Next Article in Journal
Efficient Synthesis of 2-OH Thioglycosides from Glycals Based on the Reduction of Aryl Disulfides by NaBH4
Next Article in Special Issue
Progress in C-C and C-Heteroatom Bonds Construction Using Alcohols as Acyl Precursors
Previous Article in Journal
A Modeling and Analysis Study Reveals That CaMKII in Synaptic Plasticity Is a Dominant Affecter in CaM Systems in a T286 Phosphorylation-Dependent Manner
Previous Article in Special Issue
Asymmetric Michael Addition in Synthesis of β-Substituted GABA Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Indane-1,3-Dione: From Synthetic Strategies to Applications

Aix Marseille Univ, CNRS, ICR, UMR 7273, F-13397 Marseille, France
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(18), 5976; https://doi.org/10.3390/molecules27185976
Submission received: 10 August 2022 / Revised: 1 September 2022 / Accepted: 6 September 2022 / Published: 14 September 2022
(This article belongs to the Special Issue Featured Reviews in Organic Chemistry)

Abstract

:
Indane-1,3-dione is a versatile building block used in numerous applications ranging from biosensing, bioactivity, bioimaging to electronics or photopolymerization. In this review, an overview of the different chemical reactions enabling access to this scaffold but also to the most common derivatives of indane-1,3-dione are presented. Parallel to this, the different applications in which indane-1,3-dione-based structures have been used are also presented, evidencing the versatility of this structure.

1. Introduction

Indane-1-3-dione is among one of the most privileged scaffolds in chemistry, as the derivatives of this structure can find applications in various research fields ranging from medicinal chemistry, organic electronics, photopolymerization, to optical sensing and non-linear optical (NLO) applications. One of its closest analogues, namely indanone, is commonly associated with the design of biologically active compounds [1,2,3]. The most relevant examples in this field are undoubtedly Donepezil, which is still under use for the treatment of Alzheimer’s disease [4], or Indinavir, which is used for the treatment of AIDs disease [5]. Interest for indanone derivatives is notably motivated by the fact that this structure can be found in numerous natural products (Caraphenol B isolated from Caragna sinica [6], Pterosin B isolated from marine cyanobacterium [7], another derivative extracted from filamentous marine cyanobacterium Lyngbya majuscula) [7], sustaining the interest for these compounds [8,9]. Due to the similarity of structure with indanone, indane-1,3-dione is also of high current interest, and this molecule has also been extensively studied as a synthetic intermediate for the design of many different biologically active molecules [10]. Beyond its common use in the design of biologically active molecules, indane-1,3-dione is also an electron acceptor widely used for the design of dyes for solar cells applications, photoinitiators of polymerization or chromophores for NLO applications [11]. As an interesting feature, indane-1,3-dione possesses an active methylene group, making this electron acceptor an excellent candidate for its association with electron donors by means of Knoevenagel reactions [12]. Ketone groups can also be easily functionalized with malononitrile, enabling to convert it as a stronger electron acceptor. In this review, an overview of the different chemical modifications performed on the indane-1,3-dione core is reported. Following the description of the synthetic access to the indane-1,3-dione derivatives, their uses in the design of biologically active molecules and organic dyes for various applications in organic electronics are reported. Finally, in the last part, an extensive scope of applications is detailed.

2. Chemical Modification of the Indane-1,3-Dione Core

2.1. Synthesis of Indane-1,3-Dione

Indane-1,3-dione can be synthesized following different synthetic procedures. Furthermore, the most straightforward one consists in the nucleophilic addition of alkyl acetate 2 on dialkyl phthalate 1 under basic conditions, enabling to produce the intermediate 2-(ethoxycarbonyl)-1,3-dioxo-2,3-dihydro-1H-inden-2-ide anion 3. Then, by heating under acidic conditions, this intermediate can be hydrolyzed and decarboxylated in situ, producing indane-1,3-dione 4 in ca. 50% yield for the two steps [13,14]. However, several procedures were also reported to access 4 by oxidation of indane 5, using various oxidizing systems such as N-hydroxyphthalimide (NHPI) and tert-butyl nitrite (t-BuONO) [15], H2O2 with a Mn catalyst [16], pyridinium dichromate (PCC) in the presence of Adogen 464 and sodium percarbonate (Na2CO3·1.5 H2O2) [17]. Furthermore, in these different cases, reaction yields remained often limited (17 and 18% yields) while requiring expensive reagents so that the first procedure remains undoubtedly the most popular one to obtain indane-1,3-dione 4 in acceptable yield (see Scheme 1). Recently, a two-step procedure was developed, starting from 2-ethynylbenzaldehyde 6 [18]. By means of a Cu-catalyzed intramolecular annulation reaction, 3-hydroxy-2,3-dihydro-1H-inden-1-one 7 was prepared in 87% yield and a subsequent oxidation of 7 with Jones’ reagent enabled to obtain 4 in 95% yield. As alternative, o-iodoxybenzoic acid (IBX) can also be used as an oxidant, providing 4 in similar yield (92%) [19]. Several strategies were also developed to convert phthalic anhydride 8 into 4 using diethyl malonate 9 and montmorillonite KSF clay [20] or ethyl acetoacetate 11 in the presence of acetic anhydride and triethylamine [21]. In the case of substituted indane-1,3-diones, two distinct routes were developed depending on the substituents attached on the phthalic anhydrides. Thus, in the case of electron-withdrawing groups such as nitro and chlorine (12–14), the condensation of malonic acid in pyridine proved to be a straightforward route to access 19–21 [22]. Conversely, in the case of electron-donating groups such as alkyl substituents (22, 23), a specific procedure was developed, consisting [23] in a Friedel–Craft reaction of 4-methylbenzoyl chloride 22 or 3,4-dimethylbenzoyl chloride 23 with malonyl dichloride 24, followed by an acidic treatment with concentrated hydrochloric acid furnishing after purification of the two compounds 25 and 26 in 33 and 15% yield, respectively (see Scheme 2).

2.2. Chemical Engineering around the Ketone Groups

2.2.1. Functionalization with Cyano Groups

2-(3-Oxo-2,3-dihydro-1H-inden-1-ylidene)malononitrile 28 [24,25,26,27,28] and 2,2′-(1H-indene-1,3(2H)-diylidene)dimalononitrile 29 [28,29,30] can be synthesized by a Knoevenagel reaction of malononitrile 27 on 4 in ethanol using sodium acetate or piperidine as the bases. In the two cases, an excess of malononitrile was used, and the selection between the di- and the tetracyano-substituted derivative could be obtained by controlling the reaction temperature. Thus, the dicyano compound 28 can be obtained at room temperature contrarily to the tetracyano one 29 that is synthesized by heating the reaction media. In the case of 28, reaction yields ranging between 61 and 85% were determined, whereas 29 could be obtained with reaction yields ranging from 34 to 45% yield. Functionalization of substituted indane-1,3-diones (30, 32, 34) with only one dicyanomethylene group was also examined, and several situations were found. Thus, the Knoevenagel reaction furnished a mixture of inseparable isomers 31,31′, 33,33′ and 35,35′ when 5-methyl-1H-indene-1,3(2H)-dione 30 [23], ethyl 1,3-dioxo-2,3-dihydro-1H-indene-5-carboxylate 32 [31] or 5-fluoro-1H-indene-1,3(2H)-dione 34 [32] were used as the starting materials (see Scheme 3).
Conversely, 5-alkoxy-1H-indene-1,3(2H)-diones 36a furnished selectively 37a in 63% yield as a result of the specific activation of one of the two ketones by mesomeric effects [33]. A similar behavior was also observed during the synthesis of 39 [32], 41 [34] and 43 [35], their precursors 38, 40 and 42 being substituted with halogens. The steric hindrance generated by the substituents was another strategy to control the regioselectivity, and the synthesis of 45 is a relevant example of this (see Scheme 4) [36]. Concerning the symmetrically substituted indane-1,3-diones, those bearing halogens at the 5,6-positions were the most commonly studied, as exemplified with compounds 47 [37] or 49 [32,38].

2.2.2. Self-Condensation of Indane-1,3-Dione: The Bindone Adduct

Bindone 50 is an electron acceptor widely used due to its stronger electron-withdrawing ability compared to that of 29. It can be easily synthesized by self-condensation of indane-1,3-dione 4 in basic conditions (triethylamine [25], sodium acetate [39], sodium hydride [40]), or in acidic conditions (sulfuric acid [41]) (see Scheme 5).

2.2.3. Formation of bis-Thiazoles and bis-Thiazolidinone

Indane-1,3-dione 4 and corresponding derivatives have been extensively studied for their biological activities ranging from antitumor, antibacterian and anti-inflammatory activities [13,42,43,44,45]. Parallel to this, 1,3,4-thiadiazole groups also exhibit biological activities [46,47,48,49,50,51,52,53,54] so that their combinations were examined [55]. From a synthetic viewpoint, bis-thiazoles could be obtained in two steps starting from indane-1,3-dione 4. By first reacting 4 with hydrazinecarboxamide 51 in ethanol, in the presence of triethylamine, 2,2′-((1H-indene-1,3(2H)-diylidene)bis(hydrazine-1-carboxamide) 52 could be obtained. Then, upon reaction with a number of N-aryl-2-oxopropane-hydrazonoyl chloride derivatives 53–55, bis-thiazoles 56–58 could be isolated with reaction yields ranging from 78 to 89%. Using a similar procedure, reaction of 52 with ethyl (N-arylhydrazono)chloroacetate 59–61 could furnish the corresponding bis-thiazolidinone 62–64 in high yields (79–90%) (see Scheme 6).

2.3. Chemical Engineering around the Aromatic Groups

2.3.1. Polyaromatic Structures

To improve the electron-accepting ability of 4, an alternative to the substitution of indane-1,3-dione with malononitrile consists in developing polyaromatic structures. Notably, naphthalene derivatives were prepared, following the same synthetic route to that used for 4, consisting in the condensation of ethyl acetate 66 on diethyl naphthalene-2,3-dicarboxylate 65 in solvent-free and basic conditions. After decarboxylation, 67 could be prepared in 91% yield for the two steps [12,56,57,58,59,60]. Introduction of lateral groups onto 67 is possible but involves a specific route to be developed. A Diels–Alder reaction between l,3-diphenylbenzo[c]furan 69 and cyclopent-4-ene-l,3-dione 70 furnishes the 1,3-diphenylbenzo[c]furan-cyclopent-4 ene-1,3-dione adduct 71. By dehydration in acidic conditions (HCl/H2SO4), 71 could be converted to 72 in 25% yield. Recently, the design of a helical-shaped structure 75 was reported [36]. If the structure is innovative, the synthesis is identical to that used for 4, starting from dimethyl naphthalene-1,2-dicarboxylate 73 (see Scheme 7).

2.3.2. Halogenated Indane-1,3-Diones

Halogenation of indane-1,3-dione derivatives subsequent to their synthesis is not possible such that such derivatives can only be obtained by first introducing halogens onto their corresponding precursors. Notably, as a first synthetic approach, halogenated phthalic anhydrides were converted as indane-1,3-diones using ethyl acetoacetate, and a series of halogenated indane-1,3-diones 76–82 is presented in Scheme 8 [20,21,28,37,38,40,61]. Parallel to this, the strategy previously mentioned that AlCl3-promoted acylation of a benzoyl chloride derivative (83) with malonyl chloride 24 proved to be another effective approach to design chlorinated indane-1,3-dione derivatives (84) (see Scheme 8).

2.3.3. Introduction of Various Electron-Withdrawing Groups on Aromatic Ring

Nitration

As previously mentioned for halogenation, electrophilic aromatic substitution cannot be carried out on indane-1,3-dione 4 such that a post-functionalization with nitro groups is required. To date, only few indane-1,3-diones bearing nitro groups have been reported in the literature (see Scheme 9) [21,22,62].

2.3.4. Cyanation

To the best of our knowledge, no cyano-substituted indane-1,3-dione derivatives have been reported to date. Furthermore, such acceptors are as crucial as the cyano groups, which are among the best electron-accepting groups.

2.3.5. Introduction of Alkoxy-Carbonyl Groups

Here again, only the post-functionalization of naphthalic anhydrides was used to introduce CO2R groups. An example is provided below with 32 (see Scheme 10) [31]. By using ethyl acetoacetate 11, anhydride acetic as the solvent and triethylamine as the base, 32 could be obtained from 32a in 71% yield.

2.4. Chemical Engineering around the Methylene Group

2.4.1. Knoevenagel Reaction

Due to the presence of the two ketones groups on both sides of the methylene groups, indane-1,3-dione 4 possesses a privileged group for realizing Knoevenagel reactions. In the case of indane-1,3-dione 4 and its substituted derivatives, the condensation reaction can be carried out in the conditions initially used by Knoevenagel in 1894 to condense benzaldehyde with ethyl acetoacetate 11, namely in ethanol with a catalytic amount of piperidine [61]. Typically, Knoevenagel reactions performed with 4 or 68 can be realized with reaction yields higher than 70% (see Table 1 and Scheme 11) [12]. As the main interest of this reaction, use of a highly polar solvent favors the precipitation of dyes 95–114 upon cooling, and the reaction can be carried out in green conditions since a non-dangerous solvent can be used. Additionally, the work-up can be limited to a simple filtration, avoiding the use of complicated purification processes [56,57].
When 2-(3-oxo-2,3-dihydro-1H-inden-1-ylidene)malononitrile 28 and its analogues are involved in Knoevenagel reactions, another amine should be used, and diisopropylethylamine (DIPEA), which is a none-nucleophilic base, is the most popular one. As reported in several recent works, an unexpected nucleophilic addition of secondary amines onto the cyano groups of the push–pull dyes can occur, giving rise to a cyclization reaction and producing 3-(dialkylamino)-1,2-dihydro-9-oxo-9H-indeno [2,1-c]pyridine-4-carbonitrile derivatives 115, according to the mechanism proposed in Scheme 12.
Numerous examples of undesired cyclization reactions have notably been reported with piperidine providing 3-(dialkylamino)-1,2-dihydro-9-oxo-9H-indeno [2,1-c]pyridine-4-carbonitrile derivatives 115 instead of the expected push–pull dyes [62,63,64]. In 2019, an unprecedented nucleophilic addition of piperidine on 101 was also reported, providing 102 after dehydration. The mechanism supporting the formation of this unexpected structure is depicted in Scheme 13, and the crystal structure of this molecule presented in Figure 1 undoubtedly proved its formation.
When 2,2′-(1H-indene-1,3(2H)-diylidene)dimalononitrile 29 and its derivatives are engaged in Knoevenagel reactions, the high stability of their anions in basic conditions impedes the Knoevenagel reactions to proceed. Therefore, acidic conditions should be used, and acetic anhydride is commonly used in this aim (see Scheme 14) [65,66].
Besides, several reports mention the use of the classical piperidine/ethanol conditions to condense 29 onto aromatic aldehydes, despites the strong deactivation of the 29 anion in basic conditions. A few examples of products obtained in these conditions are presented in Scheme 15 (121 [67], 123 [67] or 125 [68]).
To avoid the use of base, several authors replaced ethanol by solvents of higher boiling points such as methylethylketone (see Scheme 16) [69]. By refluxing 29 and 126 at elevated temperature, 127 could be obtained in 42% yield.

2.4.2. Oxidation Reaction

Among indane-1,3-dione derivatives, ninhydrin is well known, as it can be advantageously used as a revelator for thin layer chromatography (TLC). Over the years, several strategies have been developed to access to these structures [70]. For instance, oxidation of indane-1-one 128 with selenium oxide (SeO2) can furnish ninhydrin 138 [71], but this reaction was not limited to 4, and the oxidation reaction tolerates various substituents such as OMe, tert-Bu, Me, Br, CF3 or NO2 (see compounds 139–147, Scheme 17) [72]. Direct oxidation of 4 was also investigated to convert it, as 138 and ninhydrin 138 could be prepared in 48% yield using the dual oxidizing system SeO2/H2O2 [73] in 40% yield for the two steps using iodobenzene diacetate [74] and 94% yield using N-bromosuccinimide (NBS) in DMSO [75]. However, all attempts to oxidize 4 as 138 using sodium hypochlorite failed, and phthalic acid was isolated as the unique product of the reaction [76]. The authors also demonstrated 138 to be oxidized as phthalic acid, evidencing that sodium hypochlorite is a too strong oxidant. The oxidation reaction with N-bromosuccinimide (NBS) in DMSO was not limited to indane-1,3-dione 4, and indane-1-one 149 and indane-2-one 128 could also be oxidized using the same procedure, providing 138 in 82 and 85% yield, respectively. Photooxidation of 4 in the presence of tetrabutylammonium hexafluorophosphate and oxygen using Rose Bengal as the photosensitizer could convert 4 as 138 in 75% yield upon irradiation with a UV light [77].

2.4.3. Halogenation

α,α-Dihalogenation reactions of carbonyl compounds have been extensively studied in the literature [78,79,80,81,82,83,84,85,86,87,88], and different procedures were thus developed for the halogenation of indane-1,3-dione 4. Notably, traditional reagents of halogenation including N-chlorosuccinimide (NCS) or NBS in ethanol could provide 150 and 151 in 95% and 92% yields [89]. Several green syntheses were also developed, all based on mechanical ball milling. Using this approach, halogenating agents such as trichloroisocyanuric acid or tribromoisocyanuric acid could furnish 150 and 151 in 98 and 97% yields [90]. Similarly, mechanosynthesis of 150 could be efficiently achieved by employing sodium bromide and oxone (98% yield) [85] or ammonium bromide and oxone (see Scheme 18) [91].
The conversion of nucleophilic halogens to electrophilic ones could be realized by reacting Lewis acids such as ZnBr2 or AlCl3 with lead tetraacetate [92]. High reaction yields were also obtained during the synthesis of 150 while using KBr/KBrO3 (86% yield) [93], 1,3-dibromo-5,5-dimethylhydantoin in acetic acid (88% yield) [94]. Similarly, 151 could be obtained while reacting 4 with 1,3-dichloro-5,5-dimethylhydantoin in acetic acid (89% yield) [94]. Finally, selectfluor® (1-chloromethyl-4-fluoro-1,4-diazoniabicyclo [2.2.2]octane bis(tetrafluoroborate)) was the most widely studied fluorinated agent for fluorination of 4 in water while using a surfactant (Genapol LRO) (74% yield) [95], sodium dodecyl sulfate in water (93% yield) [96], or acetonitrile as solvent (60% yield) (see Scheme 19) [41,97]. Furthermore, numerous drawbacks concerning the electrophilic fluorination with selectfluor® were reported in the literature. Notably, parallel to the formation of the expected F+ cation, formation of radical species (F) by single electron transfer has also been proposed, even if the mechanistic studies have not fully elucidated the mechanism. Nevertheless, formation of an intermediate monofluorination state could be demonstrated [98].
In the case of indane-1,3-dione 4, fluorination was proposed as occurring by means of an attack of the double bond of enol onto selectfluor®, followed by a deprotonation with the resulting diazoniabicyclo [2.2.2]octane species. By iterating the reaction a second time, 152 could be obtained (see Scheme 20) [97]. Considering that there is still a lack of efficiency for the α,α-dibromination of 1,3-diketones, the photoredox catalysis was envisioned as a possible alternative to conventional chemistry to improve the selectivity during bromination [99,100,101,102]. Light is also a traceless reagent so that light-promoted chemistry perfectly fits with the concepts of green chemistry. For bromination, light-activated radical reactions are also extensively described in the literature, enabling to efficiently generate bromine radicals.
In 2019, an interesting reaction was developed to accelerate the formation of bromide radicals and, in this aim, N-bromosuccinimide (NBS) was irradiated in the presence of bromoacetic acid and under visible light [103]. By this unique combination, two concurrent bromination mechanisms could be evidenced, the first one consisting in an electrophilic bromination resulting from the photoassisted formation of bromine, and the second one consisting in a radical bromination process resulting from the formation of bromide radicals promoted by light (see Scheme 21).
By the presence of these double sources of brominating agents, the exceptional reaction yield of 99% could be obtained as well as accounts from the coexistence of both the electrophilic and the radical bromination reactions (see Scheme 22).

2.4.4. Cyanation

In 1977, an interesting procedure was developed to introduce a cyano group at the methylene position of indane-1,3-dione. Inspired by the synthesis of indane-1,3-dione 4 starting from dialkyl phthalate and ethyl acetate, an analogue procedure was proposed, where ethyl acetate 51 was replaced by acetonitrile (see Scheme 23) [104]. Using sodium methanoate as the base, 153 could be obtained in 78% yield.

2.4.5. Nitration

Nitration of the methylene group of indane-1,3-dione 4 can be performed in one step, by using nitric acid as the reagent [105]. Furthermore, 154 could be obtained in 78% yield (see Scheme 24).

3. Indane-1,3-Diones as Reagents for Various Chemical Transformations

3.1. Synthesis of Cyclophanes

Cyclophanes play an important role in supramolecular chemistry [106,107,108,109,110,111,112,113,114,115,116,117,118,119], especially as hosts for host–guest strategies so that a continuous effort has been made to develop new, efficient and simple synthetic routes to access to these structures [120,121,122,123,124,125,126]. Among the main methods reported in the literature, Wurtz coupling [127] carbene insertion [128,129], Ni-catalyzed Grignard coupling [130], polymerization of p-xylene [131], pyrolysis of sulfones [132] or acyloin condensation [133] can be cited as popular reactions. Conversely, Suzuki-Miyaura cross-coupling reaction [134,135,136,137] or ring-closing metathesis [138,139,140,141,142,143,144,145,146,147] have been less studied. In 2012, a combination of these two key reactions (Suzuki-Miyaura cross-coupling reaction and ring-closing metathesis) has been conceived as a concise and efficient synthetic route to access to [4,4]-cyclophane derivatives [148]. In the first step, dialkylation of indane-1,3-dione 4 with meta or para-bromobenzyl bromide furnished 155a and 155b. The functionalization of the methylene group of indane-1,3-dione 4 was not an easy task since conventional alkylation conditions such as NaH in THF totally failed. After several attempts, use of freshly prepared KF-celite in dry acetonitrile enabled to produce 155a and 155b in moderate yields, 74 and 70%, respectively. These two intermediates were subsequently functionalized with allyl groups, by a Suzuki-Miyaura cross-coupling reaction using an excess of allylboronic acid pinacol ester. Finally, metathesis reactions of 156a and 156b using the first generation of Grubbs catalyst and titanium isopropoxide delivered the two macrocycles 157a and 157b in 54 and 45% yields. Finally, hydrogenation using a catalytic amount of Pd(C) yielded 158a and 158b in 80% for the two cyclophanes (see Scheme 25).

3.2. Synthesis of Crown Ether Derivatives of Indane-1,3-Dione

Compounds capable of changing the optical properties by complexation with metal cations are at the center of numerous researches, and in this field, crown ethers consisting in a ring containing ether groups capable to bind alkali cations have been extensively studied [149,150,151]. Considering that indane-1,3-dione 4 is a strong electron-acceptor, its combination with an electron donor connected to a crown ether could make the final assembly an interesting structure for ion sensing. Such a push–pull dye was reported in 2010 by Mitewa and coworkers [152]. Thus, 159 could be prepared by a Claisen–Schmidt condensation of 2-acetyl-1,3-indandione 160 with the appropriate aldehyde 161 in acetic acid. The different attempts to use piperidine as the base failed, and the starting materials 160 and 161 were entirely recovered after reaction. Unexpectedly, the Claisen–Schmidt reaction furnished a fully conjugated molecule, resulting from a deacetylation reaction, according to the mechanism depicted in Scheme 26.
To prepare 2-(1-hydroxyethylidene)-1H-indene-1,3(2H)-dione 162, several strategies can be developed, as shown in Scheme 27. Notably, a selective cyclo-oligomerization based on a self-condensation of acetyl chloride 163 in the presence of aluminum trichloride AlCl3 followed by a cross-condensation reaction with benzoyl chloride 164 could provide 162 in 87% yield according to the mechanism depicted in Scheme 27 [153]. Parallel to this, 162 can also be synthesized using the Kilgore procedure [154], consisting in an addition–elimination process of a ketone (acetone 166) on diethyl phthalate 165 under basic conditions (sodium ethoxide) (see Scheme 27) [155].
Finally, examination of the complexation of 159 with various divalent cations revealed no drastic changes of its absorption spectrum upon complexation, except with Sr2+ and Ba2+. This unexpected result was assigned to the low involvement of the nitrogen lone pair of the crown ether in the coordination of the cations, thus weakly affecting the optical properties of the push–pull dye connected to the crown ether.

