Skip to main content
ACS Bio & Med Chem Au logoLink to ACS Bio & Med Chem Au
. 2023 Dec 13;4(1):20–36. doi: 10.1021/acsbiomedchemau.3c00059

Proteases Involved in Leader Peptide Removal during RiPP Biosynthesis

Sara M Eslami , Wilfred A van der Donk †,‡,*
PMCID: PMC10885120  PMID: 38404746

Abstract

graphic file with name bg3c00059_0010.jpg

Ribosomally synthesized and post-translationally modified peptides (RiPPs) have received much attention in recent years because of their promising bioactivities and the portability of their biosynthetic pathways. Heterologous expression studies of RiPP biosynthetic enzymes identified by genome mining often leave a leader peptide on the final product to prevent toxicity to the host and to allow the attachment of a genetically encoded affinity purification tag. Removal of the leader peptide to produce the mature natural product is then carried out in vitro with either a commercial protease or a protease that fulfills this task in the producing organism. This review covers the advances in characterizing these latter cognate proteases from bacterial RiPPs and their utility as sequence-dependent proteases. The strategies employed for leader peptide removal have been shown to be remarkably diverse. They include one-step removal by a single protease, two-step removal by two dedicated proteases, and endoproteinase activity followed by aminopeptidase activity by the same protease. Similarly, the localization of the proteolytic step varies from cytoplasmic cleavage to leader peptide removal during secretion to extracellular leader peptide removal. Finally, substrate recognition ranges from highly sequence specific with respect to the leader and/or modified core peptide to nonsequence specific mechanisms.

Keywords: protease, peptidase, RiPP, macrocyclase, metalloprotease, proteolysis, maturation, cyclic peptides, leader peptide

Introduction

Ribosomally synthesized and post-translationally modified peptides (RiPPs) comprise a subset of natural products found in all domains of life and harbor a multitude of bioactivities.15 RiPP biosynthetic gene clusters (BGCs) encode pathways that follow a common logic wherein a genetically encoded peptide is post-translationally modified by enzymes that are mostly encoded in the BGC.1 The precursor peptide is generally divided into a leader peptide (LP) region involved in enzyme recognition followed by the core peptide (CP) that undergoes post-translational modification (PTM) and is released via proteolytic cleavage (Figure 1).6

Figure 1.

Figure 1

Generic overview of RiPP biosynthesis. A precursor peptide is modified by one or more class-defining modifying enzymes that introduce the PTM(s) that are characteristic for a specific RiPP class. In some cases, accessory enzymes may introduce additional non-class-defining tailoring PTMs. The BGC often, but not always, encodes export machinery and/or one or more proteases to remove the leader peptide (LP). In some examples the protease is fused to the export proteins, whereas in other cases these two functions are in separate polypeptides. In RiPP systems, the residues at the protease cleavage site are denoted X–1 and X1,7 which are the P1 and P1′ positions in protease nomenclature.8

Most early investigations into RiPP natural products have been bioactivity-guided efforts leading to notable discoveries such as nisin, microcin B17, and darobactin.913 Advances in bioinformatic technologies, such as antiSMASH, BAGEL, RODEO, RiPPer, the tools of the Enzyme Function Initiative, and many others, have enabled rapid identification of RiPP BGCs within the ever-growing genome databases.1,1422 Consequently, renewed efforts have focused on the discovery of novel RiPPs by genome mining (e.g., refs (1) and (2335)). In addition, many studies have reported engineering of RiPP pathways to make new-to-nature structures and hybrid RiPPs.1,3639 A current bottleneck in such research is the identification of proteases that recognize the cognate RiPPs and remove the LP to yield the final, mature compound. While many studies have made use of commercially available enzymes such as the endoproteinases trypsin, GluC, LysC, AspN, Tobacco Etch Virus protease, and Factor Xa, their use requires the introduction of non-native amino acids at the end of the LP, which may not always be tolerated by the biosynthetic enzymes. Alternatively, making use of existing cleavage sites for these proteases typically leads to compounds with residues remaining from the LP. In recent years, significant progress has been made toward the identification and use of the cognate RiPP proteases or protease domains, providing access to the native compounds.

Herein, we summarize the current state of identifying and utilizing dedicated leader-peptide-removing proteases from bacterial RiPP pathways, identifying gaps in our knowledge, as well as areas for future research. We organized the review by protease class and MEROPS database classification40 when available rather than RiPP family. Table 1 connects proteases to RiPP families, demonstrating that sometimes multiple protease types are used within a certain RiPP family. As will be discussed, quite a few RiPPs are matured by two independent successive proteolytic events for reasons that are not well understood.4148 In Table 1, and throughout the review, we use the established nomenclature for RiPP precursor peptides, in which residues in the CP are numbered with positive numbers counting up from the protease cleavage site and residues in the LP are indicated with negative numbers counting down (i.e., more negative) from the cleavage site toward the N-terminus (Figure 1).7 The scope of the current review does not include proteases involved in the biosynthesis of eukaryotic RiPPs, and we refer the reader to other publications regarding research on these systems.4960 This perspective is also not focused on the molecular mechanisms of these proteases, but we do discuss substrate specificity when known.

Table 1. Proteases Involved in Leader Peptide Removal during RiPP Biosynthesis Organized by Protease Familya.

graphic file with name bg3c00059_0009.jpg

graphic file with name bg3c00059_0013.jpg

a

References (31, 41, 46, 68, 71, 76, 91, 92, 94, 97, 103, 104, 108, 109, 113, 119122, 130, 132, 133, 137, 138, 140, 141, 143, 145, 147, 148, 150, 153, 157, 158, and 160202).

Cysteine Family Proteases

Within this section, we review the involvement of cysteine family proteases in the maturation of lanthipeptides (C39), polytheonamides (C1A), and lasso peptides (C96). A description of the mechanism used by ubiquitous bifunctional peptidase-containing ATP-binding transporters (PCATs) is provided, as well as the minimal LP recognition motifs for each RiPP–cysteine protease pair when known.

Several RiPP classes harbor cysteine family proteases that cleave after a canonical double glycine-like motif (GG, GA, or GS)61,62 at the end of the LP. Of the cysteine protease family, the most common type within RiPP BGCs is the PCATs, also known as ATP-binding cassette (ABC) transporter maturation and secretion (AMS) proteins.61,62 PCATs are characterized by the presence of an N-terminal papain-type C39 protease domain followed by an ABC transporter. These enzymes remove the LP at the double Gly motif, present in some of the most common RiPP precursor peptide families that resemble proteins in nitrogen fixation (NIF11 LPs) and nitrile hydratase (proteusin LPs). In addition to these very long LPs (>70 residues) that are used in a variety of different RiPP families especially in cyanobacteria (e.g., polytheonamides, thiazole/oxazole-containing peptides, and class II lanthipeptides),6365 shorter versions ending in a double glycine motif and associated with PCATs are found in a wide variety of bacteria including Bacillota (Firmicutes), Enterobacterales, and Bacteriodales.30,61,66,67

An initial LP recognition sequence Phe(−15)-X2-Ile(−12)-X3-Leu(−7)-X2-Ile(−4) (Table 1) was identified for the PCAT ComA; ComA recognizes an amphipathic α-helical portion of the N-terminus of its substrate peptide and subsequently cleaves after a double glycine site to produce the mature ComC involved in the Streptococcus pneumoniae quorum sensing pathway.6870 Providing a molecular explanation for this minimum recognition motif, an α-helical region was observed within the LP of the class II lanthipeptide precursor peptide LahA in a cocrystal structure with the protease domain LahT147 (the N-terminal 147 amino acids of the PCAT LahT). In the cocrystal structure, hydrophobic residues in the −4, – 7, and −12 positions (typically Leu, Ile, Val, Met, or Ala) along the helix were recognized via hydrophobic pockets within the LahT enzyme.71 The LahT PCAT is encoded in a BGC that contains as many as nine putative precursor peptides with conserved leader peptides but very diverse core peptides,72 explaining why it is able to remove the LP from many noncognate substrates.71 Indeed, LahT147 (and LahT150) has become a useful tool for removing LPs for many different RiPPs from a wide variety of families.24,28,7382 Another interesting approach has been the covalent attachment of the C39 protease domain of BovT to the lanthipeptide synthetase BovM to generate the class II lanthipeptide bovicin HJ50.83 The excised LahT150 protease domain and the full length transporter LtnT involved in lacticin 3147 biosynthesis (Figure 2A) are quite tolerant in terms of both LP and CP residues as well as the form of the CP (modified or unmodified) as long as the recognition motif is present in the LP. However, other PCATs such as NukT (biosynthesis of the lanthipeptide nukacin-ISK, Figure 2B) are selective for their post-translationally modified peptide,84,85 and specific interactions between the transported peptide and the transmembrane domain (TMD) of PCATs have been reported.86 This finding may explain why excision of a protease domain from a PCAT does not always provide an active enzyme in the absence of the TMD.85,87

Figure 2.

Figure 2

Representative examples of the cleavage of LPs by Cys proteases. (A) The lanthipeptide lacticin 3147 α, (B) the lanthipeptide nukacin ISK-1, (C) cryo-EM structure of PCAT1 (gray) with PEP domains (orange) bound to the peptide CtA (blue) (PDB 6V9Z), and (D) the proteusin polytheonamide B. The leader peptides are shown in blue, and the modified residues in the core peptides are shown in red. The double Gly-like motifs as well as the hydrophobic residues in positions −4, – 7, and −12 are underlined. In panels A and B, Ala-S-Ala denotes lanthionine cross-links and Abu-S-Ala denotes methyllanthionine cross-links.

PCATs typically consist of an N-terminal C39 peptidase domain (PEP), a C-terminal nucleotide-binding domain (NBD), and a TMD. Crystallization of a full length PCAT from Clostridium thermocellum (PCAT1) illustrates the conformation of the enzyme in both ATP- and non-ATP-bound forms.87 At present, the structure of the cognate post-translationally modified substrate for PCAT1 has not been determined, but it is presumed to be a RiPP system. PCAT1 forms a dimer and exhibits an α-helical barrel structure within the cell membrane (Figure 2C). The peptidase domain is weakly associated with the transporter domain. When bound to ATP, a conformational change in PCAT1 closes the transmembrane tunnel and releases the peptidase domain from its association. Activity assays demonstrated that in this state, proteolytic activity was reduced.87 The origin of this reduced activity in the detached state has been explained through NMR studies.88 Cryo-electron microscopy (cryo-EM) studies of PCAT1 with the peptide substrate CtA that is encoded nearby provided insight into the mechanism of binding, proteolysis, and transport (Figure 2C).89 First, while the PCAT1 dimer may bind two substrates, only one peptide is conformationally primed for cleavage. The C-terminal portion of the primed substrate is inserted into the transmembrane barrel while the PEP cleaves the N-terminal LP. After ATP binding, a conformational change reorients the tunnel toward the extracellular space; this releases the core peptide to the outside and allows for dissociation of the peptidase with subsequent release of the LP. Another substrate may bind to the free PEP, and the hydrolysis of ATP reorients the PCAT toward the intracellular space where it is ready for substrate docking. PCATs are used for the biosynthesis of a wide variety of RiPPs including lanthipeptides, glycocins, spliceotides, sulfatyrotides, and graspetides (Table 1), but their use is not universal within these RiPP families.

Although the C1A family is uncommon in bacteria,90 such proteases have been found within certain RiPP BGCs (Table 1). Unlike C39-based PCATs, C1A proteases contain only a protease domain, with no additional transporter domain. The PoyH C1A protease involved in the biosynthesis of the proteusin polytheonamide (Figure 2D) has been shown to be substrate tolerant.91 PoyH processed PoyA variants bearing different PTMs as well as precursor peptides from other RiPP classes such as lanthipeptides and thiopeptides containing the LP recognition sequence LDQAAGG at the −7 to −1 positions (Table 1). Whereas PoyH showed robust activity, the C1A protease ThoK, involved in the generation of the thioamitide thioholgamide, demonstrated slow catalytic processing of its cognate (semi)modified precursor peptide.92 Such slow processing may be indicative of either a protective feature of the BGC in preventing premature proteolysis before the PTMs have been completed or the need for a zymogen activation step.93

Lasso peptides are generated by unique cysteine family proteases defined by the presence of a transglutaminase-like domain (C96 family). Proteolytic activity of McjB was confirmed via in vitro reconstitution of microcin J25 biosynthesis.94 Furthermore, an investigation into the mechanism of fusilassin (also called fuscanodin) maturation demonstrated the requirement of a RiPP recognition element (RRE) encoded separately from or fused to the protease for complete proteolysis.9597 Proteolytic removal of the LP and macrolactam formation of the new N-terminal amine with a side chain carboxylate of a Asp/Glu residue appear coupled in lasso peptide biosynthesis. Although cyclization of the linear core peptide has been reported, in vitro it is generally a less efficient substrate than the full length peptide.94,96,98 Considerable substrate tolerance has been demonstrated for the fusilassin system by using chimeric substrates composed of a cognate LP for the protease and cyclase but variant core peptides, including at the P1′ position (residue 1 of the core peptide).95,99 The peptidase activity is greatly increased by the presence of the RRE domain and covariance data and NMR spectroscopy studies were used to provide a model for the interaction of the two proteins with the LP, which also provided an explanation for an invariant Thr residue at position −2 in the LP of lasso peptides.97 A recent report also documented bifunctional protease/transporter proteins involved in lasso peptide formation.100

Serine Family Proteases

This section describes the involvement of serine family proteases in the maturation of class I, III, and IV lanthipeptides (S8, S9, and POPs) and cyanobactins (S8) (Table 1). LP recognition motifs are described for nisin-like and epilancin-like class I lanthipeptide maturation, and dual proteolytic cleavage during biosynthesis of the N-to-C cyclized cyanobactin patellamide involving serine proteases is discussed.

