Abstract
Emerging antibiotic resistance requires continual improvement in the arsenal of antimicrobial drugs, especially the critical macrolide antibiotics. Formation of the macrolactone scaffold of these polyketide natural products is catalyzed by a modular polyketide synthase (PKS) thioesterase (TE). The TE accepts a linear polyketide substrate from the termina PKS acyl carrier protein to generate an acyl-enzyme adduct that is resolved by attack of a substrate hydroxyl group to form the macrolactone. Our limited mechanistic understanding of TE selectivity for a substrate nucleophile and/or water has hampered development of TEs as biocatalysts that accommodate a variety of natural and non-natural substrates. To understand how TEs direct the substrate nucleophile for macrolactone formation, acyl-enzyme intermediates were trapped as stable amides by substituting the natural serine OH with an amino group. Incorporation of the unnatural amino acid, 1,3-diaminopropionic acid (DAP), was tested with five PKS TEs. DAP-modified TEs (TEDAP) from the pikromycin and erythromycin pathways were purified and tested with six full-length polyketide intermediates from three pathways. The erythromycin TE had permissive substrate selectivity, whereas the pikromycin TE was selective for its native hexaketide and heptaketide substrates. In a crystal structure of a native substrate trapped in pikromycin TEDAP, the linear heptaketide was curled in the active site with the nucleophilic hydroxyl group positioned 4 Å from the amide-enzyme linkage. The curled heptaketide displayed remarkable shape complementarity with the TE acyl cavity. The strikingly different shapes of acyl cavities in TEs of known structure, including those reported here for juvenimicin, tylosin and fluvirucin biosynthesis, provide insights to facilitate TE engineering and optimization.
Keywords: Macrolactonizing enzyme, Thioesterase, Polyketide synthase, Substrate trapping, X-ray crystallography, Unnatural amino acid incorporation
Introduction
The number of pathogenic bacteria and fungi with evolved resistance to antimicrobial drugs is growing rapidly due to widespread clinical and agricultural use of antibiotics. This looming crisis places great urgency on the discovery and development of new pharmaceuticals. Macrolide antibiotics are an essential part of the antimicrobial arsenal, hence continual improvement in macrolide potency is essential to fight the ongoing war with pathogenic microbes. Exploiting natural biosynthetic systems, particularly for the challenging step of macrocycle formation, is an appealing approach to this problem1.
Bacterial type I modular polyketide synthases (mPKS) are nature’s biosynthetic machines for the macrolactone scaffold of macrolide antibiotics2 as well as for many anticancer drugs, 3 fungicides4 and insecticides5. The mPKS biosynthetic pathways for macrolactones are organized as a sequence of multi-enzyme modules, each consisting of domains that carry, extend, modify and ultimately offload polyketides. An acyltransferase (AT), which may reside within or outside the module, selects a specific acyl-extender unit from the co-enzyme A (CoA) pool and transfers it to the phosphopantetheine (Ppant) prosthetic group of the module acyl carrier protein (ACP). Within the module, a ketosynthase (KS) catalyzes the decarboxylative Claisen condensation of the extender unit with the polyketide intermediate of the upstream module. Optional ketoreductase (KR), dehydratase (DH) and enoylreductase (ER) domains may modify the β-ketone to form a β-hydroxy, α -ene or fully reduced module product. Finally, at the terminus of the mPKS pathway, a thioesterase (TE) offloads the full-length polyketide product by cyclization to a macrolactone or hydrolysis to a linear carboxylic acid.
The formation of macrocycles by TEs during polyketide chain termination is of great importance to the development of biocatalysts and to other applications of pathway engineering. As catalysts of the final step in the PKS assembly line, the offloading TEs are essential for overall pathway flux and act as the final quality check prior to release of the polyketide product6. Some pathways also include a type II TE, thought to serve a proofreading function7,8. The macrolactonizing TEs form a wide range of natural products, and some can also cyclize non-natural polyketide substrates9–11. Synthetic approaches to lactonization of large substrates with numerous functional groups and chiral centers is generally challenging and low-yielding12,13. Therefore, engineering TEs to broaden substrate scope or more finely control macrolactone production represents an exciting addition to the biocatalytic toolkit.
The basis for cyclization vs. hydrolysis by mPKS TEs (Fig. 1A) is poorly understood6,14 despite published structures of three macrolactone-forming TE domains, including the DEBSIII TE (erythromycin, 14-membered macrolactone 6-deoxyerythronolide B (6-deB)) 15,16, PikAIV TE (pikromycin, 12-membered 10-deoxymethynolide (10-dml) or 14-membered narbonolide) 17,18, and PimS4 TE (pimaricin, 26-membered polyene ring) 19; one hydrolyzing TE (TmcB TE, linear product leading to tautomycetin) 20; and one decarboxylating TE (CurM TE, linear product curacin A) 21. Biochemical study of DEBSIII TE and PikAIV TE revealed some substrate selectivity in macrolactone formation10,11,17,22–25. While tolerant of variation in substrate length and methyl or hydroxy substituents, DEBSIII TE was selective for the configuration of the nucleophilic alcohol6,25. Similarly, PikAIV TE tolerated certain changes in substrate length but required a C9 ketone to generate a macrolactone via the terminal hydroxy nucleophile at C1317,22 (Fig. 1A). Furthermore, cyclization by the PikAIV TE, like DEBSIII TE, is selective for one configuration of the nucleophilic hydroxyl10. These and other results11 implicate TEs as crucial quality-control gatekeepers for module processing of natural and unnatural substrates.
Figure 1.

Macrolactone formation by offloading PKS thioesterases. A. Schematic of PikAIV TE offloading reaction with (1) macrolactone or (2) hydrolysis products. B. Perpendicular views of the active site tunnel in a PKS offloading TE dimer. The product 10-dml (stick form) is inside the acyl cavity in the surface rendering of the PikAIV TE (PDB 2HFK). The top image is viewed through the tunnel exit, and the bottom image is a cutaway from the side. C. Helices in the TE dimer interface. Helices α1-α2 form all dimer contacts and stabilize the active site tunnel through intra-subunit contacts with α6 and α7. Monomers are in contrasting colors. D. Unnatural amino acid deprotection scheme leading to DAP, followed by amide formation from a polyketide-thioester.
The offloading TEs in mPKS, nonribosomal peptide synthetase26 (NRPS) and fungal iterative PKS27 (iPKS) pathways are homologs within the α/β-hydrolase superfamily. They have remarkably similar structures even though they act on divergent substrates and have distinct macrolactone15,17, carboxylic acid28, macrolactam29, or decarboxylated21 products. The mPKS TEs are dimers with independent active sites. Within each monomer, a canonical catalytic triad (Ser or Cys nucleophile, His, Asp) is located at the center of a tunnel in which the cognate ACP can deliver the polyketide substrate through one end, and the offloaded product can exit through the opposite end (Fig. 1B). TE dimer formation is critical to the integrity of the active site tunnel. Two N-terminal α-helices (α1-α2) form all inter-subunit dimer contacts and stabilize the tunnel through intra-subunit contacts with helices α6-α7 (Fig. 1C). The tunnel architecture is unlike the monomeric TEs where a mobile lid, topologically equivalent to helices α6-α7, opens and closes over the active site, and no tunnel exists30. Helices α6-α7 have been described as a “lid” in dimeric PKS TEs6, but this lid is kept closed by helices α1-α2 in the dimer interface.
The exit end of the active site tunnel forms a spacious acyl cavity large enough to accommodate the polyketide substrate as well as the macrolactone or linear carboxylic acid product. The shape and extent of the acyl pocket (helices α2, α6, α7) is the most variable region in TE structures15,17,19, and includes several positions where sequences vary (Fig. S1A). No significant structural differences exist between mPKS TE free enzymes and their corresponding substrate, product or analog complexes17,19,31. Thus, substrate selectivity and catalytic outcome – macrolactone formation vs. hydrolysis – are not driven by conformational changes to the active site tunnel or acyl cavity but rather by other TE features such as the shape and chemical environment of the acyl cavity, the substrate structure, and the length, position and amphipathic profile of the lid region6,15,17,19. Visualization of active sites mimicking the covalent acyl-enzyme intermediate are key to a deeper understanding of the various parameters that control product outcome.
