recent advances in inverse-electron-demand · cycloaddition reactions provide quick and economic...

37
REVIEW Recent Advances in Inverse-Electron-Demand Hetero-Diels–Alder Reactions of 1-Oxa-1,3-Butadienes Aleksandra Palasz 1 Received: 21 December 2015 / Accepted: 11 April 2016 / Published online: 20 April 2016 Ó The Author(s) 2016. This article is published with open access at Springerlink.com Abstract This review is an endeavor to highlight the progress in the inverse- electron-demand hetero-Diels–Alder reactions of 1-oxa-1,3-butadienes in recent years. The huge number of examples of 1-oxadienes cycloadditions found in the literature clearly demonstrates the incessant importance of this transformation in pyran ring synthesis. This type of reaction is today one of the most important methods for the synthesis of dihydropyrans which are the key building blocks in structuring of carbohydrate and other natural products. Two different modes, inter- and intramolecular, of inverse-electron-demand hetero-Diels–Alder reactions of 1-oxadienes are discussed. The domino Knoevenagel hetero-Diels–Alder reactions are also described. In recent years the use of chiral Lewis acids, chiral organocat- alysts, new optically active heterodienes or dienophiles have provided enormous progress in asymmetric synthesis. Solvent-free and aqueous hetero-Diels–Alder reactions of 1-oxabutadienes were also investigated. The reactivity of reactants, selectivity of cycloadditions, and chemical stability in aqueous solutions and under physiological conditions were taken into account to show the potential application of the described reactions in bioorthogonal chemistry. New bioorthogonal ligation by click inverse-electron-demand hetero-Diels–Alder cycloaddition of in situ-gen- erated 1-oxa-1,3-butadienes and vinyl ethers was developed. It seems that some of the hetero-Diels–Alder reactions described in this review can be applied in bioorthogonal chemistry because they are selective, non-toxic, and can function in biological conditions taking into account pH, an aqueous environment, and temperature. & Aleksandra Palasz [email protected] 1 Department of Organic Chemistry, Jagiellonian University, Ingardena 3 St, 30-060 Krako ´w, Poland 123 Top Curr Chem (Z) (2016) 374:24 DOI 10.1007/s41061-016-0026-2

Upload: others

Post on 19-Jun-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

REVIEW

Recent Advances in Inverse-Electron-DemandHetero-Diels–Alder Reactions of 1-Oxa-1,3-Butadienes

Aleksandra Pałasz1

Received: 21 December 2015 / Accepted: 11 April 2016 / Published online: 20 April 2016

� The Author(s) 2016. This article is published with open access at Springerlink.com

Abstract This review is an endeavor to highlight the progress in the inverse-

electron-demand hetero-Diels–Alder reactions of 1-oxa-1,3-butadienes in recent

years. The huge number of examples of 1-oxadienes cycloadditions found in the

literature clearly demonstrates the incessant importance of this transformation in

pyran ring synthesis. This type of reaction is today one of the most important

methods for the synthesis of dihydropyrans which are the key building blocks in

structuring of carbohydrate and other natural products. Two different modes, inter-

and intramolecular, of inverse-electron-demand hetero-Diels–Alder reactions of

1-oxadienes are discussed. The domino Knoevenagel hetero-Diels–Alder reactions

are also described. In recent years the use of chiral Lewis acids, chiral organocat-

alysts, new optically active heterodienes or dienophiles have provided enormous

progress in asymmetric synthesis. Solvent-free and aqueous hetero-Diels–Alder

reactions of 1-oxabutadienes were also investigated. The reactivity of reactants,

selectivity of cycloadditions, and chemical stability in aqueous solutions and under

physiological conditions were taken into account to show the potential application

of the described reactions in bioorthogonal chemistry. New bioorthogonal ligation

by click inverse-electron-demand hetero-Diels–Alder cycloaddition of in situ-gen-

erated 1-oxa-1,3-butadienes and vinyl ethers was developed. It seems that some of

the hetero-Diels–Alder reactions described in this review can be applied in

bioorthogonal chemistry because they are selective, non-toxic, and can function in

biological conditions taking into account pH, an aqueous environment, and

temperature.

& Aleksandra Pałasz

[email protected]

1 Department of Organic Chemistry, Jagiellonian University, Ingardena 3 St, 30-060 Krakow,

Poland

123

Top Curr Chem (Z) (2016) 374:24

DOI 10.1007/s41061-016-0026-2

Keywords Hetero-Diels–Alder reactions � 1-Oxa-1,3-butadienes �Dihydropyrans � Domino Knoevenagel hetero-Diels–Alder reactions �Bioorthogonal cycloaddition

AbbreviationsAc Acetyl

iBu Isobutyl

n-Bu n-Butyl

t-Bu tert-Butyl

(S,S)-

t-Bu-box

(S,S)-tert-Butylbis(oxazoline)

[bmim][NO3] 1-Butyl-3-methylimidazolium nitrate

Bn Benzyl

BPin 3-Boronopinacol

Bz Benzoyl

Cb Benzyloxycarbonyl

DEA Diethylamine

DIC N,N0-Diisopropylcarbodiimide

DMAP N,N-Ddimethyl-4-aminopyridine

DMF Dimethylformamide

DMSO Dimethyl sulfoxide

EDDA Ethylene diammonium diacetate

Eu(fod)3 6,6,7,7,8,8,8-Heptafluoro-2,2-dimethyl-3,5-octanedionato

europium

IBX o-Iodoxybenzoic acid

MDO Modularly designed organocatalyst

MS Molecular sieves

PCC Pyridinium chlorochromate

PDC Pyridinium dichromate

iPr Isopropyl

TBAB tetra-n-Butylammonium bromide

TBAF Tetrabutylammonium fluoride

TBA-HS Tetrabutylammonium hydrogen sulfate

TBDMS tert-Butyldimethylsilyl

TBDPS tert-Butyldiphenylsilyl

TBS tert-Butyldimethylsilyl

Tf Trifluoromethanesulfonyl

THF Tetrahydrofuran

24 Page 2 of 37 Top Curr Chem (Z) (2016) 374:24

123

1 Introduction

Cycloaddition reactions provide quick and economic methods for the construction

of monocyclic, polycyclic and heterocyclic systems. The use of hetero-substituted

diene and dienophiles is important for the application of Diels–Alder cycloadditions

towards natural and biologically active product synthesis. Dihydro- and tetrahy-

dropyran derivatives are prevalent structural subunits in a variety of natural

compounds, including carbohydrates, pheromones, alkaloids, iridoids and polyether

antibiotics [1–8]. The abundance of carbohydrates in living cells is a reason for the

development of new synthetic procedures for the preparation of natural and

unnatural sugars. There are two synthetic routes leading to dihydropyran derivatives

via [4 ? 2] cycloadditions. The first one is the [4 ? 2] cycloaddition of the

carbonyl group of aldehydes or ketones, acting as heterodienophiles, with electron-

rich 1,3-butadienes [9–23]. The second route is the hetero-Diels–Alder (HDA)

reactions of electron-deficient a,b-unsaturated carbonyl compounds, representing an

1-oxa-1,3-butadiene system, with electron-rich alkenes. Excellent diastereoselec-

tivity is a characteristic feature of heterocycloaddition of many substituted a,b-

unsaturated carbonyl compounds. The HDA reactions of oxabutadienes also show a

high regioselectivity. These reactions have been classified as cycloadditions with

inverse-electron-demand [24]. The reviews on this topic have already been

published but they cover the literature until only 1997 [1–8, 24]. The most

comprehensive one was written by Tietze and Kettschau in Topics in Current

Chemistry in 1997 [2]. The presented review is an endeavor to highlight the

progress in the HDA reactions with inverse-electron-demand of 1-oxa-1,3-

butadienes after the year 2000.

The reactivity of a,b-unsaturated carbonyl compounds in HDA reactions is low

and the reactions must be conducted at high temperature [25–27] or under high

pressure [28–30]. The use of enol ethers as dienophiles with electron-donating

groups improves the cycloadditions but high temperature is needed and diastere-

oselectivity of these reactions is still low. Aza-substituted dienophiles have been

used more rarely than their oxygenated counterparts in the HDA reactions of 1-oxa-

1,3-butadienes. Enamines can participate in these reactions, providing entry to

highly complex molecules [31–33]. The reactivity of 1-oxa-1,3-butadiene can be

enhanced by introducing electron-withdrawing substituents [34–39]. Presence of an

electron-withdrawing group in the 1-oxadiene system lowers the lowest energy

unoccupied molecular orbital (LUMO) energy level which then can more easily

overlap with the highest energy unoccupied molecular orbital (HOMO) orbital of

the dienophile. Tietze et al. calculated the influence of various substituents on the

energy of LUMO orbitals in 4-N-acetylamino-1-oxa-1,3-butadienes using semiem-

pirical methods [40]. It was found that the energy depends on the type and position

of a substituent in the 1-oxadiene system. The cyano and trifluoromethyl groups in

the 3 position were found to have the highest influence on reactivity of 1-oxa-1,3-

butadienes in cycloadditions with enol ethers. In addition to the effect of the

substituents in the heterodiene, Lewis acid catalysts, such as ZnCl2, TiCl4, SnCl4,

EtAlCl2, Me2AlCl, LiClO4, Mg(ClO4)2, Eu(fod)3, Yb(fod)3, accelerate the HDA

Top Curr Chem (Z) (2016) 374:24 Page 3 of 37 24

123

reactions [41–53]. The choice of the Lewis acid also has influence on the

stereoselectivity of cycloadditions because this catalyst is involved in an endo or an

exo-transition structure and steric interactions are important for stereochemistry.

Inverse-electron-demand HDA reaction between a,b-unsaturated carbonyl com-

pounds and electron-rich alkenes gives an enantioselective approach to chiral

dihydropyrans which are precursors for the synthesis of carbohydrate derivatives.

To obtain optically active carbohydrate derivatives by the HDA approach, either a

chiral transformation via the use of a chiral auxiliary or a catalytic enantioselective

reaction is necessary [50–53]. Two different modes of inverse-electron-demand

HDA reactions of 1-oxa-1,3-butadienes are discussed in this paper: inter- and

intramolecular mode. The geometry of the transition structures of HDA reactions

influences the diastereoselectivity of cycloadditions. There are four different

transition states for HDA reactions of 1-oxa-1,3-butadienes, according to an endo-

or exo-orientation of the dienophile and an (E)- or (Z)-configuration of the 1-oxa-

1,3-butadiene [2]. The four transition structures for inter- and intramolecular HDA

reactions providing the two diastereomers cis and trans are showed in Figs. 1 and 2.

The orientation of the dienophile–vinyl ether, with the alkoxy group being close to

the oxygen atom in the heterodiene is called endo (Fig. 1) [2]. The opposite is called

exo. For intramolecular HDA reactions of 1-oxa-1,3-butadienes, the orientation with

the chain connecting the heterodiene and dienophile lying close to the heterodiene is

called endo (Figure 2) [2]. The cis-adduct can be formed by an endo-E or exo-

Z orientation. The trans-adduct is obtained by either an exo-E or endo-Z transition

state (Figs. 1, 2).

Tietze et al. extensively described the domino Knoevenagel hetero-Diels–Alder

reactions of unsaturated aromatic and aliphatic aldehydes with different 1,3-

dicarbonyl compounds for the synthesis of heterocycles with a pyran ring [54–68].

In the intramolecular mode, the 1-oxa-1,3-butadienes are prepared in situ by a

Knoevenagel condensation of aldehydes bearing the dienophile moiety. This

cis - cycloadductendo-E exo-Z

OR1

R2

OR3

O

R3O

H

R2

R1HH

H

O

H

H

R1 R2

HH

OR3

exo-E endo-Z

OR1

R2

OR3

O

R2

H

HR1

H

OR3

HO R2

H

R3O

R1HH

Htrans - cycloadduct

Fig. 1 The four different diastereoselective approaches of an alkene such as vinyl ether to 1-oxa-1,3-butadiene for intermolecular HDA reaction

24 Page 4 of 37 Top Curr Chem (Z) (2016) 374:24

123

method has a broad scope since a multitude of different aldehydes and 1,3-

dicarbonyl compounds can be used.

Different examples of inter- and intramolecular HDA reactions of 1-oxa-1,3-

butadienes described in literature after the year 2000 are discussed below. The

usefulness of HDA reactions of oxadienes is connected with the number of bonds

which are formed in one sequence and with the fact that complex molecules can be

obtained by this method. Thus, the HDA reactions of a,b-unsaturated carbonyl

compounds are atom economic and they allow for regio-, diastereo- and

enantioselective synthesis of multifunctional pyran derivatives from relatively

simple compounds. Therefore, these cycloadditions can be potentially applied in

bioorthogonal chemistry.

2 Inverse-Electron-Demand Hetero-Diels–Alder Reactions of 1-Oxa-1,3-Butadienes

2.1 Intermolecular Hetero-Diels–Alder Reactions of 1-Oxa-1,3-Butadienes

2.1.1 Non-catalytic Intermolecular Hetero-Diels–Alder Reactions of 1-Oxa-1,3-

Butadienes

The reactivity of heterodienes in inverse-electron-demand HDA reactions can be

enhanced by introducing electron-withdrawing substituents into the 1-oxa-1,3-

butadiene system [34–39, 69–74]. For activation of an oxabutadiene in heterocy-

cloadditions, a cyano group can serve especially well. Such examples are

cycloadditions between propenenitriles with a cyano group at C-3 of the heterodiene

cis-cycloadduct

trans-cycloadduct

endo-E exo-Z

O

H

H

exo-E endo-Z

O HH

O

H

H

O

H

H

HO

H

OH

H

Fig. 2 The four different diastereoselective approaches of an alkene to 1-oxa-1,3-butadiene forintramolecular HDA reaction

Top Curr Chem (Z) (2016) 374:24 Page 5 of 37 24

123

system [71–74]. Moreover, two papers [71, 72] describe examples of cycloaddition

reaction of enaminocarbaldehydes or enaminoketones with enol ethers, leading to

4-amino-3,4-dihydro-2H-pyrans. 4-Amino-pyrans are precursors in synthesis of

3-amino sugar derivatives which are present in various antibiotics such as

gentamycin C or adriamycin. The HDA reactions of 3-(N-acetyl-N-benzylamino)-

2-formylprop-2-enenitriles 1 with enol ethers 2 yielded cis 3 and trans 4diastereoisomers of 2-alkoxy-4-amino-3,4-dihydro-2H-pyran-5-carbonitriles in

moderate yields (Scheme 1) [71]. The reactions of 2-benzoyl-3-heteroaromat-

icprop-2-enenitriles 5 with enol ethers 2 afforded diastereoisomeric cis 6 and trans 7cycloadducts [71].

