Open access peer-reviewed chapter

Recent Developments in Selected Sesquiterpenes: Molecular Rearrangements, Biosynthesis, and Structural Relationship among Congeners

Written By

Shashikumar K. Paknikar and Kamlesh Pai Fondekar

Submitted: 17 October 2017 Reviewed: 06 February 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.74998

From the Edited Volume

Terpenes and Terpenoids

Edited by Shagufta Perveen and Areej Al-Taweel

Chapter metrics overview

1,565 Chapter Downloads

View Full Metrics

Abstract

Recent developments in selected sesquiterpenoids are reviewed for the past one decade (2005–2017) with special reference to Mechanisms of multistep molecular rearrangements of some sesquiterpenes or derivatives based on isotopic labeling studies and extensive spectroscopic analysis such as molecular rearrangement of acetyl cedrene to cedrene follower, acid catalyzed rearrangement of moreliane-based triketone, synthesis of (−)-isocomene and (−)-triquinane by acid-catalyzed rearrangement of (−)-modhephene, Total synthesis of (+)-cymbodiacetal, BF3 catalyzed molecular rearrangements of mono epoxides of α- and β-himachalenes, santonic acid: Zn-HCl-ether reduction. Insights into biosynthesis of albaflavenone, caryol-1(11)-ene-10-ol, (+)-koraiol, pogostol, patchouli alcohol and valerenadiene are discussed. Congeners for probing structure-biosynthetic relationship. This approach is discussed with the availability of very interesting results on the isolation of highly oxygenated secondary metabolites from endophytic fungi, Xylaria sp.

Keywords

  • molecular rearrangements
  • mechanisms
  • synthetic application
  • CCR
  • biosynthesis
  • labeling experiments
  • congeners

1. Introduction

Sesquiterpene carbon frameworks comprise the largest group of terpenoids or sometime referred as isoprenoids. Farnesyl diphosphate (FPP) having three olefinic linkages undergo cyclization to produce very large number major cyclic frameworks which are further modified by oxidative cleavages, molecular rearrangements, loss of carbon atoms. The aim of this chapter is to provide an overview of the recent developments in sesquiterpenes with particular reference to molecular rearrangements, biosynthesis and structural relationship among congeners. The coverage is not comprehensive but a focused review of the literature (2005–till September 2017) and only the relevant research articles having a link with the above areas are selected for discussion.

Advertisement

2. Mechanisms of multistep molecular rearrangements, insight into biosynthesis and congeners for probing structure-biosynthetic relationship of selected natural products

2.1. Molecular rearrangement of acetyl cedrene to cedrene follower

The acetylation of cedrene 1 can lead to various products depending on the reaction conditions. Paknikar et al. [1] undertook a detailed study on the acetylation of cedar wood oil (Virginia) with acetic anhydride and polyphosphoric acid in dichloromethane which leads, besides acetyl cedrene 2, also to a minor product, 1,7,7-trimethyl-2,3-(3′4’-dimethylbenzo)bicyclo[3.2.1]-octane 3, called the follower. Structural analysis of 3 (Scheme 1) shows that rings A, B, C of 2 are rearranged as B, A, C in follower 3.

Scheme 1.

Acetyl cedrene 2 and its follower 3. The numbering in the brackets is the one from acetyl cedrene.

Formation of 3 from 2 can only be explained by a multistep intramolecular rearrangement. This shows that: (i) ring C of 2 has undergone initial ring enlargement and subsequent ring contraction; (ii) cleavage of the C6–C7 bond of 2 and formation of the new C6-C2 bond; (iii) enlargement of ring A of 2 with concomitant loss of water. The mechanism for the formation of 3 from 2 when 1-13C labeled acetic anhydride was used is shown in Scheme 2.

Scheme 2.

Mechanism for the formation of follower 3 from acetyl cedrene 2.

One characteristic feature of the formation of the follower 3 is sluggish reaction rates. Density Functional Theory (DFT) calculation of B3LYP/6-31G* type using the Gaussian version 09 (Gaussian) revealed that the first neutral intermediate 4 (Scheme 2) is higher in energy than acetyl cedrene by ~20 kcal. A series of further cascade- like cationic rearrangements is involved with breaking and bond-forming intermediates.

The formation of the neutral intermediate 4 is supported by the observation that this process is the reverse pathway for the biosynthesis of α-cedrene from FPP, which has been established previously [2]. Few other feasible mechanisms for the formation of follower 3 could be devised, and only the one presented fits the observation of 13C enriched label at the C-3′ position of follower 3. Hence the key rearrangement is cyclopropylcarbinyl cation-cyclopropylcarbinyl cation rearrangement (CCR) [3, 4]. During the deuteriation of commercial acetyl cedrene, the follower was also deuterated, and it was observed that aromatic protons are exchanged. Interestingly, the product was only monodeuterated (Scheme 1) and the isotope was shared equally between the C-5′ and C-6′ positions of the follower 3. This equal distribution of one deuterium atom between C-5′ and C-6′ can be accounted for by the facile 1,2-hydride and 1,2-deuteride shifts and equilibration.

