Taxodone

Last updated
Taxodone
Taxodone.png
Taxodone Ball.png
Names
IUPAC name
6α,11-Dihydroxyabieta-7,9(11),13-trien-12-one
Systematic IUPAC name
(4bS,8aS,9S)-4,9-Dihydroxy-4b,8,8-trimethyl-2-(propan-2-yl)-5,6,7,8,8a,9-hexahydrophenanthren-3(4bH)-one
Other names
NSC122420, AC1L9XIL, CID457961
Identifiers
3D model (JSmol)
ChemSpider
PubChem CID
  • InChI=1S/C20H28O3/c1-11(2)13-9-12-10-14(21)18-19(3,4)7-6-8-20(18,5)15(12)17(23)16(13)22/h9-11,14,18,21,23H,6-8H2,1-5H3/t14-,18-,20+/m0/s1 Yes check.svgY
    Key: QEAIMIKGLGBTSA-ADLFWFRXSA-N Yes check.svgY
  • InChI=1/C20H28O3/c1-11(2)13-9-12-10-14(21)18-19(3,4)7-6-8-20(18,5)15(12)17(23)16(13)22/h9-11,14,18,21,23H,6-8H2,1-5H3/t14-,18-,20+/m0/s1
    Key: QEAIMIKGLGBTSA-ADLFWFRXBH
  • O=C1\C(=C/C=2/C(=C1/O)[C@]3(C)[C@@H]([C@@H](O)C=2)C(C)(C)CCC3)C(C)C
Properties
C20H28O3
Molar mass 316.441 g·mol−1
AppearanceGolden crystalline solid
Melting point 176 to 177 °C (349 to 351 °F; 449 to 450 K)
Insoluble
Solubility in chloroform, alcohol, hexane, etherSoluble
Related compounds
Related compounds
Taxodione
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
Yes check.svgY  verify  (what is  Yes check.svgYX mark.svgN ?)

Taxodone is a naturally occurring diterpenoid found in Taxodium distichum (bald cypress), Rosmarinus officinalis (rosemary), several salvia species and other plants, along with its oxidized rearrangement product, taxodione. Taxodone and taxodione exhibit anticancer, [1] [2] [3] antibacterial, [4] [5] [6] antioxidant, [7] antifungal, [8] insecticide, [9] and antifeedant [10] activities.

Contents

Discovery

Taxodone was first isolated in 1968 from the seeds of Taxodium distichum (Bald Cypress) by S. Morris Kupchan and coworkers. [1] They reported the structure determination and basic chemistry of taxodone and its oxidized rearrangement product, taxodione. [11] [12] Taxodone occurs naturally in the form of (+)-taxodone.

Occurrence

Taxodone and/or taxodione have been identified in several plants besides Taxodium distichum including: Rosmarinus officinalis (Rosemary), [13] Salvia barrelieri , [7] Metasequoia glyptostroboides (Dawn Redwood), [4] Salvia munzii (San Diego Sage), [14] Salvia moorcroftiana , [15] Salvia staminea , [16] Salvia clevelandii (Cleveland Sage), [17] Salvia hypargeia , [3] Salvia broussonetii , [18] Salvia montbretii, [19] [20] Salvia nipponica, [21] [22] Salvia verbenaca (Wild Clary), [23] Salvia lanigera , [24] [25] Salvia prionitis , [26] Salvia deserta , [27] Salvia phlomoides, [28] [29] and Plectranthus hereroensis [30]

Taxodone, taxodione and their reaction products have been used as archeological and geological biomarkers. [31] [32] [33] [34] [35] [36] [37]

Analogs of taxodone and taxodione have also been isolated. 2-hydroxy taxodone and 2-hydroxy-taxodione have been found in Salvia texana (Texas Sage). [38] 5,6-Didehydro-7-hydroxy-taxodone was found in Salvia munzii . [14] 7-Hydroxytaxodione, 7,7‘-bistaxodione, and 11,11‘-didehydroxy-7,7‘-dihydroxytaxodione were found in Salvia montbretti. [19] [20]

