: ribosome, proline-rich antimicrobial peptides, PrAMPs, translation, antibiotics
* To whom correspondence should be addressed.
INTRODUCTION
Antimicrobial peptides (AMPs) are genetically en-
coded components of an organism’s innate immune
system [1]. Although AMPs are synthesized by the
representatives of all kingdoms of life, the primary
sources of known AMPs are animals (78%), bacteria
(12%), and plants (8%) (Fig.  1). According to their
name, most AMPs target bacteria and fungi, though
some also exhibit antiviral, antiparasitic, and anti-
tumor properties  [2]. For potential therapeutic appli-
cations, animal-derived AMPs are the most promis-
ing[3]. Pathogens develop resistance to AMPs far less
readily than to low-molecular-weight antibiotics, and
the genetic barriers are higher, impeding the hori-
zontal transfer of resistance gene between bacterial
species [4, 5].
AMPs are structurally diverse and target both
membranes, disrupting their barrier function, and
intracellular components, inhibiting vital processes
such as DNA repair, replication, transcription, trans-
lation, protein folding, and cell division [2]. A recent
INHIBITION OF PROTEIN BIOSYNTHESIS BY PEPTIDES 1537
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
Fig. 1. Antimicrobial peptides (AMPs): sources, targets, and mechanisms of action (numerical values taken from the Anti-
microbial Peptide Database APD; https://aps.unmc.edu/ as of January 2025).
discovery has revealed the ability of polyproline
AMPs to disrupt bacterial biofilms [6]. The antibac-
terial mechanism of AMPs is often multifaceted. For
instance, amphibian AMPs can not only form pores
in bacterial membranes but also block membrane
transport by affecting the expression and assembly
of membrane proteins, inhibit efflux pump activity,
disrupt energy metabolism by targeting ATP synthase,
and inhibit function of FtsZ, a GTPase critical for cell
division that has no direct analog in eukaryotic cells
[7, 8]. Beyond direct elimination of infectious agents,
AMPs can also modulate the host immune response by
attracting and activating immune cells, which enhance
pathogen destruction and inflammation control [9].
Most AMPs function by compromising the integ-
rity of bacterial membranes, demonstrating activity
against both Gram-negative and Gram-positive bacte-
ria [10, 11]. However, these membranolytic properties
are often associated with a poor safety profile for
injectable substances, so they are primarily consid-
ered for topical use [12]. In contrast, certain classes
of AMPs can enter bacterial cells non-destructively
through porins to inhibit intracellular targets. This
mode of action makes them more promising as lead
compounds for systemic therapy [13, 14].
MAIN CHARACTERISTICS
OF PROLINE-RICH ANTIMICROBIAL PEPTIDES
Proline-rich antimicrobial peptides (PrAMPs) effi-
ciently penetrate bacterial cells without compromising
SHULENINA et al.1538
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
Fig. 2. Proline-rich peptides: synthesis and mechanisms of penetration into bacterial cells. PrAMPs are synthesized on the
eukaryotic 80S ribosomes as inactive prepropeptides (pre-pro-AMP). The signal pre-sequence is cleaved off after directing
the immature propeptide (pro-AMP) to large granules. In the extracellular space, pro-AMP is activated by peptidases and
translocated into periplasmic space via the OmpF porin. Partial export of the peptide from the periplasmic space to the
external environment can be carried out by the MacAB-TolC efflux pump. Transport of PrAMPs through the inner membrane
into the cytoplasm is mediated by the proteins SbmA, YjiL/MdtM, and YgdD. Once inside the cytoplasm, PrAMPs interact
with the target, prokaryotic 70S ribosome.
membrane integrity. They are characterized by a pro-
line content exceeding 25%, a net positive charge, an-
timicrobial activity, and a mechanism of action that
involves binding intracellular targets like the DnaK
chaperone or 70S ribosome. Structurally, most PrAMPs
contain a characteristic PRP motif, though this is not
a strict requirement [15,  16]. Beyond their primary
antibacterial role, many PrAMPs also exhibit antifun-
gal activity [17]. Furthermore, some representatives of
this class demonstrate immunomodulatory functions,
such as enhancing chemotaxis and differentiation of
dendritic cells, activating adaptive immune respons-
es, and suppressing the release of pro-inflammatory
mediators [18, 19]. The functional repertoire of these
peptides may be even broader for those that incorpo-
rate additional domains alongside the proline-rich re-
gions, such as cysteine-rich or acidic amino acid-rich
domains, potentially leading to a wider spectrum of
biological activity [15].
PrAMPs are predominantly found in arthropods,
such as beetles, bees, wasps, flies and crabs, and in
mammals, like sheep, cows, pigs, goats, dolphins.
To date, only a single PrAMP has been identified in
a plant (rapeseed) [15, 20]. The range of methods
for discovering these peptides has expended signifi-
cantly. The first PrAMP, apidaecin (Api), was identi-
fied after injecting honeybees with a sublethal dose
of Escherichia coli and isolating the induced peptide
chromatographically [21]. In contrast, modern discov-
ery relies on bioinformatic analysis of whole-genome
sequencing data, an approach that has recently led
to the identification of new insect [22] and cetacean
PrAMPs [13], and an entire family of proline-rich
cathelicidins from ruminants, named rumicidins
(Rum) [23]. Notably, humans and other primates do
not produce PrAMPs [24], making the therapeutic
prospect for supplementing this gap in the innate
immune defense particularly attractive. Additionally,
INHIBITION OF PROTEIN BIOSYNTHESIS BY PEPTIDES 1539
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
PrAMPs exhibit high stability against proteolytic deg-
radation, conferred by their increased proline con-
tent, and demonstrate no toxicity toward mammalian
cell lines [25].
In animals, PrAMPs are synthesized by eukary-
otic ribosomes as inactive precursors known as pre-
propeptides. The pre-region is a signal peptide that
directs the immature PrAMP to large granules and is
subsequently cleaved off [26, 27] (Fig.2). This biosyn-
thetic pathway ensures protection of the host’s own
ribosomes and ability to store antimicrobial compo-
nents, enabling a rapid response to bacterial attack.
The remaining propeptide consists of an inactivating
pro-sequence attached to the AMP. In mammalian
PrAMPs [28, 29] and in the fruit fly’s drosocin [30] the
pro-sequence is linked to a single antimicrobial pep-
tide. The propeptides of most insect PrAMPs contain
multiple, nearly identical AMP isoforms separated by
short, conserved oligopeptide spacers [21, 31]. Final
activation occurs through proteolytic cleavage, which
releases the mature PrAMP. The active peptide is then
transported to the extracellular space or into phago-
somes to exert its action on bacterial target [27, 32].
Many PrAMPs at their minimum inhibitory con-
centration (MIC) (Table  1), do not permeabilize the
bacterial membrane, but instead inhibit intracellu-
lar targets after passing through it [23, 33, 34]. Cur-
rent data support a multi-stage penetration model of
PrAMPs into the Gram-negative bacterial cells (Fig.2).
PrAMP electrostatically interacts with the outer mem-
brane. The peptide is then translocated into the peri-
plasmic space via the OmpF porin. Notably, a portion
of the peptide may be exported to the external envi-
ronment by the MacAB-TolC efflux pump [23]. Further
transport of the peptide through the inner membrane
into the cytoplasm is mainly carried out by the SbmA
protein [35, 36], a permease family protein mediating
the proton-dependent transfer of various antimicro-
bial agents [37, 38]. Subsequent research revealed an
alternative non-lytic uptake pathway in the E. coli
ΔsbmA cells without permeabilization and significant
decrease in MIC for the PrAMPs such as several bac-
tenecin  7 (Bac7) variants, and cetacean peptides Bal1
and Lip1 [13, 39]. To date, at least two additional in-
ner membrane PrAMP transporters have been identi-
fied: YjiL/MdtM and YgdD. The role of YjiL/MdtM has
been confirmed in the internalization of natural ru-
minant peptides Bac7(1-35) and rumicidin  1 (Rum-1),
aphalin Tur1A, and the synthetically optimized onco-
cins (Onc) Onc18 and Onc112 [23, 37, 40]. YgdD me-
diates the uptake of the crab PrAMP arasin 1(1-23),
and the insect peptides apidaecin1b (Api-1b) and pyr-
rhocoricin (Pyr) [14, 41]. Functional characterization
using individual knockout strains of each transporter
established a clear hierarchy among these transport-
er systems. SbmA emerges as the primary importer,
while YgdD and YjiL/MdtM serve as less important
secondary transporters.
To date, two intracellular targets have been iden-
tified for PrAMPs, both of which ultimately disrupt the
production of functional bacterial proteins. In 2000s,
it was demonstrated that certain PrAMPs bind to the
substrate-binding pocket of Hsp70 chaperone DnaK,
thereby preventing proper protein folding [42,  43].
However, the observed sensitivity of DnaK-deficient
lines to PrAMPs pointed to the existence of addition-
al targets [44]. Subsequent studies revealed that the
main target for PrAMPs is the bacterial ribosome [45],
like for more than half of conventional low-molec-
ular-weight antibiotics. The ribosome’s central role
in translation process and its structural conservation
across bacterial species make it an ideal target for
various antimicrobial agents [46].