2.2. Acid catalyzed rearrangement of moreliane based triketone. Characterization of keto lactone, a 1-11 seco-moreliane

An interesting molecular rearrangement has been reported by Morales and co-workers [5]. They observed that triketone 5 on treatment with p-TSA in benzene resulted in the formation of a keto lactone 6, a 1–11 seco-moreliane derivative and also the first representative of this group (Figure 1).

Figure 1.

Skeletons of longipinane, moreliane and 1–11 seco-moreliane.

The rearrangement depicted in Scheme 3 involves initial cyclobutane ring expansion of the protonated triketone, generation of carbocationic intermediate 7 which rearranges via transition state in to protonated seco-moreliane 8. These steps are supported by DFT calculations.

Scheme 3.

Acid catalyzed rearrangement of triketone 5 to 1–11 seco-moreliane derivative 6.

2.3. Synthesis of (−)-isocomene and (−)-triquinane by acid catalyzed rearrangement of (−)-modhephene

Triquinanes have received considerable attention by their unique structure as well as their reported biological activities. (−)-Modhephene 9 of established absolute stereochemistry was subjected to acid catalyzed carbocation rearrangements which led to an interesting synthesis of (−)-isocomene 10 and (−)-triquinane 11[6]. This study was extended further by preparation of (−)-modhephene 9d stereospecifically at 14β geminal methyl group. Under same experimental conditions, deuterium labeled (−)-triquinane 11d a stereospecific 1,2-migration of 7/4β methyl group was observed (Scheme 4).

Scheme 4.

Molecular rearrangement of (−)-modhephene 9 to (−)-isocomene 10 and (−)-triquinane 11.

2.4. Total synthesis of (+)-cymbodiacetal

In 2010, Hayes and his co-workers reported [7] a total synthesis of (+)-Cymbodiacetal 12 by a biomimetic route proposed earlier [8, 9] using (R)-(+)-limonene 13, the key step involves hetero Diels-Alder cycloaddition which proceeds with an endo selectivity (2:1) in a quantitative yield. Exploitation of exo-isomer with m-CPBA followed by acid catalyzed opening afforded (+)-cymbodiacetal 12 (Scheme 5). The uncertainty in absolute stereochemistry was independently established by X-ray crystallography. These studies also clarified discrepancies in the previously published work [8, 9].

Scheme 5.

Total synthesis of (+)-cymbodiacetal 12.

2.5. BF3 catalyzed molecular rearrangements of mono epoxides of α- and β-himachalenes

Previous examples of acid catalyzed rearrangements of sesquiterpenes have shown that the opening of the epoxide triggers the reaction and directs the subsequent molecular rearrangements. In practically, among all the cases the aim is to valorize the naturally occurring sesquiterpene hydrocarbons.

Manoury and co-workers [10] observed that on treatment of α-himachalene monoepoxide 14 with BF3-Et2O in CH2Cl2 at room temperature afforded a tricyclic ketone 16 (71% isolated yield) product along with an unsaturated alcohol 17 (18%). The structure 16 was unambiguously assigned to ketone based on 1H, 13C, 1H-2D NMR experiments. The proposed mechanism (Scheme 6) involves ring opening of epoxide followed by participation of terminal methylene group to generate a tricyclic bridgehead carbocation 18 by ring contraction of seven membered ring to generate intermediate 19. A stereospecific 1,4-hydride transfer is proposed in the last step to the formation of 16.

Scheme 6.

Proposed mechanism for the formation of unsaturated alcohol 17 and tricyclic ketone 16.

Inspection of molecular models of intermediate 19 shows that the proposed stereospecific 1,4-hydride shift is unlikely and therefore a different process is responsible for the formation of ketone 16.

The structure assignment 17 to the minor product, a tricyclic unsaturated alcohol is based on spectral analysis and confirmed by single crystal X-ray data. The characteristic feature of 17 is the presence of a double bond involving a bridgehead carbon.

β-Himachalene monoepoxide 15 under identical experimental conditions gave two products major product (62%) and aryl-himachalene (10%). The major product was assigned structure 20. The proposed mechanism explains formation of 20 (Scheme 7). The gross structure of this compound an allo-himachalol, a natural product isolated from Cedrus deodara [11].

Scheme 7.

Mechanism for BF3 catalyzed transformation of β-himachalenes monoepoxide 15 to ketone 20.

Compounds 16, 17 and 20 are all optically active and since the absolute stereochemistry of himachalenes are known, it is observed that C7 α-H of α-himachalenes remains intact throughout the rearrangement. The absolute stereochemistry of 16, 17 and 20 is shown in Figure 2.