The biosynthesis of the class I lanthipeptide nisin has been thoroughly investigated and reviewed.101 The nisin protease NisP that removes the LP (Figure 3A) belongs to the subtilisin-like S8 family of serine proteases. NisP is translated as a prepro-protein before secretion, autoproteolysis, and extracellular peptidoglycan anchoring take place.101 The crystal structure of NisP revealed the autocatalytic cleavage of the C-terminal portion of the enzyme, although prevention of this cleavage did not inhibit NisP activity with NisA.102 Substrate recognition by NisP-like proteases of NisA-like peptides is dependent on the presence of a specific LP motif: (Ala/Gly–5)-(Ala–4)-X2-(Arg–1) (Table 1 and Figure 3A).103 The molecular details of how the residues in this motif are recognized by the enzyme are currently not available. Substrate specificity of NisP with respect to the core peptide appears to be less stringent, with proteolytic cleavage observed for NisA peptides that did not contain (methyl)lanthionine cross-links.104 However, the catalytic efficiency (kcat/Km) with unmodified NisA as the substrate was approximately 100-fold lower than with fully modified NisA. NisP was also active toward NisA substrates bearing varying numbers of (methyl)lanthionine rings. The higher efficiency with a fully modified substrate is mostly due to a higher kcat value, suggesting that the rings are important for optimal orientation of the cleavage motif in the active site. The efficiency of NisP to remove the LP from nisin mutants in which the residue at the first position of the CP was changed to every other proteinogenic amino acid has also been investigated.105 These experiments showed that NisP was able to cleave the LP from all mutants except the I1P variant, although different efficiencies were observed.

Figure 3.

Figure 3

Representative examples of LP removal by serine proteases. (A) Nisin A and (B) epilancin 15x. The LP is shown in blue, and modifications in the CP are shown in red. Recognition sequences in the LP are underlined.

The protease ElxP, involved in the biosynthesis of the class I lanthipeptide epilancin 15x, likewise belongs to the S8 family but recognizes an alternative motif, (Asp/Glu–5)-(Leu/Val–4)-X2-(Gln–1), in the LP of ElxA-like peptides (Figure 3B).106,107 Alanine substitutions of either Gln–1 or Leu–4 reduced the ElxP activity by an order of magnitude. Additionally, when the NisP recognition motif of NisA was replaced with that of ElxP, modified NisA was proteolyzed by the noncognate protease. Unlike NisP, ElxP is a cytoplasmic protease requiring tight control over the order of PTMs to prevent unproductive cleavage of the unmodified peptide. Like with NisP, the molecular details for substrate recognition are currently unresolved.

The dual action of two subtilisin family proteases PatA and PatG is necessary for the formation of the circular cyanobactin patellamides A and C (Figure 4). While sharing over 40% similarity, PatA and PatG cleave at separate recognition sequences.108,109 During the biosynthesis of a large variety of cyclic cyanobactins,110 PatA-like enzymes first cleave after the consensus sequence G(L/V)E(A/P)S, followed by proteolysis by PatG-like enzymes before the AYDG sequence and subsequent macrocyclization (Figure 4). Given the unique proteolytic and macrocylization activities of PatG and its orthologues, these proteins are categorized as transamidating proteases distinct from other members of the subtilisin family.109,111,112 The cyanobactin macrocyclases have proven quite tolerant of the core peptide sequences and have therefore been used for engineering of a variety of cyclic peptides.108,110,113118

Figure 4.

Figure 4

Biosynthetic processing of patellamides A and C. Pentagons represent thiazoles and oxazoles installed by PatD from Cys and Ser/Thr residues, respectively (blue). Cleavage by PatA (purple) and PatG (orange) are illustrated. PatG also catalyzes the N-to-C terminal cyclization, illustrated by the orange amide bonds.

S8 family proteases were also recently discovered in the processing of class III lanthipeptides. AmyP was identified in the production of gut-microbe derived peptides from Bacillus amyloliquefaciens; subsequent bioinformatic and mutational analyses illustrated a (T/S/A)EVE motif for protease activity (Table 1).119

Prolyl-oligopeptidases (POPs) belonging to the S9 serine protease family have been implicated in class III and IV lanthipeptide maturation.120 The protease FlaP, encoded in the BGC of the class III lanthipeptide flavipeptin, was selective for the fully modified FlaA peptide, only cleaving efficiently after the Pro(−12) residue in the LP when macrocycles were installed in the CP.120 At the time, it was not clear how residues −11 to −1 would be removed after FlaP cleavage. As discussed below, later studies suggested that metalloproteases that possess both endo- and aminopeptidase activity likely accomplish that task (vide infra).

Metalloproteases

In this section, we provide an overview of the wide variety of metalloproteases involved in LP cleavage of class III and IV lanthipeptides (M1, M16, and M61), lipolanthines (M1), linear-azol(in)e-containing peptides (LAPs) (M103), pearlins (M103), and epipeptides (M50). The bifunctional endo- and exo-activities of the M1 proteases are discussed as well as the intriguing “pencil sharpener” mechanism proposed for M103 TldD/E proteases.

In 2019, AplP was identified as a bifunctional Zn-dependent protease from the M1 family that is necessary for the maturation of the class III lanthipeptide NAI-112 (Figure 5A).121 AplP displays dual activities, first cleaving the N-terminal LP after conserved E-(I/L)-(L/Q) and S-A-(S/T) sequence motifs as an endopeptidase (Table 1) followed by aminopeptidase trimming of the initial product to yield the final modified CP (Figure 5A).121 Unlike FlaP, AplP is active with both modified and unmodified precursor peptides. Whereas AplP involved in NAI-112 biosynthesis is encoded within the BGC, bioinformatic analyses revealed that genes for similar enzymes are present in the genomes of known class III and IV lanthipeptide producers but not in the biosynthetic gene loci.121 Indeed, AplP was demonstrated to cleave the LP of a series of different class III precursor peptides. More recently, an AplP-like enzyme (SflL) was also shown to be involved in the biosynthesis of the class IV lanthipeptide SflA. Although not encoded within the SflA BGC, the enzyme was active against SflA, recognizing the sequence PDLLK for initial endoproteinase activity after the Lys (Table 1), followed by further aminopeptidase trimming of the remainder of the LP.122 These findings with AplP and SflL explain why class III and IV lanthipeptides have been detected as a mixture of congeners when produced by the native organisms that differ in the number of amino acids that are removed by aminopeptidase activity.123128

Figure 5.

Figure 5

Representative examples of LP removal by metalloproteases. (A) NAI-112, (B) klebsazolicin, and (C) microcin C from Y. pseudotuberculosis (McCYps). The LP is shown in blue, and the modification sites in the CP are shown in red. Lab = labionin; MeLab = methyllabionin. Black arrows indicate sites of AplP endopeptidase cleavage.

The biosynthetic pathway toward the lipolanthine microvionin also utilizes AplP-like Zn-dependent proteases termed MicP1 and MicP2 encoded within the genome of the native producer (outside of the BGC).129 MicP2 was uncharacteristically efficient compared to similar enzymes described, with cleavage observed after acidic amino acids and to a lesser extent the N-termini of Ala-Ala sequences.129 An α-helix-forming sequence (θxx)θxxθxxθ (θ = L, I, V, M, or T) in the LP was shown to be critical for PTM installation but not for MicP activity.

A recent study established a general workflow for identifying proteases that are encoded outside of RiPP BGCs.130 A correlation network was established between lanthipeptide precursor peptides and proteases from over 20,000 BGCs, with about one-third not containing a protease in the BGC. By first establishing correlations between precursor peptide sequences and proteases that are encoded within BGCs, predictions could be made for identifying proteases encoded outside of class III lanthipeptide BGCs. This workflow was first validated by the identification of the previously unknown S8 protease involved in the biosynthesis of the known class I lanthipeptide paenilan. The methodology was next extended to the discovery of Zn-dependent heteromeric proteases (M16) involved in the biosynthesis of new class III lanthipeptide families, the bacinapeptins and paenithopeptins.130 The protease complex termed PttP1/PttP2 displayed endopeptidase activity cleaving after a QAAD motif, followed by aminopeptidase activity. PttP1/PttP2 was not able to cleave the LP from an uncorrelated class III lanthipeptide; similarly, an uncorrelated AplP-like enzyme was inactive toward the precursor PttA1. These results illustrate the efficacy of correlational networking in identifying proteases for cognate precursor peptides, especially for proteases that are encoded outside of RiPP BGCs.

The ubiquitous and conserved heterodimeric metalloprotease TldD/TldE is encoded within the genomes of several RiPP producers, often outside of their BGCs, and has been implicated in their maturation. These proteins are present in ∼60% of all bacterial genomes, and genetic screens showed that production of the LAP microcin B17 in Escherichia coli relies upon TldD/E activity for LP removal (Figure 6A).131,132 Crystallization of the protease complex revealed TldD as the only metal binder using a conserved HExxxH motif, with TldE required for active complex formation. TldE has been suggested to have evolved from TldD (18% sequence identity) with loss of its catalytic residues; coevolution of TldD/E is supported by the increased stability of the heterocomplex compared to the respective homodimers.133 The TldD/E structure displayed a novel fold and resulted in a new classification in the MEROPS database (M103). The N-terminus of the microcin B17 precursor (McbA) appears to be threaded through the protease toward the active site where successive cleavage events occur, likened to the action of a pencil sharpener (Figure 6B).133 Proteolysis proceeds until the heterocycles sterically prevent further entry into the channel, resulting in cessation of proteolysis of the 13 residues N-terminal to the first heterocycle (Figure 6). In other words, the PTMs provide a ruler for the final proteolytic event. Consistent with this model, TldD/E was determined to be fairly substrate tolerant with respect to the LP, but a modified CP was required, suggesting that the unmodified peptide has a specific structure that prevents threading it into the protease channel. Substrate recognition occurs in a sequence-independent manner through two β-sheet interactions with TldD that clamp the peptide in the channel, with the side chains of the substrate pointing away from the protease, explaining the substrate tolerance. Modeling of a linear peptide bound to the protease suggests that a minimum of 15 amino acids is required at the N-terminus to access the active site for cleavage between residues 3 and 4 (release of a tripeptide), consistent with the linear N-terminal sequence of microcin B17 (Figure 6A). However, the narrow opening into the active site places stringency in processing only unfolded peptides.133 TldD/E-type proteases are also involved in LP removal of the RiPPs klebsazolicin in Klebsiella pneumoniae and a microcin C analogue in Yersinia pseudotuberculosis (McCYps, Figures 5B and 5C).134 Once again, the cognate proteases are not encoded in the BGCs of these compounds. Consistent with this observation, TldD/E proteases are not only involved in the production of RiPPs but are believed to often have multiple other cellular functions.135 As a clear demonstration of the non-sequence-dependent substrate recognition, the TldD/E involved in microcin B17 biosynthesis was also able to remove the leader peptides from klebsazolicin and McCYps despite very diverse sequences.134,136

Figure 6.

Figure 6

Structure of microcin B17 (A) and the schematic representation of the “pencil sharpener” mechanism of TldD/E (B). The LP is shown in blue and modifications in the CP are shown in red.

The currently characterized BGCs of the pearlin RiPP class encode either TldD/E-like or M50 family proteases.137,138 Pearlins are unusual among RiPPs in that the final mature compound is usually just a single, heavily modified amino acid that is made on a scaffold peptide that serves a similar role as a canonical LP. Hence, the scaffold-peptide-removing enzymes are technically carboxypeptidases. In addition, pearlin biosynthesis reuses the scaffold peptide to make multiple pearlins,137 and hence its cleavage needs to take place intracellularly. The activity of a TldD/E homologue encoded in the ammosamide BGC has yet to be confirmed,138 but if this enzyme indeed removes the scaffold peptide it cannot function similarly to the TldD/E involved in microcin B17 biosynthesis since that would successively cleave the scaffold peptide into many pieces.

The membrane-bound M50 proteases TglG and TmoG were shown to cleave 3-thia-amino acids from the C-terminus of their respective substrate peptides.139,140 TglG was intolerant toward C-terminal extension of the peptide substrate, but it was active with substrate peptides ending in both Glu and 3-thia-Glu; meanwhile, TmoG showed activity toward the TmoA peptide ending in either methionine or 3-thia-homoleucine.139,140

Investigations into the intramembrane M50 metalloprotease EpeP demonstrated the enzyme to be sufficient for production of the antimicrobial epipeptide EpeX*.141 EpeP is thought to be responsible for both the cleavage and export of the natural product. However, deletion of the epeP gene did not completely abolish EpeX* production, highlighting as yet unknown proteolytic and export systems in the native Bacillus subtilis producer.141

Bottromycin is an unusual RiPP as it does not have a canonical LP but instead a follower peptide that is attached to the CP.142 Its removal is performed by the dinuclear zinc-dependent amidohydrolase PurAH (an orthologue of BotAH).143 Macroamidine formation was shown to precede PurAH activity, and removal of the follower peptide makes macrocycle formation irreversible. Although bottromycin does not have a canonical leader peptide, the N-terminal Met residue still needs to be removed from the final product and a dedicated enzyme BotP is encoded in the BGC. BotP cleaves the N-terminal methionine of the precursor peptide BotA, congruent with other M17 leucine aminopeptidases.144 Crystallization and in vitro cleavage studies of the M17 family protease BotP investigated its substrate tolerance.145 Moreover, modeling studies suggested the importance of the sequence MGPV in the binding of BotP, and reduced in vitro activity of synthetic peptides in which Met was replaced by Leu or Ile supports the preference for an unbranched residue at the P1 position. Replacement of the native Gly in the P1′ position demonstrated some substrate tolerance as Ala, Ser, and Ile were tolerated, but efficiency was reduced when bulkier amino acids were introduced.145

Two-Step Proteolytic Processing by Different Protease Family Enzymes

The biosynthesis of several RiPPs requires an additional proteolytic step after most of the LP is removed by a PCAT. Intriguingly, in almost all characterized examples involving lanthipeptides, the second proteolytic step removes a hexapeptide. Herein, we cover several examples of two-step proteolytic cleavage within class II lanthipeptides as well as the recently elucidated multistep maturation of autoinducing peptides.