The intrinsic lability of the TE covalent intermediate – acyl-Ser ester or acyl-Cys thioester – has been a hurdle to direct visualization of TE-polyketide interactions. We previously used diphenyl-activated polyketide phosphonates to trap the PikAIV TE active site with phosphonate adducts that mimic the tetrahedral intermediate of acyl-enzyme formation17,31. However, the bulky diphenylphosphonate prevented full modification of the nucleophilic amino acid in the TE active site tunnel17. Here we adopted a recently developed platform to incorporate an amino functionality in place of the active site Ser hydroxyl or Cys thiol32. The unnatural amino acid (UAA) 2,3-diaminopropionic acid (DAP) is an isostere of Ser and Cys that can trap acyl-enzyme intermediates as stable amides (Fig. 1D). We tested DAP replacement for the nucleophilic amino acids in the offloading TE domains of five actinobacterial mPKS pathways (Fig. S1B). Amide trapping was achieved for two TEs. A crystal structure of the PikAIV TEDAP with the natural heptaketide illustrates how the structure of both the enzyme and the substrate contribute to the high fidelity of macrolactone formation. The acyl cavities of a panel of offloading PKS TEs, including three with structures reported here, reveal an overall correlation with substrate and product structures.
Results
Panel of PKS TE domains.
The offloading TE domains of five actinobacterial type I PKS pathways were investigated for the possibility of replacing the nucleophilic Ser or Cys with the DAP isostere (Figs. 1D, S1): PikAIV TE (Streptomyces venezuelae ATCC 15439), DEBSIII TE (Saccharopolyspora erythrea), JuvEV TE (Micromonospora chalcea ssp. izumensis), TylG5 TE (Streptomyces fradiae) and FluC TE (Actinomadura vulgaris). We first purified all five wild-type TE domains (TEWT) to homogeneity and confirmed the expected dimeric state in solution (apparent molecular weight 55-65 kDa). The five TE domains are homologs with 29% – 45% pairwise sequence identity, except the JuvEV and TylG5 TEs, which produce the same tylactone product and have 64% identical sequences.
DAP incorporation in PKS TE domains.
DAP incorporation in our panel of five TEs followed by reaction with linear polyketide intermediates or macrolactone products provided a promising platform to probe the structural basis of macrolactone formation and substrate selectivity by the TEs. For DAP incorporation, the catalytic Ser or Cys codon in each TE-bearing plasmid was replaced with an amber stop codon, followed by co-expression with a plasmid encoding an evolved pyrrolysyl-tRNA synthetase (DAPRS) and a suppressor tRNA (tRNAAmb) 32. DAPRS charges the tRNAAmb with a DAP variant (doubly protected DAP-EOS-NBO, Figs. 1D, S2).
To optimize conditions for production of recombinant N-terminal His6-tagged TEDAP in E. coli cells, we screened vector backbone (pMCSG733 and derivatives pCDF7 and pTrc7), DAP-EOS-NBO concentration (1-10 mM), and culture temperature (18°C, 20°C, 25°C), and used SDS-PAGE to estimate the relative proportions of full-length TE (~30 kDa) and truncated TE (~16 kDa) resulting from non-incorporation of DAP-EOS-NBO. We identified conditions to produce useful quantities of PikAIV TEDAP and DEBSIII TEDAP (Fig. S3) with yields 10-25 times lower than the corresponding TEWT (9.1 mg/L vs. 100 mg/L for PikAIV TE; 7.2 mg/L vs. 130 mg/L for DEBSIII TE). No combination of conditions yielded sufficient quantities of TEDAP for the JuvEV, FluC or TylG5 TEs, for which yields of the wild type proteins were far below the high levels for the PikAIV and DEBSIII TEs.
In forming TEDAP, the initial UAA was the doubly protected DAP-EOS-NBO (TEDAP+297.03 Da, Fig. 1D). Deprotection involved photo cleavage (365 nm) to the singly protected DAP-EOS (TEDAP+103.99 Da) followed by spontaneous deprotection to the active DAP (Fig. S4). The spontaneous deprotection step was slow for the PKS TEDAP domains, perhaps due to slow solvent exchange in the substrate tunnel. We detected all three states of PikAIV and DEBSIII TEDAP (TEDAP-EOS-NBO, TEDAP-EOS, TEDAP) by liquid chromatography / mass spectrometry (LC/MS) analysis of the intact protein in fresh preparations (Fig. S4). Complete photo deprotection occurred with UV exposure immediately following the Ni-affinity purification step. Conditions for the second spontaneous deprotection (temperature, time, pH, addition of reducing agent) were optimized to generate purified PikAIV TEDAP and DEBSIII TEDAP in the fully deprotected state.
Covalent trapping of polyketides.
We next characterized the reactivity of PikAIV and DEBSIII TEDAP with linear polyketide substrates: thiophenol-activated forms of the pikromycin pentaketide34 and 3- methoxy hexaketide35, and evaluated the formation of stable adducts using intact-protein LC/MS. For these linear substrates, PikAIV TEDAP was fully modified with both adducts (Figs. 2A, S5A,B). DEBSIII TEDAP was fully modified by the pikromycin pentaketide and modified at a lower level by the pikromycin hexaketide (Figs. 2A, S6A,B).
Figure 2.

Polyketide modification of TEDAP. A. Quantitation of adduct formation for PikAIV and DEBSIII TEDAP with a panel of linear and cyclic polyketides in a 24-hr reaction at room temperature. Values are based on peak areas in deconvoluted mass spectra for the intact proteins. Variances are based on triplicate measurements; ND = not detected. B. PikAIV TEDAP adducts from narbonolide. TEDAP opened the 14-membered macrolactone ring, trapping the natural heptaketide as a stable, linear acyl-enzyme adduct. Deconvoluted mass spectra show the incorporation of heptaketide at DAP1196. The dehydrated pyran product resulting from spontaneous hemiketal formation also accumulated. C. DEBSIII TEDAP adduct from 6-dEB. TEDAP opened the 14-membered macrolactone ring, trapping the pyran form of the natural DEBS heptaketide as a stable acyl enzyme adduct. Deconvoluted mass spectra show the incorporation of the heptaketide at DAP3043.
In addition, we predicted that, like the wild type, TEDAP could open cyclic macrolactone products (Fig. S1B), yielding linear TEDAP-polyketide adducts (Figs. 2, S5C–E, S6C–E). Thus, we tested four macrolactones (6-dEB36, narbonolide34, tylactone37 and 10-dml) using the same protocol as for the linear substrates. PikAIV TEDAP was fully modified by narbonolide and partially modified by 10-dml, but we detected no modification using either 6-dEB or tylactone. In contrast, DEBSIII TEDAP was fully modified by opening of the non-native narbonolide, modified at lower levels with the native 6-dEB macrolactone or tylactone, but was not modified by 10-dml. The lack of reactivity of PikAIV TEDAP with 6-dEB and tylactone and of DEBSIII TEDAP with 10-dml was not due to the weaker nucleophilicity of DAP relative to the native Ser. The TEWT enzymes also had no detectable reactivity with these substrates. Upon ring opening, narbonolide, 10-dml, 6-dEB and tylactone can spontaneously convert to the corresponding pyran through reversible hemiketalization followed by irreversible dehydration, as shown for an intermediate in tylactone biosynthesis11. We detected both linear and pyran forms for the ring-opened adducts of narbonolide and 10-dml, but only the pyran forms of 6-dEB and the tylosin octaketide (Figs. 2, S5C, S6C–D).
Thus, PikAIV TEDAP was highly selective for native polyketides, but DEBSIII TEDAP exhibited a broader substrate tolerance (Fig. 2A). The reactivity of DEBSIII TEDAP is consistent with early studies demonstrating that wild-type DEBSIII TE can form macrolactones smaller and larger than the natural 14-membered 6-dEB23,38–40.
Structure of natural heptaketide adduct of PikAIV TEDAP.
PikAIV and DEBSIII TEDAP were screened in crystallization experiments, and crystals were obtained for PikAIV TEDAP. Following proteolytic cleavage of the N-terminal His6 tag, PikAIV TEDAP crystallized in a form unrelated to that of the previously reported structures17. Crystallization conditions for the new form were optimized with PikAIV TEWT and then applied to the TEDAP-heptaketide. Crystallization of PikAIV TEDAP-heptaketide followed incubation of TEDAP with narbonolide, the natural pikromycin 14-membered macrolactone product, under conditions leading to complete modification with the heptaketide (+352.49 Da), as determined by intact protein LC/MS analysis (Fig. 2A). The PikAIV TEDAP-heptaketide yielded a 2.8-Å structure (Table S1). We also solved a 3.1-Å structure of PIKAIV TEDAP-EOS-NBO from partially deprotected protein (Table S1).