Enaminocarbaldehyde 1 was found to be less reactive than propenenitriles 5 since

reactions of 5 with enol ethers 2 occurred at room temperature whereas reactions

with 1 required heating in boiling toluene.

Another interesting example of HDA reaction of 1-oxa-1,3-butadienes with vinyl

ethers was described by Klahn and Kirsch [75]. They examined dehydrogenation of

b-oxonitriles 8 by treatment with o-iodoxybenzoic acid (IBX) at room temperature

(Scheme 2). Products of the dehydrogenation–unsaturated counterparts 10 can react

in situ, undergoing rapid HDA reactions with enol ethers 9 to produce poly-

functionalized dihydropyrans 11. Cycloadducts 11 were generated in moderate to

good yields and with excellent cis-diastereoselectivity (up to[99:1).

Xing et al. described cycloadditions of fluorine-containing a,b-unsaturated

ketones 14, which are electron-poor 1-oxa-1,3-butadienes, with electron-rich olefins

15 (Scheme 3) [76]. The ketones 14 were prepared by the Knoevenagel reactions of

b-keto perfluoroalkanesulfones 12 with aromatic aldehydes 13 in presence of

ammonium acetate as catalyst.

Tetrasubstituted dihydropyrans 16 were prepared in quantitative yields. All the

products 16 were the diastereomeric mixtures.

O

NC

H

NAc

Ph

R4R3

R2 OR1O

N

OR1R2R3R4NC

AcPh

O

N

R2OR1R3R4NC

AcPh

pressure flask, toluene110oC, 30-96h

R1 = Me, Et, iBu, n-Bu, R2 = H, MeR3 = H, R4 = H, Me

1 2 cis-3 trans-4

O

NC

Ph

Ar

R2 OR1

O

Ar

OR1R2

NC

Ph O

Ar

R2OR1

NC

Ph

Ar = 2-thienyl,2-furyl, 4-NC-C6H4

5 2cis-6 trans-7

CH2Cl2, rt, 1-2 days

R1 = Et, iBuR2 = H, Me

yield 48-77% 3 and 4cis 3/trans 4 1.5:1 - 3.2:1

yield 76-88% 6 and 7cis 6/trans 7 4.5:1 - 12:1

Scheme 1 Inverse-electron-demand HDA reactions of propenenitriles 1 and 5 with enol ethers 2

24 Page 6 of 37 Top Curr Chem (Z) (2016) 374:24

123

Another example of HDA reaction of 1-oxa-1,3-butadienes is inverse-electron-demand

Diels–Alder cycloaddition of sterically hindered cycloalkylidene derivatives of benzoy-

lacetonitrile 17 and derivatives of N,N-dimethylbarbituric acid 20 with enol ethers 18and cyclic enol ether 22 (Scheme 4) [77]. Spirodihydropyrans 19, dispirodihydropyrans

23, spirouracils 21, and dispirouracils 24 were prepared. The cycloaddition reactions

of 2-cycloalkylidene-3-oxo-3-phenylpropionitriles 17 or 5-cycloalkylidene-1,3-

dimethylpyrimidine-2,4,6-triones 20 with enol ethers 18 were performed in toluene

solution at reflux and the pyrans 19 and 21 were obtained in good (78–93 %) yields

(Scheme 4). The inverse-electron-demand HDA reactions between cycloalkylidene

derivatives 17 or 20 and cyclic enol ether 22 were performed in toluene solution at 110 �Cfor 24 h and the dispiropyrans 23 and 24 were obtained in good (87–93 %) yields. For all

cycloadditions, high diastereoselectivity was observed. Products were each obtained as

one enantiomerically pure diastereoisomer. Confirmation of the experimental results by

semi-empirical AM1, PM3 methods and ab initio Hartree–Fock calculations of frontier

molecular orbital energies of heterodienes (H) and dienophiles (D) has been performed.

For reaction of ethyl-vinyl ether, the energy gaps ELUMO(H)–EHOMO(D) are slightly lower

for the cycloalkylidene derivatives of N,N-dimethylbarbituric acid than for the

cycloalkylidene derivatives of benzoylacetonitrile. The reactivity of cyclic enol ether is

comparable with the reactivity of ethyl-vinyl ether [77].

The scope of intermolecular HDA reactions of 1-oxa-1,3-butadienes with

inverse-electron-demand was expanded to cycloadditions with enecarbamate [78].

Cycloadditions of 3-aryl-2-benzoylprop-2-enenitriles and 3-phenylsulfonylbut-3-en-

2-ones 25 to N-vinyl-2-oxazolidinone 26 proceeded regio- and diastereoselectively,

yielding cis 27 and trans 28 diastereoisomers of 3,4-dihydro-2-(2-oxo-3-oxazo-

lidinyl)-2H-pyrans in 37–65 % yield (Scheme 5). Cycloadducts cis-27 were the

major products. Reactions of 5-arylidene-1,3-dimethylbarbituric acids 29 with

R2

CNR1

O

OR3IBX (1.5 equiv.)

DMSO/H2O (4:1), rt

8 9

R1 = Ph, Ph(CH2)2, 2-furanyl4-Br-C6H4, cyclopropyl, CO2Me

R2 = H, Me, Ph, n-heptyl

R3 = Et, nPr, nBu

OR1

NCR2

OR3H

yield 51-88%

1110R2

CNR1

O

OR3

Scheme 2 One-pot procedure for the conversion of b-oxonitriles 8 into dihydropyrans 11 bydehydrogenation and HDA reaction

yield 51-94%

R1O2SO

R2 ArCHO

R1 = CF3, (CH2)4ClR2 = Me, Ph

Ar = C6H5, 4-MeOC6H4, 4-MeC6H4,4-BrC6H4, 2-BrC6H4, C6H5CH=CH, 2-furyl

NH4OAc, MS

benzene, rt OR2

R1O2SAr

12 1314

yield 96-99%cis:trans 53:47 - 64:36

15OR3R3 = Et, iBu80oC

neat3-5h

16O

Ar

OR3

R1O2S

R2

Scheme 3 HDA reactions of (E)-a-perfluoroalkanesulfonyl-a,b-unsatutated ketones 14

Top Curr Chem (Z) (2016) 374:24 Page 7 of 37 24

123

dienophile 26 afforded mixtures of pyrano[2,3-d]pyrimidinediones trans 30 in

50–52 % yield and products 31 resulted from an elimination of 2-oxazolidinone.

N-Vinyl-2-oxazolidinone 26 can act as a valuable dienophile in inverse-electron-

demand heterocycloaddition. This compound was found to be less reactive than enol

ethers because similar reactions of dienes 25 and 29 with enol ethers occurred at

room temperature [71, 82] whereas reactions with 26 required heating in boiling

toluene.

R3 OR2

+toluene,110oC, 24h

pressure flask

17 18

R1 = H, Me, R2 = Me, Et, iBu, R3 = H, Me, n = 0, 1, 2, 3

19yield 83-91%

O

NC

Ph

R1

n

O

NC

PhR3

OR2

R1n

R1 = H, Me

OEt

+

yield 78-93%20 21

N

N

O

OO

R1

n

toluene,110oC, 24h

pressure flaskN

N O

R1

O

O OEt

n

18

R1 = H, Me

+toluene,110oC, 24h

pressure flask

17 22R1 = H, Me

23yield 91-93%

O

NC

Ph

R1

+

yield 87-89%20 24

toluene,110oC, 24h

pressure flask

22

O

ON

N

O

OO

R1

O

NC

Ph

R1

O

N

N O

R1

O

O

O

Scheme 4 HDA reactions of sterically hindered 2-cycloalkylidene-3-oxo-3-phenylpropionitriles 17 or5-cycloalkylidene-1,3-dimethylpyrimidine-2,4,6-triones 20 with enol ethers 18 and cyclic enol ether 22

24 Page 8 of 37 Top Curr Chem (Z) (2016) 374:24

123

2.1.2 Three-Component Domino Knoevenagel Hetero-Diels–Alder Reactions of 1-

Oxa-1,3-Butadienes

Domino Knoevenagel hetero-Diels–Alder reactions with an intermolecular cycload-

dition can be performed as a three-component reaction using a mixture of a 1,3-

dicarbonyl compound, an aldehyde, and a vinyl ether or an enamine. Any cyclic 1,3-

dicarbonyl compounds such as 1,3-cyclohexanediones, Meldrum’s acid or N,N-

O

R1

R2

R3 N O

O+

O

R1

NO

R2

R3

O

O

R1

NO

R2

R3

O+

toluenereflux 8-32h

25 26 27 28R1 = Ph, 4-NO2-C6H4, 4-CH3O-C6H4, CH3CONHR2 = CN, PhSO2

R3 = Ph, Me

cis 27/trans 28 9:1 - (>100:1)

N

N O

O

O

Ar

+ N O

O

26

toluenereflux 8-32h

O

Ar

NO

ON

N

O

O O

N

N

O

O

Ar

+

29 30 31

Ar = Ph, 4-NO2-C6H4, 4-CH3O-C6H4 30/31 1:0.14 - 1:1.5

Scheme 5 Inverse-electron-demand HDA reactions of different 1-oxa-1,3-butadienes 25 and 29 with N-vinyl-2-oxazolidinone 26

X ZY

O O

OBnR3

R1 O

HCbR2N n

EDDA, toluene

ultrasound, 50oC, 15h

3233

34

3637

X ZY

O O

R1CbR2NR3

OBnn

R1 = H, Me, iPr, iBu, BnR2 = H, Men = 0, 1, 2Cb = benzyloxycarbonyl

X ZY

O O=

N N

O O

O

O O O O

O

O OH

R3 = H, Me, Et, iPr

X ZY

O O

R1CbR2N

n OBnR3

35

Pd/C, H2, 1bar

rt, 24hn

X ZY

O O

N

R3

R1

H R2

Scheme 6 Domino sequence comprising Knoevenagel, inverse-electron-demand HDA reaction andhydrogenation starting from amino aldehydes 32, 1,3-dicarbonyl compounds 34, and enol ethers 33

Top Curr Chem (Z) (2016) 374:24 Page 9 of 37 24

123

dimethylbarbituric acid, as well as reactive acyclic 1,3-dicarbonyl compounds, can

be employed. Tietze et al. examined the multicomponent domino Knoevenagel

HDA reactions of 1,3-dicarbonyl compounds 34 with amino aldehydes 32 and enol

ethers 33, followed by a reductive amination with the formation of betaines 37which can be precipitated from the solution in high purity (Scheme 6) [79, 80].

The amino aldehydes 32 were treated with the 1,3-dicarbonyl components 34 and

benzoyl enol ethers 33 in toluene in the presence of catalytic amounts of EDDA and

trimethyl orthoformate as dehydrating agent in an ultrasonic bath. The domino

reaction sequence of Knoevenagel, HDA reaction, and hydrogenation allows rapid

access to a number of N-heterocycles of different ring sizes and with different

substituents in a betaine 37.

Radi et al. described a protocol for the multicomponent microwave-assisted

organocatalytic domino Knoevenagel HDA reaction for the synthesis of substi-

tuted 2,3-dihydropyran[2,3-c]pyrazoles [81]. The reported procedure can be used

for the fast generation of pyran[2,3-c]pyrazoles with potential anti-tuberculosis

activity.

A mixture of pyrazolone 38, aldehyde 40 and 10 equiv of ethyl-vinyl ether 39was MW irradiated and heated at 110 �C in the presence of the appropriate

organocatalyst A–F (Scheme 7). The best results were obtained in the presence of

diaryl-prolinols B and C. In the absence of the catalyst the reaction did not start at

all. Using the catalyst B and t-BuOH as the solvent, the authors obtained the

cycloadducts 41 and 42 in yields (56 and 12 %, respectively) and improved

diastereoisomeric ratio (4:1) in comparison to the results previously obtained.

Inverse-electron-demand HDA reaction of 1-oxa-1,3-butadienes was used in

synthesis of the fused uracils–pyrano[2,3-d]pyrimidine-2,4-diones [82]. This group

of uracils, as a fused heterobicyclic system, constitutes an important contribution in

medicinal chemistry and a wide variety of attractive pharmacological effects has

been attributed to them [83]. First, it was examined that 5-arylidene-N,N-

dimethylbarbituric acids 43 undergo smooth HDA reactions with enol ethers 44to afford cis-47 and trans-48 diastereoisomers of 7-alkoxy-5-aryl-2H-pyrano[2,3-

d]pyrimidine-2,4-diones in excellent yields (84–95 %; Scheme 8). Cycloadducts 47with cis-configurations were the major products. Next, three-component one-pot

MW, 110oC30 min, organocatalyst A-F

NH

NH HO

NH HO

NH

COOH NH OTMS

NH HO

OMe

MeO

A B C

D

E

F

NN O OEt

Cl

Cl

NN O OEt

Cl

Cl

2414

NN

O

Cl

Cl

CHO

OEt

38

39

40

20 mol%

Scheme 7 Microwave-assisted organocatalytic mulicomponent Knoevenagel HDA reaction

24 Page 10 of 37 Top Curr Chem (Z) (2016) 374:24

123

reactions of N,N-dimethylbarbituric acid 45, aromatic and heteroaromatic aldehydes

46, and enol ethers 44 in the presence of piperidine gave uracils cis-47 and trans-48also in very good yields (87–95 %; Scheme 8). Trace amounts of compounds 49created by a trans-diaxial-elimination of the appropriate alcohol were also obtained

in these reactions.

