Available Online at Www.sciencedirect.com

41
Tetrahedron report number 828 Synthesis of six-membered oxygenated heterocycles through carboneoxygen bond-forming reactions Igor Larrosa a , Pedro Romea b, * , Fe `lix Urpı ´ b, * a School of Biological and Chemical Sciences, Queen Mary, University of London, Mile End Road, London E1 4NS, UK b Departament de Quı ´mica Orga `nica, Universitat de Barcelona, Martı ´ i Franque ´s 1e11, 08028 Barcelona, Catalonia, Spain Received 12 November 2007 Contents 1. Introduction and scope ............................................................................... 2684 2. S N 2-Mediated cyclizations .............................................................................. 2684 3. S N 1-Mediated cyclizations .............................................................................. 2685 4. Epoxide-mediated cyclizations ......................................................................... 2686 5. Alkene-mediated cyclizations .......................................................................... 2690 5.1. Stoichiometric cyclizations ...................................... ................................. 2691 5.1.1. Mercury-mediated cyclizations .............................................................. 2691 5.1.2. Halo-mediated cyclizations ................................................................ 2692 5.1.3. Seleno-mediated cyclizations ............................................................... 2693 5.2. Catalytic cyclizations ........................................................................... 2694 5.2.1. Palladium-mediated cyclizations ................................ ............................ 2694 5.2.1.1. Catalysis by Pd(0) ............................................................... 2694 5.2.1.2. Catalysis by Pd(II) .............................................................. 2695 5.2.2. Other metal-mediated cyclizations ............................... ........................... 2699 5.3. Acid-mediated cyclizations ...................................... ................................. 2699 6. Allene-mediated cyclizations ........................................................................... 2700 7. Alkyne-mediated cyclizations .......................................................................... 2700 8. Other metal-promoted cyclizations ...................................... ................................ 2704 8.1. Cyclizations of 4-alkynols ...................................... ................................. 2704 8.2. Cyclizations of diazo compounds ................................... .............................. 2706 8.3. Intramolecular etherification of aryl and vinyl halides ................................................. 2709 9. Intramolecular conjugate additions ..................................... ................................ 2710 9.1. 6-exo Ring closures ............................................................................. 2710 9.2. 6-endo Ring closures ............................................................................ 2713 10. Conclusions ....................................................................................... 2717 Acknowledgements ............................................ ...................................... 2717 References and notes ................................................................................. 2717 Biographical sketch ............................................ ...................................... 2723 * Corresponding authors. Tel.: þ34 93 4039106 (P.R.); tel.: þ34 93 4021247 (F.U.). E-mail addresses: [email protected] (P. Romea), [email protected] (F. Urpı ´). 0040-4020/$ - see front matter Ó 2007 Elsevier Ltd. All rights reserved. doi:10.1016/j.tet.2007.11.092 Available online at www.sciencedirect.com Tetrahedron 64 (2008) 2683e2723 www.elsevier.com/locate/tet

Transcript of Available Online at Www.sciencedirect.com

Page 1: Available Online at Www.sciencedirect.com

Available online at www.sciencedirect.com

Tetrahedron 64 (2008) 2683e2723www.elsevier.com/locate/tet

Tetrahedron report number 828

Synthesis of six-membered oxygenated heterocycles throughcarboneoxygen bond-forming reactions

Igor Larrosa a, Pedro Romea b,*, Felix Urpı b,*

a School of Biological and Chemical Sciences, Queen Mary, University of London, Mile End Road, London E1 4NS, UKb Departament de Quımica Organica, Universitat de Barcelona, Martı i Franques 1e11, 08028 Barcelona, Catalonia, Spain

Received 12 November 2007

Contents

1. Introduction and scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26842. SN2-Mediated cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26843. SN1-Mediated cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26854. Epoxide-mediated cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26865. Alkene-mediated cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2690

5.1. Stoichiometric cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26915.1.1. Mercury-mediated cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26915.1.2. Halo-mediated cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26925.1.3. Seleno-mediated cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2693

5.2. Catalytic cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26945.2.1. Palladium-mediated cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2694

5.2.1.1. Catalysis by Pd(0) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26945.2.1.2. Catalysis by Pd(II) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2695

5.2.2. Other metal-mediated cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26995.3. Acid-mediated cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2699

6. Allene-mediated cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27007. Alkyne-mediated cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27008. Other metal-promoted cyclizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2704

8.1. Cyclizations of 4-alkynols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27048.2. Cyclizations of diazo compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27068.3. Intramolecular etherification of aryl and vinyl halides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2709

9. Intramolecular conjugate additions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27109.1. 6-exo Ring closures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27109.2. 6-endo Ring closures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2713

10. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2717Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2717References and notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2717Biographical sketch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2723

* Corresponding authors. Tel.: þ34 93 4039106 (P.R.); tel.: þ34 93 4021247 (F.U.).

E-mail addresses: [email protected] (P. Romea), [email protected] (F. Urpı).

0040-4020/$ - see front matter � 2007 Elsevier Ltd. All rights reserved.

doi:10.1016/j.tet.2007.11.092

Page 2: Available Online at Www.sciencedirect.com

BrO

OMe

OH

OMe

OHN

O

O

O

O

O

N

O

O

R2R1

O

O OMe O O

NON

O

H

MeO

O

O

1 leucascandrolide A 2a phorboxazole A, R1 = OH, R2 = H2b phorboxazole B, R1 = H, R2 = OH

Figure 1.

2684 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

1. Introduction and scope

Six-membered oxygenated heterocycles, or pyrans, are prob-ably one of the most common structural motifs spread acrossnatural products, from simple glucose to structurally complexmetabolites such as leucascandrolide A (1),1 phorboxazole A(2a) and B (2b),2e4 and the even more elaborated architecturespresent in palytoxin, maitotoxin, and other marine natural prod-ucts (Fig. 1).5

Due to the remarkably rich array of functionalities and chiralcenters that these heterocycles can incorporate, their stereose-lective preparation has become a continuous challenge fororganic synthesis practitioners.6e10 They can be synthesizedby means of five-membered ring expansions (e.g., BaeyereVil-liger oxidation of cyclopentanones11), cycloaddition processes(e.g., Hetero-DielseAlder reactions12), or intramolecular cycli-zations.13 Classical examples of these cyclizations are thelactonisation of d-hydroxy acids or the thermodynamicallyfavorable conversion of d-hydroxy aldehydes into the corre-sponding hemiacetals, which, in turn, can be easily modified(e.g., C-glycosidation reactions14) to provide other pyran-basedstructures. In addition to these processes, there is a set of impor-tant methodologies based on the cyclization of an oxygenatedprecursor that affords pyran structures in a highly efficientand straightforward manner.

Considering the crucial ring-forming step (or the paralleltransform in the retrosynthetic sense), the cyclization methodol-ogies can be classified into three types, represented in Figure 2.Type 1 gathers those methodologies based on O1eC2 bondformation15 (or disconnection in the retrosynthetic sense),which encompass SN2- and SN1-mediated cyclizations, metal-promoted processes, and Michael-like reactions. Likewise,Type 2 and 3 methodologies affect C2eC3 and C3eC4 bondformation, respectively, which include Prins sequences, Peta-siseFerrier rearrangements, or ring-closing metathesisreactions.

1

2

34

O

Type 1

O

Type 2

O

Type 3

2

3

Figure 2.

The aim of this report is to highlight Type 1 methodologiesaddressed to the stereoselective construction of di- and tetra-hydropyrans and their application to the synthesis of naturalproducts, with special attention being paid to the most recentcontributions.

2. SN2-Mediated cyclizations

Transformations based on an SN2-mediated cyclization ofa hydroxy precursor represent the simplest strategy leadingto di- and tetrahydropyran. They only require a suitable com-bination of base (B: tertiary amine, alkoxide, hydride; seeScheme 1) and leaving group (X: sulfonate, halide; seeScheme 1) to promote a 6-exo-tet process16 with inversion ofthe configuration at the CeX center in accordance with thepattern dictated by the Williamson reaction (see path A inScheme 1). Thus, the stereochemistry of the resulting tetrahy-dropyran relies exclusively on the configuration of the acyclicprecursor.

OH X

O X

O

OH

path A path B

– H

B

Scheme 1.

With regard to the reactivity of such systems, cyclizationsthat involve leaving groups placed on a primary positiontake place smoothly,17 whereas those located on a secondaryposition often require a more accurate control of the reactionconditions.18 Even in such a challenging situation, this strat-egy enjoys an interesting synthetic potentiality, as has beenclearly proved in the successful construction of tetrahydropyr-ans embedded in leucascandrolide A (1),1f and phorboxazoleA (2a)2d and B (2b).3b Noteworthy, either 2,6-cis-tetrahydro-pyran (see 4 and 8 in Scheme 2) or 2,6-trans-tetrahydropyran(see 6 and 10 in Scheme 2) is obtained by simple applicationof the same experimental conditions. Thus, both stereochemistries

Page 3: Available Online at Www.sciencedirect.com

PhH, 90 °C

1

S

S O

S

S OH OTs

OPMB

OMe OOTsTBSO

OMe OO

OPMB

OPMB

OPMB

OH

NaH

3 4

6

OTBS1) HF·pyr, pyr 5

75%

2) NaH, PhH, 60 °C

73%

OTBDPS

OTBDPS73

7

3

7

11

15

15

11

3

7

3

5

99

5

1115

111511

15

1115

2a

O

OO N

PMBO

OH

OPiv 1) MsCl, Et3N2) TsOH, MeOH O

O N

PMBO

OPiv

OO N

PMBOTsO

HO

OTBDPS

OO N

PMBOO

OTBDPS

PivO

7 8

OPiv9

NaH

3) NaH, PhMe, 72%

PhH, 89%

10

Scheme 2.

2685I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

are available, depending on the configuration of the reactingcenter.

Remarkably, an SN2-mediated cyclization has also beenused for the assemblage of C5eC9 tetrahydropyran of phor-boxazole A (2a) in the final stages of the synthesis.2h Thereby,alcohol 11 is easily converted into its mesylate 12, the siliconprotecting group at C5 is selectively removed and the simpleexposure of the resulting hydroxy mesylate 13 to Et3N affordsthe desired 2,6-trans-tetrahydropyran 14 in an excellent yield(Scheme 3).

OP

O

ON

O

OMe

OMeOTBDPS

MeO

Br

O

ON

O

OMe

OMeOTBDPS

MeO

Br

11 R1: H R2: TES12 R1: Ms R2: TES13 R1: Ms R2: H

MsCl, Et3N, 99%

PPTS cat., MeOH, 98%

2a

Scheme

3. SN1-Mediated cyclizations

In contrast to the above-mentioned methodologies, SN1-mediated cyclizations take advantage of the stability of allylicand benzylic carbocations to drain the process represented inScheme 1 along path B.19,20 Now, the planarity of the carbo-cationic intermediate does not exert any control on the config-uration of the new stereocenter and the stereochemicaloutcome of the cyclization relies on the stereogenic elementsof the substrate and the experimental conditions.

MB

NO

OTBDPS

OR1 OR2

OTBDPS

OPMB

O

N

O

O

OTBDPS

O

86%

Et3NMeCN,

14 OTBDPS

5

5

9

3.

Page 4: Available Online at Www.sciencedirect.com

H

OTBS

Cl

TBSO OTBS

HO

H

OTBS

Cl

TBSO OTBS

H

OTBS

Cl

TBSO OTBS

H

O

ClOTBS

OTBS

15 16 17

45%dr 5:1

12b 12b 12b

Scheme 4.

2686 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

Thereby, cyclization of benzylic allylic alcohol 15 duringan aqueous work-up is supposed to proceed through cationicintermediate 16, which undergoes ring closure followed bythe release of the silicon protecting group (Scheme 4).19d

The stereochemistry of the major diastereomer 17 arisesfrom the preferential attack of the OTBS oxygen from theopposite side of C12b methyl.

Better results are typically obtained for 2,6-cis-tetrahydro-pyrans, which are usually the most stable diastereomers. Forinstance, construction of the C22eC26 2,6-cis-tetrahydro-pyran of phorboxazole A (2a) has been achieved through anintramolecular SN1-mediated cyclization of alcohol 18, asshown in Scheme 5. Actually, the stereochemical outcome ofthis transformation is consistent with the internal capture ofa transoid allylic cation at C26 by the C22 b-methoxymethylether and dealkylation of the resulting oxonium species.2d,21

Propargylic carbocations can also participate in thesecyclizations, since they are easily generated by treatment ofCo2(CO)6ealkynol complexes with either Brønsted or Lewisacid22 and are stable enough to be intramolecularly trappedby oxygenated nucleophiles. Indeed, ring closure of diol 20affords tetrahydropyrans 21 in excellent overall yield (Scheme6).23 Remarkably, the thermodynamically more stable 2,3-trans derivative is increasingly obtained with long reactiontimes or higher temperatures.

Furthermore, epoxides can also act as nucleophiles and, inparticular, precursors containing a protected epoxy alcoholmoiety participate in highly regio- and stereoselective

OTES

OH OPMB

OMOM

N

OR

Tf2O, pyrCH2Cl2, –20 °C

55%

22

26

18

Scheme

1) BF3·OEt2, CH2Cl2, T2) CAN, acetone, 0 °CHOHO

C5H11

OBn

Co2(CO)6

20

T (°C) t (h) trans-cis Y

20–20–20

0.332.518

10:11.2:19.9:1

Scheme

cyclizations. For instance, treatment of carbonates 22a orb (Scheme 7) with BF3$OEt2 followed by protecting grouptransformations furnishes 2,6-cis-tetrahydropyran 23a or b asa sole diastereomer through a putative propargylic carboca-tion,24 which has provided a new entry to a formal synthesisof (þ)-muconin 24.25

4. Epoxide-mediated cyclizations

Since the seminal studies on the intramolecular epoxideopening carried out by Nicolaou et al.26 established the struc-tural criteria required for the regio- and stereocontrolledsynthesis of six-membered oxygenated heterocycles overrid-ing the competitive formation of the corresponding five- orseven-membered counterparts, acid-catalyzed cyclizations ofhydroxy epoxides have become a common approach to thestereoselective construction of tetrahydropyrans.27

The regioselectivity of the cyclization for g-hydroxy epox-ides greatly depends on the geometry of the epoxide. Indeed,cis-epoxides reliably afford the 5-exo ring closure probablybecause these systems do not easily assume the planar ar-rangements necessary for maximum stabilization in the transi-tion states leading to six-membered rings (Eq. 1 in Scheme 8).On the other hand, trans-epoxides enable the desired 6-endoprocess provided that a p-orbital adjacent to the epoxideunit activates the CeO bond close to it and stabilizes thedeveloping positive charge in the transition state (Eq. 2 inScheme 8). In both cases, these cyclizations are accompanied

OTES

O

N

OR

OPMB

2a

22

26

19

5.

, t

OC5H11

OBnH

HO

C5H11

OBnH

H

+

21-trans 21-cis

ield (%)

947087

6.

Page 5: Available Online at Www.sciencedirect.com

OBoc

TMS

OH

OCo2(CO)6

22a

OBocO

22b

OH

TMS Co2(CO)6BF3·OEt2, CH2Cl2

–20 °C, 4 h

O

OO

TMS

t-BuO

(CO)6Co2

O

OH

TMS

(CO)6Co2 OBoc

1) Ac2O, CH2Cl2, rt, 30 min2) CAN, acetone, rt, 10 min

O

OAc

OBoc

70%

TMS O

O O

Ot-Bu

(CO)6Co2

BF3·OEt2, CH2Cl2rt, 3 h

OTMS

(CO)6Co2

OHOH

1) CAN, acetone, rt, 30 min2) K2CO3, MeOH, rt, 6 h3) CH2(OMe)2, CSA, rt , 3 h

OO

O

55%

OO

n-C12H25OH

OHO

OH

O

23a 23b

(+)-muconin 24

Scheme 7.

6-exo7-endo (3)

(4)6-exo7-endo

favored

cis epoxide

favored

R: CH=CH2

favored

R: CH2X

RO

OHO

HO

R OHO

RH

H

H

trans epoxide

O

OHR O

HO

RH

O

HO

R

H

H

Scheme 9.

RO

OH O

HO

R

6-endo

O5-exoHO

R

OOH O

HO

R

6-endo

O5-exoHO

R

favored

R

cis epoxide

trans epoxide

favored

R: CH2Xfavored

R: CH=CH2

H

H

H

H

(1)

(2)

H

H

Scheme 8.

2687I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

by inversion of the stereochemistry at the carbon undergoingnucleophilic attack.

A similar analysis was performed for d-hydroxy epoxides.In these systems, cis-epoxides exhibit low selectivity for the7-endo ring closure irrespective of the R substituent, preferen-tially forming the corresponding tetrahydropyran (Eq. 3 inScheme 9). Conversely, the R substituent plays a crucial rolein the case of trans-epoxides. As expected on kinetic grounds,any saturated substituent favors the 6-exo process and yieldsalmost exclusively the six-membered ring, whereas, onceagain, a p-orbital close to the epoxide directs the attack tothe allylic position and produces the oxepane ring (Eq. 4 inScheme 9).

These models account for most of the synthetic approachesfound across the most recent literature.28 For instance, con-struction of the bicyclic ether core of (þ)-sorangicin A (25

in Scheme 10) illustrates some of the aforementioned trends.Exposure of epoxide 26 to aqueous acid produces tetrahydro-pyran 27 through a 6-exo ring closure (see Eq. 4 in Scheme 9),without observing alternative 7-endo or even 5-endo cycliza-tions. Furthermore, 27 can be easily converted into epoxyalcohol 28, which is subjected to an acid-mediated cyclizationthat follows a 5-exo pathway (see Eq. 1 in Scheme 8). Interest-ingly, three steps can be combined in a one-pot, three-stepsequence that delivers the bicyclic ether 29 in a 62% overallyield.29,30

Similarly, a parallel set of experimental conditions based onthe use of a Lewis acid (Ti(i-PrO)4 activation in hot benzene)have been devised for the assemblage of the C22eC26 2,6-cis-tetrahydropyran 31 from 30 in a route to phorboxazole A (2a)(Scheme 11).2c

Page 6: Available Online at Www.sciencedirect.com

HOOTs

OH OHO H3O O

OHOTs

HOO

O

OO

OH

Ph

O

OH

HO

HO

O

OO

O O

HOOC

26 27

1) 10% HCl, rt, 18 h2) 10% NaOH, rt3) 10% HCl, rt, 4 h

62%

Ph

Ph

Ph

29 28

(+)-sorangicin A25

6-exo

7-endo

5-endo

5-exo

H3O

OH

6-endo

Scheme 10.

2226

22

26HO HO OPMBO

PMBO OH

OH

O

2a

30 31

Ti(i-PrO)4

PhH,

76%

6-exo

Scheme 11.

2688 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

As pointed out in Scheme 8, alkenes adjacent to epoxidesoverride the common bias favoring the 5-exo ring-closuremode. This is the behavior observed in the intramolecularcyclization of a,b-unsaturated epoxide 33, which permits theregio- and stereoselective preparation of a densely substitutedC8eC12 tetrahydropyran directed toward the synthesis of bre-venal (32 in Scheme 12).31 Indeed, removal of the PMB groupon 33 triggers a 6-endo cyclization to a tetrahydropyran, whichis subsequently converted into TES ether 34.32

Occasionally, these methodologies are combined in a syn-thetic sequence. For instance, the tristetrahydropyran substruc-ture of diol 35, an advanced intermediate toward the synthesisof thyrsiferol and venustatriol (36a and 36b, respectively, inScheme 13), has been regio- and stereoselectively preparedthrough three intramolecular cyclizations of g- and d-hydroxyepoxides.33 The first acid-catalyzed ring closure of hydroxyepoxide 37 proceeds quantitatively to form 38 with complete6-exo regioselectivity,y and no competitive 7-endo cyclizationis observed. Next, the construction of tetrahydropyran B takesadvantage of the allylic character of a trans-epoxide. Indeed,cyclization of hydroxy diepoxide 39 affords the desired six-membered heterocycle 40 under very mild conditions withexcellent 6-endo regioselectivity. Eventually, Ti(i-PrO)4 acti-vation of hydroxy epoxide 41 produces the third ring in 58%yield.

In spite of these accomplishments, it was soon realized thatnew methodologies able to switch 5-exo to 6-endo processesirrespective of the substituent R (see Eqs. 1 and 2 in Scheme

y Actually ring closure to tetrahydropyran A is carried out on a 1:1 epimeric

mixture of 3-bromo derivatives. Interestingly, the experimental conditions pro-

vide a highly efficient kinetic resolution and just hydroxy epoxide 37 un-

dergoes the desired cyclization.

8) were highly desirable. In this context, it has been recentlyreported that epoxide openings triggered by bulky silyl triflatesin nitromethane yield the desired tetrahydropyrans regioselec-tively.34 In addition to these observations, the regioselectivetenet summarized in Scheme 8 favoring 5-exo over 6-endoring closures of g-hydroxy epoxides has recently been chal-lenged in vanadium-catalyzed mediated processes.35 As out-lined in Scheme 14, the asymmetric oxidation of unsaturateda-hydroxy esters followed by an epoxidation/cyclization se-quence using the same chiral catalyst provide regio- (6-endo/5-exo ratio ca. 4:1) and stereoselectively (dr>88:12, ee>95%)2,5-trans-tetrahydropyrans in 20e35% yield (the highest theo-retical yield is 50%).

Hence, this process is divided into two steps. The first stepinvolves the oxidative resolution (acetone/O2) of a racemicmixture of a-hydroxy esters (R1, R2sH) using a vanadiumcatalyst prepared from VO(i-PrO)3 and the chiral Schiff base42. Once the alcohol has been chemo- and stereoselectivelyoxidized, change of solvent and oxidant to CHCl3/t-BuOOHpromotes a stereoselective epoxidation followed by a fastcyclization of the resultant hydroxy epoxide. The mechanisticmodel that accounts for the stereochemical outcome of theoverall process emphasizes the coordination of the estercarbonyl to the metal. Such a coordination places the esterin the pseudoaxial position, which allows an efficient p-facialdiscrimination of the carbonecarbon double bond. Finally, theresultant short-lived epoxide undergoes a backside attack ofthe hydroxyl to the corresponding 2,5-trans tetrahydropyran.36

The pursuit for regioselective 6-endo cyclizations has foundits own Holy Grail in the synthesis of fused polycyclic ethers,a family of marine metabolites with remarkable biological ac-tivity.9 As partly suggested by the name, their molecular archi-tectures contain large and highly rigid polycyclic frameworksof all fused six- to nine-membered oxygenated rings in a well-defined stereochemical arrangement, as shown in Figure 3.

Page 7: Available Online at Www.sciencedirect.com

TBDPSOOPMB

OTBDPSO

OH Me

OTESH

89%

1) DDQ, CH2Cl2-H2O, rt

2) TESOTf, 2,6-lut, CH2Cl2, 0 °C

33

brevenal 32

8

12

8

12

8 12

34

6-endo

O

OO H H

OHH Me

OH H

MeOH

O

MeH

MeH

OHC

H

Scheme 12.

