Next Article in Journal
Chemical Constituents of the Root of Jasminum giraldii
Previous Article in Journal
A New Vanadium (III) Complex of 2,6-Bis(3,5-diphenylpyrazol-1-ylmethyl)pyridine as a Catalyst for Ethylene Polymerization
Previous Article in Special Issue
Suzuki-Miyaura Cross-Coupling in Acylation Reactions, Scope and Recent Developments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of the Suzuki-Miyaura Reaction in the Synthesis of Flavonoids

by
Mamoalosi A. Selepe
and
Fanie R. Van Heerden
*
School of Chemistry and Physics, University of KwaZulu-Natal, Private Bag X01, Scottsville 3209, South Africa
*
Author to whom correspondence should be addressed.
Molecules 2013, 18(4), 4739-4765; https://doi.org/10.3390/molecules18044739
Submission received: 19 March 2013 / Revised: 10 April 2013 / Accepted: 19 April 2013 / Published: 22 April 2013
(This article belongs to the Special Issue Suzuki Reaction)

Abstract

:
The application of the Suzuki-Miyaura reaction in the synthesis of flavonoids, an important class of natural products, is reviewed. This reaction has not only been employed to provide access to flavonoid nuclei, but has also been applied to the synthesis of dimeric flavonoids and in the synthesis of libraries of flavonoid derivatives for biological activity studies. The classes of flavonoids that are discussed are the chalcones, flavones, isoflavones, neoflavones, biflavones and derivatives of flavonoids obtained by C-C bond formation via the Suzuki-Miyaura reaction.

1. Introduction

Flavonoids are plant secondary metabolites with a C6C3C6 skeleton and can be divided into three main classes, i.e., flavonoids, isoflavonoids and neoflavonoids (4-arylcoumarins). Chalcones, the biogenetic precursor to flavonoids, are often also classified as flavonoids (Figure 1). Different oxidation states and different substituents contribute to the diversity of flavonoid structures. Flavonoids play an important role in plant physiology and are of interest to humans as a result of biological activities such as antioxidant, anticancer and estrogenic activity of individual flavonoid derivatives.
Figure 1. Different classes of flavonoids.
Figure 1. Different classes of flavonoids.
Molecules 18 04739 g001
As a result of the biological activity of flavonoids, there is an interest in the development of synthetic procedures that can conveniently give access to these molecules and their derivatives. One of the methods that has recently been employed successfully in the synthesis of flavonoid scaffolds is the Suzuki-Miyaura reaction. The Suzuki-Miyaura reaction normally involves insertion of palladium into a sp2-hybridized C-X bond and consequently the major application of this reaction is in the construction of the flavonoid nucleus of chalcones, flavones, isoflavones and neoflavones rather than in the synthesis of their reduced derivatives. Unlike other methods that have been employed in the synthesis of flavonoids, the Suzuki-Miyaura reaction often employs mild conditions that are compatible with a variety of functional groups. This enables the synthesis of flavonoids of natural origin and derivatives from precursors bearing sensitive substituents [1,2,3,4]. Moreover, the method readily offers access to a variety of flavonoids for biological activity studies by using different organoboron starting materials in the final stages of the synthesis [4,5,6,7,8,9,10]. The method is amenable to large scale synthesis due to the stability and commercial availability of a wide range of boronic acids/esters, and the ease of working up the reaction mixture [6,11,12]. In the following sections, the syntheses of the four classes of flavonoids and some of their dimeric analogues using the Suzuki-Miyaura reaction are reviewed.

2. Chalcones

2.1. Chalcone Monomers

Chalcones are readily accessible by the Claisen-Schmidt condensation of benzaldehydes with acetophenones [13,14,15,16]. Other routes that have been developed include acylation of phenols with cinnamic acids [17], Heck coupling of aryl iodides with aryl vinyl ketones [18], and palladium-catalyzed coupling of arylpropargyl alcohols with aryl halides [19,20].
The synthesis of chalcones by the Suzuki-Miyaura reaction was first demonstrated by Eddarir and co-workers in 2006 [21]. Their strategy was based on two pathways; the first one involved coupling of arylboronic acids 1 with cinnamoyl chloride (2), whereas the second pathway involved coupling of styrylboronic acid (5) with benzoyl chlorides 4 (Scheme 1). Moderate yields (41–51%) of chalcones 3 were obtained from pathway A, when using Haddach and McCarthy’s conditions [anhydrous toluene, Pd(PPh3)4, Cs2CO3] [22]. These conditions gave the chalcones in good to excellent yields (68–93%) from pathway B.
Scheme 1. Preparation of chalcones by Eddarir and co-workers [21].
Scheme 1. Preparation of chalcones by Eddarir and co-workers [21].
Molecules 18 04739 g008
Since most naturally-occurring chalcones are oxygenated on the aromatic rings, the application of route B was further extended to the synthesis of methoxylated chalcone 7 in a good yield from 3,4-dimethoxybenzoyl chloride and styrylboronic acid 6 (Scheme 2) [21]. Al-Masum et al. used a similar approach to prepare a number of unnatural chalcones by using potassium styryltrifluoroborates 8 as the starting material and performing the reaction under microwave conditions (Scheme 3) [23].
Scheme 2. Preparation of a methoxylated chalcone by the coupling of a benzoyl chloride and a styrylboronic acid [21].
Scheme 2. Preparation of a methoxylated chalcone by the coupling of a benzoyl chloride and a styrylboronic acid [21].
Molecules 18 04739 g009
Scheme 3. Preparation of chalcones by the coupling of a benzoyl chloride with a potassium styryltrifluoroborates [23].
Scheme 3. Preparation of chalcones by the coupling of a benzoyl chloride with a potassium styryltrifluoroborates [23].
Molecules 18 04739 g010
In 2008, Xin reported the synthesis of aryl ketones by the reaction of arylboronic acids with benzoic anhydride [24]. The reaction was extended to the reaction of styrylboronic acid with benzoic anhydride (9) in the presence of PdCl2 and Na2CO3 in H2O/acetone (1:1) to give the chalcone in 78% yield (Scheme 4) [24]. However, the functional group tolerance of this method is yet to be tested on chalcones with varying substitution patterns including those of natural origin.
Scheme 4. Synthesis of a chalcone by cross-coupling of benzoic anhydride and phenylboronic acid [24].
Scheme 4. Synthesis of a chalcone by cross-coupling of benzoic anhydride and phenylboronic acid [24].
Molecules 18 04739 g011
Zuo et al. obtained a number of aryl-substituted chalcones by preparing either the aryl-substituted acetophenone or benzaldehyde via the Suzuki-Miyaura reaction, followed by the traditional Claisen-Schmidt synthesis of the chalcones. Most of the synthesized compounds had moderate to strong anticancer activity against five cancer cell lines, and have NF-kB nuclear translocation inhibition activities [25]. Vieira et al. used a similar approach by using the Claisien-Schmidt reaction to prepare brominated chalcones which were subjected to the Suzuki-Miyaura reaction to yield arylated chalcones [26]. In this investigation, the emphasis was on developing reaction conditions that are environmentally benign and they introduced the use of the non-toxic polyethyleneglycol as solvent for the reaction.