Activity

Taxodone and taxodione possess in vivo activity against Walker intramuscular carcinosarcoma 256 in rats (25 and 40 mg/kg, respectively) and in vitro activity against cells derived from human carcinoma of the nasopharynx (KB) (ED50 = 0.6 and 3 ug/ml respectively). [1] Taxodone and taxodione exhibit antifungal activity against wood decay fungi, with taxodione being especially active against Trametes versicolor and Fomitopsis palustris . [8] Taxodione exhibited the highest antioxidant activity among the tested diterpenoids from the roots of Salvia barrelieri . [7] Taxodone showed potent antibacterial effects against foodborne pathogenic bacteria, such as Listeria monocytogenes ATCC 19166, Salmonella typhimurium KCTC 2515, Salmonella enteritidis KCTC 2021, Escherichia coli ATCC 8739, Escherichia coli O157:H7 ATCC 43888, Enterobacter aerogenes KCTC 2190, Staphylococcus aureus ATCC 6538 and Staphylococcus aureus KCTC 1916 [4] Taxodone showed potent termicidal activity against the subterranean termite, Reticulitermes speratus Kolbe. [9] Taxodione depresses neuronal GABAA receptor-operated Cl-current (IGABA). [39] Taxodione may have potential in treatment of cardiovascular disease. [40]

The use of taxodone and taxodione to inhibit hair growth has been patented. [41] [42] [43] Treatment of benign prostate enlargement with taxodone has also been patented. [44]

Chemistry

Taxodone was the first isolated example of a quinone methide [45] [46] [47] [48] [49] [50] with a labile hydrogen adjacent to this reactive chromophore. [1] Kupchan demonstrated that taxodone aromatizes to a catechol ketone upon exposure to mild acid. Air oxidation of this catechol ketone affords taxodione.

Taxodone-to-taxodione.png

Synthesis

Taxodone rearranges easily in the presence of mild acids and reacts readily with nucleophiles. Although taxodone shows higher anticancer and antibacterial activity than taxodione it eluded creation in the laboratory for over 25 years because of its inherent instability. During this time several different groups reported syntheses of the more stable taxodione. [51] [52] [53] [54] [55] [56] [57] [58] [59] [60] [61] [62] [63] [64] [65] [66] [67] [68] [69]

In 1993 taxodone was synthesized for the first time in a 16 step sequence utilizing a unique phenol benzylic epoxide electron reorganization in the final step. [70] [71] As taxodone readily decomposes into taxodione this synthesis of taxodone also constitutes a formal synthesis of taxodione as well.

Taxodone synth ACS.png

Since the synthesis of taxodone there have been additional syntheses of taxodione and analogs. [6] [72] [73]

See also

Related Research Articles

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

Salvinorin A is the main active psychotropic molecule in Salvia divinorum. Salvinorin A is considered a dissociative hallucinogen.

<span class="mw-page-title-main">Pauson–Khand reaction</span>

The Pauson–Khand reaction is a chemical reaction described as a [2+2+1] cycloaddition between an alkyne, an alkene and carbon monoxide to form a α,β-cyclopentenone. Ihsan Ullah Khand (1935-1980) discovered the reaction around 1970, while working as a postdoctoral associate with Peter Ludwig Pauson (1925–2013) at the University of Strathclyde in Glasgow. Pauson and Khand's initial findings were intermolecular in nature, but starting a decade after the reaction's discovery, many intramolecular examples have been highlighted in both synthesis and methodology reports. This reaction was originally mediated by stoichiometric amounts of dicobalt octacarbonyl, but newer versions are both more efficient, enhancing reactivity and yield via utilizing different chiral auxiliaries for stereo induction, main group transition-metals, and additives.

<span class="mw-page-title-main">Curtius rearrangement</span> Chemical reaction

The Curtius rearrangement, first defined by Theodor Curtius in 1885, is the thermal decomposition of an acyl azide to an isocyanate with loss of nitrogen gas. The isocyanate then undergoes attack by a variety of nucleophiles such as water, alcohols and amines, to yield a primary amine, carbamate or urea derivative respectively. Several reviews have been published.

The reduction of nitro compounds are chemical reactions of wide interest in organic chemistry. The conversion can be effected by many reagents. The nitro group was one of the first functional groups to be reduced. Alkyl and aryl nitro compounds behave differently. Most useful is the reduction of aryl nitro compounds.

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

Methylecgonidine is a chemical intermediate derived from ecgonine or cocaine.

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

Ferruginol is a natural phenol with a terpenoid substructure. Specifically, it is a diterpene of the abietane chemical class, meaning it is characterized by three fused six-membered rings and alkyl functional groups. Ferruginol was first identified in 1939 by Brandt and Neubauer as the main component in the resin of the Miro tree and has since been isolated from other conifer species in the families Cupressaceae and Podocarpaceae. As a biomarker, the presence of ferruginol in fossils, mainly resin, is used to describe the density of these conifers in that particular biosphere throughout time.

In organic chemistry, an intramolecular Diels-Alder cycloaddition is a Diels–Alder reaction in which the diene and a dienophile are both part of the same molecule. The reaction leads to the formation of the same cyclohexene-like structure as usual for a Diels–Alder reaction, but as part of a more complex fused or bridged cyclic ring system. This reaction gives rise to various natural derivatives of decalin.

