Next Article in Journal
Synthesis and Characterization of Molecularly Imprinted Polymers for Phenoxyacetic Acids
Previous Article in Journal
A Study on Physical and Chemical Properties of Cellulose Paper Immersed in Various Solvent Mixtures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Germacrene D Cyclization: An Ab Initio Investigation

by
William N. Setzer
Department of Chemistry, University of Alabama in Huntsville, Huntsville, AL 35899, USA
Int. J. Mol. Sci. 2008, 9(1), 89-97; https://doi.org/10.3390/ijms9010089
Submission received: 17 December 2007 / Accepted: 23 January 2008 / Published: 25 January 2008
(This article belongs to the Section Physical Chemistry, Theoretical and Computational Chemistry)

Abstract

:
Essential oils that contain large concentrations of germacrene D are typically accompanied by cadinane sesquiterpenoids. The acid-catalyzed cyclization of germacrene D to give cadinane and selinane sesquiterpenes has been computationally investigated using both density functional (B3LYP/6-31G*) and post Hartree-Fock (MP2/6-31G* *) ab initio methods. The calculated energies are in general agreement with experimentally observed product distributions, both from acid-catalyzed cyclizations as well as distribution of the compounds in essential oils.

1. Introduction

It is generally observed that essential oils containing large concentrations of the sesquiterpene germacrene D are typically accompanied by cadinane and muurolane sesquiterpenoids [16] and germacrene D has been suggested to serve as biogenetic precursor to a number of different sesquiterpenoid skeletons [7,8]. Bülow and König have demonstrated that germacrene D readily undergoes acid-catalyzed cyclization to give cadinane, muurolane, and amorphane sesquiterpenes [8]. In addition, there is concern that skeletal rearrangements may occur during the hydrodistillation of plant materials to obtain essential oils [912]. In this work, an ab initio investigation using density functional theory (B3LYP/6-31G*) and molecular orbital theory (MP2/6-31G**) of the acid-catalyzed cyclization of germacrene D has been carried out in order to test the hypothesis that the relative ratios of cadinane and muurolane sesquiterpenoids found in essential oil compositions as well as experimental cyclization of germacrene D reflect the energetic differences of the sesquiterpenoids and their carbocation precursors.
The essential oil compositions of Beilschmiedia [1], Cedrela [2],Croton [3], Eugenia [4], Ocotea [5], and Piper [6] species generally show large concentrations of germacrene D accompanied by smaller concentrations of γ-muurolene, α-muurolene, γ-cadinene, δ-cadinene, cadina-1,4-diene, and α-cadinene. Overall, the ratios of these materials from the essential oils above are: germacrene D (81.5%), γ-muurolene (1.2%), α-muurolene (1.1%), γ-cadinene (3.2%), δ-cadinene (12.2%), cadina-1,4-diene (0.8%), and α-cadinene (0.4%).