The two modified precursor peptides of the enterococcal two-component lanthipeptide cytolysin are first cleaved after a GS motif and exported by the PCAT CylB (Figure 7A).42 The extracellular serine protease CylA then acts on the two secreted products of CylB. Expressed as a preproenzyme, CylA undergoes signal peptide directed secretion followed by autocatalytic self-cleavage to yield the active form of the protease.146 Activated CylA then cleaves the remaining N-terminal hexapeptide GDVQAE from the two CylB products to yield the two components of bioactive cytolysin (cytolysin S and L, Figure 7A).41 CylA was also active against unmodified synthetic analogues of the precursor peptides consisting of a portion of the LP bearing the CylA cleavage site as well as the first two residues of the core peptide, demonstrating that specificity is not dependent on the PTMs.41 Furthermore, in vitro CylA was able to not only remove the hexapeptide but also the entire LP.147

Figure 7.

Figure 7

Representative examples of two-step removal of leader or follower peptides for (A) cytolysin, (B) lichenicidin, (C) mersacidin, and (D) AIP. The LP is shown in blue, and modifications in the CP are shown in red. Secondary cleavage sites are shown in purple and bolded italic font, whereas cleavage sites of follower peptides are shown in orange.

The class II lanthipeptide lichenicidin constitutes another example of such a two-step processing (Figure 7B). The PCAT LicT and serine protease LicP operate similarly to CylB and CylA, respectively. First, an initial cleavage is performed by LicT after the GG motif of both modified precursor peptides, followed by removal of the remaining hexapeptide NDVNPE from the CP of one of the peptides by the serine protease LicP.147,148 Like CylA, LicP is made as a preproenzyme with a secretion signal peptide that undergoes a self-cleavage event to generate its active form, which was characterized crystallographically. As with cytolysin, in vitro studies established complete removal of the LP when the modified peptide was incubated with the serine protease alone, demonstrating that initial removal of most of the LP by LicT is unnecessary for in vitro reconstitution of LicP activity. Similarly, since LicT can process the LicA1 peptide that does not have the NDVNPE sequence (Figure 7B), it is not clear what the function is of the two-step LP removal process. LicP has tolerance for a variety of amino acids in the P1′ position, including Ser, Thr, Cys, Ile, and Gly, and although the enzyme prefers cyclized cognate substrates, it also cleaves linear substrates.148 With the observed substrate tolerance of the enzyme, biotechnological applications have been demonstrated such as traceless LP removal from unrelated systems or removal of affinity purification or solubilization tags from proteins by insertion of the recognition motif.148,149 A closely related enzyme is CerP involved in the biosynthesis of the cericidins, which also involves a two-step removal of the LP (Table 1).150 The β-peptides of the class II two-component lanthipeptides haloduracin,43 staphylococcin C55,151 and plantaracin W152 also undergo a two-step cleavage of the LP involving a second step that removes a hexapeptide, but the enzymes involved have not been characterized.

Similarly, for the class II lanthipeptide mersacidin, the PCAT MrsT first cleaves the LP of modified MrsA after the canonical GA motif, secreting inactive pre-MrsA (Figure 7C).44 The presence of the GDMEAA sequence after the cleavage site was necessary for MrsT activity.45 Further mutational studies indicated that removal of the Gly in the P1′ position to yield the sequence DMEAA prevented MrsT cleavage next to the resulting Asp residue.45 Premersacidin is then fully processed into its active form via cleavage of the N-terminal GDMEAA sequence by the extracellular protease AprE (a subtilisin protease, Table 1) in Bacillus amyloliquefaciens.153 This second cleavage step is necessary for successful proteolytic processing next to the N-terminal two-amino-acid lanthionine ring in mersacidin as MrsT (unlike LicT) was incapable of such activity; interestingly, the structurally similar lacticin 3147 (Figure 2A) is completely processed by a single PCAT LtnT, thereby highlighting the difference of MrsT as compared to its homologue.45,154

An unusual example of multistep proteolysis was discovered in the maturation of anti-inflammatory lanthipeptides from Myxococcus fulvus (Mfu). Genome mining revealed the presence of two proteases encoded in a class II lanthipeptide BGC, a PCAT and a M61 metalloprotease MfuP.46 The PCAT protease domain (MfuT150) was demonstrated to remove the LP at the double Gly motif. However, unlike other characterized M61 family members that are aminopeptidases, the M61 endopeptidase MfuP hydrolyzed the modified core peptide of MfuA within a lanthionine-containing macrocycle. Hence, this protease is not involved in LP removal, but appears to be used for CP modification. The function of this proteolytic event is not currently completely understood. Crystallization of MfuP confirmed the presence of a Zn-binding site, with activity abolished after EDTA treatment of the enzyme. Furthermore, MfuP was specific for the fully modified, pentacyclic MfuA core peptide compared to the tetracyclic mutants. This two-step proteolysis represents the first instance of combined PCAT and metalloprotease processing in the generation of a class II lanthipeptide.

A different type of combination of cysteine protease and metalloprotease activities was recently identified in the Staphylococcus aureus agr quorum sensing pathway. Pathway recapitulation of agr autoinducing peptide (AIP) biosynthesis had, until recently, remained elusive. The first step in AIP maturation involves thiolactone formation and concomitant cleavage of the C-terminal 14-mer from the AgrD peptide (e.g., Figure 7D) by the cysteine protease AgrB (C75 family). As an integral membrane protease, in vitro reconstitution of AgrB was only possible when embedded in a lipid membrane.155 The second step of AIP biosynthesis requires cleavage of the LP by another integral membrane metalloprotease (M79) MroQ. The necessity of MroQ in AIP production was demonstrated via generation of an mroQ knockout; deletion of the protease eliminated or severely reduced group I and II AIP production, which was restored with mroQ complementation.156 In vitro studies further showed that substrate specificity of MroQ is dependent on the linker peptide connecting the N-terminal α-helix and the C-terminal thiolactone of its substrate; AIPs containing the helix and C-terminal macrocycle of one AIP group and the linker of another AIP group were efficiently processed.156 However, whereas MroQ was shown to be active with groups I and II, and by homology group IV, AIP substrates, maturation of group III AIPs remains unresolved.

Additional Protease Families

Hydrolase activities were confirmed in the biosynthesis of both crocagin and mycofactocin. The crocagin protease CgnD was annotated as a Ser esterase and belongs to the SGNH/GDSL hydrolase superfamily. In vitro data showed that the enzyme cleaved off the LP from a post-translationally modified intermediate with slow turnover rates.157 Since proteolysis occurs intracellularly, low activity may prevent premature cleavage of the precursor peptide. Meanwhile, the peptidase MftE, a creatininase homologue, catalyzes LP cleavage only after decarboxylation of the mycofactocin CP.158

Nonprotease Peptide Backbone Cleavage Reactions during RiPP Biosynthesis

We have discussed protease families involved in RiPP maturation, but it is of note that natural products belonging to pyritides and thiopeptides utilize proteins that perform heterocyclization reactions during which the LP is removed (Figure 8A).159,160 Recently, another mechanism of peptide bond cleavage that does not require a protease was reported for the YcaO-like ATP-dependent enzyme MusD involved in the formation of the linear cyanobactin muscoride A. A cryo-EM structure of MusD and in vitro reconstitution of its activity revealed ATP-mediated hydrolysis of the terminal GV residues; during this process, H2O acts as a nucleophile that reacts with an acyl phosphate intermediate (Figure 8B).161 Thus, MusD removes a C-terminal recognition sequence by phosphorolysis followed by hydrolysis as opposed to proteolysis by the PatG C-terminal to the CP that leads to cyclic cyanobactins (Figure 4). MusA substrates bearing either bis-azole or two thiazoles were accepted by MusD, indicating a level of chemo-tolerance.161

Figure 8.

Figure 8

Schematic for peptide backbone cleavage during (A) [4 + 2] heterocyclization and (B) the proposed mechanism of biosynthesis of cyanobactin muscoride D.

Summary and Outlook

Protease-dependent maturation of precursor peptides remains a relatively unexplored area within studies of RiPP biosynthesis. Surprisingly diverse mechanisms have been uncovered for leader peptide removal including single- or multistep proteolysis. Additionally, an array of protease specificities have been reported ranging from highly specific for the modified RiPP (e.g., FlaP) to broad substrate tolerance with a length dependence (e.g., TldD/E). The Cys proteases LahT150 and PoyH have great utility for the removal of double Gly type leader peptides but are not universally active. For example, LahT150 is not able to remove LPs that do not have the three hydrophobic amino acids at positions −4, – 7, and −12.28,75 Presumably, among the many PCATs in the genomes, there may be other protease domains that have activity as stand-alone peptidases that may expand the catalogue of broad-substrate enzymes that can be used for the removal of LPs with a double glycine motif. Similarly, the serine proteases LicP and CylA have been used for sequence-specific, traceless peptide bond cleavage by insertion of their recognition motifs into noncognate peptides and proteins, and AplP-like proteases have demonstrated broad utility for class III lanthipeptides.

There are ongoing questions regarding RiPP classes with putative proteases for which activity has not been confirmed in vivo or in vitro (Table 1). Furthermore, the generation of a number of RiPP natural products by their native producers is dependent upon as yet unidentified host proteases that are typically not encoded in the BGCs (Table 1). The recent correlation workflow established for lanthipeptides130 may allow identification of proteases for these other RiPP families. Finally, the mechanisms underpinning substrate recognition by the proteases of many RiPP classes have yet to be elucidated. Continued research in these areas will likely afford the natural product community greater ease of production and diversification of compounds with various bioactivities.

Acknowledgments

Our work on RiPP biosynthesis is supported by the National Institutes of Health (R37 GM058822). S.M.E. was supported in part by the NIH Chemistry-Biology Interface training program (T32-GM136629).

Author Contributions

CRediT: Sara Miriam Eslami conceptualization, visualization, writing-original draft, writing-review & editing; Wilfred A. van der Donk conceptualization, supervision, visualization, writing-original draft, writing-review & editing.

The authors declare no competing financial interest.

