Next Article in Journal
The Leaching Kinetics of Iron from Titanium Gypsum in a Citric Acid Medium and Obtain Materials by Leaching Liquid
Next Article in Special Issue
Design, Synthesis, In Vitro, and In Silico Insights of 5-(Substituted benzylidene)-2-phenylthiazol-4(5H)-one Derivatives: A Novel Class of Anti-Melanogenic Compounds
Previous Article in Journal
Ag–NHC Complexes in the π-Activation of Alkynes
Previous Article in Special Issue
Recent Advances on Small-Molecule Antagonists Targeting TLR7
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Survey on the Synthesis of Variolins, Meridianins, and Meriolins—Naturally Occurring Marine (aza)Indole Alkaloids and Their Semisynthetic Derivatives †

Institut für Organische Chemie und Makromolekulare Chemie, Heinrich-Heine-Universität Düsseldorf, Universitätsstrasse 1, D-40225 Düsseldorf, Germany
*
Author to whom correspondence should be addressed.
Dedicated to Prof. Dr. Uwe H. F. Bunz on the occasion of his 60th birthday.
Molecules 2023, 28(3), 947; https://doi.org/10.3390/molecules28030947
Submission received: 19 December 2022 / Revised: 15 January 2023 / Accepted: 16 January 2023 / Published: 18 January 2023

Abstract

:
Marine natural products are a source of essential significance due to a plethora of highly diverse biological properties. The naturally occurring (aza)indole alkaloids variolin B (1), meridianins (2), and their synthetic hybrids meriolins (3) exhibit potent kinase inhibitory activities and have aroused considerable interest in the past two decades. Therefore, the immense demand for versatile synthetic accesses to these structures has considerably increased. This review surveys the synthetic pathways to these naturally occurring alkaloids and their semisynthetic derivatives.

Graphical Abstract

1. Introduction

Marine ecosystems are an enormous source of chemical diversity. The biological environment is very versatile, and the flora and fauna are exposed to enormous environmental pressure, not only including changes in pressure, salinity, or temperature [1], but also in competition for space, deterrence of predation, and species survival [2]. This led to the development of secondary metabolites in a unique composition, with exceptional frameworks and complex substitution such as halogenation, which rarely occurs in land-based ecosystems [3]. Marine alkaloids, especially indole alkaloids, are of special interest and are produced by a variety of marine sources such as sponges, tunicates, red alga, acorn worms, and symbiotic bacteria [4]. Most isolated alkaloids are biologically active and represent relevant lead structures for pharmacological and medicinal research. The biological properties are diverse with significant antibacterial, antifungal, antiviral, antiparasitic, antitumor, anti-inflammatory, antioxidant, opioid receptor agonistic, antiplasmodial, glucose uptake stimulatory, or immunomodulatory activities [1,5,6].
There is no doubt that researchers are extremely captivated by having access to active molecules. Since the amount of naturally produced alkaloids is low, the focus is on the synthesis of (marine) natural products and analogs for biological screenings and potential applications.
Variolins are a class of compounds isolated from the Antarctic sponge Kirkpatrickia varialosa. In 1994, Blunt, Munro and coworkers reported the isolation and structural elucidation of four pyridopyrrolopyrimidine alkaloids, variolins A, B, D, and N(3′)-methyl tetrahydrovariolin B (Figure 1) [7,8].
Variolins exhibit biological activity, while variolin B (1) was the most active compound. It demonstrated in vitro antitumor activity against the P388 murine leukemia cell line and antiviral activity against Herpes simplex Type I and Polio Type I viruses, whereas variolin A demonstrated weak in vitro activity against the P388 cell line. N(3′)-methyl tetrahydrovariolin B was found to be antifungal against Saccharomyces cerevisiae and cytotoxic against the HCT 116 human colon cancer cell line. Variolin D, however, did not show any activity, indicating that the 2-aminopyrimidine substituent is necessary for the activity [7,8].
Different representatives of marine natural products are the indole alkaloids meridianins A to G, isolated from the tunicate Aplidium meridianum near the South Georgia Islands. Palermo and coworkers first described the isolation of meridianins A to E in 1998 and later published the identification of two further alkaloids, meridianin F and G [9,10]. Meridianins are brominated indole alkaloids (except for unsubstituted meridianin G) with a 2-aminopyrimidine substituent in 3-position on the indole core (Figure 2).
Meridianins 2 are structurally related to the variolin family. Not surprisingly, meridianins demonstrate a plethora of biological activities. They display cytotoxic behavior against several cell lines such as LMM3 (murine mammary adenocarcinoma) [9], PTP (foreskin fibroblast cell line), Hep2 (larynx carcinoma), HT29 (colon carcinoma), RD (rabdomyocarcinoma), U937 (myeloid leukemia) [11], vero (monkey kidney fibroblasts), HEPG2 (human hepatoma cells), or LLC-PK1 (pig kidney epithelial cells) [12]. Meridianins appeared to be potent inhibitors of various kinases, with some derivatives inhibiting cyclin-dependent kinases (CDKs) specifically CDK1/cyclin B and CDK5/p25, as well as protein kinases GSK3-β (glycogen synthase kinase-3β isoform), PKA, PKG, and CK1 in a low micromolar range, as well as cAMP- and cGMP-dependent protein kinases. Some derivatives prevent cell proliferation and induce apoptosis [11]. The derivatives with bromine substituents (meridianins B to F (2bf)) tend to be more active, with meridianin E (2e), having a bromine substituent in the 7-position, being the most active compound. Moreover, comparing meridianin B (2b) and D (2d), as well as meridianin A (2a) and G (2g), a hydroxy group in the 4-position, provides a better inhibitory activity [11]. Meridianin C (2c) and G (2g) showed antiplasmodial activity against Plasmodium falciparum, antiparasitic activity against Leishmania donovani, and antifungal activity against Cryptococcus neoformans [12], while some synthetic analogs display an even more diverse spectrum of biological properties, such as antimalarial and antitubercular [12,13].
Since variolins and meridianins share structural similarities, concerning the 2-aminopyrimidine substituent and the (aza)indole core, as well as their ability to inhibit protein kinases, it was evaluated that a synthetic hybrid structure of variolin B (1) and meridianin G (2g) would give an interesting lead structure for further biological explorations. The structures of this new compound class were named meriolins 3 (Figure 3).
Meriolin 1 (3a) has first been synthesized by Fresneda und Molina [15], while the name meriolin was shaped by Meijer and co-workers [14,16,17]. The 2-aminopyrimidine-substituted 7-azaindoles exhibit pronounced cytotoxicity against several selected cancer cell lines with IC50 values in the nanomolar range [18,19,20] and initiate cell death hallmarks such as cell cycle arrest and decrease in the mitochondrial membrane potential, ΔΨm. Meriolins activate intrinsic apoptotic pathways as well as nonapoptotic cascades leading to necrosis [17,18,21]. Just like the parental natural products, meriolins potently inhibit a broad range of CDKs and appear to be even more active than variolins and meridianins in vitro and in vivo [16,17,18,19,20,22,23,24,25,26].
Due to this versatile range of biological properties and potential pharmaceutical applications, both variolin B (1) and meridianin E (2e) as well as meriolin 3 derivatives have been submitted to preclinical studies. The CDK-inhibiting active compounds are investigated for the treatment of different types of cancer, including colon cancer or murine and myeloid leukemia [27].
For a detailed description of the biological properties, relevant review articles can be found in the literature [3,22,28]. There are overviews of some syntheses of meridianins and analogs [29,30,31] or variolins [32]. Just recently, Xiao published a brief survey on meridianin syntheses and their biological properties [33]. In 2009, Morris and coworkers summarized syntheses of variolins, meridianins, and meriolins and their biological properties in a review article [34]. Presently, over ten years later, we present an updated and detailed, and, to the best of our knowledge, comprehensive collection of the syntheses of the natural products 1 and 2, as well as the semisynthetic compounds 3. For the literature search, Google Scholar, SciFinder, Reaxys, and Web of Science have been used.

2. Synthesis

2.1. Syntheses of Variolins

With their unique pyridopyrrolopyrimidine core and pronounced biological activity, variolins, especially variolin B (1), were both engaging and challenging for synthetic chemists to provide access to this compound class for biological investigations. Prework has been conducted by Alvarez and coworkers [35,36,37], as well as Vaquero [38,39], Molina and Fresneda [40], and Morris and Anderson [41], who published syntheses towards the pyridopyrrolopyrimidine core. To date, the biosynthesis of variolins is not elucidated.

2.1.1. First Total Synthesis by Morris and Anderson

The first total synthesis of variolin B (1) was achieved by Morris and Anderson in 2001 [42]. Later in 2005, they published the full details of their synthetic strategy together with the synthesis of the synthetic analog desoxyvariolin B [43]. They recognized the C2-symmetry of intermediate 8, which is cyclized to the pyridopyrrolopyrimidine in the following key step. After halogen lithium exchange in the methylthiopyrimidine 4, the reaction with diethyl carbonate (5) gave the symmetric ketone 6. The reaction with the lithiated pyridine 7, followed by the key step tandem deoxygenation and cyclization in the presence of triethylsilane and TFA led to the variolin core structure 9. The introduction of the amino groups was achieved by oxidizing the dimethylthiol 9 with m-chloroperbenzoic acid (mCPBA) to the corresponding disulfoxide, which was reacted with p-methoybenzylamine (PMB amine) (10) to give the bisprotected amine 11. Demethylation of 11 and removal of the PMB protecting groups gave the trifluoroacetate salt of the title compound, which was neutralized with concentrated ammonia to give variolin B (1) in an eight-step synthesis and an overall yield of 11% (Figure 4) [42].

