Next Article in Journal
Multi-Matrix Approach for the Analysis of Bicalutamide Residues in Oncology Centers by HPLC–FLD
Next Article in Special Issue
Synthetic Methods for the Preparation of Conformationally Restricted Analogues of Nicotine
Previous Article in Journal
Biosynthesis of Selenium Nanoparticles (via Bacillus subtilis BSN313), and Their Isolation, Characterization, and Bioactivities
Previous Article in Special Issue
Synthesis and Biological Evaluation of Amidinourea Derivatives against Herpes Simplex Viruses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synthetic Strategies in the Preparation of Phenanthridinones

by
Rajeshwar Reddy Aleti
1,
Alexey A. Festa
1,
Leonid G. Voskressensky
1 and
Erik V. Van der Eycken
1,2,*
1
Organic Chemistry Department, Science Faculty, RUDN University, Miklukho-Maklaya St., 6, 117198 Moscow, Russia
2
Laboratory for Organic & Microwave-Assisted Chemistry (LOMAC), Department of Chemistry, University of Leuven (KU Leuven), Celestijnenlaan 200F, B-3001 Leuven, Belgium
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(18), 5560; https://doi.org/10.3390/molecules26185560
Submission received: 27 July 2021 / Revised: 7 September 2021 / Accepted: 8 September 2021 / Published: 13 September 2021
(This article belongs to the Special Issue Biologically Active Heterocyclic Compounds)

Abstract

:
Phenanthridinones are important heterocyclic frameworks present in a variety of complex natural products, pharmaceuticals and displaying wide range of pharmacological actions. Its structural importance has evoked a great deal of interest in the domains of organic synthesis and medicinal chemistry to develop new synthetic methodologies, as well as novel compounds of pharmaceutical interest. This review focuses on the synthesis of phenanthridinone scaffolds by employing aryl-aryl, N-aryl, and biaryl coupling reactions, decarboxylative amidations, and photocatalyzed reactions.

1. Introduction

The phenanthridinone skeleton is a common structural moiety found in bioactive alkaloids from many sources, such as oxynitidine [1], crinasiadine [2], narciclasine [3], narciprimine [4], phenaglydon [5], pancratistatin [6], oxoassoanine [7], pratosine [8], anhydrolycorinone [9], and lycoricidine [10]. Owing to the many structural derivatives of phenanthridines [11], their synthesis has been known since the beginning of the 20th century [12]. There are many classical methods available, e.g., via the Schmidt reaction [13,14,15,16], the Ullmann reaction [17], and the Beckmann rearrangement reaction [18,19], dienone-phenol rearrangement [20], and intramolecular rearrangement reaction [21]. In addition, phenanthridinones can be constructed through dichromate oxidation [22], internal cyclisation of a benzyne intermediate [23,24,25], and other conventional methods [26,27]. However, these classical reactions have their limitations due to the requirement of additional steps for the synthesis of the key starting materials, and the generally poor to moderate overall yields. Promptly, these techniques were replaced by using non-traditional methods employing environmentally friendly chemicals, microwave-assisted reactions and catalytic approaches. These more recent techniques offer numerous advantages like shorter reaction times, improved yields, acceleration of sluggish transformations, and cleaner reaction products.
Phenanthridinones are important heterocycles, as they are used as prospective agents for anticancer treatment [28,29,30], and for antiviral and nerve disorders [31,32]. They possess different pharmacological properties of current interest as a class of HMG-CoA reductase inhibitors [33], anti-HIV [34], anti-hepatitis C virus [35], immunomodulatory activity [36], antibacterial and antifungal activities [37], and treatment of ischemic injuries [38] and other activities [39,40]. Interestingly, they are also estrogen receptor modulators (SERMs) [41], tyrosine protein kinase inhibitors [42], and neurotrophin activity enhancers [43]. Phenanthridinones are known to be inhibitors of poly ADP-ribose polymerase (PARP) family proteins [44,45,46,47,48,49,50,51]. PJ34 [N-(6-oxo-5,6-dihydrophenanthridin-2-yl)-N,N-dimethylacetamide.HCl] is a phenanthridone drug which is a selective inhibitor of poly(ADP-ribose) polymerase [52,53,54]. PJ38 is yet another lead molecule of the phenanthridinone family, which is active as a PARP inhibitor. The synthetic bioactive analogue ARC-111 is well known for its pharmacological activities, and is reported as topoisomerase-1 targeting antitumor agent [55] (Scheme 1).
This review focuses on modern techniques like transition metal-catalyzed coupling reactions, transition metal free coupling reactions, aryne-mediated coupling reactions, amidation reactions through decarboaxylation, photo-induced reactions, oxidation reactions, and anionic cycloisomerization reactions for phenathridinone synthesis.

2. Simultaneous Aryl–Aryl and N–Aryl Coupling Reactions for Phenanthridinone Synthesis

Transition metal-mediated aryl–aryl and N–aryl bond forming reactions are powerful synthetic tools for constructing highly complex molecules [56,57]. They have been recently employed in the key steps of the total synthesis of nitrogen containing natural products, as well as for the construction of heterocycles in medicinal chemistry in a one-pot fashion or multi step processes due to high efficiency and compatibility with numerous functional groups. One-pot C–C/C–N bond forming reactions have been used in drug discovery [58]. In this section, simultaneous aryl–aryl and N–aryl coupling reactions are described.

2.1. C–C and C–N Bond Formation via a One-Pot Synthesis

Simultaneous C–C and C–N bond forming reactions played pivotal roles for the synthesis of biologically relevant phenanthridinones. The most atom efficient protocols are based on one or more C–H-activation steps, avoiding excessive use of halogenated derivatives. The Snieckus group and Rault group prepared phenanthridinones by using transition metal-catalyzed C–C and C–N bond formation coupling reactions in a two-step process [59,60]. The use of N-methoxybenzamides 2 became a fruitful direction for the development of C–H activation methodology (Scheme 2). In 2011, Wang and colleagues demonstrated the synthesis of biologically important phenanthridinones through a one-pot formation of C–C and C–N bonds using palladium-catalyzed dual C–H activation of N-methoxybenzamides using phenyliodide and silver oxide as oxidant [61]. Later, the following transformation was realized under the catalysis with binaphthyl stabilized palladium nanoparticles [62]. Further, non-halogenated arenes could be engaged in analogous phenathridinone synthesis [63]. Multiple oxidative C–H bond activation and C–C/C–N bond formation steps in a one pot fashion are the benefits of this process. The utility of the methodology was disclosed through crinasiadine natural product synthesis. In continuation, N-methoxybenzamides could be annulated with aryl boronic acids [64] or aryl silanes [65] to form phenanthridinones under rhodium-catalysis. In general, the reactions start with the formation of pallada- or rhodacycles, which further react with the coupling partner.
In 2002, Caddick and his colleagues reported the synthesis of phenanthridinones using o-halobenzamides through an intramolecular Heck reaction [66]. In 2004, Ferraccioli, Catellani and colleagues reported the synthesis of phenanthridinones 6 using o-halobenzamides 4 and aryl iodides 5 as starting materials (Scheme 3) [67]. In this case, the use of norbornene was necessary to mediate the reaction and achieve a C–H activation step. Interestingly, when benzamides without halogen in the ortho position were treated with 1-bromo-2-iodobenzenes under Pd-catalysis, the synthesis of phenanthridinones was successfully achieved in the absence of norbornene [68]. An efficient intermolecular dehydrogenative annulation of aryl iodides and aryl carbamic chlorides was reported for the synthesis of phenanthridinones using Pd-norbornene-catalysis [69].
2-Bromobenzamides 7 could be subjected to Pd-catalyzed homocoupling, delivering phenanthridinone derivatives 8 [70,71,72] (Scheme 4). In 2016, phenanthridinone synthesis was realized starting from o-halobenzamides under phosphine-free palladium catalysis in N,N-dimethylacetamide [73]. Authors also demonstrated this methodology in a scalable process by performing gram scale reactions. Later, Hu and his colleagues attempted the synthesis of N–H-phenanthridinones by using palladium-catalyzed intramolecular C–H arylation of 2-halo-2-Boc-N-arylbenzamides [74]. A variety of benzamides bearing electron-donating and electron-withdrawing groups can be coupled and produce the corresponding phenanthridinones in moderate to good yields.
A decarboxylative copper-mediated coupling of benzamides 9 with ortho-nitrobenzoic acid salts 10 was realized by Miura and co-workers (Scheme 5) [75]. The C–H-activation step was achieved with the help of a quinol-8-yl directing group. After the decarboxylative coupling, cyclization occurred through nucleophilic substitution of a nitro-group by an amide moiety. The resulting phenathridinones 11 were isolated with moderate yields. Li et al. reported a Pd-catalyzed decarboxylative ortho-arylation of N-methoxyarylamides using aryl acylperoxides [76].
In 1994, Tour and colleague reported a phenanthridinone synthesis using 2-halobenzoate and aromatic boronic acids using palladium as the catalyst [77]. However, the products were obtained with low yields and the substrate scope was not investigated. Later, Tanimoto et al. established a convenient one-step access to biologically important phenanthridinones 14 using 2-halobenzoate 12 and 2-aminophenylboronic acid 13 in a Suzuki–Miyaura cross-coupling reaction with palladium(II) acetate and 2-dicyclohexylphosphino-2′,6′-dimethoxybiphenyl (SPhos) as precatalysts (Scheme 6) [78]. Interestingly, Kuwata et al. extended this strategy to synthesize the phenanthridinone alkaloids crinasiadine and N-alkylcrinasiadines from 2-aminophenylboronic acid and 2-bromobenzoate [79]. Furthermore, 2-halobenzoates were proved to be an efficient substrate for preparing biologically active phenanthridinones [80,81,82].
Recently, Ding et al. reported the synthesis of phenanthridinones 17 through Pd-catalyzed cyclization of N-aryl-2-aminopyridine 15 with 2-iodobenzoic acid 16 via C(sp2)–H bond activation (Scheme 7) [83]. The advantage of the methodology was in low catalyst-loading, and that the reaction smoothly proceeded in water. The broad substrate scope was demonstrated with high yields comparatively with previous methods. A tentative mechanism, based on experimental results, was proposed by the authors. Primarily, the substrate 15 coordinates with Pd(OAc)2 via the pyridine nitrogen atom. A six membered palladacycle dimer complex 15-A is generated through directed C–H bond activation. Further, a carboxylate-directed oxidative addition generates a Pd(IV) species 15-B. This is followed by reductive elimination, resulting in the ortho-arylated product 15-C. Subsequently, the latter undergoes intramolecular acylation to generate 17.