2.2. Bichalcones

Bichalcones consist of two chalcone monomers linked by either an ether or a C-C bond. The latter are the most prevalent class of bichalcones, while the former type has been reported from the Rhus genus only [6]. Mihigo et al. reported the total synthesis of rhuschalcone VI (10), an antiplasmodial C-C bridged natural bichalcone which consists of two molecules of isoliquiritigenin (11) [6] (Figure 2). The synthetic protocol for 10 was based on the Suzuki reaction for the construction of the C-C biaryl linkage and solvent-free aldol condensation for the synthesis of the chalcone monomers [14].
Figure 2. Structures of a rhuschalcone and its monomer.
Figure 2. Structures of a rhuschalcone and its monomer.
Molecules 18 04739 g002
Initially, it was envisaged that rhuschalcone VI (10) and its derivatives would be prepared by the Suzuki coupling of chalconylboronate ester 12 and bromochalcone 13 followed by demethylation (Scheme 5). However, the synthesis of chalconylboronate esters from bromochalcones by lithium-halogen exchange and reaction of the lithiated species with 2-isopropoxy-4,4,5,5-tetramethyl-1,3,2-dioxaborolane proved to be challenging. Thus, the reaction sequence was changed [6].
Scheme 5. The first proposed synthetic route for rhuschalcone VI (10) [6].
Scheme 5. The first proposed synthetic route for rhuschalcone VI (10) [6].
Molecules 18 04739 g012
In the modified route (Scheme 6), a boronate ester 15 was formed in the early stages from protected bromoacetophenone 14. Coupling of 15 with bromochalcone 13 in the presence of Pd(PPh3)4 and subsequent acetyl deprotection gave arylated chalcone 16. Solvent-free Claisen-Schmidt condensation of 16 with p-methoxybenzaldehyde gave a dimeric chalcone 17, which was completely or partially demethylated to give rhuschalcone VI (10) and its analogues [14]. This synthetic strategy was further expanded to the synthesis of other unnatural arylated chalcones for structure-activity relationship studies, by reaction between boronate ester 15 and bromochalcones with different substitution patterns [6].
Scheme 6. Alternative route for the synthesis of rhuschalcone VI (10).
Scheme 6. Alternative route for the synthesis of rhuschalcone VI (10).
Molecules 18 04739 g013

3. Flavones

3.1. Flavone Monomers

Classical approaches to the synthesis of flavones are the Baker-Venkataraman rearrangement, followed by cyclization of the resulting β-diketone [27,28,29], the Allan-Robinson [30] and the Algar-Flynn-Oyamada methods [31,32]. Although other classes of flavonoids have been successfully prepared by the Suzuki-Miyaura reaction, there are not many reports on the synthesis of the flavone nucleus by this method. The main challenge with this procedure is the difficulty of accessing the prerequisite precursors for the Suzuki-Miyaura reaction, which are the 2-halochromones.
Kraus and Gupta reported the synthesis of flavones by the palladium-catalyzed coupling of 2-chlorochromone 21 and arylboronic acids 23ac (Scheme 7) [33]. Even though the Suzuki-Miyaura reaction is normally conducted with bromides and iodides, it is noteworthy that the reaction of 2-chlorochromone 21 with boronic acids under standard Suzuki conditions gave good yields of the corresponding flavones 24ac (68–72%). 2-Chlorochromone 21 was in turn prepared from a sequence that involved esterification of phenol 18 with 3,3-dichloroacrylic acid, followed by a Fries rearrangement and base-catalyzed cycloelimination of the resulting dichloroketone 19a. An attempt to prepare a 2-bromochromone from dibromoketone 19b gave an intermediate 22 which was coupled with boronic acids 23ac to give aurones 25ac (Scheme 7).
A number of authors have synthesized flavones by using classical methodologies, but have used the Suzuki-Miyaura reaction to introduce additional aryl or heterocyclic groups onto the flavone skeleton [34,35,36,37,38,39,40,41].
Scheme 7. Preparation of flavones and aurones by the Suzuki-Miyaura reaction [33].
Scheme 7. Preparation of flavones and aurones by the Suzuki-Miyaura reaction [33].
Molecules 18 04739 g014
Amentoflavone (26) consists of two apigenin (27) units connected at the C-8 and C-3' positions (Figure 3).
Figure 3. Structures of amentoflavone and its monomer.
Figure 3. Structures of amentoflavone and its monomer.
Molecules 18 04739 g003

3.2. Biflavones

The methods that have been employed for the synthesis of C-C linked dimeric flavones are the Ullmann reaction [42] and metal-catalyzed cross-coupling reactions such as the Stille and Suzuki-Miyaura reactions [43,44,45]. The application of the Suzuki-Miyaura reaction to the synthesis of biflavones was first demonstrated by Muller and Fluery in the synthesis of amentoflavone (26) and its methyl derivatives [44].
Two synthetic strategies were explored for the synthesis of derivatives 31 and 32 of 26. The first route involved coupling of 8-flavonylboronic acids 29a or 29b with 3'-iodoflavones 30a and 30b and subsequent cleavage of the protecting groups (Scheme 8) [44]. The 8-flavonylboronic acids 29a and 29b were prepared by lithiation of 8-iodoflavones 28a and 28b, respectively, followed by quenching of the lithiated species with triisopropyl borate and aqueous work up.
Scheme 8. Preparation of derivatives of amentoflavone derivatives [44].
Scheme 8. Preparation of derivatives of amentoflavone derivatives [44].
Molecules 18 04739 g015
The second route proceeded by coupling of 3'-iodoflavone 30a with 2,4,6-trimethoxyphenylboronic acid to give the arylated flavone 33 (Scheme 9). The second flavone ring was constructed by acylation of 33 with p-methoxycinnamic acid, followed by oxidative cyclization. This gave a naturally-occurring hexamethyl derivative of amentoflavone, dioonflavone (34) [44].
In 1998, Zembower and Zhang prepared an inhibitor of the hepatitis B virus, robustaflavone (39) (Scheme 10) [45]. Robustaflavone (39) differs from amentoflavone (26) in the connection between the apigenin monomers which are between C-6 and C-3' instead of C-8 and C-3'.
Scheme 9. Second route for the preparation of amentoflavone derivatives [44].
Scheme 9. Second route for the preparation of amentoflavone derivatives [44].
Molecules 18 04739 g016
Scheme 10. Preparation of robustaflavone (39) [45].
Scheme 10. Preparation of robustaflavone (39) [45].
Molecules 18 04739 g017
The key steps in the synthesis of robustaflavone (39) were TlOAc-mediated C-6 iodination of apigenin derivate 36 to give the 6-iodoflavone 37 and synthesis of the 3'-flavonylboronate (35) by palladium-catalyzed cross-coupling of 3'-iodoflavone 30a with bis(pinacolato)diboron (Scheme 10) [45]. An attempt to convert the 3'-iodoflavone 30a into the corresponding boronic acid by the procedure employed in the synthesis of amentoflavone (26) was unsuccessful. The two precursors were coupled under standard Suzuki conditions to give 38, which was demethylated with BBr3 to give the target compound robustaflavone (39).
The synthesis of C-C linked biflavones by the Suzuki-Miyaura protocol has also been extended to the preparation of unnatural biflavonoids. An example is the gem-difluoromethylenated biflavone 42, which was prepared by Zheng and co-workers by the palladium-catalyzed coupling of 3'-flavonylboronate ester 41 with 6-iodoflavone 40 (Scheme 11) [46]. In this case the regioselective iodination at position 6 was achieved by using AgOAc/I2 instead of the more toxic TlOAc.
Scheme 11. Preparation of gem-difluoromethylenated biflavones.
Scheme 11. Preparation of gem-difluoromethylenated biflavones.
Molecules 18 04739 g018
More recently, the unnatural biflavones 43 and 44 were prepared by the Suzuki-Miyaura reaction of appropriately functionalized flavones using Pd(PPh3)4 as a catalyst [43]. The biflavones 43 and 44 together with other unnatural biflavones synthesized by the Stille coupling reaction were tested for inhibition of group II secretory phospholipase A2 (SPLA2IIA) and their inhibitory activity compared with that of the natural biflavones amentoflavone (26) and ochnaflavone. Of the active compounds, 44 (Figure 4) was also found to exhibit inhibitory potency comparable to that of amentoflavone (26) whereas 43 (Figure 4) exhibited a weaker activity [43].
Figure 4. Structures of two unsubstituted biflavones.
Figure 4. Structures of two unsubstituted biflavones.
Molecules 18 04739 g004

