Next Article in Journal
One Step Formation of Propene from Ethene or Ethanol through Metathesis on Nickel Ion-loaded Silica
Next Article in Special Issue
Gold Nanoparticle-Catalyzed Environmentally Benign Deoxygenation of Epoxides to Alkenes
Previous Article in Journal
Camptothecin-20(s)-O-[N-(3’α,12’α-dihydroxy-24’-carbonyl-5’β-cholan)]-lysine, a Novel Camptothecin Analogue, Induces Apoptosis towards Hepatocellular Carcinoma SMMC-7721 Cells
Previous Article in Special Issue
Studies on the Electrochemical Behavior of Thiazolidine and Its Applications Using a Flow–Through Chronoamperometric Sensor Based on a Gold Electrode
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Gold-Catalyzed Cyclizations of Alkynol-Based Compounds: Synthesis of Natural Products and Derivatives

by
Benito Alcaide
1,*,
Pedro Almendros
2,* and
José M. Alonso
1
1
Grupo de Lactamas y Heterociclos Bioactivos, Departamento de Química Orgánica I, Unidad Asociada al CSIC, Facultad de Química, Universidad Complutense de Madrid, 28040-Madrid, Spain
2
Instituto de Química Orgánica General, CSIC, Juan de la Cierva 3, 28006-Madrid, Spain
*
Authors to whom correspondence should be addressed.
Molecules 2011, 16(9), 7815-7843; https://doi.org/10.3390/molecules16097815
Submission received: 15 August 2011 / Revised: 5 September 2011 / Accepted: 8 September 2011 / Published: 13 September 2011
(This article belongs to the Special Issue Gold Catalysts)

Abstract

:
The last decade has witnessed dramatic growth in the number of reactions catalyzed by gold complexes because of their powerful soft Lewis acid nature. In particular, the gold-catalyzed activation of propargylic compounds has progressively emerged in recent years. Some of these gold-catalyzed reactions in alkynes have been optimized and show significant utility in organic synthesis. Thus, apart from significant methodology work, in the meantime gold-catalyzed cyclizations in alkynol derivatives have become an efficient tool in total synthesis. However, there is a lack of specific review articles covering the joined importance of both gold salts and alkynol-based compounds for the synthesis of natural products and derivatives. The aim of this Review is to survey the chemistry of alkynol derivatives under gold-catalyzed cyclization conditions and its utility in total synthesis, concentrating on the advances that have been made in the last decade, and in particular in the last quinquennium.

Graphical Abstract

1. Introduction

Organic synthesis has as one of its major points of interest the study of naturally occurring substances, and this remains both a source of information and an intellectual challenge. Thus, a crucial target for organic chemists is to find the appropriate reaction conditions, allowing functional group compatibility and providing high efficiency and atom economy. During the last years, gold-catalyzed cycloisomerization of alkynol-based systems has emerged as a useful tool in this area, allowing the synthesis of different structures such as furans, dihydrofurans, pyrans, furanones or ketals, among many other heterocyclic systems and naturally occurring structures [1,2,3].
This overview focuses on the most recent achievements in gold-catalyzed cycloisomerization reactions, for the synthesis of natural products and related compounds. In particular, carbon-carbon and carbon-heteroatom cyclization processes will be considered, paying special attention to reports from the last five years.

2. Cycloisomerization Processes Involving Carbon-Carbon Bond Formation

Gold-catalyzed cycloisomerization reactions involving C−C bond formation have recently emerged as an effective methodology to build hydrocarbon rings. Four, five and six membered cyclic structures, as well as medium sized rings are accessible in good yields and under interesting mild reaction conditions using gold salts and gold complexes. Fused bicyclic compounds can also be produced, leading therefore to an attractive series of natural occurring skeletons.
Benzofurans represent a recurring motif among natural products. Particularly, 2-substituted and 2,7-disubstituted benzofurans and their derivatives are known to show many different biological activities such as antineoplastic, antiviral, antioxidative or anti-inflammatory properties. Although many routes for the preparation of 2-substituted systems have been developed [4,5], 2,7-substituted benzofurans remain almost unexplored (Figure 1) [6,7,8].
Figure 1. 7-Aryl benzofuran structure core in different bioactive compounds.
Figure 1. 7-Aryl benzofuran structure core in different bioactive compounds.
Molecules 16 07815 g001
Hashmi et al. have recently proposed an efficient route leading to 7-aryl benzo[b]furans 2 through a gold-catalyzed rearrangement of 3-silyloxy-1,5-enynes [9]. The considerable effort that went into this work, involving a first catalyst screening for substrate 1a and finding the optimal conditions for the dual catalyst system [IPrAuCl]/AgNTf2 is noteworthy. Thus, an easy methodology was performed, using mild conditions, open-air systems and remarkable short reaction times, providing an interesting family of different substituted 7-aryl benzofurans (Scheme 1).
Scheme 1. Gold(I)-catalyzed rearrangement of 3-silyloxy-1,5-enynes.
Scheme 1. Gold(I)-catalyzed rearrangement of 3-silyloxy-1,5-enynes.
Molecules 16 07815 g002
Approximately a quarter of biologically active known compounds come from fungi, and among their wide range of properties, the antibiotic activity has attracted much interest [10,11]. Guanacastepene A (Figure 2) is extremely active against methicillin-resistant strains of Staphylococcus aureus and vancomicyn-resistant E. faecalis, two drug-resistant common pathogens which have generated major concern [12,13,14].
Figure 2. Rings A, B and C in natural terpene guanacastepene A.
Figure 2. Rings A, B and C in natural terpene guanacastepene A.
Molecules 16 07815 g003
It has been stated an approach to ring A of guanacastepene, by an unusual gold(I)-catalyzed cycloisomerization of alkynol-based 1,5-enynes [15]. According to the proposed retrosynthesis, most of the functionalities present in the natural terpene would be early introduced, while the presence of the cyclopropyl fused ring could allow the further generation of ring B (Scheme 2).
Scheme 2. Retrosynthesis of guanacastepene ring A.
Scheme 2. Retrosynthesis of guanacastepene ring A.
Molecules 16 07815 g004
Many 1,5-enynes were tested, and an unexpected pattern of reactivity depending on the substituents in substrates 4 was found. Thus, the desired bicyclo[3.1.0] system 3 was obtained only when the reaction was performed with the syn-enynes 4a and 4b, yielding 3a and 3b with good conversions and notable diastereoselectivity. Anti-isomers, or any stereochemical change on the starting 1,5-enynes, resulted in the opposite diastereoselectivity (systems 5), or a dramatic change on the course of the reaction, leading to alkylidene-cyclopentenes 6, cyclohexadienes 7, or α,β-unsaturated aldehydes 8 (Scheme 3).
Gold-catalyzed isomerization has been also employed in the search of an appropriate route to (−)-thujopsanone, a derivative of the natural terpene (−)-thujopsene, widely employed in cosmetics [16]. Although the first aim of the authors remained unachieved, and the obtained compound 9 did not exhibit the appreciated properties of the initial target [17], the chemistry developed merits further consideration (Scheme 4) [18]. Thus, it was observed that enynol 10 provided the unexpected ether 11 in the presence of different gold catalysts, in amounts similar to those produced by some other metal salts such as copper or platinum complexes. Interestingly, when the corresponding acetate derivative 12 reacted in the presence of AuCl3, a tandem cycloisomerization/[1,2]-acyl shift took place, leading to adduct 13, precursor of the previously mentioned adduct 9, and a close system to (−)-thujopsanone. Moreover, when the process was tested in the presence of (tBuXPhos)AuNTf2 as catalyst, an unprecedented rearrangement/cycloaddition leading to the tricyclic system 14 was reported (Scheme 5).
Scheme 3. Divergent reactivity for the gold-catalyzed reaction of 1,5-enynes.
Scheme 3. Divergent reactivity for the gold-catalyzed reaction of 1,5-enynes.
Molecules 16 07815 g005
Scheme 4. Projected synthetic route to terpene (−)-thujopsanone.
Scheme 4. Projected synthetic route to terpene (−)-thujopsanone.
Molecules 16 07815 g006
Scheme 5. Cycloisomerization of enynol 10 and [1,2]-acyl shift rearrangement of acetate 12, respectively.
Scheme 5. Cycloisomerization of enynol 10 and [1,2]-acyl shift rearrangement of acetate 12, respectively.
Molecules 16 07815 g007
Gold-catalyzed cycloisomerization methodology has also been applied to the construction of medium sized rings. Allocolchicinoids, presenting a seven membered ring, are structures related to (−)-colchicine, a natural product with important antimitotic activity (Figure 3). Many of these derivatives also show this kind of mitosis arrest, by inhibiting tubulin polymerization [19,20,21]. N-acetylcolchinol 15, for instance, is described to bind to tubulin more strongly than colchicine itself. Thus, many reports have appeared describing the synthesis of these structures [22,23,24,25,26,27,28].
Figure 3. Colchicine and allocolchicinoids systems.
Figure 3. Colchicine and allocolchicinoids systems.
Molecules 16 07815 g008
Hanna et al. reported the synthesis of derivative 17 [29]. In the proposed sequence, the seven membered ring is formed by a gold(I)-catalyzed 1,2-O-acyl shift, followed by a cyclopropanation step which leads to the fused three member ring (Scheme 6). Thus, gold-catalyzed cyclization of alkynol-based systems has also been stated in this work as a useful tool to create medium sized rings, through an easy methodology providing high yields under mild reaction conditions.
Scheme 6. Synthesis of allocolchicinoid 17 and proposed mechanism for the gold-catalyzed cycloisomerization step.
Scheme 6. Synthesis of allocolchicinoid 17 and proposed mechanism for the gold-catalyzed cycloisomerization step.
Molecules 16 07815 g009
Indole systems are ubiquitous in Nature, appearing in many different alkaloid families. Their wide range of biological activities, and their intriguing chemistry, makes these compounds a target of special interest, and a recurring topic in many studies [30,31,32,33,34,35,36,37,38,39,40,41]. For instance, the first enantioselective approach to (−)-mersicarpine, an alkaloid isolated from Kopsia plants and exhibiting an unusual tetracyclic structure has been reported. The proposed retrosynthetic analysis included the reaction of an alkynol-based intermediate in the presence of a gold salt, although only the alkyne functional group showed reactivity under these conditions, preserving the hydroxylic group for a further oxidation [42].
More interestingly, the reactivity of alkynol-based systems as formal organic synthons has been also explored in the indole chemistry. It has been established the synthesis of the non-natural skeleton 2,3-indoline-fused cyclobutane through a cascade process, including both C−C and C−O bond formation catalyzed by the same gold salt [43].
On the other hand, Echavarren et al. described in an exhaustive report about inter- and intramolecular gold-catalyzed reaction of alkynes and indoles some examples starting from alkynols and alkynol-based systems. Carbazole-like systems and related structures were therefore achieved (Scheme 7) [44].
Scheme 7. Alkynol-based reactivity in indole chemistry.
Scheme 7. Alkynol-based reactivity in indole chemistry.
Molecules 16 07815 g010
Inspired by the results of the Echevarren group, Liu et al. described the synthesis of dihydrocyclohepta[b]indoles 26 from (Z)-enynols 27 and indole, through an interesting domino sequence including a first gold(0)-catalyzed Friedel-Craft reaction, followed by a hydroarylation step [45]. The resulting products are of considerable interest, as much as they form the key subunits of several alkaloids, like ambiguine, silicine, caulerpin or caulersin. The reported work includes the optimization of the process, by testing different gold salts and solvents, leading to high reaction conversions through mild conditions (Scheme 8 and Scheme 9).
Scheme 8. Synthesis of dihydrocyclohepta[b]indoles, and related natural structures.
Scheme 8. Synthesis of dihydrocyclohepta[b]indoles, and related natural structures.
Molecules 16 07815 g011

