Next Article in Journal
Preparation of Chitosan from Brine Shrimp (Artemia urmiana) Cyst Shells and Effects of Different Chemical Processing Sequences on the Physicochemical and Functional Properties of the Product
Previous Article in Journal
Density Functional Theory Study of the Trans-Trans-Cis (TTC)→Trans-Trans-Trans (TTT) Isomerization of a Photochromic Spiropyran Merocyanine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Diterpenes from the Brown Alga Dictyota crenulata

by
Joel Campos De-Paula
1,2,
Ludmila Bomeny Bueno
3,
Diana Negrão Cavalcanti
4,
Yocie Yoneshigue-Valentin
5 and
Valéria Laneuville Teixeira
4,*
1
Departamento de Biologia Vegetal, Instituto de Biologia Roberto Alcântara Gomes, Universidade do Estado do Rio de Janeiro, Rio de Janeiro, RJ, 20550-013, Brazil
2
Programa de Pós-Graduação em Ciências Biológicas (Botânica), Museu Nacional, Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, 20940-040, Brazil
3
Programa de Pós-Graduação em Química Orgânica, Instituto de Química, Universidade Federal Fluminense, Niterói, RJ, Brazil
4
Departamento de Biologia Marinha, Instituto de Biologia, Universidade Federal Fluminense, CP 100. 644, Niterói, RJ, 24001-970, Brazil
5
Departamento de Botânica, Instituto de Biologia, Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, Brazil
*
Author to whom correspondence should be addressed.
Molecules 2008, 13(6), 1253-1262; https://doi.org/10.3390/molecules13061253
Submission received: 30 April 2008 / Revised: 2 June 2008 / Accepted: 3 June 2008 / Published: 4 June 2008

Abstract

:
The crude extract of the Brazilian brown alga Dictyota crenulata was analyzed by NMR spectroscopy and HRGC-MS techniques. Seven diterpenes were identified: pachydictyol A, dictyodial, 4β-hydroxydictyodial A, 4β-acetoxydictyodial A, isopachy-dictyol A, dictyol C and dictyotadiol. Xeniane diterpenes have previously been found in D. crenulata from the Pacific Ocean. The results characterize D. crenulata as a species that provides prenylated guaiane (group I) and xeniane diterpenes (group III), thus making it a new source of potential antiviral products.