4. Isoflavones

4.1. Isoflavones

There are two long-established procedures for the preparation of isoflavones which are still widely used, i.e., the deoxybenzoin route and the chalcone route [47,48,49]. Other methods which have been developed include reductive cleavage of isoxazoles, intramolecular ketene cycloaddition followed by decarboxylation, rearrangement and cyclisation of chalcone epoxides and rearrangement of flavanones [49,50]. The methods that have been developed more recently are the Wacker-Cook tandem conversion of α-methylene deoxybenzoins into isoflavones [51] and the Cu(I)-mediated cyclization of 3-(2-bromophenyl)-3-oxopropanal [52]. Regardless of the many new synthetic approaches presented, the application of many of them has not been demonstrated in the synthesis of polyhydroxylated isoflavones and isoflavones bearing other naturally-occurring substitution patterns.
In 1988, Suzuki and co-workers were the first to demonstrate the versatility of the palladium-catalyzed cross-coupling reactions in the synthesis of isoflavones from 3-bromochromones 45 and arylboronic acids/esters (Scheme 12) [53]. Following this, the Suzuki-Miyaura cross-coupling reaction has been applied in several instances to the synthesis of isoflavones in the presence of palladium(0) or palladium(II) catalysts [3,11,12,54,55]. Examples of catalysts which have successfully facilitated this C-C bond formation are Pd(PPh3)4[1,53], (C) [11,12], trans-[PdCl2(2-ethyl-2-oxazoline-к1N)2] [55], Pd(dppf)2Cl2 [3,56], benzothiazole-oxime-based Pd(II) catalyst [54], and Pd(OAc)2 in the presence of poly(ethyleneglycol) [56,57] or 2-(2,6-dimethoxybiphenyl)dicyclohexylphosphane (SPhos) [4].
Scheme 12. Preparation of isoflavones by the Suzuki reaction [53].
Scheme 12. Preparation of isoflavones by the Suzuki reaction [53].
Molecules 18 04739 g019
Unlike the 2-halochromone derivatives required in the synthesis of flavones by the Suzuki reaction, the 3-halochromone precursors for isoflavones can be conveniently prepared by Gammill’s protocol [58], which involves condensation of appropriately substituted 2'-hydroxyacetophenones 46 with DMF-DMA to form enaminoketones 47. Halogen-mediated ring closure of the enaminoketones gives the corresponding 3-halochromones 48 [58] (Scheme 13). Alternatively, the 3-halochromones can be obtained by direct halogenation of chromones [4].
Scheme 13. Preparation of 3-halochromones by Gammill’s protocol [58].
Scheme 13. Preparation of 3-halochromones by Gammill’s protocol [58].
Molecules 18 04739 g020
In a variation of the Suzuki-Miyaura reaction, Tsoi et al. reported the formation of an unsubstituted isoflavone by the palladium-catalyzed oxidative cross-coupling of a diazochromone (49) with phenylboronic acid in 60% yield (Scheme 14) [59].
Scheme 14. Oxidative cross coupling of phenylboronic acid with a diazochromone [59].
Scheme 14. Oxidative cross coupling of phenylboronic acid with a diazochromone [59].
Molecules 18 04739 g021

4.2. Preparation of Isoflavone Analogues for Biological Activity Studies

As a result of the readily availability of 3-halochromones and their potential to be elaborated into a wide range of compounds by coupling with different boronic acids in the final stages of the synthesis, several groups have taken advantage of the Suzuki coupling reaction for the synthesis of libraries of novel compounds based on the isoflavone scaffold for biological activity studies [4,60,61]. For example, Vasselin et al. synthesized a series of fluoro, methoxy, nitro and amino isoflavones (50ao) from appropriately substituted 3-iodochromones and arylboronic acids (Scheme 15) [61]. However, an attempt to prepare 5,7-dibenzyloxy-3-iodochromone from 4,6-dibenzyloxy-2-hydroxyacetophenone resulted in an inseparable mixture of products while a poor yield (29%) was obtained for 3-iodo-5,7-dimethoxychromone from the corresponding enaminoketone.
The synthesized isoflavones were tested for in vitro growth inhibition of human breast (MDA-MB-468 and MCF-7) and colon (HT29 and HT-116) cancer cell lines. The isoflavones 50d, 50f, 50h, 50k, 50l and 50o showed pronounced growth inhibition of MDA-MD-468 cells co-incubated with TBDD, a powerful inducer of cytochrome P450 (CYP)-1A1 activity. This suggested that the isoflavone derivatives were potential substrates for (CYP)-1A1 bioactivation [61].
In 2008, Wei and Yu synthesized 26 isoflavone glycoside derivatives based on the structure of a potent inhibitor of α-glucosidases of rat liver microsomes, 7-O-α-D-arabinofuranosyloxy-4',8-dihydroxyisoflavone (53a), also called A-76202 [4]. The analogues of 53a were prepared by coupling 3-bromochromones 52 to differently substituted arylboronic acids in the presence of Pd(OAc)2, SPhos and K2CO3 in a water-based solvent (H2O/acetone) (Scheme 16). However, in certain instances the reaction conditions were slightly altered and coupling was performed in the absence of a ligand, with NaOAc as base and MeOH as solvent to accommodate substrates which were more susceptible to degradation.
Scheme 15. Preparation of a series of isoflavones by Vasselin and co-workers [61].
Scheme 15. Preparation of a series of isoflavones by Vasselin and co-workers [61].
Molecules 18 04739 g022

50RR'
aH4-Cl
bH4-Br
cH3-NO2
dH3-OMe
eH4-OMe
fH3,4-di-OMe
g6-F4-OMe
h6-F3,4-di-OMe
i6-F3,4-OCH2O
j7-F4-OMe
k7-F3,4-di-OMe
l8-F3,4-di-OMe
m6,7-di-F3,4-di-OMe
n6,8-di-F3,4-di-OMe
o7,8-di-F3,4-di-OMe
The 3-bromochromones 52 were in turn prepared by direct bromination of the chromones 51 at C-3 by treatment with PhI(OAc), TMSBr and pyridine [4]. The synthetic natural product 53a and its derivatives were evaluated for α-glucosidase inhibition [4]. The results showed that the stereochemistry of the α-D-arabinofuranosyl unit and the 8-hydroxy group in the A-ring are essential for the activity, whereas modifications at the B-ring did not adversely affect the α-glucosidase inhibitory activity of the isoflavone 7-O-glycosides [4]. It is noteworthy that the 3-OMe and 4-NMe2 derivatives (53e and 53k, respectively) were three fold more active than the parent compound 53a (Scheme 16) [4].
Scheme 16. Synthesis of isoflavone glycosides by the Suzuki-Miyaura reaction [4].
Scheme 16. Synthesis of isoflavone glycosides by the Suzuki-Miyaura reaction [4].
Molecules 18 04739 g023

62RR'R''IC50 (μM) *
aOHH4-OH0.018
bOHHH0.040
cOHH4-OMe0.022
dOHH2-OMe0.015
eOHH3-OMe0.006
fOHH4-Me0.035
gOHH2-Me0.020
hOHH3-Me0.050
iOHH3-OH0.017
jOHH4-F0.015
kOHH4-NMe20.008
lOHH4-NHBoc0.020
mOHH4-CF30.030
nOHH4-SiMe3(35% at 0.050)
oHOHH10
pHOH4-OMe8
qHOH2-OMe10
rHOH3-OMe10
sHOH4-Me7
tHOH3-Me5
uHOH4-OH15
vHOH3-OH(33% at 50)
* Inhibition of α-glucosidase.
Matin and co-workers [5] also prepared a series of isoflavones by coupling of 3-iodochromone 54 and a variety of arylboronic acids in the presence of Pd(C) [11], followed by cleavage of the THP protecting group (Scheme 17). The isoflavones were screened together with other compounds (chalcones, flavones, flavanones, isoflavones and pyrazole derivatives) for dual PPARα and γ agonism. Of the 77 tested compounds, the isoflavones 55a, 55c, 55e and 55i were identified as novel potent dual PPARα and γ agonists, which could serve as future leads in PPAR-related disorders that include type II diabetes mellitus and metabolic syndrome [5].
In 2007, Kigoshi’s group reported the total synthesis glaziovianin A (60), a metabolite of Astelia glazioviana that exhibited cytotoxicity against HL-60 cells [60,62]. As illustrated in Scheme 18, the synthesis of glaziovianin A (60) commenced by preparing the 3-iodochrome 57 from acetophenone 56 by Gammill’s procedure [58], and the boronic acid 59 from aryl derivative 58. Coupling of 3-iodochromone 57 with phenylboronic acid 59 gave glaziovianin A (60) [60,62].
Scheme 17. Preparation of isoflavones by Matin and co-workers [5].
Scheme 17. Preparation of isoflavones by Matin and co-workers [5].
Molecules 18 04739 g024

