Skip to main content
Applied and Environmental Microbiology logoLink to Applied and Environmental Microbiology
. 2022 Nov 23;88(23):e01617-22. doi: 10.1128/aem.01617-22

Incorporation of Non-Canonical Amino Acids into Antimicrobial Peptides: Advances, Challenges, and Perspectives

Yuhui Du a,b,#, Li Li c,#, Yue Zheng d, Jiaheng Liu c, Julia Gong e, Zekai Qiu d, Yanni Li d, Jianjun Qiao d,, Yi-Xin Huo a,
Editor: Charles M Dozoisf
PMCID: PMC9746297  PMID: 36416555

ABSTRACT

The emergence of antimicrobial resistance is a global health concern and calls for the development of novel antibiotic agents. Antimicrobial peptides seem to be promising candidates due to their diverse sources, mechanisms of action, and physicochemical characteristics, as well as the relatively low emergence of resistance. The incorporation of noncanonical amino acids into antimicrobial peptides could effectively improve their physicochemical and pharmacological diversity. Recently, various antimicrobial peptides variants with improved or novel properties have been produced by the incorporation of single and multiple distinct noncanonical amino acids. In this review, we summarize strategies for the incorporation of noncanonical amino acids into antimicrobial peptides, as well as their features and suitabilities. Recent applications of noncanonical amino acid incorporation into antimicrobial peptides are also presented. Finally, we discuss the related challenges and prospects.

KEYWORDS: antimicrobial peptides, noncanonical amino acids, antimicrobial resistance

INTRODUCTION

The misuse of traditional antibiotics and slow development of new antibiotics has resulted in antimicrobial resistance (AMR). It has been predicted that, without urgent intervention in controlling resistance, the estimated global deaths resulting from AMR would reach 10 million per year by 2050 and cost the world up to 100 trillion USD (1). It is a global health and economic issue which requires novel antimicrobial drugs with different killing mechanisms (2). Antimicrobial peptides (AMPs) are linear or cyclic peptides consisting of 10 to 100 amino acid residues (35). AMPs have been discovered in most life forms, making their classification a non-trivial task (6). AMPs can be grouped according to their source organisms, such as bacteriocins, fungal peptide antibiotics, plant thionins and defensins, insect defensins and cecropins, amphibian magainins and temporins, and defensins from higher vertebrates (79). They can also be classified based on their biosynthetic mechanisms, e.g., ribosomal synthesized or nonribosomal synthesized AMPs. Nevertheless, AMPs are most commonly categorized into four classes based on their secondary structures, i.e., linear α-helical peptides, β-sheet-containing peptides, peptides involving α- and β-elements, and linear extended structures (10). As of 13 August 2022, 6,032 natural and synthetic AMPs have been reported by the DRAMP (Data Repository of Antimicrobial Peptides, http://dramp.cpu-bioinfor.org/) database, and 77 AMPs have been developed as drug candidates (preclinical or clinical stage) (11).

Their mechanisms of action are variable, including disruption of cell membranes, inhibition of metabolic and translational elements, formation of nanonets, and stimulation of antimicrobial immune activity (1216). The stereotypical mechanism of AMPs is interacting with and disrupting bacterial cell membranes. Bacterial cytoplasmic membranes are rich in anionic groups, such as the phospholipids phosphatidylglycerol, cardiolipin, and phosphatidylserine. Lipopolysaccharides (LPS) in the outer membranes of Gram-negative bacteria and teichoic acids in the cell walls of Gram-positive bacteria also provide electronegative charges to the bacterial surface. Most AMPs exhibit either a cationic, hydrophobic, or amphipathic nature, which facilitates interaction with the negatively charged bacterial membrane (1720).This interaction results in perturbation and disintegration of the cell membrane, leading to the loss of nutrients and cell death. The development of resistance against AMPs was initially considered unlikely due to their rapid bactericidal effects, multiple potential targets, and avoidance of SOS or rpoS bacterial stress pathways (2125). This is especially the case for AMPs such as teixobactin, which was recently isolated from the undescribed soil microorganism Eleftheria terrae and specifically targets conserved substrates (e.g., lipids II and III [26]), because it is relatively difficult to alter the membrane without harming its function and integrity (27, 28). Although some studies have reported resistance mechanisms against AMPs, including charge modification of the membrane surface, secretion of proteases, and enhanced activity of efflux pumps (12), AMPs remain promising candidates to overcome the increasing AMR crisis due to their broad antimicrobial activities and the decline in discoveries of new classes of antibiotics (29, 30).

Noncanonical amino acids (ncAAs) are amino acids beyond the 20 canonical amino acids (cAAs), also named unnatural amino acids (uAAs), nonstandard amino acids (nsAAs), and non-natural amino acids (nAAs) (31). The majority of ncAAs are precursors, analogues, or metabolic intermediates of cAAs, and they are widely found in nature. ncAAs can contain special side chains which are chemically and structurally distinctive from those of cAAs. The past decades have witnessed rapid growth in the diversity and scope of ncAA application (3234). The incorporation of ncAAs can introduce novel characteristics into proteins, including alteration of enzyme catalytic activity and specificity (35, 36), functional assays of enzymes (37), and the development of novel nanomaterials (38). Moreover, some ncAAs can be utilized in cosmetics, such as l-homoserine, N-hydroxyglycine, and N-hydroxyserine, which exhibit higher water uptake compared to conventional moisturizing additives (38). ncAAs have also been introduced into AMPs. The ncAA-containing AMPs have several advantages over natural AMPs, such as bioavailability, organism selectivity, metabolic stability, overall toxicity, and control of charge distribution (3941).

Here, we describe the basis of common approaches to ncAA incorporation and review the progress toward ncAA-containing modification of AMPs, with a primary focus on exciting advances within the past 5 years. For earlier studies, readers are referred to Baumann et al. (42) and Budisa (43). Furthermore, the choice of incorporation strategies is discussed. The challenges of ncAA incorporation into AMPs and the emerging applications of ncAA-containing AMPs are also presented.

METHODS FOR NCAAS INCORPORATION INTO AMPS

Solid-phase peptide synthesis.

To date, solid-phase peptide synthesis (SPPS) is the major chemical strategy for peptide synthesis (4446). It is a stepwise assembly of a peptide chain attached to an insoluble resin support (Fig. 1A). Peptide chain synthesis starts from the carboxyl end of the N terminus of the peptide. The incoming amino acid is protected, commonly by 9-fluorenylmethoxycarbonyl (Fmoc) or tert-butyloxycarbonyl (tBoc), at the N terminus, and is coupled with the preceding chain after activation of the C terminus (47). Then, the N terminus is deblocked and the synthesis process goes to the next cycle. After full assembly, the peptide is cleaved from the resin by treatment with trifluoroacetic acid (TFA), which also removes the side chain-protecting groups. The Fmoc and tBoc groups can be removed by piperidine and diluted TFA, respectively. The crude linear peptide can be purified by one-step high-pressure liquid chromatography (47, 48). A series of AMPs containing ncAAs has been synthesized using SPPS, generating broad structural and functional diversities (4851).

FIG 1.

FIG 1

Schemes of four noncanonical amino acid (ncAA) incorporation approaches. (A) Solid-phase peptide synthesis (SPPS). Fmoc, 9-fluorenylmethoxycarbonyl; tBoc, tert-butyloxycarbonyl; TFA, trifluoroacetic acid. (B) Cell-free protein synthesis (CFPS). aaRS, aminoacyl-tRNA synthetase. (C) Selective pressure incorporation (SPI). Blue and gray circles represent 20 canonical amino acids (cAAs). Blue circle with star represents specific cAA analog. (D) Genetic code expansion (GCE).

Chemical synthesis of peptides is limited by synthesis length and speed. SPPS of longer peptides (>50 amino acids) is a challenge due to inefficient coupling and side reactions. Furthermore, synthesis speeds of minutes to hours per amide bond make SPPS a time-consuming strategy. The development of automated flow-based approaches highly improved the efficiency of SPPS, allowing for the chemical synthesis of long peptides (up to 164 amino acids) with a cycle time of 2.5 min per amino acid (52, 53). However, SPPS could not cover the chemical complexity of some AMPs, such as lanthipeptides (also called lantibiotics for those which display antimicrobial activity), a major group of ribosomally synthesized and post-translationally modified peptides (RiPPs) produced by microorganisms, because of the difficulty of chemically mimicking post-translational modification (PTM) (32) (Table 1).

TABLE 1.

Comparison of ncAA incorporation approachesa

Approach Advantages Disadvantages
SPPS Well-established and standardized Time-consuming
Complete control of peptide assembly Limited PTM systems
Multi-type and multi-site incorporation of ncAAs High cost of resins, ncAAs, and coupling reagents
High-throughput synthesis is available Low efficiency for synthesis of long peptides
High yield and purity
CFPS Well-established systems such as E. coli Limited PTM systems in certain platforms
Simple and fast lysate prepn procedure Limited platforms
Simple modification of the reaction system Development of a new platform is laborious and technically demanding
Eliminate the cytotoxicity of AMPs
Easy to scale up
Available for nonproteinogenic AMPs
Combination with display technologies
SPI Fast incorporation of cAA derivatives at multiple sites Demand for auxotrophic host
Simple procedure Cytotoxicity resulting in changes to the whole proteome
Laborious purification
Limited types of ncAAs, only cAA derivatives available
GCE Site-specific incorporation of ncAAs Low efficiency of incorporation
Well-established biotechniques in typical hosts Identification and modification of suitable orthogonal aaRS-tRNA pairs
Coexpression of valuable products is available Laborious purification
Reduced cost through coexpression of synthetic pathways of ncAAs Heterologous expression of PTM enzymes and immune apparatus
a

ncAA, noncanonical amino acid; SPPS, solid-phase peptide synthesis; PTM, post-translational modification; CFPS, cell-free protein synthesis; AMP, antimicrobial peptide; SPI, selective pressure incorporation; cAA, canonical amino acid; GCE, genetic code expansion; aaRS, aminoacyl-tRNA synthetase.

Cell-free protein synthesis.

In some cases, ncAA incorporation in vivo causes cytotoxicity or poor delivery of ncAAs (33). Cell-free protein synthesis (CFPS) has become a promising strategy for producing toxic and membrane proteins which are difficult to express. CFPS was first obtained from Escherichia coli to decipher its genetic code (54) and has allowed successful incorporation of ncAAs into proteins and peptides (55). It is beneficial due to elimination of the cellular barrier, which enables direct access to enzymes, amino acids, and other reaction conditions. CFPS begins with the preparation of crude extracts from chassis strains. After depletion of endogenous DNA and mRNA, the energy sources, cofactors, and other components are added to the lysate to mimic the bacterial environment. Then, a suitable template is added to initiate the transcription and translation reaction at an appropriate temperature (56) (Fig. 1B). A wide range of CFPSs have been reported, including E. coli, Bacillus subtilis, Archaea, fungi, plants, insects, and mammals (56, 57). At present, E. coli without release factor 1 (RF1) is the most efficient platform for ncAA incorporation.