3.3. Synthesis of Tetracycline Heterocyclic Analogues

Tetracyclines are an extended family of compounds discovered and developed in the mid-1940s for their promising therapeutic properties [156]. Notably, tetracyclines are efficient as antibiotics and anti-malarial drugs, and these structures have also been identified as being beneficial of various pathologies such as cancers or Parkinson’s disease. Due to the increasing resistance of microbes to antimicrobials used from long ago, analogues to tetracyclines are actively researched, and a series of pentacyclines comprising the indane-1,3-dione motif have been synthesized [25]. The synthesis was relatively straightforward since the combination of malononitrile 27, salicylaldehyde 167 and indane-1,3-dione 4 in a one-pot procedure furnished 6-amino-7-imino-7H-indeno [2′,1′:5,6]pyrano [3,4-c]chromen-13(13bH)-one 168 in 80% yield. Imine group in 168 could be easily cleaved in acidic conditions, yielding 169 (see Scheme 28).
The one-pot synthesis of pentacyclines was not limited to 4 as the starting material, and another derivative 172 was also prepared starting from 28 following two different synthetic routes in order to confirm its chemical structure (see Scheme 29). Thus, following a first step consisting in the reaction of 28 with the diazonium salts 170 and 173, the two products 171 and 174 were subjected to an intramolecular cyclization reaction by reflux in ethanol in the presence of a catalytic amount of piperidine, providing 172 and 175 in 85 and 87% yields, respectively. Finally, hydrolysis of imine 175 under acidic conditions furnished 172 in 90% yield and confirmed the formation of this compound by the similitude of the 1H NMR spectra.
On the basis of the two synthetic routes, other tetracycline derivatives comprising three fused cycles were also prepared, as exemplified with the indenopyrane derivatives 177 and 179 or the indeno-[2,1-c]-pyridazines 181 and 182 listed below (see Scheme 30).

3.4. Synthesis of Indane-1,3-Dione Derivatives via Tert-Butylisocyanide Insertion

Isocyanide insertion into palladium–carbon bonds has emerged as an effective strategy to form C-N, C-O or C-C bonds [157,158,159,160,161,162,163], and numerous derivatives such as quinazolinones [164], 6H-isoindolo [2,1-a]indol-6-ones or indenoindolones [165] have been prepared using this strategy. In this context, the chemoselective insertion of tert-butyl isocyanide 183 to form C-C has been examined to form indane-1,3-dione derivatives starting from 1-(2-bromophenyl)-2-phenylethanone 184 [166]. This reaction tolerated various substituents since good reaction yields were obtained with 1-(2-bromophenyl)-2-phenylethanone substituted with various electron rich and electron deficient groups (see Scheme 31). After hydrochloric acid hydrolysis, indane-1,3-diones substituted at the 2-position with various aromatic rings could be obtained with reaction yields ranging from 61% to 75% (185–192). Starting from 1-(2-bromophenyl)ethan-1-one derivatives, indane-1,3-dione with aliphatic substitutes at the 2-position could also be prepared (190, 191). Finally, tolerance of this reaction to the cyano group was also evidenced (192). In this last case, 192 was not hydrolyzed as in the other case.

3.5. Copper-Catalyzed Sulfonamidation of Benzylic C-H Bonds

The transformation of a C-H bond into a C-N bond is an important reaction in organic chemistry, and the amination of primary benzylic groups has not been excluded from this interest [167]. Especially, sulfonamidation of primary benzylic groups with primary and secondary sulfonamides could give access to molecules of biological interest. However, methodologies for sulfonamidation of primary benzylic groups remain scarce in the literature [168,169], and most of the reactions only afford dramatic low yields (9–30%) [168,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187]. Conscious of the paucity of methodologies available, Powell et al. proposed in 2010 after substantial efforts, optimized conditions for the sulfonamidation of toluene 193 with N,4-dimethylbenzenesulfonamide 194 at room temperature [188]. A reaction yield as high as 70% could be obtained for the synthesis of 196 while selecting the copper catalyst Cu(CH3CN)4.PF6, the oxidant tert-butyl 3-(trifluoromethyl)benzoperoxoate 195 and the ligand, namely indane-1,3-dione 4 (see Scheme 32). Choice of the ligand was determined as being crucial, other ligands such as bathophenanthroline or 1,3-diphenylpropanedione only allowing a conversion of 33% and 52%, respectively. If indane-1,3-dione 4 was capable to enhance the overall yield, its exact role remained unclear, 4 being only capable to act as a monodentate ligand for copper. The substitution pattern of the oxidant was also determined as being of prime importance, and the presence of electron-withdrawing groups such as the trifluoromethyl group on tert-butyl 3-(trifluoromethyl)benzoperoxoate 195 could weaken the perester bond, thereby facilitating the formation of the tert-butoxy radical [189].

3.6. Michael Addition on β-Substituted meso-Tetraphenylporphyrins

meso-Tetraarylporphyrins are extensively studied due to their facile synthesis, high molar extinction coefficients, high photoluminescence quantum yields, excellent photochemical stability, but also for their ability to chelate a wide range of metal cations in their inner core [190,191].
An efficient strategy to finely tune their optical properties consists in the modification of the porphyrin core by means of changing the π-conjugation length [192,193,194], the introduction of peripheral groups [195,196,197,198,199,200,201] or breaking the planarity [202,203,204,205]. In this field, insertion of a nitro group at the β-position porphyrin macrocycle proved to be an effective approach to introduce various substituents on the porphyrin core. Notably, 2-nitro-5,10,15,20-tetraphenylporphyrin 197 and 198 could undergo a variety of nucleophilic aromatic substitution, the macrocycle behaving as a Michael acceptor. While using indane-1,3-dione 4 as the nucleophile, the nucleophilic substitution could undergo a variety of metalated and non-metalated porphyrins (175 and 176) according to the mechanism depicted in Scheme 33 [206].
As a result of the addition of indane-1,3-dione 4 onto porphyrins, meta-chlorins that differ from porphyrins by the reduction of one of the pyrrole ring could be obtained [207,208]. trans-chlorins are extensively studied due to their red-shifted absorption compared to their porphyrin analogues, enabling to design dyes with an infrared absorption [194,209,210,211,212,213]. The six chlorins could be prepared with reaction yields ranging from 67 to 87% yield. Lastly, the same authors reported an unprecedented ring-fusion of trans-chlorins bearing 1,3-indanedione functionalities. Upon addition of Ni cations inside the porphyrin core, trans-chlorins could undergo a skeletal rearrangement of the porphyrin macrocycle, and trans-chlorins could be converted to fused metalloporphyrins by elimination of one indane-1,3-dione unit (as shown in Scheme 34) [214]. By controlling the reaction time, the metalation of trans-chlorins (15–20 min) or the nickel insertion, followed by an indane-1,3-dione elimination and a ring-fusion (3–4 h) furnished in turn the fused metalloporphyrins. To clarify the role of the nickel cation in this mechanism, reflux in DMF of trans-chlorins 199-Zn, 199-Cu, 199-Ni, 200-Zn, 200-Cu, 200-Ni, 201-Zn, 201-Cu, 201-Ni with nickel acetate converted all trans-chlorins to the fused metalloporphyrins (202-Ni, 203-Ni and 204-Ni). Metalloporphyrins containing other divalent cations were also prepared, by first demetalating 202-Ni, 203-Ni and 204-Ni and then remetalating with the appropriate metal acetates (Co, Cu, Zn).

3.7. Synthesis of 1,4-Isochromandione

1,4-isochromandione is an important heterocyclic compound, as this molecule is the starting material for the synthesis of numerous biologically active compounds such as parvaquone, which is an antiprotozoal agent marketed as Clexon [215,216], or atovaquone, which is an anti-pneumocystic agent marketed as Mepron [217]. In this context, various strategies have been developed to access this elemental building block. The first report mentioning the synthesis of 1,4-isochromandione 208 was reported in 1966 by Holt et al. using indane-1,3-dione 4 as the starting materials, and 208 could be obtained in 86% yield [218]. Following this pioneering work, several improvements were performed in order to improve the reaction yield. Typically, by diazotation of 4 with tosyl azide [219], 2-diazo-1H-indene-1,3(2H)-dione 210 can be obtained with reaction yields ranging from 59% [220] in ethanol to 80% [221] in triethylamine, 88% [222] in THF and finally 93% for the best conditions in ethanol [223]. By treating first 210 in basic conditions, 211 could be obtained, and treatment in a second step with sulfuric acid could furnish 208 in 83% for the last step (see Scheme 35) [224,225].

3.8. Synthesis of Benzofurans by Electrooxidation of Hydroquinone Derivatives

Benzofurans are important compounds, as these structures are widely used for the treatment of cardiac arythmias, and amiodarone is a relevant example of benzofurans used for this purpose [226,227]. However, benzofurans are not restricted to these applications, and benzofurans are also reported as having numerous pharmaceutical applications so that the synthesis of these derivatives has been widely studied [228]. More generally, benzofurans can also be used as fluorescent probes [229], antioxidants or brightening agents [230,231]. In 2015, an electrochemical synthesis of 212 and 213 was reported by Ameri et al. consisting in oxidizing in situ hydroquinones 214 and 215 as benzoquinones in a phosphate buffer (pH = 7), thus acting as a Michael acceptor (see Scheme 36 and Scheme 37) [232]. Depending on the substitution of hydroquinone, one or two 1,4-michael additions could occur, according to the mechanism depicted in Scheme 37.
Compounds 212 and 213 could be isolated in high yields, 86 and 84%, respectively. By combining several electrochemical techniques, the presence of an ECECECEC and of an ECEC mechanism was proven for 212 and 213, respectively.

3.9. Combination of Knoevenagel Condensation and Michael Addition Reactions

In a purpose of synthetizing complex organic structures involving affordable building-blocks, multicomponent reactions (MCRs) have played an important role over the last twenty years. With the emergence of green chemistry purposes, the synthetic strategies developed to access to complex structures have to be shortened and the use of metal catalysts and organic solvents limited. To address these issues, MCRs constitute a powerful strategy but also an expeditious method enabling to rapidly generate a vast library of molecules by systematically changing one of the three reactants involved in this convergent approach [233]. Several examples of multicomponent reactions making use of indane-1,3-dione 4 as one of these substantial building blocks for MCRs have been reported during the last decades. Typically, MRCs consisted in the combination of Knoevenagel condensation followed by a cyclisation reaction resulting from a Michael addition. Considering the similarity of structures of all these cyclized compounds obtained during these different works in the presence of an identical moiety, i.e., indane-1,3-dione 4, these structures can be combined under the generic name of “indeno-fused structures” (see Scheme 38). From a structure viewpoint, interest for these heterocycles was notably motivated by their interesting properties in medicinal chemistry, these molecules possessing anti-bacterial, anticancer or cardiovascular activities [226,228,234,235]. Biological applications of these different indeno-fused structures are discussed in the section devoted to the different applications of indane-1,3-dione derivatives.
As first examples of MRCs are those that were devoted to the synthesis of quinolinone derivatives (see Scheme 38, structures 216, 217). In this work, Sandaroos and coworkers used iron triflate (Fe(OTf)3) as the Lewis acid catalyst, and the reaction could be conducted in solvent free-conditions [234]. Choice of iron triflate as the catalyst was supported by the weak nucleophilic character of the triflate anion, making the metal cation a stronger Lewis acid. The reactions performed at 90 °C for 4 h could provide the products with reaction yields ranging from 80 to 92% after purification. Control experiments performed without Fe(OTf)3 also revealed the MRCs not to proceed, highlighting the crucial role of the Lewis acid in the activation process. The authors could reuse the metal catalyst without any loss of catalytic activity, but no precision about the number of cycles examined is given. In an attempt to optimize the catalytic activity, several other metal triflates such as Zn(OTf)2, Cu(OTf)2 were examined, but iron triflate remained the most effective one. To obtain a deeper insight into the mechanism, the reaction could be successfully decomposed into two steps, either by mixing indane-1-3-dione 4 and the aldehyde (225) but also the aldehyde and the amine (226) in the first step (see Scheme 39). Although both synthetic pathways remain possible, the expected compound 217b could be obtained in the two cases.
Iron plays without contest an important role in catalysis, not only as a heterogenous catalyst but also for the design magnetic nanocomposites. Iron can also be used not only for its remarkable reactivity but also for the easy recovery of the metal catalyst. Such a strategy has been reported in an MCR synthesis of indenoquinoline-8-one derivatives. The Lewis acid developed in this work, namely sulfonic acid-functionalized cellulose-coated Fe3O4 (Fe3O4@cellulose-SO3H) nanoparticles, could be easily removed from the reaction media by use of an external magnet (see Scheme 38, structures 218) [235]. The authors optimized this synthesis be rendering it applicable in solvent-free conditions but also in water media, with sulfonic groups covering the metal particles for water compatibility. Using these magnetic Fe3O4 nanoparticles, the desired products could be obtained within 5 min. at 40 °C. Additionally, after removal of the magnetic particles with a magnet, the final product could be purified by a simple recrystallization in EtOH. The metal catalyst proved to be reusable, but a reduction of the reaction yield was nevertheless noticed. Thus, by repeating the synthesis of 218d, the reaction yield decreased from 95 to 82% yield after five runs. Nevertheless, contrarily to what was observed for the 216/217 series, an aromatization of the structure was observed, leading to the formation of 4-azafluorenones (see Scheme 38, structures 218). Here again, the role of iron particles was essential, by activating the aromatization reaction (see Scheme 40).
Similarly to the strategy applied for the easy recovery of the Fe3O4@cellulose-SO3H particles, Fe3O4@NCs/Cu(II) particles have also been developed for the synthesis of another family of indeno-fused structures, namely 219a–219o [236]. This bio-based catalyst showed remarkable efficiencies in EtOH at 60 °C since reaction yields ranging from 79 up to 97% could be obtained within only 5 min. of reaction. Here again, the recovery was easy since only a magnet and washing with EtOH was required to recover the catalyst in pure form. A good recyclability was found since the catalyst could be reused without significant loss of its catalytic activity, even after 4 runs. Thus, the reaction yield decreased from 95 to 79% yield after 4 runs for 219a.
Even though these metal catalysts were highly efficient, it is always desirable not to use catalysts or to use catalysts that can operate in homogeneous phase. In this field, several examples of ionic liquids (ILs) have been proposed as green and reusable catalysts (see Scheme 33, molecules 220 and 221) [237,238]. Ionic liquids have been proposed as an interesting alternative to the traditional catalysts in a variety of chemical reactions, as these molecules are often less pollutant than metals and can also act both as solvents and catalysts [239,240,241,242]. In the case of these reactions, ILs have thus a dual role of solvent and catalyst. A first example of IL is 1,2-dimethyl-N-butanesulfonic acid imidazolium hydrogen sulfate [DMBSI]HSO4, which was not used as a solvent [237]. Reaction conditions were optimized in ethylene glycol, and the best temperature for the MCR was determined as being 120 °C. In these conditions, the reaction was relatively fast since it could be finished within 4 min. Feasibility of the reaction in solvent-free conditions was examined, but lower reaction yields and longer reaction times were found compared to the results obtained in ethylene glycol. Recyclability of the catalyst was also examined, and after three successive runs, no significant reduction of the reaction yield was found. A few years later, another interesting example was proposed with 1-hexyl-3-methyl-imidazolium iodide [HMIM]I as the ionic liquid. It may be mentioned that in this case, the reaction could be performed in water with high reaction yields (up to 95%) while using sonication as the activation mode [238]. Sonochemistry is not widely used in organic chemistry to activate a large variety of organic reactions due to its appealing features: shorter reaction times, higher reaction yields, less byproduct formed, milder reaction conditions. Even though the kinetic is a bit slower than with the previous IL [DMBSI]HSO4 (reaction performed with reaction time ranging from 4 to 20 min.), the use of water as the solvent and the energy economy achieved while using sonochemistry as the activating mode turned out to be an attractive improvement. However, ILs exhibit a major drawback for large scale syntheses, namely, their relatively high costs. This is the reason why cheaper catalysts are continuously researched. Lastly, indeno-fused structures have been successfully synthesized by mechanosynthesis while simply using p-toluenesulfonic acid (PTSA) as the catalyst [243,244]. One of the most important principles of Green Chemistry consists in the development of environmentally benign synthetic methodologies enabling to avoid the use of solvents, to use environmentally friendly solvents or to reduce the quantity of solvent used. In this field, mechanochemistry is a promising alternative to conventional methods, notably for the synthesis of indeno-fused structures. While using PTSA as the catalyst, the reaction could be successfully performed in solvent-free conditions in a mortar by grinding at room temperature. Reaction yield ranging from 70 to 86% could be determined (see Scheme 33, molecules 224) [243]. PTSA was also used as the catalyst for MCR reactions performed in water at 90 °C for 2.5 h, providing the targeted compounds with higher reaction yields compared to that obtained by mechanosynthesis (see Scheme 33, molecules 224) [244]. Although the reaction yields and the reaction time may be a bit less interesting than that obtained for the different examples presented before, interest of PTSA is that this catalyst need not be prepared, contrarily to the different iron particles or ionic liquids previously mentioned.
A slightly different type of indeno-fused structures deserves to be mentioned in this section, namely the indenopyrimidine derivatives (see Scheme 33, molecules 224) [245]. Even if the central core of pyrimidine contains two nitrogen atoms instead of one for the previous structures, the strategy used to prepare these structures remains the same versus those previously described, consisting in a Knoevenagel reaction activated by the presence of the catalyst, followed by a Michael addition, a cyclization reaction and finally, an aromatization as classically observed for azafluorenones. In 2018, a heterogeneous catalyst, Ag2O–ZrO2, was proposed to catalyze the MCR, where zirconia was used as the support to immobilize Ag2O, which was the key part of this catalyst. Notably, Ag2O was capable to coordinate the aldehyde and favor the condensation of indane-1,3-dione 4 (see Scheme 41). Recyclability of the catalyst was also quite interesting since a reaction yield of 90% could still be obtained after six cycles (starting from 96% yield for the first run). Although the reaction was performed in ethanol for 30 min, at room temperature, the catalyst only needed to be washed with acetone and dried at 100 °C for 3 h before being reused. Parallel to this, all compounds could be purified by recrystallization in ethanol, evidencing once again the compatibility of the green protocols for the synthesis of complex structures, as exemplified with these indenopyrimidine derivatives.

3.10. Indane-1,3-Dione: Versatile Building Block for Spirocyclic Compounds Synthesis

Spirocyclic compounds are chemical structures where two cyclic rings are linked by at least one atom. Such a configuration is strongly present in natural compounds [246]. Spiroindanediones moieties are also present in numerous bioactive compounds [247], and these molecules are efficient as antitumor and antibiotic compounds [248] and antiproliferative molecules [249]. Synthesis of these spiroindanediones can be relatively complex, since the formation of these spiro compounds can led to a wide range of isomers depending on the regioselectivity, the diastereoselectivity and even the enantioselectivity of the reaction. Even if complex mixtures can be obtained during their syntheses, this approach remains however the only one to form elaborated structures. Cycloaddition, domino reaction, and multi-component reactions (MCRs) are the main synthetic pathways leading to spiroindanediones, and these different reactions are described successively. Finally, the other synthetic routes giving access to spiroindanediones are briefly described in the final part (see Scheme 42).

