ACS Publications. Most Trusted. Most Cited. Most Read
Divergent Evolution of Lanthipeptide Stereochemistry
My Activity

Figure 1Loading Img
  • Open Access
Articles

Divergent Evolution of Lanthipeptide Stereochemistry
Click to copy article linkArticle link copied!

  • Raymond Sarksian
    Raymond Sarksian
    Department of Chemistry and Howard Hughes Medical Institute, University of Illinois at Urbana-Champaign, Urbana, Illinois 61822, United States
  • Wilfred A. van der Donk*
    Wilfred A. van der Donk
    Department of Chemistry and Howard Hughes Medical Institute, University of Illinois at Urbana-Champaign, Urbana, Illinois 61822, United States
    Carl R. Woese Institute for Genomic Biology, University of Illinois at Urbana-Champaign, Urbana, Illinois 61822, United States
    *Email: [email protected]. Tel: 217 244 5360.
Open PDFSupporting Information (2)

ACS Chemical Biology

Cite this: ACS Chem. Biol. 2022, 17, 9, 2551–2558
Click to copy citationCitation copied!
https://doi.org/10.1021/acschembio.2c00492
Published August 24, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

The three-dimensional structure of natural products is critical for their biological activities and, as such, enzymes have evolved that specifically generate active stereoisomers. Lanthipeptides are post-translationally modified peptidic natural products that contain macrocyclic thioethers featuring lanthionine (Lan) and/or methyllanthionine (MeLan) residues with defined stereochemistry. In this report, we compare two class I lanthipeptide biosynthetic gene clusters (BGCs), coi and olv, that represent two families of lanthipeptide gene clusters found in Actinobacteria. The precursor peptides and BGCs are quite similar with genes encoding a dehydratase, cyclase, and methyltransferase (MT). We illustrate that the precursor peptide CoiA1 is converted by these enzymes into a polymacrocyclic product, mCoiA1, that contains an analogous ring pattern to the previously characterized post-translationally modified OlvA peptide (mOlvA). However, a clear distinction between the two BGCs is an additional Thr-glutamyl lyase (GL) domain that is fused to the MT, CoiSA, which results in divergence of the product stereochemistry for the coi BGC. Two out of three MeLan rings of mCoiA1 contain different stereochemistry than the corresponding residues in mOlvA, with the most notable difference being a rare d-allo-l-MeLan residue, the formation of which is guided by CoiSA. This study illustrates how nature utilizes a distinct GL to control natural product stereochemistry in lanthipeptide biosynthesis.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2022 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

Natural products typically recognize their targets with exquisite affinity and selectivity. (1) In the course of their evolution, the structures of these molecules have been optimized to bind to the usually chiral environments of biological targets. This high level of recognition is often achieved by the rich stereochemistry of natural products, which have made them privileged ligands. (1) In this study, we report the discovery of two lanthipeptides that have very similar ring patterns but in which the stereochemistry of two of the three macrocycles differs. We show that the acquisition of one new enzyme has driven the divergence of the two compound groups and that this new activity has resulted in coevolution of other enzymes in the pathway.
Lanthipeptides represent one of the largest classes of ribosomally synthesized and posttranslationally modified peptides (RiPPs). (2−4) They exhibit a wide range of activities, including antimicrobial, antiviral, morphogenetic, and antifungal, (3,5−10) and are defined by the presence of lanthionine (Lan) or methyllanthionine (MeLan) residues (Figure 1). Maturation of lanthipeptides features the dehydration of Ser/Thr residues followed by subsequent Michael-type addition of a Cys thiol onto the dehydroamino acids to form (Me)Lan (Figure 1). (2−4)

Figure 1

Figure 1. (A) Maturation of lanthipeptides proceeds through dehydration of Ser/Thr residues to generate the corresponding dehydroamino acids dehydroalanine and dehydrobutyrine (Dha, Dhb). Cyclization of Cys side chains onto Dha/Dhb yields (Me)Lan. (B) Four possible MeLan diastereomers can be generated as shown.

For class I lanthipeptides, a lanthipeptide dehydratase LanB first catalyzes the transesterification of a glutamate group from glutamyl-tRNAGlu to the side chain of Ser/Thr. (11−14) An elimination reaction of the glutamylated intermediate generates dehydroalanine (Dha) from Ser or dehydrobutyrine (Dhb) from Thr (Figure 1A). A LanC cyclase then catalyzes the formation of thioether rings by the addition of a Cys thiol to the dehydroamino acid intermediates. (15,16)
The stereochemical configuration of (Me)Lan residues has been shown to be important for the biological activities of lanthipeptides. (17,18) Three MeLan diastereomers have been discovered to date, (2S,3S,6R)-, (2R,3R,6R)-, and (2S,3R,6R)-MeLan, hereafter referred to as dl-, ll-, and d-allo-l-MeLan (Figure 1B). (4,19−25) Both dl- and ll-MeLan are believed to form through the anti-elimination of Thr residues to yield (Z)-Dhb residues followed by an anti-addition of Cys across the (Z)-Dhb. (4) Facial selectivity of the cyclization event dictates whether dl- or ll-MeLan is formed from the (Z)-Dhb. (20)d-allo-l-MeLan was recently reported for the morphogenetic class I lanthipeptide SapT. (24)d-allo-l-MeLan is thought to be formed through syn-elimination of glutamylated Thr residues followed by subsequent anti-addition of Cys across the (E)-Dhb. (24) The SapT biosynthetic gene cluster (BGC) features a split dehydratase made up of SptBa and SptBb that carries out dehydration. SptBa catalyzes glutamylation of Ser/Thr residues, and SptBb is a glutamyl lyase (GL) that catalyzes syn-elimination. (24)
In this study, we compare the coi BGC from Streptomyces coelicolor A3(2) with the olv BGC from Streptomyces olivaceus NRRL B-3009 (Figure 2A). These BGCs are representative examples of two groups of gene clusters in Actinobacteria (Figure 2A), with only the olv BGC previously investigated in depth. (26) Both BGCs encode a canonical class I dehydratase (CoiB and OlvB) and cyclase (CoiC and OlvC), and an O-methyltransferase (MT) that is the most widespread auxiliary enzyme in class I lanthipeptide BGCs. (26−28) The sequences of their precursor peptides are also quite similar (Figure 2B). The ring pattern and stereochemistry of the (Me)Lan residues of the olv product have been determined by nuclear magnetic resonance (NMR) spectroscopy and comparison with synthetic standards. (26) The coi BGC mainly differs from the olv BGC in that the MT CoiSA has a fused GL domain and encodes a protein of unknown function CoiH. (24,26)

Figure 2

Figure 2. (A) Comparison of coi and olv class I lanthipeptide BGCs. The coi BGC encodes an additional GL and protein of unknown function when compared to the olv BGC. (B) Sequence alignment of the precursor peptides encoded in both BGCs. Residues that are known to be modified in OlvA are colored. (C) Sequence alignment of GL domains illustrates that the additional GL domain in CoiSA is similar to SptBb.

The presence of two GLs in the coi BGC is unusual. The GL domains in both CoiB and CoiSA contain conserved Arg residues that are important for the recognition of the γ-carboxylate of the glutamylated peptide intermediate in lanthipeptide dehydratases (Figure 2C). (11−14) However, only CoiB contains the catalytic His base and Arg residue that are critical for glutamate elimination activity in canonical anti-GLs (Figures 2C and S1). (11−14) The GL domain in CoiSA in contrast contains similar putative active site residues as SptBb. (24)
Here, we investigate the regioselectivity of the two distinct GL domains in CoiB and CoiSA (Figure 2). Both enzymes catalyze glutamate elimination; however, they are proposed to generate two different Dhb isomers during the maturation of CoiA1. (24) We demonstrate that the 3-fold dehydrated and cyclized product, mCoiA1, contains three different MeLan diastereomers in a ring pattern that is very similar to the olv product. However, elucidation of the stereochemical configuration for each MeLan residue showed that two of the three residues have different stereochemistry compared to that found in the olv product. Furthermore, our data show that CoiC catalyzes cyclization with both (Z)- and (E)-Dhb residues but only when these isomers are at their native location, suggesting coevolution of the cyclase with the product stereochemistry. These findings provide intriguing insights into the divergent evolution of two widespread natural products by acquisition (or deletion) of a GL that results in different stereochemistry.

Results

Click to copy section linkSection link copied!