2.2. C–C and C–N Bond Formation via Aryne-Mediated Reactions

In 2002, Lu et al. described an aryne-mediated synthetic methodology of substituted o-halobenzamides through a palladium-catalyzed annulation for the synthesis of N-substituted phenanthridinones in good yields. The advantage of this methodology is the preparation of phenanthridinones in a single step via simultaneous C–C and C–N bond formation, under relatively mild reaction conditions, tolerating a wide variety of functional groups [84]. Later, Jeganmohan and colleagues applied this concept on methyl or methoxy benzamides 18 and reported an aryne mediated cyclization using o-(trimethylsilyl)aryl triflate 19 in the presence of Pd(OAc)2, organic acid and K2S2O8 in CH3CN yielding tricyclic phenanthridinone derivatives 20 in good yields. The scope of the catalytic reaction was examined with a variety of substituted N-methoxy benzamides (Scheme 8) [85]. Initially, a five membered palladacycle 18-A forms by the coordination of N-methoxy benzamide 18 to the Pd(0) species followed by ortho-metalation. Coordinative insertion of the benzyne intermediate 19-A into intermediate 18-A forms a seven membered palladacycle 18-B. This undergoes reductive elimination, as well as C–N bond formation using RCOOH, K2S2O8 with regeneration of the Pd catalyst (Scheme 9). Furthermore, the utility of benzyne precursors was reported using palladium, copper, and nickel catalysts [86,87,88,89,90].

2.3. C–C and C–N Bond Formation via Carbonylative and Carboxylative Reactions

In 2013, Zhang [91] and Zhu [92] concurrently described a phenanthridinone synthesis through carbonylation of biphenyl-2-amines 21 employing a combination of Pd(II) and Cu(II) salts under a carbon monoxide atmosphere (Scheme 10). Zhang and co-workers succeeded in developing a catalytic system that can avoid the formation of urea as side product, using a combination of Pd(OAc)2 and Cu(II) trifluoroacetate in 2,2,2-trifluoroethanol. The substrate scope of the reaction was verified using various electron rich and electron deficient biphenylamines. Electronic rich biphenylamines gave higher yields compared to electronic deficient biphenylamines [91]. In the case of Zhu’s work, 1 equiv of TFA was needed for smooth conversion [92]. Moreover, Chuang and colleagues used only a palladium catalyst for oxidative insertion of carbon monoxide on N-sulfonyl-2-aminobiaryls to furnish phenanthridinone 22 [93]. Interestingly, DMF could also be used as a one-carbon-source instead of toxic carbon monoxide [94]. Ling and colleagues reported a CoCl2-catalyzed carbonylation of ortho-arylanilines in the presence of diazadicarboxylates as a carbonyl source and oxygen as oxidant [95]. 10-Hydroxy-10,9-borazarophenanthrenes could be converted into phenanthridinones, including phenaglydon, through Pd(II)-catalyzed carbonylation [96].
The use of non-toxic CO2 as a carbonyl source for phenanthridine formation could be advantageous. For example, o-arylanilines 21 could be transformed into phenanthridinones 22 under a CO2 atmosphere using transition metal free conditions (Scheme 11) [97]. Substrates bearing an electron-donating group on the aryl moiety, demonstrated higher reactivity to furnish 22 with higher yields, while electron-withdrawing groups gave lower yields. Based on experimental results, a plausible mechanism was proposed. Initially, the o-arylaniline 21 reacted with the adduct of 1,5,7-triazabicyclo[4.4.0] dec-5-ene (TBD) and CO2 to produce the carbamate 21-A with the aid of MeOTf. Release of MeOH generated isocyanate 21-B at high temperature. Subsequently, internal cyclization of intermediate 21-C resulted in the formation of 21-D, which upon hydrolysis gave phenanthridinone 22. In 2021, aryisocyanates 21-B were directly transformed into phenathridinones under the action of catalytic amounts of MeOTf in DCE. That approach was used for the synthesis of amaryllidaceae alkaloids [98].
In 2019, Gao et al. developed an amino group-assisted C–H carboxylation of 2-arylanilines with CO2 through Rh(I) catalysis under redox neutral condition (Scheme 12) [99]. The reaction was carried out in the presence of a phosphine ligand and tBuOK as base. The advantage of the methodology is that no external oxidant was needed. The reaction started with the oxidative addition of Rh(I), followed by a reductive elimination of HX to generate a rhodacycle, which underwent carboxylation with CO2. The resulting Rh carboxylate cyclized into the target product. Only N-unsubstituted 2-phenylanilines 21 participated successfully in the reaction.

3. C–C Coupling Reactions for Phenanthridinone Synthesis

Arylation of arenes has emerged as a powerful tool for the synthesis of biaryl compounds, which led to a variety of bioactive molecules [100,101]. Biaryl coupling has shown a prominent strategy to prepare bioactive phenanthridinones. The synthesis of phenthridinones through C–C coupling reactions is discussed in this section.
Benzanilides with halogen substituent in the ortho-position were extensively used for the preparation of phenathridinones [102,103,104,105,106,107,108]. Later, a dehydrogenative approach with halogen-free compounds was developed. Dong and co-workers reported an intramolecular ortho-arylation of benzanilides 23 [109]. The reaction was performed in DCE under the action of palladium (II) acetate (10 mol%) and Na2S2O8 as oxidant. The addition of TFA was crucial for the transformation. The methodology allowed the synthesis of N-methylcrinasiadine natural product in 30% yield. Further, Murakami’s group used molecular oxygen as terminal oxidant for the palladium-catalyzed arylation of benzanilides 23 towards the synthesis of phenanthridinones 24 via oxidative C–H coupling (Scheme 13) [110]. Interestingly, non-substituted benzanilide afforded the phenanthridinone in 89% yield. Chloro- and bromo-groups remained intact, while cyano-substituents were not tolerated under the reaction conditions. Later, Du, Zhao, and-coworkers developed a Cu(II)-promoted conversion of N-benzoylated enaminones into 3-hydroxyphenanthridinones [111]. Interestingly, N-formyl-2-arylanilines could be cyclized into phenathridinones under Cu(0)-catalysis in the presence of stoichiometric amounts of selectfluor [112].
The following mechanism was proposed by the authors (Scheme 14) [110]. The palladium (II) species coordinates to the carbonyl oxygen of the amide group of the anilide via ortho-palladation, resulting in the formation of the aryl palladium species 23-A. Un-coordination of the amide bond in intermediate 23-A was followed by the rotation of the amide bond and placing the ortho-position of the benzoyl moiety in close proximity to the palladium center, leading to the second aromatic palladation to generate intermediate 23-B. Reductive elimination gives phenanthridinone 24 and palladium (0). Palladium (II) was regenerated by the action of molecular oxygen and benzoic acid.
Benzanilides 25 could be cyclized into phenathridinones 26 under transition metal free conditions. For instance, benzanilide 25 (X=H) underwent the desired transformation in the presence of phenyliodine (III)-bis (trifluoroacetate) (PIFA) (Scheme 15) [113]. The reaction was possible for electron-rich substrates. Bhakuni et al. developed a transition metal free aryl arene coupling of halo benzamides 25 using KOtBu, 1,10-phenathroline as ligand and AIBN as a radical initiator [114]. In 2016, Sharma et al. developed a transition metal-free microwave-assisted methodology utilizing vasicine 27 (a natural product) as a catalyst for the synthesis of phenanthridinones. The reaction proceeded through intramolecular C–H arylation with aryl halides in the presence of KOtBu base, under microwave irradiation in sulfolane as the solvent. Remarkably, the reaction worked smoothly with less reactive aryl chlorides and reached completion after 15 min [115]. Bromides could be analogously cyclized in the presence of 1-(2-hydroxyethyl)-piperazine (28) and DMAP in mesitylene (Scheme 15) [116].
Lautens and colleagues reported a synthesis of remarkable ortho-aminated phenanthridinones 31 to be produced by reacting wide scope of benzamides 29 with O-benzoyl-hydroxylamine 30 under palladium-norbornene catalysis. Various hydroxylamines could also be used, although less efficiently (Scheme 16) [117].