Based on their mechanism of action, PrAMPs
are categorized into two classes that target different
Table 1. Minimum inhibitory concentration (MIC) val-
ues of full-length and truncated PrAMPs against E.coli
strain BW25113
Peptide MIC, µM References
Bac7(1-35) 3 [40]
Bac7(1-16) 8 [39]
Bac5(1-25) 1 [62]
Bac5(1-15) 64 [63]
Bac5(3-25) 4 [63]
Onc72 7 [37]
Onc112 2 [37]
Rum-1 2 [23]
Rum-1(1-22) 16 [23]
Rum-1(9-29) 8 [23]
Rum-1(11-29) 8 [23]
Tur1A 1.2 [40]
Tur1A(1-16) 4 [40]
Bal1 1 [13]
Lip1 1 [13]
Api88 0.4 [37]
Api137 2 [37]
Api137(1-17) 29 [37]
SHULENINA et al.1540
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
stages of the translation cycle. ClassI peptides, which
include oncocins and Bac7-like cathelicidins, bind to
the initiation ribosomal complex and act during the
elongation by sterically preventing binding of the first
aminoacyl-tRNA in the A-site. In contrast, classII pep-
tides, which include apidaecins and drosocin, bind to
the ribosome after release of the synthesized poly-
peptide chain. They prevent dissociation of release
factors RF1 and RF2 from the termination complex
[20,  47,  48]. Both classes of PrAMPs bind within the
nascent polypeptide exit tunnel (NPET) of the large
ribosomal subunit. The tunnel is primarily formed by
the negatively charged 23S rRNA, and peptide binding
is stabilized by a network of hydrogen bonds, stacking
interactions, and van der Waals forces [23, 40, 49-51].
This multi-point binding mechanism significantly re-
duces the risk of bacterial resistance developing
through single-point mutations in the 23S rRNA [52,
53]. Structural data demonstrate that the majority of
proline-rich peptides adopt a rigid, typeII polyproline
helix conformation [54]. This structure is critical for
their function: it allows efficient translocation through
the bacterial membranes without causing disruption
of the lipid structure [55] and facilitates penetration
into the NPET. Moreover, due to this conformation,
PrAMPs poorly bind to the active site of proteases,
which also contributes to their stability [56].
INTERACTION OF CLASS I PrAMPs
WITH THE RIBOSOME
Class I PrAMPs bind to the NPET in reverse orien-
tation relative to the growing polypeptide chain, i.e.,
N-terminus of the PrAMP is directed toward the pep-
tidyltransferase center (PTC), and C-terminus extends
deep into the NPET.
The N-terminal part of all class  I peptides pro-
trudes into the A-site tRNA binding pocket, howev-
er the number and composition of amino acids, as
well as contacts formed with 23S rRNA, significantly
vary between the arthropod and mammalian PrAMPs
(Fig. 3). The insect peptides (Onc, Pyr, metalnikowin)
are characterized by the short, conserved N-terminal
part consisting of 2 amino acids, valine and aspar-
tate, whose deletion disrupts binding to the bacterial
ribosome, indicating key role of the N-terminal regions
in antibacterial activity [49]. The N-terminal part of
mammalian peptides (Bac7, PR-39, Tur1A) is extend-
ed to 4-5 amino acids and consists mainly of argi-
nine residues. In  vitro studies have shown that two
N-terminal arginine residues are required for efficient
penetration into bacterial cells, and their removal
reduces antibacterial activity of the peptides [57].
Deletion of the first 4 amino acids almost complete-
ly inactivates the peptide, disrupting its interaction
with the main target – the ribosome [52, 58]. Despite
the increased length of N-terminus of the mammali-
an peptides, it does not extend deeper into the A-site
pocket than the insect peptides. High arginine content
in this region allows formation of a compact, posi-
tively charged loop, ideally suited in size and charge
for anchoring interactions with the groove formed by
the nucleotides C2452, A2453, and G2454 on one side
and nucleotides U2493 and G2494 of 23S rRNA on the
other, which is likely a characteristic feature of the
mammalian PrAMPs aimed at increasing the strength
of peptide-ribosome interaction [40,  52]. Thus, the
N-terminal parts of the peptides are fundamentally
important for antibacterial activity of all character-
ized representatives of class I PrAMPs, except for ru-
micidins, whose interaction with the ribosome will be
discussed in detail below.
Central part of the class I PrAMPs, corresponding
to the residues 3-11 of Onc and 6-14 of Bac7, is locat-
ed mainly in the A-site cleft but also occupies a small
area of the A-site pocket and the upper part of the
tunnel. This is a highly conserved so-called consensus
sequence (R/K)XX(R/Y)LPRPR containing the invariant
PRP motif. Conformation of the consensus part of all
structurally characterized peptides in the NPET of
70S ribosome complex is almost completely identical
[23, 40, 49-52].
The residues Lys3 and Arg9, common to the insect
PrAMPs (Onc112, metalnikowin, Pyr), form hydrogen
bonds with A2453 and U2584, respectively, and Arg9
additionally participates in stacking interactions with
C2610. The residues 5-7 of the insect PrAMPs corre-
spond to the residues 8-10 of the mammalian PrAMPs.
Tyr6 (Onc112, metalnikowin, Pyr) and the correspond-
ing residue Arg9 of Bac7 and Tyr9 of Tur1A form
stacking interactions with C2452 of 23S rRNA. The
mammalian PrAMPs contain more arginine residues
than the insect PrAMPs, establishing more stacking in-
teractions in the NPET: Arg15 and Arg16 of the Tur1A
peptide with U2586 and His69 of L4, respectively, and
Arg12, Arg14, and Arg16 of the Bac7 peptide with
C2610 (similar to Arg9 of insect PrAMPs), U2586, and
A2062, respectively [40, 49, 51, 52].
Replacement of any residue in the Arg9-Leu10-
Pro11 region of Bac7(1-16) drastically reduces its
potential to inhibit bacterial protein synthesis and
growth of E. coli cells, while substitutions in other
regions have little effect on its activity [39]. Addition-
ally, the use of deep mutational scanning, which al-
lowed analysis of antibacterial potential of more than
600,000 variants of Bac7(1-23), demonstrated high im-
portance of preserving the native sequence of Bac7
at positions 6 and 9-14, with the most pronounced
invariance in the 9-11 segment [59].
The C-terminal residues of class I PrAMPs are
considered non-conserved and less important for
INHIBITION OF PROTEIN BIOSYNTHESIS BY PEPTIDES 1541
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
Fig. 3. Alignment of amino acid sequences and spatial models of class I PrAMPs (Rum-2, PDB ID: 9D89; Bac7, PDB ID: 5HAU;
Onc112, PDB ID: 4Z8C; and Tur1A, PDB ID: 6FKR). The consensus region is indicated by the dashed line.
ribosome binding and inhibitory activity. In particu-
lar, the C-terminal region of Bac7 can be truncated
to form peptides of 35, 23, and even 16 amino ac-
ids with minimal loss of antibacterial activity [60].
Similarly, antimicrobial activity of Tur1A was only
slightly reduced when the peptide was truncated at
the C-terminus to 16 amino acids [40]. In both cases,
central part of the peptides, which ensures strong in-
teraction with the ribosome, remained intact. Further
truncation of Bac7 by just one amino acid to the 1-15
fragment [60], as well as deletion of 7 C-terminal res-
idues from the initially shorter insect peptide to form
the OncΔ7 peptide [51], resulted in complete loss of
peptide activity, likely due to the critical proximity of
the C-terminus to the consensus region. Additionally,
it is worth noting that in the majority of available
structures of the class I PrAMPs, C-terminal residues
are poorly resolved [40, 49-52], which could indicate
structural heterogeneity.
Against the background of uniformity of the se-
quences and structures of the above-described pep-
tides, it is interesting to focus on the recently char-
acterized representatives of the class I PrAMPs, which
have a number of unique features.
Fixation of rumicidins in the constriction area
of the nascent polypeptide exit tunnel. The search
for cathelicidin genes in the whole-genome sequenc-
ing database of Cetartiodactyla using a fragment of
the cathelicidin-like domain of the Bac7 precursor led
to the discovery of CATHL(3L2/8)-like genes encod-
ing an entire family of relatively short (28–30  ami-
no acids) ruminant PrAMPs, named rumicidins [23].
Despite the high degree of sequence conservation
among the rumicidins, the greatest variability was
found in the N-terminal region, which also dramat-
ically differed from the N-termini of the previously
described PrAMPs. The N-terminal region of rumici-
dins consists of 10 amino acids, exceeding the N-ter-
mini of the peptides from other mammals by at least
5 amino acids (Fig. 3). Although all rumicidins begin
with an arginine residue, unlike the previously char-
acterized mammalian peptides, none of the rumici-
din representatives have this residue in the second
position, which could indicate a somewhat different
mechanism of penetration into bacterial cell. Anoth-
er key difference between the rumicidins and other
class I PR-AMP representatives is the change in the
consensus sequence (R/K)XX(R/Y)LPRPR, which is nec-
essary for strong interaction with the NPET [13, 39]:
in all representatives of the new class, the key func-
tional site (Arg/Tyr)-Leu is replaced by His-Arg. Sev-
eral representatives of the rumicidins, demonstrating
the greatest differences in the N-terminal sequence,
were selected for structural and functional characteri-
zation: rumicidin 1 (Rum-1) from the Tibetan antelope
chiru Pantholops hodgsonii, rumicidin2 (Rum-2) from
the African antelope hirola Beatragus hunteri, and
rumicidin 3 (Rum-3) from the American pronghorn
Antilocapra americana [23].
The in vivo experiments demonstrated absence
of lytic activity of rumicidins against the bacteri-
al membrane and involvement of the porin- and
SHULENINA et al.1542
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
Fig. 4. Interaction of rumicidin 2 (Rum-2) with the bacterial ribosome (PDB ID: 9D89). Rum-2 (shown in red) is located in
the NPET of the 70S initiation complex containing the initiator tRNA in the P-site (shown in green), sterically preventing
binding of tRNA in the A-site (tRNA model shown in purple). Rum-2 creates a “spacer” at the constriction site of the tunnel,
which is formed by the rRNA nucleotides and loops of the ribosomal proteins uL4 (shown in orange) and uL22 (shown
in light blue): Pro22 forms a hydrogen bond with Arg61 of the ribosomal protein uL4; Trp23 forms a stacking interaction
with the nucleotide A751 of 23S rRNA (shown in blue) and hydrogen bonds with Lys90 of the ribosomal protein uL22; a
CH–π- interaction occurs between Phe24 and Thr65 of ribosomal protein uL4.