2. Results and Discussion

The DFT (B3LYP/6-31G*) and post-HF (MP2/6-31G* *) relative enthalpies (H(0K)) and free energies (G°, from the DFT calculations, or G, based on HF 6-31G** calculated entropies at 25°C) for germacrene D and the sesquiterpene hydrocarbons are summarized in Table 1. The relative energies of the carbocation intermediates are listed in Table 2. The boat-chair-chair conformation of the germacrenyl carbocation (D1) that leads to the cadinenyl cation (G) is lower in energy than either of the conformations that lead to muurolenyl (D2 → H) or amorphenyl (D3 → J) carbocations by 0.65 and 1.25 kcal/mol, respectively. Cadinenyl carbocation G is slightly higher in energy than muurolenyl carbocation H (0.75 kcal/mol). The cadinenyl carbocation manifold leads, by loss of a proton, directly to α-cadinene, δ-cadinene, and γ-cadinene, with δ-cadinene the lowest energy, followed by α-cadinene (2.90 kcal/mol higher) and γ-cadinene (4.96 kcal/mol). Bülow and König [8] had found that acid-catalyzed cyclizations of germacrene D generally give a preponderance of δ-cadinene, followed by γ-cadinene, and lesser amounts of α-cadinene. The abundant δ-cadinene is consistent with the ab initio calculations, but the calculated energies of α-cadinene and γ-cadinene are not in agreement with the experimental results, and would predict α-cadinene to be more abundant than γ-cadinene. An analysis of a number of essential oils from many different families, e.g., Apiaceae [13], Asteraceae [14-19], Cistaceae [20], Clusiaceae [2125], Cupressaceae [2629], Euphorbiaceae [3], Heteropyxidaceae [30,31], Lamiaceae [3237], Lauraceae [1,5,3840], Meliaceae [2], Myrtaceae [4,41], Pinaceae [42], and Piperaceae [6], reveals that δ-cadinene is also the most abundant cadinane sesquiterpene found. In addition, γ-cadinene is more abundant in these essential oils than α-cadinene. α-Cadinene has been shown to undergo acid-catalyzed rearrangement to give β-cadinene [43], which in turn, has been found to isomerize to ω-cadinene [44-46], in agreement with the calculated energies; β-cadinene and ω-cadinene are lower in energy than α-cadinene by 1.88 kcal/mole and 2.90 kcal/mol, respectively. Bülow and König [8] reported that ω-cadinene can be formed from δ-cadinene in a 1.4:1 ratio, consistent with the nearly equal calculated energies.
The muurolenyl carbocation, H, leads directly to α-muurolene, γ-muurolene, and α-copaene. Of these, α-muurolene is lowest in energy, followed by γ-muurolene (4.21 kcal/mol higher) and α-copaene (14.13 kcal/mol higher). In the acid-catalyzed cyclizations, γ-muurolene generally predominates over α-muurolene, but the ratio depends on the acidic conditions used [8]. Notably, γ-muurolene also predominates over α-muurolene in essential oils, but we have found α-muurolene to generally predominate over γ-muurolene in the Lauraceae [1,5] while γ-muurolene predominates in the Piperaceae [6]. Of the essential oils examined in this study, δ-cadinene predominates (49% of the cadinane sesquiterpenes), followed by γ-muurolene and γ-cadinene (17% and 15%, respectively), then α-muurolene (9%), α-cadinene (3%) and cadina-1,4-diene (3%); very similar to the average percentages observed in the acid-catalyzed cyclizations of germacrene D (δ-cadinene, 42%; γ-muurolene, 17%; γ-cadinene, 19%; α-muurolene, 11%; α-cadinene, 8%; and cadina-1,4-diene, 3%). This would suggest that the relative distributions of cadinane sesquiterpenes observed, both from acid-catalyzed cyclization and present in essential oils represent equilibrium concentrations and depend on the energies of the respective compounds, and the only real inconsistency is the calculated energy of α-cadinene.
Amorphenes are generally present only in small amounts, if at all, in essential oils. Bülow and König [8], however, did find relatively high concentrations of δ-amorphene, α-amorphene, and γ-amorphene, in the acid-catalyzed cyclizations of germacrene D; up to 6% of each, depending on the conditions. The smaller amounts of amorphenes, compared to cadinenes and muurolenes reflects the relatively high energy of the amorphenyl carbocation compared to the cadinenyl and muurolenyl carbocations; about 6 kcal/mol higher. Reaction of germacrene D with strong Lewis acids (AlCl3 or ZnCl2) produces large amounts of epizonarene (as much as 80%) and zonarene (up to 25%). These two compounds are calculated to be the thermodynamically most stable of the cadinane sesquiterpenes in this investigation, with epizonarene slightly lower in energy (about 0.4 kcal/mol) than zonarene. Although these are the most stable of the cadinane sesquiterpenes, these materials are generally found only in small quantities in plant essential oils. Zonarene is, however, the major hydrocarbon component of the brown seaweed Dictyopteris zonarioides [47], and epizonarene is somewhat abundant in Cedrela fissilis leaf oil (2.5%) [48], the hexane extract of Tanacetum longifolium aerial parts (14.7%) [49], and the essential oil of Teucrium leucocladum (4.5%) [50].
Protonation of C(1) of germacrene D leads to the selinane sesquiterpenes by way of carbocation D4 with ring closure to give selinyl carbocations K and L. The calculated energies of these carbocation intermediates are much higher than those leading to the cadinane sesquiterpenes. δ-Selinene, however, is lower in energy than any of the cadinenes, muurolenes, or amorphenes, and this is consistent with the experimental observation; longer reaction times lead to increasing concentrations of δ-selinene at the expense of the cadinanes [8].

3. Computational Methods

All calculations were carried out using SPARTAN ′06 for Windows [51]. The hybrid B3LYP functional [52,53] and the 6–31G* basis set [54] were used for the optimization of all stationary points in the gas phase. Single-point Hartree-Fock ab initio energies were calculated using the DFT geometries (above) at the 6–31G* * [54] level, followed by a correlation energy calculation using the second-order Møller-Plesset model (MP2) [54]. Frequency calculations were used to characterize stationary points as minima. All enthalpies reported are zero-point (ZPE) corrected with unscaled frequencies, but with no thermal corrections; they are, therefore, H(0K). Entropies were calculated using the linear harmonic oscillator approximation.

