Polyketide synthase

Last updated

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. [1] [2]

Contents

The PKS genes for a certain polyketide are usually organized in one operon or in gene clusters. Type I and type II PKSs form either large modular protein complexes or dissociable molecular assemblies; type III PKSs exist as smaller homodimeric proteins. [3] [4]

Classification

Reaction mechanisms of type I, II and III PKSs. Decarboxylation of malonyl unit followed by thio-Claisen condensation. a) (cis-AT) type I PKS with acyl carrier protein (ACP), keto synthase (KS) and acyl transferase (AT) domains covalently bound to another . b) Type II PKS with KSa-KSb heterodimer and ACP as separate proteins. c) ACP-independent Type III PKS. PKS Overview.jpg
Reaction mechanisms of type I, II and III PKSs. Decarboxylation of malonyl unit followed by thio-Claisen condensation. a) (cis-AT) type I PKS with acyl carrier protein (ACP), keto synthase (KS) and acyl transferase (AT) domains covalently bound to another . b) Type II PKS with KSα-KSβ heterodimer and ACP as separate proteins. c) ACP-independent Type III PKS.

PKSs can be classified into three types:

Modules and domains

Biosynthesis of the doxorubicin precursor, ie-rhodomycinone. The polyketide synthase reactions are shown on top. Anthracycline aglycone biosyn1.png
Biosynthesis of the doxorubicin precursor, є-rhodomycinone. The polyketide synthase reactions are shown on top.

Each type I polyketide-synthase module consists of several domains with defined functions, separated by short spacer regions. The order of modules and domains of a complete polyketide-synthase is as follows (in the order N-terminus to C-terminus):

Domains:

The polyketide chain and the starter groups are bound with their carboxy functional group to the SH groups of the ACP and the KS domain through a thioester linkage: R-C(=O)O H + H S-protein <=> R-C(=O)S-protein + H2O.

The ACP carrier domains are similar to the PCP carrier domains of nonribosomal peptide synthetases, and some proteins combine both types of modules.

Stages

The growing chain is handed over from one thiol group to the next by trans-acylations and is released at the end by hydrolysis or by cyclization (alcoholysis or aminolysis).

Starting stage:

Elongation stages:

Termination stage:

Pharmacological relevance

Polyketide synthases are an important source of naturally occurring small molecules used for chemotherapy. [15] For example, many of the commonly used antibiotics, such as tetracycline and macrolides, are produced by polyketide synthases. Other industrially important polyketides are sirolimus (immunosuppressant), erythromycin (antibiotic), lovastatin (anticholesterol drug), and epothilone B (anticancer drug). [16]

Polyketides are a large family of natural products widely used as drugs, pesticides, herbicides, and biological probes. [17]

There are antifungal and antibacterial polyketide compounds, namely ophiocordin and ophiosetin.[ citation needed ]

And are researched for the synthesis of biofuels and industrial chemicals. [18]

Ecological significance

Only about 1% of all known molecules are natural products, yet it has been recognized that almost two thirds of all drugs currently in use are at least in part derived from a natural source. [19] This bias is commonly explained with the argument that natural products have co-evolved in the environment for long time periods and have therefore been pre-selected for active structures. Polyketide synthase products include lipids with antibiotic, antifungal, antitumor, and predator-defense properties; however, many of the polyketide synthase pathways that bacteria, fungi and plants commonly use have not yet been characterized. [20] [21] Methods for the detection of novel polyketide synthase pathways in the environment have therefore been developed. Molecular evidence supports the notion that many novel polyketides remain to be discovered from bacterial sources. [22] [23]

See also

Related Research Articles

<span class="mw-page-title-main">Oxytetracycline</span> Antibiotic

Oxytetracycline is a broad-spectrum tetracycline antibiotic, the second of the group to be discovered.

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

Mupirocin, sold under the brand name Bactroban among others, is a topical antibiotic useful against superficial skin infections such as impetigo or folliculitis. It may also be used to get rid of methicillin-resistant S. aureus (MRSA) when present in the nose without symptoms. Due to concerns of developing resistance, use for greater than ten days is not recommended. It is used as a cream or ointment applied to the skin.

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.