<span class="mw-page-title-main">Megaphone (molecule)</span> Chemical compound

Megaphone is a cytotoxic neolignan obtained from Aniba megaphylla, a flowering plant of Laurel family which gave the compound its name. Megaphone has also been prepared synthetically.

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

Totarol is a naturally produced diterpene that is bioactive as totarol. It was first isolated by McDowell and Easterfield from the heartwood of Podocarpus totara, a conifer tree found in New Zealand. Podocarpus totara was investigated for unique molecules due to the tree's increased resistance to rotting. Recent studies have confirmed totarol's unique antimicrobial and therapeutic properties. Consequently, totarol is a candidate for a new source of drugs and has been the goal of numerous syntheses.

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

Carnosic acid is a natural benzenediol abietane diterpene found in rosemary and common sage. Dried leaves of rosemary and sage contain 1.5 to 2.5% carnosic acid.

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

Callystatin A is a polyketide natural product from the leptomycin family of secondary metabolites. It was first isolated in 1997 from the marine sponge Callyspongia truncata which was collected from the Goto Islands in the Nagasaki Prefecture of Japan by the Kobayashi group. Since then its absolute configuration has been elucidated and callystatin A was discovered to have anti-fungal and anti-tumor activities with extreme potency against the human epidermoid carcinoma KB cells (IG50 = 10 pg/ml) and the mouse lymphocytic leukemia Ll210 cells (IG50 = 20 pg/ml).

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

A quinone methide is a type of conjugated organic compound that contain a cyclohexadiene with a carbonyl and an exocyclic methylidene or extended alkene unit. It is analogous to a quinone, but having one of the double bonded oxygens replaced with a carbon. The carbonyl and methylidene are usually oriented either ortho or para to each other. There are some examples of transient synthetic meta quinone methides.

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

In organic chemistry, a xylylene (sometimes quinone-dimethide) is any of the constitutional isomers having the formula C6H4(CH2)2. These compounds are related to the corresponding quinones and quinone methides by replacement of the oxygen atoms by CH2 groups. ortho- and para-xylylene are best known, although neither is stable in solid or liquid form. The meta form is a diradical. Certain substituted derivatives of xylylenes are however highly stable, such as tetracyanoquinodimethane and the xylylene dichlorides.

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

Abietane is a diterpene that forms the structural basis for a variety of natural chemical compounds such as abietic acid, carnosic acid, and ferruginol which are collectively known as abietanes or abietane diterpenes.

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

Carnosol is a phenolic diterpene found in the herbs rosemary and Mountain desert sage.

<i>beta</i>-Araneosene Chemical compound

β-Araneosene is a molecule first isolated in 1975 from the mold Sordaria araneosa by Borschberg. This unprecedented diterpene framework was given the name “araneosene”. In 1976, the skeletal class was renamed to “dolabellane” due to the isolation of several compounds containing this framework found from the sea hare Dolabella californica. Since their initial discovery, there are now more than 150 known dolabellanes, mostly isolated from marine sources.

Fétizon oxidation is the oxidation of primary and secondary alcohols utilizing the compound silver(I) carbonate absorbed onto the surface of celite also known as Fétizon's reagent first employed by Marcel Fétizon in 1968. It is a mild reagent, suitable for both acid and base sensitive compounds. Its great reactivity with lactols makes the Fétizon oxidation a useful method to obtain lactones from a diol. The reaction is inhibited significantly by polar groups within the reaction system as well as steric hindrance of the α-hydrogen of the alcohol.

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

Fascaplysin is a marine alkaloid based on 12H-pyrido[1–2-a:3,4-b′]diindole ring system. It was first isolated as a red pigment from the marine sponge Fascaplysinopsis bergquist collected in the South Pacific near Fiji in 1988. Fascaplysin possesses a broad range of in vitro biological activities including analgesic, antimicrobial, antifungal, antiviral, antimalarial, anti-angiogenic, and antiproliferative activity against numerous cancer cell lines.

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

Sugiol is a phenolic abietane derivative of ferruginol and can be used as a biomarker for specific families of conifers. The presence of sugiol can be used to identify the Cupressaceae s.1., podocarpaceae, and Araucaraiaceae families of conifers. The polar terpenoids are among the most resistant molecules to degradation besides n-alkanes and fatty acids, affording them high viability as biomarkers due to their longevity in the sedimentary record. Significant amounts of sugiol has been detected in fossil wood dated to the Eocene and Miocene periods, as well as a sample of Protopodocarpoxylon dated to the middle Jurassic.