55XR55XR
aC3,5-di-OMejC2-OMe, 3,5-di-F
bC4-OMekC3,4,5-tri-F
cC3,4-OCH2O-lN2,4-di-OMe
dC4-CF3mC4-Cl
eC4-FnC3-F
fC3,4-di-OMeoC2-OMe
gC3,4-O(CH2)2-OpC3-OCF3
hC2,4-di-FqC3-OBn
iC3-OMerC3,4,5-tri-OMe
Scheme 18. Total synthesis of glaziovianin A (60) [62].
Scheme 18. Total synthesis of glaziovianin A (60) [62].
Molecules 18 04739 g025
Thereafter, they employed the same strategy to prepare glaziovianin analogues 61ai by altering the substituents on the A- and B-rings (Scheme 19) [60]. The synthesized compounds were tested for cytotoxicity against HeLa S3 cells. Of the screened compounds, the 7-O-allyl derivative 61i (IC50 = 0.19 μM) was found to be more cytotoxic than the parent compound 60 (IC50 = 0.59 μM) [60]. Moreover, 61i was found to be a more potent M-phase inhibitor [60].
Scheme 19. Preparation of glaziovianin A derivatives [60].
Scheme 19. Preparation of glaziovianin A derivatives [60].
Molecules 18 04739 g026

61RR'
a6,7-di-OMe2',5'-di-OMe, 3',4'-OC(CH3)2O-
b6,7-di-OMe2',3',4',5'-tetra-OMe
c6,7-di-OMe3,4-OCH2O-
d6-OMe, 7-OTHP2',5'-di-OMe, 3',4'-OCH2O-
e5,6,7-tri-OMe2',5'-di-OMe, 3',4'-OCH2O-
f6-OMe, 7-OH2',5'-di-OMe, 3',4'-OCH2O-
g6-OMe, 7-OBn2',5'-di-OMe, 3',4'-OCH2O-
h6-OMe, 7-O-propargyl2',5'-di-OMe, 3',4'-OCH2O-
i6-OMe, 7-O-allyl2',5'-di-OMe, 3',4'-OCH2O-

4.3. Synthesis of Soy Isoflavones

The soy isoflavones consist of genistein (62), daidzein (63), formononetin (55b), biochanin A (64) and the less common glycitein (65) [50,57,63,64] (Figure 5). They normally exist as 7-O-glycosides, which are metabolized into the aglycones [50,65]. These isoflavones, particularly genistein, are frequently referred to as phytoestrogens, because they are non-steroidal plant-derived compounds with the ability to bind to estrogen receptors and modulate estrogenic responses [57,63,65,66]. The estrogenic properties of the soy isoflavonoids have been extensively studied with regards to health benefits [50,63,65,66,67]. The consumption of phytoestrogen-rich food has been linked to protection against hormone-dependent breast cancer and prostate cancer, alleviation of postmenopausal disorders, osteoporosis as well as cardiovascular protection [50,63,66,67].
Figure 5. Structures of soy isoflavones.
Figure 5. Structures of soy isoflavones.
Molecules 18 04739 g005
The synthesis of genistein (62) by the Suzuki-Miyaura reaction was first reported in 2010 [2,64], while that of formononetin was reported by Matin and co-workers in 2009 in the preparation of a series of isoflavones for biological activity studies [5]. Priefer’s group synthesized genistein (62) in five steps as shown in Scheme 20, which involved preparation of 3-iodo-5,7-dimethoxymethoxychromone (67) from phloroacetophenone (66) by Gammill’s procedure and coupling of the 3-iodochromone 67 with commercially available 4-hydroxyphenylboronic acid using Pd(OAc)2, poly(ethylene glycol) (PEG10000) and Na2CO3. The MOM-protecting groups on the resulting isoflavone 68 were removed with HCl to give genistein (62) [64].
Scheme 20. Preparation of genistein (62) by Priefer’s group [64].
Scheme 20. Preparation of genistein (62) by Priefer’s group [64].
Molecules 18 04739 g027
In the same year Selepe et al. concurrently reported the synthesis of genistein from the MOM-protected 3-iodochromone 67 (Scheme 21) [2]. Their synthesis involved preparation of the boronic acid 70 from aryl iodide 69 in a one-pot sequence that involved addition of n-BuLi to a solution of aryl iodide 69 and triisopropyl borate in THF/Et2O (1:2), followed by the hydrolysis of the boronate ester with an NH4Cl solution. Heterogeneous Pd(C)-assisted cross-coupling of boronic acid 70 with 3-iodochromone 67, prepared by Gammill’s protocol and removal of the MOM protecting groups gave genistein (62).
Scheme 21. Preparation of genistein (62) by Selepe et al. [2].
Scheme 21. Preparation of genistein (62) by Selepe et al. [2].
Molecules 18 04739 g028
The syntheses of daidzein (63) and its methyl derivatives isoformononetin (72) and dimethyldaidzein (73) were also reported by Priefer’s group under similar conditions to those employed for genistein (62) (Scheme 22) [57]. However, the main precursor for the C-C bond formation was 3-iodo-7-methoxychromone (71), which was coupled to 4-hydroxyphenylboronic acid and 4-methoxyphenylboronic acid to give isoformononetin (72) and dimethyldaidzein (73), respectively. Demethylation of 72 and 73 with HI in refluxing chloroform gave daidzein (63) [57].
Scheme 22. Preparation of daidzein and its methyl derivatives by Priefer’s group [57].
Scheme 22. Preparation of daidzein and its methyl derivatives by Priefer’s group [57].
Molecules 18 04739 g029