3. Cycloisomerization Processes Involving Carbon-Heteroatom Bond Formation

Heterocyclic natural occurring motifs such as furans, pyrans or spiroketals can be easily achieved through heterocyclization processes performed on alkynol-based systems. Gold promoted methodologies provide a convenient route to these structures, allowing mild reaction conditions and high yields. Total synthesis and the preparation of related derivatives have been recently described using both C−N and C−O bond formation.

3.1. Cycloisomerization on Alkynol-Based Systems

Chromones are natural heterocycles showing a wide range of biological properties. Thus, many strategies like iodocyclizations [46], metal-catalyzed cycloadditions [47], or O-arylation processes [48] have appeared for the synthesis of these oxacyclic systems. Gold catalyzed cycloisomerization of alkynol based structures 28 have been also stated for the generation of chromones 29 [49]. Interestingly, reaction proceeded with a further migration of group R1, leading to highly functionalized skeletons. Unluckily, only moderate yields were achieved (Scheme 10), inasmuch as isomerization processes competed with the expected Au-based cycloisomerization.
Scheme 9. Proposed reaction mechanism for the tandem gold catalyzed-Friedel-Crafts arylation/hydroarylation process.
Scheme 9. Proposed reaction mechanism for the tandem gold catalyzed-Friedel-Crafts arylation/hydroarylation process.
Molecules 16 07815 g012
Scheme 10. Synthesis of chromones by gold-catalyzed cycloisomerization.
Scheme 10. Synthesis of chromones by gold-catalyzed cycloisomerization.
Molecules 16 07815 g013
A similar approach has been developed for the synthesis of aurone skeletons [50], natural flavonoids, by an easy three step sequence. Aurones exhibit several biological properties [51,52,53,54,55], and its importance had led to several groups to develop convenient synthetic routes [56,57,58,59,60,61,62]. Among them, gold-catalyzed oxycyclization provided the best results, as milder reaction conditions and excellent selectivities, avoiding the formation of flavones as byproducts, were achieved (Scheme 11) [63]. In this case, high yields and complete regioselectivity were obtained.
Scheme 11. Synthesis of aurone skeleton by gold-catalyzed cycloisomerization.
Scheme 11. Synthesis of aurone skeleton by gold-catalyzed cycloisomerization.
Molecules 16 07815 g014
Moreover, the present methodology was used for the structural revision of two natural products, (Z)-4’-chloroaurone 30 [64], and (Z)-2’-hydroxyaurone 32 [65], proving that the assumed structures were not the correct ones. Thus, flavonoid systems 30 and 32 could be prepared by the above three step strategy which revealed that their spectral data did not match with the previously reported data of the natural isolated ones. Therefore, the isocumarin 31 and the flavone 33 were prepared and probed as the real structures for these natural products (Figure 4).
Figure 4. New assignation of structures 31 and 33 by comparison with the prepared by the gold catalysis aurone systems 30 and 32.
Figure 4. New assignation of structures 31 and 33 by comparison with the prepared by the gold catalysis aurone systems 30 and 32.
Molecules 16 07815 g015
Trost et al. recently completed the total synthesis of bryostatin 16 [66,67], a structurally complex macrolide which exhibits a wide range of biological activities [68,69,70,71]. Focusing on the proposed 26 step sequence (in the longest linear path, and 39 steps as the total), the gold-catalyzed 6-endo-dig oxycyclization of alkynol 34 to generate the inner dihydropyran cycle D in macrocyclic precursor 35 in 65% yield deserves special attention (Scheme 12).
Scheme 12. Gold-based synthesis of dihydropyran ring D in bryostatin total synthesis.
Scheme 12. Gold-based synthesis of dihydropyran ring D in bryostatin total synthesis.
Molecules 16 07815 g016
(+)-Cephalostatin 1 is another complex macrolide with interesting biological activity. It has been reported to be a promising anticancer agent for the p16 tumor suppressor gene, exhibiting high activity and high selectivity between cancer cells and normal cells [72,73]. Because of the small amounts of cephalostatin available from its natural marine sources, a synthetic approach has emerged as the sole viable tool to provide enough material for biological testing [74,75,76,77,78]. On the other hand, the structural complexity of cephalostatin makes this macrocycle an interesting target to develop new skills in organic synthesis.
Fortner et al. have recently described a total synthesis of cephalostatin, involving the construction of both its eastern and western fragments and their further coupling [79]. Along the high quality chemistry developed for this synthesis, we would like to focus on the dihydrofuran ring E construction on compound 36. Thus, gold-catalyzed cycloisomerization emerge again as a useful methodology to convert alkynol systems in oxacyclic skeletons, crucial and recurring motifs for total synthesis. Moreover, the efficiency of gold catalysis to promote a 5-endo-dig process with an 88% conversion, on what is a hindered internal alkyne 37, deserves special consideration (Scheme 13).
Scheme 13. Synthesis of ring E on the eastern fragment of cephalostatin.
Scheme 13. Synthesis of ring E on the eastern fragment of cephalostatin.
Molecules 16 07815 g017
Other natural occurring motifs such as oxazoles and isoxazoles have also been assembled through gold-catalyzed cycloisomerization. Thus, it has been recently established a general method for the synthesis of highly functionalized isoxazoles from alkynyl oxime ethers [80], or an intermolecular alkyne oxidation leading to 2,5-disubstituted oxazoles [81]. Nevertheless, while gold-based alkyne-oxygen cycloisomerization has recently become a hot topic in organic synthesis, only a few examples for alkyne-nitrogen coupling have been described [82,83,84,85,86,87,88,89,90]. Regarding the synthesis of natural products and derivatives, Chan et al. have recently described the synthesis of highly substituted indole skeletons [29], from readily available 2-tosylamino-phenylprop-1-yn-3-ols 38 [91]. The reported work shows a versatile approach to these natural occurring motifs, and develops a fascinating study concerning the chemical reactivity of these substrates under gold-catalyzed conditions. Thus, starting in every case from a 5-exo-dig cycloaddition which led to vinyl gold species 39, different reaction pathways were observed depending on the substituent group R1. It was stated that when R1 = aryl, reaction proceeded through a Friedel-Craft process, giving indenyl-fused indoles 40. On the other hand, changing to R1 = H, a protodeauration/1,3-allylic alcohol isomerization took place, leading to indoles 41. The presence of a nucleophile in the reaction media gave place mainly to systems 42, and for R1 = CHR2R3, a more facile protodeauration and dehydratation step delivered systems 43 (Scheme 14).
Scheme 14. Indole synthesis from gold-catalyzed cycloisomerization of 2-tosylamino-phenylprop-1-yn-3-ols.
Scheme 14. Indole synthesis from gold-catalyzed cycloisomerization of 2-tosylamino-phenylprop-1-yn-3-ols.
Molecules 16 07815 g018
Another example of gold-based C−N cyclization on alkynol systems for the total synthesis of (+)-andrachcinidine (44) has been established [92]. This natural alkaloid receives its name from its natural source, the beetle Andrachne aspera, and it has been shown to be an interesting chemical defense agent [93]. The proposed reaction sequence started with commercial ketal 45, which yielded after six steps the nitrogen-containing alkynol 46. Gold-catalyzed cyclization of 46 provided the piperidine system 47 as a single diastereomer in 89% isolated yield. The reaction mechanism is proposed to follow a first gold-based alkyne hydration providing ketone 48. Methoxy group cleavage would then generate the corresponding α,β-unsaturated system, which could undergo nucleophile addition building the expected 6-membered heterocycle (Scheme 15). It is noteworthy that no competition between nitrogen and oxygen attack was found, which would led to the less favoured 8-membered heterocycle.
Scheme 15. Synthesis of (+)-andrachcinidine.
Scheme 15. Synthesis of (+)-andrachcinidine.
Molecules 16 07815 g019