References and Notes

  1. Setzer, WN; Haber, WA. Leaf essential oil composition of five species of Beilschmiedia from Monteverde, Costa Rica. Nat Prod Commun 2007, 2, 79–83. [Google Scholar]
  2. Eason, HM; Setzer, WN. Bark essential oil composition of Cedrela tonduzii C. DC. (Meliaceae) from Monteverde, Costa Rica. Rec Nat Prod 2007, 1, 24–27. [Google Scholar]
  3. Setzer, WN. Chemical compositions of the bark essential oils of Croton monteverdensis and Croton niveus from Monteverde, Costa Rica. Nat Prod Commun 2006, 1, 567–572. [Google Scholar]
  4. Cole, RA; Haber, WA; Setzer, WN. Chemical composition of essential oils of seven species of Eugenia from Monteverde, Costa Rica. Biochem Syst Ecol 2007, 35, 877–886. [Google Scholar]
  5. Takaku, S; Haber, WA; Setzer, WN. Leaf essential oil composition of 10 species of Ocotea (Lauraceae) from Monteverde, Costa Rica. Biochem Syst Ecol 2007, 35, 525–532. [Google Scholar]
  6. Setzer, WN; Park, G; Agius, BR; Stokes, SL; Walker, TM; Haber, WA. Chemical compositions and biological activities of leaf essential oils of twelve species of Piper from Monteverde, Costa Rica. Molecules 2008, 13. (submitted). [Google Scholar]
  7. Yoshihara, K; Ohta, Y; Sakai, T; Hirose, Y. Germacrene D, a key intermediate of cadinene group compounds and bourbonenes. Tetrahedron Lett 1969, 2263–2264. [Google Scholar]
  8. Bülow, N; König, WA. The role of germacrene D as a precursor in sesquiterpene biosynthesis: investigations of acid catalyzed, photochemically and thermally induced rearrangements. Phytochemistry 2000, 55, 141–168. [Google Scholar]
  9. Bartley, JP; Foley, P. Supercritical fluid extraction of Australian-grown ginger (Zingiber officinale). J Sci Food Agric 1994, 66, 365–371. [Google Scholar]
  10. Asfaw, N; Storesund, HJ; Skattebøl, L; Aasen, AJ. Coexistence of chrysanthenone, filifolone, and (Z)-isogeranic acid in hydrodistillates. Artefacts! Phytochemistry 2001, 58, 489–492. [Google Scholar]
  11. Babu, KGD; Kaul, VK. Variation in essential oil composition of rose-scented geranium (Pelargonium sp.) distilled by different distillation techniques. Flavour Fragr J 2005, 20, 222–231. [Google Scholar]
  12. Teixeira, S; Mendes, A; Alves, A; Santos, L. Simultaneous distillation-extraction of high-value volatile compounds from Cistus ladanifer L. Anal Chim Acta 2007, 584, 439–446. [Google Scholar]
  13. Palá-Paúl, J; Pérez-Alonso, MJ; Velasco-Negueruela, A; Vadaré, J; Villa, AM; Sanz, J; Brophy, JJ. Essential oil composition of the different parts of Eryngium bourgatii Gouan from Spain. J Chromatogr A 2005, 1074, 235–239. [Google Scholar]
  14. Roussis, V; Tsoukatou, M; Petrakis, PV; Chinou, I; Skoula, M; Harborne, JB. Volatile constituents of four Helichrysum species growing in Greece. Biochem Syst Ecol 2000, 28, 163–175. [Google Scholar]
  15. Juteau, F; Masotti, V; Bessière, JM; Viano, J. Compositional characteristics of the essential oil of Artemisia campestris var glutinosa. Biochem Syst Ecol 2002, 30, 1065–1070. [Google Scholar]
  16. Maia, JGS; Zoghbi, MdGB; Andrade, EHA; da Silva, MH; Luz, AIR; da Silva, JD. Essential oils composition of Eupatorium species growing wild in the Amazon. Biochem Syst Ecol 2002, 30, 1071–1077. [Google Scholar]
  17. Chericoni, S; Flamini, G; Campeol, E; Cioni, PL; Morelli, I. GC-MS analyses of the essential oil from the aerial parts of Artemisia verlotiorum: variability during the year. Biochem Syst Ecol 2004, 32, 423–429. [Google Scholar]