Apparently, this approach is suitable for ribosomally synthesized AMPs. In the natural translation process, an amino acid is charged to the 3′ end of a tRNA by aminoacyl-tRNA synthetase (aaRS) using a specific anticodon. For the nonproteinogenic AMPs, a combination of flexizymes and genetic code reprogramming system, namely, the flexible in vitro translation (FIT) system, was conducted (58). Flexizymes are small (~45 nucleotides) ribozyme aaRSs which facilitate the aminoacylation of tRNAs. Unlike aaRSs, flexizymes can recognize the 3′ end of tRNA without depending on anticodon sequences. Three flexizymes (dFx, eFx, and aFx) have been developed (59). Simple mixing of flexizymes with the desired tRNAs and acid substrates, followed by a few hours’ incubation on ice, can lead to the charged tRNAs.

One obvious advantage of CFPS is its speed. The process of CFPS is faster than that of in vivo protein synthesis because it has no need for cloning steps. Furthermore, CFPS is useful for the incorporation of toxic ncAAs (60). Another advantage of CFPS is the open feature which makes the chemical environment easy to modify. Moreover, the barrier-free system allows lower concentrations of ncAAs, which significantly reduces costs (61). In the envelope-free system, no signal peptides/leader sequences or transporter systems are needed to secrete the mature peptide (62). In some cases, the AMPs’ antibacterial nature makes them potentially fatal to the chassis strains. Fusion strategies with a partner protein (63, 64) or coexpression of immunity proteins (65, 66) were developed to neutralize this toxicity. CFPS, on the other hand, bypasses this limitation. In addition, as a homogeneous catalyzed reaction, CFPS is relatively simple to scale up (67). However, the application of CFPS is limited by finite platforms. Although many platforms have been reported, the E. coli system is most commonly used for ncAA incorporation (68, 69). It is economical and relatively simple to prepare. Recent efforts have further improved the efficiency of CFPS using extracts from a genetically modified E. coli strain, C321.ΔA, in which 321 TAG amber codons were recoded to TAA ochre codons and release factor 1 (RF1) was deleted (55). Nonetheless, E. coli has limited PTMs, which leads to immature or nonactive AMPs. Recently, CFPS has been shown to be successfully applied to AMPs belonging to the lasso peptide class, a class of RiPPs with a topology resembling a threaded lasso or a slipknot, which are very difficult to obtain and impossible to synthesize (70, 71). This allowed the production of a huge amount of sequence-diverse variants affording the proof of concept of using CFPS to create large libraries of difficult RiPPs.

Selective pressure incorporation.

In vivo incorporation of ncAAs has been a topic of interest. One method for in vivo incorporation of ncAAs is selective pressure incorporation (SPI), which relies on competitive incorporation. The endogenous aaRS of cAAs has substrate promiscuity. It can charge structural and chemical analogs to the cognate tRNA and leads to ribosomal incorporation of the ncAAs. SPI relies on the use of an auxotrophic host, whose growth requires the addition of particular amino acids. The auxotrophic host strains are first cultured in defined medium supplied with a limited amount of cAA to be substituted. After depletion of the corresponding cAA, the ncAA is added. Then, production of the target peptide/protein is induced (Fig. 1C). Several studies have been reported to successfully incorporate ncAAs into AMPs using SPI (7275).

SPI is a residue-specific approach which can easily achieve multi-site incorporation. With the appropriate auxotroph strain, SPI can install ncAAs into AMPs via an endogenous translation apparatus in a relatively simple protocol. However, it is only suitable for ncAAs which share structural and chemical similarities with cAAs, which limits the applicable scope of this method. Furthermore, SPI inevitably results in a proteomic change. Although the enzymes produced during early growth can maintain the growth of the host and expression of the target peptide, SPI is not suitable for AMPs which take a long time to synthesize.

Genetic code expansion.

Another in vivo approach for incorporation of ncAAs is genetic code expansion (GCE), which is in a site-specific manner. This strategy is similar to the natural translation process. An aaRS charges an ncAA onto the tRNA with a specific anticodon. Then, the charged tRNA is delivered to the ribosome, leading to the incorporation of the corresponding amino acid (Fig. 1D). GCE relies on the use of an orthogonal aaRS-tRNA pair. Four aaRS-tRNA pairs are most common used for ncAA incorporation: the tyrosyl-tRNA synthetase (MjTyrRS)-tRNACUA pair from Methanococcus jannaschii, tyrosyl-tRNA synthetase (EcTyrRS)-tRNACUA pair from E. coli, leucyl-tRNA synthetase (EcLeuRS)-tRNACUA pair from E. coli, and pyrrolysyl-tRNA synthetase (PylRS)-tRNACUA pair from Methanosarcina spp. (76). Each of these orthogonal pairs can be used in certain organisms. To date, the suppression of nonsense codons, usually the amber codon TAG, has been successfully applied to multiple organisms (7782). Although stop-codon suppression (SCS) has become increasingly sophisticated, its competition with RF1-mediated termination of peptide chain and low heterogeneous enzyme activity result in low integration efficiency. Several studies have been performed on improving its efficiency, such as elimination of RF1 (83, 84), directed evolution of aaRSs (85, 86) and ribosome (87), optimization of the orthogonal system (88), and modification of elongation factor Tu (89). A recent study reported a genetically recoded E. coli C321.ΔA in which all 321 amber codons were recoded to ochre codons and RF1 was deleted (90). Two synonymous serine codons were further removed to create a new chassis using 59 codons to encode the 20 cAAs (91). The reassignment of sense codons allows incorporation of multiple ncAAs. In addition, quadruplet codons have been utilized to encode ncAAs, which, in principle, might provide 256 bland codons (92, 93). Rodriguez et al. (94) successfully introduced multiple ncAAs in a neuroreceptor expressed in vivo through stop codon and quadruplet codon suppression. Quadruplet codons have also been applied to Caenorhabditis elegans, paving the way toward in vivo multi-incorporation of ncAAs in a multicellular organism (95). Over the past 2 decades, GCE has been extensively used to incorporate ncAAs into AMPs. For example, four ncAAs have been used to replace four positions in the lasso peptide microcin J25, a 21-residue RiPP produced by E. coli strains (96), via SCS (97). Using this approach, ncAAs have also been incorporated into lanthipeptides (78, 98), thiopeptides (a group of sulfur-containing macrocyclic peptides [81, 99]), and cyanobactins (a group of ribosomal cyclic peptides produced by cyanobacteria [100, 101]). These efforts led to a batch of AMP derivatives with novel chemical and/or functional activities.

As described above, the incorporation efficiency of GCE is dependent on several parameters and requires specific genetic engineering of the host strain. At present, E. coli is the most investigated platform for GCE, and is available for heterologous expression of ncAA-incorporating AMPs. However, it is still challenging because the corresponding PTM enzymes and immune proteins must be coexpressed. Moreover, some proteases in the E. coli cytoplasm can degrade heterologous proteins, which may impede proper folding and biological activity of the proteins (102, 103). A recent study reported the incomplete dehydration of the heterologously expressed NisA in E. coli BL21(DE3). The endonuclease rne and the proteases ompT and lon were deleted to ensure the complete modification of NisA (104). Because AMPs have been successfully heterologously expressed in different host strains, such as Lactococcus, Bacillus, and yeast, some of which showed better AMP expression than E. coli (105107), it is necessary to develop new chassis, based on the principles of GCE, for suitable bioproduction of ncAA-incorporating AMPs. In addition, the in vivo GCE approaches are only available for non-toxic ribosomal synthesized peptides. To address this limitation, in vitro platforms combining CFPS and GCE have been developed (108).

RECENT DEVELOPMENTS OF NCAAS INCORPORATION INTO AMPS

Expansion of the physicochemical diversity.

In nature, the diversity of AMPs is limited due to the conservative set of amino acids and modification systems. With the development of ncAA incorporation strategies, a large number of studies have been performed to expand the repertoire of AMPs (Table 2). One example is the incorporation of the α-chloroacetamide-containing ncAA into nisin A, a well-known lanthibiotic, which resulted in novel macrocyclic topologies. Although no variants retained antimicrobial activity, this strategy could be extended to other ring junctions for structural diversification (109). Furthermore, ncAAs have also been successfully incorporated into other AMPs, such as lasso peptides (110), macrocyclic peptides (111), thiopeptides (81), and the tetracyclic polypeptide cinnamycin (112).

TABLE 2.

Recent applications of ncAA incorporation into AMPsa

Application and AMP Class(es) Strategy Modifications and consequences Reference
Expansion of physicochemical diversity
 Nisin A Lanthipeptide isolated from bacteria SPI Incorporation of 6 proline analogs into nisin A; obtained 5 variants with retained bioactivity. 75
GCE Incorporation of several ncAAs, including an α-chloroacetamide-containing ncAA which enabled nisin A variants with novel macrocyclic topologies. 109
 Macrocyclic peptide Macrocyclic α/β-peptide foldamer CFPS/FIT Incorporation of consecutive stereoisomeric cyclic β2,3-amino acids and construction of a macrocyclic peptide library. 115
Exotic macrocyclic peptide CFPS/FIT Assembly of a macrocyclic peptide library with 28 ncAAs, obtained at least 2 high-affinity ligands of IL-6. 175
Alteration/enhancement of bioactivity or stability
 Amphipathic peptidomimetic Peptidomimetic SPPS Synthesis of several amphipathic peptidomimetics containing Fmoc-triazine amino acids; obtention of a trimer, BJK-4, with enhanced proteolytic stability and no hemolytic activity. 49
 Pep05 Histatin from human saliva SPPS Incorporation of d-AA and ncAAs enhanced stability toward proteases. 50
 Coralmycin Nonproteinogenic acyl pentapeptide SPPS Incorporation of an unusual 4-amino-2-hydroxy-3isopropoxybenzoic acid motif, obtained a desmethoxy analog with higher potent antimicrobial activity. 121
 Cinnamycin Lanthipeptide isolated from bacteria GCE Incorporation of 3 pyrrolysine analogues into cinnamycin, obtention of 5 deoxycinnamycin derivatives. Incorporation of H-Lys-Alloc-OH (Alk) in position 2 increased bioactivity. 112
 Nisin A and lacticin 481 Lanthipeptide isolated from bacteria GCE Incorporation of 4 phenylalanine derivatives at 3 positions of lacticin 481, and 2 phenylalanine derivatives at one position of nisin A. Five lacticin 481 variants were obtained with slightly higher activities. 104
 HPA3NT3-A2 Cecropin-like bactericidal peptide isolated from insects SPPS Replacement of l-Lys with d-enantiomer increased stability in serum. 137
 W3R6 Chensinin-1 family peptide isolated from amphibian skin SPPS Replacement of l-Arg with d-enantiomer increased resistance to proteases, but slightly reduced antimicrobial activity. 176
 Polybia-CP Novel peptide isolated from wasp venom SPPS Replacement of all the l-amino acids with d-enantiomers increased resistance to trypsin and chymotrypsin; partial d-Lys-substituted derivative was resistant to trypsin only. 177
 P-113 Histatin from human saliva SPPS Incorporation of ncAAs into P-113 altered the antimicrobial activity, bacterial membrane permeability, antibiotic synergism, and supernatant LPS-neutralizing activities. 178
 Anti-Zika peptides Macrocyclic peptide SPPS Incorporation of 3-(2-cyano-4-pyridyl)-alanine facilitated the condensation reaction with cysteine, resulting in a cyclized anti-Zika peptide with high affinity and proteolytic stability. 179
AMP screening through in vivo fluorescent imaging
 Nisin A Lanthipeptide isolated from bacteria SPI Incorporation of 4 methionine analogues with various chemical handles, allowing “click” reactions for fluorescent label. Some variants displayed increased selectivity. 73
 LL37 and melittin Cathelicidin isolated from human, peptide isolated from bee venom SPPS Incorporation of an azido-lysine at C terminus of LL37 and melittin enabled single-molecule imaging of AMP-bacterium interactions. 163
 LL37, cecropin B, and melittin Cathelicidin isolated from human, cecropin isolated from insects, peptide isolated from bee venom SPPS Incorporation of an azido-lysine to measure the affinity of 3 AMPs to lipid A extracted from 4 Gram-negative bacteria. 162
 BODIPY-cPAF26 Rationally designed cyclic fluorogenic peptide SPPS Incorporation of a Fmoc-Trp (C2-BODIPY)-OH amino acid into AMP; real-time imaging of Aspergillus fumigatus. 180
 Apo-15 Rationally designed cyclic amphipathic peptide SPPS Incorporation of a Trp-BODIPY fluorophore into Apo-15 allowed quantification and imaging of drug-induced apoptosis. 181
 Teixobactin Novel RiPP isolated from E. terrae SPPS Incorporation of an amine-reactive rhodamine fluorophore into teixobactin enabled fluorescent imaging of Gram-positive bacteria. 182
a

ncAA, noncanonical amino acid; AMP, antimicrobial peptide; SPI, selective pressure incorporation; GCE, genetic code expansion; CFPS, cell-free protein synthesis; FIT, flexible in vitro translation; SPPS, solid-phase peptide synthesis; LPS, lipopolysaccharide; RiPP; post-translationally modified peptide.

