Next Article in Journal
A Method of Hepatocyte Extraction Conjugated with HPLC is Established for Screening Potential Active Components in Chinese Medicines—Probing Herba Artemisiae Scopariae as an Exemplifying Approach
Next Article in Special Issue
Ergosterol-Induced Sesquiterpenoid Synthesis in Tobacco Cells
Previous Article in Journal
The Sesquiterpenes β-Caryophyllene and Caryophyllene Oxide Isolated from Senecio salignus Act as Phytogrowth and Photosynthesis Inhibitors
Previous Article in Special Issue
Four Eremophil-9-en-8-one Derivatives from Cremanthodium stenactinium Samples Collected in China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Communic Acids: Occurrence, Properties and Use as Chirons for the Synthesis of Bioactive Compounds

by
Alejandro F. Barrero
1,*,
M. Mar Herrador
1,*,
Pilar Arteaga
1,
Jesús F. Arteaga
2 and
Alejandro F. Arteaga
3
1
Departamento de Química Orgánica, Facultad de Ciencias, Instituto de Biotecnología, Universidad de Granada, Campus de Fuente Nueva, s/n. 18071 Granada, Spain
2
Departamento de Ingeniería Química, Química Fisíca y Químíca Orgánica, Facultad de Ciencias Experimentales, Universidad de Huelva, Campus el Carmen, s/n, 21071, Huelva, Spain
3
Departamento de Ingeniería Química, Facultad de Ciencias, Universidad de Granada, Campus de Fuente Nueva, s/n. 18071 Granada, Spain
*
Authors to whom correspondence should be addressed.
Molecules 2012, 17(2), 1448-1467; https://doi.org/10.3390/molecules17021448
Submission received: 25 November 2011 / Revised: 31 January 2012 / Accepted: 31 January 2012 / Published: 6 February 2012
(This article belongs to the Special Issue Terpenoids)

Abstract

:
Communic acids are diterpenes with labdane skeletons found in many plant species, mainly conifers, predominating in the genus Juniperus (fam. Cupresaceae). In this review we briefly describe their distribution and different biological activities (anti- bacterial, antitumoral, hypolipidemic, relaxing smooth muscle, etc.). This paper also includes a detailed explanation of their use as chiral building blocks for the synthesis of bioactive natural products. Among other uses, communic acids have proven useful as chirons for the synthesis of quassinoids (formal), abietane antioxidants, ambrox and other perfume fixatives, podolactone herbicides, etc., featuring shorter and more efficient processes.

Graphical Abstract

1. Introduction

Communic acids are a group of diterpenic natural products [1,2,3,4] with a labdane skeleton containing three double bonds and a carboxyl group at position 19 (Figure 1). Five communic acids have been described to date that differ in the location of the double bonds and the orientation of the carboxyl group: trans-communic acid (1) with the double bonds located in positions 8(17), 12 and 14, with Δ12 double bond E stereochemistry, and axial carboxyl group orientation, cis-communic acid (2) the Z isomer of the former, mirceocommunic acid (3), also named isocommunic acid, regioisomer of the former, where the Δ12 double bond moves to Δ13(16), 4-epi-trans-communic acid (4), a C4 epimer of 1 and ent-trans-communic acid (5) is the (−) enantiomer of 1. Of these, the most abundant in Nature is 1.
Figure 1. Structure of the communic acids.
Figure 1. Structure of the communic acids.
Molecules 17 01448 g001

2. Sources

Communic acids are widely distributed in Cupresaceae species, especially in the genus Juniperus. Although there are species that contain several of them, the most common case is the presence of only one. A tertiary mixture of 13 is found in Juniperus nana Willd. [5], J. communis [6] and J. oxycedrus [7]. A binary mixture of 12 is found in J. chinensis Linn [8,9,10], J. phoenicea [11], J. thurifera var. africana [11], J. foetidissima [12], J. sabina [13], Cryptomeria japonica [14], Platycladus orientalis [15], Sabina vulgaris [16], Podocarups imbricatus BI [17], Agathis vitiensis, A. macrophylla and A. lanceolata [18], Thuja occidentalis L. [19], and Hermas villosa [20] whereas a mixture of 34 is found in J. excelsa [21]. Trans-communic acid (1) was isolated from Entada abyssinica [22], Thujopsis dolabrata [23,24,25], Pinus luchuensis [26], Chamaecyparis obtusa Endl. [27,28,29], Thuja standishii [30,31], Araucaria angustifolia [32], Chamaecyparis formosensis [33], Porella navicularis [34], J. oxycedrus [35], J. drupaceae Labill [36], Sciadopitys verticillata [37], Fritillaria thunbergii [38], Cunninghamia unicanaliculata var. pyramidalis [39], Chromolaena collina [40], Cupressus sempervirens [41,42], J. communis [43], Chloranthus spicatus [44], Sabina vulgaris Antoine [45], Torreya jackii [46], Dacrydium pierrei [47], J. phoenicea [48], Calocedrus formosana [49], Fleischmannia multinervis [50], Cretan propolis [51], Libocedrus chevalieri [52], Pinus densiflora [53], and Mikania aff. jeffreyi [54], Chamaecyparis lawsoniana [55]. Cis-communic acid (2) was detected in Larix dahurica [56], Pseudotsuga wilsoniana [57], and Cladonia rangiferina L. Web. [58].
Myrceocommunic acid (3) was isolated from Juniperus oxycedrus [59]. Moreover the main component of diterpene acids in Cunninghamia lanceolata (Lamb.) Hook was 4-epi-trans-communic acid (5) [60]. Additionally polymers of 1 and of their derivatives have been found in resins of different Agathis species [61,62] and in sandarac resin [63].
Although these acids have been isolated from different parts of the plant (fruits, wood, bark, leaves, roots, etc.), they are mainly founded in leaves, fruits, and bark.

3. Biological Activity

The three communic acids 13 exhibited strong cytotoxic activity in a brine shrimp bioassay (LD50 0.16 μg/mL) [46]. Trans-communic acid (1) and cis-communic acid (2) and plant extracts containing them were also active against different microorganisms such as Staphylococcus aureus, both standard ATCC strain and clinical isolates [55,64,65,66,67,68,69], S. epidermidis ATCC 12228 [70], Aspergillus fumigatus and Candida albicans [62]. Moreover, both acids have shown cytotoxic activity against BSC-1 cells [71]. Other activities described for 1 are: antimycobacterial (Mycobacterium aurum, M. phlei, M. fortuitum and M. smegmatis) [72], antitumoral [35,73], relaxant [74], hypolipidemic [75], testosterone 5α-reductase inhibitory [76], anti-inflammatory and antioxidant [77].