SbmA-mediated transport for penetration into bacte-
rial cells, while the MacAB-TolC pump was involved
in the export of the peptide from the periplasmic
space (Fig.  2). The antibacterial effect applies to a
number of Gram-negative bacteria belonging to the
“ESKAPE” group of pathogens and some mycobacteria.
A panel of in  vitro studies showed that addition of
rumicidins led to ribosome stalling at the start codon,
prevented formation of the first peptide bond, and,
similar to Bac7(1-22), reduced the level of GFP pro-
tein fluorescence in the coupled transcription/trans-
lation system. Moreover, the structure of one of the
class representatives, Rum-2, bound to the 70S initi-
ation complex containing fMet-tRNA
fMet
in the P-site,
showed that rumicidins bind to the NPET in an orien-
tation opposite to the nascent peptide chain (Fig. 4).
Thus, it was convincingly shown that rumicidins, de-
spite their differences from the previously described
peptides, belong to the class I PrAMPs and inhibit
translation by preventing elongation [23].
Quality of the Coulomb potential map detect-
ed in the NPET allowed unambiguous fitting of the
rumicidin  2 residues 12-27. Density for the first 11
amino acids of the N-terminal part of Rum-2 was not
observed, likely due to the high mobility of this re-
gion [23]. However, analysis of the sequence of this
region of rumicidin suggests some spatial similarity
to the previously characterized mammalian peptides
[40, 49, 52]. The arginine-rich sequence of the first 6
residues can form a structure resembling the compact
N-terminal loop of Bac7 and Tur1A, but this struc-
ture is likely somewhat extended into the space of
the A-site pocket due to the presence of two prolines
at positions 7 and 8, found in all rumicidins.
Truncation of the N-terminus of Bac7 by just 4
amino acids (RRIR) led to the significant (32-fold and
more) decrease in antimicrobial activity against E. coli
and Salmonella enterica strains [58]. For the most com-
parable in length to Bac7(5-35) truncated rumicidins,
Rum-1(9-29) and Rum-1(11-29), the MIC values against
E. coli BW25113 increased only 4-fold (8µM) compared
to the full-length Rum-1 (2  µM) (Table  1). Truncation
of the N-terminus of Rum-1 by less than 9 amino ac-
ids had virtually no effect (a 2-fold increase in MIC
values compared to the full form of the peptide) [23].
Notably, the effect on protein synthesis inhibition in
the in vitro transcription/translation system differed
significantly for the tested variants of Rum-1(9-29) and
Rum-1(11-29). Disruption of the reporter protein pro-
duction was observed at 2  µM concentration of the
full-length peptide and 4  µM of Rum-1(9-29), while de-
letion of two additional amino acid residues increased
the concentration required to achieve a similar effect
for the truncated peptide (Rum-1(11-29)) by 16-fold
(32  µM), indicating a potentially important role of
amino acid residues at positions 9 and 10 of rumi-
cidins in inhibiting bacterial ribosomes. Meanwhile,
functional significance of the first 8 amino acids of
the N-terminus of rumicidins remains unclear, as their
deletion did not lead to pronounced disruption of the
peptide transport into the cell or significant inhibition
of protein biosynthesis. It can be assumed that the
unique elongated N-terminal region of rumicidins is
important at the stages preceding penetration of the
peptide into bacterial cell. Negative effect of N-termi-
nal deletions could also be compensated by the addi-
tional features of rumicidins that are absent in other
representatives of mammalian PrAMPs.
INHIBITION OF PROTEIN BIOSYNTHESIS BY PEPTIDES 1543
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
All rumicidins have two amino acid substitu-
tions in the consensus fragment (R/K)XX(R/Y)LPRPR:
the highly conserved (Arg/Tyr)-Leu dyad is replaced
by the His-Arg pair, which is not found at these posi-
tions in other class I PrAMPs. Meanwhile, the central
part of Rum-2 (12-19 amino acids) aligns well with
the residues of the consensus sequence of the struc-
turally characterized mammalian peptides and forms
similar contacts with rRNA [40, 49, 52] (Fig. 3). Nota-
bly, correcting the consensus sequence of rumicidins
to the canonical form had no effect on the MIC val-
ues and efficiency of the protein synthesis inhibition
[23]. While several independent studies have shown
that the presence of the Arg-Leu dyad in the Bac7
sequence is crucial for antibacterial activity of the
peptide [39, 59].
A unique feature of the rumicidin 2 structure
was a clear density of the peptide in the upper
and the beginning of the middle part of the NPET,
allowing identification of the contacts of deeply lo-
cated amino acid residues in the NPET [23] (Fig. 4).
The most interesting are the contacts between the
peptide and constriction area of the NPET, which
is formed not only by rRNA nucleotides but also by
the loops of ribosomal proteins uL4 and uL22. Pro22
forms a hydrogen bond with Arg61 of the ribosomal
protein uL4, Trp23 forms a stacking interaction with
the nucleotide A751 of helix H34 of 23S rRNA and
hydrogen bonds with Lys90 of the ribosomal protein
uL22. A CH–π-interaction occurs between Phe24 and
Thr65 of the ribosomal protein uL4. Thus, Rum-2 cre-
ates a “spacer” at the constriction site of the NPET
[23]. Previously, a similar structure was shown at the
constriction site of the NPET for the class II PrAMP
Api137: Tyr7 formed an interaction with A751 of the
helix H34 of 23S rRNA, and Arg10 formed a hydrogen
bond with Arg61 of the ribosomal protein uL4 [53].
It is important to note that for the truncated vari-
ant of rumicidin Rum-1(1-22), concentration leading
to inhibition of the in  vitro transcription/translation
was 32  µM (16 times higher than for Rum-1), and
the MIC against E. coli BW25113 was 16  µM (8 times
higher than for Rum-1). Meanwhile, the C-terminal
truncated form of Bac7, Bac7(1-16), in the similar
experiment demonstrated high inhibition efficien-
cy at concentration of 5  µM, and the MIC against
E. coli BW25113 was 8  µM (2.7 times higher than for
Bac7(1-35)) [39]. The MIC values for Bac7(1-16) against
E. coli and S. enterica strains increased only 2-4 times
compared to the full-length form [58]. Thus, it can
be concluded that for maintaining activity of rumi-
cidins both in  vitro and in vivo, the C-terminal part
and the N-terminal part starting from the 9th amino
acid (in the case of Rum-1) are necessary, while ac-
tivity of Bac7 is ensured by the presence of the full
N-terminal part.
Importance of the C-terminal part of rumicidins
is determined by the presence of a pair of conserved
amino acids Trp23 and Phe24 in all representatives
of this family, as demonstrated by the N-terminal
truncated peptide variant with modifications at these
positions. Concentration of Rum-1(9-29, W23A, F24A)
required to inhibit protein synthesis in the in  vitro
system was 16 times higher than that of Rum-1 and
8 times higher than of Rum-1(9-29). It is interesting to
note that aromatic amino acid residues at the C-ter-
minus are present in many PrAMPs: in the insect
peptides (Onc, Pyr), it is tyrosine; in the mammalian
peptides, it is one (Bac7, Tur1A) or two (RP-39) phe-
nylalanines; moreover, the Trp-Phe pair was found in
the cetacean PrAMP, Lip1 [13]. However, functional
role of the Trp-Phe dyad in Lip1, as well as other
C-terminal aromatic amino acid residues, has not yet
been determined, as most studies have focused on
substitutions of amino acid residues in the N-terminal
and central parts of the peptides [39, 41]. In this re-
gard, the results of deep mutational scanning studies
are extremely interesting, as they revealed that the
C-terminal part of Bac7 has great potential for optimi-
zation. The study showed that insertion of a tyrosine
residue is one of the most significant approaches for
increasing activity of the peptide, especially at posi-
tions 18 and 19 [59]. Given the presence of a phenyl-
alanine residue at position 20 of Bac7, the appearance
of tyrosine in the preceding positions may lead to for-
mation of an aromatic “spacer” that forms specific
contacts with amino acid residues and ribonucleotides
on the surface of the NPET in the constriction area,
similar to rumicidins.
Thus, representatives of the new rumicidin family
have an extremely long N-terminal part, differences
in the consensus region, and presence of an aromatic
“spacer” in the C-terminal region. It can be assumed
that fixation of the peptide in the constriction area
of the NPET increases the strength of its interaction
with the ribosome and can compensate for the partial
deletion of N-terminus, which, in the absence of key
interaction in the constriction area of the NPET, dra-
matically reduces antimicrobial effect of the peptide.
Non-canonical positioning of Bac5 in the ribo-
somal tunnel. Another unusual representative of
PrAMPs is Bac5. This peptide (43 amino acids long)
was isolated from the bovine neutrophils and con-
sists of an arginine-rich N-terminal region followed
by Arg-Pro-Pro-Ile/Phe repeats [26]. For manifestation
of antibacterial effect of the peptide, it was crucial to
preserve not only the length of the N-terminal region,
as truncation of the N-terminus of Bac5 by 4 amino
acids led to the loss of its inhibitory properties, but
also presence of arginine at the N-terminus or in its
immediate vicinity [61]. The variants of Bac5 with de-
letion of the C-terminal region retained antimicrobial
SHULENINA et al.1544
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
Fig. 5. Interaction of the class II PrAMPs with the termination ribosomal complex(a) (PDB ID: 8AKN, 5O2R), ribosome re-
lease complex(b) (PDB ID: 6YSU), and elongation ribosomal complex (c) (PDB ID: 8AM9). P-site tRNA (P-tRNA) is shown in
green; release factor RF1 is shown in blue; ribosome-rescue factor ArfB is shown in yellow; A-site tRNA (A-tRNA) is shown
in purple; PrAMPs Api137 and Dro1 are shown in orange.
activity, which is a consequence of translation inhibi-
tion [16, 62, 63]. Unique sequence of the peptide, con-
sisting of amino acid repeats, shows complete absence
of the consensus sequence and even of the previously
considered invariant PRP motif. Structural studies of
Bac5(1-17) in complex with a vacant bacterial ribo-
some determined that the peptide adopts an orienta-
tion opposite to that of the nascent polypeptide chain.
