Next Article in Journal
Novel Polyprenylated Phloroglucinols from Hypericum sampsonii
Next Article in Special Issue
Multitarget Molecular Hybrids of Cinnamic Acids
Previous Article in Journal
Post-Polymerization Modification of Poly(L-glutamic acid) with D-(+)-Glucosamine
Previous Article in Special Issue
Natural Cinnamic Acids, Synthetic Derivatives and Hybrids with Antimicrobial Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bioinspired Syntheses of Dimeric Hydroxycinnamic Acids (Lignans) and Hybrids, Using Phenol Oxidative Coupling as Key Reaction, and Medicinal Significance Thereof †

Department of Chemistry, University of Patras, Patras 26504, Greece
*
Author to whom correspondence should be addressed.
Dedicated in memory of the late Prof. Derek H. R. Barton.
Submission received: 3 November 2014 / Revised: 20 November 2014 / Accepted: 21 November 2014 / Published: 28 November 2014
(This article belongs to the Special Issue Cinnamic Acids Hybrids with Biological Interest)

Abstract

:
Lignans are mainly dimers of 4-hydroxycinnamic acids (HCAs) and reduced analogs thereof which are produced in Nature through phenol oxidative coupling (POC) as the primary C-C or C-O bond-forming reaction under the action of the enzymes peroxidases and laccases. They present a large structural variety and particularly interesting biological activities, therefore, significant efforts has been devoted to the development of efficient methodologies for the synthesis of lignans isolated from natural sources, analogs and hybrids with other biologically interesting small molecules. We summarize in the present review those methods which mimic Nature for the assembly of the most common lignan skeleta by using either enzymes or one-electron inorganic oxidants to effect POC of HCAs and derivatives, such as esters and amides, or cross-POC of pairs of HCAs or HCAs with 4-hydrocycinnamyl alcohols. We, furthermore, provide outlines of mechanistic schemes accounting for the formation of the coupled products and, where applicable, indicate their potential application in medicine.

Graphical Abstract

1. Introduction

CinA and its 4-hydroxysubstituted derivatives (HCAs), namely CouA, CafA, FerA and SinA, form an important family of natural products ubiquitous in the plant kingdom generally known as cinnamates [1]. Cinnamates are derived biogenetically from ShiA through the intermediacy of the amino acid Phe (Scheme 1).
Scheme 1. Outline of the biosynthetic route to naturally occurring cinnamates and selected reduced derivatives.
Scheme 1. Outline of the biosynthetic route to naturally occurring cinnamates and selected reduced derivatives.
Molecules 19 19769 g020
HCAs and derivatives are powerful natural antioxidants, abundantly found in a variety of foods and drinks, which could be therefore used for the prevention and/or therapy of oxidative stress-associated diseases, like atherosclerosis, inflammatory injury and cancer [2,3]. The side chain of cinnamates is biogenetically further reduced thus leading to a variety of reduced derivatives such as the corresponding aldehydes (e.g., p-coumaraldehyde), alcohols (e.g., ConAl) and alkenes (e.g., isoeugenol). All these compounds are coined as phenylpropanoids and form the C6-C3 class of ShiA metabolites [1].
Interestingly, these phenypropanoid compounds are further dimerized or polymerized in Nature and thus provide access to two further families of natural products also widely distributed in plants, collectively coined as lignans and lignin, respectively. Lignans are mainly dimeric in Nature and are produced biosynthetically by the dimerization of HCAs and their reduced derivatives HCAls. The role of lignans in plants seems to be primarily plant defense [4]. Lignin is a high molecular weight complex polymer, derived from the polymerization of phenylpropanoid alcohols (called monolignols), such as ConAl. Lignin constitutes an integral part of the secondary cell walls of plants conferring mechanical strength to the cell wall and, by extension, to plant as a whole.
Dimerization of phenolic compounds in Nature is effected using POC as key reaction, which allows the connection of the monomers through the formation of new C-C or C-O bonds. Although this concept had been long recognized, it was the seminal investigation by Barton and Cohen on the structure of Pummerer’s ketone, produced by the ferricyanide-mediated POC of p-cresol in alkaline solution, which established the relation of such reactions in the laboratory with biosynthetic pathways [5]. This concept was applied in the elegant biomimetic two-steps synthesis of (±)-usnic acid by Barton and coworkers using the POC of methylphloracetophenone as key step for the assembly of the skeleton (Scheme 2) [6].
Scheme 2. Barton’s biomimetic synthesis of usnic acid based on POC as key step. The primary bond connecting the two monomers is shown in blue and bolded and the secondary one in red and bolded.
Scheme 2. Barton’s biomimetic synthesis of usnic acid based on POC as key step. The primary bond connecting the two monomers is shown in blue and bolded and the secondary one in red and bolded.
Molecules 19 19769 g021
According to the mechanism put forward, one-electron oxidation of the corresponding phenolate leads to a radical species which, due to the resonance effect exhibits additional mesomeric forms, with the unpaired electron residing on the ortho and para carbons. Radical pairing creates a new C-C bond connecting the two monomers and the thus obtained intermediate, following aromatization, undergoes a base-induced intramolecular Michael addition thus producing the tricyclic usnic acid hydrate in one pot. Dehydration may then be effected directly by treating with concentrated sulfuric acid or indirectly by treating with acetic anhydride in the presence of a catalytic quantity of sulfuric acid followed by sulfuric acid-mediated hydrolysis of the thus obtained usnic acid diacetate.
Although radical pairing seems to be the prevailing mechanism in POC, other mechanisms might be applicable in certain cases, e.g., two electron oxidation of one phenolate to create an onium ion followed by an electrophilic aromatic substitution on the other electron-rich phenolate, also followed by a prototropic shift leading to aromatization, in case one or both of the coupled carbon atoms bear a hydrogen atom. If this were not the case, external or better internal nucleophiles attack the intermediate cross-conjugated spirodienone or a dienone-phenol type rearrangement takes place creating stable products, which can further lose small molecule, e.g. water, to give the final product [1]. POC is also the key reaction through which the plethora of the naturally occurring lignans is biosynthetically produced from monomeric phenylpropanoid compounds [7,8,9,10,11,12,13], under the action of the enzymes peroxidases or laccases. The two families of enzymes mainly differ in the metallic cation present in their active site, Fe3+ in the former and Cu2+ in the latter. Due to their side chain, one electron oxidation of phenolate anions of phenylpropanoids creates through resonance an additional site (C-8 or C-β) for radical coupling (see canonical form V in Scheme 3).
Scheme 3. One electron oxidation of 4-hydroxyphenylpropanoid compounds creating the mesomeric forms IV of the phenoxide redical. Numbering of the aromatic ring and the alternative numbering of the side-chain is the one suggested by IUPAC.
Scheme 3. One electron oxidation of 4-hydroxyphenylpropanoid compounds creating the mesomeric forms IV of the phenoxide redical. Numbering of the aromatic ring and the alternative numbering of the side-chain is the one suggested by IUPAC.
Molecules 19 19769 g022
Pairing of two such radicals creates the new C-C or C-O bond connecting the two monomers. There are several combinations of such radicals leading to different regioisomeric dimers (Scheme 4).
According to the 2000 IUPAC recommendations, lignans are the dimers in which the new primary C-C bond is formed between the C-8 (or C-β) of one and the C-8' (or C-β) of the other monomer [14]. All other lignans in which the monomers are connected with a bond other than the 8-8' (or β-β) bond are collectively coined neolignans (NLs). NLs in which the two monomers are connected through a primary C-O bond are particularly called oxyneolignans. The ratio of the various possible regioisomers actually formed in a particular POC depends on the stereoelectronic effects operating in the phenoxide radical, the oxidant or oxidant system and the reaction conditions. The most abundant structures in lignans are constructed by primary 8-8', 8-5' and 8-Ο-4' bonds. Coupling at position 5 is only possible when this position is unoccupied. On the other hand, coupling between O atoms or between C atoms both in position 1 (1-1' coupling) have not been observed in lignans because in the former case it would create a highly unstable peroxy dimer, whereas in the latter case it could not be effected due to steric hindrance because both monomers bear a propanoid side chain in position 1 [15].
Scheme 4. Possible combinations of the phenoxide radicals from 4-hydroxypropanoids. The primary bond formed in POC is shown in blue and bold.
Scheme 4. Possible combinations of the phenoxide radicals from 4-hydroxypropanoids. The primary bond formed in POC is shown in blue and bold.
Molecules 19 19769 g023
Many lignans are produced in Nature in optically pure form, while others are produced as mixtures of enantiomers. Even different plants may produce the same lignan but with different enantiomeric composition [16]. The multitude of the pharmacological properties of lignans has been recently summarized [17], and include antioxidant, anti-inflammatory, antitumor and anti-HIV activities to name a few. Taking into consideration the many potential applications of lignans in medicine, a variety of synthetic methods have been devised which allow ease access to natural lignans and derivatives suitable for structure-activity relationship studies [18,19,20].
This review presents an overview of the application of POC on HCAs and derivatives, such esters and amides, as a key reaction in the efficient assembly of the skeleton of the most abundant lignan structures in nature and of analogs of potential medicinal interest. Particular attention is paid to POCs in which the anticipated regioisomer is being produced in significant excess to unwanted other regioisomers and oligomers. Interestingly, although a large array of oxidants, inorganic and enzymes, have been already developed for performing POC [21], a few of them have been proved particularly useful in producing regioselectively and/or stereoselectively the desired lignan isomer or an appropriate key intermediate.

