Indolocarbazole

Last updated
Structural formula of staurosporine, the first indolocarbazole to be isolated Staurosporine.svg
Structural formula of staurosporine, the first indolocarbazole to be isolated

Indolocarbazoles (ICZs) are a class of compounds that are under current study due to their potential as anti-cancer as well as antimicrobial drugs and the prospective number of derivatives and uses found from the basic backbone alone. First isolated in 1977, [1] a wide range of structures and derivatives have been found or developed throughout the world. Due to the extensive number of structures available, this review will focus on the more important groups here while covering their occurrence, biological activity, biosynthesis, and laboratory synthesis.

Contents

Chemical classification

Indolocarbazoles belong to the alkaloid sub-class of bisindoles. The most frequently isolated indolocarbazoles are Indolo(2,3-a)carbazoles; the most common subgroup of the Indolo(2,3-a)carbazoles are the Indolo(2,3-a)pyrrole(3,4-c)carbazoles. These can be divided into two major classes - halogenated (chlorinated) with a fully oxidized C-7 carbon with only one indole nitrogen containing a β-glycosidic bond and the second class consists of both indole nitrogen glycosylated, non-halogenated, and a fully reduced C-7 carbon.

Occurrence

The first isolated ICZ, dubbed staurosporine (STA) was in 1977 from a culture of Streptomyces staurosporeus found in a soil sample from Iwate Prefecture, Japan. [2] The proper stereochemistry was not proven until 1994. Over the course of the next decade, further study of the compound showed some fungal inhibition, hypotensive activity, and most importantly, a broad protein kinase inhibitor. The next landmark discovery came with the detection of rebeccamycin (REB) in a sample of Lechevalieria aerocolonigenes, again in soil, but this time in a sample from Panama. REB was found to act against leukemia and melanoma in mice, and also against human adenocarcinoma cells. [3]

Since 1977, ICZs have been discovered all over the world in actinomycetes, bacteria commonly found in soil. Numerous forms have tested positive for anti-tumor activity, such as 7-hydroxy-STA and 7-oxo-STA2. Some of the strains from which ICZ compounds have been found are Actinomadura melliaura in Bristol Cove, San Diego County, California, Streptomyces hygroscopicus in Numazu Prefecture, Japan, Micromonospora sp. L-31-CLO-002 from Fuerteventura Island, Canary Islands, Spain, and Actinomadura sp. Strain 007 from Jiaozhou Bay, Shandong, China. The wide distribution of the various strains that produce these compounds is not surprising due to the number of properties these compounds can take on with limited functionalization on the species's part.

In addition to actinomycetes, ICZs have been found in slime molds (myxomycetes), blue-green algae (cyanobacteria, and marine invertebrates. Like the ones derived from actinomycetes, the ones found in myxomycetes cover an expansive range of derivatives and functionalizations. Two of the more important ones to date have been Arcyriacyanin A, which was found to inhibit a panel of human cancer cells by effecting PKC and protein tyrosine kinase, and lycogalic acid dimethyl ester A (found in Tokushima, Japan from Lycogala epidendrum ), which showed strong antiviral activity. A few of the strains of myxomycetes studied are Arcyria ferruginea and Arcyria cinerea, both from Kochi Prefecture, Japan. [1]

Three species of cyanobacteria has been found to produce ICZ compounds. Nostoc sphaericum from Manoa Hawaii, Tolypothrix tjipanasensis from Vero Beach, Florida, and Fischerella ambigua strain 108b from Leggingen, Switzerland. An interesting note on the first two is that many of the ICZ derived from them do not have the annelated pyrrolo[3,4-c] unit. [4]

The final major group in which ICZs are found are various marine invertebrates. Three species of tunicate, one mollusk, one flatworm, and one sponge have been discovered in places ranging from Micronesia to New Zealand. Testing for further invertebrate production is ongoing by both genetic and phylum-based studies. [1]

Biological activity

Indolocarbazoles have been found to exhibit a wide range of activities, which makes their range of presence in nature unsurprising. Because of this variety, the following section will examine their modes of action in bacterial and mammalian cells independently, with special attention paid to cancer cell effects.[ citation needed ]