The conformation of Bac5(1-17) in the NPET differs
from that of class I (Bac7, Onc112, metalnikowin,
Pyr, Tur1A) and class II (Api, drosocin) PrAMPs [16].
Positioning of the N-terminus of the peptide in the
PTC suggests potential overlap not only with the CCA-
end of the A-site tRNA, as previously demonstrated
for all class I peptides, but also with the nucleotide
A76 of the initiator tRNA
fMet
. This suggests that Bac5
may have a different mechanism of action and, in
addition to disrupting accommodation of the A-site
tRNA, could interfere with the correct placement of
the initiator tRNA in the P-site of the PTC [16].
INTERACTION OF THE CLASS II PrAMPs
WITH THE RIBOSOME
Class II PrAMPs bind within the NPET in an
orientation corresponding to that of the synthesized
polypeptide chain: their C-terminus extends toward
the A-site, while their N-terminus is oriented toward
the tunnel exit [48, 53]. Apidaecin-like peptides associ-
ate with the ribosome immediately after hydrolysis of
the ester bond between the peptide and tRNA and the
release of the newly synthesized polypeptide chain
[48,53]. Once bound within the NPET, classII PrAMPs
obstruct the dissociation of release factors RF1 and
RF2. This stabilization of the termination complex
effectively halts the translation cycle by preventing
ribosome recycling [48, 53, 64] (Fig. 5). While the
overall stability of Api137, a synthetic derivative of
natural apidaecin 1b, is mediated by its interaction
with the NPET surface, the specific placement of its
C-terminal residues Arg17 and Leu18 into the A-site
cleft is crucial for translation inhibition [53, 65]. The
side chain of Arg17 is coordinated between the resi-
dues U2506, G2505, and C2452 of 23S rRNA and Gln235
of the GGQ motif of RF1. Simultaneously, Leu18 res-
idue of the peptide establishes interaction with A76
of the deacylated P-site tRNA [53]. The Api137 peptide
containing mutation at position 17, and the truncated
Api137(1-16) had significantly lower inhibitory activ-
ity compared to the values obtained for the original
peptide sequence, emphasizing the key role of the
C-terminal Arg17 [45]. Contacts in the constriction site
of the NPET are also important: Tyr7 forms a stacking
interaction with A751, and Arg10 forms a hydrogen
bond with Arg61 of the protein uL4 [53]. This binding
mode bears a notable resemblance to the engagement
of the aromatic Phe23-Trp24 dyad of rumicidin2 with
the NPET [23]. Additionally, screening of spontaneous
E. coli mutants resistant to Api137 showed that mu-
tations in the proteins uL22, uL4, and substitutions
of the 23S rRNA nucleotides A2059, A2503 increase
resistance to Api137 [53].
Drosocin arrests translating ribosomes at the stop
codon by forming a water-mediated interaction be-
tween its Arg18 residue and Gln235 of the GGQ motif
of RF1 [48, 66], which resembles interaction between
Arg17 of Api137 and Gln235 of RF1 [53]. Notably, in
drosocin, the functionally critical amino acid residues
are distributed along the entire sequence, with some
located in the N-terminal part. Any substitution of
its key arginine residues (Arg9, Arg15, and Arg18),
INHIBITION OF PROTEIN BIOSYNTHESIS BY PEPTIDES 1545
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
aswell as Lys2 and Pro16, results in a significant de-
crease of antibacterial activity [48]. Drosocin’s high
potency is also dependent on a unique modification:
O-glycosylation at Thr11 [30, 67, 68], which establishes
multiple interactions with rRNA [48].
The class II PrAMPs can retain not only release
factors RF1/RF2 but also the ribosome-rescue factor
ArfB on the ribosome [69] (Fig.5). The N-terminal do-
main of ArfB contains a conserved GGQ motif neces-
sary for catalyzing hydrolysis of peptidyl-tRNA in the
P-site [70], and in vitro tests confirmed that Api137
traps ArfB after a single peptide hydrolysis round,
similar to what occurs with RF1/RF2 [69]. Structural
data of an ArfB-ribosome complex with Api137 in the
NPET and deacylated tRNA in the P-site confirm that,
as in the case of RF1/RF2, Api137 mimics the nascent
polypeptide chain. C-terminal Leu18 of Api137 inter-
acts with the ribose of A76 of the P-site tRNA, and the
guanidine group of Arg17 interacts with the carbonyl
group of the side chain of Gln28 in ArfB’s GGQ motif,
thereby stabilizing both the tRNA and ArfB [69].
Thus, apidaecin-like peptides disrupt protein syn-
thesis by imposing complex inhibition at the termina-
tion stage. Their direct action stalls a small fraction of
ribosomes bound to release factors RF1/RF2. However,
because the number of ribosomes in an E. coli bacte-
rial cell far exceeds the number of available release
factors [71], this stall at stop codons has a profound
domino effect. It leads to the sequestration of the lim-
ited RF pool, causing ribosomes queuing behind the
stalled complexes on the mRNA [53,  72]. The release
factors pool depletion has two outcomes for the cell:
ribosomes that reach a stop codon either stall in the
pre-hydrolysis state, or they misread the stop codon
by incorporating a closely related aminoacyl-tRNA.
The latter results in the expression of the C-terminally
extended aberrant proteins [53]. By allowing transla-
tion to continue to the 3′-end of mRNA transcripts,
the apidaecin-like peptides activate the ribosome res-
cue system. However, the rescue system is rendered
ineffective due to a combination of the depleted ter-
mination factors pool, highly efficient blocking of
the stop codon of the arfA gene mRNA encoding the
release factor ArfA, as well as direct capture of the
RF-independent ribosome-rescue factor ArfB on the
ribosome [69, 72].
Unexpectedly, structural studies of the droso-
cin-bound ribosomal complexes revealed a small pop-
ulation of 70S ribosomes containing both a deacylated
initiator tRNA in the P-site and a dipeptidyl-tRNA in
the A-site [66] (Fig. 5). This finding suggests that the
classII PrAMPs may also play a role in inhibiting the
very first step of elongation. Although the density for
the dipeptide moiety of fMet-Leu-tRNA
Leu
was poorly
resolved, the overall configuration indicated a shift
of the dipeptide toward Arg18 due to the steric clash
of fMet and Val19 residue of drosocin. This configu-
ration could impair the first round of translocation
and prevent ribosome progression along the mRNA
[66]. This proposed effect of drosocin on early elon-
gation, is supported by in vivo data. Ribosome profil-
ing and proteomic analysis demonstrated that Api137
contributed to the ribosome stalling at the start codon
of some genes [72]. Among the most affected were
genes like zntA, srmB, cysS, tyrS, and yhbL. Notably,
the second codon in each of these mRNAs encodes
polar uncharged serine and threonine and hydropho-
bic leucine, alanine, and isoleucine, respectively [72].
These amino acids are similar in size and structure,
and Api137 may specifically stabilize complexes with
fMet-Ser/Thr/Leu/Ala/Ile-tRNA in the A-site, mirroring
the interaction observed between drosocin and the
ribosome in the presence of fMet-Leu dipeptide.
Additional binding sites of apidaecins on the
ribosome. Recent structural studies have revealed
that synthetic apidaecin 1b derivatives, Api137 and
Api88 [73], have two additional binding sites on the
50S ribosomal subunit: near the NPET exit pore (both
peptides) and inside domain III of 23S rRNA (only
Api88) [25] (Fig. 6). In the first case, the N-terminal
part of the peptides is located close to the ribosomal
proteins uL29 and uL23 and is attached to the sol-
vent-exposed nucleotides of 23S rRNA A1321, A507,
A508, A63, A92, and A93, while their C-terminus is
oriented into the NPET. In the second additional site
of apidaecin Api88 inside the domain III of 23S rRNA,
the N-terminus of the peptide is oriented toward the
NPET, and C-terminus points toward the 50S sub-
unit’s core. The central Pro9-Arg10-Pro11-Arg12 mo-
tif of Api88 is nestled between the helices H51 and
H49 of 23S rRNA, ensuring stable positioning of the
peptide [25]. The role of Api88’s binding inside do-
main III of 23S rRNA remains unclear, while location
of Api137 and Api88 near the exit pore may be crucial
for efficient penetration of apidaecin into the NPET.
To perform its function, apidaecin must bind to the
upper part of the NPET immediately after release of
the synthesized peptide but before dissociation of
the release factors RF1/RF2 or the ribosome-rescue
factor ArfB from the ribosome. Prepositioning near
the tunnel entrance likely increases the probability
of this molecular event. However, binding of apidae-
cins in the vestibule of the NPET may have a broader,
more strategic function. The NPET exit is formed by
ribosomal proteins uL29 and uL23, which serve as
a primary binding platform for the trigger factor –
a ribosome-associated molecular chaperone involved
in the early stages of folding of newly synthesized
peptides [74,  75], and the signal recognition particle,
which directs membrane proteins for insertion into
the plasma membrane [76,  77]. By occupying this re-
gion, PrAMPs, could form contacts with these essential
SHULENINA et al.1546
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
Fig. 6. Binding sites of apidaecins (using Api88 as an example) with the bacterial ribosome (PDB ID: 8RPZ, 8RQ0, 8RQ2).
Site 1 – canonical binding site of PrAMPs inside the nascent polypeptide exit tunnel. Site 2 – area near the exit pore
of the nascent polypeptide exit tunnel, N-terminal part of Api88 is located near the ribosomal proteins uL29 and uL23,
the C-terminus is oriented into the tunnel. Site 3 is located inside domain III of 23S rRNA, the central region of Api88
is located between the helices H51 and H49 of 23S rRNA, and the C-terminus points toward the 50S subunit’s core.