4.4. Synthesis of Prenylated Isoflavonoids

Prenylated isoflavonoids are attained by C- or O-prenylation. Prenylation is often carried out after construction of the isoflavonoid framework [68,69,70]. More complex prenylated isoflavonoids are obtained by cyclization of the prenyl side chains to adjacent hydroxy groups to give furano or pyrano rings, or by modification of phenolic A- and B-rings as well as the C-ring via oxidation or incorporation of other additional substituents [2,3,71,72,73,74].
In 2005, Ito and colleagues reported the first total synthesis of kwakhurin (74) (Figure 6), a 6'-prenyl-phytoestrogen that was isolated from Pueraria mirifica (Leguminosae) [1]. The initial synthetic strategy for 74, which was based on the deoxybenzoin route, failed to give the targeted compound in the last step that involved deprotection of the isopropyl protecting groups. Thus, an alternative route which enabled the use of easily removable protecting groups was sought [1].
Figure 6. Structure of kwakhurin.
Figure 6. Structure of kwakhurin.
Molecules 18 04739 g006
As shown in Scheme 23, the first steps in the alternative route were preparation of the MEM-protected 3-bromochromone 75 and boronic acid 76. These were coupled using Pd(PPh3)4 and TBAB in benzene and the formyl group was converted into a hydroxy group by Baeyer-Villiger oxidation and alkaline hydrolysis to give isoflavone 77 [1]. The prenyl group was introduced on the isoflavone 77 via two pathways. The first one involved O-prenylation and montmorillonite KSF-assisted 1,3-migration of the prenyl group. This gave the expected C-prenylated isoflavone 79 in a poor yield of 4%. The C-prenylated isoflavone 79 was obtained in a good yield upon O-propargylation of 77 followed by reduction of the propargyl ether and thermal rearrangement of the resulting 1,1-dimethylallyl ether 78. The last steps were methylation of the 5'-hydroxy group and deprotection of the 2'- and 4'-hydroxy groups under acid conditions to give kwakhurin (74) [1].
Scheme 23. Total synthesis of kwakhurin (74).
Scheme 23. Total synthesis of kwakhurin (74).
Molecules 18 04739 g030
In 2005, Felpin and co-workers developed a method for the Pd(C)-catalyzed Suzuki-Miyaura cross-coupling of iodocycloenones with arylboronic acids [11]. The use of the heterogeneous Pd(C) presents many advantages compared to other palladium-based catalysts. The reaction is conducted in the absence of additives and ligands, which are often expensive, and air and moisture sensitive. Furthermore, the catalyst can be easily removed from the reaction mixture by filtration, can be reused and is also compatible with water based solvents [11]. Felpin et al. further demonstrated the versatility of the Pd(C) chemistry in C-C bond forming reactions that led to the synthesis of geranylated isoflavones conrauinone D (80), 7-O-geranylformononentin (81) and griffonianone D (82) [12] (Figure 7).
Figure 7. Structures of geranylated isoflavones.
Figure 7. Structures of geranylated isoflavones.
Molecules 18 04739 g007
As illustrated in Scheme 24 [12], Pd(C)-catalyzed reaction of THP protected 3-iodochromone 54 with an 4-hydroxyphenylboronic acid or 4-methoxyphenylboronic acid gave isoflavone derivatives which could be used as precursors for the synthesis of the targeted compounds.
Scheme 24. Synthesis of conrauinone D (80), 7-O-geranylformononentin (81) and griffonianone D (82) by Felpin et al. [12].
Scheme 24. Synthesis of conrauinone D (80), 7-O-geranylformononentin (81) and griffonianone D (82) by Felpin et al. [12].
Molecules 18 04739 g031
Thus, conrauinone D (80) was prepared by O-alkylation of 83 with geranyl bromide using K2CO3 as base, followed by cleavage of the THP protecting group under mild conditions. O-Geranylation of the isoflavone 85, obtained by the THP-deprotection of 84 gave 7-O-geranylformononentin (81). The last target compound griffonianone D (82) was prepared by Mitsunobu reaction of dihydroxylated geraniol derivative 86 and subsequent cleavage of the acetonide [12].
The first total syntheses two biologically-active pyranoisoflavones, the anti-impotence pyranoisoflavone kraussianone 1 (92) and the anti-fungal pyranoisoflavone eriosemaone D (90) was reported by Selepe et al. in 2010 [2]. The key steps involved the Suzuki-Miyaura reaction for the construction of the isoflavone core and the regioselective formation of the dimethylpyran scaffolds to the phloroglucinol (A-ring) and resorcinol (B-ring) moieties (Scheme 25). The synthesis commenced by preparing the boronic acid 88 from aryl iodide 87 by an “in situ quench” procedure described for boronic acid 70. Coupling of boronic acid 88 with 3-iodochromone 67 under Felpin’s conditions and subsequent removal of the MOM protecting groups gave a pyranoisoflavone 89, which was transformed into eriosemaone D (90) by debenzylation with BCl3. Kraussianone 1 (92) was prepared by base-catalyzed aldol-type condensation of 91 with prenal followed by removal of the benzyl protecting group with BCl3 (Scheme 25) [2].
Scheme 25. Total synthesis of kraussianone 1 (92) and eriosemaone D (90) [2].
Scheme 25. Total synthesis of kraussianone 1 (92) and eriosemaone D (90) [2].
Molecules 18 04739 g032
The application of Suzuki-Miyaura reaction has also been demonstrated in the synthesis of other subclasses of isoflavonoids that include coumaronochromones. Zheng and Shen reported the total synthesis of hirtellanine A (98) [3], a coumaronochromone derivative that exhibits immunosuppresive activity (Scheme 26) [75]. The main precursors were 3-iodochromone bearing dimethylchromene scaffold 95 and a boronic ester 96. The 3-iodochromone 95 was prepared in a sequence of steps, which involved regioselective formation the chromene scaffold by a Thom-Harfenist [76] rearrangement of the propargyl ether [77], prepared by O-alkylation of the 7-hydroxy group of chromone 93 with 3-chloro-3-methylbut-1-yne. Iodination of the resulting pyranochromone 94 at C-3 following Gammill’s protocol rendered 95. The boronic ester 96 on the other hand was prepared in three steps from 1,2,4-trihydroxybenzene. The Suzuki coupling of 95 with 96 and subsequent oxidative cleavage of the p-methoxybenzyl protecting groups gave a quinone 97, which upon treatment with acetic acid gave hirtellanine A (98) [3].
Scheme 26. Total synthesis of hirtellanine A (98) [3].
Scheme 26. Total synthesis of hirtellanine A (98) [3].
Molecules 18 04739 g033
Another pyranocoumaronochromone lupinalbin H (102) was synthesized by Selepe and co-workers in steps that involved preparation of 2'-hydroxygenistein (100) from 3-iodochromone 67 and a boronic acid 99, followed by cyclodehydrogenation to lupinalbin A (101) [72]. The final step was the regioselective introduction of the dimethylpyran moiety to the A-ring of 101 via an aldol-type condensation with 3-methyl-2-butenal and 6π-electrocyclization (Scheme 27) [72].
Scheme 27. Total synthesis of lupinalbin H (102) [72].
Scheme 27. Total synthesis of lupinalbin H (102) [72].
Molecules 18 04739 g034

5. Neoflavones

5.1. Synthesis of Neoflavones

The neoflavones (4-arylcoumarins) constitute the largest subclass of neoflavonoids. Traditional procedures for the preparation of neoflavones include the Pechmann or Perkin reactions and the Kostanecki acylation of 2-hydroxybenzophenones followed by base-catalyzed ring closure [77]. Other methods that have been developed include the Wittig reaction of 2-benzophenones [77] and metal-catalyzed cross-coupling reactions such as the Negishi-type, Stille-type and Suzuki-type reactions [8,9,77,78,79,80]. More recently, 4-arylcoumarins have been synthesized by direct arylation by the palladium-catalyzed oxidative Heck coupling of arylboronic acids to coumarins [81,82].
The synthesis of neoflavones by the Suzuki reaction was first reported by Donnelly and co-workers in 1996 [83]. Their approach involved coupling of 4-trifluoromethanesulfonyloxycoumarins with arylboronic acids in the presence of Pd(PPh3)4 and copper(1) iodide as a co-catalyst. The same research group later extended the procedure to the synthesis of hydroxylated neoflavones from benzyloxyboronic acids 104 and 4-trifluoromethanesulfonyloxycoumarins 103 as shown in Scheme 28 [84]. The benzyl protecting group was removed in the late stage by hydrogenation of 105 in the presence of Pd(C) in THF and AcOH to give 106.
Subsequently, similar conditions have been applied in several instances in the syntheses of a series of neoflavones for the investigation of their biological activities [7,8,10]. For instance, neoflavones with a substitution pattern similar to the combretastatins have been prepared and evaluated for pharmacological activities that include, amongst others, cytotoxicity against CEM leukemia and HBL 100 epithelium cell lines [8,10] and antiprotozoal activity against Plasmodium falciparum and Leishmania donovani [7].
Scheme 28. Synthesis of hydroxylated neoflavones by Donnelly and co-workers [84].
Scheme 28. Synthesis of hydroxylated neoflavones by Donnelly and co-workers [84].
Molecules 18 04739 g035

5.2. Synthesis of Arylated Neoflavones

Arylated derivatives of neoflavones have been conveniently prepared by the palladium-catalyzed chemoselective cross-coupling of coumarins 107 activated at the different positions [79,80]. Zhang and colleagues prepared 3-arylneoflavones by a stepwise coupling of 3-bromo-4-trifloxycoumarin 108 or 3-bromo-4-tosyloxycoumarin 111 with differently substituted boronic acids (Scheme 29) [79].
Several palladium catalysts were screened under different conditions to determine the optimal conditions for the preparation of the bisarylated coumarins 110 and 113. The order of reactivity of the active sites was found to be 4-OTf > 3-Br > 4-OTs [79]. Thus, 108 was monoarylated at C-4 under optimized conditions [Pd(PhCN)2Cl2 (5 mol%), 1 M NaHCO3, MeOH, rt, 15–30 min] to give 3-bromoneoflavones 109 in good yields. Subsequent coupling of 3-bromoneoflavones 109 with arylboronic acids using Pd(OAc)2 in the presence of PCy3 and K2HPO4.3H2O in MeOH gave 3-arylneoflavones 110. In contrast, the 3-bromo-4-tosyloxycoumarin 111 reacted in the reversed order to give 112, whereby the 3-bromo substituent was found to be more reactive than the tosyloxy group. An attempt to prepare the heteroarylcoumarins from the intermediates 108 and 111 in a one-pot sequential procedure gave the diarylcoumarins in low yields [79].
Scheme 29. Synthesis of 3-arylcoumarins [79].
Scheme 29. Synthesis of 3-arylcoumarins [79].
Molecules 18 04739 g036

R1R2Product
4-MeOC6H44-MeOC6H4110a
4-MeOC6H44-CF3OC6H4110b
4-MeOC6H44-MeC6H4110c
4-MeOC6H4C6H5110d
4-MeOC6H44-NO2C6H4110e
4-MeOC6H44-MeO2CC6H4110f
C6H5C6H5110g
C6H54-MeOC6H4110h
4-NO2C6H44-MeOC6H4110i
4-MeO2CC6H44-MeOC6H4110j
4-MeOC6H44-MeOC6H4110a
4-MeOC6H44-CF3OC6H4113a
4-MeOC6H44-MeC6H4113b
4-MeOC6H4C6H5110h
4-MeOC6H44-MeO2CC4H6110j
4-MeC6H44-MeC6H4113c
4-MeC6H44-MeOC6H4110c
More recently, Akrawi and co-workers chemoselectively prepared 6-arylated neoflavones 115ae by the reaction of 6-bromo-4-trifluoromethylsulfonyloxycoumarin (114) with arylboronic acids in the presence of Pd(PPh3)4 and K3PO4 as a base using a mixture of toluene/dioxane as solvent (Scheme 30) [80]. The reaction proceeded preferentially at C-4. This enabled a sequential one-pot reaction of 14 with two differently substituted phenylboronic acids to give heteroarylated compounds 115ae in good yields (Scheme 30).
Scheme 30. One-pot synthesis of 6-arylneoflavones from coumarin 114.
Scheme 30. One-pot synthesis of 6-arylneoflavones from coumarin 114.
Molecules 18 04739 g037

115Ar1Ar2
a4-MeOC6H44-ClC6H4
b4-MeOC6H43-ClC6H4
c3,5-Me2C6H34-MeOC6H4
d4-ClC6H44-MeOC6H4
e4-ClC6H43-FC6H4

6. Conclusions

The Suzuki-Miyaura reaction has been successfully applied in the synthesis of a variety of flavonoids. The main advantage of this reaction is the mild reaction conditions which are compatible with different functional groups and thus allow the synthesis of flavonoids containing a number of sensitive substituents.