Figure 2.

Absolute stereochemistries of ketone 16 alcohol 17 and ketone 20.

2.6. Santonic acid: Zn-HCl-ether reduction

Santonic acid 21 (the diketocarboxylic acid obtained from santonin on digestion with aq. alkali) was subjected to reduction with the Zn-HCl-ether system [12] with an aim to obtain the previously prepared pinacol 22via intramolecular pinacolisation primarily because of conformational structure of santonic acid with close proximity of the 1,4-diketone system. Under these conditions santonic acid 21 did not afford the pinacol 22, but yielded a 60:40 mixture (GCMS, 1H NMR) of succinic anhydride derivatives 23 and 24. It is clear that the reaction proceeds via pinacol 22, which, under strong acidic conditions, undergoes further rearrangement to give anhydrides 23 and 24 (Scheme 8).

Scheme 8.

Mechanistic pathway for the conversion of santonic acid 21 to bicyclo[3.3.0] octanes 23 and 24.

2.7. Biosynthesis of albaflavenone

The tricyclic sesquiterpene antibiotic albaflavenone 25 isolated from the gram positive soil bacteria Streptomyces coelicolor A3 and Streptomyces albidoflavus is biosynthesized by enzymes encoded in a two-gene operm [13]. Initially, the sesquiterpene epi-isozizaene synthase catalyzes the cyclization of 2E, 6E-farnesyl diphosphate (FPP) to (+)-epi-isozizaene 26. A two-step allylic oxidation of 26 catalyzed by a single cytochrome P450170A1 (crP170A1) results in the formation of (+)-albaflavenone 25via an epimeric mixture of (5S)-albaflavenol 27 and (5R)-albaflavenol 28 intermediates (Scheme 9) [14].

Scheme 9.

Biosynthetic pathway of albaflavenone 25.

The mechanism and stereochemistry of FPP to epi-isozizaene 26via (3R)-nerolidyl diphosphate 29 has been conclusively established by labeling studies [15]. The entire biosynthetic process from FPP to epi-isozizaene is shown (Scheme 10). A two-step chemical synthesis of albaflavenone 25 from epi-isozizaene 26 was reported in this study.

Scheme 10.

Mechanism of the cyclization of E,E-FPP to epi-isozizaene 26via (3R)-nerodilyl diphosphate 29.

Ito and co-workers [16] reported a concise nine step total synthesis of albaflavenone without use of any protecting groups. Moreover, the absolute configuration of naturally occurring (+)-albaflavenone has been unambiguously established as 1S, 7S and 8R.

2.8. The biosynthesis of caryol-1(11)-ene-10-ol: on the mechanism of the formation of caryolene: a putative biosynthetic precursor to caryol-1(11)-ene-10-ol

In 2013, Nguyen and Tantillo [17] investigated the mechanism of the formation of caryolene 30, a putative biosynthetic precursor to caryol-1(11)-ene-10-ol 31 by DFT calculations (Figure 3).

Figure 3.

Structures of caryolene 30 and caryol-1(11)-en-10-ol 31.

Quantum chemical calculations indicated the mechanism involving a secondary carbocation intermediate 32 is not energetically viable. They proposed two mechanisms for caryolene 30 formation (pathway a and b). The pathway involves a base catalyzed deprotonation/reprotonation sequence and a tertiary carbocation minima (more likely) whereas pathway b involves intramolecular proton transfer and the generation of a secondary carbocation minima. Both mechanisms are predicted to involve concerted suprafacial/suprafacial [2 + 2] cycloaddition, whose asynchomicity allows them to avoid the constrains of orbital symmetry (Scheme 11).

Scheme 11.

Proposed mechanisms for the formation of 1,10-caryolene 30.

2.9. Biosynthesis of (+)-koraiol

As an outcome of Tantillo’s mechanism for caryolene 30 [17], biosynthetic pathway for koraiol 31 becomes evident (Scheme 12).

Scheme 12.

Biosynthesis of (+)-koraiol 31.

9-epi-E-Caryophyllene 32, caryophyllene 33 and (+)-koraiol 31 were identified by Dickschat and co-workers [18, 19] who carried out investigation on the volatiles of Fusarium fujikuroi by the use of CLSA-GCMS. The sesquiterpenoids were divided in to two groups based on their proposed biosynthetic pathways. Volatile sesquiterpenoids produced by sesquiterpene cyclase Ffsc4 were characterized as β-caryophyllene and an optically active alcohol (+)-koraiol 31. The structure 31 was assigned by extensive spectral analysis. The relative configuration of (+)-koraiol was elucidated by NOESY experiments. The cis fusion of rings A and B was deduced from the NOESY couplings of the bridge head hydrogen atoms 1H and 9H with each other with methyl protons 15-H and the pro-5-methylene protons 3-H. Interestingly, Khan et al. isolated (+)-koraiol, [α]D + 31.7° from the oleoresin of Korean pine (Pinus koraiensis Sieb.). The relative stereochemistry as shown in 31 has been established by X-ray analysis [20]. The absolute stereostructure of the rare sesquiterpene (+)-9-epi-E-caryophyllene, an enantiomer of 32 was isolated from Dacrydium cupressinum by Weavers and co-workers [21] (Figure 4).