2. Lignan Skeleton Assembly with β-β Bond Formation as the Key Step

2.1. Syntheses of Lignans Based on the Key Intermediate 4-cis,8-cis-Diaryl-3,7-dioxabicyclo[3.3.0]-octane-2,6-dione

Dilactones of the general formula 1, which are produced by POC of HCAs and halogenated derivatives, are key intermediates in the synthesis of several subtypes of classical lignans (CLs), such as the dibenzylbutyrolactone (CL2), arylnaphthalene (CL3), 2,5-diaryltetrahydrofuran (CL5a) and 2,6-diaryl- furofuran (CL6) subtypes. For the classification of CLs and NLs in subtypes see [20]. Dilactones 1 are formed through intramolecular nucleophilic attack of the bis-p-quinonemethide intermediates 2, primarily produced through POC, by the carboxyl functions (Scheme 5).
Scheme 5. Dilactone 1 as key intermediate in the synthesis of CLs.
Scheme 5. Dilactone 1 as key intermediate in the synthesis of CLs.
Molecules 19 19769 g024
Thus, POC of FerA mediated by FeCl3 and O2, produced dilactone 1a in 30% yield (Scheme 6) [22,23]. Dilactone 1a, on treatment with methanolic HCl, rearranges to aryldihydronaphthalene 3, which in turn can be converted through several steps into a mixture of arylnaphthalene lactones 4a and 4b [24]. Alternatively, catalytic hydrogenation of 1a, followed by dehydration and finally partial reduction, leads to matairesinol (5) [24]. On the other hand, LAH-mediated reduction of the diacetylated or dimethylated dilactones 1b or 1c, followed by acid-mediated cyclization of the thus obtained tetraols, provided pinoresinol (6) and eudesmin (7), respectively. Alternatively, lignans 6 and 7 may be obtained from dilactones 1b and 1c by partial reduction with DIBAL to the corresponding dilactols, followed by tosylation and LAH-mediated reduction [25]. The conversion of the dilactone 1a to the arylnaphthalene derivative 3 can be envisaged to take place through transesterification, followed by double dehydration and finally electrophilic aromatic substitution on the intermediate p-quinone-methide (Scheme 7).
Scheme 6. Synthesis of lignans 37 through POC of FerA and using dilactone 1a as key intermediate.
Scheme 6. Synthesis of lignans 37 through POC of FerA and using dilactone 1a as key intermediate.
Molecules 19 19769 g025
Scheme 7. A plausible mechanism for the conversion of dilactone 1a to arylnaphtalene 3 on treatment with HCl/MeOH.
Scheme 7. A plausible mechanism for the conversion of dilactone 1a to arylnaphtalene 3 on treatment with HCl/MeOH.
Molecules 19 19769 g026
Scheme 8. Synthesis of lignans 8–12 through POC of SinA and using dilactone 1d as key intermediate.
Scheme 8. Synthesis of lignans 8–12 through POC of SinA and using dilactone 1d as key intermediate.
Molecules 19 19769 g027
On the other hand POC of SinA mediated by FeCl3 and O2, produced dilactone 1d in 73% yield (Scheme 8) [26]. Dilactone 1d, on treatment with aqueous acid, rearranges to ThoA (8), whereas with methanolic HCl, 1d is converted to its corresponding diester 9a [27,28]. The dimethyl ether 9b of the latter can be hydrogenated to produce a mixture of the diastereomeric aryltetralins 10a and 10b, of which the former was reduced to (±)-lyoniresinol dimethyl ether (11) [29,30]. On the other hand 9b or the corresponding dibenzyl ether 9c can be transformed to either thomasic acid dimethyl ether (12b) or thomasic acid (12a), respectively, through a multi-step procedure involving LAH-mediated reduction, selective oxidation of the allylic alcohol function with MnO2, followed by further selective oxidation of the thus obtained aldehyde to the corresponding methyl ester with MnO2/HCN/MeOH, and finally either saponification or hydrogenolysis followed by saponification [27,28].
Dilactones substituted by halogens in the aromatic ring have been synthesized by either applying POC to appropriately halogenated CinAs [31,32] or by direct halogenation of acetylated dilactones [33]. Such halogen substituted dilactones like 1e and 1f have been used for the synthesis of lignans of the CL5a subtype, e.g., of galbelgin (13a) and grandisin (13b) (Scheme 9), through acid mediated rearrangement which however takes a different route (Scheme 10) than the one depicted in Scheme 7, precisely due to the presence of the halogen [33].
Scheme 9. Application of halogenated lactones 1e and 1f to lignans of the CL5a subtype.
Scheme 9. Application of halogenated lactones 1e and 1f to lignans of the CL5a subtype.
Molecules 19 19769 g028
The presence of halogen is such dilactones may, in some other cases and depending on the reaction conditions, serve to produce aryltetralins, such as 15a and 15b (Scheme 11), with different oxygenation pattern than normal, by blocking the normal ring closure position [18].
Interestingly, dilactone 1a have been converted to the arylnaphtalene lignan 17 and the lignan 18 of the 2,3-bis(arylmethylene)succinic acid type (Scheme 12) [34]. Thus, alkali treatment of 1a gave as key intermediate the trans-γ-lactone 16 upon acidification. This compound, upon treatment with HCl rearranges (Scheme 13) to arylnaphthalene diacid 17. On the other hand, selective methylation of the carboxyl function with diazo(trimethylsilyl)methane, followed by DBU-mediated opening (a β-elimination) of the γ-lactone and finally saponification provided lignan 18.
Scheme 10. A plausible mechanism for the conversion of dilactones 1e/1f to tetrahydrofuran derivatives 14a/b on treatment with HCl/MeOH.
Scheme 10. A plausible mechanism for the conversion of dilactones 1e/1f to tetrahydrofuran derivatives 14a/b on treatment with HCl/MeOH.
Molecules 19 19769 g029
Scheme 11. Application of halogenated lactone 1g to the synthesis of lignans of the CL3 subtype with a different oxygenation pattern.
Scheme 11. Application of halogenated lactone 1g to the synthesis of lignans of the CL3 subtype with a different oxygenation pattern.
Molecules 19 19769 g030
Scheme 12. Application of dilactone 1a from the dehydrodimerization of FerA to the synthesis of lignans of the arylnaphtalene and 2,3-bis(arylmethylene)succinic acid types.
Scheme 12. Application of dilactone 1a from the dehydrodimerization of FerA to the synthesis of lignans of the arylnaphtalene and 2,3-bis(arylmethylene)succinic acid types.
Molecules 19 19769 g031
Although the dilactones 1a and 1d from the dehydrodimerization of the FerA and SinA respectively, are readily available through FeCl3-mediated POC [23], the corresponding dilactone 1g from the dehydrodimerization of CafA cannot be obtained using this one-electron oxidant. However, Kumada and coworkers reported that this transformation can be effected either enzymatically or non-enzymatically, using CuCl2 as the oxidant, in 64% yield [35]. On the other hand, Jin and coworkers demonstrated that all three dilactones may be obtained using a single oxidant, namely Ag2O, in yields however ranging from 14% to 34% [36]. In the case of FerA, a FerA trimer (19) was also isolated and characterized (Scheme 14). Evaluation of the antioxidant activity of monomers and dimers showed that the most potent were the ones bearing adjacent phenolic functionalities, that is CafA and dilactone 1g and that dimerization results in remarkable enhancement of activity.
Scheme 13. Plausible mechanisms for the conversion of the dilactone 1a to arylnaphtalene 17 on treatment with NaOH first and then HCl.
Scheme 13. Plausible mechanisms for the conversion of the dilactone 1a to arylnaphtalene 17 on treatment with NaOH first and then HCl.
Molecules 19 19769 g032
Dilactones 1a and 1d have been also obtained electrochemically by the anodic oxidation of FerA and SinA, respectively, in MeOH along with NLs of the asatone (20)- and isoasatone (21)-types (Scheme 15) [37]. Interestingly, at low concentrations of HCAs only the NLs were formed.
Scheme 14. Ag2O-mediated dehydrodimerization of HCAs and the structure of the FerA trimer 19.
Scheme 14. Ag2O-mediated dehydrodimerization of HCAs and the structure of the FerA trimer 19.
Molecules 19 19769 g033
Scheme 15. Electrochemical dehydrodimerization of HCAs.
Scheme 15. Electrochemical dehydrodimerization of HCAs.
Molecules 19 19769 g034
These later compounds are plausibly formed through the Diels-Alder reaction of the cyclohexadienones 22, which are obtained through two single-electron oxidations of the corresponding phenolates (Scheme 16) [37].
Interestingly HCAs, such as FerA, can be cross-coupled with the corresponding allylic alcohol (ConAl, 23) to give the anticipated heterodimer, that is monolactone 24. This compound is, however, formed in mixture with the homodimers, that is the dilactone 1a and furofuran lignan pinoresinol (25) (Scheme 17) [38]. The latter was actually isolated as the corresponding diacetate 26. Monolactone 24, isolated from the reaction mixture through CC, is a potent germination inhibitor.
Scheme 16. Plausible mechanisms for the electrochemical dimerization of HCAs.
Scheme 16. Plausible mechanisms for the electrochemical dimerization of HCAs.
Molecules 19 19769 g035
Scheme 17. Cross POC between FerA and ConAl (23).
Scheme 17. Cross POC between FerA and ConAl (23).
Molecules 19 19769 g036
Dilactones of the general type 1 can be also obtained by the so-called non-POC of electronic rich aromatics [21], e.g., of ACAs. Thus, treatment of ACAs with thallium(III) trifluoroacetate in CH2Cl2/CF3CO2H in the presence of BF3.Et2O produced the corresponding dilactones 27 (Scheme 18) in yields ranging from 12% to 54%, based on recovered starting material [39]. Lactone 27a has been partially reduced with DIBAL by Pelter and coworkers in order to prepare the lignan 4,8-dihydroxysesamin (28) [25]. Alternatively, CoF3 may be used for such dimerizations. Although the reactions are slower and less consumption of starting materials and poorer yields are observed with this reagent, its use is associated with several advantages over Th(OCOCF3)3. For example, POC may also be performed with CoF3 but not with Th(OCOCF3)3. Accordingly, CoF3-mediated dehydrodimerization of FerA produced dilactone 1a in 20% yield.
Scheme 18. Thallium(III) trifluoroacetate-mediated dehydrodimerization of ACAs.
Scheme 18. Thallium(III) trifluoroacetate-mediated dehydrodimerization of ACAs.
Molecules 19 19769 g037
Scheme 19. Synthesis of lignans 31a and 31b through oxidative dimerization of the amide (29) of O-methylsinapic acid with L-Pro.
Scheme 19. Synthesis of lignans 31a and 31b through oxidative dimerization of the amide (29) of O-methylsinapic acid with L-Pro.
Molecules 19 19769 g038
The non-POC concept was further exploited by Mori and coworkers in the dehydrodimerization of the N-(3,4,5-trimethoxycinnamoyl)-l-Pro (29). This resulted in the stereoselective assembly of the skeleton of lignans of the 2,6-diarylfurofuran (CL6) subtype, such as yangambin (31a) and caruilignan A (31b) (Scheme 19) with a variety of pharmacological activities [40]. The method involved oxidation of the electron-rich aromatic ring of compound 29 with PbO2/TFA/Celite in CH2Cl2 at −78 °C to the corresponding benzylic cation radical which, through a cyclic radical, was dimerized and finally hydrolyzed to give furanofuranone 30 in 64% yield and e.e. >95% [40]. l-Pro acted as the chiral auxilliary inducing chirality in the dimerization step. Intermediate 30 was then readily transformed to the desired final products 31a and 31b.

