Next Article in Journal
Integrated Multi-Omics Analysis Reveals Mountain-Cultivated Ginseng Ameliorates Cold-Stimulated Steroid-Resistant Asthma by Regulating Interactions among Microbiota, Genes, and Metabolites
Previous Article in Journal
An IL-5 Single-Nucleotide Polymorphism Influences Neuroinflammation and Prospective Disease Activity in Multiple Sclerosis
Previous Article in Special Issue
Venom Ex Machina? Exploring the Potential of Cell-Free Protein Production for Venom Biodiscovery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cell-Free Systems: Ideal Platforms for Accelerating the Discovery and Production of Peptide-Based Antibiotics

1
School of Interdisciplinary Bioscience and Bioengineering (I-Bio), Pohang University of Science and Technology, Pohang 37673, Republic of Korea
2
Department of Chemical Engineering, Pohang University of Science and Technology, Pohang 37673, Republic of Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2024, 25(16), 9109; https://doi.org/10.3390/ijms25169109 (registering DOI)
Submission received: 8 July 2024 / Revised: 12 August 2024 / Accepted: 19 August 2024 / Published: 22 August 2024
(This article belongs to the Special Issue Versatility of Protein Synthesis in a Test Tube)

Abstract

:
Peptide-based antibiotics (PBAs), including antimicrobial peptides (AMPs) and their synthetic mimics, have received significant interest due to their diverse and unique bioactivities. The integration of high-throughput sequencing and bioinformatics tools has dramatically enhanced the discovery of enzymes, allowing researchers to identify specific genes and metabolic pathways responsible for producing novel PBAs more precisely. Cell-free systems (CFSs) that allow precise control over transcription and translation in vitro are being adapted, which accelerate the identification, characterization, selection, and production of novel PBAs. Furthermore, these platforms offer an ideal solution for overcoming the limitations of small-molecule antibiotics, which often lack efficacy against a broad spectrum of pathogens and contribute to the development of antibiotic resistance. In this review, we highlight recent examples of how CFSs streamline these processes while expanding our ability to access new antimicrobial agents that are effective against antibiotic-resistant infections.

1. Introduction

In an era when antibiotic resistance is rapidly becoming one of the greatest threats to global health, the race to discover and develop new antimicrobial agents has never been more urgent [1]. Recently, the World Health Organization (WHO) issued a warning on the potential consequences of a “post-antibiotic era” [2,3,4] in which common infections and minor injuries that have been treatable for decades could become dangerous. This alarming scenario has reignited interest in alternative therapeutic strategies, particularly PBAs, which show promise due to their potential to overcome traditional resistance pathways.
This renewed focus on novel antibiotics recalls one of the most monumental moments in medical history, the discovery of penicillin produced by the fungus Penicillium notatum in 1928, which revolutionized medicine by providing an effective treatment against bacterial infections. Structurally, penicillin contains a cyclic peptide (beta-lactam) backbone that is derived from a tripeptide scaffold, δ-L-(α-aminoadipoyl)-L-cysteinyl-D-valine [5,6,7]. This underscores that PBAs, including AMPs and their synthetic variants, can offer potent activity against resistant pathogens through mechanisms that differ from those of traditional antibiotics. Notably, the ability to incorporate monomers in addition to the 20 canonical amino acids with a repetitive polymeric backbone provides significant diversity, yielding products with unique mechanisms of action, effectiveness against multidrug resistant pathogens, and a low likelihood of cross-resistance [8]. As a result, this versatility enables them to target various cellular components, including membranes [9,10,11,12,13], biomolecules (enzymes [14,15], nucleic acids [16,17], and metabolites [18]), often with greater efficiency than small-molecule antibiotics (Table 1).
Recent cutting-edge technologies in high-throughput DNA sequencing and artificial intelligence (AI)-based bioinformatics tools have identified unique natural product gene clusters [38,39,40]. Continued exploration for PBAs using these techniques has also increased the possibility of discovering novel antibiotics with enhanced efficacy and reduced resistance [41,42]. A recently updated database, the Antimicrobial Peptide Database (APD3), which lists over 3000 PBAs from diverse natural sources (https://aps.unmc.edu, accessed on 1 February 2024), suggests an untapped potential for further discovery of novel bioactive materials. Despite their potential, the conventional methods for discovering and producing PBAs are often slow and limited by the complexities of production systems, such as isolation from natural sources or chemical synthesis of their derivatives. This limitation is evident in the fact that only 12 new antibiotics have entered the market from 2017 to 2021 [43]. In the 2021 WHO annual analysis, only 27 new antibiotics were in clinical trials against critical pathogens [4], and of these 27, only 6 were considered to be capable of overcoming antibiotic resistance. In contrast, more than 1300 cancer drugs were in clinical trials in 2020 [44]. This is why CFSs have emerged as a groundbreaking platform, offering unprecedented speed, flexibility, and efficiency in the discovery and synthesis of PBAs.
Unlike traditional in vivo systems, CFSs mediate protein translation via the ribosome in a controlled, open environment, enabling rapid and flexible experimentation [45]. As researchers continue to explore the full potential of CFSs [46,47], these systems have become poised to play a pivotal role in addressing challenges of rapid discovery, design, screening, and biomanufacturing of medically important peptides [48,49,50,51]. While CFSs have been extensively used for research on peptides, proteins, and enzymes, reviews focusing on PBAs integrated with CFSs are lacking. In this review, we begin by categorizing natural PBAs (Section 2) and providing a brief overview of CFSs (Section 3). Next, we present key advancements in CFSs used for the identification of mechanisms of action of PBAs (Section 4.1), characterization of biosynthetic pathways (Section 4.2), high-throughput screening (Section 4.3), and direct, rapid synthesis of PBAs with complex structures (Section 4.4). Overall, CFSs are powerful tools with the potential to rapidly create and diversify PBAs, bridging gaps in the traditional technologies used for the identification, characterization, and production of antibiotics.

2. PBAs Are Produced through Various Biosynthetic Pathways

Based on the molecular mechanisms involved in the synthesis of bioactive peptides in microorganisms [9,52,53,54], we categorize them in two broad groups: ribosomally (Group I) or non-ribosomally (Group II) produced peptides (Table 2). Ribosomally produced peptides can be further divided into two different subgroups. The first subgroup of Group I, shown in Table 2(a), is produced by the direct polymerization of the 20 canonical amino acid monomers via the translational machinery (aminoacyl tRNA synthetases (aaRSs), elongation factors, and ribosome). These PBAs can be rapidly designed and synthesized both in vivo and in vitro because they are directly synthesized by the ribosome using the cell’s genetic code in messenger RNA (mRNA) as a template [55,56,57,58]. This allows for the efficient production of diverse variants that may selectively target one or multiple pathogenic bacteria [59,60,61,62], thereby making them promising candidates for therapeutic applications [63].
The folded states of such PBAs fall into three broad structural categories: α-helical peptides (e.g., LL37, melittin) [9,64]; extended structures abundant in proline (Api-137) [63,65], tryptophan (indolicidin) [66], or arginine (bactenecin) [67]; and macrocyclic loops with one or multiple disulfide bonds (SFTI-1, kalata B1) [68,69,70]. In general, α-helical peptides disrupt cell membranes. These helical structures provide amphipathic properties, allowing them to insert into microbial cell membranes effectively and destabilize the lipid bilayers [71,72], ultimately causing cell lysis and death. Compared to α-helical peptides, extended structures rich in specific amino acids (e.g., Pro, Trp, and Arg) utilize their unique molecular characteristics. For example, proline-rich AMPs (PrAMPs) adopt a rigid conformation that facilitates membrane penetration and intracellular targeting with ribosomes, leading to inhibition of protein synthesis [73,74]. Tryptophan-rich AMPs (WrAMPs) leverage their aromatic side chains for strong membrane interactions, leading to pore formation in the membrane. Arginine-rich peptides (RrAMPs) exploit their positive charge to interact with membranes and nucleic acids, disrupting essential cellular processes. The cyclic architectures formed by disulfide bonds create a rigid structure, resulting in high-affinity interactions with biological targets due to reduced entropic effects and resistance to proteolytic degradation [75,76,77,78,79,80].
Table 2. Classification of peptide-based antibiotics (PBAs).
Table 2. Classification of peptide-based antibiotics (PBAs).
GroupMachinery for BiosynthesisTypes of PeptidesRepresentative Compounds
I(a) Translational apparatusPeptides only containing the 20 canonical amino acidsα-Helical peptides [9], PrAMPs [63,65], WrAMPs [66], RrAMPs [67], disulfide-rich peptides [69,70]
(b) Translational apparatus and post-translational enzymesRibosomally synthesized and
post-translationally modified peptides (RiPPs)
  • Linear architecture
    (azole-containing peptides (LAPs) [81], cyanobactin [82], linaridin [83])
  • Macrocyclic architecture
    (lanthipeptides [84], thiopeptide [85], lasso peptides [86])
II(c) Non-ribosomal peptide synthetases (NRPSs)Non-ribosomal peptides (NRPs)Lipopeptide surfactants [87], cyclodecapeptides [88], glycopeptide antibiotics [89], mycobactins [90], malleobactins [91]
The second subgroup of Group I, shown in Table 2(b), consists of PBAs that contain one or multiple non-canonical amino acids (ncAAs), thereby acquiring novel bioactivities. A representative example is ribosomally synthesized and post-translationally modified peptides (RiPPs). They differ from the first subgroup in that they have diverse structures beyond the linear peptide backbone and include more than the 20 canonical amino acids. This structural diversity arises from various post-translational modifications (PTMs), such as cyclization, cyclodehydration (ring-closing condensation), glycosylation, N-alkylation, and others (Table 3). Cyclization is a key modification in the production of lanthipeptides and lasso peptides [86,92], involving mechanisms such as thioether bridge formation, cysteine reduction, or head-to-tail cyclization. Cyclodehydration is another crucial process, forming biologically active motifs by producing heterocycles such as thiazoles and oxazoles through the dehydration of serine, threonine, or cysteine residues and subsequent ring-closing condensation with the peptide backbone [93,94,95]. Glycosylation is a significant PTM catalyzed by various glycosyltransferases that attach carbohydrate moieties (glycans or oligosaccharides) to specific amino acid residues within ribosomally produced peptides. This modification enhances the structural stability, solubility, and bioactivity of RiPPs [96]. N-alkylation, most commonly N-methylation, is catalyzed by N-methyltransferases, which transfer a methyl group from a donor molecule, typically S-adenosylmethionine (SAM), to the nitrogen atoms of amino acid residues within peptides [97,98]. Methylation helps maintain the structural integrity of peptides under various stress conditions, such as changes in pH or temperature [99].
Group II PBAs, shown in Table 2(c), include non-ribosomally synthesized peptides produced through a complex enzymatic process that operates independently of the translational apparatus and mRNA templates [52,100,101,102]. This process is mediated by non-ribosomal peptide synthetases (NRPSs), which are large, multi-modular enzyme complexes that function like an assembly line to incorporate a wide variety of substrates. The modules typically include domains for adenylation (A), thiolation (T), condensation (C), and thioesterase (TE) [98]. In the A domain, the enzyme selects and activates specific amino acids by forming an aminoacyl-adenylate intermediate. The activated amino acid is then transferred to the T domain, where it is covalently attached to a 4′-phosphopantetheine (Ppant) prosthetic group, creating a thioester bond. This thioester linkage secures the growing peptide chain to the enzyme, allowing it to be carried to the C domain. In the C domain, the enzyme catalyzes the peptide bond formation between the incoming amino acid and the nascent peptide chain. The TE domain is crucial for releasing the completed peptide from the NRPS complex. It catalyzes the cyclization or hydrolysis of the thioester bond, thus liberating the final peptide product. This sequential addition and elongation process ensures that amino acids are added in a specific and ordered manner, resulting in the precise assembly of the peptide product (Figure 1) [98,103]. Due to this characteristic, NRPS allows the incorporation of diverse non-canonical substrates, not being limited to the 20 standard amino acids, producing peptides with unique structures and biochemical properties, including increased stability, enhanced biological activity, and resistance to enzymatic degradation [104]. Various non-canonical substrates incorporated by NRPS include D-amino acids [105], hydroxyacids [106], N-alkylated amino acids [107], β-amino acids [108], and others [109]. Subsequent chemo-enzymatic modifications of side chains after peptide synthesis result in even more unique structures, such as cyclic [110] and branched [111] structures, or with multiple rings and cross-links [112]. These attributes make NRPs effective and selective in their biological activities, often offering antimicrobial, anticancer, and immunosuppressive properties [113,114]. Such versatility and robustness highlight the potential of NRPs in therapeutic applications and drug development.
An ideal platform for the development of PBAs with the characteristics discussed above would be a comprehensive system that integrates several key functions essential for efficient and effective discovery and production. First, it would facilitate the identification of the mechanism of action of PBA. This involves advanced tools such as high-resolution imaging [9], biochemical assays [115], and computational modeling [116] to understand how PBAs interact with bacterial targets at the molecular level. By elucidating the specific interactions and pathways through which PBAs exert their antimicrobial effects, researchers can pinpoint the sequences responsible for efficacy and optimize the composition of PBAs for enhanced activity. Second, the platform would enable the elucidation of biosynthetic pathways. Understanding these pathways is critical for revealing the step-by-step process by which these peptides are assembled, modified, and matured, providing insights into how to manipulate and optimize these pathways for better yield and functionality. Third, the system would allow for the selection of efficient candidates from a library. This requires a robust method for screening a vast array of potential peptide sequences to identify those with the most desirable properties, such as high antimicrobial activity, stability, and low toxicity. High-throughput screening techniques and machine learning algorithms could play a significant role in this selection process. Finally, the platform should support the rapid production of complex structures. This involves the ability to synthesize these peptides quickly and accurately, both in vitro and in vivo, using non-canonical substrates. Cell-free systems exemplify such an ideal platform, offering the flexibility and efficiency needed for each of these stages, from identification to production of diverse PBAs.