The advantages of these reactions are: the excellent yields, short reactions

times, and the fact that cycloadditions do not require drastic conditions, but can be

carried out at room temperature. The described reactions give easy and rapid

access to both cis-47 as trans-48 diastereoisomers of uracils and pure

diastereoisomers can be very easily isolated by column chromatography. Also,

solvent-free HDA reactions of 5-arylidene derivatives of barbituric acids 50 with

ethyl vinyl ether 51 were investigated at room temperature and pyrano[2,3-

d]pyrimidines 52 and 53 were obtained in excellent yields (Scheme 9) [84]. Three-

component one-pot syntheses of fused uracils were performed in aqueous

suspensions. ‘‘On water’’ reactions of barbituric acids 50, aldehydes 54, and

ethyl vinyl ether 51 where carried out at ambient temperature, whereas the one-pot

synthesis with barbituric acids 50, aldehydes 54, and styrene or N-vinyl-2-

oxazolidinone 56 required the heating of aqueous suspensions at 60 �C(Scheme 9). Formation of the unexpected side products 55 can be explained as

the result of three-component reactions of barbituric acids and acetaldehyde which

were produced from reaction of etyl vinyl ether and water, and ethyl vinyl ether

51. Described ‘‘on water’’ cycloadditions were characterized by higher diastere-

oselectivity in contrast to reactions carried out in homogenous organic media

(dichloromethane, toluene, Scheme 9). They allowed the cis adducts 57 to be

obtained preferentially or exclusively. Green methods presented in this study avoid

the use of catalysts, the heating of reaction mixtures for long time at high

temperature, and the use of organic solvents.

NH

N

N O

O

O

Ar

R2 OR1

N

N O

O

O

Ar

CHO

R2 OR1

N

N O

O

O

Ar

OR1R2

N

N O

O

O

Ar

R2OR1

N

N O

O

O

Ar

R2

43

4445

46

cis-47 trans-48

49

44

CH2Cl2rt, 1-5h

Ar = Ph, 4-CH3O-C6H44-NO2-C6H4, 2-thienyl,2-furyl, 4-pyridinyl

R1 = Me, Et, iBuR2 = H, Me

cis-47/trans-48 1.1:1 - 3.6:1

CH2Cl2rt, 2-7h

Scheme 8 Inverse-electron-demand HDA reactions of 5-arylidene-N,N-dimethylbarbituric acids 43 withenol ethers 44. Three-component one-pot synthesis of fused uracils – pyrano[2,3-d]-pyrimidine-2,4-diones cis-47 and trans-48

Top Curr Chem (Z) (2016) 374:24 Page 11 of 37 24

123

2.1.3 Catalytic Hetero-Diels–Alder Reactions of 1-Oxa-1,3-Butadienes

with Achiral Lewis Acids

It was mentioned in the Introduction that Lewis acids accelerate the HDA reactions

of 1-oxa-1,3-butadienes [41–53]. Lewis acids can also improve regioselectivity and

diastereoselectivity of these reactions. The example of catalytic HDA reaction are

cycloadditions of a-keto-b,c-unsaturated phosphonates 58 and 65 with cycloalke-

nes: cyclopentadiene 59, cyclohexadiene 62, dihydrofuran, and dihydropyran 66,

described by Hanessian and Compain (Scheme 10) [85]. The reactions led to the

formation of the hetero-Diels–Alder products 60, 63 and 67 in addition to the

normally expected Diels–Alder cycloadducts 61 and 64. Hetero-Diels–Alder

cycloadducts with the endo product as the major isomer were the main products

in the presence of SnCl4 as a Lewis acid. The effect of substituents on

stereochemistry of these reactions can be explained by considering steric

interactions in the transition state. Increasing the bulk of the ester moiety lowered

the ratio of hetero to normal Diels–Alder products while geminal substitution

favored the product formed by HDA reaction. In the reactions of dialkyl a-

cis-52/ trans-53 1:1 - 3.4:1

N

N

R3

O

O

XR2

R1

OEt

N

N O

O

X

R1

R2

R3

OEt

50 51 cis-52 trans-53

N

N O

O

X

R1

R2

R3

OEt

yield 52 and 53 88-96%

neat, rt, 1-5h

(10 equiv.)

R1 = R2 = H, Me X = O,SR3 = Ph, 4-MeOC6H4, 4-NO2C6H4, 4-NCC6H4, 4-ClC6H4, 4-BrC6H4, 2-thienyl, 2-furyl, 4-pyridyl

N

N O

O

XR2

R1 R3CHO

OEt

H2O, rt, 1-3hcis-52

trans-53

N

N O

O

X

R1

R2

Me

OEt

yield 52 and 53 88-96%

R1 = R2 = H, Me X = O,SR3 = Ph, 4-MeOC6H4, 4-NO2C6H4, 4-NCC6H4, 4-ClC6H4, 4-BrC6H4, 2-thienyl

50 51

54

55

N

N O

O

XR2

R1 R3CHO

OR4

H2O, 60oC, 1-3h

50

54

56

N

N O

O

X

R1

R2

R3

R4

cis-57

R1 = R2 = H, Me X = O,SR3 = Ph, 4-NO2C6H4 R4 = 4-MeO, 2-oxo-oxazolidin-3-yl

yield cis-57 89-95%

cis-52/ trans-53 4.1:1 - 8.5:1

Scheme 9 Solvent free HDA reactions of 5-arylidene-N,N-dimethylbarbituric acids 50 with ethyl vinylether 51. Three-component 50, 54 and 51 or 56 one-pot synthesis of fused uracils in water

24 Page 12 of 37 Top Curr Chem (Z) (2016) 374:24

123

R4

R1 P(OR3)2

O

OR2

O P(OR3)2

O

R2

R1 R4

0.6 equiv SnCl4CH2Cl2, -78oC

R1

R2

P(OR3)2O

R4

58 59 60 61R1 = Me, Ph R2 = H, MeR3 = Me, Et, iPr R4 = H, Me

yield 52-77%endo/exo 2.4:1 - 47:160/61 2.4:1 - (>30:1)

R1 P(OR3)2

O

OR2

1 equiv SnCl4CH2Cl2, -78oC

R1 = Me, Ph R2 = H, Me R3 = Me, Et58

O P(OR3)2

O

R2

R1

R1

P(OR3)2

R2

O62 63 64yield 43-71%endo/exo 17:1 - (>50:1)63/64 1:1 - 14:1

yield 68-78%endo/exo 25:1 - 36:1

R1 P(OR2)2

O

OR2

R3O

n O P(OR3)2

O

R2

R1

1 equiv SnCl4CH2Cl2, -78oC

65 6667R1 = Me, Ph R2 = Me, Et R3 = H, Me n = 0, 1

Scheme 10 Lewis acid SnCl4 promoted HDA reactions of a-ketophosphonoenoates 58 and 65 withdienes 59, 62 and 66

3 equiv.

3 equiv.

OH O

O

O

Ph

O

O

O

Ph

OOMe

O

PhOH

O

Ph

O

OH

Ph

yield 70%cis selectivity >97:3facial selectivity 93:7

1) 20 mol% Eu(fod)3, CH2Cl2, 48h, reflux2) filtration

1) LiAlH4, THF/Et2O, 12h, rt2) Na2SO4 aq3) filtration

O

Ph

O

OMe68

69

70

71

72

73

O

Ph

O

HO

1) DIC, DMAP, DMF, 12h, rt2) filtration

Scheme 11 Eu(fod)3-catalyzed solid-phase HDA reaction of benzylidenepyruvate 70 with (S)-(?)-O-vinyl mandelate 71

Top Curr Chem (Z) (2016) 374:24 Page 13 of 37 24

123

crotonylphosphonates 65 with dihydrofuran and dihydropyran 66, the only isolable

products were heterocycloadducts with high endo/exo ratios and good yields.

Compatibility of the carrier and the linker with the Lewis acid is a main criterion

for a successful application of Lewis acid catalysts on solid supports in asymmetric

[4 ? 2] heterocycloadditions. Dujardin et al. demonstrated the usefulness of a

Wang resin-bound heterodiene benzylidenepyruvate 70 for Eu(fod)3-catalyzed

inverse-electron-demand HDA reactions with (S)-(?)-O-vinyl mandelate 71(Scheme 11) [86]. The solid-phase sequence allowed an unprecedented reuse of

the catalyst in the presence of excess dienophile in solution. Also, attempts with

ethyl vinyl ether as an achiral dienophile gave positive results.

Gong et al. examined asymmetric inverse-electron-demand HDA reaction of

trisubstituted chiral enol ether 75 derived from (R)-mandelic acid (Scheme 12) [87].

Chiral 1,2,3,5-substituted tetrahydropyrans were synthesized by a three-step

sequence with a remarkable and unprecedented endo and facial stereocontrol. The

key step involved the Eu(fod)3-catalyzed HDA reaction of a trisubstituted chiral

enol ether 75 and an activated heterodiene 74. The stereoselective hydrogenation of

the heteroadducts 1-alkoxydihydropyrans 76 was optimized by using Pd on charcoal

and diisopropylethylamine, leading to a unique isomer [87].

Another example of inverse-electron-demand HDA reaction of 1-oxa-1,3-butadienes is

the [4 ? 2] acido-catalyzed heterocycloaddition between b-substituted N-vinyl-1,3-

oxazolidin-2-ones 78 and unsaturated a-ketoesters 77 (Scheme 13) [88, 89]. Cycload-

ditions afforded dihydropyrans 79 and 80 with high levels of endo and facial selectivities.

A complete reversal of facial differentiation was achieved by using a different Lewis

acid, leading to the stereoselective formation of either endo-a 79 or endo-b 80 adducts.

The endo-a adduct 79 was obtained with using Eu(fod)3 as the catalyst and endo-badducts 80 was the main product if the promoter was SnCl4 (Scheme 13) [88, 89].

2.1.4 Enantioselective Approach: Catalytic Enantioselective Hetero-Diels–Alder

Reactions of 1-Oxa-1,3-Butadienes with Chiral Lewis Acids

The catalytic enantioselective HDA reactions of 1-oxa-1,3-butadienes with chiral

Lewis acids were widely explored reactions. The chiral bisoxazoline copper(II)

complexes have been shown to be effective catalysts for inverse-electron-demand

HDA reactions. The reactions of a,b-unsaturated acyl phosphonates 81 and 84 and

O

OR1

OMe

O

R2

R2

O

O

OR3Ph

R1 = Bn, t-BuR2 = H, MeR3 = Me, t-Bu

5 mol% Eu(fod)3

petroleum ether reflux O

OR1

OMe

O

R2

O

R2

Ph

O

OR3

74 75 76

Scheme 12 Eu(fod)3 catalyzed HDA reaction of alkylidenepyruvate 74 with O-isopropenyl mandelicesters 75

24 Page 14 of 37 Top Curr Chem (Z) (2016) 374:24

123

b,c-unsaturated a-keto esters and amides 87 with enol ethers and sulfides 82 and 85as dienophiles were described by Evans et al. (Scheme 14) [90]. The products were

prepared with high diastereo- and enantioselectivity. The selectivities of reactions

exceeded 90 % even at room temperature. The synthesis of bicyclic adducts 83 in

high diasteromeric and enantiomeric excess proved that cyclic enol ethers 82 can be

excellent dienophiles. The derived cycloadducts were transformed to useful chiral

building blocks such as desymmetrized glutaric acid derivatives or highly

functionalized tetrahydropyran products. The authors examined that the high

yield 44-83%yield 57-77%selectivity endo/exo for 79 and 80 80:20 - (>98:2)

O

R1

MeO2C

R2

NO

O

O

R1

N

R2

O

O

MeO2CO

R1

N

R2

O

O

MeO2Cendo βendo α

77 7879 80

R2 = H, Me, Ph

SnCl4 50 mol%, -78oC

CH2Cl2

Eu(fod)3 5 mol%, reflux

cyclohexane

R1 = Ph,

MeO

MeO

MeO

Scheme 13 Lewis acid tuned facial stereodivergent HDA reactions of b-substituted N-vinyl-1,3-oxazolidin-2-ones 78

O(MeO)2P

O

Me

Y

Z O

Me

Y

X(MeO)2P

Oendo/exo product 1:1 - (>99:1)ee 64 - 95%yield 31-100%

Y

Z

=O O OR SEt

R = Me, t-Bu, TBS

10 mol%

CH2Cl2, -78oC

R = t-Bu, i-Pr, Bn, PhX = OTf, SbF6

CuN

O

N

O

Me Me

RR

2+

2X -

81 82 83

O(MeO)2P

O

R1

R2

Z O

R1

Z(MeO)2P

O

R2

R1 = Me, Ph, i-Pr, OEtR2 = H, Me

Z = OEt, SEt

[Cu((S,S)-t-Bu-box)(H2O)2](OTf)25 mol%

THF, 0oC, MS

endo/exo product 16:1 - 44:1ee 90 - 97%yield 75-98%

84 85 86

OYOC

R

Z

R = Me, Et, i-Pr, OMe, OEt, Ph, SBnY = OEt, N(Me)OMe

[Cu((S,S)-t-Bu-box)(H2O)2](OTf)22 mol%

THF, 0oC, MSO

R

ZYOC

Z = OEt, SEtendo/exo product 10:1 - 64:1ee 95 - 99%yield 87-99%

87 85 88

Scheme 14 HDA reactions of acyl phosphonates 81, 84 and b,c-unsaturated a-keto esters and amides87 with enol ethers and sulfides 82 and 85 catalyzed by bis(oxazoline)-Cu(II) complexes

Top Curr Chem (Z) (2016) 374:24 Page 15 of 37 24

123

diastereoselectivity for catalyzed HDA reactions is a result of the endo orientation

of dienophiles.