4. Chemical Reactivity

Years ago, Pascual-Teresa et al. [78,79] described two studies based on the oxidation of the lateral chain of methyl esters of communic acids 1a3a. First, the functionalization of the side chain by selective epoxidation [78], and later the singlet oxygen addition [79] was studied. In both cases the relative reactivity of the three double bonds was determined for each compound. Epoxidation of 1a and 2a with m-chloroperbenzoic acid mainly afforded a mixture of 12,13-epopxy derivatives 69 together with a mixture of 14- and 15-m-chlorobenzoates 1011 (Figure 2). The epoxidation of 3a with mCPBA gave methyl (8R)-8,17-epoxy-8,17-dihydromirceocommunate (12, 5%) and methyl 13,16-epoxy-13,16-dihydromirceocommunate (13, 19%), recovering 24% yield of 3a. This result indicates greater reactivity at the trisubstituted double bond for 1a and 2a and terminal double bond on the side chain for 3a.
Figure 2. Structures of the products obtained by epoxidation and singlet oxygen oxidation of 1a3a.
Figure 2. Structures of the products obtained by epoxidation and singlet oxygen oxidation of 1a3a.
Molecules 17 01448 g002
The singlet oxygen addition to 1a led principally to the 12-hydroxyderivatives 12R (15, 19%) and 12S (16, 5%) together with minor proportions of the 12,15-dioxyderivative 14 (6%) and the tertiary alcohol 17 (4%), whereas in the case of 3a afforded the only the 15,16-dioxyderivative 18 (12%).
Compound 1a preferably underwent ene-reactions of the singlet oxygen on the trisubstituted double bond with syn stereospecificity, in accordance the with point established by Schulte-Elte [80]. Thus, the reaction produced mainly alcohols 1517 and a minor proportion of the 12,15-dioxyderivative 14, coming from a Diels-Alder reaction. In the case of methyl isocommunate 3a, which does not have trisubstituted double bond and where the monosubstituted dienic system adopts the cisoid conformation with relative ease, the reaction that takes place with singlet oxygen is the Diels-Alder cycloaddition, slowly yielding a small amount of 15,16-dioxyderivative 18 due to the tendency of 3a to polymerize.
Furthermore, another oxygenation procedure, i.e., the oxymercuration-demercuration (OM-DM) reaction of methyl esters of trans- and cis-communic acids (1a2a) was studied [81,82,83]. Treatment of 1a with mercuric acetate (1:2) in THF/H2O and the subsequent reduction of mercurials with NaBH4 afforded compounds 1922 (Scheme 1).
Scheme 1. OM-OD reaction for compound 1a. OD reaction with NaBH4.
Scheme 1. OM-OD reaction for compound 1a. OD reaction with NaBH4.
Molecules 17 01448 g003
Compound 19 is the product corresponding to the OM-DM at C14-C15 double bond. The formation mechanism of compounds 20, 21 is shown in Scheme 2. The formation of tetrahydrofuran derivatives 2021 from 1acan be explained by two routes, both converging at intermediate A and evolving to 20, 21via radical processes. In the first route, A results from the formation of mercurinium ion on the 14,15 double bond, followed by 1,4 addition of water at C12, and heterocyclization by attack of the hydroxy group on the other mercurinium ion formed on the 8,17 double bond. In a second possible route, A is obtained by the hydration of the 8,17 double bond on the β face, followed by attack of the hydroxy group on carbon C12 on the mercurinium ion of the monosubstituted double bond. Both routes converge at the organomercurial A, whose reduction with NaBH4 in basic medium leads to the formation of a bis-radical intermediate, that by direct cyclization between carbons C13 and C17 originates 20, and by reaction with atmospheric oxygen leads to 21.
Scheme 2. Mechanism of formation of compounds 20, 21.
Scheme 2. Mechanism of formation of compounds 20, 21.
Molecules 17 01448 g004
When the OM-OD reaction of compound 1a was carried out using Na(Hg) as the demercuriating agent (Scheme 3), the products obtained were 19, 2324 and there was no evidence of the formation of either pimarane 20 or endoperoxide 21. That is due to the fast reduction of the intermediate radicals coming from the corresponding type A organomercurials by sodium amalgam (Scheme 3).
Scheme 3. OM-OD reaction for compound 1a. OD reaction with Na(Hg).
Scheme 3. OM-OD reaction for compound 1a. OD reaction with Na(Hg).
Molecules 17 01448 g005
Another interesting reaction from the synthetic point of view is the oxidative degradation of the C12,C13 double bond of either cis-, trans-communic acids or their methyl esters. This transformation opens the possibility of using them in the preparation of bioactive molecules. In order to find appropriate experimental conditions for regioselective oxidative cleavage of the C12,C13 double bond in presence of the 8(17) and 14,15 ones, two methods of double bond cleavage were tried on 1a2a: Ozonolysis and OsO4/NaIO4 treatment [84,85]. First, ozonolysis of 1a was performed under different conditions, such as type of solvent (hexane, methanol, CH2Cl2), temperature (room temperature, 0 °C, −78 °C) and different ozone stream flows. Better selectivity towards the C12,C13 double bond degradation was observed when the reaction was carried out at −78 °C in CH2Cl2 yielding aldehyde-esters 25 and 26 (Scheme 4). The ozonolysis of isomer 2a under the same conditions also led to preferential attack on the C12,C13 double bond giving rise to the same products (Scheme 4).
Scheme 4. Ozonolysis of 1a2a.
Scheme 4. Ozonolysis of 1a2a.
Molecules 17 01448 g006
The outcome of the reaction of 1a2a with OsO4/NaIO4 is, however, strongly dependent on experimental conditions. Thus, when the temperature was kept at 0 °C to 10 °C, only 26 was detected, whereas mixtures of 25 and 26 were isolated when the temperature was 25 °C or higher (Scheme 5).
Scheme 5. Oxidation of 1a2a with OsO4/NaIO4.
Scheme 5. Oxidation of 1a2a with OsO4/NaIO4.
Molecules 17 01448 g007