The linear heptaketide adduct at DAP1196 is fully resolved in the active sites of both monomers in the TEDAP-heptaketide crystals with no indication of the pyran form, which may have been disfavored by the high pH (8.5) of crystallization (Figs. 3A, S7). The heptaketide adopts a U-shaped conformation in the acyl cavity, with the distal hydroxyl (O13) ideally situated for attack at the acyl enzyme linkage, here a stable amide. Hydroxyl O13 is within ~4 Å of the DAP amide analog of the TEWT-heptaketide ester. The natural heptaketide is positioned similarly to the phosphonate pentaketide mimic we previously trapped in the PikAIV TE active site (Figs. 3B, S7D) 17. The side chain hydroxyl of Thr1125 forms the sole hydrogen bond to the O9 carbonyl of the heptaketide mimic and the O9 hydroxyl of the pentaketide mimic. The acyl cavity is filled with the heptaketide, with no space for water of hydrolysis. This contrasts with the pentaketide phosphonate analog, where the free end of the adduct exhibited greater flexibility (weaker electron density) within the acyl cavity17.
Figure 3.

Crystal structure of PikAIVDAP-heptaketide. The heptaketide is curled in the acyl cavity with the substrate nucleophile (O13) near the acyl-enzyme linkage, here a stable amide. A. Electron density (3σ Fo-Fc omit) for the heptaketide adduct at DAP1196. TEDAP is rendered in cartoon form with the catalytic triad (DAP1196, Asp1224, His1316) as sticks with atomic coloring (white C for DAP-heptaketide, yellow C for TE, red O, blue N). B. Overlay of PikAIV TEDAP-heptaketide (yellow protein) with TE-pentaketide phosphonate mimic of the tetrahedral intermediate of acyl-enzyme formation (cyan protein, PDB 2HFJ). C. Shape complementarity of TE acyl cavity and heptaketide substrate. The image is a slice through the widest region of the acyl cavity (entrance from the left, exit to the right) illustrating the surfaces of enzyme and substrate.
The acyl cavity was virtually unchanged by formation of the heptaketide adduct. Only a few side chains (Tyr1073, Leu1245, His1316, Phe1317) shifted slightly (<2 Å) from positions in the free enzyme to make space for the adduct. A remarkable hand-in-glove shape complementarity of the natural heptaketide and the acyl cavity (Fig. 3C) explains the high fidelity of macrolactone formation by the TE. This also explains the selectivity of PikAIV TE for natural substrates (Fig. 2A, Fig. S5). For example, the bulky Tyr1073 side chain would interfere with binding of the DEBS pathway heptaketide with its methyl substituent at C10. The tight fit also explains the stereoselectivity of the TE for the heptaketide O13 hydroxyl nucleophile as well as the inability of the tylosin octaketide, conformationally restricted by three double bonds in resonance, to occupy the active site. Contacts of the TE with the natural pikromycin heptaketide are predominantly hydrophobic (Fig. S7A,B; Tyr1073 and Leu1077 in helix α2; Ala1126 in the TE core domain; Gln1231, Ile1234, Leu1241 and Leu1245 in helix α6; Met1261, Ala1265 and Leu1268 in helix α7; and Phe1317 adjacent to catalytic triad His1316). In contrast to these hydrophobic amino acids, the exit end of the TE tunnel is replete with polar side chains (Asp1071, Asp1078, Gln1231, Glu1235, Ser1238, Arg1266). These side chains form an effective “hydrophilic barrier” to an extended conformation of the of the heptaketide17 (Fig. S7C).
Acyl cavities in TE domains.
For the PikAIV and DEBSIII TEs, several high-resolution crystal structures have been reported15–17, but these are lacking for the other three TEs in our panel. In order to compare acyl cavities, we solved crystal structures for the other three wild-type TEs, JuvEV TEWT, TylG5 TEWT and FluC TEWT (Table S1). These TEs possess the canonical α/β-hydrolase fold and other features common to PKS TE domains15,17,19,20 (Fig. 4). As in other offloading PKS TEs, two N-terminal α-helices form the dimer interface, and the catalytic triad is located at the center of a tunnel allowing ACP entry through one end and macrolactone or linear product formation in an acyl cavity at the opposite end. The JuvEV and TylG5 TE active sites contain a Ser-His-Asp catalytic triad (JuvEV Ser1637, Asp1664, His1756; TylG5 Ser1681, Asp1708, His1800), while the FluC TE has a rare Cys nucleophile (Cys2243, Asp2270, His2367).
Figure 4.

Structures of mPKS offloading TE domains determined in this study. A. JuvEV TE. B. FluC TE. C. TylG5 TE. Each TE is viewed along the dimer twofold with spheres indicating the polypeptide termini (blue N-terminus, red C-terminus). Amino acid side chains of each catalytic triad are shown as spheres (gray C, red O, blue N, yellow S).
We compared the acyl cavities of the seven offloading PKS TEs of known structure, including the five mentioned above and those from biosynthetic pathways for pimaricin (PimS4 TE) 19 and tautomycetin (TmcB TE) 20 (Fig. 5). Only the PikAIV and PimS4 TE crystal structures included a substrate or product. For the other cyclizing TEs, we positioned small-molecule crystal structures of the macrocyclic products41–43 in the acyl cavities of the corresponding TEs. Each acyl cavity has a shape and surface generally complementary to the product shape and surface, although the fits are not as snug as for the heptaketide trapped in PikAIV TEDAP. The interior surfaces of the acyl cavities are predominantly hydrophobic near the catalytic nucleophiles (Ser or Cys) and more polar at the tunnel exits. With the exception of the PimS4 TE, the macrocycle-forming TEs (Fig. 5A–E) have acyl cavities wide enough for the a linear polyketide to curl and bring the internal nucleophile near the acyl-enzyme linkage. Variability among the amino acid sequences and small differences in helix positions relative to the catalytic triad create slightly different shapes and orientations of the acyl cavities within these otherwise highly similar TE structures. The JuvEV and TylG5 TEs both produce tylactone (Fig. 5C,D), but a few sequence differences create somewhat different cavity shapes and, unlike the TylG5 TE, the exit half of the JuvEV TE acyl cavity has varying degrees of flexibility among the four independent views of the polypeptide in the crystal structure.
Figure 5.

Acyl cavities in structures of mPKS offloading TEs. A. PikAIV TEDAP (this study). B. DEBSIII TE (PDB 5D3K, modeled 6-dEB). C. JuvEV TE (this study, modeled tylactone). D. TylG5 TE (this study, modeled tylactone). E. FluC TE (this study, modeled fluvirucin aglycone). F. PimS4 TE (PDB 7VO5), a polyene TE. G. TmcB TE (PDB 3LCR, modeled product), which produces a linear polyketide. Cut-away views of the TE monomer structures show the active site tunnels in the viewing plane, with entrance from the left and exit to the right. The dimer interface is at the top. Acyl cavities are highlighted as semi-transparent surfaces Pik heptaketide in A and macrocyclic products in B-F. Products are adjacent to each TE; tylactone is the product of both JuvEV TE and TylG5 TE.
In contrast to the other cyclizing TEs, the PimS4 TE acyl cavity accommodates the 26-membered macrolactone product, but is too narrow for the polyene substrate to enter in an extended conformation and curl within the cavity (Fig. 5F). A pyranose ring at C13-C17 creates a bend in the substrate, suggesting it may enter the active site in a pre-curled form. The TmcB TE, the only TE in our panel that forms a linear product, has the narrowest acyl cavity and could not fit a curled substrate (Fig. 5G). Compared to the other PKS offloading TEs, the acyl cavities in the pimaricin and tautomycetin TEs are longer due to the absence of a bend at the start of helix α6 and narrower due to a shifted positions for helices α2 and α6. In the PimS4 TE, the product protrudes from the cavity exit, as would the undecaketide TmcB TE substrate.
Discussion
Although first demonstrated with an NRPS TE, our study demonstrates the broad utility of the DAP unnatural amino acid to probe the substrate binding and selectivity of enzymes where a covalent intermediate forms at a side-chain hydroxyl (Ser, Thr) or thiol (Cys) nucleophile. The DAP incorporation reported herein enabled the first view of a PKS TE native-substrate acyl-enzyme. The structure of the near-natural catalytic intermediate of PikAIV TE provides mechanistic information about the critical cyclization step in macrolide antibiotic biosynthesis. The structure of the PikAIV TEDAP-heptaketide also illustrates potential challenges in developing/adapting biocatalysts to form macrolactones.