Introduction

It has long been recognized that algae from the genus Dictyota Lamouroux present difficulties in the establishment of clear limits of separation between species. In previous studies, our group has indicated that the diterpenes from Dictyota species may have an important role as taxonomic markers [1,2,3,4,5,6,7]. In 2007, for example, a new species of the genus was established for the Brazilian coast based on morphological and chemical (diterpenes) characteristics [7].
Based on a revised biogenetic scheme proposed by Teixeira and Kelecom [1], the diterpenes have been distributed into three groups (I-III), depending on the structure of the products resulting from the the first formal cyclization of the geranyl-geraniol precursor. Thus, group I compounds result from a first cyclization of geranyl-geraniol between positions 1 and 10; group II compounds derive from a first cyclization of the precursor between positions 1 and 11, and group III compounds from either a first formal cyclization between positions 2 and 10 or by ring contraction of the prenylated germacrane.
The geographic distribution of D. crenulata is probably limited to the Atlantic and the western coast of Middle America [8]. De Clerck [8] noted that the basis on which Taylor separated D. jamaicensis from D. crenulata has never been clear, and several authors have questioned the status of D. jamaicensis [9,10,11,12]. Following the suggestion of these authors, D. jamaicensis is treated as a synonym of D. crenulata. D. crenulata is characterized by a multilayered medulla near the base of the thallus and spathulate apices.
Figure 1. D. crenulata from Aracruz (ES), Brazil: A) External morphology; B) Detail of the apex with parallel ramification; C) Triangular to crenulate dentation with concentrated sporangia located in the center of the thallus; D) Superficial view of a sporangial sori; E) Superficial view of an oogonial sorus; F) Surface view of antheridial sorus.
Figure 1. D. crenulata from Aracruz (ES), Brazil: A) External morphology; B) Detail of the apex with parallel ramification; C) Triangular to crenulate dentation with concentrated sporangia located in the center of the thallus; D) Superficial view of a sporangial sori; E) Superficial view of an oogonial sorus; F) Surface view of antheridial sorus.
Molecules 13 01253 g001
D. crenulata has proven to be a rich source of diterpenes. Previous investigations of this alga collected from Hawaii yielded pachydictyol A, dictyodial, 4β-hydroxydictyodial A, 18, O-dihydro-4β-hydroxydictyodial A, 18,O-dihydro-4β-hydroxydictyodial A 18-acetate, and β-crenulal [13,14,15], whereas specimens from the Gulf of California yielded pachydictyol A and acetoxycrenulide [16], and specimens from Easter Island gave five prenylated guaianes and dictyocrenulol [17,18] (compounds 1-16, Table 1).
Four unsaturated diterpene dialdehydes: dictyodial (10) and 4β-hydroxydictyodial (11), previously isolated from D. dichotoma [19], and (6R)-6-hydroxydichotoma-3,14-diene-1,17-dial (17) and its acetate derivative (6R)-6-acetoxydichotoma-3,14-diene-1,17-dial (18), isolated from D. menstrualis [20] showed inhibitory effects against HIV-RT (reverse transcriptase enzyme) activity. The discovery and characterization of new anti-HIV agents and new sources with either novel structures or mechanism(s) of action and low toxicity to the host remain a priority [21]. The dichotomane diterpenes 17-18 (Figure 2) inhibited virus replication with EC50 values of 40 µM (12 µg/mL) and 70 µM (24 µg/mL), respectively. Furthermore, 17 and 18 reduced HIV-1-RT activity with IC50 values of 10 µM and 35 µM, respectively. Considering that neither 17 nor 18 affected either cell viability or proliferation, the synthesis of proviral DNA is a specific target for the anti-HIV-1 activity of the diterpenes [20]. The antiviral activities described for 17 and 18 are as effective as those described for the previously isolated diterpenes [19], dictyodial (10) and 4β-hydroxydictyodial (11), that showed EC50 values of 4.3 and 9.2 µg/mL, respectively. Thus, the presence of α,β-unsaturated dialdehydes might be a determinant for the expression of the inhibitory activity of these substances.
Figure 2.
Figure 2.
Molecules 13 01253 g002

Results and Discussion

Table 1 presents the diterpene composition of the D. crenulata collections from Hawaii [13,14,15], the Gulf of California, Mexico [16], from Easter Island [17,18] and from the Município de Aracruz, Espirito Santo State, Brazil. All the D. crenulata samples produced diterpenes from groups I (prenylated guaianes) and III (xenianes and its derivatives).
Table 2 presents the similarity matrix between the populations of D. crenulata. The highest indices were obtained for samples that produced prenylated guaiane (group I) and xeniane diterpenes (group III): Gulf of California and Brazil (S = 0.55); Hawaii and Gulf of California (S = 0.44); and Hawaii and Brazil (S = 0.57).
Table 1. Diterpenes from various collections of D. crenulata Molecules 13 01253 i001
Table 1. Diterpenes from various collections of D. crenulata Molecules 13 01253 i001
A) Group I : HawaiiGulf of CaliforniaEaster IslandBrazil
[13,14,15][16][17,18][this study]
pachydictyol A (1)XXXX
isopachydictyol A (2) X
dictyotadiol(3) X
dictyol C (4) X
9-epidictyol B acetate (5) X
9-epidictyol B diacetate (6) X
dictyotriol A diacetate (7) X
dictyotriol A triacetate (8) X
dictyocrenulol (9) X
B) Group III:
dictyodial A (10)XX X
4β-hydroxydictyodial A (11)X X
4β-acetoxydictyodial A (12) X
18-O-dihydro-4β-hydroxydictyodial A (13)X
18- O-dihydro-4β-hydroxydictyodial A 18 acetate (14)X
β-crenulal or sanadaol (15)X
acetoxycrenulide (16) X X
The lowest similarity indices (S = 0.14, 0.17 and 0.22) were obtained when the population from Easter Island was compared with the remaining populations. From the calculated similarity indices, it may be stated that the population from Easter Island is different and does not belong to the same species as the remaining populations. Therefore this sample needs to be reinvestigated for taxonomic purposes.
Table 2. Similarity matrix between populations of D. crenulata.
Table 2. Similarity matrix between populations of D. crenulata.
1234
Hawaii11.00
Gulf of California20.441.00
Easter Island30.170.221.00
Brazil40.570.550.141.00
This is the first chemotaxonomic study of D. crenulata. With the exception of the Easter Island samples, the material collected from Pacific and Atlantic areas produced diterpenes from the chemical groups I and III. Specimens of this alga led to the isolation of six diterpenoids, five possessing a perhydroazulene skeleton.
4β-Acetoxydictyodial A (12) is the major product from Brazilian D. crenulata and acetylated 4β-hydroxydictyodial (11) from the Hawaiian population [15]. The results characterize D. crenulata as a species that furnishes prenylated guaiane (group I) and xeniane diterpenes (group III) and confirms the chemical similarity between Pacific and Atlantic populations and enlarges the geographical distribution of the xeniane diterpenes. The present data place D. crenulata chemically closer to others tropical western Atlantic species of Dictyota, such as D. menstrualis (Hoyt) Schnetter, Hörnig and Weber-Peukert and D. ciliolata Sonder ex Kützing. These two species differ chemically from D. crenulata by the majority production of pachydictyol A (1), dichotomane and cycloxeniane diterpenes [5] from D. menstrualis, and dictyol B acetate from D. ciliolata [23]. Since characterization of a species is the first step for its scientific use, the utilization of NMR spectroscopy and HRGC/MS techniques may be an excellent alternative for chemotaxonomic, seasonal variation, ecological and biogeographical studies.