3.10.1. Synthesis of Spiroindanediones by Cycloaddition

Spiroindanediones can be synthesized by various types of cycloadditions. Azomethine ylide is a versatile reactant capable to easily reacting with arylidene-1,3-indanedione in 1,3-dipolar cycloadditions or in [3+2] cycloadditions. Azomethine ylide can also be generated by various procedures, such that this reactant was involved in numerous reactions [250].
It was notably used to synthesize spiropyrrolidines starting from 2-ferrocenylidene-2(H)-indane-1,3-dione 230 [251]. 2-Ferrocenylidene-2(H)-indane-1,3-dione 230 could be synthesized by a Knoevenagel reaction between ferrocene-carboxyaldehyde 229 and indane-1,3-dione 4 (see Scheme 43). Then, by mixing sarcosine 232 and indoline-2,3-dione (isatin) 231, sarcosine 232 could condense with 231 and after decarboxylation, give rise to the azomethine ylide 234. This ylide can thus react with 230 in a [3+2] cycloaddition furnishing the ferrocenyl spiropyrrolidine adduct 235 (see Scheme 44). Using the same strategy, various compounds (236, 238, 240 and 242) could be synthesized in good yields (higher than 75% yields) by refluxing the methanol solutions for 12 h (see Scheme 45).
Single crystal X-ray diffraction analyses performed on different reaction products revealed how this reaction could give the product as a unique regio and stereo isomer. Notably, isatin 231 can react with different amines and give rise to more complex molecules, for example, by reaction with the azomethine ylide resulting from the reaction between ethyl glucinate hydrochloride 243 and dimethyl but-2-ynedioate 244 (see Scheme 46). The azomethine ylide 245 thus obtained is capable to react with an arylidene indanedione 246, as shown in Scheme 46. Such a reactivity was notably used to form various dihydro-spiro[indene-2,3′-pyrrolidines] 247 by a one-pot reaction at room temperature and in polar solvent, using triethylamine as the base. The different products could be obtained with reaction yields ranging between 56% and 69% (see Scheme 47) [252]. Examination of the single-crystal X-ray diffraction patterns and the 2D NMR spectra revealed that one diastereoisomer was mostly formed. Such a diastereoselectivity was assigned to a steric effect induced by the ester group and occurred during the reaction of the azomethine ylide with the indane-1,3-dione adduct during the concerted cycloaddition (see Scheme 48).
Azomethine ylide can also be formed by a decarboxylative condensation of isatin 231 with 1,3-thiazolane-4-carboxylic acid 248 [252]. Such ylides, when formed in situ, can react with various derivatives of 2-arylidene-1,3-indanediones 249 to give bispiro compounds 250 (see Scheme 49) that were tested as inhibitors for M. tuberculosis H37Rv. The reaction yields ranged between 60 and 92% depending on the aryl group. This reaction was also regioselective, furnishing only one diastereoisomer.
Azomethine imine can also be synthesized by condensation of the commercially available 3-pyrazolidinone 251 and benzaldehyde 252, as exemplified with 253 (see Scheme 50) [253]. The resulting azomethine imines 254 could react with different arylidene indane-1,3-diones 255 at room temperature using triethylamine as the organocatalyst. Reaction yields ranging between 65 and 98% could be determined, depending on the substituents [254]. The reaction can operate with a good diastereoselectivity (in most cases (>20/1)), (see Table 2, Scheme 51).
The reaction is also tolerant to a wide range of substituents, and the electronic effects induced by the substituents were determined as having no influence on the cyclization reaction, the mechanism involving a π-π stacking interaction between the two reactants (see Scheme 52).
Iminodiesters can also be used as precursors for the synthesis of azomethine ylides, and an example is given in Scheme 53. This synthesis can be base-directed, giving after a [3+2] cycloaddition of 258 with the arylideneindane-1,3-dione 260 and a lactonization reaction, the chromenopyrrolidine 259 [255]. When DMAP was used as the base, the reaction was driven by electronics factors so that chromeno [3,4-b]pyrrolidines were obtained in these conditions irrespective of the aryl derivatives (see Scheme 54). The substrate was determined as only slightly influencing the reaction yields. Furthermore, when 257 was substituted with fluorine atoms, lower reactions yields were obtained compared to the other substituents (Br, Cl, OMe, Me, NO2, …) (see Table 3 and Scheme 55).
Conversely, when 1,1,3,3-tetramethylguanidine (TMG) was used as the base, chromeno [3,4-c]pyrrolidine 262 was obtained instead of 259, as the steric factors became predominant during the cyclization reaction (see Scheme 54). The reaction products 259 and 262 differ by the intermediate anions 258 and 261 formed by reaction of 257 with DMAP and TMG. During this reaction driven by steric factors, the reaction yield was lower when 257 was substituted with methoxy groups (see Table 4 and Scheme 56).
Influence of the catalysts on the regioselectivity of the reaction was not clearly demonstrated. Furthermore, the formation of H-bonds between reactants was suggested as the key element supporting the formation of a unique isomer. When TMG was used as the catalyst, H-bonding interactions were more pronounced than with the other bases. Various chromeno [3,4-b]pyrrolidines 259 (see Table 3) and chromeno [3,4-c]pyrrolidines 262 (see Table 4) were synthesized using these procedures.
Other 1,3-dipolar cycloadditions were also performed with trifluoropyruvate imines in order to introduce fluorine groups [256]. Starting from ethyl 3,3,3-trifluoropyruvate 263, formation of the imine 264 led to a dipolarophile, which could be used for cycloaddition reactions (see Scheme 57).
After optimization of the reaction conditions, addition of a phase transfer catalyst (benzyltriethylammonium bromide (TEBAB)) 266 and at least one equivalent of LiOH as the base could lead to a complete conversion of the reactants 255 and 265 with an excellent diastereoselectivity as 267 (see Scheme 58).
This procedure was notably tested on various substrates, and the reaction proved to tolerate a wide range of functional groups such as chloride, nitro, methoxy and fluoride groups, even if in this case, the reaction yield was lower. The reaction tolerates other aromatic groups such as thiophene, furane and naphthalene (see Table 5 and Scheme 59).
Tamura cycloaddition is a common cycloaddition reaction occurring with an enolisable anhydride is used. A version where an anhydride 270 can react with an arylidene-indane-1,3-dione 255 was developed that allowed for the development of many spiro-indanedione derivatives 271 [247]. In this procedure, a base was used to promote the cycloaddition between 270 and 255, and various arylidene-indane-1,3-dione 255 could be converted as spiro-indanedione derivatives 271 (see Table 6 and Scheme 60).
Coumarin is a common functional group present in various natural products [257]. In order to obtain new coumarin-based molecules with biological activities, a [3+2] cycloaddition procedure between arylidene-indane-1,3-diones 255 and a coumarin-based 1,3-dipole precursor 272 was developed [258]. Such a cycloaddition could be catalyzed by an organic catalyst, and the best conditions were found while using 10 mol% of 1-phenyl-3-((1R)-(6-methoxyquinolin-4-yl)((2R,4S,5R)-5-vinylquinuclidin-2-yl)methyl)thiourea QN-T 273 in dichloromethane at 30 °C. Using these conditions, various substrates could be screened as reactants. Thus, 21 coumarin-indanedione derivatives 274 could be synthesized in moderate to high yields, as shown in Table 7 and Scheme 61.
The mechanism of cycloadditions could be determined due to several control reactions. In the mechanism, the coumarin first interacts with the catalyst due to H-bonding. After this initial step, the adduct and the arylidene-indane-1,3-dione can react via a [3+2] cycloaddition concerted step furnishing the spiro compounds (see Scheme 62).
Organocatalysis was also used to achieve asymmetric cycloadditions between 2-arylidene-indane-1,3-diones 255 and Morita–Baylis–Hillman carbonates 275 [259]. In this procedure, a thiourea-phosphine organocatalyst (276) was used to achieve the synthesis of various cyclopentene spiro-indanedione derivatives 277 (see Table 8 and Scheme 63). The mechanism proposed by the authors was the following: the intermediates, once activated by the organocatalyst, can release CO2, and tBuOH can react in a cycloaddition reaction with arylidene-indane-1,3-dione 255. After releasing the organocatalyst 276, the spiro-compounds 277 can be obtained (see Scheme 64).
[3+2] Cycloadditions enabling to prepare spiro-indane-1,3-diones can also be achieved with metal catalysts. Palladium is a common metal capable of catalyzing various cycloaddition reactions. Palladium was notably used as a catalyst in the [3+2] cycloaddition of vinylaziridine 278 and indane-1,3-dione derivatives 255 [260]. Such a procedure used a palladium (0) catalyst and a ligand (279) whose structures were carefully selected among a series of eleven ligands investigated. The optimal conditions were found to be 2.5 mol% of the palladium catalyst Pd2bda3 (tris(dibenzylideneacetone)dipalladium(0)), 6.7 mol% of ligand 279, in THF and at room temperature for two days (see Table 9 and Scheme 65). The scope of applicability of this reaction was examined with a wide range of substrates. Even if the reaction tolerated a wide range of substituents attached to the arylidene-indane-1,3-diones 255, the enantiomeric ratio greatly changed with the substrates, as shown in Table 9 and Scheme 65.
When a cyclohexyl group was used instead of an aromatic group to prepare the indane-1,3-dione adducts, no reaction could take place, demonstrating the importance of starting from arylidene-indane-1,3-diones. The proposed mechanism involves the formation of a pi-allyl-palladium complex, obtained by the oxidative addition of Pd (0) to vinylaziridine 278. Then, after an aza Michael addition, the resulting adducts can form H-bonds with the amide ligand, giving rise to two reversible transition states. Then, an enolate ring closure can occur, and on the basis of steric factors, only one diastereoisomer forms (see Scheme 66).
Vinylethylene carbonate 282 can also react with arylidene-indane-1,3-diones 255 in [3+2] cycloadditions using a palladium catalyst to give tetrahydrofuran-fused spirocyclic 1,3-indandiones 284 and 285 [261]. Such cycloaddition involves the formation of carbon dioxide and proceeds in the optimized conditions, with a phosphoramidite ligand L 283, in chloroform at 0 °C for two days. This reaction could be tested at gram scale. Here again, tolerance of the reaction to the substitution pattern of carbonates 282 and arylidene-indane-1,3-diones 255 was remarkable. A unique product could be obtained in high yield in all cases and with a high enantioselectivity (see Table 10 and Scheme 67).
Spirovinylcyclopropaneindanedione (VCP) 287 is a reactive spiro compound, usually obtained by reaction of indanedione 4 with 1,4-dibromobut-2-ene 286 in basic media (see Scheme 68) [262].
This spirovinylcyclopropaneindanedione 287 was notably used in a cycloaddition with palladium (0) as the catalyst with a sulfonyl-activated imine 288 [263]. Such a procedure can be advantageously used to increase the size of cycles of spiroindanedione compounds. An example of procedure giving access to the five-membered spiroindanedione 289 in 96% yield and with a diastereomeric ratio of 22:1 is presented in Scheme 69.
Spirovinylcyclopropaneindanedione (VCP) 287 is also capable of reacting with enals, as exemplified with cinnamaldehyde 290, enabling to increase by a [3+2] cycloaddition the size of the cycle of the spiro derivatives 292 from three to five carbons (see Scheme 70) [264].
The scope of application of this cycloaddition was rapidly studied, showing that aryl derivatives (293) or naphthyl-based enals (294) were compatible with this reaction, giving the cycloadducts 295 and 296 in 98% and 92% yields, respectively (see Scheme 71).
Spirovinylcyclopropaneindanedione (VCP) 287 can also react with nitroalkenes 297 in cycloaddition reactions, producing spiro-cyclopentane-indane-1,3-diones 298. Such a reaction typically operates according to a two-step procedure [262]. In a first step, the palladium will oxidatively add to VCP 287, opening the cyclopropane ring and giving an anion and a π-allylpalladium complex. The anion formed by ring opening is thus sufficiently nucleophile to add on the alkene 297. The nitro substituent present on nitroalkenes 297 is capable of stabilizing the carbanion, and this anion can give rise to an intramolecular cyclization, regenerating the catalyst (see Scheme 72, 298).
Such a procedure was tested with various nitroalkene derivatives, and the substitution pattern of the aryl substituents (presence of electron-donating or -withdrawing groups) did not impact the reaction process. The different products could be obtained in good yields (>80%). Only when the substituent was o-CF3C6H4, the product was obtained with the lowest reaction yield of the series (75%) and with the worse diastereoselectivity (14:1). Even with other aromatic, heterocycle rings and alkane groups, the reaction yields remained good as well as the diastereoselectivity ratio and the enantiomeric excess (see Table 11 and Scheme 73).
[3+2] Cycloaddition can also be used to construct oxindole-fused spiropyrazolidine compounds 302 starting from spirovinylcyclopropaneindanedione 287 (VCP) catalyzed by palladium (0) [265]. 1-Benzyl-3-diazoindolin-2-one 300 could react with 289 in the presence of Pd2dba3 using a chiral (P,N) ligand 301 in toluene at 0 °C, affording various spiropyrazolidines 302 with reaction yields ranging from 39% to 99% (see Table 12 and Scheme 74).
The mechanism of cyclization was similar to the mechanism previously depicted for the cycloaddition of VCP 289 with nitroalkenes 297. Thus, the palladium catalyst opens the cyclopropane ring, giving a zwitterionic species with an anion and a π-allylpalladium complex. Then, a [3+2] cycloaddition can occur between the anion and the diazo group, generating the cyclopentane ring after cyclization (see Scheme 75).
Palladium was also used in cooperation with an organic base to promote the annulation of vinylcyclopropanes indane-1,3-dione 287 with para-quinone methides 303 via a [3+2] cycloaddition [266]. Association of 1,8-diazabicyclo [5.4.0]undec-7-ene (DBU), a thiourea-based ligand 304, and Pd2(dba)3 could catalyze the intramolecular annulation reaction between para-quinone methides 303 and 2-vinylspiro[cyclopropane-1,2′-indene]-1′,3′-dione 287 giving the product 305 in 61% yield with a diastereoisomeric ratio of 93:7 (see Scheme 76).
Palladium catalysis was also used in synergy with diphenylphosphoric acid 306 [267]. Such a mixture could catalyze the reaction between VCP 287 and various imines 307, producing 308 (see Table 13 and Scheme 77). The different imines used in these reactions were prepared in situ.
Other metals can also be used for the cycloaddition reactions. Thus, copper was notably used to catalyze the [3+2] cycloaddition of 2-arylideneindane-1,3-diones 255 with ketoxime acetates 309 [268]. In a first study, the authors optimized the reaction conditions by screening various copper catalysts such as CuI, CuCN, CuOAc and even solid copper with various additives, at various temperatures and in various solvents. The best conditions to produce 310 were determined as being CuCN as the catalyst while using NaHSO3 as the base and dichloroethane as the solvent at 100 °C. Then, a screening of different arylidene-indane-1,3-diones 255 was realized, as shown in Table 14 and Scheme 78.
The screening showed that the reaction could work with various aryl derivatives, giving the different products with reaction yields higher than 80%. When the methoxy group was attached in para-position of the aromatic ring, the reaction yield was lower than with the other groups. However, when the substituent was another aromatic group such as naphthyl, furyl or thienyl groups, the reaction yield decreased slightly, giving the products in around 50% yield. When the substituent was an alkyl group, such as an isopropyl group, the yield was relatively low (21% yield), demonstrating the importance of an aryl ring on the ketoxime acetates to obtain high reaction yields. The reaction mechanism was investigated by the authors, and the following one was proposed. Thus, after an oxidative addition of the copper catalyst 311 on the oxime 309, a copper enamide 312 formed. Then, this enamide 312 could undergo an intramolecular annulation with 255, releasing acetic acid as a by-product and producing 313. Then, the metallocycle could undergo a reductive elimination, giving the final product 310 and regenerating the catalyst 311 (see Scheme 79).
Cobalt (II) was also used as a metal catalyst to initiate 1,3-dipolar cycloadditions between azomethine ylides 314 and 2-arylidene-indane-1,3-diones 255 [269]. This metal cation can be chelated with two phenylalanine units, forming a planar complex of cobalt (II). After a careful screening of the reaction conditions, cesium carbonate was determined as being the best base to produce 315, and the optimal reaction conditions were determined as being 10 mol% cobalt(L-phenylalanine)2, 10 mol% cesium carbonate in dichloromethane at −5 °C for two days. Using these conditions, a wide range of azomethine ylides 314 and arylidene-indane-1,3-diones 255 could be tested in these conditions (see Table 15 and Scheme 80). The reaction tolerates all substituents examined, except that higher reaction yields were obtained while introducing electron withdrawing groups on the azomethine ylides 314.
A reaction mechanism was also proposed to explain the stereoselectivity of the reactions. In this mechanism, the azomethine ylides 314 chelate to the cobalt metal, leading to an octahedral complex. This complex will hide the Re face of the azomethine ylides, allowing arylidene-indane-1,3-diones 255 to attack the deprotonated azomethine ylides by the Si face exclusively. This stereoselectivity is directly related to the fact that the two aromatic rings (one of L-phenylalanine and the other on the azomethine ylides) will shield the Re face of the azomethine ylides 255 (see Scheme 81).
Ruthenium complexes were also used in combination with visible light to catalyze the [3+2] cycloaddition of 2-arylidene-indane-1,3-diones 255 with methyl 2-(3,4-dihydroisoquinolin-2-yl)acetate 316 [270]. The visible-light catalyzed reaction was tested with various methyl 2-(3,4-dihydroisoquinolin-2-yl)acetate 316, the substituent varying from ester to nitrile groups on the dihydroisoquinoline core and by testing various groups attached to the aromatic ring of 2-arylidene-indane-1,3-diones 255 (see Table 16 and Scheme 82). The nature of the aromatic group on the 2-arylidene-indane-1,3-diones 255 strongly influenced the structure of the final product 317. Thus, when the aromatic ring was substituted with electron-withdrawing groups, the reaction gave spiro[indene-2,1′-pyrrolo [2,1-a]isoquinoline]s 317, whereas when the substituent on the aromatic ring was an electron-releasing group, 3′-arylspiro[indene-2,2′-oxirane]-1,3-diones 318 were obtained.
Such a difference of products (317 or 318) can be explained due to the two plausible mechanisms proposed by the authors and that are depicted in Scheme 83 where an azomethine ylide is produced due to the photoredox catalyst. If the double bond of 2-arylideneindane-1,3-dione 255 is electronically depleted due to the electron withdrawing group and the ester group is not too big in terms of size, the azomethine can react with 2-arylideneindane-1,3-dione 255 via a cycloaddition reaction. Conversely, if the double bond is electronically enriched by electron-donating groups and the ester group is bulky (for instance with tert-butyl groups), then only the epoxidation reaction will occur.
This difference of reactivity can be advantageously used to produce various 3′-arylspiro[indene-2,2′-oxirane]-1,3-diones 318 (see Table 17 and Scheme 84).
[3+2] Cycloaddition is not the only possible cycloaddition capable of produce spiroindanediones. The [2+2+2] cycloaddition is a smart and inventive procedure to construct aromatic rings, or heterocycles. These reactions are often metal-catalyzed [271]. 2,2-Di-2-propynyl-1,3-indandione 320 can be synthesized from indane-1,3-dione 4 and propargyl bromine 319 (see Scheme 85), and the resulting adduct 320 is a common substrate used in numerous cycloaddition reactions.
Using Molybdenum hexacarbonyl, it is also possible to realize cyclotrimerization. By reaction of 2,2-di-2-propynyl-1,3-indanedione 320 with propargyl bromide 319 upon catalysis with Mo(CO)6, a complex mixture of products was obtained, showing the strong reactivity of the alkyne [272]. If the desired product 321 was obtained in 34% yield, the self-dimerized product 323 was also produced as well as another side-product 322 whose origin was determined due to the elucidation of the mechanism (see Scheme 86).
After further investigations, conventional heating was replaced by microwaves heating, giving a better selectivity for the product, and at the same time, use of acetonitrile as the solvent could increase the reaction yield. The scope of application of the reaction was also studied, and various spiroindanediones 325 were obtained by [2+2+2] cycloadditions of 320 and 324 (see Table 18 and Scheme 87).
The team of Ratovelomanana-Vidal and coworkers used another metal, namely ruthenium, to catalyze [2+2+2]cycloadditions, enabling to form pyridines [273,274,275,276,277]. Notably, ruthenium complexes were used to undergo a cycloaddition reaction of 2,2-di-2-propynyl-1,3-indanedione 320 with cyanamide 326 (see Scheme 88). This reaction could give the expected spiro-compound 327 in 97% yield within 5 min [277]. A similar procedure could be used for another cyanamide, i.e., pyrrolidine-1-carbonitrile 328 (see Scheme 89), and the reaction conditions could be greatly improved while using microwave irradiation (see Scheme 90). Using these improved conditions, all compounds (333 and 334) could be obtained in high to almost quantitative yields [274,276].
The procedure described before used Cp*Ru(CH3CN)3PF6 as the catalyst. However, another procedure was also proposed using RuCl3 as a more accessible catalyst and allowing the formation of the spiro compound 335 using a cheaper synthetic approach (see Scheme 91) [275]
Selenocyanates such as 336 and 337 could also be used as reactants with 2,2-di-2-propynyl-1,3-indandione 320 to form the spiro compounds 338 and 339 according to the procedure shown in Scheme 92 [273].
Other metals were also employed as exemplified with cobalt, which can be used to catalyze the [2+2+2] cycloaddition of 2,2-di-2-propynyl-1,3-indanedione 340 with a methyl indole derivative, namely 341 (see Scheme 93) [278]. Rhodium catalysis is also interesting, and in this case, an amide group can act as a directing group to ensure the cycloaddition and allow for the synthesis of polycyclic molecules bearing indanedione moieties such as 344 (see Scheme 94) [279].
The mechanisms involved in these two types of catalyzed cycloaddition reactions are similar. The C-H activation allows for the formation of a carbon-metal bond with the indole compound, and after coordination of the diyne, a migratory insertion of the two alkynes gave a metallocycle, which furnished the product after reductive elimination (see Scheme 95).
Phosphabenzenes are heterocycles containing one phosphorous atom in the cycle, and such heterocycles are only poorly described in the literature. The synthesis of these molecules involve multi-step reactions, cycloadditions and reversible cycloadditions sequences [280], with dangerous sylilated compounds, constituting a major impediment for their developments [281]. To address this issue, iron was notably used to perform the synthesis of phosphabenzene 346 in safe conditions, involving in one reaction, a diyne 320, a phosphaalkyne 345 and iron diiodide in xylene (see Scheme 96) [282].
However, spiro-indanedione can also be obtained by other cycloaddition reactions that are also metal-catalyzed. For example, by starting from a cyclic carbonate 347 and by using the same procedure of that reported by Guo, various derivatives of 349 could be prepared [261]. Cyclic carbonate 347 can react with various 2-arylidene-indane-1,3-dione 255 in palladium-catalyzed cycloadditions and give six-membered spiro compounds 349 (see Table 19 and Scheme 97).
Nickel can catalyze the formation of spiro compounds containing an eight-membered ring cycle [283]. Due to the association of nickel with a NHC ligand 351, such spiro cyclization of 2,2-di-2-butyn-1-yl-1H-indene-1,3(2H)-dione 332 and 1-benzhydrylazetidin-3-one 350 could be achieved in toluene after 8 h at 0 °C, furnishing 352 in 88% yield (see Scheme 98).
A base-catalyzed [4+1] cycloaddition was also described in the literature, enabling the synthesis of spiroindanes 354 bearing a para-phenol moiety (see Scheme 99) [260]. Rhodium was also used in metal-catalyzed intramolecular [4+3] cycloadditions of dienyltriazoles 355 to give the spiro-indanedione compound 356 (see Scheme 100).
o-Quinodimethanes (o-QDM) can be used to generate various compounds containing spiroindanediones moieties. Such molecules were obtained through a [4+4] cycloaddition, producing dibenzocycloooctadiene structures (see Scheme 101) [284]. Starting from indanone containing benzo[c]oxepines such as compounds 357 and 361, and by heating in a polar solvent, it was possible to create in situ o-QDM 358 and 362, producing after [4+4] cycloaddition the different spiroindanediones 359, 360, 363 and 364.
Asymmetric cross [10+2] cycloadditions were also successfully achieved by opposing electron-deficient alkenes such as 365 and 2-arylidene-indane-1,3-dione 249 by phase transfer catalysis [285]. Such reactions can furnish spirofused polycyclic structures such as 367 as shown in Scheme 102.
To conclude, cycloaddition reactions can lead to a large variety of spiro-indanediones, even if [3+2] and [2+2+2] cycloadditions are the two privileged routes to synthesize spiro-indanedione moieties. Other cycloaddition reactions can give promising results as shown in this paragraph.