3.2. Cycloisomerization on Alkynediol-Based Systems

Ketals are important key structures, and crucial targets in organic synthesis [94,95,96,97,98]. Fused, bicyclic and spiroketals are recurring motifs in natural compounds, and their preparation is a key step in many total syntheses. In particular, spiroketals represent a structural feature of many biomedically relevant natural and non-natural systems [99,100,101,102]. Several methods have been developed for the synthesis of spiroketals, the most common being perhaps the cyclocondensation of ketone diols [103,104]. Nevertheless, gold catalyzed cycloisomerization on alkynediols has emerged as an efficient strategy to build complex ketal systems in just one step, offering specific advantages. For example, Au-catalyzed cycloisomerization of alkynediols are more exotermic, atom economical, and more compatible than ketones under a number of several reaction conditions. Thus, many groups have recently incorporated the present methodology for the synthesis of several natural compounds and derivatives [105,106,107,108].
Li et al. have described the preparation of the bisbenzannelated spiroketal core of rubromycins [109]. These natural occurring structures exhibit different biological activities, such as inhibition of DNA polymerase, inhibition of the reverse transcriptase of HIV I, or inhibition of DNA helicase [110,111,112,113]. Scheme 16 shows the basic structure motif shared by natural isolated compounds like γ-rubromycin, purpuromycin, or heliquinomycin. According to the described work, easily prepared alkynediols 49 underwent cycloisomerization in the presence of gold catalysis to yield spiroketals 50 with moderate yields, but mainly together with notable amounts of the corresponding benzofuran 51.
Scheme 16. Synthesis of spiroketal motif of rubromycins.
Scheme 16. Synthesis of spiroketal motif of rubromycins.
Molecules 16 07815 g020
A more effective spiroketalization process was found for the synthesis of cephalosporolides. Concretely, cephalosporolide H 52 is a natural spiroketal isolated from the culture broth of the marine fungus Penicillium sp. This compound presents anti-inflammatory properties by virtue of its inhibitory activity against 3α-hydroxysteroid dehydrogenase [114,115]. Dudley et al. developed a method for cephalosporolide total synthesis based on gold-catalyzed spiroketal generation [116,117]. Starting from pantolactone 53, alkynediol-based system 54 was obtained after a nine step sequence. Gold treatment of 54 yielded the desired structure 55, with an excellent 88% yield. The main inconvenient of the proposed strategy lied on the obtention of 55 as a 1:1 mixture of spiroketal epimers, although further treatment upon zinc chloride chelation provided the expected isomer in 20:1 dr (Scheme 17).
Scheme 17. Cephalosporolide H; structure and proposed synthesis.
Scheme 17. Cephalosporolide H; structure and proposed synthesis.
Molecules 16 07815 g021
Azaspiracid 56 belongs to a family of marine toxins, responsible for human poisoning and diverse chronic effects on liver, pancreas and thymus [118,119]. Its complete structure has been widely studied [120], and several methods for its synthesis have been reported [121,122]. Forsyth et al. have reported the synthesis of the F−I azaspiracid fragment 57 [123]. In particular, we would like to focus on the construction of F and G rings by a one step gold-catalyzed spiroketalization. Alkynediol-based system 58 was obtained by coupling of subunits 59 and 60, prepared from simple precursors. Treatment of 58 with AuCl provided the desired structure 57 with a high 75% yield as a sole isomer. The reaction mechanism is proposed to follow an initial syn addition of the C6 hydroxy group and the π-activated gold-alkyne complex to build ring F. Protodeauration and protonation of the resultant enol ether at C11 would promote the attack of methoxy oxygen to C10, generating therefore ring G (Scheme 18).
Okadaic acid 61 is a complex natural structure isolated from marine sponges [124,125]. Its biological activities [126,127,128], together with its attractive chemical structure have attracted much interest among organic chemists. In particular, the presence of several spiroketal motifs in this structure makes it a real challenge from the retrosynthetical point of view. An efficient synthesis of the C15-C38 fragment has been reported, based on the high activity and selectivity of AuCl for the synthesis of spiroketals 62 and 63, starting from alkynediols 64 and 66 respectively [129] (Scheme 19).
Bridged-bicyclic ketals have been also produced through gold-catalyzed cycloisomerization of alkynediols. Based on platensimycin structure, a natural inhibitor of microbial fatty acid biosynthesis [130,131,132], Corey et al. reported the total synthesis of the near-structural mimic 68 [133]. This new structure presents evidence in the literature suggesting excellent antimicrobial properties [134,135]. Thus, easily achieved alkynediol 69 reacted under gold(III) catalysis delivering ketone 70, which contains the tricyclic core of 68, with an excellent 85% yield and >98% ee (Scheme 20). The route to the desired target is completed in just nine steps, providing a facile and quick methodology to the mentioned bioactive structure.
Scheme 18. Synthesis of rings F and G in azaspiracid and reaction mechanism.
Scheme 18. Synthesis of rings F and G in azaspiracid and reaction mechanism.
Molecules 16 07815 g022
Scheme 19. Synthesis of C15-C27 and C28-C38 fragments of okadaic acid.
Scheme 19. Synthesis of C15-C27 and C28-C38 fragments of okadaic acid.
Molecules 16 07815 g023
Scheme 20. Synthesis of platensimycin-derived structure.
Scheme 20. Synthesis of platensimycin-derived structure.
Molecules 16 07815 g024

4. Conclusions

In this overview we have collected the most recent advances in gold-catalyzed cycloisomerization of alkynol and alkynediol-based systems for the preparation of natural products and derivatives. This type of process has become an established methodology for accessing a large number of both carbocyclic and heterocyclic structures, containing different sized skeletons. Three to seven-membered carbon rings, such as furan, pyrans, piperidines, and different ketal and spiroketal systems are therefore accessible through this strategy. The reactions discussed herein demonstrate the high synthetic potential of alkynol-based compounds undergoing gold catalyzed cyclization. On the other hand, the efficiency of gold salts and gold complexes have been also documented, allowing mild reaction conditions and great functional group compatibility, specially compared to related thermal or basic rearrangements. In addition, the extremely large number of natural bioactive compounds containing these type of structural motifs, readily available through gold-catalyzed conditions, will certainly provide a renewed and continuous topic of investigation in this field.