References

  1. Montalbán-López M.; Scott T. A.; Ramesh S.; Rahman I. R.; van Heel A. J.; Viel J. H.; Bandarian V.; Dittmann E.; Genilloud O.; Goto Y.; Grande Burgos M. J.; Hill C.; Kim S.; Koehnke J.; Latham J. A.; Link A. J.; Martínez B.; Nair S. K.; Nicolet Y.; Rebuffat S.; Sahl H.-G.; Sareen D.; Schmidt E. W.; Schmitt L.; Severinov K.; Süssmuth R. D.; Truman A. W.; Wang H.; Weng J.-K.; van Wezel G. P.; Zhang Q.; Zhong J.; Piel J.; Mitchell D. A.; Kuipers O. P.; van der Donk W. A. New developments in RiPP discovery, enzymology and engineering. Nat. Prod. Rep. 2021, 38 (1), 130–239. 10.1039/D0NP00027B. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Li Y.; Rebuffat S. The manifold roles of microbial ribosomal peptide-based natural products in physiology and ecology. J. Biol. Chem. 2020, 295 (1), 34–54. 10.1074/jbc.REV119.006545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Cao L.; Do T.; Link A. J. Mechanisms of action of ribosomally synthesized and posttranslationally modified peptides (RiPPs). J. Ind. Microbiol. Biotechnol. 2021, 48 (3–4), kuab005 10.1093/jimb/kuab005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Ongpipattanakul C.; Desormeaux E. K.; DiCaprio A.; van der Donk W. A.; Mitchell D. A.; Nair S. K. Mechanism of action of ribosomally synthesized and post-translationally modified peptides. Chem. Rev. 2022, 122 (18), 14722–14814. 10.1021/acs.chemrev.2c00210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Rebuffat S. Ribosomally synthesized peptides, foreground players in microbial interactions: recent developments and unanswered questions. Nat. Prod. Rep. 2022, 39 (2), 273–310. 10.1039/D1NP00052G. [DOI] [PubMed] [Google Scholar]
  6. Yang X.; van der Donk W. A. Ribosomally synthesized and post-translationally modified peptide natural products: new insights into the role of leader and core peptides during biosynthesis. Chem. Eur. J. 2013, 19, 7662–77. 10.1002/chem.201300401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Arnison P. G.; Bibb M. J.; Bierbaum G.; Bowers A. A.; Bugni T. S.; Bulaj G.; Camarero J. A.; Campopiano D. J.; Challis G. L.; Clardy J.; Cotter P. D.; Craik D. J.; Dawson M.; Dittmann E.; Donadio S.; Dorrestein P. C.; Entian K. D.; Fischbach M. A.; Garavelli J. S.; Göransson U.; Gruber C. W.; Haft D. H.; Hemscheidt T. K.; Hertweck C.; Hill C.; Horswill A. R.; Jaspars M.; Kelly W. L.; Klinman J. P.; Kuipers O. P.; Link A. J.; Liu W.; Marahiel M. A.; Mitchell D. A.; Moll G. N.; Moore B. S.; Müller R.; Nair S. K.; Nes I. F.; Norris G. E.; Olivera B. M.; Onaka H.; Patchett M. L.; Piel J.; Reaney M. J.; Rebuffat S.; Ross R. P.; Sahl H. G.; Schmidt E. W.; Selsted M. E.; Severinov K.; Shen B.; Sivonen K.; Smith L.; Stein T.; Süssmuth R. D.; Tagg J. R.; Tang G. L.; Truman A. W.; Vederas J. C.; Walsh C. T.; Walton J. D.; Wenzel S. C.; Willey J. M.; van der Donk W. A. Ribosomally synthesized and post-translationally modified peptide natural products: overview and recommendations for a universal nomenclature. Nat. Prod. Rep. 2013, 30 (1), 108–160. 10.1039/C2NP20085F. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Schechter I.; Berger A. On the size of the active site in proteases. I. Papain. Biochem. Biophys. Res. Commun. 1967, 27 (2), 157–62. 10.1016/S0006-291X(67)80055-X. [DOI] [PubMed] [Google Scholar]
  9. Rogers L. A. The inhibiting effect of Streptococcus lactis on Lactobacillus bulgaricus. J. Bacteriol. 1928, 16, 321–5. 10.1128/jb.16.5.321-325.1928. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Tagg J. R.; Dajani A. S.; Wannamaker L. W. Bacteriocins of gram-positive bacteria. Bacteriol. Rev. 1976, 40 (3), 722–756. 10.1128/br.40.3.722-756.1976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Asensio C.; Pérez-Díaz J. C.; Martínez M. C.; Baquero F. A new family of low molecular weight antibiotics from enterobacteria. Biochem. Biophys. Res. Commun. 1976, 69 (1), 7–14. 10.1016/S0006-291X(76)80264-1. [DOI] [PubMed] [Google Scholar]
  12. Yorgey P.; Davagnino J.; Kolter R. The maturation pathway of microcin B17, a peptide inhibitor of DNA gyrase. Mol. Microbiol. 1993, 9 (4), 897–905. 10.1111/j.1365-2958.1993.tb01747.x. [DOI] [PubMed] [Google Scholar]
  13. Imai Y.; Meyer K. J.; Iinishi A.; Favre-Godal Q.; Green R.; Manuse S.; Caboni M.; Mori M.; Niles S.; Ghiglieri M.; Honrao C.; Ma X.; Guo J. J.; Makriyannis A.; Linares-Otoya L.; Böhringer N.; Wuisan Z. G.; Kaur H.; Wu R.; Mateus A.; Typas A.; Savitski M. M.; Espinoza J. L.; O’Rourke A.; Nelson K. E.; Hiller S.; Noinaj N.; Schäberle T. F.; D’Onofrio A.; Lewis K. A new antibiotic selectively kills Gram-negative pathogens. Nature 2019, 576 (7787), 459–464. 10.1038/s41586-019-1791-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Medema M. H.; Blin K.; Cimermancic P.; de Jager V.; Zakrzewski P.; Fischbach M. A.; Weber T.; Takano E.; Breitling R. antiSMASH: rapid identification, annotation and analysis of secondary metabolite biosynthesis gene clusters in bacterial and fungal genome sequences. Nucleic Acids Res. 2011, 39 (Suppl. 2), W339–W346. 10.1093/nar/gkr466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Gerlt J. A.; Bouvier J. T.; Davidson D. B.; Imker H. J.; Sadkhin B.; Slater D. R.; Whalen K. L. Enzyme Function Initiative-Enzyme Similarity Tool (EFI-EST): A web tool for generating protein sequence similarity networks. Biochim. Biophys. Acta 2015, 1854 (8), 1019–37. 10.1016/j.bbapap.2015.04.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Tietz J. I.; Schwalen C. J.; Patel P. S.; Maxson T.; Blair P. M.; Tai H. C.; Zakai U. I.; Mitchell D. A. A new genome-mining tool redefines the lasso peptide biosynthetic landscape. Nat. Chem. Biol. 2017, 13 (5), 470–478. 10.1038/nchembio.2319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. van Heel A. J.; de Jong A.; Song C.; Viel J. H.; Kok J.; Kuipers O. P. BAGEL4: a user-friendly web server to thoroughly mine RiPPs and bacteriocins. Nucleic Acids Res. 2018, 46 (W1), W278–w281. 10.1093/nar/gky383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Merwin N. J.; Mousa W. K.; Dejong C. A.; Skinnider M. A.; Cannon M. J.; Li H.; Dial K.; Gunabalasingam M.; Johnston C.; Magarvey N. A. DeepRiPP integrates multiomics data to automate discovery of novel ribosomally synthesized natural products. Proc. Natl. Acad. Sci. U.S.A. 2020, 117 (1), 371–380. 10.1073/pnas.1901493116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Agrawal P.; Amir S.; Deepak; Barua D.; Mohanty D. RiPPMiner-Genome: A web resource for automated prediction of crosslinked chemical structures of RiPPs by genome mining. J. Mol. Biol. 2021, 433 (11), 166887 10.1016/j.jmb.2021.166887. [DOI] [PubMed] [Google Scholar]
  20. Moffat A. D.; Santos-Aberturas J.; Chandra G.; Truman A. W. A user guide for the identification of new RiPP biosynthetic gene clusters using a RiPPer-based workflow. Methods Mol. Biol. 2021, 2296, 227–247. 10.1007/978-1-0716-1358-0_14. [DOI] [PubMed] [Google Scholar]
  21. Mohimani H.; Kersten R. D.; Liu W. T.; Wang M.; Purvine S. O.; Wu S.; Brewer H. M.; Pasa-Tolic L.; Bandeira N.; Moore B. S.; Pevzner P. A.; Dorrestein P. C. Automated genome mining of ribosomal peptide natural products. ACS Chem. Biol. 2014, 9 (7), 1545–51. 10.1021/cb500199h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Lee Y. Y.; Guler M.; Chigumba D. N.; Wang S.; Mittal N.; Miller C.; Krummenacher B.; Liu H.; Cao L.; Kannan A.; Narayan K.; Slocum S. T.; Roth B. L.; Gurevich A.; Behsaz B.; Kersten R. D.; Mohimani H. HypoRiPPAtlas as an Atlas of hypothetical natural products for mass spectrometry database search. Nat. Commun. 2023, 14 (1), 4219. 10.1038/s41467-023-39905-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Russell A. H.; Truman A. W. Genome mining strategies for ribosomally synthesised and post-translationally modified peptides. Comput. Struct. Biotechnol. J. 2020, 18, 1838–1851. 10.1016/j.csbj.2020.06.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Hubrich F.; Bösch N. M.; Chepkirui C.; Morinaka B. I.; Rust M.; Gugger M.; Robinson S. L.; Vagstad A. L.; Piel J. Ribosomally derived lipopeptides containing distinct fatty acyl moieties. Proc. Natl. Acad. Sci. U S A 2022, 119 (3), 1–10. 10.1073/pnas.2113120119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Clark K. A.; Seyedsayamdost M. R. Bioinformatic atlas of radical SAM enzyme-modified RiPP natural products reveals an isoleucine-tryptophan crosslink. J. Am. Chem. Soc. 2022, 144 (39), 17876–17888. 10.1021/jacs.2c06497. [DOI] [PubMed] [Google Scholar]
  26. Wang S.; Lin S.; Fang Q.; Gyampoh R.; Lu Z.; Gao Y.; Clarke D. J.; Wu K.; Trembleau L.; Yu Y.; Kyeremeh K.; Milne B. F.; Tabudravu J.; Deng H. A ribosomally synthesised and post-translationally modified peptide containing a β-enamino acid and a macrocyclic motif. Nat. Commun. 2022, 13 (1), 5044. 10.1038/s41467-022-32774-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Malit J. J. L.; Wu C.; Liu L. L.; Qian P. Y. Global genome mining reveals the distribution of diverse thioamidated RiPP biosynthesis gene clusters. Front. Microbiol. 2021, 12, 635389 10.3389/fmicb.2021.635389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Ayikpoe R. S.; Shi C.; Battiste A. J.; Eslami S. M.; Ramesh S.; Simon M. A.; Bothwell I. R.; Lee H.; Rice A. J.; Ren H.; Tian Q.; Harris L. A.; Sarksian R.; Zhu L.; Frerk A. M.; Precord T. W.; van der Donk W. A.; Mitchell D. A.; Zhao H. A scalable platform to discover antimicrobials of ribosomal origin. Nat. Commun. 2022, 13 (1), 6135. 10.1038/s41467-022-33890-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Zdouc M. M.; van der Hooft J. J. J.; Medema M. H. Metabolome-guided genome mining of RiPP natural products. Trends Pharmacol. Sci. 2023, 44 (8), 532–541. 10.1016/j.tips.2023.06.004. [DOI] [PubMed] [Google Scholar]
  30. Fernandez-Cantos M. V.; Garcia-Morena D.; Yi Y.; Liang L.; Gómez-Vázquez E.; Kuipers O. P. Bioinformatic mining for RiPP biosynthetic gene clusters in Bacteroidales reveals possible new subfamily architectures and novel natural products. Front. Microbiol. 2023, 14, 1219272 10.3389/fmicb.2023.1219272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Ren H.; Dommaraju S. R.; Huang C.; Cui H.; Pan Y.; Nesic M.; Zhu L.; Sarlah D.; Mitchell D. A.; Zhao H. Genome mining unveils a class of ribosomal peptides with two amino termini. Nat. Commun. 2023, 14 (1), 1624. 10.1038/s41467-023-37287-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Ayikpoe R. S.; Zhu L.; Chen J. Y.; Ting C. P.; van der Donk W. A. Macrocyclization and backbone rearrangement during RiPP biosynthesis by a SAM-dependent domain-of-unknown-function 692. ACS Cent. Sci. 2023, 9 (5), 1008–1018. 10.1021/acscentsci.3c00160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. An J. S.; Lee H.; Kim H.; Woo S.; Nam H.; Lee J.; Lee J. Y.; Nam S. J.; Lee S. K.; Oh K. B.; Kim S.; Oh D. C. Discovery and biosynthesis of cihunamides, macrocyclic antibacterial RiPPs with a unique C-N linkage formed by CYP450 catalysis. Angew. Chem., Int. Ed. 2023, 62 (26), e202300998 10.1002/anie.202300998. [DOI] [PubMed] [Google Scholar]
  34. Lima S. T.; Ampolini B. G.; Underwood E. B.; Graf T. N.; Earp C. E.; Khedi I. C.; Pasquale M. A.; Chekan J. R. A widely distributed biosynthetic cassette is responsible for diverse plant side chain cross-linked cyclopeptides. Angew. Chem., Int. Ed. 2023, 62 (7), e202218082 10.1002/anie.202218082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Choi B.; Link A. J. Discovery, function, and engineering of graspetides. Trends Chem. 2023, 5 (8), 620–633. 10.1016/j.trechm.2023.04.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Wu C.; van der Donk W. A. Engineering of new-to-nature ribosomally synthesized and post-translationally modified peptide natural products. Curr. Opin. Biotechnol. 2021, 69, 221–231. 10.1016/j.copbio.2020.12.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Fu Y.; Xu Y.; Ruijne F.; Kuipers O. P. Engineering lanthipeptides by introducing a large variety of RiPP modifications to obtain new-to-nature bioactive peptides. FEMS Microbiol. Rev. 2023, 47 (3), fuad017 10.1093/femsre/fuad017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Guo E.; Fu L.; Fang X.; Xie W.; Li K.; Zhang Z.; Hong Z.; Si T. Robotic construction and screening of lanthipeptide variant libraries in Escherichia coli. ACS Synth. Biol. 2022, 11 (12), 3900–3911. 10.1021/acssynbio.2c00344. [DOI] [PubMed] [Google Scholar]