Heterologous Production and Characterization of mCoiA1

The coi cluster is present in the genome of one of the most widely studied strains of Streptomyces, S. coelicolor A3(2), (29−32) but its product has not been previously observed despite considerable genome mining studies. (33−37) Therefore, in this study, we used heterologous expression to investigate its product. CoiA1 was previously shown to undergo 3-fold dehydration and cyclization to yield mCoiA1 when coexpressed with CoiB, the elimination domain of CoiSA (CoiSA(ED)), and CoiC in Escherichia coli. (24) In this work, the coexpressed product was isolated and treated with endoproteinase GluC (Figure 3). High-resolution mass spectrometry (HRMS) analysis confirmed that the product was dehydrated three times (Figure 3A). Tandem MS of the GluC-digested peptide suggests that it contains an N-terminal MeLan ring and two C-terminal overlapping MeLan rings (Figure 3B, Table S3). This ring pattern is similar to that determined by tandem MS and NMR spectroscopy for the olv product, (26) consistent with the conservation of the positions of the Ser/Thr and Cys residues in the CoiA1 and OlvA precursor peptides (Figure 2B).

Figure 3

Figure 3. Liquid chromatography-MS (LC-MS) analysis of the GluC digestion product of mCoiA1 (mCoiA1GluC). (A) High-resolution electrospray ionization-MS (ESI-MS) analysis. (B) Tandem ESI-MS analysis. Fragmentation results are consistent with an N-terminal MeLan and two C-terminal overlapping MeLan.

Glutamate Elimination Activity

Next, we performed experiments to gain insight into the elimination activity of both CoiB and CoiSA(ED). Coexpression of CoiA1 with CoiB and CoiC led to at most one dehydration along with intermediates that were glutamylated once or twice (Figure 4). This finding suggests that the C-terminal lyase domain of CoiB was only able to perform one elimination of the Ser/Thr residues that were glutamylated by the N-terminal domain of CoiB. In contrast, when CoiA1 was coexpressed with CoiSA(ED), CoiC, and the CoiB-H994A mutant, in which the lyase activity of the C-terminal domain of CoiB was inactivated but the glutamylation activity was retained, (12,14) a 3-fold dehydrated product was observed (Figure S2). Thus, CoiSA(ED) was able to eliminate glutamate at all three Thr residues.

Figure 4

Figure 4. Matrix-assisted laser desorption ionization time-of-flight (MALDI-TOF) MS analysis of coexpression experiments in E. coli. Coexpressed proteins are listed on each panel. Asterisks indicate deamination products that are commonly generated in MALDI-TOF mass spectra at these masses.

An N-ethylmaleimide (NEM) alkylation assay (38) to test for the presence of free Cys suggested that the product peptide of the latter experiment was not fully cyclized (Figure S3), implying that the 3-fold dehydrated peptide generated by CoiSA(ED) is not a competent substrate for CoiC. These results show that both CoiB and CoiSA(ED) are required for correct dehydration and cyclization of CoiA1. Finally, we investigated whether CoiC is necessary to obtain a 3-fold dehydrated product. In the case of some lanthipeptides such as microbisporicin A1, (13) the lanthipeptide cyclase is required to obtain full dehydration since select (Me)Lan rings must form prior to the next dehydration event. Coexpression of CoiA1 with CoiB and CoiSA(ED) resulted in a 3-fold dehydrated product (Figure 4). Therefore, CoiC and the formation of (Me)Lan rings are not necessary for combined CoiB and CoiSA(ED) activity.

Stereochemical Analysis of MeLan Residues

mCoiA1 was previously demonstrated to contain dl-, ll-, and d-allo-l-MeLan residues (Figure 5A). (24) Assignment of the stereochemistry to specific rings was not reported. Since mCoiA1 only contains MeLan rings, Thr residues involved in ring formation were individually mutated to Ser residues to determine the stereochemistry for each MeLan ring. (25) If successful, this approach would remove a single MeLan residue (by conversion to a Lan residue) in each variant and thus allow assignment of stereochemistry to that MeLan.

Figure 5

Figure 5. Gas chromatography-MS (GC-MS) analysis of derivatized MeLan residues (m/z = 379) obtained from mCoiA1 mutant peptides using a chiral stationary phase. Coinjections with stereochemically pure derivatized dl- and ll-MeLan confirm the presence of dl-MeLan in mCoiA1-T43S and ll-MeLan in mCoiA1-T50S. (A) Derivatized MeLan from WT mCoiA1 (top), synthetic standards, and coinjections of the products of mCoiA1 with the synthetic standards. (B) Derivatized MeLan from mCoiA1-T43S (top), synthetic standards, and coinjections of the products of mCoiA1-T43S with the synthetic standards. (C) Derivatized MeLan from mCoiA1-T50S (top), synthetic standards, and coinjections of the products of mCoiA1-T50S with the synthetic standards. (D) Structures of derivatized dl- and ll-MeLan standards.

CoiA1-T43S, CoiA1-T50S, and CoiA1-T57S variants were generated by site-directed mutagenesis and coexpressed with CoiB, CoiSA(ED), and CoiC followed by isolation of the products by Ni-affinity chromatography. mCoiA1-T43S and mCoiA1-T50S underwent 3-fold dehydration and were unreactive toward NEM, suggesting that the peptides were cyclized (Figure S4). HRMS and tandem MS confirmed that the cyclization patterns of these two mutant peptides were not altered from WT mCoiA1 (Figures S5 and S6). In contrast, mCoiA1-T57S was obtained in multiple dehydration states and was not fully processed (Figure S4). Hence, only mCoiA1-T43S and -T50S were hydrolyzed in acid and the resulting amino acids were derivatized for stereochemical analysis.
Stereochemical analysis was performed by gas chromatography-mass spectrometry (GC-MS) on a chiral stationary phase and comparison to authentic dl- and ll-MeLan standards. (25,39,40) mCoiA1-T43S contained two MeLan, as anticipated (Figure 5B). Coinjections with stereochemically pure dl- and ll-MeLan confirmed the presence of dl-MeLan and absence of ll-MeLan (Figure 5B). Therefore, WT mCoiA1 must contain an N-terminal ll-MeLan ring derived from Thr43. Using the same approach, mCoiA1-T50S also revealed peaks in the GC-MS corresponding to two MeLan (Figure 5C). Coinjections confirmed one of the MeLan peaks to consist of MeLan with the ll-configuration and dl-MeLan was absent in the sample. By the same logic, WT mCoiA1 must contain a dl-MeLan ring derived from Thr50.
The second-eluting MeLan peak from mCoiA1-T43S and mCoiA1-T50S that did not match either dl- or ll-MeLan was anticipated to be the rare allo isomer as previously detected in WT mCoiA1. (24) The two possible allo-MeLan stereoisomers were previously shown to be inseparable by GC-MS. (24) Hydrolysis of mCoiA1-T43S and mCoiA1-T50S and derivatization of the amino acids with Marfey’s reagent, Nα-(2,4-dinitro-5-fluorophenyl)-l-alaninamide (l-FDAA), followed by comparison to d-allo-d/l-MeLan and l-allo-d/l-MeLan standards (24) confirmed the presence of d-allo-d/l-MeLan in both peptides by liquid chromatography-mass spectrometry (Figure S7). mCoiA1 was next demonstrated to contain d-allo-l-MeLan and not d-allo-d-MeLan by isolation of MeLan and reductive desulfurization, which would form d-Ala from d-allo-d-MeLan or l-Ala from d-allo-l-MeLan (Figure S8). (9) The desulfurization product consisted of l-Ala (Figure S8) confirming the presence of d-allo-l-MeLan. Because this isomer is also seen in both mCoiA1-T43S and mCoiA1-T50S, the MeLan derived from Thr57 in mCoiA1 must be d-allo-l-MeLan. Based on the sequence similarity of CoiSA(ED) with SptBb, the rare d-allo-l-MeLan is generated through the involvement of CoiSA(ED). The poor conversion of CoiA1-T57S suggests that mutation of Thr57 to Ser is not well tolerated by CoiSA(ED) and may imply that some syn-GLs are specific for Thr.