A highly enantioselective approach for the synthesis of optically active

carbohydrate derivatives by inverse-electron-demand HDA reaction of a,b-unsat-

urated carbonyl compounds with electron-rich alkenes catalyzed by combination of

chiral bisoxazolines and Cu(OTf)2 as the Lewis acid was also presented by

Jorgensen et al. [91]. The reaction of unsaturated a-keto esters 89 and 92 with vinyl

ether 90 and various types of cis-disubstituted alkenes 93 proceeded in good yield,

high diastereoselectivity, and excellent enantioselectivity (Scheme 15). The poten-

tial of the reaction was demonstrated by the synthesis of optically active

carbohydrates such as spiro-carbohydrates, an ethyl b-D-mannoside tetraacetate,

and acetal-protected C-2-branched carbohydrates [91].

Catalytic enantioselective HDA reaction of 1-oxa-1,3-butadiene with inverse-

electron-demand was used in synthesis of the marine neurotoxin-(?)-azaspiracid

[92]. Cycloaddition between two components of the HDA reaction 95 and 96proceeded readily using 2 mol% loadings of the hydrated copper complex 97(Scheme 16). Catalyst 97 was dehydrated with molecular sieves prior to use.

Diethyl ether was the optimal solvent for this HDA reaction (97 % ee, dr 94:6). The

desired cycloadduct 98 was isolated in 84 % yield as a single isomer.

The tridentate (Schiff base) chromium complex has been identified as a highly

diastereoselective and enantioselective catalyst in HDA reactions between aldehy-

des and mono-oxygenated 1,3-diene derivatives [93]. Jacobsen et al. examined if

use of this chiral catalyst can be evaluated for the reactions of conjugated aldehydes

[94]. The inverse-electron-demand HDA reactions of crotonaldehyde and the wide

range of a,b-unsaturated aldehydes 99 bearing b substituents and vinyl ether 100

919089de 71-99%ee 87-99%yield 93-99%

O

Ph

OEt

O

MeOO

Ph

O

MeOOEt

0.5-10 mol%

R = t-Bu, PhX = OTf, SbF6

CuN

O

N

O

Me Me

RR

2+

2X -

Et2O

949392de > 99ee 66-99.5%yield 41-76%

R3 = OEt, OBnR1 = Ph, OEt, OBnR2 = Me, Et

O

R1

R3

OAc

O

R2OR3

OAc

O

R1

O

R2O

[Cu(S,S)-t-Bu-box](OTf)210-20 mol%Et2O, rt

Scheme 15 HDA reactions of c-substituted b,c-unsaturated a-keto esters 89 and 92 with vinyl ether 90and cis-disubstituted alkenes 93 catalyzed by bis(oxazoline)-Cu(II) complexes

24 Page 16 of 37 Top Curr Chem (Z) (2016) 374:24

123

proceeded in the presence of molecular sieves MS with 5 mol% chiral catalyst 101at room temperature to provide cycloadducts 102 with excellent diastereoselectivity,

enantioselectivity, and in high yield (Scheme 17).

A dramatic improvement was observed in reactions carried out under solvent-free

conditions and excess ethyl vinyl ether 100. Usage of solvents generally resulted in

significantly lower enantioselectivity in the cycloaddition. As the steric bulk of the

alkyl group of dienophile was increased, the selectivity and reactivity decreased.

The optimal dienophile was ethyl vinyl ether. In the solid state, catalyst 101 exists

as a dimeric structure, bridged through a single water molecule and bearing one

terminal water ligand on each chromium center. Opening of a coordination site by

dissociation of the terminal water molecule for complexation of the aldehyde

substrate explains the important role of molecular sieves in these reactions [95, 96].

Asymmetric inverse-electron-demand HDA reaction of 1-oxa-1,3-butadienes was

a key step in synthesis of several members of the bioactive styryllactone family

[97]. Treatment of ethyl vinyl ether 104 with 3-boronoacrolein pinacolate 103 in the

97% eedr 94:6yield 84%

EtO

Me Me

OEt

O

O

+

95 96

OEtO

Me Me

O

OEt

98

97 2 mol%

MS, Et2O, -40oC

O

N

Me Me

NO

Me3C CMe3Cu

H2O OH2

2+OTf2

Scheme 16 Enantioselective HDA approach to the synthon of azaspiracid

5-10 mol%

R1 = H, Br, MeR2 = Me, Et, i-Pr, n-Pr, n-Bu, Ph, 4-MeO-C6H4, 4-NO2-C6H42-NO2-C6H4, CH2OBn, CH2OTBS, CO2Et, OBz

de > 95%ee 89-98%yield 70-95%

(10 equiv)99 100

101

102

(1S, 2R)

MS, neat, 24-120h+

OEtO

R2

R1

O

R1

OEt

R2

O

HR2

R1

Cr

CrClO

N O

Me

Scheme 17 HDA reactions catalyzed by (Schiff base)Cr(III) catalyst 101. Model of one-point binding ofan a,b-unsaturated aldehydes 99 to a Lewis acid center

Top Curr Chem (Z) (2016) 374:24 Page 17 of 37 24

123

presence of Jacobsen’s [(Schiff base)chromium(III)] complex 105 resulted in the

formation of cycloadduct 106 in good yield (85 %) and high enantioselectivity

(96 % ee; Scheme 18). This strategy can be applicable to the synthesis of different

stereoisomers by taking into account both isomers of mandelic acid and the different

chromium(III) complexes.

2.1.5 Enantioselective Approach: Catalytic Hetero-Diels–Alder Reactions of 1-

Oxa-1,3-Butadienes with Chiral Organocatalysts

In recent years, organocatalysis has been established as a very powerful tool for the

synthesis of functional molecules. Asymmetric versions of HDA utilizing chiral

organocatalysts have been developed. Asymmetric aminocatalytic strategies

involving HOMO activation of the dienophile constitute an important alternative

to classical LUMO-lowering pathways [98–100]. Albrech et al. developed the first

H-bond-directed inverse-electron-demand HDA proceeding via a dienamine

intermediate [101, 102]. They evaluated the organocatalytic reaction between

various b,c-unsaturated a-ketoesters 107 and (E)-4-phenylbut-2-enal 108 in the

presence of various aminocatalysts (Scheme 19). The H-bond-directing dienamine

catalyst 109 promoted the inverse-electron-demand HDA reactions.

In most of the cycloadditions of 107 and 108, good yields and high regio- and

stereoselectivities were obtained. High stereoselectivities were observed by employ-

ing a bifunctional squaramide-containing aminocatalyst 109. The authors postulated

that dienamine intermediate is formed by condensation of aminocatalyst 109 with the

a,b-unsaturated aldehyde 108, and the next heterodiene 107 in s-trans conformation is

recognized by the catalyst. Two cycloreactants 107 and 108 are activated through

H-bond interactions and are positioned to facilitate the cycloaddition step.

Most recently, there was considerable interest in applying self-assembled

organocatalysts/modularly designed organocatalysts (MDO) in catalytic reactions.

Zhao et al. demonstrated that MDO self-assembled from proline derivatives and

cinchona alkaloid derived thioureas are highly efficient catalysts for inverse-

electron-demand HDA reactions [103]. They developed highly enantioselective

HDA reactions of electron-deficient enones 111 and aldehydes 112 by using MDO

O

BPin

H OEt O

BPin

OEt

1 mol%

(1S, 2R)

molecular sieves, rt, 2h

CrClO

N O

Me

yield 85%103 104

105

106

Scheme 18 Asymmetric inverse-electron-demand HDA reaction of 3-boronoacrolein pinacolate 103 andethyl vinyl ether 104 as a key step in synthesis of several members of the styryllactone family

24 Page 18 of 37 Top Curr Chem (Z) (2016) 374:24

123

(Scheme 20). Quinidine and (2S, 3aS, 7aS)-octahydro-1H-indole-2-carboxylic acid

(OHIC) are both poor catalysts for the inverse-electron-demand HDA reactions

between aldehydes and electron-deficient enones. However, forming MDO by their

self-assembly with cinchona alkaloid-derived thioureas can improve the efficiency,

reactivity and stereoselectivity of these catalysts. Various aldehydes 112, including

long-chain and branched aldehydes, were found to be excellent substrates for the

MDO-catalyzed HDA reactions (Scheme 20).

The high yield and enantioselectivity of the reactions was restored (up to 95 %

yield and 95 % ee). The ester alkyl group of b,c-unsaturated a-ketoesters 111 has

almost no influence on either the reactivity or enantioselectivity. Similarly, the

substituent on the phenyl ring of the enones 111 has minimal effects on the

reactivity and the asymmetric induction of these reactions. b,c-Unsaturated a-

ketophosphonates 111 may also be applied in these reactions if a higher loading of

the precatalyst modules (10 mol%) is used. The authors proposed a plausible

transition state on the basis of the product 116 stereochemistry and the MDO

structure [103]. They showed that the aldehyde 112 reacts with the OHIC moiety of

the MDO to form an (E)-enamine. Next, the thiourea moiety of the MDO forms

OtBuO2C

R Ph

O107 108 110

OO

N

CF3

CF3

H

N

NH

H 109

DEA, THF, rt, 24-120hO

R

Ph

OtBuO2C

R = Ph, 4-Br-C6H4, 3-Br-C6H4, 2-Br-C6H42-F-C6H4, 4-CF3-C6H4, 4-CO2Me-C6H4,4-NO2-C6H4, 2,6-Cl2-C6H4, 4-CH3-C6H43-pyridyl, Me

yield 42-82%dr 3:1 - (>20:1)ee 79-93%

16 mol%

Scheme 19 HDA reactions between b,c-unsaturated a-ketoesters 107 and (E)-4-phenylbut-2-enal 108 inthe presence of aminocatalyst

O

R2

R1

R3 R4

CHO OR1 OH

R3R2

R4 PCC67-80%

OR1

R3R2

R4

O

R1 = CO2Me, CO2Et, CO2iPr, P(O)(OMe)2, P(O)(OEt)2, P(O)(OiPr)2

R2 = Ph, 4-OMeC6H4, 4-MeC6H4, 4CF3C6H4, 4-ClC6H4, 2-ClC6H4, MeR3 = Me, nPr, nC10H21, Bn, iPr, PhR4 = H, Me

111 112 115 116

toluene, rt, 2-48h

MDO 5 mol%113

114NCOOH

H

H H

N

NH

OMe

N

S

Ar

NH

Scheme 20 MDO catalyzed HDA reactions of electron-deficient enones 111 and aldehydes 112

Top Curr Chem (Z) (2016) 374:24 Page 19 of 37 24

123

hydrogen bonds with the enone 111 and directs to enamine from the front. The

attack of the enone 111 onto the Re face of the enamine in an endo transition state

leads to the formation of the observed (4S, 5R)-product 116.

2.1.6 Enantioselective Approach: Hetero-Diels–Alder Reactions of 1-Oxa-1,3-

Butadienes with Chiral Auxiliaries

Inverse-electron-demand HDA reaction between 1-oxa-1,3-butadienes and electron-

rich alkenes represents one of the most direct approaches for the synthesis of

optically active carbohydrate derivatives. To obtain optically active dihydropyrans

derivatives by the HDA approach, either a catalytic enantioselective reaction or a

chiral transformation via the use of a chiral auxiliary is necessary. The

enantioselective HDA reaction requires chiral 1-oxa-1,3-butadienes or optically

active alkene. The HDA reaction of the a,b-unsaturated ketone 118 prepared in situ

from protected D-xylose 117 was used as the key step for the synthesis of a C10

higher carbon sugar 119 in a one-pot multi-step route (Scheme 21) [104].

Two molecules of a,b-unsaturated ketone 118 undergo the HDA reaction affording

the 10 carbon sugar 119. Reduction and catalytic hydrogenation of cycloadduct 119gave stereoselectively a single product 121 in an excellent yield.

Recently, it was shown that fused uracils, such as pyrano[2,3-d]pyrimidines with an

aryl substituent at carbon C(5) in the ring system can be efficiently synthesized by HDA

reactions of 5-arylidene derivatives of barbituric acids with vinyl ethers [82]. To

increase the potential pharmacological activity of the fused uracil, a sugar moiety can be

introduced instead of an aryl group at the C(5) position of pyrano[2,3-d]pyrimidine.