3. CFSs as Ideal Platforms for PBA Research

CFSs provide ideal platforms as they meet the essential requirements for studying PBAs. The platforms have two main types: lysate-based and reconstituted systems, which offer distinct advantages [45]. Briefly, lysate-based systems, as shown in Figure 2a, are derived from cell extracts obtained through relatively simple methods, such as sonication [117,118], high-pressure homogenization [119,120], and enzymatic lysis [121]. These systems are typically produced using prokaryotic cells, such as Escherichia coli (E. coli) [122,123] or other bacterial cells [124,125]. The minimalistic nature of prokaryotic cells, which lack complex organelles and extensive PTM pathways, simplifies the synthesis of target proteins with high productivity (~g/L/h) [126,127]. In contrast, PTM pathways in eukaryotic cells involve a series of enzymatic reactions that modify proteins after translation, enabling them to become fully functional. These modifications include glycosylation [119], phosphorylation [128], ubiquitination [129,130], methylation [131], acetylation [132], and the formation of disulfide bonds [133], among others. Each of these modifications requires specific enzymes and cellular machinery, such as the endoplasmic reticulum and Golgi apparatus. Therefore, eukaryotic cell-free systems provide an optimal balance between the simplicity of prokaryotic expression systems and the complexity of whole-cell eukaryotic systems by harnessing the cellular machinery required for PTMs and protein folding without the need to maintain living cells. Eukaryotic cell-free systems use extracts from yeast [134,135], wheat germ [136], or Chinese hamster ovary (CHO) cells containing the enzymes and molecular chaperones required for performing PTMs and ensuring correct protein folding [137]. This enables the production of high-value proteins, such as therapeutic antibodies that require precise folding and glycosylation, to function effectively.
Reconstituted systems (Figure 2b) consist of a defined set of purified components that are essential for transcription and translation [138]. These typically include ribosomes, tRNAs, aaRS, nucleoside triphosphates (NTPs), and translation factors for initiation, elongation, and termination. These systems offer a high degree of control over the protein synthesis process by using individually purified components, which allows researchers to eliminate undesired biochemical reactions associated with crude extracts, such as proteolysis [139] and nucleic acid degradation [140]. The PURExpress® (New England Biolabs) system is one of the most well-known commercially available reconstituted CFSs [141]. Recently, methods to construct reconstituted systems have been developed, enabling researchers to efficiently produce their in-house systems in the laboratory. Due to this, reconstituted CFSs are often an ideal choice for research on PBAs. Gabant et al. synthesized multiple active bacteriocins in parallel by using a standardized synthetic gene library in PURExpress®, demonstrating its potential for the direct, fast, and accurate synthesis of active PBAs [142]. The flexible in vitro translation (FIT) system developed by Hiroaki Suga exemplifies a reconstituted system that enables the synthesis of non-standard peptides by incorporating ncAAs through a reprogrammed ribosomal translation process [143]. Genetic code reprogramming is achieved by a ribozyme, flexizyme, which creates misacylated-tRNAs with ncAAs. Flexizyme facilitates the attachment of a wide range of ncAAs to tRNAs, enabling their incorporation into peptides during in vitro translation [50,144,145]. This technology greatly expands the potential of peptide synthesis, allowing for the production of novel PBAs with unique properties [146]. Another key factor of the FIT system is its use of open reaction characteristics of CFSs. For PBA synthesis, additional enzymes are supplemented in the reaction, which enhances the biosynthetic process or PTMs in vitro.
However, reconstituted systems are relatively difficult to construct compared to lysate-based systems, due to the extensive purification steps required for the minimal components. Importantly, maintaining the stability and activity of these purified components can be challenging, further complicating the construction and operation of reconstituted systems.

4. Main Research on PBAs Using CFSs

Utilizing the benefits of CFSs for PBA studies, researchers have made significant efforts in four broad areas: understanding the molecular mechanisms of antimicrobial activity, elucidating biosynthetic pathways, developing new methods for selection, and enabling large-scale synthesis of various PBAs (Table 4) [147]. In our review, we categorized the four major areas of PBA mentioned above into separate sections, introduced the corresponding research in each, and focused on summarizing those studies (Table 5).

4.1. CFSs Enable the Elucidation of the Mechanism of Action of PBAs

PBAs combat bacterial infections through various mechanisms (Table 1). These mechanisms include cell membrane disruption [173], pore formation [174], inhibition of protein synthesis [175], interference with nucleic acid functions [176], immune modulation [177], biofilm disruption [178], and reactive oxygen species production [179]. Mechanistic studies aim to understand how PBAs exert their effects at the molecular level, which in turn provides insight for designing PBAs with enhanced specificity and altered toxicity. Thus, various methods combining biochemical [9,180], biophysical [181], and molecular biology [182,183] techniques have been developed. Advanced techniques such as spectroscopy (nuclear magnetic resonance or circular dichroism) [184,185] and microscopy (electron microscopy or confocal fluorescence microscopy) [9,156] reveal structural details of peptide interactions with membranes and intracellular targets.
Recent studies using CFSs have contributed to a better understanding of these mechanisms of action. Toe-printing assays (Figure 3a), also known as primer extension inhibition assays, are applied in vitro to study the interactions between PBAs and the ribosomes during protein synthesis [186,187]. In these assays, a chemical probe-labeled primer is extended by reverse transcriptase until it encounters a ribosome stalled on the mRNA, resulting in a “toe-print” that indicates the position of the ribosome. This technique provides detailed and direct insights into how PBAs affect the intracellular translation reaction at specific sites along the mRNA [188]. A key to success was using CFSs in which enzymes (e.g., DNases, RNases, and proteases) or other cellular components unrelated to toe-printing that could interfere with ribosome stalling [156,189,190] were selectively eliminated. Similarly, Koller et al. further demonstrate how O-glycosylated drosocins (type II PrAMPs) enhance their antimicrobial activity (Figure 3b) at the molecular level using a CFS [156]. They revealed that type II glycosylated drosocins do not disrupt the cellular membrane, but instead enter the intracellular space and arrest ribosomes at stop codons by binding to the ribosome’s exit tunnel and release factor 1 (RF1) in the A-site. This insight helped them understand the mechanism of inhibition: the binding of drosocin to the ribosome and RFs deplete free RFs in the cellular environment, preventing the ribosomes from releasing completed proteins at stop codons, thereby causing ribosome stalling and misincorporation or frameshifting.
Another recently elucidated novel mechanism involves PBAs targeting the phase transitions of cytosolic nucleic acids. Using machine learning sequence analysis, Sneideris et al. identified new PBAs that readily form liquid-like condensates with nucleic acids through phase separation (Figure 3c) [16]. These PBAs often include charged and polar amino acid residues, which interact strongly with nucleic acids. This results in the sequestration and compaction of nucleic acids in living cells, ultimately leading to bacterial cell death. Notably, this capability correlates with their inhibition of prokaryotic transcription and translation, as well as their interaction with bacterial nucleic acids in vivo, suggesting that designing PBAs capable of forming condensates could be a key to discovering new PBAs.

4.2. CFSs Facilitate the Characterization of Biosynthetic Pathways of PBA

Identifying biosynthetic gene clusters (BGCs) is also a crucial step in the discovery of new peptides, which can inform strategies for PBA production using enzymes. Recent studies have employed genome mining and bioinformatics tools to identify BGCs that encode target PBAs [62,191,192]. Investigating BGCs using CFSs provides a comprehensive framework for studying biosynthetic pathways [149,150,151] (Table 4), as each step of the biosynthesis can be independently controlled with a specific enzyme in separate reactions. This approach allows for the exploration of enzyme promiscuity [157,158,159,160,161,162,163,171,172] and kinetics [152] required for each biosynthetic step. Importantly, by using this method, the range of substrate analogs for an enzyme can be determined and enzymes can be further engineered with mutations at catalytic sites. This ultimately contributes to the discovery of novel bioactive compounds and the diversification of products that might be inaccessible through in vivo methods.
Thiopeptides, linear azole-containing peptides (LAPs), and lanthipeptides feature complex structures beyond simple linear peptide backbones. These structures are formed post-translationally by multiple enzymes encoded within their gene clusters [40]. These diverse classes of RiPPs contain cyclic motifs, such as thiazoles with macrocyclic (loop-like) structures, heterocyclic azolines, and thioether bonds [193]. Recently, Vinogradov et al. demonstrated the synthesis of a thiopeptide, known as lactazole A, using an FIT-Laz system. In this system, the key enzymes necessary to form the unique structures during the biosynthetic pathways are either directly produced from a DNA template or produced separately in vivo and supplemented as purified enzymes [157]. This approach enabled the identification of the minimal lactazole scaffold required for bioactivity. Similarly, a recent study identified the specific roles of enzymes required for intricate biosynthesis using a FIT platform, which synthesizes unique structural elements, such as an N-terminal acetyl moiety, azoles, and dehydroalanines (Figure 4a). By supplementing each enzyme in various combinations within a one-pot, cell-free protein synthesis system, the authors clarified the sequential order of modifications and facilitated the production of goadsporin analogs [158].
Building on the success of elucidating PTM reactions and discovering their cognate enzymes, CFSs have further demonstrated their potential for the overproduction of novel lanthipeptides [152]. Lanthipeptides are promising PBAs with significant therapeutic applications [131]; however, their production faces challenges at an industrial scale due to their toxicity to the host bacterial cells and the complexity of biosynthesis pathways. Recently, Liu et al. presented a novel CFS that enables the discovery and production of lanthipeptides (Figure 4b) [139]. The authors analyzed gene cluster through genome mining and identified essential enzymes (NisP, NisB, and NisC) that facilitate the bottleneck steps of biosynthetic pathways. Supplementation with the purified enzymes significantly increased the yield (180 IU/mL) of active lanthipeptides. By leveraging the enzymes’ promiscuity toward various substrates, they successfully discovered four novel lanthipeptide analogs that are not toxic to host cells. Based on these findings, they implemented the systems into host cells (E. coli) and produced the lanthipeptide analogs (260.1 ± 10.5 IU/mL).

4.3. CFSs Enable the Selection of PBAs from a Diverse Library

High throughput screening (HTS) is a highly efficient technique that allows researchers to quickly evaluate and identify potential PBAs from vast libraries of compounds. By automating the testing process, HTS enables the simultaneous analysis of thousands to millions of samples, significantly accelerating the discovery of novel PBAs with desired biological activities [133]. Building on these achievements, CFSs provide a cost-effective platform for screening large peptide libraries and enable the customization of peptide synthesis components, further enhancing the efficiency of the screening process [168,169].
In the study on thiopeptide analogs discussed in the previous section, the authors further demonstrated the capability of the CFS to select the most efficient lanthipeptides from a library. In a new CFS, in which the genes encoding proteases or peptidases from the E. coli genome were removed, the authors produced ~3000 highly active variants of the salivaricin B gene [139] (Figure 5a). Among these, a salivaricin B mutant with 4-fold higher antibacterial activity than salivaricin B against B. subtilis, M. luteus, and L. lactis (Figure 5b) was identified through minimum inhibitory concentration (MIC) testing (1 µM, 0.5 µM, and 1 µM, respectively) [139]. These findings highlight the broad applicability of the CFS platform for selecting new variants with enhanced antimicrobial activity.
HTS assays also help pinpoint optimal reaction conditions for high-yield production [164,165,194]. Recent studies have demonstrated that CFSs can be used to reconstruct large NRPSs and directly produce natural products using template DNA supplemented in the reaction [164,166]. The NRPS system typically requires customized reaction conditions, such as glucose concentration, enzyme ratios, cofactors, and reaction time (Figure 6a), which are essential for the optimal synthesis of target PBAs. The optimal conditions identified in the CFS enabled efficient NRPS synthesis, thereby facilitating the in situ production of natural products, such as diketopiperazine [165] and valinomycin [164] (Figure 6b). These results demonstrate the capability of CFSs to overcome the limitations of traditional cell-based expression systems that lack the flexibility to finely control the synthetic conditions.
Returning to the identification of the minimal lactazole (a type of RiPP) scaffold discussed in the previous section, Vinogradov et al. further explored expanding the range of substrates acceptable for the enzymes used in lactazole biosynthesis [171,172]. Using a reactivity-profiling mRNA display assay, which establishes a genotype–phenotype link through mRNA display techniques and indirectly identifies amino acid sequences via sequencing the conjugated mRNA [171], the authors identified alternative substrate preferences in the biosynthetic pathway (Figure 7a). Subsequently, amino acid sequences scored by next generation sequencing (NGS) were used to train a deep generative model that generated a diverse array of lactazole-forming sequences based on minimal lactazole scaffolds. This assay not only confirmed known aspects of lactazole biosynthesis but also provided new insights into enzyme fitness, enabling the construction of combinatorial thiopeptides, and leveraging the capabilities of CFSs for high-throughput analysis. The model also predicted the promiscuity of lactazole RiPP enzymes, aiding in the design of thiopeptide derivatives (Figure 7b).
Combined with CFSs, deep learning models have significantly expanded our ability to discover new PBAs [195,196]. These models typically learn from existing PBAs to propose novel protein sequences, enabling the creation of thousands of new designs much faster than traditional methods. A recent study showed that this platform could generate multiple de novo PBAs from synthetic DNA fragments within 30 min, followed by direct minimum inhibitory concentration (MIC) testing of these PBA mutants within the CFS (Figure 8a) [168]. This integration further accelerates the discovery and validation of new antimicrobial agents at a low cost.
Microfluidic systems also play a crucial role in accelerating the generation and selection of PBAs for high-throughput screening. Vesicle structures with lipid membranes have shown the ability to screen PBAs based on membrane specificity. Nuti et al. demonstrated that double emulsion droplets can be generated at high frequency by microfluidic systems for PBA selection (Figure 8b) [170]. In this system, diverse PBAs are synthesized within large vesicles that co-encapsulate smaller droplets containing self-quenched fluorescent dyes. When PBAs disrupt the membranes, the dye is released, producing a fluorescent signal. These fluorescent droplets are then sorted, and the DNA template within each droplet is sequenced for PBA identification.