4. N–Aryl Coupling Reactions for Phenanthridinone Synthesis

An N–Aryl coupling reaction opened avenues for the synthesis of heterocycles, that are abundant in natural products [118]. In addition, N–aryl coupling reactions proved to be a good strategy to prepare pharmaceutically important phenanthridinone derivatives. In this section, phenanthridinone synthesis, using metal catalyzed N–aryl coupling reactions and transition metal free N–aryl coupling reactions, is discussed.
Oxidative intramolecular palladium-catalyzed C–H amination of arenes for the synthesis of phenanthridin-6-one in 63% yield was reported in 2012 [60], employing the procedure for the synthesis of quinolones (Scheme 17) [119]. The removal of the tosyl group took place at the same time. Later, Gui and colleagues developed a copper-catalyzed approach for the preparation of phenanthridin-6-ones from 2-phenylbenzamides 32 (R=H) through an intramolecular N-aryl bond-forming process under basic conditions [120]. Ortho-arylation of benzamides proceeded smoothly in the presence of CuI (10 mol%), KOtBu (2 equiv), and a phosphine ligand (20 mol%). Heating the reactant in o-xylene at 120 °C for 18 h was needed to produce the target molecules in 40%–92% yield. Only N-unsubstituted amides could be used. Later, hypervalent iodine-mediated processes for the cyclization of N-methoxybenzamides 32 (R-OMe) were developed. Jiang, Xue, and-coworkers used catalytic amounts of iodobenzene and peroxy acids as oxidants [121,122]. The reactions were performed in 1,1,1,3,3,3-hexafluoroisopropanol (HFIP) under an air atmosphere. These consecutive C–H functionalization reactions can be used efficiently to construct 2-substituted-phenanthridinones at room temperature with moderate to high yields. Interestingly, when the reaction was carried out in the presence of TBAB, a bromide was introduced in para-position of nitrogen [121]. In addition, it has been recently shown that N-methoxy amides could be also cyclized into phenathridinones under electrochemical conditions [123].
Wang et al. reported an oxidative arene C(sp2)–H amidation for the synthesis of phenanthridinones using amide 33 and NIS as iodinating reagent (Scheme 18) [124]. Primarily, N-iodosuccinimide reacts with amide 33 to form N-iodinated compound 33-A. It undergoes an N–I homolytic cleavage providing amide N-radical 33-B under thermal conditions. Subsequently, 5-exo-cyclization on to the aromatic ring of intermediate 33-B afforded intermediate 33-C, or 6-endo cyclization gave benzolactam intermediate 33-D. Benzolactams could also be obtained through a ring expansion via C–N bond migration. Further rearomatization gave product 34. When the substrate was bearing OTBS-group, the intermediate 29-C delivered spirolactam 35. Later, Verma et al. reported a cyclization of 2-phenylbenzamides for the preparation of phenanthridinones using iodine, succinimide, and di-tert-butylperoxide (DTBP) as oxidant [125]. It has been shown that 2′-iodo-[1,1′-biphenyl]-2-carboxamide could be cyclized into phenathridinone under Ni-catalysis in the presence of boronic acid [126].
Other useful substrates for the synthesis of phenathridinones via a C–N bond forming route are 2-arylbenzonitriles 36. Fabis and-coworkers reported a potassium hydroxide mediated anionic ring closure of fluorine-substituted benzonitriles 36 (X = F) [127,128]. Later, Chen et al. reported a copper-catalysis of annulation of bromide-substituted benzonitriles 36 (X = Br) (Scheme 19) [129]. Furthermore, the utility of copper catalysis was extended to the synthesis of phenathridinone containing bioactive natural products [130,131]. Benzonitriles with an alkyne moiety in ortho-position could undergo an anionic cycloaromatization when treated with sodium methylate in methanol, delivering condensed phenathridinones [132,133].

5. Decarboxylative Reactions for Phenanthridinone Synthesis

Efficient C–C bond formation via transition-metal-catalyzed decarboxylation of aromatic carboxylic acids is a known strategy in organic synthesis [134]. This approach can afford unconventional alternatives to perform various types of C–C and C–heteroatom bond-forming reactions. In 2012, Shen and colleagues have achieved an efficient protocol for forming biaryl compounds through palladium-catalyzed intramolecular decarboxylative coupling of arene carboxylic acids with aryl bromides, yielding various cyclized heterocycles, including phenanthridinones [135]. Later, intramolecular decarboxylative amidation of unactivated arenes was achieved under transition metal free conditions (Scheme 20) [136]. Na2S2O8-promoted decarboxylative cyclization of biaryl-2-oxamic acids 37, gave phenanthridinones 38 with high efficiency. It was observed that the wide scope of the desired products was obtained in good to excellent yields. As electron-withdrawing groups on the aromatic ring did not lower the reaction efficiency, a radical reaction pathway is followed.

6. Photo-Mediated Reactions for Phenanthridinone Synthesis

Recently, photocatalyzed reactions proved to be a valuable method for the synthesis of heterocycles and biologically active compounds [137]. Firstly, Mondon et al. investigated the photocyclization of benzanilides to furnish unsymmetrically substituted phenanthridinones [138]. Later, Prabhakar et al. reported an improved phenanthridinone synthesis by using the photocylization of boron complexes, followed by hydrolysis [139]. In 2017, a facile photo reductive protocol was developed to remove benzyl o-protective groups from phenathridin-6-ol via blue light irradiation [140]. Furthermore, Moon and colleagues synthesized phenanthridinones 40 through a visible-light-promoted direct oxidative C–H amidation [141]. In this photocatalytic system, oxidative intramolecular C–H amidation of benzamide 39 was achieved through amidyl radicals. Those could be generated by homolysis of the N–H bond of simple amide precursors via single-electron transfer under blue LED irradiation. The scope of the methodology was further extended to different derivatives of benzamides. Interestingly, halides such as fluoro, chloro, bromo, and iodo substituents were tolerated under the reaction conditions (Scheme 21). Subsequently, the applicability of the method was verified with respect to substrates bearing p-methoxyphenyl (PMP) groups on the amide nitrogen. This worked well with the optimized system and provided the desired products easily. When meta-substituted substrates were employed, the formation of the C–N bond took place exclusively at the more sterically accessible C–H bond of the aryl ring, affording only one regioisomer.
Recently, Itoh et al. developed a transition metal free alternative for visible light-mediated cyclization of benzamides 39 into phenathridinones 40 (Scheme 22) [142]. As a photocatalyst, 1-chloroanthraquinone 41 was employed. The reaction was performed under air atmosphere. The proposed mechanism is similar to the transformation on Scheme 21. Photoexcited catalyst oxidizes benzamide 39 to generate an amidyl radical 39-A, capable of intramolecular cyclization. The subsequent radical 39-B traps oxygen, giving peroxy radical 39-C. The latter is reduced and subsequently aromatized by elimination of hydroperoxide anion. In comparison with the Ir-catalyzed reaction, this transformation tolerates alkyl-substituted amides (R3 = Ar, Alk).

7. Miscellaneous Reactions

In 2017, Yaragorla and colleagues developed a Ca(II)-catalyzed one-pot reaction between oxindole 41 and styrenes, furnishing phenanthridinones 42 in moderate to good yields (Scheme 23) [143]. The reaction starts with a dehydrative cross-coupling, delivering the isolable intermediate 43. A sigmatropic rearrangement in the latter leads to the formation of allene 44, capable of intramolecular cyclization into compound 45. This spiro-cycle 45 could be transformed into the desired phenathridinone 42 under oxidation-induced ring rearrangement.
A multicomponent approach for phenathridinone synthesis was developed by Alzaydi and co-workers [144]. The reaction between ortho-aminoacetophenone 46, ethyl cyanoacetate, sulfur and alkyne was carried out in the presence of AcOH/NH4OAc under increased pressure in a Q-tube (Scheme 24).
It was shown that unsaturated carbonyl compounds 47 could react with dimethyl glutaconate 48 to produce phenathridinones 49 in excellent yields (Scheme 25) [145]. The reaction took place in methanol at rt under the action of sodium hydroxide. The sequence started with the Michael addition of glutaconate 48 to the activated double bond of 47, giving intermediate 50. Double bond migration, followed by intramolecular cyclization gave cyclohexadiene 51, which underwent an intramolecular cyclization into a non-aromatic phenanthridinone. Spontaneous aromatization delivered the desired product 49.

8. Conclusions

Phenanthridinones have attracted the attention of organic and medicinal chemists because of the diverse pharmacological and biological properties of their alkaloids. Due to promising pharmacological profiles and unique structural features, coupled with low natural abundance, many synthetic methodologies have been developed. Here, we reviewed the synthetic strategies that can achieve phenanthridinone formation. Due to serious efforts in the development of modern synthetic methods, the construction of the phenanthridinone framework through C–C and C–N bond formation reactions via dual C–H bond activation has evolved. Still, the created approaches suffer from the need for pre functionalized substrates, limited substrate scope and scalability, leaving space for new discoveries. We believe that the most crucial developments in sustainable and efficient syntheses of phenathridinones might evolve from the fields of photocatalyzed or electrochemical C–H functionalizations, conceivably combined with the flow chemistry.