2.1.2. Synthesis by Molina and Fresneda

The next synthetic approach was conducted by Molina and Fresneda, who published their syntheses of 1 in 2002 [44] and a modified synthetic route together with the synthesis of an analog in 2003 [45]. This approach starts with the synthesis of the 7-azaindole 16. Aldehyde 13 was condensed with azidoacetate 14 and the resulting vinyl azide 15 cyclized to azaindole 16 via a nitrene insertion. After N-protection with 2-(trimethylsilyl)ethoxymethyl (SEM), the chloride key intermediate 19 was synthesized in a two-step procedure (Figure 5).
Next, two different approaches are reported (Figure 6). Aldehyde 19 was similarly condensed as aldehyde 13 to give vinyl azide 20. After N-SEM-deprotection, a Staudinger reaction with triphenylphosphane led to iminophosphorane 21 in a one-pot reaction. Reaction with benzyl isocyanate (22) in the key aza-Wittig reaction gave a non-isolable carbodiimide that subsequently cyclized to the desired pyridopyrrolopyrimidine moiety 23. The cyclization is highly regioselective. In contrast to β-(indol-2-yl)vinyl heterocumulenes, in the present reaction step, no electrocyclic ring closure to the γ-carboline occurred. Instead, the heterocycle is formed by a nucleophilic attack of the amino group on the central carbon atom of the heterocumulene.
Molina and Fresneda developed a second approach to obtain the tricyclic variolin core without the ester group at C-7. After the N-SEM-deprotection of 19, a nitroaldol condensation with nitromethane led to the formation of 25. Treatment with lithium aluminum hydride gave the corresponding 2-(2-aminoethyl)-7-azaindole, which was sequentially converted to the urea derivative 26 with benzyl isocyanide (22) without isolation. The 26 was dehydrated to the carbodiimide, which subsequently cyclized to the dihydropyrimidine 27 using the Appel reagent (CCl4/PPh3/NEt3). Applying both synthetic approaches, an oxygen substituent is placed at C-4 and a nitrogen substituent at C-9. The next step was to introduce the 2-aminopyrimidine ring at C-5, consequently leading both approaches to the acylated intermediate 31. The reaction of 23 with phosphorus oxychloride and N,N-dimethylacetamide (DMA) (28) allowed the direct introduction of an acetyl group at C-5. Ester hydrolysis led to the carbonic acid 30, and the thermal treatment forced the formation of intermediate 31 by decarboxylation. The route starting from 27 began with the introduction of a bromine substituent at C-5 and the reaction of bromine 32 with n-tributyltin(1-ethoxyvinyl)stannane (33) in the presence of dichlorobis (triphenylphosphine)-palladium(II) introduced to the acetyl group at C-5. Oxidation with 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) gave the intermediate 31 (Figure 7).
The 2-aminopyrimidine substituent was synthesized using a protocol developed by Bredereck (Figure 8) [46]. Enaminone 36 was synthesized from 31 with N,N’-dimethylformamide di-tert-butylacetal (35) in DMF. Condensation with guanidine hydrochloride (37) led to ring closure and formed the desired 2-aminopyrimidine 38. After demethylation of the hydroxy group at C-4 and the amino group at C-9, variolin B (1) was obtained in a 13-step synthesis and an overall yield of 7% [44,45].

2.1.3. Variolin B Approach by Alvarez

In 2003, Alvarez published the synthesis of variolin B (1) and the synthetic analog desoxyvariolin B [47,48,49]. Starting from 4-methoxy-7-azaindole (40), a lithium carboxylate was used as an N-protecting group as well as an ortho-directing substituent to form a 2-lithio-7-azaindole with a protocol by Katritzky [50]. Reaction with 2-(1,3-dioxoisindolin-2-yl)acetaldehyde (41) gave the alcohol 42 that was protected with dihydropyran. N-deprotection of 43 by hydrazinolysis gave the aminoacetal 44. Ring closure was achieved by the reaction with N-tosylcarbonimidic dichloride (45) and diisopropylethylamine (DIPEA) giving 46 in a diasteriometric mixture in a ratio of 1:1. Removal of the O-tetrahydropyran (THP) protecting group and elimination of the resulting hydroxy group by the formation of its mesylate and treatment with triethylamine afforded the pyridopyrrolopyrimidine scaffold (48). Regioselective iodination with N-iodosuccinimide (NIS) gave the key intermediate 49 (Figure 9).
A Stille reaction of 49 and 2-acetylamino-4-trimethylstannylpyrimidine (50) in the presence of tris(dibenzylideneacetone)dipalladium(0) afforded 51. The O-demethylation and N-acetyl-deprotection were achieved by the treatment of 51 with hydrobromic acid, and after reductive photolysis with hydrazine as a reducing agent and 1,4-dimethoxybenzene as an electron source, the tosyl group was cleaved to give variolin B (1) in a 10-step synthesis with an overall yield of 1% (Figure 10).

2.1.4. Synthesis of Variolin B by Burgos and Vaquero

The 2008 approach by Burgos and Vaquero to synthesize variolin B (1) followed the strategy to design the highly functionalized trihalo-substituted pyridopyrrolopyrimidine core 55 and introduce the substituents via palladium-mediated cross-coupling reactions [51,52]. The functionalized 7-azaindole 53 was synthesized from 7-azaindole in six single steps [39]. The 53 was reacted with N-tosylmethyl dichloroformimide (54) under phase-transfer conditions in the two-phase system LiOH (aq., 30%)/CH2Cl2 (1:1) with tetrabutylammonium chloride to give the trihalo-substituted compound 55. The C-9 amino substituent was introduced by a palladium-mediated C-N bond formation, using lithium bis(trimethylsilyl)amide (LiHMDS) and triphenylsilylamine as an ammonia source. The reaction required the use of the ligand [1,1′-biphenyl]-2-yldi-tert-butylphosphane (JohnPhos). After N-acetyl-protection, 56 was obtained (Figure 11).
Next, in a debromination-iodination process, tris(trimethylsilyl)silane (TTMSS) and azobisisobutyronitrile (AIBN) and subsequently NIS were used to exchange the bromo compound 56 to the more reactive iodo derivative 57. In a palladium-catalyzed cross-coupling reaction with the pyrimidyl stannyl reagent 58, the C-C bond at C-5 was formed and the deprotection of both amino groups led to 59. Then, in a palladium-promoted C-O coupling microwave (MW) reaction with sodium tert-butoxide, the tert-butyl group was introduced at C-4 to give the tert-butyl ether 60, and in a final step, the tert-butyl moiety was cleaved to give variolin B (1) (Figure 12). Starting from 53, variolin B was synthesized in seven steps with an overall yield of 5% [51,52].

2.2. Syntheses of Meridianins

Shortly after their discovery in 1998, it was demonstrated that meridianins exhibit strong protein kinase inhibitory properties [9,11]. Therefore, chemists quickly tried to approach the indolyl-2-aminopyridine scaffold to investigate the biological properties and for derivatization. The publication of meridianins F and G followed in 2007 [10]. Just like in the case of variolins, the biosynthetic pathway of meridianins has not been published yet.

2.2.1. First Total Synthesis of Meridianins D and G by Jiang and Yang

In the early 2000s, Jiang and Yang published a straightforward synthesis of meridianins D and G. Starting from the corresponding indolyl boronic acid, 61 with 4-chloropyrimidine 62a is the key reaction in this meridianin synthesis to furnish protected meridians 63 (Figure 13). After N-tosyl-deprotection of compounds 63 with sodium hydroxide, meridianin G (2g) is obtained in an overall yield of 63%, and meridianin D (2d) in an overall yield of 40% in this two-step synthesis [53]. Since meridianin G did not appear in the literature until 2004, Jiang and Yang referred to it as debromomeridianin D.

2.2.2. Synthesis of Meridianins by Fresneda and Molina

Shortly after the publication of the first meridianin syntheses, Fresneda and Molina developed a facile two-step synthesis of meridianins, starting from N-tosyl-3-acetylindoles 64. The reaction of 64 with dimethylformamide dimethylacetal (DMF-DMA) gave the enaminone 66. The key step was the formation of the 2-aminopyrimidine ring by condensation of 66 with guanidine hydrochloride (37). Molina and Fresneda described the synthesis of meridianin D (2d, 65% overall yield) as well as the first total synthesis of meridianin C (2c, 59% overall yield) and O-benzyl-protected derivative 2h. After O-deprotection and dehalogenation of 2h with hydrogen and palladium on carbon, meridianin A was synthesized for the first time (2a, 31% overall yield) or respectively by treating 2h with the milder deprotecting agent; no dehalogenation occurred to give the first total synthesis of meridianin E (2e, 24% overall yield) (Figure 14) [15,40].
To date, this method by Molina and Fresneda is the most commonly used strategy to prepare meridianin derivatives. Radwan and El-Sherbiny synthesized G amongst synthetic cyano and amidrazone analogs and investigated their cytotoxic properties [54]. Simon and Corbel used the described method to synthesize meridianins C and D together with N-H and N-alkylated analogs [55].
The first synthesis of meridianin F was achieved by Sperry [56] and in the same year, the synthesis of meridianin A was published by Baker alongside some analogs that were tested for CNS and antimalarial activity [57]. Bharate and Vishwakarma synthesized meridianins C and G and analogs and their antimalarial, antileishmanial, antibacterial, and antifungal activities were evaluated [12]. Hong and Lee used the protocol for the synthesis of meridianin C and derivatives substituted at the C-5 position on the indole core for the investigation of antiproliferative and pim kinase inhibitory activity [58]. A series of N-1 and C-5 functionalized derivatives, as well as meridianins C, D, and G have been synthesized by Wang, Liu, and Wang and studied for their antiviral and fungicidal properties [59], whereas Zhou and Chen synthesized meridianin C and derivatives that have been functionalized at the 2-aminopyrimidine moiety, to elucidate kinase inhibitory properties [60].