Structure Elucidation

The acetone extract of D. crenulata was analyzed by NMR spectroscopy. The NMR spectra of the crude extract and fractions revealed the presence of pachydictyol A (1), isopachydictyol A (2), dictyotadiol (3), dictyol C (4). dictyodial (10), 4β-hydroxydictyodial A (11), and 4β-acetoxydictyodial A (12). The latter was identified as the principal component in the mixture, along with dictyodial (10) and 4β-hydroxydictyodial A (11). The identification of 4β-acetoxydictyodial A (12) was based on 1H- and 13C-NMR spectroscopic data comparison with literature values [22 and 15, respectively]. 4β-Hydroxydictyodial A (11) was identified in the same way [15]. We established that the OAc group is attached β to C-4, since the proton on C-4 (δ 5.30) shows small to nil couplings (0-4 Hz) to all of its vicinal neighbours. The presence of the alcohol 3 in the mixture reinforces the β-orientation position of the OAc group.
Finally, after several attempts of purification of the xeniane diterpenes, substances 1-3 were transformed into crenulide diterpene products. The photoisomerization of xeniane into the crenulatane skeleton was previously reported [24]. The extract from D. crenulata was analyzed by HRGC-MS. This analysis revealed the presence of peaks in the mass spectra compatible with the fragmentation patterns of xeniane, 10-12, and prenylated guaiane diterpenes 1-4. The mass spectra of the diterpenes are summarized in Table 3.
Table 3. Principal peaks observed in the mass spectra (m/z in parenthesis) of diterpenes from D. crenulata.
Table 3. Principal peaks observed in the mass spectra (m/z in parenthesis) of diterpenes from D. crenulata.
CompoundMolecular formulaRt *Molecular MassPrincipal peaks
Pachydictyol A (1)C20H32O12.88288288 (M+), 270 (M+-H2O), 255 (M+-H2O-CH3), 227, 188, 177, 175, 159 (M+-H2O-side chain), 121, 107, 91, 69 (C5H9, base peak), 55, 41.
Isopachydictyol A (2)C20H32O13.14288288 (M+), 270 (M+-H2O), 255 (M+-H2O-CH3), 227, 188, 173, 159 (M+-H2O-side chain, base peak), 145, 119, 107, 91, 69 (C5H9), 55, 41.
Dictyotadiol (3)C20H32O216.40304304 (M+), 286 (M+-H2O), 271 (M+-H2O-CH3), 243, 219, 201, 187, 175, 159 (M+-2H2O-side chain), 133, 121, 109, 95, 82 (base peak), 69 (C5H9), 55.
Dictyol C (4)C20H34O215.82306306 (M+), 288 (M+-H2O), 270 (M+-2H2O), 255 (M+-2H2O-CH3), 245, 203, 185, 177, 173, 159 (M+-2H2O-side chain), 145, 121, 119, 81, 69 (C5H9,base peak), 55, 41.
Dictyodial (10)C20H30O218.49302302 (M+), 284 (M+-H2O), 269 (M+-H2O-CH3), 187, 173, 109, 107, 93, 69 (C5H9, base peak), 55.
4β-Hydroxydictyodial A (11)C20H30O320.10318300 (M+ - H2O), 201, 189, 177, 176, 165, 161, 147, 135, 109, 95, 91, 82, 69 (C5H9, base peak), 67, 55.
4β-Acetoxydictyodial A (12)C22H32O420.98360300 (M+-AcOH), 282 (M+-AcOH-H2O), 201, 189, 177, 176, 161, 147, 135, 109, 95, 93, 91, 69 (C5H9, base peak), 55.
Acetoxycrenulide (16) 24.27 316, 298, 281, 253, 235, 217, 190, 109 (base peak), 91, 69 (C5H9), 55.
* Retention time in minutes

