Enzymatic Carbon–Sulfur Bond Formation in Natural Product

Apr 18, 2017 - Daniel H. Scharf received his Diploma degree (M.S. equiv) in Biology from the Friedrich Schiller University Jena, Germany, in 2008. ...
76 downloads 12 Views 12MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Review pubs.acs.org/CR

Enzymatic Carbon−Sulfur Bond Formation in Natural Product Biosynthesis Kyle L. Dunbar,† Daniel H. Scharf,‡ Agnieszka Litomska,† and Christian Hertweck*,†,§ †

Department of Biomolecular Chemistry, Leibniz Institute for Natural Product Research and Infection Biology (HKI), Beutenbergstrasse 11a, 07745 Jena, Germany ‡ Life Sciences Institute, University of Michigan, 210 Washtenaw Avenue, Ann Arbor, Michigan 48109-2216, United States § Friedrich Schiller University, 07743 Jena, Germany ABSTRACT: Sulfur plays a critical role for the development and maintenance of life on earth, which is reflected by the wealth of primary metabolites, macromolecules, and cofactors bearing this element. Whereas a large body of knowledge has existed for sulfur trafficking in primary metabolism, the secondary metabolism involving sulfur has long been neglected. Yet, diverse sulfur functionalities have a major impact on the biological activities of natural products. Recent research at the genetic, biochemical, and chemical levels has unearthed a broad range of enzymes, sulfur shuttles, and chemical mechanisms for generating carbon−sulfur bonds. This Review will give the first systematic overview on enzymes catalyzing the formation of organosulfur natural products.

CONTENTS 1. Introduction 2. C−S Bond-Forming Hydrolases 2.1. Endoproteases 2.1.1. Autoinducer Peptides 2.2. Thioesterases 2.2.1. Thiocoraline 2.3. (Ketosynthase-like) Acyltransferases 2.3.1. Calicheamicin 3. S-Transferases 3.1. S-Methyltransferases (SAM-dependent) 3.1.1. Bismethylgliotoxin 3.1.2. Collismycin 3.1.3. Echinomycin 3.1.4. Lincomycin A 3.1.5. Thiocoraline 3.1.6. Brassinin 3.1.7. S-Methylated RiPPs 3.2. S-Glycosyltransferases 3.2.1. Glycocins 3.2.2. Glucosinolates 3.2.3. Lincomycin A and Celesticetin 3.3. Glutathione-S-transferases 3.3.1. Glutathionyl and Cysteinyl Leukotrienes 3.3.2. Epipolythiodiketopiperazines 3.3.3. Glucosinolates 3.3.4. Allicin/Alliin 3.3.5. Pseurotin (trans to cis Isomerization) 3.4. S-Transferases Involved in Conjugate Additions

© 2017 American Chemical Society

3.4.1. Tropodithietic Acid and Roseobacticide A 3.4.2. Thienamycin 3.4.3. Leinamycin 4. C−S Bond-Forming Cyclases 4.1. Lanthipeptide Cyclases 4.2. NRPS Heterocyclization Domains 5. ATP-Dependent C−S Bond-Forming Enzymes 5.1. YcaO-like enzymes 5.1.1. Thiazoles and Thiazolines in RiPPs 5.1.2. Thioviridamide 5.2. Adenine Nucleotide Alpha Hydrolase Enzymes 5.2.1. 6-Thioguanine 5.2.2. Thiolactomycin 5.3. Adenylating Protein/Sulfur Carrier Protein Systems 5.3.1. Thioquinolobactin 5.3.2. BE-7585A 6. Oxygenases 6.1. Cytochrome P450 Monooxygenases 6.1.1. Camalexin 6.1.2. Cyclobrassinin and Spirobrassinin 6.1.3. Thienodolin 6.1.4. Griseoviridin 6.1.5. Ustiloxin

5522 5522 5522 5523 5524 5524 5525 5525 5525 5525 5526 5526 5527 5527 5527 5528 5528 5529 5529 5530 5530 5531 5532 5533 5533 5534 5534

5535 5536 5538 5539 5539 5540 5542 5542 5542 5543 5543 5543 5543 5544 5544 5545 5545 5545 5545 5546 5548 5548 5548

Special Issue: Unusual Enzymology in Natural Products Synthesis

5535

Received: October 11, 2016 Published: April 18, 2017 5521

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews 6.2. Nonheme Iron-Dependent Enzymes 6.2.1. Penicillins and Cephalosporins 6.2.2. Ergothioneine and Ovothiol 6.3. Flavoenzymes 6.3.1. Coelimycin 6.3.2. Sulfadixiamycins 6.4. Tyrosinases 6.4.1. Pheomelanins 6.4.2. Grixazones 6.4.3. Grape Reaction Product 7. Radical S-Adenosylmethionine Enzymes 7.1. Thioether-Forming rSAMs 7.1.1. Albomycin 7.1.2. Sactipeptides 7.1.3. γ-Subunit of Quinohemoprotein 8. Nonenzymatic C−S Bond Formations 8.1. Nonenzymatic Conjugate Addition 8.1.1. Enediyne Warhead Activation 8.1.2. Urdamycin E and BE-7585A 8.1.3. Ralfuranone D 8.2. Nonenzymatic Addition 8.2.1. Cyslabdan 8.2.2. Paulomycin S-Conjugates 8.3. Photochemical and Radical-Mediated Thioconjugation 8.3.1. Panphenazines 9. Conclusions and Outlook Author Information Corresponding Author Notes Biographies Acknowledgments References

Review

sulfur moieties that are pivotal for biological function (Figure 1), and, as might be expected, varied enzymatic mechanisms have evolved for the formation of the carbon−sulfur bonds in these molecules. Over the past decade, our understanding of enzymatic C−S bond formations in natural products has dramatically improved. Studies at the genetic, biochemical, and chemical levels have elucidated the sources of sulfur, the reaction mechanisms, and the types of biocatalysts involved. As in primary metabolism, persulfidic sulfur (R−S−SH) and thiocarboxylate groups on sulfur-donor proteins represent major ionic sulfur sources. In addition, thiols of cysteine and glutathioneand even sulfur dioxidemay serve as S-donors in secondary metabolism. Besides substitution reactions with sulfur nucleophiles, (conjugate) additions, and radical reactions are frequently observed. The aim of this Review is to provide the first overview of the current knowledge on biosynthetic pathways leading to sulfurbearing natural products. To systematize the many avenues leading to covalent C−S bonds, the Review is organized into sections on the different classes of biocatalysts involved. The reader will note that in some cases (e.g., glycosinolates, lincomycin, thiocoralin, gliotoxin, and phytoalexins) the biosynthetic pathways were dissected as multiple C−S bondforming enzymes are involved during biosynthesis. In some sections, C−S linkages of unknown biosynthetic origin are discussed. Beyond these open questions, it has not been possible to assign an exact reaction mechanism for the formation of every C−S bond; thus, some examples were only tentatively assigned to a particular section. Finally, select nonenzymatic C−S bond formations in secondary metabolism are highlighted.

5549 5549 5551 5553 5553 5553 5553 5553 5554 5555 5556 5556 5556 5557 5557 5558 5559 5559 5560 5561 5561 5561 5562 5562 5562 5562 5564 5564 5564 5564 5564 5564

2. C−S BOND-FORMING HYDROLASES In both primary and secondary metabolite pathways, thioester bonds play an important role. Activated biosynthetic building blocks and intermediates are often provided as thioesters bound to coenzyme A and phosphopantetheinylated carrier proteins.10,12 In these cases, their formation typically requires ATPdependent ligases, and frequently transesterifications are observed. It is common knowledge that such thioesters are energy-rich moieties that are easily hydrolyzed or transformed into the corresponding esters or amides. However, there are remarkable biosynthetic pathways where thermodynamically stable esters and amides are converted into high-energy thioesters. Even more surprising is the fact that these energetically disfavored reactions are mediated by enzymes that belong to the hydrolase family: biocatalysts known to cleave amide and ester bonds.

1. INTRODUCTION Sulfur is a ubiquitously distributed element that is essential to all known living species. According to the iron−sulfur world hypothesis, it even played a key role in the evolution of life.1 Numerous essential biochemical processes in prokaryotic and eukaryotic cells are tightly linked to this particular element. Enzymes not only harbor iron sulfur clusters but also depend on cofactors that contain sulfur, such as thiamine, molybdopterin, biotin, and lipoic acid.2 Furthermore, many biosynthetic building blocks are activated as coenzyme A thioesters, and important detoxification processes, such as the conversion of cyanide to isothiocyanate3 and the conjugation of electrophilic toxins to glutathione and related compounds,4 depend on sulfur. Thio modifications are also important for stabilizing the tRNA structure and for accurate and efficient translation.5 Sulfur’s prime position in primary metabolism is indisputable, which is reflected by a large body of knowledge that has been presented in various review articles.6−9 In contrast, the role of sulfur in natural product biosynthesis has been somewhat neglected. One may recall that intermediates of fatty acids, polyketides, and nonribosomal peptides are tethered to phosphopantetheinylated carrier proteins through thioester bonds,10 as well as that S-adenosylmethionine is an important cofactor of methylations.11 Even so, apart from cysteine and methionine residues in peptides, secondary metabolites are mainly composed of carbon, hydrogen, oxygen, and nitrogen whereas sulfur atoms are scarce. However, there are innumerable examples of natural products containing diverse

2.1. Endoproteases

Proteases are ubiquitous in nature and are central to many biological functions. While most proteases catalyze the hydrolytic cleavage of amide linkages, a subset of proteases has been characterized that catalyzes ligation reactions. Such proteolytic ligases resolve the acyl-enzyme intermediate with a nonsolvent nucleophile to afford the ligated product. Ligating proteases are critical for biological processes, including the anchoring of proteins to the bacterial cell wall by sortase and peptidoglycan biosynthesis by penicillin binding proteins.13,14 In addition to the roles that ligating proteases play in primary cellular pathways, select family members are also used in natural product biosynthesis.15−19 These members catalyze the macro5522

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Figure 1. Representative sulfur-bearing natural products.

There are four known AIPs produced by S. aureus strains. The peptides vary in length (7−9 amino acids) and in sequence but are linked by the presence of a conserved thiolactone macrocycle spanning five amino acids.21 Genetic analyses revealed that AIPs are members of the ribosomally synthesized and posttranslationally modified peptide (RiPP) natural product class.20,22 AIPs are biosynthesized from a larger precursor peptide (AgrD) by the iterative action of two proteases (Scheme 1). First, the C-terminal tail of the peptide is removed (herein referred to as a follower peptide) by AgrB, a transmembrane cysteine endoprotease.23,24 After thiolysis of the peptide bond, the enzyme-bound thioester intermediate is resolved by the internal cysteine of AgrD to generate the thiolactone macrocycle. Although this enzyme-bound thioester has never been directly observed, mutation of the AgrD cysteine residue to a serine was found to cause a buildup of covalently linked AgrD and AgrB.25 Because this cross-linked species was not observed with a catalytically inactive version of

cyclization of a peptide substrate, thus endowing the natural product with a more rigid structure and enhanced proteolytic stability. Because the first step of such proteolytic macrocyclizations is the cleavage of an amide bond, it is perhaps not surprising that most characterized members ultimately catalyze the formation of an enthalpically neutral macrolactam linkage. One notable exception is found in the biosynthesis of peptidic quorum-sensing molecules. 2.1.1. Autoinducer Peptides. Autoinducing peptides (AIPs) are important quorum-sensing molecules that are widely produced by Gram-positive bacteria. In Staphylococcus aureus, AIP is the signal peptide of the accessory gene regulator (agr) system where its recognition initiates a signal cascade that leads to the global expression of virulence factors.20,21 Because of the essential role that this quorum sensing system plays in pathogenesis, AIP has been the subject of extensive investigation. 5523

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

biosynthetic gene cluster demonstrated that the thioester linkage is formed between a Cys residue on the precursor peptide and an indolic acid derivative (Scheme 2).31 Notably, a member of the alpha−beta hydrolase superfamily (NosK) is predicted to install the thiolactone ring; however, this proposed function requires experimental verification.

Scheme 1. Thiolactone Formation Catalyzed by Transmembrane Cysteine Endoprotease AgrB Is a Key Step in the Biosynthesis of Autoinducing Peptides (AIPs) (LP, Leader Peptide; FP, Follower Peptide)

2.2. Thioesterases

Many polyketides, peptides, and hybrids thereof are produced by multimodular type I polyketide synthases (PKSs) and nonribosomal peptide synthetases (NRPSs) that resemble miniaturized assembly lines.32 A hallmark of these so-called thiotemplate systems is that the growing intermediates are covalently tethered as thioesters to thiolation (T) domains, i.e., acyl or peptidyl carrier proteins (ACPs or PCPs, respectively). Thioesterase (TE) domains are typically located at the end of the assembly lines and are responsible for off-loading the mature polyketide/peptide chains.32,33 The full-length product is first transferred from the T domain onto a serine residue of the TE domain. Second, depending on the nature of the TE domain, the ester bond may be hydrolyzed to yield the linear product or cyclized into a lactone or lactam ring by intramolecular attack of an amino or hydroxyl group, respectively. While many examples of lactone- and lactamforming TE domains have been discovered, it was only recently that a thiolactone-forming TE domain was identified. 2.2.1. Thiocoraline. This peculiar C−S bond-forming reaction is involved in the biosynthesis of thiocoraline, a sulfur-rich metabolite of marine actinomycetes. Thiocoraline belongs to the family of chromodepsipeptides, which comprise several nonribosomally produced pseudosymmetrical peptidolactones and peptidothiolactones.34 Chromodepsipeptides are equipped with heteroaromatic substituents that enable DNAbis-intercalation, endowing many members of this family with cytotoxic, antiviral, and antibiotic activity. Thiocoraline is a promising anticancer agent and is structurally intriguing, as it is composed of two tetrapeptide chains fused into a bicyclic system by one disulfide and two thiolactone bonds (Scheme 3). Analysis of the thiocoraline (tio) biosynthetic gene cluster indicated that the tetrapeptide chains are assembled by an iteratively acting tetramodular NRPS (TioR and TioS) and implicated a TE domain in the macrothiolactonization.35 The proposed macrothiolactonization reaction was successfully reconstituted using the heterologously produced thiolation− thioesterase didomain (T−TE) of module 4 and synthetic substrates.36 As predicted, the T−TE didomain catalyzed both the ligation of tetrapeptidyl−thioesters and the subsequent

AgrB (Cys84Ser) and was sensitive to pH and temperature, this species is believed to be the enzyme-bound thioester intermediate. Following thiolactone formation, the peptide is secreted from the cell and the N-terminal leader peptide is removed by the signal peptidase, SpsB, to afford the mature natural product (Scheme 1).26 While these studies provided a mechanism for the biosynthesis of AIP, they did not provide an explanation for how AgrB catalyzes this thermodynamically unfavorable reaction. Reconstitution of liposome-incorporated AgrB in vitro demonstrated that the reaction does not require an exogenous energy source, ruling out the coupling of thiolactone formation with ATP hydrolysis.27 Rather, the thermodynamic sink is overcome by the partitioning of the thiolactone intermediate into the membrane and rapid proteolytic degradation (t1/2 ≈ 10 s) of the follower peptide. Apart from AIP, thiolactone macrocycles are also present in nosiheptide and glycothiohexide, members of the thiopeptide subclass of RiPPs.28−30 The discovery of the nosiheptide Scheme 2. Structures of Thioester-Containing Thiopeptides

5524

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

macrocyclization of the octapeptidyl intermediate. In the presence of a substrate analogue with a hydroxyl nucleophile in lieu of the Cys thiol group, macrolactones instead of the native macrothiolactones were obtained.

Scheme 3. Macrothiolactonization Mediated by a Thioesterase Domain of the Thiocoraline Assembly Line

2.3. (Ketosynthase-like) Acyltransferases

Canonical ketosynthases catalyze the Claisen condensation of acylthioesters and malonylthioesters to produce a diverse array of natural products.32,37 Recent work has demonstrated that select ketosynthase homologues function as acyltransferases that form ester or amide linkages rather than C−C bonds.38−43 Such an enzyme has been proposed to form a thioester in calicheamicin biosynthesis. 2.3.1. Calicheamicin. Formation of the thioester in calicheamicin, the DNA-cleaving enediyne, likely involves an unusual type of acyltransferase. Analysis of the calicheamicin (cal) biosynthetic gene cluster of Micromonospora echinospora ssp. calichensis revealed a gene (calO4) that was initially believed to encode a ketoacyl-ACP synthase.44 Homologues of CalO4 are involved in the biosynthesis of glycosidic natural products such as chlorobiocin (CloN2),45 chlorothricin (ChlB3),46 and cervimycin (CerJ).47 In all cases, sugar residues are linked to rare building blocks via ester bonds. A phylogenetic analysis showed that CalO4 falls into a clade of ACP shuttle enzymes that are closely related to FabH-like enzymes, ketosynthase III homologues involved in the priming of fatty acid and polyketide synthases.47 Even so, functional analyses in vivo and in vitro unequivocally demonstrated that these enzymes function as acyltransferases that load activated acyl groups onto hydroxyl substituents of sugars. By analogy, it appears plausible that CalO4 transfers the acyl residue onto a thiol group of a thiosugar (Scheme 4). The thiol group of the sugar building block may be introduced by attack of a sulfur nucleophile onto a carbonyl group, followed by a lyase reaction.48 However, the proposed mechanism for the transacylation and the biosynthetic origin of the thiosugar require verification.

3. S-TRANSFERASES The biosynthesis of many sulfur-bearing natural products involves designated S-transferases, which catalyze the attack of a thiol to an activated carbon without the need of ATP. This reaction is similar to the thioester formations described in section 2 but does not involve the cleavage of activated ester or thioester moieties. There are two types of S-transferases, Smethylstransferases and S-glycosyltransferases, which are better known for their ability to form C−O, C−N, and C−C bonds using S-adenosylmethionine (SAM) and activated sugars as cosubstrates.11,49,50 The third S-transferase presented herein, glutathione-S-transferase, loads thiol groups onto different types of activated carbons. In addition to these three wellcharacterized enzyme families, there are several S-transferases that do not fall into these categories but share the ability to promote Michael additions of thio nucleophiles to diverse α,βunsaturated acceptor systems.

Scheme 4. Proposed Mechanism of Thioester Formation in Calicheamicin Biosynthesis

3.1. S-Methyltransferases (SAM-dependent)

From a chemical perspective, methylation of a thiol represents one of the simplest types of C−S bond formations. It is textbook knowledge that S-methylation is the final step in methionine biosynthesis from homocysteine.51 This reaction is often catalyzed by a cobalamine-dependent methionine synthase, which utilizes 5-methyltetrahydrofolate (N 5 MeTHF) as the methyl donor. SAM-dependent S-methylations 5525

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

are also found in thiopurine detoxification pathways and as posttranslational modifications.50,52 In addition, the stepwise Smethylation of sulfide is a possible route to dimethylsulfide (Scheme 5), the most abundant organosulfur compound in the Scheme 5. SAM-Dependent Methylation of Hydrogen Sulfide and Organic Thiols

Figure 2. Crystal structure of TmtA in complex with Sadenosylhomocysteine (PDB entry code: 5EGP).

atmosphere, which is often referred to as the “smell of the sea”. Yet, this route has only been demonstrated in vitro with cellfree extracts of the ciliate Tetrahymena thermophila using SAM and sulfide or methanethiol as substrates.53 Although SAMdependent O-, N-, and C-methyltransferases are frequently associated with natural product biosynthetic pathways, until recently no genuine enzyme for a C−S bond formation was known for secondary metabolites. 3.1.1. Bismethylgliotoxin. An unusual S-methyltransferase involved in natural product modification was discovered in the context of a self-resistance mechanism against gliotoxin, a potential virulence factor of the human-pathogenic fungus Aspergillus f umigatus.54 Gliotoxin exerts its damaging effects through protein conjugation and redox-cycling, which requires the formation of free thiol groups. Two independent mechanisms confer resistance toward gliotoxin. First, the epidithiol may be oxidized into the dithio bridge by a flavin adenine dinucleotide (FAD)-dependent oxidase (GliT);55 deletion mutants of gliT showed significant higher sensitivity toward gliotoxin in comparison to the wild type.56 Second, long-term cultures of A. f umigatus produce bismethylgliotoxin (Scheme 6), which indicates that the thiol groups could be

are known, it seems that this back-up protection strategy is widespread among ETP-producing fungi. Notably, a similar scenario has been proposed for bacteria, specifically for holomycin biosynthesis and resistance in Streptomyces clavuligerus. Formation of the cyclic ene−disulfide involves a thioredoxin oxidoreductase-like enzyme (HlmI) that is homologous to GliT, and a hlmI deletion mutant produces Smethylated holomycin derivates.60 However, the corresponding enzymes catalyzing the S-alkylation have not yet been identified. 3.1.2. Collismycin. Whereas S-methylation of gliotoxin may be regarded as a detoxification pathway, several S-methyltransferases (S-MTs) are involved in the assembly of biologically active natural products. Collismycins are rare 2,2′bipyridyl natural products produced by a Streptomyces sp. that were initially identified as inhibitors of dexamethasone− glucocorticoid receptor binding.61 Since then, neuroprotective, antibacterial, and antifungal bioactivities have been attributed to the collismycins.62−64 2,2′-Bipyridyl natural products are biosynthesized by a hybrid PKS/NRPS from malonyl-CoA, lysine, cysteine, and leucine (Scheme 7).65−67 Following the formation of the bipyridyl ring system, the terminal leucine residue is cleaved by an amide hydrolase and the newly formed carboxylic acid is converted to an oxime by a multienzyme pathway. Although the collismycin biosynthetic pathway has not been fully elucidated, a SAMdependent S-methyltransferase is known to play a critical role. When a putative methyltransferase (clmM1) in the collismycin A producer Streptomyces sp. CS40 was inactivated, the resultant strain was unable to produce collismycin A but instead produced collismycin SN (Scheme 7).65 This intermediate was successfully converted into the corresponding S-methylated derivative, collismycin SC, by a mutant strain lacking the downstream-acting amide hydrolase, ClmAH. On the basis of these results, collismycin SN was proposed to be the product of bipyridyl ring formation, and ClmM1-catalyzed S-methylation was predicted to facilitate the cleavage of the S−N bond to yield collismycin SC.65 An alternative interpretation of the data is that S-methylation is required for recognition of collismycin SC by ClmAH and that collismycin SN is a byproduct formed from the spontaneous oxidation of the free thiol intermediate. When fed to the ClmAH-deficient strain, the reduction of collismycin SN would provide the free thiol for S-methylation by ClmM1. While it is clear that ClmM1 is responsible for Smethylation in collismycin A maturation, further work will be required to determine the exact biosynthetic route to collismycin A.

Scheme 6. Enzymatic Inactivation of the Reactive Epidithiol Form of Gliotoxin: (Top) GliT-Mediated Oxidation into the Epidithio Bridge (Impermanent) and TmtA-Catalyzed bis-SMethylation (Permanent); (Bottom) Analogous Route for Holomycin Processing

permanently inactivated by bisalkylation. In vitro and in vivo experiments showed that the S-methyltransferase TmtA, which is not encoded in the gli gene locus, is responsible for gliotoxin S-methylation.57,58 Analysis of the crystal structure of TmtA in complex with S-adenosylhomocysteine showed that one substrate and one cofactor binding pocket are present in each monomer (Figure 2), which suggested that the bisthiomethylation of gliotoxin occurs sequentially.59 Because many dimethylated epipolythiodioxopiperazine (ETP) derivatives 5526

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 7. Proposed Pathway for S-Methylation in Collismycin Biosynthesis; Two Routes for ClmM1-Catalyzed S-Methylation Are Shown

3.1.3. Echinomycin. An S-methyltransferase that rearranges a disulfide bond has been discovered in the context of echinomycin biosynthesis. Echinomycin islike thiocoraline (see section 2.2.1)an important member of the chromodepsipeptide family and can also intercalate into duplex DNA.34 Echinomycin differs from thiocoraline mainly in the pair of quinoxazoline residues and the substitution of the thiolactone moieties for lactones. In addition, echinomycin features a rare thioacetal bridge, which endows the bicyclic core with greater stability compared to the labile disulfide bond. Biosynthetic studies with the echinomycin producer Streptomyces lasaliensis revealed that this unique chemical motif is obtained by conversion of the disulfide bond of triostin Aand that this reaction is catalyzed by a single SAM-dependent methyltransferase (Ecm18) (Scheme 8).68 The crystal structure of Ecm18 in complex with echinomycin and S-adenosylhomocysteine (Figure 3) indicated that the disulfide−thioacetal conversion occurs in two stages. First, one sulfur atom is methylated; then the carbon adjacent to the sulfonium is deprotonated, yielding an ylide that rearranges into the thioacetal via a charge-delocalized transition state (Scheme 8).69 This model is supported by the finding that a singly Smethylated product is produced when the reduced form of triostin A is used as a substrate in the Ecm18 enzyme assay. 3.1.4. Lincomycin A. Besides formation through disulfide conversion, a thioacetal may be formed by alkylation of an anomeric thiol group as in the biosynthetic pathways leading to lincosamide antibiotics. This small family of ribosome-targeting compounds comprises lincomycin A, Bu-2545, and celesticetin (see section 3.2.3), each bearing an unusual 1-thiooctose sugar (Scheme 9).70 Early biosynthetic studies demonstrated that the methylthiolincosamide and propylhygric acid moieties of lincomycin A are produced independently and then linked; however, a biosynthetic route for the incorporation of the αlinked methylthiol moiety could not be provided.48,71−75 Only recently, the highly complex pathway has been investigated in detail (for sulfur incorporation route, see section 3.2.3). In the course of these studies, an S-methyltransferase (LmbG) was discovered that alkylates the thiol group of the octose (Scheme 9).76,77 3.1.5. Thiocoraline. A bifunctional enzyme with Smethyltransferase and amino acid adenylation activities has recently been identified that is involved in thiocoraline

Scheme 8. Conversion of a Disulfide Bond to the Thioacetal Group of Echinomycin by a SAM-Dependent Methyltransferase

biosynthesis (see section 2.2.1). In addition to the unusual thiolactone bonds, thiocoraline contains two extraordinary Smethylated L-cysteine units. Gene inactivation and biochemical analyses showed that TioN plays an essential role in the biosynthesis and activation of this rare amino acid.78 TioN is a stand-alone enzyme harboring an adenylation domain that is interrupted by the conserved SAM-binding region of SAM methyltransferases. TioN is remarkable in that it is the only interrupted, stand-alone A domain identified to date; all other 5527

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 10. Activation and S-Methylation of Cysteine to Produce and Incorporate the Rare S-Methyl Cysteine Residues of Thiocoraline

Figure 3. Crystal structure of Ecm18 in complex with echinomycin and S-adenosylhomocysteine (SAH; PDB entry code: 4NEC). The histidine residue thought to be responsible for the deprotonation of the β-carbon is displayed.

interrupted adenylation domains are components of NRPS modules.79 In vitro studies revealed that TioN adenylates Lcysteine, S-methylates the thiol group, and loads the amino acid onto its cognate thiolation (T) domain in module 4 (Scheme 10).78 Kinetic characterization of TioN demonstrated that the catalytic efficiency of the adenylation reaction is ∼500 times greater for L-Cys than for S-Me-L-Cys, indicating that adenylation precedes methylation. While TioN was able to catalyze methylation prior to or after T domain loading in vitro, the order of methylation and loading is not known in vivo. 3.1.6. Brassinin. Brassinin is a plant defense compound produced by cruciferous vegetables. This important member of the indole sulfur phytoalexin class of natural products has antifungal and anticancer activities and is a key intermediate in the biosynthesis of additional indole sulfur phytoalexins (section 6.1.2).80,81 Labeling experiments demonstrated that brassinin is biosynthesized from indole glucosinolate, likely through an isothiocyanate intermediate.82 Leveraging the observation that phytoalexins from Brassica rapa are only produced upon pathogen exposure, a recent study partially characterized the brassinin biosynthetic pathway.83 Plant leaves were challenged with Pseudomonas syringae pv maculicola and RNA sequencing was performed to identify genes that were upregulated following exposure. Among others, the expressions of a pyridoxal 5′-phosphate (PLP)-dependent C−S lyase (SUR1) and a SAM-dependent methyltransferase (DTCMT.a) were increased. Reconstitution of both enzymes in vitro demonstrated that SUR1 converts indole isothiocyanate cysteine to indole dithiocarbamate, which is subsequently methylated by DTC-MT.a to yield brassinin (Scheme 11). This biosynthetic proposal was further verified when brassinin was produced by tobacco leaves coexpressing the B. rapa glucosinolate and brassinin biosynthetic genes.

A SUR1 homologue (AtSUR1) is found in Arabidopsis thaliana, where it was previously implicated in glucosinolate biosynthesis (section 3.3.3).84 Notably, reconstitution of the activity of AtSUR1 in vitro demonstrated that it can use both cysteine−isothiocyanate conjugates and S-alkyl thiohydroximates as substrates.83 While further studies will be required to understand the remarkable substrate promiscuity of SUR1, the current data demonstrate that the enzyme plays a central role in both glucosinolate and phytoalexin biosynthesis. 3.1.7. S-Methylated RiPPs. S-Methylated RiPPs are quite rare; to the best of our knowledge, S-methylations have only been observed in select members of the thiopeptide natural products (Scheme 12).85−87 In each of these cases, the biosynthetic gene clusters, and thus the enzymes required for C−S bond formation, are unknown. Recently, a member of the proteusin RiPP class was proposed to be S-methylated, and the responsible S-methyltransferase was partially characterized. The proteusins are an emerging class of RiPP natural products grouped by their unusually large leader peptide and the presence of radical SAM enzymes in the biosynthetic gene cluster.22 The only known member of the class, polytheonamide, is produced by an unculturable bacterial symbiont of the marine sponge Theonella swinhoei.88 The genome sequence of the organism, named Candidatus Entotheonella factor, indicates that it has the potential to produce a large number of natural products, including a second proteusin.89 Although the corresponding natural product has not been isolated, the posttranslational modifications installed on the core peptide were partially characterized by coexpressing the precursor peptide (PtyA) and select tailoring genes in Escherichia coli.89,90 When PtyA was coexpressed with a methyltransferase

Scheme 9. S-Methylation of an Anomeric Thiol That Leads to the Thioacetal Residue in Lincomycina

a

An analogous modification occurs in the biosynthesis of Bu-2545; however, the responsible enzyme has not been characterized. 5528

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 11. Model for Brassinin Biosynthesis in B. rapa (L.R., Lossen Rearrangement; GST, Glutathione S-Transferase)

Scheme 12. Structures of Known S-Methylated RiPP Natural Products

the family of glycocins, RiPP natural products that are posttranslationally glycosylated at select serine and cysteine residues.22 All characterized glycocins have antibacterial activities, but the mode of action and the role of the glycosylation in the bioactivity are still under investigation.96 The enzyme responsible for C−S bond formation is a glycosyltransferase 2 (GTase-2) protein family member.92,93 Of the known examples, only the glycosyltransferases involved in the biosynthesis of sublancin and the thurandacins, SunS and ThuS, respectively, have been biochemically characterized.92,96,97 Consistent with characterized GTase-2 family members, SunS and ThuS catalyze the transfer of a nucleotide diphosphate-activated glucose molecule to the precursor peptide (Scheme 13). In both cases glycosylation occurs with inversion of the anomeric center, which suggests an SN2 displacement.94,98 On the basis of the structure of glycocin F and enterocin F4− 9 (a glycocin member containing only O-glycosyl residues), it appears that this mechanism is likely conserved in the glycocin family.99,100 Both SunS and ThuS display a remarkable substrate tolerance in terms of the peptide substrates and the sugar donor cosubstrates that can be transferred.92,94,97 SunS is particularly notable among RiPP biosynthetic enzymes in that it does not require the leader peptide for substrate recognition.97,101 Enzyme reactions carried out with truncated versions of the precursor peptide demonstrated that the enzyme recognizes an α-helix N-terminal to the site of modification; however, the molecular details of this interaction are unknown.97 Moreover, ThuS is not chemoselective and can

homologue (PtyS), the peptide was methylated up to two times.90 The modifications were assigned to the thiol groups of cysteine residues in the core peptide by iodoacetamide labeling and mass spectrometry. The biosynthetic gene cluster also contains a gene encoding a bifunctional lanthipeptide dehydratase/cyclase (PtyM). Coexpression of PtyA and PtyM in E. coli resulted in the installation of up to three lanthionine rings on the core peptide of PtyA.89 Moreover, when PtyA was coexpressed with both PtyA and PtyM, a product bearing three lanthionine rings and a single S-methylated cysteine was produced.90 While these data suggest that the proteusin-like RiPP will be S-methylated, the natural product will need to be isolated before a final verdict can be reached. 3.2. S-Glycosyltransferases

In synthetic chemistry, thioglycosides are well-known as potent glycosylation agents.91 In contrast, natural products bearing Sglycosyl residues are rare, likely because glycosyltransferases preferentially load activated sugars such as UDP-glucose onto O-, N-, and C-nucleophiles.49 Even so, S-glycosylations are involved in the biosynthetic pathways of important biologically active natural products such as S-linked glycopeptides, glucosinolates, and lincosamide antibiotics. 3.2.1. Glycocins. Whereas peptidyl O-glycosylation is a very common posttranslational modification, only five examples of S-glycosylated peptides are known: three isolated from a bacterial source (sublancin, glycocin F, and ASM1) and two generated by the in vitro maturation of a precursor peptide (thurandacins A and B).92−95 All of these examples belong to 5529