The development of the FIT system expands the generation of numerous peptide analogues with diverse structural features. In particular, integration of the FIT system and display technologies has been used for the high-throughput discovery of ligands with high target affinities, especially for macrocyclic peptides (113, 114). For example, Katoh et al. (115) reported the incorporation of consecutive derivatives from 2-aminocyclohexanecarboxylic acid and 2-aminocyclopentanecarboxylic acid and construction of a macrocyclic peptide library containing 1012 members via the FIT system. By using the random nonstandard peptide integrated discovery (RaPID) system, they identified a variation in mRNA display against human factor XIIa and interferon gamma receptor 1, an anti-hFXIIa macrocyclic peptide with high inhibitory activity and serum stability.

Alteration/enhancement of bioactivities.

The potential for alternatives to antibiotics has always been the most popular topic concerning AMPs. A large number of studies have been performed to alter/enhance their antimicrobial activity, selectivity, and antibacterial spectrum (Table 2). Previous reports have shown that modification at certain sites with certain types of amino acids could alter the bioactivity of AMPs (116). For example, saturation mutagenesis of the lantibiotics nisin A, mersacidin, and nukacin allowed access to several derivatives with enhanced activity (117119). Furthermore, site-directed mutagenesis of nisin Z generated two mutants with enhanced activity against Gram-negative bacteria (120). Based on this preliminary work, researchers have been committed to incorporating ncAAs into AMPs (32, 42). Recently, Deng et al. (73) incorporated four Met analogs into four sites of nisin A. Several variants showed improved activity against a specific target strain with reduced activity against other strains, suggesting increased antimicrobial selectivity. Another example is coralmycin, a nonproteinogenic acyl pentapeptide produced by the myxobacterium Corallococcus coralloides and exhibiting potent activity against a range of Gram-negative bacteria via inhibition of DNA gyrase (121). The incorporation of the unusual 4-amino-2hydroxy-3isopropoxybenzoic acid motif enabled the variant to have higher potent bioactivity.

As previously mentioned, the type of amino acid is an important factor in AMP modification. The charges of naturally identified AMPs vary from 0 to +20, facilitating interaction with negatively charged bacterial membranes (122). Studies have shown that enhancing cationicity without drastically altering other physicochemical parameters increases the antimicrobial activity and spectrum (123, 124). Almaaytah et al. (125) designed a scorpion venom-derived analog (A3) with enhanced cationicity. Because bacterial membranes display a negative charge while normal mammalian cells are usually zwitterionic, the derived peptide exhibited increased antimicrobial activity and reduced cytotoxicity due to its enhanced selectivity against bacterial membranes. Among the 20 cAAs, lysine and arginine are commonly used for AMP modification due to their cationicity (126). Besides the net charge, the hydrophobicity of AMPs is also a key factor in bioactivity (127, 128). The incorporating site is another factor. For example, consider the well-studied AMP nisin A. Nisin A is composed of five lanthionine rings and a hinge region which is located between the first three and the last two rings (129). Previous studies have shown that the hinge region is a more promising site for alteration of bioactivity (130). However, for AMPs with few fundamental studies, the biosynthesis pathways, structures, and antibiotic mechanisms still need further exploration.

Improvement of proteolytic resistance.

Another area of research interest involving AMP modification is increasing their proteolytic resistance. Susceptibility to host and microbial proteolytic degradation is a great challenge in transforming AMPs into therapeutics. Many efforts have been taken to improve the proteolytic resistance of AMPs, such as encapsulation, pegylation, acetylation, amidation, cyclization, and the substitution of ncAAs (131, 132). Among these, ncAA substitution (including d-amino acids) and cyclization via reactive side chains of ncAAs are potential strategies to improving the stability of AMPs.

Replacement of some l-amino acids with d-enantiomer residues is a common strategy to increase proteolytic resistance, as d-residues are not recognized by proteases which only catalyze bonds formed by l-residues. For example, MSI-238, an all-d analogue of the amphibian AMP magainin 2 isolated from the African clawed frog (Xenopus laevis), exhibited increased proteolytic resistance (133135). Another example is the modification of HPA3NT3-A2, an analog of the cecropin-like bactericidal peptide HP (2-20) derived from Helicobacter pylori; cecropin is a positively charged AMP originally isolated from insects and cecropia moths (136). Substitution of l-lysine in HPA3NT3-A2 with d-residues increased its half-life and maintained its antimicrobial activity (137). Likewise, incorporation of certain ncAAs can also enhance the stability of AMPs. Lu et al. (50) demonstrated that installation of some d- and unnatural amino acids could enhance the proteolytic stability of the peptides. Furthermore, the incorporation of a single alpha-isobutyric acid residue significantly increased plasma stability and activity. Another example is that of Gunasekaran et al. (49), who successfully synthesized a series of amphipathic peptidomimetics containing Fmoc-triazine amino acids and obtained a trimer, BJK-4, with enhanced proteolytic stability. Furthermore, incorporation of N-alkyl amino acids, α-substituted α-amino acids, β-substituted α-amino acids, β- (and γ-) amino acids, and proline analogues have also been reported to improve peptide stability (138). Generally speaking, decreasing peptide-like features leads to enhanced stability toward proteases.

Cyclization is another potential strategy to improve the proteolytic stability of AMPs (139). The most popular connections for peptide cyclization are lactam, lactone, and disulfide bonds. Other covalent connections, such as thioether bonds and carbon-carbon bonds, have also been used for peptide cyclization (140144). Cross-linking of amino acid side chains is an effective method to cyclize peptides. Cysteine, lysine, and tyrosine are widely used cAAs for such conjugation (145, 146). ncAAs with reactive side chains have also been used for cyclization of AMPs (147, 148) (Table 2). A recent study reported that cyclization of buforin II was facilitated by the introduction of aldehyde and hydrazide groups. A series of peptides with single-, double-, and triple-loop was obtained with increased activity (149). Cyclic efficiency may be affected by the residues around the cyclization site. Amino acids with more flexibility (such as Gly) and smaller space requirements are preferred (150). Cyclization may also influence the activity, selectivity, and hemolytic activity of AMPs (140, 141, 151153), although the relation is complex. In summary, the introduction of ncAAs provides increased potential for AMP cyclization, which is an effective strategy to improve the stability of linear AMPs. The effects of ring size, amphiphilicity, and peptide sequence, and the potential side effects, still require more investigation.

Screening of AMPs through in vivo fluorescent imaging.

Fluorescent imaging of AMPs is of great contribution in understanding their in vivo antimicrobial mechanisms, and requires precise recognition, localization, and quantification (154, 155). The common labeling method of fluorescent protein fusion may disturb the structure and activity of short peptides. Conjugation with fluorophores through click chemistry is a promising tool for in vivo molecule imaging due to its high specificity, quick reaction rate, and stability (156). Incorporation of ncAAs enables the site-specific installation of a suitable handle/arm, allowing bio-orthogonal labeling of AMPs by taking advantage of specific reactive groups (such as azides, alkynes, and alkenes) of ncAAs. A recent study by Deng et al. (73) successfully installed Met analogs with bio-orthogonal reactive handles into nisin A and enabled fluorescent detection. Advances in biomolecule imaging technology have now enabled the investigation of subcellular protein organization (157). Since the first reported single-molecule imaging of fluorescently labeled ncAA done by Pantoja et al. (158), super-resolution techniques utilizing ncAA incorporation have been applied to several proteins and AMPs (159163) (Table 2). These advances pave the way to investigating interactions between AMPs and pathogens at the molecular level.

In some cases, fluorescent labeling can also alter the conformation of peptide biological properties. To diminish this effect, some fluorescent ncAAs have been developed. Analogues of tryptophan, tyrosine, and phenylalanine are common candidates because they are themselves fluorescent. A range of non-natural fluorescent amino acids was also generated by either appending fluorescent moieties to amino acids or de novo construction with integrated chromophores. For detailed information on fluorescent amino acids, researchers are referred to a recent review by Cheng et al. (164).

Conclusion and future prospects.

Over the past few years, intensive research has been conducted to introduce novel properties into AMPs through ncAA installation (Table 2). This review provides a brief overview of common strategies for ncAA incorporation into AMPs and recent applications. Each of these technologies has different features and suitabilities (Table 1). At present, SPPS is still the primary approach for peptide synthesis, allowing multi-type and multi-site incorporation of ncAAs. In particular, SPPS is a more appropriate strategy for synthesizing toxic AMPs or ncAAs with ambiguous permeability and solubility. CFPS is another efficient platform to synthesize toxic AMPs. Specifically, with the development of flexizymes and display techniques, CFPS has a unique advantage in construction and screening of AMP libraries. However, with the available chassis, ncAAs, and/or orthogonal pairs, in vivo SPI or GCE is preferred due to their relatively simple procedures.

One important value of a promising biotechnology is the implementation of industrial application. Currently, ncAA modification of AMPs is more of a theoretical verification or mechanistic investigation. One obstacle is the high costs of most ncAAs. Although many ncAAs are commercially available, their cultural expenses are still high because a high concentration of ncAAs is required for efficient incorporation. In addition, many ncAAs are not yet commercialized and rely on chemical synthesis, resulting in even higher costs. Coupling GCE and certain biosynthetic pathways, through which ncAAs can be produced from inexpensive substrates, offers a promising way to reduce the cost of ncAA incorporation (165, 166). In addition, for industrial production of potent variants, a standardized and relatively simple protocol, inexpensive substrates, safety, and efficiency are essential. Over the past few decades, massive research has been done to synthesize AMPs through biological approaches. Microbial engineering has been done via (i) heterologous expression of orthogonal aaRS-tRNA pairs for ncAA incorporation (78, 104, 109); (ii) reprogramming the genetic code of the microbial chassis to improve incorporation efficiency (83, 84, 90, 91); (iii) directed evolution of the translational apparatus (8589); and (iv) mining and overexpressing the synthesis pathways of ncAAs to reduce production costs (167, 168). While the most-understood model organism E. coli has become an optimal chassis for heterologous expression, the engineering of the original production strains is still necessary, considering some characteristics of the original hosts. For example, we previously reported the co-production of nisin Z and γ-aminobutyric acid in the nisin Z production strain Lactococcus lactis, which exhibited a synergistic effect of food preservation. Because L. lactis is a probiotic strain, its freeze-dried product can be directly used, eliminating the process of purification (169). In addition, it also excludes the trouble of heterologous expression of multiple sets of synthetic and immune enzymes. Indeed, complete reprogramming of every host is an impractical task. It is a compromise strategy for prior modification of essential or growth-related genes which may reduce the effect of ncAA introduction. Furthermore, the identification of new orthogonal pairs which can recognize novel ncAAs is desired.