2.2.3. Meridianin Synthesis by Müller via Carbonylative Alkynylation

In 2005, Karpov et al. published a concise synthesis of meridianins C, D, and G. Tert-butoxycarbonyl(Boc)-protected indoles (67) reacted in a palladium-catalyzed three-component carbonylative alkynylation with TMS-protected acetylene (TMSA) (68) to the TMS-alkynones 69. Subsequent cyclocondensation with guanidine (37) concluded the meridianin synthesis as both the TMS- and the Boc-group are cleaved under the chosen reaction condition. Meridianin G was obtained with an overall yield of 45%, and meridianin C and D could be isolated with 50% overall yield (Figure 15). By choosing substituted alkynes, the C-6 of the 2-aminopyridine moiety can be easily functionalized [61]. This, and the higher yields, are favorable towards the enaminone pathway by Molina and Fresneda.

2.2.4. Meridianin Synthesis by Penoni via Indolozation of Nitrosoarenes

Efficiently, Penoni approached the synthesis of meridianins C and G. In a one-pot process, the corresponding nitrosobenzene 70 was reacted with 2-amino-4-ethynylpyrimidine (71) to give the natural products 2c (28%) and 2g (41%) (Figure 16) [62]. Penoni and coworkers could use their method to synthesize derivatives that are functionalized on the indole moiety as well as the pyrimidine substituent.

2.2.5. Synthesis of Meridianins via One-Pot Masuda Borylation-Suzuki Coupling Sequence by Müller

In 2011 and 2022, Müller and coworkers published a different synthetic strategy addressing meridianins. In a palladium-catalyzed Masuda borylation-Suzuki coupling (MBSC), one-pot procedure meridianins C, D, F and G, as well as the meridianin A precursor O-methyl meridianin A (2i), could be synthesized. 3-Iodo-N-protected indoles 67 react with pinacolyl borane (HBpin) (72) and without the isolation of the resulting boronic acid ester, the subsequent Suzuki coupling with 2-aminopyridine (62a) leads to the formation of O-methyl meridianin A (2i) and meridianins D (2d) and G (2g). The Boc-protecting group is cleaved under the Suzuki conditions. In contrast to this, N-tosyl-protected indoles 73 require an additional deprotecting step that can be implemented in the one-pot process. Treatment with potassium hydroxide leads to meridianins C (2c), F (2f), and G (2g) (Figure 17). The additional deprotecting step does not influence the overall yields. The synthesized examples have been isolated with yields ranging from 25 to 80% [63,64].
When O-methyl meridianin A (2i) was melted with pyridinium hydrochloride, the natural product 2a could be isolated in 85% (Figure 18) [63].

2.2.6. Synthesis of Meridianin F by Grainger

Grainger and coworkers worked on the regioselective dibromination of methyl indole-3-carboxylate and its application in the synthesis of indole building blocks. In this context, the synthesis of meridianin F was performed. Dibrominated indole 74 was reacted with N,O-dimethylhydroxylamine (75) to form the corresponding Weinreb amide 76. Treatment with lithium(trimethylsilyl)acetylide (77) led to the formation of alkynone 78. In the aftermath, meridianin F (2f) was furnished by cyclocondensation with guanidine (37) according to the aforementioned protocol by Müller (Section 2.2.3) in an overall yield of 37% (Figure 19) [61,65].

2.2.7. Domino Amino-Palladation Reaction for the Synthesis of Meridianins C and G by Morris

Morris and coworkers came up with a modified Cacchi protocol [66] to synthesize meridianins from readily available monocyclic precursors in a catalytic domino amino-palladation reaction [67]. The four-step synthesis starts with a Sonogashira coupling of 2-iodoaniline 79 and TMSA (68) to give 2-alkynyl anilines 80 followed by N-mesylation to give the activated sulfonamide 82. The reaction of 82 with the N-Boc-protected 4-iodo-2-aminopyrimidine 83 in a Cacchi-type protocol led to the formation of the protected meridianin precursors 84. The global deprotection was achieved in a one-pot acid/base process and furnished meridianins C (2c, 31% overall yield) and G (2g, 45% overall yield) in four steps (Figure 20). This protocol was also suitable for the generation of synthetic analogs.

2.3. Syntheses of Meriolins

Since variolins and meridianins not only share similar structural motifs, both natural product classes also exhibit biological activity, especially strong kinase inhibitory properties. Therefore, meriolins (3) have been rationally designed and can be understood as a hybrid structure of variolin B (1) and meridianins (2).

2.3.1. First Synthesis of Meriolin 1 by Molina and Fresneda

With their protocol for the synthesis of meridianins in hand, Molina and Fresneda were able to extend their strategy to 7-azaindoles, leading to the first synthesis of meriolin 1 (3a). 7-Azaindole (85a) was treated with acetyl chloride (86) in the presence of tin (IV) tetrachloride, which afforded 3-acetyl-7-azaindole (87). After N-tosyl-protection, the enaminone 90 was furnished after the reaction of 89 with DMF-DMA (65) similarly to the meridianin synthesis. Cyclocondensation with guanidine (37) led to the 2-aminopyrimidine formation and meriolin 1 (3a) was obtained in the 16% overall yield (Figure 21) [15].

2.3.2. Synthesis of Meriolin Derivatives by Joseph and Meijer

Joseph, Meijer, and coworkers used the strategy by Molina and Fresneda for the synthesis of meriolin 1 and in addition to that were able to generate a large substance library of novel meriolin derivatives. Starting from substituted 7-azaindoles 85, acylation in 3-position was achieved by treatment with acetic anhydride (91) and trifluoroacetic acid. N-protection with benzenesulfonyl chloride (91) afforded the derivatives 94 (Figure 22) [14,16].
In the case of 85g, an alternative pathway was chosen to prevent O-demethylation under acidic conditions. After its iodination, the resulting 95 was first treated with benzenesulfonyl chloride (93) to give the N-protected intermediate 96 that was reacted with 33 in a palladium-mediated Stille cross-coupling reaction to form 94f. Treatment of the 4-methoxy derivative 92b with dimethyl sulfate (97) gave the N-methylated intermediate 94 g (Figure 23).
The functionalized 7-azaindoles 94 were then transformed to the corresponding enaminones 97 according to the Molina and Fresneda protocol, and after cyclocondensation with guanidine (37), meriolins 3–7 (3c, 3e–g, 3b’) and 9–11 (3h, 3b, 3d) were obtained. Meriolin 7 (3b’) was isolated as a side product in the synthesis of meriolin 10 (3b), where a nucleophilic substitution of the chlorine substituent took place. Treatment of 97b with 2-methyl-2-thiopseudourea sulfate (98) led to the formation of the 2-methylthiopyrimidine-substituted meriolin 12 (3i). The meriolins 3–7 and 9–12 were isolated in overall yields ranging from 12 to 37% starting from the corresponding 7-azaindole 85 (Figure 24).
The 4-methoxy-substituted meriolins 3c, 3h, and 3j could be transformed to the corresponding 4-hydroxy-substituted meriolins 2 (3k, 26% overall yield), 8 (3l, 31% overall yield), and 13 (3m, 22% overall yield) by O-demethylation with hydrobromic acid in acetic acid (Figure 25) [14,16].

2.3.3. Meriolin Syntheses by Müller via Carbonylative Alkynylation

The Müller approach to meridianins via carbonylative alkynylation and subsequent pyrimidine synthesis could be transferred to the synthesis of meriolins. A small library of potential kinase inhibitors has been synthesized for screenings, among them meriolin derivatives 3a and 3n. Therefore, 3-iodo-N-Boc-7-azaindole (99a) was transformed to the alkynones 101 in a Sonogashira coupling with TMSA (68) or 1-hexyne (100). Alkynones 101 were then cyclized with guanidine (37), either in a mixture of tert-butanol and acetonitrile or in DMF, to the meriolin derivatives 3a (37% overall yield) and 3n (51% overall yield) (Figure 26) [61,68].

2.3.4. Three-Component Glyoxylation Decarbonylative Alkynylation Synthesis of Alkynones by Müller

Another approach by Merkul et al. to address meriolins was performed via a three-component glyoxylation alkynylation reaction, leading to N-benzylated and N-methylated meriolins 3o and 3p. Starting from 7-azaindoles, 85 in the first step reaction with oxalyl chloride (102) furnished the indole-3-glyoxyl chlorides. These reactive synthetic equivalents of acid chlorides were directly transformed to alkynones 104 in a decarbonylative Sonogashira coupling with 1-hexyne (100) or phenylacetylene (103). The cyclocondensation reaction with guanidine (37) went similarly as in the previously described strategy, which gave meriolins 3o and 3p in 51 and 52% overall yield (Figure 27) [69].