The natural linear heptaketide substrate in PikAIV TEDAP is poised for macrolactone formation (Fig. 3). The substrate binds in a curled conformation that is highly complementary to the shape of the acyl cavity. This exquisite fit places the substrate nucleophile (O13 hydroxyl) in an ideal position for attack on the acyl enzyme linkage and subsequent macrolactone formation.
We found a variety of shapes and orientations for the acyl cavities in other offloading PKS TEs (Fig. 5), consistent with the sequence variability among amino acids lining the cavity and slight differences in helix positions relative to the catalytic triads. All the acyl cavities have a strongly hydrophobic surface in the region adjacent to the catalytic triad despite the variation in shape and orientation. In TEs that form 12-16-membered macrocycles (PikAIV, DEBSIII, TylG5, JuvEV, FluC), a bent helix α6 and a cluster of polar side chains at the acyl cavity exit induce the linear polyketide substrate to curl, bringing the internal nucleophile near the acyl-enzyme linkage (Fig. 5A–E). The cyclized products of these TEs fit within the non-polar region of the acyl cavities. Thus, we conclude that a “hydrophilic barrier” at the cavity exit17 assists catalysis where a flexible substrate must curl to position the internal nucleophile near the catalytic center. The shapes of the macrocyclic products also differ according to their double-bond content, substituents and chiral centers. Among the TEs that form 12-16-membered macrocycles, the seemingly subtle differences in acyl cavity shape must be well matched to the substrate to ensure efficient cyclization and avoid hydrolysis to the linear product. In contrast, the 26-membered polyene macrolactone product fills the acyl cavity and protrudes from the tunnel exit of the PimS4 TE. Here the pyranose ring may pre-form the substrate for cyclization (Fig. 5F) 19. The tautomycetin TE produces a linear product and has an acyl cavity that is too slender for the polyketide to curl towards the acyl-enzyme linkage (Fig. 5G) 20. The four hydroxyl substituents at the thioester-distal end of the TmcB TE polyketide are compatible with substrate extension into the polar exit end of the acyl cavity.
PikAIV TE can form a 14-membered macrolactone from the natural heptaketide intermediate as well as a 12-membered macrolactone from the natural hexaketide (Fig. S1B), both in vitro and in the Streptomyces producer44. The tight fit of natural substrate to acyl cavity may account for this rare dual-macrolactonization capability, but it also limits the substrate scope for macrolactone formation. The cavity shape appears unable to accommodate additional substituents (Fig. 3C), for example the methyl at C10 in the DEBSIII TE substrate (Fig. S5D). Previous studies with variants of the natural hexaketide revealed an exquisite sensitivity to aspects of the substrate structure, leading to characterization of PikAIV TE as a gatekeeper for pathway fidelity10. Removal of methyl or ethyl substituents reduced the proportion of macrolactone relative to linear product, and inversion of either or both of the C12 (methyl substituent) and C13 (hydroxyl) chiral centers resulted in only linear products10. Reduction of the C9 carbonyl relaxes a conformational constraint (resonance of the carbonyl with the double bond at C10), but abolished macrolactone formation22.
In contrast to the PikAIV TEDAP, the DEBSIII TEDAP had a broader substrate scope (Fig. 2A). DEBSIII TEDAP reacted with the natural 6-dEB product and with non-natural macrolactones from the Pik and Tyl systems as well as Pik pathway intermediates, consistent with its reported production of macrocycles of various size23,38–40. Nevertheless, the DEBSIII TE reactivity is limited, as the linear product predominated if the chirality of the substrate nucleophile was inverted in a simplified unnatural substrate25. Additionally, we detected no ability of DEBSIII TEWT or TEDAP to catalyze rebound ring opening of the 12-membered macrolactone of the PikAIV TE product 10-dml (Fig. S6E).
TE formation of a macrolactone as opposed to a linear seco-acid product is the result of an internal substrate nucleophile out-competing water to resolve the acyl-enzyme. Poor positioning of the substrate nucleophile or substrate dynamics in the active site would disadvantage macrolactone formation. The observation of reduced or abolished macrolactone formation with non-natural substrates that are slightly misfit or more flexible in the acyl cavity highlights the kinetics of the competition with water – which nucleophile reaches the acyl-enzyme linkage first? In a previous study of the PikAIV TE12, substitution of the catalytic Ser with the stronger nucleophile Cys (TECys) increased the reaction rate (kcat) revealing a role for kinetics in reaction outcome12. Computed reaction trajectories for TECys indicated an altered kinetic pathway compared to TEWT12. The increased kcat of TECys tipped the balance towards cyclization and away from hydrolysis for a non-natural stereodivergent (hexaketide C-11 OH) substrate where TEWT catalyzed little or no cyclization12,45. The observation that greater speed promotes cyclization suggests that in the early enzyme-substrate complex the polyketide is curled in the active site with the substrate nucleophile poised exclusively for macrocyclization (outcompeting water).
Conclusion
Our study highlights the influence of acyl cavity shape on substrate selectivity and catalytic outcome for TEs that catalyze macrolactonization, thereby identifying the challenge of matching a substrate-enzyme pair to a desired macrolactone product. Structures of acylated forms of additional enzymes together with biochemical data and the use of a stronger Cys nucleophile could address/mitigate this challenge. This study is an important step towards realizing the promise of TE catalysts in development of improved macrolide antibiotics using native or engineered enzymes.
Methods
Construction of expression plasmids (Table S2).
DNA encoding JuvEV TE (amino acids 1508-1778, ARW71487) was subcloned from pET21b-JuvEV37 and inserted into pMCSG733 using ligation independent cloning (LIC) to create expression plasmid pTMM19. PikAIV TE (1057-1346, AAC69332) was subcloned from pCAte222 and inserted into pMCSG7 using Gibson assembly to create expression plasmid pTMM40. DNA encoding DEBSIII TE (2904-3167, AAV51822) was subcloned from pRSG3346 and inserted into pMCSG7 using Gibson assembly to create pTMM39. Synthetic DNA encoding the FluC TE (2106-2395, AFU66012) 47 was inserted into pMCSG7 using Gibson assembly to create expression plasmid pTMM41. DNA encoding TylG5 TE (1549-1841, AAB66508) was subcloned from pDHS300748 and inserted into pMCSG7 using Gibson assembly to create expression plasmid pTMM42.
For unnatural amino acid (UAA) incorporation, amber stop codons (UAG) were introduced into all TE expression plasmids in place of codons for the catalytic Ser or Cys residues using the QuikChange XL Site-Directed Mutagenesis kit (Agilent), optimized for GC-rich sequences (JuvEV TE S1637amb: pTMM47, PikAIV TE S1196amb: pTMM48, DEBSIII TE S3030amb: pTMM49, FluC TE C2243amb: pTMM50, TylG5 TE S1681amb: pTMM51). For small-scale optimization experiments with the UAA, sequences encoding all TE domains were amplified from amber codon-containing expression plasmids (pTMM47-51) and then inserted into pCDF7 (Novagen) vectors with identical construct boundaries (pTMM52-56), again using Gibson assembly. The pCDF7 expression vector encodes an N-terminal His6 tag but with a different selection marker (spectinomycin) and an origin of replication (p15A) compatible with the UAA system (pSF-DAPRS-PylT).
All primers are listed in Table S3. All constructs and mutations were verified by using the University of Michigan Sanger Sequencing Core, or by nanopore sequencing (Plasmidsaurus).
Production of wild type proteins.
All wild type TE domains were produced and purified in a similar manner. JuvEV TE, PikAIV TE and DEBSIII TE were purified in HEPES buffer. FluC TE and TylG5 TE were purified in Tris buffer. Escherichia coli BL21(DE3) cells were transformed with expression plasmids for PikAIV TE, DEBSIII TE or FluC TE. E. coli strain BL21(DE3) pRare249 was transformed with the JuvEV TE plasmid and E. coli BL21(DE3) pGro750 with the TylG5 TE plasmid. Cells were grown in 0.5 L of TB media with 100 μg/mL ampicillin to an OD600=1.5. Cultures producing JuvEV TE also contained 50 μg/mL spectinomycin (pRare2), and those producing TylG5 TE also contained 35 μg/mL chloramphenicol (pGro7). After reaching OD600=1.5, cultures were cooled to 20°C over 1 hr, induced with 200 μM IPTG and grown for 16 hr. Cultures producing JuvEV TE or TylG5 TE also had 1 g/L L-(+)-arabinose at induction. Cells were harvested by 30-min centrifugation at 8,000 rpm and cell pellets were stored at −20°C.