Funding

This paper has been supported by the RUDN University Strategic Academic Leadership Program. The research was funded by RFBR and Moscow city Government according to the project № 21-33-70007.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kametani, T.; Kigasawa, K.; Hiiragi, M.; Kusama, O. Studies on the syntheses of heteroeyelie compounds. Part DVI. Synthesis of oxynilidine and nitidine. J. Heterocycl. Chem. 1973, 10, 31–33. [Google Scholar] [CrossRef]
  2. Van Otterlo, W.A.L.; Green, I.R. A Review on Recent Syntheses of Amaryllidaceae Alkaloids and Isocarbostyrils (Time period mid-2016 to 2017). Nat. Prod. Commun. 2018, 13, 255–257. [Google Scholar] [CrossRef] [Green Version]
  3. Fuganti, C.; Mazza, M. The absolute configuration of narciclasine: A biosynthetic approach. J. Chem. Soc. Chem. Commun. 1972, 239. [Google Scholar] [CrossRef]
  4. Bozkurt, B.B.; Zencir, S.; Somer, N.U.; Kaya, G.I.; Önür, M.A.; Bastida, J.; Berenyi, A.; Zupko, I.; Topcu, Z. The Effects of Arolycoricidine and Narciprimine on Tumor Cell Killing and Topoisomerase Activity. Rec. Nat. Prod. 2012, 6, 381. [Google Scholar]
  5. Wurz, G.; Hofer, O.; Greger, H. Structure and synthesis of phenaglydon, a new quinolone derived phenanthridine alkaloid from Glycosmis cyanocarpa. Nat. Prod. Lett. 1993, 3, 177–182. [Google Scholar] [CrossRef]
  6. Banwell, M.G.; Cowden, C.J.; Gable, R.W. Lycoricidine and pancratistatin analogues from cyclopentadiene. J. Chem. Soc. Perkin Trans. I 1994, 3515–3518. [Google Scholar] [CrossRef]
  7. Meyers, A.; Hutchings, R. A total synthesis of the pyrrolophenthridone alkaloid oxoassoanine. Tetrahedron Lett. 1993, 34, 6185–6188. [Google Scholar] [CrossRef]
  8. Ganton, M.D.; Kerr, M.A. A domino amidation route to indolines and indoles: Rapid syntheses of anhydrolycorinone, hippadine, oxoassoanine, and pratosine. Org. Lett. 2005, 7, 4777–4779. [Google Scholar] [CrossRef]
  9. Wolkenberg, S.E.; Boger, D.L. Total synthesis of anhydrolycorinone utilizing sequential intramolecular Diels–Alder reactions of a 1,3,4-oxadiazole. J. Org. Chem. 2002, 67, 7361–7364. [Google Scholar] [CrossRef]
  10. Ugarkar, B.G.; Dare, J.; Schubert, E.M. Improved synthesis of lycoricidine triacetate. Synthesis 1987, 8, 715–716. [Google Scholar] [CrossRef]
  11. Smith, P.A. The schmidt reaction: Experimental conditions and mechanism. J. Am. Chem. Soc. 1948, 70, 320–323. [Google Scholar] [CrossRef]
  12. Oyster, L.; Adkins, H. The Preparation Of 9 (10)-Phenanthridone From Phenanthrene. J. Am. Chem. Soc. 1921, 43, 208–210. [Google Scholar] [CrossRef]
  13. Walls, L.P. 337. Researches in the phenanthridine series. Part IV. Synthesis of plasmoquin-like derivatives. J. Chem. Soc. 1935, 1405–1410. [Google Scholar] [CrossRef]
  14. Pan, H.L.; Fletcher, T.L. 6(5H)-phenanthridinones. II. Preparation of substituted 6 (5h)-phenanthridinones from 9-oxofluorenes. J. Heterocycl. Chem. 1970, 7, 313–321. [Google Scholar] [CrossRef]
  15. Woodroofe, C.C.; Zhong, B.; Lu, X.; Silverman, R.B. Anomalous Schmidt reaction products of phenylacetic acid and derivatives. J. Chem. Soc. Perkin Trans. 2 2000, 55–59. [Google Scholar] [CrossRef]
  16. Mosby, W.L. 2-Bromophenanthridone. J. Am. Chem. Soc. 1954, 76, 936–937. [Google Scholar] [CrossRef]
  17. Banwell, M.G.; Lupton, D.W.; Ma, X.; Renner, J.; Sydnes, M.O. Synthesis of Quinolines, 2-Quinolones, Phenanthridines, and 6 (5 H)-Phenanthridinones via Palladium [0]-Mediated Ullmann Cross-coupling of 1-Bromo-2-nitroarenes with β-Halo-enals,-enones, or-esters. Org. Lett. 2004, 6, 2741–2744. [Google Scholar] [CrossRef] [PubMed]
  18. Horning, E.; Stromberg, V.; Lloyd, H. Beckmann rearrangements. An investigation of special cases. J. Am. Chem. Soc. 1952, 74, 5153–5155. [Google Scholar] [CrossRef]
  19. Guy, A.; Guette, J.-P.; Lang, G. Utilization of polyphosphoric acid in the presence of a co-solvent. Synthesis 1980, 1980, 222–223. [Google Scholar] [CrossRef]
  20. Guo, X.; Xing, Q.; Lei, K.; Zhang-Negrerie, D.; Du, Y.; Zhao, K. A Tandem Ring Opening/Closure Reaction in A BF3-Mediated Rearrangement of Spirooxindoles. Adv. Synth. Catal. 2017, 359, 4393–4398. [Google Scholar] [CrossRef]
  21. Corsaro, A.; Librando, V.; Chiacchio, U.; Pistarà, V.; Rescifina, A. Cycloaddition of nitrile oxides to aza-analogues of phenanthrene. Tetrahedron 1998, 54, 9187–9194. [Google Scholar] [CrossRef]
  22. Gilman, H.; Eisch, J. The Chemistry and Synthetic Applications of the Phenanthridinone System. J. Am. Chem. Soc. 1957, 79, 5479–5483. [Google Scholar] [CrossRef]
  23. Hey, D.; Leonard, J.A.; Rees, C.W. 1003. Internuclear cyclisation. Part XX. Synthesis of spiro-dienones through benzyne intermediates. J. Am. Chem. Soc. 1963, 5266–5270. [Google Scholar] [CrossRef]
  24. González, C.; Guitián, E.; Castedo, L. Synthesis of phenanthridones, quinolinequinones and 7-azasteroids. Tetrahedron 1999, 55, 5195–5206. [Google Scholar] [CrossRef]
  25. Sanz, R.; Fernandez, Y.; Castroviejo, M.P.; Perez, A.; Fananas, F.J. Functionalized Phenanthridine and Dibenzopyranone Derivatives through Benzyne Cyclization–Application to the Total Syntheses of Trisphaeridine and N-Methylcrinasiadine. Eur. J. Org. Chem. 2007, 2007, 62–69. [Google Scholar] [CrossRef]
  26. Narasimhan, N.; Paradkar, M.; Alurkar, R. Synthetic application of lithiation reactions—IV: Novel synthesis of linear furoquinoline alkaloids and a synthesis of edulitine. Tetrahedron 1971, 27, 1351–1356. [Google Scholar] [CrossRef]
  27. Radhakrishnan, H.; Krishnamoorthy, M.; Chien-Hong, C. Rhodium(III)-Catalyzed ortho-Arylation of Anilides with Aryl Halides. Adv. Synth. Catal. 2015, 357, 366–370. [Google Scholar]
  28. Fukuda, H.; Karaki, F.; Dodo, K.; Noguchi-Yachide, T.; Ishikawa, M.; Hashimoto, Y.; Ohgane, K. Phenanthridin-6-one derivatives as the first class of non-steroidal pharmacological chaperones for Niemann-Pick disease type C1 protein. Bioorganic Med. Chem. Lett. 2017, 27, 2781–2787. [Google Scholar] [CrossRef]
  29. Achary, R.; Mathi, G.R.; Lee, D.H.; Yun, C.S.; Lee, C.O.; Kim, H.R.; Park, C.H.; Kim, P.; Hwang, J.Y. Novel 2, 4-diaminopyrimidines bearing fused tricyclic ring moiety for anaplastic lymphoma kinase (ALK) inhibitor. Bioorganic Med. Chem. Lett. 2017, 27, 2185–2191. [Google Scholar] [CrossRef]
  30. Ruchelman, A.L.; Kerrigan, J.E.; Li, T.-K.; Zhou, N.; Liu, A.; Liu, L.F.; LaVoie, E.J. Nitro and amino substitution within the A-ring of 5H-8,9-dimethoxy-5-(2-N,N-dimethylaminoethyl) dibenzo [c,h][1,6] naphthyridin-6-ones: Influence on topoisomerase I-targeting activity and cytotoxicity. Bioorg. Med. Chem. 2004, 12, 3731–3742. [Google Scholar] [CrossRef]
  31. Kralj, A.; Kurt, E.; Tschammer, N.; Heinrich, M.R. Synthesis and biological evaluation of biphenyl amides that modulate the US28 receptor. Chem. Med. Chem. 2014, 9, 151–168. [Google Scholar] [CrossRef]
  32. Stoica, B.A.; Loane, D.J.; Zhao, Z.; Kabadi, S.V.; Hanscom, M.; Byrnes, K.R.; Faden, A.I. PARP-1 inhibition attenuates neuronal loss, microglia activation and neurological deficits after traumatic brain injury. J. Neurotrauma 2014, 31, 758–772. [Google Scholar] [CrossRef] [PubMed]
  33. Karaki, F.; Ohgane, K.; Fukuda, H.; Nakamura, M.; Dodo, K.; Hashimoto, Y. Structure–activity relationship study of non-steroidal NPC1L1 ligands identified through cell-based assay using pharmacological chaperone effect as a readout. Bioorg. Med. Chem. 2014, 22, 3587–3609. [Google Scholar] [CrossRef] [PubMed]
  34. Patil, S.; Kamath, S.; Sanchez, T.; Neamati, N.; Schinazi, R.F.; Buolamwini, J.K. Synthesis and biological evaluation of novel 5 (H)-phenanthridin-6-ones, 5 (H)-phenanthridin-6-one diketo acid, and polycyclic aromatic diketo acid analogs as new HIV-1 integrase inhibitors. Bioorg. Med. Chem. 2007, 15, 1212–1228. [Google Scholar] [CrossRef]
  35. Nakamura, M.; Aoyama, A.; Salim, M.T.A.; Okamoto, M.; Baba, M.; Miyachi, H.; Hashimoto, Y.; Aoyama, H. Structural development studies of anti-hepatitis C virus agents with a phenanthridinone skeleton. Bioorg. Med. Chem. 2010, 18, 2402–2411. [Google Scholar] [CrossRef] [PubMed]