Acknowledgments

The University of KwaZulu-Natal and the National Research Foundation, Pretoria are thanked for financial assistance.

References

  1. Ito, F.; Iwasaki, M.; Watanabe, T.; Ishikawa, T.; Higuchi, Y. The first total synthesis of kwakhurin, a characteristic component of a rejuvenating plant, “kwao keur”: toward an efficient synthetic route to phytoestrogenic isoflavones. Org. Biomol. Chem. 2005, 3, 674–681. [Google Scholar] [CrossRef]
  2. Selepe, M.A.; Drewes, S.E.; van Heerden, F.R. Total synthesis of the pyranoisoflavone kraussianone 1 and related isoflavones. J. Nat. Prod. 2010, 73, 1680–1685. [Google Scholar] [CrossRef]
  3. Zheng, S.Y.; Shen, Z.W. Total synthesis of hirtellanine A. Tetrahedron Lett. 2010, 51, 2883–2887. [Google Scholar] [CrossRef]
  4. Wei, G.; Yu, B. Isoflavone glycosides: Synthesis and evaluation as α-glucosidase inhibitors. Eur. J. Org. Chem. 2008, 3156–3163. [Google Scholar]
  5. Matin, A.; Gavande, N.; Kim, M.S.; Yang, N.X.; Salam, N.K.; Hanrahan, J.R.; Roubin, R.H.; Hibbs, D.E. 7-Hydroxy-benzopyran-4-one derivatives: A novel pharmacophore of peroxisome proliferator-activated receptor α and -γ (PPARα and γ) dual agonists. J. Med. Chem. 2009, 52, 6835–6850. [Google Scholar]
  6. Mihigo, S.O.; Mammo, W.; Bezabih, M.; Andrae-Marobela, K.; Abegaz, B.M. Total synthesis, antiprotozoal and cytotoxicity activities of rhuschalcone VI and analogs. Bioorg. Med. Chem. 2010, 18, 2464–2473. [Google Scholar] [CrossRef]
  7. Pierson, J.-T.; Dumetre, A.; Hutter, S.; Delmas, F.; Laget, M.; Finet, J.-P.; Azas, N.; Combes, S. Synthesis and antiprotozoal activity of 4-arylcoumarins. Eur. J. Med. Chem. 2010, 45, 864–869. [Google Scholar] [CrossRef]
  8. Combes, S.; Barbier, P.; Douillard, S.; McLeer-Florin, A.; Bourgarel-Rey, V.; Pierson, J.-T.; Fedorov, A.Y.; Finet, J.-P.; Boutonnat, J.; Peyrot, V. Synthesis and biological evaluation of 4-arylcoumarin analogues of combretastatins. Part 2. J. Med. Chem. 2011, 54, 3153–3162. [Google Scholar] [CrossRef]
  9. Ganina, O.G.; Daras, E.; Bourgarel-Rey, V.; Peyrot, V.; Andresyuk, A.N.; Finet, J.-P.; Fedorov, A.Y.; Beletskaya, I.P.; Combes, S. Synthesis and biological evaluation of polymethoxylated 4-heteroarylcoumarins as tubulin assembly inhibitor. Bioorg. Med. Chem. 2008, 16, 8806–8812. [Google Scholar]
  10. Bailly, C.; Bal, C.; Barbier, P.; Combes, S.; Finet, J.P.; Hildebrand, M.P.; Peyrot, V.; Wattez, N. Synthesis and biological evaluation of 4-arylcoumarin analogues of combretastatins. J. Med. Chem. 2003, 46, 5437–5444. [Google Scholar] [CrossRef]
  11. Felpin, F.-X. Practical and efficient Suzuki-Miyaura cross-coupling of 2-iodocycloenones with arylboronic acids catalyzed by recyclable Pd(0)/C. J. Org. Chem. 2005, 70, 8575–8578. [Google Scholar] [CrossRef]
  12. Felpin, F.-X.; Lory, C.; Sow, H.; Acherar, S. Practical and efficient entry to isoflavones by Pd(0)/C-mediated Suzuki-Miyaura reaction. Total synthesis of geranylated isoflavones. Tetrahedron 2007, 63, 3010–3016. [Google Scholar] [CrossRef]
  13. Claisen, L.; Claparède, A. Condensationen von Ketonen mit Aldehyden. Chem. Ber. 1881, 14, 2460–2468. [Google Scholar] [CrossRef]
  14. Satyanarayana, M.; Tiwari, P.; Tripathi, B.K.; Srivastava, A.K.; Pratap, R. Synthesis and antihyperglycemic activity of chalcone based aryloxypropanolamines. Bioorg. Med. Chem. 2004, 12, 883–889. [Google Scholar] [CrossRef]
  15. Daskiewicz, J.B.; Comte, G.; Barron, D.; Di Pietro, A.; Thomasson, F. Organolithium mediated synthesis of prenylchalcones as potential inhibitors of chemoresistance. Tetrahedron Lett. 1999, 40, 7095–7098. [Google Scholar] [CrossRef]
  16. Wang, Z. Claisen-Schmidt Condensation. Comprehensive Organic Name Reactions and Reagents. Available online: http://dx.doi.org/10.1002/9780470638859.conrr145 (accessed on 15 September 2010).
  17. Daskiewicz, J.-B.; Barron, D. Substituent effects in the BF3-mediated acylation of phenols with cinnamic acids. Nat. Prod. Res. 2003, 17, 347–350. [Google Scholar] [CrossRef]
  18. Bianco, A.; Cavarischia, C.; Farina, A.; Guiso, M.; Marra, C. A new synthesis of flavonoids via Heck reaction. Tetrahedron Lett. 2003, 44, 9107–9109. [Google Scholar] [CrossRef]
  19. Liao, W.-W.; Mueller, T.J.J. Sequential coupling-isomerization-coupling reactions - A novel three-component synthesis of aryl chalcones. Synlett 2006, 3469–3473. [Google Scholar]
  20. Braun, R.U.; Ansorge, M.; Mueller, T.J.J. Coupling-isomerization synthesis of chalcones. Chem. Eur. J. 2006, 12, 9081–9094. [Google Scholar] [CrossRef]
  21. Eddarir, S.; Cotelle, N.; Bakkour, Y.; Rolando, C. An efficient synthesis of chalcones based on the Suzuki reaction. Tetrahedron Lett. 2003, 44, 5359–5363. [Google Scholar] [CrossRef]
  22. Haddach, M.; McCarthy, J.R. A new method for the synthesis of ketones: The palladium-catalyzed cross-coupling of acid chlorides with arylboronic acids. Tetrahedron Lett. 1999, 40, 3109–3112. [Google Scholar] [CrossRef]
  23. Al-Masum, M.; Ng, E.; Wai, M.C. Palladium-catalyzed direct cross-coupling of potassium styryltrifluoroborates and benzoyl chlorides - a one step method for chalcone synthesis. Tetrahedron Lett. 2011, 52, 1008–1010. [Google Scholar] [CrossRef]
  24. Xin, B.-W. Synthesis of aryl ketones by cross-coupling reaction of arylboronic acids with carboxylic anhydrides in aqueous phase. Synth. Commun. 2008, 38, 2826–2837. [Google Scholar] [CrossRef]
  25. Zuo, Y.; Yu, Y.; Wang, S.; Shao, W.; Zhou, B.; Lin, L.; Luo, Z.; Huang, R.; Du, J.; Bu, X. Synthesis and cytotoxicity evaluation of biaryl-based chalcones and their potential in TNF alpha-induced nuclear factor-kappa B activation inhibition. Eur. J. Med. Chem. 2012, 50, 393–404. [Google Scholar] [CrossRef]
  26. Vieira, L.C.C.; Paixão, M.W.; Corrêa, A.G. Green synthesis of novel chalcone and coumarin derivatives via Suzuki coupling reaction. Tetrahedron Lett. 2012, 53, 2715–2718. [Google Scholar]
  27. Baker, W. 322. Molecular rearrangement of some o-acyloxyacetophenones and the mechanism of the production of 3-acylchromones. J. Chem. Soc. 1933, 1381–1389. [Google Scholar] [CrossRef]