  39. Do T.; Link A. J. Protein engineering in ribosomally synthesized and post-translationally modified peptides (RiPPs). Biochemistry 2023, 62 (2), 201–209. 10.1021/acs.biochem.1c00714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Rawlings N. D.; Barrett A. J.; Thomas P. D.; Huang X.; Bateman A.; Finn R. D. The MEROPS database of proteolytic enzymes, their substrates and inhibitors in 2017 and a comparison with peptidases in the PANTHER database. Nucleic Acids Res. 2018, 46 (D1), D624–D632. 10.1093/nar/gkx1134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Booth M. C.; Bogie C. P.; Sahl H. G.; Siezen R. J.; Hatter K. L.; Gilmore M. S. Structural analysis and proteolytic activation of Enterococcus faecalis cytolysin, a novel lantibiotic. Mol. Microbiol. 1996, 21 (6), 1175–1184. 10.1046/j.1365-2958.1996.831449.x. [DOI] [PubMed] [Google Scholar]
  42. Gilmore M. S.; Segarra R. A.; Booth M. C. An HlyB-type function is required for expression of the Enterococcus faecalis hemolysin/bacteriocin. Infect. Immun. 1990, 58 (12), 3914–23. 10.1128/iai.58.12.3914-3923.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. McClerren A. L.; Cooper L. E.; Quan C.; Thomas P. M.; Kelleher N. L.; van der Donk W. A. Discovery and in vitro biosynthesis of haloduracin, a two-component lantibiotic. Proc. Natl. Acad. Sci. U. S. A. 2006, 103 (46), 17243–8. 10.1073/pnas.0606088103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Viel J. H.; Jaarsma A. H.; Kuipers O. P. Heterologous expression of mersacidin in Escherichia coli elucidates the mode of leader processing. ACS Synth. Biol. 2021, 10 (3), 600–608. 10.1021/acssynbio.0c00601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Viel J. H.; Kuipers O. P. Mutational studies of the mersacidin leader reveal the function of its unique two-step leader processing mechanism. ACS Synth. Biol. 2022, 11 (5), 1949–1957. 10.1021/acssynbio.2c00088. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Wang X.; Chen X.; Wang Z.; Zhuang M.; Zhong L.; Fu C.; Garcia R.; Müller R.; Zhang Y.; Yan J.; Wu D.; Huo L. Discovery and characterization of a myxobacterial lanthipeptide with unique biosynthetic features and anti-inflammatory activity. J. Am. Chem. Soc. 2023, 145 (30), 16924–16937. 10.1021/jacs.3c06014. [DOI] [PubMed] [Google Scholar]
  47. Begley M.; Cotter P. D.; Hill C.; Ross R. P. Identification of a novel two-peptide lantibiotic, lichenicidin, following rational genome mining for LanM proteins. Appl. Environ. Microbiol. 2009, 75 (17), 5451–5460. 10.1128/AEM.00730-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Dischinger J.; Josten M.; Szekat C.; Sahl H. G.; Bierbaum G. Production of the novel two-peptide lantibiotic lichenicidin by Bacillus licheniformis DSM 13. PLoS One 2009, 4 (8), e6788 10.1371/journal.pone.0006788. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Nguyen G. K.; Wang S.; Qiu Y.; Hemu X.; Lian Y.; Tam J. P. Butelase 1 is an Asx-specific ligase enabling peptide macrocyclization and synthesis. Nat. Chem. Biol. 2014, 10 (9), 732–8. 10.1038/nchembio.1586. [DOI] [PubMed] [Google Scholar]
  50. Harris K. S.; Durek T.; Kaas Q.; Poth A. G.; Gilding E. K.; Conlan B. F.; Saska I.; Daly N. L.; van der Weerden N. L.; Craik D. J.; Anderson M. A. Efficient backbone cyclization of linear peptides by a recombinant asparaginyl endopeptidase. Nat. Commun. 2015, 6, 10199. 10.1038/ncomms10199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Yoshimi A.; Umemura M.; Nagano N.; Koike H.; Machida M.; Abe K. Expression of ustR and the Golgi protease KexB are required for ustiloxin B biosynthesis in Aspergillus oryzae. AMB Express 2016, 6 (1), 9. 10.1186/s13568-016-0181-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Czekster C. M.; Ludewig H.; McMahon S. A.; Naismith J. H. Characterization of a dual function macrocyclase enables design and use of efficient macrocyclization substrates. Nat. Commun. 2017, 8 (1), 1045. 10.1038/s41467-017-00862-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Ongpipattanakul C.; Nair S. K. Biosynthetic proteases that catalyze the macrocyclization of ribosomally synthesized linear peptides. Biochemistry 2018, 57 (23), 3201–3209. 10.1021/acs.biochem.8b00114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Luo S.; Dong S. H. Recent advances in the discovery and biosynthetic study of eukaryotic RiPP natural products. Molecules 2019, 24 (8), 1541. 10.3390/molecules24081541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Rehm F. B. H.; Jackson M. A.; De Geyter E.; Yap K.; Gilding E. K.; Durek T.; Craik D. J. Papain-like cysteine proteases prepare plant cyclic peptide precursors for cyclization. Proc. Natl. Acad. Sci. U. S. A. 2019, 116 (16), 7831–7836. 10.1073/pnas.1901807116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Umemura M. Peptides derived from Kex2-processed repeat proteins are widely distributed and highly diverse in the Fungi kingdom. Fungal Biol. Biotechnol. 2020, 7, 11. 10.1186/s40694-020-00100-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Vogt E.; Sonderegger L.; Chen Y. Y.; Segessemann T.; Künzler M. Structural and functional analysis of peptides derived from KEX2-processed repeat proteins in Agaricomycetes using reverse genetics and peptidomics. Microbiol. Spectr. 2022, 10, e0202122 10.1128/spectrum.02021-22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Ozaki T.; Minami A.; Oikawa H. Recent advances in the biosynthesis of ribosomally synthesized and posttranslationally modified peptides of fungal origin. J. Antibiot. 2023, 76 (1), 3–13. 10.1038/s41429-022-00576-w. [DOI] [PubMed] [Google Scholar]
  59. Matabaro E.; Song H.; Sonderegger L.; Gherlone F.; Giltrap A.; Liver S.; Gossert A.; Künzler M.; Naismith J. H. Molecular insight into the enzymatic macrocyclization of multiply backbone N-methylated peptides. bioRxiv 2023, 10.1101/2022.07.21.500988. [DOI] [Google Scholar]
  60. Ludewig H.; Czekster C. M.; Oueis E.; Munday E. S.; Arshad M.; Synowsky S. A.; Bent A. F.; Naismith J. H. Characterization of the fast and promiscuous macrocyclase from plant PCY1 enables the use of simple substrates. ACS Chem. Biol. 2018, 13 (3), 801–811. 10.1021/acschembio.8b00050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Håvarstein L. S.; Diep D. B.; Nes I. F. A family of bacteriocin ABC transporters carry out proteolytic processing of their substrates concomitant with export. Mol. Microbiol. 1995, 16 (2), 229–40. 10.1111/j.1365-2958.1995.tb02295.x. [DOI] [PubMed] [Google Scholar]
  62. van Belkum M. J.; Worobo R. W.; Stiles M. E. Double-glycine-type leader peptides direct secretion of bacteriocins by ABC transporters: colicin V secretion in Lactococcus lactis. Mol. Microbiol. 1997, 23 (6), 1293–301. 10.1046/j.1365-2958.1997.3111677.x. [DOI] [PubMed] [Google Scholar]
  63. Li B.; Sher D.; Kelly L.; Shi Y.; Huang K.; Knerr P. J.; Joewono I.; Rusch D.; Chisholm S. W.; van der Donk W. A. Catalytic promiscuity in the biosynthesis of cyclic peptide secondary metabolites in planktonic marine cyanobacteria. Proc. Natl. Acad. Sci. U.S.A. 2010, 107 (23), 10430–5. 10.1073/pnas.0913677107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Haft D. H.; Basu M. K.; Mitchell D. A. Expansion of ribosomally produced natural products: a nitrile hydratase- and Nif11-related precursor family. BMC Biol. 2010, 8, 70. 10.1186/1741-7007-8-70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Freeman M. F.; Gurgui C.; Helf M. J.; Morinaka B. I.; Uria A. R.; Oldham N. J.; Sahl H. G.; Matsunaga S.; Piel J. Metagenome mining reveals polytheonamides as posttranslationally modified ribosomal peptides. Science 2012, 338, 387–90. 10.1126/science.1226121. [DOI] [PubMed] [Google Scholar]
  66. Dirix G.; Monsieurs P.; Dombrecht B.; Daniels R.; Marchal K.; Vanderleyden J.; Michiels J. Peptide signal molecules and bacteriocins in Gram-negative bacteria: a genome-wide in silico screening for peptides containing a double-glycine leader sequence and their cognate transporters. Peptides 2004, 25 (9), 1425–40. 10.1016/j.peptides.2003.10.028. [DOI] [PubMed] [Google Scholar]
  67. Cole T. J.; Parker J. K.; Feller A. L.; Wilke C. O.; Davies B. W. Evidence for widespread class II microcins in Enterobacterales genomes. Appl. Environ. Microbiol. 2022, 88 (23), e0148622 10.1128/aem.01486-22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Ishii S.; Yano T.; Ebihara A.; Okamoto A.; Manzoku M.; Hayashi H. Crystal structure of the peptidase domain of Streptococcus ComA, a bifunctional ATP-binding cassette transporter involved in the quorum-sensing pathway. J. Biol. Chem. 2010, 285 (14), 10777–85. 10.1074/jbc.M109.093781. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Ishii S.; Yano T.; Hayashi H. Expression and characterization of the peptidase domain of Streptococcus pneumoniae ComA, a bifunctional ATP-binding cassette transporter involved in quorum sensing pathway. J. Biol. Chem. 2006, 281 (8), 4726–31. 10.1074/jbc.M512516200. [DOI] [PubMed] [Google Scholar]
  70. Kotake Y.; Ishii S.; Yano T.; Katsuoka Y.; Hayashi H. Substrate recognition mechanism of the peptidase domain of the quorum-sensing-signal-producing ABC transporter ComA from Streptococcus. Biochemistry 2008, 47 (8), 2531–8. 10.1021/bi702253n. [DOI] [PubMed] [Google Scholar]
  71. Bobeica S. C.; Dong S. H.; Huo L.; Mazo N.; McLaughlin M. I.; Jimenéz-Osés G.; Nair S. K.; van der Donk W. A. Insights into AMS/PCAT transporters from biochemical and structural characterization of a double glycine motif protease. eLife 2019, 8, e42305 10.7554/eLife.42305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Huo L.; Zhao X.; Acedo J. Z.; Estrada P.; Nair S. K.; van der Donk W. A. Characterization of a dehydratase and methyltransferase in the biosynthesis of ribosomally synthesized and post-translationally modified peptides in Lachnospiraceae. ChemBioChem. 2020, 21 (1–2), 190–199. 10.1002/cbic.201900483. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Fujinami D.; Garcia de Gonzalo C. V.; Biswas S.; Hao Y.; Wang H.; Garg N.; Lukk T.; Nair S. K.; van der Donk W. A. Structural and mechanistic investigations of protein S-glycosyltransferases. Cell Chem. Biol. 2021, 28 (12), 1740–1749. 10.1016/j.chembiol.2021.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Arias-Orozco P.; Inklaar M.; Lanooij J.; Cebrián R.; Kuipers O. P. Functional expression and characterization of the highly promiscuous lanthipeptide synthetase SyncM, enabling the production of lanthipeptides with a broad range of ring topologies. ACS Synth. Biol. 2021, 10 (10), 2579–2591. 10.1021/acssynbio.1c00224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Bothwell I. R.; Caetano T.; Sarksian R.; Mendo S.; van der Donk W. A. Structural analysis of class I lanthipeptides from Pedobacter lusitanus NL19 reveals an unusual ring pattern. ACS Chem. Biol. 2021, 16 (6), 1019–1029. 10.1021/acschembio.1c00106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Ramesh S.; Guo X.; DiCaprio A. J.; De Lio A. M.; Harris L. A.; Kille B. L.; Pogorelov T. V.; Mitchell D. A. Bioinformatics-guided expansion and discovery of graspetides. ACS Chem. Biol. 2021, 16 (12), 2787–2797. 10.1021/acschembio.1c00672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Nguyen N. A.; Lin Z.; Mohanty I.; Garg N.; Schmidt E. W.; Agarwal V. An obligate peptidyl brominase underlies the discovery of highly distributed biosynthetic gene clusters in marine sponge microbiomes. J. Am. Chem. Soc. 2021, 143 (27), 10221–10231. 10.1021/jacs.1c03474. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Paoli L.; Ruscheweyh H. J.; Forneris C. C.; Hubrich F.; Kautsar S.; Bhushan A.; Lotti A.; Clayssen Q.; Salazar G.; Milanese A.; Carlström C. I.; Papadopoulou C.; Gehrig D.; Karasikov M.; Mustafa H.; Larralde M.; Carroll L. M.; Sánchez P.; Zayed A. A.; Cronin D. R.; Acinas S. G.; Bork P.; Bowler C.; Delmont T. O.; Gasol J. M.; Gossert A. D.; Kahles A.; Sullivan M. B.; Wincker P.; Zeller G.; Robinson S. L.; Piel J.; Sunagawa S. Biosynthetic potential of the global ocean microbiome. Nature 2022, 607 (7917), 111–118. 10.1038/s41586-022-04862-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Nguyen N. A.; Cong Y.; Hurrell R. C.; Arias N.; Garg N.; Puri A. W.; Schmidt E. W.; Agarwal V. A silent biosynthetic gene cluster from a methanotrophic bacterium potentiates discovery of a substrate promiscuous proteusin cyclodehydratase. ACS Chem. Biol. 2022, 17 (6), 1577–1585. 10.1021/acschembio.2c00251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Zhang S. S.; Xiong J.; Cui J. J.; Ma K. L.; Wu W. L.; Li Y.; Luo S.; Gao K.; Dong S. H. Lanthipeptides from the same core sequence: characterization of a class II lanthipeptide synthetase from Microcystis aeruginosa NIES-88. Org. Lett. 2022, 24 (11), 2226–2231. 10.1021/acs.orglett.2c00573. [DOI] [PubMed] [Google Scholar]