With the development of next-generation sequencing and computational tools, various algorithms have been established to discover novel AMPs (170172). It is foreseeable that more AMP-related synthesis pathways, physiochemical properties, and structures could be discovered. The machine-learning technique has been used to guide the optimization and designation of AMPs with ncAAs (173, 174). Given advances in computational system biology and structural biology, rational design and prediction of novel AMPs will become practical in the coming years.

ACKNOWLEDGMENTS

The present work was supported by the National Key R&D Program of China (2019YFA0904104) and the China Postdoctoral Science Foundation (2019M660475).

Contributor Information

Jianjun Qiao, Email: jianjunq@tju.edu.cn.

Yi-Xin Huo, Email: huoyixin@bit.edu.cn.

Charles M. Dozois, INRS

REFERENCES

  • 1.de Kraker ME, Stewardson AJ, Harbarth S. 2016. Will 10 million people die a year due to antimicrobial resistance by 2050? PLoS Med 13:e1002184. doi: 10.1371/journal.pmed.1002184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Ovchinnikov KV, Chi H, Mehmeti I, Holo H, Nes IF, Diep DB. 2016. Novel group of leaderless multipeptide bacteriocins from Gram-positive bacteria. Appl Environ Microbiol 82:5216–5224. doi: 10.1128/AEM.01094-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Wang Y, Wang M, Shan A, Feng X. 2020. Avian host defense cathelicidins: structure, expression, biological functions, and potential therapeutic applications. Poult Sci 99:6434–6445. doi: 10.1016/j.psj.2020.09.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Ageitos JM, Sánchez-Pérez A, Calo-Mata P, Villa TG. 2017. Antimicrobial peptides (AMPs): ancient compounds that represent novel weapons in the fight against bacteria. Biochem Pharmacol 133:117–138. doi: 10.1016/j.bcp.2016.09.018. [DOI] [PubMed] [Google Scholar]
  • 5.Lazzaro BP, Zasloff M, Rolff J. 2020. Antimicrobial peptides: application informed by evolution. Science 368:eaau5480. doi: 10.1126/science.aau5480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Zasloff M. 2002. Antimicrobial peptides of multicellular organisms. Nature 415:389–395. doi: 10.1038/415389a. [DOI] [PubMed] [Google Scholar]
  • 7.Sang Y, Blecha F. 2008. Antimicrobial peptides and bacteriocins: alternatives to traditional antibiotics. Anim Health Res Rev 9:227–235. doi: 10.1017/S1466252308001497. [DOI] [PubMed] [Google Scholar]
  • 8.Conlon BP, Nakayasu ES, Fleck LE, LaFleur MD, Isabella VM, Coleman K, Leonard SN, Smith RD, Adkins JN, Lewis K. 2013. Activated ClpP kills persisters and eradicates a chronic biofilm infection. Nature 503:365–370. doi: 10.1038/nature12790. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Andrä J, Berninghausen O, Leippe M. 2001. Cecropins, antibacterial peptides from insects and mammals, are potently fungicidal against Candida albicans. Med Microbiol Immunol 189:169–173. doi: 10.1007/s430-001-8025-x. [DOI] [PubMed] [Google Scholar]
  • 10.Koehbach J, Craik DJ. 2019. The vast structural diversity of antimicrobial peptides. Trends Pharmacol Sci 40:517–528. doi: 10.1016/j.tips.2019.04.012. [DOI] [PubMed] [Google Scholar]
  • 11.Kang X, Dong F, Shi C, Liu S, Sun J, Chen J, Li H, Xu H, Lao X, Zheng H. 2019. DRAMP 2.0, an updated data repository of antimicrobial peptides. Sci Data 6:148. doi: 10.1038/s41597-019-0154-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Bechinger B, Gorr SU. 2017. Antimicrobial peptides: mechanisms of action and resistance. J Dent Res 96:254–260. doi: 10.1177/0022034516679973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Florin T, Maracci C, Graf M, Karki P, Klepacki D, Berninghausen O, Beckmann R, Vázquez-Laslop N, Wilson DN, Rodnina MV, Mankin AS. 2017. An antimicrobial peptide that inhibits translation by trapping release factors on the ribosome. Nat Struct Mol Biol 24:752–757. doi: 10.1038/nsmb.3439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Gagnon MG, Roy RN, Lomakin IB, Florin T, Mankin AS, Steitz TA. 2016. Structures of proline-rich peptides bound to the ribosome reveal a common mechanism of protein synthesis inhibition. Nucleic Acids Res 44:2439–2450. doi: 10.1093/nar/gkw018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Loth K, Vergnes A, Barreto C, Voisin SN, Meudal H, Da Silva J, Bressan A, Belmadi N, Bachère E, Aucagne V, Cazevielle C, Marchandin H, Rosa RD, Bulet P, Touqui L, Delmas AF, Destoumieux-Garzón D. 2019. The ancestral N-terminal domain of big defensins drives bacterially triggered assembly into antimicrobial nanonets. mBio 10:e01821-19. doi: 10.1128/mBio.01821-19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Yang D, Biragyn A, Hoover DM, Lubkowski J, Oppenheim JJ. 2004. Multiple roles of antimicrobial defensins, cathelicidins, and eosinophil-derived neurotoxin in host defense. Annu Rev Immunol 22:181–215. doi: 10.1146/annurev.immunol.22.012703.104603. [DOI] [PubMed] [Google Scholar]
  • 17.Epand RM, Epand RF. 2009. Lipid domains in bacterial membranes and the action of antimicrobial agents. Biochim Biophys Acta 1788:289–294. doi: 10.1016/j.bbamem.2008.08.023. [DOI] [PubMed] [Google Scholar]
  • 18.Jenssen H, Hamill P, Hancock RE. 2006. Peptide antimicrobial agents. Clin Microbiol Rev 19:491–511. doi: 10.1128/CMR.00056-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Buda De Cesare G, Cristy SA, Garsin DA, Lorenz MC. 2020. Antimicrobial peptides: a new frontier in antifungal therapy. mBio 11:e02123-20. doi: 10.1128/mBio.02123-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Boparai JK, Sharma PK. 2020. Mini review on antimicrobial peptides, sources, mechanism and recent applications. Protein Pept Lett 27:4–16. doi: 10.2174/0929866526666190822165812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Mahlapuu M, Håkansson J, Ringstad L, Björn C. 2016. Antimicrobial peptides: an emerging category of therapeutic agents. Front Cell Infect Microbiol 6:194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Fjell CD, Hiss JA, Hancock RE, Schneider G. 2011. Designing antimicrobial peptides: form follows function. Nat Rev Drug Discov 11:37–51. doi: 10.1038/nrd3591. [DOI] [PubMed] [Google Scholar]
  • 23.Rodríguez-Rojas A, Makarova O, Rolff J. 2014. Antimicrobials, stress and mutagenesis. PLoS Pathog 10:e1004445. doi: 10.1371/journal.ppat.1004445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Spohn R, Daruka L, Lázár V, Martins A, Vidovics F, Grézal G, Méhi O, Kintses B, Számel M, Jangir PK, Csörgő B, Györkei Á, Bódi Z, Faragó A, Bodai L, Földesi I, Kata D, Maróti G, Pap B, Wirth R, Papp B, Pál C. 2019. Integrated evolutionary analysis reveals antimicrobial peptides with limited resistance. Nat Commun 10:4538. doi: 10.1038/s41467-019-12364-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Yu G, Baeder DY, Regoes RR, Rolff J. 2018. Predicting drug resistance evolution: insights from antimicrobial peptides and antibiotics. Proc R Soc B 285:20172687. doi: 10.1098/rspb.2017.2687. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Karas JA, Chen F, Schneider-Futschik EK, Kang Z, Hussein M, Swarbrick J, Hoyer D, Giltrap AM, Payne RJ, Li J, Velkov T. 2020. Synthesis and structure-activity relationships of teixobactin. Ann N Y Acad Sci 1459:86–105. doi: 10.1111/nyas.14282. [DOI] [PubMed] [Google Scholar]
  • 27.Nguyen LT, Haney EF, Vogel HJ. 2011. The expanding scope of antimicrobial peptide structures and their modes of action. Trends Biotechnol 29:464–472. doi: 10.1016/j.tibtech.2011.05.001. [DOI] [PubMed] [Google Scholar]
  • 28.Ling LL, Schneider T, Peoples AJ, Spoering AL, Engels I, Conlon BP, Mueller A, Schäberle TF, Hughes DE, Epstein S, Jones M, Lazarides L, Steadman VA, Cohen DR, Felix CR, Fetterman KA, Millett WP, Nitti AG, Zullo AM, Chen C, Lewis K. 2015. A new antibiotic kills pathogens without detectable resistance. Nature 517:455–459. doi: 10.1038/nature14098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Magana M, Pushpanathan M, Santos AL, Leanse L, Fernandez M, Ioannidis A, Giulianotti MA, Apidianakis Y, Bradfute S, Ferguson AL, Cherkasov A, Seleem MN, Pinilla C, de la Fuente-Nunez C, Lazaridis T, Dai T, Houghten RA, Hancock REW, Tegos GP. 2020. The value of antimicrobial peptides in the age of resistance. Lancet Infect Dis 20:e216–e230. doi: 10.1016/S1473-3099(20)30327-3. [DOI] [PubMed] [Google Scholar]
  • 30.Ting DSJ, Beuerman RW, Dua HS, Lakshminarayanan R, Mohammed I. 2020. Strategies in translating the therapeutic potentials of host defense peptides. Front Immunol 11:983. doi: 10.3389/fimmu.2020.00983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Rezhdo A, Islam M, Huang M, Van Deventer JA. 2019. Future prospects for noncanonical amino acids in biological therapeutics. Curr Opin Biotechnol 60:168–178. doi: 10.1016/j.copbio.2019.02.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Karbalaei-Heidari HR, Budisa N. 2020. Combating antimicrobial resistance with new-to-nature lanthipeptides created by genetic code expansion. Front Microbiol 11:590522. doi: 10.3389/fmicb.2020.590522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Jin X, Park OJ, Hong SH. 2019. Incorporation of non-standard amino acids into proteins: challenges, recent achievements, and emerging applications. Appl Microbiol Biotechnol 103:2947–2958. doi: 10.1007/s00253-019-09690-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Chin JW. 2017. Expanding and reprogramming the genetic code. Nature 550:53–60. doi: 10.1038/nature24031. [DOI] [PubMed] [Google Scholar]
  • 35.Poreba M, Salvesen GS, Drag M. 2017. Synthesis of a HyCoSuL peptide substrate library to dissect protease substrate specificity. Nat Protoc 12:2189–2214. doi: 10.1038/nprot.2017.091. [DOI] [PubMed] [Google Scholar]