2689I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

The synthesis of these large and complex structures hasraised much interest in strategies rooted on biogeneticgrounds. Unfortunately, the biosynthesis of these metabolitesis not well understood and current working hypotheses arerather speculative. In this context, the CaneeCelmereWestleymodel, originally proposed for the synthesis of polyether anti-biotics,37 has inspired most of those hypotheses, which envis-age that the fused polycyclic ether frameworks might arisefrom the cyclization of a polyepoxide precursor through acascade of SN2 ring openings.38 Therefore, any biomimeticstrategy must be able to provide a synthetic route to a chiral

OAc

O

HOBr

Br

O

Br

OOTBS

O

O

Br

OOH

O

OH

OH

100%

40 mol% CSAEt2O,rt, 0.5 h

70%

1 equiv PCHCl3, rt, 5

0.9 PhM

37

40

41

A

B

A

B

O

Br

O

O

A

B C

Scheme

ROR1

O

OH R2

HO

R2

R1O

O[V]t-BuOO CO2R

HO

2,5-trans-tetrahydropyran

acetone, 30 °C, 30–48 h

20–35%dr > 88:12ee > 95%

R1

R2

25

10 mol% VO(i-PrO)3, 11 mol%O2

Scheme

polyepoxide with the structural elements to allow for theregio- and stereoselective 6-endo ring closures required forthe fused tetrahydropyran arrays. Otherwise, alternative path-ways based on 5-exo cyclizations leading to tetrahydrofuranswould predominate (Scheme 15).39

In this context, a biomimetic approach takes advantage ofendo-selective cyclizations of methoxymethyl epoxy alcoholscontrolled by chelation with a suitable Lewis acid. Indeed,the asymmetric epoxidation of polyene 43 leads to 44 in 56%isolated yield, and removal of the silicon protecting group pro-vides quantitatively triepoxide alcohol 45 (Scheme 16). Then,

OAc

OOH

O

Br

OHO

OTBSO

O

Br

O

O

OH

OH

A

PTS min

58%

equiv Ti(i-PrO)4e, 70–80 °C, 3 h

38

39

35

A

B

A

B

C

thyrsiferol, 36a

R

OHR : OH35

R : OOH

HHO

venustatriol, 36b R : OOH

HHO

13.

ROR1

O

OH R2

10 mol% VO(i-PrO)3, 11 mol% 42

OH

R1

O[V]R2

H

OROt-BuO

OR OH

NOH

t-Bu

t-Bu

t-Bu

42

t-BuOOH, CHCl3, rt, 24–72 h

42

14.

Page 8: Available Online at Www.sciencedirect.com

O

( ) OO

R1 R2

R3 R4n

n = 1, 2, 3, or 4

R1, R2, R3, R4 : H, Me

trans trans

cis

Figure 3.

2690 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

a cascade cyclization catalyzed by La(OTf)3 proceeds in a step-wise mechanism to afford the desired tricyclic ether 46 in 9%yield.40

Lewis acid-mediated endo oxacyclization of polyepoxidesaffords cis- or trans-fused polyethers depending on the natureof the terminal nucleophile. The rationale for such dependencerelies on the mechanism depicted in Scheme 17. Initially,Lewis acid activation of a first epoxide triggers the generationof a bridged onium intermediate, which can evolve througha tight ion pair to a tertiary carbocation. Therefore, the termi-nal nucleophile may intercept different species. It has beenproposed that the cis-fused products arise by trapping of an in-termediate close to the carbocation, whereas the trans-fusedderivatives are mainly obtained with a better nucleophile(i.e., a tertiary carbamate) by adding to an intermediate struc-turally close to the epoxonium ion. This is the case for dime-thylcarbamate epoxide 47 prepared through stereoselectiveSharpless and Shi epoxidations of the corresponding alkenes.Then, BF3$OEt2-promoted oxacyclization of 47 produces thedesired polyether 48 in 31% isolated yield.41

O

OO

HO

H

H

H 9%

46

43

MeO

MeO

MeO

OR

O

OO

OO

O, Oxone

MeCN, (MeO)2CH2, 0 °C

56%R: TBDPS TBA

1

OMe

OMe

MeOO

OHO

R

H

HOMe

OMe

Scheme

OO

Nu

O

Nu O

OO

Nu

O

Nu

O5-exo ring closure

6-endo ring closure

Scheme 15.

A different strategy employs epoxysilanes as starting mate-rials for the cyclizations. It takes advantage of trimethylsilylgroups to achieve highly regioselective openings of the epox-ides, and, immediately after the ring closure, the silicon groupis removed by fluoride-promoted protodesilylation.42 As out-lined in Scheme 18, this approach allows for the straightfor-ward conversion of triepoxysilane 49 into the largetetrahydropyran array 50. Indeed, chiral epoxide 49, preparedfrom a reagent-controlled Shi epoxidation (dr 90:10), un-dergoes a cascade of ring openings mediated by a Brønstedbase and a fluoride source in a hydroxylic solvent. Finally,the resultant alcohol is acylated to furnish 50 in 20% overallyield.

5. Alkene-mediated cyclizations

Closely related to the epoxide-mediated cyclizations, thereis an important number of methodologies based on the intra-molecular attack of an oxygenated nucleophile on an olefinactivated by electrophiles. The most common precursors ared-hydroxy alkenes, which afford the corresponding pyransthrough highly regioselective 6-exo ring closures (Scheme19). The rationale for these cyclizations assumes the reversibleformation of a p-complex, which can proceed either directlyto the corresponding heterocycle by reaction with the nucleo-phile or indirectly by first collapsing to an onium intermediatebefore undergoing nucleophilic attack.43 In any case, cycliza-tion involves the attack of the oxygenated nucleophile on theopposite face of the electrophile. Therefore, the success ofthis strategy mostly relies on the stereocontrolled electrophilicaddition to the alkene.

These cyclizations can be classified as stoichiometric or cat-alytic, according to the amount of electrophile engaged in theprocess. The stoichiometric electrophile-induced cyclizationsroutinely involve mercury(II) salts and iodo or seleno reagentsas activators of the carbonecarbon double bond. Otherwise,palladium chemistry dominates the catalytic counterparts, al-though other metals are being increasingly employed.

1.1 equiv La(OTf)30.3 equiv La2O35.5 equiv H2O

CH2Cl2, rt44

MeO

MeO

MeO

OR

R: TBDPS

45 R: HF, THF00%

O

O O O

R

OMe

La

OH

O

OH

OMe

OR

OMe

H

La

16.

Page 9: Available Online at Www.sciencedirect.com

47

OO O

Me2N

O

OO

O

O

OO

OH

Me H

H Me0.01 M BF3·OEt2

CH2Cl2, 20 °C

31%48

OOR

O

OBF3

OOR

O

OBF3

OOR

O

OBF3

OR

O

O OOR

O

O O BF3BF3

tertiary carbocation ion pair epoxonium ion

OO

O OBF3R

H

Me

OO

O OBF3R

H

Mecis-fused trans-fused

Scheme 17.

2691I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

5.1. Stoichiometric cyclizations

5.1.1. Mercury-mediated cyclizationsThe electrophiles employed in these processes are mercu-

ry(II) salts and the resultant carbonemercury bonds are usu-ally reduced with NaBH4 or n-Bu3SnH.44 The course ofthese intramolecular oxymercuriations depends on the stabilityof the cationic intermediates and, with the exception of oneexample,45 they are substrate-controlled processes that usuallylead to the thermodynamically more stable tetrahydropyran.Thus, the hydroxy alkene must contain the structural elementsrequired to achieve the desired p-facial discrimination of thecarbonecarbon double bond. This is the case for the d-hy-droxy alkene 52, an intermediate in the total synthesis of theantibiotic X-206 (51 in Scheme 20).44a Indeed, steric and con-formational arguments would be responsible for the outstand-ing stereocontrolled electrophilic addition of Hg(OAc)2 to theolefin in such a way that the resulting onium intermediate 53 issuitably prepared for the intramolecular cyclization leading to2,6-cis-tetrahydropyran 54 quantitatively. Eventually, treatment

OTMS

O

O

TMS

TMS

O

H

HOH

1) Cs2CO3, C

2) Ac2O, DM

49

Scheme

OH

OH

+ E

OH

E

onium interme

complex

6-ex

– H

Scheme

of 54 with n-Bu3SnH effects the desired demercuriation to 55in excellent yield.

The intramolecular oxymercuriation has also been appliedto the assemblage of the C22eC26 tetrahydropyran of phor-boxazole B (2b). As shown in Scheme 21, treatment of diol56 with Hg(OAc)2 followed by reaction of the resultant orga-nomercuriate with iodine furnishes a 5:1 mixture of the 2,6-cis-tetrahydropyran 57-cis and the corresponding 2,6-transisomer. Then, chromatographic purification of the mixture per-mits the isolation of the desired 57-cis in 71% yield.

Given that the success of such a strategy relies on the p-facial discrimination of an alkene, any methodology able tofulfill this requirement should allow for a complete stereochem-ical control. In this context, mercury(II)-mediated electrophilicring-opening reactions of d-hydroxy cyclopropylcarbinol deriv-atives represent an appealing entry into the asymmetric synthe-sis of tetrahydropyrans. In this approach, the cyclopropane unitplays the role of an activated olefin. These highly diastereo-selective processes usually occur with anchimeric assistanceby the closest alcohol and subsequent nucleophilic backside

AcO O OH

OHH

H

O

HH

H HsF, MeOH,

AP, pyr, CH2Cl2

20%50

18.

E OE

6-exo

– H

O– E

diate

o

19.

Page 10: Available Online at Www.sciencedirect.com

3 7

3 7 3 7OH

COXc

OBOM

R OHXcCOH

H

H HgOAc

H ROBOM

O OHO

O OH

OHO

OHOH

O OO

OH

OH

Hg(OAc)2

CH2Cl2, rt

NaCl aq.O

COXc

OBOM

RX99%

54 X: HgCl55 X: H n-Bu3SnH

X-20651

52 53

PhMe94%

Scheme 20.

2692 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

attack of the d-OH, which causes an inversion in the configura-tion of the reacting center (Scheme 22). Hence, the stereochem-ical outcome of these cyclizations mostly depends upon theconfiguration of the cyclopropane ring.46

This strategy has been applied to the construction of the C3eC7 tetrahydropyran of zincophorin (58 in Scheme 23).47 Thecyclopropane ring of the advanced intermediate 59 is installedat the beginning of the synthesis by means of the catalyticasymmetric cyclopropanation of an allylic diazoacetate.Then, simple treatment of 59 with Hg(OCOCF3)2 followed byaqueous work-up cleanly produces 2,6-trans-tetrahydropyran60. In turn, this organomercuric bromide is subjected to a reduc-tive demercuriation, which eventually provides 61 in 85% yieldand 93:7 dr.

5.1.2. Halo-mediated cyclizationsHalo-mediated cyclizations are similar to the processes

based on mercury(II) salts. Most of the intramolecular halo-etherification reactions take advantage of the use of iodineor N-iodosuccinimide (NIS),48e50 reagents mild enough tobe used on structurally complex substrates. For instance,NIS-mediated iodoetherification plays a crucial role in the as-semblage of the G-ring of azaspiracid-1 (62 in Scheme 24),a prominent member of a family of marine metabolites respon-sible for human poisoning resulting from the consumption oftainted shellfish. Initially reported in 1998, azaspiracids wereshrouded in structural ambiguity that partially arises from

1) Hg(OAc)2, PhMe, 0 °C,

2) I2, rt, 12 h

56

OH

OHOO 71%

22

26

Scheme

OHHOHgX2

OHHO

Hg

X X

–H

Scheme

the complex and intricate array of sensitive functional groupsembedded in their structure and, only in 2004, have Nicolaouet al. reported the correct structure of azaspiracid-1 after con-siderable synthetic work. It is, then, not surprising that theendeavors applied to the construction of such challengingcompounds have spurred the development of new strategies(see also Schemes 57 and 91). As represented in Scheme 24,the intramolecular iodoetherification of the structurally com-plex alcohol 63 affords iodoether 64, which is reduced withan excess of n-Bu3SnH in toluene to remove the superfluousiodine at C29. Thus, the highly advanced intermediate 65 isobtained through this two-step sequence in 53% yield.

Iodoetherification on substrates containing an allylic alco-hol turns out to be highly dependent on conformational issues.Indeed, diol 66a undergoes a facile cyclization to 2,6-cis-tetra-hydropyran 67a, while the diastereomer 66b fails to give theexpected bicyclic system 67b (Scheme 25).48e The rationalefor such different behavior relies on the conformational anal-ysis of the allylic moiety. Cyclizations in which the nucleo-phile is in the R group proceed rapidly (or slowly) throughthe favored p-complex 68a (or the disfavored p-complex68b) when CeOH (or CeH) eclipses the carbonecarbon dou-ble bond. This is the situation for 66a (or 66b). Thus, cycliza-tion of 66b requires a more reactive iodinating agent such asbis(sym-collidine)2IPF6, prepared in situ from bis(sym-collidi-ne)2AgPF6 and iodine in CH2Cl2. This reagent converts 66binto the desired tetrahydropyran 67b in 80% yield.

8 h

57-cis

O

OHOO

I

2b

22

26

21.

OHO

HgX

X OHO

reductive

demercuriation

22.

Page 11: Available Online at Www.sciencedirect.com

I2

HO

HO

O

H

H O

I2, NaHCO3

MeCN0 °C to rt, 2.5 h

66a

HO

HO

O

H

H O

66b (collidine)2AgPF6, I2, CH2Cl2, rt, 2

80%

HO

OH

H

R

OO

I2

I2, NaHCO3 HO

H

OH

R

OO

MeCN

Scheme

60

61 R: H

ON2

O

O

O

CH2Cl2,91% ee > 95%

0.1% Rh2[5R-MEPY]4

BzOO OH

O

59

2) KBr aq.

85% dr 93:7

1) Hg(OCOCF3)2, CH2Cl2, rtBzO

O OH

O

R

R: HgBrn-Bu3SnH, PhMe-THF

–60 °C to rtO

O

HO

OH OH OH

OH

zincophorin 58

NO HCO2Me

RhRhN ON O

O N

Rh2[5R-MEPY]4

H

3 7

Scheme 23.

OHO

OO

O

OO

OHOH

HO

OO

NH

H

H

azaspiracid-162

H

I

F

G

C D

A

B

OAcO

OO

OO

OH

OHO

O

TeocN

H

H

63

OTBS

NIS, NaHCO3THF, 0 °C, 12 h 62%

OAcO

OO

O OOH

OO

O

TeocN

H

H

64

OTBS

XX: I

cat. Et3B, n-Bu3SnH-PhMe 1:2, 0 °C, 20 min86%

65 X: H

29

Scheme 24.

2693I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

A conceptually different approach to electrophile-inducedcyclizations takes advantage of the nucleophilicity of the car-bonyl oxygen of a ketone. As outlined in Scheme 26, if thecarbonyl of a d-alkenyl ketone could trap the p-complex ob-tained by treatment of the olefin with an electrophile, thenthe resultant oxocarbenium intermediate would easily lead tothe corresponding acetal.

This idea has been put into practice for the stereoselectiveconstruction of the C1eC5 ring of a new antitumor agent,FR-901464 (69 in Scheme 27). With this target in mind, theNBS-mediated cyclizations of two d-alkenyl ketones, 70 and71, have been tested and, in both cases, a sole diastereomerhas been isolated in good yield.51a Unfortunately, none of theresultant C4eC5 stereochemistries in 72 and 73 match withthat in FR-901464 and its final total synthesis has recentlybeen achieved following an alternative strategy.51b Despitethis failure, the key concept of this methodology maintainsa high appeal, as has been proved in a parallel process in thearena of alkynols, which has been commented further upon.52,53

5.1.3. Seleno-mediated cyclizationsCyclizations based on selenium reagents are not as com-

mon as the mercury(II)- or halo-mediated methodologies.54,55

Moreover, and contrary to these methods, the stereochemical

70%

O

HO

O

H

H O

67a

I

H

.5 h

O

HO

O

H

H O

67b

I

H

< 5%

HOH

H

R

I2

favored complex

68a

HH

R

HOI2

disfavored complex

68b

25.

Page 12: Available Online at Www.sciencedirect.com

EO R O RE

O RE R1OH

–H

OE R

OR1

Scheme 26.

NBSO

TESOO

MeOH, MeCN

59%

O

TESOO

Br OMe

70 72

NBSO

TESOO

MeOH, THF

47%

O

TESOO

BrOMe

71 73

OO O

ONH

AcO

HO

OH15

5

4

5

4 FR-90146469

Scheme 27.

2694 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

outcome of the selenoetherification reactions mostly relies onthe coordination of electrophiles such as phenylselenyl chlo-ride (PhSeCl) or N-phenylselenophthalimide (NPSP) withthe olefin, which is then followed by the ring-closure step.Therefore, the stereoselectivity of these cyclizations reflectsthe kinetic selectivity for the formation of the onium interme-diate, which may give access to 2,6-trans-tetrahydropyrans.This is the trend observed for hydroxy alkene 74. As shownin Scheme 28, 2,6-trans-tetrahydropyran 75-trans is onlyobtained through a selenium-mediated cyclization, whereas75-cis results from the other methods.50

A more selective reagent has been used for the assemblage ofthe C11eC15 2,6-trans-tetrahydropyran of leucascandrolide A(1) from diol 76 (Scheme 29).1d Unexpectedly, iodine-basedelectrophiles show little diastereoselectivity, but this is dramat-ically improved by the use of selenium reagents. Remarkably,the bulky 2,4,6-triisopropylphenylselenyl bromide (TIPPSeBr)z

affords 2,6-trans-tetrahydropyran 77 in 74% yield and gooddiastereomeric ratio (dr 88:12). Eventually, a tin-free radicalreduction of 77 provides the desired advanced intermediate 78in 85% yield.

5.2. Catalytic cyclizations

5.2.1. Palladium-mediated cyclizationsIn spite of the achievements mentioned above, the use and

removal of stoichiometric amounts of often toxic elementshave fueled research into alternative activators of olefins thatallow the desired intramolecular cyclizations under mild con-ditions and in a catalytic fashion. In this context, catalyticmethodologies based on the activation of olefins are domi-nated by the palladium chemistry, which has achieved promi-nent levels of maturity.13e,56 In this arena, the construction ofheterocycles can proceed through carbonecarbon or carboneheteroatom bond formation, depending upon the nucleophileinvolved in the ring-closure step. In particular, the synthesis

z TIPPSeBr is prepared in situ from (TIPPSe)2 and bromine.

of pyrans relies on the intramolecular cyclization of hydroxyalkenes, which implies the suitable activation of the olefinwith a palladium catalyst and the subsequent intramolecularattack by the alcohol.

Since the reactivity of palladium catalysts is dominated bytheir oxidation states, Pd(0) and Pd(II) complexes are consid-ered independently. Indeed, Pd(0) complexes are fairly nucleo-philic and the cyclizations of d-hydroxy alkenes leading topyrans progress via the p-allylpalladium cations (TsujieTrostreaction). Otherwise, parallel processes involving Pd(II) com-plexes take advantage of their electrophilic character andcommonly proceed through the fast and reversible formationof p-complexes with the carbonecarbon double bond, whichnext undergo an intramolecular attack by the nucleophile.

5.2.1.1. Catalysis by Pd(0). Palladium(0) complexes catalyzethe cyclizations of d-hydroxy alkenes containing allylic leav-ing groups (X: OAc, OCO2R, OPO(OR)2, OAr, Cl, Br). Asshown in Scheme 30, the initial p-complex evolves througha h3-complex to a p-allyl cation, a highly reactive intermedi-ate that undergoes intramolecular attack of the hydroxyl toprovide the corresponding pyran. The regioselectivity of thiscyclization is rarely a problem, since 6-exo ring closures aremore favored than their competitive 8-endo counterparts.

From a stereochemical point of view, substrate-controlledcyclizations usually proceed with retention of configurationat the reacting center. Coordination of the palladium to thecarbonecarbon double bond occurs on the less-hindered faceopposite to the leaving group and the nucleophile adds tothe cationic p-allylpalladium complex on the opposite faceto the metal. Thus, a net retention of the stereochemistry isobserved.57 Studies on the assemblage of the tetrahydropyranmoiety embedded in zampanolide (79 in Scheme 31) havedemonstrated how highly diastereoselective this strategy canbe. Indeed, Pd(0)-mediated cyclizations of both 80a and 80bafford, respectively, 2,6-cis- and 2,6-trans-tetrahydropyrans81 as a single diastereomer in good yields under very mildconditions.58

Palladium(0) catalysts also offer the opportunity for enan-tioselective cyclizations. Trost et al. have proved that

Page 13: Available Online at Www.sciencedirect.com

HO

OBn

O

OBn+

74 75-cis 75-trans

OMe OMe

XO

OBnOMe

X

OTBDPS OTBDPS OTBDPS

Conditions

X: Br, I, or SePh

Conditions cis-trans Yield (%)Hg(OAc)2, Br2, THF, rt 100:0 71 NBS, DMF, rt 100:0 77 NIS, CH2Cl2, rt 2.3:1NPSP, CSA, CH2Cl2, rt 1:1.8 87

Scheme 28.

OTBS

OH

OH

OH O

OO

CO2Me

OTBS

O OH O

CO2Me

, AIBN, hexane, 76 77

TBSMeO OMe

85%78

X: TIPPSe

X: H

1

TIPPSeBr, DTBMPCH2Cl2, –78 °C, 2 h

74% dr 88:12

XOH

OO

11

15

Scheme 29.

X

OH

Pd(0)L2 X

OH

PdL2

OH

PdL2X

OH

PdL2

O PdL2O–Pd(0)L2

3-complex-allyl cation

X

–HX

Scheme 30.

2695I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

intramolecular asymmetric allylic alkylations59 of hydroxyalkenes can be highly enantioselective, with the nucleophilicaddition of the alcohol to the p-allylic cation being the enan-tio-determining step of the process.60 As a case study, treat-ment of the hydroxy alkene 82 with a palladium catalyst

OBn

BnO

HO

OCO2EtBnO

O5 mol% Pd(PPh3)4THF, rt

71%

OBn

BnO

HO

OCO2EtBnO

O5 mol% Pd(PPh3)4THF, rt

68%

80a

80b

Scheme

generated by mixing Pd2(dba)3$CHCl3, chiral diphosphine83, and Et3N in CH2Cl2 at 0 �C provides the tetrahydropyran84 in 80% yield and 94% ee (Scheme 32). Furthermore, thestereocontrol imparted by the catalyst is absolute on themore elaborated chiral substrate 85, and the tetrahydropyran86-trans or the isomer 86-cis is exclusively formed by switch-ing the configuration of 83.61

High stereochemical control has also been observed ind-hydroxy alkenes containing other allylic leaving groups(Scheme 33). For instance, Pd(0)-catalyzed cyclization of car-bonate 87 affords the tetrahydropyran 88 in high diastereomericratio, whereas 90 is obtained with complete stereocontrol fromester 89 in 96% yield.62,63

5.2.1.2. Catalysis by Pd(II). As previously mentioned, palla-dium(II) catalysts exhibit an electrophilic character that ac-counts for the formation of p-complexes with carbonecarbon double bonds, which easily undergo the addition ofnucleophiles. In this manner, the Pd(II)-mediated intramolecu-lar cyclization of hydroxy alkenes provides a straightforward

O O

O

O

NH

HOO

zampanolide79

OBn

OBn

81-cis

81-trans

31.