CoiSA(ED) Mutational Analysis

Mutational analysis was next performed on CoiSA(ED) to decipher the importance of putative active site residues. Based on sequence analysis, CoiB is very similar to the nisin dehydratase NisB and related class I lanthipeptide dehydratases that catalyze anti-elimination (Figure S1). (11−14) GLs that catalyze anti-elimination have been well characterized both biochemically and structurally. In contrast, syn-GLs have only recently been discovered, and their putative active sites diverge from anti-GLs (Figures 2C and S1). CoiSA(ED), SptBb, and related homologs contain a highly conserved Lys residue that when mutated in CoiSA(ED) (CoiSA(ED)-K46A) resulted in the accumulation of glutamylated peptides implying its importance for glutamate elimination. (24)
In addition to the differences observed between syn- and anti-GLs, some key sequence similarities are also found. In NisB, Arg784 and Arg786 bind to the γ-carboxylate of the glutamylated peptide intermediate and are critical for elimination activity. (12−14) These residues are also conserved in CoiSA(ED), and Ala mutants were generated to determine the importance of activity. Both CoiSA(ED)-R51A and -R53A mutants were coexpressed with CoiA1, CoiB, and CoiC. For both mutants, a 3-fold dehydrated product was obtained, along with a 2-fold dehydrated product (Figure 6). Therefore, in contrast to NisB, the Arg residues do not seem to be absolutely critical as elimination activity is not severely diminished. Finally, we generated CoiSA(ED)-E89A. This Glu is highly conserved across all class I lanthipeptide GLs (Figure S1; Glu823 in NisB), and based on a calculated structure of SptBb, the residue points toward the putative active site and may play a role in catalysis. (24) Coexpression experiments of CoiA1, CoiB, and CoiC with the CoiSA(ED)-E89A mutant revealed that dehydration activity was altered, but a 3-fold dehydrated peptide was still generated (Figure 6). Thus, Glu89 is also not critical for catalysis by CoiSA(ED).

Figure 6

Figure 6. MALDI-TOF MS analysis of CoiA1 coexpression with CoiB, CoiC, and CoiSA(ED) WT or CoiSA(ED) mutants. Asterisks indicate deamination products that are commonly generated in MALDI-TOF mass spectra at these masses.

Bioinformatic Analysis and Comparison of anti- and syn-GLs

Two key differences exist between anti- and syn-GLs investigated thus far based on sequence and mutational analysis. A highly conserved Arg residue (Arg826 in NisB) is present in anti-GLs (Figure 2C) that interacts with the carbonyl oxygen of glutamylated Thr to lower the pKa of the α-proton and facilitate elimination. (14) For SptBb, CoiSA(ED), and related homologs, this residue is replaced by a hydrophobic residue (Figure 2C). In addition, SptBb and CoiSA(ED) contain a highly conserved Lys residue that is important for elimination activity that is a Tyr in anti-GLs.
We surveyed all GLs involved in the BGCs of lanthipeptides and related RiPPs using these sequence differences. A sequence-similarity network (SSN) for GLs was previously generated using the tools of the Enzyme Function Initiative (Figure 7). (41,42) In this study, the genomic context of each group was inspected and the GLs in the class I lanthipeptide BGCs were selected for sequence alignments to determine whether they belong to anti-GLs or syn-GLs (Figure 7 and Table S6). This analysis suggested grouping of the GLs into three broad categories. The largest group of GLs is the full-length LanB proteins that are all predicted to catalyze anti-elimination (Figures 7, black, and S1). Both CoiB and OlvB are found within this group. Smaller groups of anti-GLs are part of split LanB systems with dedicated glutamyl transferases and glutamate lyases (Figures 7, blue, and S1). Consistent with a previous study, (24) a significant portion of the network also revealed syn-GLs as (1) part of a split LanB system, (2) fused to a methyltransferase domain such as CoiSA, or (3) present as an additional stand-alone GL domain, with the latter two in BGCs that also contain a full-length dehydratase (Figure 7, purple). All other groups containing more than two members were inspected and found to be part of BGCs of other RiPPs such as thiopeptides (Figure 7, gray, and Table S6). Analysis of GLs within these BGCs suggests that the associated GLs are similar to anti-GLs and are likely generating (Z)-Dhb and/or Dha residues as either intermediates and/or in the final products. Hence, the occurence of syn-GLs seems to be limited in the currently sequenced genomes to class I lanthipeptide BGCs. The SSN also provided the opportunity to assess whether (E)-Dhb and/or allo-MeLan isomers are present in any previously reported class I lanthipeptides for which stereochemistry has generally not been determined. The active site residues of GLs involved in the biosynthesis of these previously characterized family members were analyzed (Figure S9). These GLs were found to be very similar to anti-GLs suggesting that (E)-Dhbs and/or allo-MeLan isomers are likely not present in these lanthipeptides.

Figure 7

Figure 7. SSN analysis of GLs. All colored groups indicate GLs within class I lanthipeptide BGCs. Full-length LanB dehydratases that catalyze anti-elimination are shown in black, and anti-GLs within split LanB systems in blue. Syn-GLs that are part of a split LanB system that are fused to a methyltransferase or that are present as stand-alone proteins are depicted in purple. GLs within BGCs of other nonlanthipeptide or hybrid RiPPs are in gray. A select number of characterized GLs are labeled including the thiopeptide GLs TbtC and LazF. For brief summaries of the biosynthetic genes in each group, see Table S6. The cytoscape file for the SSN is provided as the Supporting Information.

Discussion

Click to copy section linkSection link copied!

The overall structure of mCoiA1 in terms of the stereochemical configuration of (Me)Lan residues is the most complex of any lanthipeptide characterized thus far. mCoiA1 contains an N-terminal ll-MeLan ring and overlapping C-terminal dl- and d-allo-l-MeLan rings (Figure 8A). Except for an additional Lan in mOlvA, mCoiA1 and mOlvA contain (Me)Lan rings at equivalent positions. It is interesting to note that the Ser residue that is involved in this additional N-terminal Lan in OlvA is also present in CoiA1 but escapes dehydration. The corresponding Cys residue to generate the Lan ring is missing in CoiA1 and replaced by an Asn residue (Figures 2B and 8A).

Figure 8

Figure 8. (A) Comparison of the structures of GluC-digested mCoiA1 and mOlvA to highlight differences and similarities in ring patterns and stereochemistry. Abu, 2-aminobutyric acid. The N-terminal Lan ring in mOlvA is absent in mCoiA1, although the corresponding Ser residue (gold) is present. (B) Sequence LOGO made using orthologs of the core peptide region of OlvA and CoiA1 identified by BLAST analysis (45) to highlight conserved residues (OlvA total sequences = 43, CoiA1 total sequences = 301).