  81. Arias-Orozco P.; Yi Y.; Ruijne F.; Cebrián R.; Kuipers O. P. Investigating the specificity of the dehydration and cyclization reactions in engineered lanthipeptides by Synechococcal SyncM. ACS Synth. Biol. 2023, 12 (1), 164–177. 10.1021/acssynbio.2c00455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Nguyen N. A.; Agarwal V. A leader-guided substrate tolerant RiPP brominase allows Suzuki-Miyaura cross-coupling reactions for peptides and proteins. Biochemistry 2023, 62 (12), 1838–1843. 10.1021/acs.biochem.3c00222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Wang J.; Ge X.; Zhang L.; Teng K.; Zhong J. One-pot synthesis of class II lanthipeptide bovicin HJ50 via an engineered lanthipeptide synthetase. Sci. Rep. 2016, 6, 38630. 10.1038/srep38630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Kuipers A.; Meijer-Wierenga J.; Rink R.; Kluskens L. D.; Moll G. N. Mechanistic dissection of the enzyme complexes involved in biosynthesis of lacticin 3147 and nisin. Appl. Environ. Microbiol. 2008, 74 (21), 6591–7. 10.1128/AEM.01334-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Nishie M.; Shioya K.; Nagao J.; Jikuya H.; Sonomoto K. ATP-dependent leader peptide cleavage by NukT, a bifunctional ABC transporter, during lantibiotic biosynthesis. J. Biosci. Bioeng. 2009, 108 (6), 460–464. 10.1016/j.jbiosc.2009.06.002. [DOI] [PubMed] [Google Scholar]
  86. Rahman S.; McHaourab H. S. ATP-dependent interactions of a cargo protein with the transmembrane domain of a polypeptide processing and secretion ABC transporter. J. Biol. Chem. 2020, 295 (43), 14678–14685. 10.1074/jbc.RA120.014934. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Lin D. Y.; Huang S.; Chen J. Crystal structures of a polypeptide processing and secretion transporter. Nature 2015, 523 (7561), 425–30. 10.1038/nature14623. [DOI] [PubMed] [Google Scholar]
  88. Bhattacharya S.; Palillo A. Structural and dynamic studies of the peptidase domain from Clostridium thermocellum PCAT1. Protein Sci. 2022, 31 (2), 498–512. 10.1002/pro.4248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Kieuvongngam V.; Olinares P. D. B.; Palillo A.; Oldham M. L.; Chait B. T.; Chen J. Structural basis of substrate recognition by a polypeptide processing and secretion transporter. eLife 2020, 9, e51492 10.7554/eLife.51492. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Rawlings N. D.; Barrett A. J. Introduction: the clans and families of cysteine peptidases. Handbook of proteolytic enzymes 2013, 1743. 10.1016/B978-0-12-382219-2.00404-X. [DOI] [Google Scholar]
  91. Helf M. J.; Freeman M. F.; Piel J. Investigations into PoyH, a promiscuous protease from polytheonamide biosynthesis. J. Ind. Microbiol. Biotechnol. 2019, 46 (3–4), 551–563. 10.1007/s10295-018-02129-3. [DOI] [PubMed] [Google Scholar]
  92. Sikandar A.; Lopatniuk M.; Luzhetskyy A.; Müller R.; Koehnke J. Total in vitro biosynthesis of the thioamitide thioholgamide and investigation of the pathway. J. Am. Chem. Soc. 2022, 144 (11), 5136–5144. 10.1021/jacs.2c00402. [DOI] [PubMed] [Google Scholar]
  93. Verma S.; Dixit R.; Pandey K. C. Cysteine proteases: modes of activation and future prospects as pharmacological targets. Front. Pharmacol. 2016, 7, 107. 10.3389/fphar.2016.00107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Yan K. P.; Li Y.; Zirah S.; Goulard C.; Knappe T. A.; Marahiel M. A.; Rebuffat S. Dissecting the maturation steps of the lasso peptide microcin J25 in vitro. ChemBioChem. 2012, 13 (7), 1046–52. 10.1002/cbic.201200016. [DOI] [PubMed] [Google Scholar]
  95. DiCaprio A. J.; Firouzbakht A.; Hudson G. A.; Mitchell D. A. Enzymatic reconstitution and biosynthetic investigation of the lasso peptide fusilassin. J. Am. Chem. Soc. 2019, 141, 290–297. 10.1021/jacs.8b09928. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Koos J. D.; Link A. J. Heterologous and in vitro reconstitution of fuscanodin, a lasso peptide from Thermobifida fusca. J. Am. Chem. Soc. 2019, 141 (2), 928–935. 10.1021/jacs.8b10724. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Kretsch A. M.; Gadgil M. G.; DiCaprio A. J.; Barrett S. E.; Kille B. L.; Si Y.; Zhu L.; Mitchell D. A. Peptidase activation by a leader peptide-bound RiPP recognition element. Biochemistry 2023, 62 (4), 956–967. 10.1021/acs.biochem.2c00700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Zyubko T.; Serebryakova M.; Andreeva J.; Metelev M.; Lippens G.; Dubiley S.; Severinov K. Efficient in vivo synthesis of lasso peptide pseudomycoidin proceeds in the absence of both the leader and the leader peptidase. Chem. Sci. 2019, 10 (42), 9699–9707. 10.1039/C9SC02370D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Si Y.; Kretsch A. M.; Daigh L. M.; Burk M. J.; Mitchell D. A. Cell-free biosynthesis to evaluate lasso peptide formation and enzyme-substrate tolerance. J. Am. Chem. Soc. 2021, 143 (15), 5917–5927. 10.1021/jacs.1c01452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Hegemann J. D.; Jeanne Dit Fouque K.; Santos-Fernandez M.; Fernandez-Lima F. A bifunctional leader peptidase/ABC transporter protein is involved in the maturation of the lasso peptide cochonodin I from Streptococcus suis. J. Nat. Prod. 2021, 84 (10), 2683–2691. 10.1021/acs.jnatprod.1c00514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Lubelski J.; Rink R.; Khusainov R.; Moll G. N.; Kuipers O. P. Biosynthesis, immunity, regulation, mode of action and engineering of the model lantibiotic nisin. Cell. Mol. Life Sci. 2008, 65 (3), 455–476. 10.1007/s00018-007-7171-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Xu Y.; Li X.; Li R.; Li S.; Ni H.; Wang H.; Xu H.; Zhou W.; Saris P. E.; Yang W.; Qiao M.; Rao Z. Structure of the nisin leader peptidase NisP revealing a C-terminal autocleavage activity. Acta Crystallogr. D Biol. Crystallogr. 2014, 70 (Pt 6), 1499–505. 10.1107/S1399004714004234. [DOI] [PubMed] [Google Scholar]
  103. Siezen R. J.; Rollema H. S.; Kuipers O. P.; de Vos W. M. Homology modelling of the Lactococcus lactis leader peptidase NisP and its interaction with the precursor of the lantibiotic nisin. Protein Eng. Des. Sel. 1995, 8 (2), 117–125. 10.1093/protein/8.2.117. [DOI] [PubMed] [Google Scholar]
  104. Lagedroste M.; Smits S. H. J.; Schmitt L. Substrate specificity of the secreted nisin leader peptidase NisP. Biochemistry 2017, 56 (30), 4005–4014. 10.1021/acs.biochem.7b00524. [DOI] [PubMed] [Google Scholar]
  105. Lagedroste M.; Reiners J.; Smits S. H. J.; Schmitt L. Systematic characterization of position one variants within the lantibiotic nisin. Sci. Rep. 2019, 9 (1), 935. 10.1038/s41598-018-37532-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Ortega M. A.; Velásquez J. E.; Garg N.; Zhang Q.; Joyce R. E.; Nair S. K.; van der Donk W. A. Substrate specificity of the lanthipeptide peptidase ElxP and the oxidoreductase ElxO. ACS Chem. Biol. 2014, 9 (8), 1718–25. 10.1021/cb5002526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Velásquez J. E.; Zhang X.; van der Donk W. A. Biosynthesis of the antimicrobial peptide epilancin 15X and its N-terminal lactate moiety. Chem. Biol. 2011, 18, 857–867. 10.1016/j.chembiol.2011.05.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Lee J.; McIntosh J.; Hathaway B. J.; Schmidt E. W. Using marine natural products to discover a protease that catalyzes peptide macrocyclization of diverse substrates. J. Am. Chem. Soc. 2009, 131 (6), 2122–4. 10.1021/ja8092168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Agarwal V.; Pierce E.; McIntosh J.; Schmidt E. W.; Nair S. K. Structures of cyanobactin maturation enzymes define a family of transamidating proteases. Chem. Biol. 2012, 19 (11), 1411–22. 10.1016/j.chembiol.2012.09.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Donia M. S.; Hathaway B. J.; Sudek S.; Haygood M. G.; Rosovitz M. J.; Ravel J.; Schmidt E. W. Natural combinatorial peptide libraries in cyanobacterial symbionts of marine ascidians. Nat. Chem. Biol. 2006, 2 (12), 729–35. 10.1038/nchembio829. [DOI] [PubMed] [Google Scholar]
  111. Koehnke J.; Bent A.; Houssen W. E.; Zollman D.; Morawitz F.; Shirran S.; Vendome J.; Nneoyiegbe A. F.; Trembleau L.; Botting C. H.; Smith M. C. M.; Jaspars M.; Naismith J. H. The mechanism of patellamide macrocyclization revealed by the characterization of the PatG macrocyclase domain. Nat. Struct. Mol. Biol. 2012, 19, 767–72. 10.1038/nsmb.2340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Koehnke J.; Bent A. F.; Houssen W. E.; Mann G.; Jaspars M.; Naismith J. H. The structural biology of patellamide biosynthesis. Curr. Opin. Struct. Biol. 2014, 29, 112–21. 10.1016/j.sbi.2014.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Donia M. S.; Ravel J.; Schmidt E. W. A global assembly line for cyanobactins. Nat. Chem. Biol. 2008, 4 (6), 341–3. 10.1038/nchembio.84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. McIntosh J. A.; Robertson C. R.; Agarwal V.; Nair S. K.; Bulaj G. W.; Schmidt E. W. Circular logic: nonribosomal peptide-like macrocyclization with a ribosomal peptide catalyst. J. Am. Chem. Soc. 2010, 132 (44), 15499–501. 10.1021/ja1067806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Houssen W. E.; Koehnke J.; Zollman D.; Vendome J.; Raab A.; Smith M. C.; Naismith J. H.; Jaspars M. The discovery of new cyanobactins from Cyanothece PCC 7425 defines a new signature for processing of patellamides. ChemBioChem. 2012, 13 (18), 2683–9. 10.1002/cbic.201200661. [DOI] [PubMed] [Google Scholar]
  116. Houssen W. E.; Bent A. F.; McEwan A. R.; Pieiller N.; Tabudravu J.; Koehnke J.; Mann G.; Adaba R. I.; Thomas L.; Hawas U. W.; Liu H.; Schwarz-Linek U.; Smith M. C.; Naismith J. H.; Jaspars M. An efficient method for the in vitro production of azol(in)e-based cyclic peptides. Angew. Chem., Int. Ed. 2014, 53, 14171–4. 10.1002/anie.201408082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Oueis E.; Jaspars M.; Westwood N. J.; Naismith J. H. Enzymatic Macrocyclization of 1,2,3-Triazole Peptide Mimetics. Angew. Chem., Int. Ed. 2016, 55 (19), 5842–5. 10.1002/anie.201601564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Oueis E.; Stevenson H.; Jaspars M.; Westwood N. J.; Naismith J. H. Bypassing the proline/thiazoline requirement of the macrocyclase PatG. Chem. Commun. 2017, 53 (91), 12274–12277. 10.1039/C7CC06550G. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Zhang Y.; Hong Z.; Zhou L.; Zhang Z.; Tang T.; Guo E.; Zheng J.; Wang C.; Dai L.; Si T.; Wang H. Biosynthesis of gut-microbiota-derived lantibiotics reveals a subgroup of S8 family proteases for class III leader removal. Angew. Chem. 2022, 134 (6), e202114414 10.1002/ange.202114414. [DOI] [PubMed] [Google Scholar]