  28. Mahal, H.S.; Venkataraman, K. 387. Synthetical experiments in the chromone group. Part XIV. The action of sodamide on 1-acyloxy-2-acetonaphthones. J. Chem. Soc. 1934, 1767–1769. [Google Scholar] [CrossRef]
  29. Sharma, D.; Kumar, S.; Makrandi, J.K. Solid phase Baker–Venkataraman rearrangement under solvent-free condition using grinding technique. Green Chem. Lett. Rev. 2009, 2, 53–55. [Google Scholar] [CrossRef]
  30. Li, J.J. Allan-Robinson reaction. In Name Reactions Name Reactions—A Collection of Detailed Reaction Mechanisms; Springer: Berlin/Heidelberg, Germany, 2006; pp. 8–9. [Google Scholar]
  31. Bennett, M.; Burke, A.J.; O’Sullivan, W.I. Aspects of the Algar-Flynn-Oyamada (AFO) reaction. Tetrahedron 1996, 52, 7163–7178. [Google Scholar] [CrossRef]
  32. Bose, G.; Mondal, E.; Khan, A.T.; Bordoloi, M.J. An environmentally benign synthesis of aurones and flavones from 2'-acetoxychalcones using n-tetrabutylammonium tribromide. Tetrahedron Lett. 2001, 42, 8907–8909. [Google Scholar] [CrossRef]
  33. Kraus, G.A.; Gupta, V. Divergent approach to flavones and aurones via dihaloacrylic acids. Unexpected dependence on the halogen atom. Org. Lett. 2010, 12, 5278–5280. [Google Scholar] [CrossRef]
  34. Dahlen, K.; Wallen, E.A.A.; Grotli, M.; Luthman, K. Synthesis of 2,3,6,8-tetrasubstituted chromone scaffolds. J. Org. Chem. 2006, 71, 6863–6871. [Google Scholar] [CrossRef]
  35. Arai, M.A.; Sato, M.; Sawada, K.; Hosoya, T.; Ishibashi, M. Efficient synthesis of chromone and flavonoid derivatives with diverse heterocyclic units. Chem.-Asian J. 2008, 3, 2056–2064. [Google Scholar] [CrossRef]
  36. Larsen, L.; Yoon, D.H.; Weavers, R.T. Synthesis of a range of polyhydroxy 8-aryl flavones. Synth. Commun. 2009, 39, 2935–2948. [Google Scholar] [CrossRef]
  37. Deodhar, M.; Black, D.S.; Chan, D.S.-H.; Kumar, N. Synthesis of some new biheterocycles by a one-pot Suzuki-Miyaura coupling reaction. Heterocycles 2010, 80, 1267–1274. [Google Scholar] [CrossRef]
  38. Malik, I.; Hussain, M.; Hung, N.T.; Villinger, A.; Langer, P. Synthesis of 7,8-diarylflavones by site-selective Suzuki-Miyaura reactions. Synlett 2010, 2244–2246. [Google Scholar]
  39. Che, H.; Lim, H.; Kim, H.P.; Park, H. A chrysin analog exhibited strong inhibitory activities against both PGE(2) and NO production. Eur. J. Med. Chem. 2011, 46, 4657–4660. [Google Scholar] [CrossRef]
  40. Joshi, V.; Hatim, J.G. Facile synthesis of new substituted 3-aryl/heteroarylflavones by Suzuki-Miyaura coupling of 3-bromoflavone with substituted aryl/heteroarylboronic acids. Indian J. Heterocycl. Chem. 2011, 21, 111–116. [Google Scholar]
  41. Eleya, N.; Malik, I.; Reimann, S.; Wittler, K.; Hein, M.; Patonay, T.; Villinger, A.; Ludwig, R.; Langer, P. Efficient synthesis of arylated flavones by site-selective Suzuki–Miyaura cross-coupling ceactions of the cis(triflate) of 5,7- and 7,8-dihydroxyflavone. Eur. J. Org. Chem. 2012, 2012, 1639–1652. [Google Scholar]
  42. Nakazawa, K.; Ito, M. Synthesis of ginkgetin. Tetrahedron Lett. 1962, 317–319. [Google Scholar] [CrossRef]
  43. Chen, J.J.; Chang, H.W.; Kim, H.P.; Park, H. Synthesis of phospholipase A2 inhibitory biflavonoids. Bioorg. Med. Chem. Lett. 2006, 16, 2373–2375. [Google Scholar] [CrossRef]
  44. Muller, D.; Fleury, J.P. A new strategy for the synthesis of biflavonoids via arylboronic acids. Tetrahedron Lett. 1991, 32, 2229–2232. [Google Scholar] [CrossRef]
  45. Zembower, D.E.; Zhang, H.P. Total synthesis of robustaflavone, a potential anti-hepatitis B agent. J. Org. Chem. 1998, 63, 9300–9305. [Google Scholar] [CrossRef]
  46. Zheng, X.; Meng, W.D.; Qing, F.L. Synthesis of gem-difluoromethylenated biflavonoid via the Suzuki coupling reaction. Tetrahedron Lett. 2004, 45, 8083–8085. [Google Scholar] [CrossRef]
  47. Ollis, W.D. The Isoflavonoids. In The Chemistry of Flavonoids; Geissman, T.A., Ed.; Pergamon Press Inc.: Oxford, UK, 1962; pp. 353–440. [Google Scholar]
  48. Baker, W.; Chadderton, J.; Harborne, B.J.; Ollis, D.W. A new synthesis of isoflavones. J. Chem. Soc. 1953, 1852–1854. [Google Scholar]
  49. Dewick, P.M. Isoflavonoids. In The Flavonoids: Advances in Research since 1986; Harborne, J.B., Ed.; Chapman and Hall: London, UK, 1994; pp. 117–238. [Google Scholar]
  50. Dixon, R.A.; Ferreira, D. Genistein. Phytochemistry 2002, 60, 205–211. [Google Scholar] [CrossRef]
  51. Granados-Covarrubias, E.H.; Maldonado, L.A. A Wacker-Cook synthesis of isoflavones: formononetine. Tetrahedron Lett. 2009, 50, 1542–1545. [Google Scholar] [CrossRef]
  52. Li, Q.-L.; Liu, Q.-L.; Ge, Z.-Y.; Zhu, Y.-M. A novel synthesis of isoflavones via copper(I)-catalyzed intramolecular cyclization reaction. Helv. Chim. Acta 2011, 94, 1304–1309. [Google Scholar] [CrossRef]
  53. Hoshino, Y.; Miyaura, N.; Suzuki, A. Novel synthesis of isoflavones by the palladium-catalyzed cross-coupling reaction of 3-bromochromones with arylboronic acids or their esters. Bull. Chem. Soc. Jpn. 1988, 61, 3008–3010. [Google Scholar] [CrossRef]
  54. Dawood, K.M. Microwave-assisted Suzuki-Miyaura and Heck-Mizoroki cross-coupling reactions of aryl chlorides and bromides in water using stable benzothiazole-based palladium(II) precatalysts. Tetrahedron 2007, 63, 9642–9651. [Google Scholar] [CrossRef]
  55. Eisnor, C.R.; Gossage, R.A.; Yadav, P.N. Oxazoline Chemistry. Part 11: Syntheses of natural and synthetic isoflavones, stilbenes and related species via C-C bond formation promoted by a Pd-oxazoline complex. Tetrahedron 2006, 62, 3395–3401. [Google Scholar] [CrossRef]
  56. Ding, K.; Wang, S.M. Efficient synthesis of isoflavone analogues via a Suzuki coupling reaction. Tetrahedron Lett. 2005, 46, 3707–3709. [Google Scholar] [CrossRef]
  57. Biegasiewicz, K.F.; St Denis, J.D.; Carroll, V.M.; Priefer, R. An efficient synthesis of daidzein, dimethyldaidzein, and isoformononetin. Tetrahedron Lett. 2010, 51, 4408–4410. [Google Scholar] [CrossRef]