P-site tRNA (P-tRNA) is shown in green; Api88 is shown in blue.
factors or alter their interaction with the ribosome,
disrupting fine regulation of the functional protein
formation after biosynthesis [25].
Beyond their action on mature 70S ribosomes
and 50S ribosomal subunits, apidaecins interact with
precursors of 50S subunits, disrupting the ribosome
particle assembly process within the cell [78]. Treat-
ment of bacterial cells with Api137 leads to the ap-
pearance of an additional peak on the ribosomal pro-
file, which was attributed to the pre-50S state. This
effect was not observed with the structurally similar
Api88 or the class  I PrAMP Onc112 [79]. Structural
studies reveal that Api137 leads to the appearance
of four distinct pre-50S intermediates, differing in
structural completeness both in terms of proteins and
structural regions of 23S rRNA [78]. These intermedi-
ates are largely assembled but share a critical defect:
the functional core, including the PTC, located inside
the central protuberance of domain  V, and adjacent
rRNA elements, remain unstructured. In many precur-
sors, proteins uL22 and uL29 were absent, which led
to the change in the 50S assembly pathway. Notably,
Api137 does not bind within the NPET of these prema-
ture subunits, as apparently, the tunnel must be ful-
ly formed for peptide binding. However, Api137 was
found in the additional site [25], occupying a position
near the exit pore of the NPET in close proximity to
the sites of uL22 and uL29, creating steric hindrance
for the inclusion of these proteins [78].
STRENGTH OF PrAMP INTERACTION
WITH THE RIBOSOME
Class II PrAMPs interact with diverse ribosom-
al complexes, forming contacts only with the NPET
surface at the initial stages of elongation [66, 72] or
additionally with the ligands (P-site tRNA, RF1, ArfB)
during termination or ribosome rescue process [53,
66, 72]. The formation of complexes with different
peptide-binding modes was confirmed experimental-
ly using both ribosomal extracts containing additional
protein factors and purified ribosome preparations.
The dissociation constants (K
d
) values obtained us-
ing the fluorescence polarization method [80] using
5(6)-carboxyfluorescein-labeled Api137 and ribosomal
extract ranged from 0.34 to 0.62 µM [45, 81, 82], while
those determined using purified 50S and 70S prepa-
rations were 2.2  ±  0.1  µM and 4.7  ±  0.3  µM, respec-
tively (Table 2) [25]. The K
d
values for Api88 were
comparable in the ribosomal extract (0.7-1.31  µM) [45,
81, 82] and in the purified 50S and 70S preparations
(0.58  ±  0.05  µM and 1.82  ±  0.08  µM, respectively) [25].
The obtained values are consistent with the described
models of apidaecin-ribosome interactions [25, 53, 72].
Api137 is stabilized within the NPET by additional
factors, which explains its order-of-magnitude stron-
ger binding to ribosomes in the extract compared to
isolated 70S ribosomes. In contrast to Api137, Api88
has a weak effect on termination [25]. It appears
INHIBITION OF PROTEIN BIOSYNTHESIS BY PEPTIDES 1547
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
that additional factors do not contribute significantly
to its stabilization on the ribosome, so its K
d
values
in the extract and for purified 70S ribosomes are
similar and higher than those measured for Api137.
Notably, the additional binding sites for Api137 and
Api88 have been identified on the 50S subunit [25],
and therefore, the observed K
d
values represent a su-
perposition of affinities for each individual site. The
authors suggest the potential existence of an even
greater number of sites, which may differ between
the 50S and 70S particles. This may explain the 2-3-
fold differences in K
d
values observed between 50S
and 70S ribosome [25]. Further studies aimed at
characterizing the interactions of these peptides with
the trigger factor and signal recognition particle, as
well as investigating how PrAMPs binding alters the
affinity of these protein complexes to the ribosome
and impairs their function, will be crucial to deter-
mine whether binding to additional sites is essential
for their antibacterial activity. For such investigations,
50S subunits or 70S ribosomes lacking proteins uL22
and uL29, which appear to be necessary for PrAMPs
binding at the canonical site, could be employed [25].
In contrast to the apidaecin-like peptides, all
class I PrAMPs share a conserved binding mode, po-
sitioning their N-terminus within the A-site pocket of
the upper part of NPET. The remarkable consistency of
K
d
values for Onc112 determined using ribosomal ex-
tract (0.023-0.093  µM) [45, 81, 82] and purified 50S and
70S particles (0.050  ±  0.001  µM and 0.060  ±  0.005 µM,
respectively) [25], is likely due to the property of
Onc112 to bind to the ribosome without involvement
of additional stabilizing factors. This high affinity
of Onc112, which is comparable to that of conven-
tional NPET-binding antibiotics (for example, strep-
togramins B antibiotics quinupristin and linopristin
bind to the 70S ribosome with a K
d
of 0.04  ±  0.02  µM
and 0.040  ±  0.006  µM, respectively [83]), likely results
from extensive interactions within the tunnel. How-
ever, deletion of two N-terminal residues of oncocin
led to disruption of the peptide binding to the ribo-
some [49], indicating predominant importance of the
N-terminal region of the peptide. Structural studies
of mammalian peptides also revealed formation of a
compact, positively charged loop anchoring PrAMPs
in the A-site pocket region [40, 52]. Such non-specific
interaction could lead to the significant stabilization
of the peptide in the NPET due to formation of an
occluded state and could be more important than
the contacts within the tunnel. Along with the N-ter-
minal “anchor,” the aromatic “spacer” of rumicidins,
forming specific contacts with amino acid residues
of the proteins uL4 and uL22 and ribonucleotides on
the surface of the NPET in the constriction area [23],
could be a key element ensuring even stronger inter-
action of some class I PrAMPs members with ribo-
somal complexes. However, determining contribution
of these peptide elements requires further research.
CONCLUSION AND PERSPECTIVES
PrAMPs open up great perspectives as an alter-
native to traditional antibiotics. Their mechanism of
action based on disruption of bacterial translation –
one of the fundamental processes of life – empha-
sizes high therapeutic potential of these compounds.
The main advantages of PrAMPs are their broad
spectrum of antimicrobial action, low probability of
resistance development, and minimal toxicity to ani-
mal and human cells. However, there are several ob-
stacles to the clinical use of PrAMPs. These include
rapid peptide degradation in  vivo, low bioavailability
upon oral administration, complexity and high cost
of industrial synthesis, as well as insufficient clinical
data confirming their safety. At the current stage of
PrAMPs introduction into clinical practice, targeted
optimization of the peptide structures is necessary to
enhance their stability and boost antimicrobial activ-
ity. Along with mutational screening, expanding the
“library” of PrAMPs by discovering new members of
this class and conducting studies of the poorly in-
vestigated peptides, appear promising. Work in this
direction has led to the discovery of a unique fea-
ture of rumicidins – formation of a strong interaction
Table  2. K
d
values of PrAMPs. Preparations labeled
as 30S, 50S, 70S are purified ribosomal subunits and
ribosomes
Peptide K
d
, µM Preparation References
Onc112
0.023-0.093 70S extract [45, 81, 82]
0.060 ± 0.005 70S [25]
0.050 ± 0.001 50S [25]
1.14 ± 0.06 30S [25]
Api137
0.34-0.62 70S extract [45, 81, 82]
4.7 ± 0.3 70S [25]
2.2 ± 0.1 50S [25]
20 ± 8 30S [25]
Api88
0.7-1.31 70S extract [45, 82]
1.82 ± 0.08 70S [25]
0.58 ± 0.05 50S [25]
14 ± 3 30S [25]
SHULENINA et al.1548
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
with the ribosome due to location of the conserved
Trp23-Phe24 dyad in the constriction site of the exit
tunnel. This finding could serve as the foundation for
designing peptides that surpass natural ones in anti-
bacterial properties.
Thus, PrAMPs represent one of the most promis-
ing avenues in the development of new antibacterial
agents. Their further study and optimization offer ex-
tensive opportunities for transforming approaches to
treating infectious diseases.
Abbreviations
AMP antimicrobial peptide
MIC minimum inhibitory concentration
NPET nascent polypeptide exit tunnel
PrAMP proline-rich antimicrobial peptide
PTC peptidyltransferase center
Api apidaecin
Bac bactenecin
Onc oncocin
Pyr pyrrhocoricin
Rum rumicidin
Acknowledgments
We thank V. Paleskava for assistance in preparing il-
lustrations.
Contributions
O. V. Shulenina and A. Paleskava – writing the text;
E. A. Tolstyko – preparation of illustrations; A. L. Kone-
vega – editing the article text.
Funding
This work was supported by the Russian Science Foun-
dation (grant no.25-14-00253).
Ethics approval and consent to participate
This work does not contain any studies involving hu-
mans and animal subjects performed by any of the
authors.
Conflict of interest
The authors of this work declare that they have
noconflicts of interest.
REFERENCES
1. Dini,I., De Biasi, M.G., and Mancusi,A. (2022) Anover-
view of the potentialities of antimicrobial peptides
derived from natural sources, Antibiotics (Basel), 11,
1483, https://doi.org/10.3390/antibiotics11111483.
2. Huan,Y., Kong,Q., Mou,H., and Yi,H. (2020) Antimi-
crobial peptides: classification, design, application and
research progress in multiple fields, Front. Microbiol.,
11, 582779, https://doi.org/10.3389/fmicb.2020.582779.
3. Valdez-Miramontes, C.E., De Haro-Acosta,J., Aréchiga-
Flores, C. F., Verdiguel-Fernández, L., and Rivas-
Santiago, B. (2021) Antimicrobial peptides in domes-
tic animals and their applications in veterinary med-
icine, Peptides, 142, 170576, https://doi.org/10.1016/
J.PEPTIDES.2021.170576.