  120. Völler G. H.; Krawczyk B.; Ensle P.; Süssmuth R. D. Involvement and unusual substrate specificity of a prolyl oligopeptidase in class III lanthipeptide maturation. J. Am. Chem. Soc. 2013, 135 (20), 7426–9. 10.1021/ja402296m. [DOI] [PubMed] [Google Scholar]
  121. Chen S.; Xu B.; Chen E.; Wang J.; Lu J.; Donadio S.; Ge H.; Wang H. Zn-dependent bifunctional proteases are responsible for leader peptide processing of class III lanthipeptides. Proc. Natl. Acad. Sci. U. S. A. 2019, 116 (7), 2533–2538. 10.1073/pnas.1815594116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Ren H.; Shi C.; Bothwell I. R.; van der Donk W. A.; Zhao H. Discovery and characterization of a class IV lanthipeptide with a nonoverlapping ring pattern. ACS Chem. Biol. 2020, 15 (6), 1642–1649. 10.1021/acschembio.0c00267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Völler G. H.; Krawczyk J. M.; Pesic A.; Krawczyk B.; Nachtigall J.; Süssmuth R. D. Characterization of new class III lantibiotics-erythreapeptin, avermipeptin and griseopeptin from Saccharopolyspora erythraea, Streptomyces avermitilis and Streptomyces griseus demonstrates stepwise N-terminal leader processing. ChemBioChem. 2012, 13, 1174–83. 10.1002/cbic.201200118. [DOI] [PubMed] [Google Scholar]
  124. Krawczyk J. M.; Völler G. H.; Krawczyk B.; Kretz J.; Brönstrup M.; Süssmuth R. D. Heterologous expression and engineering studies of labyrinthopeptins, class III lantibiotics from Actinomadura namibiensis. Chem. Biol. 2013, 20 (1), 111–22. 10.1016/j.chembiol.2012.10.023. [DOI] [PubMed] [Google Scholar]
  125. Krawczyk B.; Völler G. H.; Völler J.; Ensle P.; Süssmuth R. D. Curvopeptin: a new lanthionine-containing class III lantibiotic and its co-substrate promiscuous synthetase. ChemBioChem. 2012, 13 (14), 2065–71. 10.1002/cbic.201200417. [DOI] [PubMed] [Google Scholar]
  126. Zhang Q.; Doroghazi J. R.; Zhao X.; Walker M. C.; van der Donk W. A. Expanded natural product diversity revealed by analysis of lanthipeptide-like gene clusters in actinobacteria. Appl. Environ. Microbiol. 2015, 81 (13), 4339–4350. 10.1128/AEM.00635-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Iftime D.; Jasyk M.; Kulik A.; Imhoff J. F.; Stegmann E.; Wohlleben W.; Süssmuth R. D.; Weber T. Streptocollin, a type IV lanthipeptide produced by Streptomyces collinus Tü 365. ChemBioChem. 2015, 16 (18), 2615–23. 10.1002/cbic.201500377. [DOI] [PubMed] [Google Scholar]
  128. Jungmann N. A.; van Herwerden E. F.; Hügelland M.; Süssmuth R. D. The supersized class III lanthipeptide stackepeptin displays motif multiplication in the core peptide. ACS Chem. Biol. 2016, 11 (1), 69–76. 10.1021/acschembio.5b00651. [DOI] [PubMed] [Google Scholar]
  129. Wiebach V.; Mainz A.; Schnegotzki R.; Siegert M. J.; Hügelland M.; Pliszka N.; Süssmuth R. D. An amphipathic alpha-helix guides maturation of the ribosomally-synthesized lipolanthines. Angew. Chem., Int. Ed. 2020, 59, 16777. 10.1002/anie.202003804. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Xue D.; Older E. A.; Zhong Z.; Shang Z.; Chen N.; Dittenhauser N.; Hou L.; Cai P.; Walla M. D.; Dong S.; Tang X.; Chen H.; Nagarkatti P.; Nagarkatti M.; Li Y.; Li J. Correlational networking guides the discovery of unclustered lanthipeptide protease-encoding genes. Nat. Commun. 2022, 13 (1), 1647. 10.1038/s41467-022-29325-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Rodríguez-Sáinz M. C.; Hernández-Chico C.; Moreno F. Molecular characterization of pmbA, an Escherichia coli chromosomal gene required for the production of the antibiotic peptide MccB17. Mol. Microbiol. 1990, 4 (11), 1921–32. 10.1111/j.1365-2958.1990.tb02041.x. [DOI] [PubMed] [Google Scholar]
  132. Allali N.; Afif H.; Couturier M.; Van Melderen L. The highly conserved TldD and TldE proteins of Escherichia coli are involved in microcin B17 processing and in CcdA degradation. J. Bacteriol. 2002, 184 (12), 3224–31. 10.1128/JB.184.12.3224-3231.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Ghilarov D.; Serebryakova M.; Stevenson C. E. M.; Hearnshaw S. J.; Volkov D. S.; Maxwell A.; Lawson D. M.; Severinov K. The origins of specificity in the microcin-processing protease TldD/E. Structure 2017, 25 (10), 1549–1561. 10.1016/j.str.2017.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Tsibulskaya D.; Mokina O.; Kulikovsky A.; Piskunova J.; Severinov K.; Serebryakova M.; Dubiley S. The product of Yersinia pseudotuberculosis mcc operon is a peptide-cytidine antibiotic activated inside producing cells by the TldD/E protease. J. Am. Chem. Soc. 2017, 139 (45), 16178–16187. 10.1021/jacs.7b07118. [DOI] [PubMed] [Google Scholar]
  135. Vobruba S.; Kadlcik S.; Janata J.; Kamenik Z. TldD/TldE peptidases and N-deacetylases: A structurally unique yet ubiquitous protein family in the microbial metabolism. Microbiol. Res. 2022, 265, 127186 10.1016/j.micres.2022.127186. [DOI] [PubMed] [Google Scholar]
  136. Travin D. Y.; Metelev M.; Serebryakova M.; Komarova E. S.; Osterman I. A.; Ghilarov D.; Severinov K. Biosynthesis of translation inhibitor klebsazolicin proceeds through heterocyclization and N-terminal amidine formation catalyzed by a single YcaO enzyme. J. Am. Chem. Soc. 2018, 140 (16), 5625–5633. 10.1021/jacs.8b02277. [DOI] [PubMed] [Google Scholar]
  137. Ting C. P.; Funk M. A.; Halaby S. L.; Zhang Z.; Gonen T.; van der Donk W. A. Use of a scaffold peptide in the biosynthesis of amino acid-derived natural products. Science 2019, 365 (6450), 280–284. 10.1126/science.aau6232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Daniels P. N.; Lee H.; Splain R. A.; Ting C. P.; Zhu L.; Zhao X.; Moore B. S.; van der Donk W. A. A biosynthetic pathway to aromatic amines that uses glycyl-tRNA as nitrogen donor. Nat. Chem. 2022, 14 (1), 71–77. 10.1038/s41557-021-00802-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Zhang Z.; van der Donk W. A. Nonribosomal peptide extension by a peptide amino-acyl tRNA ligase. J. Am. Chem. Soc. 2019, 141 (50), 19625–19633. 10.1021/jacs.9b07111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Yu Y.; van der Donk W. A. Biosynthesis of 3-thia-α-amino acids on a carrier peptide. Proc. Natl. Acad. Sci. U. S. A. 2022, 119 (29), e2205285119 10.1073/pnas.2205285119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Popp P. F.; Friebel L.; Benjdia A.; Guillot A.; Berteau O.; Mascher T. The epipeptide biosynthesis locus epeXEPAB is widely distributed in Firmicutes and triggers intrinsic cell envelope stress. Microb. Physiol. 2021, 31 (3), 1–12. 10.1159/000516750. [DOI] [PubMed] [Google Scholar]
  142. Franz L.; Kazmaier U.; Truman A. W.; Koehnke J. Bottromycins - biosynthesis, synthesis and activity. Nat. Prod. Rep. 2021, 38 (9), 1659–1683. 10.1039/D0NP00097C. [DOI] [PubMed] [Google Scholar]
  143. Sikandar A.; Franz L.; Melse O.; Antes I.; Koehnke J. Thiazoline-specific amidohydrolase PurAH Is the gatekeeper of bottromycin biosynthesis. J. Am. Chem. Soc. 2019, 141 (25), 9748–9752. 10.1021/jacs.8b12231. [DOI] [PubMed] [Google Scholar]
  144. Matsui M.; Fowler J. H.; Walling L. L. Leucine aminopeptidases: diversity in structure and function. Biol. Chem. 2006, 387, 1535–1544. 10.1515/BC.2006.191. [DOI] [PubMed] [Google Scholar]
  145. Mann G.; Huo L.; Adam S.; Nardone B.; Vendome J.; Westwood N. J.; Müller R.; Koehnke J. Structure and substrate recognition of the bottromycin maturation enzyme BotP. Chembiochem 2016, 17 (23), 2286–2292. 10.1002/cbic.201600406. [DOI] [PubMed] [Google Scholar]
  146. Segarra R. A.; Booth M. C.; Morales D. A.; Huycke M. M.; Gilmore M. S. Molecular characterization of the Enterococcus faecalis cytolysin activator. Infect. Immun. 1991, 59 (4), 1239–46. 10.1128/iai.59.4.1239-1246.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Tang W.; Bobeica S. C.; Wang L.; van der Donk W. A. CylA is a sequence-specific protease involved in toxin biosynthesis. J. Ind. Microbiol. Biotechnol. 2019, 46 (3–4), 537–549. 10.1007/s10295-018-2110-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Tang W.; Dong S.-H.; Repka L. M.; He C.; Nair S. K.; van der Donk W. A. Applications of the class II lanthipeptide protease LicP for sequence-specific, traceless peptide bond cleavage. Chem. Sci. 2015, 6, 6270–6279. 10.1039/C5SC02329G. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Barbosa J. C.; Mösker E.; Faria R.; Süssmuth R. D.; Mendo S.; Caetano T. Class II two-peptide lanthipeptide proteases: exploring LicTP for biotechnological applications. Appl. Microbiol. Biotechnol. 2023, 107 (5–6), 1687–1696. 10.1007/s00253-023-12388-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Wang J.; Zhang L.; Teng K.; Sun S.; Sun Z.; Zhong J. Cerecidins, novel lantibiotics from Bacillus cereus with potent antimicrobial activity. Appl. Environ. Microbiol. 2014, 80 (8), 2633–2643. 10.1128/AEM.03751-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Navaratna M. A.; Sahl H. G.; Tagg J. R. Identification of genes encoding two-component lantibiotic production in Staphylococcus aureus C55 and other phage group II S. aureus strains and demonstration of an association with the exfoliative toxin B gene. Infect. Immun. 1999, 67 (8), 4268–71. 10.1128/IAI.67.8.4268-4271.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Holo H.; Jeknic Z.; Daeschel M.; Stevanovic S.; Nes I. F. Plantaricin W from Lactobacillus plantarum belongs to a new family of two-peptide lantibiotics. Microbiology 2001, 147 (3), 643–651. 10.1099/00221287-147-3-643. [DOI] [PubMed] [Google Scholar]
  153. Viel J. H.; van Tilburg A. Y.; Kuipers O. P. Characterization of leader processing shows that partially processed mersacidin is activated by AprE after export. Front. Microbiol. 2021, 12, 765659 10.3389/fmicb.2021.765659. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Suda S.; Cotter P. D.; Hill C.; Ross R. P. Lacticin 3147-biosynthesis, molecular analysis, immunity, bioengineering and applications. Curr. Prot. Pept. Sci. 2012, 13 (3), 193–204. 10.2174/138920312800785021. [DOI] [PubMed] [Google Scholar]
  155. Wang B.; Zhao A.; Novick R. P.; Muir T. W. Key driving forces in the biosynthesis of autoinducing peptides required for staphylococcal virulence. Proc. Natl. Acad. Sci. U. S. A. 2015, 112 (34), 10679–10684. 10.1073/pnas.1506030112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Zhao A.; Bodine S. P.; Xie Q.; Wang B.; Ram G.; Novick R. P.; Muir T. W. Reconstitution of the S. aureus agr quorum sensing pathway reveals a direct role for the integral membrane protease MroQ in pheromone biosynthesis. Proc. Natl. Acad. Sci. U. S. A. 2022, 119 (33), e2202661119 10.1073/pnas.2202661119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Adam S.; Zheng D.; Klein A.; Volz C.; Mullen W.; Shirran S. L.; Smith B. O.; Kalinina O. V.; Müller R.; Koehnke J. Unusual peptide-binding proteins guide pyrroloindoline alkaloid formation in crocagin biosynthesis. Nat. Chem. 2023, 15 (4), 560–568. 10.1038/s41557-023-01153-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Bruender N. A.; Bandarian V. The Creatininase Homolog MftE from Mycobacterium smegmatis Catalyzes a Peptide Cleavage Reaction in the Biosynthesis of a Novel Ribosomally Synthesized Post-translationally Modified Peptide (RiPP). J. Biol. Chem. 2017, 292 (10), 4371–4381. 10.1074/jbc.M116.762062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Zheng Q.; Fang H.; Liu W. Post-translational modifications involved in the biosynthesis of thiopeptide antibiotics. Org. Biomol. Chem. 2017, 15, 3376–3390. 10.1039/C7OB00466D. [DOI] [PubMed] [Google Scholar]
  160. Nguyen D. T.; Le T. T.; Rice A. J.; Hudson G. A.; van der Donk W. A.; Mitchell D. A. Accessing Diverse Pyridine-Based Macrocyclic Peptides by a Two-Site Recognition Pathway. J. Am. Chem. Soc. 2022, 144 (25), 11263–11269. 10.1021/jacs.2c02824. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Zheng Y.; Nair S. K. YcaO-mediated ATP-dependent peptidase activity in ribosomal peptide biosynthesis. Nat. Chem. Biol. 2023, 19 (1), 111–119. 10.1038/s41589-022-01141-0. [DOI] [PubMed] [Google Scholar]