Page 14: Available Online at Www.sciencedirect.com

NH HN

PPh2

O

Ph2P

O

83

OH

82

CH2Cl2, 0 °C80% ee 94%

84

DMPS

OAr

2 mol% Pd2(dba)3·CHCl36 mol% 83

Et3N

O

DMPS

OH

85

CH2Cl2, 0 °C60% dr > 97:3

TMS

OAr

86-trans

OPr

O

O

TMS

OPr

O

O

TMS

OPr

O

2 mol% Pd2(dba)3·CHCl3, 6 mol% 83, Et3N

CH2Cl2, 0 °C58% dr > 97:3

2 mol% Pd2(dba)3·CHCl3, 6 mol% ent-83, Et3N

86-cis

Ar: 4-NO2Ph

Ar: 4-NO2Ph

Scheme 32.

OBnHO

OBnOPd(0)

THF, rt

ligandPPh383

ent-83

3:21:2

10:187 88

dr

OAcOHN

O

Cbz

ON

O

Cbz

THF, rt96%

1.3 mol% Pd2(dba)3·CHCl35.3 mol% ent-83

89 90

EtO2CO

Scheme 33.

OHR

HOR

PdII

H

HOR

HPdII

ORPdII

ORPdII

2,6-cis

2,6-trans

Pd(II)

Pd(II)

1,3-axial interactionsδ

Scheme 35.

2696 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

entry to oxygenated heterocycles. The regioselectivity of suchcyclizations is commonly determined by the ring size, whichfacilitates the acquisition of pyrans from d-hydroxy alkenesthrough a 6-exo ring closure (Scheme 34).64e66 The nucleo-philic attack of the hydroxyl on the corresponding p-complexusually occurs anti to the metal to produce a s-alkylpalladiu-m(II) intermediate, which next undertakes a b-hydride elimi-nation. Importantly, Pd(0) is produced in the final step and,consequently, a mild re-oxidant to transform Pd(0) back intoPd(II) without affecting the substrates or products is requiredto obtain a process catalytic in palladium; otherwise, theprocess must be carried out using stoichiometric amounts ofthe Pd(II) complex.67

The rationale for the stereochemical outcome of thesecyclizations usually hinges on the relative steric demands ofthe reversibly formed palladium intermediates.65 Attention isthen focused on the steric interactions engaged in the alterna-tive approaches outlined in Scheme 35. Hence, 2,6-cisarrangements are often preferred, especially in the case whereall the substituents are placed in equatorial positions or clearly

OHPdIILnX2

OH

-complex

Pd(II)LnX2

–HX

Scheme

unfavorable axial interactions are developed on the way to2,6-trans isomers.

The success of such a strategy is intimately related to theregioselectivity of the b-hydride elimination step. Studies onthis issue have proved that it is highly solvent dependent,DMSO being the best choice,68 but, in spite of these accom-plishments, these cyclizations have been scarcely applied tothe synthesis of natural products.69 This limitation has beenovercome through the introduction of a hydroxy group at theallylic position.70 As represented in Scheme 36, the key fea-ture of this process entails the coordination of the allylic alco-hol, a0-OH, to the metal. Then, the initial Pdep-complex isselectively formed on the same face of the double bond asthe a0-OH. This complex determines the approach of the d-OH to the reacting center (syn attack) and, consequently, thestereochemical outcome of the cyclization. Eventually, the re-sultant s-complex undergoes a selective Pd(II) eliminationleading to a new olefinic bond. Hence, the overall sequencemay be considered as a Pd-mediated SN20 process that can pro-ceed without any oxidant.

In this context, a chiral starting material can give access todisubstituted tetrahydropyrans stereoselectively (Scheme 37).

OPdIILnX

O

-alkylpalladium

–Pd(0)Ln

–HX

34.

Page 15: Available Online at Www.sciencedirect.com

OH OH

OH

ClPd Cl

Cl

PdClHO

OH

PdCl2

OH

OClPd

HOO–HCl –Pd(OH)Cl

π-complexes σ-complex

α

α'

δ

α' α'

α α

δδ

Scheme 36.

2697I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

Indeed, chiral diols 91a,b afford 2,6-cis- or 2,6-trans-tetrahy-dropyrans 92 as a single isomer in excellent yields, dependingupon the configuration of the allylic stereocenter, a0, with per-fect 1,3-chirality transfer. According to the mechanistic guide-lines presented in Scheme 37, the p-face selectivity observedin these cyclizations can be rationalized by considering con-formations on the a0-stereocenter, so that the most favoredapproaches avoid 1,3-allylic strain.71

This strategy has been successfully applied to the construc-tion of pyran rings embedded in the structure of laulimalide(93 in Scheme 38). Moreover, it has been subjected to thePd(0)-mediated methodology described in Section 5.2.1 forthe synthesis of the C23eC27 dihydropyran 95. Both the Pdcomplexes catalyze the 6-exo-like cyclization of 94 to 95 asa sole isomer in good-to-excellent yields. Furthermore, the6-endo-trig cyclization of 96 occurs through a syn-SN20 pro-cess to give exclusively the desired 2,6-trans-dihydropyran97 in 60% yield.70a

As shown in Scheme 34, the Pd-mediated intramolecularaddition of a hydroxyl group to an alkene leads to a s-

OH

HO

HOTHF, 0 °C

10 mol% PdCl2(MeCN)2

minimisation of 1,3-allylic strai91a

OH

HOTHF, 0 °C

10 mol% PdCl2(MeCN)2

minimisation of 1,3-allylic strai91b

S

R

S

S

Scheme 37. Note: Pd atoms hav

OPMB

OO

OR

OH

PhMe, 80 °C59%

20 mol% Pd2(dba)3, neocuproine

via Pd(0)-carbonate R: CO2Me

via Pd(II)-hydroxyl R: H

THF, 0 °C89%

10 mol% PdCl2(MeCN)2

94

OPMB

OH

OH

OBz

96

23

27

59 THF, –5 °C60%

15 mol% PdCl2(MeCN)2, benzoquinone

Scheme

alkylpalladium(II) intermediate. This intermediate can thenbe trapped by CO and the resultant acylpalladium species iseasily converted into the corresponding methyl ester (Scheme39). Thus, the overall process is a 1,2-alkoxycarbonylation thatavoids any b-elimination. In the case of chiral centers, the in-sertion of CO into the carbonepalladium bond usually retainsthe configuration that arises from the cyclization step, andso pyran-containing ester groups can be stereoselectively ob-tained as a result of these transformations.72,73

This methodology has been applied to the synthesis of leu-cascandrolide A (1)1a and phorboxazole A (2a).2g As shown inScheme 40, the intramolecular alkoxycarbonylation of diol 98in 1:1 MeOHePhCN (the use of benzonitrile facilitatesa cleaner and more efficient process) provides the desired2,6-cis-tetrahydropyran 99 in 75% yield and dr >10:1. Note-worthy is the fact that the cyclization is highly selective andthe two alcohols and two alkenes in 98 are completely differ-entiated, which simplifies the protecting-group strategy. Fur-thermore, the C22eC26 tetrahydropyran (101) of 2a isprepared in high yield through the Pd(OAc)2-mediated

H

HO

HH72%

n

O

92-cis

79%n

O

92-trans

OHH

H H

OH

e been omitted for clarity.

O

OOOH

O OH

O

laulimalide93

OPMB

OO

O

95

OPMB

O

OBz

97

23

27

23

27

59

38.

Page 16: Available Online at Www.sciencedirect.com

OPdIILnX

-alkylpalladium

O

CO

O PdIILnX

O MeOH–Pd(0)Ln

–HXOMe

O

Scheme 39.

2698 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

cyclization of hydroxy alkene 100, although 2.5 equiv of thePd(II) reagent is needed for complete conversion because ofits reduction by CO during the course of the reaction (acetoni-trile is used to minimize such deactivation).74 In turn, the par-allel ring closure of 102 has been carried out in the presence ofa large excess of p-benzoquinone to afford 2,6-cis-tetrahydro-pyran 103 in 58% isolated yield.

The s-alkylpalladium(II) intermediates can also be trappedthrough a Heck reaction. This possibility is nicely illustratedby a palladium-catalyzed sequence along the synthesis of a-to-copherol (104 in Scheme 41).75 Indeed, the reaction of 105

3

7 OTBDPS

O OH OH

98

MeOH-PhCN 1:1, rt75% dr > 10:1

10 mol% PdCl2, 4 equivCO (1 atm)

MeOH-MeCN 1:1, rt, 3 d86%

100

OTIPS

HO

N

O

OTBDPS

HO

O

NCl

MeOH-MeCN 1:1, rt, 158%

5.5 equiv benzoquinon

102

CO (1 atm)2.5 equiv Pd(OAc)2

10 mol% PdCl2(MeCNCO (1 atm)

22

26

15

11

Scheme

105

10 mol% Pd(OCOCF3)2

CH2Cl2, rt, 3.5 d

84% 96% ee

OH

MeO

p-benzoquinone

107

O

MeOO

OMe

-tocopherol

O

MeO

104

40 mol% 106

PdLn(OCOCFoxypalladatio

-HPdLn(O-hydride e

H2C=CHCO2Me

Scheme

with methyl acrylate in the presence of catalytic amounts ofPd(OCOCF3)2, the chiral ligand 106, and p-benzoquinoneaffords chroman 107 with 96% ee in 84% yield. This two-step sequence is assumed to proceed via the enantioselectiveoxypalladation of 105. The resultant s-alkylpalladium com-plex then undergoes a Heck reaction with methyl acrylateand, finally, a b-hydride elimination delivers the desiredproduct.

A conceptually different strategy to the synthesis of 2,3-di-hydro-4H-pyranones such as 109aec takes advantage of thereactivity of enones. As represented in Scheme 42, it relieson the oxidative cyclization of b-hydroxy enones 108aecwith PdCl2. Thus, this methodology represents a new approachto the enantioselective assemblage of these useful intermedi-ates, based on a 6-endo ring closure of a Pd(II)ep-complexfollowed by a completely regioselective b-hydride elimina-tion, without affecting the integrity of the existingstereocenters.76

3

7

1

OTBDPS

O OH O

CO2Me99

, 2 h CuCl2,

101

OTIPS

O

N

O

CO2Me

de

OTBDPS

O

O

NCl

CO2Me103

2a

)2

22

26

11

15

40.

O

MeO3)2n PdLn(OCOCF3)

O

MeO

OMe

O

PdLn(OCOCF3)

COCF3)limination

OMe

Oinsertion

N

O

N

O106

41.

Page 17: Available Online at Www.sciencedirect.com

OH O

O

O

108a

109a

Ph

O2, 10 mol% PdCl2,10 mol% CuCl,10 mol% Na2HPO4DME, 50 °C

Ph

Ph

OH O

108b

Ph

OH O

108c

O

O

Ph109b

O

O

Ph109c

74% 50%61%

Scheme 42.

Cl2CHCHCl2, 70 °C, 24 h60%

OHBn0.5 mol% [PtCl2(H2C=CH2)]2

1 mol% P(4-CF3Ph)3OBn

5 mol% Sn(OTf)4

MeNO2, Δ, 1 h77%

OH

Ph

O Ph

5 mol% AgOTfCl2CHCHCl2, 83 °C, 15 h

87%

OOH

Ph

Ph

112 113

114 115

116 117

δ

γ

γ

Scheme 44.

O

O

OHHO OH

OMe

HO OH

O

5 mol% Ph3PAuCl5 mol% AgSbF6

wet CH2Cl2, 35 °C, 2 d

118 120

119

97%

2699I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

A related cyclization involving a 6-endo ring closure of ana,b-unsaturated ketone has been used to build the dihydropyr-anyl enone unit of several Alstonia alkaloids, such as alsto-phylline (110 in Scheme 43).77 Although the detailedmechanism is not clear, the process is believed to involvethe formation of a Pdep-complex on 111, removal of thesilyl group, and nucleophilic attack of the deprotected alcoholon the palladium complex. Finally, b-hydride eliminationprovides 110, while the Pd(0) is re-oxidized to Pd(II) byt-BuOOH.

Scheme 45.

5.2.2. Other metal-mediated cyclizationsBesides palladium chemistry, there is a handful of emerging

catalytic methodologies for the cyclization of d- and g-hy-droxy alkenes, based on the activation of the olefin by othermetals, e.g., platinum,78 tin,79 silver,80 or cerium81 (Scheme44). The mechanism that accounts for such processes involvesthe electrophilic activation of olefin according to the guide-lines illustrated in Scheme 19. The regioselectivity is welldefined for d-hydroxy alkenes, whereas ring closure of g-hydr-oxy alkenes 114 and 116 is apparently ruled by the Markovni-kov tenet, in which the stabilization provided by substituentsof the olefin seems to support the formation of tetrahydropyr-ans over their tetrahydrofuran counterparts. Synthetically, thescope of these approaches is still limited and they cannot com-pete with the performance of palladium catalysts, but theyoffer new avenues to the construction of tetrahydropyransand constitute a promising area of research.

Remarkably, gold catalysts can participate in a cascade se-quence that presumably involves a 6-exo ring closure of a hydroxyenone intermediate through a metal activation of the carbonecar-bon double bond.82 Indeed, treatment of homopropargylic ether

8 mol% NNaOAc, NN

Me

MeOMe OTIPS

O

H

H

H

H1:3:3 AcOH-

80 °C

55111

Scheme

118 with Ph3PAuCleAgSbF6 in wet CH2Cl2 gives the enoneintermediate 119 that cyclizes to 2,6-cis-tetrahydropyran 120 inexcellent yield (Scheme 45).

5.3. Acid-mediated cyclizations

Acid-mediated cyclizations of hydroxy alkenes representa particular class of processes, as they are promoted by stoi-chiometric or catalytic amounts of Brønsted acids. Althoughthe mechanistic issues of these cyclizations are rather specula-tive, their regioselectivity is usually controlled by the Markov-nikov tenet and most of them involve strong Brønsted acids.Actually, they are commonly carried out in the presence oflarge quantities of these reagents. This is the situation ob-served during the synthesis of platensimycin (121 in Scheme46),83 in which treatment of a 2:1 mixture of diastereomers122a and 122b with TFA (TFAeCH2Cl2 2:1) gives thecage-like structure 123 in 87% yield, based on the isomer122a present in the mixture. Interestingly, the undesired dia-stereomer 122b is recovered unchanged.84

It also occurs that these cyclizations can be carried outunder catalytic conditions.79,85 For instance, the ring closureof hydroxy vinylsilanes such as 124 to 125 using a catalyticamount of Brønsted acids has been thoroughly studied

a2PdCl4t-BuOOH NN

Me

MeOMe O

O

H

H

H

Halstophylline

110

H2O-dioxane, 6 h

%

43.

Page 18: Available Online at Www.sciencedirect.com

0 °C, 1.5 hTFA-CH2Cl2 2:1

O

NH

OOOH

OHHOOC

platensimycin121

OH

O

O

HO

H

H

122b

122a

O

O

12387%

Scheme 46.

2700 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

(Scheme 47). A plausible mechanism for this cyclization in-volves the formation of a b-silylcarbenium ion intermediateby protonation of the sp2-carbon a to the silyl group and thesubsequent intramolecular attack of the hydroxy oxygen onthe cationic center.86

5 mol% TsOHCHCl3, 60 °C, 4 d

86%

OHDMPS

ODMPS

124 125

Scheme 47.

Eventually, Lewis acids can also be involved in the intra-molecular hydroxyalkoxylation of unactivated olefins, leadingto tetrahydropyrans. These processes presumably proceedthrough the formation of a Lewis acideOH complex thatenhances the acidity of the hydroxyl proton.86,87 Furthermore,polycyclic compounds such as (�)-caparrapi oxide and (þ)-8-epicaparrapi oxide (126 and 127, respectively, in Scheme 48)have been obtained from hydroxy alkene 128 in the presenceof stoichiometric amounts of chiral LBA (Lewis acid-assistedchiral Brønsted Acid) complexes 129 and ent-129. Indeed, theasymmetric ring closure of 128 induced by 129 gives an 81:19mixture of 130 (>99% ee) and 131 (21% ee), whereas the par-allel cyclization induced by ent-129 provides a 14:86 mixtureof 130 (27% ee) and 131 (98% ee). Eventually, optically pure130 and 131 are isolated in 74 and 73% yield by columnchromatography.88

2 equiv 129

CH2Cl2-PrCl 1:1 –80 °C, 2 d

74% dr 81:19 > 99% ee

OH128

OCOBn

2 equiv ent-129

CH2Cl2-PrCl 1:1 –80 °C, 2 d

73% dr 86:14 98% ee

O

O

Ar

OMe

SnCl4H

Ar: 2-FPh 129

Scheme

6. Allene-mediated cyclizations

Allenes containing a nucleophile can undergo intramolecu-lar cyclizations on treatment with transition metal catalysts.89

In particular, gold-catalyzed cycloisomerization of b- or d-hy-droxyallene gives access to six-membered oxygenated hetero-cycles in a straightforward manner. As occurs for alkenes (seeScheme 19), the initial coordination of the gold catalyst to oneof the terminal double bonds of a b-hydroxyallene such as 132gives rise to the formation of an intermediate, which, uponnucleophilic attack of the oxygen, is converted into a s-goldcomplex through a 6-endo-trig ring closure (Scheme 49).Subsequent protodemetallation of this complex furnishes thecorresponding dihydropyran 133.90 Likewise, d-hydroxyallene134 undergoes a 6-exo-trig hydroalkoxylation to form the cor-responding 2-alkenyl tetrahydropyran derivative 135 in excel-lent yield.91 Remarkably, gold-mediated hydroalkoxylation of134 in the presence of the chiral diphosphine 136 affords 135in 96% yield and 88% ee.92

7. Alkyne-mediated cyclizations

A conceptually related set of approaches based on the elec-trophilic activation of carbonecarbon triple bonds have alsoproved to be a highly efficient entry into the synthesis ofsix-membered oxygenated heterocycles.93 Indeed, the cata-lytic metal-mediated intramolecular cyclizations of precursorscontaining alkynes and different oxygenated nucleophiles canbe exploited for the construction of a wide array of oxygen-containing heterocycles.

Surprisingly, the simplest strategy based on the use of hy-droxy alkynes had been scarcely applied to date, but recentreports on the cyclization of 4- and 5-alkynol have establishedthe synthetic potential of such an approach. In this context,palladium94,95 and iridium96 complexes catalyze the 6-endo-dig ring closure of 2-alkynylbenzyl alcohols with internaltriple bonds to the corresponding isochromenes in good yields.The mechanism for this transformation is outlined in Scheme50. As for alkenes or allenes, initial coordination of the metalto the carbonecarbon triple bond enhances its electrophilicityand facilitates the attack of the hydroxyl. Then, a protodeme-tallation step renders the cyclic enol ether and regenerates thecatalytic species.

(–)-Caparrapi oxide126

(+)-8-Epicaparrapi oxide127

O

H

O

H

O

H

O

H

OCOBn

OCOBn130

131

48.

Page 19: Available Online at Www.sciencedirect.com

OBu

HOBu

5 mol% Ph3PAuCl5 mol% AgBF4

PhMe, rt, 1 d32%

132 133

HOBu

LnAu

OBu

LnAu

H6-endo-trig

HO

134

PhPh

5 mol% Au[P(t-Bu)2(o-biphenyl)]Cl5 mol% AgOTs

PhMe, rt, 5 min HO

PhPh

LnAu

6-exo-trigO

98%

135

MeOPAr2MeOPAr2

Ar: 3,5-(t-Bu)2-4-MeOPh

2.5 mol% Au2[136]Cl2, 5 mol% AgOTs

PhMe, –20 °C, 18 h

96% 88% ee

136

β

δ

Scheme 49.

MLn

R

OH

R

OH

LnM

O

LnMR

HO

HR

O H

LnMR

O

HR

–MLn

–MLn5-exo-dig

6-endo-dig

Scheme 50.

2701I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

As shown in Scheme 50, an alternative pathway involvinga 5-exo-dig ring closure threatens the success of such method-ology. A comprehensive analysis of the regioselectivity of thiscyclization for palladium catalysts has revealed that the6-endo-dig is favored over the 5-exo-dig mode by alkyl, ratherthan aryl, substitution on the triple bond, low solvent polarity(dioxane), high temperatures (80e100 �C), and dilute condi-tions.94,97 Keeping these limitations in mind, both palladiu-m(II) and iridium(III) catalysts provide a good entry intoisochromenes 138a,b from 2-alkynylbenzyl alcohols 137a,b(Scheme 51).

Working in the 4-alkynol arena, a tandem palladium(II)-catalyzed reaction offers an appealing entry into the synthesisof dihydropyrans.98 As represented in Scheme 52, the palla-dium-catalyzed addition of a terminal alkyne to a hydroxy

OH

Bu

dioxane, 80 °C, 3 h

2 mol% PdI2, 4 mol% KI

63%137a

O

138a

Bu

OH

Pr

CHCl3, 80 °C, 2 h

4 mol%

77%137b

O

138b

Pr

OIr

PPh3

PPh3

HO

Scheme 51.

ynoate initially affords a conjugated adduct, which undergoesa 6-endo-dig cyclization. Remarkably, the resultant dihydro-pyran contains an exocyclic carbonecarbon double bondwith a well-defined geometry (for an outstanding applicationof this palladium-based tandem synthesis and a related ruthe-nium process, see Scheme 89).

Furthermore, a palladium(II)-catalyzed 6-endo-dig cycliza-tion of b-hydroxy ynones, similar to that outlined in Scheme42, has recently been disclosed.99 Differently, the s-alkenylpal-ladium intermediate obtained in this reaction undergoes inser-tion of ethyl acrylate to form trisubstituted dihydropyranonesfeaturing an alkenyl substituent a to the carbonyl (Scheme53). Thus, the overall process entails a cascade WackereHeck reaction that involves the chemoselective coupling oftwo electron-deficient reactants. This WackereHeck sequencedoes not affect the stereochemical integrity of the starting b-hy-droxy ynones 139a,b and affords the corresponding dihydro-pyranones 140a,b in good yields.

Otherwise, 5-alkynols with a terminal triple bond undergocompletely regioselective 6-exo-dig ring closures to exocyclicenol ethers in the presence of palladium,100 iridium,101 plati-num,102 or gold102,103 catalysts (Scheme 54). Their mechanismis also based on the electrophilic activation of the alkyne ac-cording to the guidelines previously described, but, differently,the competitive 7-endo-dig pathway is never observed.

Interestingly, the resultant cyclic enol ethers can be the start-ing point for further transformations. For instance, treatment of

Page 20: Available Online at Www.sciencedirect.com

CO2R1

R

H

+

10 mol% Pd(OAc)24 mol% TDMPPPhH, rt, 2–3.5 d

HO

CO2R1

R2

OH R2RO

CO2R1

R2R

R: alkyl, PhR1: Me, Et R2: alkyl

50–60%

Scheme 52.

Ph

OH O

Ph139a

Ph

OH O

139b

O Ph

O

Ph

CO2Et

O

OCO2Et

Ph

20 equiv H2C=CHCO2Et, O210 mol% [PdCl2(CH3CN)2]

Cu(OAc)2·H2O, PPh3, 20 mol% LiBrDME, 20 h, 65 °C

140a 47%

140b 47% de > 99%

PdLnX2R1

OH O

R3 OH

O

R3R1

PdLnX2

–HXO

O

R1 R3

PdLnX

O

O

R1 R3

CO2EtCO2Et

Wacker Heck

R2

R2 R2 R2

Scheme 53.

H

OH

H

OH

MLnMLnO

–MLn

OMLnH

6-exo-dig

Scheme 54.