5. Use of Communic Acids as Starting Materials for the Synthesis of Compounds of High Added Value

Communic acids 1–3 possess a labdane diterpene structure functionalised with a carboxylic group at C19, an exocyclic methylene at C8,C17 and a side chain dienic system appropriate for the preparation of a great variety of bioactive terpenoids, such as perfume fixatives [ambrox (30) and ambracetal (40)], antitumoral quassinoids [bruceantin (52)], antifungal podolactones [nagilactone F (63) and oidiolactone C (69)], and abietanes [19-hydroxyferruginol (76) a target for tolerance after transplant and in autoimmune diseases], and sugikurojin (80)] (Scheme 6).
Ambrox (30) and ambracetal (40) are perfume fixatives with a powerful amber-type aroma. Their syntheses were carried out alternatively from methyltrans-communate (1a) or methyl cis-communate (2a) or a mixture of the two [86,87]. Two different routes to ambrox from 1a/2a are showed in Scheme 7 and Scheme 8. The key steps of these syntheses are selective degradation of the side chains, stereoselective formation of the tetrahydrofuran ring and reduction of the axial methoxycarbonyl group. In the first synthesis the transformation of 1a and/or 2a to aldehyde 25 was done using two different methods: (a) carefully controlled ozonolysis of 1a and/or 2a at low temperature or (b) Δ14 selective hydrogenation with diimide, followed by a C12–C13 degradation of the resulting 14,15-hydrogenated derivative with OsO4/NaIO4. Oxidation of 25 with Jones reagent followed of cyclization with p-TsOH in toluene at reflux stereoselectively yielded the γ-lactone 27 with the most stable cis interannular linkage. Its reduction with LiAlH4 followed by kinetically controlled cyclization with p-TsOH/CH3NO2 at room temperature gave the tetrahydrofurane derivative 28 with the suitable trans stereochemistry. The conversion of the hindered methoxycarbonyl group into the methyl group was carried out in three steps by reduction of ester 28, oxidation of the resulting alcohol to aldehyde 33 and finally reduction under Huang-Minlon conditions led to the target 30 (Scheme 7).
Scheme 6. Compounds synthesized from communic acids 13.
Scheme 6. Compounds synthesized from communic acids 13.
Molecules 17 01448 g008
Scheme 7. Synthesis of ambrox 30.
Scheme 7. Synthesis of ambrox 30.
Molecules 17 01448 g009
In the second route hydroxyolefin 31, obtained by reductive ozonolysis from 1a/2a, was treated with p-TsOH in CH3NO2 at room temperature and subsequently with LiAlH4 to give the alcohol 33. Oxidation of 33 with Jones reagent led to the aldehyde 29 whose reduction under Huang Minlon conditions yielded ambrox (30). This route was improved and shortened by direct conversion of 1a/2a into diol 32 by reductive ozonolysis followed of cyclization with p-TsOH in CH3NO2 to yield the alcohol 33 (Scheme 8).
Scheme 8. Synthesis of ambrox.
Scheme 8. Synthesis of ambrox.
Molecules 17 01448 g010
Mixtures of 1–3 from Juniperus communis fruits are of great interest because they are byproducts of gin manufacturing. Scheme 8 and Scheme 9 show the syntheses of ambrox and ambracetal from a mixture of methyl esters of 13. The key intermediate in both processes is methyl ketone 34. This compound was obtained efficiently by a chemoselective reduction of the dienic system of a mixture of 1a3a with Na/t-BuOH at room temperature and subsequent oxidation with OsO4/NaIO4. The transformation of 34 to trihydroxy derivative 35 was carried out by stereoselective epoxidation with m-CPBA at room temperature followed by reduction with LiAlH4 in THF at reflux. Stereo-selective cyclization of 35 with p-TsOH/CH3NO2 at room temperature led to 36, which was transformed in ambrox 30 following the experimental procedure outlined in Scheme 7.
Scheme 9. Synthesis of ambrox.
Scheme 9. Synthesis of ambrox.
Molecules 17 01448 g011
For ambracetal (40) synthesis, treatment of methyl ketone 34 with a catalytic amount of OsO4 in a refluxing mixture of t-BuOH/pyridine/H2O and trimethylamine oxide as co-oxidant, afforded the tetracyclic ester 37 (Scheme 10). Conversion of the methoxycarbonyl group into the methyl group was carried out as shown in Scheme 6.
Scheme 10. Synthesis of ambracetal.
Scheme 10. Synthesis of ambracetal.
Molecules 17 01448 g012
An approach to compound 51, an intermediate in the synthesis of the antitumor agent bruceantin (52) has been developed from the communic acids 13 (Scheme 11 and Scheme 12) [88] via the methyl ketone 34.
Scheme 11. Synthesis of the tetracyclic intermediate 45.
Scheme 11. Synthesis of the tetracyclic intermediate 45.
Molecules 17 01448 g013
Scheme 12. Synthesis of the intermediate 51 (precursor of bruceantin 52).
Scheme 12. Synthesis of the intermediate 51 (precursor of bruceantin 52).
Molecules 17 01448 g014
Allylic oxidation of 34 at C7 with SeO2 at 60 °C and subsequent protection of the alcohol obtained with TBSCl yielded keto-ester 37 with high stereoselectivity. Subsequent condensation of the kinetic enolate of 41 with glyoxal dimethylacetal followed by mesylation and elimination with DBU led to the α,β-unsaturated ketone 42. Chemoselective reduction of 42 with Raney nickel and subsequent ozonolysis afforded diketone 43. At this point, an intramolecular aldol condensation gave the tricyclic ketone 44, whose hydrocyanation with potassium cyanide, diethylaluminium cyanide and 18-crown-6 ether led with high stereoselectivity to an epimer mixture of acetals (45ab) (6:1) (Scheme 11). Isomer 45a was used to complete the synthetic sequence (Scheme 11). Thus, reduction of 45a, first with DIBAL and then with NaBH4 afforded the diol 47, which was acetylated yielding 48. Exposure of 48 to thiophenol and boron trifluoride etherate in CH2Cl2 at room temperature yielded thioacetal 49. This compound was obtained as an epimeric mixture and the thioether groups were sequentially removed with mercury (II) chloride and mercury oxide in acetonitrile/methanol (1:1) at room temperature. Compound 51 was finally obtained as an epimer mixture after reductive desulfurization of 50 using nickel boride (Scheme 12).