The general modes of action found in mammalian cells are inhibition of protein kinases, inhibition of eukaryotic DNA topoisomerase, and intercalative binding to DNA. The number of protein kinases thought to exist in the human genome exceeds six hundred, [5] making a nanomolar inhibitor such as STA extremely useful for both treatment of various diseases and study of protein kinases in a variety of functions. Since this discovery, a vast effort has been undergone to make highly specific STA and REB derivatives. [6] One of the major lessons learned from initial research on STA was the development of the pharmacophore model for a protein kinase inhibitor in which a bidentate hydrogen donating system flanked by various hydrophobic groups inserts into the binding site. The information derived from this original pharmacophore has led to the synthesis of highly specific inhibitors against a number of protein kinases, including PKC, cyclin-dependent kinases, G-protein coupled receptor kinases, tyrosine kinase, and cytomegalovirus pUL97 protein. [7]

Topoisomerase I and II cleave and relegate one and two sides of a DNA strand, respectively, and are consequently vital parts of cell reproduction. Studies have found that in REB-like structures, the imide function of the pyrrole segment acts to interact with Topoisomerase I, the main carbon backbone acts as an intercalative inhibitor, and the sugar moiety undergoes DNA groove binding. [6] The latter two actually act in unison due to the three-dimensional structure of a glycosylated REB molecule. [3] The Top1 inhibitor section binds to cleavable DNA-Top1 complexes so as to prevent the relegation step. Because of this, sensitivity is based on quantity of Top1 present, making cells undergoing constant reproduction and growth (namely tumor cells) most vulnerable.

At this point, bacterial inhibition of Top1 has not been found using ICZs. Because of this, it is thought that most of the anti-cell growth function of ICZs comes from inhibition of various protein kinase groups and intercalative DNA binding. [8] [9] Studies on Streptomyces griseus with in vitro protein labelling have led to inhibition of a wide range of cellular functions. This led to the theory that there were several eukaryotic protein kinases present required for secondary metabolism. [10]

Some indolocarbazoles possess antimicrobial activity and act on bacteria in both a direct [11] and host-directed manner [12] [13] . The indolocarbazole GW296115X (also known as 3744W) has shown activity against intracellular pathogens, including human cytomegalovirus, Staphylococcus aureus and Mycobacterium abcessus [12] [13] [7] .

Biosynthesis

Unfortunately, only biosynthesis of REB, STA, and K252a have been studied in depth. This section will emphasize the REB pathway due to how well studied it is. The pathway begins with the modification of L-tryptophan to 7-chloro-L-tryptophan. This is done by catalysis using RebH in vitro halogenation and RebF (a flavin reductase) to provide FADH2 for the halogenase. [14] RebO (a tryptophan oxidase) then deaminates, after which it is further reacted with another one of itself and RebD (a heme containing oxidase). This forms the majority of the carbon backbone, which then undergoes decarboxylative ring closure using RebC and RebP. A glycosylation occurs using RebG and NDP-D-glucose, which finally goes through methylation by RebM. [15] [16] [17] [18] [19] [20] These latter tailoring enzymes have been noted as permissive in terms of both aglycons/acceptors and glycosyl/alkyl donors. [21] [22] [23] A parallel pathway has been put forth for the structurally related disaccharide-substituted indolocarbazole AT2433, the aminopentose of which is also found appended to the 10-membered enediyne calicheamicin. [24]

Information for this pathway, along with those of K252a and STA, [25] was derived from information on known genes, enzymes, and intermediates. The two types of studies done on these pathways are in vivo studies of gene disruption of L. aerocolonigenes or recombinant strains of S. albus. [1] The second type of experiment consisted of in vitro experiments done on cell extracts.