2.2. Syntheses of Lignans Based on the Key Intermediate Dialkyl 1,2-dihydro-1-arylnaphthalene-2,3-dicarboxylate

Treatment of Me-Sin (32) with FeCl3·6H2O in acetone-water produced a mixture of compounds, the major component of which was the tetralol 33a. This compound was obtained in 61% yield following crystallization. The other components of the reaction mixture were separated with CC from the residue of the mother liquid. These included the dimethyl trans-1,2-dihydro-1-arylnaphthalene-2,3-dicarboxylate (dimethyl thomasidioate, 9a), lactone 34, and the diastereomer 33b received in mixture with a quantity of unprecipitated 33a. This mixture was first methylated and then rechromatographed to provide pure methyl ether 33c (Scheme 20) [30].
Scheme 20. FeCl3-mediated POC of Me-Sin.
Scheme 20. FeCl3-mediated POC of Me-Sin.
Molecules 19 19769 g039
Compound 33a was dehydrated to arylnaphthalene 9a under treatment with perchloric acid in acetic acid. However, this conversion could not be effected under the acidic conditions (FeCl3, pH ca. 2) of the POC experiment described above. This indicates that the 4-hydroxyarylnaphthalenes 33a and 33b are formed through a different mechanism (Scheme 21) than the one producing the dehydrated product 9a [29]. Interestingly, CAN-mediated POC of 32 produced a similar product mixture to that from FeCl3-mediated POC, whereas the use of potassium ferricyanide in acetone-water produced dimethyl thomasidioate (9a) in 16% yield as the only fluorescent compound under UV irradiation. On the other hand, treatment of 32 with silver carbonate produced a complex reaction mixture of non-fluorescent products.
Scheme 21. Outline of the plausible mechanisms for the formation of products 33a,b and 9a from POC of Me-Sin.
Scheme 21. Outline of the plausible mechanisms for the formation of products 33a,b and 9a from POC of Me-Sin.
Molecules 19 19769 g040
Similarly, 5-Br or -I substituted Me-Fer (35a and b) can be directly converted to dihalogenated aryl dihydronaphthalenes (36a and 36b) in 23% and 75% crude yields respectively, through POC, by using FeCl3 as one-electron oxidant. Dimer 36b was further used in a synthesis of (±)-isolariciresinol dimethyl ether 37 (Scheme 22) [32].
Setälä and coworkers also obtained the aryldihydronaphthalene 9a, in 41% yield, directly through POC of Me-Sin (32) but using hydrogen peroxide as oxidant and the enzyme HRP as catalyst at pH 4 in acetone-water. In earlier work, the authors had observed that in low pH-values the formation of oligomers/polymers is suppressed, thus favouring higher yields of the desired dimers. Interestingly, when the same reaction was performed in aqueous methanol dimer 9a was produced in only 14% yield, the main product (49% yield) being a mixture of the spiro dimers 38a and 38b (Scheme 23 and Scheme 24) [41]. This mixture failed to produce the aryldihydronaphthalene 9a upon attempted acid-mediated dienone-phenol rearrangement, which however led to the isomeric aryldihydronaphthalene 39 (Scheme 25). Thus, dimers 38 cannot be intermediates in the formation of 9a.
Scheme 22. FeCl3-mediated POC of 5-halogenated Me-Fers.
Scheme 22. FeCl3-mediated POC of 5-halogenated Me-Fers.
Molecules 19 19769 g041
Scheme 23. Biomimetic POC of Me-Sin with the system HRP-H2O2.
Scheme 23. Biomimetic POC of Me-Sin with the system HRP-H2O2.
Molecules 19 19769 g042
Scheme 24. Outline of the plausible mechanisms for the formation of products 9a and 38a,b.
Scheme 24. Outline of the plausible mechanisms for the formation of products 9a and 38a,b.
Molecules 19 19769 g043
Scheme 25. Outline of the plausible mechanism for the conversion of products 38a,b to 39.
Scheme 25. Outline of the plausible mechanism for the conversion of products 38a,b to 39.
Molecules 19 19769 g044
Bunzel and coworkers synthesized dimethyl thomasidioate (9a) from SinA in a two-steps one-pot sequence involving treatment of SinA with acetyl chloride in methanol followed by treatment of the thus obtained in situ Me-Sin (32) with aqueous FeCl3.6H2O. That way, the dimer 9a was obtained in 52% yield. Finally, saponification of diester 9a with 2M NaOH produced ThoA (8) in 60% yield (Scheme 26) [42].
Scheme 26. Synthesis of ThoA (8) from SinA.
Scheme 26. Synthesis of ThoA (8) from SinA.
Molecules 19 19769 g045
Neudorffer and coworkers electrochemically generated the phenoxy radical from Et-Sin (40), in acetonitrile containing lithium perchlorate as the supporting electrolyte and a stoichiometric amount of tetramethylammonium hydroxide, which led to the synthesis of diethyl thomasidioate (41) in 42% yield (Scheme 27), along with 12% unreacted ester 40 [43].
Scheme 27. Electrochemical POC of Et-Sin (40).
Scheme 27. Electrochemical POC of Et-Sin (40).
Molecules 19 19769 g046
When the corresponding Et-Fer (42) was subjected to anodic electrolysis under the above described reaction conditions, the corresponding dimer 43 was obtained in 15% yield along with 28% of unreacted ester 42 and three additional dimers. These included one lignan 44 of the bis(arylmethylene)succinic acid type in 8% yield, one neolignan 45 of the dihydrobenzofuran type in 13% and one oxyneolignan 46 with an β-O-4 bond in 10% (Scheme 28) [43].
Zoia and coworkers developed an enantioselective biomimetic approach to synthesize ThoA (8) [44]. They subjected amides 47ac of SinA with chiral amines, such as (S)-Phe ethyl ester, (S)-methylbenzylamine and (S)-2-phenyloxazolidinone, the latter acting as chiral auxiliaries, to POC using hydrogen peroxide as oxidant and HRP as catalyst in a dioxane-water, buffered at pH 4, solution. The amides were obtained from the coupling of SinA and the corresponding amines in 40%–50% yields. Although POC of Me-Sin with this system produced a racemic mixture of the dimethyl trans-thomasidioates 9a and 9a' in 70% yield, the presence of the chiral auxilliaries secured the creation of mixtures of the corresponding two diastereomeric bisamides I and II (48ac) of trans-ThoA with d.e. in the range of 40%–70%. The total yields (both diastereomers) of the thus obtained chiral amides of ThoA (8) were in the range 40%–60% (Scheme 29). Interestingly, best d.e. (70%) was obtained with (S)-2-phenyloxazolidinone, as the chiral auxilliary, whereas best yield (60%) with (S)-Phe ethyl ester, as the chiral auxilliary. Amides I (48ac) could be hydrolyzed to ThoA (8) with LiOOH in THF-H2O in 10%–60% yields, the highest one obtained with the amide (48aI) from (S)-2-phenyloxazolidinone. Computational methods showed that enantio- and diastereoselectivity are controlled by the β-β oxidative coupling step and the stability of the reactive conformation of the intermediate quinonemethide, respectively [44].
Scheme 28. Electrochemical POC of Et-Fer. The primary bond connecting the monomers is shown in blue and bolded line.
Scheme 28. Electrochemical POC of Et-Fer. The primary bond connecting the monomers is shown in blue and bolded line.
Molecules 19 19769 g047
Cilliers and Singleton subjected CafA to oxidation with O2 in an alkaline (KOH) aqueous solution (pH 8.5). Although POC on free HCAs usually leads to dilactones 1, in this case, a mixture of mainly dimeric products (see also Section 3.1 and Section 3.3) and unreacted CafA was obtained, the main component of which was caffeicin E (49), that is a lignan of the substituted 1,2-dihydronaphthalene (CL3) type (Scheme 30) [45].
Maeda and coworkers reported that POC of the HCA methyl ester 51, readily derived from the coumarin daphnetin 50, mediated by iron(III) chloride heptahydrate in aqueous acetone solution, gave 1,2-dihydronaphthalene 52 in 16% yield (Scheme 31) and 41% recovery of starting material [46]. Changing the oxidant to silver oxide or potassium hexacyanoferrate(III) in the presence of Na2CO3, resulted in very low yields (1%–5%) of 52, isolated as the corresponding tetra-acetate (53), along with other dimeric products. Compound 52 was shown to be a potent inhibitor of lipid peroxidation. In an earlier study, the same authors obtained the isomeric 1,2-dihydronaphalene, isolated as the corresponding tetra-acetate 56, as the main product from POC of methyl ester 55 using either silver oxide (28%), or potassium hexacyanoferrate/sodium acetate (20%) or sodium carbonate (13%), as the oxidant [47]. Ester 55 was readily obtained from the coumarin esculetin (54).
Scheme 29. Biomimetic synthesis of ThoA (8) through asymmetric POC of chiral amides of SinA.
Scheme 29. Biomimetic synthesis of ThoA (8) through asymmetric POC of chiral amides of SinA.
Molecules 19 19769 g048
Scheme 30. POC of CafA using O2 as oxidant.
Scheme 30. POC of CafA using O2 as oxidant.
Molecules 19 19769 g049
Scheme 31. POC of HCA methyl esters 51 and 55, obtained from coumarins.
Scheme 31. POC of HCA methyl esters 51 and 55, obtained from coumarins.
Molecules 19 19769 g050
Agata and coworkers isolated from Rabdosia japonica Hara a CafA tetramer, which they named rabdosiin (57). The aboveground part of this plant is used in Japan as a common household medicine for gastrointestinal disorders, called “emeiso”. Compound is a diester (hybrid) of the lignan 58 of dihydronaphthalene type with the 1R,2S configuration and the (R)-α-hydroxyacid 59. In order to prove its structure, this research group subjected Me-Caf (60) into POC using FeCl3 as the oxidant. That way, they obtained racemic lignan 61 in 55% yield. Rabdosiin is thought to be biogenetically produced through the oxidative coupling of two molecules of (R)-(+)-RosA (62) in a way analogous to the dehydrodimerization of Me-Caf towards the lignan skeleton 61 (Scheme 32) [48,49]. (−)-Rabdosiin and particularly its monopotassium and monosodium salts have potent anti-HIV activity [50]. It has been also found to be a strong inhibitor of DNA topoisomerases I and II [51]. Interestingly, natural rabdosiin has been isolated in both diastereomeric forms, namely the (−)-(1R,2S) and the (+)-(1S,2R) forms, although in both diastereomers the Dpl ester groups are R.
(R)-(+)-RosA (62) is a major constituent of the Chinese medicinal herb Danshen which has been used for the treatment of heart disease. This compound was first isolated from extracts of Rosmarinus officinalis and exhibits anti-inflammatory, antioxidative, antihistamine, antibacterial, antiviral and antihormonal effects [52]. Bogucki and Charlton developed a chemical synthesis to (S)-(−)-RosA (62a) from (S)-Tyr and CafA in 9% overall yield, which could be also applicable to (R)-Tyr [52]. In the latter case, (R)-(+)-RosA could be produced. Although attempts to oxidatively couple 62a to obtain rabdosiin by the same authors were unsuccessful, a triallylated derivative 62b of 62a was indeed converted to (−)(57)- and (+)(57a)-rabdosiin, following deprotection, albeit in low yields with the (−)-isomer predominating (Scheme 33).
The same research group subjected the ester 64 of SinA with methyl (R)-mandelate into POC, using a similar method to that used by Wallis to dehydrodimerize the corresponding Me-Sin, and obtained three main products, namely two 1,2-trans-thomasidioate esters 65 and 66 in 53 and 23% yields and one 1,2-cis diastereomer 67 in 8% yield. It is therefore evident that chiral esters of HCAs lead to diastereoselective production of coupled-cyclized dimers (Scheme 34) [52].
Scheme 32. POC of Me-Caf (60) using FeCl3 as oxidant and the structures of rabdosiin (57) and RosA (62).
Scheme 32. POC of Me-Caf (60) using FeCl3 as oxidant and the structures of rabdosiin (57) and RosA (62).
Molecules 19 19769 g051
Scheme 33. POC of a triallylated derivative 62b of (S)-(−)-RosA (62a) en route to (−)- and (+)-rabdosiin (57 and 57a, respectively).
Scheme 33. POC of a triallylated derivative 62b of (S)-(−)-RosA (62a) en route to (−)- and (+)-rabdosiin (57 and 57a, respectively).
Molecules 19 19769 g052
Scheme 34. POC of ester 64 utilizing Wallis' methodology.
Scheme 34. POC of ester 64 utilizing Wallis' methodology.
Molecules 19 19769 g053
Figure 1. Structures of biologically interesting esters and oligomers of RosA.
Figure 1. Structures of biologically interesting esters and oligomers of RosA.
Molecules 19 19769 g001
It should be noted that syntheses of the methyl and butyl esters 68 and 69 of both enantiomers (R)-(+)(62)- and (S)-(−)(62a)-RosA have been published by Reimann and Pflug [53] and Huang and coworkers [54], respectively (Figure 1). Esters 69a and 69b showed interesting antioxidant and antitumour activity [54]. (R)-(+)-RosA has been recently shown to be a potent inhibitor against snake venom [55]. Interestingly, oligomerization of RosA in nature produces complex NLs isolated from the Indinesian plant Helictora isora possessing mild inhibitory activity against the avian myeloblastosis virus, such as helicterins A (70a) and B (70b), helisorin (71) and helisterculin A (72) (Figure 1). Syntheses for such compounds have been recently disclosed by Snyder and Kontes [56].
Examples of recently reported interesting natural products of the ester (depsides) or the amide type containing the lignan dihydronaphthalene nucleus are the mono-ester 73 of caffeicin E, named SalA-R and the tris(tyramide) thoreliamide C (74) (Figure 2), the latter also incorporating a β-O-4 bond, isolated from Origanum dictamnus L. by Exarchou and coworkers [57] and from Mitrephora thorelii by Ge and coworkers [58], respectively.
Figure 2. Structures of naturally occurring lignans incorporating a 1,2-dihydronaphthalene substructure.
Figure 2. Structures of naturally occurring lignans incorporating a 1,2-dihydronaphthalene substructure.
Molecules 19 19769 g002
Daquino and coworkers subjected Me-Caf (60) and CAPE (75) to various oxidants and studied the products of the oxidative coupling reactions [59]. Ester 75 is a natural product contained in propolis with anti-inflammatory, antioxidant and antitumour activity [60]. Thus, POC of CAPE using MnO2 as oxidant in CH2Cl2 produced an unusual lignan 76 of the benzo[kl]xanthene type and the dihydronaphthalene 78 in 48% and 16.5% yields, respectively, along with residual CAPE (Scheme 35). Highest yield (72%) in 76 was obtained when CHCl3 was used as the solvent, whereas highest yield (23.7%) in 78 was obtained with n-hexane/CH2Cl2 (1:1) as the solvent. Under analogous reaction conditions and using CHCl3 as the solvent, Me-Caf gave the corresponding products 77 and 79 in 51% and 7% yields, respectively. CAPE was also oxidatively dimerized in the presence of Mn(OAc)3 in CHCl3 and produced, with 89% conversion, 76 and 78 in 71% and 22% yields, respectively. This then appears to be the best method to synthesize dimer 76. On the contrary, CafA itself failed to produce the natural benzoxanthene lignan mongolicumin (80) in satisfactory yield (<10%) under these reaction conditions. However, lignan 80 could be obtained from dimer 77 through partial alkaline hydrolysis, acid-mediated cyclization and finally mild basic hydrolysis in 40% overall yield based on CafA as starting material. Benzoxanthenes have shown potential antitumour activity and, because of their strong fluorescence, may be of considerable interest in the design of new fluorescent probes for biomedical applications [59].
Scheme 35. POC of CafA esters utilizing MnO2 or Mn(OAc)3 as the oxidant.
Scheme 35. POC of CafA esters utilizing MnO2 or Mn(OAc)3 as the oxidant.
Molecules 19 19769 g054
Scheme 36. Plausible mechanism for the oxidative dimerization of CafA esters towards lignans of the benzoxanthene type.
Scheme 36. Plausible mechanism for the oxidative dimerization of CafA esters towards lignans of the benzoxanthene type.
Molecules 19 19769 g055
The proposed, by the authors, mechanism for the formation of benzoxanthane lignans 76 and 77 is outlined in Scheme 36.
It should be noted that Maeda and coworkers isolated a benzoxanthene lignan (81a), as the corresponding triacetate (81b) during the POC of the methyl ester 55 (Scheme 37), in particular when the oxidant used was silver oxide, along with the dihydronaphthalene 56 and other compounds [47]. Formation of compound 81a may be also explained with a similar mechanism to that outlined in Scheme 36. Both compounds (56 and 81b) showed potent inhibition of lipid peroxidation.
Scheme 37. POC of ester 55 utilizing Ag2O as the oxidant. Stretched structures are used, in the proposed mechanism outlining formation of product, for the sake of drawing clearly electron movement and for avoiding overcrowding of groups.
Scheme 37. POC of ester 55 utilizing Ag2O as the oxidant. Stretched structures are used, in the proposed mechanism outlining formation of product, for the sake of drawing clearly electron movement and for avoiding overcrowding of groups.
Molecules 19 19769 g056
Asymmetric syntheses of aryltetralin lignans attract considerable interest due to the fact that the clinically important antitumor drugs Etoposide and Teniposide are derived from the natural lignan podophyllotoxin (82) (Scheme 38) [61].
Taking into consideration the fact that podophyllotoxin is biosynthesized by the oxidative cyclisation of a lignan of the dibenzylbutyrolactone (CL2) type, such as matairesinol (83) or yatein (84) (Scheme 38) [62], attempts were made to effect such cyclizations using laboratory reagents. They were, however, unsuccessful as they invariably lead to the formation of an eight-membered ring rather than the anticipated six-membered ring [19]. Interestingly, Cambie and coworkers managed to cyclize yatein (84) by non-POC with formation of the six-membered ring, thus producing deoxyisopodophyllotoxin (85), using thallium(III) oxide as oxidant in TFA (Scheme 39) [63].
Scheme 38. Outline of the proposed biosynthetic pathway to podophyllotoxin (82).
Scheme 38. Outline of the proposed biosynthetic pathway to podophyllotoxin (82).
Molecules 19 19769 g057
Scheme 39. non-POC of yatein (84) to deoxyisopodophyllotoxin (85).
Scheme 39. non-POC of yatein (84) to deoxyisopodophyllotoxin (85).
Molecules 19 19769 g058
Non-POC on dibenzylbutyrolactones (e.g., 86 and 87) has been also effected with other reagents, e.g. DDQ and Ru(OCOCF3)4, with formation of an eight-membered ring, thus providing access to lignans of the dibenzocyclooctadiene (CL4) type and in particular isosteganes, such as 88 and 89, respectively (Scheme 40) [19].
Scheme 40. non-POC of the dibenzylbutyrolactones to isosteganes.
Scheme 40. non-POC of the dibenzylbutyrolactones to isosteganes.
Molecules 19 19769 g059
Hypervalent iodine reagents, such as PIDA and PIFA, have been used by Ward and coworkers to effect biomimetic oxidative couplings in phenolic trans (e.g., 90 and 91)- or cis-dibenzylbutyrolactones and which result to isosteganes (93 and 94, respectively) directly or indirectly through spirodienones (e.g., 92) with acid-mediated dienone-phenol rearrangement (Scheme 41) [19].
Scheme 41. Use of PIFA in the biomimetic POC of phenolic dibenzylbutyrolactones to isosteganes.
Scheme 41. Use of PIFA in the biomimetic POC of phenolic dibenzylbutyrolactones to isosteganes.
Molecules 19 19769 g060

2.3. Syntheses of Lignans Based on the Key Intermediate Dialkyl Bis(arylmethyle-ne)succinate

Sarkanen and Wallis subjected cinnamate ester 95 into POC coupling using alkaline potassium ferricyanide as oxidant and obtained a mixture of the bisquinonemethides 96a and 96b, in the ratio 65:35, in 72% yield. These compounds can be tautomerized to the same dimethyl bis(arylmethylene)succinate (44a), whereas upon catalytic hydrogenation they are transformed to the corresponding diarylbutanes 97a and 97b. Furthermore, diesteres 97a and 97b could be converted to the corresponding diastereomeric tetrahydrofurans 98a and 98b upon reduction with LAH, followed by acid-mediated ring closure with dehydration (Scheme 42) [64].
It should be noted that the diethyl bis(arylmethylene)succinate 44 (Scheme 28) was identified by Neudorffer and coworkers as the minor product of the electrochemical POC of Et-Fer [43]. On the other hand, Bunzel and coworkers synthesized diethyl bis(arylmethylene)succinate 44b through POC of Et-Sin, using Mn(OAc)3.2H2O as oxidant in pyridine (Scheme 43) [42]. They isolated its corresponding diacetate 44c in 85% yield. From this compound, free acid 44d was obtained, in a two-step saponification, however in low yield (30%) along with unsaponified diester 44b.
Wang and coworkers subjected Et-Fer (42) into an alkanine potassium ferricyanide-mediated POC in PhH-H2O to obtain the corresponding diethyl bis(aryl-methylene)succinate 44e, however as a minor product of the reaction in only 9% yield (Scheme 44). The main product of the reaction was the benzofuran 99 which was obviously formed through the alternative primary β-5 coupling [65].
Scheme 42. POC of cinnamate ester 95 producing bisquinonemethides 96.
Scheme 42. POC of cinnamate ester 95 producing bisquinonemethides 96.
Molecules 19 19769 g061
Scheme 43. POC of Et-Sin (40) using Mn(OAc)3·2H2O.
Scheme 43. POC of Et-Sin (40) using Mn(OAc)3·2H2O.
Molecules 19 19769 g062
Scheme 44. POC of Et-Fer (42) using K3Fe(CN)6. In blue solid line the primary bonds and in red the secondary one.
Scheme 44. POC of Et-Fer (42) using K3Fe(CN)6. In blue solid line the primary bonds and in red the secondary one.
Molecules 19 19769 g063
When, however, they blocked the position 5 by a tert-butyl group, they succeeded to synthesize the β-β' coupling product 44f in excellent yield from the Et-Fer derivative 100 (Scheme 45). That way, alternative dimerizations involving position C-5 or even O-4, due to steric hindrance, can be avoided.
Scheme 45. POC of ethyl 5-(t-butyl)ferulate (100) using K3Fe(CN)6.
Scheme 45. POC of ethyl 5-(t-butyl)ferulate (100) using K3Fe(CN)6.
Molecules 19 19769 g064
Compound 44f is a key intermediate in the synthesis of a variety of lignans of the dibenzylbutane (CL1), the 3,4-dibenzyltetrahydrofuran (CL5c) and the arylnaphtalene (CL3) type. Examples of lignans of these subtypes are the dibenzylbutanediols meso-secoisolariciresinol (101a) and (±)-secoisolariciresinol (101b), the corresponding fully reduced derivatives, the dibenzylbutanes meso-dihydroguaiaretic acid 102a and (±)-dihydroguaiaretic acid 102b, the 3,4-dibenzyltetrahydrofurans meso-divanillyltetrahydrofuran 103a and (±)-divanillyltetrahydrofuran 103b and the arylnaphalene derivative 104 (Scheme 46). The secoisolariciresinols exhibit interesting cytotoxic or immunosuppressive activities [66].
Thus, compound 44f could be either reduced catalytically to provide access to compounds 101103, following chromatographic separation of the thus obtained diastereomeric diesters, or treated with AlCl3 to produce the 7,8-dihydro-7-arylnaphalene diester 104. (Scheme 46) [65]. AlCl3 not only mediated the required intramolecular electrophilic aromatic substitution, obviously through a p-quinonemethide intermediate, but also performed the removal of the tert-butyl protecting group.
Scheme 46. Synthesis of lignans 101104 through oxidative dimerization of Et-Fer, bearing a tert-butyl blocking group at position 5. Stretched structures are used, in the proposed mechanism outlining formation of product, for the sake of clearly showing the electron movement.
Scheme 46. Synthesis of lignans 101104 through oxidative dimerization of Et-Fer, bearing a tert-butyl blocking group at position 5. Stretched structures are used, in the proposed mechanism outlining formation of product, for the sake of clearly showing the electron movement.
Molecules 19 19769 g065
Moon and coworkers recently isolated a novel cytotoxic lignan from the seeds of Trichosanthes kirilowii, identified it as the monoester of (−)-(2R,3R)-secoisolariciresinol with FerA and designated it as hanultarin (105). This compound and the already known diester 1,4-O-diferuloyl-secoisolariciresinol (106) (Figure 3), also isolated from the seeds, showed comparable cytotoxic effects to cisplatin against several cancer cell lines [67]. Lee and coworkers synthesized, with other than the POC route, for the first time the above compounds and analogs and evaluated their cytotoxicity against several cancer cell lines [68]. They found that, although neither FerA nor secoisolariciresinol are cytotoxic, their conjugation leads to cytotoxic hybrids with the most potent being the diferuloyl ester 106 of secolariciresinol. They also proposed that the observed cytotoxic effect might originate from the inhibitory effect of the polymerization of actin or microtubule fibers.
Figure 3. The structures of (−)- and (+)-secoisolariciresinol, hanultarin (105) and 1,4-O-diferuloyl-secoisolariciresinol (106).
Figure 3. The structures of (−)- and (+)-secoisolariciresinol, hanultarin (105) and 1,4-O-diferuloyl-secoisolariciresinol (106).
Molecules 19 19769 g003
It is interesting to note that amides of the bis(arylmethylene)succinic acid have been isolated from plants. For example, Ma and coworkers isolated from the seeds of Hyoscyamus niger four lignanamides, that is hyoscyamide (107), cannabisin G (108), cannabisin D (109) and grossamide (110) (Figure 4) [69]. Hyoscyamide and cannabisin G are dimers of (Z)- and (E)-N-feruloyltyramine, respectively. On the other hand, grossamide and cannabisin D are bistyramides of dimers of FerA of the benzofuran neolignan and the arylnaphtalene lignan types, respectively. Grossamide and cannabisins D and G showed moderate cytotoxicity in cultured LNCaP human prostate cancer cells.
Figure 4. Structures of lignanamides from seeds of Hyoscyamus niger.
Figure 4. Structures of lignanamides from seeds of Hyoscyamus niger.
Molecules 19 19769 g004
Tomosaka and coworkers also isolated cannabisin G (108), along with (±)-lyoniresinol (11a) (Scheme 8), from the root bark of Berberis vulgaris L. Although both compounds exhibited antioxidant activity in a hydroxyl radical-scavenging assay, only canabbisin G showed cytoprotective activity in cultured MCF-7 cells modulated by hydrogen peroxide [70]. Li and coworkers have recently published a concise synthesis of cannabisin G using as key reaction the alkaline ferricyanide-mediated oxidative dimerization of ethyl 5-(t-butyl)ferulate (100) [71].