Experimental

General

All solvents were HPLC grade. Analytical thin-layer chromatography (TLC) separations were carried out on Merck silica gel 60 F-254 (0.2 mm) precoated aluminum plates. Once developed, plates were visualized by spraying with 2% ceric sulphate in sulfuric acid, followed by gentle heating. Silica gel 60 (Merck, 70-230 and 230-400 mesh) was used for column chromatography. NMR spectra were recorded in CDCl3 (100% Aldrich) on a Varian Unity Plus 300 spectrometer using TMS as internal standard. The preliminary gas chromatographic analysis (HRGC) was carried out on an HP 5890 CG machine equipped with an SE-54 glass capillary column (15 m x 0.25 mm; film thickness 0.25 µm) using a FID detector. The extract was analyzed by HRGC-MS on a HP 6890 series GC system, coupled to a HP 5973 mass selective detector in the electron impact mode (70 eV) equipped with an HP-1 MS capillary column (30 m x 0.25 mm; film thickness 0.25 µm). Injector and detector temperatures were set at 270 oC and 290 oC, respectively. The temperature program was kept at 160 oC, then programmed to 260 oC at a rate of 4o C/min and finally raised at a rate of 15 oC/min to 290 oC for 15 min. Hydrogen was the carrier gas at a flow rate of 1 mL/min. Diluted samples were injected manually in the split mode (1:10 or 1:20). Data were obtained from Frd area percent values. The chemical components were identified based on mass spectral comparison with those of standards and/or literature data, by co-injection in HRGC of these samples and from the Wiley 275 library data of the HRGC/MS system.

Sample Collection

Specimens of D. crenulata were collected at Praia dos Padres, Portocel, Município de Aracruz, Espirito Santo State (lat. 19o49’24”S, long. 40o03’00”W), Brazil during June 2003. The seaweeds were washed with local sea water, separated from sediments, epiphytes and other associated organisms. The algae were collected by Dra. Cristina Nassar (Universidade Federal do Rio de Janeiro) and identified by Dra. Cristina Nassar and Dr. Joel Campos de Paula. Voucher specimens are deposited at the herbarium of the Universidade do Estado do Rio de Janeiro (HRJ 10326).

Extraction and Isolation

Air-dried D. crenulata (100 g, dry weight) was extracted with acetone (100%) at room temperature. Evaporation of the solvent under reduced pressure yielded the crude extract (1.2 g) as a brownish residue. The partial crude extract (951 mg) was subjected to silica gel column chromatography (3 x 40 cm) eluting with n-hexane/acetone (9:1) to give a total of 37 fractions. The fractions yielded three compounds, the major being identified as 4β-acetoxydictyodial (12) and dictyodial (10) and 4β-hydroxydictyodial (11) as minor components.

Similarity Matrix

The degree of affinity between the algal populations of each area of collection was deduced from the similarity index (S) following Sørensen [25]. This index was calculated using the diterpene skeletons that are common to both considered populations, and using for each skeleton the lowest number of common diterpenes. This index was preferred over others since it attributes a higher weight to positive characters (here diterpene skeletons) than to negative ones.