Acknowledgments

Support for this work by the DGI-MICINN (Project CTQ2009-09318), Conunidad Autónoma de Madrid (CAM) (Project S2009/PPQ-1752), and SANTANDER-UCM (Project GR35/10-A) are gratefully acknowledged. J. M. A. thanks CAM and Fondo Social Europeo for a postdoctoral contract.

Conflict of Interest

The authors declare no conflict of interest.

References

  1. Hashmi, A.S.K.; Rudolph, M. Gold Catalysis in Total Synthesis. Chem. Soc. Rev. 2008, 37, 1766–1775. [Google Scholar] [CrossRef]
  2. Fürstner, A. Gold and Platinum Catalysis-A Convenient Tool for Generating Molecular Complexity. Chem. Soc. Rev. 2009, 28, 3208–3221. [Google Scholar] [CrossRef]
  3. Alcaide, B.; Almendros, P.; Alonso, J.M. Gold Catalyzed Oxycyclizations of Alkynols and Alkynediols. Org. Biomol. Chem. 2011, 9, 4405–4416. [Google Scholar] [CrossRef]
  4. Zeni, G.; Larock, R.C. Synthesis of Heterocycles via Palladium n-Olefin and n-Alkyne Chemistry. Chem. Rev. 2004, 104, 2285–2309. [Google Scholar] [CrossRef]
  5. Alonso, F.; Beletskaya, I.P.; Yus, M. Transition-Metal-Catalyzed Addition of Heteroatom-Hydrogen Bonds to Alkynes. Chem. Rev. 2004, 104, 3079–3159. [Google Scholar] [CrossRef]
  6. Michael, P.; Gerd, R.; Theo, S. (Eli Lilly Co.) Imidazoline Derivatives for the Treatment of Diabetes, Especially Type II Diabetes. WO Patent 2000078726 A1, 28 December 2000. [Google Scholar]
  7. Tagami, K.; Yoshimura, H.; Nagai, M.; Hibi, S.; Kikuchi, K.; Sato, T.; Okita, M.; Okamoto, Y.; Nagasaka, Y.; Kobayashi, N.; Hida, T.; Tai, K.; Tokuhara, N.; Kobayashi, S. (Eisai Co.,Ltd.) Preparation of Fused-ring Carboxylic Acid Compounds as Retinoic Acid Receptor Agonists. WO Patent 9734869 A1, 25 September 1997. [Google Scholar]
  8. Masaki, S.; Mitsunori, K.; Yuhei, M.; Masakuni, K. (Takeda Pharmaceutical) Amide Compound. WO Patent 2010018874 A1, 18 February 2010. [Google Scholar]
  9. Hashmi, A.S.K.; Yang, W.; Rominger, F. Gold(I)-Catalyzed Formation of Benzo[b]furans from 3-Silyloxy-1,5-enynes. Angew. Chem. Int. Ed. 2011, 50, 5762–5765. [Google Scholar] [CrossRef]
  10. Henkel, T.; Brunne, R.M.; Reichel, F.; Muller, H. Statiscal Investigation into the Structural Complementarity of Natural Products and Synthetic Compounds. Angew. Chem. Int. Ed. 1999, 38, 643–647. [Google Scholar] [CrossRef]
  11. Pearce, C. Biologically Active Fungal Metabolites. Adv. Appl. Microbiol. 1997, 44, 1–80. [Google Scholar] [CrossRef]
  12. For isolation and structural determination of guanacastepene, see: Brady, S.F.; Singh, M.P.; Janso, J.E.; Clardy, J. Guanacastepene, a Fungal-Derived Diterpene Antibiotic with a New Carbon Skeleto. J. Am. Chem. Soc. 2000, 122, 2116–2117. [Google Scholar] [CrossRef]
  13. For selected studies about the mentioned pathogens, see: Neu, H.C. The Crisis in Antibiotic Resistance. Science 1992, 257, 1064–1073. [Google Scholar]
  14. Swartz, M.N. Hospital-Acquired Infections: Diseases with Increasingly Limited Therapies. Proc. Natl. Acad. Sci. USA 1994, 91, 2420–2427. [Google Scholar] [CrossRef]
  15. Gagosz, F. Unusual Gold(I)-Catalyzed Isomerization of 3-Hydroxylated 1,5-Enynes: Highly Substrate-Dependent Reaction Manifolds. Org. Lett. 2005, 7, 4129–4132. [Google Scholar] [CrossRef]
  16. Ohloff, G.; Strickler, H.; Willhalm, B.; Borer, C.; Hinder, M. En-Syntheses with Singlet Oxygen. II. Dye-Sensitized Photooxygenation of (−)-Thujopsene and Stereochemistry of the Prepared Thujopsanols. Helv. Chim. Acta 1970, 53, 623–637. [Google Scholar] [CrossRef]
  17. Hatsui, T.; Suzuki, N.; Takeshita, H. Dicyanoanthracene-Sensitized Photooxygenation of Thujopsene. Chem. Lett. 1985, 639–642. [Google Scholar]
  18. Fehr, C.; Vuagnoux, M.; Buzas, A.; Arpagaus, J.; Sommer, H. Gold- and Copper-Catalyzed Cycloisomerizations towards the Synthesis of Thujopsanone-Like Compounds. Chem. Eur. J. 2011, 17, 6214–6220. [Google Scholar] [CrossRef]
  19. Graening, T.; Schmalz, H.-G. Total Synthesis of Colchicine in Comparison: A Journey through 50 Years of Synthetic Organic Chemistry. Angew. Chem. Int. Ed. 2004, 43, 3230–3256. [Google Scholar] [CrossRef]
  20. Boyé, O.; Brossi, A. The Alkaloids; Brossi, A., Cordell, G.A., Eds.; Academic Press: New York, NY, USA, 1992; Volume 41, p. 125. [Google Scholar]
  21. Jordan, M.A.; Wilson, L. Microtubules as a Target for Anticancer Drugs. Nat. Rev. Cancer 2004, 4, 253–265. [Google Scholar] [CrossRef]
  22. Vorogushin, A.V.; Wulff, W.D.; Hansen, H.-J. Central-to-Axial Chirality Transfer in the Benzannulation Reaction of Optically Pure Fischer Carbene Complexes in the Synthesis of Allocolchicinoids. Tetrahedron 2008, 64, 949–968. [Google Scholar] [CrossRef]
  23. Besong, G.; Jarowski, K.; Kocienski, P.J.; Sliwinski, E.; Boyle, F.T. Synthesis of (S)-(−)-N-Acetylcolchinol Using Intramolecular Biaryl Oxidative Coupling. Org. Biomol. Chem. 2006, 4, 2193–2207. [Google Scholar] [CrossRef]
  24. Leblanc, M.; Fagnou, K. Allocolchicinoid Synthesis via Direct Arylation. Org. Lett. 2005, 7, 2849–2852. [Google Scholar] [CrossRef]
  25. Büttner, F.; Bergemann, S.; Guénard, D.; Gust, R.; Seitz, G.; Thoret, S. Two Novel Series of Allocolchicinoids with Modified Seven Membered B-Rings: Design, Synthesis, Inhibition of Tubulin Assembly and Citotoxicity. Bioorg.Med. Chem. 2005, 13, 3497–3511. [Google Scholar] [CrossRef]
  26. Wu, T.R.; Chong, J.M. Asymmetric Synthesis of Propargylamines via 3,3’-Disubstituted Binaphtol-Modified Alkynylboronates. Org. Lett. 2006, 8, 15–18. [Google Scholar] [CrossRef]
  27. Vorogushin, A.V.; Predeus, A.V.; Wulff, W.D.; Hansen, H.-J. Diels-Alder Reaction-Aromatization Approach towards Functionalized Ring C Allocolchicinoids. Enantioselective Total Synthesis of (−)-7S-Allocolchicine. J. Org. Chem. 2003, 64, 5826–5831. [Google Scholar]
  28. For a reported synthesis of 16 see:Boyer, F.-D.; Hanna, I. Synthesis of Allocolchicines Using Sequential Ring-Closing Enyne Methathesis-Diels-Alder Reactions. Org. Lett. 2007, 9, 715–718. [Google Scholar] [CrossRef]