OH

RO

143a R: THP143b R: TBS OR

HO

144a R: THP144b R: TBS

OO

O

O

1) 1 mol% [PtCl2(CH2=CH2)], Et2O, rt, 30 min2) 5 mol% CSA, MeCN-H2O, rt, 4 h

3) MgSO4, rt, 5 h

+

145 146

Alkynol145:146Yield (%)143a

143b

144a

144b

30:120:19:1

3.7:1

75836058

Scheme 56.

2702 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

alkynols 141 (Scheme 55) with platinum or gold catalysts leadsto the enol ethers, which undergo a Prins-type cyclizationthrough an oxocarbenium intermediate that eventually affordsthe eight-membered carbocycles 142 as a sole diastereomerin excellent yields, irrespective of the catalyst.102

Internal alkynols can also participate in highly regioselectivecycloisomerizations, provided that the resultant enol ethers aresatisfactorily trapped by other nucleophiles. For instance, plat-inum(II)-catalyzed hydroxyalkoxylation of alkynols 143a,band 144a,b deliver selectively spiroketal 145 through initial6-exo-dig or 6-endo-dig ring-closure modes, respectively(Scheme 56).104e106

A similar gold-catalyzed 6-exo-dig cyclization plays a cru-cial role in the synthesis of the trioxadispiroketal of azaspira-cids (62 in Scheme 24). As represented in Scheme 57 (see also

2 mol% MLn

H

OR

H

141

OR

H OR

MLn: PtCl2(cod), PtCl4, AuCl3 R: Me, Bu, i-Pr, t-Bu, allylR1OH: solvent (MeOH, EtOH, MeCOOH, EtCOOH)

Scheme

Scheme 91), the assemblage of the AeD rings starting fromalkynol 147 relies on the addition of the C6 hydroxy groupacross the C10eC11 alkyne followed by the engagement ofthe C13 ketal exocyclic oxygen atom to form the desired spi-roketal.107 Remarkably, treatment of 147 with catalyticamounts of AuCl and PPTS furnishes 148 in 75% yield.

A closely related methodology takes advantage of the hy-droxyl-like character of hemiacetals. Yamamoto et al. have re-ported that the simultaneous activation of a carbonyl and

R1OH

OR1

R O

142

88–96%

R O

55.

Page 21: Available Online at Www.sciencedirect.com

O

O

HMeO

OH

8 mol% AuCl, PPTSMeOH, rt, 20 min

O OO

OC D

A

B

OPMB

147

148

75%

O

O

HMeO

OAu

H

O OO

O

PMBO

Me

–Au+H

MeO

H

–H , –Me2O

PMBO

OTBDPS

PMBO

OTBDPS OTBDPS

OTBDPS

610

11

62

Scheme 57.

149 150

O

OR1

R

dioxane, rt, 0.5–1 h

74–90%

R: alkyl, aryl R1: Me, Et, i-Pr

Pd(OAc)2

O

H

R

Pd(OAc)2

R1OHOH

OR1

R(AcO)2Pd

O

OR1

RPdOAc

–AcOH

AcOH–Pd(OAc)2

5 mol% Pd(OAc)2,2 equiv R1OH, benzoquinone

O

Ph

O

Ph

Ph

O

79%

151

152

85%

SnBu3

BF3·OEt2

Ph

OTMS

BF3·OEt2

O

H

R

Scheme 58.

2703I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

a carbonecarbon triple bond by palladium(II) catalysts allowsthe construction of cyclic alkenyl ethers from acetylenic alde-hydes. Indeed, cyclization of 2-alkynylbenzaldehydes 149 pro-ceeds smoothly in the presence of Pd(OAc)2 to provide acetals150 in high yields (Scheme 58).108 The authors proposea mechanism in which the palladium catalyst initially acts asa Lewis acid, promoting the formation of the correspondinghemiacetal. Then, coordination of the metal to the alkyne fa-cilitates the anti attack of the hydroxyl from the side oppositeto the metal via a 6-endo-dig pathway.x The resultant vinylpal-ladium intermediate is subsequently protonated to afford thefinal alkenyl acetal with complete regioselectivity. Noteworthyis the fact that the high reactivity of the benzylic acetal in 150(R: Ph, R1: Me) makes possible the introduction of carbonnucleophiles via HosomieSakurai allylation and Mukaiyamaaldol reactions to afford 151 and 152 in high yields.109

Other oxygenated nucleophiles such as carboxylic acidderivatives and carbonyls have also been utilized in this typeof cyclization.93a,110 A step forward in this area relies on theelectrophilic activation of the carbonecarbon triple bond of

x It is worth pointing out that the aromatic ring is crucial to the observed re-

gioselectivity. The reaction of the equivalent carbon-tethered acetylenic alde-

hydes, 4-alkynals, mainly yields the corresponding five-membered acetals

through 5-exo-dig cyclizations.

2-alkynylbenzaldehydes and related systems. Now, the nucle-ophilic attack of the carbonyl oxygen on the electronicallydeficient alkyne produces a 2-benzopyrylium-like intermedi-ate, which can undergo further transformations that ultimatelyoffer a rich assortment of structures.111e114

This methodology has been applied to the synthesis of (�)-S-15183a (153 in Scheme 59), a member of the azaphilones,a family of natural products containing a highly oxygenatedbicyclic core and a quaternary stereocenter.115 The reportedsynthesis takes advantage of a gold-mediated cycloisomeriza-tion of 2-alkynylbenzaldehyde 154 followed by oxidation ofthe resultant 2-benzopyrylium salt. As outlined in Scheme59, a plausible mechanism for this transformation entails thecoordination of the gold catalyst to the triple bond followedby cyclization and protodemetallation to furnish the 2-benzo-pyrylium salt. Eventually, oxidation of this salt affords in 84%overall yield the desired oxygenated heterocycle 155 as a race-mic mixture.

Having developed a straightforward route to the bicycliccore, attention was focused on the asymmetric oxidation ofthe benzopyrylium salt. However, all attempts were unsuccess-ful and a new strategy, based on the initial oxidation of thearomatic ring, was disclosed. Remarkably, the Cu2[(�)-spar-teine]2O2-mediated enantioselective oxidative dearomatizationof 154 provides a hydroxy derivative, which undergoes the de-sired cycloisomerization under very mild conditions to afford(�)-155 in 98% ee and 84% overall yield (Scheme 60). It is

Page 22: Available Online at Www.sciencedirect.com

HO

OH

R

O

H

O

RO

OHO

O

RO

ORCOO

1) 5 mol% Au(OAc)3ClCH2CH2Cl–TFA 10:1, rt

HO

OH

R

O

H

LnAuHO

OH

O

RLnAu

HO

OH

O

RH

2) IBX, 5 mol% TBAI, rt

84%

RCOCli-Pr2NEt, DMAP

CH2Cl2, rt

61%

(±)-S-15183a(±)-153

154 (±)-155

AuLn

R: –(CH2)6Me

TFA–AuLn

CF3COO

oxidation

Scheme 59.

CH2Cl2, –78 to –10 °C

O

OHO

O

RO

O

O2, 2.2 equiv Cu(MeCN)PF6

2.4 equiv (–)-sparteinei-Pr2NEt, DMAP

MeCN, rt

84% 98% ee(–)-155

154

OH

R KH2PO4–K2HPO4buffer pH 7.2

HO

Scheme 60.

2704 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

worth emphasizing that the 6-endo-dig ring closure is carriedout without any metal at pH 7.2.116,117

Other systems available for this chemistry are 4-propargyl-1,3-cyclopentanediones 156 represented in Scheme 61. Metal-mediated cyclization of these 1,3-diketones admits catalystsbased on palladium,118 platinum, tungsten, and mercury.119

Noteworthy is the fact that the platinum-mediated catalysisof terminal and substituted alkynes affords exclusively thebicycles 157 in high yields at room temperature. These6-endo-dig cyclizations are supposed to proceed under kineticcontrol by coordination of the catalyst to the triple bond. Then,the alkyneemetal complex undergoes anti attack of the car-bonyl oxygen and the resultant vinylplatinum intermediate isprotonated to afford the oxabicycle 157. Since related 1,3-cy-clohexadienones are prone to 5-exo-dig ring closures, the ex-cellent regioselectivity observed for 156 is presumably dueto subtle intrinsic strain on the way to the exo- or endo-bicyclicsystems.

As well as these catalytic processes, there are some stoi-chiometric metal-free methodologies based on the iodine-mediated electrophilic activation of the substituted carbonecarbon triple bond of 2-alkynylbenzaldehydes that allow forthe highly efficient construction of six-membered oxygenated

OO

ROO

RTHF–dioxane 2:1

10 mol% PtCl2

76–85%

156 157

R: H, Me

OHO

R

PtCl2

PtCl2

6-endo-digO

HOR

PtCl2

–PtCl2

rt, 24–72 h

Scheme 61.

heterocycles120,121 and other carbocycles.122 For instance,addition of the iodonium ion generated from Ipyr2BF4/HBF4

to the triple bond of 2-alkynylbenzaldehydes 149 leads to aniodobenzopyrylium salt, which, in turn, reacts with nucleo-philes and offers a straight and useful access to highly elabo-rated oxygenated heterocycles 158e161 (Scheme 62).120

Importantly, most of these transformations can be carried outfollowing a procedure based on the use of iodine.121

8. Other metal-promoted cyclizations

In addition to the methodologies just described, metals canalso catalyze other cyclizations that require the in situ gener-ation of transition metalecarbenoid species. Such transforma-tions encompass the cycloisomerization of 4-alkynols viametal vinylidene complexes and the intramolecular additionof oxygenated nucleophiles to metal carbenes obtained fromdiazo precursors.

This chapter also takes into account a less studied transfor-mation: the intramolecular etherification of aryl and vinylhalides catalyzed by palladium and copper.

8.1. Cyclizations of 4-alkynols

The transition metal-mediated isomerization of hydroxyalkynes containing a terminal carbonecarbon triple bond rep-resents a highly valuable route for preparing 3,4-dihydro-2H-pyrans (see Eq. 5 in Scheme 63).123 Unlike the methodologiespreviously discussed, these processes proceed through a metalvinylidene intermediate that undergoes intramolecular attackby an alcohol. Thus, the key element of such cyclizations relieson providing metal vinylidene intermediates reactive enough tosustain a catalytic process. These species arise from the initial

Page 23: Available Online at Www.sciencedirect.com

149

CH2Cl2, 0 °C to rt

R: alkyl, aryl, alkenyl

O HR

IIpyr2BF4, HBF4

OPh

I

O OMePh

I

O PhPh

IO

OPh

I

NMe2

O HR

TMS

PhOTMS

PhNMe2

MeOH

158

159

160

161

63%

45%

55%

42%

Scheme 62.

MLnOH

H O

alkynol metal vinylidene

metal carbene

3,4-dihydro-2H-pyran

OR

N2 OR

MLnMLnO

diazo precursor

(5)

(6)–N2

H

MLnHO

tetrahydropyran

Scheme 63.

MLnOH

H

O

OMLn

O

–MLn

-complex

OHH

MLn

HO•MLn

vinylidene intermediate

H

O MLn O MLnor

–MLn

Scheme 65.

2705I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

coordination of a metal to a terminal alkyne (Scheme 64).Then, the resultant p-complex undergoes either a 1,2-hydro-gen migration over the triple bond or an oxidative additionof the C(sp)eH bond to the metal center and subsequent1,3-shift of the hydride to the alkyne ligand.124,125

The a-carbon in the Cb]Ca]MLn moiety exhibits elec-trophilic properties and undertakes additions of nucleophilesthat eventually lead to formal anti-Markovnikov products.Importantly, the intramolecular attack of the hydroxyl on theCa-center of the vinylidene intermediate derived from 4-alkynols must override a competitive pathway arising fromthe simple addition to the activated p-complex (Scheme 65).Thus, the structural elements on the substrate, the catalyst,and the reaction conditions have to be carefully chosen toavoid such a 5-exo-dig ring closure. These conditions arerather stringent and few metals form vinylidene intermediatesthat are able to take part in catalytic sequences. To date, thislimited group includes rhodium,126 ruthenium,127 and mostly,tungsten complexes.123,128

Ruthenium(II) complexes [CpRuCl(PR3)3] catalyze thehighly selective synthesis of lactones and cyclic enol ethers,

HR HRMLnMLn

-complex

oxidative addition

Scheme

depending on the phosphine and the amount of N-hydroxysuc-cinimide employed (Scheme 66).127a Indeed, electron-richarylphosphines such as P(4-MeOPh)3 and a large excess ofN-hydroxysuccinimide produce six-membered lactones,whereas electron-poor arylphosphines such as P(4-FPh)3 andsmall amounts of the sodium salt of N-hydroxysuccinimideprovide the corresponding dihydropyrans. Unfortunately,none of these conditions can be applied to substrates bearingoxygenated substituents at the propargylic position. Despitethis limitation, the efficiency of ruthenium(II) catalysts hasbeen proved for the assemblage of structurally complex fusedpolyethers such as 163e166 from 162.127b

Tungsten-like chemistry has attracted much attention.123

The most common tungsten complex used for the cycloisome-rization of 4-alkynols is W(CO)6, which is photochemicallyactivated in situ to generate the true catalytic species, presum-ably W(CO)5.129,130 The cyclization is routinely carried out

MLnRH H

MLn

R

1,3-hydride shift

1,2-hydride shift

64.

Page 24: Available Online at Www.sciencedirect.com

O

BnO

DMF, 85 °C, 20 h60%

162

BnO

OHH

40 mol% P(4-MeOPh)3, 30 mol% Bu4NPF6

10 mol% CpRuCl[P(4-MeOPh)3]3

NOH

OO , 2 equiv NaHCO36 equiv O

BnO

163

BnO

O O

H H

H

20 mol% P(4-FPh)3, 15 mol% Bu4NPF6

5 mol% CpRuCl[P(4-FPh)3]3

NONa

OO0.5 equiv

DMF, 85 °C, 26 h70%

O

BnO

164

BnO

O

H H

H

O

BnO

165

BnO

O

H H

H

H

H

OHH

O

BnO

166

BnO

O

H H

H

H

H

Oas previously

65%

Scheme 66.

2706 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

with 5e25 mol % W(CO)6 and an excess of a tertiary amine(DABCO often gives better results than Et3N) in THF or tolu-ene under constant irradiation at 350 nm. Remarkably, thesetungsten-based catalysts tolerate substituents at the propargylicposition and are compatible with a wide variety of functionalgroups such as alcohols, ethers, amides, and carbamates.131

This methodology has been successfully applied to the con-struction of the trisaccharide component of digitoxin (167 inScheme 67). Thereby, iterative alkynol cycloisomerizationbased on W(CO)6 constitutes the most important feature ofthe synthetic sequence that provides glycal 171 and eventuallyallows for the total synthesis of digitoxin.132

In addition to W(CO)6, a preformed THF$W(CO)5 complexhas been employed.133,134 Moreover, a recent communicationhas disclosed a new catalytic system that does not require pho-tochemical activation, based on (methoxymethylcarbene)pen-tacarbonyl tungsten (172 in Scheme 68).135 The optimal

O OH H

TBSOTBSO

OH

OTBS H

THF, h , 65 °C

96%TB

1) 25 mol% W(COTHF, h , 6

2) Ac2O, Et3N, CH

81%

O OH H

HOHO

O

OTBS

O

H

HO

TBSO

HO

OHO

HO

O O

HO

O O

HO

25 mol% W(CO)6DABCO

168

171

Scheme

results are obtained with 25 mol % of oxacarbene 172 in thepresence of 10 equiv of Et3N by simple warming to 40 �C inTHF. This experimental procedure has been successfullyapplied to the synthesis of the altromycin disaccharide 173via 174 and 175.

8.2. Cyclizations of diazo compounds

Metal carbenes are easily obtained from diazo compoundsin the presence of copper and rhodium complexes. Theaccepted mechanism for this transformation combines theLewis acid character of the metal complex and the basicproperties of the carbon center of the diazo precursor. Hence,formation of an acidebase adduct is followed by back dona-tion of electron density from the metal with concomitant lossof nitrogen, thus producing the metal carbene (Scheme69).136,137

O OH H

SOTBSO

O

OTBS

O OH H

HOHO

O

OTBS

O

H

HOH

TBSO

)6, DABCO 5 °C

2Cl2H

H

H

OH

O O

digitoxin167

169

701

67.

Page 25: Available Online at Www.sciencedirect.com

O

OMOMOTIPS

O

O

CbzHNO

OMOMOMe

O

O

Me2NO

OMOMOTIPS

O

HO

CbzHN

H 25 mol% 172

10 equiv Et3NTHF, 40 °C, 12 h

83%

173174 175

Et3N + W(CO)6(Et3N)W(CO)5Et3N +MeO

W(CO)5

CO

172

h

OMe

Scheme 68.

N2

R2

MLn R1

R2

N2

MLn –N2

MLn

R2

diazo compound metal carbene

R1 R1

Scheme 69.

2707I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

Stabilized diazo compounds, particularly a-diazo ketones oresters, are suitable precursors for metal carbenes that exhibitelectrophilic properties at the carbon center, which allowthem to undergo attack of nucleophiles with the eventual releaseof the metal. In particular, catalytic conversion of the diazogroups into the corresponding metal carbenes followed by intra-molecular reaction with alcohols, ethers, or carbonyls affordsa wide variety of oxygenated heterocycles (see Eq. 6 in Scheme63). As the reacting roles are clearly defined, the regioselectiv-ity of such cyclization is not at all troublesome and the onlyconcern lies in the configuration of the resultant stereocenters.

The use of alcohols as the nucleophilic source for thesecyclizations has been seldom exploited for the synthesis ofpyran-like structures. Probably, more synthetic applicationsare lacking because the mechanism of the process is not fullyunderstood. Studies on the insertion of metal carbenes intoa polar OeH bond suggest that this transformation proceedsthrough a concerted pathway, as for parallel reactions withnonpolar CeH or SieH bonds (Scheme 70).136,138 However,evidence supporting a stepwise mechanism has also beenreported.139 The latter model entails the formation of an oxo-nium ylide, either free or associated to the metal, followed bya rapid [1,2]-hydrogen shift.

This methodology has been applied to the synthesis of2-deoxy-b-KDO (176 in Scheme 71). Metalecarbene ring

MLn

–MLn

–MLn

concerted insertion

stepwise insertion

associatedoxonium ylide

freeoxonium ylide

OH

MLnOH

MLnOH

OH

O H

Scheme 70.

closure of the polyoxygenated a-diazo ester 177 produces2,6-trans tetrahydropyran 179a as the sole diastereomer inalmost quantitative yield when the reaction is carried outwith Rh2(OAc)4 in benzene at room temperature.140 Other-wise, a 4:1 mixture of epimers 179a and 179b is obtainedwhen benzene is replaced by chloroform or wet benzene.The rationale for this behavior assumes that the intramolecularattack proceeds on two alternative chair-like conformations178a and 178b, the population of which is strongly solventdependent.141

Ethers can also react with metal carbenes via an oxoniumylide intermediate to give six-membered oxygenated hetero-cycles. Although it is still unclear whether an associated orfree oxonium ylide is the true intermediate for such transfor-mation, it is generally accepted that these species can evolvethrough a [1,2]- or [2,3]-shift for allyl ethers (Scheme 72).Irrespective of the pathway, copper catalysts usually providebetter results than their rhodium counterparts, presumably be-cause they favor ylide formation over competitive CeH bondinsertion.

In this context, copper-catalyzed cyclizations of a-diazoketones containing an allyl ether have already been appliedto the synthesis of natural products.142,143 As an example, ste-reoselective oxonium ylide [2,3]-rearrangements have beenused in an iterative approach for the construction of fused pol-yethers.142 As shown in Scheme 73, treatment of allyl a-diazoketone 180 with Cu(tfacac)2 provides the bicyclic compound181 in 80% yield and excellent diastereomeric ratio. Epimeri-zation of 181 to the more stable equatorial isomer needed foreventual iteration and reduction of the resulting mixturedelivers the alcohol 182 in 86% yield. Then, further func-tional-group manipulations afford the a-diazo ketone 183,ready for a new cyclizationeepimerizationereduction se-quence leading via 184 to polyether 185.

The use of carbonyls as oxygenated nucleophiles for theintramolecular addition to metal carbenes has received muchattention. Attack of the oxygen atom of the C]O bond ona metal carbene produces a transient carbonyl ylide that canparticipate in 1,3-dipolar cycloadditions (Scheme 74). Thus,ring closure based on the nucleophilic addition of a carbonylto a metal carbene represents the first step in a cascadesequence that gives access to structurally complex bicyclicsystems. In this context, rhodium complexes afford higheryields than the copper adducts, because they enable diazo de-composition under milder conditions. Thus, cyclization

Page 26: Available Online at Www.sciencedirect.com

OH

CO2Et

N2

R

OO Rh2(OAc)4

96%OH

RhLn

CO2Et

ORO

O

OO

RHH

CO2Et

OHCO2Et

RhLn

ORO

O

OO

RHH

CO2Et

OHH

CO2H

HOOH

HOHO

178a

178b

179a

179b

2-deoxy- -KDO177

176OOR:

PhH, rt15 min

Scheme 71.

O

OH

H

N2

O

5 mol% Cu(tfacac)2CH2Cl2, reflux

80% dr 30:1180

O

OH

H O

181

1) DBU2) LiAlH4

86%O

OH

HOH

182

O

OH

HO

183

O

N2

5 mol% Cu(tfacac)2

CH2Cl2, reflux

80%O

OH

HO

184

OH

H

H

HO

OH

HO

185

OHH

H

1) DBU2) LiAlH4

86%

Scheme 73.

MLn [1,2]–shift

[2,3]–shift

OR

MLnOR

OR

O R

MLnO MLnO O O

Scheme 72.

MLnO MLnO OR R A BR A BO

R

carbonyl ylide 1,3-dipolar cycloadditioncyclization

Scheme 74.

2708 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

methodologies that take advantage of the carbonyl ylide reac-tivity currently employ rhodium(II) catalysts.144e146

The power of the carbonyl ylide cycloaddition methodol-ogy for the rapid assemblage of bicyclic systems has beenproved in the stereoselective synthesis of structurally complexnatural products such as zaragozic acid C147 and polygalolideA148 (186 and 187, respectively, in Fig. 4).149

The first example entails the intermolecular trapping of thecarbonyl ylide intermediate from a-diazo ester 188 with 3-bu-tyn-2-one to construct the cycloadduct 189 as a single diaste-reomer in 72% yield (Scheme 75).147 The second example

deals with an intramolecular version of such a cascade se-quence from a-diazo ketone 190, in which a simple olefinacts as dipolarophile to obtain diastereoselectively the tricyclicadduct 191 in 73% yield.148 Surprisingly, the synthetic mate-rial shows a higher specific rotation than the samples obtainedfrom natural sources, which suggests that polygalolide Amight be biosynthesized through a poorly stereoselectivepathway.150

Since the cycloaddition step can be faster than metaldecomplexation from the carbonyl ylide, enantioselective tan-dem sequences can also be achieved by positioning chiralligands on the rhodium catalyst.151 This approach has beenapplied to the synthesis of pseudolaric acid A (192 in Scheme76).152 Indeed, a substrate-controlled cascade from the a-diazoketone 193 in the presence of achiral rhodium catalysts isfound to favor the formation of the oxatricycle 194a, whereasa chiral Rh2[(S)-bptv]4 catalyst smoothly delivers a mixturecontaining the desired diastereomer 194b as the major compo-nent in 82% overall yield.