Although the ring patterns of mOlvA and mCoiA1 are similar, the stereochemical configurations of two of the three (Me)Lan rings differ. In the C-terminal region of these peptides, mOlvA contains an ll-Lan, whereas mCoiA1 contains a rare d-allo-l-MeLan residue (Figure 8A). A critical distinction between the coi and olv BGCs is the presence of the additional GL domain in CoiSA, which is similar to SptBb that was previously proposed to catalyze syn-elimination of glutamylated Thr to generate (E)-Dhb that leads to the observed d-allo-l-MeLan in SapT. (24) The second deviation in stereochemistry is that mCoiA1 contains an N-terminal ll-MeLan ring where mOlvA contains a dl-MeLan ring at the equivalent position (Figure 8A). Unlike the change in stereochemistry at the C-terminus, which is accounted for by the presence of a gene encoding an additional syn-GL in the BGC, the origin of the change in the stereochemistry of the N-terminal MeLan is less clear. The conformational landscape of lanthipeptides has been shown to be important for the cyclization reaction. (43,44) Therefore, one possibility is that the additional N-terminal ll-Lan ring in mOlvA is formed early during the biosynthetic process and that this ring conformationally biases the peptide toward forming an alternative diastereomer for the adjacent ring. Alternatively, the change in the stereochemistry of the C-terminal MeLan because of the recruitment of a new GL could in turn also influence the stereochemistry of the N-terminal ring if the d-allo-l-MeLan is formed early in the maturation process. Regardless of the molecular explanation of the change in stereochemistry, in all investigated examples, engineered changes in stereochemistry have led to the abolishment of the original bioactivity of the lanthipeptide. (17,18) Therefore, it is likely that the differences in stereochemistry between the products of the widespread coi and olv gene cluster families have functional consequences.
Several examples have been reported wherein RiPP biosynthesis requires an obligate order of post-translational modifications. (3) The investigation of the coi BGC provides another example of high coordination of the post-translational modification reactions. Based on the stereochemistry of the final product, it is likely that CoiB converts Thr43 and Thr50 into (Z)-Dhb residues, which are then the substrates for CoiC-catalyzed cyclization events that provide the ll- and dl-MeLan residues, respectively. The elimination domain of CoiSA likely dehydrates Thr57 to (E)-Dhb, which CoiC then converts to d-allo-l-MeLan. When CoiSA(ED) was coexpressed with CoiA1 and CoiC and a variant of CoiB that can still glutamylate but not eliminate, three dehydrations were still observed, but the cyclase was unable to form the three thioether macrocycles, presumably because the peptide now contained three (E)-Dhb residues. Thus, CoiC is only able to accept (E)-Dhb at position 57 and requires (Z)-Dhb at positions 43 and 50 for cyclization activity. The observation that CoiC is capable of cyclization of Cys51 onto (E)-Dhb57 but apparently cannot catalyze the addition of Cys47 and Cys55 to (E)-Dhb43 and 50, respectively, is suggestive of coevolution of CoiC with the appearance of CoiSA in the pathway. Conversely, when only CoiB was coexpressed with CoiA1 and CoiC, the dehydration process stalled at a single dehydration and one or two glutamylations. These findings suggest that either the syn-elimination of Thr57 by CoiSA(ED) is required for CoiB to complete its dehydration and/or that CoiC must first form the d-allo-l-MeLan for full CoiB activity. The data also suggest that after CoiSA/CoiC act on Thr57, glutamylation of Thr43 and Thr50 by the N-terminal domain of CoiB is followed by faster glutamate elimination by the GL domain of CoiB than by CoiSA(ED). These findings therefore suggest a highly choreographed set of biosynthetic reactions to make a complex ring pattern with high fidelity.
Orthologs of the CoiA1 peptide are much more common than orthologs of the OlvA peptide in the currently sequenced genomes (Figure 8B, Tables S7 and S8). Whether the coi-like BGCs evolved from the olv-like clusters by recruitment of a new GL domain or by gene duplication or whether the olv-like BGCs lost the gene for the syn-GL is a difficult question. Based on the preponderance of anti-GLs in diverse RiPP BGCs (Figure 7), it is likely that the ancestral enzyme catalyzed anti-elimination, but this hypothesis cannot be unambiguously verified at present. What is clear is that during evolution in Actinobacteria, two distinct BGCs with a common origin diverged and that stereochemistry was very likely a key determining factor. Determining the function of the coi and olv BGC products and how stereochemistry may alter the biological activities of these compounds is therefore of great interest. Investigations to answer these questions will first need to determine the cleavage site between the leader and core peptide, which is hampered by the absence of any reports detecting the products of these BGCs in their native producing organisms and the absence of a protease in the BGC that could provide insight regarding the start position of the final product.

Conclusions

Click to copy section linkSection link copied!

The coi BGC is highly similar in architecture to the previously investigated olv BGC with a key distinction being the presence of an additional GL domain that is fused to the MT CoiSA. We illustrate that mCoiA1 has a similar ring pattern to mOlvA with one less Lan ring and that it has different stereochemistry for two out of the three MeLan rings. The most pronounced difference between the two products is the recently discovered rare d-allo-l-MeLan diastereomer that is the result of the additional GL domain of CoiSA. This study illustrates an example of divergent evolution driven by stereochemistry, which in turn is likely to be correlated to the function of the final products.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acschembio.2c00492.

  • Glutamyl lyase SSN (ZIP)

  • Experimental procedures; extended experimental data and bioinformatic analysis; materials and methods; primers used in this study; observed and calculated m/z ratios for fragments of GluC-digested mCoiA1; NEM alkylation assay for CoiA1 co-expressed with CoiB-H994A, CoiC, and CoiSA(ED); LC-MS analysis of the GluC digestion product of mCoiA1 and variants; description table of orthologs obtained from PSI-BLAST of OlvA (PDF)

Accession Codes

NCBI protein accessions: CoiA1, WP_011031313.1; CoiB, WP_011031312.1; CoiC, WP_011031311.1; and CoiSA, WP_011031310.1.

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Wilfred A. van der Donk - Department of Chemistry and Howard Hughes Medical Institute, University of Illinois at Urbana-Champaign, Urbana, Illinois 61822, United StatesCarl R. Woese Institute for Genomic Biology, University of Illinois at Urbana-Champaign, Urbana, Illinois 61822, United StatesOrcidhttps://orcid.org/0000-0002-5467-7071 Email: [email protected]
  • Author
    • Raymond Sarksian - Department of Chemistry and Howard Hughes Medical Institute, University of Illinois at Urbana-Champaign, Urbana, Illinois 61822, United StatesOrcidhttps://orcid.org/0000-0001-6940-6699
  • Funding

    This study was supported by the National Institutes of Health (Grant R37GM058822 to W.A.v.d.D.). The Bruker UltrafleXtreme MALDI TOF/TOF mass spectrometer was purchased in part with a grant from the National Center for Research Resources, National Institutes of Health (S10 RR027109 A). W.A.v.d.D. is an investigator of the Howard Hughes Medical Institute (HHMI).

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

The authors thank A. Ulanov at the Carver Biotechnology Center for help with GC-MS experiments and D. Sarlah (UIUC) for the use of a Kinetex Biphenyl column. They thank M. Simon for help with acquiring LC-MS data. This study is subject to HHMI’s Open Access to Publications policy. HHMI laboratory heads have previously granted a nonexclusive CC BY 4.0 license to the public and a sublicensable license to HHMI in their research articles. Pursuant to those licenses, the author-accepted manuscript of this article can be made freely available under a CC BY 4.0 license immediately upon publication.

References

Click to copy section linkSection link copied!

This article references 45 other publications.