Podolactones are nor-or bisnorditerpenic compounds isolated mainly from different plants of the genus Podocarpus (family Podocarpaceae) [89], and filamentous fungi (Oidodendrum truncatum [90], Aspergillus wentii [91], and Acrostalamus sp. [92]). These molecules present a wide range of biological activity, including antitumoral, insecticidal, antifeedant, allelopathic, and fungicidal activities, special attention being paid to their the antifungal activity. In this regard, LL-Z1271α (62) and oidolactone C (69) exhibited potent antifungal activities [93,94].
Considering their interesting properties, the podolactones nagilactone F (63) and LL-Z1271α (62) have been synthesized from a mixture of 1, 2 (Scheme 13 and Scheme 14) [95]. Now the key steps are a δ-lactonization in order to form the C ring, γ-lactonization and finally 14-hydroxylation.
Scheme 13. Synthesis of nagilactone F, LL-Z1271γ and LL-Z1271α.
Scheme 13. Synthesis of nagilactone F, LL-Z1271γ and LL-Z1271α.
Molecules 17 01448 g015
The synthesis begins with the degradation of the side chain of the acids 1,2 by a different procedure to those previously described. Thus, oxidation with m-CPBA of the starting material and subsequent treatment of the crude product with HIO4 led to the aldehyde 25 with good yield (73%). Compound 25 was better obtained by potassium permanganate oxidation and subsequent periodic degradation (80%). Oxidation of 25 to a carboxylic acid and esterification with CH2N2 followed by treatment with mercuric acetate (2.0 equiv.) in toluene at reflux gave the derivative 53 as an 8:1 mixture (Δ87). This mixture was reduced with NaBH4/DMF in the presence of an excess of bubbling O2, producing lactone 54 (75%), dienolide 55 (15%) and the starting product 56 (5%). This mixture was dehydrogenated with DDQ and PTSA to give an 8:3:1 mixture of 5759.
The methyl ester 57 was hydrolyzed almost quantitatively with concentrated sulphuric acid to obtain the acid 60. The treatment of 60 with lead tetraacetate under argon atmosphere and ten with SeO2 led to the δ-hydroxylactone 61 permitting firstly γ-lactone closure and subsequently allylic oxidation at C14. Then the antibiotic LL-Z1271α (62) was prepared by treatment of 61 with methanol acidified with a drop of sulphuric acid. Moreover, treatment of 61 with isopropylmagnesium bromide at 0 °C yielded 83% of condensation products, being the most of the α isomer (90%), nagilactone F (63).
Related with the above-mentioned podolactone syntheses, the first synthesis of the antifungal oidiolactone C (69) was carried out from trans-communic acid (1) (Scheme 14) [96,97]. The key step of the synthesis is a new bislactonization reaction catalyzed by Pd(II), giving rise to the podolactone-type tetracyclic skeleton from a norlabdadienedioic acid. This synthetic scheme was also used by the authors to improve podolactone LL-Z1271α synthesis.
Scheme 14. Synthesis of nagilactone F, LL-Z1271γ and LL-Z1271α.
Scheme 14. Synthesis of nagilactone F, LL-Z1271γ and LL-Z1271α.
Molecules 17 01448 g016
The selective ozonolysis of 1 and subsequent oxidation with Jones reagent, double esterification with diazomethane and allylic oxidation with SeO2/t-BuOOH yielded 36% of the hydroxydiester 64. Elimination of the trifluoroacetate of 64 with Pd(PPh3)4 led to diene 56, whose hydrolysis with sodium propanethiolate afforded diacid 65. Two different procedures were employed to carry out the double lactonization. First, the selective methylation of the carboxyl group at C12 with MeOH in the presence of 1,1′-carbonyldiimidazole and then iodolactonization under Barrett’s conditions after strict deoxygenation of the reaction medium furnished the iodo derivative 67 (80% yield) along with a 20% yield of dilactone 66. Iodo derivative 67 was exclusively converted in dilactone 66 by reaction with AgNO3/H2O/acetone (84% yield). Dilactone 66 was directly obtained from diacid 65 through a novel dilactonization process by treatment with substoichiometric Pd(II) (25%) and p-benzoquinone in a mixture of acetic acid and acetone as solvent (56%). The 9,11 double bond in diene-dilactone 68 was obtained, via the corresponding lithium enolate of 66 after adding phenylselenenyl chloride, and oxidation of the 11α-phenylseleno derivative to corresponding selenoxide by hydrogen peroxide with concomitant syn-elimination. Treatment of 68 with dimethyldioxirane afforded the natural oidiolactone C (69). Additionally, 62 was prepared by allylic oxidation as indicated in Scheme 15.
Scheme 15. Synthesis of oidiolactone C and LL-Z1271α.
Scheme 15. Synthesis of oidiolactone C and LL-Z1271α.
Molecules 17 01448 g017
Synthesis of the phenol abietane diterpenes 19-hydroxyferruginol (76), isolated from Podocarpus ferrugineus [98], and sugikurojin A (80), isolated from Cryptomeria japonica [99], from trans-communic acid (1) is shown in Scheme 16 and Scheme 17, respectively [100]. The key steps of these procedures are the side chain degradation and the elaboration of the aromatic C ring by Mn(III) cyclization.
Scheme 16. Synthesis of 19-hydroxyferruginol (76).
Scheme 16. Synthesis of 19-hydroxyferruginol (76).
Molecules 17 01448 g018
Epoxidation of ester 1a by mCPBA followed by treatment with HIO4 in THF led to aldehyde 25, whose treatment with MeMgBr and further oxidation with Jones reagent gave methylketone 71. Reaction of 71 with Me2CO3 and NaH in benzene afforded the β-ketoester 72. Treatment of 72 with Mn(OAc)3·2H2O (4.0 equiv.) and LiCl (3.0 equiv.) in Ac2O at 120 °C for 12 h led to the methyl O-acetyl salicylate 73 (74% yield). Transformation of 73 in abietane 74 was carried out by the addition of MeMgBr in excess. When this compound was treated with Et3SiH and CF3COOH was obtained silylether 75, whose treatment with LiAlH4 in THF at reflux afforded 19-hydroxyferruginol (76) (Scheme 16).
Heating of silylether 75 with Na2CrO4 and NaOAc in Ac2O-AcOH led to 7-oxoderivative 77. Compound 77 was refluxed with LiAlH4 in THF giving sugikurojin A (80). An alternative route to compound 80 from 75 involved the removal of the silyl group and further acetylation and oxidation to obtain ketone 79, which was then transformed into 80 (Scheme 17).
Scheme 17. Synthesis of sugikurojin (80).
Scheme 17. Synthesis of sugikurojin (80).
Molecules 17 01448 g019