2.4. Syntheses of Lignans Based on the Key Intermediate 3,4-Disubstituted 2,5-Di-aryltetrahydrofuran

As outlined above (Scheme 9), lignans of the 2,5-diaryltetrahydrofuran (CL5a) type can be obtained indirectly through acid-mediated rearrangement of halogenated bilactones (e.g., 1e, 1f). However, Ahmed and coworkers also managed to obtain directly such a compound through POC of suitable halogenated HCAs, such as 111, using FeCl3 as oxidant in the presence of hydrochloric acid in low concentration. That way, they isolated the tetrahydrofuran intermediate 112 in 13% yield, along with unreacted starting material (Scheme 47).
Scheme 47. Synthesis of (±)-veraguensin (113).
Scheme 47. Synthesis of (±)-veraguensin (113).
Molecules 19 19769 g066
This compound could be converted into the lignan (±)-veraguensin (113), in several steps including O-methylation, removal of the bromine atoms with catalytic hydrogenolysis, LAH-mediated reduction, tosylation and again LAH-mediated reduction [32]. However, attempts to convert directly the mono-brominated HCA 111a to the corresponding dimer 112a were unsuccessful [33]. Formation of dimer 112 was explained through HCl-catalyzed interception by water of the initially formed intermediate bisquinonemethide 114 (Scheme 48).
Scheme 48. Outline of the proposed mechanism for the formation of the 2,5-diaryl- tetrahydrofuran lignan 112 by POC of HCA 111.
Scheme 48. Outline of the proposed mechanism for the formation of the 2,5-diaryl- tetrahydrofuran lignan 112 by POC of HCA 111.
Molecules 19 19769 g067