4. Yu,G., Baeder, D.Y., Regoes, R.R., and Rolff,J. (2018)
Predicting drug resistance evolution: Insights from
antimicrobial peptides and antibiotics, Proc. R. Soc. B
Biol. Sci., 285, https://doi.org/10.1098/rspb.2017.2687.
5. Kintses, B., Méhi, O., Ari, E., Számel, M., Györkei, Á.,
Jangir, P.K., Nagy, I., Pál, F., Fekete, G., Tengölics,R.,
Nyerges, Á., Likó, I., Bálint, A., Molnár, T., Bálint, B.,
Vásárhelyi, B.M., Bustamante,M., Papp,B., and Pál,C.
(2019) Phylogenetic barriers to horizontal transfer of
antimicrobial peptide resistance genes in the human
gut microbiota, Nat. Microbiol., 4, 447-458, https://
doi.org/10.1038/s41564-018-0313-5.
6. De los Santos,L., Beckman, R.L., DeBarro,C., Keener,
J.E., Torres, M.D.T., de la Fuente-Nunez,C., Brodbelt,
J. S., and Fleeman, R. M. (2024) Polyproline peptide
targets Klebsiella pneumoniae polysaccharides to col-
lapse biofilms, Cell Rep. Phys. Sci., 5, 101869, https://
doi.org/10.1016/j.xcrp.2024.101869.
7. Xuan,J., Feng,W., Wang,J., Wang,R., Zhang,B., Bo,L.,
Chen, Z. S., Yang, H., and Sun, L. (2023) Antimicro-
bial peptides for combating drug-resistant bacterial
infections, Drug Resist. Updates, 68, 100954, https://
doi.org/10.1016/j.drup.2023.100954.
8. Canè, C., Tammaro, L., Duilio, A., and Di Somma, A.
(2024) Investigation of the mechanism of action
of AMPs from amphibians to identify bacterial
protein targets for therapeutic applications, An-
tibiotics (Basel), 13, 1076, https://doi.org/10.3390/
antibiotics13111076.
9. Kumar, P., Kizhakkedathu, J. N., and Straus, S. K.
(2018) Antimicrobial peptides: diversity, mecha-
nism of action and strategies to improve the activ-
ity and biocompatibility in vivo, Biomolecules, 8, 4,
https://doi.org/10.3390/biom8010004.
10. Eckert, R. (2011) Road to clinical efficacy: challeng-
es and novel strategies for antimicrobial peptide
development, Fut. Microbiol., 6, 635-651, https://
doi.org/10.2217/FMB.11.27.
11. Yeung, A. T. Y., Gellatly, S. L., and Hancock, R. E. W.
(2011) Multifunctional cationic host defence peptides
and their clinical applications, Cell. Mol. Life Sci., 68,
2161, https://doi.org/10.1007/S00018-011-0710-X.
12. Chi,Y., Peng,Y., Zhang,S., Tang,S., Zhang,W., Dai,C.,
and Ji, S. (2024) A rapid in vivo toxicity assessment
method for antimicrobial peptides, Toxics, 12, 387,
https://doi.org/10.3390/TOXICS12060387.
13. Sola,R., Mardirossian,M., Beckert,B., De Luna, L.S.,
Prickett, D., Tossi, A., Wilson, D. N., and Scocchi, M.
(2020) Characterization of cetacean proline-rich
antimicrobial peptides displaying activity against
INHIBITION OF PROTEIN BIOSYNTHESIS BY PEPTIDES 1549
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
eskape pathogens, Int.J. Mol. Sci., 21, 7367, https://
doi.org/10.3390/ijms21197367.
14. Paulsen, V. S., Mardirossian, M., Blencke, H. M.,
Benincasa, M., Runti, G., Nepa, M., Haug, T.,
Stensvåg,K., and Scocchi,M. (2016) Inner membrane
proteins YgdD and SbmA are required for the complete
susceptibility of Escherichia coli to the proline-rich
antimicrobial peptide arasin 1(1-25), Microbiology,
162, 601-609, https://doi.org/10.1099/mic.0.000249.
15. Stączek, S., Kunat-Budzyńska, M., Cytryńska, M., and
Zdybicka-Barabas,A. (2024) Proline-rich antimicrobi-
al peptides from invertebrates, Molecules, 29, 5864,
https://doi.org/10.3390/molecules29245864.
16. Raulf,K., Koller, T.O., Beckert,B., Lepak,A., Morici,M.,
Mardirossian,M., Scocchi,M., Bange,G., and Wilson,
D. N. (2025) The structure of the Vibrio natriegens
70S ribosome in complex with the proline-rich anti-
microbial peptide Bac5(1-17), Nucleic Acids Res., 53,
gkaf324, https://doi.org/10.1093/NAR/GKAF324.
17. Ciociola, T., Giovati, L., Giovannelli, A., Conti, S.,
Castagnola, M., and Vitali, A. (2018) The activity of
a mammalian proline-rich peptide against Gram-neg-
ative bacteria, including drug-resistant strains, relies
on a nonmembranolytic mode of action, Infect. Drug
Resist., 11, 969, https://doi.org/10.2147/IDR.S165179.
18. Thakur, A., Sharma, A., Alajangi, H.K., Jaiswal, P.K.,
Lim, Y.B., Singh,G., and Barnwal, R.P. (2022) Inpur-
suit of next-generation therapeutics: Antimicrobial
peptides against superbugs, their sources, mechanism
of action, nanotechnology-based delivery, and clini-
cal applications, Int. J. Biol. Macromol., 218, 135-156,
https://doi.org/10.1016/J.IJBIOMAC.2022.07.103.
19. Mahlapuu, M., Björn, C., and Ekblom, J. (2020) Anti-
microbial peptides as therapeutic agents: opportuni-
ties and challenges, Crit. Rev. Biotechnol., 40, 978-992,
https://doi.org/10.1080/07388551.2020.1796576.
20. Graf, M., and Wilson, D. N. (2019) Intracellular an-
timicrobial peptides targeting the protein synthesis
machinery, Adv. Exp. Med. Biol., 1117, 73-89, https://
doi.org/10.1007/978-981-13-3588-4_6.
21. Casteels, P., Ampe, C., Jacobs, F., Vaeck, M., and
Tempst, P. (1989) Apidaecins: antibacterial peptides
from honeybees, EMBO J., 8, 2387-2391, https://
doi.org/10.1002/J.1460-2075.1989.TB08368.X.
22. Huang,W., Baliga,C., Aleksandrova, E.V., Atkinson,G.,
Polikanov, Y.S., Vázquez-Laslop,N., and Mankin, A.S.
(2024) Activity, structure, and diversity of typeII pro-
line-rich antimicrobial peptides from insects, EMBO
Rep., 25, 5194-5211, https://doi.org/10.1038/s44319-
024-00277-5.
23. Panteleev, P. V., Pichkur, E. B., Kruglikov, R. N.,
Paleskava, A., Shulenina, O. V., Bolosov, I. A., Bog-
danov, I. V., Safronova, V. N., Balandin, S. V., Mari-
na, V.I., Kombarova, T.I., Korobova, O.V., Shamova,
O. V., Myasnikov, A. G., Borzilov, A. I., Osterman,
I. A., Sergiev, P. V., Bogdanov, A. A., Dontsova, O. A.,
Konevega, A. L., and Ovchinnikova, T. V. (2024) Ru-
micidins are a family of mammalian host-defense
peptides plugging the 70Sribosome exit tunnel, Nat.
Commun., 15, 8925, https://doi.org/10.1038/s41467-
024-53309-y.
24. Wang, G. (2014) Human antimicrobial peptides and
proteins, Pharmaceuticals (Basel), 7, 545-594, https://
doi.org/10.3390/ph7050545.
25. Lauer, S.M., Reepmeyer,M., Berendes,O., Klepacki,D.,
Gasse, J., Gabrielli, S., Grubmüller, H., Bock, L. V.,
Krizsan, A., Nikolay, R., Spahn, C. M. T., and Hoff-
mann, R. (2024) Multimodal binding and inhibition
of bacterial ribosomes by the antimicrobial peptides
Api137 and Api88, Nat. Commun., 15, 3945, https://
doi.org/10.1038/s41467-024-48027-4.
26. Zanetti, M., Litteri, L., Gennaro, R., Horstmann, H.,
and Romeo, D. (1990) Bactenecins, defense poly-
peptides of bovine neutrophils, are generated from
precursor molecules stored in the large granules,
J. Cell Biol., 111, 1363-1371, https://doi.org/10.1083/
JCB.111.4.1363.
27. Zanetti, M., Litteri, L., Griffiths, G., Gennaro, R., and
Romeo,D. (1991) Stimulus-induced maturation of pro-
bactenecins, precursors of neutrophil antimicrobial
polypeptides, J. Immunol., 146, 4295-4300.
28. Storici, P., and Zanetti, M. (1993) A novel cDNA se-
quence encoding a pig leukocyte antimicrobial peptide
with a cathelin-like pro-sequence, Biochem. Biophys.
Res. Commun., 196, 1363-1368, https://doi.org/10.1006/
BBRC.1993.2403.
29. Zanetti, M., Del Sal, G., Storici, P., Schneider, C., and
Romeo,D. (1993) The cDNA of the neutrophil antibi-
otic Bac5 predicts a pro-sequence homologous to a
cysteine proteinase inhibitor that is common to oth-
er neutrophil antibiotics, J.Biol. Chem., 268, 522-526,
https://doi.org/10.1016/s0021-9258(18)54182-x.
30. Bulet, P., Dimarcq, J. L., Hetru, C., Lagueux, M.,
Charlet, M., Hegy, G., Van Dorsselaer, A., and Hoff-
mann, J. A. (1993) A novel inducible antibacterial
peptide of Drosophila carries an O-glycosylated sub-
stitution, J.Biol. Chem., 268, 14893-14897, https://
doi.org/10.1016/s0021-9258(18)82417-6.