4.4. CFSs Enable the Synthesis of PBAs with Complex Structures

PBAs found in nature often have complex molecular structures that enhance their ability to bind accurately to targets [113], increasing their effectiveness at low concentrations and reducing the risk of off-target effects [114]. These intricate structures also provide significant stability, making PBAs highly resistant to proteolytic degradation in cellular environments due to their unique topological features, similar to the stability observed in macrocyclic structures [197,198]. In addition, the process of stapling residues can further improve the bioavailability of PBAs by increasing their structural rigidity and locking the peptides into active conformations, thereby resulting in higher affinities for their targets [199,200].
The most common covalent bond providing structural stability in nature is the disulfide bond formed between the thiol groups of cysteine residues [78]. Wu et al. recently introduced a method to improve the folding and reduce the aggregation of disulfide-rich peptides and proteins produced in vitro (Figure 9a) [153]. They used a modified E. coli-based CFS to enable the co-translational capture of peptides using an affinity matrix. This facilitated accurate oxidative folding through disulfide bonds and recycling of misfolded states. This approach enhances the yield (4~5-fold) of peptides with complex macrocyclic architectures and multiple disulfide bonds, indicating that CFSs could be applied to various peptides that need to maintain their structure under denaturing conditions, such as high temperature, acids, or bases.
Lasso peptides are a prevalent class of RiPPs produced in bacteria (e.g., E. coli, B. caledonica, S. aureus, B. subtilis, and P. aeruginosa) [86,201,202]. Their lariat-like threaded conformation imparts unique anticancer [203] and antibacterial activities [204], along with resistance to thermal and proteolytic degradation [202]. Despite these significant advantages, the production efficiency of lasso peptides is low due to the lack of an appropriate synthesis platform [205,206]. Understanding the BGCs of these peptides is essential for identifying the regions critical to their unique conformation, which is key to discovering and synthesizing exciting structures [86]. Si et al. recently demonstrated that a wide variety of lasso peptides could be synthesized using a CFS based on the rational design of precursor substrates containing a leader sequence. The identification of a promiscuous enzyme, lasso cyclase, was another critical factor in facilitating lasso formation (Figure 9b) [154]. Using a CFS, they achieved a 200-fold increase in production efficiency compared to traditional heterologous expression methods.
A CFS developed by Fleming et al. for thiopeptide synthesis demonstrates the potential for integrating additional chemo-enzymatic reactions and ncAAs to synthesize even more complex structures [167]. This CFS enables the incorporation of a selenocysteine residue, which undergoes oxidative elimination with hydrogen peroxide to generate dehydroalanines (Figure 9c). The addition of a PTM enzyme, TclM, facilitates an enzymatic cycloaddition reaction between the dehydroalanines, converting a linear structure into a lasso structure featuring a six-membered heterocyclic pyridine motif.
Table 5. The list of CFS-mediated PBAs covered in this article.
Table 5. The list of CFS-mediated PBAs covered in this article.
SectionPBAsStructural FeaturesTarget Microorganisms
Section 4.1Drosocin
  • O-glycosylation
Gram-negative bacteria (E. coli)
Os-C-Bacteria
P113
  • C-terminus amidation: (CO2H → CONH2)
Fungi, Bacteria, Yeast
Buforin-2-Bacteria
Section 4.2Goadsporin
  • Azole formation
  • Dehydroalanine formation
  • Leader peptide digestion
  • N-acetylation
Nisin Z
  • Macrocyclization (thioether bond)
  • Non-standard amino acids (e.g., D-alanine)
  • Dehydroalanine, dehydrobutyric acid formation
E. coli, B. subtilis
Section 4.3Salivaricin B
  • Dehydroalanine, dehydrobutyrine, and
  • α-aminobutyric acid formation
Bacteria (S. aureus RN4220)
Valinomycin
  • Non-standard amino acids
    (e.g., D-valine, L-lactate, D-hydroxyvalerate)
  • Cyclic depsipeptide formation
Fungi, Gram-positive bacteria
Gramicidin S
  • Non-standard amino acids
    (e.g., α-aminobutyric acid)
  • Cyclic depsipeptide formation
Fungi, Bacteria
De novo designed PBAs-E. coli (MG1655), B. subtilis (PY79), E. faecium
S. aureus (DSM 11729), K. pneumoniae (DSM 30104), A. baumannii, P. aeruginosa (DSM 1117), Enterobacter spp., Y. pestis (EV76), B. anthracis Sterne, S. pneumoniae (D39)
Meucin-25-Fungi, Bacteria
Cathelicidin-BF-Fungi, Bacteria
δ-lysin-Gram-positive bacteria
Section 4.4SFTI-1
  • Disulfide bond formation
Fungi
Kalata B1
  • Disulfide bond formation
Gram-positive bacteria
AA139
  • Disulfide bond formation
Gram-negative bacteria
HT-1
  • Disulfide bond formation
Bacteria
Capistruin
  • Macrocyclization
  • Leader peptide digestion
Gram-negative bacteria
Thiocillin
  • Dehydroalanine formation
  • Cycloaddition
Fungi, Gram-positive bacteria

5. Concluding Remarks

In this review, we have briefly highlighted several key aspects that, in our view, represent the recent major trends in the development of PBAs (Table 5). We acknowledge that our perspective is influenced by our research interests, and we recognize that some significant milestones in the field may not have been covered. However, our primary goal was to convey the rapid and extensive progress driven by CFSs in the area of PBAs, an area we believe not only lacks comprehensive summaries but also requires further research. Given that CFSs significantly contribute to the advancement of synthetic biology, particularly through their ability to rapidly prototype and screen ideal candidates in a high-throughput manner, this capability is poised to facilitate the identification and optimization of biosynthetic pathways for the effective synthesis of PBAs. Additionally, new chemo-enzymatic reactions tailored to these pathways have been developed, which enabled the successful synthesis of diverse PBAs that were previously unattainable through traditional methods with unique structures and biological properties. These new PBAs are already being utilized in medical practice [44], and many are in the pipelines of several pharmaceutical companies, promising to sustain the momentum for the future.
A promising area of research involves developing PBAs that target multiple pathogens. Significant efforts have been made to create broad-spectrum PBAs. For instance, some studies have focused on designing PBAs that target bacterial membranes, which, despite species-specific differences, share common structural features that can be exploited. However, the effectiveness of membrane-targeting PBAs can be limited by the variability in membrane compositions across different bacterial species. More recent research has shifted towards PBAs that bypass the membrane entirely, targeting intracellular components universally conserved among bacteria, such as ribosomes [63,148,156,207,208] or specific enzymes [209] involved in essential processes like DNA replication [15] or cell wall synthesis [210]. One notable example is the development of antimicrobial peptides that can permeate bacterial membranes and specifically target the bacterial ribosome [211], potentially inhibiting protein synthesis in a manner similar to that of tetracycline, but with broader efficacy against multiple pathogens.
The combination of computational approaches may significantly accelerate the discovery of PBA candidates that meet the desired criteria. Specifically, computational tools can be used to analyze large datasets of peptide sequences, structures, and antimicrobial activities to identify patterns associated with broad-spectrum efficacy. Machine learning models can predict new peptide sequences likely to be effective against multiple pathogens based on existing data. These predictions can then guide the design of de novo peptides, which can be synthesized and tested using CFSs.
Despite the potential of computational models, several challenges remain. For example, ncAAs predicted by computational methods must be both chemically synthesizable and compatible with the wild-type translational machinery or NRPSs. Moreover, if these ncAAs are not accepted by the cellular system, it may be necessary to engineer components of the translational machinery, such as aaRSs, tRNAs, elongation factors, and ribosomes. Additionally, ensuring that PBAs can effectively reach and interact with intracellular targets across different pathogens remains a critical hurdle. In the long-term perspective, these issues can be addressed through the principles of synthetic biology; specifically, the design–build–test–learn approach, particularly when applied within CFSs. These systems provide a flexible platform to refine PBAs for improved intracellular delivery and targeted interaction. Ultimately, we anticipate that integrating these advanced technologies will create a powerful platform for discovering new broad-spectrum PBAs, which are critically needed to address the growing issue of antibiotic resistance. Moreover, these studies will also deepen our understanding of protein synthesis and contribute valuable tools to the field of synthetic biology.