Nonribosomal peptides (NRP) are a class of peptide secondary metabolites, usually produced by microorganisms like bacteria and fungi. Nonribosomal peptides are also found in higher organisms, such as nudibranchs, but are thought to be made by bacteria inside these organisms. While there exist a wide range of peptides that are not synthesized by ribosomes, the term nonribosomal peptide typically refers to a very specific set of these as discussed in this article.

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

Mycolactone is a polyketide-derived macrolide produced and secreted by a group of very closely related pathogenic Mycobacteria species that have been assigned a variety of names including, M. ulcerans, M. liflandii, M. pseudoshottsii, and some strains of M. marinum. These mycobacteria are collectively referred to as mycolactone-producing mycobacteria or MPM.

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

Chalcone synthase or naringenin-chalcone synthase (CHS) is an enzyme ubiquitous to higher plants and belongs to a family of polyketide synthase enzymes (PKS) known as type III PKS. Type III PKSs are associated with the production of chalcones, a class of organic compounds found mainly in plants as natural defense mechanisms and as synthetic intermediates. CHS was the first type III PKS to be discovered. It is the first committed enzyme in flavonoid biosynthesis. The enzyme catalyzes the conversion of 4-coumaroyl-CoA and malonyl-CoA to naringenin chalcone.

<span class="mw-page-title-main">Biosynthesis of doxorubicin</span>

Doxorubicin (DXR) is a 14-hydroxylated version of daunorubicin, the immediate precursor of DXR in its biosynthetic pathway. Daunorubicin is more abundantly found as a natural product because it is produced by a number of different wild type strains of streptomyces. In contrast, only one known non-wild type species, streptomyces peucetius subspecies caesius ATCC 27952, was initially found to be capable of producing the more widely used doxorubicin. This strain was created by Arcamone et al. in 1969 by mutating a strain producing daunorubicin, but not DXR, at least in detectable quantities. Subsequently, Hutchinson's group showed that under special environmental conditions, or by the introduction of genetic modifications, other strains of streptomyces can produce doxorubicin. His group has also cloned many of the genes required for DXR production, although not all of them have been fully characterized. In 1996, Strohl's group discovered, isolated and characterized dox A, the gene encoding the enzyme that converts daunorubicin into DXR. By 1999, they produced recombinant Dox A, a Cytochrome P450 oxidase, and found that it catalyzes multiple steps in DXR biosynthesis, including steps leading to daunorubicin. This was significant because it became clear that all daunorubicin producing strains have the necessary genes to produce DXR, the much more therapeutically important of the two. Hutchinson's group went on to develop methods to improve the yield of DXR, from the fermentation process used in its commercial production, not only by introducing Dox A encoding plasmids, but also by introducing mutations to deactivate enzymes that shunt DXR precursors to less useful products, for example baumycin-like glycosides. Some triple mutants, that also over-expressed Dox A, were able to double the yield of DXR. This is of more than academic interest because at that time DXR cost about $1.37 million per kg and current production in 1999 was 225 kg per annum. More efficient production techniques have brought the price down to $1.1 million per kg for the non-liposomal formulation. Although DXR can be produced semi-synthetically from daunorubicin, the process involves electrophilic bromination and multiple steps and the yield is poor. Since daunorubicin is produced by fermentation, it would be ideal if the bacteria could complete DXR synthesis more effectively.

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

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

Psymberin, also known as irciniastatin A, is a cytotoxin derived from sea sponges. It was discovered by two independent research groups, one led by Dr. Phil Crews and one led by Dr. Jean Schmidt, in 2004. Psymberin was found to be highly bioactive as it showed LC50s at nanomolar concentrations against various types of tumors.

Streptogramin A is a group of antibiotics within the larger family of antibiotics known as streptogramins. They are synthesized by the bacteria Streptomyces virginiae. The streptogramin family of antibiotics consists of two distinct groups: group A antibiotics contain a 23-membered unsaturated ring with lactone and peptide bonds while group B antibiotics are depsipeptides. While structurally different, these two groups of antibiotics act synergistically, providing greater antibiotic activity than the combined activity of the separate components. These antibiotics have until recently been commercially manufactured as feed additives in agriculture, although today there is increased interest in their ability to combat antibiotic-resistant bacteria, particularly vancomycin-resistant bacteria.

