Atrop-abyssomicin C

Last updated
Atrop-abyssomicin C
Atrop aby c.png
Names
IUPAC name
12,14a,3-(Epoxymethyno)-2H-1-benzoxacyclododecin-2,4,8(5H,10aH)-trione, 6,7,11,12,13,14-hexahydro-11-hydroxy-5,7,13-trimethyl-, (5R,7S,9E,10aR,11R,12R,13R,14aR)
Other names
Atrop-abyssomicin C
Identifiers
3D model (JSmol)
ChEMBL
ChemSpider
PubChem CID
  • InChI=1S/C19H22O6/c1-8-6-9(2)14(21)13-17-19(25-18(13)23)7-10(3)16(24-17)15(22)11(19)4-5-12(8)20/h4-5,8-11,15-16,22H,6-7H2,1-3H3/b5-4+/t8-,9+,10+,11?,15+,16?,19+/m0/s1
    Key: FNEADFUPWHAVTA-WDQYZCCLSA-N
  • C[C@@H]1C[C@]23OC(=O)C4=C2OC1[C@H](O)C3\C=C\C(=O)[C@@H](C)C[C@@H](C)C4=O
Properties
C19H22O6
Molar mass 346.38 g/mol
Density 1.34±0.1 g/cm3 (Predicted)
Melting point 180 °C (decomp)
Boiling point 597.5±50.0 °C (Predicted)
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

Atrop-abyssomicin C is a polycyclic polyketide-type natural product that is the atropisomer of abyssomicin C. It is a spirotetronate that belongs to the class of tetronate antibiotics, which includes compounds such as tetronomycin, agglomerin, and chlorothricin. [1] In 2006, the Nicolaou group discovered atrop-abyssomicin C while working on the total synthesis of abyssomicin C. [2] Then in 2007, Süssmuth and co-workers isolated atrop-abyssomicin C from Verrucosispora maris AB-18-032, a marine actinomycete found in sediment of the Japanese sea. They found that atrop-abyssomicin C was the major metabolite produced by this strain, while abyssomicin C was a minor product. The molecule displays antibacterial activity by inhibiting the enzyme PabB (4-amino-4-deoxychorismate synthase), thereby depleting the biosynthesis of p -aminobenzoate. [3] [4]

Contents

Structure

Structure of Atrop-abyssomicin C and Abyssomicin C. Atrop and abyssomicin c.png
Structure of Atrop-abyssomicin C and Abyssomicin C.

Atrop-abyssomicin C has a complex, yet intriguing structural topography. The compound contains an oxabicyclo[2.2.2]octane system fused to the tetronate moiety. The 11-membered macrocyclic ring carries an α,β-unsaturated ketone that was proposed to be the reactive center. [5] Despite being a strained macrocycle, there exist an atropisomer, abyssomicin C. The atropisomerism arise due to a structural deviation in the α,β-unsaturated ketone region of the molecule. The orientation of the carbonyl in atrop-abyssomicin C is cisoid, whereas the conformation in abyssomicin C is transoid. [6] The enone moiety of atrop-abyssomicin C has a higher degree of the conjugation, which makes it a more active Michael acceptor. [7]

Biosynthesis

The biosynthesis of atrop-abyssomicin C begins with the synthesis of a linear polyketide chain in a PKS I system that consist of one loading and six extension modules. The polyketide chain is made from five acetates, two propionates, and the glycolytic pathway metabolite. D-1,3-bisphosphoglycerate, the glycolytic metabolite, is transferred to AbyA3 (an acyl-carrier protein) by AbyA2 to generate the glyceryl-ACP. AbyA1 facilitates the attachment of the glyceryl-ACP to the polyketide chain and the detachment of the polyketide from the polyketide synthase to form intermediate 2. [7] [8] [9]

Biosynthesis of linear polyketide precursor. The AbyB1, AbyB2, and AbyB3 genes code for the seven-module polyketide synthase complex that assembled the polyketide backbone. Next, the linear polyketide precursor fused with glyceryl-ACP to form intermediate 2. Atrop Abyssomicin C Biosynthesis Module.png
Biosynthesis of linear polyketide precursor. The AbyB1, AbyB2, and AbyB3 genes code for the seven-module polyketide synthase complex that assembled the polyketide backbone. Next, the linear polyketide precursor fused with glyceryl-ACP to form intermediate 2.