  1. 1
    Harvey, A. L.; Edrada-Ebel, R.; Quinn, R. J. The re-emergence of natural products for drug discovery in the genomics era. Nature Rev. Drug Discovery 2015, 14, 111129,  DOI: 10.1038/nrd4510
  2. 2
    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, 3144,  DOI: 10.1016/j.chembiol.2015.11.012
  3. 3
    Montalbán-López, M.; Scott, T. A.; Ramesh, S.; Rahman, I. R.; van Heel, A. J.; Viel, J. H.; Bandarian, V.; Dittmann, E.; Genilloud, O.; Goto, Y. New developments in RiPP discovery, enzymology and engineering. Nat. Prod. Rep. 2021, 38, 130239,  DOI: 10.1039/D0NP00027B
  4. 4
    Repka, L. M.; Chekan, J. R.; Nair, S. K.; van der Donk, W. A. Mechanistic understanding of lanthipeptide biosynthetic enzymes. Chem. Rev. 2017, 117, 54575520,  DOI: 10.1021/acs.chemrev.6b00591
  5. 5
    Kodani, S.; Hudson, M. E.; Durrant, M. C.; Buttner, M. J.; Nodwell, J. R.; Willey, J. M. The SapB morphogen is a lantibiotic-like peptide derived from the product of the developmental gene ramS in Streptomyces coelicolor. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 1144811453,  DOI: 10.1073/pnas.0404220101
  6. 6
    Kodani, S.; Lodato, M. A.; Durrant, M. C.; Picart, F.; Willey, J. M. SapT, a lanthionine-containing peptide involved in aerial hyphae formation in the Streptomycetes. Mol. Microbiol. 2005, 58, 13681380,  DOI: 10.1111/j.1365-2958.2005.04921.x
  7. 7
    Willey, J. M.; Willems, A.; Kodani, S.; Nodwell, J. R. Morphogenetic surfactants and their role in the formation of aerial hyphae in Streptomyces coelicolor. Mol. Microbiol. 2006, 59, 731742,  DOI: 10.1111/j.1365-2958.2005.05018.x
  8. 8
    Férir, G.; Petrova, M. I.; Andrei, G.; Huskens, D.; Hoorelbeke, B.; Snoeck, R.; Vanderleyden, J.; Balzarini, J.; Bartoschek, S.; Brönstrup, M. The lantibiotic peptide labyrinthopeptin A1 demonstrates broad anti-HIV and anti-HSV activity with potential for microbicidal applications. PLoS One 2013, 8, e64010  DOI: 10.1371/journal.pone.0064010
  9. 9
    Mohr, K. I.; Volz, C.; Jansen, R.; Wray, V.; Hoffmann, J.; Bernecker, S.; Wink, J.; Gerth, K.; Stadler, M.; Müller, R. Pinensins: the first antifungal lantibiotics. Angew. Chem., Int. Ed. 2015, 54, 1125411258,  DOI: 10.1002/anie.201500927
  10. 10
    Smith, T. E.; Pond, C. D.; Pierce, E.; Harmer, Z. P.; Kwan, J.; Zachariah, M. M.; Harper, M. K.; Wyche, T. P.; Matainaho, T. K.; Bugni, T. S. Accessing chemical diversity from the uncultivated symbionts of small marine animals. Nat. Chem. Biol. 2018, 14, 179185,  DOI: 10.1038/nchembio.2537
  11. 11
    Garg, N.; Salazar-Ocampo, L. M. A.; van der Donk, W. A. In vitro activity of the nisin dehydratase NisB. Proc. Natl. Acad. Sci. U.S.A. 2013, 110, 72587263,  DOI: 10.1073/pnas.1222488110
  12. 12
    Ortega, M. A.; Hao, Y.; Zhang, Q.; Walker, M. C.; van der Donk, W. A.; Nair, S. K. Structure and mechanism of the tRNA-dependent lantibiotic dehydratase NisB. Nature 2015, 517, 509512,  DOI: 10.1038/nature13888
  13. 13
    Ortega, M. A.; Hao, Y.; Walker, M. C.; Donadio, S.; Sosio, M.; Nair, S. K.; van der Donk, W. A. Structure and tRNA specificity of MibB, a lantibiotic dehydratase from Actinobacteria involved in NAI-107 biosynthesis. Cell Chem. Biol. 2016, 23, 370380,  DOI: 10.1016/j.chembiol.2015.11.017
  14. 14
    Bothwell, I. R.; Cogan, D. P.; Kim, T.; Reinhardt, C. J.; van der Donk, W. A.; Nair, S. K. Characterization of glutamyl-tRNA-dependent dehydratases using nonreactive substrate mimics. Proc. Natl. Acad. Sci. U.S.A. 2019, 116, 1724517250,  DOI: 10.1073/pnas.1905240116
  15. 15
    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, 14641467,  DOI: 10.1126/science.1121422
  16. 16
    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, 2116921175,  DOI: 10.1074/jbc.M701802200
  17. 17
    Mukherjee, S.; Huo, L.; Thibodeaux, G. N.; van der Donk, W. A. Synthesis and bioactivity of diastereomers of the virulence lanthipeptide cytolysin. Org. Lett. 2016, 18, 61886191,  DOI: 10.1021/acs.orglett.6b03246
  18. 18
    Knerr, P. J.; van der Donk, W. A. Chemical synthesis of the lantibiotic lacticin 481 reveals the importance of lanthionine stereochemistry. J. Am. Chem. Soc. 2013, 135, 70947097,  DOI: 10.1021/ja4014024
  19. 19
    Tang, W.; van der Donk, W. A. The sequence of the enterococcal cytolysin imparts unusual lanthionine stereochemistry. Nat. Chem. Biol. 2013, 9, 157159,  DOI: 10.1038/nchembio.1162
  20. 20
    Tang, W.; Jiménez-Osés, G.; Houk, K. N.; van der Donk, W. A. Substrate control in stereoselective lanthionine biosynthesis. Nat. Chem. 2015, 7, 5764,  DOI: 10.1038/nchem.2113
  21. 21
    Tang, W.; Thibodeaux, G. N.; van der Donk, W. A. The enterococcal cytolysin synthetase coevolves with substrate for stereoselective lanthionine synthesis. ACS Chem. Biol. 2016, 11, 24382446,  DOI: 10.1021/acschembio.6b00397
  22. 22
    Garg, N.; Goto, Y.; Chen, T.; van der Donk, W. A. Characterization of the stereochemical configuration of lanthionines formed by the lanthipeptide synthetase GeoM. Biopolymers 2016, 106, 834842,  DOI: 10.1002/bip.22876
  23. 23
    Lohans, C. T.; Li, J. L.; Vederas, J. C. Structure and biosynthesis of carnolysin, a homologue of enterococcal cytolysin with D-amino acids. J. Am. Chem. Soc. 2014, 136, 1315013153,  DOI: 10.1021/ja5070813
  24. 24
    Sarksian, R.; Hegemann, J. D.; Simon, M. A.; Acedo, J. Z.; van der Donk, W. A. Unexpected methyllanthionine stereochemistry in the morphogenetic lanthipeptide SapT. J. Am. Chem. Soc. 2022, 144, 63736382,  DOI: 10.1021/jacs.2c00517
  25. 25
    Bothwell, I. R.; Caetano, T.; Sarksian, R.; Mendo, S.; van der Donk, W. A. Structural analysis of class I lanthipeptides from Pedobacter lusitanus NL19 reveals an unusual ring pattern. ACS Chem. Biol. 2021, 16, 10191029,  DOI: 10.1021/acschembio.1c00106
  26. 26
    Acedo, J. Z.; Bothwell, I. R.; An, L.; Trouth, A.; Frazier, C.; van der Donk, W. A. O-methyltransferase-mediated incorporation of a β-amino acid in lanthipeptides. J. Am. Chem. Soc. 2019, 141, 1679016801,  DOI: 10.1021/jacs.9b07396
  27. 27
    Marsh, A. J.; O’Sullivan, O.; Ross, R. P.; Cotter, P. D.; Hill, C. In silico analysis highlights the frequency and diversity of type 1 lantibiotic gene clusters in genome sequenced bacteria. BMC Genomics 2010, 11, 679  DOI: 10.1186/1471-2164-11-679
  28. 28
    Walker, M. C.; Eslami, S. M.; Hetrick, K. J.; Ackenhusen, S. E.; Mitchell, D. A.; van der Donk, W. A. Precursor peptide-targeted mining of more than one hundred thousand genomes expands the lanthipeptide natural product family. BMC Genomics 2020, 21, 387  DOI: 10.1186/s12864-020-06785-7
  29. 29
    Bentley, S. D.; Chater, K. F.; Cerdeno-Tarraga, A. M.; Challis, G. L.; Thomson, N. R.; James, K. D.; Harris, D. E.; Quail, M. A.; Kieser, H.; Harper, D. Complete genome sequence of the model actinomycete Streptomyces coelicolor A3(2). Nature 2002, 417, 141147,  DOI: 10.1038/417141a
  30. 30
    Challis, G. L. Exploitation of the Streptomyces coelicolor A3(2) genome sequence for discovery of new natural products and biosynthetic pathways. J. Ind. Microbiol. Biotechnol. 2014, 41, 219232,  DOI: 10.1007/s10295-013-1383-2
  31. 31
    Ozaki, T.; Nishiyama, M.; Kuzuyama, T. Novel tryptophan metabolism by a potential gene cluster that is widely distributed among actinomycetes. J. Biol. Chem. 2013, 288, 99469956,  DOI: 10.1074/jbc.M112.436451
  32. 32
    Nodwell, J. R. Microbe profile: Streptomyces coelicolor: a burlesque of pigments and phenotypes. Microbiology 2019, 165, 953955,  DOI: 10.1099/mic.0.000821
  33. 33
    Lautru, S.; Deeth, R. J.; Bailey, L. M.; Challis, G. L. Discovery of a new peptide natural product by Streptomyces coelicolor genome mining. Nat. Chem. Biol. 2005, 1, 265269,  DOI: 10.1038/nchembio731
  34. 34
    Song, L.; Barona-Gomez, F.; Corre, C.; Xiang, L.; Udwary, D. W.; Austin, M. B.; Noel, J. P.; Moore, B. S.; Challis, G. L. Type III polyketide synthase beta-ketoacyl-ACP starter unit and ethylmalonyl-CoA extender unit selectivity discovered by Streptomyces coelicolor genome mining. J. Am. Chem. Soc. 2006, 128, 1475414755,  DOI: 10.1021/ja065247w
  35. 35
    Lin, X.; Hopson, R.; Cane, D. E. Genome mining in Streptomyces coelicolor: molecular cloning and characterization of a new sesquiterpene synthase. J. Am. Chem. Soc. 2006, 128, 60226023,  DOI: 10.1021/ja061292s
  36. 36
    Corre, C.; Song, L.; O’Rourke, S.; Chater, K. F.; Challis, G. L. 2-Alkyl-4-hydroxymethylfuran-3-carboxylic acids, antibiotic production inducers discovered by Streptomyces coelicolor genome mining. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 1751017515,  DOI: 10.1073/pnas.0805530105
  37. 37
    Wu, C.; Ichinose, K.; Choi, Y. H.; van Wezel, G. P. Aromatic polyketide GTRI-02 is a previously unidentified product of the act gene cluster in Streptomyces coelicolor A3(2). ChemBioChem 2017, 18, 14281434,  DOI: 10.1002/cbic.201700107
  38. 38
    Li, B.; Cooper, L. E.; van der Donk, W. A. In Vitro Studies of Lantibiotic Biosynthesis. In Methods Enzymol.; Elsevier, 2009; Vol. 458, pp 533558.
  39. 39
    Küsters, E.; Allgaier, H.; Jung, G.; Bayer, E. Resolution of sulphur-containing amino acids by chiral phase gas chromatography. Chromatographia 1984, 18, 287293,  DOI: 10.1007/BF02259079
  40. 40
    Liu, W.; Chan, A. S. H.; Liu, H.; Cochrane, S. A.; Vederas, J. C. Solid supported chemical syntheses of both components of the lantibiotic lacticin 3147. J. Am. Chem. Soc. 2011, 133, 1421614219,  DOI: 10.1021/ja206017p
  41. 41
    Zallot, R.; Oberg, N.; Gerlt, J. A. The EFI web resource for genomic enzymology tools: Leveraging protein, genome, and metagenome databases to discover novel enzymes and metabolic pathways. Biochemistry 2019, 58, 41694182,  DOI: 10.1021/acs.biochem.9b00735
  42. 42
    Gerlt, J. A.; Bouvier, J. T.; Davidson, D. B.; Imker, H. J.; Sadkhin, B.; Slater, D. R.; Whalen, K. L. Enzyme Function Initiative-Enzyme Similarity Tool (EFI-EST): A web tool for generating protein sequence similarity networks. Biochim. Biophys. Acta, Proteins Proteomics 2015, 1854, 10191037,  DOI: 10.1016/j.bbapap.2015.04.015
  43. 43
    Le, T.; Fouque, K. J. D.; Santos-Fernandez, M.; Navo, C. D.; Jiménez-Osés, G.; Sarksian, R.; Fernandez-Lima, F. A.; van Der Donk, W. A. Substrate sequence controls regioselectivity of lanthionine formation by ProcM. J. Am. Chem. Soc. 2021, 143, 1873318743,  DOI: 10.1021/jacs.1c09370
  44. 44
    Bobeica, S. C.; Zhu, L.; Acedo, J. Z.; Tang, W.; van der Donk, W. A. Structural determinants of macrocyclization in substrate-controlled lanthipeptide biosynthetic pathways. Chem. Sci. 2020, 11, 1285412870,  DOI: 10.1039/D0SC01651A
  45. 45
    Altschul, S. F.; Madden, T. L.; Schaffer, A. A.; Zhang, J.; Zhang, Z.; Miller, W.; Lipman, D. J. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997, 25, 33893402,  DOI: 10.1093/nar/25.17.3389