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

Chamaecydin is a chemical compound with the molecular formula C30H40O3. It is made up of three six-membered rings and two five-membered rings and has one polar hydroxyl functional group. It is well preserved in the rock record and is only found in a specific family of conifers, the swamp cypress subfamily. The presence and abundance of chamaecydin in the rock record can reveal environmental changes in ancient biomes.

References

  1. 1 2 3 4 Kupchan, S. M.; Karim, A; Marcks, C. (1968). "Tumor inhibitors. XXXIV. Taxodione and taxodone, two novel diterpenoid quinone methide tumor inhibitors from Taxodium distichum". J. Am. Chem. Soc. 90 (21): 5923–5924. doi:10.1021/ja01023a061. PMID   5679178.
  2. Zaghloul A. M.; Gohar A. A.; Naiem Z. A.; Abdel Bar F. M. (2008). "Taxodione, a DNA-binding compound from Taxodium distichum L. (Rich.)". Z. Naturforsch. C . 63 (5–6): 355–360. doi: 10.1515/znc-2008-5-608 . PMID   18669020. S2CID   23956301.
  3. 1 2 Ayhan Ulubelen, Gülaçti Topçu, Hee-Byung Chai and John M. Pezzuto (1999). "Cytotoxic Activity of Diterpenoids Isolated from Salvia hypargeia". Pharmaceutical Biology . 37 (2): 148–151. doi:10.1076/phbi.37.2.148.6082.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  4. 1 2 3 Vivek K. Bajpai & Sun Chul Kan (2010). "Antibacterial abietane-type diterpenoid, taxodone from Metasequoia glyptostroboides Miki ex Hu". Journal of Biosciences . 35 (4): 533–538. doi:10.1007/s12038-010-0061-z. PMID   21289435. S2CID   25656295.
  5. Vivek K. Bajpai; Minkyun Na; Sun Chul Kang (2010). "The role of bioactive substances in controlling foodborne pathogens derived from Metasequoia glyptostroboides Miki ex Hu". Food and Chemical Toxicology . 48 (7): 1945–1949. doi:10.1016/j.fct.2010.04.041. PMID   20435080.
  6. 1 2 Tada M.; Kurabe J.; Yoshida T.; Ohkanda T.; Matsumoto Y. (2010). "Syntheses and antibacterial activities of diterpene catechol derivatives with abietane, totarane and podocarpane skeletons against methicillin-resistant Staphylococcus aureus and Propionibacterium acnes". Chem Pharm Bull . 58 (6): 818–824. doi: 10.1248/cpb.58.818 . PMID   20522992.
  7. 1 2 3 Ufuk Kolak; Ahmed Kabouche; Mehmet Öztürk; Zahia Kabouche; Gülaçtl Topçu; Ayhan Ulubelen (2009). "Antioxidant diterpenoids from the roots of Salvia barrelieri". Phytochemical Analysis . 20 (4): 320–327. Bibcode:2009PChAn..20..320K. doi:10.1002/pca.1130. PMID   19402189.
  8. 1 2 Norihisa Kusumoto, Tatsuya Ashitani, Tetsuya Murayama, Koichi Ogiyama and Koetsu Takahashi (2010). "Antifungal Abietane-Type Diterpenes from the Cones of Taxodium distichum Rich". Journal of Chemical Ecology . 36 (12): 1381–1386. doi:10.1007/s10886-010-9875-2. PMID   21072573. S2CID   11861719.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  9. 1 2 Norihisa Kusumoto, Tatsuya Ashitani, Yuichi Hayasaka, Tetsuya Murayama, Koichi Ogiyama and Koetsu Takahashi (2009). "Antitermitic Activities of Abietane-type Diterpenes from Taxodium distichum Cones". Journal of Chemical Ecology . 35 (6): 635–642. doi:10.1007/s10886-009-9646-0. PMID   19475449. S2CID   42622420.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  10. M. C. Ballesta-Acosta1, M. J. Pascual-Villalobos and B. Rodríguez (2008). "Short communication. The antifeedant activity of natural plant products towards the larvae of Spodoptera littoralis". Spanish Journal of Agricultural Research . 6 (1): 85–91. doi: 10.5424/sjar/2008061-304 .