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

Pikromycin was studied by Brokmann and Hekel in 1951 and was the first antibiotic macrolide to be isolated. Pikromycin is synthesized through a type I polyketide synthase system in Streptomyces venezuelae, a species of Gram-positive bacterium in the genus Streptomyces. Pikromycin is derived from narbonolide, a 14-membered ring macrolide. Along with the narbonolide backbone, pikromycin includes a desosamine sugar and a hydroxyl group. Although Pikromycin is not a clinically useful antibiotic, it can be used as a raw material to synthesize antibiotic ketolide compounds such as ertythromycins and new epothilones.

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

Apratoxin A - is a cyanobacterial secondary metabolite, known as a potent cytotoxic marine natural product. It is a derivative of the Apratoxin family of cytotoxins. The mixed peptide-polyketide natural product comes from a polyketide synthase/non-ribosomal peptide synthase pathway (PKS/NRPS). This cytotoxin is known for inducing G1-phase cell cycle arrest and apoptosis. This natural product's activity has made it a popular target for developing anticancer derivatives.

<span class="mw-page-title-main">Ketoacyl synthase</span> Catalyst for a key step in fatty acid synthesis

Ketoacyl synthases (KSs) catalyze the condensation reaction of acyl-CoA or acyl-acyl ACP with malonyl-CoA to form 3-ketoacyl-CoA or with malonyl-ACP to form 3-ketoacyl-ACP. This reaction is a key step in the fatty acid synthesis cycle, as the resulting acyl chain is two carbon atoms longer than before. KSs exist as individual enzymes, as they do in type II fatty acid synthesis and type II polyketide synthesis, or as domains in large multidomain enzymes, such as type I fatty acid synthases (FASs) and polyketide synthases (PKSs). KSs are divided into five families: KS1, KS2, KS3, KS4, and KS5.

Fostriecin is a type I polyketide synthase (PKS) derived natural product, originally isolated from the soil bacterium Streptomyces pulveraceus. It belongs to a class of natural products which characteristically contain a phosphate ester, an α,β-unsaturated lactam and a conjugated linear diene or triene chain produced by Streptomyces. This class includes structurally related compounds cytostatin and phoslactomycin. Fostriecin is a known potent and selective inhibitor of protein serine/threonine phosphatases, as well as DNA topoisomerase II. Due to its activity against protein phosphatases PP2A and PP4 which play a vital role in cell growth, cell division, and signal transduction, fostriecin was looked into for its antitumor activity in vivo and showed in vitro activity against leukemia, lung cancer, breast cancer, and ovarian cancer. This activity is thought to be due to PP2A's assumed role in regulating apoptosis of cells by activating cytotoxic T-lymphocytes and natural killer cells involved in tumor surveillance, along with human immunodeficiency virus-1 (HIV-1) transcription and replication.

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

Borrelidin is an 18-membered polyketide macrolide derived from several Streptomyces species. First discovered in 1949 from Streptomyces rochei, Borrelidin shows antibacterial activity by acting as an inhibitor of threonyl-tRNA synthetase and features a nitrile moiety, a unique functionality in natural products., Borrelidin also exhibits potent angiogenesis inhibition, which was shown in a rat aorta matrix model. Other studies have been performed to show that low concentrations of borrelidin can suppress growth and induce apoptosis in malignant acute lymphoblastic leukemia cells. Borredlidin's antimalarial activity has also been shown in vitro and in vivo.

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

Swinholides are dimeric 42 carbon-ring polyketides that exhibit a 2-fold axis of symmetry. Found mostly in the marine sponge Theonella, swinholides encompass cytotoxic and antifungal activities via disruption of the actin skeleton. Swinholides were first described in 1985 and the structure and stereochemistry were updated in 1989 and 1990, respectively. Thirteen swinholides have been described in the literature, including close structural compounds such as misakinolides/bistheonellides, ankaraholides, and hurgholide A It is suspected that symbiotic microbes that inhabit the sponges rather than the sponges themselves produce swinholides since the highest concentration of swinholides are found in the unicellular bacterial fraction of sponges and not in the sponge fraction or cyanobacteria fraction that also inhabit the sponges.

