Carpanone

Last updated
Carpanone
Carpanone.png
Names
Other names
Cupanone
Identifiers
3D model (JSmol)
ChemSpider
PubChem CID
UNII
  • InChI=1S/C20H18O6/c1-9-3-11-13(21)5-17-20(25-8-24-17)19(11)18(10(9)2)12-4-15-16(23-7-22-15)6-14(12)26-20/h3-6,9-10,18-19H,7-8H2,1-2H3/t9-,10+,18+,19+,20?/m0/s1 Yes check.svgY
    Key: WTXORUUTAZJKSN-JMAAQRFFSA-N Yes check.svgY
  • InChI=1/C20H18O6/c1-9-3-11-13(21)5-17-20(25-8-24-17)19(11)18(10(9)2)12-4-15-16(23-7-22-15)6-14(12)26-20/h3-6,9-10,18-19H,7-8H2,1-2H3/t9-,10+,18+,19+,20?/m0/s1
    Key: WTXORUUTAZJKSN-JMAAQRFFBX
  • CC1C=C2C3C(C1C)c4cc5c(cc4OC36C(=CC2=O)OCO6)OCO5
Properties
C20H18O6
Molar mass 354.343 g/mol
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
Yes check.svgY  verify  (what is  Yes check.svgYX mark.svgN ?)
Infobox references

Carpanone is a naturally occurring lignan-type natural product most widely known for the remarkably complex way nature prepares it, and the similarly remarkable success that an early chemistry group, that of Orville L. Chapman, had at mimicking nature's pathway. [1] [2] Carpanone is an organic compound first isolated from the carpano trees (Cinnamomum sp.) of Bougainville Island by Brophy and coworkers, trees from which the natural product derives its name. [1] [3] The hexacyclic lignan is one of a class of related diastereomers isolated from carpano bark as mixtures of equal proportion of the "handedness" of its components (i.e., racemic mixtures), and is notable in its stereochemical complexity, because it contains five contiguous stereogenic centers. The route by which this complex structure is achieved through biosynthesis involves a series of reactions that, almost instantly, take a molecule with little three-dimensionality to the complex final structure. Notably, Brophy and coworkers isolated the simpler carpacin, a phenylpropanoid with a 9-carbon framework, recognized its substructure as being dimerized within the complex carpanone structure, [4] and proposed a hypothesis of how carpacin was converted to carpanone in plant cells:

Contents

Carpacin, an ortho-methoxystyrene, and a more common type of phenolic plant phenylpropanoid whose structure was recognized as being dimerized in carpanone Carpacin.png
Carpacin, an ortho-methoxystyrene, and a more common type of phenolic plant phenylpropanoid whose structure was recognized as being dimerized in carpanone

Remarkably, within two years, Chapman and coworkers were able to chemically design a route to mimic this proposed biosynthetic route, and achieved the synthesis of carpanone from carpacin in a single "pot", in about 50% yield. [1] [2]

Carpanone itself is limited in its pharmacologic and biologic activities, but related analogs arrived at by variations of the Brophy-Chapman approach have shown activities as tool compounds relevant to mammalian exocytosis and vesicular traffic, [5] and provided therapeutic "hits" in antiinfective, antihypertensive, and hepatoprotective areas. [3]

The original Chapman design and synthesis is considered a classic in total synthesis, and one that highlights the power of biomimetic synthesis. [1] [6]

Total synthesis

The first total synthesis of carpanone was the biomimetic approach published by Chapman et al. in 1971. The required desmethylcarpacin (2-allyl sesamol), shown below as the starting molecule in the scheme, is acquired in two high-yield steps involving three transformations:

This procedure is one of several that gives the required desmethylcarpacin (carpacin with the methyl of its methoxy group removed). [3] Though oxidative dimerizations of phenols normally used a 1-electron oxidant, Chapman then followed a precedent involving a 2-electron oxidant and treated desmethylcarpacin with PdCl2 in the presence of sodium acetate (e.g., dissolved in a mixture of methanol and water); [1] [3] the reaction was perceived to proceed via a complexation of a pair of carpacins to the Pd(II) metal via their phenolic anions (as shown in scheme, below right), [6] followed by a classic 8-8' (β-β') oxidative phenolic coupling of the two olefin tails—shown crossing in the image—to give a dimeric trans-ortho-quinone methide-type of lignan intermediate. A particular conformation of this dimer then places a 4-electron enone of one ring over the 2-electron enol of the other (shown adjacent in image for clarity), setting the state for a variant of the Diels-Alder reaction termed an inverse demand Diels-Alder reaction (see curved arrows in image), which closes the 2 new rings and generates the 5 contiguous stereocenters. The carpanone is produced in yields of ≈50% by the original method, and in yields >90% in modern variants (see below). [1] [2] [3] The synthesis of a single diastereomer was confirmed in the original Chapman work, using X-ray crystallography.