  11. Kupchan, S. M.; Karim, A; Marcks, C. (1969). "Tumor inhibitors. XLVIII. Taxodione and taxodone, two novel diterpenoid quinone methide tumor inhibitors from Taxodium distichum". J. Org. Chem. 34 (12): 3912–3918. doi:10.1021/jo01264a036. PMID   5357534.
  12. Hanson, R. C.; Lardy, H. A.; Kupchan, S. M. (1970). "Inhibition of phosphofructokinase by quinone methide and alpha-methylene lactone tumor inhibitors". Science . 168 (3929): 378–380. Bibcode:1970Sci...168..378H. doi:10.1126/science.168.3929.378. PMID   4244949. S2CID   86616738.
  13. El-Lakany, Abdalla M. (2004). "Chlorosmaridione; A Novel Chlorinated Diterpene Quinone Methide from Rosemarinus officinalis L". Natural Product Sciences . 10 (2): 59–62.
  14. 1 2 Luis, J. G.; Grillo, T. A. (1993). "New diterpenes from Salvia munzii: chemical and biogenetic aspects". Tetrahedron . 49 (28): 6277–6284. doi:10.1016/S0040-4020(01)87965-5.
  15. Simoes, F.; Michavila, A.; Rodriguez, B.; Garcia Alvarez, M. C.; Mashooda, H. (1986). "A quinone methide diterpenoid from the root of Salvia moorciuftiana". Phytochemistry . 25 (3): 755–756. Bibcode:1986PChem..25..755S. doi:10.1016/0031-9422(86)88043-8.
  16. Gulacti Topcu1, Esra N. Altiner, Seyda Gozcu, Belkis Halfon, Zeynep Aydogmus, J. M. Pezzuto, Bing-Nan Zhou, David G. I. Kingston (2003). "Studies on Di- and Triterpenoids from Salvia staminea with Cytotoxic Activity". Planta Med. 69 (5): 464–467. doi:10.1055/s-2003-39705. PMID   12802732.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  17. Iván C. Guerrero; Lucía S. Andrés; Leticia G. León; Rubén P. Machín; José M. Padrón; Javier G. Luis & José Delgadillo (2006). "Abietane Diterpenoids from Salvia pachyphylla and S. clevelandii with Cytotoxic Activity against Human Cancer Cell Lines". J. Nat. Prod. 69 (12): 1803–1805. doi:10.1021/np060279i. PMID   17190465.
  18. M. Fraga; Carmen E. Díaz; Ana Guadaño & Azucena González-Coloma (2005). "Diterpenes from Salvia broussonetii Transformed Roots and Their Insecticidal Activity". J. Agric. Food Chem. 53 (13): 5200–5206. doi:10.1021/jf058045c. PMID   15969497.
  19. 1 2 Ayhan Ulubelen; Gülaçti Topcu (1996). "New Abietane Diterpenoids from Salvia montbretii". J. Nat. Prod. 55 (4): 441–444. doi:10.1021/np50082a006.
  20. 1 2 Ayhan Ulubelen; Gülaçti Topcu (1996). "Abietane and Rearranged Abietane Diterpenes from Salvia montbretii". J. Nat. Prod. 59 (8): 734–737. doi:10.1021/np9602224.
  21. Ikeshiro Y.; Mase I.; Tomita Y. (1991). "Abietane-Type Diterpene Quinones from Salvia nipponica". Planta Med. 57 (6): 588. doi:10.1055/s-2006-960219. PMID   17226213.
  22. Hsiu-Hui Chana, Tsong-Long Hwangb, Chung-Ren Sua, Mopur Vijaya Bhaskar Reddya and Tian-Shung Wu (2011). "Anti-inflammatory, anticholinesterase and antioxidative constituents from the roots and the leaves of Salvia nipponica Miq. var. formosana". Phytomedicine . 18 (2–3): 148–150. doi:10.1016/j.phymed.2010.06.017. PMID   21115331.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  23. A. Kabouche, Z. Kabouche, R. Touzani and C. Bruneau (2008). "Diterpenes and sterols from the roots of Salvia verbenaca subsp. clandestina". Chemistry of Natural Compounds . 44 (6): 824–825. doi:10.1007/s10600-009-9204-6. S2CID   28852886.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  24. Ik-Soo Lee; Norito Kaneda; Rutt Suttisri; Abdalla M. El-Lakany; Nawal Sabri & A. Douglas Kinghorn (1998). "New Orthoquinones from the Roots of Salvia lanigera". Planta Med. 64 (7): 632–634. doi:10.1055/s-2006-957536. PMID   17253304.