Figure 4.

Structures of 9-epi-E-Caryophyllene 32, caryophyllene 33 and (+)-koraiol 31.

It is tempting to speculate (+)-koraiol 31 is biosynthesized from 9-epi-E-caryophyllene 32.

2.10. Biosynthesis of Pogostol

Biosynthesis of pogostol 34 by the endophytic fungus Geniculosporium was investigated by Dickschat and co-workers [22]. In this study, six 13C labeled isotopomers of mevalonolactone were synthesized and used in feeding experiments with the endophytic fungus Geniarlosperium. Feeding experiments with 35a and 35b gave insights into the stereochemical course of the terpene cyclization. The methyl group of the mevalonolactone that is labeled in these two isotopomers is converted into terminal (z)-methyl group of FPP (C-13). Both feeding experiments showed that the deprotonation step leading to germacrene A 36 proceeds with stereospecific deprotonation of C-13 and not C-12 of FPP (Figure 5).

Figure 5.

Biosynthesis of Pogostol 34 using isotopomers of mevalonolactone.

The volatile fraction was extracted by closed loop stripping apparatus followed by direct 13CNMR analysis (CLSA-NMR) newly developed by the same group. The biosynthesis of pogostol 34 proceeds through initial formation of germacrene-A 36. Protonation of 4,5 double bond initiates a second cyclization to cation which gets neutralized with water to give pogostol 34 (Scheme 13).

Scheme 13.

Mechanism of pogostol 34 formation from FPP.

In view of correlation of (−)-pogostol 37 with (+)-bulnesol 38 with known absolute stereochemistry, (−)-pogostol be represented by the stereostructure 37 [2325]. The stereostructure 34 thus represents (+)-pogostol (Figure 6).

Figure 6.

Absolute stereochemistry of (−)-pogostol 37—correlation of (−)-pogostol 37 and (+)-bulnesol 38.

2.11. Biosynthesis of patchouli alcohol (patchoulol)

The history of patchouli alcohol 39 from its isolation till date has narrated in a recent exhaustive review article [26]. Biosynthetic pathways were proposed based on experimental work for the conversion of FPP to patchouli alcohol 39 (Scheme 14).

Scheme 14.

Mechanism proposed for cyclization and rearrangement of FPP to patchoulol 39.

Croteau et al. [27] and Akhila et al. [28] proposed biosynthetic pathways for the conversion of FPP to patchouli alcohol 39 based on experimental work. Croteau et al. reported the 1,3-shift for conversion of 40 to 41 while Akhila et al. proposed two consecutive 1,2-hydride shifts for the same conversion (Scheme 15).

Scheme 15.

Biosynthetic pathways for the conversion of [2-2H1]-FPP to patchoulol isotopomer.

The recent isotopic labeling studies of Coates and colleagues [29] unrevealed the biosynthetic pathways for 39 which confirmed the 1,3-hydride shift across the five membered ring ruling out two consecutive 1,2-hydride shifts (Scheme 16).

Scheme 16.

Proposed biosynthesis of patchouliol 39 from deuterated FPP.

Incubation of isotopically pure [2-2H1] (E,E)-farnesyldisulfate with recombinant patchoulol synthase (rPTS) from Pogostemon cablin afforded a 65:35 mixture of monodeuterated and di-dueterated patchouliols and several hydrocarbons of which eight have been identified. This is confirmed by extensive NMR analysis on the labeled patchouliol mixture and comparison with those of unlabeled patchouliol. Deuterium label was located at position C5 (both isotopomers ca. 100%) and at C12 (minor isotopomer, 30–35%). The formation of [5,12-2H2] patchouliol is rationalized through an unknown (so far) hydrocarbon 42 which could incorporate deuterium at C12. This significant observation may have implication on the biosynthesis of nor-patchouliol 43 a congener of patchouliol, the biosynthesis is based on the earlier work [26] (Figure 7).

Figure 7.

Structures of nor-patchouliol 43, α-guaiene 44, α-bulnesene 45, (+)-guaiol 46 and (+)-bulnesol 38.

The interesting observation which can be made on the patchouli oil constituents that though α-guaine 44 and α-bulnesene 45 are genuine natural products [26], (+)-guaiol 46 and (+)-bulnesol 38 has never been reported to be present in patchouli oil.

