S. Joe
Russell†
a,
Catherine R.
Back†
b,
Christopher
Perry
a,
Kaiman A.
Cheung
a,
Laurence
Maschio
b,
Sacha N.
Charlton
a,
Nicholas R.
Lees
a,
Monserrat
Manzo-Ruiz
bc,
Martin A.
Hayes
d,
Marc W.
van der Kamp
b,
Paul R.
Race
*c and
Christine L.
Willis
*a
aSchool of Chemistry, University of Bristol, Cantock's Close, Bristol BS8 1TS, UK. E-mail: chris.willis@bristol.ac.uk
bSchool of Biochemistry, University Walk, University of Bristol, Bristol, BS8 1TD, UK
cSchool of Natural and Environmental Sciences, Newcastle University, Newcastle Upon Tyne NE1 7RU, UK. E-mail: paul.race1@newcastle.ac.uk
dDiscovery Sciences, Biopharmaceuticals R&D, AstraZeneca, Pepparedsleden 1, SE-431 83, Mölndal, Sweden
First published on 26th August 2025
The tetrodecamycins are tetracyclic natural products that exhibit potent antimicrobial activity against a multitude of drug-resistant pathogens. These compounds are structurally distinguished by the presence of a tetronate ring and trans-decalin with six contiguous asymmetric centres united by a seven-membered oxygen heterocycle. Herein we describe the first total synthesis of the antibiotic (−)-13-deoxytetrodecamycin. Our strategy is predicated on an enantioselective [4 + 2]-cycloaddition catalysed by the FAD-dependent Diels–Alderase TedJ, forming the trans-decalin with concomitant creation of two rings and four contiguous stereocenters with exquisite selectivity under mild conditions. In complementary studies, in vitro enzyme assays, X-ray crystallography and computational modelling are used to provide molecular insights into the TedJ catalysed reaction. These studies illustrate the power of adopting a chemoenzymatic approach for the enantioselective synthesis of a target compound which would be difficult to achieve using non-biological methods and provide a practical demonstration of the use of Diels–Alder biocatalysts in total synthesis. This approach has potentially widespread value in the global challenge of discovery and development of new antibiotics.
![]() | ||
Scheme 1 (a). Structures of three tetrodecamycins. (b). Outline of our initially proposed synthetic strategy to 13-deoxytetrodecamycin 3. |
The unusual structure of tetrodecamycin and its biological activity have attracted the attention of synthetic chemists and studies towards its synthesis have been reported.3,4 However, there has been only one total synthesis of tetrodecamycin 1 described in the literature,5 with analogues having also been prepared for SAR studies.6 In 2015, a further member of the tetrodecamycin family, 13-deoxytetrodecamycin 3, was isolated in low titres from extracts of Streptomyces WAC04657.7 The structure was investigated using spectroscopic methods and comparison with NMR data for tetrodecamycin. 13-Deoxytetrodecamycin 3 exhibits potent activity against Gram-positive organisms including MSRA and multi drug-resistant S. aureus ATCC BAA-44 in the range MIC 1 to 8 μg ml−1.7 There is an urgent need to develop new antibiotics against Gram-positive and Gram-negative pathogens and efficient methods for their synthesis are required. Hence our goal was to develop a flexible strategy for the total synthesis of the antibiotic 13-deoxytetrodecamycin 3.
An outline of our initially proposed synthetic strategy to 13-deoxytetrodecamycin 3 is shown in Scheme 1b. A key step is the selective intramolecular cycloaddition to convert 6 to the novel trans-decalin 7. The Diels–Alder reaction, involving the [4 + 2]-cycloaddition of a dienophile and conjugated diene, is one of the most valuable methods for the synthesis of six-membered rings generating up to four new stereocenters in a single step.8 This transformation commonly requires heat, pressure or Lewis acid catalysis, and achieving high enantioselectivity remains a challenge.9 It has long been speculated that Diels–Alder reactions are important in biosynthetic pathways forming structurally diverse natural products,10 but it is only recently that enzymes which catalyse [4 + 2]-cycloadditions have been isolated and characterised e.g. AbyU11 and AbmU12 in the biosynthesis of the class I spirotetronate antibiotics abyssomicin C and neoabyssomicin B respectively, as well as the Class II spirotetronates which contain an additional trans-decalin moiety e.g. versipelostatin13 and chlorothricin (Scheme 2).14
The tetrodecamycin biosynthetic gene cluster (BGC) has been identified, and Nodwell and co-workers proposed a biosynthetic route by analogy to pathways to other tetronate natural products in which the carbon skeleton is assembled via a type I polyketide synthase and the tetronate formed by a glycerol-derived C3 unit.2,15 Interestingly the tetrodecamycin pathway enzyme TedJ shows homology to previously identified Diels–Alderases16,17 and Nodwell and co-workers proposed that it may catalyse a [4 + 2]-cycloaddition to selectively form the decalin moiety. More recent studies by Jiao, Ge, Zhang and co-workers are in accord with this proposal and they reported that the cytochrome P450 TedH catalyses epoxidation, ether bridge formation and 13-hydroxylation in tetrodecamycin biosynthesis.18
Natural Diels–Alderases, including those found in spirotetronate biosynthetic pathways, have rarely been successfully deployed in the sustainable synthesis of bioactive compounds. This is despite their considerable potential to provide ‘green’, atom efficient routes to target molecules.19 Herein we describe the first total synthesis of 13-deoxytetrodecamycin 3, which has been achieved using a chemoenzymatic strategy incorporating the natural Diels–Alderase TedJ. Furthermore, structural, functional and computational studies have been used to provide a molecular framework for the TedJ catalysed reaction, establishing a foundation for future studies seeking to employ [4 + 2]-cyclases in asymmetric synthesis, accessed either via enzyme discovery, rational engineering, directed evolution, or a combination thereof.