Based on the observation made for the biosynthesis of agglomerin, it has been proposed that AbyA4 acetylates intermediate 2 and AbyA5 catalyzes the elimination of acetic acid to form the exocyclic double bond in intermediate 4. [1] An intramolecular Diels-Alder was proposed to take place between the exocyclic olefin and the conjugated diene at the tail end of the polyketide to form the macrocyclic ring. [7] It has been reported that the previously unidentified Abycyc gene could code for an enzyme that carries out the Diels-Alder cycloaddition. [10] Following the Diels-Alder reaction, an epoxide ring is formed and then opened by the tetronate hydroxyl group to form atrop-abyssomicin C. It has been postulated that the AbyE monooxygenase catalyzes epoxide formation. [8]

Cycloaddition to form atrop-abyssomicin C. Intermediate 2 undergo an acetylation and elimination step to form the exocyclic olefin. An intramolecular Diels-Alder reaction is carried out to form the macrocyclic ring. Next, an oxygenation step follows by a ring opening reaction leads to atrop-abyssomicin C formation. Biosynthesis of Atrop Abyssomycin C.png
Cycloaddition to form atrop-abyssomicin C. Intermediate 2 undergo an acetylation and elimination step to form the exocyclic olefin. An intramolecular Diels–Alder reaction is carried out to form the macrocyclic ring. Next, an oxygenation step follows by a ring opening reaction leads to atrop-abyssomicin C formation.

Related Research Articles

<span class="mw-page-title-main">Diels–Alder reaction</span> Chemical reaction

In organic chemistry, the Diels–Alder reaction is a chemical reaction between a conjugated diene and a substituted alkene, commonly termed the dienophile, to form a substituted cyclohexene derivative. It is the prototypical example of a pericyclic reaction with a concerted mechanism. More specifically, it is classified as a thermally-allowed [4+2] cycloaddition with Woodward–Hoffmann symbol [π4s + π2s]. It was first described by Otto Diels and Kurt Alder in 1928. For the discovery of this reaction, they were awarded the Nobel Prize in Chemistry in 1950. Through the simultaneous construction of two new carbon–carbon bonds, the Diels–Alder reaction provides a reliable way to form six-membered rings with good control over the regio- and stereochemical outcomes. Consequently, it has served as a powerful and widely applied tool for the introduction of chemical complexity in the synthesis of natural products and new materials. The underlying concept has also been applied to π-systems involving heteroatoms, such as carbonyls and imines, which furnish the corresponding heterocycles; this variant is known as the hetero-Diels–Alder reaction. The reaction has also been generalized to other ring sizes, although none of these generalizations have matched the formation of six-membered rings in terms of scope or versatility. Because of the negative values of ΔH° and ΔS° for a typical Diels–Alder reaction, the microscopic reverse of a Diels–Alder reaction becomes favorable at high temperatures, although this is of synthetic importance for only a limited range of Diels-Alder adducts, generally with some special structural features; this reverse reaction is known as the retro-Diels–Alder reaction.

Polyketides are a class of natural products derived from a precursor molecule consisting of a chain of alternating ketone (or reduced forms of a ketone) and methylene groups: (-CO-CH2-). First studied in the early 20th century, discovery, biosynthesis, and application of polyketides has evolved. It is a large and diverse group of secondary metabolites caused by its complex biosynthesis which resembles that of fatty acid synthesis. Because of this diversity, polyketides can have various medicinal, agricultural, and industrial applications. Many polyketides are medicinal or exhibit acute toxicity. Biotechnology has enabled discovery of more naturally-occurring polyketides and evolution of new polyketides with novel or improved bioactivity.

<span class="mw-page-title-main">Epothilone</span> Class of chemical compounds

Epothilones are a class of potential cancer drugs. Like taxanes, they prevent cancer cells from dividing by interfering with tubulin, but in early trials, epothilones have better efficacy and milder adverse effects than taxanes.