2.3.5. Synthesis of Meriolins with a Suzuki Coupling as a Key Reaction by Huang

To investigate their kinase inhibitory effects, Huang and coworkers established a synthetic route to derivatize meriolins via a nucleophilic substitution on the pyrimidine moiety, as well as by functionalizing the N-1 and C-2 position on the azaindole moiety. 7-Azaindole (85a) was N-protected by treatment with benzenesulfonyl chloride (93) before it was selectively brominated in the C-3 position. Brominated and N-protected 106 was then transformed to the boronic acid ester 108 in a Pd(dppf)2Cl2-catalyzed Miyaura borylation with bis(pinacolato)diboron (107). Suzuki coupling with 2,4-dichloropyrimidine (109) in the presence of Pd(PPh3)4 gave 110 in superior regioselectivity. The chlorine substituent on the pyrimidine ring could then be substituted by various amines 111 in a nucleophilic aromatic substitution. Two equivalents of amine were used, as one equivalent was consumed by the concurrent cleavage of the benzenesulfonyl group. This furnished 15 meriolin derivatives (3q-ae) with overall yields ranging from 48 to 58% (Figure 28) [70].
For the installment of solubilizing amino side chains, derivative 3ae was treated with methanesulfonyl chloride (112), and the mesylate 3af was obtained. After treatment with different amines 113, the meriolin derivatives 3ag-ai have been isolated in overall yields of 76–86% (starting from 3ae) (Figure 29) [70].
To assess the role of the NH group of the 7-azaindole unit in CDK1 binding, N-functionalization was anticipated. Compound 3ab was treated with potassium tert-butoxide before it reacted with different electrophiles 114 to give the derivatives 3aj-am (Figure 30) [70].
Lastly, compound 110 was treated with lithiumdiisopropylamine (LDA) and methyl iodide (115) to introduce a methyl group in the C-2 position. After the reaction with amine 117, an additional deprotection step was added, since the C-2 methyl group caused the N-benzenesulfonyl group to be stable under hot aminolysis conditions. This furnished meriolin 3an in 6% overall yield starting from 110 (Figure 31) [70].

2.3.6. Meriolin Synthesis via the Masuda borylation-Suzuki Coupling Sequence by Müller

The Müller group could show the versatility of the MBSC sequence by transferring their meridianin protocol to the synthesis of meriolins and other biaryl systems [18,63,64,71]. In a one-pot-process, 7-azaindoles 99 were transformed to the corresponding pinacolyl boronic acid esters in a palladium-mediated Masuda borylation. In the sense of a sequentially catalyzed reaction, a subsequent Suzuki coupling with arylhalide 62 follows (Figure 32). Under Suzuki conditions, the Boc group is concomitantly cleaved, leading to meriolin 1 (3a), meriolins 3ao-au, and the N-benzylated meriolins 3av and 3aw, with yields ranging from 37 to 96% (Table 1). If the reaction sequence is started with N-tosylated azaindoles 99b a subsequent deprotection step with a hydroxide base is required. This step can be included in the one-pot process, which led to meriolins 3a and 3ax-bi with overall yields ranging from 40 to 91% starting from 7-azaindoles 99b (Figure 32) [18,63].

2.3.7. Domino Amino-Palladation Reaction for the Synthesis of Meriolins by Morris

Morris and coworkers tried to adapt their meridianin protocol to the synthesis of meriolins starting from iodinated aminopyridines 119 to give 3-alkynylated 2-amino pyridines 120. In contrast to anilines (vide supra), it was not possible to prepare the monomesylated aminopyridines directly, which required treatment with trifluoroacetyl anhydride (TFAA) (121) first to furnish trifluoroacetamides 122. Reaction with mesyl chloride (112) led to the desired intermediate 123 that could be converted in the optimized domino reaction with N-Boc-4-iodopyrimidine-2-amine (83) and subsequent acid/base deprotection protocol to give meriolin 1 (3a) in 34% overall yield, as well as the 5-bromo meriolin 3bj in 31% overall yield (Figure 33) [67].

2.3.8. Metal-Free CH-Activation of a Pyrimidine and an Indolylboronic Ester by Singh

In 2016, Singh presented a metal-free CH-activation approach toward the synthesis of meriolin 1. The group reported cross-coupling between diazines and related electron-deficient heteroarenes with organoboron species. Treatment of N-Boc-protected boronic acid ester 124 with 2-aminopyrimidine (125) and potassium persulfate in an acetone-water mixture led to the formation of meriolin 1 in 35% yield (Figure 34). The proposed mechanism includes the formation of a sulfate anion radical that activates the boronic acid ester and generates an azaindolyl radical. The radical reacts with the in situ-formed pyrimidinyl salt to form a radical cation. After it undergoes single electron transfer, the protonated form of the desired product is obtained [72].

2.3.9. Functionalization of Meriolins via Suzuki Coupling or Nucleophilic Substitution Reactions by Singh and Malik

A different approach by Singh in cooperation with Malik was more pragmatic to synthesize a large library of meriolins to establish structure-activity-relationships and determine their potency against CDKs. It was elaborated that functionalization in the C-5 position and N-1 position on the 7-azaindole unit, as well as on the pyrimidine ring, should be accomplished. Starting from 5-bromo-7-azaindole (85j) in a Suzuki coupling with several boronic acids, 126 led to functionalized 7-azaindoles 127 (Figure 35). After iodination and protection with Boc-anhydride (128) to give 3-iodo-7-azaindoles 129, a Masuda borylation with pinacolyl borane (72) and subsequent Suzuki coupling with 4-chloropyrimidine-2-amine (62) gave the meriolin derivatives 3bk-cc in 25–46% overall yield (Figure 35, Table 2) [73].
To derivatize the pyrimidine ring, iodinated and N-protected 7-azaindole 124 was transformed to the corresponding pinacolyl boronic ester and reacted with 130 in a Suzuki coupling to give compound 3cd. The thiomethyl group was oxidized to give the sulfone 3ce using m-CPBA. Nucleophilic substitution by several primary 131 or secondary amines 132 furnished meriolins 3bk-cc in 29–39% overall yield (Figure 36, Table 3). To vary the substituents in the N-1 position, at first the synthesis of meriolin 1 was approached using a Masuda borylation and subsequent Suzuki coupling. After treatment with sodium hydride, reaction with different sulfonyl chlorides 133 gave meriolins 3cn-ct in 39–47% overall yield (Figure 36, Table 4).

2.3.10. Meriolin Synthesis via Friedel Crafts Acylation by Grädler

Grädler and coauthors started their approach on meriolins from 5-bromo-7-azaindole (85j) with a Friedel Crafts acylation using aluminium chloride and acid chloride 134. The intermediate 135 was reacted in a cyclocondensation with guanidine carbonate (136), which furnished meriolin 3cu in 40% overall yield. The bromine atom in C-5 position was then employed for further derivatization. Suzuki coupling with Boc-protected pinacolyl boronic acid ester 137 and subsequent Boc-deprotection with hydrochloric acid gave meriolin 3cv in 41% yield (Figure 37). The overall yield after three steps is 17% [68]. Using this method, as well as the carbonylative alkynylation by Müller [61], several derivatives have been synthesized and tested for their PDK1 inhibitory properties [68].

3. Conclusions

The naturally occurring alkaloids, variolin B (1) and meridianins 2, as well as their semisynthetic hybrids meriolins 3, represent highly active compounds with a broad spectrum of biological properties. To date, four total syntheses of variolin B (1), seven approaches to naturally occurring meridianins 2 and ten different strategies to meriolins 3, are known. The described methods include cyclocondensation reactions of alkynones and enaminones with guanidine in pyrimidine syntheses and metal-free couplings, via radical reactions or the indolization of nitrosoarenes. Transition metal-catalyzed cross-coupling reactions are prominently described, among them especially Suzuki or Stille couplings. The presented strategies allow the synthesis of known compounds as well as in many cases the adaption to unknown, synthetic analogs. The natural compounds are lead structures for the synthesis of more potent synthetic analogs, such as the meriolin hybrid structures. Due to their promising and diverse biological properties, some of the described compounds are employed in pre-clinical studies, thus being potential candidates for future cancer treatments.