Synthesis

Laboratory synthesis of ICZs has been a topic of great interest since their discovery. Unfortunately, due to the somewhat complex nature of the molecule and the high level of reactivity of carbons on indole molecules, a facile high yield synthesis has yet to be found. Despite this, there have been many ways found to produce this compound in its various forms. [26] Of special interest is one of the better REB syntheses, found in 1999. The process begins by producing 7-chloroindole-3-acetamide by treating 7-chloroindole with a series of reagents, shown farther down. This molecule is then glycosylated and reacted with methyl 7-chloroindole-3-glyoxylate to produce an intermediate that goes on to stabilize into the final product. While this process is one of the better ones to date, it is still work and time intensive, going through 12 total steps and only yielding 12%. [27]

Further developments

Ever since the birth of ICZ research in the late seventies, the field has been burgeoning with continued advances in both technology and organic chemistry techniques. While only a handful of ICZ based compounds have made it past stage II clinical trials, the sheer variety that these molecules can take on leaves much still unexplored territory. Of particular recent interest in synthesis techniques is the use of palladium based catalysts, which have been found to be excellent activators for use in formation of carbon-carbon bonds. [28] [29]

Related Research Articles

<span class="mw-page-title-main">Nicotinamide adenine dinucleotide</span> Chemical compound which is reduced and oxidized

Nicotinamide adenine dinucleotide (NAD) is a coenzyme central to metabolism. Found in all living cells, NAD is called a dinucleotide because it consists of two nucleotides joined through their phosphate groups. One nucleotide contains an adenine nucleobase and the other, nicotinamide. NAD exists in two forms: an oxidized and reduced form, abbreviated as NAD+ and NADH (H for hydrogen), respectively.

<span class="mw-page-title-main">Chemical biology</span> Scientific discipline

Chemical biology is a scientific discipline between the fields of chemistry and biology. The discipline involves the application of chemical techniques, analysis, and often small molecules produced through synthetic chemistry, to the study and manipulation of biological systems. Although often confused with biochemistry, which studies the chemistry of biomolecules and regulation of biochemical pathways within and between cells, chemical biology remains distinct by focusing on the application of chemical tools to address biological questions.

<span class="mw-page-title-main">Indole-3-acetic acid</span> Chemical compound

Indole-3-acetic acid is the most common naturally occurring plant hormone of the auxin class. It is the best known of the auxins, and has been the subject of extensive studies by plant physiologists. IAA is a derivative of indole, containing a carboxymethyl substituent. It is a colorless solid that is soluble in polar organic solvents.

<span class="mw-page-title-main">Glycosyltransferase</span> Class of enzymes

Glycosyltransferases are enzymes that establish natural glycosidic linkages. They catalyze the transfer of saccharide moieties from an activated nucleotide sugar to a nucleophilic glycosyl acceptor molecule, the nucleophile of which can be oxygen- carbon-, nitrogen-, or sulfur-based.

<span class="mw-page-title-main">Lipid signaling</span> Biological signaling using lipid molecules

Lipid signaling, broadly defined, refers to any biological cell signaling event involving a lipid messenger that binds a protein target, such as a receptor, kinase or phosphatase, which in turn mediate the effects of these lipids on specific cellular responses. Lipid signaling is thought to be qualitatively different from other classical signaling paradigms because lipids can freely diffuse through membranes. One consequence of this is that lipid messengers cannot be stored in vesicles prior to release and so are often biosynthesized "on demand" at their intended site of action. As such, many lipid signaling molecules cannot circulate freely in solution but, rather, exist bound to special carrier proteins in serum.

p21 Protein

p21Cip1, also known as cyclin-dependent kinase inhibitor 1 or CDK-interacting protein 1, is a cyclin-dependent kinase inhibitor (CKI) that is capable of inhibiting all cyclin/CDK complexes, though is primarily associated with inhibition of CDK2. p21 represents a major target of p53 activity and thus is associated with linking DNA damage to cell cycle arrest. This protein is encoded by the CDKN1A gene located on chromosome 6 (6p21.2) in humans.

Topoisomerase inhibitors are chemical compounds that block the action of topoisomerases, which are broken into two broad subtypes: type I topoisomerases (TopI) and type II topoisomerases (TopII). Topoisomerase plays important roles in cellular reproduction and DNA organization, as they mediate the cleavage of single and double stranded DNA to relax supercoils, untangle catenanes, and condense chromosomes in eukaryotic cells. Topoisomerase inhibitors influence these essential cellular processes. Some topoisomerase inhibitors prevent topoisomerases from performing DNA strand breaks while others, deemed topoisomerase poisons, associate with topoisomerase-DNA complexes and prevent the re-ligation step of the topoisomerase mechanism. These topoisomerase-DNA-inhibitor complexes are cytotoxic agents, as the un-repaired single- and double stranded DNA breaks they cause can lead to apoptosis and cell death. Because of this ability to induce apoptosis, topoisomerase inhibitors have gained interest as therapeutics against infectious and cancerous cells.