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 8 publications.

  1. Emily K. Desormeaux, Garrett J. Barksdale, Wilfred A. van der Donk. Kinetic Analysis of Cyclization by the Substrate-Tolerant Lanthipeptide Synthetase ProcM. ACS Catalysis 2024, 14 (24) , 18310-18321. https://doi.org/10.1021/acscatal.4c06216
  2. Ross Rayne Vermeulen, Anton Du Preez van Staden, Tracey Ollewagen, Leonardo Joaquim van Zyl, Youran Luo, Wilfred A. van der Donk, Leon Milner Theodore Dicks, Carine Smith, Marla Trindade. Initial Characterization of the Viridisins’ Biological Properties. ACS Omega 2024, 9 (29) , 31832-31841. https://doi.org/10.1021/acsomega.4c03149
  3. Youran Luo, Shuyun Xu, Autumn M. Frerk, Wilfred A. van der Donk. Facile Method for Determining Lanthipeptide Stereochemistry. Analytical Chemistry 2024, 96 (4) , 1767-1773. https://doi.org/10.1021/acs.analchem.3c04958
  4. Guannan Zhong. Cytochromes P450 Associated with the Biosyntheses of Ribosomally Synthesized and Post-translationally Modified Peptides. ACS Bio & Med Chem Au 2023, 3 (5) , 371-388. https://doi.org/10.1021/acsbiomedchemau.3c00026
  5. Li Cao, Truc Do, Angela Zhu, Jianshu Duan, Nathan Alam, A. James Link. Genome Mining and Discovery of Imiditides, a Family of RiPPs with a Class-Defining Aspartimide Modification. Journal of the American Chemical Society 2023, 145 (34) , 18834-18845. https://doi.org/10.1021/jacs.3c03991
  6. Emily K. Desormeaux, Wilfred A. van der Donk. Kinetic Analysis of Lanthipeptide Cyclization by Substrate-Tolerant ProcM. 2024https://doi.org/10.1101/2024.05.16.594612
  7. Youran Luo, Shuyun Xu, Wilfred A. van der Donk. A Facile Method for Determining Lanthipeptide Stereochemistry. 2023https://doi.org/10.1101/2023.10.26.564210
  8. Christopher J. Thibodeaux. The conformationally dynamic structural biology of lanthipeptide biosynthesis. Current Opinion in Structural Biology 2023, 81 , 102644. https://doi.org/10.1016/j.sbi.2023.102644
  9. Li Cao, Truc Do, Angela D. Zhu, Nathan Alam, A. James Link. Genome Mining and Discovery of Imiditides, a Novel Family of RiPPs with a Class-defining Aspartimide Modification. 2023https://doi.org/10.1101/2023.04.07.536058
  10. Hongshuang Liu, Senbiao Fang, Lin Zhao, Xiao Men, Haibo Zhang. A Single Active-Site Mutagenesis Confers Enhanced Activity and/or Changed Product Distribution to a Pentalenene Synthase from Streptomyces sp. PSKA01. Bioengineering 2023, 10 (3) , 392. https://doi.org/10.3390/bioengineering10030392
  11. Raymond Sarksian, Lingyang Zhu, Wilfred A. van der Donk. syn -Elimination of glutamylated threonine in lanthipeptide biosynthesis. Chemical Communications 2023, 59 (9) , 1165-1168. https://doi.org/10.1039/D2CC06345J

ACS Chemical Biology

Cite this: ACS Chem. Biol. 2022, 17, 9, 2551–2558
Click to copy citationCitation copied!
https://doi.org/10.1021/acschembio.2c00492
Published August 24, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

3635

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. (A) Maturation of lanthipeptides proceeds through dehydration of Ser/Thr residues to generate the corresponding dehydroamino acids dehydroalanine and dehydrobutyrine (Dha, Dhb). Cyclization of Cys side chains onto Dha/Dhb yields (Me)Lan. (B) Four possible MeLan diastereomers can be generated as shown.

    Figure 2

    Figure 2. (A) Comparison of coi and olv class I lanthipeptide BGCs. The coi BGC encodes an additional GL and protein of unknown function when compared to the olv BGC. (B) Sequence alignment of the precursor peptides encoded in both BGCs. Residues that are known to be modified in OlvA are colored. (C) Sequence alignment of GL domains illustrates that the additional GL domain in CoiSA is similar to SptBb.

    Figure 3

    Figure 3. Liquid chromatography-MS (LC-MS) analysis of the GluC digestion product of mCoiA1 (mCoiA1GluC). (A) High-resolution electrospray ionization-MS (ESI-MS) analysis. (B) Tandem ESI-MS analysis. Fragmentation results are consistent with an N-terminal MeLan and two C-terminal overlapping MeLan.

    Figure 4

    Figure 4. Matrix-assisted laser desorption ionization time-of-flight (MALDI-TOF) MS analysis of coexpression experiments in E. coli. Coexpressed proteins are listed on each panel. Asterisks indicate deamination products that are commonly generated in MALDI-TOF mass spectra at these masses.

    Figure 5

    Figure 5. Gas chromatography-MS (GC-MS) analysis of derivatized MeLan residues (m/z = 379) obtained from mCoiA1 mutant peptides using a chiral stationary phase. Coinjections with stereochemically pure derivatized dl- and ll-MeLan confirm the presence of dl-MeLan in mCoiA1-T43S and ll-MeLan in mCoiA1-T50S. (A) Derivatized MeLan from WT mCoiA1 (top), synthetic standards, and coinjections of the products of mCoiA1 with the synthetic standards. (B) Derivatized MeLan from mCoiA1-T43S (top), synthetic standards, and coinjections of the products of mCoiA1-T43S with the synthetic standards. (C) Derivatized MeLan from mCoiA1-T50S (top), synthetic standards, and coinjections of the products of mCoiA1-T50S with the synthetic standards. (D) Structures of derivatized dl- and ll-MeLan standards.