6. Conclusions

This paper reveals the occurrence of the communic acids in fam. Cupresaceae especially in genus Juniperus. Furthermore they constitute appropriate building blocks for the efficient preparation of interesting bioactive natural products as ambrox, nagilactone F, bruceantin, 19-hydroxyferruginol and others.

References

  1. Dictionary of Terpenoids; Connolly, J.D.; Hill, R.A. (Eds.) Chapman & Hall: London, UK, 1991; Volume 2.
  2. Hanson, J.R. Diterpenoids of terrestrial origen. Nat. Prod. Rep. 2011, 28, 1755–1772. [Google Scholar] [CrossRef]
  3. Peters, R.J. Two rings in them all: The labdane-related diterpenoids. Nat. Prod. Rep. 2010, 27, 1521–1530. [Google Scholar] [CrossRef]
  4. Hanson, J.R. Diterpenoids. Nat. Prod. Rep. 2004, 21, 312–320. [Google Scholar] [CrossRef]
  5. Sakar, M.K.; Er, N.; Dilek, E.; Del Olmo, E.; San Feliciano, A. (–)-Desoxypodophyllotoxin and diterpenoids from Juniperus nana Willd. berries. Acta Pharm. Turcica 2002, 44, 213–219. [Google Scholar]
  6. De Pascual Teresa, J.; San Feliciano, A.; Barrero, A.F. Composition of Juniperus communis (common juniper) fruit. I. An. Quim. 1973, 69, 1065–1067. [Google Scholar]
  7. De Pascual Teresa, J.; San Feliciano, A.; Miguel del Corral, J.M. Components of Juniperus oxycedrus fruits. An. Quim. 1974, 70, 1015–1019. [Google Scholar]
  8. Fang, J.-M.; Chen, Y.-C.; Wang, B.-W.; Cheng, Y.-S. Terpenes from heartwood of Juniperus chinensis. Phytochemistry 1996, 41, 1361–1365. [Google Scholar] [CrossRef]
  9. Fang, J.M.; Sou, Y.C.; Chiu, Y.H.; Cheng, Y.S. Diterpenes from the bark of Juniperus chinensis. Phytochemistry 1993, 34, 1581–1584. [Google Scholar] [CrossRef]
  10. Chang, C.-I.; Chen, W.-C.; Shao, Y.-Y.; Yeh, G.-R.; Yang, N.-S.; Chiang, W.; Kuo, Y.-H. A new labdane-type diterpene from the bark of Juniperus chinensis Linn. Nat. Prod. Res. 2008, 22, 1158–1162. [Google Scholar] [CrossRef]
  11. Barrero, A.F.; Quilez del Moral, J.F.; Herrador, M.M.; Akssira, M.; Bennamara, A.; Akkad, S.; Aitigri, M. Oxygenated diterpenes and other constituents from Moroccan Juniperus phoenicea and Juniperus thurifera var. africana. Phytochemistry 2004, 65, 2507–2515. [Google Scholar] [CrossRef]
  12. Sakar, M.K.; San Feliciano, A. Diterpenoids of Juniperus foetidissima unripe berries. Fitoterapia 1994, 65, 304–306. [Google Scholar]
  13. Barrero, A.F.; Sanchez, J.F.; Altarejos, J. Resin acids in the woods of Juniperus sabina L. and Juniperus oxycedrus L. Ars Pharm. 1987, 28, 449–457. [Google Scholar]
  14. Su, W.-C.; Fang, J.-M.; Cheng, Y.-S. Diterpenoids from leaves of Cryptomeria japonica. Phytochemistry 1996, 41, 255–261. [Google Scholar]
  15. Wang, Y.-Z.; Tang, C.-P.; Ke, C.-Q.; Weiss, H.-C.; Gesing, E.-R.; Ye, Y. Diterpenoids from the pericarp of Plactycladus orientalis. Phytochemistry 2008, 69, 518–526. [Google Scholar] [CrossRef]
  16. Fang, S.; Gu, Y.; Yu, H.; Musadillin, S. The chemical nature of antitumor compounds from Sabina vulgaris Ant. Zhiwu Xuebao 1989, 31, 382–388. [Google Scholar]
  17. Gu, Y.; Xu, Y.; Fang, S.; He, Q. The chemical constituents from Podocarpus imbricatus. Zhiwu Xuebao 1990, 32, 631–636. [Google Scholar]
  18. Smith, R.M.; Marty, R.A.; Peters, C.F. The diterpene acids in the bled resins of three Pacific kauri, Agathis vitiensis, A. lanceolata, and A. macrophylla. Phytochemistry 1981, 20, 2205–2207. [Google Scholar] [CrossRef]
  19. Hafez, S.S. New labdane-type diterpenoids from the stem bark of Thuja occidentalis L. J. Pharm. Sci. 2004, 20, 34–47. [Google Scholar]
  20. Bohlmann, F.; Zdero, C. Naturally occurring terpene derivatives. XXXIII. Constituents of Hermas villosa. Chem. Ber. 1974, 107, 1416–1419. [Google Scholar] [CrossRef]
  21. Topcu, G.; Erenler, R.; Cakmak, O.; Johansson, C.B.; Celik, C.; Chai, H.-B.; Pezzuto, J.M. Diterpenes from the berries of Juniperus excelsa. Phytochemistry 1999, 50, 1195–1199. [Google Scholar] [CrossRef]
  22. Tchinda, A.T.; Fuendjiep, V.; Mekonnen, Y.; Ngo, B.B.; Dagne, E. A bioactive diterpene from Entada abyssinica. Nat. Prod. Commun. 2007, 2, 9–12. [Google Scholar]
  23. Takahashi, K.; Nagahama, S.; Nakashima, T.; Suenaga, H. Chemotaxonomy on the leaf constituents of Thujopsis dolabrata Sieb. et Zucc.-analysis of acidic extracts. Biochem. Syst. Ecol. 2003, 31, 723–738. [Google Scholar] [CrossRef]
  24. Hasegawa, S.; Hirose, Y. Terpenoids from the seed of Thujopsis dolabrata. Phytochemistry 1981, 20, 508–510. [Google Scholar] [CrossRef]
  25. Nagahama, S.; Tajima, M.; Nishimura, K. Chemotaxonomy of ate (Thujopsis dolabrata) of Noto, Ishikawa Prefecture. Mokuzai Gakkaishi 1996, 42, 698–702. [Google Scholar]
  26. Minami, T.; Wada, S.; Tokuda, H.; Tanabe, G.; Muraoka, O.; Tanaka, R. Potential antitumor-promoting diterpenes from the cones of Pinus luchuensis. J. Nat. Prod. 2002, 65, 1921–1923. [Google Scholar] [CrossRef]
  27. Fukushima, J.; Yatagai, M.; Ohira, T. Abietane-type and labdane-type diterpenoids from the cones of Chamaecyparis obtusa. J. Wood Sci. 2002, 48, 326–330. [Google Scholar] [CrossRef]
  28. Hanari, N.; Yamamoto, H.; Kuroda, K. Comparison of terpenes in extracts from the resin and the bark of the resinous stem canker of Chamaecyparis obtusa and Thujopsis dolabrata var. hondae. J. Wood Sci. 2002, 48, 56–63. [Google Scholar] [CrossRef]
  29. Yamamoto, H.; Asano, N.; Sawano, C.; Sone, T.; Gasha, T.; Ono, Y. Diterpenes Isolated from the Resin of the Resinous Stem Canker of Japanese Cypress, Chamaecyparis obtusa. Mokuzai Gakkaishi 1997, 43, 558–565. [Google Scholar]