Funding

The authors cordially thank Deutsche Forschungsgemeinschaft (270650915/RTG 2158-1,2) and the Fonds der Chemischen Industrie for financial support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. França, P.H.B.; Barbosa, D.; da Silva, D.; Ribeiro, Ê.N.; Santana, A.; Santos, B.; Barbosa-Filho, J.; Quintans, J.; Barreto, R.; Quintans-Júnior, L.J.; et al. Indole Alkaloids from Marine Sources as Potential Leads against Infectious Diseases. BioMed Res. Int. 2014, 2014, 375423. [Google Scholar] [CrossRef] [PubMed]
  2. König, G.M.; Wright, A.; Sticher, O.; Angerhofer, C.; Pezzuto, J. Biological Activities of Selected Marine Natural Products. Planta Med. 1994, 60, 532–537. [Google Scholar] [CrossRef] [PubMed]
  3. Pauletti, P.M.; Cintra, L.S.; Braguine, C.G.; Filho, A.A.D.S.; Silva, M.L.A.E.; Cunha, W.R.; Januário, A.H. Halogenated Indole Alkaloids from Marine Invertebrates. Mar. Drugs 2010, 8, 1526–1549. [Google Scholar] [CrossRef] [Green Version]
  4. Kobayashi, J.I.; Murayama, T.; Ishibashi, M.; Kosuge, S.; Takamatsu, M.; Ohizumi, Y.; Kobayashi, H.; Ohta, T.; Nozoe, S.; Takuma, S. Hyrtiosins A and B, new indole alkaloids from the Okinawan marine sponge Hyrtios erecta. Tetrahedron 1990, 46, 7699–7702. [Google Scholar] [CrossRef]
  5. Ban, Y.; Murakami, Y.; Iwasawa, Y.; Tsuchiya, M.; Takano, N. Indole alkaloids in medicine. Med. Res. Rev. 1988, 8, 231–308. [Google Scholar] [CrossRef]
  6. Abdelmohsen, U.R.; Balasubramanian, S.; Oelschlaeger, T.; Grkovic, T.; Pham, N.; Quinn, R.; Hentschel, U. Potential of marine natural products against drug-resistant fungal, viral, and parasitic infections. Lancet Infect. Dis. 2017, 17, e30–e41. [Google Scholar] [CrossRef]
  7. Perry, N.B.; Ettouati, L.; Litaudon, M.; Blunt, J.; Munro, M.; Parkin, S.; Hope, H. Alkaloids from the antarctic sponge Kirkpatrickia varialosa: Part 1: Variolin b, a new antitumour and antiviral compound. Tetrahedron 1994, 50, 3987–3992. [Google Scholar] [CrossRef]
  8. Trimurtulu, G.; Faulkner, D.; Perry, N.; Ettouati, L.; Litaudon, M.; Blunt, J.; Munro, M.; Jameson, G. Alkaloids from the antarctic sponge Kirkpatrickia varialosa. Part 2: Variolin A and N(3′)-methyl tetrahydrovariolin B. Tetrahedron 1994, 50, 3993–4000. [Google Scholar] [CrossRef]
  9. Franco, L.H.; Joffé, E.K.; Puricelli, L.; Tatian, M.; Seldes, A.; Palermo, J. Indole Alkaloids from the Tunicate Aplidium meridianum. J. Nat. Prod. 1998, 61, 1130–1132. [Google Scholar] [CrossRef]
  10. Seldes, A.M.; Brasco, M.R.; Franco, L.H.; Palermo, J. Identification of two meridianins from the crude extract of the tunicate Aplidium meridianum by tandem mass spectrometry. Nat. Prod. Res. 2007, 21, 555–563. [Google Scholar] [CrossRef]
  11. Gompel, M.; Leost, M.; De Kier Joffe, E.; Puricelli, L.; Franco, L.; Palermo, J.; Meijer, L. Meridianins, a new family of protein kinase inhibitors isolated from the ascidian Aplidium meridianum. Bioorg. Med. Chem. Lett. 2004, 14, 1703–1707. [Google Scholar] [CrossRef] [PubMed]
  12. Bharate, S.B.; Yadav, R.; Khan, S.; Tekwani, B.; Jacob, M.; Khan, I.; Vishwakarma, R. Meridianin G and its analogs as antimalarial agents. MedChemComm 2013, 4, 1042–1048. [Google Scholar] [CrossRef]
  13. Yadav, R.R.; Khan, S.; Singh, S.; Khan, I.; Vishwakarma, R.; Bharate, S. Synthesis, antimalarial and antitubercular activities of meridianin derivatives. Eur. J. Med. Chem. 2015, 98, 160–169. [Google Scholar] [CrossRef]
  14. Bettayeb, K.; Tirado, O.; Marionneau-Lambot, S.V.; Ferandin, Y.; Lozach, O.; Morris, J.; Mateo-Lozano, S.; Drueckes, P.; Schächtele, C.; Kubbutat, M.; et al. Meriolins, a New Class of Cell Death–Inducing Kinase Inhibitors with Enhanced Selectivity for Cyclin-Dependent Kinases. Cancer Res. 2007, 67, 8325–8334. [Google Scholar] [CrossRef] [Green Version]
  15. Fresneda, P.M.; Molina, P.; Bleda, J. Synthesis of the indole alkaloids meridianins from the tunicate Aplidium meridianum. Tetrahedron 2001, 57, 2355–2363. [Google Scholar] [CrossRef]
  16. Echalier, A.; Bettayeb, K.; Ferandin, Y.; Lozach, O.; Clément, M.; Valette, A.; Liger, F.; Marquet, B.; Morris, J.; Endicott, J.; et al. Meriolins (3-(Pyrimidin-4-yl)-7-azaindoles): Synthesis, Kinase Inhibitory Activity, Cellular Effects, and Structure of a CDK2/Cyclin A/Meriolin Complex. J. Med. Chem. 2008, 51, 737–751. [Google Scholar] [CrossRef]
  17. Jarry, M.; Lecointre, C.; Malleval, C.; Desrues, L.; Schouft, M.-T.; Lejoncour, V.; Liger, F.; Lyvinec, G.; Joseph, B.; Loaëc, N.; et al. Impact of meriolins, a new class of cyclin-dependent kinase inhibitors, on malignant glioma proliferation and neo-angiogenesis. Neuro-Oncology 2014, 16, 1484–1498. [Google Scholar] [CrossRef] [Green Version]
  18. Drießen, D.; Stuhldreier, F.; Frank, A.; Stark, H.; Wesselborg, S.; Stork, B.; Müller, T. Novel meriolin derivatives as rapid apoptosis inducers. Bioorg. Med. Chem. 2019, 27, 3463–3468. [Google Scholar] [CrossRef]
  19. Alexander, A.; Karakas, C.; Chen, X.; Carey, J.; Yi, M.; Bondy, M.; Thompson, P.; Cheung, K.; Ellis, I.; Gong, Y.; et al. Cyclin E overexpression as a biomarker for combination treatment strategies in inflammatory breast cancer. Oncotarget 2017, 8, 14897–14911. [Google Scholar] [CrossRef] [Green Version]
  20. Akli, S.; Van Pelt, C.; Bui, T.; Meijer, L.; Keyomarsi, K. Cdk2 is Required for Breast Cancer Mediated by the Low-Molecular-Weight Isoform of Cyclin E. Cancer Res. 2011, 71, 3377–3386. [Google Scholar] [CrossRef]
  21. Chashoo, G.; Singh, U.; Singh, P.; Mondhe, D.; Vishwakarma, R. A Marine-Based Meriolin (3-Pyrimidinylazaindole) Derivative (4ab) Targets PI3K/AKT /mTOR Pathway Inducing Cell Cycle Arrest and Apoptosis in Molt-4 Cells. Clin. Cancer Drugs 2019, 6, 33–40. [Google Scholar] [CrossRef]
  22. Mou, J.; Chen, D.; Deng, Y. Inhibitors of Cyclin-Dependent Kinase 1/2 for Anticancer Treatment. Med. Chem. 2020, 16, 307–325. [Google Scholar] [CrossRef] [PubMed]
  23. Dorsch, D.; Sirrenberg, C.; Müller, T. Preparation of 4-(Pyrrolopyridinyl)pyrimidinyl-2-amines as Antitumor Agents. PCT International Application WO 2007/107221 A1, 27 September 2007. [Google Scholar]
  24. Dorsch, D.; Wuchrer, M.; Burgdorf, L.; Sirrenberg, C.; Esdar, C.; Mueller, T.J.J.; Merkul, E. 6-(Pyrrolopyridinyl)-pyrimidine-2-yl-amine Derivatives and Their Use for the Treatment of Cancers and Aids. PCT International Application WO 2008155000 A1, 24 December 2008. [Google Scholar]
  25. Dorsch, D.; Sirrenberg, C.; Müller, T.; Merkul, E. Preparation of 3-(4-Pyridinyl)-1H-pyrrolo[2,3-b]pyridines as Anti-Tumor Agents. Ger. Offen. DE 102008025751 A1, 3 December 2009. [Google Scholar]
  26. Dorsch, D.; Sirrenberg, C.; Müller, T.; Merkul, E.; Karapetyan, G. 7-Azaindole Derivatives as Kinase Inhibitors and Their Preparation and Use in the Treatment of Tumors. PCT International Application WO 2012104007 A2, 9 August 2012. [Google Scholar]
  27. Łukasik, P.; Baranowska-Bosiacka, I.; Kulczycka, K.; Gutowska, I. Inhibitors of Cyclin-Dependent Kinases: Types and Their Mechanism of Action. Int. J. Mol. Sci. 2021, 22, 2806. [Google Scholar] [CrossRef]
  28. Bharate, S.B.; Sawant, S.; Singh, P.; Vishwakarma, R. Kinase Inhibitors of Marine Origin. Chem. Rev. 2013, 113, 6761–6815. [Google Scholar] [CrossRef]
  29. Bharate, S.B.; Yadav, R.; Battula, S.; Vishwakarma, R. Meridianins: Marine-Derived Potent Kinase Inhibitors. Mini-Rev. Med. Chem. 2012, 12, 618–631. [Google Scholar] [CrossRef]