  18. Gauvin, A; Smadja, J. Essential oil composition of four Psiadia species from Reunion Island: A chemotaxonomic study. Biochem Syst Ecol 2005, 33, 705–714. [Google Scholar]
  19. Saroglou, V; Dorizas, N; Kypriotakis, Z; Skaltsa, H. Analysis of the essential oil composition of eight Anthemis species from Greece. J Chromatogr A 2006, 1104, 313–322. [Google Scholar]
  20. Demetzos, C; Angelopoulou, D; Perdetzoglou, D. A comparative study of the essential oils of Cistus salviifolius in several populations of Crete (Greece). Biochem Syst Ecol 2002, 30, 651–665. [Google Scholar]
  21. Nogueira, PCdL; Bittrich, V; Shepherd, GJ; Lopes, AV; Marsaioli, AJ. The ecological and taxonomic importance of flower volatiles of Clusia species (Guttiferae). Phytochemistry 2001, 56, 443–452. [Google Scholar]
  22. Schwob, I; Bessière, JM; Viano, J. Composition of the essential oils of Hypericum perforatum L. from southeastern France. C R Biol 2002, 325, 781–785. [Google Scholar]
  23. Schwob, I; Bessiere, JM; Masotti, V; Viano, J. Changes in essential oil composition in Saint John's wort (Hypericum perforatum L.) aerial parts during its phenological cycle. Biochem Syst Ecol 2004, 32, 735–745. [Google Scholar]
  24. Petrakis, PV; Couladis, M; Roussis, V. A method for detecting the biosystematic significance of the essential oil composition: The case of five Hellenic Hypericum L. species. Biochem Syst Ecol 2005, 33, 873–898. [Google Scholar]
  25. Saroglou, V; Marin, PD; Rancic, A; Veljic, M; Skaltsa, H. Composition and antimicrobial activity of the essential oil of six Hypericum species from Serbia. Biochem Syst Ecol 2007, 35, 146–152. [Google Scholar]
  26. Adams, RP. Systematics of the one seeded Juniperus of the eastern hemisphere based on leaf essential oils and random amplified polymorphic DNAs (RAPDs). Biochem Syst Ecol 2000, 28, 529–543. [Google Scholar]
  27. Salido, S; Altarejos, J; Nogueras, M; Sánchez, A; Pannecouque, C; Witvrouw, M; De Clercq, E. Chemical studies of essential oils of Juniperus oxycedrus ssp. badia. J Ethnopharmacol 2002, 81, 129–134. [Google Scholar]
  28. Cavaleiro, C; Salgueiro, LR; da Cunha, AP; Figueiredo, AC; Barroso, JG; Bighelli, A; Casanova, J. Composition and variability of the essential oils of the leaves and berries from Juniperus navicularis. Biochem Syst Ecol 2003, 31, 193–201. [Google Scholar]
  29. Vichi, S; Riu-Aumatell, M; Mora-Pons, M; Guadayol, JM; Buxaderas, S; López-Tamames, E. HS-SPME coupled to GC/MS for quality control of Juniperus communis L. berries used for gin aromatization. Food Chem 2007, 105, 1748–1754. [Google Scholar]
  30. Sibanda, S; Chigwada, G; Poole, M; Gwebu, ET; Noletto, JA; Schmidt, JM; Rea, AI; Setzer, WN. Composition and bioactivity of the leaf essential oil of Heteropyxis dehniae from Zimbabwe. J Ethnopharmacol 2004, 92, 107–111. [Google Scholar]
  31. Van Vuuren, SF; Viljoen, AM; Özek, T; Demirci, B; Başer, KHC. Seasonal and geographic variation of Heteropyxis natalensis essential oil and the effect thereof on the antimicrobial activity. S Afr J Bot 2007, 73, 441–448. [Google Scholar]
  32. Pereira, SI; Santos, PAG; Barroso, JG; Figueiredo, AC; Pedro, LG; Salgueiro, LR; Deans, SG; Scheffer, JJC. Chemical polymorphism of the essential oils from populations of Thymus caespititius grown on the island S. Jorge (Azores). Phytochemistry 2000, 55, 241–246. [Google Scholar]