2.12. Biosynthesis of Valerenadiene

Pyle et al. [30] reported the first enzymatic synthesis of valerena-4,7(11)-diene 47 (numbering used for valarenic acid) by a unique TPS from Valeriana officinalis. They identified two TPS’s VoTPS1 and VoTPS2. Transgenic yeast expressing VoTPS1 produced germacrene B 48, germacrene C 49 and germacrene D 50. On the other hand, VoTPS 2 produced valerena-4,7(11)-diene 47 as a major compound was substantiated by 13CNMR and GC–MS comparison with the synthetic standard. Minor products were identified as bicyclogermacrene 51 and alloaromadendrene 52. The proposed mechanism involves ring contraction of germacrane ring to a nine-membered intermediate having isobutenyl side chain. Cyclization gives valerena-4,7(11)-diene 47 (Scheme 17).

Scheme 17.

Biosynthesis pathway for valerena-4,7(11) diene 47 and other sesquiterpenes from VoTPS1 and VoTPS2.

Yeo et al. [31] proposed a mechanism wherein the isobutyl side chain is derived by the intermediacy of a caryophyllenyl carbocation 53. A 1,2-hydride shift followed by opening of the cyclobutyl ring. In this way the two methylene carbons of the isobutenyl side chain are predicted to arise from C1 and C11 of the originating FPP and therefore should become labeled when [1-13C] acetate is incorporated into FPP by mevalonate pathway operating in yeast (Scheme 18).

Scheme 18.

Three biosynthetic pathways for valerena-4,7(11) diene 47 and other sesquiterpenes from VoTPS1.

Valerina-1-10-diene 47 and related sesquiterpenes retain an isobutyl side chain whose origin has been recognized as enigmatic because a chemical rationalization for their biosynthesis has not been obvious. They identified seven Valeriana officinalis, terpene synthase genes (VoTPSs) and two were functionally characterized as sesquiterpene synthase VoTPS1 and VoTPS7. VoTPS7 encodes for a synthase that biosynthesizes germacrene C 49 (90%) whereas VoTPS 1 catalyzes conversion of E,E-FPP to valerena-1-10-diene 47. Overexpression of VoTPS produced valarena-1-10-diene 47 on the basis of one and two dimensional NMR analysis, further confirmed by comparison with published spectral data, GC retention time and EIMS fragmentation pattern. The most characteristic feature of the [1-13C] acetate is the FPP derived from the incorporation of [1-13C] acetate had labels located at C1, C3, C5, C7, C9 and C11 as expected using a yeast expression system, specific labeled [1-13C] acetate. FPP was catalytically cyclized (using VoTPS1) and produce valeriana-1,10-diene 47 whose 13C labels were found at C3, C5, C7, C9, C1 and C11. Of these C1 and C11 were adjacent carbons of the isobutyl side chain. The proposed mechanism involves an intermediate of a caryophyllenyl carbocation 53, 1,2-hydride shift followed by cleavage of C10-C11 bond generates a neutral monocyclic triene 54. The proposed scheme also indicates formation of other sesquiterpenes through intermediates tamariscenyl cation 55 and valerenyl cation 56.

Based on the experimental labeling data of Pyle et al. [30] and Yeo et al. [31], Paknikar et al. [4] proposed a new alternate biosynthetic route (Scheme 19) from IPP to valerenadiene 47 which fits the unusual 13C labeling found in valerian and avoids the previously unreported triene 54.

Scheme 19.

A cyclopropropane route to valerenadiene 47 (numbering based on FPP).

In Scheme 19, the 2-1-10-11 sequence of carbons in the first cyclic intermediate 57 from E,E-FPP becomes 2-10-1-11 in valerenadiene 47 which fits the 13C labeling pattern formed from [1-13C] acetate [4]. The biosynthetic pathway involves one neutral intermediate; bicyclogermacrene 36 found in valerian [32]. The key reaction is a cyclopropylcarbinyl cation-cyclopropylcarbinyl cation rearrangement (CCR) analogues to a key reaction in the biosynthesis of squalane from resqualene [3]. Structure interrelationships of the congeners of valerenadiene 47 including bicyclogermacrene 36, aromadendrene 51, germacrene C 49, germacrene D 50, α-gurjunene 58 and malliol 59 were considered in this alternate pathway.

Bicyclogermacrene 36 appears also to be an intermediate in the biosynthesis of related set of sesquiterpene with different stereochemistry found in Valeriana officinalis, including tamariscene 60, pacifigorgiol 61 and (+)-pacifigorgia-1,10-diene 62 (Scheme 20). In this scheme also the key reaction is again cyclopropylcarbinyl cation-cyclopropylcarbinyl cation rearrangement (CCR) with this time with a different stereoisomer.

Scheme 20.

Biosynthetic pathway of tamariscene 60, pacifigorgiol 61 and (+)-pacifigorgia-1,10-diene 62 from bicyclogermacrene 36.