  36. Weltin, D.; Picard, V.; Aupeix, K.; Varin, M.; Oth, D.; Marchal, J.; Dufour, P.; Bischoff, P. Immunosuppressive activities of 6 (5H)-phenanthridinone, a new poly (ADP-ribose) polymerase inhibitor. Int. J. Immunopharmacol. 1995, 17, 265–271. [Google Scholar] [CrossRef]
  37. Hu, J.; Shi, X.; Chen, J.; Mao, X.; Zhu, L.; Yu, L.; Shi, J. Alkaloids from Toddalia asiatica and their cytotoxic, antimicrobial and antifungal activities. Food Chem. 2014, 148, 437–444. [Google Scholar] [CrossRef]
  38. Ferraris, D.; Ko, Y.-S.; Pahutski, T.; Ficco, R.P.; Serdyuk, L.; Alemu, C.; Bradford, C.; Chiou, T.; Hoover, R.; Huang, S.; et al. Design and synthesis of poly ADP-ribose polymerase-1 inhibitors. 2. Biological evaluation of aza-5 [H]-phenanthridin-6-ones as potent, aqueous-soluble compounds for the treatment of ischemic injuries. J. Med. Chem. 2003, 46, 3138–3151. [Google Scholar] [CrossRef]
  39. Ceriotti, G. Narciclasine: An antimitotic substance from Narcissus bulbs. Nature 1967, 213, 595–596. [Google Scholar] [CrossRef] [PubMed]
  40. Cabral, V.; Luo, X.; Junqueira, E.; Costa, S.S.; Mulhovo, S.; Duarte, A.; Couto, I.; Viveiros, M.; Ferreira, M.-J.U. Enhancing activity of antibiotics against Staphylococcus aureus: Zanthoxylum capense constituents and derivatives. Phytomedicine 2015, 22, 469–476. [Google Scholar] [CrossRef] [PubMed]
  41. Grese, T.A.; Adrian, M.D.; Phillips, D.L.; Shetler, P.K.; Short, L.L.; Glasebrook, A.L.; Bryant, H.U. Photochemical synthesis of N-arylbenzophenanthridine selective estrogen receptor modulators (SERMs). J. Med. Chem. 2001, 44, 2857–2860. [Google Scholar] [CrossRef]
  42. Dow, R.L.; Chou, T.T.; Bechle, B.M.; Goddard, C.; Larson, E.R. Identification of tricyclic analogs related to ellagic acid as potent/selective tyrosine protein kinase inhibitors. J. Med. Chem. 1994, 37, 2224–2231. [Google Scholar] [CrossRef] [PubMed]
  43. Karra, S.; Xiao, Y.; Chen, X.; Liu-Bujalski, L.; Huck, B.; Sutton, A.; Goutopoulos, A.; Askew, B.; Josephson, K.; Jiang, X.; et al. SAR and evaluation of novel 5H-benzo [c][1,8] naphthyridin-6-one analogs as Aurora kinase inhibitors. Bioorganic Med. Chem. Lett. 2013, 23, 3081–3087. [Google Scholar] [CrossRef] [PubMed]
  44. Ishida, J.; Hattori, K.; Yamamoto, H.; Iwashita, A.; Mihara, K.; Matsuoka, N. 4-Phenyl-1,2,3,6-tetrahydropyridine, an excellent fragment to improve the potency of PARP-1 inhibitors. Bioorganic Med. Chem. Lett. 2005, 15, 4221–4225. [Google Scholar] [CrossRef] [PubMed]
  45. Fang, Y.; Tranmer, G.K. Continuous flow photochemistry as an enabling synthetic technology: Synthesis of substituted-6 (5H)-phenanthridinones for use as poly (ADP-ribose) polymerase inhibitors. Med. Chem. Comm. 2016, 7, 720–724. [Google Scholar] [CrossRef]
  46. Jagtap, P.; Szabó, C. Poly (ADP-ribose) polymerase and the therapeutic effects of its inhibitors. Nat. Rev. Drug Discov. 2005, 4, 421–440. [Google Scholar] [CrossRef]
  47. Holl, V.; Coelho, D.; Weltin, D.; Hyun, J.W.; Dufour, P.; Bischoff, P. Modulation of the antiproliferative activity of anticancer drugs in hematopoietic tumor cell lines by the poly (ADP-ribose) polymerase inhibitor 6 (5H)-phenanthridinone. Anticancer Res. 2000, 20, 3233–3241. [Google Scholar]
  48. Weltin, D.; Holl, V.; Hyun, J.W.; Dufour, P.; Marchal, J.; Bischoff, P. Effect of 6 (5H)-phenanthridinone, a poly (ADP-ribose) polymerase inhibitor, and ionizing radiation on the growth of cultured lymphoma cells. Int. J. Radiat. Biol. 1997, 72, 685–692. [Google Scholar] [CrossRef]
  49. Li, J.-H.; Serdyuk, L.; Ferraris, D.V.; Xiao, G.; Tays, K.L.; Kletzly, P.W.; Li, W.; Lautar, S.; Zhang, J.; Kalish, V.J. Synthesis of substituted 5 [H] phenanthridin-6-ones as potent poly (ADP-ribose) polymerase-1 (PARP1) inhibitors. Bioorganic Med. Chem. Lett. 2001, 11, 1687–1690. [Google Scholar] [CrossRef]
  50. Aldinucci, A.; Gerlini, G.; Fossati, S.; Cipriani, G.; Ballerini, C.; Biagioli, T.; Pimpinelli, N.; Borgognoni, L.; Massacesi, L.; Moroni, F.; et al. A key role for poly (ADP-ribose) polymerase-1 activity during human dendritic cell maturation. J. Immunol. 2007, 179, 305–312. [Google Scholar] [CrossRef]
  51. Bellocchi, D.; Macchiarulo, A.; Costantino, G.; Pellicciari, R. Docking studies on PARP-1 inhibitors: Insights into the role of a binding pocket water molecule. Bioorganic Med. Chem. 2005, 13, 1151–1157. [Google Scholar] [CrossRef] [PubMed]
  52. Scott, G.S.; Kean, R.B.; Mikheeva, T.; Fabis, M.J.; Mabley, J.G.; Szabo, C.; Hooper, D.C. The therapeutic effects of PJ34 [N-(6-oxo-5,6-dihydrophenanthridin-2-yl)-N, N-dimethylacetamide. HCl], a selective inhibitor of poly (ADP-ribose) polymerase, in experimental allergic encephalomyelitis are associated with immunomodulation. J. Pharmacol. Exp. Ther. 2004, 310, 1053–1061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Lehtiö, L.; Jemth, A.-S.; Collins, R.; Loseva, O.; Johansson, A.; Markova, N.; Hammarström, M.; Flores, A.; Holmberg-Schiavone, L.; Weigelt, J.; et al. Structural basis for inhibitor specificity in human poly (ADP-ribose) polymerase-3. J. Med. Chem. 2009, 52, 3108–3111. [Google Scholar] [CrossRef] [PubMed]
  54. Yam-Canul, P.; Chirino, Y.I.; Sánchez-González, D.J.; Martínez-Martínez, C.M.; Cruz, C.; Pedraza-Chaverri, J. PJ34, a Poly Adenosine Diphosphate-Ribose Polymerase Inhibitor, Attenuates Chromate-Induced Nephrotoxicity. Basic Clin. Pharmacol. Toxicol. 2008, 102, 483–488. [Google Scholar] [CrossRef] [PubMed]
  55. Li, T.-K.; Houghton, P.J.; Desai, S.D.; Daroui, P.; Liu, A.A.; Hars, E.S.; Ruchelman, A.L.; LaVoie, E.J.; Liu, L.F. Characterization of ARC-111 as a novel topoisomerase I-targeting anticancer drug. Cancer Res. 2003, 63, 8400–8407. [Google Scholar]
  56. Baunach, M.; Ding, L.; Bruhn, T.; Bringmann, G.; Hertweck, C. Regiodivergent N-C and N-N aryl coupling reactions of indoloterpenes and cycloether formation mediated by a single bacterial flavoenzyme. Angew. Chem. Int. Ed. Engl. 2013, 52, 9040–9043. [Google Scholar] [CrossRef]
  57. Rayadurgam, J.; Sana, S.; Sasikumar, M.; Gu, Q. Palladium catalyzed C–C and C–N bond forming reactions: An update on the synthesis of pharmaceuticals from 2015–2020. Org. Chem. Front. 2021, 8, 384–414. [Google Scholar] [CrossRef]
  58. Siddiqui, M.A.; Snieckus, V. The directed metalation connection to aryl-aryl cross coupling. Regiospecific synthesis of phenanthridines, phenanthridinones and the biphenyl alkaloid ismine. Tetrahedron Lett. 1988, 29, 5463–5466. [Google Scholar] [CrossRef]
  59. Cailly, T.; Fabis, F.; Rault, S. A new, direct, and efficient synthesis of benzonaphthyridin-5-ones. Tetrahedron 2006, 62, 5862–5867. [Google Scholar] [CrossRef]
  60. Péron, F.; Fossey, C.; Cailly, T.; Fabis, F. N-Tosylcarboxamide as a Transformable Directing Group for Pd-Catalyzed C–H Ortho-Arylation. Org. Lett. 2012, 14, 1827–1829. [Google Scholar] [CrossRef]
  61. Wang, G.-W.; Yuan, T.-T.; Li, D.-D. One-Pot Formation of C–C and C–N Bonds through Palladium-Catalyzed Dual C–H Activation: Synthesis of Phenanthridinones. Angew. Chem. Int. Ed. 2011, 50, 1380–1383. [Google Scholar] [CrossRef] [PubMed]
  62. Saha, R.; Sekar, G. Stable Pd-nanoparticles catalyzed domino CH activation/CN bond formation strategy: An access to phenanthridinones. J. Catal. 2018, 366, 176–188. [Google Scholar] [CrossRef]
  63. Karthikeyan, J.; Cheng, C.-H. Synthesis of Phenanthridinones from N-Methoxybenzamides and Arenes by Multiple Palladium-Catalyzed C–H Activation Steps at Room Temperature. Angew. Chem. Int. Ed. 2011, 50, 9880–9883. [Google Scholar] [CrossRef] [PubMed]
  64. Karthikeyan, J.; Haridharan, R.; Cheng, C.H. Rhodium (III)-Catalyzed Oxidative C–H Coupling of N-Methoxybenzamides with Aryl Boronic Acids: One-Pot Synthesis of Phenanthridinones. Angew. Chem. 2012, 124, 12509–12513. [Google Scholar] [CrossRef]
  65. Senthilkumar, N.; Parthasarathy, K.; Gandeepan, P.; Cheng, C.-H. Synthesis of Phenanthridinones from N-Methoxybenzamides and Aryltriethoxysilanes through RhIII-Catalyzed C–H and N-H Bond Activation. Chem. Asian J. 2013, 8, 2175–2181. [Google Scholar] [CrossRef]