3. Neolignan Skeleton Assembly

3.1. Synthesis of NLs of the Benzofuran Type—Formation of β-5 Bond as the Key Step

Cilliers and Singleton treated an alkaline aqueous solution of CafA with O2 and identified a series of POC products designated as caffeicins. One of them, designated caffeicin F (115), is a neolignan of the 2,3-dihydrobenzofuran (NL1) type [45]. Although no relative stereochemistry is provided for 115, the coupling constant between protons 2 and 3 (J = 7.2 Hz) indicates a trans configuration (Figure 5).
Chioccara and coworkers subjected Me-Fer (42a) to POC using HRP-H2O2 in MeOH (10% or 90% v/v)-aqueous buffer at pH 3 to obtain the benzofuranoid dimer 116 in 30% yield after crystallization (Scheme 49) [72]. Higher pH values facilitate oligomers formation. A very minute amount of the oxyneolignan (β-Ο-4 dimer) 117 was also produced in the reaction in both erythro and threo diastereomers in nearly 1:1 ratio.
Figure 5. A neolignan (115) of the 2,3-dihydrobenzofuran type (NL1).
Figure 5. A neolignan (115) of the 2,3-dihydrobenzofuran type (NL1).
Molecules 19 19769 g005
Scheme 49. Synthesis of neolignan 116 through oxidative dimerization of Me-Fer using HRP-H2O2 as oxidant system.
Scheme 49. Synthesis of neolignan 116 through oxidative dimerization of Me-Fer using HRP-H2O2 as oxidant system.
Molecules 19 19769 g068
The same reaction could be performed using either an excess of iodosylbenzene in the presence of Mn(III)TPPOAc as catalyst or a stoichiometric amount of iodosylbenzene with MnTTPCl or even excess H2O2 with MnTPPCl to obtain benzofuran 116 in 22%, 36% and ca. 25% yields, respectively.
Scheme 50. Outline of the proposed mechanism for the formation of the neolignan 116 and the oxyneolignan 117.
Scheme 50. Outline of the proposed mechanism for the formation of the neolignan 116 and the oxyneolignan 117.
Molecules 19 19769 g069
Compounds 116 and 117 are thought to be formed through the common intermediate 118 which is intercepted either intramolecularly from the adjacent hydroxyl group or intermolecularly from methanol (Scheme 50).
Maeda and coworkers reported that POC of the HCA methyl ester 51, mediated by silver oxide in benzene-acetone, produced the benzofuranoid dimer 119a as the main product, isolated as its corresponding triacetate 119b in 15% yield (Scheme 51) [46] and 21% recovery of starting material as the corresponding diacetate. No such product was isolated by using either potassium hexacyanoferrate(III)/Na2CO3 or ferric chloride as oxidant. Compound 119b exhibited potent inhibition of lipid peroxidation both in rat brain homogenate and rat liver microsomes. In earlier studies, the same authors synthesized the benzofuranoids 116 and 121 from Me-Fer (42a) and its regioisomer 120, respectively, and converted the former into the lignin schizotenuin D (122) and related compounds with potent inhibitory effects on lipid peroxidation [73,74]. The common structural characteristic of all starting materials used in these POCs producing benzofuranoid NLs is that their 5 position is unhindered.
Scheme 51. POC of Me-Fer (42a) and of HCA methyl esters 51 and 120, obtained from coumarins, producing benzofuranoid NLs.
Scheme 51. POC of Me-Fer (42a) and of HCA methyl esters 51 and 120, obtained from coumarins, producing benzofuranoid NLs.
Molecules 19 19769 g070
Bolzacchini and coworkers subjected the amide 123 of FerA with ethyl (S)-alaninate, used as chiral auxilliary, to POC employing HRP-H2O2 as oxidant system in dioxane-aqeous buffer pH 3. That way, a mixture of the two diastereomeric benzofuranoid dimers 124 and 125 was obtained in 70% yield and d.e. 65%, which could be separated by silica gel FCC (Scheme 52) [75]. It should be reminded that, without a chiral auxilliary, the oxidant system HRP-H2O2 produces racemic mixtures. The major diastereomer had the 2S,3S absolute stereochemistry, as this was shown by LiOOH-mediated hydrolysis, esterification with diazomethane, reduction with LiBH4 and finally identification with the known 2S,3R stereoisomer 126 of dehydrodiconiferyl alcohol.
Scheme 52. Diastereoselective POC of ferulamide 123 using HRP-H2O2.
Scheme 52. Diastereoselective POC of ferulamide 123 using HRP-H2O2.
Molecules 19 19769 g071
Ralph and coworkers developed a simple and cheap method for the preparation of diethyl diferulate 45 (Scheme 28), in gram quantities and in reasonable yields (ca. 50% yield), from Et-Fer (42) utilizing the biomimetic HRP-H2O2 system in an aqueous acetate buffer pH 4.0 [76].
Considering the natural neolignan 3',4-di-O-methylcedrusin (139) with known inhibitory activity of cell proliferation as lead compound, Pieters and coworkers synthesized a series of 19 related dihydrobenzofuran lignans and benzofurans, suitable for SARS, by utilizing a biomimetic reaction sequence involving as key reaction the POC of p-coumaric (127), caffeic (60) or ferulic (42a) acid methyl esters with silver oxide in anhydrous PhH-acetone. The anticipated dihydrobenzofurans 128, 115 and 116 were obtained in 23%–50% yields. Compound 116 was O-methylated giving the diferulate analog 129. This compound was oxidized with DDQ to the corresponding benzofuran 130, which was then catalytically reduced leading to benzofuran analog 131. On the other hand, dihydrobenzofurans 115116 and 128129 were catalytically reduced giving analogs 132135, which were further reduced with LAH leading to the diols 136139. Furthermore, catalytic hydrogenation of 135 caused opening of the dihydrofuran ring giving alcohol 140 (Scheme 53). All synthesized compounds were tested for potential anticancer activity in 60 human tumor cell lines. The dihydrobenzofuran dimer 115 from Me-Caf was the most potent. It should be noted that the two enantiomers of racemic 115 were separated by chiral HPLC and that the 2R,3R enantiomer was the more active one. Furthermore, the two enantiomers of racemic neolignan 129 were also separated in order to obtain 3',4-di-O-methylcedrusin (139), with the 2R,3S assigned configuration, a minor constituent of the traditional medicament “dragon’s blood”. Surprisingly, no activity was exhibited by this particular compound. Leukemia cell lines and breast cancer lines were relatively more sensitive to the cytotoxic dihydrobenzofuran NLs. These studies established that the NLs of the dihydrobenzofuran type form a new group of antimitotic and potential antitumor agents which inhibit tubulin polymerization [77].
Scheme 53. Synthesis of NLs of the benzofuran type using as key reaction the Ag2O-mediated POC of methyl esters of HCAs.
Scheme 53. Synthesis of NLs of the benzofuran type using as key reaction the Ag2O-mediated POC of methyl esters of HCAs.
Molecules 19 19769 g072
Carunchio and coworkers subjected FerA to POC using the enzyme laccase in EtOH-aqueous buffer pH 6.0 and the reaction was monitored by HPLC-MS. They identified the two main products initially formed as the benzofuranoid dimer 141 and the β-O-4 coupling product 142 (Scheme 54) [78]. They also found that enzymatic activity is inhibited by dipeptides of the type Gly-X due to the formation of complexes with Cu(II), known to play a central role in the laccase, single electron transfer, reaction mechanism.
Scheme 54. Dimers from laccase-mediated POC of FerA.
Scheme 54. Dimers from laccase-mediated POC of FerA.
Molecules 19 19769 g073
A dihydrobenzofuran dimer, namely compound 45 was also isolated by Neudorffer and coworkers, as the second most abundant compound (13% yield) of the mixture obtained by the electrochemical POC of Et-Fer (Scheme 28) [43].
Kuo and Wu reported a biomimetic multistep synthesis of salvinal (143) [79], a benzofuranoid NL isolated from the root of Salvia miltiorrhizae Bunge, whose aqueous extracts have been extensively used in Asia in the treatment of cardiovascular disorders and cancer. Salvinal is a novel Adenosine A1 receptor ligand and a novel microtubule inhibitor with antimitotic activity in multidrug-sensitive and -resistant human tumor cells [80]. Key step in the synthesis of salvinal was POC of Me-Fer by ferric chloride in aqueous acetone (Scheme 55) producing the known dihydrobenzofuran dimer 116 in 34% yield.
Scheme 55. Synthesis of salvinal (143) through FeCl3-mediated POC of Me-Fer (42a).
Scheme 55. Synthesis of salvinal (143) through FeCl3-mediated POC of Me-Fer (42a).
Molecules 19 19769 g074
Scheme 56. Ag2O-mediated POC of Me-Fer and application to the synthesis of the dihydrobenzofuran-gallic acid hybrid 147.
Scheme 56. Ag2O-mediated POC of Me-Fer and application to the synthesis of the dihydrobenzofuran-gallic acid hybrid 147.
Molecules 19 19769 g075
Rakotondramanana and coworkers subjected Me-Fer (42a) to POC using Ag2O as oxidant in toluene-acetone (2:1) and obtained as the main product the known dihydrobenzofuran 116 in 45% yield, along with the corresponding β-Ο-4 dimer 144 in 5% yield. Compound 116 was protected with the TBDMS group and then selectively reduced to produce alcohol 145. This compound, upon esterification with the gallic acid derivative 146 and finally deprotection gave the dihydrobenzofuran-gallic acid hybrid 147 (Scheme 56) [81]. Biological evaluation of these compounds for potential antiatherogenic, antiplasmodial and cytotoxic activities showed that dimer 144 presented the best antiatherogenic effect (antioxidant activity and cytoprotective effect) whereas dihydrobenzofuran 116 exhibited the best activity against murine P388 leukemia cells. On the other hand, hybrid 147 showed a very good antiplasmodial potency coupled also with an antioxidant one.
Zhang and coworkers performed radical coupling reactions between equimolar quantities of Et-Fer (42) and ConAl in the presence of HRP and H2O2-urea complex as oxidant in acetone-aqueous buffer pH 5.0. They isolated and characterized a series of cross-coupled products, two of which (148 in 6.4% yield and 149, as the corresponding acetate 149a, in 1.5% yield) (Figure 6) were NLs of the dihydrobenzofuran type, along with a series of homo-coupled products [82]. Homodimers of the same type, such as dehydrodiconniferyl alcohol 150 and diethyl dehydrodiferulate 45, were also identified in the reaction mixture.
Figure 6. Cross- and homo-coupled products from the cross-POC of Et-Fer with ConAl utilizing the HRP/H2O2-urea oxidant system.
Figure 6. Cross- and homo-coupled products from the cross-POC of Et-Fer with ConAl utilizing the HRP/H2O2-urea oxidant system.
Molecules 19 19769 g006
Subbaraju and coworkers synthesized the neolignan tiruneesiin (151) in racemic form using as key reaction the POC of Me-Fer (42a) with Ag2O in dry PhH-acetone, which produced dihydrobenzofuran 116 in 29% yield. Catalytic hydrogenation of the double bond, LAH-mediated reduction of the ester functions, followed by O-benzylation of the aromatic hydroxyl, bisacetylation of the aliphatic hydroxyl groups and finally catalytic hydrogenolysis afforded (±)-151 in 12.6% total yield (Scheme 57) [83]. It should be noted that tiruneesiin, identified in Justicia neesii Ramamoorthy as the (−)-151 enantiomer, belongs to the group of dihydrobenzofuranoid NLs, which are known to be inhibitors of tubulin polymerization [77].
Scheme 57. Total synthesis of (±)-tiruneesiin (151) utilizing Ag2O-mediated POC of Me-Fer (42a) as key step.
Scheme 57. Total synthesis of (±)-tiruneesiin (151) utilizing Ag2O-mediated POC of Me-Fer (42a) as key step.
Molecules 19 19769 g076
Antus and coworkers have shown that Ag2O-mediated POC of Et-Caf (152) leads to the dihydrobenzofuran 153 as the main product isolated as its corresponding triacetate (153a), which was used for the synthesis of the NL Americanin-D (154) [84]. When a mixture of 152 and ConAl (23) was treated with either silver oxide or silver carbonate in benzene-acetone, the cross-coupled product 155 was isolated in 6%–8% yield, along with other dimers [85]. Compound 155 was used by the same research group as starting material for the synthesis of racemic cedrusin (156) (Scheme 58) [86].
Scheme 58. Total synthesis of (±)-americanin-D (154) and -cedrusin (156) utilizing Ag2O-mediated POC of either Et-Caf (152) alone or in mixture with ConAl (23), respectively, as key step.
Scheme 58. Total synthesis of (±)-americanin-D (154) and -cedrusin (156) utilizing Ag2O-mediated POC of either Et-Caf (152) alone or in mixture with ConAl (23), respectively, as key step.
Molecules 19 19769 g077
On the other hand, the same research group oxidatively dimerized Me-Fer (42a) with Ag2O in benzene-acetone to obtain dihydrobenzofuran 116 in 38% yield. This compound was employed as starting material for the synthesis of racemic balanophonin (157) and its corresponding methyl ester 158 [87] as well as the monomethyl ether 159 of cedrusin [86] (Scheme 59). It should be noted that the NL (−)-balanophonin and its corresponding methyl ester were isolated from Balanophora japonica Mikino and Ziziphus jujuba Mill, respectively, and showed significant PGI2 inducing effect, whereas (+)-cedrusin and its monomethyl ether were isolated from Cedrus deodara and Euconomia ulmoides Oliv., respectively. The latter plant has been applied in China as hypotensive drug.
Scheme 59. Total synthesis of NLs 157–159 utilizing Ag2O-mediated POC of Me-Fer (42) as key step.
Scheme 59. Total synthesis of NLs 157–159 utilizing Ag2O-mediated POC of Me-Fer (42) as key step.
Molecules 19 19769 g078
Daquino and coworkers subjected CAPE (75) into POC using Ag2O as oxidant in CH2Cl2 and obtained the corresponding dihydrobenzofuran 160 in 58% yield (Scheme 60) [59] when a ratio of CAPE:Ag2O = 1:2 was used. Interestingly, when POC was repeated in the presence of Mn(ACAC)2 no dihydrobenzofuran was formed, the main product being the benzoxanthene lignan 76 in 54% yield.
Bruschi and coworkers oxidized Evans’ 2-oxazolidinone amides (161164) of FerA utilizing either HRP-H2O2 in dioxane or acetone/aqueous buffer pH 3.5 or Ag2O in CH2Cl2 as oxidants and obtained the corresponding diastereomeric dihydrobenzofurans 165172. These compounds were further reduced with LiBH4 to the corresponding known enantiomeric dihydrodiconiferyl alcohols 173 and 174 (Scheme 61). The major enantiomer was identified in each case and the e.e. was calculated. Best enantioselectivities (e.e. 59%–62%) were secured with HRP-H2O2 in acetone-aqueous buffer as the oxidant system and (R)- and (S)-5-phenyl-2-oxazolidinones as the chiral auxiliaries [88].
Scheme 60. POC of CAPE (75) using Ag2O alone or Ag2O in the presence of Mn(ACAC)2 as oxidant.
Scheme 60. POC of CAPE (75) using Ag2O alone or Ag2O in the presence of Mn(ACAC)2 as oxidant.
Molecules 19 19769 g079
Scheme 61. Stereoselective POC of amides 161165 of FerA with 2-oxazilidinones using either Ag2O or HRP-H2O2 as oxidant.
Scheme 61. Stereoselective POC of amides 161165 of FerA with 2-oxazilidinones using either Ag2O or HRP-H2O2 as oxidant.
Molecules 19 19769 g080
Snyder and Kontes subjected Me-Fer (42a) into POC utilizing silver acetate as the oxidant in toluene and obtained a complex mixture of products from which the dihydrobenzofuran 116 was isolated in 20% yield (Scheme 62) [56].
Scheme 62. POC of Me-Fer using AgOAc as oxidant.
Scheme 62. POC of Me-Fer using AgOAc as oxidant.
Molecules 19 19769 g081
Arrieta-Baez and Stark subjected pairs of HCAs in cross-POC utilizing H2O2 in combination with either APP or HRP, as the oxidant systems in a phosphate buffer (pH 6.1) and studied the formation of products from cross-coupling and dimerization reactions. They found that only the mixtures of CafA and FerA and of CafA and SinA produced cross-coupled products in addition to dimerization products. In combinations of CouA and any of the other HCAs, dimerization products were only observed from the other HCAs but not CouA. Also, from the mixture FerA and SinA dimerization products were only observed from SinA. Both enzymes produced identical major products. The various major products formed were isolated by HPLC and identified. They had predominantly β-β γ-lactone (175176) and β-5 benzofuran molecular frameworks (177178) and one of them (179) was of the oxyneolignan 1,4-benzodioxane type (Figure 7) [89].
Figure 7. Major cross-coupled products from POC of pairs of HCAs using peroxidase-H2O2 as oxidant.
Figure 7. Major cross-coupled products from POC of pairs of HCAs using peroxidase-H2O2 as oxidant.
Molecules 19 19769 g007
Saliu and coworkers also performed cross-coupling reactions in pairs of equimolar methyl esters (60, 42a and 32) of the HCAs CafA, FerA and SinA and of the amides 180 and 181 of FerA and SinA, respectively) with the chiral auxiliary (R)-methylbenzylamine, utilizing the HRP-H2O2 oxidant system in dioxane-aqueous buffer pH 3.2 or MnO2 in CH2Cl2. The two oxidant systems gave, in general, comparable results. Cross-coupling products, along with homodimers, were essentially obtained in 12%–24% yields from the combinations of 32 with 60 and of 32 with 42a. They were predominantly of the benzofuran type (181182). In one case, a 1,4-benzodioxane product (183) was also formed in 9% yield. The combination of 60 and 42a produced as main product in 32% yield the homodimer benzofuran 116 from 42a. Homodimers obtained were either of the benzofuran (116) or the dihydronaphthalene (184) type (Figure 8) [90].
Figure 8. Major dimers obtained from POC of pairs of methyl esters of HCAs using peroxidase-H2O2 or MnO2 as oxidant.
Figure 8. Major dimers obtained from POC of pairs of methyl esters of HCAs using peroxidase-H2O2 or MnO2 as oxidant.
Molecules 19 19769 g008
On the other hand, the combination of amides 180 and 181 produced mainly a benzofuran type cross-coupled product 185 in 27% and two homodimers, one of the benzofuran type (186) and the other of the dihydronaphthalene type (187) in 6 and 15% yields, respectively (Scheme 63). Product 185 was a mixture of diastereomers with a d.e. evaluated to be 38%.
Scheme 63. Products from POC of chiral amides of HCAs using peroxidase-H2O2 as oxidant.
Scheme 63. Products from POC of chiral amides of HCAs using peroxidase-H2O2 as oxidant.
Molecules 19 19769 g082
Figure 9. The structure of schizotenuins.
Figure 9. The structure of schizotenuins.
Molecules 19 19769 g009
Interestingly, conjugation of NLs of the benzofuran type with other natural products can lead to hybrids with interesting biological properties, such as the dihydrobenzofuran-gallic acid hybrid 147 (Scheme 56) mentioned above. Other worth mentioning such hybrids are the so called schizotenuins A (188), C1 (189) and C2 (190) (Figure 9) which were isolated by Matsuta and coworkers from the terrestrial part of Schizonepeta tenuifolia BRIQ. This plant has been used in traditional Chinese medicines as anti-inflammatory crude drug. Hybrids 188190 are actually the diester and the two possible monoesters of the benzofuran 192 (CafA dimer) with DplA and act by inhibiting the 3α-hydroxysteroid dehydrogenase. Compound 192 was prepared from the fully protected dihydrobenzofuran 191, readily available using Ag2O-mediated POC of Me-Fer (42a) as key reaction, by DDQ-mediated oxidation, followed by deprotection steps (Scheme 64) [73].
Scheme 64. Synthesis of the benzofuran core (192) of schizotenuins.
Scheme 64. Synthesis of the benzofuran core (192) of schizotenuins.
Molecules 19 19769 g083
A dihydrobenzofuran unit constitutes the central substructure of the bicyclic Spm alkaloid (S,S,S)-aphelandrine (194) which is present in Aphelandra plants (Acanthaceae). Nezbedovà and coworkers has obtained this alkaloid in preparative yield through a stereoselective intramolecular POC of the Spm-CouA hybrid (S)-dihydroxyverbacine (193) in the presence of the soluble protein fraction of barley seedlings (Hordeum vulgare, Gramineae) as catalyst and O2 as oxidant (Scheme 65) [91]. The PA alkaloid (+)-hordatine A (196), isolated from barley (Hordeum vulgare L.), has a similar benzofuranoid substructure. It had been previously synthesized enzymatically from coumaroylagmatine (195) using cell free extracts from the shoots of barley seedlings [92].
Another interesting, naturally occurring, dihydrobenzofuranoid hybrid is (+)-LitA (197) (Figure 10), in particular because of its potent and non-toxic anti-HIV activity that results from inhibition of HIV-1 integrase [93]. Several total syntheses of 197 have been already reported [93,94,95,96,97]. This compound can be considered as a hybrid of the (2S,3S)-dimer 201 of CafA with (R)-DplA. Dimer 201 could be assembled in nature through the unusual coupling (β-2) of the C-centered radicals 198 and 199, followed by the usual Michael addition on the p-quinonemethide intermediate 200. Alternatively, LitA might be considered as the product (hybrid) of cross-POC between RosA and CafA, involving the same type of C-C bond formation as key step for the assembly of the molecular skeleton.
Scheme 65. The PA alkaloids 194 and 196, incorporating dihydrobenzofuranoid substructures, arising from enzyme-catalyzed POC of CouA moieties. The PA moieties are drawn in blue.
Scheme 65. The PA alkaloids 194 and 196, incorporating dihydrobenzofuranoid substructures, arising from enzyme-catalyzed POC of CouA moieties. The PA moieties are drawn in blue.
Molecules 19 19769 g084
Figure 10. The structure of (+)-LitA (197) and outline of plausible mechanism for the formation of core dihydrobenzofuranoid structure by either POC of CafA or cross-POC of CafA with RosA.
Figure 10. The structure of (+)-LitA (197) and outline of plausible mechanism for the formation of core dihydrobenzofuranoid structure by either POC of CafA or cross-POC of CafA with RosA.
Molecules 19 19769 g010
RosA and LitA along with dimers and trimers of the former, in which the monomeric units are joined through dihydrobenzofuran moieties, were isolated and identified by Ly and coworkers from the methanol exctract of dried leaves of Celastrus hindsii Benth a species of the Celastraceae family, which is used in Vietnam as traditional medicine for treating ulcers, tumors and inflammation. These compounds showed antioxidative activity whose effectiveness could be correlated to the number of the phenolic hydroxyl groups in each molecule [98]. One of the dimers, namely LitA-B (202) may be considered as a hybrid of neolignan 201 with two molecules of DplA (Figure 11). It is thought of being formed in nature through POC coupling of two molecules of RosA. LitA-B is also a potent and nontoxic inhibitor of HIV-1 replication. The pharmaceutically interesting acids RosA, LitA and LitA-B have been also found in some other plants [98].
Figure 11. The structure of LitA-B (202) and outline of the possible mechanism of its formation by oxidative cyclization through POC between two molecules of RosA.
Figure 11. The structure of LitA-B (202) and outline of the possible mechanism of its formation by oxidative cyclization through POC between two molecules of RosA.
Molecules 19 19769 g011
Figure 12. The structure of thorielamide A (203).
Figure 12. The structure of thorielamide A (203).
Molecules 19 19769 g012
Another example of a dihydrobenzofuranoid hybrid is thoreliamide A (203) (Figure 12) which has been isolated in optically inactive form from the stems of the plant Mitrephora thorelii (Annonaceae) by Ge and coworkers [58]. This natural product may be considered as the hybrid of the unusual dihydrobenzuran 204 with two molecules of tyramine. Compound 204 might arise from cross-POC of FerA and SinA or HFeA with further functional group elaboration on the aromatic SinA/HFeA side of the molecule.