Moreover, the stereochemical outcome of these transforma-tions can also rely on chiral Lewis acids affecting the cycload-dition step. Although the scope of these transformations issometimes narrow, significant levels of enantioselectivityhave been reported for the reaction of carbonyl ylides arisingfrom a-diazo ketones 195 with substrates able to form chelateswith chiral Pyboxelanthanide triflate catalysts (Scheme77).153

Page 27: Available Online at Www.sciencedirect.com

O

O

N2

O

5 mol% Rh2(OAc)4

PhCF3, 100 °C, 5 min73%

OTBDPS

OPMB

O OO

PMBO

190191

O

O

OPMB

O187

OTBDPSOTBDPS

t-BuO2COTBDPS

N2 O

TMSO CO2t-Bu

OMOM

O

PhH, reflux, 1 hOR1

R1

TMSO OTBDPS

COMe

R1: CO2t-Bu R2: (CH2)2OMOM

OO

COMe

R1R1

TMSOOTBDPS

R2

72%

186

188 189

OR2

5 mol% Rh2(OAc)43 equiv HC CCOMe

Scheme 75.

MeO

HOO

O

OO

OH

OOHOOC

HOOCHO COOH

OHO

O

PhOAc

Phzaragozic acid C

186

polygalolide A187

Figure 4.

Rh2(OAc)4CH2Cl2, 0 °C

61% dr 3:1O O

N2PMBO

OO

OO

O

OPMB

OH

OO

O

OPMB

OH

3 mol% Rh2[(S)-bptv]4PhCF3, –40 °C

82% dr 1.6:1193

194a

194b 192pseudolaric acid A

OO

RhRhRh2[(S)-bptv]4

iPrH

NO

O

O

O

HOOCOAc

Scheme 76.

2709I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

8.3. Intramolecular etherification of aryl and vinyl halides

Although the coupling of oxygenated nucleophiles witharyl halides is a long-established method for the preparationof aryl ethers, the harsh reaction conditions commonly re-quired to effect this transformation have restricted its syntheticscope. Indeed, there are relatively few examples for theconstruction of oxygen heterocycles through the intramolecu-lar etherification of aryl halides.

This lack of synthetic methodologies was covered a fewyears ago by a procedure based on intramolecular palladium-catalyzed carboneoxygen bond formation.154 The reactionproceeds under mild conditions using bulky and electron-richo-biarylphosphines in the presence of weak bases. Remarkably,the choice of the phosphine turns out to be crucial, since its stericbulk accelerates the reductive elimination step and minimizesthe competitive b-hydride elimination pathway (Scheme 78).

Aryl chlorides and bromides are suitable substrates for thistransformation. Then, primary as well as secondary substrates

can be cyclized in the presence of bulky dialkylphosphino-biaryl ligands such as 203a and 203b (Scheme 79). Moreover,aryl bromides in which the hydroxyl-bearing chain containsstereocenters and additional functional groups such as amines,amides, or esters also afford the corresponding heterocycles inhigh yields without erosion of the optical purity.

The potential of this transformation has been proved in thesynthesis of several heterocyclic compounds with differentbiological activities.154a,155 For instance, the intramolecularpalladium-mediated etherification of the aryl bromide 204generates the chroman ring 205, which is easily transformedinto (S)-equol (206), a metabolite with a high estrogenic activ-ity (Scheme 80).155d

In spite of the success of this palladium-catalyzed cycliza-tion, the pursuit of easy-to-handle reagents and easier proce-dures is fueling the development of new methodologies. Arecent report on a copper-catalyzed intramolecular couplingof vinyl bromides is a good example of this emerging chemis-try.156 As illustrated in Scheme 81, the vinyl bromide 207 is

Page 28: Available Online at Www.sciencedirect.com

CH2Cl2, –10 °C

O

OR

N2O

195

O

OR

O

O

O

MeO

OBnO

O

O

MeO

O CO2Bn

O

MeO

O

NO O

O

O

p-BrPhCH2O

O

NO O

O

201 *

200 *

94% exo/endo 82:18

ee (exo) 96%

60% exo/endo > 1:99ee (endo) 96%

98%exo/endo 93:7ee (exo) 94%

199 * 96%exo/endo 12:88ee (endo) 91%

10 mol% 196-Sc(OTf)3BnOCH2CHO

CH2Cl2, –25 °C10 mol% 196-Sc(OTf)3

MeCOCO2Bn, 10 mol% TFA

CH2Cl2, –10 °C

CH2Cl2, rt

CH2Cl2

Rh2(OAc)4

2 mol%

202 *

NN N

OO

R1

R2R2

R1

196

197

198

R1: i-PrR1: Ph R1: Ph

R2: HR2: HR2: Ph

*absolute configuration has not been determined

20 mol% 198-Yb(OTf)3

ON

OO

10 mol% 197-Yb(OTf)3

ON

OO

Scheme 77.

base

reductive

elimination

-hydride elimination

and

reductive elimination

Pd(0)LnX

OH

PdXLn

OH

PdLn

O

O

O

H

X: Cl, Br

Scheme 78.

2710 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

easily transformed into the tetrahydropyran 208 in the pres-ence of a copper(I) salt and 1,10-phenanthroline.

9. Intramolecular conjugate additions

The venerable Michael reaction has found a fertile field ofapplication in the synthesis of six-membered oxygenatedheterocycles. As outlined in Scheme 82, intramolecular nucleo-philic attack of an alcohol on the electronically deficientC(sp2) or C(sp) center of a,b-unsaturated systems proceedssmoothly through exo and endo ring-closure modes to providethe corresponding pyrans.

9.1. 6-exo Ring closures

The most common procedure for the synthesis of tetrahydro-pyrans based on the Michael reaction involves 6-exo-trig cycli-zations of a,b-unsaturated hydroxy ketones or esters (see Eq. 7

in Scheme 82). These transformations are usually carried out inthe presence of stoichiometric, or even catalytic, amounts ofbase under thermodynamic control, which implies that the con-figuration of the new stereocenter can be predicted from confor-mational analysis of the resultant heterocycle. Therefore,considering that the most stable conformer of saturated six-membered heterocycles adopts the chair form as a rule, it isin general anticipated that these cyclizations mainly pro-vide 2,6-cis-disubstituted tetrahydropyrans.157 As representedin Scheme 83, this methodology has been largely appliedto the synthesis of leucascandrolide A (1),1d,g,j phorboxazoleA (2a),2c spongistatin 1 (209), and other natural products.158,159

However, careful optimization of the experimental condi-tions enables the acquisition of the kinetically favored 2,6-trans diastereomers.160,161 Thereby, cyclization of hydroxyester 210 affords either 211-cis or 211-trans tetrahydropyranunder basic conditions (Scheme 84).160b The all-equatorial211-cis diastereomer is assumed to be the most stable isomer

Page 29: Available Online at Www.sciencedirect.com

X

R

OH

X: Cl, Br R: H, Me

2–3 mol% Pd(OAc)22.5–3.5 mol% 203

1.5 equiv Cs2CO3

O RPhMe, 50 °C, 24 h

72–85%

P(t-Bu)2 P(t-Bu)2

203a 203b

Scheme 79.

Br

2.5 mol% Pd(OAc)22.5 mol% 203a

1.5 equiv Cs2CO3

PhMe, 50 °C, 22 h O

46%

MeO OH

Ar: 4-MeOPh

Ar Ar

204 205

O(S)-equol

206

MeO

HO

OH

pyr·HCl

66%

150 °C, 28 h

Scheme 80.

10 mol% CuI20 mol% 1,10-phenanthroline

2 equiv Cs2CO3

PhMe, , 20 h

60%Br OH

Ph

O207 208

Ph

Scheme 81.

6-exo cyclizations

EWGO

6-exo-trig

6-endo cyclizations

O

6-endo-trig

O

6-endo-dig

O O

O

OH

OH

O

EWGOH

(7)

(8)

(9)

O

6-endo-trig

OH(10)

EWG EWG

Scheme 82.

2711I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

and is predominantly obtained at room temperature and longreaction times. Otherwise, low temperatures and shorter reac-tion times supply 211-trans in good diastereomeric ratios.

The synthesis of (þ)-SCH-351448 (212 in Scheme 85)takes advantage of analogous procedures.162 Indeed, treatment

of a,b-unsaturated ester 213 with a catalytic amount of baseunder equilibrating conditions (THF, 0 �C) provides themore stable 2,6-cis-disubstituted tetrahydropyran 214-cis,whereas low temperatures favor the cyclization to the 2,6-trans derivative 214-trans. Similarly, assemblage of the 2,6-cis-tetrahydropyran 216-cis from a,b-unsaturated ester 215 iscarried out at �35 �C, because the 2,6-trans isomer 216-transpredominates when the reaction is conducted at �78 �C.

Furthermore, the stereochemical outcome of oxa-Michaelcyclizations on a,b-unsaturated esters containing an additionalhydroxy group at the allylic position is strongly influenced bythe protecting group and the alkene geometry.163e166 In thismanner, kinetically controlled cyclization in aprotic basicconditions of a,b-unsaturated esters 217 affords 2,3-cis- or2,3-trans-disubstituted tetrahydropyrans 218, depending on thegeometry of the carbonecarbon double bond. As shown inScheme 86, tetrahydropyran 218-trans is almost exclusivelyobtained from the (Z)-ester 217-Z, whereas the 218-cis isomeris prepared in high diastereomeric ratios from (E)-ester 217-E.In both cases, NaH provides good results, but potassium basessuch as KHMDS are the most convenient regarding bothstereoselectivity and rate.164a,b

It has also been shown that (E)-a,b-unsaturated ester 219cyclizes quantitatively to 2,3-cis-2,6-cis-tetrahydropyran 220(Scheme 87), whereas (Z)-a,b-unsaturated ester 221 pro-vides 2,3-trans-2,6-trans-tetrahydropyran 222a, which canbe equilibrated to the thermodynamically more stable 2,3-cis-2,6-cis 222b.167,168 Theoretical calculations on the intra-molecular Michael addition on all these systems emphasizethe important role of the cation, the conformationalanalysis of the a,b-unsaturated esters, and the reversibilityof the process for the stereochemical outcome of suchcyclizations.

One of the most appealing features of Michael-mediatedcyclizations is that they can participate in cascade se-quences.169,170 This is the case for the double-cascade cycliza-tion triggered by a 6-exo-trig oxa-Michael addition on theadvanced intermediate 224 en route to tetronasin 1 (223 inScheme 88). Indeed, treatment of 224 with KHMDS in tolueneat 0 �C gives exclusively the 2,6-cis-disubstituted tetrahydro-pyran 225 in an impressive 67% yield.171

Moreover, a new route to 2,6-cis-tetrahydropyrans based ona tandem alkyneeenone coupling and a Michael addition hasrecently been disclosed.172 As illustrated in Scheme 89, theeneeyne coupling of b,g-unsaturated ketone 226 and TMS-protected homopropargylic alcohol 227 presumably providesa hydroxy enone, which undergoes a Michael-like cyclizationto afford the C11eC15 2,6-cis-disubstituted tetrahydropyran228 in good yield and diastereomeric ratio.173 Mechanistically,the preference for the 2,6-cis diastereomer seems to arise froma late transition state, in which the relative stabilities wouldplay a significant role. This ruthenium-based sequence is com-bined with a palladium-catalyzed dihydropyran synthesis pre-viously reported (see Scheme 52). In this case, coupling ofterminal alkyne 229 with hydroxy ynoate 230 in the presenceof palladium(II) catalysts delivers the C19eC23 dihydropyran231 in 55% yield. Both processes cover key transformations

Page 30: Available Online at Www.sciencedirect.com

OH OH

OTBS

CO2Et

O OH O

OTBS

O

CO2Et

10 mol% t-BuOK

THF, 0 °C, 10 min

80%dr 9:1

1

OHCO2Et

OTIPS

ON

1.1 equiv NaHMDSTHF, –78 °C, 1 h

87%

2a

O

PMBO

PMBO OO

OPMBO

OPMB

OHO

OPMB

O

PMB

86% dr > 95:5

O

AcO

O OO O

OMeHO

HOAcO

OO

O

O

OHOH

HO

OHH

OH

Cl

spongistatin 1209

1) AcOH, THF–H2Ort, 48 h

2) KOH, MeOHrt, 24 h

TBSO TBSO

O

OTIPS

ON

BocBoc

CO2Et

Scheme 83.

210

OH

OH

S

S

2.2 equiv NaH

MeO O

THF

78% dr 81:19

–78 to 0 °C2.5 h

O

OH

S

S OMe

O

211-trans

1 equiv NaHTHF

61% dr 98:2

–40 °C to rt7 h

O

OH

S

S OMe

O

211-cis

Scheme 84.

OH O

CO2Bn

O O

CO2Bn

O O

CO2Bn

213214-trans 214-cis

10 mol% t-BuOKTHF, –78 °C, 20 min

86% dr 13:1

10 mol% t-BuOK

THF, 0 °C, 10 min

90%

OH

CO2Me

O

CO2Me

O

CO2Me

215216-trans 216-cis

60 mol% t-BuOKTHF, –78 °C

74% dr 12.5:1

60 mol% t-BuOKTHF, from –78 to –35 °C

80%

O

CO2H

OHO

O

O

OH

HO

O O

O

O

NaO2C

OH

(+)-SCH-351448, 212

Scheme 85.

2712 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

for the total synthesis of 232, a ring-expanded analogue ofbryostatin 1.

Ring closure of the putative intermediate represented inScheme 89 is believed to be catalyzed by the ruthenium Lewisacid. Indeed, Michael-mediated cyclizations can also proceedunder Brønsted acid conditions.174 Although poorly exploited,

this sort of reaction has been applied to 6-exo-trig cyclizationson a,b-unsaturated ketones for the synthesis of complex natu-ral products. For instance, deprotection of 1,3-diol in 233 inacetic acid triggers the formation of 2,6-trans-tetrahydropyran234 in excellent yield and moderate diastereoselectivity(Scheme 90);175 reaction of lactol 235 with a stabilized

Page 31: Available Online at Www.sciencedirect.com

OHCO2Me

OBz

O

OBz

CO2Me

H

H

2,3-cis-tetrahydropyran

OH

OBz

O

OBz

CO2Me

H

H

2,3-trans-tetrahydropyran

CO2Me

217-E 217-Z218-cis 218-trans

+

Ester Base Solvent T (ºC) dr (cis-trans)

217-E

217-E

217-Z

217-Z

KHMDSNaH

KHMDSNaH

PhMeTHF

PhMeTHF

–78–7800

9:19:11:201:11

Scheme 86.

t-BuOKTHF, 0 °C

100% dr > 95:5

PGOOH

OPG

CO2R1

PGOO

OPG

CO2R1

219 220

O

Ph

R1:

t-BuOKPhMe-t-BuOH, rt

76% dr > 98:2

PGOOH

OPG

PGOO

OPG

CO2R2

221 222a

O

Ph

R2:

CO2R2 t-BuOKPhMe, 0 °C

100% dr 95:5

PGOO

OPG

CO2R2

222b

PG: TBDPS 2,3-cis-2,6-cis

2,3-cis-2,6-cis2,3-trans-2,6-transPG: TBDPS

26

3

26

3

26

3

Scheme 87.

2713I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

phosphorane provides enone 236, which is cyclized by treat-ment with CSA to furnish 2,6-cis-tetrahydropyran 237 as thesole diastereomer in 73% yield over two steps176 and, eventu-ally, removal of isopropylidene acetal and silicon protectinggroups in the advanced intermediate 238 en route to bryosta-tins with HF leads to the C1eC16 fragment 239 containingtwo 2,6-cis-tetrahydropyrans.177,178

As well as all of these cyclizations, a 6-exo-dig ring-closuremode based on a Michael reaction is, in fact, possible. How-ever, there is no clear example in the literature, because theenone generated in this cyclization process usually undergoesa second Michael addition, as can be appreciated in the syn-thesis of azaspiracids (62 in Scheme 24, see also Scheme57). In this case, the retrosynthetic analysis of the C27eC40domain followed by Forsyth et al. relies on a formal Double

EtO2C

MeO2CHO O

OM

EtO2CO O

OMe

CO2Me

H

H

KHMDS PhMe, 0 °C, 30 min

67% single diastereom

224

225

Scheme

Intramolecular Hetero Michael Addition (DIHMA) based ona 6-exo-dig cyclization followed by a stereoselective 6-exo-trig ring closure.179 In this manner, the acid-mediated doubleintramolecular oxa-Michael addition on dihydroxy ynone240 gives the bicyclic ketal 241 as a single diastereomer inexcellent yield (Scheme 91).180 Similarly, treatment of ynone242 with TBAF promotes silyl ether cleavage and addition ofthe corresponding oxygen atoms upon C28 to provide in highyield the targeted FG-ring system 243 present inazaspiracids.181

9.2. 6-endo Ring closures

Contrary to the Michael-mediated 6-exo-dig cyclizations,parallel 6-endo-dig processes have been identified as a useful

e

er

O OOMe

H

H

tetronasin 1223

HO

O

OO

ONa

88.

Page 32: Available Online at Www.sciencedirect.com

+acetone, rt, 40 h

226 227

[CpRu(MeCN)3]PF6

10 mol%

OH

PMBO

TMS

O OPMB

O

OTBDPSO

OH

PMBO

O

RTMS

O

PMBO

OTMS

OPMB

O

OTBDPSO

22856%dr 9:1

CO2Me

OOOH

H

OTBS

229 230

+

1) 4 mol% Pd(OAc)24 mol% TDMPP

Ph, rt, 1 dO

CO2Me

OTBS

O

O

231

O OMeO2C

MeOOAc

O

CO2Me

AcO

OH H

OH

OH

O

O(H2C)3

55%

2) 8 mol% Pd(OCOCF3)22 d

analogue of bryostatin 1232

19 23

11

15

11

15

19 23

Scheme 89.

O

SEMO

SEMO OO

OBnO

OSEMO

OBn

OHO

SEM

AcOHTHF–H2O, rt, 19 h

95% dr: 2.5:1233 234

O OH

BnO OBn

OH

BnO OBn

O

O

BnO OBn

HO10 mol% CSA

CH2Cl2, rt, 1 h

73%235 236 237

TMS

O

OBn

OOBn OTBSO

OTBS

OO

OH

MeOOBn

OH

1) 56% HF, MeCN, rt, 24 h

2) HC(OMe)3, MeOH, cat PPTS, rt, 3 h

34%238 239 OBn

Scheme 90.

2714 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

tool to gain access to pyranones (see Eq. 8 in Scheme 82). Inparticular, 6-endo-dig cyclizations of o-alkynoylphenolsprovide an appealing entry into the synthesis of the heterocy-clic structures present in many natural products. The

TBDPSO

O OO

24

O

OTBSO

NBoc OTBS

THF

85%

TBA

24

0

2

TBDPSO

OH OHO

6-exo-dig

1) 1CH2C

2) 1

Scheme

selectivity of these reactions turns out to be highly dependentupon the experimental conditions. Indeed, base-catalyzedcyclization of o-alkynoylphenols can proceed through a5-exo-dig or 6-endo-dig pathway to afford, respectively,

O

O

OTBDPS

41

OO

O

H

HO

NBoc

, rtF

2

243

O

O

HO

OTBDPS

O

6-exo-trig

0 mol% CSAl2-MeOH 5:1, rt, 3.5 h

0 mol% TsOHPhH, rt, 4 h

90%

91.

Page 33: Available Online at Www.sciencedirect.com

2715I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

benzofuranones or benzopyranones. According to the experi-mental results, theoretical calculations on a model ketoneshown in Scheme 92 have revealed that the transition statesfor both processes are very close in energy. Nevertheless,the 6-endo-dig product is thermodynamically favored andcan be obtained under equilibrating conditions in the absenceof proton donors.182

This approach has been successfully applied to the synthe-sis of several members of a family of antibiotics containing ananthrapyran core.183,184 As represented in Scheme 93, basictreatment of ynone 244 under aprotic conditions promotesa highly selective 6-endo-dig cyclization to the tetracycliccompound 245 that is further elaborated into the antibioticAH-1763 IIa (246).184a Alternatively, this type of cyclizationcan also be carried out under acidic conditions, but the mech-anism is still unclear.185

Overshadowed by 6-exo-trig cyclizations, the related 6-endo-trig ring-closure mode (see Eq. 9 in Scheme 82) hadtraditionally attracted not much attention. However, the in-creasing interest in the asymmetric synthesis of di- and tetra-hydropyranones has modified this situation, since this type ofcyclization emerges as an appealing alternative to the hetero-

244

Cs2CO3

O

O

O OH

O

OBn

O

O

O

acetone, rt, 30 min

71%

Scheme

OH

O

O

O

O

O

OO

O

O6-endo-dig5-exo-dig

OO

B HH

Scheme 92.

247

5

Cl

O

Cl

OBIpc2

Claulimalide

93

O

O249

Et2O, 0 °C1 h

(+)-Ipc2BCli-Pr2NEt O

BzO

Scheme

DielseAlder reaction. Initial reports on 6-endo-trig cycliza-tions of a,b-unsaturated ketones are due to Paterson and Os-borne.186 In particular, the aldol adducts obtained throughstereoselective reactions from b-chloro vinyl ketones undergoLewis acid-mediated cyclizations that eventually give dihydro-pyranones.187 Application of this methodology to methyl ke-tone 247 and 3-benzoyloxypropanal supplies aldol 248,which is easily converted into the desired dihydropyranone249 (Scheme 94). Further recrystallization provides an enan-tiomerically pure material that has been employed for the totalsynthesis of swinholide A,188a scytophycin C,188b and, morerecently, laulimalide (93 in Scheme 38).188c

Aldol adducts lacking the b-chlorine atom can also takepart in these cyclizations to afford the tetrahydropyranones.A comprehensive survey of the ability of Brønsted and Lewisacids as well as palladium(II) complexes to enhance the elec-trophilicity of enones 250 has identified Amberlyst-15, Al-(ClO4)3$9H2O, and [Pd(MeCN)4](BF4)2 as suitable catalyststo carry out the desired 6-endo-trig cyclization (250a,b to251a,b) under mild conditions (Scheme 95).76b,189 As antici-pated for a thermodynamically controlled reaction, Amber-lyst-15 affords the all-equatorial products after extendedreaction times. Interestingly, the palladium-mediated cycliza-tions are irreversible and give access to the other isomers inexcellent diastereomeric ratios.

Another aldol reaction is at the center of a different method-ology. Aldol compounds arising from b-keto esters can partic-ipate in an acid-mediated Knoevenagel condensation thatgenerates a highly reactive a,b-unsaturated keto ester, whichundergoes a 6-endo-trig cyclization based on a reversibleoxa-Michael addition. Finally, decarboxylation of the resultantsystem provides the 2,6-cis-disubstituted tetrahydropyra-nones.190,191 This methodology has been applied to the prepa-ration of (�)-centrolobine (252 in Scheme 96). The reportedsequence takes advantage of a Mukaiyama aldol reaction to

O

O

OBn

O

O

OH O

O

OH

245 AH-1763 IIa 246

93.

6% 80% ee248

Cl

O OH

OBz

TMSOTf, i-Pr2NEtH2Cl2, –78 °C to rt, 1.5 h

61% (upgrade to >98% ee by recrystallization)

HC(CH2)2OBz78 to –20 °C

20 h

94.