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

potentially harmful plant metabolites. By screening of a cDNA library of B. napus, the first thiohydroximate S-glycosyltransferase implicated in glucosinolate biosynthesis was discovered. The deduced amino acid sequence shows a highly conserved motif for a glucose-binding domain and is overall highly similar to glucosyltransferases characterized in other species. The heterologously produced enzyme was partly characterized in vitro, inferring UDP-glucose/thiohydroximate S-glucosyltransferase activity by measuring glucose incorporation.104 A functionally related putative thiohydroximate S-glucosyltransferase (UGT74B1) was identified in Arabidopsis thaliana.105 In vitro analyses using recombinant UGT74B1 showed its ability to S-glucosylate phenylacetothiohydroximic to yield the corresponding desulfoglucosinolate (Scheme 14). Kinetic analyses of UGT74B1 with thiohydroximates and the decrease in glucosinolate production in mutants lacking a functional ugt74b1 gene suggested that the enzyme represents a designated S-glucosyltransferase in glucosinolate biosynthesis. Notably, UGT74B1 proved to be a versatile biocatalyst for the chemoenzymatic synthesis of desulfoglycosinolates.106 3.2.3. Lincomycin A and Celesticetin. A unique enzymatic S-glycosylation−transglycosylation sequence takes place in lincomycin biosynthesis. The longstanding mystery as to how the C−S bond is formed in lincomycin A was recently solved.107 A functionally uncharacterized gene, lmbE, with homology to mycothiol (MSH)-S-conjugate amidase was inactivated in the lincomycin biosynthetic gene cluster in Streptomyces lincolensis. The resulting mutant strain produced an α-S-linked MSH−lincomycin conjugate (Scheme 15). Accordingly, reconstitution of LmbE activity in vitro demonstrated that this enzyme catalyzed the removal of the sugar moieties from the MSH−lincomycin conjugate. To identify the enzyme responsible for attaching the MSH moiety to the lincosamide sugar, two additional functionally uncharacterized genes, lmbV and lmbT, with homology to MSH-dependent isomerase and glycosyltransferase genes, respectively, were inactivated. Both mutant strains were unable to produce lincomycin; however, in the lmbV mutant strain, a new analogue was produced that bore a β-linkage between

Scheme 13. Peptide S-Glycosylation in the Biosynthesis of Glycocins, e.g., Sublancin

also catalyze the glycosylation of serine and threonine residues; however, S-glycosylation is faster, potentially due to the increased nucleophilicity of the acceptor.94 Given this high level of promiscuity, glycocin glycosyltransferases represent an attractive target for natural product bioengineering efforts. 3.2.2. Glucosinolates. S-Glycosylation is an essential process in the biosynthesis of glucosinolates, which play an important role in plant defense of crucifers (e.g., cabbage and rape/canola).102 These structurally unusual sulfur compounds are also valued as the precursors of the spicy mustard oil components of wasabi and radish, and of sulforaphane, the cancer-protective agent from broccoli. The thioglycosides are actually only storage forms for chemically reactive defense compounds. Upon wounding or infection of the plant, thioglycosidases (myrosinases) cleave the glycosidic C−S bond, thus liberating aglycones with free thiols that readily undergo Lossen-type rearrangements into the corresponding isothiocyanates (Scheme 14). In oilseed rape (Brassica napus), the glucosinolate-derived thiocarbamate progoitrin has implications for food toxicology because it impairs the production of thyroid hormones.103 Thus, glucosinolate biosynthesis in rape has been a target for molecular breeding approaches to reduce the content of

Scheme 14. General Glucosinolate Biosynthetic Pathway Involving S-Glycosylation and O-Sulfation (by Sulfotransferase, SOT)a

a

The thioglycosidic bond is cleaved by a myrosinase, thus triggering a Lossen rearrangement to afford the isothiocyanate. Structures of representative isothiocyanates from plants are shown. Goitrin is formed from the uncatalyzed cyclization of the corresponding isothiocyanate. 5530

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 15. S-Glycosylation and Transglycosylation Steps in Lincosamide Biosynthesis

SAM-dependent S-methyltransferase (LmbG; see Scheme 9).76,77 In celesticetin biosynthesis, the LmbF homologue, CcbF, catalyzes the oxidative decarboxylation of the cysteine side chain to afford the aldehyde derivative (Scheme 15). This intermediate is subsequently modified to afford the salicylic ester. It should be noted that the concept of sulfur introduction by a substitution reaction with cysteine or a cysteinyl-bearing larger molecule and its subsequent decomposition to a thiol will be revisited in numerous other pathways, specifically in those employing glutathione S-transferases.

lincomycin and ergothioneine (EGT; Scheme 15). Reconstitution of CcbV (the LmbV homologue from the celesticetin BGC) activity in vitro demonstrated that the enzyme catalyzed the SN2 displacement of EGT with MSH to form the α-linked C−S bond found in lincomycin A. Reconstitution of the GTase, LmbT, demonstrated that the enzyme catalyzed the inverting glycosylation of EGT with GDP-activated lincosamide (Scheme 15). The EGT−lincosamide conjugate is then linked to the propylhygric acid by the action of LmbC, LmbN, and LmbD. While these results demonstrated how the α-linked C−S bond was formed, it did not provide an explanation for the maturation of the deglycosylated MSH−lincomycin analogue into lincomycin A. Reconstitution of LmbF, an aspartate aminotransferase fold type I (AAT_1) superfamily member, demonstrated that these enzymes catalyze the pyridoxal 5′phosphate (PLP)-dependent degradation of the cysteine residue to form desmethyl lincomycin A, the substrate of the

3.3. Glutathione-S-transferases

Glutathione-S-transferases (GSTs) are best known for their involvement in xenobiotic-detoxification processes where they yield water-soluble conjugates.108,109 However, related thioconjugate-forming enzymes have also been implicated in biosynthetic pathways, most notably in the construction of the epidithio bridges of fungal diketopiperazines that may serve 5531

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 16. Nucleophilic Epoxide Opening and S-Conjugation by a Glutathione S-Transferase Leads to the Slow-Reacting Substance of Anaphylaxis; Downstream Peptidases Convert Leukotriene C4 into Leukotriene E4

Scheme 17. Incorporation of a Dithiol Moiety into the Glutathione Diketopiperazine Core by a Specialized GST

as virulence factors. They have also been implicated in leukotriene, glucosinolate, and allicin biosynthesis, as well as in transient C−S bond formation in E-/Z-isomerizations. 3.3.1. Glutathionyl and Cysteinyl Leukotrienes. The glutathione-detoxification pathway in combination with the

eicosanoid pathway gives rise to a group of sulfur-bearing leukotrienes, LTC4, LTD4, and LTE4 (Scheme 16). These compounds are also known as the “slow-reacting substance of anaphylaxis” (SRS-A), as they represent proinflammatory mediators that are produced during acute asthma attacks as 5532

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 18. Detailed Mechanism for the Thiol-Liberating C−S Lyase Reaction in the Gliotoxin Pathway

Scheme 19. Model for Glucosinolate Biosynthesis Involving a Glutathione-S-Transferase (GST) (CYP, Cytochrome P450; SOT, Sulfotransferase)

Almost five decades ago, isotopic labeling experiments showed that L-(35S)-cysteine is a sulfur source for sporidesmin biosynthesis in P. chartarum.125 Yet, only recently, the mechanistic details for C−S bond formation in ETPs have been elucidated in A. f umigatus. A nonribosomal peptide synthetase, GliP, assembles the DKP scaffold, which is subsequently hydroxylated by the cytochrome P450 (CYP) monooxygenase GliC at the α-carbon position. According to the current biosynthetic model, dehydration would yield an acyliminium intermediate. This electrophilic intermediate is believed to be the cosubstrate of a specialized GST, GliG, which catalyzes the nucleophilic attack of glutathione (Scheme 17).126,127 To degrade the glutathione backbone and liberate the thiol, three distinct enzymes are required. A γ-glutamyl-transferase (GliK) cleaves the isopeptide bond, releasing pyroglutamic acid in the process. The resulting Cys-Gly dipeptide is hydrolyzed by an unusual metal-dependent dipeptidase (GliJ).128 The final step of the degradation cascade is catalyzed by the PLPdependent C−S lyase GliI (Scheme 18).129 The resulting epidithiol is then oxidized by a flavin adenine dinucleotide (FAD)-dependent oxidase GliT to produce the disulfide bridge.55,56,94 On the basis of sequence homologies, the pathway for the introduction of dithiol residues appears to be highly conserved in ETP biosynthesis. Notably, this strategy of oxygenation followed by nucleophile addition mirrors phase I/II xenobioticdetoxification pathways130 found in diverse organisms and is involved in glucosinolate production in crucifers (see section 3.3.3). Given the widespread nature of this detoxification pathway, it is conceivable that additional secondary metabolites are synthesized in an analogous way. 3.3.3. Glucosinolates. As briefly mentioned in section 3.2.2, glusosinolates (often also referred to as mustard oil

well as after allergen and exercise challenge.110 In particular, cysteinyl leukotriene (LTEA) is one of the most potent bronchoconstrictors known to date.111 The glutathionyl-Sconjugate (LTC4) has been identified as the major trigger of stress-induced oxidative DNA damage.112 LTC4, LTD4, and LTE4 are produced de novo from membrane phospholipids in the context of the leukotriene route of the arachidonic acid cascade.113 The epoxide leukotriene A4, which results from leukotriene lipoxygenation,114 is the precursor of leukotrienes C4 and D4.115 “Leukotriene C4 synthase” was found to be a designated glutathione S-transferase that catalyzes the epoxide ring opening by the nucleophilic glutathione thiol. 113 The glutathionyl adduct is then degraded into leukotriene D4 by means of a γ-glutamyl transferase.116,117 Finally, leukotriene D4 is converted into leukotriene E4 by a dipeptidase (Scheme 16).118 A similar merger of detoxification and biosynthetic enzymes has also taken place in the biosynthetic pathways to ETPs and glucosinolates described below. 3.3.2. Epipolythiodiketopiperazines. Numerous natural products that are characterized by diketopiperazine (DKP) cores and transanullar disulfide bridges constitute the large family of epipolythiodiketopiperazines (ETPs).119 The unusual sulfur functionality of ETPs is mainly responsible for their diverse biological activities, because the dithio group promotes redox cycling, formation of reactive oxygen species, and protein conjugation.120,121 Infamous mycotoxins belonging to the ETPs are found among human, animal, and plant pathogens. For example, gliotoxin is a potential virulence factor of Aspergillus f umigatus,54 sporidesmin from Pithomyces chartarum has been shown to be involved in the development of facial eczema in sheep,122,123 and sirodesmin is a virulence factor of the canola pathogen Leptosphaeria maculans.124 5533

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 20. Model of Allicin Biosynthesis

transform the oxime hydroxyl into the required leaving group for the aforementioned Lossen rearrangement.149 3.3.4. Allicin/Alliin. An S-glutathionylation−degradation route similar to those observed in the biosyntheses of cysteinyl leukotrienes, ETPs, and glucosinolates has been implicated in the biosynthesis of allicin (diallylthiosulfinate). This sulfurcontaining defense molecule from garlic (Allium sativum) and other Allium species possesses various biological activities and is responsible for the characteristic odor of freshly cut or crushed garlic, which actually indicates an induced plant defense mechanism.150 Damage of the plant tissue activates an enzymatic trigger mechanism, which leads to the formation of allicin from the nonproteinogenic amino acid alliin (S-allyl-Lcysteine sulfoxide).151−153 This C−S cleavage reaction is catalyzed by alliinase, a PLP-dependent lyase, and yields ammonia, pyruvate, and an unstable sulfenic acid that readily forms alliin by dimerization and dehydratation.154 On the basis of precursor feeding, radiolabeling experiments, and isolation of γ-glutamyl-S-allyl cysteine, it was assumed that S-allyl cysteine derives from glutathione S-conjugates.155−157 In addition, labeling experiments with 14C-valine yielded radiolabeled methacrylic acid, a possible source of the allylic residue of alliin and plausible Michael acceptor (Scheme 20).158 Yet, the exact reaction and the involvement of a glutathione Stransferase have remained elusive.159 Recently, enzymes for the cleavage of the glutamyl group of the proposed glutathione-derived intermediate have been identified. Three S-allyl-cysteine-forming γ-glutamyl transpeptidases (AsGGT1−3) with different kinetic properties and subcellular localizations were characterized.160 The next step in the pathway is the stereoselective S-oxygenation of S-allyl cysteine by a FAD-dependent monooxygenase (AsFMO1) to yield alliin.161 The preference of AsFMO1 for S-allyl L-cysteine over γ-glutamyl-S-allyl L-cysteine established the order in which the reactions occur in garlic. 3.3.5. Pseurotin (trans to cis Isomerization). GSTs have also been implicated in the biosynthesis of fungal polyketides that lack a C−S bond. The biosynthesis of the pseurotin class of fungal natural products requires the formation of a high-energy cis double bond between carbons 12 and 13; however, the hybrid polyketide synthase−nonribosomal peptide synthetase involved in pseurotin biosynthesis releases the precursor in the E-conformation.162,163 Through a combination of gene-deletion studies and in vitro enzyme-reconstitution assays, a GST

glycosides) are a hallmark of plants in the order Brassicales and contribute to the pungent taste of cabbages and radishes.102 In the natural environment, these structurally intriguing thioglycosides serve as important defense compounds against herbivore attack and microbial infection. Upon attack of the plant by pathogens, the thioglycoside linkage of the glucosinolates is cleaved by specialized β-thioglucosidases called myrosinases.131 The unstable aglycones undergo various rearrangement reactions. Glucosinolate diversity results from the incorporation of various aliphatic and aromatic amino acids.132 Free amines are uniformly converted into aldoxime moieties by CYP monooxygenases.133−137 After decarboxylation, CYPs transform the aldoximes into nitrile oxides, which may serve as electrophiles for C−S bond formation (Scheme 19).138−142 Whereas these reactive intermediates may react with different thiols in vitro, classical plant-feeding studies suggested that cysteine is the preferred sulfur source in vivo.143 However, three recent lines of evidence indicate that glutathione rather than cysteine is the true sulfur donor. First, mutants lacking glutathione biosynthetic genes were not able to induce indolic glucosinolate production upon fungal or herbivore attack.144,145 Second, two GSTs were found to be coexpressed with glucosinolate biosynthetic genes in Arabidopsis.146 Third, GSH adducts were found in Nicotiana benthaminia expressing glucosinolate pathway genes.147 In analogy to gliotoxin biosynthesis, a γ-glutamyl peptidase (GGP1) was identified that could catalyze the first cleavage step to degrade the GSH backbone (Scheme 19). An Arabidopsis double mutant lacking ggp1 as well as a redundant gene ggp3 showed reduced levels of glucosinolates.148 The next known step in the pathway is catalyzed by the C−S lyase SUR1, which has been shown to be essential for glucosinolate production in Arabidopsis.84 SUR1 was recently reconstituted in vitro and was shown to cleave S-cysteine thiohydroximate to thiohydroximate.83 Although the enzyme was not tested for the ability to process the Cys-Gly dipeptide linked aldoxime, these data strongly suggest that the glycine residue is cleaved off prior to SUR1 processing. As outlined in section 3.2.2, the SUR1 reaction product is S-glycosylated by specialized glucosyltransferases to produce desulfoglucosinolates.105 The last step in glucosinolate biosynthesis is catalyzed by 3′-phosphoadenosine5′-phosphosulfate (PAPS)-dependent sulfotransferases, which 5534

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 21. Glutathione-S-Transferase Mediates E/Z-Isomerization of a Double Bond in Pseurotin Biosynthesis; GSH, Glutathione

3.4. S-Transferases Involved in Conjugate Additions

(PsoE) and a bifunctional methyltransferase/FAD-dependent monooxygenase (PsoF) were shown to be required for the olefin isomerization.164,165 When the 12,13-E-configured intermediate, presynerazol, was incubated with glutathione and NADPH in the presence of PsoE and PsoF, the 12,13-Zconfigured compound synerazol was obtained.164 Importantly, neither enzyme was capable of catalyzing the isomerization alone. The isomerization is proposed to proceed through a Sconfigured glutathione−presynerazol conjugate at C-13 that is subsequently oxidized to the sulfoxide by PsoF (Scheme 21). Although these products have not been isolated, compounds with a mass/charge ratio consistent with these conjugates were detected in the enzyme reactions. The sulfoxide intermediate is thought to undergo a PsoF-catalyzed pericyclic syn-elimination to yield Z-configured presynerazol, which is further oxidized to synerazol by PsoF. The crystal structure of PsoE with bound presynerazol and GSH (Figure 4) shows the characteristic GST

In several biosynthetic pathways, enzymes or domains catalyze the formation of C−S bonds by conjugate addition of a sulfur nucleophile to enones or α,β-unsaturated acids and thioesters. Yet, these biocatalysts are only distantly related to wellcharacterized enzymes such as GSTs and/or use different substrates. In contrast to the cyclases outlined below (section 4.1), which also mediate conjugate additions, the S-transferases included in this section do not lead to intramolecular cyclizations. Notably, mechanistically related reactions can also take place without enzyme catalysis (section 8.1). 3.4.1. Tropodithietic Acid and Roseobacticide A. A GST homologue (TdaB) has been implicated in the biosynthesis of unusual sulfur-substituted tropolone derivatives that regulate the symbiotic interaction of marine alpha-proteobacteria (clade Roseobacter) with their algal hosts. As long as the microalgae (Emiliania huxleyi) provide the bacteria with food, the bacteria (Phaeobacter inhibens) produce tropodithietic acid (TDA), an antibiotic that protects the host. However, the algae release p-coumaric acid upon senescence, which triggers P. inhibens to synthesize algicidal roseobacticides.167 Stable isotope-labeling studies indicated that the carbon backbone TDA is derived from phenylacetic acid and that sulfur amino acid metabolism serves as the source for the sulfur.168,169 On the basis of in silico analyses of the TDA biosynthetic gene cluster, isotope labeling, and mutational studies, a plausible sulfurization pathway was proposed (Scheme 22).170 First, phenylacetic acid is transformed into a seven-membered carbacycle by oxidative ring cleavage. Then, dehydrogenation (by TdaE) and dehydration (by TdaC) yield the tropolone ring. The GST homologue TdaB is believed to catalyze the conjugate addition of S-thiocysteine, rather than glutathione, to nascent tropolone ring. Subsequently, the flavoprotein TdaF could mediate an oxidative cleavage of the disulfide adduct. The thioaldehyde byproduct was proposed to be converted to cysteamine and subsequently recycled in primary metabolism;170 however, it could also spontaneously degrade to hydroxypyruvate, ammonia, and hydrogen sulfide. The reaction cycle is repeated at the adjacent carbon to afford the dithiol. The vicinal thiol groups could undergo either spontaneous oxidation by molecular oxygen or enzymatic-catalyzed oxidation, as in the gliotoxin and holomycin pathways. However, a homologue of GliT or HlmI is not encoded in the TDA gene cluster.

Figure 4. Crystal structure of PsoE in complex with presynerazol− glutathione conjugate (PDB entry code: 5F8B). A zoomed-in view of the active site shows that the conjugate is bound in a shallow, solventexposed pocket, which may promote transfer to PsoF.

fold, yet the sequence similarity of PsoE to other GSTs is low. PsoE has specific structural characteristics that might favor a fast release of the glutathione conjugate and modification by PsoF. A similar GST-mediated isomerization has been implicated in the biosynthesis of hypothemycin; however, the GST in this case is only able to produce a mixture of 85% E and 15% Z products.166 This low level of conversion suggests that an additional enzyme may be required to catalyze the isomerization effectively. To date, an auxiliary enzyme has not been identified and mechanistic insights into this reaction remain missing. 5535

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 22. Model for the Biosynthesis of Tropodithietic Acid and Roseobacticide Involving a GST-Like Enzyme That Transfers S-Thiocysteine; Box: Biosynthetic Origin of Sulfur and Formation of S-Thiocysteine by PLP-Dependent Degradation of L-Cystine

Scheme 23. Selected Structures of Carbapenem Antibiotics and Model for the Biosynthesis of the Carbapenem Core

been shown to be a source of sulfur that is incorporated into TDA and roseobacticide via cysteine (Scheme 22). The oxidized, dimerized form, cystine, is the substrate for a PLPdependent C−S lyase (PatB), which cleaves the molecule into ammonia, pyruvate, and S-thiocysteine.170,172 This process is analogous to the cysteinyl-degradation pathway studied in gliotoxin biosynthesis (see section 3.3.2). 3.4.2. Thienamycin. Another complex scenario has been observed for the thiolation of the carbapenem core structure en route to thienamycin and related peptidoglycan-targeting

In contrast to TDA formation, only a single TdaB/TdaF reaction cycle is required for roseobacticide biosynthesis. The monothiol intermediate could then be fused to a derivative of p-coumarate, which is released by the senescing algal host, and further modified to afford the algicidal agent (Scheme 22).171,172 The S-methyltransferase responsible for forming the second C−S bond in roseobacticide biosynthesis has not been identified. It is noteworthy that the microalga provides the bacteria food in the form of dimethylsulfoniopropionate (DMSP), which has 5536

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 24. Two Proposed Routes for the Origin of the Cysteamine Side Chain of Thienamycin; Box: Current Proposals for C− S Bond Formation

antibiotics (Scheme 23). The Streptomyces cattleya metabolite thienamycinthe first carbapenem isolatedinspired the development of many potent members of this family of antibacterial agents.173,174 It should be highlighted that, among the β-lactams currently in clinical use, the carbapenems stand out because they are relatively resistant to hydrolysis by most βlactamases.174 Early stable isotope-labeling experiments indicated that the biosynthesis of thienamycinas all carbapenemsdiffers markedly from the assembly of the structurally related β-lactam ring systems of penicillins and cephalosporins.175 Biochemical analyses showed that the carbapenem scaffold is assembled by a crotonase-like carboxymethylproline synthase (CarB/ThnE) that fuses glutamate semialdehyde (or its cyclized form) and malonyl-CoA (Scheme 23).176−179 The resulting proline derivative is then cyclized into the β-lactam by an ATP-dependent enzyme (CarA/ThnM).178−180 In the case of simple carbapenem antibiotics, the carbapenem ring is epimerized and desaturated by a 2-oxoglutarate-dependent nonheme iron oxygenase (CarC) to give carbapenem carboxylic acid.179,181,182 For complex carbapenems, such as thienamycin, the remaining biosynthetic steps are largely unknown; however, significant progress has been made in recent years. The ethyl side chain at position C2 is introduced by two successive SAM-dependent methylations catalyzed by a cobalamin-dependent radical SAM enzyme (ThnK)175,183,184 and is subsequently oxidized by a nonheme iron-dependent oxygenase (ThnG).185,186 Various late-stage enzymatic diversification steps have been identified, including S-oxygenation and N-acetylation.185−187 However, both the source of the

cysteaminyl side chain and the mechanism of C−S bond formation remain unclear. There are currently two hypotheses regarding the source of the sulfhydryl side chain of thienamycin. Radioisotope-labeling experiments conducted with 35S-cystine and 35S-pantethine demonstrated that, while both could be incorporated into thienamycin, cystine incorporation was significantly more efficient.175 This led to the proposal that the cysteaminyl side chain originated from the conjugate addition and decarboxylation of cysteine. However, the discovery of the OA-6129 series of carbapenems (Scheme 24) in a different Streptomyces sp.188 suggested that the sulfhydryl compound used could be pantetheine. Moreover, a randomly generated mutant of the epithienamycin producer, Streptomyces f ulvoviridis, lost the ability to produce the N-acetylcysteamine-substituted carbapenems and instead formed OA-6129 carbapenems.189 This mutant lacked A933 acylase, which, following purification from S. f ulvoviridis, was shown to cleave the pantetheinyl side chain of OA-6129 carbapenems to afford the cysteaminyl derivatives.190−192 Taken together, these findings suggested that OA6129 group carbapenems are the precursors of the cysteaminesubstituted congeners. Subsequently, in vitro studies demonstrated that three enzymes encoded in the thienamycin biosynthetic gene cluster catalyze the stepwise degradation of coenzyme A into cysteamine.187 It was shown that ThnR (a member of the Nudix hydrolase family) and ThnH (another hydrolase) successively cleave coenzyme A to give pantetheine, and that ThnT hydrolyzes the amide bond between β-alanine and cysteamine (Scheme 24).187,193,194 While these experiments provide support for pantetheine being the sulfur source, insertional inactivation of thnR and thnT did not affect 5537

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 25. Model of Online Thiol Conjugate Addition and C−S Lyase Reaction on PKS-Bound Intermediates in the Biosynthesis of Leinamycin (DUF, Domain of Unknown Function; SH, C−S Lyase Domain)

is needed to solve the riddle of C−S bond formation in thienamycin and other thioether-containing carbapenems. 3.4.3. Leinamycin. Another obscure conjugate addition of a sulfur nucleophile takes place in leinamycin biosynthesis in Streptomyces atroolivaceus S-140. Leinamycin (LNM) is a sulfurcontaining cytotoxin that is characterized by an unusual 1,3dioxo-1,2-dithiolane moiety spiro-fused to a thiazole-containing 18-membered lactam ring (Scheme 25).199 The 1,3-dioxo-1,2dithiolane moiety is essential for the antitumor activity of leinamycin, as it alkylates DNA through an episulfonium ion intermediate.200,201 The LMN biosynthetic gene cluster has been identified,202−204 and major parts of the NRPS−PKS pathway to the 18-membered macrolactam have been elucidated. Structures of intermediates and shunt products from mutant strains suggested that sulfur is introduced during polyketide chain elongation, yet after β-alkylation at C-3.205−209 In silico analyses identified two uncommon domains within PKS module 8, a domain of unknown function (DUF) and a tentative lyase (SH) domain.210 It was proposed that the DUF could catalyze the conjugate addition of a sulfur nucleophile (e.g., L-cysteine) to an α,β-unsaturated linear polyketide intermediate. This reaction would be analogous to the recently studied online Michael additions of carbon and oxygen nucleophiles to ACP-bound, α,β-unsaturated thioester intermediates.211 Whereas designated PKS domains have been shown to introduce alkyl branches212−214 and pyran rings215−217 by conjugate addition, biochemical support for the proposed vinylogous S-transfer by the DUF domain is missing. However, using L-cysteine and Smodified L-cysteine analogues as in vitro substrates, the SH domain was shown to act as a PLP-dependent C−S bond lyase that catalyzes the β-elimination of the L-cysteinyl residue to give pyruvate, ammonia, and the corresponding thiol group (Scheme 25).210 It should be noted that the still-enigmatic second C−S bond formation is not essential for bioactivity. The thiol-substituted macrolactam (leinamycin E1), the immediate

theinamycin production under standard cultivation conditions.195 Since the sequence of the thienamycin biosynthetic gene cluster was determined in 2003,196 there have been multiple proposals regarding the installation of the C2 side chain onto carbapenem-3-carboxylic acid (Scheme 24). Initially it was thought that the carbapenem ring would be oxidized to the corresponding carbapenem derivative, which could serve as a substrate for the conjugate addition of the thiol. This conjugate addition was thought to be catalyzed by a glutathione-Stransferase homologue (ThnV) encoded in the gene cluster.196 However, thnV homologues are not found in the biosynthetic gene clusters of all sulfur-containing carbapenem antibiotics.186,197 Mutational analysis of the thn gene locus in S. cattleya revealed that, among other genes, thnN and thnO are essential for thienamycin biosynthesis.196 Crossfeeding experiments indicated that ThnN (and likely also ThnO) perform reactions before methylation of the carbapenem ring, yet after the bicyclic core has been produced.195 Thus, it was hypothesized that these two enzymes may be involved in the conjugate addition. Yet, homologues of ThnN, which contains an amino acid adenylation domain, and ThnO, likely a member of the NAD(P)-dependent aldehyde dehydrogenase superfamily, were found to be responsible for the reduction of a carboxylic acid to an aldehyde in the grixazone biosynthetic pathway (see section 6.4.2), casting doubt on this model.198 A third model for C−S bond formation implicates two additional cobalamin-dependent radical SAM enzymes encoded in the biosynthetic gene cluster (ThnL and ThnP).184 In this proposal, sulfur addition happens prior to oxidation of the carbapenem ring and occurs through a radical mechanism. Consistent with this hypothesis, mutant strains lacking thnL and thnP are unable to produce thienamycin and appear to be blocked early in the biosynthetic pathway.195,196 Moreover, the thnP mutant accumulates carbapenem-3-carboxylic acid.195 Thus, while there have been intriguing clues, further research 5538

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 26. Lanthionine, Labionine, and Aminovinyl Cysteine Ring Formation by Intramolecular Conjugate Addition of a Cysteine Thiol

eCys), and labionin (Lab) (Scheme 26).22,219,220 The biosynthesis of each of these rings begins with the dehydration of individual serine and threonine residues on the precursor peptide by the action of the lanthipeptide dehydratase. To date, three classes of lanthipeptide dehydratases have been characterized. While the classes are not homologous, all three catalyze the activation of the side-chain oxygen followed by elimination of the activated species (Scheme 26). The resultant dehydroalanine (Dha, from serine) and dehydrobutyrine (Dhb, from threonine) residues are further modified by the lanthipeptide cyclase to generate the thioether linkages. For Lan, MeLan, and Lab rings, a cysteine thiol is appended to select Dha and Dhb residues in a Michael-like addition (Scheme 26). Avi(Me)Cys rings are formed in an analogous fashion from the addition of an enethiolate, which is generated from the oxidative decarboxylation of a C-terminal cysteine residue.221−223 There are two classes of lanthipeptide cyclases that are evolutionarily related but are differentiated by the presence or absence of a Zn metal center. Zn-dependent lanthipeptide

product of the lmn assembly line, actually serves as a prodrug that can be activated by reactive oxygen species.218

4. C−S BOND-FORMING CYCLASES To date, two classes of C−S bond-forming enzymes that catalyze intramolecular cyclizations in an ATP- and oxidationindependent fashion have been identified. These cyclases are responsible for the formation of thiazoline rings on NRPS assembly lines and the conjugate additions that form the eponymous lanthionine rings in lanthipeptide biosynthesis. In both cases the cyclizations are important for biological activity because they confer rigidity to the peptide natural products and, in select cases, lead to installation of pharmacophoric groups (e.g., the installation of DNA-intercalating bisthiazoles). 4.1. Lanthipeptide Cyclases

The lanthipeptides are a diverse RiPP subclass containing various S-heterocycles that include the nonproteinogenic amino acids lanthionine (Lan), methyl lanthionine (MeLan), aminovinyl cysteine (AviCys), aminovinylmethyl cysteine (AviM5539

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

cyclases ligate a Zn2+ ion with a conserved cysteine−cysteine− histidine/cysteine triad.224−227 The cysteine thiol of the precursor peptide is proposed to ligate the Zn, thereby activating the thiol for the subsequent Michael-type addition. A conserved histidine and aspartate located in the active site are thought to perform the requisite acid/base chemistry during thioether bridge formation.225,228 In contrast, the mechanism of Zn-independent lanthipeptide cyclases is largely uncharacterized. Biochemical experiments demonstrated that the enzymes do not require metals for catalysis suggesting that the enzymes perform thioether formation with acid/base catalysis.229,230 After nucleophilic attack, the enolate intermediate can be resolved in one of two ways. In the majority of lanthipeptides, the enolate is protonated to form the Lan/MeLan/Avi(Me)Cys macrocycle (Scheme 26). However, the enolate intermediate can also catalyze a second Michael-like addition, which, after protonation, results in a Lab macrocycle.229−231 While it is not clear how a subset of lanthipeptide cyclases form Lab rings, it is noteworthy that characterized Lab-forming enzymes are exclusively zinc-independent cyclases. Many lanthipeptides contain multiple Dha/Dhb residues and have multiple macrocycles. While it was originally thought that the regioselectivity of the lanthipeptide cyclase governed the final ring topology of the natural product, work with two divergent lanthipeptide cyclases demonstrated that the naturally occurring macrocycles are formed independent of the cyclase used.232 Moreover, it was demonstrated that the configuration of the thioether ring is governed by the precursor peptide sequence surrounding the acceptor Dha or Dhb.233 Together these results have led to a new model for lanthipeptide cyclization where the regio- and stereoselectivity of the transformation is dictated by the sequence of the precursor peptide rather than by the cyclase. Notably, a similar mechanism is thought to be involved in the sactipeptide biosynthesis (see section 7.1.2) and might be a more general trend in RiPP biosynthesis.234−236 For further reading on the biosynthesis and properties of lanthipeptides, we refer you to recent reviews.219,220,237 In addition to the lanthipeptides, AviCys rings are also found in the linaridin class of RiPP natural products (Scheme 27).22,238 While linaridins were originally thought to be a subclass of lanthipeptide natural products because of the presence of Dha and Dhb residues in addition to the AviCys ring, the identification of the biosynthetic gene cluster for the founding member of the family, cypemycin, demonstrated that neither a lanthipeptide dehydratase nor a lanthipeptide cyclase is involved in biosynthesis.238 Moreover, the precursor peptide sequence demonstrated that the acceptor residue for thioenolate attack is a cysteine in cypemycin instead of a serine as found in AviCys-containing lanthipeptides. Although mutagenesis studies have identified enzymes important for the biosynthesis of linaridins, the enzyme responsible for catalyzing thioether bond formation remains unknown.