Cell pellets were resuspended in 35 mL lysis buffer (50 mM HEPES pH 7.4 or 50 mM Tris pH 7.4, 300 mM NaCl, 10% (v/v) glycerol, 20 mM imidazole, 0.1 mg mL−1 lysozyme, 0.05 mg mL−1 DNase, and 2 mM MgCl2), incubated on ice for 45 minutes, lysed by sonication, and centrifuged at 38,760 x g for 30 min at 4°C. The soluble fraction was filtered through a 0.45 μm syringe filter, loaded onto a 5 mL HisTrap column (Cytiva Life Sciences), washed with 10 column volumes of buffer A (50 mM HEPES pH 7.4 or 50 mM Tris pH 7.4, 300 mM NaCl, 10% (v/v) glycerol), and eluted with a linear gradient of 20-400 mM imidazole in buffer A.
For crystallization screening, the N-terminal His6 tags were removed from JuvEV TE, PikAIV TE and DEBSIII TE. For those samples, nickel affinity eluate containing enriched TE domain was pooled, augmented with tobacco etch virus (TEV) protease (1:30 protease:TE) in buffer B with 2 mM DTT, and dialyzed overnight at 4°C in buffer B (50 mM HEPES pH 7.4 or Tris pH 7.4, 150 mM NaCl, 10% (v/v) glycerol). Tag-free TEs were separated from any uncleaved TE and TEV protease by nickel affinity chromatography.
Following nickel affinity purification, all TEs were concentrated by filter centrifugation (Amicon) and further purified by size-exclusion chromatography (HiLoad 16/60 Superdex S75, Cytiva) in buffer B. All TE domains eluted at apparent molecular weights of 55-65 kDa, consistent with expected dimer molecular weights for each TE domain +/− the N-terminal His6-tag (~3 kDa).
DAP-EOS-NBO synthesis.
The doubly protected form of the unnatural amino acid (2S)-2,3- diaminopropionic acid (DAP), (2S)-2-amino-3-{[(2-{[1-(6-nitrobenzo[d][1,3]dioxol-5-yl)ethyl]thio}ethoxy)carbonyl]amino}propanoic acid (DAP-EOS-NBO), was synthesized in house and by the Vahlteich Medicinal Chemistry Core (University of Michigan) using a published protocol32. The purity of the synthesized DAP-EOS-NBO was verified by LC/ESI-MS and 1H NMR (Fig. S2).
DAP-EOS-NBO incorporation and TEDAP purification.
Extensive small-scale expression optimization was conducted to maximize incorporation of the DAP-EOS-NBO amino acid and TEDAP production. E. coli BL21(DE3) cells were co-transformed with TE expression plasmids having an amber codon substituted for the catalytic Ser or Cys and with pSF-DAPRS-PylT32, which encodes an evolved Methanosarcina barkeri (Mb) pyrrolysyl-tRNA synthetase/suppressor tRNA pair for the DAP-EOS-NBO amino acid and was kindly provided by Jason Chin (MRC Laboratory of Molecular Biology). 5 mL cell cultures were grown in TB media in 50 mL conical tubes with 100 μg mL−1 ampicillin (pTMM47-51) or 50 μg mL−1 spectinomycin (pTMM52-56), 50 μg mL−1 kanamycin (pSF-DAPRS-PylT), and 1 mM DAP-EOS-NBO resuspended in 0.5 M NaOH. The cultures were grown in the dark at 37°C to OD600=1.2, cooled to 20°C over 1 hr, induced with 200 μM IPTG, and grown for 16 hr. Growth in the dark prevented premature photo-deprotection of the DAP-EOS-NBO. Cells were harvested by 30-min 4°C centrifugation at 4,000 rpm and cell pellets were stored at −20°C.
Harvested cells from 5 mL cultures were resuspended in 1 mL HEPES or Tris buffer A, lysed for 15 min at room temperature using CelLytic Express (Sigma) following the manufacturer’s instructions, and centrifuged for 15 min at 4°C and 14,000 rpm. The lysate supernatant was incubated with 50 μL Ni-NTA resin (ThermoFisher) pre-equilibrated in buffer A at 4°C for 2 hr. Nickel resin was washed three times with 100 μL HEPES or Tris buffer A and bound protein was eluted with HEPES or Tris buffer A with 400 mM imidazole. Production of TEDAP was evaluated using SDS-PAGE. Suitable amounts of DEBSIII TEDAP were obtained from pTMM39 or pTMM54 expression, and PikAIV TEDAP from pTMM40 or pTMM53 expression. Expression tests for other TEDAP proteins (JuvEV, FluC, TylG5) did not yield sufficient protein for analysis.
For large-scale production of TEDAP domains, E. coli BL21(DE3) cells were co-transformed with pSF-DAPRS-PylT and expression plasmids for PikAIV TEDAP (pTMM53) or DEBSIII TEDAP (pTMM39) and grown at 37°C in 0.25 L TB media with 50 μg mL−1 kanamycin, 100 μg mL−1 ampicillin (pTMM39) or 50 μg mL−1 spectinomycin (pTMM53). DAP-EOS-NBO (1 mM in 0.5 M NaOH) was added 2-3 hr into cell growth. Cells were grown to OD600=1.2, cooled to 20°C over 1 hr, induced with 200 μM IPTG, and grown for 16 hr. Flasks were wrapped in aluminum foil for culture growth and expression to prevent photo deprotection. Cells were harvested by 30-min centrifugation at 4°C and 8,000 rpm and cell pellets were stored at −20°C.
PikAIV TEDAP and DEBSIII TEDAP were purified as for the wild type proteins, with a few modifications. DEBSIII TEDAP was purified in Tris buffer. For photo-deprotection, the eluate from Ni-affinity purification (TEDAP-EOS-NBO) was pooled in a 50 mL conical tube and irradiated with UV light (365 nm) for 4 min at room temperature to remove the photolabile protecting group. N-terminal His tags were removed as for the wild type proteins. In the final size-exclusion chromatography step, PikAIV TEDAP (MW 31.1 kDa) eluted at apparent molecular weight ~60 kDa, and DEBSIII TEDAP (28.6 kDa) at ~58 kDa (Fig. S3). TEDAP-EOS was incubated 1 week at 4°C in pH 8.0 buffer to promote spontaneous loss of the remaining protecting group, yielding TEDAP, and deprotection was confirmed by mass spectrometry (see below, Fig. S4).
Formation of TEDAP acyl-enzyme adducts.
Pikromycin thiophenol-pentaketide was synthesized and used with PikAIII and PikAIV for in vitro biocatalysis of narbonolide as previously described34. Pikromycin C3-hydroxymethyl-thiophenol-hexaketide substrates was also synthesized as previously described35. Macrocyclic 6-deoxyerythronolide B (6-dEB) 36 and tylactone37 were chemoenzymatically synthesized in vitro.
TEDAP-polyketide adducts were formed by 24-hr incubation at room temperature of a 5 μL reaction mixture (40 μM PikAIV TEDAP or DEBSIII TEDAP, 160 μM polyketide substrates/products, 50 mM HEPES pH 8.0, 150 mM NaCl, 10% (v/v) glycerol). Reactions were initiated by addition of the polyketide substrate or product. Reaction mixtures were diluted to 4 μM TEDAP, centrifuged 15 min at 4°C and 13,000 rpm, and analyzed by LC-MS as described below.
LC-MS analysis of TEs.
An Agilent 6545 Q-TOF mass spectrometer operated in positive-ion mode with an Agilent 1290 HPLC system was used to assess the deprotection state of PikAIV TEDAP and DEBSIII TEDAP. 2 μL of TEDAP (4 μM) underwent reverse-phase HPLC (Agilent PLRP-S reversed phase column 3.0 μM, 50 x 2.10 mm) in H2O with 0.2% (v/v) formic acid at a flow rate of 0.2 mL min−1. Protein was eluted over a 8 min linear gradient of 5-100% acetonitrile with 0.2% (v/v) formic acid. MS conditions were as follows: fragmentor voltage, 225 V; skimmer voltage, 25 V; nozzle voltage, 1000 V; sheath gas temperature, 350°C; drying gas temperature, 325°C. Data were processed using the MassHunter BioConfirm 10.0 software (Agilent).
Crystallography.
Crystals of all TE domains were grown by sitting-drop vapor diffusion at 20°C. All crystallization conditions included a cryoprotectant, and crystals were flash cooled in liquid N2 directly from their growth solutions.
JuvEV TE (residues 1508-1778) crystals grew in a 1 μL:1 μL mixture of protein (~10 mg/mL in HEPES buffer B without glycerol) and reservoir solution (20 mM MgCl2, 22% (w/v) polyacrylic acid 5100, 100 mM HEPES pH 7.5). Large, thin, rod-shaped crystals formed in 1 week.