3.10.2. Synthesis of Spiro-Indane-1,3-Diones by Domino Reaction

Domino reactions are defined as chemical processes where the final product comes from a sequence of reactions. The product of a first reaction become the reactant of another one. In contrast to a one-pot procedure or a multi-component procedure, the reaction conditions cannot be changed after the beginning of the reaction, and no additional compounds can be introduced in the reaction mixture.
One of the first examples of domino reaction described in the literature concerned the synthesis of spiro-compounds 378 by means of a domino Knoevenagel/Diels–Alder /Epimerization sequence [286]. This domino reaction was performed at room temperature in methanol for four days (see Table 20 and Scheme 103).
The mechanism supporting the chemical structure of the final products was suggesting as proceeding according to the following steps (see Scheme 104): First, the amine, i.e., L-proline, catalyzes the classical Knoevenagel condensation between 4 and 368, providing 255. In a second step, L-proline reacts with the Michael acceptor 369, generating the diene 370. Then, the diene 370 can react with the dienophile 255, regenerating the amine. Finally, by epimerization still in the presence of L-proline, the final product 370 can be obtained.
This strategy was notably applied to the construction of chiral spiro[indane-1,3-dione-tetrahydrothiophenes] 373. For this reaction, a tertiary amine-thiourea organocatalyst 372 was used, enabling a sulfa Michael/Michael sequence to occur [287]. By mixing indane-1,3-dione 4, an aldehyde 368 and a thiol, the aforementioned organocatalyst 372 and molecular sieves in toluene at −20 °C, spiro-compounds could be obtained. The scope of application of this reaction was examined with various aromatic groups (see Table 21 and Scheme 105).
X-ray structure of one of the substrates could be obtained and enabled to determine how the organocatalyst was interacting with the substrate and could promote the reaction at the Si face of the substrate (see Figure 2).
More recently, the scope of application has been expanded to the synthesis of spiro tetrahydrothiophene-indan-1,3-diones 376 starting from 1,4-dithiane-2,5-diol 374. The reaction was performed at room temperature in dichloromethane [288]. The scope of application of this reaction was studied, as shown in Table 22 and Scheme 106.
Lewis acids were also used to synthesize spiro compounds by domino reactions, and the combination of a Knoevenagel condensation and a 1,3-dipolar cycloaddition was notably examined. ZnCl2 was used as the Lewis acid-based catalyst and bromonitrile oxide 377 as the main reactant. Starting from indane-1,3-dione 4, aromatic aldehydes 368 and dibromonitrile oxide 377 in basic conditions in THF at 45 °C, spiro-compounds comprising an isoxazole moiety 378 could be obtained (see Table 23 and Scheme 107) [289].
The same procedure was also applied to the synthesis of spiro-compounds 380 by replacing the former aromatic aldehyde 368 with benzoimidazoles 379 (see Table 24 and Scheme 108), thiazole 382 or benzothiazole 383 (see Scheme 109). These two reactions can lead to interesting compounds since the two families of products (380, 384 and 385) were tested as ligands for coupling reactions.
Other domino reactions involved a Michael addition followed by a 1,3-dipolar cycloaddition of 2-arylidene-1,3-indanediones 255 and 5-aryl-1,3,4-oxathiazol-2-ones 388 in toluene (see Scheme 110) [290]. A mechanism involving the formation of a benzonitrile sulfide intermediate was proposed by the authors (see Scheme 111).
Phosphine can also act as an initiator for the synthesis of spiro compounds containing cyclopentanones [291]. In this aim, 2-arylidene-indane-1,3-diones 255 and ynones 391 were mixed in EtOH. The reaction was studied for different ynones 391 and 2-arylidene-indane-1,3-diones 255 (see Table 25 and Scheme 112). Since 2-arylidene-indane-1,3-diones 255 can be synthesized by a Knoevenagel reaction, a tentative of one-pot procedure involving indane-1,3-dione 4, an aromatic aldehyde 393 and an ynone 394 revealed than the one-pot procedure was possible and that the separated synthesis of 2-arylidene-indane-1,3-diones 255 was not necessary (see Scheme 113). Moreover, a mechanism was proposed, where the phosphine binds to the alkyne 394 in first step, creating a α,β-unsaturated ketone 396. Then, the phosphorous will stabilize the oxygen, allowing for the formation of a carbanion in the methyl in α-position of the ketone (397), and the resulting conjugated system 397 can react with 2-arylidene-indane-1,3-diones 255. The carbanion 398 formed during the addition is stabilized by the two ketones of indane-1,3-dione and can undergo an addition at the C3-position of the α,β unsaturated ketone, generating a cyclopentanedione 399. The phosphine is then regenerated, providing the product 395 (see Scheme 114).
Silver can be used as a catalyst in association with diphenylphosphine oxide 402 and a magnesium salt as an additive to synthesize spiro-compounds 403 [292]. The mechanism of reaction involves the formation of a radical that can undergo a cyclisation process (see Scheme 115). Such a reaction was used to synthesize spiro-compounds with indane-1,3-dione (see Scheme 116).
Gold was also used as a catalyst for domino reactions. Starting from indane-1,3-dione 4, a double propargylation of indane-1,3-dione produced 320, which was converted as 404 by the mono functionalization of one of the two propargyl groups by a Sonogashira reaction. Compound 404 was later used for the gold-catalyzed enediyne cyclization (see Scheme 117).
Then, a domino intramolecular cyclization could be carried out with 404 leading to 5′-hydroxy-6′-methyl-1′,3′-dihydro-2,2′-spirobi[indene]-1,3-dione 406 using 405 as the catalyst (see Scheme 118).
In 2020, a cascade Michael addition/cycloaddition reaction between 2-arylidene-indane-1,3-dione 255 and allenoates 407 was reported [293]. Various spiro-derivatives 408 could be obtained by this domino reaction, the overall reaction being catalyzed by a phosphine (see Table 26 and Scheme 119). The mechanism is the combination of a Michael addition of activated allenoates (A) on 2-arylidene- indane-1,3-diones 255. After a proton migration, a second 2-arylidene-indane-1,3-dione is attacked, forming (D). Then, by an intramolecular cycloaddition, (E) is formed, and by regeneration of the phosphine catalyst, the final product (408) can be obtained. This reaction also led to the by-product 409 coming from the [4+2] cycloaddition of the activated allenoates to the arylidene indanedione, furnishing 4aa (see Scheme 120).
A domino process was also reported with 2-arylidene-indane-1,3-diones 255 and N-alkoxyacrylamides 410 in the presence of a base and allowing for the formation of spiro-compounds 411 bearing an indane-1,3-dione moiety and a lactam group [294]. In this process, twenty different derivatives were obtained (see Table 27 and Scheme 121).
A mechanism was tentatively proposed by the authors, where the base activates first N-alkoxyacrylamide 410, and the resulting anion 412 can thus react with 2-arylidene-indane-1,3-dione 255 in an aza-Michael reaction followed by an intramolecular Michael addition, forming a spiro compound 414 containing an enol. By reaction of this enol 414 with another N-alkoxyacrylamide 410, the final product 411 can be obtained (see Scheme 122).
In order to achieve the synthesis of more complex molecules that can exhibit biological properties, a cascade double Michael addition/acetalization proved to be an interesting approach to synthesize complex spiro indane-1,3-dione derivatives 416 and 417. The reaction between a 2-hydroxyarylidene-indane-1,3-dione 260 and hexenedione derivatives 415 could lead to various spiro compounds 416 and 417, when the reaction was catalyzed by DABCO (see Table 28 and Scheme 123) [295]. A mechanism was notably proposed: firstly, the enolate of the dione 415 is formed by deprotonation of the amine (see Scheme 124). Then, the enolate can proceed to the nucleophilic attack onto the Michael acceptor, i.e., 2-arylideneindane-1,3-dione 260 (see Scheme 124). Then, the enone part of the hexenedione 261 can act as a second Michael acceptor. The indane-1,3-dione anion can also attack the Michael acceptor. An acetalization reaction can occur when the deprotonated hydroxyl group of indane-1,3-dione attacks the enolisable ketone, and after protonation, DABCO is regenerated, and the product 416 is formed (see Scheme 124).
A quadruple cascade reaction was also reported to create complex molecules starting from 2-arylidene-indane-1,3-diones 260 and conjugated enals 421. An iminium–enamine–iminium–enamine sequential activation followed by an oxo-Michael addition showed that it was possible to synthesize complex molecules bearing indanedione moieties (see Scheme 125) [296]. The mechanism proposed by the authors was the following one: after condensation of proline on the α,β-unsaturated aldehyde, the resulting iminium 424 could be attacked on its Re face by the alcohol of 2-arylidene-indane-1,3-dione. This oxo-Michael reaction forms an enamine 425 that can react via an intramolecular Michael addition, giving a nucleophilic iminium 426 capable of reacting with another equivalent of α,β-unsaturated aldehyde. The enamine 427 thus obtained can react in an intramolecular condensation, giving the final product 423 and releasing the catalyst (see Scheme 126).

3.10.3. Synthesis of Spiro-Indane-1,3-Diones by MCR

Multicomponent reaction (MCR) is a process where more than two chemical reagents react together to form one product. Such processes are interesting, since they reduce the number of steps to form the final product, facilitate in the purification of the product by avoiding the presence of side-products and, enable the perfect control of the stereoisomeric parameters at the same time.
With regard to the interest of MRC, numerous spiro-indane-1,3-diones 428 were prepared with this procedure. By use of pyridine as the base, spiro-indane-1,3-diones 428 could be easily synthesized starting from indane-1,3-dione 4, an aromatic aldehyde 429 and a pyridinium ylide 431 [297]. The pyridinium ylide 431 could be synthesized from an aromatic ketone 432 and pyridine, furnishing in a first way the pyridinium salt 433. Then, this pyridinium salt 433 can be converted as a pyridinium ylide 431 due to a base. Parallel to this, indane-1,3-dione 4 can react with the aromatic aldehyde 429 in a Knoevenagel condensation, furnishing 2-arylidene-indane-1,3-dione 430. Then, the pyridinium ylide 431 can add on the 2-arylidene-indane-1,3-dione adduct 430, furnishing in turn 428 (see Scheme 127). Due to this procedure, various aldehydes were tested, and moderate to good yields were obtained during the screening of the different aldehydes (see Table 29 and Scheme 128).
Benzothiazole can also be used to achieve the synthesis of spiro compounds starting from indane-1,3-dione 4 [298]. By using indane-1,3-dione, dimethyl but-2-ynedioate 434 and substituted benzothiazoles (435 or 436), spiro compounds 437 and 438 could be respectively obtained with 2-methylbenzo[d]thiazole 435 and 2,5-dimethylbenzo[d]thiazole 436 (see Scheme 129). The mechanism of the reaction was not clearly established, but it seems to proceed via the formation first of a nitrogen-carbon bond between benzothiazole and the alkyne, and in a second step of the formation of a carbon-carbon bond between the alkyne and indane-1,3-dione 4 (see Scheme 130).
Synthesis of spiro compounds can also be obtained through a microwave and catalyst-free procedure. By mixing proline 439, an aromatic aldehyde 368 and 2-arylidene-indane-1,3-diones 255 and by using a microwave-assisted synthesis, spiro-N-fused indanedione compounds 441 and 442 could be successfully prepared [299]. In this procedure, the combination of a condensation, a decarboxylation and a 3+2 cycloaddition, could produce two isomers 441 and 442, as shown in Scheme 131. The scope of application of this reaction was studied with various aromatic substrates (see Table 30 and Scheme 132).
A one-pot five-component reaction was also developed, associated with a CuAAC (Copper catalyzed alkyne azide cycloaddition), a [3+2] cycloaddition and a condensation, furnishing triazole-containing spiro-indane-1,3-diones 435 [300]. In this procedure, five components were used, namely indane-1,3-dione 4, an aromatic azide 444, an aromatic aldehyde 368, 1-(prop-2-yn-1-yl)indoline-2,3-dione 443 and sarcosine 232 that could react with copper sulfate, sodium ascorbate as the catalyst in PEG 400 as the solvent. This reaction proved to be versatile since various aromatic azides 444 or aromatic aldehydes 368 could be used (see Table 31 and Scheme 133). The mechanism proposed by the authors demonstrated that two products can be obtained. However, for unexpected reasons, the reaction proved to be regioselective, and 445 was obtained as the unique product of the reaction.
MCR conditions were also used for the design of spiro-indandiones exhibiting medicinal properties [249]. Indane-1,3-dione 4, aromatic aldehydes 368 and methyl enones 446 were mixed together with an organocatalyst 447, allowing for the formation of spiro compounds 448 containing a halogenated aromatic ring capable of reacting subsequently in a Suzuki cross-coupling reaction, and allowing for the formation of various derivatives 450, which were tested for biological applications (see Table 32 and Scheme 134).
Nanoparticles were also used as catalysts in one-pot three-component reactions [301]. By using proline-functionalized Fe3O4 particles (LPSF) and DABCO as the base, spiro-cyclopropanes 452 could be synthesized starting from indane-1,3-dione 4, aromatic aldehydes 368 and aromatic ketones 451, bearing a bromine in β-position. Various derivatives were prepared due to this strategy (see Table 33 and Scheme 135). Another advantage of this strategy is that iron-based nanoparticles were magnetic, allowing for an easy recovery of the catalyst. The mechanism was supposed to work as depicted below: the nanoparticles due to the amine groups present on the external core of the nanoparticles can interact with substituted benzaldehyde 368 and indane-1,3-dione 4, promoting the Knoevenagel condensation and providing 2-arylidene-indane-1,3-dione 255. Parallel to this, the ketone 451 can interact with nanoparticles (interaction between the amine and the ketone), and then DABCO can easily react with the bromine atom, forming a zwitterion, composed of the quaternary amine and a carbanion in beta position of the ketone. The zwitterion thus prepared can undergo an addition onto 2-arylidene-indane-1,3-dione 255. Then, an intramolecular cycloaddition reaction can give the desired product 452 (see Scheme 136).
If cycloadditions, Domino processes and multicomponent reactions can lead to numerous spiro compounds; several other procedures were also developed to access to spiro compounds.

3.10.4. Synthesis of Spiro-Indane-1,3-Diones by Miscellaneous Way

In this part, all the other synthetic procedures leading to spiro-compounds are discussed. pH-switchable compounds belong to a recent concept that is born with molecular machines [302]. Using 5-(2-bromoethyl)phenanthridin-5-ium bromide 453 and indane-1,3-dione 4, a simple reaction in basic media could lead to the formation of a spiro compound, where the spiro compound can be in a closed/opened position 454a/454b depending on the pH (see Scheme 137). Such pH switchable compounds have the advantage to be easily tunable in terms of absorption and emission maxima [303].
Indane-1,3-dione 4 is an interesting scaffold that was extensively used in organic photovoltaics cells. In this way, it is interesting to synthesize new molecules for optoelectronic applications. Synthesis of spirobisindanedione was successfully obtained through a multi-step synthesis. The same team succeeded in synthesizing bindone 50 as well as spiroindanedione 455-457 (see Scheme 138) [40]. Indane-1,3-dione 4 by reacting with phthaloyl dichloride 458 and sodium hydride, and depending on the reaction conditions used, can give 50 or 456 (see Scheme 139). The reaction of oxidative cleavage of 456 with sodium periodate and ruthenium trichloride can give the spirobisindanedione 455 (see Scheme 140). A similar version of these molecules was also performed with a bromine atom attached to indane-1,3-dione (see Scheme 141).
Iron can be used as a catalyst in an alkene hydrofunctionnalization with 2-arylidene-indane-1,3-diones 255 [304]. In this process, an iron (III) catalyst, the selected 2-arylidene-indane-1,3-dione 255 could react with a terminal bromo-alkene 462. Various substrates were tested (see Table 34 and Scheme 142), and a mechanism was proposed by the authors to support the synthesis of 463 (see Scheme 143).
Copper was also used as a catalyst for the synthesis of spiro compounds. In a [3+2] radical cycloaddition catalyzed by copper, 2-arylidene-indane-1,3-dione 255 could react in the presence of N-acetyl enamides 464 at high temperature [305]. Many N-acetyl enamides 464 and 2-arylidene-indane-1,3-diones 255 were mixed, providing a wide range of spiro structures 465 (see Table 35 and Scheme 144). The plausible mechanism involves an air oxidation of the copper (I) complex to a copper (II) complex, producing in the meantime a radical (see Scheme 145).
Copper was also used in combination with oximes to produce spiro indanedione derivatives 467 (see Table 36 and Scheme 146) [306]. A copper (I) salt was used as the catalyst, and the mechanism was similar to the previous mechanism, in which the copper catalyst is first oxidized, allowing for the formation of a radical. Then, the copper complex forms a metallocycle, where the oxidation state of copper is Cu(III). Being unstable and by reductive elimination, the product 467 can be formed and the copper catalyst regenerated (see Scheme 147).
Seyferth–Gilbert reagent 468 is a reagent classically used to transform carbonyl groups into alkynes. Such a reactant was notably used in association with indanedione derivatives 255 to form spiropyrazolineindane-1,3-diones 458 using CsF as the base [307]. When the 2-arylidene-indane-1,3-diones 255 and the Seyferth–Gilbert reagent 468 were mixed with sodium hydroxide, 3-pyrazolylphthalide 470 could be obtained. Therefore, the possibility to design various structures by modifying the quantities and the nature of the base was demonstrated, as shown in Scheme 148. By using CsF in 0.1 equivalent, the authors could synthesize a wide range of spiropyrazolineindane-1,3-diones 469 (see Table 37 and Scheme 149).
To sum up, spiroindane-1,3-dione are interesting structures that can be obtained using various synthetic procedures. Due to their similitudes with natural compounds, these compounds are promising structures for the design of biologically active compounds, active pharmaceutical ingredients or biomimetic compounds. If cycloaddition remains one of the privileged ways to produce spiroindane-1,3-diones, highly complex structures could also be prepared and imply the development of appropriate synthetic procedures such as domino or multicomponent reactions. The list of the synthetic strategies developed to access to spiro compounds has been exhaustively detailed. Furthermore, other reviews were written on these topics, specially devoted to the synthesis of spiro compounds [308]. The importance of such moieties shows that this field is clearly not unveiled, and future discovery will allow for the design of highly biologically active spiroindane-1,3-diones.

4. Applications of Indane-1,3-Dione-Based Structures

4.1. Photopolymerization

During the past decades, substantial efforts have been devoted to develop photopolymerization processes under visible light and low light intensity [309,310,311,312,313,314,315]. Visible light is a safe spectral range of irradiation for the manipulators, and light is also a traceless reagent, making photopolymerization a green approach for the design of polymeric materials. Several parameters govern the photoinitiating ability of the photosensitizers such as their molar extinction coefficients, their redox properties and notably their easiness to be oxidized or reduced, depending of the additives used in the photoinitiating system [316,317,318,319,320,321,322,323,324,325]. Additionally, efficiency of the polymerization process is also highly dependent on the rate constant of interaction with the additives [326]. With the aim at developing dyes with high molar extinction coefficients, push–pull dyes are the most favorable structures, as the careful selection of the electron-donating and electron-accepting moieties connected at both ends of the π-conjugated spacer can efficiently tune the broadness but also the position of the intramolecular charge transfer band [327,328,329,330]. In this field, indane-1,3-dione and its derivatives have been extensively studied as photoinitiators, and a selection of structures (470 [331], 471–475 [332], 476 [333], 477 [334], 478–497 [335]) is presented in Scheme 150. Among the most interesting findings, photoinitiating ability was demonstrated as being directly related to the solvatochromic properties of the push–pull dyes. Only dyes for which linear correlations using empirical solvent polarity scales (Bakhshiev’s [336], Kawski−Chamma−Viallet’s [337], Lippert−Mataga [338], McRae’s [339], and Suppan’s [340] solvatochromic scales) could be established to initiate a polymerization process. If the direct relation existing between solvatochromic properties and photoinitiating abilities could be demonstrated, no clear explanations could be provided to support this unexpected behavior. Intrigued by these results and considering that little exploration as to the scope of push–pull dyes as photoinitiators has been disclosed, in 2020, Lalevée and coworkers examined this point with a new series of 21 dyes 478–497 in which the optical properties were finely tuned by modification of the electron-accepting core through an extended π-conjugation or by converting 4 and 68 into stronger electron acceptors.
All dyes showed an excellent ability to initiate the free radical polymerization of acrylate (Ebecryl 40) upon irradiation with a light-emitting diode (LED) at 405 nm, which is the wavelength currently under use for 3D printers [341,342,343]. As an appealing feature, beyond simply initiating a polymerization process, dyes 486 could also exhibit excellent photobleaching properties, what is rarely observed and what is actively researched for visible light photoinitiators, considering that these dyes are highly colored compounds often imposing their own colors to the final coating [344,345,346].

4.2. Non-Linear Optical Properties

Push–pull dyes exhibiting a large ground state dipole moment usually have high molecular non-linear optical (NLO) efficiencies [347,348]. Considering this, indanedione-1,3-dione 4, by its electron-withdrawing ability, is an excellent candidate for the design of dyes exhibiting such a property [349,350]. To present large optical nonlinearities, molecules should be organized in a none-centrosymmetric arrangement, and different strategies have been developed to stabilizing the dipole orientation. Thus, introduction of indane-1,3-dione derivatives into polymer composites (498–500) [351] or formation of Langmuir-Blodgett films with amphiphilic derivatives (501–504) [352,353] proved to be effective strategies to address this issue (see Scheme 115). In this last case, introduction of an additional double bond did not significantly modify the structure of the Langmuir–Blodgett films. Highest hyperpolarizability was obtained with 501, indicating the high order of the LB film obtained with this molecule. Parallel to this, an increase in the number of layers enhanced the NLO signal. Covalent linkages to polymers proved as being another strategy to retain the molecular orientation obtained by electrical poling [354]. By heating the polymer 505 at a temperature higher than its Tg and upon application of an intense electric field, an orientation of the push–pull dyes connected to the polymer backbone could be obtained. While maintaining the polymer film at a temperature lower than the polymer Tg, relaxation of the chromophore alignment could be efficiently slowed down, maintaining the molecular orientation of the dyes over time. Improvement of the molecular hyperpolarizability could also be obtained by improving the electron withdrawing of the acceptor, as exemplified with the dye 506 based on 2-methylidene-3-(dicyano-methylidene)-1-indanone [24]. Comparison with a reference compound, i.e., disperse red 1, revealed 243 to exhibit an hyperpolarizability as high as 1558 × 10−30 esu·D, greatly higher than that of the reference compound (814 × 10−30 esu·D). Most of the dyes prepared for NLO applications are synthesized by means of a Knoevenagel reaction, as exemplified by the selection of molecules 498–506 presented in Scheme 151.

4.3. Fluorescent Chemosensors and Chemodosimeters

The detection of metal cations, halide ions, cyanides and even of neutral species has been an active research field so that two types of detectors were developed [355]. The first category of fluorescent sensors are those comprising a binding site giving rise to an irreversible chemical reaction with ions, and these first types of compounds are named chemodosimeters. The second type of optical sensors are those capable of initiate a communication mechanism between the binding site and the ions, but in a reversible way. In this last case, these compounds are thus named fluorescent chemosensors. Indane-1,3-dione 4 has been at the origin of the elaboration of an efficient chemodosimeter for cyanide detection. Cyanides are extremely toxic anions so that a concentration as low as 2.7 µM is tolerated in drinking water [356,357]. In 2009, a chemodosimeter based on calix[4]pyrrole with an indane-1,3-dione unit at the β-pyrrolic position was reported by Lee and coworkers [213]. A dependence of the dye discoloration with the cyanide concentration was clearly evidenced, and a disappearance of the yellow color of 507 upon addition of cyanides could be easily detected with the naked eye. Chemodosimeter proved also to be highly selective since only cyanide ions were detected even when hidden within other ions. The remarkable selectivity and affinity for cyanide anions was explained by the fast equilibrium process involved in the complexation followed by the reaction induced by this extremely nucleophilic anion. After complexation of cyanide anions in the binding site of the calix[4]pyrrole, a nucleophilic addition at the β-position of the indane-1-3-dione group could occur, as shown in Scheme 152.
Based on the same nucleophilic addition of cyanides onto the double bond of push–pull dyes developed with indane-1,3-dione, 508 could be used as a dosimeter enabling to combine two detection modes [358]. Thus, upon addition of cyanide anions on 508, a clear discoloration of the dye could be detected with the naked eye. Parallel to this, a complete quenching of luminescence could be evidenced by photoluminescence measurements. If the detection of cyanides anions in THF was extremely fast, in water, kinetic of addition was considerably reduced, assigned to a solvatation effect of water molecules around the cyanide anions (see Scheme 153). Here again, 508 proved to be selective for the detection of cyanide anion, and this is again related to the small size of cyanides but also to their averred nucleophilic character. A similar strategy was developed with diketopyrrolopyrrole 509 [359]. A dual mode of detection was also observed with this compound. In this last case, a mechanism of exciplex formation was proposed to support the fluorescence quenching observed experimentally (see Scheme 153 and Figure 3).

4.4. Solar Cells

Organic solar cells have received significant attention for the possibility to convert photons to electrons. Among the most widely studied electron-acceptors, [6,6]-phenyl C61 butyric acid methyl ester (PC61BM) is the most popular one [99,360,361,362]. Due to a severe phase segregation upon aging of the device, the low solubility of PC61BM in most of the common organic solvents and non-fullerene acceptors has been identified as promising alternatives to address the segregation issue and the solubility issue. Furthermore, the photon-to-electron conversions remain low, lingering around 3–6% for non-fullerene acceptors [363,364,365,366,367,368,369,370,371,372,373,374,375,376,377,378], justifying the constant efforts to develop new structures. To create strong electron acceptors that could replace PC61BM, indane-1,3-dione 4 was notably combined with dibenzosilole [379]. A photon-to-current conversion efficiency (PCE) of 2.76% could be obtained for the solution-processed devices comprising a poly(3-hexylthiophene) (P3HT)/510 blend as the active layer. By replacing dibenzosilole by a naphthalimide unit, encouraging results were obtained with 511, with a PCE reaching 3.52% in the same conditions [380]. However, other authors also tested the opposite situation, using the dibenzosilole-based compounds as electron donors [381]. End-groups on 512 and 513 were determined as drastically affecting the morphology of the active layer, more than the photophysical properties of the dyes. Even after solvent vapor annealing, the power conversion efficiency of solar cells fabricated with the 512/PC61BM blend remained low, peaking around 0.5% contrarily to 6.6% for devices fabricated with the 513/PC61BM blend. The excellent electron-donating ability of 513 can also be assigned to the presence of the thiophene moiety, reported as improving both the optical and photovoltaic properties of the dyes comprising this group [382]. This trend was confirmed with 514 [11] and 515a [383] or 515b [384], with which a PCE of 2.4%, 6.46% and 8.22% were determined with solar cells of similar structures than that used for 513. In the case of 515b, improvement of the photovoltaic properties was assigned to the more balanced charge transportation in the devices and the use of copper isothiocyanate acting as a hole transport/injection layer. However, counterexamples exist, as exemplified with the asymmetric structure 516 [385]. Photovoltaic properties of 516 in bulk heterojunction solar cells remained poor, and a PCE of 1.7% was obtained for an active layer composed of a 1:1 515/PC61BM blend. This low efficiency was notably assigned to the low hole mobility of 516 and adverse charge transport within the active layer (see Scheme 154). Finally, a similar low power conversion efficiency was obtained with 517 and 518, still based on an asymmetric structure [386]. Here again, the low efficiency (0.24–0.33%) obtained with 517 and 518 was assigned to an unfavorable morphology of the active layer as well as its high roughness.