Scheme 27. Model for the Biosynthesis of Cypemycin, a Linaridin-Type RiPP

respectively,242,243 the cytotoxic metalloantibiotic bleomycin from Streptomyces verticillus,244,245 and the antimitotic epothilone from Sorangium cellulosum (Scheme 28).246 The biosynthesis of S- and O-heterocycles in nonribosomal peptides has been reviewed extensively.239,241,247 In brief, the formation of the five-membered heterocycles requires the enzyme-mediated cyclocondensation of X-Cys dipeptide residues, catalyzed by Cy (cyclization) or HC (heterocyclization) domains. While these domains are typically embedded in the NRPS module, an example of a trans-acting Cy domain is known.248 To date, multiple Cy domains have been biochemically characterized including, but not limited to, those involved in yersiniabactin, pyochelin, and epothilone biosynthesis.249−251 These studies demonstrated that Cy domains are bifunctional enzymes that catalyze both the condensation reaction and heterocycle formation. As uncyclized intermediates have been observed in reactions performed with Cy domain mutants and substrate derivatives, amide formation is thought to occur first through a mechanism identical to canonical condensation domains (Scheme 28).252−255 Following condensation, the thiazoline heterocycle is formed from the nucleophilic attack of the side chain and subsequent dehydration of the tetrahedral intermediate. NRPS Cy domains have a conserved DxxxxDxxS motif that was originally predicted to be the active site of the protein; however, two recent crystal structures of Cy domains demonstrate that these residues play a structural role (Figure 5).256,257 Instead, an Asp-Thr dyad was implicated in catalysis and shown to be highly conserved in diverse Cy domains. In a subset of biosynthetic clusters, the nascent thiazoline ring is either oxidized to a thiazole or reduced to a thiazolidine by an FMN-dependent oxidase (Ox) or NADPH-dependent reductase (R) domain, respectively (Scheme 28).258−260 As with the Cy domains, Ox and R domains can be encoded within the NRPS module or be stand-alone enzymes. Beyond thiazoline formation, an NRPS Cy domain has been implicated in the ring opening of a thiirane intermediate to form the thiotetronate system of thiolactomycin (see section

4.2. NRPS Heterocyclization Domains

Thiazolidines, thiazolines, and thiazoles are important moieties in many nonribosomal peptide natural products. In addition to increasing the proteolytic stability of the peptide, these sulfurcontaining heterocycles are often critical for the biological activity of the compounds (e.g., metal binding and target recognition).239−241 Notable examples of S-heterocyclic nonribosomal peptides are the siderophores yersiniabactin and pyochelin from Yersinia pestis and Pseudomonas spp., 5540

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 28. Structures of Select Natural Products Bearing Thiazolidine, Thiazoline, and Thiazole Rings and General Mechanisms for Heterocyclizations

Figure 5. Crystal structures of a NRPS condensation domain (left; PDB entry code: 1L5A) and Cy domain (right; PDB entry code: 5T7Z). The active site HHxxxD motif of the condensation domains is replaced by a structural DxxxxDxxS motif in Cy domains (colored cyan). The putative AspThr catalytic dyad in the Cy domain is indicated (colored green).

5.2.2)yet another avenue to form a thioester moiety (Scheme 29).261 The Cy domain has the typical DxxxxDxxS structural motif of the heterocyclization domains involved in thiazoline/ oxazoline formation. However, mechanistic details for the

thiolactonization with concomitant shifting of the double bond

need to be clarified. 5541

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

and Archaea.262 The family is named after the Escherichia coli homologue, which has been implicated in the thiomethylation of the S12 protein of the ribosome.263 For over two decades YcaO protein family members have been known to be involved in the biosynthesis of azol(in)e-containing RiPPs; however, they were thought to play a regulatory role.264 Recently, it was demonstrated that YcaO proteins catalyze a range of ATPdependent cyclodehydrations in RiPPs and may promote other reactions (e.g., thioamide formation). 5.1.1. Thiazoles and Thiazolines in RiPPs. As in NRPSderived peptides, thiazole and thiazoline heterocycles are also found in many ribosomally produced natural products, mainly in members of the cyanobactin (e.g., patellamide), thiopeptide (e.g., thiocillin), linear azol(in)e-containing peptides (e.g., plantazolicin), and bottromycin (e.g., bottromycin A2) subclasses of RiPPs (Scheme 30).22 Although the enzymes responsible for azol(in)e heterocycle biosynthesis have not been characterized in all instances, current evidence suggests a universal strategy for C−S bond formation catalyzed by a cyclodehydratase belonging to the YcaO superfamily of proteins.265,266 This enzyme is not chemoselective and is also responsible for cyclizing serine and threonine residues to generate the (methyl)oxazoline moieties also commonly found in these natural product classes. Isotope-labeling experiments demonstrated that the YcaO cyclodehydratase catalyzes the nucleophilic addition of a cysteine side-chain thiol and the subsequent ATP-dependent activation of the amide carbonyl

Scheme 29. Model for the Thiirane Ring-Opening Reaction Mediated by a Cyclase (Cy) Domain in Thiolactomycin Biosynthesis

5. ATP-DEPENDENT C−S BOND-FORMING ENZYMES In many biosynthetic pathways that do not involve substitution reactions with good leaving groups, the formation of covalent C−S bonds may require ATP. In most cases, ATP is essential to activate a carbonyl and transform an intermediary hydroxyl into a leaving group. However, the classes of the enzymes, the types of sulfur nucleophiles involved, and the resulting functionalities are highly diverse. 5.1. YcaO-like enzymes

YcaO proteins, also known as DUF181 proteins and formerly annotated as docking proteins, are widely dispersed in Bacteria

Scheme 30. Examples of Thiazol(in)e-Containing RiPPs and Model of Heterocycle Biosynthesis

5542

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

oxygen (Scheme 30).265−267 Subsequently, the tetrahedral intermediate is resolved by elimination of the carbonyl oxygen as either phosphate or adenosine monophosphate, depending on the specific enzyme, to afford the thiazoline residue.265,267,268 Similar to the biosynthesis of NRPS-derived natural products, the thiazoline heterocycle is further oxidized to yield the thiazole by either a flavin mononucleotide-dependent dehydrogenase or a CYP-dependent decarboxylase (Scheme 30).22,264,269 In all cases where the enzymatic activity has been reconstituted in vitro, a third protein is required for efficient azol(in)e biogenesis. This third protein is responsible for binding the leader peptide region of the precursor peptide, presenting the peptide substrate to the cyclodehydratase and the dehydrogenase, and potentiating the activity of the cyclodehydratase.267,270−272 As is common for RiPP maturation proteins, characterized cyclodehydratases are remarkably promiscuous and are able to accept diverse peptide substrates.264,273 5.1.2. Thioviridamide. In addition to genetic loci coding for azol(in)e-containing RiPPs, a gene encoding a YcaO protein family member was found in the thioviridamide biosynthetic gene cluster.274 Thioviridamide is an unusual natural product in that it contains thioamide linkages (Scheme 31).275 In analogy to thioamide formation in 6-thioguanine biosynthesis (see section 5.2.1), it is thought that the thioamides of thioviridamide are formed through the activation of the carbonyl oxygen and subsequent nucleophilic attack of sulfide. Given the mechanistic similarities to thiazoline formation in RiPPs, the YcaO protein encoded in the thioviridamide gene cluster has been proposed to perform this transformation (Scheme 31).272,274 However, further research will be needed to determine if this biosynthetic proposal is correct.

member. As members of this family are known to catalyze C−S bond formation in thionucleoside biosynthesis, YcfA was proposed to carry out thioamide formation in 6TG biosynthesis. On the basis of the model of the catalytic cycle of the 2thiouridylase MnmA and the 4-thiouridylase ThiI,7,283,284 YcfA is thought to catalyze the adenylation of guanine followed by the nucleophilic addition of sulfide and elimination of carbonyl oxygen as AMP (Scheme 32).281 The highly reactive persulfide Scheme 31. Model of Thioamide Formation Involving a YcaO-like Enzyme in Thioviridamide Biosynthesis

moiety required for YcfA-mediated thiolation is likely provided by a sulfur-relay system that includes a cysteine desulfurylase and sulfur carrier proteins. Mutational studies of the three cysteines in YcfA indicated that only Cys113 was required for 6TG formation when the cluster was expressed in E. coli. It is thought that this cysteine accepts sulfide from the sulfur-relay system as a cysteine persulfide and is responsible for sulfur transfer to the adenylated nucleobase. After C−S bond formation, AMP elimination would afford 6TG. It is noteworthy that, in thiouridine formation, sulfur release from the persulfide requires the formation of an active site disulfide with a second cysteine residue.283,284 This disulfide is subsequently reduced to regenerate the active enzyme. As only a single cysteine residue has been shown to be required for YcfA activity, it is unclear how this AANH-like family member carries out the critical C−S bond-forming step. 5.2.2. Thiolactomycin. Surprisingly, a YcfA homologue catalyzes a key step for introducing sulfur into thiolactomycinbut by a strikingly different mechanism and a highly specialized sulfur-relay system. Thiolactomycin is a bacterial thiotetronate antibiotic that selectively inhibits type II fatty acid synthases.285−287 Radioisotope-labeling experiments using 35SL-cysteine in a Nocardia sp. showed that the sulfur atom of thiolactomycin derives from L-cysteine, likely by a process similar to the IscS-catalyzed sulfur relay in E. coli. A biosynthetic model involving a thiirane intermediate was proposed that could explain thiolactone formation with concomitant double-bond shift.288 The recent identification of tlm biosynthetic gene clusters in Streptomyces thiolactonus and Salinispora pacif ica confirmed this proposal and shed more light on the details of thiolactomycin assembly.261,289 The backbone of the polyketide is assembled by an iterative PKS−NRPS.261,289,290 Mutational analyses

5.2. Adenine Nucleotide Alpha Hydrolase Enzymes

Adenine nucleotide alpha hydrolase enzymes catalyze critical transformations in a diverse set of primary metabolic pathways. Select members of the protein family catalyze sulfur introduction into nucleobases as a posttranscriptional RNA modification.276,277 These thiolation reactions are ATP-dependent and proceed through a complex sulfur-relay system. Interestingly, similar strategies involving thiouridylase-like enzymes and sulfur relay systems have been identified in two natural product biosynthetic pathways. 5.2.1. 6-Thioguanine. The first enzyme responsible for thioamide formation in a natural product was identified in the context of 6-thioguanine biosynthesis. 6-Thioguanine (6TG) is a thiopurine in clinical use as an anticancer agent.278,279 Although 6TG has been known as a synthetic antimetabolite since the 1950s, it was not until the late 1980s that the compound was isolated from a natural source, the plant pathogen Erwinia amylovora.280 E. amylovora is the causative agent of fire blight, a devastating disease affecting trees in the Rosaceae family (mainly apples and pears). Only recently, it was shown that 6TG plays a major role as an E. amylovora virulence factor in fire blight.281 Gene inactivation in E. amylovora and heterologous expression in E. coli revealed that 6TG biosynthesis is encoded by the “yellow compound formation” (ycf) gene cluster and that four genes (ycfABCD) are responsible for 6TG biosynthesis.281,282 Through bioinformatics and three-dimensional structure predictions, YcfA was identified as an adenine nucleotide alpha hydrolase-like (AANH-like) superfamily 5543

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 32. Thioamide Formation in 6-Thioguanine Biosynthesis Parallels Thiouridine Formation

revealed the essential role of a CYP monooxygenase (TlmD1) in TLM biosynthesis.261 According to the current model, TlmD1 would epoxidize the ACP-bound full-length polyketide. Then, an ATP-dependent thiouridylase (TlmJ) would deliver a persulfide that converts the epoxide into the postulated thiirane intermediate. Analogous to the mechanism of the thiouridylase enzymes and proposed mechanism for YcfA, TlmJ is thought to catalyze the adenylation of the hydroxyl group to facilitate thiirane formation (Scheme 33). The persulfide group in TlmJ is likely regenerated by the pathway-specific cysteine desulfurase TlmS.261 The cysteine substrate of TlmS is likely activated by the adenylation (A) domain and tethered to the PCP domain of the PKS−NRPS enzyme, which would allow a fast and continuous supply of activated sulfur. After sulfur transfer, the adjacent thioesterase domain could cleave PCP-bound alanine, regenerating the PCP for a new reaction cycle (Scheme 33). The final step, the cyclization into the thiotetronate system, is likely catalyzed by the cyclization (Cy) domain (see section 4.2). It is noteworthy that this peculiar thiotetronate biosynthetic pathway is encoded in various bacteria across distinct genera, including marine bacteria.261,289

Scheme 33. Model for the Biosynthesis of Thiolactomycin Involving a Thiirane Intermediate

5.3. Adenylating Protein/Sulfur Carrier Protein Systems

In primary metabolism, sulfur is often transferred via thiocarboxy adenylate sulfur carrier protein systems, in which a carboxy residue is activated by ATP and transformed into the corresponding thiocarboxy group by attack of a persulfide.7,291,292 The source of the persulfide varies but is provided by a cysteine desulfurase in many cases, including the biosynthesis of the important cofactors molybdopterin and thiamine. Remarkably, this strategy of C−S bond formation has been adopted by at least two natural product biosynthetic pathways. 5.3.1. Thioquinolobactin. Thioquinolobactin (TQB) is an unusual Pseudomonas f luorescens metabolite featuring a rare thiocarboxylate residue,293 which is essential for the antifungal activity of the molecule.294 However, the thiocarboxylate moiety is prone to rapid hydrolysis, yielding quinolobactin, an iron chelator.293 The heterocyclic core results from the oxidative cleavage of the tryptophan indole and cyclization of the intermediary hydroxykynurenine.295 Although the mechanism of C−S bond formation has not fully been elucidated, incorporation of sulfur has been found to be remarkably similar to mechanisms found in primary metabolism.296 In such pathways, the C-terminus of a small sulfur carrier protein is adenylated by a member of the E1 superfamily. This adenylated intermediate is subsequently transformed into a thiocarboxylate by transfer of sulfur from a cysteine desulfurase bound persulfide.7,291,292 In agreement with this model, the TQB biosynthetic gene cluster in P.

f luorescens codes for a small sulfur carrier protein (QbsE) and a didomain protein composed of an E1 domain and a sulfurtransferring rhodanese domain (QbsC).296 Reconstitution of thiocarboxylate formation in vitro demonstrated that an additional protein, a metallopeptidase (QbsD), is required for activity. This enzyme was shown to cleave off the final two amino acids on the C-terminus of QbsE (-GGCF) to generate the signature digylcine motif that is conserved in the sulfur 5544

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

ThiS.6,301−303 As the BE-7585A biosynthetic gene cluster lacks a ThiS homologue, the sulfur carrier proteins from primary metabolic pathways in A. orientalis were screened as sulfur sources for the BexX reaction.304 While ThiS from A. orientalis was not a competent sulfur donor, BexX reactions performed with the thiocarboxylate-bearing sulfur carrier proteins from cysteine (AoCysO) and molybdopterin (AoMoaD2) pathways afforded 2-thioglucose-6-phosphate. Thiocarboxylate formation on AoCysO and AoMoaD2 required the addition of MoeZ, a bifunctional E1 and rhodanese homologue. In analogy to the QbsC reaction in THN biosynthesis, thiosulfate was the sulfur source used in vitro. A mechanism for 2-thiosugar formation has been proposed (Scheme 35). First, a covalent adduct between BexX and G6P is formed that tautomerizes to the 2-keto-bearing sugar. This adduct was directly detected by mass spectrometry and was also present in a cocrystal structure of BexX and G6P (Figure 6).300,304 The C-2 ketone is attacked by the sulfur from the sulfur carrier protein. A cocrystal structure of BexX and AoCysO demonstrates that the C-terminus of the sulfur carrier protein is in close proximity to the G6P binding site (Figure 6). The resultant tetrahedral intermediate is subsequently dehydrated and tautomerized to form the imine. Finally, hydrolysis of the imine linkage releases the product from BexX.299

carrier protein family. Following proteolysis by QbsD, QbsE was transformed into the thiocarboxylate by QbsC using thiosulfate as a sulfur source. In analogy to characterized thiocarboxylate-forming enzymes, a mechanism for the QbsC-catalyzed sulfur transfer was proposed (Scheme 34). The steps leading to sulfur transfer Scheme 34. Model for the Biosynthesis of Thioquinolobactin; Red, Reduction

6. OXYGENASES Oxygenases are often indirectly involved in the formation of C−S bonds, as they set the stage for the attack of a sulfur nucleophile. Thus, as has been outlined above, S-transferring enzymes often come in pairs with oxygenases; for examples, see the biosynthetic pathways of gliotoxin (CYP and GST, section 3.3.2), glucosinolates (CYP and GST, section 3.3.3), and thiolactomycin (CYP and thiouridylase, section 5.2.2). In camalexin biosynthesis, a CYP enzyme generates a reactive intermediate that forms an S-conjugate, yet the potentially involved S-transferase is unknown. In this section a variety of oxygenases are presented that mediate the formation of C−S bonds. In some cases one may question whether these biocatalysts catalyze the covalent linkage of carbon and sulfur atoms or just produce highly reactive compounds that spontaneously react with sulfur-bearing molecules.

from the QbsE-thiocarboxylate to quinolobactin are currently unknown. However, it is thought that quinolobactin is activated by a bifunctional AMP-ligase/methyltransferase (QbsL) and subsequent sulfur transfer is catalyzed by an Acyl-CoA transferase homologue (QbsK). 295 In support of this biosynthetic proposal, homologues of qbsL and qbsK are found in the biosynthetic clusters of two additional thiocarboxylate-containing natural products: the structurally similar compound 2,6-pyridinedicarbothioic acid, produced by Pseudomonas stutzeri, and yatakemycin, produced by Streptomyces sp. TP-A0356.297,298 Moreover, it is noteworthy that the biosynthetic gene cluster of 2,6-pyridinedicarbothioic acid encodes homologues of QbsC, QbsD, and QbsE.298 While the functions of these proteins have not been verified, their presence suggests that this strategy for sulfur mobilization is used in the biosynthesis of additional natural products. 5.3.2. BE-7585A. A related, yet even more complex pathway has been unraveled for the biosynthesis of BE7585A, an antibiotic containing a rare 2-thiosugar. Analysis of BE-7585A biosynthesis in Amycolatopsis orientalis subsp. vinearia implicated BexX, homologue of the thiamine thiazole synthase ThiG, in 2-thiosugar biosynthesis.299 Initial BexX in vitro assays demonstrated that the enzyme could form a covalent adduct with the presumed precursor of the thiosugar, glucose-6-phosphate (G6P); however, the sulfur donor remained enigmatic.300 The mechanism of ThiG has been heavily investigated, and it is known that the sulfur donor for the reaction is the thiocarboxylate of the sulfur carrier protein

6.1. Cytochrome P450 Monooxygenases

Cytochrome P450 monooxygenases are widespread enzymes that are well-known for their ability to transform a broad range of substrates.305 Most commonly, they catalyze hydroxylations and epoxidations, but also more unusual reactions such as sixelectron oxidations of methyl groups into carboxyl groups, aryl couplings, and the formation of nitrile oxide (see section 3.3.3) have been identified. Several CYPs have been implicated in C− S bond-forming reactions leading to the thiazole of camalexin, the S-heterocycles of cyclobrassinin, spirobrassinin, thienodolin, and griseoviridin, and the sulfoxide-substituted ansa-compound ustiloxin. 6.1.1. Camalexin. In addition to glucosinolates, various plants from the crucifer family, e.g., Arabidopsis thaliana, produce camalexin, a thiazole-containing phytoalexin with antifungal activity.306,307 Notably, the route to the thiazole heterocycle markedly differs from the well-known cyclocondensation reactions observed for ribosomal and nonribosomal peptides (see sections 4.2 and 5.1.1). In vitro assays revealed that camalexin biosynthesis in A. thaliana requires the combination of three CYP monoox5545

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 35. Mechanism for BexX-Catalyzed 2-Thioglucose-6-phosphate Formationa

a

Thiocarboxylate formation on the sulfur carrier protein is shown in the box. Two possibilities for the source of sulfur are indicated. CD, cysteine desulfurase; SCP, sulfur carrier protein

ygenases.308 First, in analogy to glycosinolate pathways, CYP79B2/3 catalyzes the formation of indole-3-acetaldoxime from tryptophan (Scheme 36).309 The oxime is then processed into indole-3-acetonitrile by CYP71A13 or CYP71A12, apparently redundant enzymes, followed by hydroxylation to give a cyanohydrin, in analogy to the biosynthesis of cyanogenic glycosides. In fact, indole-3-carboxyaldehyde has been observed as a side product of cyanhydrin decomposition (Scheme 36). However, in the presence of thiols such as glutathione and cysteine, the enzyme assay gave the corresponding S-adducts. One may conceive that the hydroxylated indole-3-acetonitrile forms an iminium species that could be attacked by sulfur nucleophiles. Although this key step may take place without enzyme catalysis, the slow rate for the formation of the cysteine−indole-3-acetonitrile adduct suggests that a designated transferase (e.g., a GST) may be requiredas in the gliotoxin and glucosinolate biosynthetic pathways. The final oxidative heterocyclization to the thiazole is promoted by CYP71B15. 6.1.2. Cyclobrassinin and Spirobrassinin. The S-heterocycles cyclobrassinin and spirobrassinin from Chinese cabbage (Brassica rapa) are remarkable examples of glucosinolatederived phytoalexins. The dithiocarbamoyl brassinin, which derives from the corresponding isothiocyanate (section 3.1.6), has been identified as the precursor to both cyclobrassinin and

Figure 6. Crystal structures of Amycolatopsis orientalis BexX/CysO complex and BexX covalent intermediate. (A) The BexX/CysO complex (PDB entry code: 4N6E) is displayed. BexX and CysO are colored gray and orange, respectively. A zoomed-in view of the active site is shown to highlight the close proximity of the CysO C-terminus and the lysine responsible for glucose-6-phosphate tethering. (B) A zoomed-in view of the active site of BexX cocrystallized with glucose6-phosphate (PDB entry code: 4N6F) is displayed.

spirobrassinin (Scheme 37).82 RNA sequencing facilitated the identification of upregulated phytoalexin biosynthetic genes in 5546

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 36. Model for Camalexin Biosynthesisa

a

The C−S bond is also formed nonenzymatically in vitro.

Scheme 37. CYP-Mediated S-Heterocyclizations of Brassinin into Spirobrassinin and Cyclobrassinina

a

Wasalexin is produced from brassinin by an unknown pathway.

Scheme 38. CYP-Mediated S-Heterocyclization in the Thienodolin Pathway

cabbage leaves elicited with Pseudomonas syringae pv maculicola. Various CYP monooxygenase genes were selected and individually expressed in a Saccharomyces cerevisiae strain equipped with the CYP reductase ATR1. Using brassinin as the substrate, the microsomal fraction of CYP71CR1- and CYP71CR2-expressing strains catalyzed the formation of cyclobrassinin and spirobrassininol, respectively. On the basis of labeling experiments, it was concluded that brassinin is epoxidized at the C2,C3 position. Depending on the site of

nucleophilic oxirane ring opening, either a six-membered or a five-membered S-heterocycle is formed, followed by dehydration (likely spontaneous) or CYP-mediated oxidation of the resulting alcohols (Scheme 37).310 Brassinin was also identified as an intermediate for the biosynthesis of wasalexin, a phytoalexin with antifungal activity produced by salt cress (Thellungiella salsuginea) and first isolated from wasabi (Eutrema japonica).311 Although the genes for wasalexin production are unknown, brassinin was 5547

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

established as an intermediate by labeling studies and retrobiosynthetic analysis.312 6.1.3. Thienodolin. In a similar way, a CYP may catalyze the formation of thienodolin (THN), a bacterial S-heterocycle that inhibits nitric oxide synthases.313 The gene cluster responsible for the production of THN was recently discovered in the genome of Streptomyces albogriseolus.314 On the basis of gene inactivation experiments and in vitro analyses, a biosynthetic model was proposed that involved the sequential formation of two C−S bonds (Scheme 38). The pathway is initiated by the halogenation of tryptophan.314 Then, aminotransferase ThnJ converts 6-chlorotryptophan into 6-chloroindole-3-pyruvic acid,315 into which sulfur may be introduced by an unknown mechanism, possibly catalyzed by the concerted action of a dehydrogenase, an isopeptidase, and a protein of unknown function encoded in the biosynthetic gene cluster (ThnD−F). The second C−S bond may be formed by a sequence that is analogous to cyclobrassinin biosynthesis.310 The C2−C3 of the indole could be oxidized by a CYP monooxygenase to afford an epoxide, which is opened by nucleophilic attack of the thiol. Subsequently, dehydration and oxidation of the resultant S-containing heterocycle would provide 6-chlorothieno[2,3]indole-2-carboxylic acid (Scheme 38).315 Finally, an amidotransferase (ThnA) introduces the carboxamide function of THN.314 6.1.4. Griseoviridin. Another type of CYP-mediated Sheterocyclization produces the unusual thioene ring of the macrolide griseoviridin, a ribosome-targeting antibiotic from Streptomyces griseoviridis.316 Through analysis of the griseoviridin biosynthetic gene cluster, targeted knockouts, and mutant complementation, it was shown that a CYP monooxygenase (SgvP) is required for the formation of the C−S bond.317,318 The structure of the PKS−NRPS-derived product suggests that the S-heterocycle is in part derived from cysteine. Surprisingly, a mutant lacking SgvP produces a shunt product with a dehydroalanine residue (Scheme 39). Two possible mechanisms for the CYP-mediated cyclization were proposed: a radical mechanism involving a sulfur- and an α-carbon-centered radical and a nucleophilic mechanism invoking an epoxide intermediate (Scheme 39). This latter route resembles the proposed pathways for the brassinin derivatives and thienodolin. 6.1.5. Ustiloxin. A hallmark of the ustiloxins is an unusual aryl sulfoxide linkage formed by the combination of a CYP and a flavoenzyme (Scheme 40). This structurally remarkable group of antimitotic phytotoxins is produced by the causative agent of rice false smut, Ustilaginoidea virens.319−321 A recent report demonstrated that ustiloxins are also produced by Aspergillus f lavus, which facilitated elucidating the molecular basis of their biosynthesis.322 Although the peptide structures of these natural ansa-compounds initially suggested a nonribosomal origin, the discovery of the ustiloxin biosynthetic gene cluster in A. f lavus demonstrated that ustiloxins are members of the RiPP class of natural products.22,323,324 Deletion mutagenesis experiments performed on the ustiloxin biosynthetic gene cluster in A. f lavus demonstrated that the macrocycle is derived from a stretch of four amino acids within a gene-encoded precursor peptide.324 The precursor peptide encodes multiple copies of a small structural peptide that are separated by protease recognition sites.323−325 Proteolytic digestion of the precursor peptide affords the tetrapeptide, which is then posttranslationally modified by eight proteins to give the final natural product ustiloxin B (Scheme 40).324,326

Scheme 39. CYP-Mediated Heterocyclization in Griseoviridin Biosynthesisa

a

Two possible mechanisms for SgvP-catalyzed C−S bond formation are displayed.

The deletion of the ustiloxin maturation enzymes in A. f lavus lead to the production of biosynthetic intermediates that facilitated the characterization of the biosynthetic pathway.326 Deletion of ustC, which encodes a cytochrome P450 family protein, resulted in the production of an intermediate lacking the sulfinyl amino acid on the phenyl ring. Furthermore, the deletion of ustF1, which encodes a flavin-dependent monooxygenase, resulted in a strain producing an intermediate bearing a cysteine moiety on the phenyl ring. The activity of UstF1 was reconstituted in vitro, and the enzyme was shown to oxidize the cysteinyl thioether to the sulfonyl derivative. While the activity of UstC has yet to be reconstituted in vitro, these data suggest that this cytochrome P450 homologue is responsible for C−S bond formation in ustiloxin biosynthesis. Additional support for this functional assignment is provided by the absence of a ustC homologue in the biosynthetic gene clusters of similar fungal RiPPs that lack the arylsulfoxide moiety (e.g., phomopsin, Scheme 40).327 In analogy to the aforementioned examples, UstC likely catalyzes C−S bond formation through an epoxide intermediate (Scheme 40). A similar thioether linkage is formed during the biosynthesis of the potent fungal toxins α-amanitin (Scheme 40) and phalloidin.328 These macrocyclic RiPPs are produced by members of the phylum Basidomycota and harbor a tryptathionine moiety originating from core peptide-encoded Cys and Trp. While the biosynthetic gene clusters for each of these natural products are known and the enzymes responsible for core peptide macrocyclization have been characterized, it is still unclear how the tryptathionine cross-bridge is formed.17,329−331 A ustC homologue is not found in the local 5548

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 40. Arylsulfoxide Formation in Ustiloxin Biosynthesisa

a

A possible reaction mechanism for the UstC-mediated C−S bond-forming reaction is displayed in the box.

genomic region of the precursor peptide and peptidase in either instance. Further work will be required to identify the mechanism of C−S bond formation in ustiloxin biosynthesis and to identify the enzyme responsible for the analogous modification in α-amanitin and phalloidin maturation.

Scheme 41. General Scheme for the Biosynthesis of Penicillins and Cephalosporins from an NRPS-Derived Tripeptide; ACVS, Aminoadipyl-Cysteinyl-Valine Synthetase

6.2. Nonheme Iron-Dependent Enzymes

Like CYPs, nonheme iron-dependent oxygenases are best known to oxygenate various substrates though the introduction of molecular oxygen. In addition, these versatile enzymes catalyze a broad range of reactions including C−C cleavage and halogenation.332,333 By catalyzing the formation of C−S bonds, nonheme iron-dependent oxygenases also play key roles in the biosynthetic pathways of penicillins, cephalosporin, ergothioneine, and ovothiol. In particular, the detailed analysis of the biosynthesis of penicillins and ergothioneine greatly enhanced our understanding of radical sulfurization reactions. 6.2.1. Penicillins and Cephalosporins. Penicillins and cephalosporins are important β-lactam antibiotics in clinical use. Whereas penicillins are characterized by thiazolidine rings, a dihydrothiazine is typical for the cephalosporins. The Sheterocycles of all naturally occurring penicillins and cephalosporins derive from a single tripeptide, D-(L-α-aminoadipyl)-Lcysteinyl-D-valine, which is assembled by a nonribosomal peptide synthetase (ACVS) (Scheme 41).334−336 In diverse penicillin-producing fungi and bacteria, isopenicillin N synthase (IPNS), a nonheme iron-dependent enzyme, mediates the formation of the canonical penicillin core structure, the fourmembered β-lactam ring fused to the five-membered thiazolidine.335−338 The reaction mechanism of IPNS has been explored in detail through kinetic, spectroscopic, structural, and computational studies.335,339 The combined results from these investigations

were used to formulate a reaction mechanism (Scheme 42). First, the deprotonated thiol group of the tripeptide substrate (ACV) binds to the iron center, which facilitates the binding of molecular oxygen and subsequent formation of a ferricsuperoxo species. The Fe(III)-superoxo complex extracts a hydrogen atom from C-3 of the cysteine, yielding a ferroushydroperoxy intermediate and a thioaldehyde. The valine amide nitrogen is deprotonated by the hydroperoxy moiety, which sets the stage for the formation of the β-lactam ring via nucleophilic attack at the thiocarbonyl. The ferryl-oxo species formed during this process abstracts a hydrogen atom from the β-carbon of valine, and the resulting isopropyl radical reacts with the thiolate sulfur to afford the thiazolidine ring. Recently, additional support for this mechanism was obtained using ACV substrate analogues. The presence of a radical at the β-carbon of valine was probed by use of substrate derivatives in which valine was replaced by amino acids 5549

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 42. Model for Radical-Based C−S Bond Formation in Penicillins Catalyzed by Isopenicillin N Synthase

Figure 7. Crystal structure of isopenicillin N synthase from Aspergillus nidulans in complex with δ-(L-α-aminoadipoyl)-L-cysteinyl-D-valine, iron, and nitric oxide (PDB entry code: 1BLZ). The zoomed-in view shows the active site. Fe-ligating residues are colored cyan.

containing cyclopropylmethyl groups.340 Consistent with the proposed reaction mechanism, radical clock intermediates resulting from the rearrangement of the cyclopropylcarbinyl radical were observed in reactions performed with these substrates. Furthermore, the first spectroscopic evidence for the ferric-superoxo and ferryl-oxo species was obtained using deuterated substrates that displayed significant primary kinetic isotope effects. As previous work had demonstrated that both C−H bond-cleavage events are rate-limiting,339 these substrates were used to accumulate the hydrogen-abstracting species, which was then characterized by Mössbauer spectroscopy.341 The crystal structure of IPNS shows that the iron center is coordinated by two histidine residues and an aspartic acid residue.342 Crystals obtained anaerobically demonstrate that the remaining ligation sites on the iron center are occupied by ACV and two water molecules.343 Although a crystal structure of IPNS bound to AVS and oxygen has not been obtained, a surrogate structure was solved with the oxygen analogue nitric oxide (Figure 7). The nitric oxide binds opposite of the aspartate ligand and immediately adjacent to the substrate. As oxygen would likely bind in a similar manner, this configuration would place the superoxo species in the appropriate position to extract the C-3 hydrogen of cysteine. The cephalosporin β-lactam dihydrothiazine scaffold results from an oxidative expansion of the penicillin ring system, which affords the formation of a new C−S bond (Scheme 43). This committed step in the biosynthesis of all cephalosporin antibiotics is catalyzed by deacetoxycephalosporin C synthase (DAOCS), a 2-oxoglutarate-dependent nonheme iron-depend-

ent enzyme.335 Following ring expansion, the remaining methyl group of deacetoxycephalosporin C (DAOC) is hydroxylated. While this hydroxylation is performed by a second 2oxoglutarate-dependent nonheme iron oxygenase in bacteria (DACS), fungal DAOCS homologues are bifunctional (denoted DAOC/DACS) and catalyze both ring expansion and hydroxylation in a single active site. Interestingly, DAOC/ DACS hydroxylation activity was abolished by the mutation of an active site residue (M306I), converting the bifunctional enzyme into an expandase.344 Early insights into the mechanism of DAOCS were obtained using isotopically labeled substrates. These studies determined the fate of each of the methyl groups of penicillin N and provided evidence of a long-lived radical species during the transformation.345−348 Together with structural and biochemical data,349−354 a reaction mechanism was proposed that follows the consensus mechanism for 2-oxoglutarate-dependent oxygenases (Scheme 43). First 2-oxoglutarate and molecular oxygen react to form a ferryl-oxo species, which extracts a hydrogen from penicillin N. The initial methyl radical is thought to rearrange to an episulfide radical intermediate, which subsequently opens to afford a tertiary carbon radical at C3. The C3 radical is resolved by transfer of a hydrogen radical to the metal center, yielding deacetoxycephalosporin C. Although the initial structural studies suggested that DAOCS operated through a ping-pong mechanism where succinate release would precede penicillin N binding (Figure 8),350,351 recent presteady-state kinetics and NMR-/MS-based binding studies indicate that a ternary (DAOCS·Fe(II)·2-oxoglutarate· 5550

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 43. Mechanism of DAOCS-Catalyzed Ring Expansion of Penicillin N into Cephalosporin Precursors; 2-OG, 2Oxoglutarate

penicillin N) complex is formed.354 For additional information on the biosynthesis of β-lactam antibiotics, readers are referred to extensive reviews on the topic.335,336 6.2.2. Ergothioneine and Ovothiol. Ergothioneine (EGT) is a histidine derivative with a thiol substituent at the C-2 of the imidazole ring. EGT has been found in a broad range of bacteria, fungi, plants, and animals, where it likely plays a role in cellular redox homeostasis.4,355 Notably, only fungi and bacteria biosynthesize EGT; therefore, other organisms must acquire it through symbioses with producing organisms or through their dietary sources. Humans possess a highly specific transporter (ETT) that promotes the uptake of EGT, which cannot diffuse across the plasma membrane due to its zwitterionic nature.356 The biosynthetic pathway for EGT was identified in Mycobacterium smegmatis by genome mining for a SAMdependent methyltransferase specific to EGT producers and in a local genomic region containing a PLP-dependent decarboxylase.357 Reconstitution of the methyltransferase (EgtD) demonstrated that the enzyme catalyzes the permethylation of the histidine amino group to afford hercynine. Hercynine serves as the substrate for the nonheme iron-dependent enzyme EgtB, which is responsible for C−S bond formation (Scheme 44).357,358 The cosubstrate for the EgtB reaction, γ-glutamyl cysteine, is provided by EgtA, which catalyzes the condensation of glutamate and cysteine.357 The crystal structure of an EgtB homologue from Mycobacterium thermoresistible in complex with manganese, γglutamyl cysteine, and the substrate analogue N-α-dimethyl histidine has been solved.358 The metal center is coordinated by three histidine residues, the imidazole ring of hercynine, the

Figure 8. Crystal structure of cephalosporin synthase from Streptomyces clavuligerus. (Top) Catalytic center of DAOCS in complex with Fe(II) and 2-oxoglutarate (2OG; PDB entry code: 1UOB). (Bottom) Catalytic center of DAOCS in complex with Fe(II) and the substrate penicillin G (penG; PDB entry code: 1UOF). In these structures the binding sites of penicillin G and 2OG overlap. Feligating residues are colored cyan.