  29. Boyer, F.-D.; Le Goff, X.; Hanna, I. Gold(I)-Catalyzed Cycloisomerization of 1,7- and 1,8-Enynes: Application to the Synthesis of a New Allocolchicinoid. J. Org. Chem. 2008, 74, 5163–5166. [Google Scholar]
  30. Carbone, M.; Yan, L.; Irace, C.; Mollo, E.; Castelluccio, F.; Di Pascale, A.; Cimino, G.; Santamaria, R.; Guo, Y.-W.; Gavagnin, M. Structure and Cytotoxicity of Phidianidines A and B: First Finding of 1,2,4-Oxadiazole System in a Marine Natural Product. Org. Lett. 2011, 13, 2516–2519. [Google Scholar]
  31. Sunderhaus, J.D.; Sherman, D.H.; Williams, R.M. Studies on the Biosynthesis of the Stephadin and Notoamide Natural Products: A Stereochemical and Genetic Coundrum. Israel J. Chem. 2011, 51, 442–452. [Google Scholar] [CrossRef]
  32. Wu, M.; Wu, P.; Xie, H.; Wu, G.; Wei, X. Monoterpenoid Indole Alkaloid Mediating DNA Strand Scission from Turpina Arguta. Planta Medica 2011, 77, 284–286. [Google Scholar] [CrossRef]
  33. Ozcelik, B.; Kartal, M.; Orhan, I. Cytotoxicity, Antiviral and Antimicrobial Activities of Alkaloids, Flavonoids and Phenolic Acid. Pharm. Biol. 2011, 49, 396–402. [Google Scholar] [CrossRef]
  34. Yap, W.-S.; Gan, C.-Y.; Low, Y.-Y.; Choo, Y.-M.; Etoh, T.; Hayashi, M.; Komiyama, K.; Kam, T.-S. Grandilodines A-C, Biologically Active Indole Alkaloids from Kopsia Grandifolia. J. Nat. Prod. 2011, 74, 1309–1312. [Google Scholar] [CrossRef]
  35. Chung, Y.-M.; Lan, Y.-H.; Hwang, T.-L.; Leu, Y.-L. Anti-Inflammatory and Antioxidant Components from Hygroyza Aristata. Molecules 2011, 16, 1917–1927. [Google Scholar] [CrossRef]
  36. Foo, K.; Newhouse, T.; Mori, I.; Takayama, H.; Baran, P.S. Total Synthesis Guided Structure Elucidation of (+)-Psychotetramine. Angew. Chem. Int. Ed. 2011, 50, 2716–2719. [Google Scholar] [CrossRef]
  37. Finlayson, R.; Pearce, A.; Norrie, A.; Page, M.J.; Kaiser, M.; Bourguet-Kondracki, M.-L.; Harper, J.; Webb, V.; Copp, B. Didemnidines A and B, Indole Spermidine Alkaloids from the New Zealand Ascidian Didemnum sp. J. Nat. Prod. 2011, 74, 888–892. [Google Scholar] [CrossRef]
  38. Nakadate, S.; Nozawa, K.; Horie, H.; Fuji, Y.; Yaguchi, T. New Type Indole Diterpene, Eujindoles, From Eupenicillium Javanicum. Heterocycles 2011, 83, 351–356. [Google Scholar] [CrossRef]
  39. Ruiz-Sanchis, P.; Savina, S.; Albericio, F.; Alvarez, M. Structure, Bioactivity and Synthesis of Natural Products with Hexahydropyrrolo[2,3-b]Indole. Chem. Eur. J. 2011, 17, 1388–1408. [Google Scholar] [CrossRef]
  40. Yamada, Y.; Kitajima, M.; Kogure, N.; Wongseripipatana, S.; Takayama, H. Seven New Monoterpenoid Indole Alkaloids from Gelsemium Elegans. Chem. Asian J. 2011, 6, 166–173. [Google Scholar] [CrossRef]
  41. Palmisano, G.; Penoni, A.; Sisti, M.; Tiblietti, F.; Tollari, S.; Nicholas, K. Synthesis of Indole Derivatives with Biological Activity by Reactions between Unsaturated Hydrocarbons and N-Aromatic Precursors. Curr. Org. Chem. 2010, 14, 2409–2441. [Google Scholar] [CrossRef]
  42. Nakajima, T.O.; Satoshi, Y.; Fukuyama, T. Total Synthesis of (−)-Mersicarpine. J. Am. Chem. Soc. 2010, 132, 1236–1237. [Google Scholar] [CrossRef]
  43. Zhang, L. Tandem Au-catalyzed 3,3-Rearrangement-[2+2] Cycloadditions of Propargylic Esters: Expeditious Access to Highly Functionalized 2,3-Indoline-Fused Cyclobutanes. J. Am. Chem. Soc. 2005, 127, 16804–16805. [Google Scholar] [CrossRef]
  44. Ferrer, C.; Amijs, C.H.M.; Echavarren, A.M. Intra- and Intermolecular Reactions of Indoles with Alkynes Catalyzed by Gold. Chem. Eur. J. 2007, 13, 1358–1373. [Google Scholar] [CrossRef]
  45. Lu, Y.; Du, X.; Jia, X.; Liu, Y. Gold Catalyzed Intermolecular Reactions of (Z)-Enynols with Indoles for the Construction of Dihydrocyclohepta[b]indole Skeletons through a Cascade Friedel-Crafts/Hydroarylation Sequence. Adv. Syth. Catal. 2009, 351, 1517–1522. [Google Scholar] [CrossRef]
  46. Raffa, G.; Belot, S.; Balme, G.; Monteiro, N. Iodocyclization versus Diiodination in the Reaction of 3-Alkynyl-4-methoxycoumarins with Iodine: Synthesis of 3-Iodofuro[2,3-b]chromones. Org. Biomol. Chem. 2011, 9, 1474–1478. [Google Scholar] [CrossRef]
  47. Wang, L.; Peng, S.; Wang, J. Palladium-Catalyzed Cascade Reactions of Coumarins with Alkynes: Synthesis of Highly Substituted Cyclopentadiene Fused Chromones. Chem. Commun. 2011, 47, 5422–5424. [Google Scholar] [CrossRef]
  48. Zhao, J.; Zhao, Y.; Fu, H. Transition-Metal-Free Intramolecular Ullmann-Type O-Arylation: Synthesis of Chromone Derivatives. Angew. Chem. Int. Ed. 2011, 50, 3769–3773. [Google Scholar] [CrossRef]
  49. Renault, J.; Qian, Z.; Uriac, P.; Goualt, N. Electrophilic Carbon Transfer in Gold Catalysis: Synthesis of Substituted Chromones. Tetrahedron Lett. 2011, 52, 2476–2479. [Google Scholar] [CrossRef]
  50. For a review, see: Boumendjel, A. Aurones: A Subclass of Flavone with Promising Biological Potential. Curr. Med. Chem. 2003, 10, 2621–2330. [Google Scholar] [CrossRef]
  51. Brooks, C.J.; Watson, D.G. Phytoalexins. Nat. Prod. Rep. 1985, 427–459. [Google Scholar]
  52. Morimoto, M.; Fukumoto, H.; Nozoe, T.; Hagiwara, A.; Komai, K. Synthesis and Insect Antifeedant Activity of Aurones against Spodoptera Litura Larvae. J. Agric. Food. Chem. 2007, 55, 700–705. [Google Scholar] [CrossRef]
  53. Okombi, S.; Rival, D.; Bonnet, S.; Mariotte, A.-M.; Perrier, E.; Boumendjel, A. Discovery of Benzylidenezofuran-3(2H)-one (Aurones) as Inhibitors of Tyrosinase Derived from Human Melanocytes. J. Med. Chem. 2006, 49, 329–333. [Google Scholar] [CrossRef]
  54. Venkateswarlu, S.; Panchagnula, G.K.; Subbaraju, G.V. Synthesis and Antioxidative Activity of 3’,4’,6,7-tetrahydroxyaurone, a Metabollite of Biden Frondosa. Biosci. Biotechnol. Biochem. 2004, 68, 2183–2185. [Google Scholar] [CrossRef]
  55. Auf’mkolk, M.; Koerhle, J.; Hesch, R.D.; Cody, V. Inhibition of Rat Liver Iodothyronine Deiodinase. Interaction of Aurones with the Iodothyronine Ligand-Binding Site. Biol. Chem. 1986, 261, 11623–11630. [Google Scholar]
  56. Donnelly, J.A.; Fox, M.J.; Sharma, T.C. α-Halo Ketones. XI Generation of the Wheeler Aurone Synthesis. Tetrahedron 1979, 35, 875–879. [Google Scholar] [CrossRef]