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.

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

Phoslactomycin (PLM) is a natural product from the isolation of Streptomyces species. This is an inhibitor of the protein serine/threonine phosphatase which is the protein phosphate 2A (PP2A). The PP2A involves the growth factor of the cell such as to induce the formation of mitogen-activated protein interaction and playing a role in cell division and signal transduction. Therefore, PLM is used for the drug that prevents the tumor, cancer, or bacteria. There are nowsaday has 7 kinds of different PLM from PLM A to PLM G which differ the post-synthesis from the biosynthesis of PLM.

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

Prescopranone is a key intermediate in the biosynthesis of scopranones. Prescopranone is the precursor to scopranone A, scopranone B, and scopranone C, which are produced by Streptomyces sp. BYK-11038.

Andrimid is an antibiotic natural product that is produced by the marine bacterium Vibrio coralliilyticus. Andrimid is an inhibitor of fatty acid biosynthesis by blocking the carboxyl transfer reaction of acetyl-CoA carboxylase (ACC).

References

  1. Khosla, C.; Gokhale, R. S.; Jacobsen, J. R.; Cane, D. E. (1999). "Tolerance and Specificity of Polyketide Synthases". Annual Review of Biochemistry. 68: 219–253. doi:10.1146/annurev.biochem.68.1.219. PMID   10872449.
  2. Jenke-Kodama, H.; Sandmann, A.; Müller, R.; Dittmann, E. (2005). "Evolutionary Implications of Bacterial Polyketide Synthases". Molecular Biology and Evolution. 22 (10): 2027–2039. doi: 10.1093/molbev/msi193 . PMID   15958783.
  3. Weng, Jing-Ke; Noel, Joseph P. (2012). "Structure–Function Analyses of Plant Type III Polyketide Synthases". Natural Product Biosynthesis by Microorganisms and Plants, Part A. Methods in Enzymology. Vol. 515. pp. 317–335. doi:10.1016/B978-0-12-394290-6.00014-8. ISBN   978-0-12-394290-6. PMID   22999180.
  4. Pfeifer, Blaine A.; Khosla, Chaitan (March 2001). "Biosynthesis of Polyketides in Heterologous Hosts". Microbiology and Molecular Biology Reviews. 65 (1): 106–118. doi:10.1128/MMBR.65.1.106-118.2001. PMC   99020 . PMID   11238987.
  5. Sattely, Elizabeth S.; Fischbach, Michael A.; Walsh, Christopher T. (2008). "Total biosynthesis: in vitro reconstitution of polyketide and nonribosomal peptide pathways". Natural Product Reports. 25 (4): 757–793. doi:10.1039/b801747f. PMID   18663394.
  6. Weissman, Kira J. (2020). "Bacterial Type I Polyketide Synthases". Comprehensive Natural Products III: 4–46. doi:10.1016/b978-0-12-409547-2.14644-x. ISBN   9780081026915. S2CID   201202295.
  7. Helfrich, Eric J. N.; Piel, Jörn (2016). "Biosynthesis of polyketides by trans-AT polyketide synthases". Natural Product Reports. 33 (2): 231–316. doi:10.1039/c5np00125k. PMID   26689670.
  8. "The polyketide metabolites". General Pharmacology: The Vascular System. 23 (6): 1228. November 1992. doi:10.1016/0306-3623(92)90327-g.
  9. Hertweck, Christian; Luzhetskyy, Andriy; Rebets, Yuri; Bechthold, Andreas (2007). "Type II polyketide synthases: gaining a deeper insight into enzymatic teamwork". Nat. Prod. Rep. 24 (1): 162–190. doi:10.1039/B507395M. PMID   17268612.
  10. Sattely, Elizabeth S.; Fischbach, Michael A.; Walsh, Christopher T. (2008). "Total biosynthesis: in vitro reconstitution of polyketide and nonribosomal peptide pathways". Natural Product Reports. 25 (4): 757–793. doi:10.1039/b801747f. PMID   18663394.