  30. Sandtorv, A.H. Chapter 5 - Chemical Synthesis of Meridianins and Related Derivatives. Stud. Nat. Prod. Chem. 2017, 53, 143–166. [Google Scholar] [CrossRef]
  31. Stanovnik, B.; Svete, J. The Synthesis Aplysinopsins, Meridianines, and Related Compounds. Mini-Rev. Org. Chem. 2005, 2, 211–224. [Google Scholar] [CrossRef]
  32. Popowycz, F.; Routier, S.; Joseph, B.; Mérour, J.-Y. Synthesis and reactivity of 7-azaindole (1H-pyrrolo[2,3-b]pyridine). Tetrahedron 2007, 63, 1031–1064. [Google Scholar] [CrossRef]
  33. Xiao, L. A Review: Meridianins and Meridianins Derivatives. Molecules 2022, 27, 8714. [Google Scholar] [CrossRef]
  34. Walker, S.R.; Carter, E.; Huff, B.; Morris, J. Variolins and Related Alkaloids. Chem. Rev. 2009, 109, 3080–3098. [Google Scholar] [CrossRef]
  35. Alvarez, M.; Fernández, D.; Joule, J. Synthesis of 3-Aryl- and 3-Heteroaryl-7-azaindoles. Synthesis 1999, 1999, 615–620. [Google Scholar] [CrossRef]
  36. Alvarez, M.; Fernández, D.; Joule, J. Synthesis of 1,2-dihydropyrrolo [1,2-c] pyrimidin-1-ones. J. Chem. Soc. Perkin Trans. 1 1999, 1999, 249–256. [Google Scholar] [CrossRef]
  37. Álvarez, M.; Fernández, D.; Joule, J. Synthesis of deoxyvariolin B. Tetrahedron Lett. 2001, 42, 315–317. [Google Scholar] [CrossRef]
  38. Mendiola, J.; Minguez, J.; Alvarez-Builla, J.; Vaquero, J. Reaction of 2-Bromomethylazoles and TosMIC:  A Domino Process to Azolopyrimidines. Synthesis of Core Tricycle of the Variolins Alkaloids. Org. Lett. 2000, 2, 3253–3256. [Google Scholar] [CrossRef] [PubMed]
  39. Mendiola, J.; Baeza, A.; Alvarez-Builla, J.; Vaquero, J. Reaction of Bromomethylazoles and Tosylmethyl Isocyanide. A Novel Heterocyclization Method for the Synthesis of the Core of Marine Alkaloids Variolins and Related Azolopyrimidines. J. Org. Chem. 2004, 69, 4974–4983. [Google Scholar] [CrossRef]
  40. Fresneda, P.M.; Molina, P.; Delgado, S.; Bleda, J. Synthetic studies towards the 2-aminopyrimidine alkaloids variolins and meridianins from marine origin. Tetrahedron Lett. 2000, 41, 4777–4780. [Google Scholar] [CrossRef]
  41. Anderson, R.J.; Morris, J. Studies toward the total synthesis of the variolins: Rapid entry to the core structure. Tetrahedron Lett. 2001, 42, 311–313. [Google Scholar] [CrossRef]
  42. Anderson, R.J.; Morris, J. Total synthesis of variolin B. Tetrahedron Lett. 2001, 42, 8697–8699. [Google Scholar] [CrossRef] [Green Version]
  43. Anderson, R.J.; Hill, J.; Morris, J. Concise Total Syntheses of Variolin B and Deoxyvariolin B. J. Org. Chem. 2005, 70, 6204–6212. [Google Scholar] [CrossRef]
  44. Molina, P.; Fresneda, P.; Delgado, S.; Bleda, J. Synthesis of the potent antitumoral marine alkaloid variolin B. Tetrahedron Lett. 2002, 43, 1005–1007. [Google Scholar] [CrossRef]
  45. Molina, P.; Fresneda, P.; Delgado, S. Carbodiimide-Mediated Preparation of the Tricyclic Pyrido[3′,2′:4,5]pyrrolo[1,2-c]pyrimidine Ring System and Its Application to the Synthesis of the Potent Antitumoral Marine Alkaloid Variolin B and Analog. J. Org. Chem. 2003, 68, 489–499. [Google Scholar] [CrossRef] [PubMed]
  46. Bredereck, H.; Effenberger, F.; Botsch, H.; Rehn, H. Synthesen in der heterocyclischen Reihe, V: Umsetzungen von vinylogen Carbonsäureamiden zu Heterocyclen. Chem. Ber. 1965, 98, 1081–1086. [Google Scholar] [CrossRef]
  47. Ahaidar, A.; Fernández, D.; Pérez, O.; Danelón, G.; Cuevas, C.; Manzanares, I.; Albericio, F.; Joule, J.; Álvarez, M. Synthesis of variolin B. Tetrahedron Lett. 2003, 44, 6191–6194. [Google Scholar] [CrossRef]
  48. Ahaidar, A.; Fernández, D.; Danelón, G.; Cuevas, C.; Manzanares, I.; Albericio, F.; Joule, J.; Álvarez, M. Total Syntheses of Variolin B and Deoxyvariolin B1. J. Org. Chem. 2003, 68, 10020–10029. [Google Scholar] [CrossRef]
  49. Fernández, D.; Ahaidar, A.; Danelón, G.; Cironi, P.; Marfil, M.; Pérez, O.; Cuevas, C.; Albericio, F.; Joule, J.; Álvarez, M. Synthesis of Polyheterocyclic Nitrogen-Containing Marine Natural Products. Monatsh. Chem. 2004, 135, 615–627. [Google Scholar] [CrossRef]
  50. Katritzky, A.R.; Akutagawa, K. Carbon dioxide: A reagent for simultaneous protection of nucleophilic centers and the activation of alternative locations to electrophilic attack. V. Activation of the 2-alkyl group of a 2-alkylindole toward proton loss and subsequent electrophilic substitution. J. Am. Chem. Soc. 1986, 108, 6808–6809. [Google Scholar] [CrossRef]
  51. Baeza, A.; Mendiola, J.; Burgos, C.; Alvarez-Builla, J.; Vaquero, J. Palladium-mediated C–N, C–C, and C–O functionalization of azolopyrimidines: A new total synthesis of variolin B. Tetrahedron Lett. 2008, 49, 4073–4077. [Google Scholar] [CrossRef] [Green Version]
  52. Baeza, A.; Mendiola, J.; Burgos, C.; Alvarez-Builla, J.; Vaquero, J. Application of Selective Palladium-Mediated Functionalization of the Pyrido[3′,2′:4,5]pyrrolo[1,2-c]pyrimidine Heterocyclic System for the Total Synthesis of Variolin B and Deoxyvariolin B. Eur. J. Org. Chem. 2010, 2010, 5607–5618. [Google Scholar] [CrossRef]
  53. Jiang, B.; Yang, C.-G. Synthesis of Indolylpyrimidiness via Cross-Coupling of Indolylboronic Acid with Choropyrimidines: Facile Synthesis of Meridianin D. Heterocycles 2000, 53, 1489–1498. [Google Scholar] [CrossRef]
  54. Radwan, M.A.A.; El-Sherbiny, M. Synthesis and antitumor activity of indolylpyrimidines: Marine natural product meridianin D analogues. Bioorg. Med. Chem. 2007, 15, 1206–1211. [Google Scholar] [CrossRef]
  55. Corbel, B.; Michaud, F.; Meijer, L.; Simon, G.; Couthon-Gourves, H.; Haelters, J.-P.; Kervarec, N. Towards the syntheses of N-H and N-alkylated derivatives of meridianins. J. Heterocycl. Chem. 2007, 44, 793–801. [Google Scholar] [CrossRef]
  56. Sperry, J. A concise synthesis of meridianin F. Tetrahedron Lett. 2011, 52, 4537–4538. [Google Scholar] [CrossRef]
  57. Lebar, M.D.; Hahn, K.; Mutka, T.; Maignan, P.; McClintock, J.; Amsler, C.; van Olphen, A.; Kyle, D.; Baker, B. CNS and antimalarial activity of synthetic meridianin and psammopemmin analogs. Bioorg. Med. Chem. 2011, 19, 5756–5762. [Google Scholar] [CrossRef] [PubMed]
  58. More, K.N.; Jang, H.; Hong, V.; Lee, J. Pim kinase inhibitory and antiproliferative activity of a novel series of meridianin C derivatives. Bioorg. Med. Chem. Lett. 2014, 24, 2424–2428. [Google Scholar] [CrossRef]
  59. Dong, J.; Huang, S.-S.; Hao, Y.-N.; Wang, Z.-W.; Liu, Y.-X.; Li, Y.-Q.; Wang, Q.-M. Marine-natural-products for biocides development: First discovery of meridianin alkaloids as antiviral and anti-phytopathogenic-fungus agents. Pest Manag. Sci. 2020, 76, 3369–3376. [Google Scholar] [CrossRef] [PubMed]
  60. Han, S.; Zhuang, C.; Zhou, W.; Chen, F. Structural-Based Optimizations of the Marine-Originated Meridianin C as Glucose Uptake Agents by Inhibiting GSK-3β. Mar. Drugs 2021, 19, 149. [Google Scholar] [CrossRef]
  61. Karpov, A.S.; Merkul, E.; Rominger, F.; Müller, T. Concise Syntheses of Meridianins by Carbonylative Alkynylation and a Four-Component Pyrimidine Synthesis. Angew. Chem. Int. Ed. 2005, 44, 6951–6956. [Google Scholar] [CrossRef]
  62. Tibiletti, F.; Simonetti, M.; Nicholas, K.; Palmisano, G.; Parravicini, M.; Imbesi, F.; Tollari, S.; Penoni, A. One-pot synthesis of meridianins and meridianin analogues via indolization of nitrosoarenes. Tetrahedron 2010, 66, 1280–1288. [Google Scholar] [CrossRef]
  63. Merkul, E.; Schäfer, E.; Müller, T. Rapid synthesis of bis(hetero)aryls by one-pot Masuda borylation-Suzuki coupling sequence and its application to concise total syntheses of meridianins A and G. Org. Biomol. Chem. 2011, 9, 3139–3141. [Google Scholar] [CrossRef]