Therefore, 5-ylidene barbituric acids bearing the carbohydrate substituent were

constructed. A convenient and efficient procedure for the preparation of fused uracils

containing a sugar moiety was described [105]. The reaction sequence was:

Knoevenagel condensation of unprotected sugars and barbituric acid in water,

acetylation of C-glycosides and HDA reaction. The cycloaddition reactions of O-

yield 81%

OOH

BzO

OO

PDC, MeCN, Ac2O

80 oC, 6h

117

O

OO

O

118

OO

OO

O

OO

O

119

NaBH4, EtOH rt, 1h

OO

OO

O

HOO

O

120

OO

OO

O

HOO

O

121

H2/Pd-C, EtOH

40oC, 3h

Scheme 21 Stereoselective synthesis of higher carbon sugar 121 from protected D-xylose 117 by HDAreaction of the a,b-unsaturated ketone 118

24 Page 20 of 37 Top Curr Chem (Z) (2016) 374:24

123

acetylated 1,3-dimethyl-2,4,6-trioxo-pyrimidin-5-ylidene alditols 123, representing an

1-oxa-1,3-butadiene system, with enol ethers 124 were performed in the absence of

solvent at room temperature for 2–5 min and enantiomerically pure cis and trans

diastereoisomers of pyrano[2,3-d]pyrimidines 125–128 with an alditol moiety were

obtained in good 80–87 % yields (Scheme 22). It is worth noting that barbituric acid

5-ylidene alditols 123 are extremely reactive because they underwent smooth HDA

reactions at a temperature of -80 �C as well as at room temperature. Observed

diastereoselectivity of the HDA reactions of 123 and 124 changed in the range from

6.6:1 to 2.1:1. Cycloadducts cis were the major products in all reactions except those of

D-galactose derivatives (Scheme 22). The inverse-electron-demand HDA reactions of

O-acetylated 5-ylidene derivatives 123 with a tenfold excess of cyclic enol ether 129were performed in solvent-less conditions at room temperature for 30 min, and

pyrano[30,20:5,6]pyrano[2,3-d]pyrimidines 130 and 131 were obtained in good

76–78 % yields (Scheme 23). O-Acetylated 1,3-dimethyl-2,4,6-trioxo-pyrimidin-5-

ylidene alditols 123 can act as active heterodienes in HDA reactions and their use in

cycloadditions allows preparation of the enantiomerically pure diastereoisomers of

pyrano[2,3-d]pyrimidines with a sugar moiety.

In the field of pericyclic reactions, the development of new cycloreactants is a

continuous challenge. Dimedone enamines were applied as new dienophiles in HDA

reactions with inverse-electron-demand of 1-oxa-1,3-butadienes [106]. Cycloaddi-

tions of barbituric acid 5-ylidene alditols 132, representing a 1-oxa-1,3-butadiene

system, with dimedone enamines 133 were performed in dichloromethane at room

temperature for 3 days, and fused uracils–chromeno[2,3-d]pyrimidine-2,4-diones

134 were prepared in good (73–87 %) yields (Scheme 24). Only one enantiomer-

ically pure stereoisomer was obtained in each studied cycloaddition. Analysis of

122 minor

+

OR10

L-Xylo, R1=R4=R5=R8=OAc, R2=R3=R6=R7=R9=HL-Arabino, R1=R4=R6=R7=R9=H, R2=R3=R5=R8=OAcD-Gluco, R1=R4=R5=R8=H, R2=R3=R6=R9=OAc, R7=CH2OAcD-Galacto, R1=R4=R6=R8=H, R2=R3=R5=R9=OAc, R7=CH2OAcD-Ribo, R1=R3=R5=R7=R9=H, R2=R4=R6=R8=OAc

N

N OO

R2 R1R4 R3R6 R5

R8R9R7

O

123 124R10 = Et, iBu

N

NO

O

O

Na

O

R2

R5

R6 R1

R4

R3

R7

7R

major

N

N OO

R2 R1R4 R3R6 R5

O

R9 R8

OR10

R7

H

H

trans 126cis 125

5S

7R

N

N OO

R2 R1R4 R3R6 R5

O

R9 R8

OR10

R7

H

H+5R

L-Xylo, L-Arabino, D-Gluco, D-Ribo

or

Ac2O, ZnCl224h, rt

neat, rt2-5 min

5R

minormajortrans 127 cis 128

+N

N OO

R2 R1R4 R3R6 R5

O

R9 R8

OR10

R7

H

H5S

7S

N

N OO

R2 R1R4 R3R6 R5

O

R9 R8

OR10

R7

H

H7S

D-Galacto

Scheme 22 Acetylation of C-glycosides 122 and HDA reactions of 5-ylidene derivatives 123 with enolethers 124. Synthesis of pyrano[2,3-d]pyrimidines 125-128 with a sugar moiety

Top Curr Chem (Z) (2016) 374:24 Page 21 of 37 24

123

proton nuclear magnetic resonance ( 1H NMR) and two-dimensional (2D) NMR

spectra allowed for the determination that cycloadducts 134 exist in solution as a

mixture of the neutral form 134 NF and dipolar ion 134 DI. The prepared fused

uracils, possessing both amine and enol functional groups, share amphiprotic

properties and are zwitterions in solid state. Important for biological interaction,

groups such as different sugar moieties, enol moieties and different amino groups

can be introduced into fused uracil systems by this simple HDA reaction. It was also

shown that different alkenes can be used as dienophiles towards barbituric acid

5-ylidene alditols 132; for example, styrene or 1-amino-2-thiocarbamoyl-cyclopent-

1-ene [106].

The application of stereoselective inverse-electron-demand HDA reaction of

1-oxa-1,3-dienes and chiral allenamides in natural product synthesis was described

by Song et al. [107]. They used this reaction as a key step in synthesis of the C1–C9

subunit of (?)-zincophorin (Scheme 25).

neat, rt, 30 min

O

129

+

L-Xylo, R1=R4=R5=R8=OAc, R2=R3=R6=R7=R9=HL-Arabino, R1=R4=R6=R7=R9=H, R2=R3=R5=R8=OAcD-Gluco, R1=R4=R5=R8=H, R2=R3=R6=R9=OAc, R7=CH2OAC

123

N

N OO

R2 R1R4 R3R6 R5

R8R9R7

O

+

trans-cis-(5,5a:5a,9a)-131cis-cis-(5,5a:5a,9a)-130

9aR5aS

5RN

N OO

R2 R1R4 R3R6 R5

O

R9 R8R7

OH

H

9aR5aS5SN

N OO

R2 R1R4 R3R6 R5

O

R9 R8R7

OH

H

Scheme 23 HDA reactions of O-acetylated 1,3-dimethyl-2,4,6-trioxo-pyrimidin-5-ylidene alditols 123with cyclic enol ether 129

134 NF

N

N OO

X

X

R2 R1R4 R3R6 R5

R8R9R7

O

132 134

N

N OO

X

X

R2 R1R4 R3R6 R5

R8R9R7

O OH

NH Y

O

NH

Y

+

133

CH2Cl2rt, 3 days

yield 73-87%L-Arabino, R1=R4=R6=R7=R9=H, R2=R3=R5=R8=OAc, X=MeL-Xylo, R1=R4=R5=R8=OAc, R2=R3=R6=R7=R9=H, X=MeD-Gluco, R1=R4=R5=R8=H, R2=R3=R6=R9=OAc, R7=CH2OAc, X=MeD-Gluco, R1=R4=R5=R8=H, R2=R3=R6=R9=OAc, R7=CH2OAc, X=HD-Galacto, R1=R4=R6=R8=H, R2=R3=R5=R9=OAc, R7=CH2OAc, X=Me

X = H, MeY-NH = 4-MeC6H4-NH, 4-MeOC6H4-NH, 4-NO2C6H4-NH

MeMe

NH

MeMe

NH

MeMeMe

MeNH

MeMe

NH

OHNO

134 DI

N

N OO

X

X

R2 R1R4 R3R6 R5

R8R9R7

O O

NH Y

H

Scheme 24 HDA reactions of alditols 132 with dimedone enamines 133. Synthesis of fused uracils -chromeno[2,3-d]pyrimidine-2,4-diones 134

24 Page 22 of 37 Top Curr Chem (Z) (2016) 374:24

123

Both reactions 136 with 135 and 137 with 135 provided respectively pyrans 138and 139 in 58 and 54 % yields, as single isomers, after heating in a sealed tube at

85 �C for 48 h in acetonitrile as the solvent (Scheme 25).

2.2 Intramolecular Hetero-Diels–Alder Reactions of 1-Oxa-1,3-Butadienes

2.2.1 Two-Component Domino Knoevenagel Hetero-Diels–Alder Reactions of 1-

Oxa-1,3-Butadienes with an Intramolecular Cycloaddition

The domino Knoevenagel intramolecular hetero-Diels–Alder reaction is one of the

most powerful synthetic routes for the synthesis of various heterocycles and natural

products. This reaction can be used in dihydropyran synthesis [54–68]. In

intramolecular cycloaddition, the 1-oxa-1,3-butadienes are prepared in situ by

Knoevenagel condensation of aldehydes possessing the dienophile moiety and a 1,3-

dicarbonyl compound. A lot of different aldehydes and 1,3-dicarbonyl compounds

such as barbituric acids, Meldrum’s acid, 1,3-cyclohexanedione, dimedone,

4-hydroxycoumarin, indiandiones, pyrazolones, or isooxazolones can be used.

Cis-fused cycloadducts are the main products in intramolecular HDA reactions of

oxabutadienes obtained from aromatic aldehydes [54, 55]. Reactions of oxabuta-

dienes derived from aliphatic aldehydes result in the trans-fused cycloadducts [56,

57]. In recent years, intramolecular HDA reactions of 1-oxa-1,3-butadienes have

been used widely in numerous reactions in organic synthesis due to their economical

and stereo-controlled nature. These reactions allow the formation of two or more

rings at once, avoiding sequential chemical transformations. Therefore, the scope of

the intramolecular HDA reactions of 1-oxadienes was expanded recently. The

influence of an electron-withdrawing group at C-3 in 1-oxa-1,3-butadienes on the

intramolecular HDA reaction was studied. First, the influence of cyano, carbonyl,

and ethoxycarbonyl groups was examined [108]. Next, it was demonstrated that

O

OP

P = TBDPS135

C

NH

N

Ph

O

R

S

CH3CN, 85oC, 48h58%

136137

138139

CH3CN, 85oC, 48h54%

O HN

N

O

Ph

OPS

R

C

NH

N

Ph

OR

S

O HN

N

O

Ph

OP

S

R

dr > 95:5dr > 95:5for (+)-zincophorin

Scheme 25 HDA reactions of chiral enone 135 with chiral allenamides 136 and 137. The key step insynthesis of the C1-C9 subunit of (?)-zincophorin

Top Curr Chem (Z) (2016) 374:24 Page 23 of 37 24

123

sulfur-containing substituents incorporated into 1-oxa-1,3-butadienes positively

influence the results of the cycloaddition. The intramolecular HDA reactions of

sulfenyl-, sulfinyl-, and sulfonyl-activated methylene compounds 140 with

2-alkenyloxy aromatic aldehydes 141 were conducted (Scheme 26) [109].

Knoevenagel condensations of 1-(phenylsulfenyl)-, 1-(phenylsulfinyl)-, and

1-(phenylsulfonyl)-2-propanones 140 with 2-alkenyloxy aromatic aldehydes 141yielded the corresponding condensation products 142 which in turn underwent

intramolecular HDA reactions during heating in boiling toluene or xylene

(Scheme 26). Cis-fused 2H-pyran derivatives 143 were the major products. An

increase of the reactivity and a decrease of the diastereoselectivity of the HDA

reactions were observed in order: PhS derivative, PhSO2 derivative and compounds

containing PhSO group [108, 109].

The most widely used 1-oxa-1,3-butadienes in intramolecular HDA reactions are

usually those where the double bond is placed between the symmetrical 1,3-dicarbonyl

compounds. Shanmugasundaram et al. studied the heterocycloaddition in which the

alkene part was flanked by a keto carbonyl and a lactone carbonyl [110]. The reactions

of 4-hydroxy coumarin and its benzo-analogues 144 with O-prenylated aromatic

aldehydes 145 were examined (Scheme 27). Pyrano fused polycyclic compounds 147and 148 were prepared with a high degree of chemoselectivity by the application of

microwave irradiation. These reactions offers an easy access to pyrano[3,2-

c]coumarin 147 which is a structural element of many natural products.

Chemoselectivity was achieved with the reduction in reaction time because the

cycloadducts 147 and 148 formed in the ratios ranging from 79:21–95:5 when the

reactions were carried out under microwave irradiation for 10–150 s. Reactions of

unsymmetrical 1,3-diones 144 with citronellal were also described [110].

Surprising formation of a 2,3-dihydro-4H-pyran containing 14-membered macro-

cycle 151 by sequential olefin cross metathesis and a highly regiospecific intramolecular

HDA reaction of 1-oxa-1,3-dienes was described by Prasad and Kumar (Scheme 28)

[111]. They studied the reaction of a hydroxydienone 149 derived from tartaric acid with

Grubbs’ second generation catalyst. Presence of the unprotected hydroxyl group in the

hydroxyenone led to the formation of macrocycle 151. Protection of the hydroxyl group

resulted in the ring-closing metathesis product 150.

The authors made the experiment to show that obtaining macrocycle 151involves the formation of intermediate 155. Dimerization of hydroxyamide 152 by

PhS(O)n

OR1

CHO

O

R2 R3

O

OR1

PhS(O)n

R2 R3

O

OR1

PhS(O)n

R2R3

H

HEDDA, MeCN

rt

toluene or xylene

reflux, 10-24h

n = 0, 1, 2R1 = Me, Ph

cis/trans 2.5:1 - (>100:1)

140 141 341241R2 = Me, PhR3 = Me, H

Scheme 26 Domino Knoevenagel intramolecular HDA reactions of sulfenyl, sulfinyl and sulfonylactivated methylene compounds 140 with 2-alkenyloxy aromatic aldehydes 141. Synthesis of polycyclic2H-pyran derivatives 143

24 Page 24 of 37 Top Curr Chem (Z) (2016) 374:24

123

olefin cross metathesis with Grubbs second generation catalyst gave the bis-amide

153 (Scheme 28). Protection of the two hydroxyl groups in 153 as the bis-silyl ether

154 and then the reaction with 2-propenylmagnesium bromide resulted in formation

of the macrocycle 156. Deprotection of the silyl ethers in 156 furnished the

macrocycle 151 in 85 % yield. These studies represent the first example of a tandem

olefin cross metathesis HDA reaction sequence.