3.2. Synthesis of Oxyneolignans of the Ether Type—Formation of β-O-4 Bond as the Key Step

As we noted above, Carunchio and coworkers subjected FerA to POC using the enzyme laccase and identified two main products initially formed, a benzofuranoid dimer 141 and the β-O-4 coupling product 142 (Scheme 54) [78]. The latter product is an example of the NL3 subtype of NLs (Figure 13). Οn the other hand, Neudorffer and coworkers isolated the oxyneolignan ether 46 (Scheme 28 and Figure 13) in 10% yield, among other lignans, from the reaction mixture of the electrochemical POC of Et-Fer [43]. Furthermore, Rakotondramanana and coworkers isolated the corresponding dimethyl ester 144 (Scheme 56 and Figure 13) in very low yield, along with the dihydrobenzofuranoid FerA dimer 116, from the Ag2O-mediated POC of Me-Fer [81].
Figure 13. The structure of FerA and Et-Fer dimers 142, 46 and 144 respectively, as examples of the NL3 subtype.
Figure 13. The structure of FerA and Et-Fer dimers 142, 46 and 144 respectively, as examples of the NL3 subtype.
Molecules 19 19769 g013
Cross (e.g., 205a,b)- and homo (206a,b and 46)-coupling products (Figure 14) of the β-Ο-4 ether type from the cross-POC of Et-Fer and ConAl (23) were isolated and/or identified, among other lignans, by Zhang and coworkers [82]. Compounds 205 and 206 are obviously formed through interception of the p-benzoquinonoid intermediates 207 by water.
Lu and coworkers subjected Et-Fer in POC utilizing the complex CuCl(OH)-TMEDA as catalyst and oxygen as oxidant in acetonitrile and obtained a mixture composed of β-O-4, 8-5, 8-8 (cyclic and noncyclic) and 5-5 coupled dehydrodiferulates, from which the β-O-4 dimer 46 was obtained through FCC along with the β-5 (dihydrobenzofuranoid) dimer 45 (the structure of dimers 45 and 46 can be seen in Scheme 28). Crystallization of the mixture resulted to pure crystalline 45 and left 46 in mother liquor. From the latter, pure dimer 46 was obtained through simple treatment of mother liquor with TBAF, which converted the remaining dimer 45 to a noncyclic β-5 dimer, and finally rechromatography [99].
Figure 14. Cross (205)- and homo (206)-coupled products of the β-Ο-4 ether type from the cross-POC of Et-Fer with ConAl and outline of mechanism of their formation.
Figure 14. Cross (205)- and homo (206)-coupled products of the β-Ο-4 ether type from the cross-POC of Et-Fer with ConAl and outline of mechanism of their formation.
Molecules 19 19769 g014
Interestingly, several oxyneolignan hybrids with interesting biological activities have been isolated from plants. For example, Agata and coworkers isolated from the overground part of Melissa officinalis L., a labiate plant used as folkmedicine in Europe for the treatment of chronic bronchial catarrh, feverish cold, headaches and tension, in addition to RosA two new compounds consisted of three CafA units which they named MelA-A (207) and -B (208) (Figure 15) [100]. The former compound may be thought of deriving from RosA and CafA through cross-POC with a β-Ο-4 bond connecting the two molecules. Althernatively, compound 207 may be considered as a hybrid of β-O-4 dimeric CafA and DplA. MelA-B is a product of an intramolecular dehydration of MelA-A.
Lu and Foo isolated from Salvia officinalis (Sage), a popular herb which has been used since ancient times for its health giving properties and for treating all kinds of ailments, in addition to RosA two compounds, namely the known SalA-K (209) and a novel cyclobutane derivative named sagerinic acid (210) (Figure 15) [101]. The former might be formed by cross-POC of RosA with CafA during which the primary β-Ο-4 bond is formed, followed by a stereoselective interception of the derived p-benzoquinone intermediate by water. In the latter, the two RosA units may be first connected by a Cβ-Cβ bond through POC, probably followed by an intramolecular Michael addition to create the second C-C bond with ring closure and finally reduction of the intermediate o-benzoquinone formed. From the same plant, Lu and coworkers isolated another novel CafA trimer, which they named sagecoumarin (211) (Figure 15), and methyl melitrate A (212) and proposed that MelA-A, its ester 212 and 211 are all synthesized from SalA-K [102].
Another SalA, namely SalA-B (213) (Figure 15) has been isolated from Salvia miltiorrhiza Bge, a well-known Chinese medicine for treating and preventing aging diseases for throusands of years, and has been shown to present interesting cardiovascular effects [103] and to be a potential chemoprotective agent for head and neck squamous cell cancer [104]. SalA-B is structurally related to LitA.
Figure 15. The structures of melitric and salvianolic acids as well as of sagerinic acid and sagecoumarin.
Figure 15. The structures of melitric and salvianolic acids as well as of sagerinic acid and sagecoumarin.
Molecules 19 19769 g015
Figure 16. The structures of SalA-Z and SalA-S and of the methyl ester 214 of the former.
Figure 16. The structures of SalA-Z and SalA-S and of the methyl ester 214 of the former.
Molecules 19 19769 g016
Exarchou and coworkers isolated from Origanum dictamnus L. (Lamiaceae), an endemic plant of Crete which was considered by ancient Greeks as Panacea and used in many cases for its healing effects, in addition to a new dihydronaphthalene hybrid SalA-R (73, Figure 2), RosA and MelA-A three new hybrids 213215 of dimeric CafA, with an unusual connection (β-Ο-3), and DplA (Figure 16). Interestingly, SalA-S (215) was the most potent antioxidant, followed by RosA, in the DPPH assay used [57].

3.3. Synthesis of Oxyneolignans of the 1,4-Benzodioxane Type—Formation of β-O-4 Bond as the Key Step

As we noted above, Antus and coworkers oxidatively dimerized Et-Caf (152) using Ag2O as oxidant and then acetylated the crude reaction product to obtain pure triacetate 153a in 29% yield following crystallization (Scheme 58). From the mother liquor, the authors isolated through silica gel CC a mixture of the isomeric oxyneolignans 216 and 217 (Figure 17) of the 1,4-benzodioxane NL2 type in 6% yield [84].
Figure 17. Structure of oxyneolignans 216 and 217 and outline of the mechanism of formation of regioisomer 216. Isomer 217 might be formed from the alternative phenoxy radical II.
Figure 17. Structure of oxyneolignans 216 and 217 and outline of the mechanism of formation of regioisomer 216. Isomer 217 might be formed from the alternative phenoxy radical II.
Molecules 19 19769 g017
The same research group subjected a mixture of Et-Caf (152) and ConAl (23) (molar ratio = 1:1.1) in cross-POC using silver oxide or silver carbonate as oxidant in benzene/acetone (2:1) and obtained a mixture from which the various products were isolated by preparative TLC. Major product of the coupling reaction was the 3-aryl-1,4-benzodioxane derivative 218a, which crystallized out from a mixture with the alternative regioisomer 219a. Both had the trans relative stereochemistry. Their corresponding cis isomers 218b and 219b were also obtained in much lower yields (Scheme 66). Silver carbonate resulted in higher regioselectivity (218a:219a = 25:1) than that obtained with silver oxide (218a:219a = 19:1). On the other hand, the use of hexacyanoferrate(III)/sodium carbonate as oxidizing agent in acetone/water (ca. 1:3) reversed the regioselectivity of the reaction and produced 219a as the major isomer. The cis isomers were also isolated as crystalline compounds in lower yields.
The 218 (a + b):219(a + b) ratio was ca. 1:9 [85].
Scheme 66. Cross-POC between Et-Caf and ConAl (23).
Scheme 66. Cross-POC between Et-Caf and ConAl (23).
Molecules 19 19769 g085
POC of CafA with O2 in an alkaline (KOH) aqueous solution produced, in addition to caffeicins E and F (see sections 2.2 and 3.1 above), dimeric CafA compounds of the 1,4-benzodioxane type which were named caffeicin A–D. The two substituents of the heterocyclic ring had the cis relative configuration (see e.g., compounds 220 and 221) (Figure 18) based on 1H-NMR coupling constants considerations [45]. The cis formed is preferred since the bulky groups are both in the equatorial position with the dioxane ring adopting the boat form.
As mentioned above, when a mixture of CafA and SinA was subjected in cross-POC utilizing H2O2 in combination with either APP or HRP, an oxyneolignan (179) of the 1,4-benzodioxane type was obtained (Figure 7 and Figure 18) in modest yield [89].
The reported coupling constant (J) for protons 7 and 8 was 2.5 Hz. Therefore, the relative stereochemistry in compound 179 should be cis (Figure 18). On the other hand, cross-POC of a pair of Me-Caf and Me-Sin utilizing the HRP-H2O2 oxidant system produced, among other lignans, the 1,4-benzodioxane product 183 in 9% yield (Figure 8) [90]. Based on coupling constant arguments (JH7–H8 = 2.3 Hz), this compound should also have the cis relative stereochemistry (Figure 18).
An interesting optically inactive 1,4-benzodioxane hybrid 222 (Figure 18) was isolated, along with other lignanamides (see Section 2.2 and Section 3.1) from the stems of Mitrephora thorelii [58]. This compound, named thoreliamide B, can be considered as the hybrid of acid 223 and tyramine. The former may then obviously be a cross-POC product of SinAl with CafA. Based on coupling constant arguments (JH7–H8 = 8.0 Hz), this racemic compound should have the trans relative stereochemistry.
Figure 18. Structures of oxyneolignans of the 1,4-benzodioxane type.
Figure 18. Structures of oxyneolignans of the 1,4-benzodioxane type.
Molecules 19 19769 g018

3.4. Synthesis of Other NLs—Formation of β-1 or 5-5 Bond as the Key Step

Setälä and coworkers performed a cross-POC reaction with an equimolar mixture of Me-Sin (32) and alcohol 224, a syringyl lignin model compound, utilizing HRP-H2O2 as oxidant system in acetone-aqueous buffer solution pH 3.5. Acetylation of the crude reaction mixture, followed first by silica gel CC and then by preparative HPLC separation, lead to a ca. 50% recovery of alcohol 224 and the isolation of two products, namely the spiro cross-coupling product 225 in 19% yield and the sinapate dimer 226 in ca. 4% yield (Scheme 67) [105]. Treatment of spiro compound 225 with a catalytic quantity of pTSA monohydrate in methanol gave, with loss of the side chain, a single diastereomer of compound 227, which was isolated in 33% yield as the corresponding peracetate 228.
As we noted above, Lu and coworkers subjected Et-Fer in POC utilizing the complex CuCl(OH)-TMEDA as catalyst and O2 as oxidant in acetonitrile and obtained a mixture composed of β-O-4, 8-5, 8-8 (cyclic and noncyclic) and 5-5 coupled dehydrodiferulates. From this mixture, they isolated not only the β-Ο-4 and the 8-5 coupled dehydrodiferulates (see Scheme 28) but also the 5-5 coupled diferulate (229) (Figure 19) and the 8-8 coupled cyclic (43) and noncyclic (44) dehydrodiferulates (see Scheme 28) by flash chromatographic fractionation [99].
Scheme 67. Cross-POC between Me-Sin and alcohol 224.
Scheme 67. Cross-POC between Me-Sin and alcohol 224.
Molecules 19 19769 g086
Figure 19. The structure of 5-5 coupled dimethyl dehydrodiferulates (229).
Figure 19. The structure of 5-5 coupled dimethyl dehydrodiferulates (229).
Molecules 19 19769 g019

4. Conclusions

We have presented in this review an overview of past and recent applications of an old reaction, namely POC, in the bioinspired syntheses of a variety of CLs, and NLs with potential medicinal significance. These involve, in particular, lignans of the dilactone, dihydronaphthalene, bis(aryl-methylene)succinic acid, 1,4-benzodioxane, tetrahydrofuran, dihydrobenzofuran, aryl β-cinnamyl ether, and benzoxanthene subtypes. POC utilizes simple starting materials and experimental procedures and creates in one-pot experiments relatively complex structures often with good yields, which would otherwise require several steps to be assembled. Simple, commercially available or easily synthesized, building blocks, such as naturally occurring HCAs and HCAls and synthetic analogs as well as esters and amides of HCAs, are usually employed. Coupling between the same or different (cross-POC), in some cases, building blocks is effected by the action of one-electron inorganic oxidants or enzymes (peroxidases, usually HRP and less often laccases) as catalysts with H2O2 or O2 as oxidant, which mimic even closer the natural conditions for performing such dehydrodimerizations. Suitable inorganic oxidants have been also used in the so-called non-POC which uses ACAs as starting materials and has been used successfully, for example, in the synthesis of lignans of the isostegane type. Although POC usually produces mixtures of regio- and diastereoisomers and therefore chromatographic techniques are necessary for their separation, it is still a very attractive alternative to other multi-step methodologies, in particularly in cases where a single regioisomer and/or diastereoiosmer is formed as the main product, because of the simplicity in its performance and the fact that creates complex structures in one-pot reactions. The regioselectivity of POC has been improved by selectively blocking positions which could be potentially involved in bond-forming reactions through radical coupling. One such group is the tert-butyl group which is readily introduced and finally removed after the POC reaction. Interestingly, the primary products of POC can be readily transformed through simple reactions in a multitude of other medicinally significant lignans and hybrids. A serious drawback of POC is that it provides the dimers in racemic mixtures, even when enzymes are used as catalysts. However, this problem has been faced with considerable success by subjecting to POC chiral esters or amides of HCAs, readily produced by condensing HCAs with chiral alcohols or amines, e.g., esters of chiral α-amino acids, used as chiral auxilliaries. That way, the thus produced diastereomeric compounds are resolved with chromatographic techniques and, if necessary, the chiral auxilliary is removed, for example, by alkaline hydrolysis or LiOOH-mediated hydrolysis. It is reasonable to assume that POC will continue to be used for the ready assembly of the skeleta of naturally occurring and biologically interesting lignans of several types and that it will be further developed towards the direction of improving its regio- and the stereoselectivity by the thoughtful choice of combinations of substrates, oxidant systems, reaction conditions and chiral auxilliaries.

Abbreviation

ACAs
alkoxycinnamic acids
APP
anionic potato peroxidase
CafA
caffeic acid
CAN
ceric ammonium nitrate
CAPE
phenethyl caffeate
CC
column chromatography
CinA
cinnamic acid
CL
classical lignan
ConAl
coniferyl alcohol
CouA
p-coumaric acid
DBU
1,8-diazabicyclo[5.4.0]undec-7-ene
DCC
N,N'-dicyclohexylcarbodiimide
DDQ
2,3-dichloro-5,6-dicyano-p-benzoquinone
DIBAL
diisobutylaluminium hydride
DMAP
4-(dimethylamino)pyridine
DplA
3-(3,4-dihydroxyphenyl)lactic acid
DPPH
2,2-diphenyl-1-picrylhydrazyl
Et-Caf
ethyl caffeate
Et-Fer
ethyl ferulate
Et-Sin
ethyl sinapate
FCC
flash column chromatography
FerA
ferulic acid
Gly
glycine
HCAs
4-hydroxycinnamic acids
HCAls
4-hydroxycinnamyl alcohols
HFeA
5-hydroxyferulic acid
HRP
horse radish peroxidase
LAH
lithium aluminium hydride
LitA
lithospermic acid
Me-Caf
methyl caffeate
Me-Fer
methyl ferulate
MelA
melitric acid
Me-Sin
methyl sinapate
Mn(ACAC)2
manganese(II) acetylacetonate
Mn(III)TPPOAc
tetraphenylporphyrinatomanganese(III) acetate
NL
neolignan
non-POC
“non-phenolic” oxidative coupling
PA
polyamine
Phe
Phenylalanine
PIDA
phenyliodonium diacetate
PIFA
phenyliodonium bis(trifluoroacetate)
POC
phenol oxidative coupling
Pro
proline
pTSA
p-toluenesulfonic acid
RosA
rosmarinic acid
SalA
Salvianolic acid
SARS
structure-activity relationship studies
ShiA
shikimic acid
SinA
sinapic acid
SinAl
sinapyl alcohol
Spm
spermine
TBAF
tetrabutylammonium fluoride
TBDMS
tert-butyldimethylsilyl
ThoA
thomasidioic acid
TLC
thin layer chromatography
TMEDA
tetramethylethylenediamine
Tyr
tyrosine

Acknowledgments

We thank the Research Committee of the University of Patras (Greece) for funding the present work on the form of a postdoctoral fellowship (G.E.M.)