Polyketide synthases (PKSs) are a family of multi-domain enzymes or enzyme complexes that produce polyketides, a large class of secondary metabolites, in bacteria, fungi, plants, and a few animal lineages. The biosyntheses of polyketides share striking similarities with fatty acid biosynthesis.

<span class="mw-page-title-main">Aminodeoxychorismate synthase</span>

In enzymology, an aminodeoxychorismate synthase is an enzyme that catalyzes the chemical reaction

<span class="mw-page-title-main">Macrophomic acid</span> Chemical compound

Macrophomic acid is a fungal metabolite isolated from the fungus Macrophoma commelinae. The enzyme macrophomate synthase converts 5-acetyl-4-methoxy-6-methyl-2-pyrone to 4-acetyl-3-methoxy-5-methyl-benzoic acid through an unusual intermolecular Diels-Alder reaction. The pathway to formation of macrophomic acid suggests that the enzyme is a natural Diels-Alderase. Formation of this type of aromatic ring compound normally proceeds via the shikimate and polyketide pathways; however, the production of macrophomic acid by macrophomate synthase proceeds totally differently. Learning about the production of macrophomic acid by a possible natural Diels-Alderase enzyme is important in understanding enzyme catalytic mechanisms. This knowledge can then be applied to organic synthesis.

<span class="mw-page-title-main">Sporolides</span> Chemical compound

Sporolides A and B are polycyclic macrolides extracted from the obligate marine bacterium Salinispora tropica, which is found in ocean sediment. They are composed of a chlorinated cyclopenta[a]indene ring and a cyclohexenone moiety. They were the second group of compounds isolated from Salinispora, and were said to indicate the potential of marine actinomycetes as a source of novel secondary metabolites. The structures and absolute stereochemistries of both metabolites were elucidated using a combination of NMR spectroscopy and X-ray crystallography.

<span class="mw-page-title-main">Torreyanic acid</span> Group of chemical compounds

Torreyanic acid is a dimeric quinone first isolated and by Lee et al. in 1996 from an endophyte, Pestalotiopsis microspora. This endophyte is likely the cause of the decline of Florida torreya, an endangered species that is related to the taxol-producing Taxus brevifolia. The natural product was found to be cytotoxic against 25 different human cancer cell lines with an average IC50 value of 9.4 µg/mL, ranging from 3.5 (NEC) to 45 (A549) µg/mL. Torreyanic acid was found to be 5-10 times more potent in cell lines sensitive to protein kinase C (PKC) agonists, 12-o-tetradecanoyl phorbol-13-acetate (TPA), and was shown to cause cell death via apoptosis. Torreyanic acid also promoted G1 arrest of G0 synchronized cells at 1-5 µg/mL levels, depending on the cell line. It has been proposed that the eukaryotic translation initiation factor EIF-4a is a potential biochemical target for the natural compound.

<span class="mw-page-title-main">Codinaeopsin</span> Chemical compound

Codinaeopsin is an antimalarial isolated from a fungal isolate found in white yemeri trees (Vochysia guatemalensis) in Costa Rica. It is reported to have bioactivity against Plasmodium falciparum with an IC50 = 2.3 μg/mL (4.7 μM). Pure codinaeopsin was reported to be isolated with a total yield of 18 mg/mL from cultured fungus. The biosynthesis of codinaeopsin involves a polyketide synthase-nonribosomal peptide synthetase (PKS-NRPS) hybrid.

In Lewis acid catalysis of organic reactions, a metal-based Lewis acid acts as an electron pair acceptor to increase the reactivity of a substrate. Common Lewis acid catalysts are based on main group metals such as aluminum, boron, silicon, and tin, as well as many early and late d-block metals. The metal atom forms an adduct with a lone-pair bearing electronegative atom in the substrate, such as oxygen, nitrogen, sulfur, and halogens. The complexation has partial charge-transfer character and makes the lone-pair donor effectively more electronegative, activating the substrate toward nucleophilic attack, heterolytic bond cleavage, or cycloaddition with 1,3-dienes and 1,3-dipoles.

Prosolanapyrone-III cycloisomerase is an enzyme with systematic name prosolanapyrone-III:(-)-solanapyrone A isomerase. This enzyme catalyses the following chemical reaction

<span class="mw-page-title-main">Marinone</span> Chemical compound

Marinone is an antibiotic made by marine actinomycetes.