  57. Bose, G.; Mondal, E.; Khan, A.T.; Bordoloi, M.J. An Environmentally Benign Synthesis of Aurones and Flavones from 2’-Acetoxychalcones using n-Tetrabutylammonium Tribromide. Tetrahedron Lett. 2001, 42, 8907–8909. [Google Scholar] [CrossRef]
  58. Lévai, A.; Tökés, A.L. Synthesis of Aurones by the Oxidative Rearrangement of 2’-Hydroxychalcones with Thallium(III) Nitrate. Synth. Commum. 1982, 12, 701–707. [Google Scholar]
  59. Imafuku, K.; Honda, M.; McOmie, J.F.W. Cyclodehydrogenation of 2’-Hydroxychalcones with 2,3-Dichloro-5,6-dicyano-p-benzoquinone: A Simple Route for Flavones and Aurones. Synthesis 1987, 199–201. [Google Scholar]
  60. Sekizaki, H. Synthesis of 2-Benzylidene-3(2H)-benzofuran-3-ones (Aurones) by Oxidation of 2’-Hydroxychalcones with Mercury(II) Acetate. Bull. Chem. Soc. Jpn. 1988, 61, 1407–1409. [Google Scholar] [CrossRef]
  61. Thakkar, K.; Cushman, M. A Novel Oxidative Cyclization of 2’-Hydroxychalcones to 4,5-Dialkoxyaurones by Thallium(III) Nitrate. J. Org. Chem. 1995, 60, 6499–6510. [Google Scholar] [CrossRef]
  62. An, Z.-W.; Catellani, M.; Chiusoli, G.P. Palladium-catalyzed Synthesis of Aurone from Salicyloyl Chloride and Phenylacetylene. J. Organomet. Chem. 1990, 397, 371–373. [Google Scholar] [CrossRef]
  63. Harkat, H.; Blanc, A.; Weibel, J.-M.; Pale, P. Versatile and Expeditious Synthesis of Aurones via AuI-Catalyzed Cyclization. J. Org. Chem. 2008, 73, 1620–1623. [Google Scholar] [CrossRef]
  64. Atta-ur-Rahman; Choudhary, M.I.; Hayat, S.; Khan, A.; Ahmed, A. Two New Aurones from Marine Brown Alga Spatoglossum Variable. Chem. Pharm. Bull. 2001, 49, 105–107. [Google Scholar] [CrossRef]
  65. Kobayashi, S.; Miyase, T.; Noguchi, H. Polyphenolic Glycosides and Oligosacharide Multiesters from the Roots of Polygala Dalmaisiana. J. Nat. Prod. 2002, 65, 319–328. [Google Scholar] [CrossRef]
  66. Trost, B.M.; Dong, G. Total Synthesis of Bryostatin 16 Using a Pd-Catalyzed Diyne Coupling as Macrocyclization Method and Synthesis of C20-epi-Bryostatin 7 as a Potent Anticancer Agent. J. Am. Chem. Soc. 2010, 132, 16403–16416. [Google Scholar] [CrossRef]
  67. Pettit, G.R.; Gao, F.; Blumberg, P.M.; Herald, C.L.; Coll, J.C.; Kamano, Y.; Lewin, N.E.; Schmidt, J.M.; Chapuis, J.-C. Antineoplastic Agents. 340. Isolation and Structural Elucidation of Bryostatins 16-18. J. Nat. Prod. 1996, 59 For more information of bryostatin 16, see:, 286–289. [Google Scholar] [CrossRef]
  68. Hale, K.J.; Hummersome, M.G.; Manaviazar, S.; Frigerio, M. The Chemistry and Biology of the Bryostatin Antitumour Macrolides. Nat. Prod. Rep. 2002, 19, 413–453. [Google Scholar] [CrossRef]
  69. Newman, D.J.; Cragg, G.M. Marine Natural Products and related Compounds in Clinical and Advanced Preclinical Trials. J. Nat. Prod. 2004, 67, 1216–1238. [Google Scholar] [CrossRef]
  70. Hale, K.J.; Manaviazar, S. New Approaches to the Total Synthesis of Bryostatin Antitumor Macrolides. Chem. Asian J. 2010, 5, 704–754. [Google Scholar] [CrossRef]
  71. Manaviazar, S.; Hale, K.J. Total Synthesis of Bryostatin 1: A Short Route. Angew. Chem. Int. Ed. 2011. [Google Scholar] [CrossRef]
  72. Serrano, M. The Tumour Suppressor Protein p16INK4a. Exp. Cell. Res. 1997, 237, 7–13. [Google Scholar] [CrossRef]
  73. Pettit, G.R.; Inoue, M.; Kamano, Y.; Herald, D.L.; Arm, C.; Dufresne, C.; Christie, N.D.; Schmidt, J.M.; Doubek, D.L.; Krupa, T.S. Antineoplastic Agent. 174. Isolation and Styructure of the Cytostatic Depsipeptide Dolastatin 13 from the Sea Hare Dolabella Auricularia. J. Am. Chem. Soc. 1989, 110, 2006–2007. [Google Scholar]
  74. Jeong, J.U.; Sutton, S.C.; Kim, S.; Fuchs, P.L. Biomimetic Total Syntheses of (+)-Cephalostatin 12, and (+)-Ritterazine K. J. Am. Chem. Soc. 1995, 117, 10157–10158. [Google Scholar] [CrossRef]
  75. LaCour, T.G.; Guo, C.; Bhandatu, S.; Fuchs, P.L.; Boyd, M.R. Interphylal Product Splicing: The First Total Syntheses of Cephalostatin 1, the North Hemisphere of Ritterazine G, and the High Active Hybrid Analog, Ritterostatin GN1N. J. Am. Chem. Soc. 1998, 120, 692–707. [Google Scholar]
  76. Jeong, J.U.; Guo, C.; Fuchs, P.L. Synthesis of the South Unit of Cephalostatin. 7. Total Syntheses of (+)-Cephalostatin 7, (+)-Cephalostatin 12, and (+)-Ritterazine K. J. Am. Chem. Soc. 1999, 121, 2071–2084. [Google Scholar] [CrossRef]
  77. Kim, S.; Sutton, S.C.; Guo, C.; LaCour, T.G.; Fuchs, P.L. Synthesis of the North 1 Unit of the Cephalostatin Family from Hecogenin Acetate. J. Am. Chem. Soc. 1999, 121, 2056–2070. [Google Scholar]
  78. Lee, S.; Fuchs, P.L. The First Total Synthesis of (Corrected) Ritterazine M. Org. Lett. 2002, 4, 317–318. [Google Scholar] [CrossRef]
  79. Fortner, K.C.; Kato, D.; Tanaka, Y.; Shair, M.D. Enantioselective Synthesis of (+)-Cephalostatin 1. J. Am. Chem. Soc. 2010, 132, 275–280. [Google Scholar] [CrossRef]
  80. Ueda, M.; Sato, A.; Ikeda, Y.; Miyoshi, T.; Naito, T.; Miyata, O. Direct Synthesis of Trisubstituted Isoxazoles through Gold-Catalyzed Domino Reaction of Alkynyl Oxime Ethers. Org. Lett. 2010, 11, 2594–2597. [Google Scholar]
  81. He, W.; Li, C.; Zhang, L. An Efficient [2+2+1] Synthesis of 2,5-Disubstututed Oxazoles via Gold-Catalyzed Intermolecular Alkyne Oxidation. J. Am. Chem. Soc. 2011, 133, 8482–8485. [Google Scholar] [CrossRef]
  82. Hashmi, A.S.K.; Schuster, A.M.; Zimmer, M.; Rominger, F. Synthesis of 5-Halo-4H-1,3-oxazine-6-amines by Copper-Mediated Domino Reaction. Chem. Eur. J. 2011, 17, 5511–5515. [Google Scholar] [CrossRef]
  83. Gao, X.; Pan, Y.-M.; Li, M.L.; Chen, L.; Zhan, Z.-P. Facile One-Pot Synthesis of Three Different Substituted Thiazoles from Propargylic Alcohols. Org. Biomol. Chem. 2010, 8, 3259–3266. [Google Scholar] [CrossRef]
  84. Yoshimatsu, M.; Matsui, M.; Yamamoto, T.; Sawa, A. Convinient Preparation of 4-Arylmethyl- and 4-Hetarylmethyl Thiazoles by Regioselective Cycloaddition Reactions of 3-Sulfanyl- and Selenylpropargyl Alcohols. Tetrahedron Lett. 2010, 66, 7975–7987. [Google Scholar]