Wada et al. developed a new type of intramolecular HDA reaction of 1-oxa-1,3-

butadienes–tandem transetherification-intramolecular HDA reaction. Heterodienes

were obtained in situ by a transetherification under thermal conditions from b-

alkoxy-substituted a,b-unsaturated carbonyl compounds bearing an electron-with-

drawing substituent and d,e-unsaturated alcohols [112]. This tandem reaction

proceeded stereoselectively to afford trans-fused hydropyranopyrans. Next, chiral

C

144 145146

147

without MI: yield 40-75%147/148 55:45 - 80:20

with MI: yield 74-92%147/148 79:21 - 95:5

O O

OO

H

H

A

B

148

O O

OH

A

B

O

CHO

CEtOH, reflux, 4-8h orMicrowave Irradiation (MI) 10-150s

O O

O O

A

B

C

O O

O OH

H

A

B

C

Scheme 27 Microwave accelerated domino Knoevenagel intramolecular HDA reactions of 4-hydroxycoumarin and its benzo-analogues 144 with O-prenylated aromatic aldehydes 145

149

O

O

O

OH

O

O

O

OH

Grubbs' secondgen. cat. (1-5 mol%)

CH2Cl2

151

O

O

O

HO OH

O

O

MeO

Me

150

O

O

O

OH

NMe2

O

O

O

NMe2

OH OH

O

NMe2

O

OGrubbs' secondgen. cat. (5 mol%)

CH2Cl2, 50oC6h, 80%

TBDMSClimidazole, DMAP

CH2Cl2reflux, 6h, 87%

O

O

O

NMe2

RO OR

O

NMe2

O

O

R = TBDMS

TBAF, THFrt, 12h, 85% O

O

O

RO OR

O

O

MeO

Me

152 153 154

155

2-Propenyl magnesium bromideTHF, -15oC, 0.5h, 83%

156

O

O

O

RO OR

O

O

MeO

Me

151

O

O

O

HO OH

O

O

MeO

Me

Scheme 28 Synthesis of macrocycle 151 by sequential cross metathesis and HDA reaction

Top Curr Chem (Z) (2016) 374:24 Page 25 of 37 24

123

Lewis acid catalysts were used in this new type of transformation. Wada et al.

examined the catalytic asymmetric tandem transetherification-intramolecular HDA

reaction of methyl (E)-4-methoxy-2-oxo-3-butenoate 157 with d,e-unsaturated

alcohols 158 (Scheme 29) [113]. The optically active catalyst derived from the

(S,S)-tert-Bu-bis(oxazoline) and Cu(SbF6)2 in presence of molecular sieves was a

highly effective Lewis acid catalyst. The trans-fused hydropyranopyran derivatives

160 were prepared in yields up to 83 % and with high enantiomeric excess up to

98 %. In order to prevent the acid-induced cyclization, molecular sieves were used

as a dehydratation agent.

Yadav et al. presented the synthesis of carbohydrate analogues, cis-fused chiral

polyoxygenated (tricyclic, tetracyclic, and pentacyclic) heterocycles by domino

Knoevenagel intramolecular HDA reactions [114]. TheO-prenyl derivative of a sugar

aldehyde 161 derived from D-glucose underwent reactions with 1,3-diones 162, 164,

166 and 168 in presence of sodium acetate in acetic acid at 80 �C (Scheme 30). The

reactions were highly stereoselective affording exclusively cis-fused furopyranopy-

rans 163, 165, 167 and 169 in 70–82 % yields. The authors suggested that the

cycloadditions proceeded in a concerted manner via an endo-E-syn transition state.

2.2.2 Catalytic Intramolecular HDA Reaction of 1-Oxa-1,3-Butadienes and Alkynes

Due to the lower reactivity of alkynes in comparison to the corresponding alkenes,

no HDA reaction of 1-oxa-1,3-butadienes with alkynes has been reported. Recently,

different Lewis acids have provided new opportunities for various catalytic alkyne

reactions. Some of the most frequently used transition metal catalysts are

copper(I) compounds. Khoshkholgh et al. studied the intramolecular HDA reaction

of 1-oxa-1,3-butadiene and an alkyne in the presence of CuI [115]. The Williamson

reaction of propargyl bromide 171 and salicylaldehydes 170 afforded compounds

172 (Scheme 31). The 1-oxa-1,3-butadienes 174 were prepared through Knoeve-

nagel reaction of O-propargylated salicylaldehyde derivatives 172 and barbituric

acids 173 with yields between 75 and 94 %. Intramolecular HDA reactions were

CO2Me

O

MeO OHR1 R1

R2 R2

157 158- MeOH

CuN

O

N

O

RR

2+

2X - R = t-Bu, X = OTfR = t-Bu, X = SbF6R = Ph, X = SbF6

R1 = H, MeR2 = H, Me

Chiral Cat. 10 mol%, rt transetherification

160 yield 75 - 83 % ee 3 -98 %

159

160

O

O

R2

R2

R1R1

CO2Me

Chiral Cat. 10 mol%CH2Cl2, rt, 1 - 96 hMS

O

O

R2

R2

R1R1

CO2MeH

H

Scheme 29 Catalylic enantioselective tandem transetherification-intramolecular HDA reaction ofmethyl (E)-4-methoxy-2-oxo-3-butenoate 157 with d,e-unsaturated alcohols 158

24 Page 26 of 37 Top Curr Chem (Z) (2016) 374:24

123

carried out in the presence of CuI (40 mol%) and water as the solvent and

tetracyclic uracils 175 were prepared in 70–84 % yields. The authors explained that

the initial step of cycloaddition was probably the formation of a p-complex with CuI

since copper(I) salts can act as a p-electrophilic Lewis acid. This complexation of

alkyne increases activity toward an 1-oxa-1,3-butadiene system (Scheme 31).

Yadav et al. developed a novel method for the synthesis of sugar-annulated

tetracyclic- and pentacyclic-heterocycles in a single-step operation [116]. The O-

O

O

O

162

168

169

O

R1

OR2

R1 = R2 = MeR1 = Me, R2 = OMe

O

O

R

R

R = H, Me

164

N

N O

O

O

Me

Me166

O

O

H

O

O

O

161

AcONa 3 equiv., AcOH 80oC, 6.5 h

AcONa 3 equiv., AcOH 80oC, 8.5 h

AcONa 3 equiv., AcOH 80oC, 5 - 6 h

AcONa 3 equiv., AcOH 80oC, 5 - 6 h

O

O O

OO

OOH

H

163

O O

OO

O

H

R1 O

R2

H

O O

OO

OOH

R

R

H

165

N

N

O O

OO

OOH

O

Me

H

Me

167

yield 82 %

yield 80-82 %

yield 72 %

yield 70-79 %

Scheme 30 Domino Knoevenagel HDA reactions of sugar aldehyde 161 with 4-hydroxycoumarin 162,cyclohexa-1,3-dione or dimedone 164, 1,3-dimethylbarbituric acid 166 and 1,3-diones 168

6-28 h,

BrK2CO3

DMF, rt

CHO

OHR1

CHO

OR1

H2Oi or ii

N

N

O

OO

R2

R2

O

R1R1 = 5-Br, 5-NO2, 3-MeO, naphthaldehyde

i: R2 = H, 10% HClii: R2 = Me, 20% (NH4)2HPO4

N

N

O

OO

R2

R2

O

R1

A B

C

D

N

N

O

OO

R2

R2

CHO

OR1

170 171 172

173 174 175172

CuI 40 mol%

H2O, reflux

yield 70-84%

Scheme 31 Synthesis of O-propargylated aldehydes 172 and fused uracils - oxa-helicene derivatives 175by intramolecular HDA reaction

Top Curr Chem (Z) (2016) 374:24 Page 27 of 37 24

123

propargyl derivative of a sugar aldehyde 176 derived from D-glucose undergoes

smooth domino Knoevenagel intramolecular HDA reactions with 1,3-diketones 177,

179, 181 and 1-aryl-pyrazol-5-ones 183 in the presence of CuI/Et3N in refluxing

methanol (Scheme 32).

Furopyranopyrans 178, 180, 182 and 184 were prepared in 75–90 % yields. The

authors assumed that these cycloadditions proceed in a concerted way via an endo-

E-syn transition state [116]. Acyclic 1,3-diketones such as acetyl acetone and ethyl

acetoacetate can’t be used in the reaction.

Another example of catalytic intramolecular HDA reaction of 1-oxa-1,3-

butadienes and an alkyne is domino Knoevenagel intramolecular HDA reaction

of indolin-2-thiones 185 and O-propargylated salicylaldehyde derivatives 186 in the

presence of ZnO (Scheme 33) [117].

The major advantage of this reaction is the fact that pentacyclic indole

derivatives 188 can be isolated by filtration from the reaction mixture. This method

also has advantages such as the use of commercially available, non-toxic and

inexpensive ZnO as catalyst, low loading of catalyst, and high yields of products.

Balalaie et al. described domino Knoevenagel intramolecular HDA reactions of

O-propargylated salicylaldehyde 190 with active methylene compounds 189 in the

presence of ZrO2-nanopowder (NP) as a Lewis acid in ionic liquid and different

organic solvents (Scheme 34) [118]. The reactions were carried out for different

active methylene compounds 189 such as barbituric acid, N,N-dimethyl barbituric

acid, indandione, Meldrum’s acid, and pyrazolone.

The solutions of aldehyde 190 and appropriate 1,3-dicarbonyl compound 189, ZrO2

and base in ionic liquid or organic solvent were stirred for 5–40 min at room temperature

O

OH

O

167 178

O

O

H

O

O

O

176

O

O

R

R

R = H, Me179 180

N

N O

O

O

Me

Me

181 182

NN

OAr

Me

Ar = C6H5, 4-MeC6H4, 4-FC6H4, 3-ClC6H4

183yield 75-80 %

O

O O

OO

OOH

O O

OO

OOH

R

R

N

N

O O

OO

OOH

O

Me

Me

184

yield 75%

yield 84-90 %

yield 85%

O O

OO

O

HN

N

Me

Ar

CuI 30 mol%/Et3N 100 mol%

MeOH, reflux, 4 h

CuI 30 mol%/Et3N 100 mol%

MeOH, reflux, 3.5-4 h

CuI 30 mol%/Et3N 100 mol%

MeOH, reflux, 4.5 h

CuI 30 mol%/Et3N 100 mol%

MeOH, reflux, 4.5-5.5 h

Scheme 32 Domino Knoevenagel HDA reactions of sugar aldehyde 176 with 4-hydroxycoumarin 167,cyclohexa-1,3-dione or dimedone 179, 1,3-dimethylbarbituric acid 181 and 1-aryl-pyrazol-5-ones 183

24 Page 28 of 37 Top Curr Chem (Z) (2016) 374:24

123

and desired products were obtained in 80–95 % yields. The best results were obtained

for 5-nitro-O-propargylated salicylaldehyde and 1-butyl-3-methylimidazolium nitrate

[bmim][NO3] as the reaction medium. Balalaie et al. also used ionic liquid [bmim][NO3]

in the presence of 30 mol% CuI in the domino Knoevenagel HDA reactions of O-

propargylated salicylaldehydes with some active methylene compounds [119].

2.2.3 Two-Component Domino Knoevenagel Hetero-Diels–Alder Reactions of 1-

Oxa-1,3-Butadienes with Intramolecular Cycloaddition in Water or Solvent-

Free

Water is the solvent of choice for nature to carry out syntheses of complex organic

molecules. Water is a clean, inexpensive, environment friendly reaction medium.

Therefore, the choice of water as the solvent for organic reactions in the laboratory

synthesis is obvious. However, water as a solvent was ruled out from organic

reactions. It has been changed in 1980 by the pioneering work of Breslow and

Rideout, who demonstrated that Diels–Alder reactions of water-soluble reagents

would be greatly accelerated in aqueous solution [120]. In 2005, Sharpless et al.

demonstrated that the Diels–Alder reaction of the water-insoluble reactants showed

substantial rate acceleration in aqueous suspension over homogeneous solution

[121, 122]. Some examples of domino Knoevenagel HDA reactions of 1-oxa-1,3-

butadienes with intramolecular cycloaddition in water as the solvent are described

below. Moghaddam et al. examined domino Knoevenagel HDA of 4-hydroxy-

dithiocoumarin 194 and O-acrylated salicylaldehyde derivatives 193 in water

(Scheme 35) [123]. Pentacyclic heterocycles 196 and 197 were formed by a

catalyst-free method in good yields and with high regio- and stereo-selectivity.

Aldehydes 193 underwent the Knoevenagel condensation with 4-hydroxy-

dithiocoumarin 194 in water at reflux to give the intermediates 195 in which two

different heterodiene fragments were presented. The thiocarbonyl group of the

thioester 195 reacted as heterodiene. The cycloadducts were obtained as a mixture

of cis- and trans-isomers. The authors observed the influence of the substituent R2

on reaction diastereoselectivity. The trans-isomer 196 was the main product for

some reactions whereas for others, the products 197 were formed with the

predominance of the cis-isomers (Scheme 35).

The importance of quinoline and its fused derivatives prompted Baruah and Bhuyan

to study the domino Knoevenagel intramolecular HDA reactions of 3-formyl quinoline

NR1

S

OR2

NS

O

R1

R2

R1 = Me, Et, Ph R2 = H, Br, OMe, NO2 yield 85-95%

185 186 187 188

NR1

S

CHO

O

R2

ZnO 10 mol%, MeCN3h, reflux

Scheme 33 ZnO-catalyzed synthesis of indole-thiopyrano-chromene derivatives 188 via dominoKnoevenagel intramolecular HDA reactions

Top Curr Chem (Z) (2016) 374:24 Page 29 of 37 24

123

containing a dienophile moiety [124]. Appropriate 1-oxa-1,3-butadienes were prepared

from acetanilides 198 by treatment with Vilsmeier reagent (Scheme 36).