Biomimetic transformation of desmethylcarpacin into carpanone in one pot, via a tandem oxidative coupling-Diels Alder reaction sequence. Note, in the second image in the scheme, the two lines crossing at the top are the two molecules overlapping each other (and do not imply chemical bonds). In this scheme, Pd (II) is shown forming a complex between two monomers of carpacin, then mediating oxidative 8-8' (b-b') phenolic coupling of their alkene tails to generate a dimer, a trans-ortho-quinone methide intermediate, followed immediately by an endo-selective inverse electron-demand hetero-Diels-Alder reaction (see Diels-Alder reaction#Mechanism), to close the rings and generates the stereocenters. Carpanone synthesis.png
Biomimetic transformation of desmethylcarpacin into carpanone in one pot, via a tandem oxidative coupling–Diels Alder reaction sequence. Note, in the second image in the scheme, the two lines crossing at the top are the two molecules overlapping each other (and do not imply chemical bonds). In this scheme, Pd (II) is shown forming a complex between two monomers of carpacin, then mediating oxidative 8-8' (β-β') phenolic coupling of their alkene tails to generate a dimer, a trans-ortho-quinone methide intermediate, followed immediately by an endo-selective inverse electron-demand hetero-Diels-Alder reaction (see Diels–Alder reaction#Mechanism), to close the rings and generates the stereocenters.

For the elegance of its "one-pot construction of a tetracyclic scaffold with complete stereocontrol of five contiguous stereo centers", [1] the original Chapman design and synthesis is "[n]ow considered a classic in total synthesis" that "highlights the power of biomimetic synthesis". [1] [6]

Extensions of the system

The Chapman approach has been applied in a variety of ways since its original report, varying substrates, oxidants, [7] and other aspects (and so synthesis of carpanone has subsequently been achieved by "quite a few research groups"); [1] [3] the actual mechanism of Pd(II) action is likely more complex than the original conjecture, and there is evidence that the mechanism, broadly speaking, depends on actual conditions (specific substrate, oxidant, etc.). [3] Various groups, including the laboratories of Steve Ley, Craig Lindley, and Matthew Shair, have succeeded in extending the Chapman method to solid-supported synthesis, i.e., phenolic starting materials on polymeric supports, thus allowing the generation of libraries of carpanone analogs. [1] [5] A hetero-8-8' oxidative coupling system akin to the Chapman approach has been developed that uses IPh(OAC)2, and that allows for preparation of more electron rich homodimers, and for hetero-tetracyclic analogs of carpanone. [8]

References and notes

  1. 1 2 3 4 5 6 7 8 9 10 11 C.W. Lindsley, C.R. Hopkins & G.A. Sulikowski, 2011, Biomimetic synthesis of lignans, In "Biomimetic Organic Synthesis" (E. Poupon & B. Nay, Eds.), Weinheim: Wiley-VCH, ISBN   9783527634767, see , accessed 4 June 2014.
  2. 1 2 3 O.L. Chapman, M.R. Engel, J.P. Springer & J.C. Clardy, 1971, Total synthesis of carpanone, J. Am. Chem. Soc.93:6697–6698.
  3. 1 2 3 4 5 6 7 F. Liron, F. Fontana, J.-O. Zirimwabagabo, G. Prestat, J. Rajabi, C. La Rosa & G. Poli, 2009, A New Cross-Coupling-Based Synthesis of Carpanone, Org. Lett., 11(19):4378–4381, DOI: 10.1021/ol9017326, see "Archived copy" (PDF). Archived from the original (PDF) on 2014-06-07. Retrieved 2014-06-06.{{cite web}}: CS1 maint: archived copy as title (link) or , accessed 4 June 2014
  4. G.C. Brophy, J. Mohandas, M. Slaytor, T.R. Watson & L.A. Wilson, 1969, Novel lignans from a Cinnamomum sp. from Bougainville, Tetrahedron Lett. 10:5159-5162.
  5. 1 2 Brian C. Goess, Rami N. Hannoush, Lawrence K. Chan, Tomas Kirchhausen, and Matthew D. Shair, 2006, Synthesis of a 10,000-Membered Library of Molecules Resembling Carpanone and Discovery of Vesicular Traffic Inhibitors, J. Am. Chem. Soc.128(16): 5391–5403, DOI: 10.1021/ja056338g, see , accessed 4 June 2014.
  6. 1 2 3 4 Nicolaou, K. C.; E. J. Sorensen (1996). Classics in Total Synthesis . Weinheim, Germany: VCH. pp.  95–97. ISBN   978-3-527-29284-4.
  7. Per Lindsley et al., see following, oxidant systems, generally including dioxygen, adventitious or otherwise, include azobisisobutyronitrile, Ag2O, M(II) salen systems (M=Co, Mn, Fe), singlet oxygen (hν, Rose Bengal), dibenzoyl peroxide, and IPh(OAC)2.
  8. C.W. Lindsley, L.K. Chan, B.C. Goess, R. Joseph & M.D. Shair, 2001, Solid-phase biomimetic synthesis of carpanone-like molecules, J. Am. Chem. Soc. 122, 422–423.