Author Contributions

Conceptualization, H.P., H.J. and J.L.; visualization: H.P., H.J., D.K. and J.L.; writing—original draft preparation, H.P., H.J. and D.K.; writing—review and editing, H.P., H.J., D.K. and J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ministry of Trade, Industry & Energy (MOTIE), grant number 20020231; optimization of structure-based mRNA vaccine production and efficacy evaluation.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors thank the members of the chemical biology laboratory for their critical reading.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Miller, W.R.; Arias, C.A. Eskape pathogens: Antimicrobial resistance, epidemiology, clinical impact and therapeutics. Nat. Rev. Microbiol. 2024, 1–19. [Google Scholar] [CrossRef] [PubMed]
  2. Reardon, S. Who warns against ‘post-antibiotic’ era. Nature 2014, 15, 135–138. [Google Scholar] [CrossRef]
  3. Kwon, J.H.; Powderly, W.G. The post-antibiotic era is here. Science 2021, 373, 471. [Google Scholar] [CrossRef]
  4. World Health Organization. 2021 Antibacterial Agents in Clinical and Preclinical Development: An Overview and Analysis; World Health Organization: Geneva, Switzerland, 2022. [Google Scholar]
  5. Baldwin, J.E.; Abraham, E. The biosynthesis of penicillins and cephalosporins. Nat. Prod. Rep. 1988, 5, 129–145. [Google Scholar] [CrossRef]
  6. Gunsior, M.; Breazeale, S.D.; Lind, A.J.; Ravel, J.; Janc, J.W.; Townsend, C.A. The biosynthetic gene cluster for a monocyclic beta-하lactam antibiotic, nocardicin A. Chem. Biol. 2004, 11, 927–938. [Google Scholar] [CrossRef]
  7. Fleming, A. On the antibacterial action of cultures of a penicillium, with special reference to their use in the isolation of B. Influenzæ. Br. J. Exp. Pathol. 1929, 10, 226–236. [Google Scholar] [CrossRef]
  8. MacNair, C.R.; Rutherford, S.T.; Tan, M.W. Alternative therapeutic strategies to treat antibiotic-resistant pathogens. Nat. Rev. Microbiol. 2024, 22, 262–275. [Google Scholar] [CrossRef]
  9. Chen, E.H.; Wang, C.H.; Liao, Y.T.; Chan, F.Y.; Kanaoka, Y.; Uchihashi, T.; Kato, K.; Lai, L.; Chang, Y.W.; Ho, M.C.; et al. Visualizing the membrane disruption action of antimicrobial peptides by cryo-electron tomography. Nat. Commun. 2023, 14, 5464. [Google Scholar] [CrossRef]
  10. Pierre, J.F.; Peters, B.M.; La Torre, D.; Sidebottom, A.M.; Tao, Y.; Zhu, X.; Cham, C.M.; Wang, L.; Kambal, A.; Harris, K.G.; et al. Peptide yy: A paneth cell antimicrobial peptide that maintains candida gut commensalism. Science 2023, 381, 502–508. [Google Scholar] [CrossRef]
  11. Gidalevitz, D.; Ishitsuka, Y.; Muresan, A.S.; Konovalov, O.; Waring, A.J.; Lehrer, R.I.; Lee, K.Y. Interaction of antimicrobial peptide protegrin with biomembranes. Proc. Natl. Acad. Sci. USA 2003, 100, 6302–6307. [Google Scholar] [CrossRef]
  12. Stephani, J.C.; Gerhards, L.; Khairalla, B.; Solov’yov, I.A.; Brand, I. How do antimicrobial peptides interact with the outer membrane of gram-negative bacteria? Role of lipopolysaccharides in peptide binding, anchoring, and penetration. ACS Infect. Dis. 2024, 10, 763–778. [Google Scholar] [CrossRef]
  13. Melcrova, A.; Maity, S.; Melcr, J.; de Kok, N.A.W.; Gabler, M.; van der Eyden, J.; Stensen, W.; Svendsen, J.S.M.; Driessen, A.J.M.; Marrink, S.J.; et al. Lateral membrane organization as target of an antimicrobial peptidomimetic compound. Nat. Commun. 2023, 14, 4038. [Google Scholar] [CrossRef] [PubMed]
  14. Mukhopadhyay, J.; Sineva, E.; Knight, J.; Levy, R.M.; Ebright, R.H. Antibacterial peptide microcin j25 inhibits transcription by binding within and obstructing the rna polymerase secondary channel. Mol. Cell 2004, 14, 739–751. [Google Scholar] [CrossRef]
  15. Michalczyk, E.; Hommernick, K.; Behroz, I.; Kulike, M.; Pakosz-Stepien, Z.; Mazurek, L.; Seidel, M.; Kunert, M.; Santos, K.; von Moeller, H.; et al. Molecular mechanism of topoisomerase poisoning by the peptide antibiotic albicidin. Nat. Catal. 2023, 6, 52–67. [Google Scholar] [CrossRef] [PubMed]
  16. Sneideris, T.; Erkamp, N.A.; Ausserwoger, H.; Saar, K.L.; Welsh, T.J.; Qian, D.; Katsuya-Gaviria, K.; Johncock, M.; Krainer, G.; Borodavka, A.; et al. Targeting nucleic acid phase transitions as a mechanism of action for antimicrobial peptides. Nat. Commun. 2023, 14, 7170. [Google Scholar] [CrossRef]
  17. Schmidt, N.W.; Jin, F.; Lande, R.; Curk, T.; Xian, W.; Lee, C.; Frasca, L.; Frenkel, D.; Dobnikar, J.; Gilliet, M.; et al. Liquid-crystalline ordering of antimicrobial peptide-DNA complexes controls tlr9 activation. Nat. Mater. 2015, 14, 696–700. [Google Scholar] [CrossRef] [PubMed]
  18. Karathanasi, G.; Bojer, M.S.; Baldry, M.; Johannessen, B.A.; Wolff, S.; Greco, I.; Kilstrup, M.; Hansen, P.R.; Ingmer, H. Linear peptidomimetics as potent antagonists of staphylococcus aureus agr quorum sensing. Sci. Rep. 2018, 8, 3562. [Google Scholar] [CrossRef] [PubMed]
  19. Hirano, M.; Saito, C.; Yokoo, H.; Goto, C.; Kawano, R.; Misawa, T.; Demizu, Y. Development of antimicrobial stapled peptides based on magainin 2 sequence. Molecules 2021, 26, 444. [Google Scholar] [CrossRef]
  20. Giacometti, A.; Cirioni, O.; Del Prete, M.S.; Paggi, A.M.; D’Errico, M.M.; Scalise, G. Combination studies between polycationic peptides and clinically used antibiotics against gram-positive and gram-negative bacteria. Peptides 2000, 21, 1155–1160. [Google Scholar] [CrossRef]
  21. Galdiero, E.; Siciliano, A.; Gesuele, R.; Di Onofrio, V.; Falanga, A.; Maione, A.; Liguori, R.; Libralato, G.; Guida, M. Melittin inhibition and eradication activity for resistant polymicrobial biofilm isolated from a dairy industry after disinfection. Int. J. Microbiol. 2019, 2019, 4012394. [Google Scholar] [CrossRef]
  22. Lima, W.G.; de Brito, J.C.M.; Cardoso, V.N.; Fernandes, S.O.A. In-depth characterization of antibacterial activity of melittin against staphylococcus aureus and use in a model of non-surgical mrsa-infected skin wounds. Eur. J. Pharm. Sci. 2021, 156, 105592. [Google Scholar] [CrossRef] [PubMed]
  23. Diederen, B.M.; van Duijn, I.; Willemse, P.; Kluytmans, J.A. In vitro activity of daptomycin against methicillin-resistant staphylococcus aureus, including heterogeneously glycopeptide-resistant strains. Antimicrob. Agents Chemother. 2006, 50, 3189–3191. [Google Scholar] [CrossRef] [PubMed]
  24. Feng, X.; Sambanthamoorthy, K.; Palys, T.; Paranavitana, C. The human antimicrobial peptide ll-37 and its fragments possess both antimicrobial and antibiofilm activities against multidrug-resistant acinetobacter baumannii. Peptides 2013, 49, 131–137. [Google Scholar] [CrossRef]
  25. Turner, J.; Cho, Y.; Dinh, N.N.; Waring, A.J.; Lehrer, R.I. Activities of ll-37, a cathelin-associated antimicrobial peptide of human neutrophils. Antimicrob. Agents Chemother. 1998, 42, 2206–2214. [Google Scholar] [CrossRef] [PubMed]
  26. Neshani, A.; Zare, H.; Eidgahi, M.R.A.; Kakhki, R.K.; Safdari, H.; Khaledi, A.; Ghazvini, K. Ll-37: Review of antimicrobial profile against sensitive and antibiotic-resistant human bacterial pathogens. Gene Rep. 2019, 17, 100519. [Google Scholar] [CrossRef]
  27. Kim, M.K.; Kang, N.; Ko, S.J.; Park, J.; Park, E.; Shin, D.W.; Kim, S.H.; Lee, S.A.; Lee, J.I.; Lee, S.H.; et al. Antibacterial and antibiofilm activity and mode of action of magainin 2 against drug-resistant acinetobacter baumannii. Int. J. Mol. Sci. 2018, 19, 3041. [Google Scholar] [CrossRef]
  28. Gu, H.; Kato, T.; Kumeta, H.; Kumaki, Y.; Tsukamoto, T.; Kikukawa, T.; Demura, M.; Ishida, H.; Vogel, H.J.; Aizawa, T. Three-dimensional structure of the antimicrobial peptide cecropin p1 in dodecylphosphocholine micelles and the role of the c-terminal residues. ACS Omega 2022, 7, 31924–31934. [Google Scholar] [CrossRef]
  29. Rengaraj, R.; Mariappan, S.; Sekar, U.; Kamalanadhan, A. Detection of vancomycin resistance among enterococcus faecalis and staphylococcus aureus. J. Clin. Diagn. Res. 2016, 10, DC04-6. [Google Scholar] [CrossRef] [PubMed]
  30. Campbell, M.L.; Marchaim, D.; Pogue, J.M.; Sunkara, B.; Bheemreddy, S.; Bathina, P.; Pulluru, H.; Chugh, N.; Wilson, M.N.; Moshos, J.; et al. Treatment of methicillin-resistant staphylococcus aureus infections with a minimal inhibitory concentration of 2 mug/ml to vancomycin: Old (trimethoprim/sulfamethoxazole) versus new (daptomycin or linezolid) agents. Ann. Pharmacother. 2012, 46, 1587–1597. [Google Scholar] [CrossRef] [PubMed]
  31. AlQahtani, H.; Baloch, S.; Aleanizy, F.S.; Alqahtani, F.Y.; Tabb, D. Influence of the minimum inhibitory concentration of daptomycin on the outcomes of staphylococcus aureus bacteraemia. J. Glob. Antimicrob. Resist. 2021, 24, 23–26. [Google Scholar] [CrossRef] [PubMed]
  32. Karyne, R.; Curty Lechuga, G.; Almeida Souza, A.L.; Rangel da Silva Carvalho, J.P.; Simoes Villas Boas, M.H.; De Simone, S.G. Pan-drug resistant acinetobacter baumannii, but not other strains, are resistant to the bee venom peptide mellitin. Antibiotics 2020, 9, 178. [Google Scholar]
  33. Zavascki, A.P.; Goldani, L.Z.; Li, J.; Nation, R.L. Polymyxin B for the treatment of multidrug-resistant pathogens: A critical review. J. Antimicrob. Chemother. 2007, 60, 1206–1215. [Google Scholar] [CrossRef]
  34. Shariati, A.; Arshadi, M.; Khosrojerdi, M.A.; Abedinzadeh, M.; Ganjalishahi, M.; Maleki, A.; Heidary, M.; Khoshnood, S. The resistance mechanisms of bacteria against ciprofloxacin and new approaches for enhancing the efficacy of this antibiotic. Front. Public Health 2022, 10, 1025633. [Google Scholar] [CrossRef]
  35. Jakobsen, L.; Sandvang, D.; Jensen, V.F.; Seyfarth, A.M.; Frimodt-Moller, N.; Hammerum, A.M. Gentamicin susceptibility in escherichia coli related to the genetic background: Problems with breakpoints. Clin. Microbiol. Infect. 2007, 13, 830–832. [Google Scholar] [CrossRef]
  36. Grossman, T.H. Tetracycline antibiotics and resistance. Cold Spring Harb. Perspect. Med. 2016, 6, a025387. [Google Scholar] [CrossRef]
  37. De Oliveira, D.M.P.; Forde, B.M.; Phan, M.D.; Steiner, B.; Zhang, B.; Zuegg, J.; El-Deeb, I.M.; Li, G.; Keller, N.; Brouwer, S.; et al. Rescuing tetracycline class antibiotics for the treatment of multidrug-resistant acinetobacter baumannii pulmonary infection. MBio 2022, 13, e0351721. [Google Scholar] [CrossRef]
  38. Sadybekov, A.V.; Katritch, V. Computational approaches streamlining drug discovery. Nature 2023, 616, 673–685. [Google Scholar] [CrossRef]
  39. Montalban-Lopez, M.; van Heel, A.J.; Kuipers, O.P. Employing the promiscuity of lantibiotic biosynthetic machineries to produce novel antimicrobials. Fems Microbiol. Rev. 2017, 41, 5–18. [Google Scholar] [CrossRef] [PubMed]
  40. Hemmerling, F.; Piel, J. Strategies to access biosynthetic novelty in bacterial genomes for drug discovery. Nat. Rev. Drug Discov. 2022, 21, 359–378. [Google Scholar] [CrossRef]
  41. Antimicrobial Resistance, C. Global burden of bacterial antimicrobial resistance in 2019: A systematic analysis. Lancet 2022, 399, 629–655. [Google Scholar]
  42. Mookherjee, N.; Anderson, M.A.; Haagsman, H.P.; Davidson, D.J. Antimicrobial host defence peptides: Functions and clinical potential. Nat. Rev. Drug Discov. 2020, 19, 311–332. [Google Scholar] [CrossRef]
  43. Sayburn, A. Only 12 new antibiotics entered the market in five years, who warns in review. Pharm. J. 2023, 310, 7971. [Google Scholar]
  44. PhRMA. Medicines in Development for Cancer 2020 Report; PhRMA: Washington, DC, USA, 2020. [Google Scholar]
  45. Garenne, D.; Haines, M.C.; Romantseva, E.F.; Freemont, P.; Strychalski, E.A.; Noireaux, V. Cell-free gene expression. Nat. Rev. Method Prim. 2021, 1, 49. [Google Scholar] [CrossRef]
  46. Silverman, A.D.; Karim, A.S.; Jewett, M.C. Cell-free gene expression: An expanded repertoire of applications. Nat. Rev. Genet. 2020, 21, 151–170. [Google Scholar] [CrossRef]
  47. Kofman, C.; Lee, J.; Jewett, M.C. Engineering molecular translation systems. Cell Syst. 2021, 12, 593–607. [Google Scholar] [CrossRef]
  48. Inoue, S.; Nguyen, D.T.; Hamada, K.; Okuma, R.; Okada, C.; Okada, M.; Abe, I.; Sengoku, T.; Goto, Y.; Suga, H. De novo discovery of pseudo-natural prenylated macrocyclic peptide ligands. Angew. Chem. Int. Ed. 2024, e202409973. [Google Scholar] [CrossRef]
  49. Kawakami, T.; Ohta, A.; Ohuchi, M.; Ashigai, H.; Murakami, H.; Suga, H. Diverse backbone-cyclized peptides via codon reprogramming. Nat. Chem. Biol. 2009, 5, 888–890. [Google Scholar] [CrossRef]
  50. Lee, J.; Schwarz, K.J.; Kim, D.S.; Moore, J.S.; Jewett, M.C. Ribosome-mediated polymerization of long chain carbon and cyclic amino acids into peptides in vitro. Nat. Commun. 2020, 11, 4304. [Google Scholar] [CrossRef]
  51. Lee, K.; Willi, J.A.; Cho, N.; Kim, I.; Jewett, M.C.; Lee, J. Cell-free biosynthesis of peptidomimetics. Biotechnol. Bioprocess Eng. 2023, 28, 905–921. [Google Scholar] [CrossRef]
  52. Walsh, C. Antibiotics: Actions, Origins, Resistance; Harvard Medical School: Boston, MA, USA, 2003. [Google Scholar]
  53. Schwarzer, D.; Finking, R.; Marahiel, M.A. Nonribosomal peptides: From genes to products. Nat. Prod. Rep. 2003, 20, 275–287. [Google Scholar] [CrossRef]
  54. Marahiel, M.A.; Stachelhaus, T.; Mootz, H.D. Modular peptide synthetases involved in nonribosomal peptide synthesis. Chem. Rev. 1997, 97, 2651–2674. [Google Scholar] [CrossRef]
  55. Zasloff, M. Magainins, a class of antimicrobial peptides from xenopus skin: Isolation, characterization of two active forms, and partial cdna sequence of a precursor. Proc. Natl. Acad. Sci. USA 1987, 84, 5449–5453. [Google Scholar] [CrossRef]
  56. Alalwani, S.M.; Sierigk, J.; Herr, C.; Pinkenburg, O.; Gallo, R.; Vogelmeier, C.; Bals, R. The antimicrobial peptide ll-37 modulates the inflammatory and host defense response of human neutrophils. Eur. J. Immunol. 2010, 40, 1118–1126. [Google Scholar] [CrossRef]
  57. Nizet, V.; Ohtake, T.; Lauth, X.; Trowbridge, J.; Rudisill, J.; Dorschner, R.A.; Pestonjamasp, V.; Piraino, J.; Huttner, K.; Gallo, R.L. Innate antimicrobial peptide protects the skin from invasive bacterial infection. Nature 2001, 414, 454–457. [Google Scholar] [CrossRef]
  58. Hoelscher, M.P.; Forner, J.; Calderone, S.; Kramer, C.; Taylor, Z.; Loiacono, F.V.; Agrawal, S.; Karcher, D.; Moratti, F.; Kroop, X.; et al. Expression strategies for the efficient synthesis of antimicrobial peptides in plastids. Nat. Commun. 2022, 13, 5856. [Google Scholar] [CrossRef]