Armed with this knowledge, the [4 + 2]-cycloaddition of unsaturated aldehyde 4 was investigated and our optimised reaction conditions were with Me2AlCl at low temperature giving trans-decalin 8 in 67% isolated yield (Scheme 3). These results were consistent with our DFT calculations that indicated a lower kinetic barrier for the conversion of 4 to 8 compared with cyclisation of 6 to 7. The structure of 8 was confirmed by NMR and was consistent with an endo-selective [4 + 2]-cycloaddition.4e Coupling aldehyde 8 with tetronate 5 followed by oxidation of the resultant diastereomeric alcohols with DMP gave the novel trans-decalin 7 in 64% yield over the two steps.
Next, we turned our attention to the enzyme catalysed [4 + 2]-cycloaddition of 6. This biotransformation would be of particular value as it would be expected to provide a single enantiomer of 7 for use as an intermediate in the total synthesis of (−)-13-deoxytetradecamycin 3. Based on a comparative sequence analysis with known decalin forming [4 + 2]-cyclases, the tedJ gene of the Streptomyces WAC04657 tetrodecamycin biosynthetic gene cluster was identified as encoding a probable flavin dependent Diels–Alderase. This gene was subsequently cloned and an N-terminally hexa-histidine tagged variant of TedJ recombinantly over-expressed in E. coli, followed by purification to homogeneity (Fig. S1, SI). Purified TedJ was light yellow in colour and exhibited an absorbance spectrum consistent with that of a flavoprotein (Fig. S3, SI). Denaturation of TedJ by heating to 60 °C in MeOH for 30 min, followed by mass spectrometric (MS) analysis of the resulting material, revealed the presence of FAD (Fig. S4 and S5, SI), confirming its identity as a non-covalently bound cofactor in TedJ. Flavoenzymes which catalyse reactions without a change in the oxidation level of the cofactor have been reported20,21 and it seems likely that FAD in TedJ may play a solely structural role in the [4 + 2]-cycloaddition, as has been proposed for PyrE3, which catalyses decalin formation in pyrroindomycin biosynthesis.16,17
For in vitro assays, tetronate 6 (0.2 mM) in 10% methanol co-solvent was incubated with purified recombinant TedJ at 25 °C in Tris-HCl buffer at pH 7.0 for 3 h. Analysis of the resultant mixture by reverse-phase LC-MS indicated the presence of starting material 6 and a new product with the same retention time as the synthetic standard 7 (Fig. 1). There was no evidence for cyclisation occurring in control experiments without enzyme. Under the same assay conditions except in DMSO co-solvent, complete consumption of substrate 6 with TedJ was observed after 3 h. These results are likely attributed to an improved substrate solubility in DMSO.
Without the need for any added cofactors, the in vitro reaction was scaled up using substrate 6 and TedJ and the product purified by flash column chromatography. NMR analysis confirmed that TedJ catalysed the intramolecular [4 + 2]-cycloaddition of 6, producing trans-decalin 7 (17 mg, 44% isolated yield) with the creation of two carbocyclic rings and four contiguous stereocenters, a transformation which failed in our hands under standard chemical conditions (heat or with Lewis acids). Importantly, not only had a single diastereomer been formed, but comparison of the product 7 with racemic material by chiral HPLC using a 0.2:
99.8 EtOAc
:
hexane isocratic method confirmed that a single enantiomer, (−)-7, [α]D −112 (c 0.54 in CHCl3) had been isolated from the enzymatic reaction (Fig. S6, SI).