  11. Abe, Ikuro; Morita, Hiroyuki (2010). "Structure and function of the chalcone synthase superfamily of plant type III polyketide synthases". Natural Product Reports. 27 (6): 809–838. doi:10.1039/b909988n. PMID   20358127.
  12. Shen, B (April 2003). "Polyketide biosynthesis beyond the type I, II and III polyketide synthase paradigms". Current Opinion in Chemical Biology. 7 (2): 285–295. doi:10.1016/S1367-5931(03)00020-6. PMID   12714063.
  13. Wong, Chin Piow; Morita, Hiroyuki (2020). "Bacterial Type III Polyketide Synthases". Comprehensive Natural Products III: 250–265. doi:10.1016/b978-0-12-409547-2.14640-2. ISBN   9780081026915. S2CID   195410516.
  14. Shimizu, Yugo; Ogata, Hiroyuki; Goto, Susumu (3 January 2017). "Type III Polyketide Synthases: Functional Classification and Phylogenomics". ChemBioChem. 18 (1): 50–65. doi: 10.1002/cbic.201600522 . PMID   27862822. S2CID   45980356.
  15. Koehn, F. E.; Carter, G. T. (2005). "The evolving role of natural products in drug discovery". Nature Reviews Drug Discovery. 4 (3): 206–220. doi:10.1038/nrd1657. PMID   15729362. S2CID   32749678.
  16. Wawrik, B.; Kerkhof, L.; Zylstra, G. J.; Kukor, J. J. (2005). "Identification of Unique Type II Polyketide Synthase Genes in Soil". Applied and Environmental Microbiology. 71 (5): 2232–2238. Bibcode:2005ApEnM..71.2232W. doi:10.1128/AEM.71.5.2232-2238.2005. PMC   1087561 . PMID   15870305.
  17. Pankewitz, Florian; Hilker, Monika (May 2008). "Polyketides in insects: ecological role of these widespread chemicals and evolutionary aspects of their biogenesis". Biological Reviews. 83 (2): 209–226. doi:10.1111/j.1469-185X.2008.00040.x. PMID   18410406. S2CID   27702684.
  18. Cai, Wenlong; Zhang, Wenjun (1 April 2018). "Engineering modular polyketide synthases for production of biofuels and industrial chemicals". Current Opinion in Biotechnology. 50: 32–38. doi:10.1016/j.copbio.2017.08.017. PMC   5862724 . PMID   28946011.
  19. Von Nussbaum, F.; Brands, M.; Hinzen, B.; Weigand, S.; Häbich, D. (2006). "Antibacterial Natural Products in Medicinal Chemistry—Exodus or Revival?". Angewandte Chemie International Edition. 45 (31): 5072–5129. doi:10.1002/anie.200600350. PMID   16881035.
  20. Castoe, T. A.; Stephens, T.; Noonan, B. P.; Calestani, C. (2007). "A novel group of type I polyketide synthases (PKS) in animals and the complex phylogenomics of PKSs". Gene. 392 (1–2): 47–58. doi:10.1016/j.gene.2006.11.005. PMID   17207587.
  21. Ridley, C. P.; Lee, H. Y.; Khosla, C. (2008). "Chemical Ecology Special Feature: Evolution of polyketide synthases in bacteria". Proceedings of the National Academy of Sciences. 105 (12): 4595–4600. Bibcode:2008PNAS..105.4595R. doi: 10.1073/pnas.0710107105 . PMC   2290765 . PMID   18250311.
  22. Metsä-Ketelä, M.; Salo, V.; Halo, L.; Hautala, A.; Hakala, J.; Mäntsälä, P.; Ylihonko, K. (1999). "An efficient approach for screening minimal PKS genes from Streptomyces". FEMS Microbiology Letters. 180 (1): 1–6. doi:10.1016/S0378-1097(99)00453-X. PMID   10547437.
  23. Wawrik, B.; Kutliev, D.; Abdivasievna, U. A.; Kukor, J. J.; Zylstra, G. J.; Kerkhof, L. (2007). "Biogeography of Actinomycete Communities and Type II Polyketide Synthase Genes in Soils Collected in New Jersey and Central Asia". Applied and Environmental Microbiology. 73 (9): 2982–2989. Bibcode:2007ApEnM..73.2982W. doi:10.1128/AEM.02611-06. PMC   1892886 . PMID   17337547.