31. Casteels-Josson, K., Capaci, T., Casteels, P., and
Tempst, P. (1993) Apidaecin multipeptide precur-
sor structure: a putative mechanism for amplifica-
tion of the insect antibacterial response., EMBO J.,
12, 1569-1578, https://doi.org/10.1002/J.1460-2075.
1993.TB05801.X.
32. Scocchi,M., Skerlavaj,B., Romeo,D., and Gennaro,R.
(1992) Proteolytic cleavage by neutrophil elastase
converts inactive storage proforms to antibacterial
bactenecins, Eur. J. Biochem., 209, 589-595, https://
doi.org/10.1111/J.1432-1033.1992.TB17324.X.
33. Casteels, P., and Tempst, P. (1994) Apidaecin-type
peptide antibiotics function through a nonpore-
forming mechanism involving stereospecificity,
SHULENINA et al.1550
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
Biochem. Biophys. Res. Commun., 199, 339-345,
https://doi.org/10.1006/bbrc.1994.1234.
34. Scocchi, M., Tossi, A., and Gennaro, R. (2011) Pro-
line-rich antimicrobial peptides: converging to
a non-lytic mechanism of action, Cell. Mol. Life
Sci., 68, 2317-2330, https://doi.org/10.1007/S00018-
011-0721-7.
35. Runti,G., del Lopez Ruiz, M.C., Stoilova,T., Hussain,R.,
Jennions, M., Choudhury, H. G., Benincasa, M.,
Gennaro, R., Beis, K., and Scocchi, M. (2013) Func-
tional characterization of SbmA, a bacterial inner
membrane transporter required for importing the
antimicrobial peptide Bac7(1-35), J. Bacteriol., 195,
5343-5351, https://doi.org/10.1128/JB.00818-13.
36. Ghilarov, D., Inaba-Inoue, S., Stepien, P., Qu, F.,
Michalczyk,E., Pakosz,Z., Nomura,N., Ogasawara,S.,
Walker, G. C., Rebuffat, S., Iwata, S., Heddle, J. G.,
and Beis,K. (2021) Molecular mechanism of SbmA, a
promiscuous transporter exploited by antimicrobial
peptides, Sci. Adv., 7, eabj5363, https://doi.org/10.1126/
SCIADV.ABJ5363.
37. Krizsan, A., Knappe, D., and Hoffmann, R. (2015) In-
fluence of the yjiL-mdtM gene cluster on the antibac-
terial activity of proline-rich antimicrobial peptides
overcoming Escherichia coli resistance induced by
the missing SbmA transporter system, Antimicrob.
Agents Chemother., 59, 5992, https://doi.org/10.1128/
AAC.01307-15.
38. Slotboom, D. J., Ettema, T. W., Nijland, M., and
Thangaratnarajah, C. (2020) Bacterial multi-solute
transporters, FEBS Lett., 594, 3898-3907, https://
doi.org/10.1002/1873-3468.13912.
39. Mardirossian, M., Sola, R., Beckert, B., Valencic, E.,
Collis, D. W. P., Borišek, J., Armas, F., Di Stasi, A.,
Buchmann, J., Syroegin, E. A., Polikanov, Y. S.,
Magistrato,A., Hilpert,K., Wilson, D.N., and Scocchi,M.
(2020) Peptide inhibitors of bacterial protein synthe-
sis with broad spectrum and SbmA-independent bac-
tericidal activity against clinical pathogens, J. Med.
Chem., 63, 9590-9602, https://doi.org/10.1021/acs.
jmedchem.0c00665.
40. Mardirossian, M., Pérébaskine, N., Benincasa, M.,
Gambato, S., Hofmann, S., Huter, P., Müller, C.,
Hilpert, K., Innis, C. A., Tossi, A., and Wilson, D. N.
(2018) The dolphin proline-rich antimicrobial pep-
tide Tur1A inhibits protein synthesis by targeting
the bacterial ribosome, Cell Chem. Biol., 25, 530,
https://doi.org/10.1016/J.CHEMBIOL.2018.02.004.
41. Lai, P.K., Tresnak, D.T., and Hackel, B.J. (2019) Identi-
fication and elucidation of proline-rich antimicrobial
peptides with enhanced potency and delivery, Biotech-
nol. Bioeng., 116, 2439-2450, https://doi.org/10.1002/
bit.27092.
42. Knappe,D., Zahn,M., Sauer,U., Schiffer,G., Sträter,N.,
and Hoffmann, R. (2011) Rational design of oncocin
derivatives with superior protease stabilities and
antibacterial activities based on the high-resolu-
tion structure of the oncocin-DnaK complex, Chem-
BioChem, 12, 874-876, https://doi.org/10.1002/CBIC.
201000792.
43. Zahn, M., Berthold, N., Kieslich, B., Knappe,D., Hoff-
mann,R., and Sträter, N. (2013) Structural studies on
the forward and reverse binding modes of peptides
to the chaperone DnaK, J. Mol. Biol., 425, 2463-2479,
https://doi.org/10.1016/J.JMB.2013.03.041.
44. Scocchi, M., Lüthy, C., Decarli, P., Mignogna, G.,
Christen, P., and Gennaro, R. (2009) The proline-rich
antibacterial peptide Bac7 binds to and inhibits
invitro the molecular chaperone DnaK, Int.J. Peptide
Res. Ther., 15, 147-155, https://doi.org/10.1007/s10989-
009-9182-3.
45. Krizsan, A., Volke, D., Weinert, S., Sträter, N.,
Knappe, D., and Hoffmann, R. (2014) Insect-derived
proline-rich antimicrobial peptides kill bacteria by
inhibiting bacterial protein translation at the 70Sri-
bosome, Angewandte Chemie, 53, 12236-12239, https://
doi.org/10.1002/anie.201407145.
46. Wilson, D. N. (2014) Ribosome-targeting antibiot-
ics and mechanisms of bacterial resistance, Nat.
Rev. Microbiol., 12, 35-48, https://doi.org/10.1038/
nrmicro3155.
47. Polikanov, Y. S., Aleksashin, N. A., Beckert, B., and
Wilson, D. N. (2018) The mechanisms of action of
ribosome-targeting peptide antibiotics, Front. Mol.
Biosci., 5, 48, https://doi.org/10.3389/fmolb.2018.00048.
48. Mangano, K., Klepacki, D., Ohanmu, I., Baliga, C.,
Huang, W., Brakel, A., Krizsan, A., Polikanov, Y. S.,
Hoffmann, R., Vázquez-Laslop, N., and Mankin,
A. S. (2023) Inhibition of translation termination
by the antimicrobial peptide Drosocin, Nat. Chem.
Biol., 19, 1082-1090, https://doi.org/10.1038/s41589-
023-01300-x.
49. Gagnon, M. G., Roy, R. N., Lomakin, I. B., Florin, T.,
Mankin, A. S., and Steitz, T. A. (2016) Structures of
proline-rich peptides bound to the ribosome re-
veal a common mechanism of protein synthesis in-
hibition, Nucleic Acids Res., 44, 2439-2450, https://
doi.org/10.1093/nar/gkw018.
50. Roy, R. N., Lomakin, I. B., Gagnon, M. G., and Steitz,
T. A. (2015) The mechanism of inhibition of pro-
tein synthesis by the proline-rich peptide oncocin,
Nat. Struct. Mol. Biol., 22, 466-469, https://doi.org/
10.1038/nsmb.3031.
51. Seefeldt, A.C., Nguyen,F., Antunes,S., Pérébaskine,N.,
Graf, M., Arenz, S., Inampudi, K. K., Douat, C.,
Guichard, G., Wilson, D. N., and Innis, C. A. (2015)
Theproline-rich antimicrobial peptide Onc112 inhib-
its translation by blocking and destabilizing the ini-
tiation complex, Nat. Struct. Mol. Biol., 22, 470-475,
https://doi.org/10.1038/nsmb.3034.
52. Seefeldt, A. C., Graf, M., Pérébaskine, N., Nguyen, F.,
Arenz,S., Mardirossian,M., Scocchi,M., Wilson, D.N.,
INHIBITION OF PROTEIN BIOSYNTHESIS BY PEPTIDES 1551
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
and Innis, C. A. (2016) Structure of the mammalian
antimicrobial peptide Bac7(1-16) bound within the
exit tunnel of a bacterial ribosome, Nucleic Acids Res.,
44, 2429, https://doi.org/10.1093/NAR/GKV1545.
53. Florin,T., Maracci,C., Graf,M., Karki,P., Klepacki,D.,
Berninghausen,O., Beckmann,R., Vázquez-Laslop,N.,
Wilson, D.N., Rodnina, M.V., and Mankin, A.S. (2017)
An antimicrobial peptide that inhibits translation
by trapping release factors on the ribosome, Nat.
Struct. Mol. Biol., 24, 752-757, https://doi.org/10.1038/
nsmb.3439.
54. Niidome,T., Mihara,H., Oka,M., Hayashi,T., Saiki,T.,
Yoshida,K., and Aoyagi,H. (1998) Structure and prop-
erty of model peptides of proline/arginine-rich region
in bactenecin 5, Peptide Res., 51, 337-345, https://
doi.org/10.1111/J.1399-3011.1998.TB01224.X.
55. Ohgita, T., Takechi-Haraya, Y., Okada, K., Matsui, S.,
Takeuchi, M., Saito, C., Nishitsuji, K., Uchimura, K.,
Kawano, R., Hasegawa, K., Sakai-Kato, K., Akaji, K.,
Izutsu, K.I., and Saito, H. (2020) Enhancement of di-
rect membrane penetration of arginine-rich peptides
by polyproline II helix structure, Biochim. Biophys.
Acta, 1862, 183403, https://doi.org/10.1016/j.bbamem.
2020.183403.
56. Shinnar, A. E., Butler, K. L., and Park, H. J. (2003)
Cathelicidin family of antimicrobial peptides: Pro-
teolytic processing and protease resistance, Bioorg.