5551

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 44. Biosynthesis of Ergothioneine in Bacteria and Fungi; Box: Product of the EGT C−S Lyasea

a In the absence of a reductant, the hypothesized 2-sulfenic acid intermediate disproportionates to form EGT and EGT-2-sulfinic acid. In the presence of a reductant, EGT is formed directly.

thiol of γ-glutamyl cysteine, and a water molecule in an octahedral geometry (Figure 9). Combined with assays

hydroperoxo species. The thiyl radical could then react with hercynine to form the iminyl radical, which is subsequently rearomatized through deprotonation and a ligand-to-metal electron transfer. Finally, S-sulfoxidation completes the reaction cycle. The thiol residue of ergothioneine is liberated by means of at least two pathway-specific enzymes. First, aminohydrolase EgtC cleaves glutamate from the γ-glutamyl cysteine sulfoxide conjugate (Scheme 44).357,359 Second, a PLP-dependent lyase (EgtE) cleaves the C−S bond of the cysteine-derived moiety, producing pyruvate, ammonia, and EGT (Scheme 44).360 This reaction is reminiscent of the GliI-mediated C−S-cleavage reaction in ETP biosynthesis (see section 3.3.2). When EgtE reactions are performed in the absence of a reductant (e.g., dithiothreitol), the production of EGT-2-sulfinic acid is also observed (Scheme 44). EGT-2-sulfinic acid is proposed to originate from the production of the corresponding sulfenic acid. When a reductant is present, this sulfenic acid intermediate can be reduced nonenzymatically to afford ergothioneine. Notably, the lyase also accepts the corresponding thioether as substrate, thus eliminating the need for a reductant in the transformation. In comparison to the four-step mycobacterial pathway, EGT biosynthesis in fungi is composed of just three steps (Scheme 44).361−365 In characterized fungal pathways, the homologues of EgtD and EgtB are fused in a single protein (Egt1). In vivo and in vitro characterization of Egt1 demonstrated that this

Figure 9. Crystal structure of EgtB from Mycobacterium thermoresistible in complex with γ-glutamyl cysteine (γGC), N-α-dimethyl histidine (DAH), and manganese (PDB entry code: 4X8D). The Tyr residue implicated in the catalytic cycle is indicated. Mn-binding residues are colored cyan.

performed on EgtB point mutants, a mechanism for C−S bond formation was proposed.358 In the presence of O2, the iron center of the substrate-bound complex forms an iron(III)superoxo species (Scheme 45). A single electron transfer from the thiol of γ-glutamyl cysteine and protonation of the resultant peroxide anion by an active site tyrosine affords the iron(III)5552

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

48).378,379 Heterologous overexpression of the xia biosynthetic gene cluster in Streptomyces albus lead to the identification of the sulfadixiamycins.380 The structures of the sulfadixiamycins suggest that two equivalents of the pentacyclic xiamycin could capture sulfur dioxide, which may be formed in the course of sulfur primary metabolism, in a fashion similar to synthetic polymerization processes. Mutational analyses of the xia biosynthetic gene cluster revealed a candidate gene (xiaH) for sulfadixiamycin biosynthesis.380 A mutant strain lacking xiaH no longer produced the sulfa-dimers of xiamycin but was still able to produce the parent compound. Although the reaction could not be reconstituted in vitro, whole-cell biotransformation reactions with heterologously expressed xiaH were successful. Because XiaH also promotes the formation of N−C- and N−N-fused xiamycin dimers (dixiamycins) with the same substitution patterns,381 it was concluded that the formation of the C−S and N−S bonds proceeds through a radical mechanism. In fact, the fusion sites for dixiamycin and sulfadixiamycin biosynthesis (N-1, C-6, and C-21) can be rationalized by resonance stabilization of a xiamycin radical intermediate (Scheme 48). XiaH belongs to the large family of flavoenzymes, which mainly catalyze two-electron redox reactions. Although only rarely observed, some flavoenzymes are also capable of mediating single-electron transfers with formation of a flavin semiquinone radical. Perhaps the best-known examples are FAD-dependent monoamine oxidases, which oxidize amines to imines by one-electron oxidations that involve tyrosyl radicals.382,383 By analogy, sulfadixiamycin formation likely involves the formal abstraction of a hydrogen radical or a oneelectron oxidation to the radical cation, followed by deprotonation. The resulting resonance-stabilized radical intermediates may either pair to yield dixiamycins or react with sulfur dioxide. The sulfonyl radicals could then pair with another xiamycin radical.380

bifunctional enzyme catalyzes the methylation of histidine and subsequent sulfoxidation.361−363,365 Kinetic characterization showed that Egt1 from Neurospora crassa preferentially uses cysteine rather than the γ-glutamyl cysteine dipeptide as the sulfur donor for the sulfoxidation reaction.362 This result is consistent with a lack of an EgtC homologue in the genome of N. crassa. As with the mycobacterial EGT pathway, a C−S lyase (Egt2) removes the cysteine side chain to afford the mature compound.362,363 A similar moiety is found in ovothiol, a thiohistidine derivate initially isolated from sea urchin (Paracentrotus lividus) eggs366,367 and subsequently from various pathogenic protists such as Leishmania donovani and Trypanosoma cruzi.368,369 While the function of ovothiol is not yet understood, gene expression and metabolomic experiments in P. lividus link the molecule to the oxidative stress response during embryo development.370 Assays performed with cell-free extracts from the parasitic protist Crithidia fasciculata and radiolabeled ovothiol precursors implicated an oxygen-dependent enzyme in the conversion of cysteine and histidine to 5-histidylcysteine sulfoxide.371 In vitro reconstitution studies demonstrated that an EgtB/Egt1 homologue (OvoA) catalyzes the C-5 sulfoxidation of histidine using cysteine as the sulfur donor (Scheme 46).372 An evaluation of OvoA activity with different substrates suggested that the reaction mechanism parallels EgtB/Egt1-catalyzed C− S bond formation and indicated that the regioselectivity of OvoA (C-5 vs C-2 modification) is dictated by the structure of the sulfur acceptor.373,374 The product from the OvoA reaction is further processed to 5-thiohistidine and N-methylated at the imidazole ring to produce ovothiol, although the responsible enzymes have not been identified. 6.3. Flavoenzymes

In natural product biosynthesis, flavin-dependent oxygenases are known to promote a broad array of chemical redox transformations. The catalytic repertoire of the versatile flavin coenzymes includes both two- and one-electron transfers. Typically, dihydroflavins react with molecular oxygen to form flavin-4a-OOH adducts for various oxygenations. However, radical reactions promoted by flavoenzymes are also known.375 For the incorporation of sulfur, both avenues have been proposed. 6.3.1. Coelimycin. One or more flavoenzymes appear to be involved in the formation of the unusual 1,5-oxathiocane ring of coelimycin, an unusual yellow pigment produced by a Streptomyces coelicolor mutant.376 Bioinformatic analysis of the cpk gene locus and stable isotope-labeling experiments suggested that the coelimycin carbon backbone is assembled by a modular polyketide synthase. A plausible mechanism for the incorporation of N-actetyl cysteine would involve a bisepoxidated intermediate that could be attacked by the thiol group (Scheme 47). The epoxidation reactions could be catalyzed by the predicted gene products of cpkD, cpkH, and scF, which are all similar to known flavin-dependent epoxidases/dehydrogenases.376 6.3.2. Sulfadixiamycins. A radical-based C−S bond formation mediated by a flavoenzyme has been proposed for the biosynthesis of highly unusual diarylsulfone and sulfonamide antibiotics produced by a mangrove tree endophyte, Streptomyces sp. HKI0595. This endophytic streptomycete was previously shown to produce the indolosesquiterpene xiamycin (xia)377 by an unusual terpenoid cyclization sequence (Scheme

6.4. Tyrosinases

Tyrosinases are widely distributed type-3 copper-dependent enzymes that are mainly known to be involved in the biosynthesis of pigments.384−386 These dinuclear copper enzymes catalyze the four-electron oxidation of phenols to form ortho-quinones. In some cases these ortho-quinones can serve as electrophiles for the nucleophilic attack of thiols. It is notable that the active site of some tyrosinases contains a thioether linkage between a Cys and a copper-ligating His residue. This autoxidation reaction is proposed to conformationally restrain the His to improve copper binding.387 6.4.1. Pheomelanins. In humans and animals, tyrosinase catalyzes the hydroxylation of tyrosine into L-3,4-dihydroxyphenylalanine (L-DOPA) and its further oxidation into LDOPA quinone, a key intermediate in the biosynthesis of black/brown and red pigments named eumelanins and pheomelanins, respectively.384 The color of the pigments is strongly influenced by the incorporation of sulfur. In the presence of cysteine, the highly reactive DOPA quinone immediately reacts with the thiol to form cysteinyl-DOPA. After tyrosinase-mediated oxidation of the cysteinyl adduct into the corresponding quinone, intramolecular imine formation yields a benzothiazine heterocycle (Scheme 49). This building block and its regioisomers may polymerize to give the redcolored pheomelanins in human hair. Dimerization of the benzothiazine yields trichochromes, which are found not only in red human hair388 but also in skin, where they are in part 5553

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 45. Mechanism of Radical Sulfurization Catalyzed by EgtB

Scheme 46. Model for Ovothiol Biosynthesis

Scheme 47. Model for the Biosynthesis of Coelimycina

a

Following C−S bond formation, the opening of the second epoxide ring could proceed through a Payne rearrangement (a) or occur directly (b).

6.4.2. Grixazones. A tyrosinase homologue lacking ring hydroxylation activity enables the formation of a C−S bond in the biosynthesis of the Streptomyces griseus pigments grixazones A and B.391 These phenoxazinone derivatives are assembled from two 3-amino-4-hydroxybenzaldehyde (3,4-AHBAL) building blocks and N-acetylcysteine.392 Characterization of the grixazone (gri) biosynthetic pathway demonstrated that 3,4AHBAL is assembled from L-aspartate-4-semialdehyde and dihydroxyacetone phosphate by four proteins (Scheme 50).198,392 Two additional genes in the biosynthetic gene cluster, griF and griE, were predicted to be responsible for the

responsible for cellular photodamage as they function as endogenous UVA-photosensitizers.389 A similar heterocyclic system was found in cytotoxic red pigments (pheofungins) of an engineered mutant of Aspergillus nidulans impaired in posttranslational protein modification.390 Transcription analyses and gene knockout experiments indicated that pheofungins originate from phenolic compounds derived from orsellinic acid. Because tyrosinases are also widespread in the fungal domain, pheofungin biosynthesis likely involves the condensation of cysteine with ortho-quinones in analogy to the pheomelanin pathway.390 5554

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 48. Biosynthesis of Sulfadixiamycins A−C, Diarylsulfones, and Sulfonamides by Flavoenzyme-Mediated SO2 Capture

maturation of the 3,4-AHBAL-derived chromophore.393 The deduced gene products were homologous to tyrosinases (GriF) and tyrosinase-associated copper chaperones (GriE). Consistent with this prediction, heterologously produced GriE activated GriF by transferring copper ions, and a mutant lacking both genes accumulated 3,4-AHBAL. GriF was found to oxidize 3,4-AHBAL into the corresponding ortho-quinone imine, which condenses with a second equivalent of orthoquinone imine to form the phenoxazinone ring system. In the presence of N-acetylcysteine, grixazone A was formed by GriF in an in vitro assay. It was thus assumed that the Nacetylcysteine thiol undergoes a nonenzymatic conjugate addition to the ortho-quinone imine. The resulting thio conjugate of 3,4-AHBAL could be reoxidized by GriF and nonenzymatically coupled with another molecule of the orthoquinone imine to yield the mature grixazone (Scheme 50).393 6.4.3. Grape Reaction Product. Another prominent example of a tyrosinase involved in the formation of a C−S linkage is found in the production of the caftaric acid− glutathione conjugate, grape reaction product (GRP). During wine production, polyphenoloxidase oxidizes caftaric acid into

the corresponding ortho-quinone.394 In the presence of glutathione, the electrophilic ortho-quinone is converted to GRP (Scheme 51). Although GRP is the most abundant product, C-5 and C-6 glutathione conjugates and conjugates with other thiols have been observed.395 Consistent with this lack of specificity, in vitro experiments demonstrated that thiol addition is a nonenzymatic process.394,395 When the concentration of caftaric acid is greater than the concentration of the thiols in the grape juice, the ortho-quinone can readily react with various phenols (e.g., caftaric acid).396,397 The resultant dimers can be reoxidized by polyphenoloxidase and coupled to additional phenols, leading to the formation of brown pigments and loss of aroma compounds. This juice browning negatively impacts the quality of the wine and is commonly controlled by the addition of the polyphenoloxidase inhibitor SO2 or by the addition of a reductant (e.g., ascorbic acid) that can reduce the ortho-quinones back to caftaric acid. 5555

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 49. Model for the Biosynthesis of Eumelanins and Pheomelaninsa

a Tyrosinases generate highly reactive ortho-quinones that are readily attacked by cysteine. The fungal pigment pheofungin B may be formed by an analogous process.

Scheme 50. Model for the Biosynthesis of Grixazone

7.1. Thioether-Forming rSAMs

7. RADICAL S-ADENOSYLMETHIONINE ENZYMES Radical S-adenosylmethionine (rSAM) proteins form a highly diverse family of enzymes catalyzing a wide range of radical reactions.398 Members catalyze the reductive cleavage of SAM to methionine and the highly reactive 5′-deoxyadenosyl radical using a [4Fe-4S] cluster (Scheme 52). This radical is then used to perform the wide array of difficult transformationsoften the functionalization of unactivated C−H bondsthat are characteristic of the family. Although rSAM enzymes are involved in the biosynthesis of multiple classes of natural products, C−S bond-forming rSAM enzymes are scarce.

A subset of rSAM enzymes catalyze C−S bond formation. Characterized sulfur-inserting rSAMs are responsible for the formation of the thioether linkages of important cofactors (biotin and lipoic acid) and for the methylthiolation of tRNA and ribosomal proteins.399 A remarkable feature of such enzymes is their use of a second iron−sulfur cluster in C−S bond formation. This auxiliary cluster is thought to serve as a sulfur source for the cofactor synthesizing enzymes, whereas it is thought to bind the methylthio moiety prior to reaction with the substrate radical in methylthiolating enzymes. Recent studies have implicated rSAM proteins in C−S bond formation in two classes of natural products, namely, albomycin and the sactipeptides. In both instances the sulfur-inserting rSAM is thought to use an auxiliary iron−sulfur cluster to catalyze C−S bond formation. 7.1.1. Albomycin. Albomycin is a Trojan horse antibiotic from Streptomyces spp. that employs siderophore transport systems to enter bacterial cells.400−402 The molecule is composed of a potent iron chelator ferrichrome linked to a derivative of SB-217452, an aminoacyl-tRNA synthetase inhibitor (Scheme 53). After entering bacterial cells, the conjugate is hydrolyzed by cellular peptidases, thus releasing the bioactive nucleoside antibiotic.403 Albomycin features a tetrahydrothiophene ring that is essential for its antibiotic potency; it is remarkable that substitution of the sulfur atom for oxygen eliminates activity.404 The albomycin S-heterocycle is reminiscent of the tetrahydrothiophene ring of biotin, which is installed by the rSAM enzyme BioB. The stepwise C−S bond formation catalyzed by BioB has been studied in detail.398 In brief, BioB cleaves SAM units to generate 5′-deoxyadenosyl radicals, which abstract hydrogens from the methyl and methylene substituents of a cyclic urea intermediate, and the [2Fe-2S]-cluster bound to BioB serves as the sulfur donor (Scheme 53). It is plausible that 5556

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 51. Production of Grape Reaction Product in Winemaking; G-SH, Glutathione; PPO, Polyphenoloxidase

domain.235,236,417−419 Sactionine synthetase mutants lacking these auxiliary 4Fe-4S clusters are not able to catalyze thioether formation despite being able to perform SAM cleavage. Two roles have been proposed for the role of the auxiliary 4Fe-4S cluster(s) in thioether formation (Scheme 54). In the initial reconstitution report of the subtilosin biosynthetic enzyme, AlbA, it was demonstrated that the UV−vis absorption spectrum of the auxiliary 4Fe-4S cluster changes when the precursor peptide is present.417 This was used as evidence to propose that the cysteine thiol of the thioether linkage is ligated to the auxiliary metal center, which acts as an electron sink for the formation of the thioether between the carbon-centered radical and the cysteine thiol (Scheme 54).417−419 However, the direct ligation mechanism does not provide an explanation for the presence of both D- and L-sactionine linkages in subtilosin A and thuricin CD.410,420 The second mechanism proposes that the auxiliary 4Fe-4S cluster is involved in the oxidation of the carbon-centered radical to an N-acyliminium ion (Scheme 54).235,413 This planar intermediate is then attacked at either the si- or re-face to generate either the L- or D-sactionine linkage, respectively. The face of attack is proposed to be controlled by the flexibility of the substrate, akin to the mechanism observed in lanthionine formation (see section 4.1).233 While the current data does not rule out either mechanism, it is of note that the two auxiliary 4Fe-4S clusters in the SPASM-containing rSAM protein anaerobic sulfatasematurating enzyme (anSME) are thought to be involved in substrate oxidation.421 An identical thioether linkage is found in the cyclothiazomycin group of thiopeptides (Scheme 55).422−424 This linkage is formed between a cysteine thiol and a serine residue through an unknown mechanism. As the biosynthetic gene cluster lacks any genes with homology to rSAM enzymes, it is evident that this moiety is not installed in an analogous fashion to the sactionine linkages of the sactipeptides.424,425 The current biosynthetic proposal involves the formation of a dehydroalanine by a lanthipeptide dehydratase (see section 4.1) encoded in the biosynthetic gene cluster, followed by a Michael addition and subsequent rearrangement (Scheme 55).425 Notably, this putative Michael addition is not catalyzed by a lanthipeptide cyclase as one is not present in the biosynthetic cluster. 7.1.3. γ-Subunit of Quinohemoprotein. Outside of RiPP biosynthesis, a SPASM domain-containing rSAM enzyme is also involved in C−S bond formation in the posttranslational maturation of quinohemoprotein amine dehydrogenase (QHNDH).426−428 QHNDH is composed of three proteins, QhpA−C, that together catalyze the oxidative deamination of

sulfur could be introduced into the carbon backbone of the albomycin precursor molecule by an analogous mechanism.48 Indeed, analysis of the albomycin (abm) biosynthetic gene cluster showed two candidate genes, abmM and abmJ, coding for rSAM enzymes.405 The predicted gene product of abmM is similar to BioB and shows binding motifs for [4Fe-4S] and [2Fe-2S] clusters. Sulfur could be delivered by the putative cysteine desulfurase AbmD, which appears to represent an important link between primary and secondary sulfur metabolic pathways; generally, cysteine desulfurases are involved in the assembly of Fe−S clusters in bacteria.406 Interestingly, pathways responsible for increasing homocysteine levels were shown to positively impact albomycin production.407 Yet, the exact mechanism of sulfur incorporation remains to be elucidated. 7.1.2. Sactipeptides. The sactipeptides (sulfur-to-α carbon-linked peptides) are a subclass of RiPP natural products bearing thioether linkages between a cysteine thiol and the unactivated α-carbon of an acceptor amino acid (Scheme 54).22 In analogy to lanthionine rings, this moiety has recently been named a sactionine linkage.234 To date, six sactipeptides have been reported: subtilosin A, subtilosin A1, sporulation killing factor (SKF), thuricin CD (composed of two peptides, thuricin α and thuricin β), and thurincin H.206,299,408−412 A seventh sactipeptide, annotated as a member of the uncharacterized “six cysteines in forty-five” (SCIFF) group, was recently partially biosynthesized in vitro with recombinant enzymes.413,414 Initial studies regarding the biosynthesis of sactipeptides demonstrated that a member of the rSAM protein family was required for production.415,416 As many characterized rSAM proteins catalyze C−H bond abstraction,398 it was proposed that the rSAM enzyme found in all sactipeptide BGCs is responsible for sactionine formation.410,415 In the last five years, the activities of multiple sactionine synthases have been reconstituted in vitro.413,417−419 As predicted bioinformatically, all of the enzymes bind a 4Fe-4S cluster with a canonical triple cysteine motif (CxxxCxxC) that is a hallmark of rSAM enzymes. In all of the sactionine synthases, this rSAM 4Fe-4S cluster is used to reductively cleave SAM to form methionine and a 5′-deoxyadenosyl radical (5′-dAdo·; Scheme 54). Using peptide substrates that were deuterium-labeled at the α-carbon of the acceptor amino acid, it was demonstrated that this 5′dAdo extracts the Cα hydrogen to initiate thioether bond formation.235,236 In addition to the rSAM 4Fe-4S cluster, reconstituted sactionine synthases contain up to two additional 4Fe-4S clusters that are bound by conserved cysteine residues in their C-terminal SPASM (subtilosin, thuricin H, and SCIFF synthase) or Twitch (SKF and thuricin CD synthase) 5557

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 52. Model of Reductive SAM Cleavage by rSAM Enzymes

Scheme 53. Model for the Biosynthesis of the Trojan Horse Antibiotic Albomycin, Likely Involving Putative rSAM Enzymes AbmM and/or AbmJ; Box: Biotin Biosynthesis Involving an rSAM Enzyme with a [2Fe-2S] Cluster

8. NONENZYMATIC C−S BOND FORMATIONS

aliphatic amines. This protein complex is found in a variety of Gram-negative and Gram-positive bacteria. The γ-subunit of QHNDH, QhpC, is a small protein (∼9 kDa) bearing three thioether cross-links at the β-carbon and γ-carbon of aspartate and glutamate residues, respectively (Scheme 56).429,430 These linkages are required to provide structure to the small peptide. A recent report demonstrated that a homologue of the sactionine synthase, QhpD, is responsible for installing these linkages.428 In addition to the aforementioned thioether bridges, QhpC also contains a fourth thioether linkage formed between a cysteine thiol and the indole of an oxidized tryptophan residue, known as a cysteine tryptophylquinone (CTQ) moiety.429,430 The biosynthetic strategy for the formation of the C−S bond in CTQ is unknown.427

As outlined in the preceding sections, several enzymes involved in the biosynthesis of organosulfur natural products produce highly reactive intermediates that have the potential to scavenge sulfur-bearing molecules. Indeed, in some instances it is unclear whether the C−S bond in question is formed enzymatically or due to the innate reactivity of the biosynthetic intermediate. For example, the inability to identify a S-transferase in brassinin biosynthetic pathways suggests that glutathione conjugation may result from a spontaneous reaction with the electrophilic isothiocyanate moiety (see section 3.1.6). In this section, a selection of nonenzymatic additions to secondary metabolites will be presented. Apart from nucleophilic additions, C−S 5558

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 54. Model for rSAM-Catalyzed Thioether Formation in Sactipeptide Biosynthesisa

a

The two proposals for the role of the auxiliary 4Fe-4S clusters in C−S bond formation are displayed.

Scheme 55. Model for Thioether Formation in Cyclothiamycin Biosynthesis

bond-forming radical reactions are conceivable, which may be initiated by redox cycling or photoactivation by UV light.

mainly composed of sugar units, allows the binding of calicheamicin to DNA and intercalation of the enediyne “warhead” between base stacks. The trisulfide bond of the “prodrug” of the highly cytotoxic natural product is cleaved by reduction or attack of a nucleophile, which initiates an intramolecular conjugate addition of the remaining sulfide to the enone system (Scheme 57). The transition of sp2 to sp3 hybridization at the quaternary, sulfur-substituted carbon

8.1. Nonenzymatic Conjugate Addition

8.1.1. Enediyne Warhead Activation. Perhaps one of the best-known uncatalyzed C−S bond formations takes place in the context of the trigger reaction leading to calicheamicinmediated DNA double-bond scission.431,432 The “homing” unit, 5559

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 56. Structure of Mature QhpC

Scheme 57. Nonenzymatic Sulfide Conjugate Addition That Sets the Stage for Bergman or Myers−Saito Cyclizations of Calicheamicin and Neocarzinostatin, Respectivelya

a

The resulting diradical intermediate induces double-strand (ds) DNA scissions.

derive from methionine and that the methanethiol residue was incorporated intact. A mechanistically plausible route was proposed that would involve the Michael addition of the thiol to urdamycin A, followed by rearomatization and reoxidation to the quinone with molecular oxygen (Scheme 58).434 Analyses of the urdamycin435 and BE-7585A299 biosynthetic gene clusters did not provide any clues about candidate genes potentially required for C−S bond formation. Thus, it was also taken into consideration that the conjugate addition might not require any biocatalyst. To test this hypothesis, urdamycin A was incubated with either 2-thio-D-glucose or glucose in aqueous buffer at pH 8.0.299 HPLC-MS monitoring of the reaction revealed that only in the presence of the thiosugar a new compound was formed, which appeared to be the expected thio adduct. By analogy to the model reaction with urdamycin A and 2-thio-D-glucose, the C−S bond of BE-7585A would result from the conjugate addition of the preformed thiodisaccharide on the angucycline core (Scheme 58). It is remarkable that no other thio derivatives of urdamycin and BE-7585A have been detected in the broths of the producing organisms, not even the thioether that would result from the nucleophilic attack of a thiomonosaccharide formed during BE-7585A biosynthesis. Although is it possible that such a monosaccharide adduct could be transformed into BE-7585A after the Michael addition, the absence of any other thio

increases the strain on the enediyne ring and decreases the distance between the alkyne moieties. The highly strained intermediate readily undergoes a Bergman cyclization, thereby alleviating the torsional strain. The resulting aryl diradical abstracts hydrogen from the DNA backbone, thus inducing double-strand scissions (Scheme 57). An analogous transformation occurs in the activation of the enediyne natural product neocarzinostatin, albeit through a different mechanism.431,432 The 1,8 conjugate addition of a cellular thiol yields the highly reactive cummulene intermediate, which undergoes a Myers−Saito cyclization (Scheme 57). As with calicheamicin, the diradical product of this cyclization causes double-strand DNA breaks. Notably, neocarzinostatin can also be activated by a base-catalyzed intramolecular addition in the absence of a thiol nucleophile. 8.1.2. Urdamycin E and BE-7585A. Urdamycin E from Streptomyces f radiae433 and BE-7585A from Amycolatopsis orientalis (see section 5.3.2) are angucyclic polyketide antibiotics with unusual sulfur modifications. In both cases, C−S bonds between a benz[a]anthraquinone core and a thiol are installed. Feeding experiments with methyl-13C-labeled methionine, selenomethionine, and selenoethionine led to the formation of methyl-13C-labeled urdamycin E, seleno-urdamycin E, and its ethyl homologue, respectively. Thus, it was concluded that the sulfur and the methyl group of urdamycin E 5560

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 58. Tentatively Nonenzymatic Conjugate Additions of Thio Nucleophiles to Angucyclines Involved in the Formation of the Antibiotics Urdamycin A and BE-7585A

culture.437 The finding that the ralfuranone scaffold readily reacts with sulfur nucleophiles including glutathione may have implications for plant infections involving reactive oxygen species.

adducts suggests that the reaction is controlled in some way. It is conceivable that either appropriate S-nucleophiles are shuttled to the reactive angucycline core or that enzyme complexes of polyketide synthase components and sugartailoring enzymes prevent the attack of alternative sulfur compounds. 8.1.3. Ralfuranone D. Yet another type of Michael acceptor sets the stage for nonenzymatic thiol conjugate additions to yield thioethers in ralfuranones, a family of arylsubstituted furanones isolated from the plant pathogenic bacterium Ralstonia solanacearum (Scheme 59).436,437 Specifically, thiol and methyl thioether derivatives (e.g., ralfuranone D) of the previously characterized ralfuranone B436 were detected. Through analysis of the ralfuranone biosynthetic gene cluster,438 feeding experiments, and in vitro studies it was concluded that the C−S bond formation does not involve enzyme catalysis. Stable-isotope-labeling experiments with L[methyl-13C2H3] methionine indicated that the methyl thio group was introduced intact into ralfuranone D. Addition of various thiols to R. solanacearum cultures yielded the corresponding thio conjugates, including the cysteinyl and glutathione adducts. Furthermore, a new ralfuranone thioether derivative, an analogue of ralfuranone D, was obtained by adding ethanethiol to the crude extract of a R. solanacearum

8.2. Nonenzymatic Addition

8.2.1. Cyslabdan. Cyslabdan is a diterpene natural product bearing a N-acetylcysteine residue isolated from Streptomyces cyslabdanicus K04-0144.439−441 While cyslabdan alone does not display significant antibacterial activity, it potentiates the activity of carbapenem antibiotics against methicillin-resistant Staphylococcus aureus by >1000-fold.440−442 Genome mining for the biosynthetic gene cluster identified four genes cldA−D, which when expressed in a heterologous host lead to the production of cyslabdan A.443 Additionally, a new compound bearing an epoxide was detected. As cldC encoded a protein with homology to cytochrome P450 enzymes, a version of the cld biosynthetic gene cluster lacking this gene was created and screened for production of cyslabdan derivatives. Consistent with the functional annotation of CldC, the reduced diterpene derivative was obtained (Scheme 60). The discovery that the biosynthesis of cyslabdan A proceeded through an epoxide intermediate suggested that the C−S bond might be formed nonenzymatically likely through the addition of mycothiol to 5561

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

the highly reactive epoxide. Indeed, a deletion strain lacking the mycothiol amide hydrolase mca produced the mycothiolconjugated diterpene rather than cyslabdan A. Although an enzymatic route to C−S bond formation has not been ruled out for cyslabdan biosynthesis, the available data strongly suggest that mycothiol conjugation will occur spontaneously. The spontaneous nucleophilic attack of an epoxide is likely involved in the biosynthesis of the spithioneines. These ergothioneine-conjugated pyrrolizidine alkaloid natural products were isolated from the marine actinomycete Streptomyces spinoverrucosus strain SNB-048.444 Given their structure, it was proposed that the spithioneines arose from the conjugation of ergothioneine to the epoxide-containing precursor bohemamine (Scheme 61). Indeed, when bohemamine and ergothioneine were coincubated under basic conditions, spithioneine A was formed. As the biosynthetic gene cluster for these natural products has not been identified, it is unclear whether the conjugation reaction will proceed nonenzymatically in vivo. 8.2.2. Paulomycin S-Conjugates. Paulomycins are a group of isothiocyanate-containing antibiotics originally isolated from Streptomyces paulus.445−447 Large-scale fermentations of S. paulus facilitated the isolation and structural characterization of a series of paulomycin analogues (Scheme 62) where the isothiocyanate is conjugated to N-acetylcysteine (paldimycins and antibiotic 273a2) or hydrogen sulfide (U-77,802 and U-77,803).448−450 Recently, the paulomycin biosynthetic gene cluster was identified and the biosynthetic pathway was partially characterized.451−453 Although many aspects of biosynthesis remain unclear, the lack of an obvious candidate enzyme for isothiocyanate conjugation suggests that these congeners are spontaneously generated. In support of this proposal, the paldimycins are readily synthesized from the corresponding paulomycin in vitro with the addition of N-acetylcysteine and a mild base.448

Scheme 59. Vinylogous Thiol Addition Leads to Ralfuranone D Formation

Scheme 60. Cyslabdan Biosynthesis Proceeds through an Epoxide Intermediate; MSH; Mycothiol

Scheme 61. Possible Origin of Spithioneines

8.3. Photochemical and Radical-Mediated Thioconjugation

8.3.1. Panphenazines. Radical-based thioconjugation reactions have been observed for the phenazine chromophore, which is widespread among bacteria.454 Genome mining and metabolic profiling of the rare actinomycete Kitasatospora sp. led to the discovery of prenylated phenazines along with pantetheine-S-conjugates of phenazine-1-carboxylic acid (PCA).455,456 In vitro experiments showed that these so-called panphenazines could be formed in two ways, chemically induced and photoinduced, both involving phenazine radicals (Scheme 63). Phenazine radicals are formed during redox cycling processes in vivo, which involve single-electron transfers.454 PCA reacts with diverse biogenic thiols under radical-forming conditions, which provides a plausible model for irreversible glutathione depletion in cells in the presence of the bacterial phenazine pyocyanine.457 Furthermore, the phenazine UV absorption maxima at 250 and 370 nm enable photoexcitation by UV light. Indeed, irradiation of PCA in the presence of pantetheine led to the formation of panphenazines.456 Hence, it is conceivable that photoactivation of suitable (hetero)aromatic natural products may lead to thioconjugates.

responsible biosynthetic pathways and enzymes. Despite the diversity of the biosynthetic enzymes involved and of the linkages installed, common themes have emerged. For example, the biosynthetic logic for C−S bond formation first identified in the modification of tRNA and cofactor biosynthesis has been uncovered in the biosynthesis of secondary metabolites. Furthermore, pathways for the detoxification of reactive electrophilic compounds have been coopted by secondary metabolite biosynthetic enzymes to introduce sulfur from cellular thiols (e.g., glutathione and mycothiol). Similar approaches to C−S bond formation can be found in

9. CONCLUSIONS AND OUTLOOK Nature has evolved varied strategies to install carbon−sulfur bonds in natural products, and, over the last few decades, substantial progress has been made toward elucidating the 5562

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 62. Structures of Paulomycin Thiol Conjugates; NAC; N-acetylcysteine

Scheme 63. Structures of Naturally Occurring S-Conjugates of Phenazine-1-carboxylic Acid (PCA); Model for Radical-Based Thioconjugation of Phenazines

alliin, the trisulfide trigger of calicheamicin); however, there are many natural products that were omitted as their biosynthesis is completely enigmatic. For example, in addition to 6thioguanine and thioviridamide, there are a handful of thioamide-containing natural products for which the corresponding biosynthetic enzymes are unknown (Scheme 64).460−463 The current understanding of the biocatalysts involved in the production of organosulfur compounds may guide the investigation of these obscure pathways and provide a

taxonomically distant organisms, potentially arising independently or from horizontal gene transfer, as is the case for isopenicillin N synthase genes.458,459 Although the body of knowledge on the biosynthesis of sulfur-containing natural products has grown rapidly in recent years, many questions remain regarding the enzymes, precise mechanisms, and biosynthetic building blocks involved. Where possible, we have attempted to highlight these knowledge gaps (e.g., thiosugar biosynthesis, the origin of the allyl moiety of 5563

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Scheme 64. Thioamide-Containing Natural Products Where the Biosynthesis of the C−S Bond Remains Enigmatic

foundation for the discovery of novel sulfur-bearing natural products.464 Moreover, the many mechanisms involved in C−S bond formation may inspire synthetic chemists in the synthesis of complex thio compounds.

mining and elucidation of natural product biosynthetic pathways from phytopathogenic bacteria. Christian Hertweck is the Head of Department at the Leibniz Institute for Natural Product Research and Infection Biology (HKI), a Full Professor at the Friedrich Schiller University Jena, and an elected member of the National Academy of Sciences (Leopoldina). He obtained his Ph.D. in Bioorganic Chemistry in 1999 (supervisor Wilhelm Boland) at the University of Bonn and the Max Planck Institute for Chemical Ecology before his postdoctoral work at the University of Washington, Seattle, as a Humboldt fellow of Heinz G. Floss and Bradley S. Moore. His research focuses on the biosynthesis of microbial natural products and their roles in interspecies interactions. In 2015 he was awarded the Leibniz Prize in recognition of his contribution to the field.