  59. Deo, S.; Turton, K.L.; Kainth, T.; Kumar, A.; Wieden, H.J. Strategies for improving antimicrobial peptide production. Biotechnol. Adv. 2022, 59, 107968. [Google Scholar] [CrossRef]
  60. Hancock, R.E.W.; Sahl, H.-G. Antimicrobial and host-defense peptides as new anti-infective therapeutic strategies. Nat. Biotechnol. 2006, 24, 1551–1557. [Google Scholar] [CrossRef]
  61. Zasloff, M. Antimicrobial peptides of multicellular organisms. Nature 2002, 415, 389–395. [Google Scholar] [CrossRef]
  62. King, A.M.; Zhang, Z.; Glassey, E.; Siuti, P.; Clardy, J.; Voigt, C.A. Systematic mining of the human microbiome identifies antimicrobial peptides with diverse activity spectra. Nat. Microbiol. 2023, 8, 2420–2434. [Google Scholar] [CrossRef]
  63. Mangano, K.; Klepacki, D.; Ohanmu, I.; Baliga, C.; Huang, W.; Brakel, A.; Krizsan, A.; Polikanov, Y.S.; Hoffmann, R.; Vazquez-Laslop, N.; et al. Inhibition of translation termination by the antimicrobial peptide drosocin. Nat. Chem. Biol. 2023, 19, 1082–1090. [Google Scholar] [CrossRef]
  64. Engelberg, Y.; Landau, M. The human ll-37(17-29) antimicrobial peptide reveals a functional supramolecular structure. Nat. Commun. 2020, 11, 3894. [Google Scholar] [CrossRef] [PubMed]
  65. Florin, T.; Maracci, C.; Graf, M.; Karki, P.; Klepacki, D.; Berninghausen, O.; Beckmann, R.; Vazquez-Laslop, N.; Wilson, D.N.; Rodnina, M.V.; et al. An antimicrobial peptide that inhibits translation by trapping release factors on the ribosome. Nat. Struct. Mol. Biol. 2017, 24, 752–757. [Google Scholar] [CrossRef] [PubMed]
  66. Chan, D.I.; Prenner, E.J.; Vogel, H.J. Tryptophan- and arginine-rich antimicrobial peptides: Structures and mechanisms of action. Biochim. Biophys. Acta 2006, 1758, 1184–1202. [Google Scholar] [CrossRef] [PubMed]
  67. Lòpez-Oyama, A.B. Interaction of the cationic peptide bactencin with ddpc/dmpg phospholipid mixtures at the air-water interface. Biophys. J. 2010, 98, 84a. [Google Scholar] [CrossRef]
  68. Robinson, P.J.; Bulleid, N.J. Mechanisms of disulfide bond formation in nascent polypeptides entering the secretory pathway. Cells 2020, 9, 1994. [Google Scholar] [CrossRef] [PubMed]
  69. Nawae, W.; Hannongbua, S.; Ruengjitchatchawalya, M. Defining the membrane disruption mechanism of kalata b1 via coarse-grained molecular dynamics simulations. Sci. Rep. 2014, 4, 3933. [Google Scholar] [CrossRef]
  70. Korsinczky, M.L.; Clark, R.J.; Craik, D.J. Disulfide bond mutagenesis and the structure and function of the head-to-tail macrocyclic trypsin inhibitor sfti-1. Biochemistry 2005, 44, 1145–1153. [Google Scholar] [CrossRef] [PubMed]
  71. Mattiuzzo, M.; Bandiera, A.; Gennaro, R.; Benincasa, M.; Pacor, S.; Antcheva, N.; Scocchi, M. Role of the escherichia coli sbma in the antimicrobial activity of proline-rich peptides. Mol. Microbiol. 2007, 66, 151–163. [Google Scholar] [CrossRef]
  72. Lointier, M.; Aisenbrey, C.; Marquette, A.; Tan, J.H.; Kichler, A.; Bechinger, B. Membrane pore-formation correlates with the hydrophilic angle of histidine-rich amphipathic peptides with multiple biological activities. Biochim. Biophys. Acta-Biomembr. 2020, 1862, 183212. [Google Scholar] [CrossRef]
  73. Seefeldt, A.C.; Nguyen, F.; Antunes, S.; Perebaskine, N.; Graf, M.; Arenz, S.; Inampudi, K.K.; Douat, C.; Guichard, G.; Wilson, D.N.; et al. The proline-rich antimicrobial peptide onc112 inhibits translation by blocking and destabilizing the initiation complex. Nat. Struct. Mol. Biol. 2015, 22, 470–475. [Google Scholar] [CrossRef] [PubMed]
  74. Mardirossian, M.; Perebaskine, N.; Benincasa, M.; Gambato, S.; Hofmann, S.; Huter, P.; Muller, C.; Hilpert, K.; Innis, C.A.; Tossi, A.; et al. The dolphin proline-rich antimicrobial peptide tur1a inhibits protein synthesis by targeting the bacterial ribosome. Cell Chem. Biol. 2018, 25, 530–539.e7. [Google Scholar] [CrossRef] [PubMed]
  75. Cabalteja, C.C.; Lin, Q.; Harmon, T.W.; Rao, S.R.; Di, Y.P.; Horne, W.S. Heterogeneous-backbone proteomimetic analogues of lasiocepsin, a disulfide-rich antimicrobial peptide with a compact tertiary fold. ACS Chem. Biol. 2022, 17, 987–997. [Google Scholar] [CrossRef] [PubMed]
  76. Mourenza, A.; Ganesan, R.; Camarero, J.A. Resistance is futile: Targeting multidrug-resistant bacteria with de novo cys-rich cyclic polypeptides. RSC Chem. Biol. 2023, 4, 722–735. [Google Scholar] [CrossRef] [PubMed]
  77. Gongora-Benitez, M.; Tulla-Puche, J.; Albericio, F. Multifaceted roles of disulfide bonds. Peptides as therapeutics. Chem. Rev. 2014, 114, 901–926. [Google Scholar] [CrossRef] [PubMed]
  78. Wang, C.K.; Craik, D.J. Designing macrocyclic disulfide-rich peptides for biotechnological applications. Nat. Chem. Biol. 2018, 14, 417–427. [Google Scholar] [CrossRef]
  79. Colgrave, M.L.; Craik, D.J. Thermal, chemical, and enzymatic stability of the cyclotide kalata B1: The importance of the cyclic cystine knot. Biochemistry 2004, 43, 5965–5975. [Google Scholar] [CrossRef]
  80. Clark, R.J.; Jensen, J.; Nevin, S.T.; Callaghan, B.P.; Adams, D.J.; Craik, D.J. The engineering of an orally active conotoxin for the treatment of neuropathic pain. Angew. Chem. Int. Ed. 2010, 49, 6545–6548. [Google Scholar] [CrossRef]
  81. Travin, D.Y.; Watson, Z.L.; Metelev, M.; Ward, F.R.; Osterman, I.A.; Khven, I.M.; Khabibullina, N.F.; Serebryakova, M.; Mergaert, P.; Polikanov, Y.S.; et al. Structure of ribosome-bound azole-modified peptide phazolicin rationalizes its species-specific mode of bacterial translation inhibition. Nat. Commun. 2019, 10, 4563. [Google Scholar] [CrossRef]
  82. Parajuli, A.; Kwak, D.H.; Dalponte, L.; Leikoski, N.; Galica, T.; Umeobika, U.; Trembleau, L.; Bent, A.; Sivonen, K.; Wahlsten, M. A unique tryptophan c-prenyltransferase from the kawaguchipeptin biosynthetic pathway. Angew. Chem. Int. Ed. 2016, 55, 3596–3599. [Google Scholar] [CrossRef]
  83. Claesen, J.; Bibb, M. Genome mining and genetic analysis of cypemycin biosynthesis reveal an unusual class of posttranslationally modified peptides. Proc. Natl. Acad. Sci. USA 2010, 107, 16297–16302. [Google Scholar] [CrossRef]
  84. 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. A scalable platform to discover antimicrobials of ribosomal origin. Nat. Commun. 2022, 13, 6135. [Google Scholar] [CrossRef]
  85. Haste, N.M.; Thienphrapa, W.; Tran, D.N.; Loesgen, S.; Sun, P.; Nam, S.-J.; Jensen, P.R.; Fenical, W.; Sakoulas, G.; Nizet, V. Activity of the thiopeptide antibiotic nosiheptide against contemporary strains of methicillin-resistant staphylococcus aureus. J. Antibiot. 2012, 65, 593–598. [Google Scholar] [CrossRef] [PubMed]
  86. Hegemann, J.D.; Zimmermann, M.; Xie, X.; Marahiel, M.A. Lasso peptides: An intriguing class of bacterial natural products. Acc. Chem. Res. 2015, 48, 1909–1919. [Google Scholar] [CrossRef]
  87. Rimal, B.; Chang, J.; Liu, C.; Rashid, R.; Singh, M.; Kim, S.J. The effects of daptomycin on cell wall biosynthesis in enterococcal faecalis. Sci. Rep. 2023, 13, 12227. [Google Scholar] [CrossRef]
  88. Rivera-Pérez, C.; Ponce González, X.P.; Hernández-Savedra, N.Y. Antimicrobial and anticarcinogenic activity of bioactive peptides derived from abalone viscera (haliotis fulgens and haliotis corrugata). Sci. Rep. 2023, 13, 15185. [Google Scholar] [CrossRef] [PubMed]
  89. Nicolaou, K.; Boddy, C.N.; Bräse, S.; Winssinger, N. Chemistry, biology, and medicine of the glycopeptide antibiotics. Angew. Chem. Int. Ed. 1999, 38, 2096–2152. [Google Scholar] [CrossRef]
  90. Baltz, R.H. Function of mbth homologs in nonribosomal peptide biosynthesis and applications in secondary metabolite discovery. J. Ind. Microbiol. Biotechnol. 2011, 38, 1747. [Google Scholar] [CrossRef]
  91. Franke, J.; Ishida, K.; Ishida-Ito, M.; Hertweck, C. Nitro versus hydroxamate in siderophores of pathogenic bacteria: Effect of missing hydroxylamine protection in malleobactin biosynthesis. Angew. Chem. Int. Ed. 2013, 125, 8429. [Google Scholar] [CrossRef]
  92. Repka, L.M.; Chekan, J.R.; Nair, S.K.; van der Donk, W.A. Mechanistic understanding of lanthipeptide biosynthetic enzymes. Chem. Rev. 2017, 117, 5457–5520. [Google Scholar] [CrossRef]
  93. Ortega, M.A.; van der Donk, W.A. New insights into the biosynthetic logic of ribosomally synthesized and post-translationally modified peptide natural products. Cell Chem. Biol. 2016, 23, 31–44. [Google Scholar] [CrossRef]
  94. Shi, C.; Patel, V.A.; Mitchell, D.A.; Zhao, H. Enterolyin s, a polythiazole-containing hemolytic peptide from enterococcus caccae. ChemBioChem 2024, 25, e202400212. [Google Scholar] [CrossRef] [PubMed]
  95. Letzel, A.C.; Pidot, S.J.; Hertweck, C. Genome mining for ribosomally synthesized and post-translationally modified peptides (ripps) in anaerobic bacteria. BMC Genom. 2014, 15, 983. [Google Scholar] [CrossRef]
  96. Bednarska, N.G.; Wren, B.W.; Willcocks, S.J. The importance of the glycosylation of antimicrobial peptides: Natural and synthetic approaches. Drug Discov. Today 2017, 22, 919–926. [Google Scholar] [CrossRef] [PubMed]
  97. Rudenko, A.Y.; Mariasina, S.S.; Sergiev, P.V.; Polshakov, V.I. Analogs of s-adenosyl-l-methionine in studies of methyltransferases. Mol. Biol. 2022, 56, 229–250. [Google Scholar] [CrossRef]
  98. Winn, M.; Fyans, J.K.; Zhuo, Y.; Micklefield, J. Recent advances in engineering nonribosomal peptide assembly lines. Nat. Prod. Rep. 2016, 33, 317–347. [Google Scholar] [CrossRef] [PubMed]
  99. Brytan, W.; Padrela, L. Structural modifications for the conversion of proteins and peptides into stable dried powder formulations: A review. J. Drug Deliv. Sci. Technol. 2023, 89, 104992. [Google Scholar] [CrossRef]
  100. Sussmuth, R.D.; Mainz, A. Nonribosomal peptide synthesis-principles and prospects. Angew. Chem. Int. Ed. 2017, 56, 3770–3821. [Google Scholar] [CrossRef]
  101. Weissman, K.J. The structural biology of biosynthetic megaenzymes. Nat. Chem. Biol. 2015, 11, 660–670. [Google Scholar] [CrossRef]
  102. Brakhage, A.A. Regulation of fungal secondary metabolism. Nat. Rev. Microbiol. 2013, 11, 21–32. [Google Scholar] [CrossRef]
  103. Sieber, S.A.; Marahiel, M.A. Molecular mechanisms underlying nonribosomal peptide synthesis: Approaches to new antibiotics. Chem. Rev. 2005, 105, 715–738. [Google Scholar] [CrossRef] [PubMed]
  104. Atanasov, A.G.; Waltenberger, B.; Pferschy-Wenzig, E.M.; Linder, T.; Wawrosch, C.; Uhrin, P.; Temml, V.; Wang, L.; Schwaiger, S.; Heiss, E.H.; et al. Discovery and resupply of pharmacologically active plant-derived natural products: A review. Biotechnol. Adv. 2015, 33, 1582–1614. [Google Scholar] [CrossRef]
  105. Folger, I.B.; Frota, N.F.; Pistofidis, A.; Niquille, D.L.; Hansen, D.A.; Schmeing, T.M.; Hilvert, D. High-throughput reprogramming of an nrps condensation domain. Nat. Chem. Biol. 2024, 20, 761–769. [Google Scholar] [CrossRef]
  106. Scott, T.A.; Heine, D.; Qin, Z.; Wilkinson, B. An l-threonine transaldolase is required for l-threo-beta-hydroxy-alpha-amino acid assembly during obafluorin biosynthesis. Nat. Commun. 2017, 8, 15935. [Google Scholar] [CrossRef]
  107. Awakawa, T.; Fujioka, T.; Zhang, L.; Hoshino, S.; Hu, Z.; Hashimoto, J.; Kozone, I.; Ikeda, H.; Shin-Ya, K.; Liu, W.; et al. Reprogramming of the antimycin nrps-pks assembly lines inspired by gene evolution. Nat. Commun. 2018, 9, 3534. [Google Scholar] [CrossRef] [PubMed]
  108. Niquille, D.L.; Hansen, D.A.; Mori, T.; Fercher, D.; Kries, H.; Hilvert, D. Nonribosomal biosynthesis of backbone-modified peptides. Nat. Chem. 2018, 10, 282–287. [Google Scholar] [CrossRef]
  109. Wang, Z.; Koirala, B.; Hernandez, Y.; Zimmerman, M.; Park, S.; Perlin, D.S.; Brady, S.F. A naturally inspired antibiotic to target multidrug-resistant pathogens. Nature 2022, 601, 606–611. [Google Scholar] [CrossRef] [PubMed]
  110. Alonzo, D.A.; Schmeing, T.M. Biosynthesis of depsipeptides, or depsi: The peptides with varied generations. Protein Sci. 2020, 29, 2316–2347. [Google Scholar] [CrossRef] [PubMed]
  111. Konz, D.; Klens, A.; Schorgendorfer, K.; Marahiel, M.A. The bacitracin biosynthesis operon of bacillus licheniformis atcc 10716: Molecular characterization of three multi-modular peptide synthetases. Chem. Biol. 1997, 4, 927–937. [Google Scholar] [CrossRef] [PubMed]
  112. Wang, W.; Khojasteh, S.C.; Su, D. Biosynthetic strategies for macrocyclic peptides. Molecules 2021, 26, 3338. [Google Scholar] [CrossRef]
  113. Bonjorno, A.F.; Pavan, A.R.; Fernandes, G.F.S.; Scarim, C.B.; Castagnolo, D.; Dos Santos, J.L. Bacprotacs targeting clp protease: A promising strategy for anti-mycobacterial drug discovery. Front. Chem. 2024, 12, 1358539. [Google Scholar] [CrossRef]
  114. Upert, G.; Luther, A.; Obrecht, D.; Ermert, P. Emerging peptide antibiotics with therapeutic potential. Med. Drug. Discov. 2021, 9, 100078. [Google Scholar] [CrossRef]
  115. Shi, J.; Chen, C.; Wang, D.; Wang, Z.; Liu, Y. The antimicrobial peptide li14 combats multidrug-resistant bacterial infections. Commun. Biol. 2022, 5, 926. [Google Scholar] [CrossRef]
  116. Randall, J.R.; Vieira, L.C.; Wilke, C.O.; Davies, B.W. Deep mutational scanning and machine learning for the analysis of antimicrobial-peptide features driving membrane selectivity. Nat. Biomed. Eng. 2024, 8, 842–853. [Google Scholar] [CrossRef]
  117. Levine, M.Z.; Gregorio, N.E.; Jewett, M.C.; Watts, K.R.; Oza, J.P. Escherichia coli-based cell-free protein synthesis: Protocols for a robust, flexible, and accessible platform technology. J. Vis. Exp. 2019, 144, e58882. [Google Scholar]
  118. Kwon, Y.C.; Jewett, M.C. High-throughput preparation methods of crude extract for robust cell-free protein synthesis. Sci. Rep. 2015, 5, 8663. [Google Scholar] [CrossRef] [PubMed]
  119. Jaroentomeechai, T.; Stark, J.C.; Natarajan, A.; Glasscock, C.J.; Yates, L.E.; Hsu, K.J.; Mrksich, M.; Jewett, M.C.; DeLisa, M.P. Single-pot glycoprotein biosynthesis using a cell-free transcription-translation system enriched with glycosylation machinery. Nat. Commun. 2018, 9, 2686. [Google Scholar] [CrossRef] [PubMed]
  120. Rasor, B.J.; Karim, A.S.; Alper, H.S.; Jewett, M.C. Cell extracts from bacteria and yeast retain metabolic activity after extended storage and repeated thawing. ACS Synth. Biol. 2023, 12, 904–908. [Google Scholar] [CrossRef]