Acknowledgements

We are grateful to Conselho Nacional de Pesquisa e Desenvolvimento Tecnológico (CNPq) for financial support and Productivity Fellowships to YYV and VLT.

References and Notes

  1. Teixeira, V.L.; Kelecom, A. A chemotaxonomic study of diterpenes from marine brown algae of the genus Dictyota. Sci. Total Environ. 1988, 75, 271–283. [Google Scholar] [CrossRef]
  2. Teixeira, V.L.; Almeida, S.A.S.; Kelecom, A. Chemosystematic and biogeographic studies of the diterpenes from the marine brown alga Dictyota dichotoma. Biochem. Syst. Ecol. 1990, 18, 87–92. [Google Scholar] [CrossRef]
  3. Teixeira, V.L.; Cavalcanti, D.N.; Pereira, R.C. Chemotaxonomic study of the diterpenes from the brown alga Dictyota menstrualis. Biochem. Syst. Ecol. 2001, 29, 313–316. [Google Scholar] [CrossRef]
  4. De-Paula, J.C.; Pedrini, A.G.; Pinheiro, M.D.; Pereira, R.C.; Teixeira, V.L. Chemical similarity between the brown algae Dictyota cervicornis and D. pardalis (Dictyotales, Phaeophyta). Biochem. Syst. Ecol. 2001, 29, 425–427. [Google Scholar] [CrossRef]
  5. Cavalcanti, D.N.; Rezende, C.M.; Pinto, A.C.; Teixeira, V.L. Diterpenoid constituents from the brown alga Dictyota menstrualis (Dictyotaceae, Phaeophyta). Nat. Prod. Comm. 2006, 1, 609–611. [Google Scholar]
  6. Freitas, O.S.P.; Oliveira, A.S.; De-Paula, J.C.; Pereira, R.C.; Cavalcanti, D.N.; Teixeira, V.L. Chemical variation in the diterpenes from the Brazilian brown alga Dictyota mertensii (Dictyotaceae, Phaeophyta). Nat. Prod. Comm. 2007, 2, 13–15. [Google Scholar]
  7. De-Paula, J.C.; Bueno, L.B.; Frugulhetti, I.C.P.P.; Yoneshigue-Valentin, Y.; Teixeira, V.L. Dictyota dolabellana sp. nov. (Dictyotaceae, Phaeophyceae) based on morphological and chemical data. Bot. Mar. 2007, 50, 288–293. [Google Scholar]
  8. De Clerck, O. The genus Dictyota in the Indian Ocean. Opera Botanica Belgica vol. 13; National Botanical Garden of Belgium: Meise, 2003. [Google Scholar]
  9. Dawson, E.Y. Additions to the marine flora of Costa Rica and Nicaragua. Pac. Natur. 1962, 3, 375–395. [Google Scholar]
  10. Nizamuddin, M.; Gerloff, J. News species and news combinations in the genus Dilophus J. Agardh. Nova Hedwigia 1979, 31, 865–879. [Google Scholar]
  11. Hörnig, I.; Schnetter, R.; Prud’homme van Reine, W.F. The genus Dictyota (Phaeophyceae) in the North Atlantic. II. Key to the species. Nova Hedwigia 1992, 54, 397–402. [Google Scholar]
  12. Wysor, B.; De Clerck, O. An update and annotated list of the marine macroalgae of the Caribbean coast of the Republic of Panama. II. Phaeophyceae. Bot. Mar. 2003, 46, 151–160. [Google Scholar]
  13. Finner, J.; Clardy, J.; Fenical, W.; Minale, L.; Riccio, R.; Battaile, J.; Kirkup, M.; Moore, R.P. Structure of dictyodial and dictyolactone, unusual marine diterpenoids. J. Org. Chem. 1979, 44, 2044–2047. [Google Scholar]
  14. Kirkup, M.P.; Moore, R.E. Identity of sanadaol with β-crenulal, a diterpene from the brown alga Dictyota crenulata. Phytochemistry 1983a, 22, 2527–2529. [Google Scholar] [CrossRef]
  15. Kirkup, M.P.; Moore, R.E. Two minor diterpenes realted to dictyodial A from the brown alga Dictyota crenulata. Phytochemistry 1983b, 22, 2539–2541. [Google Scholar] [CrossRef]