  66. Caddick, S.; Kofie, W. Observations on the intramolecular Heck reactions of aromatic chlorides using palladium/imidazolium salts. Tetrahedron Lett. 2002, 43, 9347–9350. [Google Scholar] [CrossRef]
  67. Ferraccioli, R.; Carenzi, D.; Rombola, O.; Catellani, M. Synthesis of 6-Phenanthridinones and Their Heterocyclic Analogues through Palladium-Catalyzed Sequential Aryl–Aryl and N-Aryl Coupling. Org. Lett. 2004, 6, 4759–4762. [Google Scholar] [CrossRef]
  68. Banerji, B.; Chatterjee, S.; Chandrasekhar, K.; Nayan, C.; Killi, S.K. Palladium-Catalyzed Direct Synthesis of Phenanthridones from Benzamides through Tandem N–H/C–H Arylation. Eur. J. Org. Chem. 2017, 2017, 5214–5218. [Google Scholar] [CrossRef]
  69. Li, X.; Pan, J.; Song, S.; Jiao, N. Pd-catalyzed dehydrogenative annulation approach for the efficient synthesis of phenanthridinones. Chem. Sci. 2016, 7, 5384–5389. [Google Scholar] [CrossRef] [Green Version]
  70. Furuta, T.; Yamamoto, J.; Kitamura, Y.; Hashimoto, A.; Masu, H.; Azumaya, I.; Kan, T.; Kawabata, T. Synthesis of axially chiral amino acid and amino alcohols via additive–ligand-free Pd-catalyzed domino coupling reaction and subsequent transformations of the product amidoaza [5] helicene. J. Org. Chem. 2010, 75, 7010–7013. [Google Scholar] [CrossRef]
  71. Donati, L.; Michel, S.; Tillequin, F.; Porée, F.-H. Selective Unusual Pd-Mediated Biaryl Coupling Reactions: Solvent Effects with Carbonate Bases. Org. Lett. 2010, 12, 156–158. [Google Scholar] [CrossRef]
  72. Donati, L.; Leproux, P.; Prost, E.; Michel, S.; Tillequin, F.; Gandon, V.; Porée, F.H. Solvent/Base Effects in the Selective Domino Synthesis of Phenanthridinones That Involves High-Valent Palladium Species: Experimental and Theoretical Studies. Chem. Eur. J. 2011, 17, 12809–12819. [Google Scholar] [CrossRef] [PubMed]
  73. Liu, H.; Han, W.; Li, C.; Ma, Z.; Li, R.; Zheng, X.; Fu, H.; Chen, H. Practical Synthesis of Phenanthridinones by Palladium-Catalyzed One-Pot C–C and C–N Coupling Reaction: Extending the Substrate Scope to o-Chlorobenzamides. Eur. J. Org. Chem. 2016, 2016, 389–393. [Google Scholar] [CrossRef]
  74. Hu, Q.-F.; Gao, T.-T.; Shi, Y.-J.; Lei, Q.; Yu, L.-T. Palladium-catalyzed intramolecular C–H arylation of 2-halo-N-Boc-N-arylbenzamides for the synthesis of N–H phenanthridinones. RSC Adv. 2018, 8, 13879–13890. [Google Scholar] [CrossRef] [Green Version]
  75. Takamatsu, K.; Hirano, K.; Miura, M. Copper-Mediated Decarboxylative Coupling of Benzamides with ortho-Nitrobenzoic Acids by Directed C–H Cleavage. Angew. Chem. Int. Ed. 2017, 56, 1–6. [Google Scholar] [CrossRef]
  76. Li, D.; Xu, N.; Zhang, Y.; Wang, L. A highly efficient Pd-catalyzed decarboxylative ortho-arylation of amides with aryl acylperoxides. Chem. Commun. 2014, 50, 14862–14865. [Google Scholar] [CrossRef] [PubMed]
  77. Lamba, J.J.S.; Tour, J.M. Imine-Bridged Planar Poly(p-phenylene) Derivatives for Maximization of Extended π-Conjugation. The Common Intermediate Approach. J. Am. Chem. Soc. 1994, 116, 11723–11736. [Google Scholar] [CrossRef]
  78. Tanimoto, K.; Nakagawa, N.; Takeda, K.; Kirihata, M.; Tanimori, S. A convenient one-pot access to phenanthridinones via Suzuki–Miyaura cross-coupling reaction. Tetrahedron Lett. 2013, 54, 3712–3714. [Google Scholar]
  79. Kuwata, Y.; Sonoda, M.; Tanimori, S. Facile Synthesis of Phenanthridinone Alkaloids via Suzuki–Miyaura Cross-coupling. J. Heterocycl. Chem. 2017, 54, 1645–1651. [Google Scholar] [CrossRef]
  80. Chen, Z.; Wang, X. A Pd-catalyzed, boron ester-mediated, reductive cross-coupling of two aryl halides to synthesize tricyclic biaryls. Org. Biomol. Chem. 2017, 15, 5790–5796. [Google Scholar]
  81. Gandeepan, P.; Rajamalli, P.; Cheng, C.-H. Palladium-catalyzed C–H activation and cyclization of anilides with 2-iodoacetates and 2-iodobenzoates: An efficient method toward oxindoles and phenanthridones. Synthesis 2016, 48, 1872–1879. [Google Scholar] [CrossRef]
  82. Nealmongkol, P.; Calmes, J.; Ruchirawat, S.; Thasana, N. Synthesis of phenanthridinones using Cu-or Pd-mediated C–N bond formation. Tetrahedron 2017, 73, 735–741. [Google Scholar] [CrossRef]
  83. Ding, X.; Zhang, L.; Mao, Y.; Rong, B.; Zhu, N.; Duan, J.; Guo, K. Synthesis of Phenanthridinones by Palladium-Catalyzed Cyclization of N-Aryl-2-aminopyridines with 2-Iodobenzoic Acids in Water. Synlett 2020, 31, 280–284. [Google Scholar] [CrossRef]
  84. Lu, C.; Dubrovskiy, A.V.; Larock, R.C. Palladium-Catalyzed Annulation of Arynes by ortho-Halobenzamides: Synthesis of Phenanthridinones. J. Org. Chem. 2012, 77, 8648–8656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Pimparkar, S.; Jeganmohan, M. Palladium-catalyzed cyclization of benzamides with arynes: Application to the synthesis of phenaglydon and N-methylcrinasiadine. Chem. Commun. 2014, 50, 12116–12119. [Google Scholar] [CrossRef] [PubMed]
  86. Zhang, T.-Y.; Lin, J.-B.; Li, Q.-Z.; Kang, J.-C.; Pan, J.-L.; Hou, S.-H.; Chen, C.; Zhang, S.-Y. Copper-Catalyzed Selective ortho-C–H/N–H Annulation of Benzamides with Arynes: Synthesis of Phenanthridinone Alkaloids. Org. Lett. 2017, 19, 1764–1767. [Google Scholar] [CrossRef]
  87. Feng, M.; Tang, B.; Wang, N.; Xu, H.X.; Jiang, X. Ligand Controlled Regiodivergent C1 Insertion on Arynes for Construction of Phenanthridinone and Acridone Alkaloids. Angew. Chem. Int. Ed. 2015, 54, 14960–14964. [Google Scholar] [CrossRef]
  88. Yang, Y.; Huang, H.; Wu, L. Palladium-catalyzed annulation of benzynes with N-substituted-N-(2-halophenyl) formamides: Synthesis of phenanthridinones. Org. Biomol. Chem. 2014, 12, 5351–5355. [Google Scholar] [CrossRef] [PubMed]
  89. Feng, M.; Tang, B.; Xu, H.-X.; Jiang, X. Collective Synthesis of Phenanthridinone through C–H Activation Involving a Pd-Catalyzed Aryne Multicomponent Reaction. Org. Lett. 2016, 18, 4352–4355. [Google Scholar] [CrossRef]
  90. Zhao, J.; Li, H.; Li, P.; Wang, L. Annulation of Benzamides with Arynes Using Palladium with Photoredox Dual Catalysis. J. Org. Chem. 2019, 84, 9007–9016. [Google Scholar] [CrossRef]
  91. Liang, Z.; Zhang, J.; Liu, Z.; Wang, K.; Zhang, Y. Pd (II)-catalyzed C (sp2)–H carbonylation of biaryl-2-amine: Synthesis of phenanthridinones. Tetrahedron 2013, 69, 6519–6526. [Google Scholar] [CrossRef]
  92. Liang, D.; Hu, Z.; Peng, J.; Huang, J.; Zhu, Q. Synthesis of phenanthridinones via palladium-catalyzed C (sp2)–H aminocarbonylation of unprotected o-arylanilines. Chem. Commun. 2013, 49, 173–175. [Google Scholar] [CrossRef]
  93. Rajeshkumar, V.; Lee, T.-H.; Chuang, S.-C. Palladium-catalyzed oxidative insertion of carbon monoxide to N-sulfonyl-2-aminobiaryls through C–H bond activation: Access to bioactive phenanthridinone derivatives in one pot. Org. Lett. 2013, 15, 1468–1471. [Google Scholar] [CrossRef] [PubMed]
  94. Nageswar Rao, D.; Rasheed, S.; Das, P. Palladium/Silver Synergistic Catalysis in Direct Aerobic Carbonylation of C(sp2)–H Bonds Using DMF as a Carbon Source: Synthesis of Pyrido-Fused Quinazolinones and Phenanthridinones. Org. Lett. 2016, 18, 3142–3145. [Google Scholar] [CrossRef] [PubMed]
  95. Ling, F.; Zhang, C.; Ai, C.; Lv, Y.; Zhong, W. Metal-Oxidant-Free Cobalt-Catalyzed C (sp2)–H Carbonylation of ortho-Arylanilines: An Approach toward Free (NH)-Phenanthridinones. J. Org. Chem. 2018, 83, 5698–5706. [Google Scholar] [CrossRef] [PubMed]
  96. Zhou, Q.J.; Worm, K.; Dolle, R.E. 10-Hydroxy-10,9-boroxarophenanthrenes: Versatile synthetic intermediates to 3,4-benzocoumarins and triaryls. J. Org. Chem. 2004, 69, 5147–5149. [Google Scholar] [CrossRef] [PubMed]
  97. Wang, S.; Shao, P.; Du, G.; Xi, C. MeOTf-and TBD-mediated carbonylation of ortho-arylanilines with CO2 leading to phenanthridinones. J. Org. Chem. 2016, 81, 6672–6676. [Google Scholar] [CrossRef] [PubMed]
  98. Li, L.; Chen, J.J.; Kan, X.L.; Zhang, L.; Zhao, Y.L.; Liu, Q. One-Pot Synthesis of Phenanthridinones by Using a Base-Catalyzed/Promoted Bicyclization of α, β-Unsaturated Carbonyl Compounds with Dimethyl Glutaconate. Eur. J. Org. Chem. 2015, 2015, 4892–4899. [Google Scholar] [CrossRef]