![]() | ||
Scheme 4 Completing the total synthesis of 13-deoxytetrodecamycin 3via selective epoxidation of trans-decalin 7. |
Thus, for the synthesis of the required 14,15-epoxide 12, we reasoned that as with epoxidation of alkene 7, an intermediate iodonium ion would also form selectively on the α-face of the molecule. Subsequent attack by water from the β-face would give vicinal iodohydrins 10 and 11, which in turn could be converted to the required epoxide 12. Treatment of 7 with widely used N-iodosuccinimide returned starting material, indicating that a more reactive electrophile was required. Hence, N-iodosaccharin 13 was prepared as previously described22 and on reaction with 7 gave a mixture of iodohydrins 10 and 11 which was used directly in the next step (Scheme 4). The crude mixture was treated with NaH, giving the required 14,15-epoxide 12 in 64% isolated yield over the 2 steps. The 1H-NMR spectrum of 12 displayed signals at δ2.77 (d, J 4.0) and δ2.51 (d, J 4.0) assigned to 15-H and 14-H respectively. Demethylation of the methyl ether in 12 using LiCl in DMSO gave 14, which was immediately treated with HCl completing the first total synthesis of 13-deoxytetrodecamycin 3. When trans-decalin (−)-7, obtained from the biotransformation with TedJ, was carried through this same sequence of epoxide formation-demethylation-cyclisation, it gave (−)-13-deoxytetrodecamycin 3. The 1H- and 13C-NMR data of the synthetic material were in accord with those reported for the isolated natural product (Table S4, SI) and the optical rotation [α]D −14.0 (c 0.75 CHCl3) correlated well with that reported for the natural product [α]D −17.6 (c 0.65, CHCl3).7
Molecular docking was used to provide further insights into the mechanism and selectivity of TedJ. The DFT optimised endo-transition state of 6 was docked into the TedJ crystal structure and the highest ranked docking result was relaxed using molecular mechanics energy minimisation with the Enlighten2 tools,25 yielding the likely reactive pose (Fig. 2b, see SI for further details). The pose of 6 in TedJ is similar to that reported for PyrE3. This is expected due to the significant sequence and structural similarities of TedJ and PyrE3, which each catalyse the formation of trans-decalins, albeit employing structurally distinct substrates.17 A number of protein–ligand interactions are conserved between the TedJ–6 and PyrE3 complexes. The tetronate and ketonic carbonyls of 6 were found to interact with His74, Ser208 and FAD in TedJ (Fig. 2c), analogous to the interactions of His74, Met201 and FAD in PyrE3 with its substrate. Notably, in this suggested reactive pose, the side-chain of Ala210 points towards the s-cis diene such that there is good stereo-complementarity (Fig. 2c). Inspection of the TedJ–6 complex reveals that steric constraints imposed by the side chains of residues H74, L77, R218, R262, and R358, appear critical in defining the stereochemical outcome of the TedJ catalysed reaction. The relative positioning of these residues, reinforced by the rigidity of the enzyme active site, would preclude the generation of alternative diastereomers or regioisomers, an observation fully in keeping with our in vitro assay data. Analogous to the proposed role of Ala45 in PyrE3, Ala210 would disfavour the orientation of the diene in an exo-selective manner. Thus, transition state docking and relaxation provide a mechanistic explanation for the stereoselectivity of the TedJ catalysed reaction.
It is proposed that FAD plays an exclusively structural role in TedJ, with no change in redox state during the reaction. Hence, without the need for additional cofactors, the TedJ-catalysed biotransformation was readily scaled up to generate trans-decalin (−)-7 as a single enantiomer and in sufficient amounts to complete the first total synthesis of the broad-spectrum antibiotic (−)-13-deoxytetrodecamycin 3. Stereoselective formation of epoxide 12via the precursor iodohydrins 10 and 11, followed by demethylation and acid-mediated cyclisation gave 3 with spectroscopic data and optical rotation value that correlated well with those of the natural product. These studies add a valuable new enzyme to the toolbox of biocatalysts for chemoenzymatic synthesis, unlocking a reaction with exquisite regio- and stereoselectivity under mild conditions, and facilitating access to a potent antimicrobial agent of potential clinical value. Elucidation of the X-ray crystal structure of TedJ, in combination with computational modelling, has provided important insights into the catalytic mechanism of this enzyme, establishing a foundation for the further development of this Diels–Alderase as a biocatalyst for use in synthesis. In future, the versatility and utility of TedJ may be further enhanced through structure guided rational enzyme engineering or directed evolution.
Supplementary information is available and includes experimental procedures, analytical and spectral data, NMR spectra, computational methods, protein expression, purification, crystallisation and X-ray. See DOI: https://doi.org/10.1039/d5sc05480j.
Footnote |
† Joint first authors. |
This journal is © The Royal Society of Chemistry 2025 |