5. Biological Applications

Indane-1,3-dione 4 has been the focus of intense research efforts in medicinal chemistry since the demonstration in the early 1930s of the bacteriostatic activity of indane-1,3-dione and related derivatives, but also of many 1,3-diketo compounds showing interesting physiological activities [387,388]. This section provides an explicit overview of the importance of indane-1,3-dione 4 as a building block for the design of molecules with potential biological applications.

5.1. Indane-1,3-Dione as Antimicrobial Agent

Any agent that kills or slows down the growth of a micro-organism may be defined as an antimicrobial agent. The increasing demand of antimicrobial drugs resulting from a faster microbial resistance to drugs requires the development of new compounds of innovative structures. Thus, antiseptics have been designed with indane-1,3-dione 4 as soon as 1931 by Walker et al. and their biological activities improved one year later by Robinson et al., who developed 1,3-diketo systems exhibiting remarkable physiological properties In Robinson’s work, the indane-1,3-dione core was formed by a Friedel–Craft reaction, enabling to introduce various n-alkyl groups or hydroxyl groups onto the aromatic core (compounds 521) (see Scheme 155). Notably, the presence of phenol groups was identified as improving the bacteriostatic effect of the resulting compounds [388,389].
All of these compounds were tested in vitro on Gram-positive bacteria: Staphylococcus albus Rosenbach (S. epidermidis), Staphylococcus aureus (S. aureus), (Bacillus megatherium (B. megatherium), Bacillus subtilis (B. subtilis), Bacillus mycoides (B. mycoides), Gram-negative bacteria: Bacterium pyocyaneum (P. aeruginosa), Bacterium prodigiosum (B. Prodigiosum) and Escherichia coli (E coli) but also on acid-fast bacteria such as Mycobacterium phlei (M. phlei) according to the Rideal–Walker method. Even though this method is not used anymore due to its lack of reliability on phenols, it was replaced by the McFarland protocol to furnish more reproducible data and to determine the minimum inhibitory concentration (MIC) [390]. This difference of protocols does not facilitate the comparison with recent research, but the activity of other phenol compounds still allows us to obtain a general picture of the antiseptic properties of this set of compounds. Notably, Robinson’s team noticed that all the molecules they developed were particularly effective against Gram-positive bacteria but had small efficiency against Gram-negative bacteria. Concerning the acid-fast bacteria (M. phlei), some molecules could also kill these organisms but only at high concentrations. Finally, the authors demonstrated an improvement of the bactericidal activity by increasing the number of carbons in the side-chains toward Gram-positive bacteria but also on Bacterium typhosum, responsible for typhoid fever [391]. This effect reaches its maxima for three molecules (see Scheme 156: 522, 523 and 524), with an activity more than ten times higher than that of the methyl or p-n-octylphenol (expressed in equimolecular phenol coefficient of bactericidal power). Even if these molecules have not been used as antibiotics later, modern researchers may inspire for sure from these former studies to design modern drugs due to the rise in antibiotic resistance.
The place of indane-1,3-dione in this field has clearly evolved. A structure related to indane-1,3-dione, namely indan-1-one, also showed interesting antimicrobial properties so that these two structures were often studied concomitantly for the design of antimicrobial compounds [1,2,3]. Considering that the biological activity of indane-1,3-dione 4 can be greatly improved by chemical engineering, indane-1,3-dione 4 has thus been extensively used as a building block in multicomponent reactions (MCRs) to combine the properties of indanones or indane-1,3-dione with that of other structures also displaying physiological properties.
A relevant example of this strategy has been reported in 2011 with a study devoted to the antimicrobial activity of 1,3-disubstituted indeno [1,2-c]pyrazoles [392]. Pyrazoles are an important class of pharmaceutical compounds. Considering that heterocyclic systems have been the focus of intense synthetic efforts during the last decades, indenopyrazoles have thus been extensively screened. In this study, the authors have notably synthetized 16 compounds possessing a 4-substituted thiazole moiety, all prepared with indane-1,3-dione as the starting material, and two different series were developed (see Scheme 157) in order to compare the impact of the fused indenopyrazole on the antimicrobial properties.
Several micro-organisms were tested such as Gram-positive bacteria (S. aureus, B. subtilis), Gram-negative-bacteria (E. Coli) and fungus (Aspergillus niger, Candida albicans). In this series of molecules, three molecules showed noticeable activities against C. albicans, A. niger, S. aureus and E. Coli (528h, 529d, and 529h). More precisely, antimicrobial activity of these molecules proved to be on par with that of the reference compound, i.e., Norfloxacin. Nevertheless, 528h was less effective against B. subtilis, whereas the two others only showed moderate minimum inhibitory concentration (MIC) with this micro-organism. Noticeably, an improved antimicrobial activity was found for all compounds once cyclized as indenopyrazoles, and a higher activity was also determined for all molecules substituted with 4-chlorophenyl groups.
After highlighting the importance of this structure, the same team published a second work in which the synthetic strategy was modified compared to the previous work and seventeen 3-aryl-1-heteroarylindeno [1,2-c]pyrazol-4(1H)-ones 532 could be prepared with this new synthetic method [393]. Indane-1,3-dione 4 was used to perform Knoevenagel condensations with various aldehydes 368, and the formation of the pyrazole cycle could be obtained by reaction of 2-hydrazinylbenzo[d]thiazole/2-hydrazinyl6-substituted benzo[d]thiazoles 530 and the Knoevenagel adduct 255 in stoichiometric amount, according to the reaction presented in Scheme 158. Despite mild reaction conditions, the reaction yields remained low, ranging between 25 and 41%, even after optimization. Nonetheless, almost all of these new synthetized molecules showed an activity against four bacteria, Gram-positive and Gram-negative, but also against fungi.
In 2018, another multicomponent reaction involving indane-1,3-dione 4 was reported by Alsharif et al. for the design of an antimicrobial agent comprising nitrogen-based heterocyclic compounds [394]. The four-component reaction involving indane-1,3-dione 4, 9-ethyl-9H-carbazole-3-carbaldehyde 533, ethyl acetoacetate 534 and ammonium acetate 535 in the presence of a catalytic amount of piperidine could furnish ethyl 4-(9-ethyl-9H-carbazol-3-yl)-2-methyl-5-oxo-4,5-dihydro-1H-indeno [1,2-b]pyridine-3-carboxylate (ECPC, 536), following a procedure previously reported in the literature (see Scheme 159) [395]. This molecule was notably tested for its antibacterial activity, and for this study, S. aureus Streptococcus pyogenes (S. pyogenes) was selected as the Gram-positive bacteria and E. Coli as Salmonella typhimurium (S. typhimurium) as the Gram-negative bacteria. Biological activity of ECPC was also compared with that of tetracycine used as a standard for these tests. ECPC 536 showed an interesting antibacterial activity, since minimum inhibitory concentrations as low as 32 µg/mL were determined with the four bacteria, comparable to that required with the reference tetracycline.
In 2012, another series of heterocyclic molecules was synthetized using indane-1,3-dione 4 as the building block in an ionic liquid-catalyzed three-component condensation involving thiourea 537 and the appropriate aromatic aldehyde 368, and the resulting molecules 538 were tested as antimicrobial agents (see Scheme 160 and Table 38) [396]. In this work, the authors focused their attention on the pyrimidine scaffold, which is present in numerous natural physiologically active substances. The authors improved the biological activity of indane-1,3-dione derivatives by attaching a thione moiety to the pyrimidine group, this group being also known for its antifungal, antibiotic and antibacterial activities. Among the fourteen 4,6-diaryl- and 4,5-fused pyrimidine-2-thiones synthesized in this work, only three of them were prepared with indane-1,3-dione, the other compounds being variously substituted tetrahydrobenzo[h]quinazoline-2-thiones and pyrimidine-2-thiones. Among the three indane-1,3-dione derivatives, only 538c showed a better antibacterial activity toward B. subtilis, S. aureus, Pseudomonas aeruginosa (P. aeruginosa) and E. coli and a better antifungal activity toward A. niger, C. albicans, A. fumigatus than the variously substituted tetrahydrobenzo[h]quinazoline-2-thiones and pyrimidine-2-thiones. In addition, among the eleven other 4,6-diaryl- and 4,5-fused pyrimidine-2-thiones, only two of them showed similar antimicrobial properties than 538c, even if all of them have a less important zone of inhibition than the reference compound, namely Ampicillin trihydrate.
Finally, in 2016, another research group designed a series of spiro[indolo-3,10′-indeno [1,2-b]quinolin]-2,4,11′-triones 540 by means of a three-component condensation of enaminones 539, isatin 231 and indane-1,3-dione 4 using a mixture of ethanol/water as the solvent and cerium ammonium nitrate (CAN) as the catalyst (see Scheme 161) [397]. As the main advantage of this approach, spiro[indolo-3,10′-indeno [1,2-b]quinolin]-2,4,11′-triones 540 could be obtained with short reaction times, in high yields and by using a simple synthetic protocol. From this viewpoint, spiro[indolo-3,10′-indeno [1,2-b]quinolin]-2,4,11′-triones 540 reported in this work could be obtained by a greener approach than that previously reported in the literature [300,398,399,400,401]. Antimicrobial activities of these compounds have been tested on Gram-positive bacteria (S. aureus and B. subtilis) and Gram-negative bacteria (E. coli and P. aeruginosa) as well as their antifungal activity in two yeasts: Candida albicans and Saccharomyces cerevisiae. Among the twenty-two molecules synthetized in this work, all of them showed good antibacterial and antifungal activities, but none of them were effective against P. aeruginosa. Nevertheless, the best results were obtained for 540-IVc. Notably, 540-IVc exhibited the lowest MIC of 16 mg/mL against S. aureus, a MIC of 8 µg/mL against B. subtilis and a MIC of 64 µg/mL against E. coli. In the case of yeast, 540-IVc, 540-IVk and 540-IVn showed the best inhibition ability against C. albicans and S. cerevisiae with inhibition zones of 15 and 16 mm, respectively. Furthermroe, 540-IVc and 540-IVn were also determined as being more efficient against B. subtilis and S. aureus, the inhibition zone reaching 22 and 24 mm, respectively. However, despite these promising results, none of the new compounds could surpass the inhibition diameters obtained with the reference compound, i.e., ciprofloxacin (26, 24 and 25 nm against S. aureus, B. subtilis and E. Coli, respectively).

5.2. Indane-1,3-Diones as Anticancer Agents

Cancer is a generic name that regroups all diseases where the proliferation of abnormal cells is observed. Since only half of the cancer patients do not recover with systemic chemotherapy or obtain only a partial recovery, the development of anticancer agents is thus the focus of intense research efforts [402]. This research was also supported by the current incapacity to cure cancers combined with longer life expectancy but also by the fact that cancer is the second leading cause of death.
The combination of N-heterocycles and indenones has been widely studied during the last 40 years for the design of various compounds exhibiting anti-cancer properties [403,404,405]. Inspired by Onychine, which is a natural biologically active azafluorenone, numerous synthetic azafluorenones have been prepared and identified as exhibiting promising cytotoxic, phosphodiesterase inhibitory, adenosine A2a receptor antagonistic, anti-inflammatory/antiallergic, coronary-dilating and calcium-modulating properties [406,407]. Following the course of these investigations, a series of indeno-heterocycles has been designed by an MCR involving indane-1,3-dione 4 as the building block, and a library of 33 indeno-heterocycles 541 could be obtained (see Scheme 162). All these molecules were evaluated for their potential cytotoxic and apoptosis-inducing properties [407]. As specificities, all molecules reported in this work exhibited the same indenopyridine moiety in order to maintain the planarity of the structure. Planar structures are required in order for these molecules to act as DNA intercalators and topoisomerase inhibitors. Even if a similar synthetic protocol was used for the synthesis of the 33 molecules, different structures were nevertheless obtained. These compounds were notably tested for their cytotoxicity on Jurkat cells, a model for human T-cell leukemia, with Annexin-V/propidium iodide assay [408]. However, cytotoxicity of molecules 541 remained low since viability of the Jurkat cells was around 85–98% relative to the control experiments, and apoptosis induction was even lower, ranging between 1–6%. Aniline, pyrazole and triazole-based indeno-heterocycles 541 proved to be totally inactive for apoptosis induction. Only the pyrimidinedione-containing indeno-heterocycle showed a somewhat enhanced cytotoxicity, the cell viability being reduced to 80%. All attempts to mix the pyrimidinedione-containing indeno-heterocycle with the other compounds did not improve the cytotoxicity, evidencing that the pyrimidinedione moiety was the key-element to obtain an acceptable cytotoxicity. To obtain a deeper insight into the biological activity of this pyrimidinedione-containing indeno-heterocycle, this compound was compared to etoposide, which is an anti-cancer clinical drug agent known for its remarkable cytotoxic effect through topoisomerase II-dependent DNA cleavage. The two molecules could manifest good apoptosis-induction activities against Jurkat cells. Furthermore, the indeno-pyrimidine compound could demonstrate better cell-killing and apoptosis-induction activities than etoposide at low concentration (cytotoxic IC50 = 3 μM). Nevertheless, clinical tests still need to be performed with this pyrimidinedione-containing indeno-heterocycle, and the authors also highlight the important drawback of poor water solubility of this compound, which may affect its future use as anti-cancer agent.
Three years later, in the continuation of the pioneering work initiated by Manpadi et al., another series of 35 pentacycle-based indeno-heterocycles 544 and 545 was prepared in order to investigate the cell-killing and the apoptosis mechanism toward a panel of human cancer cell lines [409]. All these structures were inspired by camptothecin, which is extensively used in traditional Chinese medicine. As the main motivation of this study, the authors suggested the drug action of these molecules to operate via an intracellular oxidation of the dihydropyridine cycle of 544, leading to the formation of the pyridine moiety 545 in situ (see Scheme 163). To demonstrate this, all compounds were synthetized with a hydrogenated (“h”) and an aromatized (“a”) version of the indenopyridine core using the same three-component synthesis of that used by Manpadi. When no spontaneous oxidation was observed, an additional oxidation step with chloranil in DMF was required to obtain the aromatized structure. These molecules were further tested for their antiproliferative properties on a panel of human cancer cell lines (HeLa, Jurkat, MCF-7, A-549, Lovo, U373, SKMEL, PC3, MG-MID), representing many species of cancer such as T-cell leukemia, cervical, breast, or lung cancers. The authors demonstrated that the modification of the R group on the aromatic ring did not significantly modify the biological activity of the molecules, except that the molecule bearing R=H showed no cytotoxic activity. The biological tests also revealed that the dihydropyridines 544 do not have cytotoxic activities contrarily to the pyridine analogues 545, supporting an intracellular oxidation process of the dihydropyridines 544 to the corresponding pyridines 545.
Sulfonamides are well known to act as inhibitors of numerous human α-carbonic anhydrases [410]. Carbonic anhydrases are essential in humans, as these anhydrases are capable of catalyzing the reversible hydration of CO2 and allow for the respiration and the transportation of carbon dioxide within the human body. In 2012, the idea of Ghorab et al. was to combine the remarkable properties of sulfonamides to that of indenopyridines, which also possess anti-cancer properties [411]. To obtain these structures (549), a two-step synthesis involving indane-1,3-dione 4, 4-aminobenzenesulfonamide 546 and an aromatic aldehyde 368 as the starting materials was performed, the two steps being both realized in EtOH as the solvent. Anti-cancer activity of the resulting 18 compounds 549a-549q was tested in vivo against breast cancer cell-line (MCF 7) and their biological activities compared to that of doxorubicin, a reference drug. Among all compounds, 549d exhibited an IC50 of 4.34 μM, lower than that of the reference doxorubicin (5.40 μM), evidencing the pertinence of the approach. Furthermore, 549d, which is specifically substituted with a hydroxyphenyl group at the 4-position and a cyano group at the 3-position exhibited an increased cytotoxic activity compared to the other compounds, and this specific substitution is certainly at the origin of the higher potency of this molecule (see Scheme 164).
Related to indenopyridines, indenopyrazoles are another family of structures showing important inhibiting tyrosine kinase activities, but also anti-cancer or anti-proliferative properties [412]. Nevertheless, for this purpose, indane-1,3-dione 4 was not directly used for their synthesis but a derivative containing a close scaffold: 2-(4-methoxybenzoyl)indan-1,3-dione 552. Its synthesis is similar to that of compounds 4 and 67 described in Section 1 except that no heating was required for the last step to obtain the triketone molecule and, in a second step, the intramolecular cyclization of 552 with hydrazine hydrate could furnish the targeted indenopyrazoles 553, as shown in Scheme 165 [413]. In 2002, another library of indenopyrazoles mainly substituted at the 5-position of aniline and at the 3-position with various groups was examined as Cyclin-dependent kinase (CDK) inhibitors [414]. Cclin-dependent kinases play a key role in the cell cycle regulatory machinery and the replication process. Considering that the different indenopyrazoles can regulate the development of proliferative tumors such as cancers, these molecules have notably been tested against various cancer cell lines (HCT116, NCI-H460, PC-3, MiaPaCa-2, HT-29, HT-1080 and B16-F0). After selecting 553-k for its good activity and selectivity against kinase targets (CDK2/E: IC50 = 13 nM, CDK1/B: IC50 = 44 nM), this molecule exhibited good cytotoxicity, particularly for HCT116 cells, which died not only by CDK inhibition but also by activation of the apoptotic machinery and by inhibition of Rb phosphorylation. This strong activity combined with a remarkable selectivity of cells offers very promising therapeutic applications, explaining why several patents have been established for structures close to that of 553-k [412,415,416].

5.3. Indane-1,3-Dione as Building Block for Bioimaging Agents

Fluorescent imaging techniques play an important role in medicinal chemistry by enabling to localize tumors at the cellular level and in this aim; extensive works are performed to develop fluorescent dyes capable of target specific cells or ions [417]. With the aim at gaining a better understanding of all these complex diseases, multiple recent techniques have been developed to visualize tumors, and numerous fluorescent agents have been proposed with indane-1,3-dione, which remains a cheap and versatile building-block in chemistry.
One of the most popular imaging techniques is named aggregation-induced emission (AIE). As specificity, dyes exhibiting AIE properties are often weakly emissive in solution but highly emissive in the solid state due to a restriction of the intramolecular rotations that constitute non-radiative deactivation processes. In addition to enhancing the fluorescence intensity of dyes, AIE is also extensively used for bioimaging due to its biocompatibility, its high-fidelity imaging, high brightness and long-living excited state favorable for an efficient in-situ imaging [418]. In 2020, an interesting study demonstrated that beyond imaging, AIE dyes could also be used for the generation of reactive oxygen species (ROS), enabling to combine bioimaging and photodynamic therapy (PDT) [419]. However, the key point of this strategy relies in a perfect targeting of the infected lysosomes by AIE dyes. To achieve this goal, three molecules based on triphenylamines were designed, bearing morpholines as pendant groups. In these structures, triphenylamine was used as the electron donor for the design of the different push–pull dyes and indane-1,3-dione, malononitrile or the formyl group were used as the electron acceptors (see MPAA 554, MPAN 555 and MPAT 556 in Scheme 166).
Among the three dyes, MPAT 556 exhibited the most red-shifted absorption and emission of the series so that this dye was the focus of extensive works. The three dyes showed a remarkable fluorescence stability since upon irradiation with a laser for 20 min., a decrease of less than 10% of the initial fluorescence intensity was determined, far from the loss of 50% determined with the benchmark LysoTracker Green DND-26 [420]. Examination of the fluorescence properties of MPAA 554, MPAN 555 and MPAT 556 in water revealed these dyes to exhibit high stoke shifts ranging between 187, 160 and 188 nm for MPAA 554, MPAN 555 and MPAT 556, respectively. For MPAT 556 displaying the most red-shifted emission, fluorescence maxima at 673 nm in water and 675 nm in the solid state could be determined. Conversely, for the other dyes, emissions located at 530 and 614 nm could be respectively determined in water for MPAA 554 and MPAN 555. For these reasons, the indane-1,3-dione adduct MPAT 556 was selected for the biological tests in A549, HeLa and HepG2 cells. No significant decrease for the cells’ viability was detected after 24 h of incubation, even at high concentration (100 μM), demonstrating that MPAT 556 can be used for long-term monitoring of cells without inducing cells apoptosis. In vivo experiments in zebrafish also furnished nice lysosome images with a good dot distribution during imaging of vertebrates. Finally, the ability of MPAT 556 to promote ROS production was evidenced while using H2DCF-DA as the probe. In PBS buffer solutions, a 36-fold enhancement of the photoluminescence of H2DCF-CA was evidenced, demonstrating the production of ROS by MPAT 556 upon green light irradiation. In HeLa cells, cell viability showed a remarkable decline, the cell viability reducing to 19% at 40 μM MPAT and after 10 min. of irradiation with a green light (see Figure 4).
In 2017, Gao et al. used AIE dyes for a completely different application, namely the lipid droplet-specific (LD) imaging and the dynamic movement tracking [421]. Lipid droplets (LDs) are involved in many metabolic processes. However, conventional dyes used to stain LDs suffer from numerous drawbacks such that new fluorophores are actively researched. Several families of molecules have been tested over the years, but the different fluorophores exhibited numerous limitations such as aggregation-caused quenching (ACQ), low stoke shift (40 nm), high background noise or even difficulty of preparation [422,423,424,425,426,427,428,429,430]. Consequently, the molecule selected in this study was a classical push–pull dye made of triphenylamine, as the electron-donating group was connected to indane-1,3-dione acting as the electron-withdrawing group and prepared by means of a Knoevenagel reaction. In addition of its AIE properties, 557 also exhibits what is named a twisted intramolecular charge transfer (TICT), resulting from restricted intramolecular rotations in the solid state. Based on both AIE and TICT properties, 557 was thus identified as a promising candidate for LD-specific imaging and movement tracking due to its high two-photon absorption cross-sections in the near infrared (NIR) range (see Scheme 167). For this reason, 557 was notably used as fluorescent dye for near infrared (NIR) two-photon excited fluorescence (TPEF).
To evidence the AIE effect in 557, UV–visible absorption and emission spectra were recorded in a THF/water mixture. A raise in the water fraction in THF from 0 to 70% led to a decrease in the fluorescence but also to a significant redshift not only of the emission spectrum (from 594 to 609 nm) but also of the absorption spectrum (from 478 to 489 nm), which can be attributed to TICT effects. A further increase in the water fraction in THF up to 99% induced a further redshift of the emission maximum up to 612 nm, resulting from the formation of nanoaggregates of average diameter of 119.6 nm, measured by light scattering analyses. Finally, 557 was tested on HCC827 and A549 cells. As shown in Figure 5, 557 could efficiently stain the cells with high signal-to-noise ratios. From the different tests, several advantages were determined for 557 such as:
A fast permeability;
An excellent selectivity for LD;
A high photostability;
A low cytotoxicity.
Finally, photostability of 557 was examined upon irradiation at 514 nm with a laser (7% power) for 10 min. As shown in Figure 6, the fluorescence intensity loss was less than 20%, evidencing the good photostability of this dye.
As we saw previously, the selectivity of dyes to target the right cells is a primordial need, especially when photodynamic therapy (PDT) is envisioned in complement to cell visualization. In 2019, Mondal et al. performed an MCR using indane-1,3-dione 4 as a component to elaborate a series of spiro-heterocyclic compounds 561–564 (see Scheme 168) that was subsequently tested as chromophores for the selective detection of Zn2+ ion and cell imaging [431]. Regarding specificity, this synthesis was realized by means of a microwave-assisted multi-component reaction (MWAMCRs); the reaction was carried out in EtOH, and the products could be obtained in pure form without chromatography, making this approach highly biocompatible.
Nine of the compounds were selected for further tests aiming at evaluating the possibility to coordinate Zn2+ ion, determining the selectivity toward other ions other than Zn2+ or the fluorescence emission at physiological pH. Among the nine dyes, one of them (i.e., 564f) was even selected for further in vivo tests on human hepatocellular liver carcinoma cells (HepG2 cells). This chemosensor could be successfully used as an intracellular Zn2+ imaging agent due to its remarkable cell permeability properties, where this molecule exhibited a fluorescence comportment only when coordinated with Zn2+ even in vivo. Cytotoxicity tests also revealed 564f to be weakly toxic since the cell viability was higher than 90% at 10 µM. Thus, 564f can be effectively used in vivo as chemosensors for the detection of Zn2+ cations. A mechanism supporting the enhancement of fluorescence upon coordination to Zn2+ cations was also proposed by the authors and is depicted in Scheme 169.