  33. Skaltsa, HD; Mavrommati, A; Constantinidis, T. A chemotaxonomic investigation of volatile constituents in Stachys subsect. Swainsonianeae (Labiatae). Phytochemistry 2001, 57, 235–244. [Google Scholar]
  34. Veličković, DT; Randjelović, NV; Ristić, MS; Šmelcerović, AA; Veličkoviú, AS. Chemical composition and antimicrobial action of the ethanol extracts of Salvia pratensis L., Salvia glutinosa L. and Salvia aethiopis L. J Serb Chem Soc 2002, 67, 639–646. [Google Scholar]
  35. Couladis, M; Chinou, IB; Tzakou, O; Loukis, A. Composition and antimicrobial activity of the essential oil of Ballota pseudodictamnus L. Bentham. Phytother Res 2002, 16, 723–726. [Google Scholar]
  36. Ložiene, K; Vaièiuniene, J; Venskutonis, PR. Chemical composition of the essential oil of different varieties of thyme (Thymus pulegioides) growing wild in Lithuania. Biochem Syst Ecol 2003, 31, 249–259. [Google Scholar]
  37. Skaltsa, HD; Demetzos, C; Lazari, D; Sokovic, M. Essential oil analysis and antimicrobial activity of eight Stachys species from Greece. Phytochemistry 2003, 64, 743–752. [Google Scholar]
  38. Telascrea, M; de Araújo, CC; Marques, MOM; Facanali, R; de Moraes, PLR; Cavalheiro, AJ. Essential oil from the leaves of Cryptocarya mandioccana Meisner (Lauraceae): Composition and intraspecific chemical variability. Biochem Syst Ecol 2007, 35, 222–232. [Google Scholar]
  39. Setzer, WN; Stokes, SL; Penton, AF; Takaku, S; Haber, WA; Hansell, E; Caffrey, CR; McKerrow, JH. Cruzain inhibitory activity of leaf essential oils of Neotropical Lauraceae and essential oil components. Nat Prod Commun 2007, 2, 1203–1210. [Google Scholar]
  40. Agius, BR; Setzer, MC; Stokes, SL; Walker, TM; Haber, WA; Setzer, WN. Composition and bioactivity of essential oils of Lauraceae from Monteverde, Costa Rica. Int J Essent Oil Ther 2008, (in press). [Google Scholar]
  41. Cole, RA; Haber, WA; Lawton, RO; Setzer, WN. Leaf essential oil composition of three species of Myrcianthes from Monteverde, Costa Rica. Chem Biodiv 2008, (in press). [Google Scholar]
  42. Nicolić, B; Ristić, M; Bojoviæ, S; Marin, PD. Variability of the needle essential oils of Pinus heldreichii from different populations in Montenegro and Serbia. Chem Biodiv 2007, 4, 905–916. [Google Scholar]
  43. Fringuelli, F; Pizzo, F; Taticchi, A; Ferreira, VF; Michelotti, EL; Porter, B; Wenkert, E. Diels-Alder reactions of cycloalkenones. 4. Short syntheses of some cadinenes. J Org Chem 1985, 50, 890–891. [Google Scholar]
  44. Vlahov, R; Holub, M; Herout, V. On terpenes. CLXXXIV. Sesquiterpenic hydrocarbons from the essential oil of Mentha piperita of Bulgarian origin. Coll Czech Chem Comm 1967, 32, 822–829. [Google Scholar]
  45. Connell, DW; Hildebrand, RP; Sutherland, MD. Terpenoid chemistry XIV: the significance of the term “δ-cadinene”. Tetrahedron Lett 1968, 519–523. [Google Scholar]
  46. Nagasampagi, BA; Yankov, L; Dev, S. Sesquiterpenoids from the wood of Cedrela toona Roxb – partial synthesis of T-muurolol, T-cadinol and cubenol, structures of δ-cadinene and δ-cadinol. Tetrahedron Lett 1968, 1913–1918. [Google Scholar]
  47. Fenical, W; Sims, JJ; Wing, RM; Radlick, PC. Zonarene, a sesquiterpene from the brown seaweed Dictyopteris zonarioides. Phytochemistry 1972, 11, 1161–1163. [Google Scholar]