<span class="mw-page-title-main">Maklamicin</span> Chemical compound

Maklamicin is a spirotetronate-class polyketide natural product. Isolated from Micromonospora sp. GMKU326 found in the root of Maklam phueak, it displays antibiotic activity against Gram-positive bacterial strains Micrococcus luteus, Bacillus subtilis, Bacillius cereus, Staphylococcus aureus, and Enterococcus faecalis.

<span class="mw-page-title-main">Dihydromaltophilin</span> Chemical compound

Dihydromaltophilin, or heat stable anti-fungal factor (HSAF), is a secondary metabolite of Streptomyces sp. and Lysobacter enzymogenes. HSAF is a polycyclic tetramate lactam containing a single tetramic acid unit and a 5,5,6-tricyclic system. HSAF has been shown to have anti-fungal activity mediated through the disruption of the biosynthesis of Sphingolipid's by targeting a ceramide synthase unique to fungi.

Tylactone synthase or TYLS is a Type 1 polyketide synthase. TYLS is found in strains of Streptomyces fradiae and responsible for the synthesis of the macrolide ring, tylactone, the precursor of an antibiotic, tylosin. TYLS is composed of five large multi-functional proteins, TylGI-V. Each protein contains either one or two modules. Each module consists of a minimum of a Ketosynthase (KS), an Acyltransferase (AT), and an Acyl carrier protein (ACP) but may also contain a Ketoreductase (KR), Dehydrotase (DH), and Enoyl Reductase (ER) for additional reduction reactions. The domains of TYLS have similar activity domains to those found in other Type I polyketide synthase such as 6-Deoxyerythronolide B synthase (DEBS). The TYLS system also contains a loading module consisting of a ketosynthase‐like decarboxylase domain, an acyltransferase, and acyl carrier protein. The terminal Thioesterase terminates tylactone synthesis by cyclizing the macrolide ring. After the TYLS completes tylactone synthesis, the tylactone molecule is modified by oxidation at C-20 and C-23 and glycosylation of mycaminose, mycinose, and mycarose to produce tylosin.

Vinylcyclopropane [5+2] cycloaddition is a type of cycloaddition between a vinylcyclopropane (VCP) and an olefin or alkyne to form a seven-membered ring.

<span class="mw-page-title-main">Chlorotonil A</span> Polyketide product

Chlorotonil A is a polyketide natural product produced by the myxobacterium Sorangium cellulosum So ce1525. It displays antimalarial activity in an animal model, and has in vitro antibacterial and antifungal activity. The activity of chlorotonil A has been attributed to the gem-dichloro-1,3-dione moiety, which is a unique functionality in polyketides. In addition to its unique halogenation, the structure of chlorotonil A has also garnered interest due to its similarity to anthracimycin, a polyketide natural product with antibiotic activity against Gram-positive bacteria.

<span class="mw-page-title-main">Agglomerin</span> Bacterial metablolites

Agglomerins are bacterial natural products, identified as metabolites of Pantoea agglomerans which was isolated in 1989 from river water in Kobe, Japan. They belong to the class of tetronate antibiotics, which include tetronomycin, tetronasin, and abyssomicin C. The members of the agglomerins differ only in the composition of the acyl chain attached to the tetronate ring. They possess antibiotic activity against anaerobic bacteria and weak activity against aerobic bacteria in vitro. The structures were solved in 1990. Agglomerin A is the major component (38%), followed by agglomerin B (30%), agglomerin C (24%), and agglomerin D (8%).

<span class="mw-page-title-main">Conipyridoin E</span> Chemical compound

Conipyridoin E is a tetramic acid derivative produced by the fungus Coniochaeta cephalothecoides which was found on a Tibetan Plateau. This natural product has been shown to exhibit antibacterial and antifungal activity against a variety of bacteria, such as Staphylococcus aureus, methicillin-resistant Staphyloccusaureus, and Enterococcus faecalis with MIC50 values of around 0.97 μM. Isolation of a number of analogs of conipyridoin has been accomplished by Han et al. in order to discover novel antibiotic natural products to combat antibiotic resistance.