    Figure 6

    Figure 6. MALDI-TOF MS analysis of CoiA1 coexpression with CoiB, CoiC, and CoiSA(ED) WT or CoiSA(ED) mutants. Asterisks indicate deamination products that are commonly generated in MALDI-TOF mass spectra at these masses.

    Figure 7

    Figure 7. SSN analysis of GLs. All colored groups indicate GLs within class I lanthipeptide BGCs. Full-length LanB dehydratases that catalyze anti-elimination are shown in black, and anti-GLs within split LanB systems in blue. Syn-GLs that are part of a split LanB system that are fused to a methyltransferase or that are present as stand-alone proteins are depicted in purple. GLs within BGCs of other nonlanthipeptide or hybrid RiPPs are in gray. A select number of characterized GLs are labeled including the thiopeptide GLs TbtC and LazF. For brief summaries of the biosynthetic genes in each group, see Table S6. The cytoscape file for the SSN is provided as the Supporting Information.

    Figure 8

    Figure 8. (A) Comparison of the structures of GluC-digested mCoiA1 and mOlvA to highlight differences and similarities in ring patterns and stereochemistry. Abu, 2-aminobutyric acid. The N-terminal Lan ring in mOlvA is absent in mCoiA1, although the corresponding Ser residue (gold) is present. (B) Sequence LOGO made using orthologs of the core peptide region of OlvA and CoiA1 identified by BLAST analysis (45) to highlight conserved residues (OlvA total sequences = 43, CoiA1 total sequences = 301).

  • References


    This article references 45 other publications.