  58. Gammill, B.R. A new and efficient synthesis of 3-halogenated 4H-1-benzopyran-4-ones. Synthesis 1979, 901–903. [Google Scholar] [CrossRef]
  59. Tsoi, Y.-T.; Zhou, Z.; Chan, A.S.C.; Yu, W.-Y. Palladium-catalyzed oxidative cross-coupling reaction of arylboronic acids with diazoesters for stereoselective synthesis of (E)-α,β-diarylacrylates. Org. Lett. 2010, 12, 4506–4509. [Google Scholar] [CrossRef]
  60. Ikedo, A.; Hayakawa, I.; Usui, T.; Kazami, S.; Osada, H.; Kigoshi, H. Structure-activity relationship study of glaziovianin A against cell cycle progression and spindle formation of HeLa S3 cells. Bioorg. Med. Chem. Lett. 2010, 20, 5402–5404. [Google Scholar] [CrossRef] [Green Version]
  61. Vasselin, D.A.; Westwell, A.D.; Matthews, C.S.; Bradshaw, T.D.; Stevens, M.F.G. Structural studies on bioactive compounds. 40. Synthesis and biological properties of fluoro-, methoxyl-, and amino-substituted 3-phenyl-4H-1-benzopyran-4-ones and a comparison of their antitumor activities with the activities of related 2-phenylbenzothiazoles. J. Med. Chem. 2006, 49, 3973–3981. [Google Scholar] [CrossRef]
  62. Hayakawa, I.; Ikedo, A.; Kigoshi, H. Synthesis of glaziovianin A: A potent antitumor isoflavone. Chem. Lett. 2007, 36, 1382–1383. [Google Scholar] [CrossRef]
  63. Cornwell, T.; Cohick, W.; Raskin, I. Dietary phytoestrogens and health. Phytochemistry 2004, 65, 995–1016. [Google Scholar] [CrossRef]
  64. St Denis, J.D.; Gordon, J.S.; Carroll, V.M.; Priefer, R. Novel synthesis of the isoflavone genistein. Synthesis 2010, 1590–1592. [Google Scholar]
  65. Andersen, Ø.M.; Markham, K.R. Flavonoids; Taylor and Francis: New York, NY, USA, 2006. [Google Scholar]
  66. Rice-Evans, C.A.; Packer, L. Flavonoids in Health and Disease, 2nd ed.; Marcel Dekker: New York, NY, USA, 2003. [Google Scholar]
  67. Lin, J.-K.; Weng, M.-S. Flavonoids as Nutraceuticals. In The Science of Flavonoids; Grotewold, E., Ed.; Springer Science and Business media: New York, NY, USA, 2006. [Google Scholar]
  68. Rao, K.S.R.M.; Iyer, C.S.R.; Iyer, P.R. Synthesis of alpinum isoflavone, derrone and related pyranoisoflavones. Tetrahedron 1987, 43, 3015–3019. [Google Scholar] [CrossRef]
  69. Jain, A.C.; Jain, S.M. Synthesis of derrubone and robustone. Tetrahedron 1972, 28, 5063–5067. [Google Scholar] [CrossRef]
  70. Hastings, J.M.; Hadden, M.K.; Blagg, B.S.J. Synthesis and evaluation of derrubone and select analogues. J. Org. Chem. 2007, 73, 369–373. [Google Scholar] [CrossRef]
  71. Botta, B.; Menendez, P.; Zappia, G.; de Lima, R.A.; Torge, R.; Delle Monache, G. Prenylated isoflavonoids: Botanical distribution, Structures, Biological activities and biotechnological studies. An update (1995–2006). Curr. Med. Chem. 2009, 16, 3414–3468. [Google Scholar] [CrossRef]
  72. Selepe, M.A.; Drewes, S.E.; van Heerden, F.R. Total synthesis of the pyranocoumaronochromone lupinalbin H. Tetrahedron 2011, 67, 8654–8658. [Google Scholar] [CrossRef]
  73. Veitch, N.C. Isoflavonoids of the Leguminosae. Nat. Prod. Rep. 2007, 24, 417–464. [Google Scholar] [CrossRef]
  74. Veitch, N.C. Isoflavonoids of the Leguminosae. Nat. Prod. Rep. 2009, 26, 776–802. [Google Scholar] [CrossRef]
  75. Shou, Q.Y.; Tan, Q.; Shen, Z.W. Hirtellanines A and B, a pair of isomeric isoflavonoid derivatives from Campylotropis hirtella and their immunosuppressive activities. Bioorg. Med. Chem. Lett. 2009, 19, 3389–3391. [Google Scholar] [CrossRef]
  76. Harfenist, M.; Thom, E. Influence of structure on the rate of thermal rearrangement of aryl propargyl ethers to the chromenes. Gem-dimethyl effect. J. Org. Chem. 1972, 37, 841–848. [Google Scholar] [CrossRef]
  77. Garazd, M.M.; Garazd, Y.L.; Khilya, V.P. Neoflavones. 2. Methods for synthesizing and modifying 4-arylcoumarins. Chem. Nat. Comp. 2005, 41, 245–271. [Google Scholar] [CrossRef]
  78. Phapale, V.B.; Cardenas, D.J. Nickel-catalysed Negishi cross-coupling reactions: scope and mechanisms. Chem. Soc. Rev. 2009, 38, 1598–1607. [Google Scholar] [CrossRef]
  79. Zhang, L.; Meng, T.; Fan, R.; Wu, H. General and efficient route for the synthesis of 3,4-disubstituted coumarins via Pd-catalyzed site-selective cross-coupling reactions. J. Org. Chem. 2007, 72, 7279–7286. [Google Scholar] [CrossRef]
  80. Akrawi, O.A.; Nagy, G.Z.; Patonay, T.; Villinger, A.; Langer, P. Chemoselective Suzuki-Miyaura reactions of 4-trifluoromethylsulfonyloxy-6-bromocoumarin. Tetrahedron Lett. 2012, 53, 3206–3209. [Google Scholar] [CrossRef]
  81. Khoobi, M.; Alipour, M.; Zarei, S.; Jafarpour, F.; Shafiee, A. A facile route to flavone and neoflavone backbones via a regioselective palladium catalyzed oxidative Heck reaction. J. Chem. Soc., Chem. Commun. 2012, 48, 2985–2987. [Google Scholar] [CrossRef]
  82. Li, Y.; Qi, Z.; Wang, H.; Fu, X.; Duan, C. Palladium-catalyzed oxidative Heck coupling reaction for direct synthesis of 4-arylcoumarins using coumarins and arylboronic acids. J. Org. Chem. 2012, 77, 2053–2057. [Google Scholar] [CrossRef]
  83. Boland, G.M.; Donnelly, D.M.X.; Finet, J.P.; Rea, M.D. Synthesis of neoflavones by Suzuki arylation of 4-substituted coumarins. J. Chem. Soc. Perkin Trans. 1996, 1, 2591–2597. [Google Scholar]
  84. Donnelly, D.M.X.; Finet, J.P.; Guiry, P.J.; Rea, M.D. Synthesis of C-ring hydroxylated neoflavonoids by ligand coupling reactions. Synth. Commun. 1999, 29, 2719–2730. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Selepe, M.A.; Van Heerden, F.R. Application of the Suzuki-Miyaura Reaction in the Synthesis of Flavonoids. Molecules 2013, 18, 4739-4765. https://doi.org/10.3390/molecules18044739

AMA Style

Selepe MA, Van Heerden FR. Application of the Suzuki-Miyaura Reaction in the Synthesis of Flavonoids. Molecules. 2013; 18(4):4739-4765. https://doi.org/10.3390/molecules18044739

Chicago/Turabian Style

Selepe, Mamoalosi A., and Fanie R. Van Heerden. 2013. "Application of the Suzuki-Miyaura Reaction in the Synthesis of Flavonoids" Molecules 18, no. 4: 4739-4765. https://doi.org/10.3390/molecules18044739

Article Metrics

Back to TopTop