  16. Sun, H.H.; McEnroe, F.J.; Fenical, W. Acetoxycrenulide, a new bicyclic cyclopropane-containing diterpenoid from the brown seaweed Dictyota crenulata. J. Org. Chem. 1983, 48, 1903–1906. [Google Scholar] [CrossRef]
  17. Soto, H.; Rovirosa, J.; San Martin, A.; Argandoña, V. Metabolitos secundarios de Dictyota crenulata. Bol. Soc. Chil. Quím. 1994, 39, 173–178. [Google Scholar]
  18. Soto, H.; Rovirosa, J.; San-Martin, A. A new diterpene from Dictyota crenulata. Z. Naturforsch. 2003, 58b, 795–798. [Google Scholar]
  19. Ninomya, M.; Matsuka, S.; Kawakubo, A.; Bito, N. HIV-1 reverse transcriptase inhibitors containing hydroxydictyodial or dictyodial. Jpn. Kokai Tokkyo Koho JP 285,877 1995. [Google Scholar]
  20. Pereira, H.S.; Leão-Ferreira, L.R.; Moussatché, N.; Teixeira, V.L.; Cavalcanti, D.N.; Costa, L.J.; Diaz, R.; Frugulhetti, I.C.P.P. Antiviral activity of diterpenes isolated from the Brazilian marine alga Dictyota menstrualis against human immunodeficiency virus type 1 (HIV-1). Antiviral Res. 2004, 64, 69–76. [Google Scholar]
  21. Barreca, M.L.; Chimirri, A.; De Clercq, E.; De Luca, L.; Monforte, A.M.; Monforte, P.; Rao, A.; Zappala, M. Anti-HIV agents: design and discovery of new potent RT inhibitors. Farmaco 2003, 58, 259–263. [Google Scholar]
  22. Siamopoulou, P.; Bimplakis, A.; Iliopoulou, D.; Vagia, C.; Cos, P.; Berghe, D.V.; Roussis, V. Diterpenes from the brown algae Dictyota dichotoma and Dictyota linearis. Phytochemistry 2004, 65, 2025–2030. [Google Scholar] [CrossRef]
  23. Cronin, G.; Lindquist, N.; Hay, M.E.; Fenical, W. Effects of storage and extraction procedures on yields of lipophilic metabolites from the brown seaweeds Dictyota ciliolata and D. menstrualis. Mar. Ecol. Prog. Ser. 1995, 119, 265–273. [Google Scholar] [CrossRef]
  24. Guella, G.; Francesco, P. Photochemical conversion of xenicane into the crenulatane skeleton with diterpenoids of the brown seaweed Dictyota sp. from the coasts of Senegal. J. Chem. Soc. Chem. Commun. 1993, 1539. [Google Scholar] [CrossRef]
  25. Sørensen, T. A method of establishing groups of equal amplitude in plant sociology based on similarity of species and its application to analyses of the vegetation on Danish commons. Biol. Skrift./Kongel. Danske Videns. Selsk. 1948, 5, 1–34. [Google Scholar]
  • Sample Availability: Available from authors.

Share and Cite

MDPI and ACS Style

De-Paula, J.C.; Bueno, L.B.; Cavalcanti, D.N.; Yoneshigue-Valentin, Y.; Teixeira, V.L. Diterpenes from the Brown Alga Dictyota crenulata. Molecules 2008, 13, 1253-1262. https://doi.org/10.3390/molecules13061253

AMA Style

De-Paula JC, Bueno LB, Cavalcanti DN, Yoneshigue-Valentin Y, Teixeira VL. Diterpenes from the Brown Alga Dictyota crenulata. Molecules. 2008; 13(6):1253-1262. https://doi.org/10.3390/molecules13061253

Chicago/Turabian Style

De-Paula, Joel Campos, Ludmila Bomeny Bueno, Diana Negrão Cavalcanti, Yocie Yoneshigue-Valentin, and Valéria Laneuville Teixeira. 2008. "Diterpenes from the Brown Alga Dictyota crenulata" Molecules 13, no. 6: 1253-1262. https://doi.org/10.3390/molecules13061253

Article Metrics

Back to TopTop