5.4. Indane-1,3-Dione in Neurology Drugs

The brain is the most complex organ in humans, and its working principle is still not fully understood yet, which explains the lack of medicines specific to this organ, despite all the research performed in this field. Moreover, the eldering of the populating on Earth will cause a significant increase in the neurodegenerative diseases that usually occur after 65 years, so that the demand on treatments for these illnesses will increase in the coming decades [432,433]. Parkinson’s disease (PD) is one of the most important neurodegenerative diseases and causes an important loss of dopamine in the brain, altering motor function cognition and the mood of the patient. Common treatments to PD are antagonists of two brain receptors, namely A2a and A1. However, it is still unknown yet if either one of the two brains receptors plays a more important role than the other in the disease [434].
In this context, Shook’s team published a series of arylindenopyrimidines 565 that were developed as potential dual A2A/A1 antagonists and that could be potentially used for the treatment of AD [435]. These arylindenopyrimidines 565 were synthetized through a two-step synthesis using indane-1,3-dione 4 as the starting material (see chemical structures in Scheme 170). In vitro and in vivo tests performed with eight molecules revealed unequal results depending on the substitution pattern. Notably, the different tests revealed the necessity to let the NH2 group be unsubstituted, whereas the substitution of the pendant aromatic group had almost no incidence on the in vivo and in vitro activity. Furthermore, molecules substituted with morpholine or pyrrolidine groups showed the highest in vivo activity in mouse catalepsy.
To improve the biological activity, two different syntheses were developed to obtain substitutions on the heterocycle next to the indane group. The 12-step synthesis developed by the authors could give access to twelve compounds 577 substituted at the 9-position, and the seven-step synthesis developed in parallel could furnish 12 other compounds 584 substituted at the 8-position (see Scheme 171).
Comparison between the two series of analogs 577 and 584 revealed an improved biological activity for all molecules substituted at the 8-positions compared to those bearing a substitution at the 9-positions. Notably, a difference in the mouse catalepsy as high as 0.2 and 10.7 mg/kg for the ED50 for the regioisomers 577-(1) and 584-(1) bearing a piperidine substituent could be determined. Furthermore, several adducts, particularly those containing cyclic amines and heterocyclic functions attached with an alkyl secondary amine showed ED50 in the micro-scale for the mouse catalepsy. These adducts also showed decent in vitro activity so that these structures can be cited as promising candidates for AD treatments (see Scheme 172). Nevertheless, the lack of the commercially available substituted indane-1,3-diones allowing access to these structures is relatively complex and costly.
Furthermore, these structures remain promising, as exemplified with JNJ-40255293 (585) for which preclinical tests have been carried out (see Scheme 173) [436]. These additional tests revealed a better selectivity of the JNJ-40255293 antagonist toward the A2a receptors than the A1 ones. Unfortunately, preclinical tests were rapidly abandoned due to genotoxicity, neuronal necrosis and edema observed in several animals [437].
In 2002, more promising results were obtained with another series of indeno-fused structures, namely arylindenopyridines 588 and 593 that were patented for AD treatments (see Scheme 174) [438]. To support the biological efficiency, the authors suggested a binding of arylindenopyridines onto the A2a receptors. Anti-inflammatory activity was also identified, with phosphodiesterase (PDE)-inhibiting activity. The starting materials for these syntheses were various 1,3-indanedione adducts (586 or 590), such as 5-nitroindane-1,3-dione (see Scheme 138). The MCR used to access to these structures involved in the first step a Knoevenagel reaction between the indane-1,3-dione derivative 586 or 590 with the appropriate aldehyde 368, followed by a condensation of guanidine carbonate 588 to the push–pull compound 587 to form the indenopyrimidine 589. In the case of 491, an aromatization with DDQ was required, providing 593. The selectivity toward the A2a receptor and A1 were tested in vitro, and it turned out that some of these molecules could exhibit Ki lower than 50 nM, particularly with the A2a binding. Despite these interesting preliminary results, no in vivo tests were performed with these molecules.
In 2011, a series of tricyclic 3,4-dihydropyrimidine-2-thiones 595 was proposed as a potential A1 receptor and TRPA1 antagonists. The TRPA1 channel is implicated in numerous inflammatory or neuropathic pains [439]. These molecules were notably prepared via a Biginelli reaction with indane-1,3-dione 4 due to which the effects of the substitution of the phenyl groups along with the replacement of urea by thiourea in the cyclization reaction were studied. According to the in vitro tests performed on human and rats cells, experimental results revealed that all compounds prepared with thiourea were more biologically active than those prepared with urea. Influence of the stereochemistry is also clearly evidenced, as shown for compounds 595-(3) and 595-(4). Thus, for 595-(2), a half maximal inhibitory concentration hTRPA1 IC50 of 0.075µM was determined for the S,R-stereoisomer, whereas for the S,S-stereoisomer 595-(3), an hTRPA1 IC50 higher than 10 µM was determined (see Scheme 139). The meta-substitution on the phenyl ring could induce a significant increase in the activity (see Scheme 175).
Nevertheless, only in vitro tests were performed with these molecules, and the authors highlighted some drawbacks such as a low solubility, a poor metabolic stability and the potential toxicity of thiourea used to prepare these molecules [440].
In 2015, Ahmed et al. described a solvent-free three-component synthesis based on 1,3-indanedione as the starting material and developed a series of 14 potential anticonvulsants compounds with this strategy [243]. Seizure is a transient occurrence of signs and/or symptoms due to an abnormal excessive or synchronous neuronal activity in the brain [441]. This condition is often affiliated with epilepsy, which concerns about 0.5–1% of the population. Moreover, serious improvement needs to be performed concerning their efficiency but also to address the important issue of side effects [442]. The authors principally acted on one factor, namely the substitution on the pendant phenyl group. After synthesis, each molecule was tested via an anticonvulsant evaluation with the maximal electroshock (MES) method and compared with the results obtained with a standard drug, i.e., phenytoin, along with their toxicity with the Rotorod test. Most of the compounds designed in this study seemed to have positive effects on seizures ranging from moderate to good activity, particularly 222d, 222e, and 222j (see Scheme 176), which proved to be significantly active at a dose of 40 mg/kg. Moreover, the Rotarod tests showed no toxicity for all the compounds tested. However, none of them were as efficient as the reference compound phenytoin.
Among the neurodegenerative diseases, Alzheimer’s disease (AD) is the most common one, but this disease is also badly treated. During the last 25 years, only five medicines were approved for AD treatment with two different strategies of action. Thus, cholinesterase inhibitors (ChEIs) such as tacrine (1), donepezil (2), rivastigmine (3) and galantamine (4) were proposed in parallel to N-methyl-D-aspartate (NMDA) antagonists such as memantine (5) [443]. Nonetheless, none of them can repair the damage or delay the disease progression, which could constitute a major improvement in this field. The aggregation of a small peptide named amyloid β (Aβ) is associated with AD, and the inhibition of both cholinesterase sites (acetylcholinesterase, AChE, and butyrylcholinesterase, BChE) may prevent this aggregation [444].
In 2018, Tanoli et al. developed a series of tricyclic fused ring systems exhibiting activity against both acetylcholinesterase (AChE) and butyrylcholinesterase (BChE). Precisely, this study was supported by previous works reported in the literature, demonstrating that other tricyclic structures based on pyrimidine or quinazoline could act as AChE and BChE inhibitors [445]. In order to access the pyrimidine-fused rings, indane-1,3-dione 4 was used in a Biginelli reaction along with thiourea 597 and benzaldehyde 368 (Scheme 177).
The dihydropyrimidine (DHPM) scaffold is a potential inhibitor of cholinesterases so that in vitro tests were carried out. Interesting results were obtained with these molecules since the introduction of the indanone-pyrimidine fused ring could lead to an improved inhibition activity compared to tacrine and donepezil for both electric eel AChE (IC50 = 0.09 µM) and equine serum BChE (IC50 = 1.04 µM) assays. These results are close to that obtained with the commercially available Donepezil with IC50 eeAChE = 0.05 µM and IC50 eqBChE = 5.4 µM. Docking studies on this molecule have also exposed the well accommodation of this molecule into the bottom of the gorge and the importance of the hydrogen bonding with its environment, such as Ser200, or even the phenyl ring at the 4-position, which forms π-π stacking interactions with trp84 (see Figure 7).
In 2010, the good activity of 2-[(2-(4-chlorophenyl)hydrazinyl) methylene]-1H-indene-1,3(2H)-dione 605f as a novel inhibitor capable of prevent amyloid aggregation (IC50 = 23µM) [446] motivated the group of Campana to obtain a deeper insight into the biological activity of this molecule and in this aim a series of indane-2-arylhydrazinylmethylene-1,3-diones 605a–f and indol-2-aryldiazenylmethylene-3-ones 608a–m (see Scheme 178) [447].
In this work, the authors clearly highlight a significant increase in the biological activity with the molecules prepared with indane-1,3-dione compared to those prepared with 1,3-cyclopentanedione. Nonetheless, even if all the compounds exhibited decent to good AB aggregation processes, none of them could overcome the reference quercetin (IC50 = 0.8 µM).

5.5. Indane-1,3-Dione as Anticoagulant Drugs

One of the oldest utilizations of indanedione derivatives in medicine concerns the anticoagulant properties. Since the 1940s, several authors reported the insecticidal properties of acylated indane-1,3-diones (see Scheme 179), especially for houseflies [154]. Later on, this toxicity has been attributed to an anticoagulant property, especially marked for acylated indane-1,3-diones (see Scheme 179) [448]. However, this activity can also be used as treatment for vitamin K antagonists (VKAs), as observed for phenindione derivatives (see Scheme 179) [449]. However, pheindione is not so much used anymore and was replaced by its homologue, fluindione, or other coumarins VKA due to side effects such as hypersensitivity identified for pheindione [450]. However, despite a proven efficiency as vitamin K epoxide reductase, fluindione 609b is used almost only in France because of a lack of study data on old people [451].
The synthesis of these indane-1,3-dione derivatives is actually not performed anymore starting from the unsubstituted indane-1,3-dione, but a recent study uses a similar structure with indane-1,3-dione as the starting material [452]. Nonetheless, this molecule is only an intermediary of reaction for the synthesis of 2,3-disubstituted indoles.
Concerning the synthesis of acylindanediones 613, the standard synthetic strategy is quite similar to that used for indane-1,3-dione 4 since it consists in a Claisen condensation using an alcoolate as the base (see Scheme 180) [453]. However, an alternative was proposed in 2018 by Larsen et al. involving indane-1,3-dione 4 as the starting material with the possibility to afford 25 different C2 acylated 1,3-indandiones 616 (see Scheme 181) [454]. This strategy, which is quite efficient, makes use of an easy set up and producing the molecules in high yields using 1-ethyl-(3-(3-dimethylamino)propyl)-carbodiimide (EDCI) and 4-dimethylaminopyridine (DMAP) as the coupling agents. Furthermore, all tests made with this new strategy have only been performed at a small scale compared to the initial synthesis.
These type of indanes are more commonly used as rodenticides than as for pharmaceutical purposes. All commercially available rodenticides are either 4-hydroxycoumarins or indane-1,3-diones and even if coumarins are more widely used, a resistance has recently been evidenced toward derivatives such as Warfarin. At present, less resistances have been identified toward indane derivatives. Chlorophacinone, which belongs to the first generation of rodenticides, is still commonly used in many countries. Even if the second generation is more effective against rodents, their impact on the environment is quite significant with multiple poisoning detected in cats, dogs or even otters [455,456].

6. Conclusions-Perspectives

Indane-1,3-dione is a versatile molecule that has found applications in numerous research fields. Directly related to the broadness of applications, numerous synthetic routes have been examined first to prepare indane-1,3-dione, the corresponding derivatives bearing various substituents. If indane-1,3-dione is an excellent electron acceptor for the design of various push–pull dyes, the most popular use of indane-1,3-dione is undoubtedly in multicomponent or domino reactions to prepare polycyclic structures for biological activities. Considering the remarkable biological activity of indane-1,3-dione derivatives, future work will certainly be more focused on these biological applications than on the optoelectronics applications. Numerous electron acceptors have already been prepared with this scaffold, limiting the possibility of new developments.