Author Contributions

George E. Magoulas performed the literature search, wrote part of the paper and cooperated in the preparation of the manuscript. Dionissios Papaioannou conceived the concept and wrote part of the review and prepared all figures and schemes outlining mechanistic considerations. Both authors read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Geissman, T.A.; Crout, D.H.G. Organic Chemistry of Secondary Plant Metabolism; Freeman, Cooper & Company: San Francisco, CA, USA, 1969; Chapter V; pp. 136–166, Chapter XIV, pp. 372–400. [Google Scholar]
  2. Teixera, J.; Gaspar, A.; Garrido, E.M.; Garrido, J.; Borges, F. Hydroxycinnamic acid antioxidants: An electrochemical overview. BioMed Res. Int. 2013. [Google Scholar] [CrossRef]
  3. De, P.; Baltas, M.; Bedos-Belval, F. Cinnamic acid derivatives as anticancer agents—A review. Curr. Med. Chem. 2011, 18, 1672–1703. [Google Scholar] [CrossRef] [PubMed]
  4. Lewis, N.G.; Davin, L.B. Lignans: Biosynthesis and function. In Comprehensive Natural Products Chemistry; Barton, Sir D.H.R., Nakanishi, K., Meth-Cohn, O., Eds.; Elsevier: Oxford, UK, 1999; Volume 1, pp. 639–712. [Google Scholar]
  5. Barton, D.H.R.; Cohen, T. Some biogenetic aspects of phenol oxidation. In Festschrift Arthur Stoll; Birkhauser: Basel, Switzerland, 1957; pp. 117–143. [Google Scholar]
  6. Barton, D.H.R.; Deflorin, A.M.; Edwards, O.E. The synthesis of Usnic acid. J. Chem. Soc. 1956, 530–534. [Google Scholar]
  7. Erdtman, H. Dehydrierungen in der coniferylreihe. II. Dehydrodi-isoeugenol. Annalen 1933, 503, 283–294. [Google Scholar] [CrossRef]
  8. Erdtman, H.; Wachtmeister, C.A. Phenoldehydrogenation as a biosynthetic reaction. In Festschrift Arthur Stoll; Birkhauser: Basel, Switzerland, 1957; pp. 144–165. [Google Scholar]
  9. Erdtman, H. Recent Advances in Phytochemistry; Aplleton-Century-Crofts: New York, NY, USA, 1968; Volume 1, p. 29. [Google Scholar]
  10. Gotlieb, O.R. Chemosystematics of the lauraceae. Phytochemistry 1972, 11, 1537–1570. [Google Scholar] [CrossRef]
  11. Brown, B.R. Biochemical Aspects of Oxidative Coupling of Phenols. In Oxidative Coupling of Phenols; Taylor, W.I., Battersby, A.R., Eds.; Edward Arnold Publishers: London, UK, 1967; pp. 167–201. [Google Scholar]
  12. Scott, A.I. Oxidative coupling of phenolic compounds. Q. Rev. (Chem. Soc.) 1965, 19, 1–35. [Google Scholar] [CrossRef]
  13. Lewis, N.G.; Sarkanen, S. (Eds.) Lignin and Lignan Biosynthesis; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 1998; Volume 697.
  14. Moss, G.P. Nomenclature of lignans and neolignans. Pure Appl. Chem. 2000, 72, 1493–1523. [Google Scholar] [CrossRef]
  15. Setälä, H. Regio- and Stereoselectivity of Oxidative Coupling Reactions of Phenols: Spirodienones as Construction Units in Lignin. Ph.D. Thesis, University of Helsinki, Helsinki, Finland, September 2008. [Google Scholar]
  16. Umezawa, T.; Okunishi, T.; Shimada, M. Stereochemical diversity of lignan biosynthesis. Wood Res.: Bull. Wood Res. Inst. Kyoto Univ. 1997, 84, 62–75. [Google Scholar]
  17. Deyama, T.; Nishibe, S. Pharmacological properties of lignans. In Lignins and Lignans: Advances in Chemistry; Heitner, C., Dimmel, D.R., Schmidt, J.A., Eds.; CRC Press; Taylor & Francis Group: Boca Raton, FL, USA, 2010; pp. 585–629. [Google Scholar]
  18. Ward, R.S. The synthesis of lignans and neolignans. Chem. Soc. Rev. 1982, 11, 75–125. [Google Scholar] [CrossRef]
  19. Ward, R.S. Different strategies for the chemical synthesis of lignans. Phytochem. Rev. 2003, 2, 391–400. [Google Scholar] [CrossRef]
  20. Pan, J.-Y.; Chen, S.-L.; Yang, M.-H.; Wu, J.; Sinkkonen, J.; Zou, K. An update on lignans: Natural products and synthesis. Nat. Prod. Rep. 2009, 26, 1251–1292. [Google Scholar] [CrossRef] [PubMed]
  21. Pal, T.; Pal, A. Oxidative phenol coupling: A key step for the biommimetic synthesis of many important natural products. Curr. Sci. 1996, 71, 106–108. [Google Scholar]
  22. Erdtman, H.; Phenoldehydrierung, V.I. Dehydrierende Kupplung einiger Guajakoi-derivate. Svensk. Kem. Tidskr. 1935, 47, 223–230. [Google Scholar]
  23. Cartwright, N.J.; Haworth, R.D. Constituents of natural phenolic resins. XIX. Oxidation of ferulic acid. J. Chem. Soc. 1944, 535–537. [Google Scholar]
  24. Takei, Y.; Mori, K.; Matsui, M. Synthesis of dl-matairesinol dimethyl ether, dehydrodimethyl conidendrin and dehydrodimethyl retrodendrin from ferulic acid. Agric. Biol. Chem. (Jpn.) 1973, 37, 637–641. [Google Scholar] [CrossRef]
  25. Pelter, A.; Ward, R.S.; Watson, D.J.; Collins, P.; Kay, I.T. Synthesis of 2,6-diaryl-4,8-dihydroxy-3,7-dioxabicyclo[3.3.0]octanes. J. Chem. Soc. Perkin Trans. 1 1982, 175–181. [Google Scholar]
  26. Freudenberg, K.; Schraube, H. Synthese des syringaresinols und versuche mit sinapinalkohol. Chem. Ber. 1955, 88, 16–23. [Google Scholar] [CrossRef]
  27. Ahmed, R.; Lehrer, M.; Stevenson, R. Synthesis of thomasidioic acid. Tetrahedron Lett. 1973, 14, 747–750. [Google Scholar] [CrossRef]
  28. Ahmed, R.; Lehrer, M.; Stevenson, R. Synthesis of thomasic acid. Tetrahedron 1973, 29, 3753–3759. [Google Scholar] [CrossRef]
  29. Wallis, A.F.A. Oxidation of (E)- and (Z)-2,6-Dimethoxy-4-propenylphenol with ferric chloride—A facile route to the 2-aryl ethers of 1-Arylpropan-1,2-diols. Aust. J. Chem. 1973, 26, 585–594. [Google Scholar] [CrossRef]
  30. Wallis, A.F.A. Oxidative dimerization of methyl (E)-sinapate. Aust. J. Chem. 1973, 26, 1571–1576. [Google Scholar] [CrossRef]
  31. Kumada, Y.; Naganawa, H.; Takeuchi, T.; Umezawa, H.; Yamashita, K.; Watanabe, K. Biochemical activities of the derivatives of dehydrodicaffeic acid dilactone. J. Antibiot. 1978, 31, 105–111. [Google Scholar] [CrossRef] [PubMed]
  32. Ahmed, R.; Schreiber, F.G.; Stevenson, R.; Williams, J.R.; Yeo, H.M. Oxidative coupling of bromo- and iodo-ferulic acid derivatives: synthesis of (±)-veraguensin. Tetrahedron 1976, 32, 1339–1344. [Google Scholar] [CrossRef]
  33. Stevenson, R.; Williams, J.R. Synthesis of tetrahydrofuran lignans, (±)-galbelgin and (±)-grandisin. Tetrahedron 1977, 33, 285–288. [Google Scholar] [CrossRef]
  34. Ralph, J.; Quideau, S.; Grabber, J.H.; Hatfield, R.D. Identification and synthesis of new ferulic acid dehydrodimers present in grass cell walls. J. Chem. Soc. Perkin Trans. 1 1994, 3485–3498. [Google Scholar]
  35. Kumada, Y.; Takeuchi, T.; Umezawa, H. Characterization of the dehydrodicaffeic acid dilactone-forming enzyme and the enzymic and chemical synthesis of this mushroom product. Agric. Biol. Chem. 1977, 41, 877–885. [Google Scholar] [CrossRef]
  36. Jin, X.L.; Yang, R.T.; Shang, Y.J.; Dai, F.; Qan, Y.P.; Cheng, L.X.; Zhou, B.; Liu, Z.L. Oxidative coupling of cinnamic acid derivatives and their radical-scavenging activities. Chin. Sci. Bull. 2010, 55, 2885–2890. [Google Scholar] [CrossRef]
  37. Iguchi, M.; Nishiyama, A.; Eto, H.; Terada, Y.; Yamamura, S. Anodic oxidation of 4-hydroxycinnamic acids. Chem. Lett. 1979, 1397–1400. [Google Scholar]
  38. Cooper, R.; Gottlieb, H.E.; Lavie, D.; Levy, E.C. Lignans from Aegilops ovata L. Synthesis of a 2,4- and 2,6-diaryl monoepoxylignanolide. Tetrahedron 1979, 35, 861–868. [Google Scholar] [CrossRef]
  39. Taylor, E.C.; Andrade, J.G.; Rall, G.J.H.; Steliou, K.; Jagdmann, G.E., Jr.; McKillop, A. Thallium in Organic Synthesis. 60. 2,6-diaryl-3,7-dioxabicyclo[3.3.0]octane-4,8-dione lignans by oxidative dimerization of 4-alkoxycinnamic acids with thallium(III) trifluroacetate or cobalt(III) trifluoride. J. Org. Chem. 1981, 46, 3078–3081. [Google Scholar] [CrossRef]
  40. Mori, N.; Watanabe, H.; Kitahara, T. Simple and efficient asymmetric synthesis of furofuran lignans yangambin and caruilignan A. Synthesis 2006, 400–404. [Google Scholar]
  41. Setälä, H.; Pajunen, A.; Kilpelläinen, I.; Brunow, G. Horse radish peroxidase-catalysed oxidative coupling of methyl sinapate to give diastereomeric spiro dimers. J. Chem. Soc. Perkin Trans. 1 1994, 1163–1165. [Google Scholar]
  42. Bunzel, M.; Ralph, J.; Kim, H.; Lu, F.; Ralph, S.A.; Marita, J.M.; Hatfield, R.D.; Steinhart, H. Sinapate dehydrodimers and sinapate-ferulate heterodimers in cereal dietary fiber. J. Agric. Food Chem. 2003, 51, 1427–1434. [Google Scholar] [CrossRef] [PubMed]
  43. Neudorffer, A.; Deguin, B.; el Ha, C.; Fleury, M.-B.; Largeron, M. Electrochemical oxidative coupling of 4-hydroxycinnamic ester derivatives: A convenient methodology for the biomimetic synthesis of lignin precursors. Collect. Czechoslov. Chem. Commun. 2003, 68, 1515–1530. [Google Scholar] [CrossRef]
  44. Zoia, L.; Bruschi, M.; Orlandi, M.; Tolppa, E.-L.; Rindone, B. Assymetric biomimetic oxidations of phenols: the mechanism of the diastereo- and enantioselective synthesis of thomasidioic acid. Molecules 2008, 13, 129–148. [Google Scholar] [CrossRef] [PubMed]
  45. Cilliers, J.J.L.; Singleton, V.L. Characterization of the products of non-enzymic autoxidative phenolic reactions in a caffeic acid model system. J. Agric. Food Chem. 1991, 39, 1298–1303. [Google Scholar] [CrossRef]
  46. Maeda, S.; Masuda, H.; Tokoroyama, T. Studies on the preparation of bioactive lignans by oxidative coupling reaction. IV. Oxidative coupling reaction of methyl (E)-3-(3,4-dihydroxy-2-methoxyphenyl)propenoate and lipid peroxidation inhibitory effects of the produced lignans. Chem. Pharm. Bull. 1995, 43, 84–90. [Google Scholar] [CrossRef] [PubMed]
  47. Maeda, S.; Masuda, H.; Tokoroyama, T. Studies on the preparation of bioactive lignans by oxidative coupling reaction. IΙ. Oxidative coupling reaction of methyl (E)-3-(4,5-dihydroxy-2-methoxyphenyl)propenoate and lipid peroxidation inhibitory effects of the produced lignans. Chem. Pharm. Bull. 1994, 42, 2506–2513. [Google Scholar] [CrossRef] [PubMed]
  48. Agata, I.; Hatano, T.; Nishibe, S.; Okuda, T. Rabdosiin, a new rosmarinic acid dimer with a lignan skeleton from Rabdosia japonica. Chem. Pharm. Bull. 1988, 36, 3223–3225. [Google Scholar] [CrossRef]
  49. Agata, I.; Hatano, T.; Nishibe, S.; Okuda, T. A tetrameric derivative of caffeic acid from Rabdosia japonica. Phytochemistry 1989, 28, 2447–2450. [Google Scholar] [CrossRef]
  50. Kashiwada, Y.; Nishizawa, M.; Yamagishi, T.; Tanaka, T.; Nonaka, G.; Cosentino, L.M.; Snider, J.V.; Lee, K. Anti-AIDS agents, 18. Sodium and potassium salts of caffeic acid tetramers from Arnebia euchroma as anti-HIV agents. J. Nat. Prod. 1995, 58, 392–400. [Google Scholar] [CrossRef] [PubMed]
  51. </b>Kashiwada, Y.; Bastow, K.F.; Lee, K. Novel lignan derivatives as selective inhibitors of DNA topoisomerase II. Bioorg. Med. Chem. Lett. 1995, 5, 905–908. [Google Scholar] [CrossRef]
  52. Bogucki, D.E.; Charlton, J.L. A non-enzymatic synthesis of (S)-(−)-rosmarinic acid and a study of a biomimetic route to (+)-rabdosiin. Can. J. Chem. 1997, 75, 1783–1794. [Google Scholar] [CrossRef]
  53. Reimann, E.; Pflug, T. Synthese der enantiomeren (+)- and (−)-rosmarinsäuremethylester. Monatsch. Chem. 1998, 129, 187–193. [Google Scholar]
  54. Huang, L.-J.; Li, C.-H.; Lu, Z.-M.; Ma, Z.-B.; Yu, D.-Q. Total synthesis and biological evaluation of (+)- and (−)-butyl ester of rosmarinic acid. J. Asian Nat. Prod. Res. 2006, 8, 561–566. [Google Scholar] [CrossRef] [PubMed]
  55. Aung, H.T.; Furukawa, T.; Nikai, T.; Niwa, M.; Takaya, Y. Contribution of cinnamic acid analoues in rosmarinic acid to inhibition of snake venom induced hemorrhage. Bioorg. Med. Chem. 2011, 19, 2392–2396. [Google Scholar] [CrossRef] [PubMed]
  56. Snyder, S.A.; Kontes, F. Explorations into neolignan biosynthesis: Concise total syntheses of Helicterin B, Helisorin, and Helisterculin A from a common intermediate. J. Am. Chem. Soc. 2009, 131, 1745–1752. [Google Scholar] [CrossRef] [PubMed]
  57. Exarchou, V.; Takis, P.G.; Malouta, M.; Vervoort, J.; Karali, E.; Troganis, A.N. Four new depsides in Origanum dictamnus methanol extract. Phytochem. Lett. 2013, 6, 46–52. [Google Scholar] [CrossRef]
  58. Ge, F.; Tang, C.-P.; Ye, Y. Lignanamides and sesquiterpenoids from stems of Mitrephora thorelii. Helv. Chim. Acta 2008, 91, 1023–1030. [Google Scholar] [CrossRef]