  • 36.Tsuchiya K, Numata K. 2017. Chemoenzymatic synthesis of polypeptides containing the unnatural amino acid 2-aminoisobutyric acid. Chem Commun (Camb) 53:7318–7321. doi: 10.1039/c7cc03095a. [DOI] [PubMed] [Google Scholar]
  • 37.Yu Y, Cui C, Wang J, Lu Y. 2017. Biosynthetic approach to modeling and understanding metalloproteins using unnatural amino acids. Sci China Chem 60:188–200. doi: 10.1007/s11426-016-0343-2. [DOI] [Google Scholar]
  • 38.Clerici F, Erba E, Gelmi ML, Pellegrino S. 2016. Non-standard amino acids and peptides: from self-assembly to nanomaterials. Tetrahedron Lett 57:5540–5550. doi: 10.1016/j.tetlet.2016.11.022. [DOI] [Google Scholar]
  • 39.Bhonsle JB, Clark T, Bartolotti L, Hicks RP. 2013. A brief overview of antimicrobial peptides containing unnatural amino acids and ligand-based approaches for peptide ligands. Curr Top Med Chem 13:3205–3224. doi: 10.2174/15680266113136660226. [DOI] [PubMed] [Google Scholar]
  • 40.Venugopal D, Klapper D, Srouji AH, Bhonsle JB, Borschel R, Mueller A, Russell AL, Williams BC, Hicks RP. 2010. Novel antimicrobial peptides that exhibit activity against select agents and other drug resistant bacteria. Bioorg Med Chem 18:5137–5147. doi: 10.1016/j.bmc.2010.05.065. [DOI] [PubMed] [Google Scholar]
  • 41.Hicks RP, Bhonsle JB, Venugopal D, Koser BW, Magill AJ. 2007. De novo design of selective antibiotic peptides by incorporation of unnatural amino acids. J Med Chem 50:3026–3036. doi: 10.1021/jm061489v. [DOI] [PubMed] [Google Scholar]
  • 42.Baumann T, Nickling JH, Bartholomae M, Buivydas A, Kuipers OP, Budisa N. 2017. Prospects of in vivo incorporation of non-canonical amino acids for the chemical diversification of antimicrobial peptides. Front Microbiol 8:124. doi: 10.3389/fmicb.2017.00124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Budisa N. 2013. Expanded genetic code for the engineering of ribosomally synthetized and post-translationally modified peptide natural products (RiPPs). Curr Opin Biotechnol 24:591–598. doi: 10.1016/j.copbio.2013.02.026. [DOI] [PubMed] [Google Scholar]
  • 44.Wu Y, Tam WS, Chau HF, Kaur S, Thor W, Aik WS, Chan WL, Zweckstetter M, Wong KL. 2020. Solid-phase fluorescent BODIPY-peptide synthesis via in situ dipyrrin construction. Chem Sci 11:11266–11273. doi: 10.1039/d0sc04849f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Ross AC, McKinnie SM, Vederas JC. 2012. The synthesis of active and stable diaminopimelate analogues of the lantibiotic peptide lactocin S. J Am Chem Soc 134:2008–2011. doi: 10.1021/ja211088m. [DOI] [PubMed] [Google Scholar]
  • 46.Knerr PJ, van der Donk WA. 2012. Chemical synthesis and biological activity of analogues of the lantibiotic epilancin 15X. J Am Chem Soc 134:7648–7651. doi: 10.1021/ja302435y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Varnava KG, Sarojini V. 2019. Making solid-phase peptide synthesis greener: a review of the literature. Chem Asian J 14:1088–1097. doi: 10.1002/asia.201801807. [DOI] [PubMed] [Google Scholar]
  • 48.Behrendt R, White P, Offer J. 2016. Advances in Fmoc solid-phase peptide synthesis. J Pept Sci 22:4–27. doi: 10.1002/psc.2836. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Gunasekaran P, Kim EY, Lee J, Ryu EK, Shin SY, Bang JK. 2020. Synthesis of Fmoc-triazine amino acids and its application in the synthesis of short antibacterial peptidomimetics. Int J Mol Sci 21:3602. doi: 10.3390/ijms21103602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Lu J, Xu H, Xia J, Ma J, Xu J, Li Y, Feng J. 2020. d- and unnatural amino acid substituted antimicrobial peptides with improved proteolytic resistance and their proteolytic degradation characteristics. Front Microbiol 11:563030. doi: 10.3389/fmicb.2020.563030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Knerr PJ, Oman TJ, Garcia De Gonzalo CV, Lupoli TJ, Walker S, van der Donk WA. 2012. Non-proteinogenic amino acids in lacticin 481 analogues result in more potent inhibition of peptidoglycan transglycosylation. ACS Chem Biol 7:1791–1795. doi: 10.1021/cb300372b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Mijalis AJ, Thomas DA, 3rd, Simon MD, Adamo A, Beaumont R, Jensen KF, Pentelute BL. 2017. A fully automated flow-based approach for accelerated peptide synthesis. Nat Chem Biol 13:464–466. doi: 10.1038/nchembio.2318. [DOI] [PubMed] [Google Scholar]
  • 53.Hartrampf N, Saebi A, Poskus M, Gates ZP, Callahan AJ, Cowfer AE, Hanna S, Antilla S, Schissel CK, Quartararo AJ, Ye X, Mijalis AJ, Simon MD, Loas A, Liu S, Jessen C, Nielsen TE, Pentelute BL. 2020. Synthesis of proteins by automated flow chemistry. Science 368:980–987. doi: 10.1126/science.abb2491. [DOI] [PubMed] [Google Scholar]
  • 54.Nirenberg MW, Matthaei JH. 1961. The dependence of cell-free protein synthesis in E. coli upon naturally occurring or synthetic polyribonucleotides. Proc Natl Acad Sci USA 47:1588–1602. doi: 10.1073/pnas.47.10.1588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Martin RW, Des Soye BJ, Kwon YC, Kay J, Davis RG, Thomas PM, Majewska NI, Chen CX, Marcum RD, Weiss MG, Stoddart AE, Amiram M, Ranji Charna AK, Patel JR, Isaacs FJ, Kelleher NL, Hong SH, Jewett MC. 2018. Cell-free protein synthesis from genomically recoded bacteria enables multisite incorporation of noncanonical amino acids. Nat Commun 9:1203. doi: 10.1038/s41467-018-03469-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Zemella A, Thoring L, Hoffmeister C, Kubick S. 2015. Cell-free protein synthesis: pros and cons of prokaryotic and eukaryotic systems. Chembiochem 16:2420–2431. doi: 10.1002/cbic.201500340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Liu W-Q, Zhang L, Chen M, Li J. 2019. Cell-free protein synthesis: recent advances in bacterial extract sources and expanded applications. Biochem Eng J 141:182–189. doi: 10.1016/j.bej.2018.10.023. [DOI] [Google Scholar]
  • 58.Goto Y, Katoh T, Suga H. 2011. Flexizymes for genetic code reprogramming. Nat Protoc 6:779–790. doi: 10.1038/nprot.2011.331. [DOI] [PubMed] [Google Scholar]
  • 59.Lee J, Schwieter KE, Watkins AM, Kim DS, Yu H, Schwarz KJ, Lim J, Coronado J, Byrom M, Anslyn EV, Ellington AD, Moore JS, Jewett MC. 2019. Expanding the limits of the second genetic code with ribozymes. Nat Commun 10:5097. doi: 10.1038/s41467-019-12916-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Worst EG, Exner MP, De Simone A, Schenkelberger M, Noireaux V, Budisa N, Ott A. 2015. Cell-free expression with the toxic amino acid canavanine. Bioorg Med Chem Lett 25:3658–3660. doi: 10.1016/j.bmcl.2015.06.045. [DOI] [PubMed] [Google Scholar]
  • 61.Rosenblum G, Cooperman BS. 2014. Engine out of the chassis: cell-free protein synthesis and its uses. FEBS Lett 588:261–268. doi: 10.1016/j.febslet.2013.10.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Gabant P, Borrero J. 2019. PARAGEN 1.0: a standardized synthetic gene library for fast cell-free bacteriocin synthesis. Front Bioeng Biotechnol 7:213. doi: 10.3389/fbioe.2019.00213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Ishida Y, Inouye M. 2016. Suppression of the toxicity of Bac7 (1-35), a bovine peptide antibiotic, and its production in E. coli. AMB Express 6:19. doi: 10.1186/s13568-016-0190-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Rao XC, Li S, Hu JC, Jin XL, Hu XM, Huang JJ, Chen ZJ, Zhu JM, Hu FQ. 2004. A novel carrier molecule for high-level expression of peptide antibiotics in Escherichia coli. Protein Expr Purif 36:11–18. doi: 10.1016/j.pep.2004.01.020. [DOI] [PubMed] [Google Scholar]
  • 65.Stein T, Heinzmann S, Solovieva I, Entian KD. 2003. Function of Lactococcus lactis nisin immunity genes nisI and nisFEG after coordinated expression in the surrogate host Bacillus subtilis. J Biol Chem 278:89–94. doi: 10.1074/jbc.M207237200. [DOI] [PubMed] [Google Scholar]
  • 66.Cotter PD, Draper LA, Lawton EM, McAuliffe O, Hill C, Ross RP. 2006. Overproduction of wild-type and bioengineered derivatives of the lantibiotic lacticin 3147. Appl Environ Microbiol 72:4492–4496. doi: 10.1128/AEM.02543-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Zawada JF, Yin G, Steiner AR, Yang J, Naresh A, Roy SM, Gold DS, Heinsohn HG, Murray CJ. 2011. Microscale to manufacturing scale-up of cell-free cytokine production: a new approach for shortening protein production development timelines. Biotechnol Bioeng 108:1570–1578. doi: 10.1002/bit.23103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Hong SH, Ntai I, Haimovich AD, Kelleher NL, Isaacs FJ, Jewett MC. 2014. Cell-free protein synthesis from a release factor 1 deficient Escherichia coli activates efficient and multiple site-specific nonstandard amino acid incorporation. ACS Synth Biol 3:398–409. doi: 10.1021/sb400140t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Quast RB, Mrusek D, Hoffmeister C, Sonnabend A, Kubick S. 2015. Cotranslational incorporation of non-standard amino acids using cell-free protein synthesis. FEBS Lett 589:1703–1712. doi: 10.1016/j.febslet.2015.04.041. [DOI] [PubMed] [Google Scholar]
  • 70.Si Y, Kretsch AM, Daigh LM, Burk MJ, Mitchell DA. 2021. Cell-free biosynthesis to evaluate lasso peptide formation and enzyme-substrate tolerance. J Am Chem Soc 143:5917–5927. doi: 10.1021/jacs.1c01452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Hegemann JD, Zimmermann M, Xie X, Marahiel MA. 2015. Lasso peptides: an intriguing class of bacterial natural products. Acc Chem Res 48:1909–1919. doi: 10.1021/acs.accounts.5b00156. [DOI] [PubMed] [Google Scholar]
  • 72.Oldach F, Al Toma R, Kuthning A, Caetano T, Mendo S, Budisa N, Süssmuth RD. 2012. Congeneric lantibiotics from ribosomal in vivo peptide synthesis with noncanonical amino acids. Angew Chem Int Ed Engl 51:415–418. doi: 10.1002/anie.201106154. [DOI] [PubMed] [Google Scholar]