  121. Didovyk, A.; Tonooka, T.; Tsimring, L.; Hasty, J. Rapid and scalable preparation of bacterial lysates for cell-free gene expression. ACS Synth. Biol. 2017, 6, 2198–2208. [Google Scholar] [CrossRef]
  122. Shin, J.; Noireaux, V. Efficient cell-free expression with the endogenous E. Coli RNA polymerase and sigma factor 70. J. Biol. Eng. 2010, 4, 8. [Google Scholar] [CrossRef]
  123. Pardee, K.; Slomovic, S.; Nguyen, P.Q.; Lee, J.W.; Donghia, N.; Burrill, D.; Ferrante, T.; McSorley, F.R.; Furuta, Y.; Vernet, A. Portable, on-demand biomolecular manufacturing. Cell 2016, 167, 248–259.e12. [Google Scholar] [CrossRef] [PubMed]
  124. Des Soye, B.J.; Davidson, S.R.; Weinstock, M.T.; Gibson, D.G.; Jewett, M.C. Establishing a high-yielding cell-free protein synthesis platform derived from vibrio natriegens. ACS Synth. Biol. 2018, 7, 2245–2255. [Google Scholar] [CrossRef] [PubMed]
  125. Kelwick, R.; Webb, A.J.; MacDonald, J.T.; Freemont, P.S. Development of a bacillus subtilis cell-free transcription-translation system for prototyping regulatory elements. Metab. Eng. 2016, 38, 370–381. [Google Scholar] [CrossRef] [PubMed]
  126. Des Soye, B.J.; Gerbasi, V.R.; Thomas, P.M.; Kelleher, N.L.; Jewett, M.C. A highly productive, one-pot cell-free protein synthesis platform based on genomically recoded escherichia coli. Cell Chem. Biol. 2019, 26, 1743–1754.e9. [Google Scholar] [CrossRef] [PubMed]
  127. Bowie, J.U.; Sherkhanov, S.; Korman, T.P.; Valliere, M.A.; Opgenorth, P.H.; Liu, H. Synthetic biochemistry: The bio-inspired cell-free approach to commodity chemical production. Trends Biotechnol. 2020, 38, 766–778. [Google Scholar] [CrossRef]
  128. Thoring, L.; Dondapati, S.K.; Stech, M.; Wustenhagen, D.A.; Kubick, S. High-yield production of "difficult-to-express" proteins in a continuous exchange cell-free system based on cho cell lysates. Sci. Rep. 2017, 7, 11710. [Google Scholar] [CrossRef]
  129. Takahashi, H.; Nozawa, A.; Seki, M.; Shinozaki, K.; Endo, Y.; Sawasaki, T. A simple and high-sensitivity method for analysis of ubiquitination and polyubiquitination based on wheat cell-free protein synthesis. BMC Plant Biol. 2009, 9, 39. [Google Scholar] [CrossRef]
  130. Ramadan, A.; Nemoto, K.; Seki, M.; Shinozaki, K.; Takeda, H.; Takahashi, H.; Sawasaki, T. Wheat germ-based protein libraries for the functional characterisation of the arabidopsis e2 ubiquitin conjugating enzymes and the ring-type e3 ubiquitin ligase enzymes. BMC Plant Biol. 2015, 15, 275. [Google Scholar] [CrossRef]
  131. Cheng, Z.; He, B.B.; Lei, K.; Gao, Y.; Shi, Y.; Zhong, Z.; Liu, H.; Liu, R.; Zhang, H.; Wu, S.; et al. Rule-based omics mining reveals antimicrobial macrocyclic peptides against drug-resistant clinical isolates. Nat. Commun. 2024, 15, 4901. [Google Scholar] [CrossRef]
  132. Cheng, Y.; Qin, J.; Huang, Y.; Wang, T. The antimicrobial effects of plga microspheres containing the antimicrobial peptide op-145 on clinically isolated pathogens in bone infections. Sci. Rep. 2022, 12, 14541. [Google Scholar] [CrossRef]
  133. Hilpert, K.; Volkmer-Engert, R.; Walter, T.; Hancock, R.E. High-throughput generation of small antibacterial peptides with improved activity. Nat. Biotechnol. 2005, 23, 1008–1012. [Google Scholar] [CrossRef]
  134. Schoborg, J.A.; Hodgman, C.E.; Anderson, M.J.; Jewett, M.C. Substrate replenishment and byproduct removal improve yeast cell-free protein synthesis. Biotechnol. J. 2014, 9, 630–640. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, X.; Liu, J.; Zheng, Y.; Li, J.; Wang, H.; Zhou, Y.; Qi, M.; Yu, H.; Tang, W.; Zhao, W.M. An optimized yeast cell-free system: Sufficient for translation of human papillomavirus 58 L1 mRNA and assembly of virus-like particles. J. Biosci. Bioeng. 2008, 106, 8–15. [Google Scholar] [CrossRef]
  136. Madin, K.; Sawasaki, T.; Ogasawara, T.; Endo, Y. A highly efficient and robust cell-free protein synthesis system prepared from wheat embryos: Plants apparently contain a suicide system directed at ribosomes. Proc. Natl. Acad. Sci. USA 2000, 97, 559–564. [Google Scholar] [CrossRef] [PubMed]
  137. Martin, R.W.; Majewska, N.I.; Chen, C.X.; Albanetti, T.E.; Jimenez, R.B.C.; Schmelzer, A.E.; Jewett, M.C.; Roy, V. Development of a cho-based cell-free platform for synthesis of active monoclonal antibodies. ACS Synth. Biol. 2017, 6, 1370–1379. [Google Scholar] [CrossRef]
  138. Shimizu, Y.; Inoue, A.; Tomari, Y.; Suzuki, T.; Yokogawa, T.; Nishikawa, K.; Ueda, T. Cell-free translation reconstituted with purified components. Nat. Biotechnol. 2001, 19, 751–755. [Google Scholar] [CrossRef] [PubMed]
  139. Liu, W.Q.; Ji, X.; Ba, F.; Zhang, Y.; Xu, H.; Huang, S.; Zheng, X.; Liu, Y.; Ling, S.; Jewett, M.C.; et al. Cell-free biosynthesis and engineering of ribosomally synthesized lanthipeptides. Nat. Commun. 2024, 15, 4336. [Google Scholar] [CrossRef]
  140. Sin, S.T.; Deng, J.; Ji, L.; Yukawa, M.; Chan, R.W.; Volpi, S.; Vaglio, A.; Fenaroli, P.; Bocca, P.; Cheng, S.H.; et al. Effects of nucleases on cell-free extrachromosomal circular DNA. JCI Insight 2022, 7, e156070. [Google Scholar] [CrossRef]
  141. Tuckey, C.; Asahara, H.; Zhou, Y.; Chong, S. Protein synthesis using a reconstituted cell-free system. Curr. Protoc. Mol. Biol. 2014, 108, 16.31.1–16.31.22. [Google Scholar] [CrossRef]
  142. Gabant, P.; Borrero, J. Paragen 1.0: A standardized synthetic gene library for fast cell-free bacteriocin synthesis. Front. Bioeng. Biotechnol. 2019, 7, 213. [Google Scholar] [CrossRef]
  143. Passioura, T.; Katoh, T.; Goto, Y.; Suga, H. Selection-based discovery of druglike macrocyclic peptides. Annu. Rev. Biochem. 2014, 83, 727–752. [Google Scholar] [CrossRef]
  144. Lee, J.; Coronado, J.N.; Cho, N.; Lim, J.; Hosford, B.M.; Seo, S.; Kim, D.S.; Kofman, C.; Moore, J.S.; Ellington, A.D. Ribosome-mediated biosynthesis of pyridazinone oligomers in vitro. Nat. Commun. 2022, 13, 6322. [Google Scholar] [CrossRef] [PubMed]
  145. Lee, J.; Schwieter, K.E.; Watkins, A.M.; Kim, D.S.; Yu, H.; Schwarz, K.J.; Lim, J.; Coronado, J.; Byrom, M.; Anslyn, E.V. Expanding the limits of the second genetic code with ribozymes. Nat. Commun. 2019, 10, 5097. [Google Scholar] [CrossRef]
  146. Goto, Y.; Suga, H. The rapid platform for the discovery of pseudo-natural macrocyclic peptides. Acc. Chem. Res. 2021, 54, 3604–3617. [Google Scholar] [CrossRef] [PubMed]
  147. Li, J.; Jaitzig, J.; Theuer, L.; Legala, O.E.; Sussmuth, R.D.; Neubauer, P. Type ii thioesterase improves heterologous biosynthesis of valinomycin in escherichia coli. J. Biotechnol. 2015, 193, 16–22. [Google Scholar] [CrossRef] [PubMed]
  148. Krizsan, A.; Volke, D.; Weinert, S.; Strater, N.; Knappe, D.; Hoffmann, R. Insect-derived proline-rich antimicrobial peptides kill bacteria by inhibiting bacterial protein translation at the 70s ribosome. Angew. Chem. Int. Ed. 2014, 53, 12236–12239. [Google Scholar] [CrossRef] [PubMed]
  149. Zhang, Y.; Goto, Y.; Suga, H. Discovery, biochemical characterization, and bioengineering of cyanobactin prenyltransferases. Trends Biochem. Sci. 2023, 48, 360–374. [Google Scholar] [CrossRef]
  150. Alfi, A.; Popov, A.; Kumar, A.; Zhang, K.Y.J.; Dubiley, S.; Severinov, K.; Tagami, S. Cell-free mutant analysis combined with structure prediction of a lasso peptide biosynthetic protease b2. ACS Synth. Biol. 2022, 11, 2022–2028. [Google Scholar] [CrossRef] [PubMed]
  151. Goto, Y.; Suga, H. Engineering of ripp pathways for the production of artificial peptides bearing various non-proteinogenic structures. Curr. Opin. Chem. Biol. 2018, 46, 82–90. [Google Scholar] [CrossRef]
  152. Liu, R.; Zhang, Y.; Zhai, G.; Fu, S.; Xia, Y.; Hu, B.; Cai, X.; Zhang, Y.; Li, Y.; Deng, Z.; et al. A cell-free platform based on nisin biosynthesis for discovering novel lanthipeptides and guiding their overproduction in vivo. Adv. Sci. 2020, 7, 2001616. [Google Scholar] [CrossRef]
  153. Wu, Y.; Cui, Z.; Huang, Y.H.; de Veer, S.J.; Aralov, A.V.; Guo, Z.; Moradi, S.V.; Hinton, A.O.; Deuis, J.R.; Guo, S.; et al. Towards a generic prototyping approach for therapeutically-relevant peptides and proteins in a cell-free translation system. Nat. Commun. 2022, 13, 260. [Google Scholar] [CrossRef]
  154. 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, 5917–5927. [Google Scholar] [CrossRef] [PubMed]
  155. Taniguchi, M.; Ochiai, A.; Kondo, H.; Fukuda, S.; Ishiyama, Y.; Saitoh, E.; Kato, T.; Tanaka, T. Pyrrhocoricin, a proline-rich antimicrobial peptide derived from insect, inhibits the translation process in the cell-free escherichia coli protein synthesis system. J. Biosci. Bioeng. 2016, 121, 591–598. [Google Scholar] [CrossRef] [PubMed]
  156. Koller, T.O.; Morici, M.; Berger, M.; Safdari, H.A.; Lele, D.S.; Beckert, B.; Kaur, K.J.; Wilson, D.N. Structural basis for translation inhibition by the glycosylated drosocin peptide. Nat. Chem. Biol. 2023, 19, 1072–1081. [Google Scholar] [CrossRef] [PubMed]
  157. Vinogradov, A.A.; Shimomura, M.; Goto, Y.; Ozaki, T.; Asamizu, S.; Sugai, Y.; Suga, H.; Onaka, H. Minimal lactazole scaffold for in vitro thiopeptide bioengineering. Nat. Commun. 2020, 11, 2272. [Google Scholar] [CrossRef]
  158. Ozaki, T.; Yamashita, K.; Goto, Y.; Shimomura, M.; Hayashi, S.; Asamizu, S.; Sugai, Y.; Ikeda, H.; Suga, H.; Onaka, H. Dissection of goadsporin biosynthesis by in vitro reconstitution leading to designer analogues expressed in vivo. Nat. Commun. 2017, 8, 14207. [Google Scholar] [CrossRef]
  159. Vinogradov, A.A.; Shimomura, M.; Kano, N.; Goto, Y.; Onaka, H.; Suga, H. Promiscuous enzymes cooperate at the substrate level en route to lactazole a. J. Am. Chem. Soc. 2020, 142, 13886–13897. [Google Scholar] [CrossRef]
  160. Goto, Y.; Suga, H. In vitro biosynthesis of peptides containing exotic azoline analogues. ChemBioChem 2020, 21, 84–87. [Google Scholar] [CrossRef] [PubMed]
  161. Goto, Y.; Ito, Y.; Kato, Y.; Tsunoda, S.; Suga, H. One-pot synthesis of azoline-containing peptides in a cell-free translation system integrated with a posttranslational cyclodehydratase. Chem. Biol. 2014, 21, 766–774. [Google Scholar] [CrossRef]
  162. Goto, Y.; Suga, H. A post-translational cyclodehydratase, patd, tolerates sequence variation in the c-terminal region of substrate peptides. Chem. Lett. 2016, 45, 1247–1249. [Google Scholar] [CrossRef]
  163. Fleming, S.R.; Himes, P.M.; Ghodge, S.V.; Goto, Y.; Suga, H.; Bowers, A.A. Exploring the post-translational enzymology of paaa by mrna display. J. Am. Chem. Soc. 2020, 142, 5024–5028. [Google Scholar] [CrossRef]
  164. Zhuang, L.; Huang, S.; Liu, W.Q.; Karim, A.S.; Jewett, M.C.; Li, J. Total in vitro biosynthesis of the nonribosomal macrolactone peptide valinomycin. Metab. Eng. 2020, 60, 37–44. [Google Scholar] [CrossRef]
  165. Li, J.; Zhang, L.; Liu, W. Cell-free synthetic biology for in vitro biosynthesis of pharmaceutical natural products. Synth. Syst. Biotechnol. 2018, 3, 83–89. [Google Scholar] [CrossRef] [PubMed]
  166. Goering, A.W.; Li, J.; McClure, R.A.; Thomson, R.J.; Jewett, M.C.; Kelleher, N.L. In vitro reconstruction of nonribosomal peptide biosynthesis directly from DNA using cell-free protein synthesis. ACS Synth. Biol. 2017, 6, 39–44. [Google Scholar] [CrossRef]
  167. Fleming, S.R.; Bartges, T.E.; Vinogradov, A.A.; Kirkpatrick, C.L.; Goto, Y.; Suga, H.; Hicks, L.M.; Bowers, A.A. Flexizyme-enabled benchtop biosynthesis of thiopeptides. J. Am. Chem. Soc. 2019, 141, 758–762. [Google Scholar] [CrossRef] [PubMed]
  168. Pandi, A.; Adam, D.; Zare, A.; Trinh, V.T.; Schaefer, S.L.; Burt, M.; Klabunde, B.; Bobkova, E.; Kushwaha, M.; Foroughijabbari, Y.; et al. Cell-free biosynthesis combined with deep learning accelerates de novo-development of antimicrobial peptides. Nat. Commun. 2023, 14, 7197. [Google Scholar] [CrossRef] [PubMed]
  169. Zhang, Y.; Liu, L.-H.; Xu, B.; Zhang, Z.; Yang, M.; He, Y.; Chen, J.; Zhang, Y.; Hu, Y.; Chen, X. Screening antimicrobial peptides and probiotics using multiple deep learning and directed evolution strategies. Acta Pharm. Sin. B 2024, 14, 3476–3492. [Google Scholar] [CrossRef]
  170. Nuti, N.; Rottmann, P.; Stucki, A.; Koch, P.; Panke, S.; Dittrich, P.S. A multiplexed cell-free assay to screen for antimicrobial peptides in double emulsion droplets. Angew. Chem. Int. Ed. 2022, 61, e202114632. [Google Scholar] [CrossRef]
  171. Vinogradov, A.A.; Nagai, E.; Chang, J.S.; Narumi, K.; Onaka, H.; Goto, Y.; Suga, H. Accurate broadcasting of substrate fitness for lactazole biosynthetic pathway from reactivity-profiling mRNA display. J. Am. Chem. Soc. 2020, 142, 20329–20334. [Google Scholar] [CrossRef]
  172. Vinogradov, A.A.; Chang, J.S.; Onaka, H.; Goto, Y.; Suga, H. Accurate models of substrate preferences of post-translational modification enzymes from a combination of mrna display and deep learning. ACS Cent. Sci. 2022, 8, 814–824. [Google Scholar] [CrossRef]
  173. Henzler Wildman, K.A.; Lee, D.K.; Ramamoorthy, A. Mechanism of lipid bilayer disruption by the human antimicrobial peptide, ll-37. Biochemistry 2003, 42, 6545–6558. [Google Scholar] [CrossRef]
  174. Jean-Francois, F.; Elezgaray, J.; Berson, P.; Vacher, P.; Dufourc, E.J. Pore formation induced by an antimicrobial peptide: Electrostatic effects. Biophys. J. 2008, 95, 5748–5756. [Google Scholar] [CrossRef]
  175. Graf, M.; Wilson, D.N. Intracellular antimicrobial peptides targeting the protein synthesis machinery. Adv. Exp. Med. Biol. 2019, 1117, 73–89. [Google Scholar] [PubMed]
  176. Patil, N.A.; Thombare, V.J.; Li, R.; He, X.; Lu, J.; Yu, H.H.; Wickremasinghe, H.; Pamulapati, K.; Azad, M.A.K.; Velkov, T.; et al. An efficient approach for the design and synthesis of antimicrobial peptide-peptide nucleic acid conjugates. Front. Chem. 2022, 10, 843163. [Google Scholar] [CrossRef]
  177. Xie, Y.; Wan, H.; Zeng, X.; Zhang, Z.; Wang, Y. Characterization and antimicrobial evaluation of a new spgly-amp, glycine-rich antimicrobial peptide from the mud crab scylla paramamosain. Fish Shellfish Immunol. 2020, 106, 384–392. [Google Scholar] [CrossRef]
  178. Park, S.C.; Park, Y.; Hahm, K.S. The role of antimicrobial peptides in preventing multidrug-resistant bacterial infections and biofilm formation. Int. J. Mol. Sci. 2011, 12, 5971–5992. [Google Scholar] [CrossRef]