  48. Maia, BHLNS; de Paula, JR; Sant'Ana, J; da Silva, MFdGF; Fernandes, JB; Vieira, PC; Costa, MdSS; Ohashi, OS; Silva, JNM. Essential oils of Toona and Cedrela species (Meliaceae): Taxonomic and ecological implications. J Braz Chem Soc 2000, 11, 629–639. [Google Scholar]
  49. Mahmood, U; Kaul, VK; Singh, B. Sesquiterpene and long chain ester from Tanacetum longifolium. Phytochemistry 2002, 61, 913–917. [Google Scholar]
  50. El-Shazly, AM; Hussein, KT. Chemical analysis and biological activities of the essential oil of Teucrium leucocladum Boiss. (Lamiaceae). Biochem Syst Ecol 2004, 32, 665–674. [Google Scholar]
  51. SPARTAN ′06 for Windows; Wavefunction, Inc: Irvine, California, 2006.
  52. Becke, AD. Density-functional thermochemistry. III. The role of exact exchange. J Chem Phys 1993, 98, 5648–5652. [Google Scholar]
  53. Lee, C; Yang, W; Parr, RG. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys Rev B 1988, 37, 785–789. [Google Scholar]
  54. Hehre, WJ; Radom, L; Schleyer, PvR. Ab initio Molecular Orbital Theory; Wiley: New York, 1986. [Google Scholar]
Table 1. Ab Initio Enthalpies and Free Energies for Sesquiterpene Hydrocarbons.
Table 1. Ab Initio Enthalpies and Free Energies for Sesquiterpene Hydrocarbons.
CompoundRelative Energies (kcal/mol)CompoundRelative Energies (kcal/mol)
B3LYP/6-31G*MP2/6-31G**B3LYP/6-31G*MP2/6-31G**
H(0K)G°H(0K)GH(0K)G°H(0K)G
Ijms 09 00089 f133.8633.1841.9641.05 Ijms 09 00089 f22.091.843.743.51
Ijms 09 00089 f38.628.8610.2110.17 Ijms 09 00089 f42.492.574.133.96
Ijms 09 00089 f56.747.178.068.05 Ijms 09 00089 f66.326.287.867.78
Ijms 09 00089 f75.725.587.927.69 Ijms 09 00089 f810.5310.7811.0511.13
Ijms 09 00089 f910.6811.0011.7911.91 Ijms 09 00089 f1020.4521.4913.4214.33
Ijms 09 00089 f115.725.378.648.27 Ijms 09 00089 f1212.6712.4313.5613.39
Ijms 09 00089 f135.585.907.527.48 Ijms 09 00089 f148.048.089.449.34
Ijms 09 00089 f1520.0220.0320.6120.73 Ijms 09 00089 f162.793.342.042.32
Ijms 09 00089 f176.416.468.198.06 Ijms 09 00089 f188.849.467.137.45
Ijms 09 00089 f1920.5521.8213.1714.08 Ijms 09 00089 f2016.9117.2315.7115.98
Ijms 09 00089 f210000 Ijms 09 00089 f2214.6114.9013.3913.71
Table 2. Ab Initio Enthalpies and Free Energies for Sesquiterpenyl Carbocations.
Table 2. Ab Initio Enthalpies and Free Energies for Sesquiterpenyl Carbocations.
CarbocationRelative Energies (kcal/mol)CarbocationRelative Energies (kcal/mol)
B3LYP/6-31G*MP2/6-31G**B3LYP/6-31G*MP2/6-31G**
H(0K)G°H(0K)GH(0K)G°H(0K)G
Ijms 09 00089 f234.354.297.406.81 Ijms 09 00089 f241.681.401.551.51
Ijms 09 00089 f255.004.527.216.81 Ijms 09 00089 f260000
Ijms 09 00089 f275.605.777.757.19 Ijms 09 00089 f288.158.428.198.42
Ijms 09 00089 f2915.1915.4511.9411.58 Ijms 09 00089 f3026.3326.0822.3122.95
Ijms 09 00089 f312.431.921.941.92 Ijms 09 00089 f3223.0523.2318.9119.06

Share and Cite

MDPI and ACS Style

Setzer, W.N. Germacrene D Cyclization: An Ab Initio Investigation. Int. J. Mol. Sci. 2008, 9, 89-97. https://doi.org/10.3390/ijms9010089

AMA Style

Setzer WN. Germacrene D Cyclization: An Ab Initio Investigation. International Journal of Molecular Sciences. 2008; 9(1):89-97. https://doi.org/10.3390/ijms9010089

Chicago/Turabian Style

Setzer, William N. 2008. "Germacrene D Cyclization: An Ab Initio Investigation" International Journal of Molecular Sciences 9, no. 1: 89-97. https://doi.org/10.3390/ijms9010089

Article Metrics

Back to TopTop