Based on the results of three groups [4, 30, 31] a new consolidated mechanism for the biosynthesis of valerenadiene 47 from FPP via bicyclogermacrene 36 through alloaromadendryl cation 63 and CCR is presented which also explains formation of alloaromadendrene 64 (Scheme 21) replace alloaromadendryl cation with allo-aromadendryl cation.

Scheme 21.

Proposed new consolidated mechanism for the biosynthesis of valerenadiene 47.

2.13. Congeners of Xylaria sp.: structural interrelations

Endophytic fungi are reported to produce a number of bioactive metabolites and serve as an excellent source of highly oxygenated compounds which are likely to be potential drugs and also for the applications in crop science. The fungi belonging to genus Xylaria produces plethora of biologically related and structurally fascinating cadinenic and eudesmanic sesquiterpenes.

Liu and coworkers [33] reported isolation of highly oxygenated cadinane based compounds, three new xylaric acid A 65, xylaric acid B 66 and xylaric acid C 67 and nine known compounds xylaric acid D 68, heptelidic acid (avocetlin) [34] 69 hydroheptelidic acid 70, gliocladic acid 71, chlorheptelidic acid 72, trichoderonic acid A 73. The structure assignments are based on extensive spectral analysis. All these congeners belong to cadinane or seco-cadinane group of sesquiterpenes (Figure 8). The stereochemistry at C6 and C7 is unchanged for all the metabolites where C1 remains same for 66, 67, 69, 70, 72 and 73 and changes for 65, 68 and 71.

Figure 8.

Structural interrelations among the congeners of Xylaria sp. and the sequence of formation of isolated metabolites 6573.

Knowing the absolute stereochemistry of the congeners and their fungal origin, they belong to the “antipodal” set of compounds and they can be regarded as a result of extensive oxidative reactions of (−)-γ-cadinene 74. Recently, Rabe et al. [35] have reported isolation of several sesquiterpenes including (−)-γ-cadinene, [α]D-32.3° by incubation of FPP with six purified bacterial terpene cyclases. The results were further supported by labeling experiments with 13C labeled isotopomers of FPP. Interestingly, antipodal cadinenic sesquiterpenes with known absolute configurations have been isolated from Indian vetiver oil (Vetiveria zizanioides) [36]. Isolation of (−)-γ-cadinene 74, khusinol 75 and khusinol oxide 76 could be regarded as the precursors for the metabolites of Xylaria sp. A very clean sequence indicating a plausible order of formation of Xylaria sp. metabolites associated with the termite nest is presented (Scheme 22). We believe that this presentation will be useful while investigating the biosynthetic pathways using isotopic labeling studies.

Scheme 22.

Proposed plausible order of formation of Xylaria sp. metabolites from FPP.

Advertisement

3. Conclusions

This chapter gives overview of some of the interesting molecular rearrangements of sesquiterpenes reported over last decade. Further biosynthesis of albaflavenone, caryol-1(11)-ene-10-ol, (+)-koraiol, pogostol, patchouli alcohol and valerenadiene are also presented. The recent trends in the biosynthesis of natural products is focused on enzymatic synthesis using isotopic labeling, nevertheless discussions on structural interrelationships of various congeners provides insights in to natural occurrence of these molecules and finding their biosynthetic links.

Advertisement

Acknowledgments

We wish to dedicate this review to Professor R. B. Bates on his retirement from Research. We thank Dr. Asha D’Souza, Prof. Shailesh Shah and Rahul Chowgule for their valuable help in providing many research articles required for this review.

References

  1. 1. Paknikar SK, Kamounah FS, Hansen PE, Wadia MS. Multi-step rearrangement mechanism for acetyl cedrene to the hydrocarbon follower. Chemistry of Natural Compounds. 2017;53:1-4. DOI: 10.1007/s10600-017-2085-1
  2. 2. Nabeta K, Ara Y, Aoki Y, Miyake M. Biosynthesis of monoterpenes and sesquiterpenes in Larix leptolepis Callus from deuterated mevalonates. Journal of Natural Products. 1990;53:1241-1248. DOI: 10.1021/np50071a016
  3. 3. Blagg BSJ, Jarstfer DH, Rogers DH, Poulter CD. Recombinant squalene synthase. A mechanism for the rearrangement of presqualene diphosphate to squalene. Journal of the American Chemical Society. 2002;124:8846-8853. DOI: 10.1021/ja020411a
  4. 4. Paknikar SK, Kadam SH, Ehrlich AL, Bates RB. Alternate biosynthesis of valerenadiene and related sesquiterpenes. Natural Product Communications. 2013;8:1195-1196. PMID: 24273843
  5. 5. Morales PA, Amaro-Luis JM, Kutateladze AG. Structure determination and mechanism of formation of a seco-moreliane derivative supported by computational analysis. Journal of Natural Products. 2017;80:1210-1214. DOI: 10.1021/acs.jnatprod.7b00041
  6. 6. Joseph-Nathan P, Reyes-Trejo B, Morales-Rios MS. Molecular rearrangements of (−)-modhephene and (−)-isocomene to a (−)-triquinane. Journal of Organic Chemistry. 2006;71:4411-4417. DOI: 10.1021/jo060258t
  7. 7. Uroos M, Lewis W, Blake AJ, Hayes CJ. Total synthesis of (+)-cymbodiacetal: A re-evaluation of the biomimetic route. Journal of Organic Chemistry. 2010;75:8465-8470. DOI: 10.1021/jo101691n
  8. 8. Bottini AT, Dev V, Garfagnoli DJ, Hope H, Joshi P, Lohani H, Mathela CS, Nelson TE. Isolation and crystal structure of a novel dihemiacetal bis-monoterpenoid from Cymbopogon martini. Phytochemistry. 1987;26:2301-2302. DOI: org/10.1016/S0031-9422(00)84706-8
  9. 9. D’Souza AM, Paknikar SK, Dev V, Beauchamp PS, Kamat SP. Biogenetic-type synthesis of (+)-Cymbodiacetal, a constituent of Cymbopogon martini. Journal of Natural Products. 2004;67:700-702. DOI: 10.1021/np030338h
  10. 10. Heib AE, Benharref A, Parres-Manadie S, Manoury E, J-C Daran M, Urrutigoity M. Molecular rearrangement of epoxide derived from sesquiterpenes by Lewis acid catalysis. Tetrahedron: Asymmetry. 2010;21:1272-1277. DOI: org/10.1016/j.tetasy.2010.04.024
  11. 11. Bisarya SC, Dev S. Studies in sesquiterpenes—XXXIII: Himachalol. Tetrahedron. 1968;24:3869-3879. DOI: org/10.1016/S0040-4020(01)92593-1
  12. 12. Pai Fondekar KP, Malik BL, Paknikar SK. Santonic acid: Zn–HCl–ether reduction and ceric ammonium nitrate oxidation. Natural Product Research. 2014;28:1567-1573. DOI: org/10.1080/14786419.2014.927464
  13. 13. Zhao B, Lin X, Lei L, Lamb DC, Kelly SL, Waterman MR, Cane DE. Biosynthesis of the sesquiterpene antibiotic albaflavenone in Streptomyces coelicolor A3(2). Journal of Biological Chemistry. 2008;283:8183-8189. DOI: 10.1074/jbc.M710421200
  14. 14. Cane DE, Lin X. Biosynthesis of the sesquiterpene antibiotic albaflavenone in Streptomyces coelicolor. Mechanism and stereochemistry of the enzymatic formation of epi-isozizaene. Journal of the American Chemical Society. 2009;131:6332-6333. DOI: 10.1021/ja901313v
  15. 15. Moody SC, Zhao B, Lei L, Nelson DR, Mullins JGL, Waterman MR, Kelly SL, Lamb D. Investigating conservation of the albaflavenone biosynthetic pathway and CYP170 bifunctionality in streptomycetes. FEBS Journal. 2012;279:1640-1649. DOI: 10.1111/j.1742-4658.2011.08447.x
  16. 16. Kobayashi T, Kon Y, Abe H, Ito H. Concise total synthesis of albaflavenone utilizing sequential intramolecular aldol condensation: Determination of absolute configuration. Organic Letters. 2014;16:6397-6399. DOI: 10.1021/ol503202d
  17. 17. Nguyen QNN, Tantillo DJ. Caryolene-forming carbocation rearrangements. Beilstein Journal of Organic Chemistry. 2013;9:323-331. DOI: 10.3762/bjoc.9.37
  18. 18. Brock NL, Huss K, Tudzynski B, Dickschat JS. Genetic dissection of sesquiterpene biosynthesis by Fusarium fujikuroi. ChemBioChem. 2013;14:311-315. DOI: 10.1002/cbic.201200695
  19. 19. Citron CA, Brock NL, Tudzynsky B, Dickschat JS. Labelling studies on the biosynthesis of terpenes in Fusarium fujikuroi. Chemical Communications. 2014;50:5224-5226. DOI: org/10.1039/C3CC45982A
  20. 20. Khan VA, Gatilov YV, Dubovenko ZV, Pentegova VA. Crystal structure of koraiol—A sesquiterpene alcohol with a new type of carbon skeleton from the oleoresin of Pinus koraiensis. Chemistry of Natural Compounds. 1979;15:572-576. DOI: org/10.1007/BF00565927
  21. 21. Hinkley SFR, Perry NB, Weavers RT. Confirmation of structure and absolute stereochemistry of 9-epi-β-caryophyllene from Dacrydium cupressinum. Phytochemistry. 1994;35:1489-1494. DOI: org/10.1016/S0031-9422(00)86882-X
  22. 22. Barra L, Schulz B, Dickschat JS. Pogostol biosynthesis by the endophytic fungus Geniculosporium. ChemBioChem Communications. 2014;15:2379-2383. DOI: 10.1002/cbic.201402298
  23. 23. Amand S, Langenfeld A, Blond A, Dupont J, Ney B, Prado S. Guaiane sesquiterpenes from Biscogniauxia nummularia featuring potent antigerminative activity. Journal of Natural Products. 2012;75:798-801. DOI: 10.1021/np2009913
  24. 24. Teisseire P, Maupetit P, Corbier B. Contribution to the knowledge of Patchouliu oil. Recherches (Paris). 1974;19:8-35
  25. 25. Dolejs L, Mironov A, Sorm F. On terpenes. CXXI, structure of bulnesol and stereochemistry of guaiol, nepetalinic acids and iridomyrmecins. Collection of Czechoslovak Chemical Communications. 1961;26:1015-1020. DOI: org/10.1135/cccc19611015
  26. 26. Van Beek TA, Joulain D. The essential oil of patchouli, Pogostemon cablin: A review. Flavour and Fragrance Journal. 2018;33:6-51. DOI: 10.1002/ffj.3418
  27. 27. Crotea R, Munk SL, Akoh CC, Fisk HJ, Satterwhite DM. Biosynthesis of the sesquiterpene patchoulol from farnesyl pyrophosphate in leaf extracts of Pogostemon cablin (patchouli): Mechanistic considerations. Archives of Biochemistry and Biophysics. 1987;256:56-68. DOI: org/10.1016/0003-9861(87)90425-5
  28. 28. Akhila A, Sharma PK, Thakur RS. Biosynthetic relationships of patchouli alcohol, seychellene and cycloseychellene in Pogostemon cablin. Phytochemistry. 1988;27:2105-2108. DOI: org/10.1016/0031-9422(88)80105-5
  29. 29. Faraldos JA, Wu S, Chappell J, Coates RM. Doubly deuterium-labeled patchouli alcohol from cyclization of singly labeled [2-2H1]farnesyl diphosphate catalyzed by recombinant patchoulol synthase. Journal of the American Chemical Society. 2010;132:2998-3008. DOI: 10.1021/ja909251r
  30. 30. Pyle BW, Tran HT, Pickel B, Haslam TM, Gao Z, MacNevin G, Vederas JC, Kim S-U, Ro D-K. Enzymatic synthesis of valerena-4,7(11)-diene by a unique sesquiterpene synthase from the valerian plant (Valeriana officinalis). FEBS Journal. 2012;279:3163-3146. DOI: 10.1111/j.1742-4658.2012.08692.x
  31. 31. Yeo YS, Nybo SE, Chittiboyina AG, Weerasooriya AO, Wang YH, Gongoa-Castillo E, Vaillancourt B, Buell CR, DellaPenna D, Celiz MD, Jones AD, Wurtele ES, Ransom N, Dudareva N, Shaaban KA, Tibrewal N, Chandra S, Smillie T, Khan IA, Coates RM, Watt DS, Chappell J. Functional identification of valerena-1,10-diene synthase, a terpene synthase catalyzing a unique chemical cascade in the biosynthesis of biologically active sesquiterpenes in Valeriana officinalis. Journal of Biological Chemistry. 2013;288:3163-3173. DOI: 10.1074/jbc.M112.415836
  32. 32. Paul C, Koenig WA, Muhle H. Pacifigorgianes and tamariscene as constituents of Frullania tamarisci and Valeriana officinalis. Phytochemistry. 2001;57:307-313. DOI: org/10.1016/S0031-9422(01)00018-8
  33. 33. Yan S, Li S, Wu W, Zhao F, Bao L, Ding R, Gao H, Wen HA, Song F, Liu HW. Terpenoid and phenolic metabolites from the fungus Xylaria sp. associated with termite nests. Chemistry and Biodiversity. 2011;8:1689-1700. DOI: 10.1002/cbdv.201100026
  34. 34. Granchi C, Minutolo F. Anticancer agents that counteract tumor glycolysis. ChemMedChem. 2012;7:1318-1350. DOI: 10.1002/cmdc.201200176
  35. 35. Rabe P, Schmitz T, Dickschat JS. Mechanistic investigations on six bacterial terpene cyclases. Beilstin Journal of Organic Chemistry. 2016;12:1839-1850. DOI: 10.3762/bjoc.12.173
  36. 36. Bhatawadekar SV, Pednekar PR, Chakravarti KK, Paknikar SK. A survey of sesquiterpenoids of vetiver oil. In: Atal CK, Kapur BM, editors. Cultivation and Utilization of Aromatic Plants. Jammu-Tawi: RRL, CSIR; 1982. pp. 412-426 and references cited therein

Written By

Shashikumar K. Paknikar and Kamlesh Pai Fondekar

Submitted: 17 October 2017 Reviewed: 06 February 2018 Published: 05 November 2018