  • 73.Deng J, Viel JH, Chen J, Kuipers OP. 2020. Synthesis and characterization of heterodimers and fluorescent nisin species by incorporation of methionine analogues and subsequent click chemistry. ACS Synth Biol 9:2525–2536. doi: 10.1021/acssynbio.0c00308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Kuthning A, Durkin P, Oehm S, Hoesl MG, Budisa N, Süssmuth RD. 2016. Towards biocontained cell factories: an evolutionarily adapted Escherichia coli strain produces a new-to-nature bioactive lantibiotic containing thienopyrrole-alanine. Sci Rep 6:33447. doi: 10.1038/srep33447. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Nickling JH, Baumann T, Schmitt FJ, Bartholomae M, Kuipers OP, Friedrich T, Budisa N. 2018. Antimicrobial peptides produced by selective pressure incorporation of non-canonical amino acids. J Vis Exp 135:57551. doi: 10.3791/57551. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Davis L, Chin JW. 2012. Designer proteins: applications of genetic code expansion in cell biology. Nat Rev Mol Cell Biol 13:168–182. doi: 10.1038/nrm3286. [DOI] [PubMed] [Google Scholar]
  • 77.Wang L, Brock A, Herberich B, Schultz PG. 2001. Expanding the genetic code of Escherichia coli. Science 292:498–500. doi: 10.1126/science.1060077. [DOI] [PubMed] [Google Scholar]
  • 78.Bartholomae M, Baumann T, Nickling JH, Peterhoff D, Wagner R, Budisa N, Kuipers OP. 2018. Expanding the genetic code of Lactococcus lactis and Escherichia coli to incorporate non-canonical amino acids for production of modified lantibiotics. Front Microbiol 9:657. doi: 10.3389/fmicb.2018.00657. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Mukai T, Kobayashi T, Hino N, Yanagisawa T, Sakamoto K, Yokoyama S. 2008. Adding l-lysine derivatives to the genetic code of mammalian cells with engineered pyrrolysyl-tRNA synthetases. Biochem Biophys Res Commun 371:818–822. doi: 10.1016/j.bbrc.2008.04.164. [DOI] [PubMed] [Google Scholar]
  • 80.Hancock SM, Uprety R, Deiters A, Chin JW. 2010. Expanding the genetic code of yeast for incorporation of diverse unnatural amino acids via a pyrrolysyl-tRNA synthetase/tRNA pair. J Am Chem Soc 132:14819–14824. doi: 10.1021/ja104609m. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Luo X, Zambaldo C, Liu T, Zhang Y, Xuan W, Wang C, Reed SA, Yang PY, Wang RE, Javahishvili T, Schultz PG, Young TS. 2016. Recombinant thiopeptides containing noncanonical amino acids. Proc Natl Acad Sci USA 113:3615–3620. doi: 10.1073/pnas.1602733113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Chemla Y, Friedman M, Heltberg M, Bakhrat A, Nagar E, Schwarz R, Jensen MH, Alfonta L. 2017. Expanding the genetic code of a photoautotrophic organism. Biochemistry 56:2161–2165. doi: 10.1021/acs.biochem.7b00131. [DOI] [PubMed] [Google Scholar]
  • 83.Johnson DB, Xu J, Shen Z, Takimoto JK, Schultz MD, Schmitz RJ, Xiang Z, Ecker JR, Briggs SP, Wang L. 2011. RF1 knockout allows ribosomal incorporation of unnatural amino acids at multiple sites. Nat Chem Biol 7:779–786. doi: 10.1038/nchembio.657. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Johnson DB, Wang C, Xu J, Schultz MD, Schmitz RJ, Ecker JR, Wang L. 2012. Release factor one is nonessential in Escherichia coli. ACS Chem Biol 7:1337–1344. doi: 10.1021/cb300229q. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Sharma V, Zeng Y, Wang WW, Qiao Y, Kurra Y, Liu WR. 2018. Evolving the N-terminal domain of pyrrolysyl-tRNA synthetase for improved incorporation of noncanonical amino acids. Chembiochem 19:26–30. doi: 10.1002/cbic.201700268. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Bryson DI, Fan C, Guo LT, Miller C, Söll D, Liu DR. 2017. Continuous directed evolution of aminoacyl-tRNA synthetases. Nat Chem Biol 13:1253–1260. doi: 10.1038/nchembio.2474. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Wang K, Neumann H, Peak-Chew SY, Chin JW. 2007. Evolved orthogonal ribosomes enhance the efficiency of synthetic genetic code expansion. Nat Biotechnol 25:770–777. doi: 10.1038/nbt1314. [DOI] [PubMed] [Google Scholar]
  • 88.Zheng Y, Lewis TL, Jr., Igo P, Polleux F, Chatterjee A. 2017. Virus-enabled optimization and delivery of the genetic machinery for efficient unnatural amino acid mutagenesis in mammalian cells and tissues. ACS Synth Biol 6:13–18. doi: 10.1021/acssynbio.6b00092. [DOI] [PubMed] [Google Scholar]
  • 89.Park HS, Hohn MJ, Umehara T, Guo LT, Osborne EM, Benner J, Noren CJ, Rinehart J, Söll D. 2011. Expanding the genetic code of Escherichia coli with phosphoserine. Science 333:1151–1154. doi: 10.1126/science.1207203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Lajoie MJ, Rovner AJ, Goodman DB, Aerni HR, Haimovich AD, Kuznetsov G, Mercer JA, Wang HH, Carr PA, Mosberg JA, Rohland N, Schultz PG, Jacobson JM, Rinehart J, Church GM, Isaacs FJ. 2013. Genomically recoded organisms expand biological functions. Science 342:357–360. doi: 10.1126/science.1241459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Fredens J, Wang K, de la Torre D, Funke LFH, Robertson WE, Christova Y, Chia T, Schmied WH, Dunkelmann DL, Beránek V, Uttamapinant C, Llamazares AG, Elliott TS, Chin JW. 2019. Total synthesis of Escherichia coli with a recoded genome. Nature 569:514–518. doi: 10.1038/s41586-019-1192-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Anderson JC, Wu N, Santoro SW, Lakshman V, King DS, Schultz PG. 2004. An expanded genetic code with a functional quadruplet codon. Proc Natl Acad Sci USA 101:7566–7571. doi: 10.1073/pnas.0401517101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.de la Torre D, Chin JW. 2021. Reprogramming the genetic code. Nat Rev Genet 22:169–184. doi: 10.1038/s41576-020-00307-7. [DOI] [PubMed] [Google Scholar]
  • 94.Rodriguez EA, Lester HA, Dougherty DA. 2006. In vivo incorporation of multiple unnatural amino acids through nonsense and frameshift suppression. Proc Natl Acad Sci USA 103:8650–8655. doi: 10.1073/pnas.0510817103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Xi Z, Davis L, Baxter K, Tynan A, Goutou A, Greiss S. 2022. Using a quadruplet codon to expand the genetic code of an animal. Nucleic Acids Res 50:4801–4812. doi: 10.1093/nar/gkab1168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Wilson KA, Kalkum M, Ottesen J, Yuzenkova J, Chait BT, Landick R, Muir T, Severinov K, Darst SA. 2003. Structure of microcin J25, a peptide inhibitor of bacterial RNA polymerase, is a lassoed tail. J Am Chem Soc 125:12475–12483. doi: 10.1021/ja036756q. [DOI] [PubMed] [Google Scholar]
  • 97.Piscotta FJ, Tharp JM, Liu WR, Link AJ. 2015. Expanding the chemical diversity of lasso peptide MccJ25 with genetically encoded noncanonical amino acids. Chem Commun (Camb) (Camb) 51:409–412. doi: 10.1039/c4cc07778d. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Shi Y, Yang X, Garg N, van der Donk WA. 2011. Production of lantipeptides in Escherichia coli. J Am Chem Soc 133:2338–2341. doi: 10.1021/ja109044r. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Bagley MC, Dale JW, Merritt EA, Xiong X. 2005. Thiopeptide antibiotics. Chem Rev 105:685–714. doi: 10.1021/cr0300441. [DOI] [PubMed] [Google Scholar]
  • 100.Sivonen K, Leikoski N, Fewer DP, Jokela J. 2010. Cyanobactins: ribosomal cyclic peptides produced by cyanobacteria. Appl Microbiol Biotechnol 86:1213–1225. doi: 10.1007/s00253-010-2482-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Tianero MD, Donia MS, Young TS, Schultz PG, Schmidt EW. 2012. Ribosomal route to small-molecule diversity. J Am Chem Soc 134:418–425. doi: 10.1021/ja208278k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Baneyx F. 1999. Recombinant protein expression in Escherichia coli. Curr Opin Biotechnol 10:411–421. doi: 10.1016/s0958-1669(99)00003-8. [DOI] [PubMed] [Google Scholar]
  • 103.Baneyx F, Mujacic M. 2004. Recombinant protein folding and misfolding in Escherichia coli. Nat Biotechnol 22:1399–1408. doi: 10.1038/nbt1029. [DOI] [PubMed] [Google Scholar]
  • 104.Kakkar N, Perez JG, Liu WR, Jewett MC, van der Donk WA. 2018. Incorporation of nonproteinogenic amino acids in class I and II lantibiotics. ACS Chem Biol 13:951–957. doi: 10.1021/acschembio.7b01024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Deng T, Ge H, He H, Liu Y, Zhai C, Feng L, Yi L. 2017. The heterologous expression strategies of antimicrobial peptides in microbial systems. Protein Expr Purif 140:52–59. doi: 10.1016/j.pep.2017.08.003. [DOI] [PubMed] [Google Scholar]
  • 106.Tanhaieian A, Sekhavati MH, Ahmadi FS, Mamarabadi M. 2018. Heterologous expression of a broad-spectrum chimeric antimicrobial peptide in Lactococcus lactis: its safety and molecular modeling evaluation. Microb Pathog 125:51–59. doi: 10.1016/j.micpath.2018.09.016. [DOI] [PubMed] [Google Scholar]
  • 107.Peng H, Liu HP, Chen B, Hao H, Wang KJ. 2012. Optimized production of scygonadin in Pichia pastoris and analysis of its antimicrobial and antiviral activities. Protein Expr Purif 82:37–44. doi: 10.1016/j.pep.2011.11.008. [DOI] [PubMed] [Google Scholar]
  • 108.Liu Y, Davis RG, Thomas PM, Kelleher NL, Jewett MC. 2021. In vitro-constructed ribosomes enable multi-site incorporation of noncanonical amino acids into proteins. Biochemistry 60:161–169. doi: 10.1021/acs.biochem.0c00829. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Zambaldo C, Luo X, Mehta AP, Schultz PG. 2017. Recombinant macrocyclic lanthipeptides incorporating non-canonical amino acids. J Am Chem Soc 139:11646–11649. doi: 10.1021/jacs.7b04159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Al Toma RS, Kuthning A, Exner MP, Denisiuk A, Ziegler J, Budisa N, Süssmuth RD. 2015. Site-directed and global incorporation of orthogonal and isostructural noncanonical amino acids into the ribosomal lasso peptide capistruin. Chembiochem 16:503–509. doi: 10.1002/cbic.201402558. [DOI] [PubMed] [Google Scholar]