Page 34: Available Online at Www.sciencedirect.com

250a

8 mol% [Pd(MeCN)4](BF4)2

CH2Cl2, rt, 3 hPh

OOHO

PhO

O

H

HO PhHM

78% dr 96:4H

251a

250b

8 mol% [Pd(MeCN)4](BF4)2

CH2Cl2, rt, 3 hPh

OOHO

PhO

75% dr 97:3

251b

OMH

PhHO

H

H

Scheme 95.

O

MeO OH

(±)-centrolobine 252

TMSO

OTMSOMe

OTBS

OHC+

O

OH

O

MeO

MeO2C

253

HO

OH

O

MeO

MeO2C1) 1 equiv Yb(OTf)3CH2Cl2, –78 °C, 3 h2) 2 equiv TFA4-MeOPhCHO

rt, 5 h60%

Scheme 96.

OH O

Ot-Bu

O

Ar

O

O Ar255

256a

1) 10 mol% 254a

PhMe, –25 °C, 24–38 h2) 50 mol% TsOH80 °C, 10–50 h

N

HBnO

N

NH

NH

ArS

254a

N

H

N

OMe

254bNH

NH

ArS

OH O

Ot-Bu

O O

O

1) 20 mol% 254b

5 mol% piperidinium acetatePhMe, 4 Å MS, rt, 2 d2) 30 mol% TsOHPhMe, 80 °C, 9 h77% 80% ee

Ph

H Ph

O 257

flindersiachromanone 256b

65–94% 88–94% ee

Scheme 97.

2716 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

install the first appendage of the tetrahydropyran. Then, addi-tion of anisaldehyde and TFA to the reaction mixture affordsthe tetrahydropyranone 253 and further functional-groupmanipulations deliver (�)-centrolobine.192

Closely connected to this methodology, it has recentlybeen discovered that the chiral thiourea-based catalysts 254,shown in Scheme 97, activate a-alkylidene-b-keto esters255 and trigger intramolecular conjugate additions of phenolto the corresponding flavanones 256. Remarkably, a one-potreaction involving the simple b-keto ester 257, hydrocinna-maldehyde, piperidinium acetate, and chiral thiourea 254bin the presence of molecular sieves in toluene at room tem-perature followed by decarboxylation affords the naturalproduct, flindersiachromanone 256b, in 80% ee and 77%overall yield.193,194

Although 2,6-cis-disubstituted tetrahydropyranones are usu-ally the most stable diastereomers, other elements in the

structure can alter this situation. For instance, thermodynami-cally controlled transannular conjugate addition ona,b-unsaturated ketone 258 affords 2,6-trans-disubstituted tet-rahydropyranone 259, used as an advanced intermediate ina formal total synthesis of (�)-apicularen A (260 in Scheme98). Interestingly, exposure of epimer 258-epi to identical con-ditions yields the same tetrahydropyranone 259.195

In contrast to the above examples, some 6-endo-trig cycli-zations match the model reaction represented by Eq. 10 inScheme 82.196 For instance, chiral dienyl sulfoxides 261have been identified as suitable precursors of dihydropyrans.This process is based on a highly stereoselective conjugatedaddition (the examples shown in Scheme 99 afford a singlediastereomer) followed by subsequent transformations of theresultant allylic sulfoxides 262.197 Interestingly, this methodo-logy accepts additional stereocenters and other substituents onthe double bond of the dienyl sulfoxide.

Page 35: Available Online at Www.sciencedirect.com

Amberlyst-15

O

O

HO

OMeHO

O

O

O

OMeHO

O

O

OMeHO

HOAmberlyst-15

CDCl3, reflux, 18 h90%

selectivity > 10:1

CDCl3, reflux, 23 h78%

selectivity > 10:1258 259 258-epi

OH

O

O

OHHO

(–)- apicularen A260

N

O

H

Scheme 98.

LDASp-Tol

O

ROH

261a

261b

R: PhR: Bu

Sp-Tol

O

RO

H

H262a

262b

84%94%

10 equiv P(OMe)3

PhO

263a

50%

HO

THF, –78 °C to rt3–4 h

MeOH, 50 °C4 d

Scheme 99.

2717I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

10. Conclusions

As this report proves, the apparently simple construction ofpyrans through CeO bond formation is surprising, with an ex-tremely varied array of methodologies ranging from the vener-able SN2- or Michael-based cyclizations to the more recenttransition metal-catalyzed reactions. The scope and syntheticefficiency of such ring-forming methodologies have alreadybeen established along with the total synthesis of a large num-ber of structurally complex natural products, but, in spite ofthese outstanding achievements, the increasing demand formore selective processes to gain access in a straightforwardmanner to six-membered oxygenated heterocycles is fuelingthe development of new concepts and methodologies. Remark-able examples are the recent developments in epoxide-medi-ated cyclizations that allow selective formation of the pyransover the more favored furans or the spectacular simultaneousconstruction of up to three fused tetrahydropyran units in acascade sequence. Moreover, emerging methodologies suchas platinum- or gold-catalyzed cyclizations of hydroxy al-kenes, allenes, and alkynes will undoubtedly play a majorrole in further developments in this field.

Acknowledgements

We are grateful for the financial support from the SpanishMinisterio de Ciencia y Tecnologıa (Grant CTQ2006-13249/BQU), the Generalitat de Catalunya (2005SGR00584), andthe Universitat de Barcelona (ACES-UB2006).

References and notes

1. For total syntheses of leucascandrolide A, see: (a) Hornberger, K. R.;

Hamblett, C. L.; Leighton, J. L. J. Am. Chem. Soc. 2000, 122, 12894e

12895; (b) Kopecky, D. J.; Rychnovsky, S. D. J. Am. Chem. Soc. 2001,

123, 8420e8421; (c) Wipf, P.; Reeves, J. T. Chem. Commun. 2002,

2066e2067; (d) Fettes, A.; Carreira, E. M. J. Org. Chem. 2003, 68,

9274e9283; (e) Paterson, I.; Tudge, M. Angew. Chem., Int. Ed. 2003,

42, 343e347; (f) Williams, D. R.; Plummer, S. V.; Patnaik, S. Angew.

Chem., Int. Ed. 2003, 42, 3934e3938; (g) Crimmins, M. T.; Siliphaivanh,

P. Org. Lett. 2003, 5, 4641e4644; (h) Wang, Y.; Janjic, J.; Kozmin, S. A.

Pure Appl. Chem. 2005, 77, 1161e1169; (i) Su, Q.; Dakin, L. A.; Panek,

J. S. J. Org. Chem. 2007, 72, 2e24; (j) Ferrie, L.; Reymond, S.;

Capdevielle, P.; Cossy, J. Org. Lett. 2007, 9, 2461e2464; (k) Van Orden,

L. J.; Patterson, B. D.; Rychnovsky, S. D. J. Org. Chem. 2007, 72, 5784e5793.

2. For total syntheses of phorboxazole A, see: (a) Forsyth, C. J.; Ahmed, F.;

Cink, R. D.; Lee, C. S. J. Am. Chem. Soc. 1998, 120, 5597e5598; (b)

Smith, A. B., III; Minbiole, K. P.; Verhoest, P. R.; Schelhaas, M.

J. Am. Chem. Soc. 2001, 123, 10942e10953; (c) Pattenden, G.; Gonzalez,

M. A.; Little, P. B.; Millan, D. S.; Plowright, A. T.; Tornos, J. A.; Ye, T.

Org. Biomol. Chem. 2003, 1, 4173e4208; (d) Williams, D. R.; Kiryanov,

A. A.; Emde, U.; Clark, M. P.; Berliner, M. A.; Reeves, J. T. Proc. Natl.

Acad. Sci. U.S.A. 2004, 101, 12058e12063; (e) Smith, A. B., III; Razler,

T. M.; Ciavarri, J. P.; Hirose, T.; Ishikawa, T. Org. Lett. 2005, 7, 4399e

4402; (f) Wang, B.; Forsyth, C. J. Org. Lett. 2006, 8, 5223e5226; (g)

White, J. D.; Kuntiyong, P.; Lee, T. H. Org. Lett. 2006, 8, 6029e6042;

(h) White, J. D.; Lee, T. H.; Kuntiyong, P. Org. Lett. 2006, 8, 6043e

6046.

3. For total syntheses of phorboxazole B, see: (a) Evans, D. A.; Fitch,

D. M.; Smith, T. E.; Cee, V. J. J. Am. Chem. Soc. 2000, 122, 10033e

10046; (b) Li, D.-R.; Zhang, D.-H.; Sun, C.-Y.; Zhang, J.-W.; Yang,

L.; Chen, J.; Liu, B.; Su, C.; Zhou, W.-S.; Lin, G.-Q. Chem.dEur. J.2006, 12, 1185e1204; (c) Lucas, B. S.; Gopalsamuthiram, V.; Burke,

S. D. Angew. Chem., Int. Ed. 2007, 46, 769e772.

4. For a review on the synthesis of phorboxazoles, see: Haustedt, L. O.;

Hartung, I. V.; Hoffmann, H. M. R. Angew. Chem., Int. Ed. 2003, 42,

2711e2716.

5. (a) Kishi, Y. Pure Appl. Chem. 1998, 70, 339e344; (b) Faulkner, D. J.

Nat. Prod. Rep. 2002, 19, 1e48 and references therein.

6. For reviews on the synthesis of tetrahydropyrans, see: (a) Boivin, T. L. B.

Tetrahedron 1987, 43, 3309e3362; (b) Elliott, M. C. J. Chem. Soc., Per-

kin Trans. 1 2002, 2301e2323; (c) Clarke, P. A.; Santos, S. Eur. J. Org.

Chem. 2006, 2045e2053; (d) Tang, Y.; Oppenheimer, J.; Song, Z.; You,

L.; Zhang, X.; Hsung, R. P. Tetrahedron 2006, 62, 10785e10813.

Page 36: Available Online at Www.sciencedirect.com

2718 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

7. For the stereoselective preparation of pyrans in the context of the synthe-

sis of polyether ionophore antibiotics, see: Faul, M. M.; Huff, B. E.

Chem. Rev. 2000, 100, 2407e2473.

8. For the stereoselective preparation of pyrans in the context of the synthe-

sis of marine macrolides, see: Yeung, K.-S.; Paterson, I. Chem. Rev.2005, 105, 4237e4313.

9. For the stereoselective preparation of pyrans in the context of the synthe-

sis of polycyclic ethers, see: (a) Nakata, T. Chem. Rev. 2005, 105, 4314e4347; (b) Inoue, M. Chem. Rev. 2005, 105, 4379e4405.

10. For the stereoselective preparation of spirans, see: Aho, J. E.; Pihko,

P. M.; Rissa, T. K. Chem. Rev. 2005, 105, 4406e4440.

11. For a BaeyereVilliger oxidation of a cyclopentanone en route to phor-

boxazoles, see: Misske, A. M.; Hoffmann, H. M. R. Tetrahedron 1999,

55, 4315e4324.

12. For HDA cycloadditions en route to 1 and 2, see Refs. 1e,2a,b.

13. For the synthesis of pyrans in the context of comprehensive reactivity

reviews, see: (a) Nakamura, I.; Yamamoto, Y. Chem. Rev. 2004, 104,

2127e2198; (b) Deiters, A.; Martin, S. F. Chem. Rev. 2004, 104,

2199e2238; (c) Barluenga, J.; Santamarıa, J.; Tomas, M. Chem. Rev.2004, 104, 2259e2283; (d) Zeni, G.; Larock, R. C. Chem. Rev. 2004,

104, 2285e2309; (e) Muzart, J. Tetrahedron 2005, 61, 5955e6008.

14. For C-glycosidation transformations en route to 1 and 2, see Refs.

1a,e,3a.

15. We have systematically adopted along the review the heterocyclic

nomenclature (and not the carbohydrate-like one), which assigns number

1 to the oxygen atom.

16. Baldwin, J. E. J. Chem. Soc., Chem. Commun. 1976, 734e736.

17. (a) Smith, A. B., III; Zhu, W.; Shirakami, S.; Sfouggatakis, C.; Doughty,

V. A.; Bennett, C. S.; Sakamoto, Y. Org. Lett. 2003, 5, 761e764; (b)

Brenner, E.; Baldwin, R. M.; Tamagnan, G. Org. Lett. 2005, 7, 937e939.

18. (a) Cink, R. D.; Forsyth, C. J. J. Org. Chem. 1997, 62, 5672e5673; (b)

Chakraborty, T. K.; Reddy, V. R.; Reddy, T. J. Tetrahedron 2003, 59,

8613e8622; (d) Hicks, J. D.; Flamme, E. M.; Roush, W. R. Org. Lett.2005, 7, 5509e5512; (e) Chakraborty, T. K.; Reddy, V. R. Tetrahedron

Lett. 2006, 47, 2099e2102.

19. (a) Bolster, M. G.; Jansen, B. J. M.; de Groot, A. Tetrahedron 2002, 58,

5275e5285; (b) Smith, A. B., III; Kanoh, N.; Ishiyama, H.; Minakawa,

N.; Rainier, J. D.; Hartz, R. A.; Cho, Y. S.; Cui, H.; Moser, W. H. J. Am.

Chem. Soc. 2003, 125, 8228e8237; (c) Pichlmair, S.; Marques,

M. M. B.; Green, M. P.; Martin, H. J.; Mulzer, J. Org. Lett. 2003, 5,

4657e4659; (d) Hua, D. H.; Huang, X.; Chen, Y.; Battina, S. K.; Tamura,

M.; Noh, S. K.; Koo, S. I.; Namatame, I.; Tomoda, H.; Perchellet, E. M.;

Perchellet, J.-M. J. Org. Chem. 2004, 69, 6065e6078; (e) Prasad, K.;

Anbarasan, P. Tetrahedron 2007, 63, 1089e1092; (f) Zhou, J.; Snider,

B. B. Org. Lett. 2007, 9, 2071e2074.

20. Occasionally, an SN20 mechanism can also be proposed to rationalize this

type of cyclizations, see: (a) Ravn, M. M.; Peters, R. J.; Coates, R. M.;

Croteau, R. J. Am. Chem. Soc. 2002, 124, 6998e7006; (b) Kang, Y.; Mei,

Y.; Du, Y.; Jin, Z. Org. Lett. 2003, 5, 4481e4484.

21. Williams, D. R.; Kiryanov, A. A.; Emde, U.; Clark, M. P.; Berliner,

M. A.; Reeves, J. T. Angew. Chem., Int. Ed. 2003, 42, 1258e1262.

22. For a recent account on the Nicholas reaction, see: Dıaz, D. D.;

Betancort, J. M.; Martın, V. S. Synlett 2007, 343e359.

23. Betancort, J. M.; Martın, T.; Palazon, J. M.; Martın, V. S. J. Org. Chem.

2003, 68, 3216e3224; See also: Dıaz, D.; Martın, T.; Martın, V. S. Org.Lett. 2001, 3, 3289e3291.

24. Crisostomo, F. R. P.; Martın, T.; Martın, V. S. Org. Lett. 2004, 6, 565e

568.

25. Crisostomo, F. R. P.; Carrillo, R.; Leon, L. G.; Martın, T.; Padron, J. M.;

Martın, V. S. J. Org. Chem. 2006, 71, 2339e2345.

26. (a) Nicolaou, K. C.; Prasad, C. V. C.; Somers, P. K.; Hwang, C.-K. J. Am.

Chem. Soc. 1989, 111, 5330e5334; (b) Nicolaou, K. C.; Prasad, C. V. C.;

Somers, P. K.; Hwang, C.-K. J. Am. Chem. Soc. 1989, 111, 5335e5340.

27. For theoretical studies on the acid-mediated epoxy alcohol cyclizations,

see: (a) Na, J.; Houk, K. N.; Shevlin, C. G.; Janda, K. D.; Lerner, R. A.

J. Am. Chem. Soc. 1993, 115, 8453e8454; (b) Na, J.; Houk, K. N. J. Am.

Chem. Soc. 1996, 118, 9204e9205.

28. Conformational restrictions on the substrate may alter the model

reactions described in Schemes 7 and 8. For an example, see: Nicolaou,

K. C.; Harrison, S. T. J. Am. Chem. Soc. 2007, 129, 429e440 and refer-

ences therein.

29. Crimmins, M. T.; Haley, M. W. Org. Lett. 2006, 8, 4223e4225.

30. For other 6-exo cyclizations, see: (a) Bhatt, U.; Christmann, M.;

Quitschalle, M.; Claus, E.; Kalesse, M. J. Org. Chem. 2001, 66,

1885e1893; (b) Williams, D. R.; Ihle, D. C.; Plummer, S. V. Org.Lett. 2001, 3, 1383e1386; (c) Yoshimitsu, T.; Makino, T.; Nagaoka, H.

J. Org. Chem. 2004, 69, 1993e1998; (d) Chang, S.-K.; Paquette, L. A.

Synlett 2005, 2915e2918; (e) Kadota, I.; Abe, T.; Sato, Y.; Kabuto, C.;

Yamamoto, Y. Tetrahedron Lett. 2006, 47, 6545e6548; (f) Takahashi,

S.; Hongo, Y.; Ogawa, N.; Koshino, H.; Nakata, T. J. Org. Chem.

2006, 71, 6305e6308; (g) Takizawa, A.; Fujiwara, K.; Doi, E.; Murai,

A.; Kawai, H.; Suzuki, T. Tetrahedron 2006, 62, 7408e7435; (h) Bedore,

M. W.; Chang, S.-K.; Paquette, L. A. Org. Lett. 2007, 9, 513e516.

31. Fuwa, H.; Ebine, M.; Bourdelais, A. J.; Baden, D. G.; Sasaki, M. J. Am.

Chem. Soc. 2006, 128, 16989e16999.

32. For other 6-endo cyclizations, see: (a) Mukai, C.; Sugimoto, Y.-i.; Ikeda,

Y.; Hanaoka, M. Tetrahedron 1998, 54, 823e850; (b) Uehara, H.; Oishi,

T.; Inoue, M.; Shoji, M.; Nagumo, Y.; Kosaka, M.; Le Brazidec, J.-Y.;

Hirama, M. Tetrahedron 2002, 58, 6493e6512; (c) Sato, K.; Sasaki,

M. Angew. Chem., Int. Ed. 2007, 46, 2518e2522; (d) Shangguan, N.;

Kiren, S.; Williams, L. J. Org. Lett. 2007, 9, 1093e1096.

33. McDonald, F. E.; Wei, X. Org. Lett. 2002, 4, 593e595.

34. (a) Morimoto, Y.; Nishikawa, Y.; Ueba, C.; Tanaka, T. Angew. Chem.,

Int. Ed. 2006, 45, 810e812; (b) Morimoto, Y.; Yata, H.; Nishikawa, Y.

Angew. Chem., Int. Ed. 2007, 46, 6481e6484.

35. Blanc, A.; Toste, F. D. Angew. Chem., Int. Ed. 2006, 45, 2096e2099.

36. For a comprehensive analysis of such processes in the context of func-

tionalized tetrahydrofurans, see: Hartung, J.; Drees, S.; Greb, M.;

Schimdt, P.; Svoboda, I.; Fuess, H.; Murso, A.; Stalke, D. Eur. J. Org.

Chem. 2003, 2388e2408.

37. Cane, D. E.; Celmer, W. D.; Westley, J. W. J. Am. Chem. Soc. 1983, 105,

3594e3600.

38. (a) Fujiwara, K.; Murai, A. Bull. Chem. Soc. Jpn. 2004, 77, 2129e2146;

(b) Valentine, J. C.; McDonald, F. E. Synlett 2006, 1816e1828; (c)

Gallimore, A. R.; Spencer, J. B. Angew. Chem., Int. Ed. 2006, 45,

4406e4413.

39. The efficiency of such strategy has been proved in the synthesis of a

Cs-symmetric diastereomer of glabrescol, a polyether containing five

contiguous tetrahydrofuran rings, see: Xiong, Z.; Corey, E. J. J. Am.Chem. Soc. 2000, 122, 4831e4832.

40. Tokiwano, T.; Fujiwara, K.; Murai, A. Synlett 2000, 335e338.

41. Bravo, F.; McDonald, F. E.; Neiwert, W. A.; Do, B.; Hardcastle, K. I.

Org. Lett. 2003, 5, 2123e2126.

42. Simpson, G. L.; Heffron, T. P.; Merino, E.; Jamison, T. F. J. Am. Chem.

Soc. 2006, 128, 1056e1057.

43. Chamberlin, A. R.; Mulholland, R. L.; Kahn, S. D.; Hehre, W. J. J. Am.Chem. Soc. 1987, 109, 672e677.

44. For mercury(II)-mediated cyclizations, see: (a) Evans, D. A.; Bender,

S. L.; Morris, J. J. Am. Chem. Soc. 1988, 110, 2506e2526; (b) de Kon-

ing, C. B.; Green, I. R.; Michael, J. P.; Oliveira, J. R. Tetrahedron 2001,

57, 9623e9634; (c) Liu, B.; Zhou, W.-S. Tetrahedron Lett. 2003, 44,

4933e4935; (d) Takao, H.; Wakabayashi, A.; Takahashi, K.; Imagawa,

H.; Sugihara, T.; Nishizawa, M. Tetrahedron Lett. 2004, 45, 1079e1082; (e) Petri, A. F.; Bayer, A.; Maier, M. E. Angew. Chem., Int. Ed.

2004, 43, 5821e5823; (f) Nicolaou, K. C.; Pihko, P. M.; Bernal, F.;

Frederick, M. O.; Qian, W.; Uesaka, N.; Diedrichs, N.; Hinrichs, J.;

Koftis, T. V.; Loizidou, E.; Petrovic, G.; Rodriquez, M.; Sarlah, D.;

Zou, N. J. Am. Chem. Soc. 2006, 128, 2244e2257.

45. The cyclization of g-hydroxy-cis-alkenes promoted by mercury(II) salts

and chiral bisoxazolines as ligands affords 2-substituted tetrahydro-

furans. Application of this methodology to just a d-hydroxy-cis-alkene

provides the corresponding tetrahydropyran in excellent ee, see:

Kang, S. H.; Kim, M. J. Am. Chem. Soc. 2003, 125, 4684e4685.

46. Meyer, C.; Blanchard, N.; Defosseux, M.; Cossy, J. Acc. Chem. Res.

2003, 36, 766e772.

Page 37: Available Online at Www.sciencedirect.com

2719I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

47. Defosseux, M.; Blanchard, N.; Meyer, C.; Cossy, J. J. Org. Chem. 2004,

69, 4626e4647.

48. (a) Schultz, A. G.; Guzi, T. J.; Larsson, E.; Rahm, R.; Thakkar, K.;

Bidlack, J. M. J. Org. Chem. 1998, 63, 7795e7804; (b) Bernard, N.;

Chemla, F.; Ferreira, F.; Mostefai, N.; Normant, J.-F. Chem.dEur. J.2002, 8, 3139e3147; (c) Van der Weghe, P.; Aoun, D.; Boiteau, J.-G.;

Eustache, J. Org. Lett. 2002, 4, 4105e4108; (d) Kang, S. H.; Kang,

S. Y.; Kim, C. M.; Choi, H.-w.; Jun, H.-S.; Lee, B. M.; Park, C. M.;

Jeong, J. W. Angew. Chem., Int. Ed. 2003, 42, 4779e4782; (e) Gao,

X.; Snider, B. B. J. Org. Chem. 2004, 69, 5517e5527; (f) Dıez, D.;

Nu~nez, M. G.; Moro, R. F.; Lumeras, W.; Marcos, I. S.; Basabe, P.;

Urones, J. G. Synlett 2006, 939e941; (g) Nicolaou, K. C.; Koftis, T. V.;

Vyskocil, S.; Petrovic, G.; Tang, W.; Frederick, M. O.; Chen, D. Y.-K.;

Li, Y.; Ling, T.; Yamada, Y. M. A. J. Am. Chem. Soc. 2006, 128, 2859e2872.