  59. Daquino, C.; Rescifina, A.; Spatafora, C.; Tringali, C. Biomimetic synthesis of natural and “unnatural” lignans by oxidative coupling of caffeic esters. Eur. J. Org. Chem. 2009, 2009, 6289–6300. [Google Scholar] [CrossRef]
  60. Lin, H.-P.; Lin, C.-Y.; Huo, C.; Su, L.-C.; Chuu, C.-P. Anticancer effect of caffeic acid phenethyl ester. Pharmacologia 2012, 3, 26–30. [Google Scholar] [CrossRef]
  61. Maddaford, S.P.; Charlton, J.L. A general asymmetric synthesis of (−)-α-dimethylretrodendrin and its diastereomers. J. Org. Chem. 1993, 58, 4132–4138. [Google Scholar] [CrossRef]
  62. Broomhead, A.J.; Rahman, M.M.A.; Dewick, P.M.; Jackson, D.E.; Lucas, J.A. Matairesinol as precursor of Podophyllum lignans. Phytochemistry 1991, 30, 1489–1492. [Google Scholar] [CrossRef]
  63. Cambie, R.C.; Craw, P.A.; Rutledge, P.S.; Woodgate, P.D. Oxidative coupling of lignans III. Non-phenolic oxidative coupling of deoxypodorhizon and related compounds. Aust. J. Chem. 1988, 41, 897–918. [Google Scholar] [CrossRef]
  64. Sarkanen, K.V.; Wallis, A.F.A. Oxidative dimerization of methyl (E)-4-hydroxy-3,5-di-t-butylcinnamate with potassium ferricyanide. J. Chem. Soc. Perkin Trans. 1 1973, 1878–1881. [Google Scholar]
  65. Wang, Q.; Yang, Y.; Li, Y.; Yu, W.; Hou, Z.H. An efficient method for the synthesis of lignans. Tetrahedron 2006, 62, 6107–6112. [Google Scholar] [CrossRef]
  66. Sugahara, T.; Yamauchi, S.; Kondo, A.; Ohno, F.; Tominaga, S.; Nakashima, Y.; Kishida, T.; Akiyama, K.; Maruyama, M. First stereoselective synthesis of meso-secoisolariciresinol and comparison of its biological activity with (+) and (−) secoisolariciresinol. Biosci. Biotechnol. Biochem. 2007, 71, 2962–2968. [Google Scholar] [CrossRef] [PubMed]
  67. Moon, S.-S.; Rahman, A.A.; Kim, J.-Y.; Kee, S.-H. Hanultarin, a cytotoxic lignan as an inhibitor of actin cytoskeleton polymerization from the seeds of Trichosanthes kirilowii. Bioorg. Med. Chem. 2008, 16, 7264–7269. [Google Scholar] [CrossRef] [PubMed]
  68. Lee, E.; Ahamed, V.S.J.; Kumar, M.S.; Rhee, S.W.; Moon, S.-S.; Hong, I.S. Synthesis and evaluation of cytotoxic effect of hanultarin and its derivatives. Bioorg. Med. Chem. Lett. 2011, 21, 6245–6248. [Google Scholar] [CrossRef] [PubMed]
  69. Ma, C.-Y.; Liu, W.K.; Che, C.-T. Lignanamides and nonalkaloidal components of Hyoscyamus niger seeds. J. Nat. Prod. 2002, 65, 206–209. [Google Scholar] [CrossRef] [PubMed]
  70. Τomosaka, H.; Chin, Y.-W.; Salim, A.A.; Keller, W.J.; Chai, H.; Kingborn, A.D. Antioxidant and cytoprotective compounds from Berberis vulgaris (Barberry). Phytother. Res. 2008, 22, 979–981. [Google Scholar] [CrossRef] [PubMed]
  71. Li, D.; Li, W.; Wang, Q.; Yang, Z.; Hou, Z. Consize synthesis of cannabisin G. Bioorg. Med. Chem. Lett. 2010, 20, 5095–5098. [Google Scholar] [CrossRef] [PubMed]
  72. Chioccara, F.; Poli, S.; Rindone, B.; Pilati, T.; Brunow, G.; Pietikäinen, P.; Setälä, H. Regio- and Diastereo-selective synthesis of dimeric lignans using oxidative coupling. Acta Chem. Scand. 1993, 47, 610–616. [Google Scholar] [CrossRef]
  73. Maeda, S.; Masuda, H.; Tokoroyama, T. Studies on the preparation of bioactive lignans by oxidative coupling reaction. I. Preparation and lipid peroxidation inhibitory effect of benzofuran lignans related to schizotenuins. Chem. Pharm. Bull. 1994, 42, 2500–2505. [Google Scholar] [CrossRef] [PubMed]
  74. Maeda, S.; Masuda, H.; Tokoroyama, T. Studies on the preparation of bioactive lignans by oxidative coupling reaction. III. Synthesis of polyphenolic benzofuran and coumestan derivatives by oxidative coupling reaction of methyl (E)-3-(4-hydroxy-2-methoxyphenyl)propenoate and their inhibitory effect on lipid peroxidation. Chem. Pharm. Bull. 1994, 42, 2536–2545. [Google Scholar] [CrossRef] [PubMed]
  75. Bolzacchini, E.; Brunow, G.; Meinardi, S.; Orlandi, M.; Rindone, B.; Rummakko, P.; Setala, H. Enantioselective synthesis of a benzofuranic neolignan by oxidative coupling. Tetrahedron Lett. 1998, 39, 3291–3294. [Google Scholar] [CrossRef]
  76. Ralph, J.; Garcia Conesa, M.T.; Williamson, G. Simple preparation of 8–5 coupled diferulate. J. Agric. Food Chem. 1998, 46, 2531–2532. [Google Scholar] [CrossRef]
  77. Pieters, L.; van Dyck, S.; Gao, M.; Bai, R.; Hamel, E.; Vlietinck, A.; Lemière, G. Synthesis and biological evaluation of dihydrobenzofuran lignans and related compounds as potentia antitumor agents that inhibit tubulin polymerization. J. Med. Chem. 1999, 42, 5475–5481. [Google Scholar] [CrossRef] [PubMed]
  78. Carunchio, F.; Crescenzi, C.; Girelli, A.M.; Messina, A.; Tarola, A.M. Oxidation of ferulic acid by laccase: Identification of the products and inhibitory effect of some dipeptides. Talanta 2001, 55, 189–200. [Google Scholar] [CrossRef] [PubMed]
  79. Kuo, Y.-H.; Wu, C.-H. Synthesis of 5-(3-hydroxypropyl)-7-methoxy-2-(3'-methoxy-4'-hydroxyphenyl)-3-benzo[b]furancarbaldehyde, a novel adenosine A1 receptor ligand from the root of Salvia miltiorrhiza. J. Nat. Prod. 1996, 59, 625–628. [Google Scholar] [CrossRef]
  80. Chang, J.Y.; Chang, C.-Y.; Kuo, C.-C.; Chen, L.-T.; Wein, Y.-S.; Kuo, Y.-H. Salvinal, a novel microtubule inhibitor isolated from Salvia miltiorrhizae Bunge (Danshen), with antimitotic activity in multi-drug-sensitive and-resistant human tumor cells. Mol. Pharmacol. 2004, 65, 77–84. [Google Scholar] [CrossRef] [PubMed]
  81. Rakotondramanana, D.L.A.; Delomenede, M.; Baltas, M.; Duran, H.; Bedos-Belval, F.; Rasoanaivo, P.; Negre-Salvayre, A.; Gornitzka, H. Synthesis of ferulic ester dimers, functionalization and biological evaluation as potential antiatherogenic and antiplasmodial agents. Bioorg. Med. Chem. 2007, 15, 6018–6026. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, A.; Lu, F.; Sun, R.; Ralph, J. Ferulate-coniferyl alcohol cross-coupled products formed by radical coupling reactions. Planta 2009, 229, 1099–1108. [Google Scholar] [CrossRef] [PubMed]
  83. Subbaraju, G.; Kavitha, J.; Rajasekhar, D.; Hsu, F.-L.; Cheng, K.-T. Justicia lignans: Part 10—Synthesis of tiruneesiin, the first neolignan from Justicia species. Ind. J. Chem. 2007, 46B, 357–359. [Google Scholar]
  84. Antus, S.; Bauer, R.; Gottsegen, A.; Seligmann, O.; Wagner, H. Synthese von Americanin-D. Liebigs Ann. Chem. 1987, 1987, 357–360. [Google Scholar] [CrossRef]
  85. Antus, S.; Baitz-Gàcs, E.; Bauer, R.; Gottsegen, A.; Seligmann, O.; Wagner, H. Regioselective synthesis of 2- and 3-aryl-1,4-benzodioxanes. Liebigs Ann. Chem. 1989, 1989, 1147–1151. [Google Scholar] [CrossRef]
  86. Antus, S.; Baitz-Gàcs, E.; Bauer, R.; Gottsegen, A.; Seligmann, O.; Wagner, H. Total synthesis of cedrusin and its methyl ether. Liebigs Ann. Chem. 1990, 1990, 495–497. [Google Scholar] [CrossRef]
  87. Antus, S.; Gottsegen, A.; Kolonits, P.; Wagner, H. Total synthesis of two naturally occurring neolignans of potential biological activity. Liebigs Ann. Chem. 1989, 1989, 593–594. [Google Scholar] [CrossRef]
  88. Bruschi, M.; Orlandi, M.; Rindone, B.; Rummakko, P.; Zoia, L. Asymmetric biomimetic oxidations of phenols using oxazolidines as chiral auxiliaries: the enantioselective synthesis of (+)- and (−)-dehydrodiconiferyl alcohol. J. Phys. Org. Chem. 2006, 19, 592–596. [Google Scholar] [CrossRef]
  89. Arrieta-Baez, D.; Stark, R.E. Modeling suberization with peroxidase-catalyzed polymerization of hydroxycinnamic acids: Cross-coupling and dimerization reactions. Phytochemistry 2006, 67, 743–753. [Google Scholar] [CrossRef] [PubMed]
  90. Saliu, F.; Tolppa, E.-L.; Zoia, L.; Orlandi, M. Horseradish peroxidase catalyzed oxidative cross-coupling reactions: The synthesis of “unnatural” dihydrobenzofuran lignans. Tetrahedron Lett. 2011, 52, 3856–3860. [Google Scholar] [CrossRef]
  91. Nezbedovà, L.; Hesse, M.; Drandarov, K.; Werner, C. New reagent for oxidative phenol coupling. The transformation of the monocyclic spermine base (S)-dihydroverbacine to the bicyclic alkaloid (S,S,S)-aphelandrine by cell free extract of barley seedlings. Tetrahedron Lett. 2001, 42, 4139–4141. [Google Scholar] [CrossRef]
  92. Bird, C.R.; Smith, T.A. The biosynthesis of coumarylagmantine in barley seedings. Phytochemistry 1981, 20, 2345–2346. [Google Scholar] [CrossRef]
  93. For a very informative short introduction on the isolation, characterization, biological properties and synthesis of (+)-lithospermic acid see: Varadaraju, T.G.; Hwu, J.R. Synthesis of anti-HIV lithospermic acid by two divergent strategies. Org. Biomol. Chem. 2012, 10, 5456–5465. [Google Scholar]
  94. Jacobson, R.M.; Raths, R.A. Total synthesis of heptamethyl lithospermate. J. Org. Chem. 1979, 44, 4013–4014. [Google Scholar] [CrossRef]
  95. O’Malley, S.J.; Tan, K.L.; Watzke, A.; Bergman, R.G.; Ellman, J.A. Total synthesis of (+)-lithospermic acid by asymmetric intramolecular alkylation via catalytic C-H bond activation. J. Am. Chem. Soc. 2005, 127, 13496–13497. [Google Scholar] [CrossRef] [PubMed]
  96. Wang, D.-H.; Yu, J.-Q. Highly convergent total synthesis of (+)-lithospermic acid via a late-stage intermolecular C-H olefination. J. Am. Chem. Soc. 2011, 133, 5767–5769. [Google Scholar] [CrossRef] [PubMed]
  97. Fischer, J.; Savage, G.P.; Coster, M.J. A concise route to dehydrobenzo[b]furans: Formal total synthesis of (+)-lithospermic acid. Org. Lett. 2011, 13, 3376–3379. [Google Scholar] [CrossRef] [PubMed]
  98. Ly, T.N.; Shimoyamada, M.; Yamauchi, R. Isolation and characterization of rosmarinic acid oligomers in Celastrus hindsii Benth leaves and their antioxidative activities. J. Agric. Food Chem. 2006, 54, 3786–3793. [Google Scholar] [CrossRef] [PubMed]
  99. Lu, F.; Wei, L.; Azarpira, A.; Ralph, J. Rapid synthesis of dehydrodiferulates via biomometic radical coupling reactions of ethyl ferulate. J. Agric. Food Chem. 2012, 60, 8272–8277. [Google Scholar] [CrossRef] [PubMed]
  100. Agata, I.; Kusakabe, H.; Hatano, T.; Nishibe, S.; Okuda, T. Melitric acids A and B, new trimeric caffeic acid derivatives from Meliss officinalis. Chem. Pharm. Bull. 1993, 41, 1608–1611. [Google Scholar] [CrossRef]
  101. Lu, Y.; Foo, L.Y. Rosmarinic acid derivatives from Salvia officinalis. Phytochemistry 1999, 51, 91–94. [Google Scholar] [CrossRef]
  102. Lu, Y.; Foo, L.Y.; Wong, H. Sagecoumarin, a novel caffeic acid trimer from Salvia officinalis. Phytochemistry 1999, 52, 1149–1152. [Google Scholar] [CrossRef]
  103. Wang, J.; Xiong, X.; Feng, B. Cardiovascular effects of salvianolic acid B. Evid. Based Complement. Alternat. Med. 2013, 2013. [Google Scholar] [CrossRef]
  104. Zhao, Y.; Guo, Y.; Gu, X. Salvianolic acid B, a potential chemoprotective agent, for head and necksquamous cell cancer. J. Oncol. 2011, 2011. [Google Scholar] [CrossRef]
  105. Setälä, H.; Pajunen, A.; Rummakko, P.; Sipilä, J.; Brunow, G. A novel spiro compound formed by oxidative cross coupling of methyl sinapate with a syringyl lignin model compound. A model system for the β-1 pathway in lignin biosynthesis. J. Chem. Soc. Perkin Trans. 1 1999, 461–464. [Google Scholar]

Share and Cite

MDPI and ACS Style

Magoulas, G.E.; Papaioannou, D. Bioinspired Syntheses of Dimeric Hydroxycinnamic Acids (Lignans) and Hybrids, Using Phenol Oxidative Coupling as Key Reaction, and Medicinal Significance Thereof. Molecules 2014, 19, 19769-19835. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules191219769

AMA Style

Magoulas GE, Papaioannou D. Bioinspired Syntheses of Dimeric Hydroxycinnamic Acids (Lignans) and Hybrids, Using Phenol Oxidative Coupling as Key Reaction, and Medicinal Significance Thereof. Molecules. 2014; 19(12):19769-19835. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules191219769

Chicago/Turabian Style

Magoulas, George E., and Dionissios Papaioannou. 2014. "Bioinspired Syntheses of Dimeric Hydroxycinnamic Acids (Lignans) and Hybrids, Using Phenol Oxidative Coupling as Key Reaction, and Medicinal Significance Thereof" Molecules 19, no. 12: 19769-19835. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules191219769

Article Metrics

Back to TopTop