  • 111.Frost JR, Jacob NT, Papa LJ, Owens AE, Fasan R. 2015. Ribosomal synthesis of macrocyclic peptides in vitro and in vivo mediated by genetically encoded aminothiol unnatural amino acids. ACS Chem Biol 10:1805–1816. doi: 10.1021/acschembio.5b00119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Lopatniuk M, Myronovskyi M, Luzhetskyy A. 2017. Streptomyces albus: a new cell factory for non-canonical amino acids incorporation into ribosomally synthesized natural products. ACS Chem Biol 12:2362–2370. doi: 10.1021/acschembio.7b00359. [DOI] [PubMed] [Google Scholar]
  • 113.Huang Y, Wiedmann MM, Suga H. 2019. RNA display methods for the discovery of bioactive macrocycles. Chem Rev 119:10360–10391. doi: 10.1021/acs.chemrev.8b00430. [DOI] [PubMed] [Google Scholar]
  • 114.Passioura T, Suga H. 2017. A RaPID way to discover nonstandard macrocyclic peptide modulators of drug targets. Chem Commun (Camb) 53:1931–1940. doi: 10.1039/c6cc06951g. [DOI] [PubMed] [Google Scholar]
  • 115.Katoh T, Sengoku T, Hirata K, Ogata K, Suga H. 2020. Ribosomal synthesis and de novo discovery of bioactive foldamer peptides containing cyclic β-amino acids. Nat Chem 12:1081–1088. doi: 10.1038/s41557-020-0525-1. [DOI] [PubMed] [Google Scholar]
  • 116.Field D, Cotter PD, Ross RP, Hill C. 2015. Bioengineering of the model lantibiotic nisin. Bioengineered 6:187–192. doi: 10.1080/21655979.2015.1049781. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Field D, Connor PM, Cotter PD, Hill C, Ross RP. 2008. The generation of nisin variants with enhanced activity against specific Gram-positive pathogens. Mol Microbiol 69:218–230. doi: 10.1111/j.1365-2958.2008.06279.x. [DOI] [PubMed] [Google Scholar]
  • 118.Field D, Begley M, O'Connor PM, Daly KM, Hugenholtz F, Cotter PD, Hill C, Ross RP. 2012. Bioengineered nisin A derivatives with enhanced activity against both Gram positive and Gram negative pathogens. PLoS One 7:e46884. doi: 10.1371/journal.pone.0046884. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Islam MR, Shioya K, Nagao J, Nishie M, Jikuya H, Zendo T, Nakayama J, Sonomoto K. 2009. Evaluation of essential and variable residues of nukacin ISK-1 by NNK scanning. Mol Microbiol 72:1438–1447. doi: 10.1111/j.1365-2958.2009.06733.x. [DOI] [PubMed] [Google Scholar]
  • 120.Yuan J, Zhang ZZ, Chen XZ, Yang W, Huan LD. 2004. Site-directed mutagenesis of the hinge region of nisinZ and properties of nisinZ mutants. Appl Microbiol Biotechnol 64:806–815. doi: 10.1007/s00253-004-1599-1. [DOI] [PubMed] [Google Scholar]
  • 121.Hawkins PME, Liu DY, Linington RG, Payne RJ. 2021. Solid-phase synthesis of coralmycin A/epi-coralmycin A and desmethoxycoralmycin A. Org Biomol Chem 19:6291–6300. doi: 10.1039/d1ob01062j. [DOI] [PubMed] [Google Scholar]
  • 122.Torres MDT, Sothiselvam S, Lu TK, de la Fuente-Nunez C. 2019. Peptide design principles for antimicrobial applications. J Mol Biol 431:3547–3567. doi: 10.1016/j.jmb.2018.12.015. [DOI] [PubMed] [Google Scholar]
  • 123.Du Q, Hou X, Ge L, Li R, Zhou M, Wang H, Wang L, Wei M, Chen T, Shaw C. 2014. Cationicity-enhanced analogues of the antimicrobial peptides, AcrAP1 and AcrAP2, from the venom of the scorpion, Androctonus crassicauda, display potent growth modulation effects on human cancer cell lines. Int J Biol Sci 10:1097–1107. doi: 10.7150/ijbs.9859. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Guo X, Ma C, Du Q, Wei R, Wang L, Zhou M, Chen T, Shaw C. 2013. Two peptides, TsAP-1 and TsAP-2, from the venom of the Brazilian yellow scorpion, Tityus serrulatus: evaluation of their antimicrobial and anticancer activities. Biochimie 95:1784–1794. doi: 10.1016/j.biochi.2013.06.003. [DOI] [PubMed] [Google Scholar]
  • 125.Almaaytah A, Farajallah A, Abualhaijaa A, Al-Balas Q. 2018. A3, a scorpion venom derived peptide analogue with potent antimicrobial and potential antibiofilm activity against clinical isolates of multi-drug resistant gram positive bacteria. Molecules 23:1603. doi: 10.3390/molecules23071603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Zhou J, Zhang L, He Y, Liu K, Zhang F, Zhang H, Lu Y, Yang C, Wang Z, Fareed MS, Liang X, Yan W, Wang K. 2021. An optimized analog of antimicrobial peptide Jelleine-1 shows enhanced antimicrobial activity against multidrug resistant P. aeruginosa and negligible toxicity in vitro and in vivo. Eur J Med Chem 219:113433. doi: 10.1016/j.ejmech.2021.113433. [DOI] [PubMed] [Google Scholar]
  • 127.Wang J, Chou S, Yang Z, Yang Y, Wang Z, Song J, Dou X, Shan A. 2018. Combating drug-resistant fungi with novel imperfectly amphipathic palindromic peptides. J Med Chem 61:3889–3907. doi: 10.1021/acs.jmedchem.7b01729. [DOI] [PubMed] [Google Scholar]
  • 128.Lyu Y, Chen T, Shang L, Yang Y, Li Z, Zhu J, Shan A. 2019. Design of Trp-rich dodecapeptides with broad-spectrum antimicrobial potency and membrane-disruptive mechanism. J Med Chem 62:6941–6957. doi: 10.1021/acs.jmedchem.9b00288. [DOI] [PubMed] [Google Scholar]
  • 129.Zaschke-Kriesche J, Reiners J, Lagedroste M, Smits SHJ. 2019. Influence of nisin hinge-region variants on lantibiotic immunity and resistance proteins. Bioorg Med Chem 27:3947–3953. doi: 10.1016/j.bmc.2019.07.014. [DOI] [PubMed] [Google Scholar]
  • 130.Healy B, Field D, O'Connor PM, Hill C, Cotter PD, Ross RP. 2013. Intensive mutagenesis of the nisin hinge leads to the rational design of enhanced derivatives. PLoS One 8:e79563. doi: 10.1371/journal.pone.0079563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Yang Z, He S, Wu H, Yin T, Wang L, Shan A. 2021. Nanostructured antimicrobial peptides: crucial steps of overcoming the bottleneck for clinics. Front Microbiol 12:710199. doi: 10.3389/fmicb.2021.710199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Shao C, Zhu Y, Lai Z, Tan P, Shan A. 2019. Antimicrobial peptides with protease stability: progress and perspective. Future Med Chem 11:2047–2050. doi: 10.4155/fmc-2019-0167. [DOI] [PubMed] [Google Scholar]
  • 133.Baker MA, Maloy WL, Zasloff M, Jacob LS. 1993. Anticancer efficacy of magainin2 and analogue peptides. Cancer Res 53:3052–3057. [PubMed] [Google Scholar]
  • 134.Haimovich B, Tanaka JC. 1995. Magainin-induced cytotoxicity in eukaryotic cells: kinetics, dose-response and channel characteristics. Biochim Biophys Acta 1240:149–158. doi: 10.1016/0005-2736(95)00204-9. [DOI] [PubMed] [Google Scholar]
  • 135.Bucki R, Pastore JJ, Randhawa P, Vegners R, Weiner DJ, Janmey PA. 2004. Antibacterial activities of rhodamine B-conjugated gelsolin-derived peptides compared to those of the antimicrobial peptides cathelicidin LL37, magainin II, and melittin. Antimicrob Agents Chemother 48:1526–1533. doi: 10.1128/AAC.48.5.1526-1533.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Bylund J, Karlsson A, Boulay F, Dahlgren C. 2002. Lipopolysaccharide-induced granule mobilization and priming of the neutrophil response to Helicobacter pylori peptide Hp(2-20), which activates formyl peptide receptor-like 1. Infect Immun 70:2908–2914. doi: 10.1128/IAI.70.6.2908-2914.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Lee JK, Park Y. 2020. All d-lysine analogues of the antimicrobial peptide HPA3NT3-A2 increased serum stability and without drug resistance. Int J Mol Sci 21:5632. doi: 10.3390/ijms21165632. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Gentilucci L, De Marco R, Cerisoli L. 2010. Chemical modifications designed to improve peptide stability: incorporation of non-natural amino acids, pseudo-peptide bonds, and cyclization. Curr Pharm Des 16:3185–3203. doi: 10.2174/138161210793292555. [DOI] [PubMed] [Google Scholar]
  • 139.Gunasekera S, Muhammad T, Strömstedt AA, Rosengren KJ, Göransson U. 2020. Backbone cyclization and dimerization of LL-37-derived peptides enhance antimicrobial activity and proteolytic stability. Front Microbiol 11:168. doi: 10.3389/fmicb.2020.00168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Neubauer D, Jaśkiewicz M, Sikorska E, Bartoszewska S, Bauer M, Kapusta M, Narajczyk M, Kamysz W. 2020. Effect of disulfide cyclization of ultrashort cationic lipopeptides on antimicrobial activity and cytotoxicity. Int J Mol Sci 21:7208. doi: 10.3390/ijms21197208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Góngora-Benítez M, Tulla-Puche J, Albericio F. 2014. Multifaceted roles of disulfide bonds. Peptides as therapeutics. Chem Rev 114:901–926. doi: 10.1021/cr400031z. [DOI] [PubMed] [Google Scholar]
  • 142.Jaśkiewicz M, Neubauer D, Kamysz W. 2017. Comparative study on antistaphylococcal activity of lipopeptides in various culture media. Antibiotics (Basel) 6:15. doi: 10.3390/antibiotics6030015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.White CJ, Yudin AK. 2011. Contemporary strategies for peptide macrocyclization. Nat Chem 3:509–524. doi: 10.1038/nchem.1062. [DOI] [PubMed] [Google Scholar]
  • 144.Marsault E, Peterson ML. 2011. Macrocycles are great cycles: applications, opportunities, and challenges of synthetic macrocycles in drug discovery. J Med Chem 54:1961–2004. doi: 10.1021/jm1012374. [DOI] [PubMed] [Google Scholar]
  • 145.Zhang Y, Zhang Q, Wong CTT, Li X. 2019. Chemoselective peptide cyclization and bicyclization directly on unprotected peptides. J Am Chem Soc 141:12274–12279. doi: 10.1021/jacs.9b03623. [DOI] [PubMed] [Google Scholar]
  • 146.Ban H, Nagano M, Gavrilyuk J, Hakamata W, Inokuma T, Barbas CF, 3rd.. 2013. Facile and stabile linkages through tyrosine: bioconjugation strategies with the tyrosine-click reaction. Bioconjug Chem 24:520–532. doi: 10.1021/bc300665t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Haney CM, Horne WS. 2014. Dynamic covalent side-chain cross-links via intermolecular oxime or hydrazone formation from bifunctional peptides and simple organic linkers. J Pept Sci 20:108–114. doi: 10.1002/psc.2596. [DOI] [PubMed] [Google Scholar]