<span class="mw-page-title-main">Spirotetronate cyclase AbyU</span> An enzyme

Spirotetronate cyclase AbyU is an enzyme responsible for catalyzing the Diels-Alder reaction in the abyssomicin C biosynthetic pathway. A key step in the biosynthesis of this compound catalyzed by AbyU involves intramolecular [4+2] cycloaddition—also known as the Diels-Alder reaction—to form a heterobicyclic ring system precursor consisting of tetronic acid and a cyclohexene ring that are spiro-linked.

References

  1. 1 2 Kanchanabanca, C.; Tao, W.; Hong, H.; Liu, Y.; Hanh, F.; Samborskyy, M.; Deng, Z.; Sun, Y.; Leadlay, P.F. (2013). "Unusual acetylation-elimination in the formation of tetronate antibiotics". Angewandte Chemie International Edition. 52 (22): 5785–8. doi:10.1002/anie.201301680. PMID   23606658.
  2. Nicolaou, K.C.; Harrison, S.T. (2006). "Total Synthesis of Abyssomicin C and atrop-Abyssomicin C**". Angewandte Chemie International Edition. 45 (20): 3256–60. doi:10.1002/anie.200601116. PMID   16634106.
  3. Keller, S.; Nicholson, G.; Drahl, C.; Sorensen, E.; Fiedler, H.P.; Süssmuth, R.D. (2007). "Abyssomicins G and H and atrop-abyssomicin C from the marine Verrucosispora strain AB-18-032". Journal of Antibiotics. 60 (6): 391–4. doi: 10.1038/ja.2007.54 . PMID   17617698.
  4. Keller S, Schadt HS, Ortel I, Süssmuth RD (2007). "Action of atrop-abyssomicin C as an inhibitor of 4-amino-4-deoxychorismate synthase PabB". Angew. Chem. Int. Ed. Engl. 46 (43): 8284–6. doi:10.1002/anie.200701836. PMID   17886307.
  5. Nicolaou, K.C.; Harrison, S.T.; Chen, J.S. (2009). "Discoveries from the Abyss: The Abyssomicins and Their Total Synthesis". Synthesis. 2009 (1): 33–42. doi:10.1055/s-0028-1083259. PMC   2677807 . PMID   20047014.
  6. Nicolaou, K.C.; Harrison, S.T. (2007). "Total synthesis of abyssomicin C, atrop-abyssomicin C, and abyssomicin D: implications for natural origins of atrop-abyssomicin C". Journal of the American Chemical Society. 129 (2): 429–40. doi:10.1021/ja067083p. PMID   17212423.
  7. 1 2 3 Savic, V. (2013). "Chapter 5 – Abyssomicins: Isolation, Properties, and Synthesis". Studies in Natural Products Chemistry. 40: 133–172. doi:10.1016/B978-0-444-59603-1.00005-9.
  8. 1 2 Gottardi, E; Krawczyk, J.; von Suchodoletz, H.; Schadt, S.; Mühlenweg, A.; Uguru, G.; Süssmuth, R.D. (2011). "Abyssomicin Biosynthesis: Formation of an Unusual Polyketide, Antibiotic‐Feeding Studies and Genetic Analysis". ChemBioChem. 12 (9): 1401–1410. doi:10.1002/cbic.201100172. PMC   3625739 . PMID   21656887.
  9. Vieweg, L.; Reichau, S.; Schobert, R; Leadlay, P.F.; Süssmuth, R. (2014). "Recent advances in the field of bioactive tetronates". Natural Product Reports. 31 (11): 1554–1584. doi: 10.1039/c4np00015c . PMID   24965099.
  10. Hashimoto, T.; Hashimoto, J.; Teruya, K.; Hirano, T.; Shin-ya, K.; Ikeda, H.; Liu, H.W.; Nishiyama, M.; Kuzuyama T. (2015). "Biosynthesis of versipelostatin: identification of an enzyme-catalyzed [4+2]-cycloaddition required for macrocyclization of spirotetronate-containing polyketides". Journal of the American Chemical Society. 137 (2): 572–5. doi:10.1021/ja510711x. PMC   4308742 . PMID   25551461.