  99. Gao, Y.; Cai, Z.; Li, S.; Li, G. Rhodium (I)-Catalyzed Aryl C–H Carboxylation of 2-Arylanilines with CO2. Org. Lett. 2019, 21, 3663–3669. [Google Scholar] [CrossRef]
  100. Hussain, I.; Singh, T. Synthesis of Biaryls through Aromatic C–H Bond Activation: A Review of Recent Developments. Adv. Synth. Catal. 2014, 356, 1661–1696. [Google Scholar] [CrossRef]
  101. Hassan, J.; Sévignon, M.; Gozzi, C.; Schulz, E.; Lemaire, M. Aryl–Aryl Bond Formation One Century after the Discovery of the Ullmann Reaction. Chem. Rev. 2002, 102, 1359–1470. [Google Scholar] [CrossRef]
  102. Ames, D.E.; Opalko, A. Palladium-catalysed cyclisation of 2-substituted halogenoarenes by dehydrohalogenation. Tetrahedron 1984, 40, 1919–1925. [Google Scholar] [CrossRef]
  103. Harayama, T.; Kawata, Y.; Nagura, C.; Sato, T.; Miyagoe, T.; Abe, H.; Takeuchi, Y. Effect of oxygen substituents on the regioselectivity of the Pd-assisted biaryl coupling reaction of benzanilides. Tetrahedron Lett. 2005, 46, 6091–6094. [Google Scholar] [CrossRef]
  104. Campeau, L.-C.; Parisien, M.; Jean, A.; Fagnou, K. Catalytic direct arylation with aryl chlorides, bromides, and iodides: Intramolecular studies leading to new intermolecular reactions. J. Am. Chem. Soc. 2006, 128, 581–590. [Google Scholar] [CrossRef]
  105. Bernardo, P.H.; Fitriyanto, W.; Chai, C.L. Palladium-mediated synthesis of calothrixin B. Synlett 2007, 2007, 1935–1939. [Google Scholar] [CrossRef]
  106. Bernini, R.; Cacchi, S.; Fabrizi, G.; Sferrazza, A. A simple general approach to phenanthridinones via palladium-catalyzed intramolecular direct arene arylation. Synthesis 2008, 2008, 729–738. [Google Scholar] [CrossRef]
  107. Roman, D.S.; Takahashi, Y.; Charette, A.B. Potassium tert-butoxide promoted intramolecular arylation via a radical pathway. Org. Lett. 2011, 13, 3242–3245. [Google Scholar] [CrossRef] [PubMed]
  108. Ohno, H.; Iwasaki, H.; Eguchi, T.; Tanaka, T. The first samarium (II)-mediated aryl radical cyclisation onto an aromatic ring. Chem. Commun. 2004, 2004, 2228–2229. [Google Scholar] [CrossRef] [PubMed]
  109. Yeung, C.S.; Zhao, X.; Borduas, N.; Dong, V.M. Pd-catalyzed ortho-arylation of phenylacetamides, benzamides, and anilides with simple arenes using sodium persulfate. Chem. Sci. 2010, 1, 331–336. [Google Scholar] [CrossRef]
  110. Ishida, N.; Nakanishi, Y.; Moriya, T.; Murakami, M. Synthesis of Phenanthridinones and Phenanthridine Derivatives through Palladium-catalyzed Oxidative C–H Coupling of Benzanilides. Chem. Lett. 2011, 40, 1047–1049. [Google Scholar] [CrossRef]
  111. Yu, Q.; Zhang, N.; Tang, Y.; Lu, H.; Huang, J.; Wang, S.; Du, Y.; Zhao, K. Copper(II)-Mediated Cascade Oxidative C–C Coupling and Aromatization: Synthesis of 3-Hydroxyphenanthridinone Derivatives. Synthesis 2012, 44, 2374–2384. [Google Scholar] [CrossRef]
  112. Bao, H.; Hu, X.; Zhang, J.; Liu, Y. Cu(0)/Selectfluor system-catalyzed intramolecular Csp2-H/Csp2-Hcross-dehydrogenative coupling (CDC). Tetrahedron 2019, 75, 130533. [Google Scholar] [CrossRef]
  113. Moreno, I.; Tellitu, I.; Etayo, J.; SanMartín, R.; Domínguez, E. Novel applications of hypervalent iodine: PIFA mediated synthesis of benzo [c] phenanthiridines and benzo [c] phenanthridinones. Tetrahedron 2001, 57, 5403–5411. [Google Scholar] [CrossRef]
  114. Bhakuni, B.S.; Kumar, A.; Balkrishna, S.J.; Sheikh, J.A.; Konar, S.; Kumar, S. KOtBu Mediated Synthesis of Phenanthridinones and Dibenzoazepinones. Org. Lett. 2012, 14, 2838–2841. [Google Scholar] [CrossRef] [PubMed]
  115. Sharma, S.; Kumar, M.; Sharma, S.; Nayal, O.S.; Kumar, N.; Singh, B.; Sharma, U. Microwave assisted synthesis of phenanthridinones and dihydrophenanthridines by vasicine/KO t Bu promoted intramolecular C–H arylation. Org. Biomol. Chem. 2016, 14, 8536–8544. [Google Scholar] [CrossRef] [PubMed]
  116. Yadav, L.; Tiwari, M.K.; Shyamlal, B.R.K.; Chaudhary, S. Organocatalyst in direct C(sp2)–H arylation of unactivated arenes: [1-(2-Hydroxyethyl)-piperazine]-Catalyzed Inter-/Intra-molecular C–H bond activation. J. Org. Chem. 2020, 85, 8121–8141. [Google Scholar] [CrossRef]
  117. Abel-Snape, X.; Whyte, A.; Lautens, M. Synthesis of aminated phenanthridinones via palladium/Norbornene catalysis. Org. Lett. 2020, 22, 7920–7925. [Google Scholar] [CrossRef]
  118. Bariwal, J.; Van der Eycken, E. C–N bond forming cross-coupling reactions: An overview. Chem. Soc. Rev. 2013, 42, 9283–9303. [Google Scholar] [CrossRef]
  119. Inamoto, K.; Saito, T.; Hiroya, K.; Doi, T. Palladium-Catalyzed Intramolecular Amidation of C(sp2)–H Bonds: Synthesis of 4-Aryl-2-quinolinones. J. Org. Chem. 2010, 75, 3900–3903. [Google Scholar] [CrossRef]
  120. Gui, Q.; Yang, Z.; Chen, X.; Liu, J.; Tan, Z.; Guo, R.; Yu, W. Synthesis of Phenanthridin-6(5H)-ones via Copper-Catalyzed Cyclization of 2-Phenylbenzamides. Synlett 2013, 24, 1016–1020. [Google Scholar]
  121. Liang, D.; Sersen, D.; Yang, C.; Deschamps, J.R.; Imler, G.H.; Jiang, C.; Xue, F. One-pot sequential reaction to 2-substituted-phenanthridinones from N-methoxybenzamides. Org. Biomol. Chem. 2017, 15, 4390–4398. [Google Scholar] [CrossRef]
  122. Liang, D.; Yu, W.; Nguyen, N.; Deschamps, J.R.; Imler, G.H.; Li, Y.; MacKerell , A., Jr.; Jiang, C.; Xue, F. Iodobenzene-Catalyzed Synthesis of Phenanthridinones via Oxidative C–H Amidation. J. Org. Chem. 2017, 82, 3589–3596. [Google Scholar] [CrossRef]
  123. Subramanian, K.; Yedage, S.L.; Sethi, K.; Bhanage, B.M. Tetrabutylammonium Iodide (TBAI) Catalyzed Electrochemical C–H Bond Activation of 2-arylated N-methoxyamides for the synthesis of phenanthridinones. Synlett 2021, 32, 999–1003. [Google Scholar]
  124. Wu, L.; Hao, Y.; Liu, Y.; Wang, Q. NIS-mediated oxidative arene C(sp2)–H amidation toward 3,4-dihydro-2(1H)-quinolinone, phenanthridone, and N-fused spirolactam derivatives. Org. Biomol. Chem. 2019, 17, 6762–6770. [Google Scholar] [CrossRef]
  125. Verma, A.; Singh Banjara, L.; Meena, R.; Kumar, S. Transition-Metal-Free synthesis of N-Substituted phenanthridinones and spiro-isoindolinones: C(sp2)–N and C(sp2)–O coupling through radical pathway. Asian J. Org. Chem. 2020, 9, 105–110. [Google Scholar] [CrossRef] [Green Version]
  126. Sen, A.; Dhital, R.N.; Sato, T.; Ohno, A.; Yamada, Y.M. Switching from biaryl formation to amidation with convoluted polymeric nickel catalysis. ACS Catal. 2020, 18, 14410–14418. [Google Scholar] [CrossRef]
  127. Cailly, T.; Fabis, F.; Legay, R.; Oulyadi, H.; Rault, S. The synthesis of three new heterocycles: The pyrido [4,3 or 3,4 or 2,3-c]-1,5-naphthyridines. Tetrahedron 2007, 63, 71–76. [Google Scholar] [CrossRef]
  128. Dubost, E.; Magnelli, R.; Cailly, T.; Legay, R.; Fabis, F.; Rault, S. General method for the synthesis of substituted phenanthridin-6(5H)-ones using a KOH-mediated anionic ring closure as the key step. Tetrahedron 2010, 66, 5008–5016. [Google Scholar] [CrossRef]
  129. Chen, Y.-F.; Wu, Y.-S.; Jhan, Y.-H.; Hsieh, J.-C. An efficient synthesis of (NH)-phenanthridinones via ligand-free copper-catalyzed annulation. Org. Chem. Front. 2014, 1, 253–257. [Google Scholar] [CrossRef]
  130. Chen, W.-L.; Jhang, Y.-Y.; Hsieh, J.-C. Copper-catalyzed intramolecular C–N coupling reaction of aryl halide with amide. Res. Chem. Intermed. 2017, 43, 3517–3526. [Google Scholar] [CrossRef]
  131. Fan-Chiang, T.-T.; Wang, H.-K.; Hsieh, J.-C. Synthesis of phenanthridine skeletal Amaryllidaceae alkaloids. Tetrahedron 2016, 72, 5640–5645. [Google Scholar] [CrossRef]
  132. Wu, M.-J.; Lin, C.-F.; Chen, S.-H. Double anionic cycloaromatization of 2-(6-substituted-3-hexene-1, 5-diynyl) benzonitriles initiated by methoxide addition. Org. Lett. 1999, 1, 767–768. [Google Scholar] [CrossRef]
  133. Wu, M.-J.; Lin, C.-F.; Lu, W.-D. Anionic Cycloaromatization of 1-Aryl-3-hexen-1,5-diynes Initiated by Methoxide Addition:  Synthesis of Phenanthridinones, Benzo[c]phenanthridinones, and Biaryls. J. Org. Chem. 2002, 67, 5907–5912. [Google Scholar] [CrossRef]