  179. Cho, J.; Lee, D.G. The antimicrobial peptide arenicin-1 promotes generation of reactive oxygen species and induction of apoptosis. Biochim. Biophys. Acta 2011, 1810, 1246–1251. [Google Scholar] [CrossRef] [PubMed]
  180. Zeng, P.; Cheng, Q.; Yi, L.; Shui Yee Leung, S.; Chen, S.; Chan, K.F.; Wong, K.Y. C-terminal modification of a de novo designed antimicrobial peptide via capping of macrolactam rings. Bioorg. Chem. 2023, 130, 106251. [Google Scholar] [CrossRef]
  181. Hall, K.; Aguilar, M.I. Surface plasmon resonance spectroscopy for studying the membrane binding of antimicrobial peptides. Methods Mol. Biol. 2010, 627, 213–223. [Google Scholar] [PubMed]
  182. Uyterhoeven, E.T.; Butler, C.H.; Ko, D.; Elmore, D.E. Investigating the nucleic acid interactions and antimicrobial mechanism of buforin ii. FEBS Lett. 2008, 582, 1715–1718. [Google Scholar] [CrossRef]
  183. Chen, Y.; Guarnieri, M.T.; Vasil, A.I.; Vasil, M.L.; Mant, C.T.; Hodges, R.S. Role of peptide hydrophobicity in the mechanism of action of alpha-helical antimicrobial peptides. Antimicrob. Agents Chemother. 2007, 51, 1398–1406. [Google Scholar] [CrossRef]
  184. Baek, M.H.; Kamiya, M.; Kushibiki, T.; Nakazumi, T.; Tomisawa, S.; Abe, C.; Kumaki, Y.; Kikukawa, T.; Demura, M.; Kawano, K.; et al. Lipopolysaccharide-bound structure of the antimicrobial peptide cecropin p1 determined by nuclear magnetic resonance spectroscopy. J. Pept. Sci. 2016, 22, 214–221. [Google Scholar] [CrossRef]
  185. Fernandez, D.I.; Le Brun, A.P.; Whitwell, T.C.; Sani, M.A.; James, M.; Separovic, F. The antimicrobial peptide aurein 1.2 disrupts model membranes via the carpet mechanism. Phys. Chem. Chem. Phys. 2012, 14, 15739–15751. [Google Scholar] [CrossRef]
  186. Wilson, D.N.; Arenz, S.; Beckmann, R. Translation regulation via nascent polypeptide-mediated ribosome stalling. Curr. Opin. Struct. Biol. 2016, 37, 123–133. [Google Scholar] [CrossRef]
  187. Kohanski, M.A.; Dwyer, D.J.; Collins, J.J. How antibiotics kill bacteria: From targets to networks. Nat. Rev. Microbiol. 2010, 8, 423–435. [Google Scholar] [CrossRef]
  188. Gould, P.S.; Bird, H.; Easton, A.J. Translation toeprinting assays using fluorescently labeled primers and capillary electrophoresis. Biotechniques 2005, 38, 397–400. [Google Scholar] [CrossRef]
  189. Egorova, T.; Sokolova, E.; Shuvalova, E.; Matrosova, V.; Shuvalov, A.; Alkalaeva, E. Fluorescent toeprinting to study the dynamics of ribosomal complexes. Methods 2019, 162–163, 54–59. [Google Scholar] [CrossRef]
  190. Correia Santos, S.; Bischler, T.; Westermann, A.J.; Vogel, J. Maps integrates regulation of actin-targeting effector stec into the virulence control network of salmonella small rna pint. Cell Rep. 2021, 34, 108722. [Google Scholar] [CrossRef]
  191. Li, Y.M.; Milne, J.C.; Madison, L.L.; Kolter, R.; Walsh, C.T. From peptide precursors to oxazole and thiazole-containing peptide antibiotics: Microcin b17 synthase. Science 1996, 274, 1188–1193. [Google Scholar] [CrossRef]
  192. Dunbar, K.L.; Dell, M.; Gude, F.; Hertweck, C. Reconstitution of polythioamide antibiotic backbone formation reveals unusual thiotemplated assembly strategy. Proc. Natl. Acad. Sci. USA 2020, 117, 8850–8858. [Google Scholar] [CrossRef]
  193. 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.; et al. Ribosomally synthesized and post-translationally modified peptide natural products: Overview and recommendations for a universal nomenclature. Nat. Prod. Rep. 2013, 30, 108–160. [Google Scholar] [CrossRef]
  194. Huang, S.; Liu, Y.; Liu, W.Q.; Neubauer, P.; Li, J. The nonribosomal peptide valinomycin: From discovery to bioactivity and biosynthesis. Microorganisms 2021, 9, 780. [Google Scholar] [CrossRef]
  195. Li, L.; Li, J.; Yu, X.; Cao, R.; Hong, M.; Xu, Z.; Ren Lu, J.; Wang, Y.; Zhu, H. Antimicrobial peptides fight against pseudomonas aeruginosa at a sub-inhibitory concentration via anti-qs pathway. Bioorg. Chem. 2023, 141, 106922. [Google Scholar] [CrossRef] [PubMed]
  196. Xu, B.; Wang, L.; Yang, C.; Yan, R.; Zhang, P.; Jin, M.; Du, H.; Wang, Y. Specifically targeted antimicrobial peptides synergize with bacterial-entrapping peptide against systemic mrsa infections. J. Adv. Res. 2024. [Google Scholar] [CrossRef] [PubMed]
  197. Dhople, V.; Krukemeyer, A.; Ramamoorthy, A. The human beta-defensin-3, an antibacterial peptide with multiple biological functions. Biochim. Biophys. Acta 2006, 1758, 1499–1512. [Google Scholar] [CrossRef]
  198. Zimmermann, M.; Hegemann, J.D.; Xie, X.; Marahiel, M.A. The astexin-1 lasso peptides: Biosynthesis, stability, and structural studies. Chem. Biol. 2013, 20, 558–569. [Google Scholar] [CrossRef]
  199. Bagley, M.C.; Dale, J.W.; Merritt, E.A.; Xiong, X. Thiopeptide antibiotics. Chem. Rev. 2005, 105, 685–714. [Google Scholar] [CrossRef]
  200. Tan, S.; Ludwig, K.C.; Muller, A.; Schneider, T.; Nodwell, J.R. The lasso peptide siamycin-i targets lipid ii at the gram-positive cell surface. ACS Chem. Biol. 2019, 14, 966–974. [Google Scholar] [CrossRef]
  201. Cao, L.; Do, T.; Link, A.J. Mechanisms of action of ribosomally synthesized and posttranslationally modified peptides (ripps). J. Ind. Microbiol. Biotechnol. 2021, 48, kuab005. [Google Scholar] [CrossRef]
  202. Cheng, C.; Hua, Z.C. Lasso peptides: Heterologous production and potential medical application. Front. Bioeng. Biotechnol. 2020, 8, 571165. [Google Scholar] [CrossRef] [PubMed]
  203. Digal, L.; Samson, S.C.; Stevens, M.A.; Ghorai, A.; Kim, H.; Mifflin, M.C.; Carney, K.R.; Williamson, D.L.; Um, S.; Nagy, G.; et al. Nonthreaded isomers of sungsanpin and ulleungdin lasso peptides inhibit h1299 cancer cell migration. ACS Chem. Biol. 2024, 19, 81–88. [Google Scholar] [CrossRef]
  204. Mathavan, I.; Zirah, S.; Mehmood, S.; Choudhury, H.G.; Goulard, C.; Li, Y.; Robinson, C.V.; Rebuffat, S.; Beis, K. Structural basis for hijacking siderophore receptors by antimicrobial lasso peptides. Nat. Chem. Biol. 2014, 10, 340–342. [Google Scholar] [CrossRef]
  205. Martin-Gomez, H.; Tulla-Puche, J. Lasso peptides: Chemical approaches and structural elucidation. Org. Biomol. Chem. 2018, 16, 5065–5080. [Google Scholar] [CrossRef] [PubMed]
  206. Maksimov, M.O.; Pan, S.J.; James Link, A. Lasso peptides: Structure, function, biosynthesis, and engineering. Nat. Prod. Rep. 2012, 29, 996–1006. [Google Scholar] [CrossRef] [PubMed]
  207. Syroegin, E.A.; Flemmich, L.; Klepacki, D.; Vazquez-Laslop, N.; Micura, R.; Polikanov, Y.S. Structural basis for the context-specific action of the classic peptidyl transferase inhibitor chloramphenicol. Nat. Struct. Mol. Biol. 2022, 29, 152–161. [Google Scholar] [CrossRef] [PubMed]
  208. Wilson, D.N. Ribosome-targeting antibiotics and mechanisms of bacterial resistance. Nat. Rev. Microbiol. 2014, 12, 35–48. [Google Scholar] [CrossRef]
  209. Ma, B.; Fang, C.; Lu, L.; Wang, M.; Xue, X.; Zhou, Y.; Li, M.; Hu, Y.; Luo, X.; Hou, Z. The antimicrobial peptide thanatin disrupts the bacterial outer membrane and inactivates the ndm-1 metallo-beta-lactamase. Nat. Commun. 2019, 10, 3517. [Google Scholar] [CrossRef]
  210. Mhlongo, J.T.; Waddad, A.Y.; Albericio, F.; de la Torre, B.G. Antimicrobial peptide synergies for fighting infectious diseases. Adv. Sci. 2023, 10, e2300472. [Google Scholar] [CrossRef]
  211. Krizsan, A.; Prahl, C.; Goldbach, T.; Knappe, D.; Hoffmann, R. Short proline-rich antimicrobial peptides inhibit either the bacterial 70s ribosome or the assembly of its large 50s subunit. ChemBioChem 2015, 16, 2304–2308. [Google Scholar] [CrossRef]
Figure 1. NRPSs are large complexes composed of adenylation (A), thiolation (T), and condensation (C) domains. Non-ribosomal peptides are synthesized through sequential enzymatic reactions: the A domain activates and attaches amino acids (colored circles) to the T domain, initiating biosynthesis by forming a thioester bond with Ppant (blue). The C domain then forms peptide bonds between modules. During termination, the thioesterase (TE) domain hydrolyzes the bond between the T domain and the peptide, releasing the final peptide product.
Figure 1. NRPSs are large complexes composed of adenylation (A), thiolation (T), and condensation (C) domains. Non-ribosomal peptides are synthesized through sequential enzymatic reactions: the A domain activates and attaches amino acids (colored circles) to the T domain, initiating biosynthesis by forming a thioester bond with Ppant (blue). The C domain then forms peptide bonds between modules. During termination, the thioesterase (TE) domain hydrolyzes the bond between the T domain and the peptide, releasing the final peptide product.
Ijms 25 09109 g001
Figure 2. Two types of cell-free systems (CFSs). (a) Lysate-based CFSs (yellow) provide a cell-like environment. (b) Reconstituted CFSs (blue) contain a minimal set of 36 purified enzymes required for transcription and translation for protein synthesis. The open environment of CFSs allows the precise control of biosynthetic pathways of PBAs through the modular assembly of required components.
Figure 2. Two types of cell-free systems (CFSs). (a) Lysate-based CFSs (yellow) provide a cell-like environment. (b) Reconstituted CFSs (blue) contain a minimal set of 36 purified enzymes required for transcription and translation for protein synthesis. The open environment of CFSs allows the precise control of biosynthetic pathways of PBAs through the modular assembly of required components.
Ijms 25 09109 g002
Figure 3. Cell-free systems as ideal platforms for elucidating the intracellular mechanisms of the antimicrobial action of PBAs. (a) In a reconstituted cell-free system, deleterious enzymes (e.g., RNases, DNases, and proteases) that could reduce protein production are excluded. This allows for the use or synthesis of short DNAs, RNAs, or peptides. The toe-printing technique, which utilizes ribosome stalling, helps identify the PBA mechanisms that inhibit ribosome polymerization. The platform enables the identification of inhibited translation stages in the discovery process, especially if the PBAs target the large subunit of ribosomes. (b) The inhibition mechanism of type II PrAMPs (drosocin) is elucidated on a cell-free platform by selectively adding or removing RF1. Cryo-EM analysis shows that drosocin binds to the exit tunnel of the ribosome and interacts with RF1 located in the A-site (exit tunnel of the ribosome: blue, drosocin: red, RF1: yellow), inhibiting translation efficiency by approximately 50%. (c) The capability to form biomolecular condensates (RNA and PBA) via phase transitions is the origin of bactericidal activity.
Figure 3. Cell-free systems as ideal platforms for elucidating the intracellular mechanisms of the antimicrobial action of PBAs. (a) In a reconstituted cell-free system, deleterious enzymes (e.g., RNases, DNases, and proteases) that could reduce protein production are excluded. This allows for the use or synthesis of short DNAs, RNAs, or peptides. The toe-printing technique, which utilizes ribosome stalling, helps identify the PBA mechanisms that inhibit ribosome polymerization. The platform enables the identification of inhibited translation stages in the discovery process, especially if the PBAs target the large subunit of ribosomes. (b) The inhibition mechanism of type II PrAMPs (drosocin) is elucidated on a cell-free platform by selectively adding or removing RF1. Cryo-EM analysis shows that drosocin binds to the exit tunnel of the ribosome and interacts with RF1 located in the A-site (exit tunnel of the ribosome: blue, drosocin: red, RF1: yellow), inhibiting translation efficiency by approximately 50%. (c) The capability to form biomolecular condensates (RNA and PBA) via phase transitions is the origin of bactericidal activity.
Ijms 25 09109 g003
Figure 4. CFSs facilitate the characterization of biosynthetic pathways of PBAs. (a) The sequence of PTMs within a biosynthetic gene cluster (BGC) can be efficiently confirmed using a reconstituted CFS. Enzyme A and B of the BGC should react sequentially to produce an azole-ring and a dehydroalanine motif. Enzymatic reactions in reverse order do not yield the desired product (red x). Goadsporin (lower panel), an azole-containing linear PBA, is synthesized using a FIT system by supplementing the key PTM enzymes in a sequential order. (b) To analyze the rate-limiting step of NisZ biosynthesis mediated by NisP, NisB, and NisC, an excess amount of the purified enzyme involved in the targeted step is supplied to the in vitro reaction, while the other enzymes are expressed directly from plasmids at low concentrations. The increase in production indicates that the step catalyzed by the additionally supplied enzyme is rate limiting (upper panel). For large-scale synthesis, the key enzymes are overexpressed in plasmids, enabling E. coli to produce the target lanthipeptide analog (RL6) in large amounts (lower panel).
Figure 4. CFSs facilitate the characterization of biosynthetic pathways of PBAs. (a) The sequence of PTMs within a biosynthetic gene cluster (BGC) can be efficiently confirmed using a reconstituted CFS. Enzyme A and B of the BGC should react sequentially to produce an azole-ring and a dehydroalanine motif. Enzymatic reactions in reverse order do not yield the desired product (red x). Goadsporin (lower panel), an azole-containing linear PBA, is synthesized using a FIT system by supplementing the key PTM enzymes in a sequential order. (b) To analyze the rate-limiting step of NisZ biosynthesis mediated by NisP, NisB, and NisC, an excess amount of the purified enzyme involved in the targeted step is supplied to the in vitro reaction, while the other enzymes are expressed directly from plasmids at low concentrations. The increase in production indicates that the step catalyzed by the additionally supplied enzyme is rate limiting (upper panel). For large-scale synthesis, the key enzymes are overexpressed in plasmids, enabling E. coli to produce the target lanthipeptide analog (RL6) in large amounts (lower panel).
Ijms 25 09109 g004