Further reading

Related Research Articles

Aromatic compounds are those chemical compounds that contain one or more rings with pi electrons delocalized all the way around them. In contrast to compounds that exhibit aromaticity, aliphatic compounds lack this delocalization. The term "aromatic" was assigned before the physical mechanism determining aromaticity was discovered, and referred simply to the fact that many such compounds have a sweet or pleasant odour; however, not all aromatic compounds have a sweet odour, and not all compounds with a sweet odour are aromatic compounds. Aromatic hydrocarbons, or arenes, are aromatic organic compounds containing solely carbon and hydrogen atoms. The configuration of six carbon atoms in aromatic compounds is called a "benzene ring", after the simple aromatic compound benzene, or a phenyl group when part of a larger compound.

The Stille reaction is a chemical reaction widely used in organic synthesis. The reaction involves the coupling of two organic groups, one of which is carried as an organotin compound. A variety of organic electrophiles provide the other coupling partner. The Stille reaction is one of many palladium-catalyzed coupling reactions.

In organic chemistry, an electrocyclic reaction is a type of pericyclic rearrangement where the net result is one pi bond being converted into one sigma bond or vice versa. These reactions are usually categorized by the following criteria:

Arynes or benzynes are highly reactive species derived from an aromatic ring by removal of two substituents. The most common arynes are ortho but meta- and para-arynes are also known. o-Arynes are examples of strained alkynes.

In chemical synthesis, "click" chemistry is a class of biocompatible small molecule reactions commonly used in bioconjugation, allowing the joining of substrates of choice with specific biomolecules. Click chemistry is not a single specific reaction, but describes a way of generating products that follow examples in nature, which also generates substances by joining small modular units. In many applications, click reactions join a biomolecule and a reporter molecule. Click chemistry is not limited to biological conditions: the concept of a "click" reaction has been used in chemoproteomic, pharmacological, and various biomimetic applications. However, they have been made notably useful in the detection, localization and qualification of biomolecules.

Bamford–Stevens reaction

The Bamford–Stevens reaction is a chemical reaction whereby treatment of tosylhydrazones with strong base gives alkenes. It is named for the British chemist William Randall Bamford and the Scottish chemist Thomas Stevens Stevens (1900–2000). The usage of aprotic solvents gives predominantly Z-alkenes, while protic solvent gives a mixture of E- and Z-alkenes. As an alkene-generating transformation, the Bamford–Stevens reaction has broad utility in synthetic methodology and complex molecule synthesis.

Oxidative coupling in chemistry is a coupling reaction of two molecular entities through an oxidative process. Usually oxidative couplings are catalysed by a transition metal complex like in classical cross-coupling reactions, although the underlying mechanism is different due to the oxidation process that requires an external oxidant. Many such couplings utilize dioxygen as the stoichiometric oxidant but proceed by electron transfer.

A cascade reaction, also known as a domino reaction or tandem reaction, is a chemical process that comprises at least two consecutive reactions such that each subsequent reaction occurs only in virtue of the chemical functionality formed in the previous step. In cascade reactions, isolation of intermediates is not required, as each reaction composing the sequence occurs spontaneously. In the strictest definition of the term, the reaction conditions do not change among the consecutive steps of a cascade and no new reagents are added after the initial step. By contrast, one-pot procedures similarly allow at least two reactions to be carried out consecutively without any isolation of intermediates, but do not preclude the addition of new reagents or the change of conditions after the first reaction. Thus, any cascade reaction is also a one-pot procedure, while the reverse does not hold true. Although often composed solely of intramolecular transformations, cascade reactions can also occur intermolecularly, in which case they also fall under the category of multicomponent reactions.

A dendralene is a discrete acyclic cross-conjugated polyene. The simplest dendralene is buta-1,3-diene (1) or [2]dendralene followed by [3]dendralene (2), [4]dendralene (3) and [5]dendralene (4) and so forth. [2]dendralene (butadiene) is the only one not cross-conjugated.