49. For a recent example on the classical iodolactonization process leading

to a six-membered lactone, see: Zakarian, A.; Batch, A.; Holton, R. A.

J. Am. Chem. Soc. 2003, 125, 7822e7824.

50. For other halo-mediated cyclizations, see: Zhong, H. M.; Sohn, J.-H.;

Rawal, V. H. J. Org. Chem. 2007, 72, 386e397.

51. (a) Albert, B. J.; Koide, K. Org. Lett. 2004, 6, 3655e3658; (b) Albert,

B. J.; Sivaramakrishnan, A.; Naka, T.; Koide, K. J. Am. Chem. Soc.

2006, 128, 2792e2793.

52. For a related iodoetherification involving a ketone, see: Clarke, P. A.;

Grist, M.; Ebden, M.; Wilson, C.; Blake, A. J. Tetrahedron 2005, 61,

353e363.

53. For other cyclizations involving oxygenated nucleophiles different than

alcohols or carboxylic acids, see: (a) Liu, K.; Taylor, R. E.; Kartika,

R. Org. Lett. 2006, 8, 5393e5395; (b) Braddock, D. C. Org. Lett.

2006, 8, 6055e6058.

54. For a review on selenocyclizations, see: Petragnani, N.; Stefani, H. A.;

Valduga, C. J. Tetrahedron 2001, 57, 1411e1448.

55. (a) See Refs. 1d, 19d, 50, 51; (b) Iwasaki, K.; Nakatani, M.; Inoue, M.;

Katoh, T. Tetrahedron 2003, 59, 8763e8773.

56. For a comprehensive overview, see: Tsuji, J. Palladium Reagents andCatalysts. New Perspectives for the 21st Century; John Wiley and

Sons: Chichester, UK, 2004.

57. The stereoselectivity of the process can be eroded if p-facial selectiv-

ity is not enough effective or the equilibration of diastereomeric com-

plexes via pesep mechanism is faster than the nucleophilic

addition.

58. Hansen, E. C.; Lee, D. Tetrahedron Lett. 2004, 45, 7151e7155.

59. For reviews on stereoselective allylic alkylations, see: (a) Trost, B. M.;

Van Vranken, D. L. Chem. Rev. 1996, 96, 395e422; (b) Trost, B. M.

Acc. Chem. Res. 1996, 29, 355e364; (c) Trost, B. M.; Crawley, M. L.

Chem. Rev. 2003, 103, 2921e2943.

60. Trost, B. M.; Machacek, M. R.; Faulk, B. D. J. Am. Chem. Soc. 2006,

128, 6745e6754.

61. For further applications of this methodology, see: Trost, B. M.;

Machacek, M. R.; Tsui, H. C. J. Am. Chem. Soc. 2005, 127, 7014e7024.

62. Campbell, J. E.; Englund, E. E.; Burke, S. D. Org. Lett. 2002, 4, 2273e

2275.

63. For other Pd(0)-mediated cyclizations, see: (a) Lucas, B. S.; Burke, S. D.

Org. Lett. 2003, 5, 3915e3918; (b) Zacuto, M. J.; Leighton, J. L. Org.Lett. 2005, 7, 5525e5527.

64. The alternative 7-endo ring closures leading to oxepanes have never been

observed.

65. The intramolecular cyclization of g-hydroxy alkenes could afford

the corresponding six-membered oxygenated heterocycles through a

6-endo ring closure. However, the five-membered oxygenated counter-

parts are usually obtained, with the exception of those substrates contain-

ing E-disubstituted olefins. See: (a) Semmelhack, M. F.; Bodurow, C.

J. Am. Chem. Soc. 1984, 106, 1496e1498; (b) Semmelhack, M. F.;

Zhang, N. J. Org. Chem. 1989, 54, 4483e4485; (c) McCormick, M.;

Monahan, R., III; Soria, J.; Goldsmith, D.; Liotta, D. J. Org. Chem.1989, 54, 4485e4487.

66. The regioselectivity of Pd(II)-mediated intramolecular cyclizations of

o-allylic phenols seems to be largely dependent of palladium(II) salts.

The use of PdCl2 appears to afford six-membered heterocycles, whereas

the five-membered counterparts are preferred with Pd(OAc)2. See:

Hosokawa, T.; Murahashi, S.-I. Acc. Chem. Res. 1990, 23, 49e54.

67. Hegedus, S. L. Organometallic in Synthesis: A Manual; Schlosser, M.,

Ed.; John Wiley and Sons: Chichester, UK, 2002; Chapter 10,

pp 1123e1217.

68. Semmelhack, M. F.; Kim, C. R. K.; Dobler, W.; Meier, M. Tetrahedron

Lett. 1989, 30, 4925e4928.

69. For other Pd(II)-mediated cyclizations, see: (a) Semmelhack, M. F.; Epa,

W. R.; Cheung, A. W.-H.; Gu, Y.; Kim, C.; Zhang, N.; Lew, W. J. Am.

Chem. Soc. 1994, 116, 7455e7456; (b) Arai, M. A.; Kuraishi, M.;

Arai, T.; Sasai, H. J. Am. Chem. Soc. 2001, 123, 2907e2908; (c) Trend,

R. M.; Ramtohul, Y. K.; Ferreira, E. M.; Stoltz, B. M. Angew. Chem., Int.Ed. 2003, 42, 2892e2895; (d) Koh, J. H.; Mascarenhas, C.; Gagne, M. R.

Tetrahedron 2004, 60, 7405e7410.

70. (a) Uenishi, J.; Ohmi, M. Angew. Chem., Int. Ed. 2005, 44, 2756e2760;

(b) Kawai, N.; Lagrange, J.-M.; Ohmi, M.; Uenishi, J. J. Org. Chem.2006, 71, 4530e4537; See also: (c) Miyazawa, M.; Hirose, Y.; Narant-

setseg, M.; Yokoyama, H.; Yamaguchi, S.; Hirai, Y. Tetrahedron Lett.

2004, 45, 2883e2886.

71. This analysis has been expanded to the corresponding Z isomers. Under

the same conditions, (2S,3Z,8S) isomer also affords 92-cis as a single

product in 76% yield. However, cyclization of (2R,3Z,8S) isomer turns

out to be less stereoselective and more stringent conditions are required.

72. Semmelhack, M. F.; Kim, C.; Zhang, N.; Bodurow, C.; Sanner, M.;

Dobler, W.; Meier, M. Pure Appl. Chem. 1990, 62, 2035e2040.

73. For an analysis on the influence of stereochemistry on the intramolecular

Pd(II)-mediated alkoxycarbonylation, see: Blakemore, P. R.; Browder,

C. C.; Hong, J.; Lincoln, C. M.; Nagornyy, P. A.; Robarge, L. A.; War-

drop, D. J.; White, J. D. J. Org. Chem. 2005, 70, 5449e5460.

74. White, J. D.; Kranemann, C. L.; Kuntiyong, P. Org. Lett. 2001, 3, 4003e4006.

75. Tietze, L. F.; Sommer, K. M.; Zinngrebe, J.; Stecker, F. Angew. Chem.,

Int. Ed. 2005, 44, 257e259.

76. (a) Reiter, M.; Ropp, S.; Gouverneur, V. Org. Lett. 2004, 6, 91e94; (b)

Baker-Glenn, C.; Hodnett, N.; Reiter, M.; Ropp, S.; Ancliff, R.; Gouver-

neur, V. J. Am. Chem. Soc. 2005, 127, 1481e1486; (c) Reiter, M.; Turner,

H.; Mills-Webb, R.; Gouverneur, V. J. Org. Chem. 2005, 70, 8478e8485;

(d) Han, S. B.; Krische, M. J. Org. Lett. 2006, 8, 5657e5660.

77. Liao, X.; Zhou, H.; Wearing, X. Z.; Ma, J.; Cook, J. M. Org. Lett. 2005,

7, 3501e3504.

78. Qian, H.; Han, X.; Widenhoefer, R. A. J. Am. Chem. Soc. 2004, 126,

9536e9537.

79. Coulombel, L.; Favier, I.; Du~nach, E. Chem. Commun. 2005, 2286e

2288.

80. Yang, C.-G.; Reich, N. W.; Shi, Z.; He, C. Org. Lett. 2005, 7, 4553e4556.

81. Marotta, E.; Foresti, E.; Marcelli, T.; Peri, F.; Righi, P.; Scardovi, N.;

Rosini, G. Org. Lett. 2002, 4, 4451e4453.

82. Jung, H. H.; Floreancig, P. E. Org. Lett. 2006, 8, 1949e1951.

83. (a) Nicolaou, K. C.; Li, A.; Edmonds, D. J. Angew. Chem., Int. Ed. 2006,

45, 7086e7090; (b) Zou, Y.; Chen, C.-H.; Taylor, C. D.; Foxman, B. M.;

Snider, B. B. Org. Lett. 2007, 9, 1825e1828; (c) Kaliappan, K. P.;

Ravikumar, V. Org. Lett. 2007, 9, 2417e2419; (d) Nicolaou, K. C.;

Edmonds, D. J.; Li, A.; Tria, G. S. Angew. Chem., Int. Ed. 2007, 46,

1e5; (e) Nicolaou, K. C.; Tang, Y.; Wang, J. Chem. Commun. 2007,

1922e1923.

84. For other examples on the Brønsted acid-mediated cyclization of

hydroxy alkenes, see: (a) Nicolaides, D. N.; Gautam, D. R.; Litinas,

K. E.; Papamehael, T. J. Chem. Soc., Perkin Trans. 1 2002, 1455e

1460; (b) Linares-Palomino, P. J.; Salido, S.; Altarejos, J.; Sanchez, A.

Tetrahedron Lett. 2003, 44, 6651e6655; (c) Neighbors, J. D.; Beutler,

J. A.; Wiemer, D. F. J. Org. Chem. 2005, 70, 925e931; (d)

Alvarez-Manzaneda, E. J.; Chahboun, R.; Perez, I. B.; Cabrera, E.;

Alvarez, E.; Alvarez-Manzaneda, R. Org. Lett. 2005, 7, 1477e1480;

(e) Grutter, C.; Alonso, E.; Chougnet, A.; Woggon, W.-D. Angew.

Chem., Int. Ed. 2006, 45, 1126e1130.

85. Rosenfeld, D. C.; Shekhar, S.; Takemiya, A.; Utsunomiya, M.; Hartwig,

J. F. Org. Lett. 2006, 8, 4179e4182.

Page 38: Available Online at Www.sciencedirect.com

2720 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

86. For a recent account on these cyclizations, see: Miura, K.; Hosomi, A.

Synlett 2003, 143e155.

87. (a) Tsujimori, H.; Bando, M.; Mori, K. Eur. J. Org. Chem. 2000, 297e

302; (b) Coulombel, L.; Rajzmann, M.; Pons, J.-M.; Olivero, S.; Du~nach,

E. Chem.dEur. J. 2006, 12, 6356e6365.

88. (a) Uyanik, M.; Ishihara, K.; Yamamoto, H. Bioorg. Med. Chem. 2005,

13, 5055e5065 and references therein; (b) See also Ref. 84e.

89. For reviews on the chemistry of allenes, see: (a) Zimmer, R.; Dinesh,

C. U.; Nandanan, E.; Khan, F. A. Chem. Rev. 2000, 100, 3067e3125;

(b) Bates, R. W.; Satcharoen, V. Chem. Soc. Rev. 2002, 31, 12e21.

90. (a) Gockel, B.; Krause, N. Org. Lett. 2006, 8, 4485e4488; (b) For a

related process involving a gold-catalyzed cyclization of a b-allene ester,

see: Piera, J.; Krumlinde, P.; Strubing, D.; Backvall, J.-E. Org. Lett.

2007, 9, 2235e2237.

91. Zhang, Z.; Liu, C.; Kinder, R. E.; Han, X.; Qian, H.; Widenhoefer, R. A.

J. Am. Chem. Soc. 2006, 128, 9066e9073.

92. Zhang, Z.; Widenhoefer, R. A. Angew. Chem., Int. Ed. 2007, 46, 283e

285.

93. For reviews on the electrophilic activation of alkynes, see: (a) Alonso, F.;

Beletskaya, I. P.; Yus, M. Chem. Rev. 2004, 104, 3079e3159; (b) Beller,

M.; Seayad, J.; Tillack, A.; Jiao, H. Angew. Chem., Int. Ed. 2004, 43,

3368e3398.

94. Gabriele, B.; Salerno, G.; Fazio, A.; Pitelli, R. Tetrahedron 2003, 59,

6251e6259.

95. For a Pd(II)eCu(II) bimetallic catalytic system, see: Asao, N.; Chan,

C. S.; Takahashi, K.; Yamamoto, Y. Tetrahedron 2005, 61, 11322e

11326.

96. Li, X.; Chianese, A. R.; Vogel, T.; Crabtree, R. H. Org. Lett. 2005, 7,

5437e5440.

97. Nakamura, H.; Ohtaka, M.; Yamamoto, Y. Tetrahedron Lett. 2002, 43,

7631e7633.

98. Trost, B. M.; Frontier, A. J. J. Am. Chem. Soc. 2000, 122, 11727e11728.

99. Silva, F.; Reiter, M.; Mills-Webb, R.; Sawicki, M.; Klar, D.; Bensel, N.;

Wagner, A.; Gouverneur, V. J. Org. Chem. 2006, 71, 8390e8394.

100. For an alkoxycarbonylation process in which the s-alkenylpalladium

complex is trapped with CO, see: Marshall, J. A.; Yanik, M. M. Tetra-

hedron Lett. 2000, 41, 4717e4721.

101. Genin, E.; Antoniotti, S.; Michelet, V.; Genet, J.-P. Angew. Chem., Int.Ed. 2005, 44, 4949e4953.

102. Barluenga, J.; Dieguez, A.; Fernandez, A.; Rodrıguez, F.; Fa~nanas, F. J.

Angew. Chem., Int. Ed. 2006, 45, 2091e2093.

103. (a) Antoniotti, S.; Genin, E.; Michelet, V.; Genet, J.-P. J. Am. Chem. Soc.2005, 127, 9976e9977; (b) For a related transformation involving

epoxy-alkynes, see: Dai, L.-Z.; Qi, M.-J.; Shi, Y.-L.; Liu, X.-G.; Shi,

M. Org. Lett. 2007, 9, 3191e3194.

104. Liu, B.; DeBrabander, J. K. Org. Lett. 2006, 8, 4907e4910.

105. For a related cycloisomerization using a palladium(II) catalyst, see:

Trost, B. M.; Horne, D. B.; Woltering, M. J. Angew. Chem., Int. Ed.

2003, 42, 5987e5990.

106. For iridium- and rhodium-mediated synthesis of acetals from alkyne

diols, see: (a) Messerle, B. A.; Vuong, K. Q. Pure Appl. Chem. 2006,

78, 385e390; (b) Messerle, B. A.; Vuong, K. Q. Organometallics

2007, 26, 3031e3040.

107. Li, Y.; Zhou, F.; Forsyth, C. J. Angew. Chem., Int. Ed. 2007, 46, 279e

282.

108. (a) Asao, N.; Nogami, T.; Takahashi, K.; Yamamoto, Y. J. Am. Chem.Soc. 2002, 124, 764e765; (b) Mondal, S.; Nogami, T.; Asao, N.;

Yamamoto, Y. J. Org. Chem. 2003, 68, 9496e9498.

109. For related transformations, see: (a) Patil, N. T.; Yamamoto, Y. J. Org.

Chem. 2004, 69, 5139e5142; (b) Yao, X.; Li, C.-J. Org. Lett. 2006, 8,

1953e1955; (c) Obika, S.; Kono, H.; Yasui, Y.; Yanada, R.; Takemoto,

Y. J. Org. Chem. 2007, 72, 4462e4468.

110. (a) For the synthesis of fused a-pyrones through palladium-catalyzed

6-endo-dig cyclizations, see: Raju, S.; Batchu, V. R.; Swamy, N. K.; Dev,

R. V.; Sreekanth, B. R.; Babu, J. M.; Vyas, K.; Kumar, P. R.; Mukkanti,

K.; Annamalai, P.; Pal, M. Tetrahedron 2006, 62, 9554e9570; (b) For

the synthesis of d-alkylidene lactones through gold-catalyzed 6-exo-dig

cyclizations, see: Harkat, H.; Weibel, J.-M.; Pale, P. Tetrahedron Lett.

2006, 47, 6273e6276; (c) For related mercury-catalyzed 6-exo-dig cycli-

zations, see: Imagawa, H.; Fujikawa, Y.; Tsuchihiro, A.; Kinoshita, A.;

Yoshinaga, T.; Takao, H.; Nishizawa, M. Synlett 2006, 639e641.

111. For a silver catalyzed cyclization on 2-alkynylketones, see: Patil, N. T.;

Pahadi, N. K.; Yamamoto, Y. J. Org. Chem. 2005, 70, 10096e10098.

112. For a rhodium-mediated cyclizationecycloaddition on 2-alkynylbenz-

aldehydes, see: Shin, S.; Gupta, A. K.; Rhim, C. Y.; Oh, C. H. Chem.

Commun. 2005, 4429e4431.

113. For a tungsten-mediated cyclizationecycloaddition on 2-alkynylphenyl-

carbonyl derivatives, see: Kusama, H.; Funami, H.; Shido, M.; Hara, Y.;

Takaya, J.; Iwasawa, N. J. Am. Chem. Soc. 2005, 127, 2709e2716.

114. 2-Benzopyrylium salts can undergo formal [4þ2] benzannulation cyclo-

additions to polysubstituted aromatic compounds without any oxygen-

ated heterocycle. For representative examples, see: (a) Asao, N.;

Takahashi, K.; Lee, S.; Kasahara, T.; Yamamoto, Y. J. Am. Chem. Soc.

2002, 124, 12650e12651; (b) Asao, N.; Nogami, T.; Lee, S.; Yamamoto,

Y. J. Am. Chem. Soc. 2003, 125, 10921e10925; (c) Asao, N.; Kasahara,

T.; Yamamoto, Y. Angew. Chem., Int. Ed. 2003, 42, 3504e3506; (d)

Straub, B. F. Chem. Commun. 2004, 1726e1728; (e) Asao, N.; Sato,

K.; Menggenbateer; Yamamoto, Y. J. Org. Chem. 2005, 70, 3682e

3685; (f) Dyker, G.; Hildebrandt, D. J. Org. Chem. 2005, 70, 6093e

6096; (g) Sato, K.; Asao, N.; Yamamoto, Y. J. Org. Chem. 2005, 70,

8977e8981; (h) Asao, N.; Sato, K. Org. Lett. 2006, 8, 5361e5363; (i)

Asao, N.; Aikawa, H. J. Org. Chem. 2006, 71, 5249e5253; (j) Hilde-

brandt, D.; Dyker, G. J. Org. Chem. 2006, 71, 6728e6733.

115. Zhu, J.; Germain, A. R.; Porco, J. A., Jr. Angew. Chem., Int. Ed. 2004, 43,

1239e1243.

116. Zhu, J.; Grigoriadis, N. P.; Lee, J. P.; Porco, J. A., Jr. J. Am. Chem. Soc.

2005, 127, 9342e9343.

117. A similar strategy has been followed for the enantioselective synthesis of

(�)-mitorubrin, see: Zhu, J.; Porco, J. A., Jr. Org. Lett. 2006, 8, 5169e

5171.

118. Gulıas, M.; Rodrıguez, J. R.; Castedo, L.; Mascare~nas, J. L. Org. Lett.

2003, 5, 1975e1977.

119. Imagawa, H.; Kotani, S.; Nishizawa, M. Synlett 2006, 642e644.

120. For the use of Ipyr2BF4, see: Barluenga, J.; Vazquez-Villa, H.; Merino,

I.; Ballesteros, A.; Gonzalez, J. M. Chem.dEur. J. 2006, 12, 5790e

5805.

121. For the use of iodine, NBS, and ICl, see: (a) Zhou, C.; Dubrovsky, A. V.;

Larock, R. C. J. Org. Chem. 2006, 71, 1626e1632; (b) Yue, D.; Della

Ca, N.; Larock, R. C. J. Org. Chem. 2006, 71, 3381e3388.

122. (a) Barluenga, J.; Vazquez-Villa, H.; Ballesteros, A.; Gonzalez, J. M.

Org. Lett. 2003, 5, 4121e4123; (b) Barluenga, J.; Vazquez-Villa, H.;

Ballesteros, A.; Gonzalez, J. M. Adv. Synth. Catal. 2005, 347, 526e530.

123. (a) McDonald, F. E. Chem.dEur. J. 1999, 5, 3103e3106; (b) Miki, K.;

Uemura, S.; Ohe, K. Chem. Lett. 2005, 34, 1068e1073.

124. For reviews on the preparation and reactivity of vinylidene complexes,

see: (a) Ref. 93a; (b) Trost, B. M.; Frederiksen, M. U.; Rudd, M. T.

Angew. Chem., Int. Ed. 2005, 44, 6630e6666; (c) Bruneau, C.; Dixneuf,

P. H. Angew. Chem., Int. Ed. 2006, 45, 2176e2203.

125. For a recent contribution to this mechanism, see: Grotjahn, D. B.; Zeng,

X.; Cooksy, A. L. J. Am. Chem. Soc. 2006, 128, 2798e2799 and refer-

ences therein.

126. Trost, B. M.; Rhee, Y. H. J. Am. Chem. Soc. 2003, 125, 7482e7483.

127. (a) Trost, B. M.; Rhee, Y. H. J. Am. Chem. Soc. 2002, 124, 2528e2533;

(b) Trost, B. M.; Rhee, Y. H. Org. Lett. 2004, 6, 4311e4313.

128. Molybdenum catalysts have been used exceptionally. See: McDonald,

F. E.; Zhu, H. Y. H. Tetrahedron 1997, 53, 11061e11068.

129. McDonald, F. E.; Reddy, K. S.; Dıaz, Y. J. Am. Chem. Soc. 2000, 122,

4304e4309.

130. For a theoretical study on the activation of W(CO)6 and the mechanism

of the process, see: Sheng, Y.; Musaev, D. G.; Reddy, K. S.; McDonald,

F. E.; Morokuma, K. J. Am. Chem. Soc. 2002, 124, 4149e4157.

131. (a) Cutchins, W. W.; McDonald, F. E. Org. Lett. 2002, 4, 749e752; (b)

Parker, K. A.; Chang, W. Org. Lett. 2003, 5, 3891e3893; (c) Wipf, P.;

Graham, T. H. J. Org. Chem. 2003, 68, 8798e8807; (d) Davidson,

M. H.; McDonald, F. E. Org. Lett. 2004, 6, 1601e1603; (e) Alcazar,

E.; Pletcher, J. M.; McDonald, F. E. Org. Lett. 2004, 6, 3877e3880;

Page 39: Available Online at Www.sciencedirect.com

2721I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

(f) Moilanen, S. B.; Tan, D. S. Org. Biomol. Chem. 2005, 3, 798e803;

(g) Koo, B.; McDonald, F. E. Org. Lett. 2005, 7, 3621e3624.