  30. Minami, T.; Iwamoto, M.; Ohtsu, H.; Ohishi, H.; Tanaka, R.; Yoshitake, A. Aromatase inhibitory activities of standishinal and the diterpenoids from the bark of Thuja standishii. Planta Med. 2002, 68, 742–745. [Google Scholar] [CrossRef]
  31. Iwamoto, M.; Ohtsu, H.; Tokuda, H.; Nishino, H.; Matsunaga, S.; Tanaka, R. Anti-tumor promoting diterpenes from the stem bark of Thuja standishii (Cupressaceae). Bioorg. Med. Chem. 2001, 9, 1911–1921. [Google Scholar] [CrossRef]
  32. Fonseca, F.N.; Ferreira, A.J.S.; Sartorelli, P.; Lopes, N.P.; Floh, E.I.S.; Handro, W.; Kato, M.J. Phenylpropanoid derivatives and biflavones at different stages of differentiation and development of Araucaria angustifolia. Phytochemistry 2000, 55, 575–580. [Google Scholar]
  33. Lin, T.-C.; Fang, J.-M.; Cheng, Y.-S. Terpenes and lignans from leaves of Chamaecyparis formosensis. Phytochemistry 1999, 51, 793–801. [Google Scholar] [CrossRef]
  34. Bungert, M.; Gabler, J.; Adam, K.-P.; Zapp, J.; Becker, H. Pinguisane sesquiterpenes from the liverwort Porella navicularis. Phytochemistry 1998, 49, 1079–1083. [Google Scholar] [CrossRef]
  35. Liu, Y.; Nair, M.G. Labdane diterpenes in Curcuma mangga rhizomes inhibit lipid peroxidation, cyclooxygenase enzymes and human tumour cell proliferatio. Food Chem. 2011, 124, 527–532. [Google Scholar] [CrossRef]
  36. Sakar, M.K. Diterpenes from ripe “fruits” of Juniperus drupaceae Labill. Acta Pharm. Turcica 1987, 29, 65–68. [Google Scholar]
  37. Hasegawa, S.; Hirose, Y. Diterpenes from the seed of Sciadopitys verticillata. Phytochemistry 1985, 24, 2041–2046. [Google Scholar] [CrossRef]
  38. Kitajima, J.; Komori, T.; Kawasaki, T. Studies on the constituents of the crude drug “Fritillariae Bulbus.” III. On the diterpenoid constituents of fresh bulbs of Fritillaria thunbergii Miq. Chem. Pharm. Bull. 1982, 30, 3912–3921. [Google Scholar] [CrossRef]
  39. Ding, T.; Liu, H.; Pu, Q. Studies on the resin constituents of Taxodiaceae. I. Sesquiterpene and diterpene components of the resin from Cunninghamia unicanaliculata var. pyramidalis Yunnan Zhiwu Yanjiu 1982, 4, 307–311. [Google Scholar]
  40. Bohlmann, F.; Fiedler, L. Naturally occurring terpene derivatives. Part 131. A new communic acid derivative from Chromolaena collina (DC) K. et R. J. Indian Chem. Soc. 1978, 55, 1161–1162. [Google Scholar]
  41. Alvarez-Manzaneda Roldan, E.; Chahboun, R. Span. Procedure for isolation of trans-communic acid and derivatives from Cupressus sempervirens. Patent ES 2284341 A1 20071101, 2284. [Google Scholar]
  42. Rawat, P.; Khan, M.F.; Kumar, M.; Tamarkar, A.K.; Srivastava, A.K.; Arya, K.R.; Maurya, R. Constituents from fruits of Cupressus sempervirens. Fitoterapia 2010, 81, 162–166. [Google Scholar] [CrossRef]
  43. De Pascual Teresa, J.; Barrero, A.F.; Muriel, L.; San Feliciano, A.; Grande, M. New natural diterpene acids from Juniperus communis. Phytochemistry 1980, 19, 1153–1156. [Google Scholar]
  44. Xiao, Z.-Y.; Wang, X.-C.; Zhang, G.-P.; Huang, Z.-L.; Hu, L.-H. Terpenoids from roots of Chloranthus spicatus. Helv. Chim. Acta 2010, 93, 803–810. [Google Scholar] [CrossRef]
  45. Wang, W.; Ba, H.; Hajia, A.; Liao, L.; Duo, L. Study on chemical constituents of Sabina vulgaris Antoine. Tianran Chanwu Yanjiu Yu Kaifa 2005, 17, 588–591. [Google Scholar]
  46. Chen, R.; Zhang, Y.; Fang, S. Inhibitors of human DNA polymerase β isolated from Jack Rorreya (Torreya jackii). Zhongcaoyao 1997, 28, 707–710. [Google Scholar]
  47. Xu, Y.; Fang, S.; He, Q. The chemical constituents in Dacrydium pierrei. Zhiwu Xuebao 1991, 33, 646–648. [Google Scholar]
  48. Dawidar, A.M.; Ezmirly, S.T.; Abdel-Mogib, M. Sesquiterpenes and diterpenes from Juniperus phoenicea L. Pharmazie 1991, 46, 472–473. [Google Scholar]
  49. Lee, G.H.; Lin, C.C.; Cheng, Y.S.; Peng, S.M. Structure of methyl trans-communate. Acta Crystallogr. C Cryst. Str. 1987, C43, 1382–1384. [Google Scholar]
  50. Bohlmann, F.; Dhar, A.K.; Jakupovic, J.; King, R.M.; Robinson, H. A caryophyllene derivative from Fleischmannia pycnocephaloides. Phytochemistry 1981, 20, 1425–1426. [Google Scholar]
  51. Popova, M.P.; Chinou, I.B.; Marekov, I.N.; Bankova, V.S. Terpenes with antimicrobial activity from Cretan propolis. Phytochemistry 2009, 70, 1262–1271. [Google Scholar] [CrossRef]
  52. Zhang, Y.-J.; Litaudon, M.; Bousserouel, H.; Martin, M.-T.; Thoison, O.; Leonce, S.; Dumontet, V.; Sevenet, T.; Gueritte, F. Sesquiterpenoids and Cytotoxic Lignans from the Bark of Libocedrus chevalieri. J. Nat. Prod. 2007, 70, 1368–1370. [Google Scholar] [CrossRef]
  53. Sultan, M.Z.; Jeon, Y.-M.; Moon, S.-S. Labdane-type diterpenes active against acne from pine cones (Pinus densiflora). Planta Med. 2008, 74, 449–452. [Google Scholar] [CrossRef]
  54. Mendes, C.C.; Cruz, F.G.; Guedes, M.L.S.; Roque, N.F. Terpenes from Mikania aff. jeffreyi (Asteraceae). Zeitschrift Fuer Naturforschung B Chem. Sci. 2005, 60, 875–879. [Google Scholar]
  55. Smith, E.C.J.; Williamson, E.M.; Wareham, N.; Kaatz, G.W.; Gibbons, S. Antibacterials and modulators of bacterial resistance from the immature cones of Chamaecyparis lawsoniana. Phytochemistry 2007, 68, 210–217. [Google Scholar]
  56. Shmidt, E.N.; Pentegova, V.A. Chemical composition of Larix dahurica soft resin. Khimiya Prirodnykh Soedinenii 1974, 675–676. [Google Scholar]
  57. Hsieh, Y.-L.; Fang, J.-M.; Cheng, Y.-S. Terpenoids and flavonoids from Pseudotsuga wilsoniana. Phytochemistry 1998, 47, 845–850. [Google Scholar]