AUTHOR INFORMATION Corresponding Author

*Tel.: +49 3641 532 1100. Fax: +49 3641 532 0804. E-mail: [email protected]. Notes

The authors declare no competing financial interest. Biographies Kyle Dunbar obtained his B.Sc. in Chemical Biology from the University of California, Berkeley. In 2014, he received his Ph.D. in Chemistry from University of Illinois at Urbana−Champaign under the supervision of Douglas Mitchell. His Ph.D. work focused on the biosynthesis of azoline heterocycles in ribosomal natural products. Following a short postdoctoral position at the University of Illinois at Urbana−Champaign, Kyle joined the group of Christian Hertweck at the Leibniz Institute for Natural Product Research and Infection Biology as an Alexander von Humboldt postdoctoral fellow. His current research focuses on the molecular characterization of interspecies interactions.

ACKNOWLEDGMENTS We thank Florian Kloss for valuable discussions during the initial planning phase and Evelyn Molloy for the critical review of the manuscript. Financial support by the Alexander von Humboldt Foundation (for K.L.D.) and the National Academy of Sciences (Leopoldina) (for D.H.S.) is gratefully acknowledged. REFERENCES (1) Wächtershäuser, G. From Volcanic Origins of Chemoautotrophic Life to Bacteria, Archaea and Eukarya. Philos. Trans. R. Soc., B 2006, 361, 1787−1806. (2) Richter, M. Functional Diversity of Organic Molecule Enzyme Cofactors. Nat. Prod. Rep. 2013, 30, 1324−1345. (3) Cipollone, R.; Ascenzi, P.; Tomao, P.; Imperi, F.; Visca, P. Enzymatic Detoxification of Cyanide: Clues from Pseudomonas Aeruginosa Rhodanese. J. Mol. Microbiol. Biotechnol. 2008, 15, 199− 211. (4) Fahey, R. C. Glutathione Analogs in Prokaryotes. Biochim. Biophys. Acta, Gen. Subj. 2013, 1830, 3182−3198. (5) Shigi, N. Biosynthesis and Functions of Sulfur Modifications in tRNA. Front. Genet. 2014, 5, 67. (6) Begley, T. P.; Xi, J.; Kinsland, C.; Taylor, S.; McLafferty, F. The Enzymology of Sulfur Activation During Thiamin and Biotin Biosynthesis. Curr. Opin. Chem. Biol. 1999, 3, 623−629. (7) Mueller, E. G. Trafficking in Persulfides: Delivering Sulfur in Biosynthetic Pathways. Nat. Chem. Biol. 2006, 2, 185−194.

Daniel H. Scharf received his Diploma degree (M.S. equiv) in Biology from the Friedrich Schiller University Jena, Germany, in 2008. He then continued his Ph.D. studies at the Leibniz Institute for Natural Product Research and Infection Biology under the supervision of Axel A. Brakhage. His Ph.D. researched focused on the gliotoxin biosynthetic pathway in Aspergillus f umigatus. After a postdoc in his thesis laboratory, he joined the group of Georgios Skiniotis at the Life Sciences Institute, University of Michigan, in 2016. His current research interests concern the structural and mechanistic characterization of polyketide synthases and nonribosomal peptide synthetases. Agnieszka Litomska was born in Lodz, Poland, and received her M.Sc. degree in Biotechnology, with specialization in Molecular Biotechnology and Technical Biochemistry, from Lodz University of Technology in 2012. She is currently a Ph.D. student in the laboratory of Christian Hertweck at the Leibniz Institute for Natural Product Research and Infection Biology in Jena, Germany. Her research focuses on genome 5564

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Staphylococcal Virulence. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, 10679−10684. (28) Bagley, M. C.; Dale, J. W.; Merritt, E. A.; Xiong, X. Thiopeptide Antibiotics. Chem. Rev. 2005, 105, 685−714. (29) Pascard, C.; Ducruix, A.; Lunel, J.; Prange, T. Highly Modified Cysteine-Containing Antibiotics. Chemical Structure and Configuration of Nosiheptide. J. Am. Chem. Soc. 1977, 99, 6418−6423. (30) Northcote, P. T.; Siegel, M.; Borders, D. B.; Lee, M. D. Glycothiohexide Alpha, a Novel Antibiotic Produced by ″Sebekia″ Sp., Ll-14e605. Iii. Structural Elucidation. J. Antibiot. 1994, 47, 901−908. (31) Yu, Y.; Duan, L.; Zhang, Q.; Liao, R.; Ding, Y.; Pan, H.; WendtPienkowski, E.; Tang, G.; Shen, B.; Liu, W. Nosiheptide Biosynthesis Featuring a Unique Indole Side Ring Formation on the Characteristic Thiopeptide Framework. ACS Chem. Biol. 2009, 4, 855−864. (32) Fischbach, M. A.; Walsh, C. T. Assembly-Line Enzymology for Polyketide and Nonribosomal Peptide Antibiotics: Logic, Machinery, and Mechanisms. Chem. Rev. 2006, 106, 3468−3496. (33) Du, L.; Lou, L. Pks and Nrps Release Mechanisms. Nat. Prod. Rep. 2010, 27, 255−278. (34) Dawson, S.; Malkinson, J. P.; Paumier, D.; Searcey, M. Bisintercalator Natural Products with Potential Therapeutic Applications: Isolation, Structure Determination, Synthetic and Biological Studies. Nat. Prod. Rep. 2007, 24, 109−126. (35) Lombó, F.; Velasco, A.; Castro, A.; De la Calle, F.; Braña, A. F.; Sánchez-Puelles, J. M.; Méndez, C.; Salas, J. A. Deciphering the Biosynthesis Pathway of the Antitumor Thiocoraline from a Marine Actinomycete and Its Expression in Two Streptomyces Species. ChemBioChem 2006, 7, 366−376. (36) Robbel, L.; Hoyer, K. M.; Marahiel, M. A. Tios T-Te − a Prototypical Thioesterase Responsible for Cyclodimerization of the Quinoline- and Quinoxaline-Type Class of Chromodepsipeptides. FEBS J. 2009, 276, 1641−1653. (37) Heath, R. J.; Rock, C. O. The Claisen Condensation in Biology. Nat. Prod. Rep. 2002, 19, 581−596. (38) Abugrain, M. E.; Brumsted, C. J.; Osborn, A. R.; Philmus, B.; Mahmud, T. A Highly Promiscuous Ss-Ketoacyl-Acp Synthase (Kas) Iii-Like Protein Is Involved in Pactamycin Biosynthesis. ACS Chem. Biol. 2017, 12, 362−366. (39) Sun, L.; Wang, S.; Zhang, S.; Shao, L.; Zhang, Q.; Skidmore, C.; Chang, C. W.; Yu, D.; Zhan, J. Characterization of Three Tailoring Enzymes in Dutomycin Biosynthesis and Generation of a Potent Antibacterial Analogue. ACS Chem. Biol. 2016, 11, 1992−2001. (40) Kresovic, D.; Schempp, F.; Cheikh-Ali, Z.; Bode, H. B. A Novel and Widespread Class of Ketosynthase Is Responsible for the Head-toHead Condensation of Two Acyl Moieties in Bacterial Pyrone Biosynthesis. Beilstein J. Org. Chem. 2015, 11, 1412−1417. (41) Lv, M.; Zhao, J.; Deng, Z.; Yu, Y. Characterization of the Biosynthetic Gene Cluster for Benzoxazole Antibiotics A33853 Reveals Unusual Assembly Logic. Chem. Biol. 2015, 22, 1313−1324. (42) Proschak, A.; Zhou, Q.; Schoner, T.; Thanwisai, A.; Kresovic, D.; Dowling, A.; ffrench-Constant, R.; Proschak, E.; Bode, H. B. Biosynthesis of the Insecticidal Xenocyloins in Xenorhabdus Bovienii. ChemBioChem 2014, 15, 369−372. (43) Kwon, H. J.; Smith, W. C.; Scharon, A. J.; Hwang, S. H.; Kurth, M. J.; Shen, B. C-O Bond Formation by Polyketide Synthases. Science 2002, 297, 1327−1330. (44) Ahlert, J.; Shepard, E.; Lomovskaya, N.; Zazopoulos, E.; Staffa, A.; Bachmann, B. O.; Huang, K.; Fonstein, L.; Czisny, A.; Whitwam, R. E.; et al. The Calicheamicin Gene Cluster and Its Iterative Type I Enediyne Pks. Science 2002, 297, 1173−1176. (45) Freitag, A.; Wemakor, E.; Li, S. M.; Heide, L. Acyl Transfer in Clorobiocin Biosynthesis: Involvement of Several Proteins in the Transfer of the Pyrrole-2-Carboxyl Moiety to the Deoxysugar. ChemBioChem 2005, 6, 2316−2325. (46) He, Q. L.; Jia, X. Y.; Tang, M. C.; Tian, Z. H.; Tang, G. L.; Liu, W. Dissection of Two Acyl-Transfer Reactions Centered on Acyl-SCarrier Protein Intermediates for Incorporating 5-Chloro-6-Methyl-OMethylsalicyclic Acid into Chlorothricin. ChemBioChem 2009, 10, 813−819.

(8) Fontecave, M.; Ollagnier-de-Choudens, S.; Mulliez, E. Biological Radical Sulfur Insertion Reactions. Chem. Rev. 2003, 103, 2149−2166. (9) Kessler, D. Enzymatic Activation of Sulfur for Incorporation into Biomolecules in Prokaryotes. FEMS Microbiol. Rev. 2006, 30, 825− 840. (10) Beld, J.; Sonnenschein, E. C.; Vickery, C. R.; Noel, J. P.; Burkart, M. D. The Phosphopantetheinyl Transferases: Catalysis of a PostTranslational Modification Crucial for Life. Nat. Prod. Rep. 2014, 31, 61−108. (11) Struck, A. W.; Thompson, M. L.; Wong, L. S.; Micklefield, J. SAdenosyl-Methionine-Dependent Methyltransferases: Highly Versatile Enzymes in Biocatalysis, Biosynthesis and Other Biotechnological Applications. ChemBioChem 2012, 13, 2642−2655. (12) Franke, J.; Hertweck, C. Biomimetic Thioesters as Probes for Enzymatic Assembly Lines: Synthesis, Applications, and Challenges. Cell Chem. Biol. 2016, 23, 1179−1192. (13) Marraffini, L. A.; DeDent, A. C.; Schneewind, O. Sortases and the Art of Anchoring Proteins to the Envelopes of Gram-Positive Bacteria. Microbiol. Mol. Biol. Rev. 2006, 70, 192−221. (14) Sauvage, E.; Kerff, F.; Terrak, M.; Ayala, J. A.; Charlier, P. The Penicillin-Binding Proteins: Structure and Role in Peptidoglycan Biosynthesis. FEMS Microbiol. Rev. 2008, 32, 234−258. (15) Lee, J.; McIntosh, J.; Hathaway, B. J.; Schmidt, E. W. Using Marine Natural Products to Discover a Protease That Catalyzes Peptide Macrocyclization of Diverse Substrates. J. Am. Chem. Soc. 2009, 131, 2122−2124. (16) Barber, C. J. S.; Pujara, P. T.; Reed, D. W.; Chiwocha, S.; Zhang, H.; Covello, P. S. The Two-Step Biosynthesis of Cyclic Peptides from Linear Precursors in a Member of the Plant Family Caryophyllaceae Involves Cyclization by a Serine Protease-Like Enzyme. J. Biol. Chem. 2013, 288, 12500−12510. (17) Luo, H.; Hong, S. Y.; Sgambelluri, R. M.; Angelos, E.; Li, X.; Walton, J. D. Peptide Macrocyclization Catalyzed by a Prolyl Oligopeptidase Involved in Alpha-Amanitin Biosynthesis. Chem. Biol. 2014, 21, 1610−1617. (18) Nguyen, G. K.; Wang, S.; Qiu, Y.; Hemu, X.; Lian, Y.; Tam, J. P. Butelase 1 Is an Asx-Specific Ligase Enabling Peptide Macrocyclization and Synthesis. Nat. Chem. Biol. 2014, 10, 732−738. (19) Mylne, J. S.; Colgrave, M. L.; Daly, N. L.; Chanson, A. H.; Elliott, A. G.; McCallum, E. J.; Jones, A.; Craik, D. J. Albumins and Their Processing Machinery Are Hijacked for Cyclic Peptides in Sunflower. Nat. Chem. Biol. 2011, 7, 257−259. (20) Thoendel, M.; Horswill, A. R. Biosynthesis of Peptide Signals in Gram-Positive Bacteria. Adv. Appl. Microbiol. 2010, 71, 91−112. (21) Wang, B.; Muir, T. W. Regulation of Virulence in Staphylococcus Aureus: Molecular Mechanisms and Remaining Puzzles. Cell Chem. Biol. 2016, 23, 214−224. (22) 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. (23) Qiu, R.; Pei, W.; Zhang, L.; Lin, J.; Ji, G. Identification of the Putative Staphylococcal Agrb Catalytic Residues Involving the Proteolytic Cleavage of Agrd to Generate Autoinducing Peptide. J. Biol. Chem. 2005, 280, 16695−16704. (24) Zhang, L.; Gray, L.; Novick, R. P.; Ji, G. Transmembrane Topology of Agrb, the Protein Involved in the Post-Translational Modification of Agrd in Staphylococcus Aureus. J. Biol. Chem. 2002, 277, 34736−34742. (25) Thoendel, M.; Horswill, A. R. Identification of Staphylococcus Aureus Agrd Residues Required for Autoinducing Peptide Biosynthesis. J. Biol. Chem. 2009, 284, 21828−21838. (26) Kavanaugh, J. S.; Thoendel, M.; Horswill, A. R. A Role for Type I Signal Peptidase in Staphylococcus Aureus Quorum Sensing. Mol. Microbiol. 2007, 65, 780−798. (27) Wang, B.; Zhao, A.; Novick, R. P.; Muir, T. W. Key Driving Forces in the Biosynthesis of Autoinducing Peptides Required for 5565

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Mechanism for Formation of the 2,2′-Bipyridyl Ring. Chem. Biol. 2012, 19, 399−413. (66) Zhu, Y.; Fu, P.; Lin, Q.; Zhang, G.; Zhang, H.; Li, S.; Ju, J.; Zhu, W.; Zhang, C. Identification of Caerulomycin a Gene Cluster Implicates a Tailoring Amidohydrolase. Org. Lett. 2012, 14, 2666− 2669. (67) Qu, X.; Pang, B.; Zhang, Z.; Chen, M.; Wu, Z.; Zhao, Q.; Zhang, Q.; Wang, Y.; Liu, Y.; Liu, W. Caerulomycins and Collismycins Share a Common Paradigm for 2,2′-Bipyridine Biosynthesis Via an Unusual Hybrid Polyketide-Peptide Assembly Logic. J. Am. Chem. Soc. 2012, 134, 9038−9041. (68) Watanabe, K.; Hotta, K.; Praseuth, A. P.; Koketsu, K.; Migita, A.; Boddy, C. N.; Wang, C. C.; Oguri, H.; Oikawa, H. Total Biosynthesis of Antitumor Nonribosomal Peptides in Escherichia Coli. Nat. Chem. Biol. 2006, 2, 423−428. (69) Hotta, K.; Keegan, R. M.; Ranganathan, S.; Fang, M.; Bibby, J.; Winn, M. D.; Sato, M.; Lian, M.; Watanabe, K.; Rigden, D. J.; et al. Conversion of a Disulfide Bond into a Thioacetal Group During Echinomycin Biosynthesis. Angew. Chem., Int. Ed. 2014, 53, 824−828. (70) Spizek, J.; Novotna, J.; Rezanka, T. Lincosamides: Chemical Structure, Biosynthesis, Mechanism of Action, Resistance, and Applications. Adv. Appl. Microbiol. 2004, 56, 121−154. (71) Najmanova, L.; Kutejova, E.; Kadlec, J.; Polan, M.; Olsovska, J.; Benada, O.; Novotna, J.; Kamenik, Z.; Halada, P.; Bauer, J.; et al. Characterization of N-Demethyllincosamide Methyltransferases Lmbj and Ccbj. ChemBioChem 2013, 14, 2259−2262. (72) Novotna, J.; Olsovska, J.; Novak, P.; Mojzes, P.; Chaloupkova, R.; Kamenik, Z.; Spizek, J.; Kutejova, E.; Mareckova, M.; Tichy, P.; et al. Lincomycin Biosynthesis Involves a Tyrosine Hydroxylating Heme Protein of an Unusual Enzyme Family. PLoS One 2013, 8, e79974. (73) Kadlcik, S.; Kucera, T.; Chalupska, D.; Gazak, R.; Koberska, M.; Ulanova, D.; Kopecky, J.; Kutejova, E.; Najmanova, L.; Janata, J. Adaptation of an L-Proline Adenylation Domain to Use 4-Propyl-LProline in the Evolution of Lincosamide Biosynthesis. PLoS One 2013, 8, e84902. (74) Janata, J.; Kadlcik, S.; Koberska, M.; Ulanova, D.; Kamenik, Z.; Novak, P.; Kopecky, J.; Novotna, J.; Radojevic, B.; Plhackova, K.; et al. Lincosamide Synthetase–a Unique Condensation System Combining Elements of Nonribosomal Peptide Synthetase and Mycothiol Metabolism. PLoS One 2015, 10, e0118850. (75) Lin, C. I.; Sasaki, E.; Zhong, A.; Liu, H. W. In Vitro Characterization of Lmbk and Lmbo: Identification of Gdp-D-ErythroAlpha-D-Gluco-Octose as a Key Intermediate in Lincomycin a Biosynthesis. J. Am. Chem. Soc. 2014, 136, 906−909. (76) Wang, M.; Zhao, Q.; Zhang, Q.; Liu, W. Differences in PlpDependent Cysteinyl Processing Lead to Diverse S-Functionalization of Lincosamide Antibiotics. J. Am. Chem. Soc. 2016, 138, 6348−6351. (77) Ushimaru, R.; Lin, C. I.; Sasaki, E.; Liu, H. W. Characterization of Enzymes Catalyzing Transformations of Cysteine S-Conjugated Intermediates in the Lincosamide Biosynthetic Pathway. ChemBioChem 2016, 17, 1606−1611. (78) Al-Mestarihi, A. H.; Villamizar, G.; Fernández, J.; Zolova, O. E.; Lombó, F.; Garneau-Tsodikova, S. Adenylation and S-Methylation of Cysteine by the Bifunctional Enzyme Tion in Thiocoraline Biosynthesis. J. Am. Chem. Soc. 2014, 136, 17350−17354. (79) Labby, K. J.; Watsula, S. G.; Garneau-Tsodikova, S. Interrupted Adenylation Domains: Unique Bifunctional Enzymes Involved in Nonribosomal Peptide Biosynthesis. Nat. Prod. Rep. 2015, 32, 641− 653. (80) Pedras, M. S.; Yaya, E. E.; Glawischnig, E. The Phytoalexins from Cultivated and Wild Crucifers: Chemistry and Biology. Nat. Prod. Rep. 2011, 28, 1381−1405. (81) Chripkova, M.; Zigo, F.; Mojzis, J. Antiproliferative Effect of Indole Phytoalexins. Molecules 2016, 21, 1626. (82) Monde, K.; Takasugi, M.; Ohnishi, T. Biosynthesis of Cruciferous Phytoalexins. J. Am. Chem. Soc. 1994, 116, 6650−6657.

(47) Bretschneider, T.; Zocher, G.; Unger, M.; Scherlach, K.; Stehle, T.; Hertweck, C. A Ketosynthase Homolog Uses Malonyl Units to Form Esters in Cervimycin Biosynthesis. Nat. Chem. Biol. 2012, 8, 154−161. (48) Lin, C. I.; McCarty, R. M.; Liu, H. W. The Biosynthesis of Nitrogen-, Sulfur-, and High-Carbon Chain-Containing Sugars. Chem. Soc. Rev. 2013, 42, 4377−4407. (49) Lairson, L. L.; Henrissat, B.; Davies, G. J.; Withers, S. G. Glycosyltransferases: Structures, Functions, and Mechanisms. Annu. Rev. Biochem. 2008, 77, 521−555. (50) Liscombe, D. K.; Louie, G. V.; Noel, J. P. Architectures, Mechanisms and Molecular Evolution of Natural Product Methyltransferases. Nat. Prod. Rep. 2012, 29, 1238−1250. (51) Matthews, R. G.; Koutmos, M.; Datta, S. Cobalamin-Dependent and Cobamide-Dependent Methyltransferases. Curr. Opin. Struct. Biol. 2008, 18, 658−666. (52) Clarke, S. G. Protein Methylation at the Surface and Buried Deep: Thinking Outside the Histone Box. Trends Biochem. Sci. 2013, 38, 243−252. (53) Drotar, A.; Fall, L. R.; Mishalaine, E. A.; Tavernier, J. E.; Fall, R. Enzymatic Methylation of Sulfide, Selenide, and Organic Thiols by Tetrahymena Thermophila. Appl. Environ. Microbiol. 1987, 53, 2111− 2118. (54) Scharf, D. H.; Heinekamp, T.; Brakhage, A. A. Human and Plant Fungal Pathogens: The Role of Secondary Metabolites. PLoS Pathog. 2014, 10, e1003859. (55) Scharf, D. H.; Groll, M.; Habel, A.; Heinekamp, T.; Hertweck, C.; Brakhage, A. A.; Huber, E. M. Flavoenzyme-Catalyzed Formation of Disulfide Bonds in Natural Products. Angew. Chem., Int. Ed. 2014, 53, 2221−2224. (56) Scharf, D. H.; Remme, N.; Heinekamp, T.; Hortschansky, P.; Brakhage, A. A.; Hertweck, C. Transannular Disulfide Formation in Gliotoxin Biosynthesis and Its Role in Self-Resistance of the Human Pathogen Aspergillus Fumigatus. J. Am. Chem. Soc. 2010, 132, 10136− 10141. (57) Dolan, S. K.; Owens, R. A.; O’Keeffe, G.; Hammel, S.; Fitzpatrick, D. A.; Jones, G. W.; Doyle, S. Regulation of Nonribosomal Peptide Synthesis: Bis-Thiomethylation Attenuates Gliotoxin Biosynthesis in Aspergillus Fumigatus. Chem. Biol. 2014, 21, 999−1012. (58) Scharf, D. H.; Habel, A.; Heinekamp, T.; Brakhage, A. A.; Hertweck, C. Opposed Effects of Enzymatic Gliotoxin N- and SMethylations. J. Am. Chem. Soc. 2014, 136, 11674−11679. (59) Duell, E. R.; Glaser, M.; Le Chapelain, C.; Antes, I.; Groll, M.; Huber, E. M. Sequential Inactivation of Gliotoxin by the SMethyltransferase Tmta. ACS Chem. Biol. 2016, 11, 1082−1089. (60) Li, B.; Forseth, R. R.; Bowers, A. A.; Schroeder, F. C.; Walsh, C. T. A Backup Plan for Self-Protection: S-Methylation of Holomycin Biosynthetic Intermediates in Streptomyces Clavuligerus. ChemBioChem 2012, 13, 2521−2526. (61) Shindo, K.; Yamagishi, Y.; Okada, Y.; Kawai, H. Collismycins a and B, Novel Non-Steroidal Inhibitors of Dexamethasone-Glucocorticoid Receptor Binding. J. Antibiot. 1994, 47, 1072−1074. (62) Garcia, I.; Vior, N. M.; González-Sabín, J.; Braña, A. F.; Rohr, J.; Moris, F.; Méndez, C.; Salas, J. A. Engineering the Biosynthesis of the Polyketide-Nonribosomal Peptide Collismycin a for Generation of Analogs with Neuroprotective Activity. Chem. Biol. 2013, 20, 1022− 1032. (63) Gomi, S.; Amano, S.; Sato, E.; Miyadoh, S.; Kodama, Y. Novel Antibiotics Sf2738a, B and C, and Their Analogs Produced by Streptomyces Sp. J. Antibiot. 1994, 47, 1385−1394. (64) Stadler, M.; Bauch, F.; Henkel, T.; Mühlbauer, A.; Müller, H.; Spaltmann, F.; Weber, K. Antifungal Actinomycete Metabolites Discovered in a Differential Cell-Based Screening Using a Recombinant Topo1 Deletion Mutant Strain. Arch. Pharm. 2001, 334, 143− 147. (65) Garcia, I.; Vior, N. M.; Braña, A. F.; González-Sabin, J.; Rohr, J.; Moris, F.; Méndez, C.; Salas, J. A. Elucidating the Biosynthetic Pathway for the Polyketide-Nonribosomal Peptide Collismycin A: 5566

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

(83) Klein, A. P.; Sattely, E. S. Biosynthesis of Cabbage Phytoalexins from Indole Glucosinolate. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 1910−1915. (84) Mikkelsen, M. D.; Naur, P.; Halkier, B. A. Arabidopsis Mutants in the C-S Lyase of Glucosinolate Biosynthesis Establish a Critical Role for Indole-3-Acetaldoxime in Auxin Homeostasis. Plant J. 2004, 37, 770−777. (85) Yun, B. S.; Hidaka, T.; Furihata, K.; Seto, H. Microbial Metabolites with Tipa Promoter Inducing Activity. Iii. Thioxamycin and Its Novel Derivative, Thioactin, Two Thiopeptides Produced by Streptomyces Sp. Dp94. J. Antibiot. 1994, 47, 1541−1545. (86) Matsumoto, M.; Kawamura, Y.; Yasuda, Y.; Tanimoto, T.; Matsumoto, K.; Yoshida, T.; Shoji, J. Isolation and Characterization of Thioxamycin. J. Antibiot. 1989, 42, 1465−1469. (87) Puar, M. S.; Chan, T. M.; Hegde, V.; Patel, M.; Bartner, P.; Ng, K. J.; Pramanik, B. N.; MacFarlane, R. D. Sch 40832: A Novel Thiostrepton from Micromonospora Carbonacea. J. Antibiot. 1998, 51, 221−224. (88) Freeman, M. F.; Gurgui, C.; Helf, M. J.; Morinaka, B. I.; Uria, A. R.; Oldham, N. J.; Sahl, H. G.; Matsunaga, S.; Piel, J. Metagenome Mining Reveals Polytheonamides as Posttranslationally Modified Ribosomal Peptides. Science 2012, 338, 387−390. (89) Wilson, M. C.; Mori, T.; Ruckert, C.; Uria, A. R.; Helf, M. J.; Takada, K.; Gernert, C.; Steffens, U. A.; Heycke, N.; Schmitt, S.; et al. An Environmental Bacterial Taxon with a Large and Distinct Metabolic Repertoire. Nature 2014, 506, 58−62. (90) Helf, M. J.; Jud, A.; Piel, J. Enzyme from an Uncultivated Sponge Bacterium Catalyzes S-Methylation in a Ribosomal Peptide. ChemBioChem 2017, 18, 444−450. (91) Lian, G.; Zhang, X.; Yu, B. Thioglycosides in Carbohydrate Research. Carbohydr. Res. 2015, 403, 13−22. (92) Oman, T. J.; Boettcher, J. M.; Wang, H.; Okalibe, X. N.; van der Donk, W. A. Sublancin Is Not a Lantibiotic but an S-Linked Glycopeptide. Nat. Chem. Biol. 2011, 7, 78−80. (93) Stepper, J.; Shastri, S.; Loo, T. S.; Preston, J. C.; Novak, P.; Man, P.; Moore, C. H.; Havlicek, V.; Patchett, M. L.; Norris, G. E. Cysteine S-Glycosylation, a New Post-Translational Modification Found in Glycopeptide Bacteriocins. FEBS Lett. 2011, 585, 645−650. (94) Li, J.; Wang, C.; Zhang, Z. M.; Cheng, Y. Q.; Zhou, J. H. The Structural Basis of an Nadp(+)-Independent Dithiol Oxidase in Fk228 Biosynthesis. Sci. Rep. 2014, 4, 4145. (95) Hata, T.; Tanaka, R.; Ohmomo, S. Isolation and Characterization of Plantaricin Asm1: A New Bacteriocin Produced by Lactobacillus Plantarum a-1. Int. J. Food Microbiol. 2010, 137, 94−99. (96) Garcia De Gonzalo, C. V.; Denham, E. L.; Mars, R. A.; Stulke, J.; van der Donk, W. A.; van Dijl, J. M. The Phosphoenolpyruvate:Sugar Phosphotransferase System Is Involved in Sensitivity to the Glucosylated Bacteriocin Sublancin. Antimicrob. Agents Chemother. 2015, 59, 6844−6854. (97) Wang, H.; van der Donk, W. A. Substrate Selectivity of the Sublancin S-Glycosyltransferase. J. Am. Chem. Soc. 2011, 133, 16394− 16397. (98) Garcia De Gonzalo, C. V.; Zhu, L.; Oman, T. J.; van der Donk, W. A. Nmr Structure of the S-Linked Glycopeptide Sublancin 168. ACS Chem. Biol. 2014, 9, 796−801. (99) Maky, M. A.; Ishibashi, N.; Zendo, T.; Perez, R. H.; Doud, J. R.; Karmi, M.; Sonomoto, K. Enterocin F4−9, a Novel O-Linked Glycosylated Bacteriocin. Appl. Environ. Microbiol. 2015, 81, 4819− 4826. (100) Venugopal, H.; Edwards, P. J.; Schwalbe, M.; Claridge, J. K.; Libich, D. S.; Stepper, J.; Loo, T.; Patchett, M. L.; Norris, G. E.; Pascal, S. M. Structural, Dynamic, and Chemical Characterization of a Novel S-Glycosylated Bacteriocin. Biochemistry 2011, 50, 2748−2755. (101) Oman, T. J.; van der Donk, W. A. Follow the Leader: The Use of Leader Peptides to Guide Natural Product Biosynthesis. Nat. Chem. Biol. 2010, 6, 9−18. (102) Halkier, B. A.; Gershenzon, J. Biology and Biochemistry of Glucosinolates. Annu. Rev. Plant Biol. 2006, 57, 303−333.