Aralkylamine <i>N</i>-acetyltransferase Class of enzymes

Aralkylamine N-acetyltransferase (AANAT), also known as arylalkylamine N-acetyltransferase or serotonin N-acetyltransferase (SNAT), is an enzyme that is involved in the day/night rhythmic production of melatonin, by modification of serotonin. It is in humans encoded by the ~2.5 kb AANAT gene containing four exons, located on chromosome 17q25. The gene is translated into a 23 kDa large enzyme. It is well conserved through evolution and the human form of the protein is 80 percent identical to sheep and rat AANAT. It is an acetyl-CoA-dependent enzyme of the GCN5-related family of N-acetyltransferases (GNATs). It may contribute to multifactorial genetic diseases such as altered behavior in sleep/wake cycle and research is on-going with the aim of developing drugs that regulate AANAT function.

<span class="mw-page-title-main">Cyclin-dependent kinase 4</span> Human protein

Cyclin-dependent kinase 4 also known as cell division protein kinase 4 is an enzyme that in humans is encoded by the CDK4 gene. CDK4 is a member of the cyclin-dependent kinase family.

<span class="mw-page-title-main">Long-chain-fatty-acid—CoA ligase</span> Class of enzymes

The long chain fatty acyl-CoA ligase is an enzyme of the ligase family that activates the oxidation of complex fatty acids. Long chain fatty acyl-CoA synthetase catalyzes the formation of fatty acyl-CoA by a two-step process proceeding through an adenylated intermediate. The enzyme catalyzes the following reaction,

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

Staurosporine is a natural product originally isolated in 1977 from the bacterium Streptomyces staurosporeus. It was the first of over 50 alkaloids to be isolated with this type of bis-indole chemical structure. The chemical structure of staurosporine was elucidated by X-ray analysis of a single crystal and the absolute stereochemical configuration by the same method in 1994.

<span class="mw-page-title-main">PPP1CA</span> Enzyme

Serine/threonine-protein phosphatase PP1-alpha catalytic subunit is an enzyme that in humans is encoded by the PPP1CA gene.

<span class="mw-page-title-main">1-Aminocyclopropane-1-carboxylate synthase</span> Class of enzymes

The enzyme aminocyclopropane-1-carboxylic acid synthase catalyzes the synthesis of 1-Aminocyclopropane-1-carboxylic acid (ACC), a precursor for ethylene, from S-Adenosyl methionine, an intermediate in the Yang cycle and activated methyl cycle and a useful molecule for methyl transfer:

Strictosidine synthase (EC 4.3.3.2) is an enzyme in alkaloid biosynthesis that catalyses the condensation of tryptamine with secologanin to form strictosidine in a formal Pictet–Spengler reaction:

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

Betaenone B, like other betaenones, is a secondary metabolite isolated from the fungus Pleospora betae, a plant pathogen. Its phytotoxic properties have been shown to cause sugar beet leaf spots, which is characterized by black, pycnidia containing, concentric circles eventually leading to necrosis of the leaf tissue. Of the seven phytotoxins isolated in fungal leaf spots from sugar beet, betaenone B showed the least amount of phytotoxicity showing only 8% inhibition of growth while betaenone A and C showed 73% and 89% growth inhibition, respectively. Betaenone B is therefore not considered toxic to the plant, but will produce leaf spots when present in high concentrations (0.33 μg/μL). While the mechanism of action of betaenone B has yet to be elucidated, betaenone C has been shown to inhibit RNA and protein synthesis. Most of the major work on betaenone B, including the initial structure elucidation of betaenone A, B and C as well as the partial elucidation mechanism of biosynthesis, was presented in three short papers published between 1983 and 1988. The compounds were found to inhibit a variety of protein kinases signifying a possible role in cancer treatment.