To introduce the dienophile in compound 201, the reactions of 2-chloro-3-

formylquinolines 199 with alcohol 200 in presence of aqueous sodium hydroxide

under phase transfer catalytic conditions were used. Domino Knoevenagel HDA

reactions of 201 and active methylene compound 202, 204 and 206 in presence of

piperidine at room temperature in water gave the cis fused penta or hexacyclic

pyrano[2,3-b]quinoline derivative 203, 205 or 207 with high yield (70–80 %) and

diastereoselectivity ([99 %). The products were isolated by filtration in the pure

form from aqueous medium.

Baruah and Bhuyan also studied the HDA reactions for other 1-oxa-1,3-

butadienes (Scheme 37) [124]. These compounds possessing diene and dienophile

moieties were prepared from aldehydes 209 by treatment with N-allyl methyl amine

210 in presence of K2CO3. Obtained products 211 on treatment with 212 or 213 in

presence of piperidine in water at room temperature afforded the cycloadducts 214or 215 in 52–70 % yields.

Parmar et al. used alkenyl- and alkynyl-ether-tethered ketones instead of

aldehydes to extend the substrate scope in domino Knoevenagel intramolecular

CHO

O

X

189 190

NN O

Ar Ar = C6H5, 3-ClC6H4

O

O

solvent = MeCN, H2O, EtOH, MeOH, t = 360 - 2400 min, 192 yield 75-90% solvent = [bmim][NO3], t = 2 - 75 min, 192 yield 81-95 %X = H, 5-Me, 5-NO2, 5-Cl

solvent = MeCN, H2O, EtOH, MeOH, [bmim][NO3]base = DMAP, Et3N

189 = N

N

O

OO

H

H

N

N

O

OO

Me

Me

O

O

O

O

O

O

191 192

X

OO

O

X

O

OOZrO2-NP 20-40 mol%, rtsolventbase

Scheme 34 Domino Knoevenagel- HDA reactions of O-propargylated salicylaldehydes 190 with activemethylene compounds 189 in the presence of ZrO2-nanopowder

H2O, 3h

refluxR1 = H, OMeR2 = H, Br, NO2

R3 = Me, Ph 195O

S

O

O

R2

R1

R3S

CHO

O R3

R2

R1

O

S

OH

S

+

193 194

total yield 60-90% 196/197 4:96 - 97:3

O

S

S

OH

H

R3

H

O

R2

R1

O

S

S

OH

H

HR3

O

R2

R1

+

196 197

Scheme 35 Synthesis thiochromone-annulated thiopyranocoumarins 196 and 197 by dominoKnoevenagel HDA reaction

24 Page 30 of 37 Top Curr Chem (Z) (2016) 374:24

123

HDA reactions of 1-oxa-1,3-butadienes. They synthesized dihydropyran derivatives

219 and 220 as single stereoisomers with a tertiary ring junction carbon by solvent-

free one-pot procedure in the presence of tetrabutylammonium-hydrogensulfate

(Scheme 38) [125].

3 Application of Inverse-Electron-Demand Hetero-Diels–AlderReactions of 1-Oxa-1,3-Butadienes in Bioorthogonal Chemistry

For chemical biologists, discovering new reactions which can expand the toolbox of

bioorthogonal chemistry is a current challenge. Development of new orthogonal

methods for labeling in the biosystems is still continued, although effective

bioorthogonal reactions such as copper-free click chemistry have been developed

[126]. Reactions which can be used in bioorthogonal click chemistry should meet

the requirements: high reactivity and selectivity of reagent functional groups,

chemical stability in aqueous solutions in vivo, biocompatibility and high reaction

N

CHO

O

ROH

TBAB50% NaOHaq

198 199

200

201

N

N

O

OO

Me

Me

N O

R

O

NN

H

H

O

O

Me Me

OO

N O

R

OH

H

O

NN

Me

O Ph

N O

R

OH

H

NN

Me

Ph

202

203 yield 76 - 80%

N

H

H2O, rt, 3h

N

H

H2O, rt, 3h

N

H

H2O, rt, 4h

204

205 yield 70 - 75%

206

207 yield 74 - 80%

N

CHO

O

R

201

R

NHCOCH3

R = H, Me

N

CHO

Cl

RDMF/POCl380oC

R = H, Me

Scheme 36 Synthesis of 3-formyl quinoline 201 containing dienophile moiety. Domino KnoevenagelHDA reactions of 201 and active methylene compounds 202, 204 and 206

Top Curr Chem (Z) (2016) 374:24 Page 31 of 37 24

123

rate under physiological conditions [127–130]. Bioorthogonal ligations have been

widely used in biomedical research since they can selective labels of biomolecules

in living systems. Some of inverse-electron-demand HDA reactions 1-oxa-1,3-

butadienes developed in recent years fulfill the criteria of click chemistry compiled

by Sharpless [131, 132] and, in the future, can be used as bioorthogonal

cycloaddition. There is only one example in the literature of application of

inverse-electron-demand HDA reactions of 1-oxa-1,3-butadienes in bioorthogonal

chemistry. Lei et al. described a new bioorthogonal ligation by click HDA

cycloaddition of in situ-generated o-quinolinone quinone methides and vinyl

thioethers [133]. High selectivity and the fact that this cycloaddition can proceed

smoothly under aqueous conditions make it suitable for bioorthogonal chemistry. o-

Quinone methides represent an 1-oxa-1,3-butadiene system which can undergo

quick and selective inverse-electron-demand HDA cycloadditions. It is important

that generation of the o-quinone methides can’t be conducted in harsh reaction

conditions because it could be harmful for the organism cells. HDA cycloadditions

of photochemically generated o-naphthoquinone methides 222 with vinyl ethers or

enamines 223 as dienophiles were described by Arumugam and Popik (Scheme 39)

[134–136]. They used ultraviolet (UV) light to generate 1-oxa-1,3-butadienes 222.

Li et al. optimized both reaction partners to make the reaction suitable for

bioorthogonal ligation [133]. Introduction of more electronegative nitrogen into a

heterodiene system 221 improved its reactivity and hydrophilicity (Scheme 40). As

dienophile was used small and chemically stable in vivo vinyl thioether 227. o-

Quinolinone quinine methide 226 was prepared from 8-(hydroxymehyl)-2-

methylquinolin-7-ol 225 without use of catalyst and UV light. Cycloreactants 226

N

CHO

N

R

Me208 209

210

211

NHMe

K2CO3 DMF

N

H

H2O, rt, 5h

N

H

H2O, rt, 6h

N

N

O

OO

Me

Me

212

N

CHO

N

R

Me211N

N

Me

O Ph

213

N N

R

O

NN

H

H

O

O

Me Me

Me214 yield 65 - 70%

N N

R

OH

H

NN

Me

Ph

Me215 yield 52 - 55%

R

NHCOCH3

R = H, Me

N

CHO

Cl

RDMF/POCl380oC

R = H, Me

Scheme 37 Synthesis of 3-formyl quinoline 211 containing dienophile moiety. Domino KnoevenagelHDA reactions of 211 and active methylene compounds 212 and 213

24 Page 32 of 37 Top Curr Chem (Z) (2016) 374:24

123

and 227 underwent HDA cycloaddition under physiological conditions (37 �C,

H2O). The authors used this bioorthonogal cycloaddition for labeling of proteins and

imaging of taxol derivatives inside live cells.

Li et al. proved that HDA cycloaddition of o-quinolinone quinine methide 226and vinyl thioether 227 can be utilized for labeling of multiple biomolecules in

complex living systems when it is combined with other methods [137]. When

cycloreactants 225 and 227 were combined with 5,6-didehydro-11,12-didehy-

drodibenzo[a,e]cyclooctene 229 and (azidomethyl)benzene 230 in a mixture of

H2O/CH3CN at 37 �C, only products 228 and 231 were obtained, and no cross

reaction products were prepared (Scheme 41). 1,3-Dipolar cycloaddition of azide

230 and alkyne 229 is widely used in bioorthogonal ligation as strain-promoted

azide alkyne cycloaddition (SPAAC). The results indicated that these two ligations

proceeded simultaneously without interfering with each other.

It seems that some of the HDA reactions described in Chapter 2 can be used in

bioorthogonal chemistry in the future because they are selective, non-toxic, and can

function in biological conditions taking into account pH, an aqueous environment,

and temperature.

216

R2

R2O

O

NN

Me

H

R3

R3

R1

Me

R2

R2O

O

NN

Me

R1

MeN

N

Me

O

R1

R2

R2

O

O

R3

R3

R2

R2

O

O

217

218

219

220

R1 = Ph, 3-Cl-C6H4,4-Me-C6H4

R2 = H, NO2 R3 = H, Meyield 71-80%

yield 65-77%

ZnO 25 mol%TBA-HS 25 mol%

neat, 160oC, 2-3.5h

ZnO 25 mol%TBA-HS 25 mol%

neat, 160oC, 4-6h

R2 = H, NO2

Scheme 38 Domino Knoevenagel HDA reaction of 2-(alkenloxy)- and 2-(alkynyloxy)acetophenones217 and 218 with pyrazolones 216. Synthesis of chromeno-fused pyrazoles 219 and 220

224O

R1

R3R2

OH

OH hv

H2OO R2 R3

R1

221 222 223R1, R2 = H, R3 = OCH3

R1, R2 = H, R3 = OCH2OHR1 = H, R2 = CH3; R3 = H

yield 94-100%

R1 = n-BuH, R2 = H; R3 = CH3

R1, R2 = CH3, R3 = morpholine

Scheme 39 HDA cycloadditions of photochemically generated o-naphthoquinone methides 222 withvinyl ethers or enamines 223

Top Curr Chem (Z) (2016) 374:24 Page 33 of 37 24

123

4 Conclusion

This review article is an effort to summarize recent developments in inverse-

electron-demand HDA reactions of 1-oxa-1,3-butadienes. Some of the papers

related to the inverse-electron-demand HDA reactions of 1-oxa-1,3-butadienes

found in the literature clearly demonstrate the importance of this transformation

which opened up efficient and creative routes to different natural products

containing six-membered oxygen ring systems. This type of cycloaddition is today

one of the most important methods for the synthesis of dihydropyrans which are the

key building blocks in carbohydrate derivative synthesis. Especially, the domino

Knoevenagel HDA reactions have been frequently applied for the synthesis of

natural products. The main advantage of the inverse-electron-demand HDA reaction

of oxabutadienes is formation of dihydropyran derivatives with up to three

stereogenic centers in one step from simple achiral precursors. This transformation

characterizes the huge diversity, excellent efficiency, high regioselectivity,

diastereoselectivity, and enantioselectivity observed in many cases. In recent years,

the use of chiral Lewis acids, chiral organocatalysts, new heterodienes, or new

dienophiles have given enormous progress. Recently, HDA reactions of 1-oxabu-

tadienes conducted without a solvent or in water were developed and the results

suggested that the presented green methods may displace other methods that use

various organic solvents and that are performed at high temperature. Application of

inverse-electron-demand HDA reactions of 1-oxa-1,3-butadienes in bioorthogonal

chemistry is still challenging because there is only one example of this

bioorthogonal cycloaddition in the literature. The author of this review sincerely

hopes that this article will stimulate future research in bioorthogonal inverse-

N OH

OH

37oC

H2O

N O

SOH

N O

SOH

622522 227 228 yield 92%

Scheme 40 HDA cycloaddition of o-quinolinone quinine methide 226 and vinyl thioether 227 underphysiological conditions

N OH

OH

N O

SOH

225 227 228

SOH

N3

229 230 231

H2O/CH3CN 1:1, 37oC

NN

N

Scheme 41 HDA reaction of 225 with 227 and SPAAC of 229 with 230 proceeded simultaneouslywithout interfering with each other

24 Page 34 of 37 Top Curr Chem (Z) (2016) 374:24

123

electron-demand cycloaddition of 1-oxa-1,3-butadienes and will encourage scien-

tists to design novel bioorthogonal ligations.

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0

International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, dis-

tribution, and reproduction in any medium, provided you give appropriate credit to the original

author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were

made.