Persistent carbene In organic chemistry, a type of carbene demonstrating particular stability

A persistent carbene (also known as stable carbene) is a type of carbene demonstrating particular stability. The best-known examples and by far largest subgroup are the N-heterocyclic carbenes (NHC) (sometimes called Arduengo carbenes), for example diaminocarbenes with the general formula (R2N)2C:, where the four R moieties are typically alkyl and aryl groups. The groups can be linked to give heterocyclic carbenes, such as those derived from imidazole, imidazoline, thiazole or triazole.

Barrelene Chemical compound

Barrelene is a bicyclic organic compound with chemical formula C8H8 and systematic name bicyclo[2.2.2]octa-2,5,7-triene. First synthesized and described by Howard Zimmerman in 1960, the name derives from the resemblance to a barrel, with the staves being three ethylene units attached to two methine groups. It is the formal Diels–Alder adduct of benzene and acetylene. Due to its unusual molecular geometry, the compound is of considerable interest to theoretical chemists.

Danishefskys diene Chemical compound

Danishefsky's diene is an organosilicon compound and a diene with the formal name trans-1-methoxy-3-trimethylsilyloxy-buta-1,3-diene named after Samuel J. Danishefsky. Because the diene is very electron-rich it is a very reactive reagent in Diels-Alder reactions. This diene reacts rapidly with electrophilic alkenes, such as maleic anhydride. The methoxy group promotes highly regioselective additions. The diene is known to react with amines, aldehydes, alkenes and alkynes. Reactions with imines and nitro-olefins have been reported.

The Rubottom oxidation is a useful, high-yielding chemical reaction between silyl enol ethers and peroxyacids to give the corresponding α-hydroxy carbonyl product. The mechanism of the reaction was proposed in its original disclosure by A.G. Brook with further evidence later supplied by George M. Rubottom. After a Prilezhaev-type oxidation of the silyl enol ether with the peroxyacid to form the siloxy oxirane intermediate, acid-catalyzed ring-opening yields an oxocarbenium ion. This intermediate then participates in a 1,4-silyl migration to give an α-siloxy carbonyl derivative that can be readily converted to the α-hydroxy carbonyl compound in the presence of acid, base, or a fluoride source.

Absinthin Chemical compound

Absinthin is a naturally produced triterpene lactone from the plant Artemisia absinthium (Wormwood). It constitutes one of the most bitter chemical agents responsible for absinthe's distinct taste. The compound shows biological activity and has shown promise as an anti-inflammatory agent, and should not to be confused with thujone, a neurotoxin also found in Artemisia absinthium.

Endiandric acid C Chemical compound

Endiandric acid C, isolated from the tree Endiandra introrsa, is a well characterized chemical compound. Endiadric acid C is reported to have better antibiotic activity than ampicillin.

The retro-Diels–Alder reaction (rDA) is the microscopic reverse of the Diels–Alder reaction —the formation of a diene and dienophile from a cyclohexene. It can be accomplished spontaneously with heat, or with acid or base mediation.

Phenol oxidation with hypervalent iodine reagents leads to the formation of quinone-type products or iodonium ylides, depending on the structure of the phenol. Trapping of either product is possible with a suitable reagent, and this method is often employed in tandem with a second process.

Photoredox catalysis

Photoredox catalysis is a branch of photochemistry that uses single-electron transfer. Photoredox catalysts are generally drawn from three classes of materials: transition-metal complexes, organic dyes, and semiconductors. While organic photoredox catalysts were dominant throughout the 1990s and early 2000s, soluble transition-metal complexes are more commonly used today.

Ni(IV) organometallic complexes (Nickel(IV)-centered organometallic complexes) are a form of organonickel compounds which feature a central nickel atom at the +4 oxidation state. These complexes typically need to be supported by non-canonical ligand scaffolds which are usually developed fit-for-purpose. These types of complexes are rare, with only few reports in the chemical literature when compared to lower oxidation state Ni species. These types of high-valent nickel compounds have risen in popularity only in the past decade due to their general instability, challenging characterization, and unique reactivity with organic molecules. Organic reactions exploiting the reactivity of nickel typically occur via organometallic Ni0, NiI, NiII, and NiIII intermediate species. These types of species are well studied and have been the subject of several literature reviews, with few regarding NiIV species.

In organic chemistry, the Conia-ene reaction is an intramolecular cyclization reaction between an enolizable carbonyl such as an ester or ketone and an alkyne or alkene, giving a cyclic product with a new carbon-carbon bond. As initially reported by J. M. Conia and P. Le Perchec, the Conia-ene reaction is a heteroatom analog of the ene reaction that uses an enol as the ene component. Like other pericyclic reactions, the original Conia-ene reaction required high temperatures to proceed, limiting its wider application. However, subsequent improvements, particularly in metal catalysis, have led to significant expansion of reaction scope. Consequently, various forms of the Conia-ene reaction have been employed in the synthesis of complex molecules and natural products.