  162. Wang C. Y.; Medlin J. S.; Nguyen D. R.; Disbennett W. M.; Dawid S. Molecular determinants of substrate selectivity of a pneumococcal Rgg-regulated peptidase-containing ABC transporter. MBio 2020, 11 (1), e02502–19. 10.1128/mBio.02502-19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Paik S. H.; Chakicherla A.; Hansen J. N. Identification and characterization of the structural and transporter genes for, and the chemical and biological properties of, sublancin 168, a novel lantibiotic produced by Bacillus subtilis 168. J. Biol. Chem. 1998, 273 (36), 23134–42. 10.1074/jbc.273.36.23134. [DOI] [PubMed] [Google Scholar]
  164. Singh V.; Rao A. Distribution and diversity of glycocin biosynthesis gene clusters beyond Firmicutes. Glycobiology 2021, 31 (2), 89–102. 10.1093/glycob/cwaa061. [DOI] [PubMed] [Google Scholar]
  165. Wang X.; Wang Z.; Dong Z.; Yan Y.; Zhang Y.; Huo L. Deciphering the biosynthesis of novel class I lanthipeptides from marine Pseudoalteromonas reveals a dehydratase PsfB with dethiolation activity. ACS Chem. Biol. 2023, 18 (5), 1218–1227. 10.1021/acschembio.3c00135. [DOI] [PubMed] [Google Scholar]
  166. Scott T. A.; Verest M.; Farnung J.; Forneris C. C.; Robinson S. L.; Ji X.; Hubrich F.; Chepkirui C.; Richter D. U.; Huber S.; Rust P.; Streiff A. B.; Zhang Q.; Bode J. W.; Piel J. Widespread microbial utilization of ribosomal β-amino acid-containing peptides and proteins. Chem. 2022, 8 (10), 2659–2677. 10.1016/j.chempr.2022.09.017. [DOI] [Google Scholar]
  167. Luu D. D.; Joe A.; Chen Y.; Parys K.; Bahar O.; Pruitt R.; Chan L. J. G.; Petzold C. J.; Long K.; Adamchak C.; Stewart V.; Belkhadir Y.; Ronald P. C. Biosynthesis and secretion of the microbial sulfated peptide RaxX and binding to the rice XA21 immune receptor. Proc. Natl. Acad. Sci. U. S. A. 2019, 116 (17), 8525–8534. 10.1073/pnas.1818275116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Qiu R.; Pei W.; Zhang L.; Lin J.; Ji G. Identification of the putative staphylococcal AgrB catalytic residues involving the proteolytic cleavage of AgrD to generate autoinducing peptide. J. Biol. Chem. 2005, 280 (17), 16695–704. 10.1074/jbc.M411372200. [DOI] [PubMed] [Google Scholar]
  169. Zhao A.; Bodine S. P.; Xie Q.; Wang B.; Ram G.; Novick R. P.; Muir T. W. Reconstitution of the S. aureus agr quorum sensing pathway reveals a direct role for the integral membrane protease MroQ in pheromone biosynthesis. Proc. Natl. Acad. Sci. U.S.A. 2022, 119 (33), e2202661119 10.1073/pnas.2202661119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. van der Meer J. R.; Polman J.; Beerthuyzen M. M.; Siezen R. J.; Kuipers O. P.; de Vos W. M. Characterization of the Lactococcus lactis nisin A operon genes nisP, encoding a subtilisin-like serine protease involved in precursor processing, and nisR, encoding a regulatory protein involved in nisin biosynthesis. J. Bacteriol. 1993, 175 (9), 2578–88. 10.1128/jb.175.9.2578-2588.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Suzuki M.; Komaki H.; Kaweewan I.; Dohra H.; Hemmi H.; Nakagawa H.; Yamamura H.; Hayakawa M.; Kodani S. Isolation and structure determination of new linear azole-containing peptides spongiicolazolicins A and B from Streptomyces sp. CWH03. Appl. Microbiol. Biotechnol. 2021, 105, 93–104. 10.1007/s00253-020-11016-w. [DOI] [PubMed] [Google Scholar]
  172. Cordell G. A.; Daley S. Pyrroloquinoline quinone chemistry, biology, and biosynthesis. Chem. Res. Toxicol. 2022, 35 (3), 355–377. 10.1021/acs.chemrestox.1c00340. [DOI] [PubMed] [Google Scholar]
  173. Wang P.; Xia Y.; Li J.; Kang Z.; Zhou J.; Chen J. Overexpression of pyrroloquinoline quinone biosynthetic genes affects l-sorbose production in Gluconobacter oxydans WSH-003. Biochem. Eng. J. 2016, 112, 70–77. 10.1016/j.bej.2016.04.011. [DOI] [Google Scholar]
  174. Hölscher T.; Görisch H. Knockout and overexpression of pyrroloquinoline quinone biosynthetic genes in Gluconobacter oxydans 621H. J. Bacteriol. 2006, 188 (21), 7668–76. 10.1128/JB.01009-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Nanudorn P.; Thiengmag S.; Biermann F.; Erkoc P.; Dirnberger S. D.; Phan T. N.; Fürst R.; Ueoka R.; Helfrich E. J. N. Atropopeptides are a novel family of ribosomally synthesized and posttranslationally modified peptides with a complex molecular shape. Angew. Chem., Int. Ed. 2022, 61 (41), e202208361 10.1002/anie.202208361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Li X.; Ma S.; Zhang Q. Chemical synthesis and biosynthesis of Darobactin. Tetrahedron Lett. 2023, 116, 154337 10.1016/j.tetlet.2023.154337. [DOI] [Google Scholar]
  177. Jin M.; Wright S. A.; Beer S. V.; Clardy J. The biosynthetic gene cluster of Pantocin A provides insights into biosynthesis and a tool for screening. Angew. Chem., Int. Ed. 2003, 42 (25), 2902–5. 10.1002/anie.200351054. [DOI] [PubMed] [Google Scholar]
  178. Crone K. K.; Jomori T.; Miller F. S.; Gralnick J. A.; Elias M. H.; Freeman M. F. RiPP enzyme heterocomplex structure-guided discovery of a bacterial borosin α-N-methylated peptide natural product. RSC Chem. Biol. 2023, 4 (10), 804–816. 10.1039/D3CB00093A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Hug J. J.; Dastbaz J.; Adam S.; Revermann O.; Koehnke J.; Krug D.; Müller R. Biosynthesis of Cittilins, Unusual Ribosomally Synthesized and Post-translationally Modified Peptides from Myxococcus xanthus. ACS Chem. Biol. 2020, 15 (8), 2221–2231. 10.1021/acschembio.0c00430. [DOI] [PubMed] [Google Scholar]
  180. Ziemert N.; Ishida K.; Liaimer A.; Hertweck C.; Dittmann E. Ribosomal synthesis of tricyclic depsipeptides in bloom-forming cyanobacteria. Angew. Chem., Int. Ed. 2008, 47 (40), 7756–9. 10.1002/anie.200802730. [DOI] [PubMed] [Google Scholar]
  181. Weiz A. R.; Ishida K.; Makower K.; Ziemert N.; Hertweck C.; Dittmann E. Leader peptide and a membrane protein scaffold guide the biosynthesis of the tricyclic peptide microviridin. Chem. Biol. 2011, 18 (11), 1413–21. 10.1016/j.chembiol.2011.09.011. [DOI] [PubMed] [Google Scholar]
  182. Ahmed M. N.; Reyna-Gonzalez E.; Schmid B.; Wiebach V.; Süssmuth R. D.; Dittmann E.; Fewer D. P. Phylogenomic analysis of the microviridin biosynthetic pathway coupled with targeted chemo-enzymatic synthesis yields potent protease inhibitors. ACS Chem. Biol. 2017, 12 (6), 1538–1546. 10.1021/acschembio.7b00124. [DOI] [PubMed] [Google Scholar]
  183. Makarova K. S.; Blackburne B.; Wolf Y. I.; Nikolskaya A.; Karamycheva S.; Espinoza M.; Barry C. E. 3rd; Bewley C. A.; Koonin E. V. Phylogenomic analysis of the diversity of graspetides and proteins involved in their biosynthesis. Biol. Direct 2022, 17 (1), 7 10.1186/s13062-022-00320-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Lee H.; Park Y.; Kim S. Enzymatic cross-linking of side chains generates a modified peptide with four hairpin-like bicyclic repeats. Biochemistry 2017, 56 (37), 4927–4930. 10.1021/acs.biochem.7b00808. [DOI] [PubMed] [Google Scholar]
  185. Clark K. A.; Bushin L. B.; Seyedsayamdost M. R. Aliphatic ether bond formation expands the scope of radical SAM enzymes in natural product biosynthesis. J. Am. Chem. Soc. 2019, 141 (27), 10610–10615. 10.1021/jacs.9b05151. [DOI] [PubMed] [Google Scholar]
  186. Flühe L.; Knappe T. A.; Gattner M. J.; Schäfer A.; Burghaus O.; Linne U.; Marahiel M. A. The radical SAM enzyme AlbA catalyzes thioether bond formation in subtilosin A. Nat. Chem. Biol. 2012, 8 (4), 350–357. 10.1038/nchembio.798. [DOI] [PubMed] [Google Scholar]
  187. Ishida K.; Nakamura A.; Kojima S. Crystal structure of the AlbEF complex involved in subtilosin A biosynthesis. Structure 2022, 30 (12), 1637–1646. 10.1016/j.str.2022.10.002. [DOI] [PubMed] [Google Scholar]
  188. Rued B. E.; Covington B. C.; Bushin L. B.; Szewczyk G.; Laczkovich I.; Seyedsayamdost M. R.; Federle M. J. Quorum sensing in Streptococcus mutans regulates production of Tryglysin, a novel RaS-RiPP antimicrobial compound. mBio 2021, 12 (2), e02688–20. 10.1128/mBio.02688-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Wever W. J.; Bogart J. W.; Baccile J. A.; Chan A. N.; Schroeder F. C.; Bowers A. A. Chemoenzymatic synthesis of thiazolyl peptide natural products featuring an enzyme-catalyzed formal [4 + 2] cycloaddition. J. Am. Chem. Soc. 2015, 137 (10), 3494–7. 10.1021/jacs.5b00940. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Hudson G. A.; Zhang Z.; Tietz J. I.; Mitchell D. A.; van der Donk W. A. In vitro biosynthesis of the core scaffold of the thiopeptide thiomuracin. J. Am. Chem. Soc. 2015, 137 (51), 16012–16015. 10.1021/jacs.5b10194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. Elashal H. E.; Koos J. D.; Cheung-Lee W. L.; Choi B.; Cao L.; Richardson M. A.; White H. L.; Link A. J. Biosynthesis and characterization of fuscimiditide, an aspartimidylated graspetide. Nat. Chem. 2022, 14 (11), 1325–1334. 10.1038/s41557-022-01022-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Ma S.; Zhang Q. Linaridin natural products. Nat. Prod. Rep. 2020, 37 (9), 1152–1163. 10.1039/C9NP00074G. [DOI] [PubMed] [Google Scholar]
  193. Schramma K. R.; Bushin L. B.; Seyedsayamdost M. R. Structure and biosynthesis of a macrocyclic peptide containing an unprecedented lysine-to-tryptophan crosslink. Nat. Chem. 2015, 7 (5), 431–7. 10.1038/nchem.2237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Kenney G. E.; Rosenzweig A. C. Chalkophores. Annu. Rev. Biochem. 2018, 87, 645–676. 10.1146/annurev-biochem-062917-012300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Hudson G. A.; Burkhart B. J.; DiCaprio A. J.; Schwalen C. J.; Kille B.; Pogorelov T. V.; Mitchell D. A. Bioinformatic mapping of radical S-adenosylmethionine-dependent ribosomally synthesized and post-translationally modified peptides identifies new Cα, Cβ, and Cγ-linked thioether-containing peptides. J. Am. Chem. Soc. 2019, 141 (20), 8228–8238. 10.1021/jacs.9b01519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Caruso A.; Martinie R. J.; Bushin L. B.; Seyedsayamdost M. R. Macrocyclization via an arginine-tyrosine crosslink broadens the reaction scope of radical S-adenosylmethionine enzymes. J. Am. Chem. Soc. 2019, 141 (42), 16610–16614. 10.1021/jacs.9b09210. [DOI] [PubMed] [Google Scholar]
  197. Sugiyama R.; Suarez A. F. L.; Morishita Y.; Nguyen T. Q. N.; Tooh Y. W.; Roslan M. N. H. B.; Lo Choy J.; Su Q.; Goh W. Y.; Gunawan G. A.; Wong F. T.; Morinaka B. I. The biosynthetic landscape of triceptides reveals radical SAM enzymes that catalyze cyclophane formation on Tyr-and His-containing motifs. J. Am. Chem. Soc. 2022, 144 (26), 11580–11593. 10.1021/jacs.2c00521. [DOI] [PubMed] [Google Scholar]
  198. Nguyen T. Q. N.; Tooh Y. W.; Sugiyama R.; Nguyen T. P. D.; Purushothaman M.; Leow L. C.; Hanif K.; Yong R. H. S.; Agatha I.; Winnerdy F. R.; Gugger M.; Phan A. T.; Morinaka B. I. Post-translational formation of strained cyclophanes in bacteria. Nat. Chem. 2020, 12 (11), 1042–1053. 10.1038/s41557-020-0519-z. [DOI] [PubMed] [Google Scholar]
  199. Phan C. S.; Morinaka B. I. A Prevalent Group of Actinobacterial Radical SAM/SPASM Maturases Involved in Triceptide Biosynthesis. ACS Chem. Biol. 2022, 17 (12), 3284–3289. 10.1021/acschembio.2c00621. [DOI] [PubMed] [Google Scholar]
  200. Kostenko A.; Lien Y.; Mendauletova A.; Ngendahimana T.; Novitskiy I. M.; Eaton S. S.; Latham J. A. Identification of a poly-cyclopropylglycine-containing peptide via bioinformatic mapping of radical S-adenosylmethionine enzymes. J. Biol. Chem. 2022, 298 (5), 101881 10.1016/j.jbc.2022.101881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Gabrielsen C.; Brede D. A.; Nes I. F.; Diep D. B. Circular bacteriocins: biosynthesis and mode of action. Appl. Environ. Microbiol. 2014, 80 (22), 6854–6862. 10.1128/AEM.02284-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. Perez R. H.; Zendo T.; Sonomoto K. Circular and leaderless bacteriocins: biosynthesis, mode of action, applications, and prospects. Front. Microbiol. 2018, 9, n/a. 10.3389/fmicb.2018.02085. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from ACS Bio & Med Chem Au are provided here courtesy of American Chemical Society

RESOURCES