  85. Asanuma, Y.; Fujiwara, S.-I.; Shin-Ike, T.; Kambe, N. Selenoimidoylation of Alcohols with Selenium and Isocyanides and its Application to the Synthesis of Selenium-Containing Heterocycles. J. Org. Chem. 2004, 69, 4845–4848. [Google Scholar] [CrossRef]
  86. Wilckens, K.; Uhlemann, M.; Czekelius, C. Gold-Catalyzed endo-Cyclizations of 1,4-Diynes to Seven-Membered Ring Heterocycles. Chem. Eur. J. 2009, 15, 13323–13326. [Google Scholar]
  87. Yan, B.; Zhou, Y.; Zhang, H.; Chen, J.; Liu, Y. Highly Efficient Synthesis of Functionalized Indolizines and Indolizinones by Copper-Calayzed Cycloisomerization of Propargylic Pyridines. J. Org. Chem. 2007, 72, 7783–7786. [Google Scholar] [CrossRef]
  88. Bunnelle, E.M.; Smith, C.R.; Lee, S.K.; Singaram, S.W.; Rhodes, A.J.; Sarpong, R. Pt-Catalyzed Cyclization/Migration of Propargylic Alcohols for the Synthesis of 3(2H)-Furanones, Pyrrolones, Indolizines, and Indolizinones. Tetrahedron Lett. 2008, 64, 7008–7014. [Google Scholar]
  89. For some examples of non-metal catalyzed cycloisomerizations see: Ji, K.-G.; Zhu, H.-T.; Yang, F.; Shu, X.-Z.; Zhao, S.-C.; Liu, X.-Y.; Shaukat, A.; Liang, Y.-M. A Novel Iodine-Promoted Tandem Cyclization: An Efficient Synthesis of Substituted 3,4-Diiodoheterocyclic Compounds. Chem. Eur. J. 2010, 16, 6151–6154. [Google Scholar] [CrossRef]
  90. Zang, X.; Teo, W.T.; Chan, S.W.H.; Chan, P.W.H. Brønsted Acid Catalyzed Cyclization of Propargylic Alcohols wiyh Thioamides. Facile Synthesis of Di- and Trisubstituted Thiazoles. J. Org. Chem. 2010, 75, 6290–6293. [Google Scholar] [CrossRef]
  91. Kothandaraman, P.; Rao, W.; Foo, S.J.; Chan, P.W.H. Gold-Catalyzed Cycloisomerization Reaction of 2-Tosylamino-phenylprop-1-yn-3-ols via a Versatile Approach for Indole Synthesis. Angew. Chem. Int. Ed. 2010, 49, 4619–4623. [Google Scholar] [CrossRef]
  92. Jung, H.H.; Floreancig, P.E. Gold-Catalyzed Synthesis of Oxygen- and Nitrogen-Containing Heterocycles from Alkynyl Ethers: Application to the Total Synthesis of Andrachcinidine. J. Org. Chem. 2007, 72, 7359–7366. [Google Scholar] [CrossRef]
  93. Mill, S.; Hootelé, C. Alkaloids of Andrachne Aspera. J. Nat. Prod. 2000, 63, 762–764. [Google Scholar] [CrossRef]
  94. Forsyth, C.J. Asymmetric Synthesis, 2nd; Christmann, M., Braese, S., Eds.; Wiley-VCH Verlag Gmbh & Co.: Weinheim, Germany, 2008; pp. 271–276. [Google Scholar]
  95. Ballini, R.; Petrini, M. Nitroalkanes as Key Building Blocks for the Synthesis of Heterocyclic Derivatives. ARKIVOC 2008, 9, 195–223. [Google Scholar]
  96. Forsyth, C.J. Asymmetric Synthesis; Christmann, M., Braese, S., Eds.; Wiley-VCH Verlag Gmbh & Co.: Weinheim, Germany, 2007; pp. 256–261. [Google Scholar]
  97. Sherry, B.D.; Maus, L.; Laforteza, B.N.; Toste, D.F. Gold(I)-Catalyzed Synthesis of Dihydropyrans. J. Am. Chem. Soc. 2006, 128, 8132–8133. [Google Scholar] [CrossRef]
  98. Alcaide, B.; Almendros, P.; Carrascosa, R.; Torres, M.R. Gold/Acid-Cocatalyzed Regiodivergent Preparation of Bridged Ketals via Direct Bis-Oxycyclization of Alkynic Acetonides. Adv. Synth. Catal. 2010, 352, 1277–1283. [Google Scholar] [CrossRef]
  99. Perron, F.; Albizati, K.F. Chemistry of Spiroketals. Chem. Rev. 1989, 89, 1617–1661. [Google Scholar] [CrossRef]
  100. Aho, J.E.; Pihko, P.M.; Rissa, T.K. Nanometric Spiroketals in Natural Products: Structures, Sources, and Synthetic Strategies. Chem. Rev. 2005, 105, 4406–4440. [Google Scholar] [CrossRef]
  101. Rama Raju, B.; Saikia, A.K. Asymmetric Synthesis of Naturally Ocurring Spiroketals. Molecules 2008, 13, 1942–2038. [Google Scholar] [CrossRef]
  102. Blunt, J.W.; Copp, B.R.; Hu, W.P.; Munro, M.H.G.; Northcote, P.T.; Prinsep, M.R. Marine Natural Products. Nat. Prod. Rep. 2009, 26, 170–244. [Google Scholar] [CrossRef]
  103. Brimble, M.A.; Bryant, C.J. Synthesis of the Spiroketal-Containing Anti-Helicobacter Pylori Agents CJ-12,954 and CJ-13,014. Chem. Commun. 2006, 4506–4508. [Google Scholar] [CrossRef]
  104. Izquierdo-Cubero, I.; Plaza Lopez-Espinosa, M.T.; Kari, N. Synthesis of Optically Active Chalcogran from L-Sorbose. Carbohydr. Res. 1994, 261, 231–242. [Google Scholar] [CrossRef]
  105. Antoniotti, S.; Genin, E.; Michelet, V.; Genêt, J.-P. Highly Efficient Access to Strained Bicyclic Ketals via Gold-catalyzed Cycloisomerization of Bis-Homopropargylic Diols. J. Am. Chem. Soc. 2005, 127, 9976–9977. [Google Scholar] [CrossRef]
  106. Liu, L.-P.; Hammond, G.B. Highly Efficient and Tunable Synthesis of Dioxabicyclo[4.2.1] Ketals and Tetrahydropyrans via Gold-Catalyzed Cycloisomerization of 2-Alkynyl-1,5-diols. Org. Lett 2009, 11, 5090–5092. [Google Scholar] [CrossRef]
  107. Liu, B.; De Brabander, J.K. Metal-catalyzed Regioselective Oxy-Functionalization of Internal Alkynes: An Entry into Ketones, Acetals, and Spiroketals. Org. Lett. 2006, 8, 4907–4910. [Google Scholar] [CrossRef]
  108. Hashmi, A.S.K.; Bührle, M.; Wölfe, M.; Rudolph, M.; Wieteck, M.; Rominger, F.; Frey, W. Gold Catalysis: Tandem reactions of Diyne-Diols and External Nuchleophiles as an Easy Access to Tricyclic Cage-Like Structures. Chem. Eur. J. 2010, 16, 9846–9854. [Google Scholar]
  109. Zhang, Y.; Xue, J.; Xin, Z.; Xie, Z.; Li, Y. Gold-Catalyzed Double Intramolecular Alkyne Hydroalkoxylation: Synthesis of the Bisbenzannulated Spiroketal Core of Rubromycins. Synlett 2008, 6, 0940–0944. [Google Scholar]
  110. Brockmann, H.; Lenk, W.; Scwantje, G.; Zeeck, A. Robromycins. Tetrahedron Lett. 1966, 3525–3530. [Google Scholar]
  111. Goldman, M.E.; Salituro, G.S.; Bowen, J.A.; Williamson, J.M.; Zink, L.; Scleif, W.A.; Emini, E.A. Inhibition of Human Immunodefiency Virus-1 reverse Transcriptase Activity by Robromycins: Competitive Interaction at the Template Primer Site. Mol. Pharmacol. 1990, 38, 20. [Google Scholar]
  112. Trani, A.; Dallanoce, C.; Pranzone, G.; Ripamonti, F.; Goldstein, B.P.; Cibiatti, R. Semisynthetic Derivatives of Purpuromycin as Potential Topical Agents for Vaginal Infections. J. Med. Chem. 1997, 40, 967–971. [Google Scholar] [CrossRef]