132. McDonald, F. E.; Reddy, K. S. Angew. Chem., Int. Ed. 2001, 40, 3653e

3655; See also McDonald, F. E.; Wu, M. Org. Lett. 2002, 4, 3979e

3981.

133. For comprehensive experimental and theoretical studies on the

THF$W(CO)5 and W(CO)6-mediated cyclizations, see: (a) Barluenga,

J.; Dieguez, A.; Rodrıguez, F.; Fa~nanas, F. J. Angew. Chem., Int. Ed.2005, 44, 126e128; (b) Sordo, T.; Campomanes, P.; Dieguez, A.;

Rodrıguez, F.; Fa~nanas, F. J. J. Am. Chem. Soc. 2005, 127, 944e952;

(c) Barluenga, J.; Dieguez, A.; Rodrıguez, F.; Fa~nanas, F. J.; Sordo, T.;

Campomanes, P. Chem.dEur. J. 2005, 11, 5735e5741.

134. For the THF$W(CO)5-mediated cyclizations of 2-ethynylphenyl ketones,

see: (a) Iwasawa, N.; Shido, M.; Maeyama, K.; Kusama, H. J. Am. Chem.

Soc. 2000, 122, 10226e10227; (b) Kusama, H.; Sawada, T.; Okita, A.;

Shiozawa, F.; Iwasawa, N. Org. Lett. 2006, 8, 1077e1079.

135. Koo, B.; McDonald, F. E. Org. Lett. 2007, 9, 1737e1740.

136. Doyle, M. P.; McKervey, M. A.; Ye, T. Modern Catalytic Methods for

Organic Synthesis with Diazo Compounds. From Cyclopropanes toYlides; John Wiley and Sons: New York, NY, 1998.

137. For reviews on the synthesis and reactivity of metal carbenes, see: (a)

Hodgson, D. M.; Pierard, F. Y. T. M.; Stupple, P. A. Chem. Soc. Rev.

2001, 30, 50e61; (b) Doyle, M. P. Top. Organomet. Chem. 2004, 13,

203e222; (c) Doyle, M. P. J. Org. Chem. 2006, 71, 9253e9260; (d)

For a recent report on the mechanism of Rh-catalyzed carbene formation

from diazo compounds, see: Wong, F. M.; Wang, J.; Hengge, A. C.; Wu,

W. Org. Lett. 2007, 9, 1663e1665.

138. Qu, Z.; Shi, W.; Wang, J. J. Org. Chem. 2004, 69, 217e219.

139. Lu, C.-D.; Liu, H.; Chen, Z.-Y.; Hu, W.-H.; Mi, A.-Q. Org. Lett. 2005, 7,

83e86.

140. Lopez-Herrera, F. J.; Sarabia-Garcıa, F. Tetrahedron 1997, 53, 3325e

3346.

141. For a similar transformation, see: Sarabia, F.; Chammaa, S.; Lopez

Herrera, F. J. Tetrahedron 2001, 57, 10271e10279.

142. (a) Marmsater, F. P.; West, F. G. J. Am. Chem. Soc. 2001, 123, 5144e

5145; (b) Marmsater, F. P.; Vanecko, J. A.; West, F. G. Tetrahedron

2002, 58, 2027e2040.

143. (a) Clark, J. S.; Whitlock, G.; Jiang, S.; Onyia, N. Chem. Commun. 2003,

2578e2579; (b) Clark, J. S.; Fessard, T. G.; Whitlock, G. A. Tetrahedron

2006, 62, 73e78.

144. For reviews on the reactivity of carbonyl ylides, see: (a) Mehta, G.;

Muthusamy, S. Tetrahedron 2002, 58, 9477e9504; (b) Padwa, A.

Helv. Chim. Acta 2005, 88, 1357e1374.

145. For the use of allenes as dipolarophiles, see: (a) Hodgson, D. M.; Le

Strat, F. Chem. Commun. 2004, 822e823; (b) Zhang, X.; Ko, R. Y. Y.;

Li, S.; Miao, R.; Chiu, P. Synlett 2006, 1197e1200.

146. For a combined application of ruthenium metathesis and rhodiume

carbonyl ylide formation/1,3-dipolar cycloaddition, see: Hodgson,

D. M.; Angrish, D. Adv. Synth. Catal. 2006, 348, 2509e2514.

147. Nakamura, S.; Hirata, Y.; Kurosaki, T.; Anada, M.; Kataoka, O.;

Kitagaki, S.; Hashimoto, S. Angew. Chem., Int. Ed. 2003, 42, 5351e5355.

148. Nakamura, S.; Sugano, Y.; Kikuchi, F.; Hashimoto, S. Angew. Chem., Int.

Ed. 2006, 45, 6532e6535.

149. For other applications to the synthesis of natural products, see: (a) Dau-

ben, W. G.; Dinges, J.; Smith, T. C. J. Org. Chem. 1993, 58, 7635e7637;

(b) Graening, T.; Friedrichsen, W.; Lex, J.; Schmalz, H.-G. Angew.Chem., Int. Ed. 2002, 41, 1524e1526; (c) Hodgson, D. M.; Le Strat,

F.; Avery, T. D.; Donohue, A. C.; Bruckl, T. J. Org. Chem. 2004, 69,

8796e8803; (d) Graening, T.; Bette, V.; Neudorfl, J.; Lex, J.; Schmalz,

H.-G. Org. Lett. 2005, 7, 4317e4320.

150. Snider, B. B.; Wu, X.; Nakamura, S.; Hashimoto, S. Org. Lett. 2007, 9,

873e874.

151. (a) Kitagaki, S.; Anada, M.; Kataoka, O.; Matsuno, K.; Umeda, C.;

Watanabe, N.; Hashimoto, S.-i. J. Am. Chem. Soc. 1999, 121, 1417e1418; (b) Hodgson, D. M.; Stupple, P. A.; Pierard, F. Y. T. M.; Labande,

A. H.; Johnstone, C. Chem.dEur. J. 2001, 7, 4465e4476; (c) Hodgson,

D. M.; Labande, A. H.; Pierard, F. Y. T. M.; Exposito Castro, M. �A.

J. Org. Chem. 2003, 68, 6153e6159.

152. Geng, Z.; Chen, B.; Chiu, P. Angew. Chem., Int. Ed. 2006, 45, 6197e6201.

153. (a) Inoue, K.; Suga, H.; Inoue, S.; Sato, H.; Kakehi, A. Synthesis 2003,

1413e1418; (b) Suga, H.; Inoue, K.; Inoue, S.; Kakehi, A.; Shiro, M.

J. Org. Chem. 2005, 70, 47e56; (c) Suga, H.; Suzuki, T.; Inoue, K.;

Kakehi, A. Tetrahedron 2006, 62, 9218e9225.

154. (a) Kuwabe, S.-i.; Torraca, K. E.; Buchwald, S. L. J. Am. Chem. Soc.

2001, 123, 12202e12206; (b) Arcadi, A.; Cacchi, S.; Fabrizi, G.;

Marinelli, F.; Verdecchia, M. Synlett 2006, 909e915.

155. (a) Chae, J.; Buchwald, S. L. J. Org. Chem. 2004, 69, 3336e3339; (b)

Palucki, M.; Yasuda, N. Tetrahedron Lett. 2005, 46, 987e990; (c) Furst-

ner, A.; Davies, P. W. J. Am. Chem. Soc. 2005, 127, 15024e15025; (d)

Heemstra, J. M.; Kerrigan, S. A.; Doerge, D. R.; Helferich, W. G.;

Boulanger, W. A. Org. Lett. 2006, 8, 5441e5443.

156. Fang, Y.; Li, C. J. Am. Chem. Soc. 2007, 129, 8092e8093.

157. For equilibration of the initially formed 2,6-trans to the desired 2,6-cis-

tetrahydropyran, see: (a) Evans, D. A.; Ripin, D. H. B.; Halstead, D. P.;

Campos, K. R. J. Am. Chem. Soc. 1999, 121, 6816e6826; (b) Micalizio,

G. C.; Pinchuk, A. N.; Roush, W. R. J. Org. Chem. 2000, 65, 8730e8736; (c) Ball, M.; Baron, A.; Bradshaw, B.; Omori, H.; MacCormick,

S.; Thomas, E. J. Tetrahedron Lett. 2004, 45, 8737e8740; (d) Strand,

D.; Rein, T. Org. Lett. 2005, 7, 199e202.

158. For the synthesis of spongistatin 1, see: Paterson, I.; Chen, D. Y.-K.;

Coster, M. J.; Ace~na, J. L.; Bach, J.; Gibson, K. R.; Keown, L. E.; Oballa,

R. M.; Trieselmann, T.; Wallace, D. J.; Hodgson, A. P.; Norcross, R. D.

Angew. Chem., Int. Ed. 2001, 40, 4055e4060.

159. For other examples, see: (a) Paterson, I.; Arnott, E. A. Tetrahedron Lett.1998, 39, 7185e7188; (b) Yadav, J. S.; Bandyopadhyay, A.; Kunwar,

A. C. Tetrahedron Lett. 2001, 42, 4907e4911; (c) Robertson, J.; Meo,

P.; Dallimore, J. W. P.; Doyle, B. M.; Hoarau, C. Org. Lett. 2004, 6,

3861e3863; (d) Alonso, D.; Perez, M.; Gomez, G.; Covelo, B.; Fall,

Y. Tetrahedron 2005, 61, 2021e2026; (e) Yadav, J. S.; Prakash, S. J.;

Gangadhar, Y. Tetrahedron: Asymmetry 2005, 16, 2722e2728; (f)

Lambert, W. T.; Hanson, G. H.; Benayoud, F.; Burke, S. D. J. Org.Chem. 2005, 70, 9382e9398; (g) Paterson, I.; Anderson, E. A.; Dalby,

S. M.; Loiseleur, O. Org. Lett. 2005, 7, 4125e4128; (h) Chandrasekhar,

S.; Prakash, S. J.; Shyamsunder, T. Tetrahedron Lett. 2005, 46, 6651e

6653; (i) Son, J. B.; Kim, S. N.; Kim, N. Y.; Lee, D. H. Org. Lett.2006, 8, 661e664; (j) Nonoyama, A.; Hamajima, A.; Isobe, M. Tetrahe-

dron 2007, 63, 5886e5894.

160. (a) Schneider, C.; Schuffenhauer, A. Eur. J. Org. Chem. 2000, 73e82;

(b) Vakalopoulos, A.; Hoffmann, H. M. R. Org. Lett. 2001, 3, 177e180; (c) Oishi, T.; Suzuki, M.; Watanabe, K.; Murata, M. Tetrahedron

Lett. 2006, 47, 3975e3978.

161. Occasionally, 2,6-trans-tetrahydropyrans can be more stable than the

2,3-cis counterparts, being obtained in high diastereoselectivity. For

instance, see: Li, M.; O’Doherty, G. A. Org. Lett. 2006, 8, 6087e6090.

162. (a) Bhattacharjee, A.; Soltani, O.; De Brabander, J. K. Org. Lett. 2002, 4,

481e484; (b) Soltani, O.; De Brabander, J. K. Org. Lett. 2005, 7, 2791e2793; (c) Pan, Y.; De Brabander, J. K. Synlett 2006, 853e856.

163. For preliminar analyses on the influence of the protecting group and the

alkene geometry, see: (a) Martın, V. S.; Nu~nez, M. T.; Ramırez, M. A.;

Soler, M. A. Tetrahedron Lett. 1990, 31, 763e766; (b) Gung, B. W.;

Francis, M. B. J. Org. Chem. 1993, 58, 6177e6179; (c) Martın, V. S.;

Palazon, J. M. Tetrahedron Lett. 1992, 33, 2399e2402.

164. (a) Betancort, J. M.; Martın, V. S.; Padron, J. M.; Palazon, J. M.;

Ramırez, M. A.; Soler, M. A. J. Org. Chem. 1997, 62, 4570e4583; (b)

Ramırez, M. A.; Padron, J. M.; Palazon, J. M.; Martın, V. S. J. Org.

Chem. 1997, 62, 4584e4590; (c) Banwell, M. G.; Bissett, B. D.; Bui,

C. T.; Pham, H. T. T.; Simpson, G. W. Aust. J. Chem. 1998, 51, 9e18;

(d) Vares, L.; Rein, T. J. Org. Chem. 2002, 67, 7226e7237.

165. For a parallel analysis on a,b-unsaturated ketones, see: Avery, T. D.;

Caiazza, D.; Culbert, J. A.; Taylor, D. K.; Tiekink, E. R. T. J. Org.

Chem. 2005, 70, 8344e8351.

166. This rule of thumb could be applied to the cyclization (K2CO3 in meth-

anol) of a (Z)-unsaturated ester containing a methyl group at the allylic

position, which furnishes the corresponding 2,3-trans-2,6-cis-tetrahydro-

pyran, see Ref. 73.

Page 40: Available Online at Www.sciencedirect.com

2722 I. Larrosa et al. / Tetrahedron 64 (2008) 2683e2723

167. Strand, D.; Norrby, P.-O.; Rein, T. J. Org. Chem. 2006, 71, 1879e1891.

168. The opposite bias has been observed for the cyclization of hydroxy esters

containing an epoxide at the allylic position. see: Harvey, J. E.; Raw,

S. A.; Taylor, R. J. K. Org. Lett. 2004, 6, 2611e2614.

169. The thermodynamically controlled addition of benzaldehyde to

5-hydroxy-2,3-unsaturated esters or amides in the presence of catalytic

amounts of base might be mentioned as a good example of this concept.

Indeed, it relies on the initial formation of a metal hemiacetal that sub-

sequently undergoes an intramolecular nucleophilic attack to the conju-

gated position to provide benzylidene protected 1,3-syn diols. Thus,

although the target of such process is not the assemblage of a pyran-

like heterocycles, it illustrates the potential of oxa-Michael addition in

cascade sequences. see: (a) Evans, D. A.; Gauchet-Prunet, J. A.

J. Org. Chem. 1993, 58, 2446e2453; (b) Ahmed, Md. M.; Mortensen,

M. S.; O’Doherty, G. A. J. Org. Chem. 2006, 71, 7741e7746; (c) Evans,

D. A.; Nagorny, P.; Reynolds, D. J.; McRae, K. J. Angew. Chem., Int. Ed.2007, 46, 541e544.

170. Nicolaou, K. C.; Lim, Y. H.; Papageorgiou, C. D.; Piper, J. L. Angew.

Chem., Int. Ed. 2005, 44, 7917e7921.

171. Ley, S. V.; Brown, D. S.; Clase, J. A.; Fairbanks, A. J.; Lennon, I. C.;

Osborn, H. M. I.; Stokes, E. S. E.; Wadsworth, D. J. J. Chem. Soc.,

Perkin Trans. 1 1998, 2259e2276.

172. Trost, B. M.; Yang, H.; Wuitschik, G. Org. Lett. 2005, 7, 4761e4764.

173. Trost, B. M.; Yang, H.; Thiel, O. R.; Frontier, A. J.; Brindle, C. S. J. Am.

Chem. Soc. 2007, 129, 2206e2207.

174. Most of these cyclizations involve protic acids. For a Lewis acid-medi-

ated 6-exo-trig assemblage of 2,6-cis-tetrahydropyran from silyloxy

enones, see: Evans, P. A.; Andrews, W. J. Tetrahedron Lett. 2005, 46,

5625e5627.

175. Paterson, I.; Keown, L. E. Tetrahedron Lett. 1997, 38, 5727e5730.

176. Nicolaou, K. C.; Bunnage, M. E.; McGarry, D. G.; Shi, S.; Somers, P. K.;

Wallace, P. A.; Chu, X.-J.; Agrios, K. A.; Gunzner, J. L.; Yang, Z.

Chem.dEur. J. 1999, 5, 599e617.

177. O’Brien, M.; Taylor, N. H.; Thomas, E. J. Tetrahedron Lett. 2002, 43,

5491e5494.

178. For other examples, see: (a) Paterson, I.; Luckhurst, C. A. Tetrahedron

Lett. 2003, 44, 3749e3754; (b) Chandrasekhar, S.; Shyamsunder, T.;

Prakash, S. J.; Prabhakar, A.; Jagadeesh, B. Tetrahedron Lett. 2006,

47, 47e49; (c) Liu, J.; Yang, J. H.; Ko, C.; Hsung, R. P. Tetrahedron

Lett. 2006, 47, 6121e6123; (d) See Refs. 157c and 160c.

179. Nguyen, S.; Xu, J.; Forsyth, C. J. Tetrahedron 2006, 62, 5338e5346.

180. Aiguade, J.; Hao, J.; Forsyth, C. J. Org. Lett. 2001, 3, 979e982.

181. (a) Forsyth, C. J.; Hao, J.; Aiguade, J. Angew. Chem., Int. Ed. 2001, 40,

3663e3667; (b) Forsyth, C. J.; Xu, J.; Nguyen, S. T.; Samdal, I. A.;

Briggs, L. R.; Rundberget, T.; Sandvik, M.; Miles, C. O. J. Am. Chem.Soc. 2006, 128, 15114e15116.

182. Nakatani, K.; Okamoto, A.; Yamanuki, M.; Saito, I. J. Org. Chem. 1994,

59, 4360e4361.

183. Nakatani, K.; Okamoto, A.; Matsuno, T.; Saito, I. J. Am. Chem. Soc.

1998, 120, 11219e11225.

184. (a) Tietze, L. F.; Gericke, K. M.; Singidi, R. R. Angew. Chem., Int. Ed.2006, 45, 6990e6993; (b) Tietze, L. F.; Singidi, R. R.; Gericke, K. M.

Org. Lett. 2006, 8, 5873e5876.

185. (a) Sakamoto, K.; Honda, E.; Ono, N.; Uno, H. Tetrahedron Lett. 2000,

41, 1819e1823; (b) Uno, H.; Sakamoto, K.; Honda, E.; Fukuhara, K.;

Ono, N.; Tanaka, J.; Sakanaka, M. J. Chem. Soc., Perkin Trans. 1

2001, 229e238.

186. Paterson, I.; Osborne, S. Tetrahedron Lett. 1990, 31, 2213e2216.

187. For other intramolecular conjugate addition/elimination processes, see:

(a) Miao, H.; Yang, Z. Org. Lett. 2000, 2, 1765e1768; (b) Nicolaou,

K. C.; Xu, H.; Wartmann, M. Angew. Chem., Int. Ed. 2005, 44, 756e

761; (c) Gao, B.; Yu, Z.; Fu, Z.; Feng, X. Tetrahedron Lett. 2006, 47,

1537e1539; (d) Hjelmgaard, T.; Søtofte, I.; Tanner, D. J. Org. Chem.

2005, 70, 5688e5697; (e) See also Ref. 32c.

188. (a) Paterson, I.; Smith, J. D.; Ward, R. A. Tetrahedron 1995, 51, 9413e9436; (b) Paterson, I.; Watson, C.; Yeung, K.-S.; Wallace, P. A.; Ward,

R. A. J. Org. Chem. 1997, 62, 452e453; (c) Paterson, I.; De Savi, C.;

Tudge, M. Org. Lett. 2001, 3, 3149e3152.

189. Reiter, M.; Turner, H.; Gouverneur, V. Chem.dEur. J. 2006, 12, 7190e7203.

190. (a) Clarke, P. A.; Martin, W. H. C. Org. Lett. 2002, 4, 4527e4529; (b)

Sabitha, G.; Reddy, G. S. K. K.; Rajkumar, M.; Yadav, J. S.;

Ramakrishna, K. V. S.; Kunwar, A. C. Tetrahedron Lett. 2003, 44,

7455e7457; (c) Clarke, P. A.; Martin, W. H. C.; Hargreaves, J. M.;

Wilson, C.; Blake, A. J. Chem. Commun. 2005, 1061e1063; (d) Clarke,

P. A.; Martin, W. H. C.; Hargreaves, J. M.; Wilson, C.; Blake, A. J. Org.Biomol. Chem. 2005, 3, 3551e3563.

191. For a similar procedure, see: DiVirgilio, E. S.; Dugan, E. C.; Mulrooney,

C. A.; Kozlowski, M. C. Org. Lett. 2007, 9, 385e388.

192. Clarke, P. A.; Martin, W. H. C. Tetrahedron 2005, 61, 5433e5438.

193. Biddle, M. M.; Lin, M.; Scheidt, K. A. J. Am. Chem. Soc. 2007, 129,

3830e3831.

194. For a related process, see: Chemler, J. A.; Yan, Y.; Leonard, E.; Koffas,

M. A. G. Org. Lett. 2007, 9, 1855e1858.

195. Hilli, F.; White, J. M.; Rizzacasa, M. A. Org. Lett. 2004, 6,

1289e1292.

196. (a) Jones, D. N.; Khan, M. A.; Mirza, S. M. Tetrahedron 1999, 55,

9933e9946; (b) Mukai, C.; Yamashita, H.; Hanaoka, M. Org. Lett.2001, 3, 3385e3387; (c) Kaye, P. T.; Nocanda, X. W. J. Chem. Soc., Per-

kin Trans. 1 2002, 1318e1323; (d) Kaye, P. T.; Musa, M. A.; Nocanda,

X. W.; Robinson, R. S. Org. Biomol. Chem. 2003, 1, 1133e1138.

197. Fernandez de la Pradilla, R.; Tortosa, M. Org. Lett. 2004, 6, 2157e2160.

Page 41: Available Online at Www.sciencedirect.com

2723edron 64 (2008) 2683e2723

I. Larrosa et al. / Tetrah

Biographical sketch

Igor Larrosa graduated from the University of Barcelona in 1999. In 2004, he

completed his Ph.D. under the supervision of Dr. Felix Urpı and Dr. Pedro

Romea in the area of asymmetric C-glycosidation reactions and natural

product synthesis. Subsequently, he became a postdoctoral research associate

in Professor Anthony G. M. Barrett’s group at Imperial College London, where

he carried out research on methodology development and its applications to

natural product synthesis. In 2007, he was appointed as a lecturer at Queen

Mary, University of London.

Pedro Romea completed his B.Sc. in Chemistry in 1984 at the University of

Barcelona. That year, he joined the group of Professor Jaume Vilarrasa, at the

University of Barcelona and received his Master’s Degree in 1985, and he fol-

lowed Ph.D. studies in the same group from 1987 to 1991. Then, he joined the

group of Professor Ian Paterson at the University of Cambridge (UK), where

he participated in the total synthesis of oleandolide. Back to the University

of Barcelona, he became Associate Professor in 1993. His research interests

have focused on the development of new synthetic methodologies and their

application to the stereoselective synthesis of naturally occurring molecular

structures.

Felix Urpı received his B.Sc. in Chemistry in 1980 at the University of Bar-

celona. In 1981, he joined the group of Professor Jaume Vilarrasa, at the

University of Barcelona and received his Master’s Degree in 1981 and

Ph.D. in 1988, where he was an Assistant Professor. He then worked as an

NATO postdoctoral research associate in titanium enolate chemistry with

Professor David A. Evans, at Harvard University at Boston. He moved back

to the University of Barcelona and became Associate Professor in 1991. His

research interests have focused on the development of new synthetic methodo-

logies and their application to the stereoselective synthesis of naturally occur-

ring molecular structures.