  25. Sabri, N. N., Abou-Donia, A. A., Assad, A. M., Ghazy, N. M., El-lakany, A. M., Tempesta, M. S. and Sanson D. R. (1989). "Abietane diterpene quinones from the roots of Salvia verbenaca and S. lanigera". Planta Medica . 55 (6): 582. doi:10.1055/s-2006-962111. PMID   17262492.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  26. Li M.; Zhang J. S.; Ye Y. M.; Fang J. N. (2000). "Constituents of the roots of Salvia prionitis". J Nat Prod . 63 (1): 139–141. doi:10.1021/np990357k. PMID   10650097.
  27. Y. Tezuka; R. Kasimu; J. X. Li; P. Basnet; K. Tanaka; T. Namba; S. Kadot (1998). "Constituents of Roots of Salvia deserta SCHANG. (Xinjiang-Danshen)". Chem Pharm Bull . 46 (1): 107–112. doi: 10.1248/cpb.46.107 .
  28. Benjamın Rodrıguez (2003). "A Methoxyabietane Diterpenoid from the Root of Salvia phlomoides and Structural Correction of Another Diterpene from Cryptomeria japonica". Z. Naturforsch. 58b: 324–327.
  29. J. A. Hueso-Rodrıguez; M. L. Jimeno; B. Rodrıguez; G. Savona; M. Bruno (1983). "Abietane diterpenoids from the root of Salvia phlomoides". Phytochemistry . 22 (9): 2005–2009. Bibcode:1983PChem..22.2005H. doi:10.1016/0031-9422(83)80033-8.
  30. Olga Batista, M. Fátima Simões, José Nascimento, Sofia Riberio, Aida Duartea, Benjamín Rodríguez and Maria C. de la Torreb (1996). "A rearranged abietane diterpenoid from Plectranthus hereroensis". Phytochemistry . 41 (2): 571–573. Bibcode:1996PChem..41..571B. doi:10.1016/0031-9422(95)00646-X. PMID   8821435.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  31. A. Otto; H. Walther; W. Püttmann (1997). "Sesqui- and diterpenoid biomarkers preserved in Taxodium-rich Oligocene oxbow lake clays, Weisselster basin, Germany". Organic Geochemistry . 26 (1–2): 105–115. Bibcode:1997OrGeo..26..105O. doi:10.1016/S0146-6380(96)00133-7.
  32. A. Otto; H. Walther; W. Püttmann (1994). "Molecular composition of a leaf- and root-bearing Oligocene Oxbow Lake Clay in the Weisselster Basin, Germany". Organic Geochemistry . 22 (2): 275–286. Bibcode:1994OrGeo..22..275O. doi:10.1016/0146-6380(94)90174-0.
  33. Angelika Otto; Bernd R. T. Simoneit; William C. Rember (2003). "Resin compounds from the seed cones of three fossil conifer species from the Miocene Clarkia flora, Emerald Creek, Idaho, USA, and from related extant species". Review of Palaeobotany and Palynology . 126 (3–4): 225–241. Bibcode:2003RPaPa.126..225O. doi:10.1016/S0034-6667(03)00088-5.
  34. Maya Stefanova; Bernd R.T. Simoneit (2008). "Polar aromatic biomarkers of Miocene-aged Chukurovo resinite and correlation with a progenitor macrofossil". International Journal of Coal Geology . 75 (3): 166–174. Bibcode:2008IJCG...75..166S. doi:10.1016/j.coal.2008.05.003.
  35. A. Zdravkov; A. Bechtel; R. F. Sachsenhofer; J. Kortenski; R. Gratzer (2011). "Vegetation differences and diagenetic changes between two Bulgarian lignite deposits – Insights from coal petrology and biomarker composition". Organic Geochemistry . 42 (3): 237–254. Bibcode:2011OrGeo..42..237Z. doi:10.1016/j.orggeochem.2010.12.006.
  36. Yann Hautevelle; Raymond Michels; Fabrice Malartre; Alain Trouiller (2006). "Vascular plant biomarkers as proxies for palaeoflora and palaeoclimatic changes at the Dogger/Malm transition of the Paris Basin (France)". Organic Geochemistry . 37 (5): 610–625. Bibcode:2006OrGeo..37..610H. doi:10.1016/j.orggeochem.2005.12.010.
  37. Kenneth E. Peters; Clifford C. Walters; J. Michael Moldowan (2005). "The Biomarker Guide: Biomarkers and isotopes in petroleum systems and Earth History". 2: 546.{{cite journal}}: Cite journal requires |journal= (help)