Figure 5. Screening for salivaricin B synthesis and its derivatives obtained from the selection process. (a) Genome mining and screening platforms enable the production of a library containing mutations at multiple sites of the wild-type salivaricin B structure. The enzyme-binding sites (light blue) revealed via genome mining remained intact during mutagenesis. Using cell-free protein synthesis platforms, a novel salivaricin B derivative with high antimicrobial activity (green √) against gram-negative bacteria was discovered. (b) Through a selection process from a library containing numerous derivatives, a new salivaricin B derivative with 4-fold enhanced efficacy was found to contain a histidine residue (red).
Figure 5. Screening for salivaricin B synthesis and its derivatives obtained from the selection process. (a) Genome mining and screening platforms enable the production of a library containing mutations at multiple sites of the wild-type salivaricin B structure. The enzyme-binding sites (light blue) revealed via genome mining remained intact during mutagenesis. Using cell-free protein synthesis platforms, a novel salivaricin B derivative with high antimicrobial activity (green √) against gram-negative bacteria was discovered. (b) Through a selection process from a library containing numerous derivatives, a new salivaricin B derivative with 4-fold enhanced efficacy was found to contain a histidine residue (red).
Ijms 25 09109 g005
Figure 6. CFSs for screening targeted PBAs. (a) Biosynthesis of NRPs using NRPSs requires precise reaction conditions to achieve high yields. NRPS modules are produced separately and combined into a multi-well plate to identify optimal reaction conditions, such as glucose concentration, NRPS ratios, cofactors (NAD, CoA, and ATP), and reaction time. (b) Molecular structures of diketopiperazine and valinomycin are produced efficiently by optimizing synthetic conditions using a CFS.
Figure 6. CFSs for screening targeted PBAs. (a) Biosynthesis of NRPs using NRPSs requires precise reaction conditions to achieve high yields. NRPS modules are produced separately and combined into a multi-well plate to identify optimal reaction conditions, such as glucose concentration, NRPS ratios, cofactors (NAD, CoA, and ATP), and reaction time. (b) Molecular structures of diketopiperazine and valinomycin are produced efficiently by optimizing synthetic conditions using a CFS.
Ijms 25 09109 g006
Figure 7. Schematic of the research for investigating lactazole BGCs. (a) For the reactivity-profiling mRNA display assay, N-terminally biotinylated thiopeptide precursors are produced using a CFS by reprogramming translation initiation. These biotinylated peptides are then displayed on mRNAs, forming peptide–mRNA conjugates by the nucleophilic attack of the amino group of puromycin (purple circle) to the nascent peptide chain in the ribosome. During the assay, the conjugates are treated with a set of Laz enzymes, which convert some precursors to mature thiopeptides by cleaving off the biotin tag. A streptavidin-based pulldown step is then used to remove the reactive (biotin-cleaved, mature, black dotted box) from the unreactive (biotin-retained, only partially modified or not modified) fractions (red dotted box). (b) Peptide sequences are scored using the “Y score”, which is determined by analyzing the frequency of each peptide in both fractions of the library. This model predicts the promiscuity of lactazole RiPP enzymes, facilitating the de novo design of thiopeptide derivatives bearing a thiazole (red pentagon).
Figure 7. Schematic of the research for investigating lactazole BGCs. (a) For the reactivity-profiling mRNA display assay, N-terminally biotinylated thiopeptide precursors are produced using a CFS by reprogramming translation initiation. These biotinylated peptides are then displayed on mRNAs, forming peptide–mRNA conjugates by the nucleophilic attack of the amino group of puromycin (purple circle) to the nascent peptide chain in the ribosome. During the assay, the conjugates are treated with a set of Laz enzymes, which convert some precursors to mature thiopeptides by cleaving off the biotin tag. A streptavidin-based pulldown step is then used to remove the reactive (biotin-cleaved, mature, black dotted box) from the unreactive (biotin-retained, only partially modified or not modified) fractions (red dotted box). (b) Peptide sequences are scored using the “Y score”, which is determined by analyzing the frequency of each peptide in both fractions of the library. This model predicts the promiscuity of lactazole RiPP enzymes, facilitating the de novo design of thiopeptide derivatives bearing a thiazole (red pentagon).
Ijms 25 09109 g007
Figure 8. Diverse platforms using CFSs for screening PBAs. (a) To design PBAs de novo for new drug discovery, a deep generative model is integrated with a CFS. The model prioritizes potential PBAs and predicts target PBAs with higher accuracy. In the experimental pipeline, PBAs are rapidly synthesized from a synthetic DNA template in the CFS and directly tested against bacterial cultures. Notably, the six newly discovered de novo-designed PBAs exhibit antimicrobial activity against a wide spectrum of multidrug-resistant pathogens. (b) Microfluidic systems that generate multiple droplets containing different PBAs provide an efficient way to identify new PBAs with high membrane disruption activity. A self-quenching dye is embedded in a vesicle, which is further encapsulated in bacteria-like lipid vesicles. Increased fluorescence indicates that the PBA disrupts the inner vesicle, resulting in the dilution of the self-quenched dye within the droplet.
Figure 8. Diverse platforms using CFSs for screening PBAs. (a) To design PBAs de novo for new drug discovery, a deep generative model is integrated with a CFS. The model prioritizes potential PBAs and predicts target PBAs with higher accuracy. In the experimental pipeline, PBAs are rapidly synthesized from a synthetic DNA template in the CFS and directly tested against bacterial cultures. Notably, the six newly discovered de novo-designed PBAs exhibit antimicrobial activity against a wide spectrum of multidrug-resistant pathogens. (b) Microfluidic systems that generate multiple droplets containing different PBAs provide an efficient way to identify new PBAs with high membrane disruption activity. A self-quenching dye is embedded in a vesicle, which is further encapsulated in bacteria-like lipid vesicles. Increased fluorescence indicates that the PBA disrupts the inner vesicle, resulting in the dilution of the self-quenched dye within the droplet.
Ijms 25 09109 g008
Figure 9. CFSs enable the synthesis of more complex protein structures. (a) Resin-assisted reduction and oxidative folding. Beads, utilized to capture the tag of disulfide-rich PBAs, are supplied into a CFS to avoid the trade-off between folding and aggregation, sequestering both on-path and off-path folding intermediates. Following the capture of the tag, disulfide bonds are completely reduced. Subsequently, the oxidation of cysteine is reperformed under thermodynamic control, enabling the reshuffling of disulfide bonds. This process ensures that PBAs adopt the most stable and functional structure via multiple disulfide bonds. (b) A co-expression of enzymes required to process precursor peptides significantly increases the yield of lasso peptides by up to 200-fold compared with conventional heterologous expression strategies. (c) The open environment of the reconstituted CFS allows for the incorporation and in vitro translation of ncMs, followed by additional in situ chemo-enzymatic reactions. By excluding specific aaRSs or RFs, ncM is introduced into the peptide without competition from the endogenous tRNAs. Chemo-enzymatic reactions between non-canonical motifs incorporated into a peptide form a novel architecture (e.g., cyclic peptide), expanding the diversity of the PBAs produced through CFSs.
Figure 9. CFSs enable the synthesis of more complex protein structures. (a) Resin-assisted reduction and oxidative folding. Beads, utilized to capture the tag of disulfide-rich PBAs, are supplied into a CFS to avoid the trade-off between folding and aggregation, sequestering both on-path and off-path folding intermediates. Following the capture of the tag, disulfide bonds are completely reduced. Subsequently, the oxidation of cysteine is reperformed under thermodynamic control, enabling the reshuffling of disulfide bonds. This process ensures that PBAs adopt the most stable and functional structure via multiple disulfide bonds. (b) A co-expression of enzymes required to process precursor peptides significantly increases the yield of lasso peptides by up to 200-fold compared with conventional heterologous expression strategies. (c) The open environment of the reconstituted CFS allows for the incorporation and in vitro translation of ncMs, followed by additional in situ chemo-enzymatic reactions. By excluding specific aaRSs or RFs, ncM is introduced into the peptide without competition from the endogenous tRNAs. Chemo-enzymatic reactions between non-canonical motifs incorporated into a peptide form a novel architecture (e.g., cyclic peptide), expanding the diversity of the PBAs produced through CFSs.
Ijms 25 09109 g009
Table 1. Characteristics of various PBAs and their activity against various pathogens.
Table 1. Characteristics of various PBAs and their activity against various pathogens.
TypePeptideProposed Antibacterial MechanismsTarget
Gram-Negative Pathogens and MIC
Target
Gram-Positive Pathogens and MIC
MIC for MDR
PeptidesMagainin II
[19,20]
  • Membrane disruption
  • E. coli DH5α
    (3.125 μM)
  • P. aeruginosa
    (12.5 μM)
  • S. aureus
    (100 μM)
  • MR P. aeruginosa
    (12.5 μM)
  • MR S. aureus
    (8–64 μg/mL)
Melittin
[21,22,23]
  • Membrane disruption
  • P. aeruginosa
    (25–50 μg/mL)
  • S. saprophyticus
    ATCC 15305
    (1 μM)
  • S. haemoliticus
    (1 μg/mL)
  • A. baumanni
    (31 μg/mL)
  • MR S. aureus
    (0.12 μM)
LL-37
[24,25,26]
  • Membrane disruption
  • Immunomodulation
  • E. coli
    ATCC 11775
    (1.9 μg/mL)
  • A. baumannii-ATCC 19606
    (32 μg/mL)
  • MR A. baumannii
    (16–32 µg/mL)
  • MR P. aeruginosa
    (32–287.5 μg/mL)
Buforin II
[20,27]
  • Membrane disruption
  • Intracellular targeting
    (nucleic acid)
  • E. coli
    (0.125–2 μg/mL)
  • P. aeruginosa
    (1–16 μg/mL)
  • A. baumanni
    (8–16 µM)
  • MR S. aureus
    (1–8 μg/mL)
Cecropin P1
[20]
  • Membrane disruption
  • E. coli
    (0.125–2 μg/mL)
  • P. aeruginosa
    (2–32 μg/mL)
  • MR S. aureus
    (16–64 μg/mL)
Indolicidin
[20]
  • Membrane disruption
  • Intracellular targeting (inhibition of DNA replication and transcription, Ca2+-calmodulin interaction)
  • E. coli
    (0.50–8 μg/mL)
  • P. aeruginosa
    (4–64 μg/mL)
  • MR S. aureus
    (2–16 μg/mL)
PR-39
[23,28]
  • Intracellular targeting
    (nucleic acid)
  • E. coli K12
    (0.3 µM)
  • S. aureus
    (200 µM)
  • MR S. aureus
    (>40 μM)
Vancomycin
[29,30]
  • Inhibition of cell wall synthesis
-
  • S. aureus
    (4–8 μg/mL)
  • MR S. aureus
    (≤2 μg/mL)
Daptomycin
[23,31]
  • Membrane disruption
-
  • S. aureus
    (0.25–1 μg/mL)
  • MR S. aureus
    (0.125–1.0 μg/mL)
Polymyxin
[32,33]
  • Membrane disruption
  • P. aeruginosa
    (≤ 2 μg/mL)
  • A. spp
    (≥4 μg/mL)
  • A. baumannii 3367
    (0.25 μg/mL)
  • A. baumannii 100
    (8 μg/mL)
Small moleculesCiprofloxacin
[27,34]
  • Intracellular targeting
    (DNA gyrase)
  • E. coli
    (0.004–0.25 μg/mL)
  • E. faecium
    (0.8–25 μg/mL)
  • A. baumanni
    (128 µM)
Gentamicin
[27,35]
  • Intracellular targeting
    (ribosome)
  • E. coli
    (4 μg/mL)
  • E. species
    (1–16 µg/mL)
  • A. baumanni
    (256 µM)
Tetracycline
[36,37]
  • Intracellular targeting (ribosome)
  • E. coli K12 TB1
    (2–16 μg/mL)
  • P. aeruginosa
    (16–32 µg/mL)
  • A. baumannii
    (>128 μM)
Table 3. Four major PTMs and their reaction products.
Table 3. Four major PTMs and their reaction products.
Types of PTMsPTM ReactionsPBAs
CyclizationBridging
reactions
Thioether formation
  • Nisin A
  • Thuricin H
  • Subtilosin A
Disulfide bond formation
(Cysteine oxidation)
  • Apamin
  • Kalata B1
  • Bactenecin
Head-to-tail cyclization
  • Circularin A
  • Microcystin-LA
  • Subtilosin A
  • Kalata B1
  • Patellamide A
CyclodehydrationHeterocycle formation
  • Goadsporin
  • Thiostrepton
  • Patellamide A
  • TP-1161
AlkylationN-methylation
  • Cyclosporin A
  • Thiostrepton
  • Microcystin-LA
GlycosylationN-Glycosylation
  • Eosinophil cationic protein (ECP)
  • h-LF
  • Tyrocidine A
  • Lactoferrin
S-Glycosylation
  • Sublancin
O-glycosylation
  • Datucin
  • Diptericin
  • Formaecin
  • Drosocin
Table 4. Main research topics on PBAs using CFSs.
Table 4. Main research topics on PBAs using CFSs.
Mechanisms
(Section 4.1)
Biosynthetic Pathways
(Section 4.2)
Selections
(Section 4.3)
Synthesis
(Section 4.4)
MethodsMutagenesis of PBAs [148]Identification of biosynthetic pathways [149,150,151]Screening of multiple variants [139,152]Inducing proper folding
[153,154]
Qualitative and quantitative analysis on biomolecules [16,63,155,156]Enzyme promiscuity for substrates [157,158,159,160,161,162,163]Optimization of reaction conditions [164,165,166]Chemo-enzymatic modification [167]
Kinetic analysis of enzymes [152]
Integration of deep generative models
[168,169]
Generation of diverse environments through microfluidic system [170]
Target peptide selection using mRNA displays
[171,172]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Park, H.; Jin, H.; Kim, D.; Lee, J. Cell-Free Systems: Ideal Platforms for Accelerating the Discovery and Production of Peptide-Based Antibiotics. Int. J. Mol. Sci. 2024, 25, 9109. https://doi.org/10.3390/ijms25169109

AMA Style

Park H, Jin H, Kim D, Lee J. Cell-Free Systems: Ideal Platforms for Accelerating the Discovery and Production of Peptide-Based Antibiotics. International Journal of Molecular Sciences. 2024; 25(16):9109. https://doi.org/10.3390/ijms25169109

Chicago/Turabian Style

Park, Hyeongwoo, Haneul Jin, Dayeong Kim, and Joongoo Lee. 2024. "Cell-Free Systems: Ideal Platforms for Accelerating the Discovery and Production of Peptide-Based Antibiotics" International Journal of Molecular Sciences 25, no. 16: 9109. https://doi.org/10.3390/ijms25169109

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

Article Metrics

Back to TopTop