(103) Cartea, M. E.; Velasco, P. Glucosinolates in Brassica Foods: Bioavailability in Food and Significance for Human Health. Phytochem. Rev. 2008, 7, 213−229. (104) Marillia, E. F.; MacPherson, J. M.; Tsang, E. W.; Van Audenhove, K.; Keller, W. A.; GrootWassink, J. W. Molecular Cloning of a Brassica Napus Thiohydroximate S-Glucosyltransferase Gene and Its Expression in Escherichia Coli. Physiol. Plant. 2001, 113, 176−184. (105) Grubb, C. D.; Zipp, B. J.; Ludwig-Muller, J.; Masuno, M. N.; Molinski, T. F.; Abel, S. Arabidopsis Glucosyltransferase Ugt74b1 Functions in Glucosinolate Biosynthesis and Auxin Homeostasis. Plant J. 2004, 40, 893−908. (106) Marroun, S.; Montaut, S.; Marquès, S.; Lafite, P.; Coadou, G.; Rollin, P.; Jousset, G.; Schuler, M.; Tatibouët, A.; Oulyadi, H.; et al. Ugt74b1 from Arabidopsis Thaliana as a Versatile Biocatalyst for the Synthesis of Desulfoglycosinolates. Org. Biomol. Chem. 2016, 14, 6252−6261. (107) Zhao, Q.; Wang, M.; Xu, D.; Zhang, Q.; Liu, W. Metabolic Coupling of Two Small-Molecule Thiols Programs the Biosynthesis of Lincomycin A. Nature 2015, 518, 115−119. (108) Sheweita, S. A. Drug-Metabolizing Enzymes: Mechanisms and Functions. Curr. Drug Metab. 2000, 1, 107−132. (109) Cummins, I.; Dixon, D. P.; Freitag-Pohl, S.; Skipsey, M.; Edwards, R. Multiple Roles for Plant Glutathione Transferases in Xenobiotic Detoxification. Drug Metab. Rev. 2011, 43, 266−280. (110) Bisgaard, H. Pathophysiology of the Cysteinyl Leukotrienes and Effects of Leukotriene Receptor Antagonists in Asthma. Allergy 2001, 56, 7−11. (111) Austen, K. F. The Cysteinyl Leukotrienes: Where Do They Come From? What Are They? Where Are They Going? Nat. Immunol. 2008, 9, 113−115. (112) Dvash, E.; Har-Tal, M.; Barak, S.; Meir, O.; Rubinstein, M. Leukotriene C4 Is the Major Trigger of Stress-Induced Oxidative DNA Damage. Nat. Commun. 2015, 6, 10112. (113) Morris, H. R.; Taylor, G. W.; Jones, C. M.; Piper, P. J.; Samhoun, M. N.; Tippins, J. R. Slow Reacting Substances (Leukotrienes): Enzymes Involved in Their Biosynthesis. Proc. Natl. Acad. Sci. U. S. A. 1982, 79, 4838−4842. (114) Haeggström, J. Z.; Funk, C. D. Lipoxygenase and Leukotriene Pathways: Biochemistry, Biology, and Roles in Disease. Chem. Rev. 2011, 111, 5866−5898. (115) Hammerström, S.; Samuelsson, B. Detection of Leukotriene A4 as an Intermediate in the Biosynthesis of Leukotrienes C4 and D4. FEBS Lett. 1980, 122, 83−86. (116) Sirois, P.; Borgeat, P. From Slow Reacting Substance of Anaphylaxis (Srs-a) to Leukotriene D4 (Ltd4). Int. J. Immunopharmacol. 1980, 2, 281−293. (117) Anderson, M. E.; Allison, R. D.; Meister, A. Interconversion of Leukotrienes Catalyzed by Purified Gamma-Glutamyl Transpeptidase: Concomitant Formation of Leukotriene D4 and Gamma-Glutamyl Amino Acids. Proc. Natl. Acad. Sci. U. S. A. 1982, 79, 1088−1091. (118) Lee, C. W.; Lewis, R. A.; Corey, E. J.; Austen, K. F. Conversion of Leukotriene D4 to Leukotriene E4 by a Dipeptidase Released from the Specific Granule of Human Polymorphonuclear Leucocytes. Immunology 1983, 48, 27−35. (119) Amatov, T.; Jahn, U. Gliotoxin: Nature’s Way of Making the Epidithio Bridge. Angew. Chem., Int. Ed. 2014, 53, 3312−3314. (120) Welch, T. R.; Williams, R. M. Epidithiodioxopiperazines. Occurrence, Synthesis and Biogenesis. Nat. Prod. Rep. 2014, 31, 1376− 1404. (121) Scharf, D. H.; Heinekamp, T.; Remme, N.; Hortschansky, P.; Brakhage, A. A.; Hertweck, C. Biosynthesis and Function of Gliotoxin in Aspergillus Fumigatus. Appl. Microbiol. Biotechnol. 2012, 93, 467− 472. (122) Gallagher, C. H. Skin Lesions of Sheep Caused by Sporidesmin. Nature 1964, 201, 1293−1294. (123) Russell, G. R. Detection and Estimation of a Facial Eczema Toxin. Sporidesmin. Nature 1960, 186, 788−789. (124) Gardiner, D. M.; Cozijnsen, A. J.; Wilson, L. M.; Pedras, M. S.; Howlett, B. J. The Sirodesmin Biosynthetic Gene Cluster of the Plant 5567

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Pathogenic Fungus Leptosphaeria Maculans. Mol. Microbiol. 2004, 53, 1307−1318. (125) Towers, N. R.; Wright, D. E. Biosynthesis of Sporidesmin from Amino Acids. N. Z. J. Agric. Res. 1969, 12, 275−280. (126) Davis, C.; Carberry, S.; Schrettl, M.; Singh, I.; Stephens, J. C.; Barry, S. M.; Kavanagh, K.; Challis, G. L.; Brougham, D.; Doyle, S. The Role of Glutathione S-Transferase Glig in Gliotoxin Biosynthesis in Aspergillus Fumigatus. Chem. Biol. 2011, 18, 542−552. (127) Scharf, D. H.; Remme, N.; Habel, A.; Chankhamjon, P.; Scherlach, K.; Heinekamp, T.; Hortschansky, P.; Brakhage, A. A.; Hertweck, C. A Dedicated Glutathione S-Transferase Mediates Carbon-Sulfur Bond Formation in Gliotoxin Biosynthesis. J. Am. Chem. Soc. 2011, 133, 12322−12325. (128) Scharf, D. H.; Chankhamjon, P.; Scherlach, K.; Heinekamp, T.; Willing, K.; Brakhage, A. A.; Hertweck, C. Epidithiodiketopiperazine Biosynthesis: A Four-Enzyme Cascade Converts Glutathione Conjugates into Transannular Disulfide Bridges. Angew. Chem., Int. Ed. 2013, 52, 11092−11095. (129) Scharf, D. H.; Chankhamjon, P.; Scherlach, K.; Heinekamp, T.; Roth, M.; Brakhage, A. A.; Hertweck, C. Epidithiol Formation by an Unprecedented Twin Carbon-Sulfur Lyase in the Gliotoxin Pathway. Angew. Chem., Int. Ed. 2012, 51, 10064−10068. (130) Armstrong, R. N. Enzyme-Catalyzed Detoxication Reactions: Mechanisms and Stereochemistry. CRC Crit. Rev. Biochem. 1987, 22, 39−88. (131) Bones, A. M.; Rossiter, J. T. The Enzymic and Chemically Induced Decomposition of Glucosinolates. Phytochemistry 2006, 67, 1053−1067. (132) Fahey, J. W.; Zalcmann, A. T.; Talalay, P. The Chemical Diversity and Distribution of Glucosinolates and Isothiocyanates among Plants. Phytochemistry 2001, 56, 5−51. (133) Chen, S.; Glawischnig, E.; Jorgensen, K.; Naur, P.; Jorgensen, B.; Olsen, C. E.; Hansen, C. H.; Rasmussen, H.; Pickett, J. A.; Halkier, B. A. Cyp79f1 and Cyp79f2 Have Distinct Functions in the Biosynthesis of Aliphatic Glucosinolates in Arabidopsis. Plant J. 2003, 33, 923−937. (134) Hansen, C. H.; Wittstock, U.; Olsen, C. E.; Hick, A. J.; Pickett, J. A.; Halkier, B. A. Cytochrome P450 Cyp79f1 from Arabidopsis Catalyzes the Conversion of Dihomomethionine and Trihomomethionine to the Corresponding Aldoximes in the Biosynthesis of Aliphatic Glucosinolates. J. Biol. Chem. 2001, 276, 11078−11085. (135) Hull, A. K.; Vij, R.; Celenza, J. L. Arabidopsis Cytochrome P450s That Catalyze the First Step of Tryptophan-Dependent Indole3-Acetic Acid Biosynthesis. Proc. Natl. Acad. Sci. U. S. A. 2000, 97, 2379−2384. (136) Mikkelsen, M. D.; Hansen, C. H.; Wittstock, U.; Halkier, B. A. Cytochrome P450 Cyp79b2 from Arabidopsis Catalyzes the Conversion of Tryptophan to Indole-3-Acetaldoxime, a Precursor of Indole Glucosinolates and Indole-3-Acetic Acid. J. Biol. Chem. 2000, 275, 33712−33717. (137) Wittstock, U.; Halkier, B. A. Cytochrome P450 Cyp79a2 from Arabidopsis Thaliana L. Catalyzes the Conversion of L-Phenylalanine to Phenylacetaldoxime in the Biosynthesis of Benzylglucosinolate. J. Biol. Chem. 2000, 275, 14659−14666. (138) Bak, S.; Feyereisen, R. The Involvement of Two P450 Enzymes, Cyp83b1 and Cyp83a1, in Auxin Homeostasis and Glucosinolate Biosynthesis. Plant Physiol. 2001, 127, 108−118. (139) Bak, S.; Tax, F. E.; Feldmann, K. A.; Galbraith, D. W.; Feyereisen, R. Cyp83b1, a Cytochrome P450 at the Metabolic Branch Point in Auxin and Indole Glucosinolate Biosynthesis in Arabidopsis. Plant Cell 2001, 13, 101−111. (140) Hansen, C. H.; Du, L.; Naur, P.; Olsen, C. E.; Axelsen, K. B.; Hick, A. J.; Pickett, J. A.; Halkier, B. A. Cyp83b1 Is the OximeMetabolizing Enzyme in the Glucosinolate Pathway in Arabidopsis. J. Biol. Chem. 2001, 276, 24790−24796. (141) Hemm, M. R.; Ruegger, M. O.; Chapple, C. The Arabidopsis Ref2Mutant Is Defective in the Gene Encoding Cyp83a1 and Shows Both Phenylpropanoid and Glucosinolate Phenotypes. Plant Cell 2003, 15, 179−194.

(142) Naur, P.; Petersen, B. L.; Mikkelsen, M. D.; Bak, S.; Rasmussen, H.; Olsen, C. E.; Halkier, B. A. Cyp83a1 and Cyp83b1, Two Nonredundant Cytochrome P450 Enzymes Metabolizing Oximes in the Biosynthesis of Glucosinolates in Arabidopsis. Plant Physiol. 2003, 133, 63−72. (143) Wetter, L. R.; Chisholm, M. D. Sources of Sulfur in the Thioglucosides of Various Higher Plants. Can. J. Biochem. 1968, 46, 931−935. (144) Bednarek, P.; Pislewska-Bednarek, M.; Svatos, A.; Schneider, B.; Doubsky, J.; Mansurova, M.; Humphry, M.; Consonni, C.; Panstruga, R.; Sanchez-Vallet, A.; et al. A Glucosinolate Metabolism Pathway in Living Plant Cells Mediates Broad-Spectrum Antifungal Defense. Science 2009, 323, 101−106. (145) Schlaeppi, K.; Bodenhausen, N.; Buchala, A.; Mauch, F.; Reymond, P. The Glutathione-Deficient Mutant Pad2−1 Accumulates Lower Amounts of Glucosinolates and Is More Susceptible to the Insect Herbivore Spodoptera Littoralis. Plant J. 2008, 55, 774−786. (146) Hirai, M. Y.; Klein, M.; Fujikawa, Y.; Yano, M.; Goodenowe, D. B.; Yamazaki, Y.; Kanaya, S.; Nakamura, Y.; Kitayama, M.; Suzuki, H.; et al. Elucidation of Gene-to-Gene and Metabolite-to-Gene Networks in Arabidopsis by Integration of Metabolomics and Transcriptomics. J. Biol. Chem. 2005, 280, 25590−25595. (147) Geu-Flores, F.; Nielsen, M. T.; Nafisi, M.; Moldrup, M. E.; Olsen, C. E.; Motawia, M. S.; Halkier, B. A. Glucosinolate Engineering Identifies a Gamma-Glutamyl Peptidase. Nat. Chem. Biol. 2009, 5, 575−577. (148) Geu-Flores, F.; Moldrup, M. E.; Bottcher, C.; Olsen, C. E.; Scheel, D.; Halkier, B. A. Cytosolic Gamma-Glutamyl Peptidases Process Glutathione Conjugates in the Biosynthesis of Glucosinolates and Camalexin in Arabidopsis. Plant Cell 2011, 23, 2456−2469. (149) Piotrowski, M.; Schemenewitz, A.; Lopukhina, A.; Muller, A.; Janowitz, T.; Weiler, E. W.; Oecking, C. Desulfoglucosinolate Sulfotransferases from Arabidopsis Thaliana Catalyze the Final Step in the Biosynthesis of the Glucosinolate Core Structure. J. Biol. Chem. 2004, 279, 50717−50725. (150) Borlinghaus, J.; Albrecht, F.; Gruhlke, M. C.; Nwachukwu, I. D.; Slusarenko, A. J. Allicin: Chemistry and Biological Properties. Molecules 2014, 19, 12591−12618. (151) Jacob, C. A Scent of Therapy: Pharmacological Implications of Natural Products Containing Redox-Active Sulfur Atoms. Nat. Prod. Rep. 2006, 23, 851−863. (152) Slusarenko, A. J.; Patel, A.; Portz, D. Control of Plant Diseases by Natural Products: Allicin from Garlic as a Case Study. Eur. J. Plant Pathol. 2008, 121, 313−322. (153) Parry, R. J.; Sood, G. R. Investigations of the Biosynthesis of Trans-(+)-S-1-Propenyl-L-Cysteine Sulfoxide in Onions (Allium Cepa). J. Am. Chem. Soc. 1989, 111, 4514−4515. (154) Rabimkov, A.; Zhu, X. Z.; Grafi, G.; Galili, G.; Mirelman, D. Alliin Lyase (Alliinase) from Garlic (Allium Sativum). Biochemical Characterization and Cdna Cloning. Appl. Biochem. Biotechnol. 1994, 48, 149−171. (155) Lancaster, J. E.; Shaw, M. L. Gamma-Glutamyl-Transferase Peptides in the Biosynthesis of S-Alk(En)Yl-L-Cysteine Sulfoxides (Flavor Precursors) in Allium. Phytochemistry 1989, 28, 455−460. (156) Suzuki, T.; Sugii, M.; Kakimoto, T. Metabolic Incorporation of L-Valine-[14c] into S-(2-Carboxypropyl) Glutathione and S-(2Carboxypropyl) Cysteine in Garlic. Chem. Pharm. Bull. 1962, 10, 328−331. (157) Turnbull, A.; Galpin, I. J.; Collin, H. A. Comparison of the Onion Plant (Allium Cepa) and Onion Tissue-Culture 0.3. Feeding of C-14-Labeled Precursors of the Flavor Precursor Compounds. New Phytol. 1980, 85, 483−487. (158) Granroth, B. Biosynthesis and Decomposition of Cysteine Derivatives in Onion and Other Allium Species. Ann. Acad. Sci. Fenn., Chem. 1970, 154, 1−71. (159) Jones, M. G.; Hughes, J.; Tregova, A.; Milne, J.; Tomsett, A. B.; Collin, H. A. Biosynthesis of the Flavour Precursors of Onion and Garlic. J. Exp. Bot. 2004, 55, 1903−1918. 5568

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

(160) Yoshimoto, N.; Yabe, A.; Sugino, Y.; Murakami, S.; Sai-Ngam, N.; Sumi, S.; Tsuneyoshi, T.; Saito, K. Garlic Gamma-Glutamyl Transpeptidases That Catalyze Deglutamylation of Biosynthetic Intermediate of Alliin. Front. Plant Sci. 2015, 5.10.3389/ fpls.2014.00758 (161) Yoshimoto, N.; Onuma, M.; Mizuno, S.; Sugino, Y.; Nakabayashi, R.; Imai, S.; Tsuneyoshi, T.; Sumi, S.; Saito, K. Identification of a Flavin-Containing S-Oxygenating Monooxygenase Involved in Alliin Biosynthesis in Garlic. Plant J. 2015, 83, 941−951. (162) Maiya, S.; Grundmann, A.; Li, X.; Li, S. M.; Turner, G. Identification of a Hybrid Pks/Nrps Required for Pseurotin a Biosynthesis in the Human Pathogen Aspergillus Fumigatus. ChemBioChem 2007, 8, 1736−1743. (163) Zou, Y.; Xu, W.; Tsunematsu, Y.; Tang, M.; Watanabe, K.; Tang, Y. Methylation-Dependent Acyl Transfer between Polyketide Synthase and Nonribosomal Peptide Synthetase Modules in Fungal Natural Product Biosynthesis. Org. Lett. 2014, 16, 6390−6393. (164) Yamamoto, T.; Tsunematsu, Y.; Hara, K.; Suzuki, T.; Kishimoto, S.; Kawagishi, H.; Noguchi, H.; Hashimoto, H.; Tang, Y.; Hotta, K.; et al. Oxidative Trans to Cis Isomerization of Olefins in Polyketide Biosynthesis. Angew. Chem., Int. Ed. 2016, 55, 6207−6210. (165) Tsunematsu, Y.; Fukutomi, M.; Saruwatari, T.; Noguchi, H.; Hotta, K.; Tang, Y.; Watanabe, K. Elucidation of Pseurotin Biosynthetic Pathway Points to Trans-Acting C-Methyltransferase: Generation of Chemical Diversity. Angew. Chem., Int. Ed. 2014, 53, 8475−8479. (166) Reeves, C. D.; Hu, Z.; Reid, R.; Kealey, J. T. Genes for the Biosynthesis of the Fungal Polyketides Hypothemycin from Hypomyces Subiculosus and Radicicol from Pochonia Chlamydosporia. Appl. Environ. Microbiol. 2008, 74, 5121−5129. (167) Geng, H. F.; Bruhn, J. B.; Nielsen, K. F.; Gram, L.; Belas, R. Genetic Dissection of Tropodithietic Acid Biosynthesis by Marine Roseobacters. Appl. Environ. Microbiol. 2008, 74, 1535−1545. (168) Berger, M.; Brock, N. L.; Liesegang, H.; Dogs, M.; Preuth, I.; Simon, M.; Dickschat, J. S.; Brinkhoff, T. Genetic Analysis of the Upper Phenylacetate Catabolic Pathway in the Production of Tropodithietic Acid by Phaeobacter Gallaeciensis. Appl. Environ. Microbiol. 2012, 78, 3539−3551. (169) Teufel, R.; Gantert, C.; Voss, M.; Eisenreich, W.; Haehnel, W.; Fuchs, G. Studies on the Mechanism of Ring Hydrolysis in Phenylacetate Degradation - a Metabolic Branching Point. J. Biol. Chem. 2011, 286, 11021−11034. (170) Brock, N. L.; Nikolay, A.; Dickschat, J. S. Biosynthesis of the Antibiotic Tropodithietic Acid by the Marine Bacterium Phaeobacter Inhibens. Chem. Commun. 2014, 50, 5487−5489. (171) Seyedsayamdost, M. R.; Wang, R.; Kolter, R.; Clardy, J. Hybrid Biosynthesis of Roseobacticides from Algal and Bacterial Precursor Molecules. J. Am. Chem. Soc. 2014, 136, 15150−15153. (172) Wang, R.; Gallant, E.; Seyedsayamdost, M. R. Investigation of the Genetics and Biochemistry of Roseobacticide Production in the Roseobacter Clade Bacterium Phaeobacter Inhibens. mBio 2016, 7, e02118-15. (173) Kahan, F. M.; Kropp, H.; Sundelof, J. G.; Birnbaum, J. Thienamycin: Development of Imipenen-Cilastatin. J. Antimicrob. Chemother. 1983, 12, 1−35. (174) Papp-Wallace, K. M.; Endimiani, A.; Taracila, M. A.; Bonomo, R. A. Carbapenems: Past, Present, and Future. Antimicrob. Agents Chemother. 2011, 55, 4943−4960. (175) Williamson, J. M.; Inamine, E.; Wilson, K. E.; Douglas, A. W.; Liesch, J. M.; Albers-Schönberg, G. Biosynthesis of the Beta -Lactam Antibiotic, Thienamycin, by Streptomyces Cattleya. J. Biol. Chem. 1985, 260, 4637−4647. (176) Sleeman, M. C.; Schofield, C. J. Carboxymethylproline Synthase (Carb), an Unusual Carbon-Carbon Bond-Forming Enzyme of the Crotonase Superfamily Involved in Carbapenem Biosynthesis. J. Biol. Chem. 2004, 279, 6730−6736. (177) Hamed, R. B.; Batchelar, E. T.; Mecinovic, J.; Claridge, T. D.; Schofield, C. J. Evidence That Thienamycin Biosynthesis Proceeds Via

C-5 Epimerization: Thne Catalyzes the Formation of (2s,5s)-TransCarboxymethylproline. ChemBioChem 2009, 10, 246−250. (178) Bodner, M. J.; Li, R.; Phelan, R. M.; Freeman, M. F.; Moshos, K. A.; Lloyd, E. P.; Townsend, C. A. Definition of the Common and Divergent Steps in Carbapenem B-Lactam Antibiotic Biosynthesis. ChemBioChem 2011, 12, 2159−2165. (179) Li, R.; Stapon, A.; Blanchfield, J. T.; Townsend, C. A. Three Unusual Reactions Mediate Carbapenem and Carbapenam Biosynthesis. J. Am. Chem. Soc. 2000, 122, 9296−9297. (180) Gerratana, B.; Stapon, A.; Townsend, C. A. Inhibition and Alternate Substrate Studies on the Mechanism of Carbapenam Synthetase from Erwinia Carotovora. Biochemistry 2003, 42, 7836− 7847. (181) Phelan, R. M.; Townsend, C. A. Mechanistic Insights into the Bifunctional Non-Heme Iron Oxygenase Carbapenem Synthase by Active Site Saturation Mutagenesis. J. Am. Chem. Soc. 2013, 135, 7496−7502. (182) Stapon, A.; Li, R.; Townsend, C. A. Synthesis of (3s,5r)Carbapenam-3-Carboxylic Acid and Its Role in Carbapenem Biosynthesis and the Stereoinversion Problem. J. Am. Chem. Soc. 2003, 125, 15746−15747. (183) Houck, D. R.; Kobayashi, K.; Williamson, J. M.; Floss, H. G. Stereochemistry of Methylation in Thienamycin Biosynthesis: Example of a Methyl Transfer from Methionine with Retention of Configuration. J. Am. Chem. Soc. 1986, 108, 5365−5366. (184) Marous, D. R.; Lloyd, E. P.; Buller, A. R.; Moshos, K. A.; Grove, T. L.; Blaszczyk, A. J.; Booker, S. J.; Townsend, C. A. Consecutive Radical S-Adenosylmethionine Methylations Form the Ethyl Side Chain in Thienamycin Biosynthesis. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, 10354−10358. (185) Bodner, M. J.; Phelan, R. M.; Freeman, M. F.; Li, R.; Townsend, C. A. Non-Heme Iron Oxygenases Generate Natural Structural Diversity in Carbapenem Antibiotics. J. Am. Chem. Soc. 2010, 132, 12−13. (186) Li, R.; Lloyd, E. P.; Moshos, K. A.; Townsend, C. A. Identification and Characterization of the Carbapenem Mm 4550 and Its Gene Cluster in Streptomyces Argenteolus Atcc 11009. ChemBioChem 2014, 15, 320−331. (187) Freeman, M. F.; Moshos, K. A.; Bodner, M. J.; Li, R. F.; Townsend, C. A. Four Enzymes Define the Incorporation of Coenzyme a in Thienamycin Biosynthesis. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 11128−11133. (188) Yoshioka, T.; Kojima, I.; Isshiki, K.; Watanabe, A.; Shimauchi, Y.; Okabe, M.; Fukagawa, Y.; Ishikura, T. Structures of Oa-6129a, B1, B2 and C, New Carbapenem Antibiotics Produced by Streptomyces Sp. Oa-6129. J. Antibiot. 1983, 36, 1473−1482. (189) Fukagawa, Y.; Okabe, M.; Azuma, S.; Kojima, I.; Ishikura, T.; Kubo, K. Studies on the Biosynthesis of Carbapenem Antibiotics. I. Biosynthetic Significance of the Oa-6129 Group of Carbapenem Compounds as the Direct Precursors for Ps-5, Epithienamycins a and C and Mm 17880. J. Antibiot. 1984, 37, 1388−1393. (190) Kubo, K.; Ishikura, T.; Fukagawa, Y. Studies on the Biosynthesis of Carbapenem Antibiotics. Ii. Isolation and Functions of a Specific Acylase Involved in the Depantothenylation of the Oa6129 Compounds. J. Antibiot. 1984, 37, 1394−1402. (191) Kubo, K.; Ishikura, T.; Fukagawa, Y. Studies on the Biosynthesis of Carbapenem Antibiotics. Iii. Enzymological Characterization of the L-Amino Acid Acylase Activity of A933 Acylase. J. Antibiot. 1985, 38, 622−630. (192) Kojima, I.; Fukagawa, Y.; Okabe, M.; Ishikura, T.; Shibamoto, N. Mutagenesis of Oa-6129 Carbapenem-Producing Blocked Mutants and the Biosynthesis of Carbapenems. J. Antibiot. 1988, 41, 899−907. (193) Buller, A. R.; Freeman, M. F.; Wright, N. T.; Schildbach, J. F.; Townsend, C. A. Insights into Cisautoproteolysis Reveal a Reactive State Formed through Conformational Rearrangement. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 2308−2313. (194) Buller, A. R.; Labonte, J. W.; Freeman, M. F.; Wright, N. T.; Schildbach, J. F.; Townsend, C. A. Autoproteolytic Activation of Thnt 5569

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

(213) Heine, D.; Bretschneider, T.; Sundaram, S.; Hertweck, C. Enzymatic Polyketide Chain Branching to Give Substituted Lactone, Lactam, and Glutarimide Heterocycles. Angew. Chem., Int. Ed. 2014, 53, 11645−11649. (214) Sundaram, S.; Heine, D.; Hertweck, C. Polyketide Synthase Chimeras Reveal Key Role of Ketosynthase Domain in Chain Branching. Nat. Chem. Biol. 2015, 11, 949−951. (215) Pöplau, P.; Frank, S.; Morinaka, B. I.; Piel, J. An Enzymatic Domain for the Formation of Cyclic Ethers in Complex Polyketides. Angew. Chem., Int. Ed. 2013, 52, 13215−13218. (216) Berkhan, G.; Hahn, F. A Dehydratase Domain in Ambruticin Biosynthesis Displays Additional Activity as a Pyran-Forming Cyclase. Angew. Chem., Int. Ed. 2014, 53, 14240−14244. (217) Luhavaya, H.; Dias, M. V.; Williams, S. R.; Hong, H.; de Oliveira, L. G.; Leadlay, P. F. Enzymology of Pyran Ring a Formation in Salinomycin Biosynthesis. Angew. Chem., Int. Ed. 2015, 54, 13622− 13625. (218) Huang, S. X.; Yun, B. S.; Ma, M.; Basu, H. S.; Church, D. R.; Ingenhorst, G.; Huang, Y.; Yang, D.; Lohman, J. R.; Tang, G. L.; et al. Leinamycin E1 Acting as an Anticancer Prodrug Activated by Reactive Oxygen Species. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, 8278−8283. (219) Knerr, P. J.; van der Donk, W. A. Discovery, Biosynthesis, and Engineering of Lantipeptides. Annu. Rev. Biochem. 2012, 81, 479−505. (220) 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. (221) Kupke, T.; Kempter, C.; Gnau, V.; Jung, G.; Gotz, F. Mass Spectroscopic Analysis of a Novel Enzymatic Reaction. Oxidative Decarboxylation of the Lantibiotic Precursor Peptide Epia Catalyzed by the Flavoprotein Epid. J. Biol. Chem. 1994, 269, 5653−5659. (222) Majer, F.; Schmid, D. G.; Altena, K.; Bierbaum, G.; Kupke, T. The Flavoprotein Mrsd Catalyzes the Oxidative Decarboxylation Reaction Involved in Formation of the Peptidoglycan Biosynthesis Inhibitor Mersacidin. J. Bacteriol. 2002, 184, 1234−1243. (223) Ortega, M. A.; Cogan, D. P.; Mukherjee, S.; Garg, N.; Li, B.; Thibodeaux, G. N.; Maffioli, S. I.; Donadio, S.; Sosio, M.; Escano, J.; et al. Two Flavoenzymes Catalyze the Post-Translational Generation of 5-Chlorotryptophan and 2-Aminovinyl-Cysteine During Nai-107 Biosynthesis. ACS Chem. Biol. 2017, 12, 548−557. (224) Li, B.; Yu, J. P. J.; Brunzelle, J. S.; Moll, G. N.; van der Donk, W. A.; Nair, S. K. Structure and Mechanism of the Lantibiotic Cyclase Involved in Nisin Biosynthesis. Science 2006, 311, 1464−1467. (225) Li, B.; van der Donk, W. A. Identification of Essential Catalytic Residues of the Cyclase Nisc Involved in the Biosynthesis of Nisin. J. Biol. Chem. 2007, 282, 21169−21175. (226) Tang, W.; van der Donk, W. A. Structural Characterization of Four Prochlorosins: A Novel Class of Lantipeptides Produced by Planktonic Marine Cyanobacteria. Biochemistry 2012, 51, 4271−4279. (227) Okeley, N. M.; Paul, M.; Stasser, J. P.; Blackburn, N.; van der Donk, W. A. Spac and Nisc, the Cyclases Involved in Subtilin and Nisin Biosynthesis, Are Zinc Proteins. Biochemistry 2003, 42, 13613− 13624. (228) Helfrich, M.; Entian, K. D.; Stein, T. Structure-Function Relationships of the Lanthionine Cyclase Spac Involved in Biosynthesis of the Bacillus Subtilis Peptide Antibiotic Subtilin. Biochemistry 2007, 46, 3224−3233. (229) Wang, H.; van der Donk, W. A. Biosynthesis of the Class Iii Lantipeptide Catenulipeptin. ACS Chem. Biol. 2012, 7, 1529−1535. (230) Muller, W. M.; Schmiederer, T.; Ensle, P.; Sussmuth, R. D. In Vitro Biosynthesis of the Prepeptide of Type-Iii Lantibiotic Labyrinthopeptin A2 Including Formation of a C-C Bond as a PostTranslational Modification. Angew. Chem., Int. Ed. 2010, 49, 2436− 2440. (231) Meindl, K.; Schmiederer, T.; Schneider, K.; Reicke, A.; Butz, D.; Keller, S.; Guhring, H.; Vertesy, L.; Wink, J.; Hoffmann, H.; et al. Labyrinthopeptins: A New Class of Carbacyclic Lantibiotics. Angew. Chem., Int. Ed. 2010, 49, 1151−1154.