  38. Gonzalez, A. G.; Aguilar, Z. E.; Luis, J. G; Ravelo, A. G.; Dominguez, X. (1988). "Quinone methide diterpenoids from the roots of Salvia texana". Phytochemistry . 27 (6): 1777–1781. Bibcode:1988PChem..27.1777G. doi:10.1016/0031-9422(88)80442-4.
  39. D. Rutherford; M. Nielsen; N. Tokutomi; N. Akaike (1994). "Effects of plant diterpenes on the neuronal GABAA receptor-operated chloride current". NeuroReport . 5 (18): 2569–2572. doi:10.1097/00001756-199412000-00041. PMID   7696606.
  40. C. R. Tirapelli; S. R. Ambrosio; F. B. da Costa; A. M. de Oliveira (2008). "Diterpenes: a therapeutic promise for cardiovascular diseases". Recent Patents on Cardiovascular Drug Discovery. 3 (1): 1–8. doi:10.2174/157489008783331689. PMID   18221123.
  41. US Patent 5824665
  42. US Patent 6218435
  43. US Patent 20070203240
  44. World Patent 2004064736
  45. Karantsios, D.; Scarpa, J. S.; Eugster, C. H. (1966). "Struktur von Fuerstion". Helv. Chim. Acta . 49 (3): 1151–1172. doi:10.1002/hlca.19660490313.
  46. Gonzalez, A. G.; Fraga, B. M.; Gonzalez, C. M. (1983). "X-ray analysis of netzahualcoyone, a triterpene Quinone methide from orthosphenia mexicana". Tetrahedron . 84 (29): 3033–3036. doi:10.1016/s0040-4039(00)88088-0.
  47. Simoes, F.; Miehavila, A; Rodriguez. B.; Garcia A1varez, M. C.; Mashooda, H. (1986). "A quinone methide diterpenoid from the root of Salvia moorciuftiana". Phytochemistry . 25 (3): 755–756. Bibcode:1986PChem..25..755S. doi:10.1016/0031-9422(86)88043-8.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  48. Sankaram, A. V. B.; Murthi, M. M.; Bhaskaraiah, K.; Narasimha Rao, G. L.; Suhrahmanyam, M.; Shoolery, J. N. (1988). "Bharangin, a novel diterpenoid quinonemethide from pygmacopremna herbacea (Roxb.) moldenke". Tetrahedron Lett. 29 (2): 245–248. doi:10.1016/S0040-4039(00)80066-0.
  49. Fernando, H. C.; Gunatilaka, A. A. L.; Kumar, V.; Weexatunga, G. (1988). "Two new quinone-methides from cassine balae: Revised structure of balaenonol". Tetrahedron Lett. 29 (3): 387–390. doi:10.1016/S0040-4039(00)80104-5.
  50. Sankaram, A. V. B.; Bhaskaraiah, K.; Marthandamurthi. M.; Subrahmanyam, M. (1989). "Isobharangin, a new biogenetically significant diterpenoid quinonemethide from pygmacopremna herbacea (Roxb.) moldenke". Tetrahedron Lett. 30 (7): 867–868. doi:10.1016/S0040-4039(01)80638-9.
  51. Mori, K.; Matsui, M. (1970). "Diterpenoid total synthesis. XIII Taxodione, a quinone methide tumor inhibitor". Tetrahedron . 26 (14): 3467–3473. doi:10.1016/S0040-4020(01)92926-6. PMID   5449327.
  52. Matsumoto, T.; Tachibana, Y.; Uchida, J.; Fukui, K. (1971). "The Total Synthesis of (±)-Taxodione, A Tumor Inhibitor". Bull. Chem. Soc. Jpn. 44 (10): 2766–2770. doi: 10.1246/bcsj.44.2766 .
  53. Matsumoto, T.; Ohsuga, Y.; Fukui, K. (1974). "Synthesis of Taxodione and Methyl 11-hydroxy-12-methoxy-7-oxobieta-8,11,13-trien-18-oate". Chem. Lett. 3 (3): 297–300. doi:10.1246/cl.1974.297.
  54. Matsumoto, T.; Osbuga, Y.; Harada, S.; Fukui, K. (1977). "Synthesis of Taxodione, Royleanone, Cryptojaponol, and Methyl 11-Hydroxy-12-methoxy-7-oxoabieta-8,11)13-trien-18-oate". Bull. Chem. Soc. Jpn. 50: 266–272. doi: 10.1246/bcsj.50.266 .
  55. Matsumoto. T.; Usui, S.; Morimoto. T. (1977). "A Convenient Synthesis of (±)-Taxodione, (±)-Ferruginol, and (±)-Sugiol". Bull. Chem. Soc. Jpn. 50 (6): 1575–1579. doi: 10.1246/bcsj.50.1575 .
  56. Ohtsuka, Y.; Tahara, A. (1978). "Diterpenoids. XLVI. Syntheses of Taxodione, Royleanone and Their Analogues". Chem. Pharm. Bull. 26 (7): 2007–2013. doi: 10.1248/cpb.26.2007 .