References

1. Boger DL, Weinreb SN (1987) Hetero Diels–Alder methodology in organic synthesis. Academic

Press, San Diego

2. Tietze LF, Kettschau G (1997) Top Curr Chem 189:1–120

3. Tietze LF, Harfiel U, Hubsch T, Voss E, Wichmann J (1991) Chem Ber 124:881–888

4. Tietze LF, Voss E, Herms K, Sheldrick GM (1985) Tetrahedron Lett 26:5273–5276

5. Tietze LF (1990) J Heterocycl Chem 27:47–69

6. Tietze LF (1996) Chem Rev 96:115–136

7. Tietze LF, Beifuss U (1993) Angew Chem Int Ed Engl 32:131–163

8. Tietze LF, Beifuss U (1993) Angew Chem 105:137–170

9. Jurczak J, Tkacz M (1979) J Org Chem 44:3347–3352

10. Konował A, Jurczak J, Zamojski A (1975) Rocz Chem 42:2045–2049

11. Jung ME, Shishido K, Light L, Davis L (1981) Tetrahedron Lett 22:2045–2049

12. Danishefsky S, Kerwin JF Jr, Kobayashi S (1982) J Am Chem Soc 104:358–360

13. Daniewski WM, Kubak E, Jurczak J (1985) J Org Chem 50:3963–3965

14. Bednarski M, Danishefsky SJ (1986) J Am Chem Soc 108:7060–7067

15. Garner P, Ramakanth S (1986) J Org Chem 51:2609–2612

16. Jurczak J, Gołebiewski A, Ram A (1986) Tetrahedron Lett 27:853–856

17. Jurczak J, Gołebiewski A, Raczko J (1988) Tetrahedron Lett 29:5975–5978

18. Gołebiewski A, Raczko J, Jacobsson U, Jurczak J (1991) Tetrahedron 47:1053–1064

19. Mikami K, Motoyama Y, Terada M (1994) J Am Chem Soc 116:2812–2820

20. Lowe RF, Stoodley RJ (1994) Tetrahedron Lett 35:6351–6354

21. Lubineau A, Acrostanzo H, Queneau J (1995) Carb Chem 14:1307–1328

22. Johannsen M, Jorgensen KA (1995) J Org Chem 60:5757–5762

23. Lehmler HJ, Nieger M, Breitmaier E (1996) Synthesis 105–110

24. Desimoni G, Tacconi V (1975) Chem Rev 75:651–692

25. Snowden RL, Sonnay P, Ohloff G (1981) Helv Chim Acta 64:1247–1256

26. Woods GF, Sanders H (1946) J Am Chem Soc 68:2483–2485

27. Tietze LF (1974) Chem Ber 107:2491–2498

28. Buback M, Tost W, Hubsch T, Voss E, Tietze LF (1989) Chem Ber 122:1179–1186

29. Matsumoto K, Sera A, Uchida T (1985) Synthesis 1–26

30. Dauben WWG, Krabbenhoht HO (1977) J Org Chem 42:282–287

31. Rappoport Z (1994) The chemistry of enamines in the chemistry of functional groups. Wiley, New

York

32. Hickmott PW (1982) Tetrahedron 38:1975–2050

33. Hickmott PW (1982) Tetrahedron 38:3363–3446

34. John RA, Schmidt V, Wyler H (1987) Helv Chim Acta 70:600–606

35. Zhuo JC, Wyler H (1995) Helv Chim Acta 78:151–164

36. Dvorak D, Arnold Z (1982) Tetrahedron Lett 23:4401–4402

37. Haag-Zeino B, Schmidt RR (1990) Liebigs Ann Chem 1197–1203

38. Tietze LF, Harfiel U, Hubsch T, Voss E, Bogdanowicz-Szwed K, Wichmann J (1991) Liebigs Ann

Chem 275–281

39. Tietze LF, Voss E (1986) Tetrahedron Lett 27:6181–6184

40. Tietze LF, Fennen J, Wichmann J (1992) Chem Ber 125:1507–1511

41. Boger DL, Robarge KD (1988) J Org Chem 53:3373–3377

Top Curr Chem (Z) (2016) 374:24 Page 35 of 37 24

123

42. Sera A, Ohara M, Yamada H, Egashira E, Ueda N, Setsune J (1988) Bull Chem Soc Jpn

67:1912–1915

43. Sera A, Ohara M, Yamada H, Egashira E, Ueda N, Setsune J (1990) Chem Lett 2043–2046

44. Pale P, Bouquant J, Chuche J, Carrupt PA, Vogel P (1994) Tetrahedron 50:8035–8052

45. Wada E, Pei W, Yasuoka H, Chin U, Kanemasa S (1996) Tetrahedron 52:1205–1220

46. Merour JY, Chichereau R, Desarbre E, Gadonneix P (1996) Synthesis 519–524

47. Hiroi K, Umemura M, Tomikawa Y (1993) Heterocycles 35:73–79

48. Tietze LF, Saling P (1992) Synlett 281–282

49. Wada E, Yasuoka H, Kanemasa S (1994) Chem Lett 1637–1640

50. Thorhauge J, Johannsen M, Jorgensen KA (1998) Angew Chem Int Ed 37:2404–2406

51. Thorhauge J, Johannsen M, Jorgensen KA (1998) Angew Chem 110:2543–2546

52. Kumar K, Waldmann H, Eschenbrenner-Lux V (2014) Angew Chem Int Ed 53:11146–11157

53. Kumar K, Waldmann H, Eschenbrenner-Lux V (2014) Angew Chem 126:11326–11337

54. Tietze LF, Stegelmeier H, Harms K, Brumby T (1982) Angew Chem Int Ed 21:863–864

55. Tietze LF, Stegelmeier H, Harms K, Brumby T (1982) Angew Chem 94:868–869

56. Tietze LF, Kiedrowski G, Harms K, Clegg W, Sheldrick GM (1980) Angew Chem Int Ed

19:134–135

57. Tietze LF, Kiedrowski G, Harms K, Clegg W, Sheldrick GM (1980) Angew Chem 92:130–131

58. Tietze LF, Brumby T, Pfeiffer T (1988) Liebigs Ann Chem 9–12

59. Tietze LF, Ott C, Gerke K, Buback M (1993) Angew Chem Int Ed 32:1485–1486

60. Tietze LF, Ott C, Gerke K, Buback M (1993) Angew Chem 105:1536–1538

61. Tietze LF, Brand S, Pfeiffer T, Antel J, Harms K, Sheldrick GM (1987) J Am Chem Soc

109:921–923

62. Tietze LF, Brumby T, Pretor M, Remberg G (1988) J Org Chem 53:810–820

63. Tietze LF, Brumby T, Brand S, Bratz M (1988) Chem Ber 121:499–506

64. Tietze LF, Brand S, Brumby T, Fennen J (1990) Angew Chem Int Ed 29:665–667

65. Tietze LF, Brand S, Brumby T, Fennen J (1990) Angew Chem 102:675–677

66. Tietze LF, Kiedrowski G, Berger B (1982) Synthesis 683–684

67. Tietze LF, Bachmann J, Wichmann J, Burkhardt O (1994) Synthesis 1185–1194

68. Tietze LF, Meier H, Nutt H (1989) Chem Ber 122:643–650

69. Pałasz A, Jelska K, O _zog M, Serda P (2007) Monatsh Chem 138:481–488

70. Pałasz A (2012) Monatsh Chem 143:1175–1185

71. Pałasz A, Bogdanowicz-Szwed K (2008) Monatsh Chem 139:647–655

72. Bogdanowicz-Szwed K, Pałasz A (1997) Monatsh Chem 128:1157–1172

73. Bogdanowicz-Szwed K, Pałasz A (1995) Monatsh Chem 126:1341–1348

74. Bogdanowicz-Szwed K, Pałasz A (2001) Z Naturforsch B 56:416–422

75. Klahn P, Kirsch SF (2014) Eur J Org Chem 3149–3155

76. Xing Ch, Li X, Zhu S, Zhao J, Zhu S (2006) Tetrahedron Lett 47:4951–4955

77. Pałasz A, Pałasz T (2011) Tetrahedron 67:1422–1431

78. Pałasz A (2005) Org Biomol Chem 3:3207–3212

79. Tietze LF, Evers H, Topken E (2001) Angew Chem Int Ed 40:903–905

80. Tietze LF, Evers H, Topken E (2001) Angew Chem 113:927–929

81. Radi M, Bernardo V, Bechi B, Castagnolo D, Pagano M (2009) Tetrahedron Lett 50:6572–6575

82. Pałasz A (2008) Monatsh Chem 139:1397–1404

83. Pałasz A, Cie _z D (2015) Eur J Med Chem 97:582–611

84. Pałasz A (2010) Synthesis 23:4021–4032

85. Hanessian S, Compain P (2002) Tetrahedron 58:6521–6529

86. Dujardin G, Leconte S, Coutable L, Brown E (2001) Tetrahedron Lett 42:8849–8852

87. Gong J, Bonfand E, Brown E, Dujardin G, Michelet V, Genet JP (2003) Tetrahedron Lett

44:2141–2144

88. Gohier F, Bouhadjera K, Faye D, Gaulon K, Maisonneuve V, Dujardin G, Dhal R (2007) Org Lett

9:211–214

89. Gaulon C, Dhal R, Chapin T, Maisonneuve V, Dujardin G (2004) J Org Chem 69:4192–4202

90. Evans DA, Johnson JS, Olhava EJ (2000) J Am Chem Soc 122:1635–1649

91. Audrain H, Thorhauge J, Hazell RG, Jorgensen KA (2000) J Org Chem 65:4487–4497

92. Evans DA, Kverno L, Dunn T, Beauchemin A, Raymer B, Mulder JA, Olhava EJ, Juhl M,

Kagechika K, Favor DA (2008) J Am Chem Soc 130:16295–16309

93. Dossetter AG, Jamison TF, Jacobsen EN (1999) Angew Chem Int Ed 38:2398–2400

24 Page 36 of 37 Top Curr Chem (Z) (2016) 374:24

123

94. Dossetter AG, Jamison TF, Jacobsen EN (1999) Angew Chem 111:2549–2552

95. Gademann K, Chavez DE, Jacobsen EN (2002) Angew Chem Int Ed 41:3059–3061

96. Gademann K, Chavez DE, Jacobsen EN (2002) Angew Chem 114:3185–3187

97. Favre A, Carreaux F, Deligny M, Carboni B (2008) Eur J Org Chem 4900–4907

98. Han B, Li JL, Ma C, Zhang SJ, Chen YC (2009) Angew Chem Int Ed 48:5474–5477

99. Han B, Li JL, Ma C, Zhang SJ, Chen YC (2009) Angew Chem 121:5582–5585

100. Albrecht Ł, Dickmeiss G, Cruz-Acosta F, Rodriguez-Escrich C, Davis RL, Jorgensen KA (2012) J

Am Chem Soc 134:2543–2546

101. Albrecht Ł, Dickmeiss G, Weise ChF, Rodriguez-Escrich C, Jorgensen KA (2012) Angew Chem Int

Ed 51:13109–13113

102. Albrecht Ł, Dickmeiss G, Weise ChF, Rodriguez-Escrich C, Jorgensen KA (2012) Angew Chem

124:13286–13290

103. Sinha D, Perera S, Zhao JCG (2013) Chem Eur J 19:6976–6979

104. Liu HM, Zou DP, Zhang F, Zhu WG, Peng T (2004) Eur J Org Chem 2103–2106

105. Pałasz A, Kalinowska-Tłuscik J, Jabłonski M (2013) Tetrahedron 69:8216–8227

106. Pałasz A, Cie _z D, Musielak B, Kalinowska-Tłuscik J (2015) Tetrahedron 71:8911–8924

107. Song Z, Hsung RP, Lu T, Lohse AG (2007) J Org Chem 72:9722–9731

108. Bogdanowicz-Szwed K, Pałasz A (1999) Monatsh Chem 130:795–807

109. Bogdanowicz-Szwed K, Pałasz A (2001) Monatsh Chem 132:393–401

110. Shanmugasundaram M, Manikandan S, Raghunathan R (2002) Tetrahedron 58:997–1003

111. Prasad KR, Kumar SM (2013) Tetrahedron 69:6512–6518

112. Wada E, Kumaran G, Kanemasa S (2000) Tetrahedron Lett 41:73–76

113. Wada E, Koga H, Kumaran G (2002) Tetrahedron Lett 43:9397–9400

114. Yadav JS, Reddy BVS, Narsimhaswamy D, Lakshmi NP, Narsimulu K, Srinivasulu G, Kunwar AC

(2004) Tetrahedron Lett 45:3493–3497

115. Khoshkholgh MJ, Balalaie S, Gleiter R, Rominger F (2008) Tetrahedron 64:10924–10929

116. Yadav JS, Reddy BVS, Gopal AVH, Rao RN, Somaiah R, Reddy PP, Kunwar AC (2010) Tetra-

hedron Lett 51:2305–2308

117. Kiamehr M, Moghaddam FM (2009) Tetrahedron Lett 50:6723–6727

118. Balalaie S, Poursaeed A, Khoshkholgh MJ, Bijanzadeh HR, Wolf E (2012) C R Chim 15:283–289

119. Balalaie S, Azizian J, Shameli A, Bijanzadeh HR (2013) Synth Commun 43:1787–1795

120. Rideout DS, Breslow R (1980) J Am Chem Soc 102:7816–7817

121. Narayan S, Muldoon JM, Finn GV, Fokin V, Kolb HC, Sharpless KB (2005) Angew Chem Int Ed

44:3275–3279

122. Narayan S, Muldoon JM, Finn GV, Fokin V, Kolb HC, Sharpless KB (2005) Angew Chem

117:3339–3343

123. Moghaddam FM, Kiamehr M, Khodabakhshi MR, Mirjafary Z, Fathi S, Saeidian H (2010)

Tetrahedron 66:8615–8622

124. Baruah B, Bhuyan PJ (2009) Tetrahedron 65:7099–7104

125. Parmar NJ, Pansuriya BR, Labana BM, Sutariya TR, Kant R, Gupta VK (2012) Eur J Org Chem

5953–5964

126. Baskin JM, Prescher JA, Laughlin ST, Agard NJ, Chang PV, Miller IA, Lo A, Codelli JA, Bertozzi

CR (2007) Proc Natl Acad Sci USA 104:16793–16797

127. Ning X, Temming RP, Dommerholt J, Guo J, Blanco D, Debets MF, Wolfert MA, Boons GJ, Van

Delft FL (2010) Angew Chem Int Ed 49:3065–3068

128. Ning X, Temming RP, Dommerholt J, Guo J, Blanco D, Debets MF, Wolfert MA, Boons GJ, Van

Delft FL (2010) Angew Chem 122:3129–3132

129. Bertozzi CR, Sletten EM (2009) Angew Chem Int Ed 48:6974–6998

130. Bertozzi CR, Sletten EM (2009) Angew Chem 121:7108–7133

131. Kolb HC, Finn MG, Sharpless KB (2001) Angew Chem Int Ed 40:2004–2021

132. Kolb HC, Finn MG, Sharpless KB (2001) Angew Chem 113:2056–2075

133. Li Q, Dong T, Liu X, Lei X (2013) J Am Chem Soc 135:4996–4999

134. Arumugam S, Popik VV (2011) J Am Chem Soc 133:5573–5579

135. Arumugam S, Popik VV (2011) J Am Chem Soc 133:15730–15736

136. Arumugam S, Popik VV (2012) J Am Chem Soc 134:8408–8411

137. Li Q, Dong T, Liu X, Zhang X, Yang X, Lei X (2014) Curr Org Chem Soc 18:86–92

Top Curr Chem (Z) (2016) 374:24 Page 37 of 37 24

123