Results in Structural Reorganization Necessary for Substrate Binding and Catalysis. J. Mol. Biol. 2012, 422, 508−518. (195) Rodríguez, M.; Núñez, L. E.; Braña, A. F.; Méndez, C.; Salas, J. A.; Blanco, G. Mutational Analysis of the Thienamycin Biosynthetic Gene Cluster from Streptomyces Cattleya. Antimicrob. Agents Chemother. 2011, 55, 1638−1649. (196) Nunez, L. E.; Mendez, C.; Brana, A. F.; Blanco, G.; Salas, J. A. The Biosynthetic Gene Cluster for the Beta-Lactam Carbapenem Thienamycin in Streptomyces Cattleya. Chem. Biol. 2003, 10, 301−311. (197) Blanco, G. Comparative Analysis of a Cryptic ThienamycinLike Gene Cluster Identified in Streptomyces Flavogriseus by Genome Mining. Arch. Microbiol. 2012, 194, 549−555. (198) Suzuki, H.; Ohnishi, Y.; Horinouchi, S. Gric and Grid Constitute a Carboxylic Acid Reductase Involved in Grixazone Biosynthesis in Streptomyces Griseus. J. Antibiot. 2007, 60, 380−387. (199) Kara, M.; Asano, K.; Kawamoto, I.; Takiouchi, T.; Katsumata, S.; Takahashi, K. I.; Nakano, H. Leinamycin, a New Antitumor Antibiotic from Streptomyces - Producing Organism, Fermentation and Isolation. J. Antibiot. 1989, 42, 1768−1774. (200) Asai, A.; Hara, M.; Kakita, S.; Kanda, Y.; Yoshida, M.; Saito, H.; Saitoh, Y. Thiol-Mediated DNA Alkylation by the Novel Antitumor Antibiotic Leinamycin. J. Am. Chem. Soc. 1996, 118, 6802−6803. (201) Fekry, M. I.; Szekely, J.; Dutta, S.; Breydo, L.; Zang, H.; Gates, K. S. Noncovalent DNA Binding Drives DNA Alkylation by Leinamycin: Evidence That the Z,E-5-(Thiazol-4-Yl)-Penta-2,4-Dienone Moiety of the Natural Product Serves as an Atypical DNA Intercalator. J. Am. Chem. Soc. 2011, 133, 17641−17651. (202) Cheng, Y. Q.; Tang, G. L.; Shen, B. Identification and Localization of the Gene Cluster Encoding Biosynthesis of the Antitumor Macrolactam Leinamycin in Streptomyces Atroolivaceus S140. J. Bacteriol. 2002, 184, 7013−7024. (203) Cheng, Y. Q.; Tang, G. L.; Shen, B. Type I Polyketide Synthase Requiring a Discrete Acyltransferase for Polyketide Biosynthesis. Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 3149−3154. (204) Tang, G. L.; Cheng, Y. Q.; Shen, B. Leinamycin Biosynthesis Revealing Unprecedented Architectural Complexity for a Hybrid Polyketide Synthase and Nonribosomal Peptide Synthetase. Chem. Biol. 2004, 11, 33−45. (205) Huang, Y.; Huang, S. X.; Ju, J. H.; Tang, G. L.; Liu, T.; Shen, B. Characterization of the Inmklm Genes Unveiling Key Intermediates for Beta-Alkylation in Leinamycin Biosynthesis. Org. Lett. 2011, 13, 498−501. (206) Liu, T.; Huang, Y.; Shen, B. Bifunctional Acyltransferase/ Decarboxylase Lnmk as the Missing Link for Beta-Alkylation in Polyketide Biosynthesis. J. Am. Chem. Soc. 2009, 131, 6900−6901. (207) Lohman, J. R.; Bingman, C. A.; Phillips, G. N.; Shen, B. Structure of the Bifunctional Acyltransferase/Decarboxylase Lnmk from the Leinamycin Biosynthetic Pathway Revealing Novel Activity for a Double-Hot-Dog Fold. Biochemistry 2013, 52, 902−911. (208) Tang, G. L.; Cheng, Y. Q.; Shen, B. Polyketide Chain Skipping Mechanism in the Biosynthesis of the Hybrid Nonribosomal PeptidePolyketide Antitumor Antibiotic Leinamycin in Streptomyces Atroolivaceus S-140. J. Nat. Prod. 2006, 69, 387−393. (209) Tang, G. L.; Cheng, Y. Q.; Shen, B. Chain Initiation in the Leinamycin-Producing Hybrid Nonribosomal Peptide/Polyketide Synthetase from Streptomyces Atroolivaceus S-140 - Discrete, Monofunctional Adenylation Enzyme and Peptidyl Carrier Protein That Directly Load D-Alanine. J. Biol. Chem. 2007, 282, 20273−20282. (210) Ma, M.; Lohman, J. R.; Liu, T.; Shen, B. C-S Bond Cleavage by a Polyketide Synthase Domain. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, 10359−10364. (211) Sundaram, S.; Hertweck, C. On-Line Enzymatic Tailoring of Polyketides and Peptides in Thiotemplate Systems. Curr. Opin. Chem. Biol. 2016, 31, 82−94. (212) Bretschneider, T.; Heim, J. B.; Heine, D.; Winkler, R.; Busch, B.; Kusebauch, B.; Stehle, T.; Zocher, G.; Hertweck, C. Vinylogous Chain Branching Catalysed by a Dedicated Polyketide Synthase Module. Nature 2013, 502, 124−128. 5570

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

(232) Zhang, Q.; Yu, Y.; Velasquez, J. E.; van der Donk, W. A. Evolution of Lanthipeptide Synthetases. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 18361−18366. (233) Tang, W.; Jimenez-Oses, G.; Houk, K. N.; van der Donk, W. A. Substrate Control in Stereoselective Lanthionine Biosynthesis. Nat. Chem. 2015, 7, 57−64. (234) Ding, W.; Li, Y.; Zhang, Q. Substrate-Controlled Stereochemistry in Natural Product Biosynthesis. ACS Chem. Biol. 2015, 10, 1590−1598. (235) Benjdia, A.; Guillot, A.; Lefranc, B.; Vaudry, H.; Leprince, J.; Berteau, O. Thioether Bond Formation by Spasm Domain Radical Sam Enzymes: Calpha H-Atom Abstraction in Subtilosin a Biosynthesis. Chem. Commun. 2016, 52, 6249−6252. (236) Bruender, N. A.; Bandarian, V. Skfb Abstracts a Hydrogen Atom from Calpha on Skfa to Initiate Thioether Cross-Link Formation. Biochemistry 2016, 55, 4131−4134. (237) Asaduzzaman, S. M.; Sonomoto, K. Lantibiotics: Diverse Activities and Unique Modes of Action. J. Biosci. Bioeng. 2009, 107, 475−487. (238) Claesen, J.; Bibb, M. Genome Mining and Genetic Analysis of Cypemycin Biosynthesis Reveal an Unusual Class of Posttranslationally Modified Peptides. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 16297−16302. (239) Roy, R. S.; Gehring, A. M.; Milne, J. C.; Belshaw, P. J.; Walsh, C. T. Thiazole and Oxazole Peptides: Biosynthesis and Molecular Machinery. Nat. Prod. Rep. 1999, 16, 249−263. (240) Hider, R. C.; Kong, X. Chemistry and Biology of Siderophores. Nat. Prod. Rep. 2010, 27, 637−657. (241) Crosa, J. H.; Walsh, C. T. Genetics and Assembly Line Enzymology of Siderophore Biosynthesis in Bacteria. Microbiol. Mol. Biol. Rev. 2002, 66, 223−249. (242) Cox, C. D.; Rinehart, K. L.; Moore, M. L.; Cook, J. C. Pyochelin: Novel Structure of an Iron-Chelating Growth Promoter for Pseudomonas Aeruginosa. Proc. Natl. Acad. Sci. U. S. A. 1981, 78, 4256− 4260. (243) Drechsel, H.; Stephan, H.; Lotz, R.; Haag, H.; Zähner, H.; Hantke, K.; Jung, G. Structure Elucidation of Yersiniabactin, a Siderophore from Highly Virulent. Yersinia Strains. Liebigs Ann. Chem. 1995, 1995, 1727−1733. (244) Takita, T.; Muraoka, Y.; Yoshioka, T.; Fujii, A.; Maeda, K.; Umezawa, H. The Chemistry of Bleomycin. J. Antibiot. 1972, 25, 755− 758. (245) Umezawa, H.; Maeda, K.; Takeuchi, T.; Okami, Y. New Antibiotics, Bleomycin a and B. J. Antibiot. 1966, 19, 200−209. (246) Bollag, D. M.; McQueney, P. A.; Zhu, J.; Hensens, O.; Koupal, L.; Liesch, J.; Goetz, M.; Lazarides, E.; Woods, C. M. Epothilones, a New Class of Microtubule-Stabilizing Agents with a Taxol-Like Mechanism of Action. Cancer Res. 1995, 55, 2325−2333. (247) Hur, G. H.; Vickery, C. R.; Burkart, M. D. Explorations of Catalytic Domains in Non-Ribosomal Peptide Synthetase Enzymology. Nat. Prod. Rep. 2012, 29, 1074−1098. (248) Di Lorenzo, M.; Stork, M.; Naka, H.; Tolmasky, M. E.; Crosa, J. H. Tandem Heterocyclization Domains in a Nonribosomal Peptide Synthetase Essential for Siderophore Biosynthesis in Vibrio Anguillarum. BioMetals 2008, 21, 635−648. (249) Gehring, A. M.; Mori, I.; Perry, R. D.; Walsh, C. T. The Nonribosomal Peptide Synthetase Hmwp2 Forms a Thiazoline Ring During Biogenesis of Yersiniabactin, an Iron-Chelating Virulence Factor of Yersinia Pestis. Biochemistry 1998, 37, 11637−11650. (250) Chen, H.; O’Connor, S.; Cane, D. E.; Walsh, C. T. Epothilone Biosynthesis: Assembly of the Methylthiazolylcarboxy Starter Unit on the Epob Subunit. Chem. Biol. 2001, 8, 899−912. (251) Quadri, L. E. N.; Keating, T. A.; Patel, H. M.; Walsh, C. T. Assembly of the Pseudomonas Aeruginosa Nonribosomal Peptide Siderophore Pyochelin: In Vitro Reconstitution of Aryl-4, 2-Bisthiazoline Synthetase Activity from Pchd, Pche, and Pchf. Biochemistry 1999, 38, 14941−14954. (252) Marshall, C. G.; Burkart, M. D.; Keating, T. A.; Walsh, C. T. Heterocycle Formation in Vibriobactin Biosynthesis: Alternative

Substrate Utilization and Identification of a Condensed Intermediate. Biochemistry 2001, 40, 10655−10663. (253) Marshall, C. G.; Hillson, N. J.; Walsh, C. T. Catalytic Mapping of the Vibriobactin Biosynthetic Enzyme Vibf. Biochemistry 2002, 41, 244−250. (254) Miller, D. A.; Walsh, C. T. Yersiniabactin Synthetase: Probing the Recognition of Carrier Protein Domains by the Catalytic Heterocyclization Domains, Cy1 and Cy2, in the Chain-Initiating Hmwp2 Subunit. Biochemistry 2001, 40, 5313−5321. (255) Duerfahrt, T.; Eppelmann, K.; Müller, R.; Marahiel, M. A. Rational Design of a Bimodular Model System for the Investigation of Heterocyclization in Nonribosomal Peptide Biosynthesis. Chem. Biol. 2004, 11, 261−271. (256) Bloudoff, K.; Fage, C. D.; Marahiel, M. A.; Schmeing, T. M. Structural and Mutational Analysis of the Nonribosomal Peptide Synthetase Heterocyclization Domain Provides Insight into Catalysis. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 95−100. (257) Dowling, D. P.; Kung, Y.; Croft, A. K.; Taghizadeh, K.; Kelly, W. L.; Walsh, C. T.; Drennan, C. L. Structural Elements of an Nrps Cyclization Domain and Its Intermodule Docking Domain. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 12432−12437. (258) Schneider, T. L.; Shen, B.; Walsh, C. T. Oxidase Domains in Epothilone and Bleomycin Biosynthesis: Thiazoline to Thiazole Oxidation During Chain Elongation. Biochemistry 2003, 42, 9722− 9730. (259) Miller, D. A.; Luo, L.; Hillson, N.; Keating, T. A.; Walsh, C. T. Yersiniabactin Synthetase: A Four-Protein Assembly Line Producing the Nonribosomal Peptide/Polyketide Hybrid Siderophore of Yersinia Pestis. Chem. Biol. 2002, 9, 333−344. (260) Reimmann, C.; Patel, H. M.; Serino, L.; Barone, M.; Walsh, C. T.; Haas, D. Essential Pchg-Dependent Reduction in Pyochelin Biosynthesis of Pseudomonas Aeruginosa. J. Bacteriol. 2001, 183, 813− 820. (261) Tao, W.; Yurkovich, M. E.; Wen, S.; Lebe, K. E.; Samborskyy, M.; Liu, Y.; Yang, A.; Liu, Y.; Ju, Y.; Deng, Z.; et al. A Genomics-Led Approach to Deciphering the Mechanism of Thiotetronate Antibiotic Biosynthesis. Chem. Sci. 2016, 7, 376−385. (262) Cox, C. L.; Doroghazi, J. R.; Mitchell, D. A. The Genomic Landscape of Ribosomal Peptides Containing Thiazole and Oxazole Heterocycles. BMC Genomics 2015, 16, 778. (263) Strader, M. B.; Costantino, N.; Elkins, C. A.; Chen, C. Y.; Patel, I.; Makusky, A. J.; Choy, J. S.; Court, D. L.; Markey, S. P.; Kowalak, J. A. A Proteomic and Transcriptomic Approach Reveals New Insight into Beta-Methylthiolation of Escherichia Coli Ribosomal Protein S12. Mol. Cell. Proteomics 2011, 10, M110.005199. (264) Melby, J. O.; Nard, N. J.; Mitchell, D. A. Thiazole/OxazoleModified Microcins: Complex Natural Products from Ribosomal Templates. Curr. Opin. Chem. Biol. 2011, 15, 369−378. (265) Dunbar, K. L.; Melby, J. O.; Mitchell, D. A. Ycao Domains Use Atp to Activate Amide Backbones During Peptide Cyclodehydrations. Nat. Chem. Biol. 2012, 8, 569−575. (266) Dunbar, K. L.; Mitchell, D. A. Insights into the Mechanism of Peptide Cyclodehydrations Achieved through the Chemoenzymatic Generation of Amide Derivatives. J. Am. Chem. Soc. 2013, 135, 8692− 8701. (267) Koehnke, J.; Mann, G.; Bent, A. F.; Ludewig, H.; Shirran, S.; Botting, C.; Lebl, T.; Houssen, W. E.; Jaspars, M.; Naismith, J. H. Structural Analysis of Leader Peptide Binding Enables Leader-Free Cyanobactin Processing. Nat. Chem. Biol. 2015, 11, 558−563. (268) Koehnke, J.; Bent, A. F.; Zollman, D.; Smith, K.; Houssen, W. E.; Zhu, X.; Mann, G.; Lebl, T.; Scharff, R.; Shirran, S.; et al. The Cyanobactin Heterocyclase Enzyme: A Processive Adenylase That Operates with a Defined Order of Reaction. Angew. Chem., Int. Ed. 2013, 52, 13991−13996. (269) Crone, W. J.; Vior, N. M.; Santos-Aberturas, J.; Schmitz, L. G.; Leeper, F. J.; Truman, A. W. Dissecting Bottromycin Biosynthesis Using Comparative Untargeted Metabolomics. Angew. Chem., Int. Ed. 2016, 55, 9639−9643. 5571

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

(270) Burkhart, B. J.; Hudson, G. A.; Dunbar, K. L.; Mitchell, D. A. A Prevalent Peptide-Binding Domain Guides Ribosomal Natural Product Biosynthesis. Nat. Chem. Biol. 2015, 11, 564−570. (271) Dunbar, K. L.; Tietz, J. I.; Cox, C. L.; Burkhart, B. J.; Mitchell, D. A. Identification of an Auxiliary Leader Peptide-Binding Protein Required for Azoline Formation in Ribosomal Natural Products. J. Am. Chem. Soc. 2015, 137, 7672−7677. (272) Dunbar, K. L.; Chekan, J. R.; Cox, C. L.; Burkhart, B. J.; Nair, S. K.; Mitchell, D. A. Discovery of a New Atp-Binding Motif Involved in Peptidic Azoline Biosynthesis. Nat. Chem. Biol. 2014, 10, 823−829. (273) Sardar, D.; Schmidt, E. W. Combinatorial Biosynthesis of Ripps: Docking with Marine Life. Curr. Opin. Chem. Biol. 2016, 31, 15−21. (274) Izawa, M.; Kawasaki, T.; Hayakawa, Y. Cloning and Heterologous Expression of the Thioviridamide Biosynthesis Gene Cluster from Streptomyces Olivoviridis. Appl. Environ. Microbiol. 2013, 79, 7110−7113. (275) Hayakawa, Y.; Sasaki, K.; Nagai, K.; Shin-ya, K.; Furihata, K. Structure of Thioviridamide, a Novel Apoptosis Inducer from Streptomyces Olivoviridis. J. Antibiot. 2006, 59, 6−10. (276) Black, K. A.; Dos Santos, P. C. Shared-Intermediates in the Biosynthesis of Thio-Cofactors: Mechanism and Functions of Cysteine Desulfurases and Sulfur Acceptors. Biochim. Biophys. Acta, Mol. Cell Res. 2015, 1853, 1470−1480. (277) Iwata-Reuyl, D. An Embarrassment of Riches: The Enzymology of Rna Modification. Curr. Opin. Chem. Biol. 2008, 12, 126−133. (278) Elion, G. B. The Purine Path to Chemotherapy. Biosci. Rep. 1989, 9, 509−529. (279) Stanulla, M.; Schü nemann, H. J. Thioguanine Versus Mercaptopurine in Childhood All. Lancet 2006, 368, 1304−1306. (280) Feistner, G.; Staub, C. M. 6-Thioguanine from Erwinia Amylovora. Curr. Microbiol. 1986, 13, 95−101. (281) Coyne, S.; Chizzali, C.; Khalil, M. N.; Litomska, A.; Richter, K.; Beerhues, L.; Hertweck, C. Biosynthesis of the Antimetabolite 6Thioguanine in Erwinia Amylovora Plays a Key Role in Fire Blight Pathogenesis. Angew. Chem., Int. Ed. 2013, 52, 10564−10568. (282) Coyne, S.; Litomska, A.; Chizzali, C.; Khalil, M. N.; Richter, K.; Beerhues, L.; Hertweck, C. Control of Plant Defense Mechanisms and Fire Blight Pathogenesis through the Regulation of 6-Thioguanine Biosynthesis in Erwinia Amylovora. ChemBioChem 2014, 15, 373−376. (283) Numata, T.; Ikeuchi, Y.; Fukai, S.; Suzuki, T.; Nureki, O. Snapshots of Trna Sulphuration Via an Adenylated Intermediate. Nature 2006, 442, 419−424. (284) Wright, C. M.; Christman, G. D.; Snellinger, A. M.; Johnston, M. V.; Mueller, E. G. Direct Evidence for Enzyme Persulfide and Disulfide Intermediates During 4-Thiouridine Biosynthesis. Chem. Commun. 2006, 3104−3106. (285) Jackowski, S.; Murphy, C. M.; Cronan, J. E., Jr.; Rock, C. O. Acetoacetyl-Acyl Carrier Protein Synthase. A Target for the Antibiotic Thiolactomycin. J. Biol. Chem. 1989, 264, 7624−7629. (286) Kodali, S.; Galgoci, A.; Young, K.; Painter, R.; Silver, L. L.; Herath, K. B.; Singh, S. B.; Cully, D.; Barrett, J. F.; Schmatz, D.; et al. Determination of Selectivity and Efficacy of Fatty Acid Synthesis Inhibitors. J. Biol. Chem. 2005, 280, 1669−1677. (287) Price, A. C.; Choi, K. H.; Heath, R. J.; Li, Z.; White, S. W.; Rock, C. O. Inhibition of Beta-Ketoacyl-Acyl Carrier Protein Synthases by Thiolactomycin and Cerulenin. Structure and Mechanism. J. Biol. Chem. 2001, 276, 6551−6559. (288) Brown, M. S.; Akopiants, K.; Resceck, D. M.; McArthur, H. A.; McCormick, E.; Reynolds, K. A. Biosynthetic Origins of the Natural Product, Thiolactomycin: A Unique and Selective Inhibitor of Type Ii Dissociated Fatty Acid Synthases. J. Am. Chem. Soc. 2003, 125, 10166− 10167. (289) Tang, X.; Li, J.; Millán-Aguiñaga, N.; Zhang, J. J.; O’Neill, E. C.; Ugalde, J. A.; Jensen, P. R.; Mantovani, S. M.; Moore, B. S. Identification of Thiotetronic Acid Antibiotic Biosynthetic Pathways by Target-Directed Genome Mining. ACS Chem. Biol. 2015, 10, 2841−2849.

(290) Yurkovich, M. E.; Jenkins, R.; Sun, Y.; Tosin, M.; Leadlay, P. F. The Polyketide Backbone of Thiolactomycin Is Assembled by an Unusual Iterative Polyketide Synthase. Chem. Commun. 2017, 53, 2182−2185. (291) Hidese, R.; Mihara, H.; Esaki, N. Bacterial Cysteine Desulfurases: Versatile Key Players in Biosynthetic Pathways of Sulfur-Containing Biofactors. Appl. Microbiol. Biotechnol. 2011, 91, 47−61. (292) Webb, M. E.; Marquet, A.; Mendel, R. R.; Rebeille, F.; Smith, A. G. Elucidating Biosynthetic Pathways for Vitamins and Cofactors. Nat. Prod. Rep. 2007, 24, 988−1008. (293) Mossialos, D.; Meyer, J. M.; Budzikiewicz, H.; Wolff, U.; Koedam, N.; Baysse, C.; Anjaiah, V.; Cornelis, P. Quinolobactin, a New Siderophore of Pseudomonas Fluorescens Atcc 17400, the Production of Which Is Repressed by the Cognate Pyoverdine. Appl. Environ. Microbiol. 2000, 66, 487−492. (294) Matthijs, S.; Tehrani, K. A.; Laus, G.; Jackson, R. W.; Cooper, R. M.; Cornelis, P. Thioquinolobactin, a Pseudomonas Siderophore with Antifungal and Anti-Pythium Activity. Environ. Microbiol. 2007, 9, 425−434. (295) Matthijs, S.; Baysse, C.; Koedam, N.; Tehrani, K. A.; Verheyden, L.; Budzikiewicz, H.; Schäfer, M.; Hoorelbeke, B.; Meyer, J. M.; De Greve, H.; et al. The Pseudomonas Siderophore Quinolobactin Is Synthesized from Xanthurenic Acid, an Intermediate of the Kynurenine Pathway. Mol. Microbiol. 2004, 52, 371−384. (296) Godert, A. M.; Jin, M.; McLafferty, F. W.; Begley, T. P. Biosynthesis of the Thioquinolobactin Siderophore: An Interesting Variation on Sulfur Transfer. J. Bacteriol. 2007, 189, 2941−2944. (297) Huang, W.; Xu, H.; Li, Y.; Zhang, F.; Chen, X. Y.; He, Q. L.; Igarashi, Y.; Tang, G. L. Characterization of Yatakemycin Gene Cluster Revealing a Radical S-Adenosylmethionine Dependent Methyltransferase and Highlighting Spirocyclopropane Biosynthesis. J. Am. Chem. Soc. 2012, 134, 8831−8840. (298) Lewis, T. A.; Cortese, M. S.; Sebat, J. L.; Green, T. L.; Lee, C. H.; Crawford, R. L. A Pseudomonas Stutzeri Gene Cluster Encoding the Biosynthesis of the Ccl4-Dechlorination Agent Pyridine-2,6-Bis(Thiocarboxylic Acid). Environ. Microbiol. 2000, 2, 407−416. (299) Sasaki, E.; Ogasawara, Y.; Liu, H. W. A Biosynthetic Pathway for Be-7585a, a 2-Thiosugar-Containing Angucycline-Type Natural Product. J. Am. Chem. Soc. 2010, 132, 7405−7417. (300) Sasaki, E.; Liu, H. W. Mechanistic Studies of the Biosynthesis of 2-Thiosugar: Evidence for the Formation of an Enzyme-Bound 2Ketohexose Intermediate in Bexx-Catalyzed Reaction. J. Am. Chem. Soc. 2010, 132, 15544−15546. (301) Park, J. H.; Dorrestein, P. C.; Zhai, H.; Kinsland, C.; McLafferty, F. W.; Begley, T. P. Biosynthesis of the Thiazole Moiety of Thiamin Pyrophosphate (Vitamin B1). Biochemistry 2003, 42, 12430− 12438. (302) Begley, T. P. Cofactor Biosynthesis: An Organic Chemist’s Treasure Trove. Nat. Prod. Rep. 2006, 23, 15−25. (303) Jurgenson, C. T.; Begley, T. P.; Ealick, S. E. The Structural and Biochemical Foundations of Thiamin Biosynthesis. Annu. Rev. Biochem. 2009, 78, 569−603. (304) Sasaki, E.; Zhang, X.; Sun, H. G.; Lu, M. Y.; Liu, T. L.; Ou, A.; Li, J. Y.; Chen, Y. H.; Ealick, S. E.; Liu, H. W. Co-Opting SulphurCarrier Proteins from Primary Metabolic Pathways for 2-Thiosugar Biosynthesis. Nature 2014, 510, 427−431. (305) Podust, L. M.; Sherman, D. H. Diversity of P450 Enzymes in the Biosynthesis of Natural Products. Nat. Prod. Rep. 2012, 29, 1251− 1266. (306) Møldrup, M. E.; Geu-Flores, F.; Halkier, B. A. Assigning Gene Function in Biosynthetic Pathways: Camalexin and Beyond. Plant Cell 2013, 25, 360−367. (307) Sellam, A.; Dongo, A.; Guillemette, T.; Hudhomme, P.; Simoneau, P. Transcriptional Responses to Exposure to the Brassicaceous Defence Metabolites Camalexin and Allyl-Isothiocyanate in the Necrotrophic Fungus Alternaria Brassicicola. Mol. Plant Pathol. 2007, 8, 195−208. 5572

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

(308) Klein, A. P.; Anarat-Cappillino, G.; Sattely, E. S. Minimum Set of Cytochromes P450 for Reconstituting the Biosynthesis of Camalexin, a Major Arabidopsis Antibiotic. Angew. Chem., Int. Ed. 2013, 52, 13625−13628. (309) Glawischnig, E.; Hansen, B. G.; Olsen, C. E.; Halkier, B. A. Camalexin Is Synthesized from Indole-3-Acetaidoxime, a Key Branching Point between Primary and Secondary Metabolism in Arabidopsis. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 8245−8250. (310) Klein, A. P.; Sattely, E. S. Two Cytochromes P450 Catalyze SHeterocyclizations in Cabbage Phytoalexin Biosynthesis. Nat. Chem. Biol. 2015, 11, 837−839. (311) Pedras, M. S.; Sorensen, J. L.; Okanga, F. I.; Zaharia, I. L. Wasalexins a and B, New Phytoalexins from Wasabi: Isolation, Synthesis, and Antifungal Activity. Bioorg. Med. Chem. Lett. 1999, 9, 3015−3020. (312) Pedras, M. S.; Yaya, E. E.; Hossain, S. Unveiling the Phytoalexin Biosynthetic Puzzle in Salt Cress: Unprecedented Incorporation of Glucobrassicin into Wasalexins a and B. Org. Biomol. Chem. 2010, 8, 5150−5158. (313) Park, E. J.; Pezzuto, J. M.; Jang, K. H.; Nam, S. J.; Bucarey, S. A.; Fenical, W. Suppression of Nitric Oxide Synthase by Thienodolin in Lipopolysaccharide-Stimulated Raw 264.7 Murine Macrophage Cells. Nat. Prod. Commun. 2012, 7, 789−794. (314) Milbredt, D.; Patallo, E. P.; van Pee, K. H. A Tryptophan 6Halogenase and an Amidotransferase Are Involved in Thienodolin Biosynthesis. ChemBioChem 2014, 15, 1011−1020. (315) Wang, Y.; Wang, J.; Yu, S.; Wang, F.; Ma, H.; Yue, C.; Liu, M.; Deng, Z.; Huang, Y.; Qu, X. Identifying the Minimal Enzymes for Unusual Carbon-Sulfur Bonds Formation in the Thienodolin Biosynthesis. ChemBioChem 2016, 17, 799. (316) Johnston, N. J.; Mukhtar, T. A.; Wright, G. D. Streptogramin Antibiotics: Mode of Action and Resistance. Curr. Drug Targets 2002, 3, 335−344. (317) Xie, Y. C.; Wang, B.; Liu, J.; Zhou, J. C.; Ma, J. Y.; Huang, H. B.; Ju, J. H. Identification of the Biosynthetic Gene Cluster and Regulatory Cascade for the Synergistic Antibacterial Antibiotics Griseoviridin and Viridogrisein in Streptomyces Griseoviridis. ChemBioChem 2012, 13, 2745−2757. (318) Xie, Y.; Li, Q.; Song, Y.; Ma, J.; Ju, J. Involvement of Sgvp in Carbon-Sulfur Bond Formation During Griseoviridin Biosynthesis. ChemBioChem 2014, 15, 1183−1189. (319) Koiso, Y.; Natori, M.; Iwasaki, S.; Sato, S.; Sonoda, R.; Fujita, Y.; Yaegashi, H.; Sato, Z. Ustiloxin: A Phytotoxin and a Mycotoxin from False Smuth Balls on Rice Panicles. Tetrahedron Lett. 1992, 33, 4157−4160. (320) Koiso, Y.; Li, Y.; Iwasaki, S.; Hanaka, K.; Kobayashi, T.; Sonoda, R.; Fujita, Y.; Yaegashi, H.; Sato, Z. Ustiloxins, Antimitotic Cyclic Peptides from False Smut Balls on Rice Panicles Caused by Ustilaginoidea Virens. J. Antibiot. 1994, 47, 765−773. (321) Koiso, Y.; Morisaki, N.; Yamashita, Y.; Mitsui, Y.; Shirai, R.; Hashimoto, Y.; Iwasaki, S. Isolation and Structure of an Antimitotic Cyclic Peptide, Ustiloxin F: Chemical Interrelation with a Homologous Peptide, Ustiloxin B. J. Antibiot. 1998, 51, 418−422. (322) Umemura, M.; Koike, H.; Nagano, N.; Ishii, T.; Kawano, J.; Yamane, N.; Kozone, I.; Horimoto, K.; Shin-ya, K.; Asai, K.; et al. Middas-M: Motif-Independent De Novo Detection of Secondary Metabolite Gene Clusters through the Integration of Genome Sequencing and Transcriptome Data. PLoS One 2013, 8, e84028. (323) Tsukui, T.; Nagano, N.; Umemura, M.; Kumagai, T.; Terai, G.; Machida, M.; Asai, K. Ustiloxins, Fungal Cyclic Peptides, Are Ribosomally Synthesized in Ustilaginoidea Virens. Bioinformatics 2015, 31, 981−985. (324) Umemura, M.; Nagano, N.; Koike, H.; Kawano, J.; Ishii, T.; Miyamura, Y.; Kikuchi, M.; Tamano, K.; Yu, J.; Shin-ya, K.; et al. Characterization of the Biosynthetic Gene Cluster for the Ribosomally Synthesized Cyclic Peptide Ustiloxin B in Aspergillus Flavus. Fungal Genet. Biol. 2014, 68, 23−30. (325) Nagano, N.; Umemura, M.; Izumikawa, M.; Kawano, J.; Ishii, T.; Kikuchi, M.; Tomii, K.; Kumagai, T.; Yoshimi, A.; Machida, M.;

et al. Class of Cyclic Ribosomal Peptide Synthetic Genes in Filamentous Fungi. Fungal Genet. Biol. 2016, 86, 58−70. (326) Ye, Y.; Minami, A.; Igarashi, Y.; Izumikawa, M.; Umemura, M.; Nagano, N.; Machida, M.; Kawahara, T.; Shin-ya, K.; Gomi, K.; et al. Unveiling the Biosynthetic Pathway of the Ribosomally Synthesized and Post-Translationally Modified Peptide Ustiloxin B in Filamentous Fungi. Angew. Chem., Int. Ed. 2016, 55, 8072−8075. (327) Ding, W.; Liu, W. Q.; Jia, Y.; Li, Y.; van der Donk, W. A.; Zhang, Q. Biosynthetic Investigation of Phomopsins Reveals a Widespread Pathway for Ribosomal Natural Products in Ascomycetes. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 3521−3526. (328) Walton, J. D.; Hallen-Adams, H. E.; Luo, H. Ribosomal Biosynthesis of the Cyclic Peptide Toxins of Amanita Mushrooms. Biopolymers 2010, 94, 659−664. (329) Luo, H.; Hallen-Adams, H. E.; Scott-Craig, J. S.; Walton, J. D. Ribosomal Biosynthesis of Alpha-Amanitin in Galerina Marginata. Fungal Genet. Biol. 2012, 49, 123−129. (330) Hallen, H. E.; Luo, H.; Scott-Craig, J. S.; Walton, J. D. Gene Family Encoding the Major Toxins of Lethal Amanita Mushrooms. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 19097−19101. (331) Luo, H.; Hallen-Adams, H. E.; Walton, J. D. Processing of the Phalloidin Proprotein by Prolyl Oligopeptidase from the Mushroom Conocybe Albipes. J. Biol. Chem. 2009, 284, 18070−18077. (332) Abu-Omar, M. M.; Loaiza, A.; Hontzeas, N. Reaction Mechanisms of Mononuclear Non-Heme Iron Oxygenases. Chem. Rev. 2005, 105, 2227−2252. (333) Kovaleva, E. G.; Lipscomb, J. D. Versatility of Biological NonHeme Fe(Ii) Centers in Oxygen Activation Reactions. Nat. Chem. Biol. 2008, 4, 186−193. (334) Brakhage, A. A.; Thön, M.; Spröte, P.; Scharf, D. H.; AlAbdallah, Q.; Wolke, S. M.; Hortschansky, P. Aspects on Evolution of Fungal Beta-Lactam Biosynthesis Gene Clusters and Recruitment of Trans-Acting Factors. Phytochemistry 2009, 70, 1801−1811. (335) Hamed, R. B.; Gomez-Castellanos, J. R.; Henry, L.; Ducho, C.; McDonough, M. A.; Schofield, C. J. The Enzymes of Beta-Lactam Biosynthesis. Nat. Prod. Rep. 2013, 30, 21−107. (336) Baldwin, J. E.; Schofield, C. In The Chemistry of B-Lactams; Page, M. I., Ed.; Springer Netherlands: Dordrecht, 1992. (337) Cohen, G.; Shiffman, D.; Mevarech, M.; Aharonowitz, Y. Microbial Isopenicillin N Synthase Genes: Structure, Function, Diversity and Evolution. Trends Biotechnol. 1990, 8, 105−111. (338) Kreisberg-Zakarin, R.; Borovok, I.; Yanko, M.; Aharonowitz, Y.; Cohen, G. Recent Advances in the Structure and Function of Isopenicillin N Synthase. Antonie van Leeuwenhoek 1999, 75, 33−39. (339) Baldwin, J. E.; Bradley, M. Isopenicillin N Synthase: Mechanistic Studies. Chem. Rev. 1990, 90, 1079−1088. (340) Howard-Jones, A. R.; Elkins, J. M.; Clifton, I. J.; Roach, P. L.; Adlington, R. M.; Baldwin, J. E.; Rutledge, P. J. Interactions of Isopenicillin N Synthase with Cyclopropyl-Containing Substrate Analogues Reveal New Mechanistic Insight. Biochemistry 2007, 46, 4755−4762. (341) Tamanaha, E.; Zhang, B.; Guo, Y.; Chang, W. C.; Barr, E. W.; Xing, G.; St. Clair, J.; Ye, S.; Neese, F.; Bollinger, J. M., Jr.; et al. Spectroscopic Evidence for the Two C-H-Cleaving Intermediates of Aspergillus Nidulans Isopenicillin N Synthase. J. Am. Chem. Soc. 2016, 138, 8862−8874. (342) Roach, P. L.; Clifton, I. J.; Fulop, V.; Harlos, K.; Barton, G. J.; Hajdu, J.; Andersson, I.; Schofield, C. J.; Baldwin, J. E. Crystal Structure of Isopenicillin N Synthase Is the First from a New Structural Family of Enzymes. Nature 1995, 375, 700−704. (343) Roach, P. L.; Clifton, I. J.; Hensgens, C. M. H.; Shibata, N.; Schofield, C. J.; Hajdu, J.; Baldwin, J. E. Structure of Isopenicillin N Synthase Complexed with Substrate and the Mechanism of Penicillin Formation. Nature 1997, 387, 827−830. (344) Lloyd, M. D.; Lipscomb, S. J.; Hewitson, K. S.; Hensgens, C. M. H.; Baldwin, J. E.; Schofield, C. J. Controlling the Substrate Selectivity of Deacetoxycephalosporin/Deacetylcephalosporin C Synthase. J. Biol. Chem. 2004, 279, 15420−15426. 5573

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

(345) Townsend, C. A. The Stereochemical Fate of Chiral-Methyl Valines in Cephalosporin C Biosynthesis. J. Nat. Prod. 1985, 48, 708− 724. (346) Townsend, C. A.; Theis, A. B.; Neese, A. S.; Barrabee, E. B.; Poland, D. Stereochemical Fate of Chiral Methyl of Valine in the Ring Expansion of Penicillin N to Deacetoxycephalosporin C. J. Am. Chem. Soc. 1985, 107, 4760−4767. (347) Neuss, N.; Nash, C. H.; Baldwin, J. E.; Lemke, P. A.; Grutzner, J. B. Incorporation of (2rs,3s)-(4−13s)Valine into Cephalosporin C. J. Am. Chem. Soc. 1973, 95, 3797−3798. (348) Kluender, H.; Bradley, C. H.; Sih, C. J.; Fawcett, P.; Abraham, E. P. Synthesis and Incorporation of (2s,3s)-(4−13c)Valine into BetaLactam Antibiotics. J. Am. Chem. Soc. 1973, 95, 6149−6150. (349) Valegård, K.; van Scheltinga, A. C. T.; Lloyd, M. D.; Hara, T.; Ramaswamy, S.; Perrakis, A.; Thompson, A.; Lee, H.-J.; Baldwin, J. E.; Schofield, C. J.; et al. Structure of a Cephalosporin Synthase. Nature 1998, 394, 805−809. (350) Valegård, K.; van Scheltinga, A. C. T.; Dubus, A.; Ranghino, G.; Oster, L. M.; Hajdu, J.; Andersson, I. The Structural Basis of Cephalosporin Formation in a Mononuclear Ferrous Enzyme. Nat. Struct. Mol. Biol. 2004, 11, 95−101. (351) Oster, L. M.; van Scheltinga, A. C.; Valegård, K.; Hose, A. M.; Dubus, A.; Hajdu, J.; Andersson, I. Conformational Flexibility of the C Terminus with Implications for Substrate Binding and Catalysis Revealed in a New Crystal Form of Deacetoxycephalosporin C Synthase. J. Mol. Biol. 2004, 343, 157−171. (352) Lloyd, M. D.; Lee, H. J.; Harlos, K.; Zhang, Z. H.; Baldwin, J. E.; Schofield, C. J.; Charnock, J. M.; Garner, C. D.; Hara, T.; Terwisscha van Scheltinga, A. C.; et al. Studies on the Active Site of Deacetoxycephalosporin C Synthase. J. Mol. Biol. 1999, 287, 943−960. (353) Lee, H. J.; Lloyd, M. D.; Harlos, K.; Clifton, I. J.; Baldwin, J. E.; Schofield, C. J. Kinetic and Crystallographic Studies on Deacetoxycephalosporin C Synthase (Daocs). J. Mol. Biol. 2001, 308, 937−948. (354) Tarhonskaya, H.; Szöllössi, A.; Leung, I. K.; Bush, J. T.; Henry, L.; Chowdhury, R.; Iqbal, A.; Claridge, T. D.; Schofield, C. J.; Flashman, E. Studies on Deacetoxycephalosporin C Synthase Support a Consensus Mechanism for 2-Oxoglutarate Dependent Oxygenases. Biochemistry 2014, 53, 2483−2493. (355) Hartman, P. E. Ergothioneine as Antioxidant. Methods Enzymol. 1990, 186, 310−318. (356) Grundemann, D. The Ergothioneine Transporter Controls and Indicates Ergothioneine Activity–a Review. Prev. Med. 2012, 54 (Suppl), S71−74. (357) Seebeck, F. P. In Vitro Reconstitution of Mycobacterial Ergothioneine Biosynthesis. J. Am. Chem. Soc. 2010, 132, 6632−6633. (358) Goncharenko, K. V.; Vit, A.; Blankenfeldt, W.; Seebeck, F. P. Structure of the Sulfoxide Synthase Egtb from the Ergothioneine Biosynthetic Pathway. Angew. Chem., Int. Ed. 2015, 54, 2821−2824. (359) Vit, A.; Mashabela, G. T.; Blankenfeldt, W.; Seebeck, F. P. Structure of the Ergothioneine-Biosynthesis Amidohydrolase Egtc. ChemBioChem 2015, 16, 1490−1496. (360) Song, H.; Hu, W.; Naowarojna, N.; Her, A. S.; Wang, S.; Desai, R.; Qin, L.; Chen, X.; Liu, P. Mechanistic Studies of a Novel C-S Lyase in Ergothioneine Biosynthesis: The Involvement of a Sulfenic Acid Intermediate. Sci. Rep. 2015, 5, 11870. (361) Bello, M. H.; Barrera-Perez, V.; Morin, D.; Epstein, L. The Neurospora Crassa Mutant Ncdeltaegt-1 Identifies an Ergothioneine Biosynthetic Gene and Demonstrates That Ergothioneine Enhances Conidial Survival and Protects against Peroxide Toxicity During Conidial Germination. Fungal Genet. Biol. 2012, 49, 160−172. (362) Hu, W.; Song, H.; Sae Her, A.; Bak, D. W.; Naowarojna, N.; Elliott, S. J.; Qin, L.; Chen, X. P.; Liu, P. H. Bioinformatic and Biochemical Characterizations of C-S Bond Formation and Cleavage Enzymes in the Fungus Neurospora Crassa Ergothioneine Biosynthetic Pathway. Org. Lett. 2014, 16, 5382−5385. (363) Pluskal, T.; Ueno, M.; Yanagida, M. Genetic and Metabolomic Dissection of the Ergothioneine and Selenoneine Biosynthetic Pathway in the Fission Yeast, S. Pombe, and Construction of an Overproduction System. PLoS One 2014, 9, e97774.