Author Contributions

Conceptualization, C.P., D.B. and F.D.; methodology, C.P., D.B. and F.D.; software, C.P., D.B. and F.D.; validation, C.P., D.B. and F.D.; formal analysis, C.P., D.B. and F.D.; investigation, C.P., D.B. and F.D.; resources, F.D.; data curation, F.D.; writing—original draft preparation, C.P., D.B. and F.D.; writing—review and editing, C.P., D.B. and F.D.; visualization, C.P., D.B. and F.D.; supervision, F.D.; project administration, F.D.; funding acquisition, F.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by The Agence Nationale de la Recherche (ANR), grant number ANR-17-CE08-0010 DUALITY project for the PhD grant of Corentin Pigot. This research was funded by the Agence Innovation Defense (AID) through the PhD grant of Damien Brunel. The research was funded by Aix Marseille University and The Centre National de la Recherche Scientifique (CNRS), under the frame of permanent fundings.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No data available.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Patil, S.A.; Patil, R.; Patil, S.A. Recent Developments in Biological Activities of Indanones. Eur. J. Med. Chem. 2017, 138, 182–198. [Google Scholar] [CrossRef] [PubMed]
  2. Turek, M.; Szczęsna, D.; Koprowski, M.; Bałczewski, P. Synthesis of 1-Indanones with a Broad Range of Biological Activity. Beilstein J. Org. Chem. 2017, 13, 451–494. [Google Scholar] [CrossRef] [PubMed]
  3. Menezes, J.C.J.M.D.S. Arylidene Indanone Scaffold: Medicinal Chemistry and Structure–Activity Relationship View. RSC Adv. 2017, 7, 9357–9372. [Google Scholar] [CrossRef]
  4. Costanzo, P.; Cariati, L.; Desiderio, D.; Sgammato, R.; Lamberti, A.; Arcone, R.; Salerno, R.; Nardi, M.; Masullo, M.; Oliverio, M. Design, Synthesis, and Evaluation of Donepezil-Like Compounds as AChE and BACE-1 Inhibitors. ACS Med. Chem. Lett. 2016, 7, 470–475. [Google Scholar] [CrossRef]
  5. Vacca, J.P.; Dorsey, B.D.; Schleif, W.A.; Levin, R.B.; McDaniel, S.L.; Darke, P.L.; Zugay, J.; Quintero, J.C.; Blahy, O.M.; Roth, E. L-735,524: An Orally Bioavailable Human Immunodeficiency Virus Type 1 Protease Inhibitor. Proc. Natl. Acad. Sci. USA 1994, 91, 4096. [Google Scholar] [CrossRef]
  6. Luo, H.-F.; Zhang, L.-P.; Hu, C.-Q. Five Novel Oligostilbenes from the Roots of Caragana Sinica. Tetrahedron 2001, 57, 4849–4854. [Google Scholar] [CrossRef]
  7. Nagle, D.G.; Zhou, Y.-D.; Park, P.U.; Paul, V.J.; Rajbhandari, I.; Duncan, C.J.G.; Pasco, D.S. A New Indanone from the Marine Cyanobacterium Lyngbya Majuscula That Inhibits Hypoxia-Induced Activation of the VEGF Promoter in Hep3B Cells. J. Nat. Prod. 2000, 63, 1431–1433. [Google Scholar] [CrossRef]
  8. Kim, J.; Kim, I. Design and Synthesis of a Hybrid Framework of Indanone and Chromane: Total Synthesis of a Homoisoflavanoid, Brazilane. Org. Biomol. Chem. 2018, 16, 89–100. [Google Scholar] [CrossRef]
  9. Yang, Y.; Philips, D.; Pan, S. A Concise Synthesis of Paucifloral F and Related Indanone Analogues via Palladium-Catalyzed α-Arylation. J. Org. Chem. 2011, 76, 1902–1905. [Google Scholar] [CrossRef]
  10. Buckingham, J. Dictionary of Natural Products, Supplement 1, 1st ed.; Taylor & Francis: New York, NY, USA, 1994; ISBN 978-1-00-305992-9. [Google Scholar]
  11. Winzenberg, K.N.; Kemppinen, P.; Scholes, F.H.; Collis, G.E.; Shu, Y.; Birendra Singh, T.; Bilic, A.; Forsyth, C.M.; Watkins, S.E. Indan-1,3-Dione Electron-Acceptor Small Molecules for Solution-Processable Solar Cells: A Structure–Property Correlation. Chem. Commun. 2013, 49, 6307–6309. [Google Scholar] [CrossRef]
  12. Pigot, C.; Noirbent, G.; Bui, T.-T.; Péralta, S.; Gigmes, D.; Nechab, M.; Dumur, F. Push-Pull Chromophores Based on the Naphthalene Scaffold: Potential Candidates for Optoelectronic Applications. Materials 2019, 12, 1342. [Google Scholar] [CrossRef] [PubMed]
  13. Patel, A.; Giles, D.; Basavarajaswamy, G.; Sreedhar, C.; Patel, A. Synthesis, Pharmacological Evaluation and Molecular Docking Studies of Indanone Derivatives. Med. Chem. Res. 2012, 21, 4403–4411. [Google Scholar] [CrossRef]
  14. Karthik, R.; Jasmin Sajni, R.; Sasikumar, S.; Kalyan, S.B.; Christina, A.J.M.; Jagan, A.; Sundara Saravanan, K. Evaluation of Anti-Tubercular Activity of Some Synthesised Benz Spiro-Oxirane Derivatives of Indane-1,3-Dione. Pharmacologyonline 2008, 2, 176–191. [Google Scholar]
  15. Liu, J.; Hu, K.-F.; Qu, J.-P.; Kang, Y.-B. Organopromoted Selectivity-Switchable Synthesis of Polyketones. Org. Lett. 2017, 19, 5593–5596. [Google Scholar] [CrossRef]
  16. Mardani, H.R.; Golchoubian, H. Selective and Efficient C-H Oxidation of Alkanes with Hydrogen Peroxide Catalyzed by a Manganese(III) Schiff Base Complex. J. Mol. Catal. Chem. 2006, 259, 197–200. [Google Scholar] [CrossRef]
  17. Delaval, N.; Bouquillon, S.; Hénin, F.; Muzart, J. Use of Benzotrifluoride as Solvent for Chromium-Catalysed Oxidations with Sodium Percarbonate. J. Chem. Res. Synop. 1999, 4, 286–287. [Google Scholar] [CrossRef]
  18. He, G.; Wu, C.; Zhou, J.; Yang, Q.; Zhang, C.; Zhou, Y.; Zhang, H.; Liu, H. A Method for Synthesis of 3-Hydroxy-1-Indanones via Cu-Catalyzed Intramolecular Annulation Reactions. J. Org. Chem. 2018, 83, 13356–13362. [Google Scholar] [CrossRef]
  19. Kumar, K.; Kumar, P.; Joshi, P.; Rawat, D.S. IBX-TfOH Mediated Oxidation of Alcohols to Aldehydes and Ketones under Mild Reaction Conditions. Tetrahedron Lett. 2020, 61, 151749. [Google Scholar] [CrossRef]
  20. Marvi, O.; Giahi, M. Montmorillonite KSF Clay as Novel and Recyclable Heterogeneous Catalyst for the Microwave Mediated Synthesis of Indan-1,3-Diones. Bull. Korean Chem. Soc. 2009, 30, 2918–2920. [Google Scholar] [CrossRef]
  21. Guo, S.; Zhang, N.; Tang, X.; Mao, Z.; Zhang, X.; Yan, M.; Xuan, Y. Cyclopropanation of Active Methylene Compounds with β-Alkoxycarbonyl Vinylsulfonium Salts. Chin. Chem. Lett. 2019, 30, 406–408. [Google Scholar] [CrossRef]
  22. Berezina, G.R.; Vorob’ev, Y.G.; Mukhanova, O.N. Macroheterocyclic Compounds Based on Nitro- and Chloro-1,3-Indandiones. Russ. J. Gen. Chem. 2005, 75, 1594–1598. [Google Scholar] [CrossRef]
  23. Li, S.; Ye, L.; Zhao, W.; Zhang, S.; Mukherjee, S.; Ade, H.; Hou, J. Energy-Level Modulation of Small-Molecule Electron Acceptors to Achieve over 12% Efficiency in Polymer Solar Cells. Adv. Mater. 2016, 28, 9423–9429. [Google Scholar] [CrossRef] [PubMed]
  24. Cui, Y.; Ren, H.; Yu, J.; Wang, Z.; Qian, G. An Indanone-Based Alkoxysilane Dye with Second Order Nonlinear Optical Properties. Dyes Pigments 2009, 81, 53–57. [Google Scholar] [CrossRef]
  25. Abdelrazek, F.M.; Metz, P.; Jaeger, A. Some Reactions with Indane-1,3-Dione: A Facile Synthesis of Pentacycline Heterocyclic Analogues. J. Heterocycl. Chem. 2019, 56, 1939–1945. [Google Scholar] [CrossRef]
  26. Tirelli, N.; Amabile, S.; Cellai, C.; Pucci, A.; Regoli, L.; Ruggeri, G.; Ciardelli, F. New Terthiophene Derivatives for Ultrahigh Molecular Weight Polyethylene-Based Absorption Polarizers. Macromolecules 2001, 34, 2129–2137. [Google Scholar] [CrossRef]
  27. Shang, Y.; Wen, Y.; Li, S.; Du, S.; He, X.; Cai, L.; Li, Y.; Yang, L.; Gao, H.; Song, Y. A Triphenylamine-Containing Donor−Acceptor Molecule for Stable, Reversible, Ultrahigh Density Data Storage. J. Am. Chem. Soc. 2007, 129, 11674–11675. [Google Scholar] [CrossRef]
  28. Planells, M.; Robertson, N. Naphthyl Derivatives Functionalised with Electron Acceptor Units–Synthesis, Electronic Characterisation and DFT Calculations. Eur. J. Org. Chem. 2012, 2012, 4947–4953. [Google Scholar] [CrossRef]
  29. Yang, X.; Fox, T.; Berke, H. Synthetic and Mechanistic Studies of Metal-Free Transfer Hydrogenations Applying Polarized Olefins as Hydrogen Acceptors and Amine Borane Adducts as Hydrogen Donors. Org. Biomol. Chem. 2012, 10, 852–860. [Google Scholar] [CrossRef]
  30. Morales, A.R.; Frazer, A.; Woodward, A.W.; Ahn-White, H.-Y.; Fonari, A.; Tongwa, P.; Timofeeva, T.; Belfield, K.D. Design, Synthesis, and Structural and Spectroscopic Studies of Push–Pull Two-Photon Absorbing Chromophores with Acceptor Groups of Varying Strength. J. Org. Chem. 2013, 78, 1014–1025. [Google Scholar] [CrossRef]
  31. Matsui, M.; Tanaka, N.; Funabiki, K.; Haishima, Y.; Manseki, K.; Jin, J.; Inoue, Y.; Higashijima, S.; Kubota, Y. Application of Indoline Dyes Attached with Strongly Electron-Withdrawing Carboxylated Indan-1,3-Dione Analogues Linked with a Hexylthiophene Ring to Dye-Sensitized Solar Cells. Tetrahedron 2018, 74, 3498–3506. [Google Scholar] [CrossRef]
  32. Dai, S.; Zhao, F.; Zhang, Q.; Lau, T.-K.; Li, T.; Liu, K.; Ling, Q.; Wang, C.; Lu, X.; You, W.; et al. Fused Nonacyclic Electron Acceptors for Efficient Polymer Solar Cells. J. Am. Chem. Soc. 2017, 139, 1336–1343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Zhu, J.; Li, S.; Liu, X.; Yao, H.; Wang, F.; Zhang, S.; Sun, M.; Hou, J. Subtle Side-Chain Tuning on Terminal Groups of Small Molecule Electron Acceptors for Efficient Fullerene-Free Polymer Solar Cells. J. Mater. Chem. A 2017, 5, 15175–15182. [Google Scholar] [CrossRef]
  34. Lai, H.; Chen, H.; Zhou, J.; Qu, J.; Wang, M.; Xie, W.; Xie, Z.; He, F. 3D Interpenetrating Network for High-Performance Nonfullerene Acceptors via Asymmetric Chlorine Substitution. J. Phys. Chem. Lett. 2019, 10, 4737–4743. [Google Scholar] [CrossRef] [PubMed]
  35. Aldrich, T.J.; Matta, M.; Zhu, W.; Swick, S.M.; Stern, C.L.; Schatz, G.C.; Facchetti, A.; Melkonyan, F.S.; Marks, T.J. Fluorination Effects on Indacenodithienothiophene Acceptor Packing and Electronic Structure, End-Group Redistribution, and Solar Cell Photovoltaic Response. J. Am. Chem. Soc. 2019, 141, 3274–3287. [Google Scholar] [CrossRef]
  36. Swick, S.M.; Zhu, W.; Matta, M.; Aldrich, T.J.; Harbuzaru, A.; Lopez Navarrete, J.T.; Ponce Ortiz, R.; Kohlstedt, K.L.; Schatz, G.C.; Facchetti, A.; et al. Closely Packed, Low Reorganization Energy π-Extended Postfullerene Acceptors for Efficient Polymer Solar Cells. Proc. Natl. Acad. Sci. USA 2018, 115, E8341. [Google Scholar] [CrossRef]
  37. Cui, Y.; Yang, C.; Yao, H.; Zhu, J.; Wang, Y.; Jia, G.; Gao, F.; Hou, J. Efficient Semitransparent Organic Solar Cells with Tunable Color Enabled by an Ultralow-Bandgap Nonfullerene Acceptor. Adv. Mater. 2017, 29, 1703080. [Google Scholar] [CrossRef]
  38. Yao, H.; Cui, Y.; Yu, R.; Gao, B.; Zhang, H.; Hou, J. Design, Synthesis, and Photovoltaic Characterization of a Small Molecular Acceptor with an Ultra-Narrow Band Gap. Angew. Chem. Int. Ed. 2017, 56, 3045–3049. [Google Scholar] [CrossRef]
  39. Bürckstümmer, H.; Tulyakova, E.V.; Deppisch, M.; Lenze, M.R.; Kronenberg, N.M.; Gsänger, M.; Stolte, M.; Meerholz, K.; Würthner, F. Efficient Solution-Processed Bulk Heterojunction Solar Cells by Antiparallel Supramolecular Arrangement of Dipolar Donor–Acceptor Dyes. Angew. Chem. Int. Ed. 2011, 50, 11628–11632. [Google Scholar] [CrossRef]
  40. Wilbuer, J.; Schnakenburg, G.; Esser, B. Syntheses, Structures and Optoelectronic Properties of Spiroconjugated Cyclic Ketones. Eur. J. Org. Chem. 2016, 2016, 2404–2412. [Google Scholar] [CrossRef]
  41. Sloop, J.C.; Boyle, P.D.; Fountain, A.W.; Gomez, C.; Jackson, J.L.; Pearman, W.F.; Schmidt, R.D.; Weyand, J. Novel Fluorinated Indanone, Tetralone and Naphthone Derivatives: Synthesis and Unique Structural Features. Appl. Sci. 2012, 2, 61–99. [Google Scholar] [CrossRef]
  42. Nikulin, V.I.; Lugovskaya, N.Y.; Sveshnikov, N.N.; Pisarenko, L.M.; D’yachkovskaya, R.F. Synthesis and Antitumor Activity of 2-Arylindane-1,3-Dione Derivatives with an Alkylating Fragment. Pharm. Chem. J. 1994, 28, 146–149. [Google Scholar] [CrossRef]
  43. Abdel-Latif, F.F.; Mashaly, M.M.; El-Gawish, E.H. ChemInform Abstract: Synthesis of Heterocycles Through Reactions of Nucleophiles with Acrylonitriles. Part 15. Synthesis of Some New Functionalized Benzo(b) Pyrans and Indeno(1,2-b)Pyrans of Potential Biological Activity. ChemInform 1995, 26, 42. [Google Scholar] [CrossRef]
  44. Becker, H.-D. 2-[p-(Phenylsulfonyl) Phenyl]-1, 3-Indanedione and Its Tautomer. U.S. Patent 3,356,732, 25 December 1967. [Google Scholar]
  45. Liu, Y.; Saldivar, A.; Bess, J.; Solomon, L.; Chen, C.-M.; Tripathi, R.; Barrett, L.; Richardson, P.L.; Molla, A.; Kohlbrenner, W.; et al. Investigating the Origin of the Slow-Binding Inhibition of HCV NS3 Serine Protease by a Novel Substrate Based Inhibitor. Biochemistry 2003, 42, 8862–8869. [Google Scholar] [CrossRef] [PubMed]
  46. Siddiqui, N.; Ahuja, P.; Ahsan, W.; Pandeya, S.N.; Shamsher Alam, M. Thiadiazoles: Progress Report on Biological Activities. J. Chem. Pharm. Res. 2009, 1, 19–30. [Google Scholar]
  47. Salgın-Gökşen, U.; Gökhan-Kelekçi, N.; Göktaş, Ö.; Köysal, Y.; Kılıç, E.; Işık, Ş.; Aktay, G.; Özalp, M. 1-Acylthiosemicarbazides, 1,2,4-Triazole-5(4H)-Thiones, 1,3,4-Thiadiazoles and Hydrazones Containing 5-Methyl-2-Benzoxazolinones: Synthesis, Analgesic-Anti-Inflammatory and Antimicrobial Activities. Bioorg. Med. Chem. 2007, 15, 5738–5751. [Google Scholar] [CrossRef] [PubMed]
  48. Schenone, S.; Brullo, C.; Bruno, O.; Bondavalli, F.; Ranise, A.; Filippelli, W.; Rinaldi, B.; Capuano, A.; Falcone, G. New 1,3,4-Thiadiazole Derivatives Endowed with Analgesic and Anti-Inflammatory Activities. Bioorg. Med. Chem. 2006, 14, 1698–1705. [Google Scholar] [CrossRef]
  49. Wei, M.-X.; Feng, L.; Li, X.-Q.; Zhou, X.-Z.; Shao, Z.-H. Synthesis of New Chiral 2,5-Disubstituted 1,3,4-Thiadiazoles Possessing γ-Butenolide Moiety and Preliminary Evaluation of in Vitro Anticancer Activity. Eur. J. Med. Chem. 2009, 44, 3340–3344. [Google Scholar] [CrossRef]
  50. Mavrova, A.T.; Wesselinova, D.; Tsenov, Y.A.; Denkova, P. Synthesis, Cytotoxicity and Effects of Some 1,2,4-Triazole and 1,3,4-Thiadiazole Derivatives on Immunocompetent Cells. Eur. J. Med. Chem. 2009, 44, 63–69. [Google Scholar] [CrossRef]
  51. Siddiqui, N.; Ali, S.; Khan, S.; Drabu, S.; Rana, A.; Alam, M. Synthesis of 3-Arylamino-4-Aryl-5-(N-Arylthiocarbonylimino)-4, 5-Dihydro-1, 2, 4-Thiadiazoles as Anticonvulsant Agents. Indian J. Heterocycl. Chem. 2004, 14, 159–160. [Google Scholar]
  52. Kuş, C.; Ayhan-Kılcıgil, G.; Özbey, S.; Kaynak, F.B.; Kaya, M.; Çoban, T.; Can-Eke, B. Synthesis and Antioxidant Properties of Novel N-Methyl-1,3,4-Thiadiazol-2-Amine and 4-Methyl-2H-1,2,4-Triazole-3(4H)-Thione Derivatives of Benzimidazole Class. Bioorg. Med. Chem. 2008, 16, 4294–4303. [Google Scholar] [CrossRef]
  53. Cressier, D.; Prouillac, C.; Hernandez, P.; Amourette, C.; Diserbo, M.; Lion, C.; Rima, G. Synthesis, Antioxidant Properties and Radioprotective Effects of New Benzothiazoles and Thiadiazoles. Bioorg. Med. Chem. 2009, 17, 5275–5284. [Google Scholar] [CrossRef] [PubMed]
  54. Sharma, R.; Misra, G.P.; Sainy, J.; Chaturvedi, S.C. Synthesis and Biological Evaluation of 2-Amino-5-Sulfanyl-1,3,4-Thiadiazole Derivatives as Antidepressant, Anxiolytics and Anticonvulsant Agents. Med. Chem. Res. 2011, 20, 245–253. [Google Scholar] [CrossRef]
  55. Muhammad, Z.A.; Masaret, G.S.; Amin, M.M.; Abdallah, M.A.; Farghaly, T.A. Anti-Inflammatory, Analgesic and Anti-Ulcerogenic Activities of Novel Bis-Thiadiazoles, Bis-Thiazoles and Bis-Formazanes. Med. Chem. 2017, 13, 226–238. [Google Scholar] [CrossRef]
  56. Pigot, C.; Noirbent, G.; Peralta, S.; Duval, S.; Nechab, M.; Gigmes, D.; Dumur, F. Unprecedented Nucleophilic Attack of Piperidine on the Electron Acceptor during the Synthesis of Push-Pull Dyes by a Knoevenagel Reaction. Helv. Chim. Acta 2019, 102, e1900229. [Google Scholar] [CrossRef]
  57. Pigot, C.; Noirbent, G.; Peralta, S.; Duval, S.; Bui, T.-T.; Aubert, P.-H.; Nechab, M.; Gigmes, D.; Dumur, F. New Push-Pull Dyes Based on 2-(3-Oxo-2,3-Dihydro-1H-Cyclopenta[b]Naphthalen-1-Ylidene)Malononitrile: An Amine-Directed Synthesis. Dyes Pigments 2020, 175, 108182. [Google Scholar] [CrossRef]
  58. Feng, H.; Qiu, N.; Wang, X.; Wang, Y.; Kan, B.; Wan, X.; Zhang, M.; Xia, A.; Li, C.; Liu, F.; et al. An A-D-A Type Small-Molecule Electron Acceptor with End-Extended Conjugation for High Performance Organic Solar Cells. Chem. Mater. 2017, 29, 7908–7917. [Google Scholar] [CrossRef]
  59. Li, R.; Liu, G.; Xiao, M.; Yang, X.; Liu, X.; Wang, Z.; Ying, L.; Huang, F.; Cao, Y. Non-Fullerene Acceptors Based on Fused-Ring Oligomers for Efficient Polymer Solar Cells via Complementary Light-Absorption. J. Mater. Chem. A 2017, 5, 23926–23936. [Google Scholar] [CrossRef]
  60. Sanguinet, L.; Williams, J.C.; Yang, Z.; Twieg, R.J.; Mao, G.; Singer, K.D.; Wiggers, G.; Petschek, R.G. Synthesis and Characterization of New Truxenones for Nonlinear Optical Applications. Chem. Mater. 2006, 18, 4259–4269. [Google Scholar] [CrossRef]
  61. Knoevenagel, E. Ueber Eine Darstellungsweise Des Benzylidenacetessigesters. Berichte Dtsch. Chem. Ges. 1896, 29, 172–174. [Google Scholar] [CrossRef]
  62. Landmesser, T.; Linden, A.; Hansen, H.-J. A Novel Route to 1-Substituted 3-(Dialkylamino)-9-Oxo-9H-Indeno[2,1-c]Pyridine-4-Carbonitriles. Helv. Chim. Acta 2008, 91, 265–284. [Google Scholar] [CrossRef]
  63. Helmy, S.; Oh, S.; Leibfarth, F.A.; Hawker, C.J.; Read de Alaniz, J. Design and Synthesis of Donor–Acceptor Stenhouse Adducts: A Visible Light Photoswitch Derived from Furfural. J. Org. Chem. 2014, 79, 11316–11329. [Google Scholar] [CrossRef] [PubMed]
  64. Gao, M.; Su, H.; Lin, Y.; Ling, X.; Li, S.; Qin, A.; Tang, B.Z. Photoactivatable Aggregation-Induced Emission Probes for Lipid Droplets-Specific Live Cell Imaging. Chem. Sci. 2017, 8, 1763–1768. [Google Scholar] [CrossRef] [PubMed]
  65. Capobianco, A.; Esposito, A.; Caruso, T.; Borbone, F.; Carella, A.; Centore, R.; Peluso, A. Tuning Wavefunction Mixing in Push–Pull Molecules: From Neutral to Zwitterionic Compounds. Eur. J. Org. Chem. 2012, 2012, 2980–2989. [Google Scholar] [CrossRef]
  66. Bello, K.A.; Cheng, L.; Griffiths, J. Near-Infrared Absorbing Methine Dyes Based on Dicyanovinyl Derivatives of Indane-1,3-Dione. J. Chem. Soc. Perkin Trans. 1987, 2, 815–818. [Google Scholar] [CrossRef]
  67. Singh, A.; Lim, C.-K.; Lee, Y.-D.; Maeng, J.; Lee, S.; Koh, J.; Kim, S. Tuning Solid-State Fluorescence to the Near-Infrared: A Combinatorial Approach to Discovering Molecular Nanoprobes for Biomedical Imaging. ACS Appl. Mater. Interfaces 2013, 5, 8881–8888. [Google Scholar] [CrossRef] [PubMed]
  68. Zilinskaite, V.; Gudeika, D.; Grazulevicius, J.V.; Hladka, I. Synthesis and Cationic Polymerization of Oxyranyl-Functionalized Indandiones. Polym. Bull. 2016, 1, 229–239. [Google Scholar] [CrossRef]
  69. Gudeika, D.; Zilinskaite, V.; Grazulevicius, J.V.; Lytvyn, R.; Rutkis, M.; Tokmakov, A. 4-(Diethylamino)Salicylaldehyde-Based Twin Compounds as NLO-Active Materials. Dyes Pigments 2016, 134, 244–250. [Google Scholar] [CrossRef]
  70. Ziarani, G.M.; Lashgari, N.; Azimian, F.; Kruger, H.G.; Gholamzadeh, P. Ninhydrin in Synthesis of Heterocyclic Compounds. ARKIVOC 2015, 2015, 1–139. [Google Scholar] [CrossRef]
  71. Marminon, C.; Nacereddine, A.; Bouaziz, Z.; Nebois, P.; Jose, J.; Le Borgne, M. Microwave-Assisted Oxidation of Indan-1-Ones into Ninhydrins. Tetrahedron Lett. 2015, 56, 1840–1842. [Google Scholar] [CrossRef]
  72. Jong, J.A.W.; Moret, M.-E.; Verhaar, M.C.; Hennink, W.E.; Gerritsen, K.G.F.; van Nostrum, C.F. Effect of Substituents on the Reactivity of Ninhydrin with Urea. ChemistrySelect 2018, 3, 1224–1229. [Google Scholar] [CrossRef]
  73. Taherpour, A.; Kamal, S. 1,2,3-Trione Compounds Synthesis by Oxidation 1,3-Diketones. Asian J. Chem. 2007, 19, 4107–4109. [Google Scholar]
  74. Prakash, O.; Sharma, P.K.; Saini, N. ChemInform Abstract: A New Synthesis of Ninhydrin and Its Ketals Using Iodobenzene Diacetate. ChemInform 1995, 26, 39. [Google Scholar] [CrossRef]
  75. Tatsugi, J.; Izawa, Y. A Facile One-Pot Synthesis of Vicinal Di- and Tri-Ketones from α-Methylene Ketones by N-Bromosuccinimide-Dimethyl Sulphoxide Oxidation. J. Chem. Res. Synop. Print 1988, 11, 356–357. [Google Scholar]
  76. Vickery, B.; Kaberia, F. Reactions of Sodium Hypochlorite with Some Compounds Having Reactive Methylene Groups. Experientia 1979, 35, 299. [Google Scholar] [CrossRef]
  77. Wasserman, H.H.; Pickett, J.E. The Fluoride Ion Effect in the Reactions of Singlet Oxygen with Enols. Tetrahedron 1985, 41, 2155–2162. [Google Scholar] [CrossRef]
  78. Gao, S.; Bethel, T.K.; Kakeshpour, T.; Hubbell, G.E.; Jackson, J.E.; Tepe, J.J. Substrate Controlled Regioselective Bromination of Acylated Pyrroles Using Tetrabutylammonium Tribromide (TBABr3). J. Org. Chem. 2018, 83, 9250–9255. [Google Scholar] [CrossRef]
  79. Wengryniuk, S.E.; Weickgenannt, A.; Reiher, C.; Strotman, N.A.; Chen, K.; Eastgate, M.D.; Baran, P.S. Regioselective Bromination of Fused Heterocyclic N-Oxides. Org. Lett. 2013, 15, 792–795. [Google Scholar] [CrossRef]
  80. Pathak, T.P.; Miller, S.J. Site-Selective Bromination of Vancomycin. J. Am. Chem. Soc. 2012, 134, 6120–6123. [Google Scholar] [CrossRef]