  113. Chino, M.; Nishikawa, K.; Umekia, M.; Hayashi, C.; Yamazaki, T.; Tsuchida, T.; Sawa, T.; Hamada, M.; Takeuchi, T. Heliquinomycin, A New Inhibitor of DNA Helicase, Produced by Streptomyces sp. MJ929-SF2. Taxonomy, Production, Isolation, Physicochemical Properties and Biological Activities. J. Antibiot. 1996, 49, 752. [Google Scholar] [CrossRef]
  114. Li, X.; Yao, Y.; Zheng, Y.; Satter, I.; Lin, W. Cephalosporolides H and I, Two Novel Lactones from a Marine-Derived Fungus, Penicillium sp. Arch. Phram. Res. 2007, 30, 812–815. [Google Scholar] [CrossRef]
  115. Penning, T.M. Inhibition of 5β-Dihydrocortisone Reduction in Rat Liver Cytosol: A Rapid Spectrophotometric Screen for Nonsteroidal Antiinflammatory Drug Potency. Pharm. Sci. 1985, 74, 651–654. [Google Scholar] [CrossRef]
  116. Sami, F.; Dudley, G.B. Stereocontrol of 5,5-Spiroketals in the Synthesis of Cephalosporolide H Epimers. Org. Lett. 2010, 12, 4698–4701. [Google Scholar] [CrossRef]
  117. Sami, F.; Dudley, G.B. A Gold-Catalyzed Alkyne-Diol Cycloisomerization for the Synthesis of Oxygenated 5,5-Spiroketals. Beilstein J. Org. Chem. 2011, 7, 570–577. [Google Scholar] [CrossRef]
  118. McMahon, I.; Silke, J. Winter Toxicity of Unknown Aetiology in Mussels. Harmful Algae News 1996, 14, 2. [Google Scholar]
  119. Ito, E.; Satake, M.; Ofuji, N.; Kurita, N.; McMahon, T.; James, K.; Yasumoto, T. Multiple Organ Damage Caused by a New Toxin Azaspiracid, Isolated from Mussels Produced in Ireland. Toxicon 2000, 38, 917–930. [Google Scholar] [CrossRef]
  120. Satake, M.; Ofuji, K.; Naoki, H.; James, K.J.; Furey, A.; McMahon, T.; Silke, J.; Yasumoto, T. Azaspiracid, a New Marine Toxin Having Unique Spiro Ring Assemblies, Isolated from Irish Mussels, Myttilus Edulis. J. Am. Chem. Soc. 1998, 120, 9967–9968. [Google Scholar] [CrossRef]
  121. Nicolau, K.C.; Kotifs, T.V.; Vyskocil, S.; Petrovic, G.; Ling, T.; Yamamda, Y.M.A.; Tang, W.; Frederick, M.O. Structural Revision and Total Synthesis of Azaspiracid-1, part 2: Definition of the ABCD Domain and Total Synthesis. Angew. Chem. Int. Ed. 2004, 43, 4318–4324. [Google Scholar]
  122. Inoki, S.; Mukaiyama, T. A Convinient Method for the Stereoselective Preparation of trans-(2-Hydroxymethyl)tetrahydrofurans by the Oxidative Cyclization of 5-Hydroxy-1-alkenes with Molecular Oxygen Catalyzed by Cobalt(II) Complex. Chem. Lett. 1990, 67–70. [Google Scholar]
  123. Li, Y.; Zhou, F.; Forsyth, C.J. Gold(I)-Catalyzed Bis-Spiroketalization: Synthesis of the Trioxadispiroketal-Containing A−D Rings of Azaspiracid. Angew. Chem. Int. Ed. 2007, 46, 279–282. [Google Scholar] [CrossRef]
  124. Tachibana, K.; Scheuer, P.J.; Tsukitamni, Y.; Kikuchi, H.; Van Engen, D.; Clardy, J.; Gopichand, Y.; Schimtz, F.J. Okadaic Acid, A Cytotoxic Ployether from Two Marine Sponges of the Genus Halichondria. J. Am. Chem. Soc. 1981, 103, 2469–2471. [Google Scholar] [CrossRef]
  125. Dounay, A.B.; Forsyth, C.J. Okadaic Acid: The Archetypal Serine/Threonine Protein Phosphatase Inhibitor. Curr. Med. Chem. 2002, 9, 1939–1980. [Google Scholar]
  126. Scheuer, P. Marine Natural Products Research: A Look into the Dive Bag. J. Nat. Prod. 1995, 58, 335–343. [Google Scholar] [CrossRef]
  127. Yasumoto, T.; Murata, M.; Oshima, Y.; Sano, M.; Matsumoto, G.K.; Clardy, J. Diarrhetic Shellfish Toxins. Tetrahedron 1985, 41, 1019–1022. [Google Scholar] [CrossRef]
  128. Bialojan, C.; Takai, A. Inhibitory Effect of Marine-Sponge Toxin, Okadaic Acid, on Protein Phosphatases. Specifity and Kinetics. Biochem. J. 1988, 256, 283–290. [Google Scholar]
  129. Fang, C.; Pang, Y.; Forsyth, C.J. Formal Total Synthesis of Okadaic Acid via Regiocontrolled Gold(I)-Catalyzed Spiroketalizations. Org. Lett. 2010, 12, 4528–4531. [Google Scholar] [CrossRef]
  130. Wang, J.; Soisson, S.M.; Young, K.; Shoop, W.; Kodali, S.; Galgoci, A.; Painter, R.; Parthasarathy, G.; Tang, Y.S.; Cummings, R.; et al. Platensimycin is a Selective FabF Inhibitor with Potent Antibiotic Properties. Nature 2006, 441, 358–361. [Google Scholar]
  131. Singh, S.B.; Jayasuriya, H.; Ondeyka, J.G.; Herath, K.B.; Zhang, C.; Zink, D.L.; Tsou, N.N.; Ball, R.G.; Basilio, A.; Genilloud, O.; Diez, M.T.; Vicente, F.; Pelaez, F.; Young, K.; Wang, J. Isolation, Structure, and Absolute Stereochemistry of Platensimycin, a Broad Spectrum Antibiotic Discovered Using an Antisense Differential Sensitivity Strategy. J. Am. Chem. Soc. 2006, 128, 11916–11920. [Google Scholar]
  132. Häbich, D.; von Nussbaum, F. Platensimycin, A New Antibiotic and “Superbug Challenger” from Nature. ChemMedChem 2006, 1, 951–954. [Google Scholar] [CrossRef]
  133. Yeung, Y.-Y.; Corey, E.J. A Simple, Efficient, and Enantiocontrolled Synthesis of a Near-Structural Mimic of Platensimycin. Org. Lett. 2008, 17, 3877–3878. [Google Scholar] [CrossRef]
  134. Nicolau, K.C.; Tang, Y.; Wang, Y.; Stepan, A.F.; Li, A.; Montero, A. Total Synthesis and Antibacterial Properties of Carbaplatensimycin. J. Am. Chem. Soc. 2007, 129, 14850–14851. [Google Scholar] [CrossRef]
  135. Nicolau, K.C.; Lister, T.; Denton, R.M.; Montero, A.; Edmons, D.J. Adamantaplatensimycin: A Bioactive Analogue of Platensimycin. Angew. Chem. Int. Ed. 2007, 46, 4712–4714. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Alcaide, B.; Almendros, P.; Alonso, J.M. Gold-Catalyzed Cyclizations of Alkynol-Based Compounds: Synthesis of Natural Products and Derivatives. Molecules 2011, 16, 7815-7843. https://doi.org/10.3390/molecules16097815

AMA Style

Alcaide B, Almendros P, Alonso JM. Gold-Catalyzed Cyclizations of Alkynol-Based Compounds: Synthesis of Natural Products and Derivatives. Molecules. 2011; 16(9):7815-7843. https://doi.org/10.3390/molecules16097815

Chicago/Turabian Style

Alcaide, Benito, Pedro Almendros, and José M. Alonso. 2011. "Gold-Catalyzed Cyclizations of Alkynol-Based Compounds: Synthesis of Natural Products and Derivatives" Molecules 16, no. 9: 7815-7843. https://doi.org/10.3390/molecules16097815

Article Metrics

Back to TopTop