  64. Kruppa, M.; Sommer, G.; Müller, T. Concise Syntheses of Marine (Bis)indole Alkaloids Meridianin C, D, F, and G and Scalaridine A via One-Pot Masuda Borylation-Suzuki Coupling Sequence. Molecules 2022, 27, 2233. [Google Scholar] [CrossRef]
  65. Parsons, T.B.; Ghellamallah, C.; Male, L.; Spencer, N.; Grainger, R. Regioselective dibromination of methyl indole-3-carboxylate and application in the synthesis of 5,6-dibromoindoles. Org. Biomol. Chem. 2011, 9, 5021–5023. [Google Scholar] [CrossRef] [PubMed]
  66. Cacchi, S.; Fabrizi, G.; Marinelli, F.; Moro, L.; Pace, P. 3-Aryl-2-Unsubstituted Indoles through the Palladium-Catalysed Reaction of o-Ethynyltrifluoroacetanilide with Aryl Iodides. Synlett 1997, 12, 1363–1366. [Google Scholar] [CrossRef]
  67. Walker, S.R.; Czyz, M.; Morris, J. Concise Syntheses of Meridianins and Meriolins Using a Catalytic Domino Amino-Palladation Reaction. Org. Lett. 2014, 16, 708–711. [Google Scholar] [CrossRef] [PubMed]
  68. Wucherer-Plietker, M.; Merkul, E.; Müller, T.; Esdar, C.; Knöchel, T.; Heinrich, T.; Buchstaller, H.-P.; Greiner, H.; Dorsch, D.; Finsinger, D.; et al. Discovery of novel 7-azaindoles as PDK1 inhibitors. Bioorg. Med. Chem. Lett. 2016, 26, 3073–3080. [Google Scholar] [CrossRef] [PubMed]
  69. Merkul, E.; Oeser, T.; Müller, T. Consecutive Three-Component Synthesis of Ynones by Decarbonylative Sonogashira Coupling. Eur. J. Chem. 2009, 15, 5006–5011. [Google Scholar] [CrossRef] [PubMed]
  70. Huang, S.; Li, R.; Connolly, P.; Emanuel, S.; Middleton, S. Synthesis of 2-amino-4-(7-azaindol-3-yl)pyrimidines as cyclin dependent kinase 1 (CDK1) inhibitors. Bioorg. Med. Chem. Lett. 2006, 16, 4818–4821. [Google Scholar] [CrossRef] [PubMed]
  71. Rehberg, N.; Sommer, G.; Drießen, D.; Kruppa, M.; Adeniyi, E.; Chen, S.; Wang, L.; Wolf, K.; Tasch, B.; Ioerger, T.; et al. Nature-Inspired (di)Azine-Bridged Bisindole Alkaloids with Potent Antibacterial In Vitro and In Vivo Efficacy against Methicillin-Resistant Staphylococcus aureus. J. Med. Chem. 2020, 63, 12623–12641. [Google Scholar] [CrossRef]
  72. Thatikonda, T.; Singh, U.; Ambala, S.; Vishwakarma, R.; Singh, P. Metal free C-H functionalization of diazines and related heteroarenes with organoboron species and its application in the synthesis of a CDK inhibitor, meriolin 1. Org. Biomol. Chem. 2016, 14, 4312–4320. [Google Scholar] [CrossRef]
  73. Singh, U.; Chashoo, G.; Khan, S.; Mahajan, P.; Nargotra, A.; Mahajan, G.; Singh, A.; Sharma, A.; Mintoo, M.; Guru, S.; et al. Design of Novel 3-Pyrimidinylazaindole CDK2/9 Inhibitors with Potent In Vitro and In Vivo Antitumor Efficacy in a Triple-Negative Breast Cancer Model. J. Med. Chem. 2017, 60, 9470–9489. [Google Scholar] [CrossRef]
Figure 1. Variolins isolated from the Antarctic sponge Kirkpatrickia varialosa [7,8].
Figure 1. Variolins isolated from the Antarctic sponge Kirkpatrickia varialosa [7,8].
Molecules 28 00947 g001
Figure 2. Meridianins 2 isolated from the tunicate Aplidium meridianum [9,10].
Figure 2. Meridianins 2 isolated from the tunicate Aplidium meridianum [9,10].
Molecules 28 00947 g002
Figure 3. Meriolin 1 (3a) as a synthetic hybrid structure of variolin B (1) and meridianins 2 [14].
Figure 3. Meriolin 1 (3a) as a synthetic hybrid structure of variolin B (1) and meridianins 2 [14].
Molecules 28 00947 g003
Figure 4. First total synthesis of variolin B (1) by Morris and Anderson [42].
Figure 4. First total synthesis of variolin B (1) by Morris and Anderson [42].
Molecules 28 00947 g004
Figure 5. Synthesis of key intermediate azaindole 19. Reaction conditions for a: first: 1.4 equivs NaH, DMF, rt, 45 min. Then: 1.4 equivs SEM-Cl, rt, 12 h [44].
Figure 5. Synthesis of key intermediate azaindole 19. Reaction conditions for a: first: 1.4 equivs NaH, DMF, rt, 45 min. Then: 1.4 equivs SEM-Cl, rt, 12 h [44].
Molecules 28 00947 g005
Figure 6. Two approaches to the synthesis of the tricyclic pyridopyrrolopyrimidine structures, 23 and 27 [44].
Figure 6. Two approaches to the synthesis of the tricyclic pyridopyrrolopyrimidine structures, 23 and 27 [44].
Molecules 28 00947 g006
Figure 7. Introduction of an acetyl group at C-5 [44].
Figure 7. Introduction of an acetyl group at C-5 [44].
Molecules 28 00947 g007
Figure 8. Synthesis of the 2-aminopyrimidine ring to give access to variolin B (1) [44].
Figure 8. Synthesis of the 2-aminopyrimidine ring to give access to variolin B (1) [44].
Molecules 28 00947 g008
Figure 9. Synthesis of the key intermediate iodide 49 [47].
Figure 9. Synthesis of the key intermediate iodide 49 [47].
Molecules 28 00947 g009
Figure 10. Synthesis of variolin B (1) via Stille coupling as a key reaction step [47].
Figure 10. Synthesis of variolin B (1) via Stille coupling as a key reaction step [47].
Molecules 28 00947 g010
Figure 11. Synthesis of the trihalo core and introduction of the C-9 amino substituent [51].
Figure 11. Synthesis of the trihalo core and introduction of the C-9 amino substituent [51].
Molecules 28 00947 g011
Figure 12. Palladium-mediated synthesis of variolin B (1) [51].
Figure 12. Palladium-mediated synthesis of variolin B (1) [51].
Molecules 28 00947 g012
Figure 13. First synthesis of meridianins D and G [53].
Figure 13. First synthesis of meridianins D and G [53].
Molecules 28 00947 g013
Figure 14. Synthesis of meridianins A, C, D and E by Molina and Fresneda [15].
Figure 14. Synthesis of meridianins A, C, D and E by Molina and Fresneda [15].
Molecules 28 00947 g014
Figure 15. Synthesis of meridianins C, D, and G by Karpov et al. via carbonylative alkynylation [61].
Figure 15. Synthesis of meridianins C, D, and G by Karpov et al. via carbonylative alkynylation [61].
Molecules 28 00947 g015
Figure 16. Indolization of nitrosoarenes for the synthesis of meridianins C and G by Penoni [62].
Figure 16. Indolization of nitrosoarenes for the synthesis of meridianins C and G by Penoni [62].
Molecules 28 00947 g016
Figure 17. Synthesis of meridianins by Müller via MBSC sequence [63,64].
Figure 17. Synthesis of meridianins by Müller via MBSC sequence [63,64].
Molecules 28 00947 g017
Figure 18. Demethylation of 2i furnished meridianin A.
Figure 18. Demethylation of 2i furnished meridianin A.
Molecules 28 00947 g018
Figure 19. Synthesis of meridianin F by Grainger [65].
Figure 19. Synthesis of meridianin F by Grainger [65].
Molecules 28 00947 g019
Figure 20. Synthesis of meridianins C and G via palladium-catalyzed domino reaction by Morris [67].
Figure 20. Synthesis of meridianins C and G via palladium-catalyzed domino reaction by Morris [67].
Molecules 28 00947 g020
Figure 21. Synthesis of meriolin 1 by Molina and Fresneda [15].
Figure 21. Synthesis of meriolin 1 by Molina and Fresneda [15].
Molecules 28 00947 g021
Figure 22. Preparation of 3-acetyl-N-protected intermediates 94 [14].
Figure 22. Preparation of 3-acetyl-N-protected intermediates 94 [14].
Molecules 28 00947 g022
Figure 23. Alternative pathways to access 94f and 94g [14].
Figure 23. Alternative pathways to access 94f and 94g [14].
Molecules 28 00947 g023
Figure 24. Synthesis of meriolins 3–7 and 9–12 by Joseph and Meijer [14].
Figure 24. Synthesis of meriolins 3–7 and 9–12 by Joseph and Meijer [14].
Molecules 28 00947 g024
Figure 25. O-demethylation of meriolins 3, 9, and 14 gives 4-hydroxy-substituted meriolins 2, 8, and 13 [14].