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

Indole is an organic compound with the formula C6H4CCNH3. Indoles are derivatives of indole where one or more H's have been replaced by other groups. Indole is classified as an aromatic heterocycle. It has a bicyclic structure, consisting of a six-membered benzene ring fused to a five-membered pyrrole ring. Indoles are widely distributed in nature, most notably as amino acid tryptophan and neurotransmitter serotonin.

<span class="mw-page-title-main">BIM-1</span> Biological protein kinase C inhibitor

BIM-1 and the related compounds BIM-2, BIM-3, and BIM-8 are bisindolylmaleimide-based protein kinase C (PKC) inhibitors. These inhibitors also inhibit PDK1 explaining the higher inhibitory potential of LY33331 compared to the other BIM compounds a bisindolylmaleimide inhibitor toward PDK1.

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

The purinosome is a putative multi-enzyme complex that carries out de novo purine biosynthesis within the cell. It is postulated to include all six of the human enzymes identified as direct participants in this ten-step biosynthetic pathway converting phosphoribosyl pyrophosphate to inosine monophosphate:

Demethylrebeccamycin-D-glucose O-methyltransferase is an enzyme with systematic name S-adenosyl-L-methionine:demethylrebeccamycin-D-glucose O-methyltransferase. This enzyme catalyses the following chemical reaction

4′-Demethylrebeccamycin synthase (EC 4.3.3.5, arcyriaflavin A N-glycosyltransferase, RebG) is an enzyme with systematic name 4′-demethylrebeccamycin D-glucose-lyase. It catalyses the following chemical reaction