(364) Jones, G. W.; Doyle, S.; Fitzpatrick, D. A. The Evolutionary History of the Genes Involved in the Biosynthesis of the Antioxidant Ergothioneine. Gene 2014, 549, 161−170. (365) Sheridan, K. J.; Lechner, B. E.; Keeffe, G. O.; Keller, M. A.; Werner, E. R.; Lindner, H.; Jones, G. W.; Haas, H.; Doyle, S. Ergothioneine Biosynthesis and Functionality in the Opportunistic Fungal Pathogen, Aspergillus Fumigatus. Sci. Rep. 2016, 6, 35306. (366) Palumbo, A.; d'Ischia, M.; Misuraca, G.; Prota, G. Isolation and Structure of a New Sulfur-Containing Amino-Acid from Sea-Urchin Eggs. Tetrahedron Lett. 1982, 23, 3207−3208. (367) Turner, E.; Hager, L. J.; Shapiro, B. M. Ovothiol Replaces Glutathione Peroxidase as a Hydrogen Peroxide Scavenger in Sea Urchin Eggs. Science 1988, 242, 939−941. (368) Ariyanayagam, M. R.; Fairlamb, A. H. Ovothiol and Trypanothione as Antioxidants in Trypanosomatids. Mol. Biochem. Parasitol. 2001, 115, 189−198. (369) Spies, H. S.; Steenkamp, D. J. Thiols of Intracellular Pathogens. Identification of Ovothiol a in Leishmania Donovani and Structural Analysis of a Novel Thiol from Mycobacterium Bovis. Eur. J. Biochem. 1994, 224, 203−213. (370) Castellano, I.; Migliaccio, O.; D’Aniello, S.; Merlino, A.; Napolitano, A.; Palumbo, A. Shedding Light on Ovothiol Biosynthesis in Marine Metazoans. Sci. Rep. 2016, 6, 21506. (371) Vogt, R. N.; Spies, H. S.; Steenkamp, D. J. The Biosynthesis of Ovothiol a (N-Methyl-4-Mercaptohistidine). Identification of S-(4′-LHistidyl)-L-Cysteine Sulfoxide as an Intermediate and the Products of the Sulfoxide Lyase Reaction. Eur. J. Biochem. 2001, 268, 5229−5241. (372) Braunshausen, A.; Seebeck, F. P. Identification and Characterization of the First Ovothiol Biosynthetic Enzyme. J. Am. Chem. Soc. 2011, 133, 1757−1759. (373) Mashabela, G. T.; Seebeck, F. P. Substrate Specificity of an Oxygen Dependent Sulfoxide Synthase in Ovothiol Biosynthesis. Chem. Commun. 2013, 49, 7714−7716. (374) Song, H.; Leninger, M.; Lee, N.; Liu, P. Regioselectivity of the Oxidative C-S Bond Formation in Ergothioneine and Ovothiol Biosyntheses. Org. Lett. 2013, 15, 4854−4857. (375) Walsh, C. T.; Wencewicz, T. A. Flavoenzymes: Versatile Catalysts in Biosynthetic Pathways. Nat. Prod. Rep. 2013, 30, 175−200. (376) Gomez-Escribano, J. P.; Song, L.; Fox, D. J.; Yeo, V.; Bibb, M. J.; Challis, G. L. Structure and Biosynthesis of the Unusual Polyketide Alkaloid Coelimycin P1, a Metabolic Product of the Cpk Gene Cluster of Streptomyces Coelicolor M145. Chem. Sci. 2012, 3, 2716−2720. (377) Ding, L.; Maier, A.; Fiebig, H. H.; Lin, W. H.; Hertweck, C. A Family of Multicyclic Indolosesquiterpenes from a Bacterial Endophyte. Org. Biomol. Chem. 2011, 9, 4029−4031. (378) Xu, Z.; Baunach, M.; Ding, L.; Hertweck, C. Bacterial Synthesis of Diverse Indole Terpene Alkaloids by an Unparalleled Cyclization Sequence. Angew. Chem., Int. Ed. 2012, 51, 10293−10297. (379) Baunach, M.; Franke, J.; Hertweck, C. Terpenoid Biosynthesis Off the Beaten Track: Unconventional Cyclases and Their Impact on Biomimetic Synthesis. Angew. Chem., Int. Ed. 2015, 54, 2604−2626. (380) Baunach, M.; Ding, L.; Willing, K.; Hertweck, C. Bacterial Synthesis of Unusual Sulfonamide and Sulfone Antibiotics by Flavoenzyme-Mediated Sulfur Dioxide Capture. Angew. Chem., Int. Ed. 2015, 54, 13279−13283. (381) Baunach, M.; Ding, L.; Bruhn, T.; Bringmann, G.; Hertweck, C. Regiodivergent N-C and N-N Aryl Coupling Reactions of Indoloterpenes and Cycloether Formation Mediated by a Single Bacterial Flavoenzyme. Angew. Chem., Int. Ed. 2013, 52, 9040−9043. (382) Murray, A. T.; Dowley, M. J.; Pradaux-Caggiano, F.; Baldansuren, A.; Fielding, A. J.; Tuna, F.; Hendon, C. H.; Walsh, A.; Lloyd-Jones, G. C.; John, M. P.; et al. Catalytic Amine Oxidation under Ambient Aerobic Conditions: Mimicry of Monoamine Oxidase B. Angew. Chem., Int. Ed. 2015, 54, 8997−9000. (383) Dunn, R. V.; Munro, A. W.; Turner, N. J.; Rigby, S. E.; Scrutton, N. S. Tyrosyl Radical Formation and Propagation in Flavin Dependent Monoamine Oxidases. ChemBioChem 2010, 11, 1228− 1231. 5574

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

(384) Hearing, V. J. Determination of Melanin Synthetic Pathways. J. Invest. Dermatol. 2011, 131, E8−E11. (385) Halaouli, S.; Asther, M.; Sigoillot, J. C.; Hamdi, M.; Lomascolo, A. Fungal Tyrosinases: New Prospects in Molecular Characteristics, Bioengineering and Biotechnological Applications. J. Appl. Microbiol. 2006, 100, 219−232. (386) Fairhead, M.; Thony-Meyer, L. Bacterial Tyrosinases: Old Enzymes with New Relevance to Biotechnology. New Biotechnol. 2012, 29, 183−191. (387) Kanteev, M.; Goldfeder, M.; Fishman, A. Structure-Function Correlations in Tyrosinases. Protein Sci. 2015, 24, 1360−1369. (388) Agrup, G.; Hansson, C.; Rorsman, H.; Rosengren, A. M.; Rosengren, E. Trichochromes in Red Human Hair. Acta Derm. Venereol. 1978, 58, 357−358. (389) Wondrak, G. T.; Jacobson, M. K.; Jacobson, E. L. Endogenous Uva-Photosensitizers: Mediators of Skin Photodamage and Novel Targets for Skin Photoprotection. Photochem. Photobiol. Sci. 2006, 5, 215−237. (390) Scherlach, K.; Nützmann, H. W.; Schroeckh, V.; Dahse, H. M.; Brakhage, A. A.; Hertweck, C. Cytotoxic Pheofungins from an Engineered Fungus Impaired in Posttranslational Protein Modification. Angew. Chem., Int. Ed. 2011, 50, 9843−9847. (391) Ohnishi, Y.; Furusho, Y.; Higashi, T.; Chun, H. K.; Furihata, K.; Sakuda, S.; Horinouchi, S. Structures of Grixazone a and B, aFactor-Dependent Yellow Pigments Produced under Phosphate Depletion by Streptomyces Griseus. J. Antibiot. 2004, 57, 218−223. (392) Suzuki, H.; Ohnishi, Y.; Furusho, Y.; Sakuda, S.; Horinouchi, S. Novel Benzene Ring Biosynthesis from C(3) and C(4) Primary Metabolites by Two Enzymes. J. Biol. Chem. 2006, 281, 36944−36951. (393) Suzuki, H.; Furusho, Y.; Higashi, T.; Ohnishi, Y.; Horinouchi, S. A Novel O-Aminophenol Oxidase Responsible for Formation of the Phenoxazinone Chromophore of Grixazone. J. Biol. Chem. 2006, 281, 824−833. (394) Singleton, V. L.; Salgues, M.; Zaya, J.; Trousdale, E. Caftaric Acid Disappearance and Conversion to Products of Enzymic Oxidation in Grape Must and Wine. Am. J. Enol. Vitic. 1985, 36, 50−56. (395) Ferreira-Lima, N.; Vallverdu-Queralt, A.; Meudec, E.; Mazauric, J. P.; Sommerer, N.; Bordignon-Luiz, M. T.; Cheynier, V.; Le Guerneve, C. Synthesis, Identification, and Structure Elucidation of Adducts Formed by Reactions of Hydroxycinnamic Acids with Glutathione or Cysteinylglycine. J. Nat. Prod. 2016, 79, 2211−2222. (396) Cheynier, V.; Rigaud, J.; Souquet, J.-M.; Duprat, F.; Moutounet, M. Must Browning in Relation to the Behavior of Phenolic Compounds During Oxidation. Am. J. Enol. Vitic. 1990, 41, 346−349. (397) Ryan, D.; Robards, K.; Prenzler, P.; Jardine, D.; Herlt, T.; Antolovich, M. Liquid Chromatography with Electrospray Ionisation Mass Spectrometric Detection of Phenolic Compounds from Olea Europaea. J. Chromatogr. A 1999, 855, 529−537. (398) Broderick, J. B.; Duffus, B. R.; Duschene, K. S.; Shepard, E. M. Radical S-Adenosylmethionine Enzymes. Chem. Rev. 2014, 114, 4229− 4317. (399) Jarrett, J. T. The Biosynthesis of Thiol- and ThioetherContaining Cofactors and Secondary Metabolites Catalyzed by Radical S-Adenosylmethionine Enzymes. J. Biol. Chem. 2015, 290, 3972−3979. (400) Benz, G.; Schröder, T.; Kurz, J.; Wünsche, C.; Karl, W.; Steffens, G.; Pfitzner, J.; Schmidt, D. Konstitution Der Desferriform Der Albomycine D1, D2, E. Angew. Chem., Int. Ed. Engl. 1982, 21, 1322−1335. (401) Ferguson, A. D.; Braun, V.; Fiedler, H. P.; Coulton, J. W.; Diederichs, K.; Welte, W. Crystal Structure of the Antibiotic Albomycin in Complex with the Outer Membrane Transporter Fhua. Protein Sci. 2000, 9, 956−963. (402) Braun, V.; Pramanik, A.; Gwinner, T.; Köberle, M.; Bohn, E. Sideromycins: Tools and Antibiotics. BioMetals 2009, 22, 3−13. (403) Braun, V.; Günthner, K.; Hantke, K.; Zimmermann, L. Intracellular Activation of Albomycin in Escherichia Coli and Salmonella Typhimurium. J. Bacteriol. 1983, 156, 308−315.

(404) Paulsen, H.; Brieden, M.; Benz, G. Verzweigte Und Kettenverlängerte Zucker, Xxxi. Synthese Des Sauerstoffanalogons Der Desferriform Von Δ1-Albomycin. Liebigs Ann. Chem. 1987, 1987, 565−575. (405) Zeng, Y.; Kulkarni, A.; Yang, Z.; Patil, P. B.; Zhou, W.; Chi, X.; Van Lanen, S.; Chen, S. Biosynthesis of Albomycin Δ(2) Provides a Template for Assembling Siderophore and Aminoacyl-Trna Synthetase Inhibitor Conjugates. ACS Chem. Biol. 2012, 7, 1565−1575. (406) Ayala-Castro, C.; Saini, A.; Outten, F. W. Fe-S Cluster Assembly Pathways in Bacteria. Microbiol. Mol. Biol. Rev. 2008, 72, 110−125. (407) Kulkarni, A.; Zeng, Y.; Zhou, W.; Van Lanen, S.; Zhang, W.; Chen, S. A Branch Point of Streptomyces Sulfur Amino Acid Metabolism Controls the Production of Albomycin. Appl. Environ. Microbiol. 2016, 82, 467−477. (408) Sit, C. S.; van Belkum, M. J.; McKay, R. T.; Worobo, R. W.; Vederas, J. C. The 3d Solution Structure of Thurincin H, a Bacteriocin with Four Sulfur to Alpha-Carbon Crosslinks. Angew. Chem., Int. Ed. 2011, 50, 8718−8721. (409) Babasaki, K.; Takao, T.; Shimonishi, Y.; Kurahashi, K. Subtilosin a, a New Antibiotic Peptide Produced by Bacillus Subtilis 168: Isolation, Structural Analysis, and Biogenesis. J. Biochem. 1985, 98, 585−603. (410) Kawulka, K. E.; Sprules, T.; Diaper, C. M.; Whittal, R. M.; McKay, R. T.; Mercier, P.; Zuber, P.; Vederas, J. C. Structure of Subtilosin a, a Cyclic Antimicrobial Peptide from Bacillus Subtilis with Unusual Sulfur to Alpha-Carbon Cross-Links: Formation and Reduction of Alpha-Thio-Alpha-Amino Acid Derivatives. Biochemistry 2004, 43, 3385−3395. (411) Rea, M. C.; Sit, C. S.; Clayton, E.; O’Connor, P. M.; Whittal, R. M.; Zheng, J.; Vederas, J. C.; Ross, R. P.; Hill, C. Thuricin Cd, a Posttranslationally Modified Bacteriocin with a Narrow Spectrum of Activity against Clostridium Diff icile. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 9352−9357. (412) Lee, H.; Churey, J. J.; Worobo, R. W. Biosynthesis and Transcriptional Analysis of Thurincin H, a Tandem Repeated Bacteriocin Genetic Locus, Produced by Bacillus Thuringiensis Sf361. FEMS Microbiol. Lett. 2009, 299, 205−213. (413) Bruender, N. A.; Wilcoxen, J.; Britt, R. D.; Bandarian, V. Biochemical and Spectroscopic Characterization of a Radical SAdenosyl-L-Methionine Enzyme Involved in the Formation of a Peptide Thioether Cross-Link. Biochemistry 2016, 55, 2122−2134. (414) Haft, D. H.; Basu, M. K. Biological Systems Discovery in Silico: Radical S-Adenosylmethionine Protein Families and Their Target Peptides for Posttranslational Modification. J. Bacteriol. 2011, 193, 2745−2755. (415) Murphy, K.; O’Sullivan, O.; Rea, M. C.; Cotter, P. D.; Ross, R. P.; Hill, C. Genome Mining for Radical Sam Protein Determinants Reveals Multiple Sactibiotic-Like Gene Clusters. PLoS One 2011, 6, e20852. (416) Zheng, G.; Hehn, R.; Zuber, P. Mutational Analysis of the SboAlb Locus of Bacillus Subtilis: Identification of Genes Required for Subtilosin Production and Immunity. J. Bacteriol. 2000, 182, 3266− 3273. (417) Fluhe, L.; Knappe, T. A.; Gattner, M. J.; Schafer, A.; Burghaus, O.; Linne, U.; Marahiel, M. A. The Radical Sam Enzyme Alba Catalyzes Thioether Bond Formation in Subtilosin A. Nat. Chem. Biol. 2012, 8, 350−357. (418) Fluhe, L.; Burghaus, O.; Wieckowski, B. M.; Giessen, T. W.; Linne, U.; Marahiel, M. A. Two [4fe-4s] Clusters Containing Radical Sam Enzyme Skfb Catalyze Thioether Bond Formation During the Maturation of the Sporulation Killing Factor. J. Am. Chem. Soc. 2013, 135, 959−962. (419) Wieckowski, B. M.; Hegemann, J. D.; Mielcarek, A.; Boss, L.; Burghaus, O.; Marahiel, M. A. The Pqqd Homologous Domain of the Radical Sam Enzyme Thnb Is Required for Thioether Bond Formation During Thurincin H Maturation. FEBS Lett. 2015, 589, 1802−1806. (420) Sit, C. S.; McKay, R. T.; Hill, C.; Ross, R. P.; Vederas, J. C. The 3d Structure of Thuricin Cd, a Two-Component Bacteriocin with 5575

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

Cysteine Sulfur to Alpha-Carbon Cross-Links. J. Am. Chem. Soc. 2011, 133, 7680−7683. (421) Grell, T. A.; Goldman, P. J.; Drennan, C. L. Spasm and Twitch Domains in S-Adenosylmethionine (Sam) Radical Enzymes. J. Biol. Chem. 2015, 290, 3964−3971. (422) Aoki, M.; Ohtsuka, T.; Yamada, M.; Ohba, Y.; Yoshizaki, H.; Yasuno, H.; Sano, T.; Watanabe, J.; Yokose, K.; Seto, H. Cyclothiazomycin, a Novel Polythiazole-Containing Peptide with Renin Inhibitory Activity. Taxonomy, Fermentation, Isolation and PhysicoChemical Characterization. J. Antibiot. 1991, 44, 582−588. (423) Hashimoto, M.; Murakami, T.; Funahashi, K.; Tokunaga, T.; Nihei, K.; Okuno, T.; Kimura, T.; Naoki, H.; Himeno, H. An Rna Polymerase Inhibitor, Cyclothiazomycin B1, and Its Isomer. Bioorg. Med. Chem. 2006, 14, 8259−8270. (424) Cox, C. L.; Tietz, J. I.; Sokolowski, K.; Melby, J. O.; Doroghazi, J. R.; Mitchell, D. A. Nucleophilic 1,4-Additions for Natural Product Discovery. ACS Chem. Biol. 2014, 9, 2014−2022. (425) Wang, J.; Yu, Y.; Tang, K.; Liu, W.; He, X.; Huang, X.; Deng, Z. Identification and Analysis of the Biosynthetic Gene Cluster Encoding the Thiopeptide Antibiotic Cyclothiazomycin in Streptomyces Hygroscopicus 10−22. Appl. Environ. Microbiol. 2010, 76, 2335− 2344. (426) Ono, K.; Okajima, T.; Tani, M.; Kuroda, S.; Sun, D.; Davidson, V. L.; Tanizawa, K. Involvement of a Putative [Fe-S]-Cluster-Binding Protein in the Biogenesis of Quinohemoprotein Amine Dehydrogenase. J. Biol. Chem. 2006, 281, 13672−13684. (427) Nakai, T.; Deguchi, T.; Frebort, I.; Tanizawa, K.; Okajima, T. Identification of Genes Essential for the Biogenesis of Quinohemoprotein Amine Dehydrogenase. Biochemistry 2014, 53, 895−907. (428) Nakai, T.; Ito, H.; Kobayashi, K.; Takahashi, Y.; Hori, H.; Tsubaki, M.; Tanizawa, K.; Okajima, T. The Radical S-Adenosyl-LMethionine Enzyme Qhpd Catalyzes Sequential Formation of IntraProtein Sulfur-to-Methylene Carbon Thioether Bonds. J. Biol. Chem. 2015, 290, 11144−11166. (429) Datta, S.; Mori, Y.; Takagi, K.; Kawaguchi, K.; Chen, Z. W.; Okajima, T.; Kuroda, S.; Ikeda, T.; Kano, K.; Tanizawa, K.; et al. Structure of a Quinohemoprotein Amine Dehydrogenase with an Uncommon Redox Cofactor and Highly Unusual Crosslinking. Proc. Natl. Acad. Sci. U. S. A. 2001, 98, 14268−14273. (430) Satoh, A.; Kim, J. K.; Miyahara, I.; Devreese, B.; Vandenberghe, I.; Hacisalihoglu, A.; Okajima, T.; Kuroda, S.; Adachi, O.; Duine, J. A.; et al. Crystal Structure of Quinohemoprotein Amine Dehydrogenase from Pseudomonas Putida. Identification of a Novel Quinone Cofactor Encaged by Multiple Thioether Cross-Bridges. J. Biol. Chem. 2002, 277, 2830−2834. (431) Liang, Z. X. Complexity and Simplicity in the Biosynthesis of Enediyne Natural Products. Nat. Prod. Rep. 2010, 27, 499−528. (432) Smith, A. L.; Nicolaou, K. C. The Enediyne Antibiotics. J. Med. Chem. 1996, 39, 2103−2117. (433) Rohr, J.; Zeeck, A. Metabolic Products of Microorganisms. 240. Urdamycins, New Angucycline Antibiotics from Streptomyces Fradiae. II. Structural Studies of Urdamycins B to F. J. Antibiot. 1987, 40, 459− 467. (434) Rohr, J. Biosynthetic Formation of the S-Methyl Group of the Angucycline Antibiotic Urdamycin E. J. Chem. Soc., Chem. Commun. 1989, 492−493. (435) Trefzer, A.; Hoffmeister, D.; Künzel, E.; Stockert, S.; Weitnauer, G.; Westrich, L.; Rix, U.; Fuchser, J.; Bindseil, K. U.; Rohr, J.; et al. Function of Glycosyltransferase Genes Involved in Urdamycin a Biosynthesis. Chem. Biol. 2000, 7, 133−142. (436) Schneider, P.; Jacobs, J. M.; Neres, J.; Aldrich, C. C.; Allen, C.; Nett, M.; Hoffmeister, D. The Global Virulence Regulators Vsrad and Phca Control Secondary Metabolism in the Plant Pathogen. ChemBioChem 2009, 10, 2730−2732. (437) Pauly, J.; Spiteller, D.; Linz, J.; Jacobs, J.; Allen, C.; Nett, M.; Hoffmeister, D. Ralfuranone Thioether Production by the Plant Pathogen. ChemBioChem 2013, 14, 2169−2178. (438) Wackler, B.; Schneider, P.; Jacobs, J. M.; Pauly, J.; Allen, C.; Nett, M.; Hoffmeister, D. Ralfuranone Biosynthesis in Ralstonia

Solanacearum Suggests Functional Divergence in the Quinone Synthetase Family of Enzymes. Chem. Biol. 2011, 18, 354−360. (439) Ohtawa, M.; Hishinuma, Y.; Takagi, E.; Yamada, T.; Ito, F.; Arima, S.; Uchida, R.; Kim, Y. P.; Omura, S.; Tomoda, H.; et al. Synthesis and Structural Revision of Cyslabdan. Chem. Pharm. Bull. 2016, 64, 1370−1377. (440) Fukumoto, A.; Kim, Y. P.; Matsumoto, A.; Takahashi, Y.; Shiomi, K.; Tomoda, H.; Omura, S. Cyslabdan, a New Potentiator of Imipenem Activity against Methicillin-Resistant Staphylococcus Aureus, Produced by Streptomyces Sp. K04−0144. I. Taxonomy, Fermentation, Isolation and Structural Elucidation. J. Antibiot. 2008, 61, 1−6. (441) Fukumoto, A.; Kim, Y. P.; Hanaki, H.; Shiomi, K.; Tomoda, H.; Omura, S. Cyslabdan, a New Potentiator of Imipenem Activity against Methicillin-Resistant Staphylococcus Aureus, Produced by Streptomyces Sp. K04−0144. Ii. Biological Activities. J. Antibiot. 2008, 61, 7−10. (442) Koyama, N.; Tokura, Y.; Munch, D.; Sahl, H. G.; Schneider, T.; Shibagaki, Y.; Ikeda, H.; Tomoda, H. The Nonantibiotic Small Molecule Cyslabdan Enhances the Potency of Beta-Lactams against Mrsa by Inhibiting Pentaglycine Interpeptide Bridge Synthesis. PLoS One 2012, 7, e48981. (443) Ikeda, H.; Shin-Ya, K.; Nagamitsu, T.; Tomoda, H. Biosynthesis of Mercapturic Acid Derivative of the Labdane-Type Diterpene, Cyslabdan That Potentiates Imipenem Activity against Methicillin-Resistant Staphylococcus Aureus: Cyslabdan Is Generated by Mycothiol-Mediated Xenobiotic Detoxification. J. Ind. Microbiol. Biotechnol. 2016, 43, 325−342. (444) Fu, P.; MacMillan, J. B. Spithioneines a and B, Two New Bohemamine Derivatives Possessing Ergothioneine Moiety from a Marine-Derived Streptomyces Spinoverrucosus. Org. Lett. 2015, 17, 3046−3049. (445) Argoudelis, A. D.; Brinkley, T. A.; Brodasky, T. F.; Buege, J. A.; Meyer, H. F.; Mizsak, S. A. Paulomycins a and B. Isolation and Characterization. J. Antibiot. 1982, 35, 285−294. (446) Wiley, P. F. A New Antibiotic, U-43,120 (Nsc-163500). J. Antibiot. 1976, 29, 587−589. (447) Argoudelis, A. D.; Baczynskyj, L.; Haak, W. J.; Knoll, W. M.; Mizsak, S. A.; Shilliday, F. B. New Paulomycins Produced by Streptomyces Paulus. J. Antibiot. 1988, 41, 157−169. (448) Argoudelis, A. D.; Baczynskyj, L.; Mizsak, S. A.; Shilliday, F. B.; Spinelli, P. A.; DeZwaan, J. Paldimycins a and B and Antibiotics 273a2 Alpha and 273a2 Beta. Synthesis and Characterization. J. Antibiot. 1987, 40, 419−436. (449) Argoudelis, A. D.; Baczynskyj, L.; Mizsak, S. A.; Shilliday, F. B. O-Demethylpaulomycins a and B, U-77,802 and U-77,803, Paulomenols a and B, New Metabolites Produced by Streptomyces Paulus. J. Antibiot. 1988, 41, 1316−1330. (450) Argoudelis, A. D.; Baczynskyj, L.; Buege, J. A.; Marshall, V. P.; Mizsak, S. A.; Wiley, P. F. Paulomycin-Related Antibiotics: Paldimycins and Antibiotics 273a2. Isolation and Characterization. J. Antibiot. 1987, 40, 408−418. (451) Li, J.; Xie, Z.; Wang, M.; Ai, G.; Chen, Y. Identification and Analysis of the Paulomycin Biosynthetic Gene Cluster and Titer Improvement of the Paulomycins in Streptomyces Paulus Nrrl 8115. PLoS One 2015, 10, e0120542. (452) Gonzalez, A.; Rodriguez, M.; Brana, A. F.; Mendez, C.; Salas, J. A.; Olano, C. New Insights into Paulomycin Biosynthesis Pathway in Streptomyces Albus J1074 and Generation of Novel Derivatives by Combinatorial Biosynthesis. Microb. Cell Fact. 2016, 15, 56. (453) Li, J.; Wang, M.; Ding, Y.; Tang, Y.; Zhang, Z.; Chen, Y. Involvement of an Octose Ketoreductase and Two Acyltransferases in the Biosynthesis of Paulomycins. Sci. Rep. 2016, 6, 21180. (454) Laursen, J. B.; Nielsen, J. Phenazine Natural Products: Biosynthesis, Synthetic Analogues, and Biological Activity. Chem. Rev. 2004, 104, 1663−1686. (455) Heine, D.; Martin, K.; Hertweck, C. Genomics-Guided Discovery of Endophenazines from Kitasatospora Sp. Hki 714. J. Nat. Prod. 2014, 77, 1083−1087. 5576

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577

Chemical Reviews

Review

(456) Heine, D.; Sundaram, S.; Beudert, M.; Martin, K.; Hertweck, C. A Widespread Bacterial Phenazine Forms S-Conjugates with Biogenic Thiols and Crosslinks Proteins. Chem. Sci. 2016, 7, 4848− 4855. (457) O’Malley, Y. Q.; Reszka, K. J.; Spitz, D. R.; Denning, G. M.; Britigan, B. E. Pseudomonas Aeruginosa Pyocyanin Directly Oxidizes Glutathione and Decreases Its Levels in Airway Epithelial Cells. Am. J. Physiol. Lung Cell Mol. Physiol. 2004, 287, L94−L103. (458) Roelofs, D.; Timmermans, M. J.; Hensbergen, P.; van Leeuwen, H.; Koopman, J.; Faddeeva, A.; Suring, W.; de Boer, T. E.; Mariën, J.; Boer, R.; et al. A Functional Isopenicillin N Synthase in an Animal Genome. Mol. Biol. Evol. 2013, 30, 541−548. (459) Brakhage, A. A.; Al-Abdallah, Q.; Tüncher, A.; Spröte, P. Evolution of Beta-Lactam Biosynthesis Genes and Recruitment of Trans-Acting Factors. Phytochemistry 2005, 66, 1200−1210. (460) Lincke, T.; Behnken, S.; Ishida, K.; Roth, M.; Hertweck, C. Closthioamide: An Unprecedented Polythioamide Antibiotic from the Strictly Anaerobic Bacterium Clostridium Cellulolyticum. Angew. Chem., Int. Ed. 2010, 49, 2011−2013. (461) Behnken, S.; Lincke, T.; Kloss, F.; Ishida, K.; Hertweck, C. Antiterminator-Mediated Unveiling of Cryptic Polythioamides in an Anaerobic Bacterium. Angew. Chem., Int. Ed. 2012, 51, 2425−2428. (462) Dassama, L. M.; Kenney, G. E.; Rosenzweig, A. C. Methanobactins: From Genome to Function. Metallomics 2017, 9, 7−20. (463) Puar, M. S.; Ganguly, A. K.; Afonso, A.; Brambilla, R.; Mangiaracina, P.; Sarre, O.; MacFarlane, R. D. Sch 18640. A New Thiostrepton-Type Antibiotic. J. Am. Chem. Soc. 1981, 103, 5231− 5233. (464) Winter, J. M.; Behnken, S.; Hertweck, C. Genomics-Inspired Discovery of Natural Products. Curr. Opin. Chem. Biol. 2011, 15, 22− 31.

5577

DOI: 10.1021/acs.chemrev.6b00697 Chem. Rev. 2017, 117, 5521−5577