    1. 1
      Harvey, A. L.; Edrada-Ebel, R.; Quinn, R. J. The re-emergence of natural products for drug discovery in the genomics era. Nature Rev. Drug Discovery 2015, 14, 111129,  DOI: 10.1038/nrd4510
    2. 2
      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, 3144,  DOI: 10.1016/j.chembiol.2015.11.012
    3. 3
      Montalbán-López, M.; Scott, T. A.; Ramesh, S.; Rahman, I. R.; van Heel, A. J.; Viel, J. H.; Bandarian, V.; Dittmann, E.; Genilloud, O.; Goto, Y. New developments in RiPP discovery, enzymology and engineering. Nat. Prod. Rep. 2021, 38, 130239,  DOI: 10.1039/D0NP00027B
    4. 4
      Repka, L. M.; Chekan, J. R.; Nair, S. K.; van der Donk, W. A. Mechanistic understanding of lanthipeptide biosynthetic enzymes. Chem. Rev. 2017, 117, 54575520,  DOI: 10.1021/acs.chemrev.6b00591
    5. 5
      Kodani, S.; Hudson, M. E.; Durrant, M. C.; Buttner, M. J.; Nodwell, J. R.; Willey, J. M. The SapB morphogen is a lantibiotic-like peptide derived from the product of the developmental gene ramS in Streptomyces coelicolor. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 1144811453,  DOI: 10.1073/pnas.0404220101
    6. 6
      Kodani, S.; Lodato, M. A.; Durrant, M. C.; Picart, F.; Willey, J. M. SapT, a lanthionine-containing peptide involved in aerial hyphae formation in the Streptomycetes. Mol. Microbiol. 2005, 58, 13681380,  DOI: 10.1111/j.1365-2958.2005.04921.x
    7. 7
      Willey, J. M.; Willems, A.; Kodani, S.; Nodwell, J. R. Morphogenetic surfactants and their role in the formation of aerial hyphae in Streptomyces coelicolor. Mol. Microbiol. 2006, 59, 731742,  DOI: 10.1111/j.1365-2958.2005.05018.x
    8. 8
      Férir, G.; Petrova, M. I.; Andrei, G.; Huskens, D.; Hoorelbeke, B.; Snoeck, R.; Vanderleyden, J.; Balzarini, J.; Bartoschek, S.; Brönstrup, M. The lantibiotic peptide labyrinthopeptin A1 demonstrates broad anti-HIV and anti-HSV activity with potential for microbicidal applications. PLoS One 2013, 8, e64010  DOI: 10.1371/journal.pone.0064010
    9. 9
      Mohr, K. I.; Volz, C.; Jansen, R.; Wray, V.; Hoffmann, J.; Bernecker, S.; Wink, J.; Gerth, K.; Stadler, M.; Müller, R. Pinensins: the first antifungal lantibiotics. Angew. Chem., Int. Ed. 2015, 54, 1125411258,  DOI: 10.1002/anie.201500927
    10. 10
      Smith, T. E.; Pond, C. D.; Pierce, E.; Harmer, Z. P.; Kwan, J.; Zachariah, M. M.; Harper, M. K.; Wyche, T. P.; Matainaho, T. K.; Bugni, T. S. Accessing chemical diversity from the uncultivated symbionts of small marine animals. Nat. Chem. Biol. 2018, 14, 179185,  DOI: 10.1038/nchembio.2537
    11. 11
      Garg, N.; Salazar-Ocampo, L. M. A.; van der Donk, W. A. In vitro activity of the nisin dehydratase NisB. Proc. Natl. Acad. Sci. U.S.A. 2013, 110, 72587263,  DOI: 10.1073/pnas.1222488110
    12. 12
      Ortega, M. A.; Hao, Y.; Zhang, Q.; Walker, M. C.; van der Donk, W. A.; Nair, S. K. Structure and mechanism of the tRNA-dependent lantibiotic dehydratase NisB. Nature 2015, 517, 509512,  DOI: 10.1038/nature13888
    13. 13
      Ortega, M. A.; Hao, Y.; Walker, M. C.; Donadio, S.; Sosio, M.; Nair, S. K.; van der Donk, W. A. Structure and tRNA specificity of MibB, a lantibiotic dehydratase from Actinobacteria involved in NAI-107 biosynthesis. Cell Chem. Biol. 2016, 23, 370380,  DOI: 10.1016/j.chembiol.2015.11.017
    14. 14
      Bothwell, I. R.; Cogan, D. P.; Kim, T.; Reinhardt, C. J.; van der Donk, W. A.; Nair, S. K. Characterization of glutamyl-tRNA-dependent dehydratases using nonreactive substrate mimics. Proc. Natl. Acad. Sci. U.S.A. 2019, 116, 1724517250,  DOI: 10.1073/pnas.1905240116
    15. 15
      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, 14641467,  DOI: 10.1126/science.1121422
    16. 16
      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, 2116921175,  DOI: 10.1074/jbc.M701802200
    17. 17
      Mukherjee, S.; Huo, L.; Thibodeaux, G. N.; van der Donk, W. A. Synthesis and bioactivity of diastereomers of the virulence lanthipeptide cytolysin. Org. Lett. 2016, 18, 61886191,  DOI: 10.1021/acs.orglett.6b03246
    18. 18
      Knerr, P. J.; van der Donk, W. A. Chemical synthesis of the lantibiotic lacticin 481 reveals the importance of lanthionine stereochemistry. J. Am. Chem. Soc. 2013, 135, 70947097,  DOI: 10.1021/ja4014024
    19. 19
      Tang, W.; van der Donk, W. A. The sequence of the enterococcal cytolysin imparts unusual lanthionine stereochemistry. Nat. Chem. Biol. 2013, 9, 157159,  DOI: 10.1038/nchembio.1162
    20. 20
      Tang, W.; Jiménez-Osés, G.; Houk, K. N.; van der Donk, W. A. Substrate control in stereoselective lanthionine biosynthesis. Nat. Chem. 2015, 7, 5764,  DOI: 10.1038/nchem.2113
    21. 21
      Tang, W.; Thibodeaux, G. N.; van der Donk, W. A. The enterococcal cytolysin synthetase coevolves with substrate for stereoselective lanthionine synthesis. ACS Chem. Biol. 2016, 11, 24382446,  DOI: 10.1021/acschembio.6b00397
    22. 22
      Garg, N.; Goto, Y.; Chen, T.; van der Donk, W. A. Characterization of the stereochemical configuration of lanthionines formed by the lanthipeptide synthetase GeoM. Biopolymers 2016, 106, 834842,  DOI: 10.1002/bip.22876
    23. 23
      Lohans, C. T.; Li, J. L.; Vederas, J. C. Structure and biosynthesis of carnolysin, a homologue of enterococcal cytolysin with D-amino acids. J. Am. Chem. Soc. 2014, 136, 1315013153,  DOI: 10.1021/ja5070813
    24. 24
      Sarksian, R.; Hegemann, J. D.; Simon, M. A.; Acedo, J. Z.; van der Donk, W. A. Unexpected methyllanthionine stereochemistry in the morphogenetic lanthipeptide SapT. J. Am. Chem. Soc. 2022, 144, 63736382,  DOI: 10.1021/jacs.2c00517
    25. 25
      Bothwell, I. R.; Caetano, T.; Sarksian, R.; Mendo, S.; van der Donk, W. A. Structural analysis of class I lanthipeptides from Pedobacter lusitanus NL19 reveals an unusual ring pattern. ACS Chem. Biol. 2021, 16, 10191029,  DOI: 10.1021/acschembio.1c00106
    26. 26
      Acedo, J. Z.; Bothwell, I. R.; An, L.; Trouth, A.; Frazier, C.; van der Donk, W. A. O-methyltransferase-mediated incorporation of a β-amino acid in lanthipeptides. J. Am. Chem. Soc. 2019, 141, 1679016801,  DOI: 10.1021/jacs.9b07396
    27. 27
      Marsh, A. J.; O’Sullivan, O.; Ross, R. P.; Cotter, P. D.; Hill, C. In silico analysis highlights the frequency and diversity of type 1 lantibiotic gene clusters in genome sequenced bacteria. BMC Genomics 2010, 11, 679  DOI: 10.1186/1471-2164-11-679
    28. 28
      Walker, M. C.; Eslami, S. M.; Hetrick, K. J.; Ackenhusen, S. E.; Mitchell, D. A.; van der Donk, W. A. Precursor peptide-targeted mining of more than one hundred thousand genomes expands the lanthipeptide natural product family. BMC Genomics 2020, 21, 387  DOI: 10.1186/s12864-020-06785-7
    29. 29
      Bentley, S. D.; Chater, K. F.; Cerdeno-Tarraga, A. M.; Challis, G. L.; Thomson, N. R.; James, K. D.; Harris, D. E.; Quail, M. A.; Kieser, H.; Harper, D. Complete genome sequence of the model actinomycete Streptomyces coelicolor A3(2). Nature 2002, 417, 141147,  DOI: 10.1038/417141a
    30. 30
      Challis, G. L. Exploitation of the Streptomyces coelicolor A3(2) genome sequence for discovery of new natural products and biosynthetic pathways. J. Ind. Microbiol. Biotechnol. 2014, 41, 219232,  DOI: 10.1007/s10295-013-1383-2
    31. 31
      Ozaki, T.; Nishiyama, M.; Kuzuyama, T. Novel tryptophan metabolism by a potential gene cluster that is widely distributed among actinomycetes. J. Biol. Chem. 2013, 288, 99469956,  DOI: 10.1074/jbc.M112.436451
    32. 32
      Nodwell, J. R. Microbe profile: Streptomyces coelicolor: a burlesque of pigments and phenotypes. Microbiology 2019, 165, 953955,  DOI: 10.1099/mic.0.000821
    33. 33
      Lautru, S.; Deeth, R. J.; Bailey, L. M.; Challis, G. L. Discovery of a new peptide natural product by Streptomyces coelicolor genome mining. Nat. Chem. Biol. 2005, 1, 265269,  DOI: 10.1038/nchembio731
    34. 34
      Song, L.; Barona-Gomez, F.; Corre, C.; Xiang, L.; Udwary, D. W.; Austin, M. B.; Noel, J. P.; Moore, B. S.; Challis, G. L. Type III polyketide synthase beta-ketoacyl-ACP starter unit and ethylmalonyl-CoA extender unit selectivity discovered by Streptomyces coelicolor genome mining. J. Am. Chem. Soc. 2006, 128, 1475414755,  DOI: 10.1021/ja065247w
    35. 35
      Lin, X.; Hopson, R.; Cane, D. E. Genome mining in Streptomyces coelicolor: molecular cloning and characterization of a new sesquiterpene synthase. J. Am. Chem. Soc. 2006, 128, 60226023,  DOI: 10.1021/ja061292s
    36. 36
      Corre, C.; Song, L.; O’Rourke, S.; Chater, K. F.; Challis, G. L. 2-Alkyl-4-hydroxymethylfuran-3-carboxylic acids, antibiotic production inducers discovered by Streptomyces coelicolor genome mining. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 1751017515,  DOI: 10.1073/pnas.0805530105
    37. 37
      Wu, C.; Ichinose, K.; Choi, Y. H.; van Wezel, G. P. Aromatic polyketide GTRI-02 is a previously unidentified product of the act gene cluster in Streptomyces coelicolor A3(2). ChemBioChem 2017, 18, 14281434,  DOI: 10.1002/cbic.201700107
    38. 38
      Li, B.; Cooper, L. E.; van der Donk, W. A. In Vitro Studies of Lantibiotic Biosynthesis. In Methods Enzymol.; Elsevier, 2009; Vol. 458, pp 533558.
    39. 39
      Küsters, E.; Allgaier, H.; Jung, G.; Bayer, E. Resolution of sulphur-containing amino acids by chiral phase gas chromatography. Chromatographia 1984, 18, 287293,  DOI: 10.1007/BF02259079
    40. 40
      Liu, W.; Chan, A. S. H.; Liu, H.; Cochrane, S. A.; Vederas, J. C. Solid supported chemical syntheses of both components of the lantibiotic lacticin 3147. J. Am. Chem. Soc. 2011, 133, 1421614219,  DOI: 10.1021/ja206017p
    41. 41
      Zallot, R.; Oberg, N.; Gerlt, J. A. The EFI web resource for genomic enzymology tools: Leveraging protein, genome, and metagenome databases to discover novel enzymes and metabolic pathways. Biochemistry 2019, 58, 41694182,  DOI: 10.1021/acs.biochem.9b00735
    42. 42
      Gerlt, J. A.; Bouvier, J. T.; Davidson, D. B.; Imker, H. J.; Sadkhin, B.; Slater, D. R.; Whalen, K. L. Enzyme Function Initiative-Enzyme Similarity Tool (EFI-EST): A web tool for generating protein sequence similarity networks. Biochim. Biophys. Acta, Proteins Proteomics 2015, 1854, 10191037,  DOI: 10.1016/j.bbapap.2015.04.015
    43. 43
      Le, T.; Fouque, K. J. D.; Santos-Fernandez, M.; Navo, C. D.; Jiménez-Osés, G.; Sarksian, R.; Fernandez-Lima, F. A.; van Der Donk, W. A. Substrate sequence controls regioselectivity of lanthionine formation by ProcM. J. Am. Chem. Soc. 2021, 143, 1873318743,  DOI: 10.1021/jacs.1c09370
    44. 44
      Bobeica, S. C.; Zhu, L.; Acedo, J. Z.; Tang, W.; van der Donk, W. A. Structural determinants of macrocyclization in substrate-controlled lanthipeptide biosynthetic pathways. Chem. Sci. 2020, 11, 1285412870,  DOI: 10.1039/D0SC01651A
    45. 45
      Altschul, S. F.; Madden, T. L.; Schaffer, A. A.; Zhang, J.; Zhang, Z.; Miller, W.; Lipman, D. J. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997, 25, 33893402,  DOI: 10.1093/nar/25.17.3389
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acschembio.2c00492.

    • Glutamyl lyase SSN (ZIP)

    • Experimental procedures; extended experimental data and bioinformatic analysis; materials and methods; primers used in this study; observed and calculated m/z ratios for fragments of GluC-digested mCoiA1; NEM alkylation assay for CoiA1 co-expressed with CoiB-H994A, CoiC, and CoiSA(ED); LC-MS analysis of the GluC digestion product of mCoiA1 and variants; description table of orthologs obtained from PSI-BLAST of OlvA (PDF)

    Accession Codes

    NCBI protein accessions: CoiA1, WP_011031313.1; CoiB, WP_011031312.1; CoiC, WP_011031311.1; and CoiSA, WP_011031310.1.


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.