References

  1. 1 2 3 4 Sanchez C.; Mendez C.; Salas J. A. (2006). "Indolocarbazole natural products: occurrence, biosynthesis, and biological activity". Nat. Prod. Rep. 23 (6): 1007–1045. doi:10.1039/b601930g. PMID   17119643.
  2. Wada Y.; Nagasaki H.; Tokuda M.; Orito K. (2007). "Synthesis of N-protected staurosporinones". J. Org. Chem. 72 (6): 2008–14. doi:10.1021/jo062184r. PMID   17309305.
  3. 1 2 Long Byron H.; Rose William C.; Vyas Dolatrai M.; Matson James A.; Forenza S. (2002). "Discovery of antitumor indolocarbazoles: rebeccamycin, NSC 655649, and fluoroindolocarbazoles". Curr. Med. Chem. Anti-Cancer Agents. 2 (2): 255–266. doi:10.2174/1568011023354218. PMID   12678746.
  4. Knubel G.; Larsen L. K.; Moore R. E.; Levine I. A.; Patterson G. M. L. (1990). "Cytotoxic, antiviral indolocarbazoles from a blue-green alga belonging to the Nostocaceae". J. Antibiot. 43 (10): 1236–1239. doi: 10.7164/antibiotics.43.1236 . PMID   2175302.
  5. Manning G.; Whyte D. B.; Martinez R.; Hunter T.; Sudarsanam S. (2002). "The Protein Kinase Complement of the Human Genome". Science. 298 (5600): 1912–1916, 1933–1934. Bibcode:2002Sci...298.1912M. doi:10.1126/science.1075762. PMID   12471243. S2CID   26554314.
  6. 1 2 Prudhomme M (2004). "Biological targets of antitumor indolocarbazoles bearing a sugar moiety". Curr. Med. Chem. 4 (6): 509–521. doi:10.2174/1568011043352650. PMID   15579016.
  7. 1 2 Zimmermann A.; Wilts H.; Lenhardt M.; Hahn M.; Mertens T. (2000). "Indolocarbazoles exhibit strong antiviral activity against human cytomegalovirus and are potent inhibitors of the pUL97 protein kinase". Antiviral Res. 48 (1): 49–60. doi:10.1016/s0166-3542(00)00118-2. PMID   11080540.
  8. Iwai Y.; Omura S.; Li Z. (1994). "Indolocarbazole alkaloids produced by microorganisms. Staurosporine as the main". Kagaku to Seibutsu. 32: 463–469. doi: 10.1271/kagakutoseibutsu1962.32.463 .
  9. Pereira E. R., Fabre S., Sancelme M., Prudhomme M. (1995). "Antimicrobial activities of indolocarbazole and bis-indole protein kinase C inhibitors. II. Substitution on maleimide nitrogen with functional groups bearing a labile hydrogen". J. Antibiot. 48 (8): 863–868. doi: 10.7164/antibiotics.48.863 . PMID   7592032.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  10. Lum P. Y.; Armour C. D.; Stepaniants S. B.; Cavet G.; Wolf M. K.; Butler J. S.; Hinshaw J. C.; Garnier P.; Prestwich G. D.; Leonardson A.; Garrett-Engele P.; Rush C. M.; Bard M.; Schimmack G.; Phillips J. W.; Roberts C. J.; Shoemaker D. D. (2004). "Discovering modes of action for therapeutic compounds using a genome-wide screen of yeast heterozygotes". Cell. 116 (1): 121–137. doi: 10.1016/s0092-8674(03)01035-3 . PMID   14718172.
  11. Pereira, Elisabete Rodrigues; Fabre, Serge; Sancelme, Martine; Prudhomme, Michelle; Rapp, Maryse (1995-08-25). "Antimicrobial Activities of Indolocarbazole and Bis-indole Protein Kinase C Inhibitors II. Substitution on Maleimide Nitrogen with Functional Groups Bearing a Labile Hydrogen". The Journal of Antibiotics. 48 (8): 863–868. doi:10.7164/antibiotics.48.863. ISSN   0021-8820. PMID   7592032.
  12. 1 2 van den Biggelaar, Robin H. G. A.; Walburg, Kimberley V.; van den Eeden, Susan J. F.; van Doorn, Cassandra L. R.; Meiler, Eugenia; de Ries, Alex S.; Meijer, Annemarie H.; Ottenhoff, Tom H. M.; Saris, Anno (2024). "Identification of kinase modulators as host-directed therapeutics against intracellular methicillin-resistant Staphylococcus aureus". Frontiers in Cellular and Infection Microbiology. 14. doi: 10.3389/fcimb.2024.1367938 . ISSN   2235-2988. PMC   10999543 . PMID   38590439.
  13. 1 2 Richter, Adrian; Shapira, Tirosh; Av-Gay, Yossef (2019-12-20). "THP-1 and Dictyostelium Infection Models for Screening and Characterization of Anti-Mycobacterium abscessus Hit Compounds". Antimicrobial Agents and Chemotherapy. 64 (1). doi:10.1128/AAC.01601-19. ISSN   0066-4804. PMC   7187604 . PMID   31636068.
  14. Bitto, E; Huang, Y; Bingman, CA; Singh, S; Thorson, JS; Phillips GN, Jr (1 January 2008). "The structure of flavin-dependent tryptophan 7-halogenase RebH". Proteins. 70 (1): 289–93. doi: 10.1002/prot.21627 . PMID   17876823.