Chem., 31, 425-436, https://doi.org/10.1016/S0045-
2068(03)00080-4.
57. Guida, F., Benincasa, M., Zahariev, S., Scocchi, M.,
Berti, F., Gennaro, R., and Tossi, A. (2015) Effect of
size and N-terminal residue characteristics on bacte-
rial cell penetration and antibacterial activity of the
proline-rich peptide Bac7, J. Med. Chem., 58, 1195-
1204, https://doi.org/10.1021/JM501367P.
58. Podda, E., Benincasa, M., Pacor, S., Micali, F., Mat-
tiuzzo, M., Gennaro, R., and Scocchi, M. (2006) Dual
mode of action of Bac7, a proline-rich antibacterial
peptide, Biochim. Biophys. Acta, 1760, 1732-1740,
https://doi.org/10.1016/j.bbagen.2006.09.006.
59. Koch, P., Schmitt, S., Heynisch, A., Gumpinger, A.,
Wüthrich, I., Gysin, M., Shcherbakov, D., Hobbie,
S. N., Panke, S., and Held, M. (2022) Optimization of
the antimicrobial peptide Bac7 by deep mutational
scanning, BMC Biol., 20, 114, https://doi.org/10.1186/
s12915-022-01304-4.
60. Benincasa, M., Scocchi, M., Podda, E., Skerlavaj, B.,
Dolzani, L., and Gennaro, R. (2004) Antimicrobi-
al activity of Bac7 fragments against drug-resistant
clinical isolates, Peptides, 25, 2055-2061, https://
doi.org/10.1016/J.PEPTIDES.2004.08.004.
61. Tokunaga, Y., Niidome, T., Hatakeyama, T., and
Aoyagi, H. (2001) Antibacterial activity of bactene-
cin 5 fragments and their interaction with phospho-
lipid membranes, J.Peptide Sci., 7, 297-304, https://
doi.org/10.1002/PSC.317.
62. Mardirossian, M., Barrière, Q., Timchenko, T.,
Müller, C., Pacor, S., Mergaert, P., Scocchi, M., and
Wilsona, D.N. (2018) Fragments of the nonlytic pro-
line-rich antimicrobial peptide Bac5 kill Escherichia
coli cells by inhibiting protein synthesis, Antimi-
crob. Agents Chemother., 62, https://doi.org/10.1128/
AAC.00534-18.
63. Mardirossian, M., Sola, R., Degasperi, M., and
Scocchi, M. (2019) Search for shorter portions of
the proline-rich antimicrobial peptide fragment
Bac5(1-25) that retain antimicrobial activity by block-
ing protein synthesis, ChemMedChem, 14, 343-348,
https://doi.org/10.1002/cmdc.201800734.
64. Graf,M., Huter, P., Maracci,C., Peterek, M., Rodnina,
M.V., and Wilson, D.N. (2018) Visualization of trans-
lation termination intermediates trapped by the
Apidaecin 137 peptide during RF3-mediated recy-
cling of RF1, Nat. Commun., 9, 3053, https://doi.org/
10.1038/s41467-018-05465-1.
65. Baliga, C., Brown, T. J., Florin, T., Colon, S., Shah, V.,
Skowron, K. J., Kefi, A., Szal, T., Klepacki, D., Moore,
T. W., Vázquez-Laslop, N., and Mankin, A. S. (2021)
Charting the sequence-activity landscape of peptide
inhibitors of translation termination, Proc. Natl. Acad.
Sci. USA, 118, e2026465118, https://doi.org/10.1073/
PNAS.2026465118.
66. Koller, T.O., Morici,M., Berger,M., Safdari, H.A., Lele,
D.S., Beckert,B., Kaur, K.J., and Wilson, D.N. (2023)
Structural basis for translation inhibition by the gly-
cosylated drosocin peptide, Nat. Chem. Biol., 19, 1072-
1081, https://doi.org/10.1038/s41589-023-01293-7.
67. Bulet, P., Hetru, C., Dimarcq, J. L., and Hoffmann, D.
(1999) Antimicrobial peptides in insects, structure
and function, Dev. Compar. Immunol., 23, 329-344,
https://doi.org/10.1016/S0145-305X(99)00015-4.
68. Gobbo,M., Biondi,L., Filira,F., Gennaro,R., Beninca-
sa,M., Scolaro,B., and Rocchi,R. (2002) Antimicrobial
peptides: synthesis and antibacterial activity of lin-
ear and cyclic drosocin and apidaecin 1b analogues,
J. Med. Chem., 45, 4494-4504, https://doi.org/10.1021/
jm020861d.
69. Chan, K. H., Petrychenko, V., Mueller, C., Maracci, C.,
Holtkamp,W., Wilson, D.N., Fischer,N., and Rodnina,
M.V. (2020) Mechanism of ribosome rescue by alter-
native ribosome-rescue factor B, Nat. Commun., 11,
4106, https://doi.org/10.1038/s41467-020-17853-7.
70. Chadani,Y., Ono,K., Kutsukake,K., and Abo,T. (2011)
Escherichia coli YaeJ protein mediates a novel ribo-
some-rescue pathway distinct from SsrA- and Ar-
fA-mediated pathways, Mol. Microbiol., 80, 772-785,
https://doi.org/10.1111/j.1365-2958.2011.07607.x.
71. Adamski, F. M., McCaughan, K. K., Jørgensen, F.,
Kurland, C.G., and Tate, W. P. (1994) The concentra-
tion of polypeptide chain release factors 1 and 2 at
different growth rates of Escherichia coli, J.Mol. Biol.,
238, 302-308, https://doi.org/10.1006/jmbi.1994.1293.
SHULENINA et al.1552
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
72. Mangano, K., Florin, T., Shao, X., Klepacki, D.,
Chelysheva, I., Ignatova, Z., Gao, Y., Mankin, A. S.,
and Vázquez-Laslop, N. (2020) Genome-wide effects
of the antimicrobial peptide apidaecin on transla-
tion termination in bacteria, ELife, 9, e62655, https://
doi.org/10.7554/eLife.62655.
73. Berthold, N., Czihal, P., Fritsche, S., Sauer, U.,
Schiffer, G., Knappe, D., Alber, G., and Hoffmann, R.
(2013) Novel apidaecin1b analogs with superior serum
stabilities for treatment of infections by Gram-neg-
ative pathogens, Antimicrob. Agents Chemother.,
57, 402-409, https://doi.org/10.1128/AAC.01923-12.
74. Ferbitz, L., Maier, T., Patzelt, H., Bukau, B., Deuer-
ling, E., and Ban, N. (2004) Trigger factor in com-
plex with the ribosome forms a molecular cradle
for nascent proteins, Nature, 431, 590-596, https://
doi.org/10.1038/NATURE02899.
75. Buskiewicz, I., Deuerling, E., Gu, S. Q., Jöckel, J.,
Rodnina, M. V., Bukau, B., and Wintermeyer, W.
(2004) Trigger factor binds to ribosome-signal-recog-
nition particle (SRP) complexes and is excluded by
binding of the SRP receptor, Proc. Natl. Acad. Sci.
USA, 101, 7902-7906, https://doi.org/10.1073/PNAS.
0402231101.
76. Wild, K., Halic, M., Sinning, I., and Beckmann, R.
(2004) SRP meets the ribosome, Nat. Struct. Mol. Biol.,
11, 1049-1053, https://doi.org/10.1038/NSMB853.
77. Gu, S.Q., Peske,F., Wieden, H.J., Rodnina, M.V., and
Wintermeyer, W. (2003) The signal recognition par-
ticle binds to protein L23 at the peptide exit of the
Escherichia coli ribosome, RNA, 9, 566-573, https://
doi.org/10.1261/RNA.2196403.
78. Lauer, S. M., Gasse, J., Krizsan, A., Reepmeyer, M.,
Sprink, T., Nikolay, R., Spahn, C. M. T., and Hoff-
mann, R. (2025) The proline-rich antimicrobial pep-
tide Api137 disrupts large ribosomal subunit assem-
bly and induces misfolding, Nat. Commun., 16, 567,
https://doi.org/10.1038/s41467-025-55836-8.
79. Krizsan, A., Prahl, C., Goldbach, T., Knappe, D., and
Hoffmann, R. (2015) Short proline-rich antimicrobi-
al peptides inhibit either the bacterial 70S ribosome
or the assembly of its large 50S subunit, ChemBio-
Chem, 16, 2304-2308, https://doi.org/10.1002/CBIC.
201500375.
80. Moerke, N. J. (2009) Fluorescence polarization
(FP) Assays for monitoring peptide-protein or nu-
cleic acid-protein binding, Curr. Protocols Chem.
Biol., 1, 1-15, https://doi.org/10.1002/9780470559277.
ch090102.
81. Kolano, L., Knappe, D., Volke, D., Sträter, N., and
Hoffmann, R. (2020) Ribosomal target-binding sites
of antimicrobial peptides Api137 and Onc112 are
conserved among pathogens indicating new lead
structures to develop novel broad-spectrum antibi-
otics, ChemBioChem, 21, 2628-2634, https://doi.org/
10.1002/CBIC.202000109.
82. Ludwig, T., Krizsan, A., Mohammed, G. K., and Hoff-
mann, R. (2022) Antimicrobial activity and 70S ri-
bosome binding of apidaecin-derived Api805 with
increased bacterial uptake rate, Antibiotics, 11, 430,
https://doi.org/10.3390/antibiotics11040430.
83. Noeske, J., Huang, J., Olivier, N. B., Giacobbe, R. A.,
Zambrowski, M., and Cate, J. H. D. (2014) Synergy
of streptogramin antibiotics occurs independently
of their effects on translation, Antimicrob. Agents
Chemother., 58, 5269-5279, https://doi.org/10.1128/
AAC.03389-14.
Publishers Note. Pleiades Publishing remains
neutral with regard to jurisdictional claims in published
maps and institutional affiliations. AI tools may have
been used in the translation or editing of this article.