  134. Wang, Q.; Su, Y.; Li, L.; Huang, H. Transition-metal catalysed C–N bond activation. Chem. Soc. Rev. 2016, 45, 1257–1272. [Google Scholar] [CrossRef] [Green Version]
  135. Shen, Z.; Ni, Z.; Mo, S.; Wang, J.; Zhu, Y. Palladium-Catalyzed Intramolecular Decarboxylative Coupling of Arene Carboxylic Acids/Esters with Aryl Bromides. Chem. Eur. J. 2012, 18, 4859–4865. [Google Scholar] [CrossRef]
  136. Yuan, M.; Chen, L.; Wang, J.; Chen, S.; Wang, K.; Xue, Y.; Yao, G.; Luo, Z.; Zhang, Y. Transition-metal-free synthesis of phenanthridinones from biaryl-2-oxamic acid under radical conditions. Org. Lett. 2015, 17, 346–349. [Google Scholar] [CrossRef]
  137. Chen, J.-R.; Hu, X.-Q.; Lu, L.-Q.; Xiao, W.-J. Exploration of Visible-Light Photocatalysis in Heterocycle Synthesis and Functionalization: Reaction Design and Beyond. Acc. Chem. Res. 2016, 49, 1911–1923. [Google Scholar] [CrossRef] [PubMed]
  138. Mondon, A.; Krohn, K. Synthese des Narciprimins und verwandter Verbindungen. Chem. Ber. 1972, 105, 3726–3747. [Google Scholar] [CrossRef]
  139. Prabhakar, S.; Lobo, A.M.; Tavares, M.R. Boron complexes as control synthons in photocyclisations: An improved phenanthridine synthesis. J. Chem. Soc. Chem. Commun. 1978, 1978, 884–885. [Google Scholar] [CrossRef]
  140. Todorov, A.R.; Wirtanen, T.; Helaja, J. Photoreductive removal of O-benzyl groups from oxyarene N-heterocycles assisted by O-pyridine–pyridone tautomerism. J. Org. Chem. 2017, 82, 13756–13767. [Google Scholar] [CrossRef]
  141. Moon, Y.; Jang, E.; Choi, S.; Hong, S. Visible-light-photocatalyzed synthesis of phenanthridinones and quinolinones via direct oxidative C–H amidation. Org. Lett. 2018, 20, 240–243. [Google Scholar] [CrossRef] [PubMed]
  142. Usami, K.; Yamaguchi, E.; Tada, N.; Itoh, A. Transition-Metal-Free Synthesis of Phenanthridinones through Visible-Light-Driven Oxidative C–H Amidation. Eur. J. Org. Chem. 2020, 2020, 1496–1504. [Google Scholar] [CrossRef]
  143. Yaragorla, S.; Pareek, A.; Dada, R. Cycloisomerization of Oxindole-Derived 1,5-Enynes: A Calcium(II)-Catalyzed One-Pot, Solvent-free Synthesis of Phenanthridinones, 3-(Cyclopentenylidene)indolin-2-ones and 3-Spirocyclic Indolin-2-ones. Adv. Synth. Catal. 2017, 359, 3068–3075. [Google Scholar] [CrossRef]
  144. Alzaydi, K.M.; Abojabal, N.S.; Elnagdi, M.H. Multicomponent reactions in Q-Tubes™: One-pot synthesis of benzo[c]chromen-6-one and phenanthridin-6(5H)-one derivatives in a four-component reaction. Tetrahedron Lett. 2016, 57, 3596–3599. [Google Scholar] [CrossRef]
  145. Zou, S.; Zhang, Z.; Chen, C.; Xi, C. MeOTf-Catalyzed intramolecular acyl-cyclization of aryl isocyanates: Efficient access to phenanthridin-6(5H)-one and 3, 4-Dihydroisoquinolin-1(2H)-one derivatives. Asian J. Org. Chem. 2021, 10, 355–359. [Google Scholar] [CrossRef]
Scheme 1. Selected examples of biologically important phenanthridinones.
Scheme 1. Selected examples of biologically important phenanthridinones.
Molecules 26 05560 sch001
Scheme 2. Phenanthridinone preparation from N-methoxybenzamides.
Scheme 2. Phenanthridinone preparation from N-methoxybenzamides.
Molecules 26 05560 sch002
Scheme 3. Palladium-catalysed phenanthridinones from 2-bromobenzamides and aryl iodide.
Scheme 3. Palladium-catalysed phenanthridinones from 2-bromobenzamides and aryl iodide.
Molecules 26 05560 sch003
Scheme 4. Palladium-catalyzed homocoupling of 2-bromobenzamides.
Scheme 4. Palladium-catalyzed homocoupling of 2-bromobenzamides.
Molecules 26 05560 sch004
Scheme 5. Phenanthridinone preparation using benzamides and o-nitrobenzoic acids.
Scheme 5. Phenanthridinone preparation using benzamides and o-nitrobenzoic acids.
Molecules 26 05560 sch005
Scheme 6. Palladium-catalyzed phenanthridinones from 2-halobenzoate and 2-aminophenylboronic acid.
Scheme 6. Palladium-catalyzed phenanthridinones from 2-halobenzoate and 2-aminophenylboronic acid.
Molecules 26 05560 sch006
Scheme 7. Phenanthridinone preparation from N-aryl-2-aminopyridines and 2-iodobenzoic acids.
Scheme 7. Phenanthridinone preparation from N-aryl-2-aminopyridines and 2-iodobenzoic acids.
Molecules 26 05560 sch007
Scheme 8. Benzyne-involved preparations of phenathridinones.
Scheme 8. Benzyne-involved preparations of phenathridinones.
Molecules 26 05560 sch008
Scheme 9. Proposed mechanism for the palladium-catalyzed aryne mediate cyclization.
Scheme 9. Proposed mechanism for the palladium-catalyzed aryne mediate cyclization.
Molecules 26 05560 sch009
Scheme 10. Pd-catalyzed carbonylative synthesis of phenanthridinones.
Scheme 10. Pd-catalyzed carbonylative synthesis of phenanthridinones.
Molecules 26 05560 sch010
Scheme 11. Transition metal free carboxylative cyclization of o-arylanilines.
Scheme 11. Transition metal free carboxylative cyclization of o-arylanilines.
Molecules 26 05560 sch011
Scheme 12. Rh-catalyzed carboxylative cyclization of o-arylanilines.
Scheme 12. Rh-catalyzed carboxylative cyclization of o-arylanilines.
Molecules 26 05560 sch012
Scheme 13. Biaryl dehydrogenative coupling.
Scheme 13. Biaryl dehydrogenative coupling.
Molecules 26 05560 sch013
Scheme 14. Proposed mechanism for the biaryl dehydrogenative coupling.
Scheme 14. Proposed mechanism for the biaryl dehydrogenative coupling.
Molecules 26 05560 sch014
Scheme 15. Transition metal free aryl arene coupling of halo benzamides for phenanthridinone synthesis.
Scheme 15. Transition metal free aryl arene coupling of halo benzamides for phenanthridinone synthesis.
Molecules 26 05560 sch015
Scheme 16. Rh-catalyzed carboxylative cyclization of o-arylanilines.
Scheme 16. Rh-catalyzed carboxylative cyclization of o-arylanilines.
Molecules 26 05560 sch016
Scheme 17. C–H amidations for the synthesis of phenathridinones.
Scheme 17. C–H amidations for the synthesis of phenathridinones.
Molecules 26 05560 sch017
Scheme 18. NIS mediated oxidative arene C(sp2)–H amidation.
Scheme 18. NIS mediated oxidative arene C(sp2)–H amidation.
Molecules 26 05560 sch018
Scheme 19. Transformations of 2-arylbenzonitriles.
Scheme 19. Transformations of 2-arylbenzonitriles.
Molecules 26 05560 sch019
Scheme 20. Transition metal free synthesis of phenanthridinones from biaryl-2-oxamic acids.
Scheme 20. Transition metal free synthesis of phenanthridinones from biaryl-2-oxamic acids.
Molecules 26 05560 sch020
Scheme 21. Visible light-induced synthesis of phenanthridinones through direct oxidative C–H amidation.
Scheme 21. Visible light-induced synthesis of phenanthridinones through direct oxidative C–H amidation.
Molecules 26 05560 sch021
Scheme 22. Visible light-induced synthesis of phenanthridinones in the presence of 1- chloroanthraquinone catalyst.
Scheme 22. Visible light-induced synthesis of phenanthridinones in the presence of 1- chloroanthraquinone catalyst.
Molecules 26 05560 sch022
Scheme 23. Ca(II)-catalyzed oxidative one-pot domino sequence towards phenathridinones.
Scheme 23. Ca(II)-catalyzed oxidative one-pot domino sequence towards phenathridinones.
Molecules 26 05560 sch023
Scheme 24. Multicomponent approach towards phenathridinones.
Scheme 24. Multicomponent approach towards phenathridinones.
Molecules 26 05560 sch024
Scheme 25. Domino synthesis of phenathridinones from dimethyl glutaconate.
Scheme 25. Domino synthesis of phenathridinones from dimethyl glutaconate.
Molecules 26 05560 sch025
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Aleti, R.R.; Festa, A.A.; Voskressensky, L.G.; Van der Eycken, E.V. Synthetic Strategies in the Preparation of Phenanthridinones. Molecules 2021, 26, 5560. https://doi.org/10.3390/molecules26185560

AMA Style

Aleti RR, Festa AA, Voskressensky LG, Van der Eycken EV. Synthetic Strategies in the Preparation of Phenanthridinones. Molecules. 2021; 26(18):5560. https://doi.org/10.3390/molecules26185560

Chicago/Turabian Style

Aleti, Rajeshwar Reddy, Alexey A. Festa, Leonid G. Voskressensky, and Erik V. Van der Eycken. 2021. "Synthetic Strategies in the Preparation of Phenanthridinones" Molecules 26, no. 18: 5560. https://doi.org/10.3390/molecules26185560

Article Metrics

Back to TopTop