  58. Yoshikawa, K.; Kokudo, N.; Tanaka, M.; Nakano, T.; Shibata, H.; Aragaki, N.; Higuchi, T.; Hashimoto, T. Novel abietane diterpenoids and aromatic compounds from Cladonia rangiferina and their antimicrobial activity against antibiotics resistant bacteria. Chem. Pharm. Bull. 2008, 56, 89–92. [Google Scholar] [CrossRef]
  59. De Pascual Teresa, J.; San Feliciano, A.; Miguel del Corral, M.J. Components of Juniperus oxycedrus berries. I. An. Quim. 1972, 68, 1061–1062. [Google Scholar]
  60. Zhou, X.; Cheng, C.; Roshchin, V.I. Study on the chemical composition of substances extracted from the green crown of Cunninghamia lanceolata Hook. Linchan Huaxue Yu Gongye 1997, 17, 55–60. [Google Scholar]
  61. Carman, R.M.; Cowley, D.E. Diterpenoids. XII. Dundatholic acid. Aust. J. Chem. 1967, 20, 193–196. [Google Scholar] [CrossRef]
  62. Carman, R.M.; Cowley, D.E.; Marty, R.A. Diterpenoids. XXV. Dundathic acid and polycommunic acid. Aust. J. Chem. 1970, 23, 1655–1665. [Google Scholar] [CrossRef]
  63. Scalarone, D.; van der Horst, J.; Boon, J.J.; Chiantore, O. Direct-temperature mass spectrometric detection of volatile terpenoids and natural terpenoid polymers in fresh and artificially aged resins. J. Mass Spectrom. 2003, 38, 607–617. [Google Scholar] [CrossRef]
  64. Trusheva, B.; Popova, M.; Bankova, V.; Tsvetkova, I.; Naydenski, C.; Sabatini, A.G. A new type of European propolis, containing bioactive labdanes. Rivista Italiana EPPOS 2003, 36, 3–7. [Google Scholar]
  65. Hafez, S.S. New labdane-type diterpenoids from the stem bark of Thuja occidentalis L. J. Pharm. Sci. 2004, 20, 34–47. [Google Scholar]
  66. Muhammad, I.; Mossa, J.S.; Al-Yahya, M.A.; Ramadan, A.F.; El-Feraly, F.S. Further antibacterial diterpenes from the bark and leaves of Juniperus procera Hochst. ex Endl. Phytother. Res. 1995, 9, 584–588. [Google Scholar] [CrossRef]
  67. Samoylenko, V.; Dunbar, D.C.; Gafur, M.A.; Khan, S.I.; Ross, S.A.; Mossa, Jaber S.; El-Feraly, F.S.; Tekwani, B.L.; Bosselaers, J.; Muhammad, I. Antiparasitic, nematicidal and antifouling constituents from Juniperus berries. Phytother. Res. 2008, 22, 1570–1576. [Google Scholar] [CrossRef]
  68. Hirayama, T.; Someya, K.; Kuroyanagi, M.; Nakane, T. Natural antibacterial agents containing labdatrienoic acid and oral hygiene compositions containing them. Jpn. Kokai Tokkyo Koho 2009, JP 2009023920 A 20090205. [Google Scholar]
  69. Jeon, Y.M.; Moon, S.S.; Lee, T.H.; Cho, S.C.; Yim, B.K. Extract of pine cone having antibacterial and antifungal activity against particularly bacteria causing acne or athlete's foot, and cosmetic composition for improving skin disease without side effects on skin comprising the same or compound isolated from the same. Repub. Korean Kongkae Taeho Kongbo 2006, KR 2006104161 A 20061009. [Google Scholar]
  70. Xue, J.-J.; Fan, C.-Q.; Dong, L.; Yang, S.-P.; Yue, J.-M. Novel antibacterial diterpenoids from Larix chinensis beissn. Chem. Biodiver. 2004, 1, 1702–1707. [Google Scholar] [CrossRef]
  71. Perry, N.B.; Foster, L.M. Antitumor lignans and cytotoxic resin acids from a New Zealand gymnosperm, Libocedrus plumosa. Phytomedicine 1994, 1, 233–237. [Google Scholar] [CrossRef]
  72. Gordien, A.Y.; Gray, A.I.; Franzblau, S.G.; Seidel, V. Antimycobacterial terpenoids from Juniperus communis L. (Cuppressaceae). J. Ethnopharmacol. 2009, 126, 500–505. [Google Scholar] [CrossRef]
  73. Tanaka, R.; Ohtsu, H.; Iwamoto, M.; Minami, T.; Tokuda, H.; Nishino, H.; Matsunaga, S.; Yoshitake, A. Cancer chemopreventive agents, labdane diterpenoids from the stem bark of Thuja standishii (Gord.) Carr. Cancer Lett. 2000, 161, 165–170. [Google Scholar] [CrossRef]
  74. Ribeiro, L.A.A.; Tavares, J.F.; de Andrade, N.C.; da Silva, M.S.; da Silva, B.A. The (8)17,12E,14-labdatrien-18-oic acid (labdane 302), a labdane-type diterpene isolated from Xylopia langsdorffiana St. Hil. & Tul. (Annonaceae), relaxes the guinea pig trachea. Revista Brasileira de Farmacognosia 2007, 17, 197–203. [Google Scholar] [CrossRef]
  75. Kim, Y.-K.; Kim, W.-K. Hypolipidemic effects of Korean softwood components. Han'guk Sikp'um Yongyang Kwahak Hoechi 2001, 30, 1204–1214. [Google Scholar]
  76. Ko, J.H.; Hwang, E.I.; Lee, K.W.; Choi, K.H. Testosterone 5α-reductase inhibitor for preventing and treating diseases associated with testosterone 5α-reductase comprising extract of pine needle or compounds isolated therefrom. Repub. Korean Kongkae Taeho Kongbo 2006, KR 2006067255 A 20060619. [Google Scholar]
  77. Shimizu, M.; Tsuji, H.; Shogawa, H.; Fukumura, H.; Tanaami, S.; Hayashi, T.; Arisawa, M.; Morita, N. Anti-inflammatory constituents of topically applied crude drugs. II. Constituents and anti-inflammatory effect of Cryptomeria japonica D. Don. Chem. Pharm. Bull. 1988, 36, 3967–3973. [Google Scholar] [CrossRef]
  78. De Pascual Teresa, J.; San Feliciano, A.; Miguel del Corral, J.M.; Barrero, A.F. Transformaciones químicas en la cadena lateral de los ácidos comúnicos. 1. Epoxidaciones. Studia Chemica 1984, 9, 255–267. [Google Scholar]
  79. De Pascual Teresa, J.; Mateos, A.F.; Barrero, A.F.; Pollos, P. Adiciones de oxígeno singlete a mirceocomunato de metilo y trans-comunato de metilo. Síntesis de lambertianato de metilo. Stud. Chem. 1984, 9, 49–56. [Google Scholar]
  80. Schulte-Elte, K.H.; Muller, B.L.; Rautenstrauch, V. Preference for syn ene additions of 1O2 to trisubstituted, acyclic olefins. Helv. Chim. Acta 1978, 61, 2777–2783. [Google Scholar] [CrossRef]