Figure 25. O-demethylation of meriolins 3, 9, and 14 gives 4-hydroxy-substituted meriolins 2, 8, and 13 [14].
Molecules 28 00947 g025
Figure 26. Synthesis of meriolins 3a and 3n via carbonylative alkynylation and subsequent pyrimidine synthesis by Müller [61,68].
Figure 26. Synthesis of meriolins 3a and 3n via carbonylative alkynylation and subsequent pyrimidine synthesis by Müller [61,68].
Molecules 28 00947 g026
Figure 27. Synthesis of meriolins via three-component glyoxylation decarbonylative alkynylation by Merkul et al. [69].
Figure 27. Synthesis of meriolins via three-component glyoxylation decarbonylative alkynylation by Merkul et al. [69].
Molecules 28 00947 g027
Figure 28. Meriolin synthesis by Huang via nucleophilic substitution on the pyrimidine moiety [70].
Figure 28. Meriolin synthesis by Huang via nucleophilic substitution on the pyrimidine moiety [70].
Molecules 28 00947 g028
Figure 29. Introduction of solubilizing side chains gave the meriolin derivatives 3ag-ai [70].
Figure 29. Introduction of solubilizing side chains gave the meriolin derivatives 3ag-ai [70].
Molecules 28 00947 g029
Figure 30. N-functionalization of compound 3ab with different electrophiles gave meriolins 3aj-am [70].
Figure 30. N-functionalization of compound 3ab with different electrophiles gave meriolins 3aj-am [70].
Molecules 28 00947 g030
Figure 31. Synthesis of meriolin 3an [70].
Figure 31. Synthesis of meriolin 3an [70].
Molecules 28 00947 g031
Figure 32. MBSC sequence for the synthesis of meriolins 3a and 3ao-bi by Müller [18,63].
Figure 32. MBSC sequence for the synthesis of meriolins 3a and 3ao-bi by Müller [18,63].
Molecules 28 00947 g032
Figure 33. Synthesis of meriolins 3a and 3bj via domino amino-palladation reaction by Morris [67].
Figure 33. Synthesis of meriolins 3a and 3bj via domino amino-palladation reaction by Morris [67].
Molecules 28 00947 g033
Figure 34. Metal-free synthesis of meriolin 1 via CH-activation by Singh [72].
Figure 34. Metal-free synthesis of meriolin 1 via CH-activation by Singh [72].
Molecules 28 00947 g034
Figure 35. Functionalization of the C-5 position led to meriolins 3bk-cc [73].
Figure 35. Functionalization of the C-5 position led to meriolins 3bk-cc [73].
Molecules 28 00947 g035
Figure 36. Functionalization of the pyrimidine ring and the N-1 position gave meriolins 3cf-ct [73].
Figure 36. Functionalization of the pyrimidine ring and the N-1 position gave meriolins 3cf-ct [73].
Molecules 28 00947 g036
Figure 37. Preparation of meriolin 3cv via Friedel Crafts acylation by Grädler [68].
Figure 37. Preparation of meriolin 3cv via Friedel Crafts acylation by Grädler [68].
Molecules 28 00947 g037
Table 1. Introduced heterocycles R2 62 and the corresponding yields of the synthesized meriolins 3.
Table 1. Introduced heterocycles R2 62 and the corresponding yields of the synthesized meriolins 3.
EntryAzaindole 99R1(hetero)aryl R2 62Meriolins 3 (Yield)
199aH4-pyrimidin-2-amine (62b)3a, meriolin 1 (63%)
299aH6-pyrazin-2-amine (62c)3ao (53%)
399aH5-pyrimidin-2-amine (62d)3ap (66%)
499aH2-pyrimidin-4-amine (62e)3aq (37%)
599aH6-pyridin-2-amine (62f)3ar (81%)
699aH4-pyridin-2-amine (62g)3as (64%)
799aH2-aniline (62h)3at (74%)
899aH4-phenol (62i)3au (57%)
999cBn4-pyrimidin-2-amine (62b)3av (96%)
1099cBn4-pyridin-2-amine (62g)3aw (93%)
1199bH4-pyrimidin-2-amine (62b)3a, meriolin 1 (81%)
1299bH5-pyridin-2-amine (62j)3ax (91%)
1399bH4-pyridin-2-amine (62g)3ay (75%)
1499bHN-benzyl-5-pyridin-2-amine (62k)3az (77%)
1599bH4-(2-methoxypyrimidine) (62l)3ba (40%)
1699bH4-pyridin-2,6-diamine (62m)3bb (67%)
1799bH5-pyrimidin-2-amine (62d)3bc (47%)
1899bH4-(2-methylthiopyrimidine) (62n)3bd (83%)
1999bH4-(6-methoxypyrimidin-2-amine) (62o)3be (62%)
2099bH4-pyrimidin-2,6-diamine (62p)3bf (53%)
2199bHN-benzyl-4-pyridin-2-amine (62q)3bg (83%)
2299bH2-pyrimidin-4-amine (62e)3bh (75%)
23 a99bH5-isoquinolin (62r)3bi (82%)
a Suzuki coupling with DME/water.
Table 2. Boronic acids 126 used for the functionalization in C-5 position and the corresponding yields of meriolins 3bk-cc.
Table 2. Boronic acids 126 used for the functionalization in C-5 position and the corresponding yields of meriolins 3bk-cc.
EntryBoronic Acid R1-B(OH)2 (126)Meriolins 3 (Yield)
1(4-(trifluoromethyl)phenyl)boronic acid (126a)3bk (63%)
2(4-fluorophenyl)boronic acid (126b)3bl (65%)
3(4-chlorophenyl)boronic acid (126c)3bm (62%)
4(4-(trifluoromethoxy)phenyl)boronic acid (126d)3bn (60%)
5(4-methoxyphenyl)boronic acid (126e)3bo (59%)
6(4-(methylthio)phenyl)boronic acid (126f)3bp (54%)
7(3-fluorophenyl)boronic acid (126g)3bq (55%)
8m-tolylboronic acid (126h)3br (50%)
9(3-(trifluoromethyl)phenyl)boronic acid (126i)3bs (58%)
10(2-(methylthio)phenyl)boronic acid (126j)3bt (59%)
11(2-ethylphenyl)boronic acid (126k)3bu (48%)
12naphthalen-1-ylboronic acid (126l)3bv (60%)
13(2-methoxynaphthalen-1-yl)boronic acid (126m)3bw (55%)
14furan-3-ylboronic acid (126n)3bx (48%)
15thiophen-3-ylboronic acid (126o)3by (44%)
16pyridin-3-ylboronic acid (126p)3bz (45%)
17benzo[b]thiophen-2-ylboronic acid (126q)3ca (40%)
18benzofuran-2-ylboronic acid (126r)3cb (39%)
19(5-methoxy-1H-indol-2-yl)boronic acid (126s)3cc (45%)
Table 3. Primary 131 and secondary amines 132 used for the functionalization of the pyrimidine ring and the corresponding yields of meriolins 3cf-cm.
Table 3. Primary 131 and secondary amines 132 used for the functionalization of the pyrimidine ring and the corresponding yields of meriolins 3cf-cm.
EntryAmine R2-NH2 (131) or R2NHR3 (132)Meriolins 3 (Yield)
12-phenylethan-1-amine (131a)3cf (65%)
22-(4-methoxyphenyl)ethan-1-amine (131b)3cg (60%)
32-(3,4-dimethoxyphenyl)ethan-1-amine (131c)3ch (68%)
42-(1H-indol-3-yl)ethan-1-amine (131d)3ci (50%)
5pyrrolidine (132a)3cj (52%)
6piperidine (132b)3ck (52%)
7morpholine (132c)3cl (60%)
81-methylpiperazine (132d)3cm (54%)
Table 4. Sulfonyl chlorides 133 for the functionalization in N-1 position and the corresponding yields of meriolins 3cn-ct.
Table 4. Sulfonyl chlorides 133 for the functionalization in N-1 position and the corresponding yields of meriolins 3cn-ct.
EntrySulfonyl Chloride R4-SO2Cl (133)Meriolins 3 (Yield)
14-fluorobenzenesulfonyl chloride (133a)3cn (70%)
24-bromobenzenesulfonyl chloride (133b)3co (70%)
34-(trifluoromethyl)benzenesulfonyl chloride (133c)3cp (77%)
44-(trifluoromethoxy)benzenesulfonyl chloride (133d)3cq (80%)
54-acetamidobenzenesulfonyl chloride (133e)3cr (71%)
62,3-dihydrobenzo[b][1,4]dioxine-6-sulfonyl chloride (133f)3cs (79%)
71-methyl-1H-imidazole-5-sulfonyl chloride (133g)3ct (65%)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kruppa, M.; Müller, T.J.J. A Survey on the Synthesis of Variolins, Meridianins, and Meriolins—Naturally Occurring Marine (aza)Indole Alkaloids and Their Semisynthetic Derivatives. Molecules 2023, 28, 947. https://doi.org/10.3390/molecules28030947

AMA Style

Kruppa M, Müller TJJ. A Survey on the Synthesis of Variolins, Meridianins, and Meriolins—Naturally Occurring Marine (aza)Indole Alkaloids and Their Semisynthetic Derivatives. Molecules. 2023; 28(3):947. https://doi.org/10.3390/molecules28030947

Chicago/Turabian Style

Kruppa, Marco, and Thomas J. J. Müller. 2023. "A Survey on the Synthesis of Variolins, Meridianins, and Meriolins—Naturally Occurring Marine (aza)Indole Alkaloids and Their Semisynthetic Derivatives" Molecules 28, no. 3: 947. https://doi.org/10.3390/molecules28030947

Article Metrics

Back to TopTop