  15. Zhang G.; Shen J.; Cheng H.; Zhu L.; Fang L.; Luo S.; Muller Mark T.; Lee Gun E.; Wei L.; Du Y.; Sun D.; Wang Peng G. (2005). "Syntheses and biological activities of rebeccamycin analogues with uncommon sugars". J. Med. Chem. 48 (7): 2600–2611. doi:10.1021/jm0493764. PMID   15801850.
  16. Sanchez C.; Zhu L.; Brana A. F.; Salas A. P.; Rohr J.; Mendez C.; Salas J. A. (2005). "Combinatorial biosynthesis of antitumor indolocarbazole compounds". Proc. Natl. Acad. Sci. USA. 102 (2): 461–466. Bibcode:2005PNAS..102..461S. doi: 10.1073/pnas.0407809102 . PMC   544307 . PMID   15625109.
  17. Sanchez C.; Brana A. F.; Mendez C.; Salas J. A. (2006). "Reevaluation of the violacein biosynthetic pathway and its relationship to indolocarbazole biosynthesis". ChemBioChem. 7 (8): 1231–1240. doi:10.1002/cbic.200600029. PMID   16874749. S2CID   7668743.
  18. Nishizawa T.; Gruschow S.; Jayamaha D. H.; Nishizawa-Harada C.; Sherman D. H. (2006). "Enzymatic assembly of the bis-indole core of rebeccamycin". J. Am. Chem. Soc. 128 (3): 724–725. doi:10.1021/ja056749x. PMID   16417354.
  19. Hyun, CG; Bililign, T; Liao, J; Thorson, JS (3 January 2003). "The biosynthesis of indolocarbazoles in a heterologous E. coli host". ChemBioChem. 4 (1): 114–7. doi:10.1002/cbic.200390004. PMID   12512086. S2CID   37460017.
  20. Singh, S; McCoy, JG; Zhang, C; Bingman, CA; Phillips GN, Jr; Thorson, JS (15 August 2008). "Structure and mechanism of the rebeccamycin sugar 4'-O-methyltransferase RebM". The Journal of Biological Chemistry. 283 (33): 22628–36. doi: 10.1074/jbc.m800503200 . PMC   2504894 . PMID   18502766.
  21. Zhang, C; Albermann, C; Fu, X; Peters, NR; Chisholm, JD; Zhang, G; Gilbert, EJ; Wang, PG; Van Vranken, DL; Thorson, JS (May 2006). "RebG- and RebM-catalyzed indolocarbazole diversification". ChemBioChem. 7 (5): 795–804. doi:10.1002/cbic.200500504. PMID   16575939. S2CID   25017820.
  22. Zhang, C; Weller, RL; Thorson, JS; Rajski, SR (8 March 2006). "Natural product diversification using a non-natural cofactor analogue of S-adenosyl-L-methionine". Journal of the American Chemical Society. 128 (9): 2760–1. doi:10.1021/ja056231t. PMID   16506729.
  23. Singh, S; Zhang, J; Huber, TD; Sunkara, M; Hurley, K; Goff, RD; Wang, G; Zhang, W; Liu, C; Rohr, J; Van Lanen, SG; Morris, AJ; Thorson, JS (7 April 2014). "Facile chemoenzymatic strategies for the synthesis and utilization of S-adenosyl-(L)-methionine analogues". Angewandte Chemie International Edition in English. 53 (15): 3965–9. doi:10.1002/anie.201308272. PMC   4076696 . PMID   24616228.
  24. Gao, Q; Zhang, C; Blanchard, S; Thorson, JS (July 2006). "Deciphering indolocarbazole and enediyne aminodideoxypentose biosynthesis through comparative genomics: insights from the AT2433 biosynthetic locus". Chemistry & Biology. 13 (7): 733–43. doi: 10.1016/j.chembiol.2006.05.009 . PMID   16873021.
  25. Makino M.; Sugimoto H.; Shiro Y.; Asamizu S.; Onaka H.; Nagano S. (2007). "Crystal structures and catalytic mechanism of cytochrome P450 StaP that produces the indolocarbazole skeleton". Proc. Natl. Acad. Sci. USA. 104 (28): 11591–11596. Bibcode:2007PNAS..10411591M. doi: 10.1073/pnas.0702946104 . PMC   1913897 . PMID   17606921.
  26. Gu R.; Hameurlaine A.; Dehaen W. (2007). "Facile one-pot synthesis of 6-monosubstituted and 6,12-disubstituted 5,11-dihydroindolo[3,2-b]carbazoles and preparation of various functionalized derivatives". J. Org. Chem. 72 (19): 7207–7213. doi:10.1021/jo0711337. PMID   17696477.
  27. Faul M. M.; Winneroski L. L.; Krumrich C. A. (1999). "Synthesis of Rebeccamycin and 11-Dechlororebeccamycin". J. Org. Chem. 64 (7): 2465–2470. doi:10.1021/jo982277b.
  28. Trost B. M., Krische M. J., Berl V., Grenzer E. M. (2002). "Chemo-, Regio-, and Enantioselective Pd-Catalyzed Allylic Alkylation of Indolocarbazole Pro-aglycons". Org. Lett. 4 (12): 2005–2008. doi:10.1021/ol020046s. PMID   12049503.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  29. Saulnier M. G., Frennesson D. B., Deshpande M. S., Vyas D. M. (1995). "Synthesis of a rebeccamycin-related indolo[2,3-a]carbazole by palladium(0) catalyzed polyannulation". Tetrahedron Lett. 36 (43): 7841–7844. doi:10.1016/0040-4039(95)01644-w.{{cite journal}}: CS1 maint: multiple names: authors list (link)