  81. Barrero, A.F.; Sánchez, J.F.; Altarejos, J. Biomimetic Cyclization of Communic Acids to Pimarane Skeleton Via Organomercurial Intermediates. Tetrahedron Lett. 1988, 29, 3713–3716. [Google Scholar]
  82. Barrero, A.F.; Ramírez, A.; Salido, S.; Altarejos, J. Oxymercuriation-demedrcuriation reactions of methyl myrceocommunate. An. Quim. 1990, 86, 786–790. [Google Scholar]
  83. Barrero, A.F.; Sánchez, J.F.; Altarejos, J.; Perales, A. Oxymercuriation-demercuriation of the methyl esters of communic acids. X-ray molecular structures of methyl (8R,12R)-8,12-epoxyisopimar-15-en-19-oate. J. Chem. Soc. Perkin Trans. 1 1991, 2513–2523. [Google Scholar]
  84. Barrero, A.F.; Sánchez, J.F.; Altarejos, J. Selective ozonolysis of methyl trans-communate. Synthesis of drimanes. Tetrahedron Lett. 1989, 30, 5515–5518. [Google Scholar] [CrossRef]
  85. Barrero, A.F.; Alvarea-Manzaneda, E.J.; Altarejos, J.; Ramos, J.M.; Salido, S. Preferential oxidation reactions of the side chain of unsaturated labdanes. Bull. Soc. Chim. Fr. 1993, 130, 700–707. [Google Scholar]
  86. Barrero, A.F.; Altarejos, J.; Alvarea-Manzaneda, E.J.; Ramos, J.M.; Salido, S. Synthesis of ambrox from communic acids. Tetrahedron 1993, 49, 6251–6262. [Google Scholar]
  87. Barrero, A.F.; Altarejos, J.; Alvarea-Manzaneda, E.J.; Ramos, J.M.; Salido, S. Amber-type odorants from communic acids. Tetrahedron 1993, 49, 9525–9534. [Google Scholar] [CrossRef]
  88. Barrero, A.F.; Alvarez-Manzaneda, E.J.; Alvarez-Manzaneda, R.; Chahboun, R.; Meneses, R.; Cuerva, J.M.; Aparicio, M.; Romera, J.L. Approach to the synthesis of antitumor quassinoids from labdane diterpenes: An efficient synthesis of a picrasane-related intermediate. Org. Lett. 2001, 3, 647–650. [Google Scholar] [CrossRef]
  89. Dictionary of Terpenoids; Connolly, J.D.; Hill, R.A. (Eds.) Chapman-Hall: London, UK, 1991; Volume 2, pp. 853–857.
  90. Andersen, N.R.; Rasmunsen, P.R.; Falshaw, C.P.; King, T.J. The relative and absolute configuration of clerocidin and its cometabolites. Tetrahedron Lett. 1984, 25, 469–472. [Google Scholar] [CrossRef]
  91. Dorner, J.W.; Cole, R.J.; Springer, J.P.; Cox, R.H.; Cutler, H.; Wicklow, D.T. Isolation and identification of two new biologically active norditerpene dilactones from Aspergillus wentii. Phytochemistry 1980, 19, 1157–1161. [Google Scholar] [CrossRef]
  92. Ellestad, G.A.; Evans, R.H.; Kunstmann, M.P.; Lancaster, J.E.; Morton, G.O. Structure and chemistry of antibiotic LL-Z1271α, an antifungal carbon-17 terpene. J. Am. Chem. Soc. 1970, 92, 5483–5489. [Google Scholar] [CrossRef]
  93. Hosoe, T.; Nozawa, K.; Lumley, T.C.; Currah, R.S.; Fukushima, K.; Takizawa, K.; Miyaji, M.; Kawai, K. Tetranorditerpene lactones, potent antifungal antibiotics for human pathogenic yeasts, from a unique species of Oidiodendron. Chem. Pharm. Bull. 1999, 47, 1591–1597. [Google Scholar] [CrossRef]
  94. Ichikawa, K.; Ikunaka, M.; Kojima, N.; Nishida, H.; Yoshikawa, N. Terpenoid lactone compounds and their production process. 1999, EP 0 933 273 A1. [Google Scholar]
  95. Barrero, A.F.; Sánchez, J.F.; Elmerabet, J.; Jiménez-González, D.; Macías, F.A.; Simonet, A.M. Enantiospecific syntheses of the potent bioactives Nagilactone F and the mould metabolite LL-Z1271α. An evaluation of their allelopathic potential. Tetrahedron 1999, 55, 7289–7304. [Google Scholar] [CrossRef]
  96. Barrero, A.F.; Quilez del Moral, J.F.; Cuerva, J.M.; Cabrera, E.; Jimenez-González, D. preparation of bioactive podolactones via a new Pd-catalysed bislactonisation reaction. Synthesis of oidiolactone C. Tetrahedron Lett. 2000, 41, 5203–5206. [Google Scholar] [CrossRef]
  97. Barrero, A.F.; Arseniyadis, S.; Quilez del Moral, J.F.; Herrador, M.M.; Valdivia, M.; Jimenez, D. First synthesis of the antifungal oidiolactone C from trans-communic acid: Cytotoxic and antimicrobial activity in podolactone-related compounds. J. Org. Chem. 2002, 67, 2501–2508. [Google Scholar] [CrossRef]
  98. Cambie, R.C.; Cox, R.E.; Sidwell, D. Phenolic diterpenoids of Podocarpus ferrugineus and other podocarps. Phytochemistry 1984, 23, 333–336. [Google Scholar] [CrossRef]
  99. Arihara, S.; Umeyama, A.; Bando, S.; Imoto, S.; Ono, M.; Tani, M.; Yoshikawa, K. A new abietane and two dimeric abietane diterpenes from the black heartwood of Cryptomeria japonica. Chem. Pharm. Bull. 2004, 52, 354–358. [Google Scholar] [CrossRef]
  100. Alvarez-Manzaneda, E.; Chahboun, R.; Cabrera, E.; Alvarez, E.; Alvarez-Manzaneda, R.; Lachkar, M.; Messouri, I. Synthesis of phenol abietane diterpenes based on the oxidative radical cyclization utilizing the Mn(OAc)3/Ac2O system. Synlett 2007, 2425–2429. [Google Scholar]

Share and Cite

MDPI and ACS Style

Barrero, A.F.; Herrador, M.M.; Arteaga, P.; Arteaga, J.F.; Arteaga, A.F. Communic Acids: Occurrence, Properties and Use as Chirons for the Synthesis of Bioactive Compounds. Molecules 2012, 17, 1448-1467. https://doi.org/10.3390/molecules17021448

AMA Style

Barrero AF, Herrador MM, Arteaga P, Arteaga JF, Arteaga AF. Communic Acids: Occurrence, Properties and Use as Chirons for the Synthesis of Bioactive Compounds. Molecules. 2012; 17(2):1448-1467. https://doi.org/10.3390/molecules17021448

Chicago/Turabian Style

Barrero, Alejandro F., M. Mar Herrador, Pilar Arteaga, Jesús F. Arteaga, and Alejandro F. Arteaga. 2012. "Communic Acids: Occurrence, Properties and Use as Chirons for the Synthesis of Bioactive Compounds" Molecules 17, no. 2: 1448-1467. https://doi.org/10.3390/molecules17021448

Article Metrics

Back to TopTop