  57. Snitman, D. L.; Himmelsbach, R. J.; Haltiwanger, R. C.; Watt, D. S. (1979). "A synthesis of (±)-cryptojaponol and (±)-taxodione". Tetrahedron Lett. 20 (27): 2477–2480. doi:10.1016/S0040-4039(01)86325-5.
  58. Johnson, W. S.; Shenvi, A. B.; Boots, S. G. (1982). "An approach to taxodione involving biomimetic polyene cyclization methodology". Tetrahedron . 38 (10): 1397–1404. doi:10.1016/0040-4020(82)80219-6.
  59. Stevens, R. V.; Bisaochi, G. G. (1982). "Benzocyclobutenones as synthons for the synthesis of C-11 oxygenated diterpenoids. Application to the total synthesis of (.+-.)-taxodione". J. Org. Chem. 47 (12): 2396–2399. doi:10.1021/jo00133a032.
  60. Poirier, D.; Jean, M.; Burnell, R. H. (1983). "Alternate Syntheses of Taxodione". Synth. Commun. 13 (3): 201–205. doi:10.1080/00397918308065989.
  61. Banerjee, A K; Carrasco, M. C. (1988). "Synthetic Approaches to (±)-Taxodione". Synth. Commun. 13 (4): 281–287. doi:10.1080/00397918308066977.
  62. Banerjee, A. K; Carrasco, M. C. (1986). "Total synthesis of (±)-12-methoxyabieta-8,11,13-trien-6-one, a versatile intermediate for diterpene syntheses". J. Chem. Soc., Perkin Trans. 1 : 25–31. doi:10.1039/p19860000025.
  63. Burnell, P. H.; Jean, M.; Poirer, D. (1987). "Synthesis of taxodione". Can. J. Chem. 65 (4): 775–781. doi:10.1139/v87-132.
  64. Engler, T. A.; Sampath. U.; N aganathan, S.; Van Der Velde, D.; Takusagawa, F. (1989). "A new general synthetic approach to diterpenes: Application to syntheses of (.+-.)-taxodione and (.+-.)-royleanone". J. Org. Chem. 54 (24): 5712–5727. doi:10.1021/jo00285a018.
  65. Haslinger, E.; Michl, G. (1988). "Synthesis of (+)-taxodione from (−)-abietic acid". Tetrahedron Lett. 29 (45): 5751–5754. doi:10.1016/S0040-4039(00)82181-4.
  66. Haslinger, E.; Michl, G. (1989). "missing". Justus Liebigs Ann. Chem. 1989 (7): 677–686. doi:10.1002/jlac.198919890212.
  67. Harring, S. R; Livinghouse, T. (1992). "A concise biomimetic total synthesis of (±)-taxodione via a BF3·MeNO2 promoted cationic cascade annulation". J. Chem. Soc., Chem. Commun. (6): 502–503. doi:10.1039/C39920000502.
  68. Ruedi, P.; Eugster, C. H. (1981). "14-Hydroxytaxodion: Partialsynthese und Reaktionen". Helv. Chim. Acta . 64 (7): 2219–2226. doi:10.1002/hlca.19810640728.
  69. Matsumoto, T.; Kawashima, H.; Iyo, K. (1982). "Synthesis of 3.BETA.-hydroxytaxodione and coleons S and T". Bull. Chem. Soc. Jpn. 55 (4): 1168–1173. doi: 10.1246/bcsj.55.1168 .
  70. Sanchez, Anthony J.; Konopelski, Joseph P. (1994). "The First Total Synthesis of (±)-Taxodone". Synlett . 1994 (5): 335–336. doi:10.1055/s-1994-22844.
  71. Sanchez, Anthony J.; Konopelski, Joseph P. (1994). "Phenol Benzylic Epoxide to Quinone Methide Electron Reorganization: The Synthesis of (±)-Taxodone". J. Org. Chem. 59 (18): 5445–5452. doi:10.1021/jo00097a057.
  72. Scott R. Harring & Tom Livinghouse (1994). "Polyene cascade cyclizations mediated by BF3·CH3NO2. an unusually efficient method for the direct, stereospecific synthesis of polycyclic intermediates via cationic initiation at non-functionalized 3° alkenes. An application to the total synthesis of (±)-taxodione". Tetrahedron . 50 (31): 9229–9254. doi:10.1016/S0040-4020(01)85502-2.
  73. E. Alvarez-Manzaneda, R. Chahboun, E. Cabrera, E. Alvarez, R. Alvarez-Manzaneda, M. Lachkar and I. Messouri (2007). "First synthesis of picealactone C. A new route toward taxodione-related terpenoids from abietic acid". Tetrahedron Lett. 48 (6): 989–992. doi:10.1016/j.tetlet.2006.12.009.{{cite journal}}: CS1 maint: multiple names: authors list (link)