FluC TE (residues 2106-2395) crystals grew in a 0.5 μL:0.5 μL mixture of protein (~10 mg/mL in Tris buffer B) and reservoir solution (100 mM sodium acetate, 30% (v/v) PEG 400, 100 mM MES pH 6.5). Small hexagonal crystals grew within 2-4 days.
TylG5 TE (residues 1549-1841) crystals were grown using the sitting drop method at 20°C in a 1 μL:1 μL mixture of protein (~10 mg/mL in Tris buffer B) and reservoir solution (1.0 M sodium citrate, 100 mM sodium malonate, 100 mM Bis-Tris propane pH 6.5). Crystals grew in clusters of hexagonal rods.
PikAIV TEWT (residues 1057-1346), PikAIV TEDAP, TEDAP-EOS-NBO and TEDAP-heptaketide crystallized under identical conditions. Crystals grew in a 0.5 μL:0.5 μL mixture of protein (15-20 mg/mL in HEPES buffer B) and reservoir solution (1.2 M LiCl, 25% (w/v) PEG 4000, 100 mM HEPES pH 8.5). Large rectangular (WT) or rod-shaped (DAP) crystals formed in 3-5 days.
All diffraction data were collected at 100 K on GM/CA beamline 23ID-B or 23ID-D at the Advanced Photon Source (APS) at Argonne National Laboratory (Argonne, IL) (Table S1). Data were processed and scaled using XDS51, with the exception of the TylG5 TE data, which were indexed and integrated in MOSFLM52 and scaled in SCALA in the CCP4 suite15. All manual model building was done in Coot53.
The TylG5 TE yielded crystals apparently in Laue group 6/mmm, but the intensity statistics indicated twinning54. Automated molecular replacement in the BALBES pipeline55 detected twinning operator K, H, -L with a refined twin fraction (α) of 49%, and identified a solution in space group P65 using 42% identical PikAIV TE17 (2HFJ) as a search model. Manual model building was iterated with refinement in REFMAC556 and BUSTER57. TLS groups were manually defined based on groups proposed by the TLSMD server58.
The structure of JuvEV TE was solved by molecular replacement with Phaser59 in the Phenix60 software suite using the 63% identical TylG5 TE as a search model. Four copies of the TylG5 TE monomer were placed, and translational non-crystallographic symmetry (tNCS) was indicated by a peak at (0, ½, 0) of height 73.4% relative to the origin of the native Patterson map. The structure was completed by iterative rounds of model building in Coot and refinement in phenix.refine61. Electron density for helix 〈6 in the JuvEV TE active-site lid was poor. An AlphaFold262,63 (AF2) prediction was an excellent match to the crystal structure for most of the JuvEV TE, but helix 〈6 was predicted at low confidence (pLDDT 70-80). Nevertheless, the AF2 model had a reasonable fit to weak electron density for part of helix 〈6 in chains C and D. This helix was omitted from chains A and B where no density was visible.
The FluC TE (residues 2106-2395) crystal structure was solved by molecular replacement in Phaser using the 36% identical PikAIV TE (2H7X) 31 as a molecular replacement search model and refined in phenix.refine.
With His tags removed, wild type PikAIV TE and all DAP variants crystallized isomorphously, but in a different form than for the published PikAIV TE structures. The structure of tag-free wild-type PikAIV TE was solved by molecular replacement with Phaser using our previous His-tagged PikAIV TE structure17 (2HFK). The new PikAIV TEWT structure was then used to solve structures of PikAIV TEDAP, TEDAP-EOS-NBO and TEDAP-heptaketide. For model refinement, stereochemical restraints for DAP-EOS-NBO and DAP-heptaketide were generated using eLBOW64, and the adducts were defined as l-amino acids in the protein. Subsequent model building was performed in Coot and refinement in phenix.refine.
Structures were validated in MolProbity65 (Table S1). All structure images were generated in PyMOL66, and sequences were aligned with Clustal67 in Jalview68.
Supplementary Material
Acknowledgements
The work was supported by NIH grant R01 DK042303 and the Rita Willis Professorship to J.L.S. and by NIH grant R35 GM118101 and the Hans W. Vahlteich Professorship to D.H.S. V.C. was supported by NIH training grant T32 GM140223. We thank Jason Chin (MRC, Cambridge, UK) for pSF-DAPRS-PylT encoding the pyro-Lys tRNA synthetase, the University of Michigan Vahlteich Medicinal Chemistry Core for synthesis of doubly protected DAP, and Wendy Feng (University of Michigan Life Sciences Institute Mass Spectrometry Core) for assistance with MS analysis. Gm/CA@APS has been funded by the National Cancer Institute (ACB-12002) and the National Institute of General Medical Sciences (AGM-12006, P30GM138396). This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract No. DE-AC02-06CH11357.
Footnotes
Supporting Information
For the panel of five TE domains, multiple sequence alignment and substrate/product structures; LC/MS and NMR validation of doubly protected DAP-EOS-NBO; purification of PikAIV TEDAP and DEBSIII TEDAP; LC/MS validation of deprotection to yield PikAIV TEDAP and DEBSIII TEDAP; PikAIV TEDAP reaction with a substrate panel; DEBSIII TEDAP reaction with a substrate panel; details of the PikAIV acyl cavity with and without the heptaketide adduct; crystallographic information table; table of expression plasmids; table of PCR primers
The authors declare no competing financial interest.
References
- 1.Kittendorf JD, and Sherman DH (2009). The methymycin/pikromycin pathway: a model for metabolic diversity in natural product biosynthesis. Bioorg Med Chem 17, 2137–2146. 10.1016/j.bmc.2008.10.082. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Staunton J, and Wilkinson B (1997). Biosynthesis of erythromycin and rapamycin. Chem Rev 97, 2611–2630. 10.1021/cr9600316. [DOI] [PubMed] [Google Scholar]
- 3.Udagawa T, Yuan J, Panigrahy D, Chang YH, Shah J, and D’Amato RJ (2000). Cytochalasin E, an epoxide containing Aspergillus-derived fungal metabolite, inhibits angiogenesis and tumor growth. J Pharmacol Exp Ther 294, 421–427. [PubMed] [Google Scholar]
- 4.Robbins N, Spitzer M, Wang W, Waglechner N, Patel DJ, O’Brien JS, Ejim L, Ejim O, Tyers M, and Wright GD (2016). Discovery of ibomycin, a complex macrolactone that exerts antifungal activity by impeding endocytic trafficking and membrane function. Cell Chem Biol 23, 1383–1394. 10.1016/j.chembiol.2016.08.015. [DOI] [PubMed] [Google Scholar]
- 5.Kim HJ, Choi SH, Jeon BS, Kim N, Pongdee R, Wu Q, and Liu HW (2014). Chemoenzymatic synthesis of spinosyn A. Angew Chem Int Ed Engl 53, 13553–13557. 10.1002/anie.201407806. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Horsman ME, Hari TP, and Boddy CN (2016). Polyketide synthase and non-ribosomal peptide synthetase thioesterase selectivity: logic gate or a victim of fate? Nat Prod Rep 33, 183–202. 10.1039/c4np00148f. [DOI] [PubMed] [Google Scholar]
- 7.Claxton HB, Akey DL, Silver MK, Admiraal SJ, and Smith JL (2009). Structure and functional analysis of RifR, the type II thioesterase from the rifamycin biosynthetic pathway. J Biol Chem 284, 5021–5029. 10.1074/jbc.M808604200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Geyer K, Hartmann S, Singh RR, and Erb TJ (2022). Multiple Functions of the Type II Thioesterase Associated with the Phoslactomycin Polyketide Synthase. Biochemistry 61, 2662–2671. 10.1021/acs.biochem.2c00234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Menzella HG, Reid R, Carney JR, Chandran SS, Reisinger SJ, Patel KG, Hopwood DA, and Santi DV (2005). Combinatorial polyketide biosynthesis by de novo design and rearrangement of modular polyketide synthase genes. Nat Biotechnol 23, 1171–1176. 10.1038/nbt1128. [DOI] [PubMed] [Google Scholar]