  • 148.Lau YH, de Andrade P, Quah ST, Rossmann M, Laraia L, Skold N, Sum TJ, Rowling PJE, Joseph TL, Verma C, Hyvonen M, Itzhaki LS, Venkitaraman AR, Brown CJ, Lane DP, Spring DR. 2014. Functionalised staple linkages for modulating the cellular activity of stapled peptides. Chem Sci 5:1804–1809. doi: 10.1039/C4SC00045E. [DOI] [Google Scholar]
  • 149.Reuther JF, Goodrich AC, Escamilla PR, Lu TA, Del Rio V, Davies BW, Anslyn EV. 2018. A versatile approach to noncanonical, dynamic covalent single- and multi-loop peptide macrocycles for enhancing antimicrobial activity. J Am Chem Soc 140:3768–3774. doi: 10.1021/jacs.8b00046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Chow HY, Zhang Y, Matheson E, Li X. 2019. Ligation technologies for the synthesis of cyclic peptides. Chem Rev 119:9971–10001. doi: 10.1021/acs.chemrev.8b00657. [DOI] [PubMed] [Google Scholar]
  • 151.Gang D, Kim DW, Park HS. 2018. Cyclic peptides: promising scaffolds for biopharmaceuticals. Genes (Basel) 9:557. doi: 10.3390/genes9110557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Unger T, Oren Z, Shai Y. 2001. The effect of cyclization of magainin 2 and melittin analogues on structure, function, and model membrane interactions: implication to their mode of action. Biochemistry 40:6388–6397. doi: 10.1021/bi0026066. [DOI] [PubMed] [Google Scholar]
  • 153.Li H, Hu Y, Pu Q, He T, Zhang Q, Wu W, Xia X, Zhang J. 2020. Novel stapling by lysine tethering provides stable and low hemolytic cationic antimicrobial peptides. J Med Chem 63:4081–4089. doi: 10.1021/acs.jmedchem.9b02025. [DOI] [PubMed] [Google Scholar]
  • 154.Imran M, Revol-Junelles AM, de Bruin M, Paris C, Breukink E, Desobry S. 2013. Fluorescent labeling of nisin Z and assessment of anti-listerial action. J Microbiol Methods 95:107–113. doi: 10.1016/j.mimet.2013.07.009. [DOI] [PubMed] [Google Scholar]
  • 155.Scherer K, Wiedemann I, Ciobanasu C, Sahl HG, Kubitscheck U. 2013. Aggregates of nisin with various bactoprenol-containing cell wall precursors differ in size and membrane permeation capacity. Biochim Biophys Acta 1828:2628–2636. doi: 10.1016/j.bbamem.2013.07.014. [DOI] [PubMed] [Google Scholar]
  • 156.Takayama Y, Kusamori K, Nishikawa M. 2019. Click chemistry as a tool for cell engineering and drug delivery. Molecules 24:172. doi: 10.3390/molecules24010172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Laxman P, Ansari S, Gaus K, Goyette J. 2021. The benefits of unnatural amino acid incorporation as protein labels for single molecule localization microscopy. Front Chem 9:641355. doi: 10.3389/fchem.2021.641355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Pantoja R, Rodriguez EA, Dibas MI, Dougherty DA, Lester HA. 2009. Single-molecule imaging of a fluorescent unnatural amino acid incorporated into nicotinic receptors. Biophys J 96:226–237. doi: 10.1016/j.bpj.2008.09.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Schvartz T, Aloush N, Goliand I, Segal I, Nachmias D, Arbely E, Elia N. 2017. Direct fluorescent-dye labeling of α-tubulin in mammalian cells for live cell and superresolution imaging. Mol Biol Cell 28:2747–2756. doi: 10.1091/mbc.E17-03-0161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Sakin V, Hanne J, Dunder J, Anders-Össwein M, Laketa V, Nikić I, Kräusslich HG, Lemke EA, Müller B. 2017. A versatile tool for live-cell imaging and super-resolution nanoscopy studies of HIV-1 env distribution and mobility. Cell Chem Biol 24:635–645.e5. doi: 10.1016/j.chembiol.2017.04.007. [DOI] [PubMed] [Google Scholar]
  • 161.Neubert F, Beliu G, Terpitz U, Werner C, Geis C, Sauer M, Doose S. 2018. Bioorthogonal click chemistry enables site-specific fluorescence labeling of functional NMDA receptors for super-resolution imaging. Angew Chem Int Ed Engl 57:16364–16369. doi: 10.1002/anie.201808951. [DOI] [PubMed] [Google Scholar]
  • 162.Nelson N, Opene B, Ernst RK, Schwartz DK. 2020. Antimicrobial peptide activity is anticorrelated with lipid a leaflet affinity. PLoS One 15:e0242907. doi: 10.1371/journal.pone.0242907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Nelson N, Schwartz DK. 2018. Single-molecule resolution of antimicrobial peptide interactions with supported lipid A bilayers. Biophys J 114:2606–2616. doi: 10.1016/j.bpj.2018.04.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Cheng Z, Kuru E, Sachdeva A, Vendrell M. 2020. Fluorescent amino acids as versatile building blocks for chemical biology. Nat Rev Chem 4:275–290. doi: 10.1038/s41570-020-0186-z. [DOI] [PubMed] [Google Scholar]
  • 165.Völler JS, Budisa N. 2017. Coupling genetic code expansion and metabolic engineering for synthetic cells. Curr Opin Biotechnol 48:1–7. doi: 10.1016/j.copbio.2017.02.002. [DOI] [PubMed] [Google Scholar]
  • 166.Biava HD. 2020. Tackling Achilles’ heel in synthetic biology: pairing intracellular synthesis of noncanonical amino acids with genetic-code expansion to foster biotechnological applications. Chembiochem 21:1265–1273. doi: 10.1002/cbic.201900756. [DOI] [PubMed] [Google Scholar]
  • 167.Ehrlich M, Gattner MJ, Viverge B, Bretzler J, Eisen D, Stadlmeier M, Vrabel M, Carell T. 2015. Orchestrating the biosynthesis of an unnatural pyrrolysine amino acid for its direct incorporation into proteins inside living cells. Chemistry 21:7701–7704. doi: 10.1002/chem.201500971. [DOI] [PubMed] [Google Scholar]
  • 168.Cellitti SE, Ou W, Chiu HP, Grünewald J, Jones DH, Hao X, Fan Q, Quinn LL, Ng K, Anfora AT, Lesley SA, Uno T, Brock A, Geierstanger BH. 2011. d-Ornithine coopts pyrrolysine biosynthesis to make and insert pyrroline-carboxy-lysine. Nat Chem Biol 7:528–530. doi: 10.1038/nchembio.586. [DOI] [PubMed] [Google Scholar]
  • 169.Liu J, Meng F, Du Y, Nelson E, Zhao G, Zhu H, Caiyin Q, Zhang Z, Qiao J. 2020. Co-production of nisin and γ-aminobutyric acid by engineered Lactococcus lactis for potential application in food preservation. Front Microbiol 11:49. doi: 10.3389/fmicb.2020.00049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Vezina B, Rehm BHA, Smith AT. 2020. Bioinformatic prospecting and phylogenetic analysis reveals 94 undescribed circular bacteriocins and key motifs. BMC Microbiol 20:77. doi: 10.1186/s12866-020-01772-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Xin B, Liu H, Zheng J, Xie C, Gao Y, Dai D, Peng D, Ruan L, Chen H, Sun M. 2020. In silico analysis highlights the diversity and novelty of circular bacteriocins in sequenced microbial genomes. mSystems 5:e00047-20. doi: 10.1128/mSystems.00047-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Lee EY, Wong GCL, Ferguson AL. 2018. Machine learning-enabled discovery and design of membrane-active peptides. Bioorg Med Chem 26:2708–2718. doi: 10.1016/j.bmc.2017.07.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Hosseinzadeh P, Watson PR, Craven TW, Li X, Rettie S, Pardo-Avila F, Bera AK, Mulligan VK, Lu P, Ford AS, Weitzner BD, Stewart LJ, Moyer AP, Di Piazza M, Whalen JG, Greisen PJ, Christianson DW, Baker D. 2021. Anchor extension: a structure-guided approach to design cyclic peptides targeting enzyme active sites. Nat Commun 12:3384. doi: 10.1038/s41467-021-23609-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Wang Y, Yang YJ, Chen YN, Zhao HY, Zhang S. 2016. Computer-aided design, structural dynamics analysis, and in vitro susceptibility test of antibacterial peptides incorporating unnatural amino acids against microbial infections. Comput Methods Programs Biomed 134:215–223. doi: 10.1016/j.cmpb.2016.06.005. [DOI] [PubMed] [Google Scholar]
  • 175.Passioura T, Liu W, Dunkelmann D, Higuchi T, Suga H. 2018. Display selection of exotic macrocyclic peptides expressed under a radically reprogrammed 23 amino acid genetic code. J Am Chem Soc 140:11551–11555. doi: 10.1021/jacs.8b03367. [DOI] [PubMed] [Google Scholar]
  • 176.Li Y, Liu T, Liu Y, Tan Z, Ju Y, Yang Y, Dong W. 2019. Antimicrobial activity, membrane interaction and stability of the d-amino acid substituted analogs of antimicrobial peptide W3R6. J Photochem Photobiol B 200:111645. doi: 10.1016/j.jphotobiol.2019.111645. [DOI] [PubMed] [Google Scholar]
  • 177.Jia F, Wang J, Peng J, Zhao P, Kong Z, Wang K, Yan W, Wang R. 2017. d-Amino acid substitution enhances the stability of antimicrobial peptide polybia-CP. Acta Biochim Biophys Sin (Shanghai) 49:916–925. doi: 10.1093/abbs/gmx091. [DOI] [PubMed] [Google Scholar]
  • 178.Wu CL, Hsueh JY, Yip BS, Chih YH, Peng KL, Cheng JW. 2020. Antimicrobial peptides display strong synergy with vancomycin against vancomycin-resistant E. faecium, S. aureus, and wild-type E. coli. Int J Mol Sci 21:4578. doi: 10.3390/ijms21134578. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Nitsche C, Onagi H, Quek JP, Otting G, Luo D, Huber T. 2019. Biocompatible macrocyclization between cysteine and 2-cyanopyridine generates stable peptide inhibitors. Org Lett 21:4709–4712. doi: 10.1021/acs.orglett.9b01545. [DOI] [PubMed] [Google Scholar]
  • 180.Mendive-Tapia L, Subiros-Funosas R, Zhao C, Albericio F, Read ND, Lavilla R, Vendrell M. 2017. Preparation of a Trp-BODIPY fluorogenic amino acid to label peptides for enhanced live-cell fluorescence imaging. Nat Protoc 12:1588–1619. doi: 10.1038/nprot.2017.048. [DOI] [PubMed] [Google Scholar]
  • 181.Barth ND, Subiros-Funosas R, Mendive-Tapia L, Duffin R, Shields MA, Cartwright JA, Henriques ST, Sot J, Goñi FM, Lavilla R, Marwick JA, Vermeren S, Rossi AG, Egeblad M, Dransfield I, Vendrell M. 2020. A fluorogenic cyclic peptide for imaging and quantification of drug-induced apoptosis. Nat Commun 11:4027. doi: 10.1038/s41467-020-17772-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Morris MA, Malek M, Hashemian MH, Nguyen BT, Manuse S, Lewis K, Nowick JS. 2020. A fluorescent teixobactin analogue. ACS Chem Biol 15:1222–1231. doi: 10.1021/acschembio.9b00908. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Applied and Environmental Microbiology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES