Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Nov 4.
Published in final edited form as: J Am Chem Soc. 2022 Oct 14;144(42):19253–19257. doi: 10.1021/jacs.2c09201

Development of a C–C Bond Cleavage/Vinylation/Mizoroki–Heck Cascade Reaction: Application to the Total Synthesis of 14- and 15-Hydroxypatchoulol

Christina G Na 1, Suh Hyun Kang 2, Richmond Sarpong 3
PMCID: PMC9635941  NIHMSID: NIHMS1844677  PMID: 36240482

Abstract

A C–C bond cleavage/vinylation/Mizoroki–Heck cascade reaction has been developed to provide access to densely functionalized bicyclo[2.2.2]octane frameworks. The sequence proceeds through the coupling of dihydroxylated pinene derivatives, prepared from carvone, with gem-dichloroalkenes. The method was applied to 12-step total syntheses of both 14- and 15-hydroxypatchoulol, which provided unambiguous support for the structure of the natural products and corrects a misassignment in the isolation report.


The combination of transition-metal-catalyzed C–C bond cleavage and cross-coupling reactions has been shown to be a powerful strategy for rapidly generating molecular complexity from simple building blocks.1 When applied to total synthesis, such methods can underpin nonintuitive strategies to access various complex molecular frameworks in an expedient fashion.2 Our group has previously explored the transition-metal-catalyzed C–C bond cleavage/cross-coupling reactions of dihydroxylated pinene derivatives (e.g., 1, Figure 1A),3 prepared from carvone, to provide functionalized cyclohexenone derivatives (see 1 → 3).4 The cyclohexenones thus obtained were further elaborated to bicyclo[3.3.1]nonane frameworks (e.g., 4) using a radical cyclization approach, which enabled the total synthesis of xishacorene B.4d Overall, this strategy of “chiral pool” remodeling has enabled access to a unique chemical space starting from readily available, enantioenriched small molecules such as carvone.5

Figure 1.

Figure 1.

(A, B) C–C bond cleavage/cross-coupling tactics to access bridged bicycles and (C) natural products containing bicyclo[2.2.2]-octane frameworks.

As part of our ongoing interest in the development of C–C bond cleavage/cross-coupling reactions and their application to total synthesis, we sought to access bicyclo[2.2.2]octanes (e.g., 7, Figure 1B) in a single operation from cyclobutanol 1 and gem-dihaloalkene electrophiles (5).6 These frameworks were expected to arise from cross-coupled product 6 through a subsequent C–C bond formation with the α-carbon of the enone moiety in an intramolecular Mizoroki–Heck reaction.7 Highly substituted bicycles such as 7 bearing various functional groups present many opportunities for derivatization and therefore application in complex molecule synthesis. In one such application, we envisioned that this transformation could provide short access to the hydroxylated patchoulols (Figure 1C).

14-Hydroxypatchoulol (9) and 15-hydroxypatchoulol (10) are tricyclic sesquiterpenoids that were isolated in 2015 from Valeriana stenoptera, a plant used in traditional medicine and in the fragrance industry.8,9 These molecules contain the skeleton characteristic of patchouli alcohol (8) adorned with additional hydroxy groups. We hypothesized that our planned C–C bond cleavage/cross-coupling/Mizoroki–Heck cascade reaction would allow for a concise synthesis of 10.

Mechanistically, we envisioned a dual Pd(0)/(II) catalytic cycle, first involving insertion of a Pd(0)-complex into gem-dihaloalkene 5 to provide Pd(II)-species 12 (Figure 2A). Ligand exchange with cyclobutanol 11 would form Pd(II)-alkoxide 13, which could undergo β-carbon elimination to yield alkyl-substituted Pd(II)-complex 14. Reductive elimination would then give vinyl chloride 6 to close the first catalytic cycle. In a new catalytic cycle, oxidative addition of 6 would give 15, which could undergo an intramolecular migratory insertion into the enone moiety to provide 16 containing the bicyclo[2.2.2]octane scaffold. β-Hydride elimination would then close the second cycle to afford 7 as the desired product. We hypothesized that with an appropriate ligand—in particular, one that would provide a large steric demand—insertion at the α-carbon could be favored to position the Pd(II) complex at the less substituted β-carbon.7 Initial studies therefore commenced with a screen of various ligands using Pd(OAc)2 as the catalyst, cyclobutanol 11, dichlorostyrene 17, and cesium carbonate. From this screening of conditions, we identified Xantphos as the ligand that provided the highest conversion of 11 and yield of the desired bicyclo[2.2.2]octane (i.e., 19; see the Supporting Information for details on the ligand screen). At 70 °C, 19 was obtained in 56% yield along with a small amount of vinyl chloride 18 (Figure 2B, entry 1). The loading of Xantphos relative to Pd(OAc)2 proved to be critical for reliable results, with a lower ratio of bicyclo[2.2.2]octane 19 to vinyl chloride 18 observed at a higher ligand loading (entry 2). We speculate that when excess Xantphos is used relative to Pd, ligand association saturates the Pd center, leading to catalyst deactivation and diminished oxidative addition of 18.10 The use of excess base proved to be important, with poorer conversion to 19 observed when 2 equiv of Cs2CO3 were used instead of 4 equiv (entry 3). In this case, the excess base likely scavenges the HCl generated in the course of the reaction. At 100 °C, neither 18 nor 19 was detected due to decomposition of 17 at these higher temperatures (entry 4). At a lower temperature of 40 °C, only formation of 18 was observed along with recovered cyclobutanol 11 (entry 5). Using the more reactive dibromoalkene 20 led to alkyne adduct 21 as the major product (Figure 2C)—presumably formed from the dehydrobromination of 20 to give a Pd-acetylide species prior to the coupling.11

Figure 2.

Figure 2.

(A) Proposed mechanism. (B) Optimization table. aDetermined by 1H NMR spectroscopy of the crude reaction mixtures using 1,3,5-trimethoxybenzene as an internal standard; bIsolated yield. (C) Reaction with dibromostyrene.

Having established suitable conditions to achieve our desired C–C bond cleavage/vinylation/Mizoroki–Heck cascade reaction using styrenyl gem-dichloroalkene electrophiles, we next investigated the use of alkyl gem-dichloroalkenes that would be desired for the total syntheses of the hydroxypatchoulol congeners. Dichloroalkenes possessing an alkyl substituent (e.g., 22, Figure 3) led to tricyclo[3.2.1.0]octanes (e.g., 23) as the major product. We hypothesize that, with alkyl dichloroalkene coupling partners, two consecutive carbopalladations occur prior to β-H elimination, giving rise to 23 via 24, whereas β-H elimination following a second carbopalladation event is not possible for the aryl substituted dichloroalkene electrophiles discussed earlier (see Figure 2B). A range of alkyl-substituted dichloroalkenes (25–30) coupled with 11 under the conditions that we have established to provide the tricyclo[3.2.1.0]octane products in moderate yields.

Figure 3.

Figure 3.

Reactions with alkyl-substituted dichloroalkenes and further elaboration to the [3.2.1] framework. aDetailed reaction conditions: 1.5 equiv of electrophile, 10 mol % Pd(OAc)2, 11 mol % Xantphos, 4 equiv of Cs2CO3, 1,4-dioxane (0.2 M), 70 °C, 24 h. b100 °C instead of 70 °C. c100 °C instead of 70 °C, 20 mol % instead of 11 mol % Xantphos.

The vinylcyclopropane moiety of the resulting products underwent concomitant hydrogenation and hydrogenolysis when subjected to Pd/C and H2 to give bicyclo[3.2.1]octane products (e.g., 31 → 32). Although cleavage of the less-substituted C–C bond is usually favored in reactions of vinylcyclopropanes with metal complexes,12 we observed preferential cleavage of the more substituted C–C bond in 31, which we hypothesize results from nucleophilic attack of Pd onto the vinylcyclopropane moiety to give a π-allylpalladium species,13 which is selectively protonated and hydrogenated on the more accessible convex face to give 32.

Our observations in these initial coupling studies set the stage for a synthesis of 10, as outlined in the retrosynthesis in Scheme 1A. We envisioned the patchoulol framework arising from 33 through an intramolecular aldol reaction, and designed a synthesis involving a key coupling step between cyclobutanol 11 and dichlorovinyl ketone 34.14 Notably, 34 would introduce the carbonyl group to enable the aldol reaction to close the final ring. In addition, 34 lacks any β-hydrogens that would give rise to the competing formation of a tricyclo[3.2.1.0]octane from the coupling, as observed using dichloroalkenes 25–30. In practice, 34 polymerized under the reaction conditions. However, ketal-protected 35 proved more serviceable and provided bicyclo[2.2.2]octane 36 containing all of the requisite carbon atoms of the undecane carbon skeleton in 53% yield (Scheme 1B).

Scheme 1. Retrosynthesis and Key Stepa.

Scheme 1.

a(A) Retrosynthesis and (B) key step for the synthesis of 10.

The dioxolane protecting group in 36 was cleaved with pyridinium p-toluenesulfonate (PPTS) to provide enone 37 (Scheme 2A). Global hydrogenation of 37 was accomplished with Pd/C and H2 to give the fully saturated product (38) in excellent yield and diastereoselectivity. In the presence of excess lithium hexamethyldisilazide (LiHMDS), 38 cyclized to give the aldol product (not shown).15 We found that protection of the resulting hydroxy group was necessary to avoid a retro-aldol reaction under basic conditions. In situ protection of the alkoxide resulting from aldol cyclization of 38 with TMSCl provided the corresponding silyl ether, along with a mixture of silyl enol ethers which readily hydrolyzed to give 39 upon silica gel chromatographic purification.

Scheme 2. Synthesis of 15-Hydroxypatchoulol (10) and 14-Hydroxypatchoulol (9)a.

Scheme 2.

a(A) Synthesis of 15-hydroxypatchoulol (10). (B) Alternative perspective drawing of epi-40. (C) Synthesis of 14-hydroxypatchoulol (9).

α-Methylation of ketone 39 was accomplished in moderate yield (49% yield) albeit with a significant amount of recovered starting material (38% rsm), favoring alkylation on the convex face of the molecule to give undesired diastereomer 40 as the major product. Thermodynamic epimerization mediated by DBU at 70 °C provided epi-40 bearing the desired equatorially disposed α-methyl group (see Scheme 2B).

Removing the ketone group directly proved to be challenging. For example, attempted Wolff–Kishner reductions, including the Caglioti16 and Myers modifications,17 were unsuccessful. Attempts to convert the carbonyl group to a thiocarbonyl or thioketal prior to reduction were equally unsuccessful. Ultimately, reducing ketone epi-40 to alcohol 41 using dissolving metal conditions, followed by Barton–McCombie deoxygenation, accomplished the desired net removal of the ketone group. A subsequent cleavage of the silyl protecting groups furnished 10.

We found that the NMR data that we acquired for 10 did not match those described in the isolation report. As such, we unambiguously characterized 10 using X-ray crystallographic analysis. From our analysis, it appears that the compound characterized in the isolation report as 10 is actually a constitutional isomer of 10 that bears a primary hydroxy group, but does not contain the patchoulol skeleton. This latter statement is supported by the fact that all other hydroxymethyl derivatives of patchouli alcohol are known, and none have characterization data that match those attributed to the compound described as 10 in the isolation report.9a,18 To gain further support for our conclusions, we also prepared 14-hydroxypatchoulol (9) in an analogous manner to our synthesis of 15-hydroxypatchoulol, beginning from diastereomeric cyclobutanol 43 obtained from the reductive cyclization of epoxycarvone 42 (Scheme 2C), which provided data that also did not match those of the compound described as 10 in the isolation report.19

In conclusion, we have accomplished 12-step syntheses of both 14- and 15-hydroxypatchoulol from (R)-carvone. Key to this strategy was the development of a concise entry into substituted bicyclo[2.2.2]octanes using a Pd-catalyzed cyclobutanol C–C bond cleavage/cross-coupling/Mizoroki–Heck cascade coupling reaction with select gem-dichloroalkenes, which differs from pioneering syntheses of the patchouli sesquiterpenoids that feature cycloadditions to build the bicyclo[2.2.2]octane framework.20 In our approach, tricyclo[3.2.1.0]octanes can also be prepared when alkyl substituted dichloroalkenes are employed as coupling partners. Given the biological21 and organoleptic22 properties of patchouli alcohol, we anticipate that this synthetic approach to hydroxylated patchouli alcohol congeners will set the stage for comprehensive function studies. In addition to preparing other hydroxylated patchoulol congeners for testing, our current studies are focused on gaining insight into the observed regioselectivity of the intramolecular Mizoroki–Heck portion of the cascade sequence, as well as generalizing the reaction manifold to other cycloalkanol substrates to access other unique carbon frameworks.

Supplementary Material

SI

ACKNOWLEDGMENTS

R.S. is grateful to the National Institutes of General Medical Sciences (R35 GM130345) for financial support. C.G.N. thanks the NIH for a postdoctoral fellowship (F32 GM137490). We thank Dr. Hasan Celik and UC Berkeley’s NMR facility in the College of Chemistry (CoC-NMR) for spectroscopic assistance. Instruments in CoC-NMR are supported in part by NIH S10OD024998. We are grateful to Dr. Nicholas Settineri (UC Berkeley) for X-ray crystallographic studies of 10 and Dr. Miao Zhang (DOE Catalysis Facility, UC Berkeley) for support with the acquisition of IR data. We thank the Hartwig lab for helpful discussions.

Footnotes

The authors declare no competing financial interest.

Accession Codes

CCDC 2203888 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

Complete contact information is available at: https://pubs.acs.org/10.1021/jacs.2c09201

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.2c09201.

Experimental procedures and spectral data for all new compounds. (PDF)

Contributor Information

Christina G. Na, Department of Chemistry, University of California, Berkeley, California 94720, United States

Suh Hyun Kang, Department of Chemistry, University of California, Berkeley, California 94720, United States.

Richmond Sarpong, Department of Chemistry, University of California, Berkeley, California 94720, United States.

REFERENCES

  • (1).(a) Murakami M; Ishida N Cleavage of Carbon–Carbon σ-Bonds of Four-Membered Rings. Chem. Rev. 2021, 121, 264. [DOI] [PubMed] [Google Scholar]; (b) Fumagalli G; Stanton S; Bower JF Recent Methodologies That Exploit C–C Single-Bond Cleavage of Strained Ring Systems by Transition Metal Complexes. Chem. Rev. 2017, 117, 9404. [DOI] [PubMed] [Google Scholar]; (c) Murakami M; Ishida N Potential of Metal-Catalyzed C–C Single Bond Cleavage for Organic Synthesis. J. Am. Chem. Soc. 2016, 138, 13759. [DOI] [PubMed] [Google Scholar]; (d) Xu T; Dong G Rhodium-Catalyzed Regioselective Carboacylation of Olefins: A C – C Bond Activation Approach for Accessing Fused-Ring Systems. Angew. Chem., Int. Ed. 2012, 51, 7567. [DOI] [PubMed] [Google Scholar]; (e) Jun C-H Transition metal-catalyzed carbon–carbon bond activation. Chem. Soc. Rev. 2004, 33, 610. [DOI] [PubMed] [Google Scholar]
  • (2).For a review on C–C single bond cleavage strategies in total synthesis, see: Wang B; Perea MA; Sarpong R Transition Metal-Mediated C–C Single Bond Cleavage: Making the Cut in Total Synthesis. Angew. Chem., Int. Ed. 2020, 59, 18898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (3).Bermejo FA; Mateos AF; Escribano AM; Lago RM; Burón JM; López MR; González RR Ti(III)-promoted cyclizations. Application to the synthesis of (E)-endo-bergamoten-12-oic acids. Moth oviposition stimulants isolated from Lycopersicon hirsutum. Tetrahedron 2006, 62, 8933. [Google Scholar]
  • (4).(a) Weber M; Owens K; Masarwa A; Sarpong R Construction of Enantiopure Taxoid and Natural Product-like Scaffolds Using a C–C Bond Cleavage/Arylation Reaction. Org. Lett. 2015, 17, 5432. [DOI] [PubMed] [Google Scholar]; (b) Masarwa A; Weber M; Sarpong R Selective C–C and C–H Bond Activation/Cleavage of Pinene Derivatives: Synthesis of Enantiopure Cyclohexenone Scaffolds and Mechanistic Insights. J. Am. Chem. Soc. 2015, 137, 6327. [DOI] [PubMed] [Google Scholar]; (c) Kuroda Y; Nicacio KJ; da Silva-Jr IA; Leger PR; Chang S; Gubiani JR; Deflon VM; Nagashima N; Rode A; Blackford K; Ferreira AG; Sette LD; Williams DE; Andersen RJ; Jancar S; Berlinck RGS; Sarpong R Isolation, synthesis and bioactivity studies of phomactin terpenoids. Nat. Chem. 2018, 10, 938. [DOI] [PubMed] [Google Scholar]; (d) Kerschgens I; Rovira AR; Sarpong R Total Synthesis of (−)-Xishacorene B from (R)-Carvone Using a C–C Activation Strategy. J. Am. Chem. Soc. 2018, 140, 9810. [DOI] [PubMed] [Google Scholar]; (e) Leger PR; Kuroda Y; Chang S; Jurczyk J; Sarpong R C–C Bond Cleavage Approach to Complex Terpenoids: Development of a Unified Total Synthesis of the Phomactins. J. Am. Chem. Soc. 2020, 142, 15536. [DOI] [PMC free article] [PubMed] [Google Scholar]; (f) Lusi RF; Perea MA; Sarpong R C–C Bond Cleavage of α-Pinene Derivatives Prepared from Carvone as a General Strategy for Complex Molecule Synthesis. Acc. Chem. Res. 2022, 55, 746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (5).Brill ZG; Condakes ML; Ting CP; Maimone TJ Navigating the Chiral Pool in the Total Synthesis of Complex Terpene Natural Products. Chem. Rev. 2017, 117, 11753. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (6).(a) Pioneering work with orthogonal coupling of 1,1-dihaloalkenes: Roush WT; Moriarty KJ; Brown BB Stereoselective synthesis of (Z,E)-2-bromo-1,3-dienes via the palladium(0) catalyzed cross coupling reactions of 1,1-dibromoolefins and vinylboronic acids. Tetrahedron Lett. 1990, 31, 6509. [Google Scholar]; (b) Roush WR; Kageyama M Enantioselective synthesis of the bottom-half of chlorothricolide. Tetrahedron Lett. 1985, 26, 4327. [Google Scholar]; (c) Roush WR; Riva R Enantioselective synthesis of the bottom half of chlorothricolide. 2. A comparative study of substituent effects on the stereoselectivity of the key intramolecular Diels-Alder reaction. J. Org. Chem. 1988, 53, 710. [Google Scholar]; (d) Roush WR; Brown BB; Drozda SE A highly stereoselective synthesis of the octahydronaphthalene subunit of kijanolide and tetronolide. Tetrahedron Lett. 1988, 29, 3541. [Google Scholar]
  • (7).(a) For reviews on the Mizoroki–Heck reaction, see: Douglas CJ; Overman LE Catalytic asymmetric synthesis of all-carbon quaternary stereocenters. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 5363. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Dounay AB; Overman LE The Asymmetric Intramolecular Heck Reaction in Natural Product Total Synthesis. Chem. Rev. 2003, 103, 2945. [DOI] [PubMed] [Google Scholar]; (c) Beletskaya IP; Cheprakov AV The Heck Reaction as a Sharpening Stone of Palladium Catalysis. Chem. Rev. 2000, 100, 3009. [DOI] [PubMed] [Google Scholar]
  • (8).Dong F-W; Wu Z-K; Yang L; Zi C-T; Yang D; Ma R-J; Liu Z-H; Luo H-R; Zhou J; Hu J-M Iridoids and sesquiterpenoids of Valeriana stenoptera and their effects on NGF-induced neurite outgrowth in PC12 cells. Phytochemistry 2015, 118, 51. [DOI] [PubMed] [Google Scholar]
  • (9).(a) Compounds 9 and 10 have been renamed based on the numbering used for the patchoulol skeleton in previous reports: Aleu J; Hanson JR; Galán RH; Collado IG Biotransformation of the Fungistatic Sesquiterpenoid Patchoulol by Botrytis cinerea. J. Nat. Prod. 1999, 62, 437. [DOI] [PubMed] [Google Scholar]; (b) 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. J. Am. Chem. Soc. 2010, 132, 2998. [DOI] [PubMed] [Google Scholar]
  • (10).Herrmann WA; Brossmer C; Ofele K; Beller M; Fischer H Zum Mechanismus der Heck-Reaktion: Katalysator-Deaktivierung durch PC-Bindungsbruch. J. Organomet. Chem. 1995, 491, C1. [Google Scholar]
  • (11).Chelucci G Synthesis and Metal-Catalyzed Reactions of gem-Dihalovinyl Systems. Chem. Rev. 2012, 112, 1344. [DOI] [PubMed] [Google Scholar]
  • (12).(a) Poulter SR; Heathcock CH Catalytic hydrogenation of cyclopropyl alkenes. I. Effect of alkyl substitution on the heterogeneous hydrogenolysis reaction. Tetrahedron Lett. 1968, 9, 5339. [Google Scholar]; (b) Barrett AGM; Tam W Regioselective Ring Opening of Vinylcyclopropanes by Hydrogenation with Palladium on Activated Carbon. J. Org. Chem. 1997, 62, 7673. [Google Scholar]
  • (13).Rubin M; Rubina M; Gevorgyan V Transition Metal Chemistry of Cyclopropenes and Cyclopropanes. Chem. Rev. 2007, 107, 3117. [DOI] [PubMed] [Google Scholar]
  • (14).Barker D; Brimble MA; Do P; Turner P Addition of silyloxydienes to 2,6-dibromo-1,4-benzoquinone: an approach to highly oxygenated bromonaphthoquinones for the synthesis of thysanone. Tetrahedron 2003, 59, 2441. [Google Scholar]
  • (15).Cory RM; Bailey MD; Tse DWC A divergent approach to patchouli sesquiterpenes: Synthesis of 3-oxopatchouli alcohol, 5-oxo-7-hydroxy-13-norcycloseychellene, 6-methoxy-4,12-dehydro-13-norcycloseychellene and patchouli alcohol. Tetrahedron Lett. 1990, 31, 6839. [Google Scholar]
  • (16).Caglioti L; Magi M The reaction of tosylhydrazones with lithium aluminium hydride. Tetrahedron 1963, 19, 1127. [Google Scholar]
  • (17).Furrow ME; Myers AG Practical Procedures for the Preparation of N-tert-Butyldimethylsilylhydrazones and Their Use in Modified Wolff–Kishner Reductions and in the Synthesis of Vinyl Halides and gem-Dihalides. J. Am. Chem. Soc. 2004, 126, 5436. [DOI] [PubMed] [Google Scholar]
  • (18).Ding W-B; Lin L-D; Liu M-F; Wei X-Y Two new sesquiterpene glycosides from Pogostemon cablin. J. Asian Nat. Prod. Res. 2011, 13, 599. [Google Scholar]
  • (19). While 9 can be prepared from patchouli alcohol (8) using a hydroxy-directed C–H functionalization (private communication with Prof. John Hartwig (UC Berkeley) and co-workers), such a strategy would not provide as straightforward of an approach to 10.
  • (20).(a) For previous syntheses of patchouli alcohol, see: Xu G-Q; Lin G-Q; Sun B-F Concise asymmetric total synthesis of (−)-patchouli alcohol. Org. Chem. Front. 2017, 4, 2031. [Google Scholar]; (b) Näf F; Ohloff G A Short Stereoselective Total Synthesis of Racemic Patchouli Alcohol. Helvetica Chimica 1974, 57, 1868. [Google Scholar]; (c) Mirrington RN; Schmalzl KJ Bicyclo[2.2.2]octenes. V. Total synthesis of (±)-patchouli alcohol. J. Org. Chem. 1972, 37, 2871. [Google Scholar]; (d) Dumas D; Danishefsky S The total synthesis of racemic patchouli and epi-patchouli alcohol. Chem. Commun. 1968, 21, 1287. [Google Scholar]
  • (21).Lee H-S; Lee J; Smolensky D; Lee S-H Potential benefits of patchouli alcohol in prevention of human diseases: A mechanistic review. Int. Immunopharmacol. 2020, 89, 107056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).Barton DHR; Belœil J-C; Billion A; Boivion J; Lallemand J-Y; Mergui S Oxidation of patchouli alcohol by the ‘gif system’: Isolation and organoleptic properties of three new ketonic derivatives. Helv. Chim. Acta 1987, 70, 273. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

SI

RESOURCES