- 10.Hansen DA, Koch AA, and Sherman DH (2017). Identification of a thioesterase bottleneck in the pikromycin pathway through full-module processing of unnatural pentaketides. J Am Chem Soc 139, 13450–13455. 10.1021/jacs.7b06432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Koch AA, Schmidt JJ, Lowell AN, Hansen DA, Coburn KM, Chemler JA, and Sherman DH (2020). Probing selectivity and creating structural diversity through hybrid polyketide synthases. Angew Chem Int Ed Engl 59, 13575–13580. 10.1002/anie.202004991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Koch AA, Hansen DA, Shende VV, Furan LR, Houk KN, Jiménez-Osés G, and Sherman DH (2017). A single active site mutation in the pikromycin thioesterase generates a more effective macrocyclization catalyst. J Am Chem Soc 139, 13456–13465. 10.1021/jacs.7b06436. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Parenty A, Moreau X, and Campagne J-M (2006). Macrolactonizations in the total synthesis of natural products. Chem Rev 106, 911–939. 10.1021/cr0301402. [DOI] [PubMed] [Google Scholar]
- 14.Hari TP, Labana P, Boileau M, and Boddy CN (2014). An evolutionary model encompassing substrate specificity and reactivity of type I polyketide synthase thioesterases. Chembiochem 15, 2656–2661. 10.1002/cbic.201402475. [DOI] [PubMed] [Google Scholar]
- 15.Tsai SC, Miercke LJ, Krucinski J, Gokhale R, Chen JC, Foster PG, Cane DE, Khosla C, and Stroud RM (2001). Crystal structure of the macrocycle-forming thioesterase domain of the erythromycin polyketide synthase: versatility from a unique substrate channel. Proc Natl Acad Sci U S A 98, 14808–14813. 10.1073/pnas.011399198. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Argyropoulos P, Bergeret F, Pardin C, Reimer JM, Pinto A, Boddy CN, and Schmeing TM (2016). Towards a characterization of the structural determinants of specificity in the macrocyclizing thioesterase for deoxyerythronolide B biosynthesis. Biochim Biophys Acta 1860, 486–497. 10.1016/j.bbagen.2015.11.007. [DOI] [PubMed] [Google Scholar]
- 17.Akey DL, Kittendorf JD, Giraldes JW, Fecik RA, Sherman DH, and Smith JL (2006). Structural basis for macrolactonization by the pikromycin thioesterase. Nat Chem Biol 2, 537–542. 10.1038/nchembio824. [DOI] [PubMed] [Google Scholar]
- 18.Tsai SC, Lu H, Cane DE, Khosla C, and Stroud RM (2002). Insights into channel architecture and substrate specificity from crystal structures of two macrocycle-forming thioesterases of modular polyketide synthases. Biochemistry 41, 12598–12606. 10.1021/bi0260177. [DOI] [PubMed] [Google Scholar]
- 19.Zhou Y, Tao W, Qi Z, Wei J, Shi T, Kang Q, Zheng J, Zhao Y, and Bai L (2022). Structural and mechanistic insights into chain release of the polyene PKS thioesterase domain. ACS Catal 12, 762–776. 10.1021/acscatal.1c04991. [DOI] [Google Scholar]
- 20.Scaglione JB, Akey DL, Sullivan R, Kittendorf JD, Rath CM, Kim ES, Smith JL, and Sherman DH (2010). Biochemical and structural characterization of the tautomycetin thioesterase: analysis of a stereoselective polyketide hydrolase. Angew Chem Int Ed Engl 49, 5726–5730. 10.1002/anie.201000032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Gehret JJ, Gu L, Gerwick WH, Wipf P, Sherman DH, and Smith JL (2011). Terminal alkene formation by the thioesterase of curacin A biosynthesis: structure of a decarboxylating thioesterase. J Biol Chem 286, 14445–14454. 10.1074/jbc.M110.214635. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Aldrich CC, Venkatraman L, Sherman DH, and Fecik RA (2005). Chemoenzymatic synthesis of the polyketide macrolactone 10-deoxymethynolide. J Am Chem Soc 127, 8910–8911. 10.1021/ja0504340. [DOI] [PubMed] [Google Scholar]
- 23.Jacobsen JR, Hutchinson CR, Cane DE, and Khosla C (1997). Precursor-directed biosynthesis of erythromycin analogs by an engineered polyketide synthase. Science 277, 367–369. 10.1126/science.277.5324.367. [DOI] [PubMed] [Google Scholar]
- 24.Dunn BJ, Watts KR, Robbins T, Cane DE, and Khosla C (2014). Comparative analysis of the substrate specificity of trans- versus cis-acyltransferases of assembly line polyketide synthases. Biochemistry 53, 3796–3806. 10.1021/bi5004316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Pinto A, Wang M, Horsman M, and Boddy CN (2012). 6-Deoxyerythronolide B synthase thioesterase-catalyzed macrocyclization is highly stereoselective. Org Lett 14, 2278–2281. 10.1021/ol300707j. [DOI] [PubMed] [Google Scholar]
- 26.Gaudelli NM, and Townsend CA (2014). Epimerization and substrate gating by a TE domain in β-lactam antibiotic biosynthesis. Nat Chem Biol 10, 251–258. 10.1038/nchembio.1456. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Korman TP, Crawford JM, Labonte JW, Newman AG, Wong J, Townsend CA, and Tsai SC (2010). Structure and function of an iterative polyketide synthase thioesterase domain catalyzing Claisen cyclization in aflatoxin biosynthesis. Proc Natl Acad Sci U S A 107, 6246–6251. 10.1073/pnas.0913531107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Sasse F, Steinmetz H, Höfle G, and Reichenbach H (1995). Gephyronic acid, a novel inhibitor of eukaryotic protein synthesis from Archangium gephyra (myxobacteria). Production, isolation, physico-chemical and biological properties, and mechanism of action. J Antibiot (Tokyo) 48, 21–25. 10.7164/antibiotics.48.21. [DOI] [PubMed] [Google Scholar]
- 29.Little RF, and Hertweck C (2022). Chain release mechanisms in polyketide and non-ribosomal peptide biosynthesis. Nat Prod Rep 39, 163–205. 10.1039/d1np00035g. [DOI] [PubMed] [Google Scholar]
- 30.Bruner SD, Weber T, Kohli RM, Schwarzer D, Marahiel MA, Walsh CT, and Stubbs MT (2002). Structural basis for the cyclization of the lipopeptide antibiotic surfactin by the thioesterase domain SrfTE. Structure 10, 301–310. 10.1016/s0969-2126(02)00716-5. [DOI] [PubMed] [Google Scholar]
- 31.Giraldes JW, Akey DL, Kittendorf JD, Sherman DH, Smith JL, and Fecik RA (2006). Structural and mechanistic insights into polyketide macrolactonization from polyketide-based affinity labels. Nat Chem Biol 2, 531–536. 10.1038/nchembio822. [DOI] [PubMed] [Google Scholar]
- 32.Huguenin-Dezot N, Alonzo DA, Heberlig GW, Mahesh M, Nguyen DP, Dornan MH, Boddy CN, Schmeing TM, and Chin JW (2019). Trapping biosynthetic acyl-enzyme intermediates with encoded 2,3-diaminopropionic acid. Nature 565, 112–117. 10.1038/s41586-018-0781-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Stols L, Gu M, Dieckman L, Raffen R, Collart FR, and Donnelly MI (2002). A new vector for high-throughput, ligation-independent cloning encoding a tobacco etch virus protease cleavage site. Protein Expr Purif 25, 8–15. 10.1006/prep.2001.1603. [DOI] [PubMed] [Google Scholar]
- 34.Hansen DA, Rath CM, Eisman EB, Narayan AR, Kittendorf JD, Mortison JD, Yoon YJ, and Sherman DH (2013). Biocatalytic synthesis of pikromycin, methymycin, neomethymycin, novamethymycin, and ketomethymycin. J Am Chem Soc 135, 11232–11238. 10.1021/ja404134f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Hansen DA, Koch AA, and Sherman DH (2015). Substrate controlled divergence in polyketide synthase catalysis. J Am Chem Soc 137, 3735–3738. 10.1021/ja511743n. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Lowry B, Robbins T, Weng CH, O’Brien RV, Cane DE, and Khosla C (2013). In vitro reconstitution and analysis of the 6-deoxyerythronolide B synthase. J Am Chem Soc 135, 16809–16812. 10.1021/ja409048k. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Lowell AN, DeMars MD, Slocum ST, Yu F, Anand K, Chemler JA, Korakavi N, Priessnitz JK, Park SR, Koch AA, et al. (2017). Chemoenzymatic total synthesis and structural diversification of tylactone-based macrolide antibiotics through late-stage polyketide assembly, tailoring, and C-H functionalization. J Am Chem Soc 139, 7913–7920. 10.1021/jacs.7b02875. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Kao CM, Luo G, Katz L, Cane DE, and Khosla C (1995). Manipulation of macrolide ring size by directed mutagenesis of a modular polyketide synthase. J Am Chem Soc 117, 9105–9106. 10.1021/ja00140a043. [DOI] [Google Scholar]
- 39.Kao CM, Luo G, Katz L, Cane DE, and Khosla C (1996). Engineered biosynthesis of structurally diverse tetraketides by a trimodular polyketide synthase. J Am Chem Soc 118, 9184–9185. 10.1021/ja9617552. [DOI] [Google Scholar]
- 40.Kao CM, McPherson M, McDaniel RN, Fu H, Cane DE, and Khosla C (1997). Gain of function mutagenesis of the erythromycin polyketide synthase. 2. Engineered biosynthesis of an eight-membered ring tetraketide lactone. J Am Chem Soc 119, 11339–11340. 10.1021/ja972609e. [DOI] [Google Scholar]
- 41.Hegde VR, Patel MG, Gullo VP, Ganguly AK, Sarre O, Puar MS, and McPhail AT (1990). Macrolactams: A new class of antifungal agents. J Am Chem Soc 112, 6403–6405. 10.1021/ja00173a042. [DOI] [Google Scholar]
- 42.Omura S, Matsubara H, Nakagawa A, Furusaki A, and Matsumoto T (1980). X-ray crystallography of protylonolide and absolute configuration of tylosin. J Antibiot (Tokyo) 33, 915–917. 10.7164/antibiotics.33.915. [DOI] [PubMed] [Google Scholar]
- 43.Stang EM, and White MC (2009). Total synthesis and study of 6-deoxyerythronolide B by late-stage C-H oxidation. Nat Chem 1, 547–551. 10.1038/nchem.351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Xue Y, Zhao L, Liu HW, and Sherman DH (1998). A gene cluster for macrolide antibiotic biosynthesis in Streptomyces venezuelae: architecture of metabolic diversity. Proc Natl Acad Sci U S A 95, 12111–12116. 10.1073/pnas.95.21.12111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Adrover-Castellano ML, Schmidt JJ, Glasser CA, Qu F, and Sherman DH (2024). Directed evolution of a modular polyketide synthase thioesterase for generation of a hybrid macrocyclic ring system. ChemRxiv. 10.26434/chemrxiv-2024-02cj6. [DOI] [Google Scholar]
- 46.Gokhale RS, Hunziker D, Cane DE, and Khosla C (1999). Mechanism and specificity of the terminal thioesterase domain from the erythromycin polyketide synthase. Chem Biol 6, 117–125. 10.1016/S1074-5521(99)80008-8. [DOI] [PubMed] [Google Scholar]
- 47.Lin TY, Borketey LS, Prasad G, Waters SA, and Schnarr NA (2013). Sequence, cloning, and analysis of the fluvirucin B1 polyketide synthase from Actinomadura vulgaris. ACS Synth Biol 2, 635–642. 10.1021/sb4000355. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Yoon YJ, Beck BJ, Kim BS, Kang HY, Reynolds KA, and Sherman DH (2002). Generation of multiple bioactive macrolides by hybrid modular polyketide synthases in Streptomyces venezuelae. Chem Biol 9, 203–214. 10.1016/s1074-5521(02)00095-9. [DOI] [PubMed] [Google Scholar]
- 49.Lopanik NB, Shields JA, Buchholz TJ, Rath CM, Hothersall J, Haygood MG, Hakansson K, Thomas CM, and Sherman DH (2008). In vivo and in vitro transacylation by BryP, the putative bryostatin pathway acyltransferase derived from an uncultured marine symbiont. Chem Biol 15, 1175–1186. 10.1016/j.chembiol.2008.09.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Nishihara K, Kanemori M, Kitagawa M, Yanagi H, and Yura T (1998). Chaperone coexpression plasmids: Differential and synergistic roles of DnaK-DnaJ-GrpE and GroEL-GroES in assisting folding of an allergen of Japanese cedar pollen, Cryj2, in Escherichia coli. Appl Environ Microbiol 64, 1694–1699. 10.1128/AEM.64.5.1694-1699.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Kabsch W. (2010). Xds. Acta Crystallogr D Biol Crystallogr 66, 125–132. 10.1107/S0907444909047337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Leslie AG (2006). The integration of macromolecular diffraction data. Acta Crystallogr D Biol Crystallogr 62, 48–57. 10.1107/S0907444905039107. [DOI] [PubMed] [Google Scholar]
- 53.Emsley P, and Cowtan K (2004). Coot: Model-building tools for molecular graphics. Acta Crystallogr D Biol Crystallogr 60, 2126–2132. 10.1107/S0907444904019158. [DOI] [PubMed] [Google Scholar]
- 54.Padilla JE, and Yeates TO (2003). A statistic for local intensity differences: Robustness to anisotropy and pseudo-centering and utility for detecting twinning. Acta Crystallogr D Biol Crystallogr 59, 1124–1130. 10.1107/s0907444903007947. [DOI] [PubMed] [Google Scholar]
- 55.Long F, Vagin AA, Young P, and Murshudov GN (2008). BALBES: A molecular-replacement pipeline. Acta Crystallogr D Biol Crystallogr 64, 125–132. 10.1107/S0907444907050172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Murshudov GN, Skubák P, Lebedev AA, Pannu NS, Steiner RA, Nicholls RA, Winn MD, Long F, and Vagin AA (2011). REFMAC5 for the refinement of macromolecular crystal structures. Acta Crystallogr D Biol Crystallogr 67, 355–367. 10.1107/S0907444911001314. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Bricogne G, Blanc E, Brandl M, Flensburg C, Keller P, Paciorek W, Roversi P, Sharff A, Smart OS, and Vonrhein C (2017). BUSTER 2.10.3 (Global Phasing Ltd.). [Google Scholar]
- 58.Painter J, and Merritt EA (2006). Optimal description of a protein structure in terms of multiple groups undergoing TLS motion. Acta Crystallogr D Biol Crystallogr 62, 439–450. 10.1107/S0907444906005270. [DOI] [PubMed] [Google Scholar]
- 59.McCoy AJ, Grosse-Kunstleve RW, Adams PD, Winn MD, Storoni LC, and Read RJ (2007). Phaser crystallographic software. J Appl Crystallogr 40, 658–674. 10.1107/S0021889807021206. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Adams PD, Afonine PV, Bunkóczi G, Chen VB, Davis IW, Echols N, Headd JJ, Hung LW, Kapral GJ, Grosse-Kunstleve RW, et al. (2010). PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr D Biol Crystallogr 66, 213–221. 10.1107/S0907444909052925. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Afonine PV, Grosse-Kunstleve RW, Echols N, Headd JJ, Moriarty NW, Mustyakimov M, Terwilliger TC, Urzhumtsev A, Zwart PH, and Adams PD (2012). Towards automated crystallographic structure refinement with phenix.refine. Acta Crystallogr D Biol Crystallogr 68, 352–367. 10.1107/S0907444912001308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O, Tunyasuvunakool K, Bates R, Žídek A, Potapenko A, et al. (2021). Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589. 10.1038/s41586-021-03819-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Mirdita M, Schütze K, Moriwaki Y, Heo L, Ovchinnikov S, and Steinegger M (2022). ColabFold: making protein folding accessible to all. Nat Methods 19, 679–682. 10.1038/s41592-022-01488-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Moriarty NW, Grosse-Kunstleve RW, and Adams PD (2009). electronic Ligand Builder and Optimization Workbench (eLBOW): A tool for ligand coordinate and restraint generation. Acta Crystallogr D Biol Crystallogr 65, 1074–1080. 10.1107/S0907444909029436. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Chen VB, Arendall WB 3rd, Headd JJ, Keedy DA, Immormino RM, Kapral GJ, Murray LW, Richardson JS, and Richardson DC (2010). MolProbity: All-atom structure validation for macromolecular crystallography. Acta Crystallogr D Biol Crystallogr 66, 12–21. 10.1107/S0907444909042073. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.The PyMOL Molecular Graphics System, Version 2. Schrödinger, LLC. [Google Scholar]
- 67.Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan PA, McWilliam G , Valentin F, Wallace IM, Wilm A, Lopez R, et al. (2007). Clustal W and Clustal X version 2.0. Bioinformatics 23, 2947–2948. 10.1093/bioinformatics/btm404. [DOI] [PubMed] [Google Scholar]
- 68.Waterhouse AM, Procter JB, Martin DM, Clamp M, and Barton GJ (2009). Jalview Version 2--a multiple sequence alignment editor and analysis workbench. Bioinformatics 25, 1189–1191. 10.1093/bioinformatics/btp033. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