  81. Shi, X.; Dai, L. Mild Halogenation of Stabilized Ester Enolates by Cupric Halides. J. Org. Chem. 1993, 58, 4596–4598. [Google Scholar] [CrossRef]
  82. Yang, D.; Yan, Y.-L.; Lui, B. Mild α-Halogenation Reactions of 1,3-Dicarbonyl Compounds Catalyzed by Lewis Acids. J. Org. Chem. 2002, 67, 7429–7431. [Google Scholar] [CrossRef]
  83. Tanemura, K.; Suzuki, T.; Nishida, Y.; Satsumabayashi, K.; Horaguchi, T. A Mild and Efficient Procedure for α-Bromination of Ketones Using N-Bromosuccinimide Catalysed by Ammonium Acetate. Chem. Commun. 2004, 4, 470–471. [Google Scholar] [CrossRef]
  84. Arbuj, S.S.; Waghmode, S.B.; Ramaswamy, A.V. Photochemical α-Bromination of Ketones Using N-Bromosuccinimide: A Simple, Mild and Efficient Method. Tetrahedron Lett. 2007, 48, 1411–1415. [Google Scholar] [CrossRef]
  85. Wang, G.-W.; Gao, J. Solvent-Free Bromination Reactions with Sodium Bromide and Oxone Promoted by Mechanical Milling. Green Chem. 2012, 14, 1125–1131. [Google Scholar] [CrossRef]
  86. Meshram, H.M.; Reddy, P.N.; Vishnu, P.; Sadashiv, K.; Yadav, J.S. A Green Approach for Efficient α-Halogenation of β-Dicarbonyl Compounds and Cyclic Ketones Using N-Halosuccinimides in Ionic Liquids. Tetrahedron Lett. 2006, 47, 991–995. [Google Scholar] [CrossRef]
  87. Zou, L.-H.; Li, Y.-C.; Li, P.-G.; Zhou, J.; Wu, Z. Solvent-Controlled α-Monobromination, α,α-Dibromination or Imidation of 1,3-Diketones with N-Bromosuccinimide. Eur. J. Org. Chem. 2018, 2018, 5639–5643. [Google Scholar] [CrossRef]
  88. Saikia, I.; Borah, A.J.; Phukan, P. Use of Bromine and Bromo-Organic Compounds in Organic Synthesis. Chem. Rev. 2016, 116, 6837–7042. [Google Scholar] [CrossRef]
  89. Luo, L.; Meng, L.; Sun, Q.; Ge, Z.; Li, R. Novel Synthesis of Thiazolo/Thienoazepine-5,8-Diones from Dihalo Cyclic 1,3-Diketones and Mercaptonitrile Salts. RSC Adv. 2014, 4, 6845–6849. [Google Scholar] [CrossRef]
  90. Mishra, A.K.; Nagarajaiah, H.; Moorthy, J.N. Trihaloisocyanuric Acids as Atom-Economic Reagents for Halogenation of Aromatics and Carbonyl Compounds in the Solid State by Ball Milling. Eur. J. Org. Chem. 2015, 2015, 2733–2738. [Google Scholar] [CrossRef]
  91. Macharla, A.K.; Chozhiyath Nappunni, R.; Marri, M.R.; Peraka, S.; Nama, N. Oxidative Bromination of Ketones Using Ammonium Bromide and Oxone®. Tetrahedron Lett. 2012, 53, 191–195. [Google Scholar] [CrossRef]
  92. Kim, J.-J.; Kweon, D.-H.; Cho, S.-D.; Kim, H.-K.; Lee, S.-G.; Yoon, Y.-J. Conversion of Nucleophilic Halides to Electrophilic Halides: Efficient and Selective Halogenation of Azinones, Amides, and Carbonyl Compounds Using Metal Halide/Lead Tetraacetate. Synlett 2006, 2006, 194–200. [Google Scholar] [CrossRef]
  93. Košmrlj, J.; Kočevar, M.; Polanc, S. A New Convenient Bromination with KBrO3/KBr/Dowex®. Synth. Commun. 1996, 26, 3583–3592. [Google Scholar] [CrossRef]
  94. Spitulnik, M.J. Synthesis of 1-Aryl-4-Halo-2-Pyrazolin-5-Ones by Ascorbic Acid Reduction of 1-Aryl-4,4-Dihalo-2-Pyrazolin-5-Ones. Synthesis 2002, 1985, 299–300. [Google Scholar] [CrossRef]
  95. Stavber, G.; Zupan, M.; Jereb, M.; Stavber, S. Selective and Effective Fluorination of Organic Compounds in Water Using Selectfluor F-TEDA-BF4. Org. Lett. 2004, 6, 4973–4976. [Google Scholar] [CrossRef]
  96. Matarlo, J.S.; Evans, C.E.; Sharma, I.; Lavaud, L.J.; Ngo, S.C.; Shek, R.; Rajashankar, K.R.; French, J.B.; Tan, D.S.; Tonge, P.J. Mechanism of MenE Inhibition by Acyl-Adenylate Analogues and Discovery of Novel Antibacterial Agents. Biochemistry 2015, 54, 6514–6524. [Google Scholar] [CrossRef] [PubMed]
  97. Sloop, J.C.; Churley, M.; Guzman, A.; Moseley, S.; Stalker, S.; Weyand, J.; Yi, J. Synthesis and Reactivity of Fluorinated Cyclic Ketones: Initial Findings. Am. J. Org. Chem. 2014, 4, 1–10. [Google Scholar]
  98. Nyffeler, P.T.; Durón, S.G.; Burkart, M.D.; Vincent, S.P.; Wong, C.-H. Selectfluor: Mechanistic Insight and Applications. Angew. Chem. Int. Ed. 2005, 44, 192–212. [Google Scholar] [CrossRef]
  99. Yi, H.; Zhang, G.; Wang, H.; Huang, Z.; Wang, J.; Singh, A.K.; Lei, A. Recent Advances in Radical C-H Activation/Radical Cross-Coupling. Chem. Rev. 2017, 117, 9016–9085. [Google Scholar] [CrossRef]
  100. Shaw, M.H.; Twilton, J.; MacMillan, D.W.C. Photoredox Catalysis in Organic Chemistry. J. Org. Chem. 2016, 81, 6898–6926. [Google Scholar] [CrossRef]
  101. Wang, C.-S.; Dixneuf, P.H.; Soulé, J.-F. Photoredox Catalysis for Building C-C Bonds from C(Sp2)–H Bonds. Chem. Rev. 2018, 118, 7532–7585. [Google Scholar] [CrossRef]
  102. Li, A.Y.; Moores, A. Carbonyl Reduction and Biomass: A Case Study of Sustainable Catalysis. ACS Sustain. Chem. Eng. 2019, 7, 10182–10197. [Google Scholar] [CrossRef]
  103. Zhang, K.; Ma, R.; Wang, Y.; Shi, Z.; Lu, T.; Feng, J. Visible-Light-Promoted α,α-Dibromination in Minutes: Efficient Route for Construction of Quaternary Carbon Centers. ACS Sustain. Chem. Eng. 2019, 7, 18542–18546. [Google Scholar] [CrossRef]
  104. Buckle, D.R.; Cantello, B.C.C.; Smith, H.; Spicer, B.A. 2-Cyano-1,3-Dicarbonyl Compounds with Antiallergic Activity. J. Med. Chem. 1977, 20, 265–269. [Google Scholar] [CrossRef] [PubMed]
  105. Paquette, L.A.; Crich, D.; Fuchs, P.L.; Molander, G.A. Encyclopedia of Reagents for Organic Synthesis; Wiley: New York, NY, USA, 2009. [Google Scholar]
  106. Newman, D.J.; Cragg, G.M.; Snader, K.M. The Influence of Natural Products upon Drug Discovery. Nat. Prod. Rep. 2000, 17, 215–234. [Google Scholar] [CrossRef] [PubMed]
  107. Fürstner, A. Cover Picture: Total Syntheses and Biological Assessment of Macrocyclic Glycolipids (Eur. J. Org. Chem. 5/2004). Eur. J. Org. Chem. 2004, 2004, 933. [Google Scholar] [CrossRef]
  108. Layton, M.E.; Morales, C.A.; Shair, M.D. Biomimetic Synthesis of (−)-Longithorone A. J. Am. Chem. Soc. 2002, 124, 773–775. [Google Scholar] [CrossRef]
  109. Yoshinari, T.; Ohmori, K.; Schrems, M.G.; Pfaltz, A.; Suzuki, K. Total Synthesis and Absolute Configuration of Macrocidin A, a Cyclophane Tetramic Acid Natural Product. Angew. Chem. Int. Ed. 2010, 49, 881–885. [Google Scholar] [CrossRef]
  110. Huang, M.; Song, L.; Liu, B. Construction of the Cyclophane Core of the Hirsutellones via a RCM Strategy. Org. Lett. 2010, 12, 2504–2507. [Google Scholar] [CrossRef]
  111. Diederich, F.; Lutter, H.D. Catalytic Cyclophanes. 4. Supramolecular Catalysis of Benzoin Condensations by a Thiazolium Cyclophane. J. Am. Chem. Soc. 1989, 111, 8438–8446. [Google Scholar] [CrossRef]
  112. Driggers, E.M.; Hale, S.P.; Lee, J.; Terrett, N.K. The Exploration of Macrocycles for Drug Discovery—An Underexploited Structural Class. Nat. Rev. Drug Discov. 2008, 7, 608–624. [Google Scholar] [CrossRef]
  113. Rajakumar, P.; Mohammed Abdul Rasheed, A.; Iman Rabia, A.; Chamundeeswari, D. Synthesis and Study of Anti-Inflammatory Activity of Some Novel Cyclophane Amides. Bioorg. Med. Chem. Lett. 2006, 16, 6019–6023. [Google Scholar] [CrossRef]
  114. Rajakumar, P.; Mohammed Abdul Rasheed, A.; Balu, P.M.; Murugesan, K. Synthesis, Characterization, and Anti-Bacterial Efficacy of Some Novel Cyclophane Amide. Bioorg. Med. Chem. 2006, 14, 7458–7467. [Google Scholar] [CrossRef] [PubMed]
  115. Seel, C.; Vögtle, F. Molecules with Large Cavities in Supramolecular Chemistry. Angew. Chem. Int. Ed. Engl. 1992, 31, 528–549. [Google Scholar] [CrossRef]
  116. Zakarian, J.E.; El-Azizi, Y.; Collins, S.K. Exploiting Quadrupolar Interactions in the Synthesis of the Macrocyclic Portion of Longithorone C. Org. Lett. 2008, 10, 2927–2930. [Google Scholar] [CrossRef] [PubMed]
  117. Kotha, S.; Meshram, M.; Tiwari, A. Advanced Approach to Polycyclics by a Synergistic Combination of Enyne Metathesis and Diels–Alder Reaction. Chem. Soc. Rev. 2009, 38, 2065–2092. [Google Scholar] [CrossRef]
  118. Kotha, S.; Krishna, N.G.; Halder, S.; Misra, S. A Synergistic Approach to Polycyclics via a Strategic Utilization of Claisen Rearrangement and Olefin Metathesis. Org. Biomol. Chem. 2011, 9, 5597–5624. [Google Scholar] [CrossRef]
  119. Jimenez, L.; Diederich, F. Catalytic Cyclophanes: A Highly Efficient Model for Pyruvate Oxidase. Tetrahedron Lett. 1989, 30, 2759–2762. [Google Scholar] [CrossRef]
  120. Gleiter, R.; Hopf, H. Modern Cyclophane Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weiheim, Germany, 2004; ISBN 978-3-527-60396-1. [Google Scholar]
  121. Weber, E. Cyclophanes; Springer: Berlin/Heidelberg, Germany, 1994; Volume 172. [Google Scholar]
  122. Vögtle, F. Cyclophane Chemistry: Synthesis, Structures, and Reactions; John Wiley and Sons: Chichester, UK, 1993. [Google Scholar]
  123. Diederich, F. Cyclophanes; Royal Society of Chemistry: Cambridge, UK, 1991. [Google Scholar]
  124. Takemura, H. Cyclophane Chemistry for the 21st Century; Research Signpost: Trivandrum, India, 2002. [Google Scholar]
  125. Davis, F.; Higson, S. Macrocycles: Construction, Chemistry and Nanotechnology Applications; John Wiley and Sons Ltd.: Chichester, UK, 2011. [Google Scholar]
  126. Kotha, S.; Shirbhate, M.E.; Waghule, G.T. Selected Synthetic Strategies to Cyclophanes. Beilstein J. Org. Chem. 2015, 11, 1274–1331. [Google Scholar] [CrossRef]
  127. Cram, D.J.; Steinberg, H. Macro Rings. I. Preparation and Spectra of the Paracyclophanes. J. Am. Chem. Soc. 1951, 73, 5691–5704. [Google Scholar] [CrossRef]
  128. Cram, D.J.; Helgeson, R.C. Macro Rings. XXXIV. A Ring Expansion Route to the Higher Paracyclophanes, and Spectra-Structure Correlations of Their Derived Ketones1. J. Am. Chem. Soc. 1966, 88, 3515–3521. [Google Scholar] [CrossRef]
  129. Krois, D.; Lehner, H. [4.2]- and [4.3]-Metacyclophanes. J. Chem. Soc. Perkin 1 1982, 2369–2372. [Google Scholar] [CrossRef]
  130. Tamao, K.; Kodama, S.; Nakatsuka, T.; Kiso, Y.; Kumada, M. One-Step Preparation of Metacyclophanes and (2,6)Pyridinophanes by Nickel-Catalzyed Grignard Cyclocoupling. J. Am. Chem. Soc. 1975, 97, 4405–4406. [Google Scholar] [CrossRef]
  131. Errede, L.A.; Gregorian, R.S.; Hoyt, J.M. The Chemistry of Xylylenes. VI. The Polymerization of p-Xylylene2. J. Am. Chem. Soc. 1960, 82, 5218–5223. [Google Scholar] [CrossRef]
  132. Pechlivanidis, Z.; Hopf, H.; Ernst, L. Paracyclophanes: Extending the Bridges. Synthesis. Eur. J. Org. Chem. 2009, 2009, 223–237. [Google Scholar] [CrossRef]
  133. Allinger, N.L.; Cram, D.J. Macro Rings. IV. The Preparation of Three New Paracyclophanes1. J. Am. Chem. Soc. 1954, 76, 2362–2367. [Google Scholar] [CrossRef]
  134. Carbonnelle, A.-C.; Zhu, J. A Novel Synthesis of Biaryl-Containing Macrocycles by a Domino Miyaura Arylboronate Formation:  Intramolecular Suzuki Reaction. Org. Lett. 2000, 2, 3477–3480. [Google Scholar] [CrossRef] [PubMed]
  135. Smith, B.B.; Hill, D.E.; Cropp, T.A.; Walsh, R.D.; Cartrette, D.; Hipps, S.; Shachter, A.M.; Pennington, W.T.; Kwochka, W.R. Synthesis of [n]- and [n.n]Cyclophanes by Using Suzuki−Miyaura Coupling. J. Org. Chem. 2002, 67, 5333–5337. [Google Scholar] [CrossRef]
  136. Bodwell, G.J.; Li, J. Concise Synthesis and Transannular Inverse Electron Demand Diels−Alder Reaction of [3](3,6)Pyridazino[3](1,3)Indolophane. Rapid Access to a Pentacyclic Indoloid System. Org. Lett. 2002, 4, 127–130. [Google Scholar] [CrossRef]
  137. Bodwell, G.J.; Li, J. A Concise Formal Total Synthesis of (±)-Strychnine by Using a Transannular Inverse-Electron-Demand Diels–Alder Reaction of a [3](1,3)Indolo[3](3,6)Pyridazinophane. Angew. Chem. Int. Ed. 2002, 41, 3261–3262. [Google Scholar] [CrossRef]
  138. Smith, A.B.; Adams, C.M.; Kozmin, S.A. On the Reversible Nature of the Olefin Cross Metathesis Reaction. J. Am. Chem. Soc. 2001, 123, 990–991. [Google Scholar] [CrossRef]
  139. Locke, A.J.; Jones, C.; Richards, C.J. A Rapid Approach to Ferrocenophanes via Ring-Closing Metathesis. J. Organomet. Chem. 2001, 637–639, 669–676. [Google Scholar] [CrossRef]
  140. Fürstner, A.; Stelzer, F.; Rumbo, A.; Krause, H. Total Synthesis of the Turrianes and Evaluation of Their DNA-Cleaving Properties. Chem. Eur. J. 2002, 8, 1856–1871. [Google Scholar] [CrossRef]
  141. Martinez, V.; Blais, J.-C.; Astruc, D. A Fast Organometallic Route from P-Xylene, Mesitylene, and p-Diisopropylbenzene to Organoiron and Polycyclic Aromatic Cyclophanes, Capsules and Polymers. Angew. Chem. Int. Ed. 2003, 42, 4366–4369. [Google Scholar] [CrossRef] [PubMed]
  142. Tae, J.; Yang, Y.-K. Efficient Synthesis of Macrocyclic Paracyclophanes by Ring-Closing Metathesis Dimerization and Trimerization Reactions. Org. Lett. 2003, 5, 741–744. [Google Scholar] [CrossRef] [PubMed]
  143. Watson, M.D.; Jäckel, F.; Severin, N.; Rabe, J.P.; Müllen, K. A Hexa-Peri-Hexabenzocoronene Cyclophane:  An Addition to the Toolbox for Molecular Electronics. J. Am. Chem. Soc. 2004, 126, 1402–1407. [Google Scholar] [CrossRef] [PubMed]
  144. Martinez, V.; Blais, J.-C.; Bravic, G.; Astruc, D. Coupling Multiple Benzylic Activation of Simple Arenes by CpFe+ with Multiple Alkene Metathesis Using Grubbs Catalysts:  An Efficient Carbon−Carbon Bond Formation Strategy Leading to Polycycles, Cyclophanes, Capsules, and Polymeric Compounds and Their CpFe+ Complexes. Organometallics 2004, 23, 861–874. [Google Scholar] [CrossRef]
  145. Branowska, D.; Buczek, I.; Kalińska, K.; Nowaczyk, J.; Rykowski, A. S-Transalkylation/Ring Closing Metathesis as a Route to Azathiamacrocycles Incorporating 2,2′-Bipyridine Subunits. Tetrahedron Lett. 2005, 46, 8539–8541. [Google Scholar] [CrossRef]
  146. Ueda, T.; Kanomata, N.; Machida, H. Synthesis of Planar-Chiral Paracyclophanes via Samarium(II)-Catalyzed Intramolecular Pinacol Coupling. Org. Lett. 2005, 7, 2365–2368. [Google Scholar] [CrossRef]
  147. Branowska, D.; Rykowski, A. Ring-Closing Metathesis Approach to Symmetrical and Unsymmetrical Cycloakeno[c]Fused 2,2′-Bipyridine-Based Cyclophanes. Tetrahedron 2005, 61, 10713–10718. [Google Scholar] [CrossRef]
  148. Kotha, S.; Chavan, A.S.; Shaikh, M. Diversity-Oriented Approach to Macrocyclic Cyclophane Derivatives by Suzuki–Miyaura Cross-Coupling and Olefin Metathesis as Key Steps. J. Org. Chem. 2012, 77, 482–489. [Google Scholar] [CrossRef]
  149. Alfimov, M.V.; Fedorova, O.A.; Gromov, S.P. Photoswitchable Molecular Receptors. Photoreact. Control Photofunct. Mater. II 2003, 158, 183–198. [Google Scholar] [CrossRef]
  150. Li, J.; Yim, D.; Jang, W.-D.; Yoon, J. Recent Progress in the Design and Applications of Fluorescence Probes Containing Crown Ethers. Chem. Soc. Rev. 2017, 46, 2437–2458. [Google Scholar] [CrossRef] [PubMed]
  151. Gokel, G.W.; Leevy, W.M.; Weber, M.E. Crown Ethers:  Sensors for Ions and Molecular Scaffolds for Materials and Biological Models. Chem. Rev. 2004, 104, 2723–2750. [Google Scholar] [CrossRef] [PubMed]
  152. Ahmedova, A.; Burdzhiev, N.; Ciattini, S.; Stanoeva, E.; Mitewa, M. Synthesis, Structure, Spectral and Coordination Properties of a Crown Ether Derivative of 1,3-Indandione. A New Structural Evidence for the Versatile Reactivity of 2-Acetyl-1,3-Indandione. Comptes Rendus Chim. 2010, 13, 1269–1277. [Google Scholar] [CrossRef]
  153. Sartori, G.; Casnati, G.; Bigi, F.; Baraldi, D. Friedel-Crafts Coordinated Processes: Selective Cyclooligomerization of Acyl Chlorides. Tetrahedron Lett. 1991, 32, 2153–2156. [Google Scholar] [CrossRef]
  154. Kilgore, L.B.; Ford, J.H.; Wolfe, W.C. Insecticidal Properties of 1,3-Indandiones. Ind. Eng. Chem. 1942, 34, 494–497. [Google Scholar] [CrossRef]
  155. Teotonio, E.E.S.; Brito, H.F.; Viertler, H.; Faustino, W.M.; Malta, O.L.; de Sá, G.F.; Felinto, M.C.F.C.; Santos, R.H.A.; Cremona, M. Synthesis and Luminescent Properties of Eu3+-Complexes with 2-Acyl-1,3-Indandionates (ACIND) and TPPO Ligands: The First X-Ray Structure of Eu–ACIND Complex. Polyhedron 2006, 25, 3488–3494. [Google Scholar] [CrossRef]
  156. Griffin, M.O.; Fricovsky, E.; Ceballos, G.; Villarreal, F. Tetracyclines: A Pleitropic Family of Compounds with Promising Therapeutic Properties. Review of the Literature. Am. J. Physiol.-Cell Physiol. 2010, 299, C539–C548. [Google Scholar] [CrossRef]
  157. Li, M.; Lv, X.-L.; Wen, L.-R.; Hu, Z.-Q. Direct Solvent-Free Regioselective Construction of Pyrrolo[1,2-a][1,10]Phenanthrolines Based on Isocyanide-Based Multicomponent Reactions. Org. Lett. 2013, 15, 1262–1265. [Google Scholar] [CrossRef]
  158. Pirrung, M.C.; Sarma, K.D. Aqueous Medium Effects on Multi-Component Reactions. Multicomponent React. 2005, 61, 11456–11472. [Google Scholar] [CrossRef]
  159. Wang, S.-X.; Wang, M.-X.; Wang, D.-X.; Zhu, J. Catalytic Enantioselective Passerini Three-Component Reaction. Angew. Chem. Int. Ed. 2008, 47, 388–391. [Google Scholar] [CrossRef]
  160. Kłossowski, S.; Wiraszka, B.; Berłożecki, S.; Ostaszewski, R. Model Studies on the First Enzyme-Catalyzed Ugi Reaction. Org. Lett. 2013, 15, 566–569. [Google Scholar] [CrossRef] [PubMed]
  161. Andreana, P.R.; Liu, C.C.; Schreiber, S.L. Stereochemical Control of the Passerini Reaction. Org. Lett. 2004, 6, 4231–4233. [Google Scholar] [CrossRef] [PubMed]
  162. Kusebauch, U.; Beck, B.; Messer, K.; Herdtweck, E.; Dömling, A. Massive Parallel Catalyst Screening:  Toward Asymmetric MCRs. Org. Lett. 2003, 5, 4021–4024. [Google Scholar] [CrossRef] [PubMed]
  163. Dömling, A.; Ugi, I. Multicomponent Reactions with Isocyanides. Angew. Chem. Int. Ed. 2000, 39, 3168–3210. [Google Scholar] [CrossRef]
  164. Jiang, X.; Tang, T.; Wang, J.-M.; Chen, Z.; Zhu, Y.-M.; Ji, S.-J. Palladium-Catalyzed One-Pot Synthesis of Quinazolinones via Tert-Butyl Isocyanide Insertion. J. Org. Chem. 2014, 79, 5082–5087. [Google Scholar] [CrossRef] [PubMed]
  165. Tang, T.; Jiang, X.; Wang, J.-M.; Sun, Y.-X.; Zhu, Y.-M. Divergent Synthesis of 6H-Isoindolo[2,1-a]Indol-6-Ones and Indenoindolones: An Investigation of Pd-Catalyzed Isocyanide Insertion. Tetrahedron 2014, 70, 2999–3004. [Google Scholar] [CrossRef]
  166. Duan, H.; Chen, Z.; Han, L.; Feng, Y.; Zhu, Y.; Yang, S. Palladium-Catalyzed Chemoselective Synthesis of Indane-1,3-Dione Derivatives via Tert-Butyl Isocyanide Insertion. Org. Biomol. Chem. 2015, 13, 6782–6788. [Google Scholar] [CrossRef]
  167. Wan, J.-P.; Jing, Y. Recent Advances in Copper-Catalyzed C-H Bond Amidation. Beilstein J. Org. Chem. 2015, 11, 2209–2222. [Google Scholar] [CrossRef]
  168. Liang, C.; Collet, F.; Robert-Peillard, F.; Müller, P.; Dodd, R.H.; Dauban, P. Toward a Synthetically Useful Stereoselective C−H Amination of Hydrocarbons. J. Am. Chem. Soc. 2008, 130, 343–350. [Google Scholar] [CrossRef]
  169. Bryant, J.R.; Mayer, J.M. Oxidation of C−H Bonds by [(Bpy)2(Py)RuIVO]2+ Occurs by Hydrogen Atom Abstraction. J. Am. Chem. Soc. 2003, 125, 10351–10361. [Google Scholar] [CrossRef]
  170. Pelletier, G.; Powell, D.A. Copper-Catalyzed Amidation of Allylic and Benzylic CH Bonds. Org. Lett. 2006, 8, 6031–6034. [Google Scholar] [CrossRef] [PubMed]
  171. Smith, K.; Hupp, C.D.; Allen, K.L.; Slough, G.A. Catalytic Allylic Amination versus Allylic Oxidation:  A Mechanistic Dichotomy. Organometallics 2005, 24, 1747–1755. [Google Scholar] [CrossRef]
  172. Lu, H.; Subbarayan, V.; Tao, J.; Zhang, X.P. Cobalt(II)-Catalyzed Intermolecular Benzylic C−H Amination with 2,2,2-Trichloroethoxycarbonyl Azide (TrocN3). Organometallics 2010, 29, 389–393. [Google Scholar] [CrossRef]
  173. Badiei, Y.M.; Dinescu, A.; Dai, X.; Palomino, R.M.; Heinemann, F.W.; Cundari, T.R.; Warren, T.H. Copper–Nitrene Complexes in Catalytic C-H Amination. Angew. Chem. Int. Ed. 2008, 47, 9961–9964. [Google Scholar] [CrossRef] [PubMed]
  174. Huard, K.; Lebel, H. N-Tosyloxycarbamates as Reagents in Rhodium-Catalyzed C-H Amination Reactions. Chem. Eur. J. 2008, 14, 6222–6230. [Google Scholar] [CrossRef] [PubMed]
  175. Kalita, B.; Lamar, A.A.; Nicholas, K.M. Hydrous Zinc Halide-Catalyzed Aminosulfonation of Hydrocarbons. Chem. Commun. 2008, 36, 4291–4293. [Google Scholar] [CrossRef]
  176. Li, Z.; Capretto, D.A.; Rahaman, R.; He, C. Silver-Catalyzed Intermolecular Amination of C-H Groups. Angew. Chem. Int. Ed. 2007, 46, 5184–5186. [Google Scholar] [CrossRef]
  177. Harden, J.D.; Ruppel, J.V.; Gao, G.-Y.; Zhang, X.P. Cobalt-Catalyzed Intermolecular C-H Amination with Bromamine-T as Nitrene Source. Chem. Commun. 2007, 44, 4644–4646. [Google Scholar] [CrossRef]
  178. Bhuyan, R.; Nicholas, K.M. Efficient Copper-Catalyzed Benzylic Amidation with Anhydrous Chloramine-T. Org. Lett. 2007, 9, 3957–3959. [Google Scholar] [CrossRef]
  179. Lebel, H.; Huard, K. De Novo Synthesis of Troc-Protected Amines:  Intermolecular Rhodium-Catalyzed C−H Amination with N-Tosyloxycarbamates. Org. Lett. 2007, 9, 639–642. [Google Scholar] [CrossRef]
  180. Fiori, K.W.; Du Bois, J. Catalytic Intermolecular Amination of C−H Bonds: Method Development and Mechanistic Insights. J. Am. Chem. Soc. 2007, 129, 562–568. [Google Scholar] [CrossRef] [PubMed]
  181. Fructos, M.R.; Trofimenko, S.; Díaz-Requejo, M.M.; Pérez, P.J. Facile Amine Formation by Intermolecular Catalytic Amidation of Carbon−Hydrogen Bonds. J. Am. Chem. Soc. 2006, 128, 11784–11791. [Google Scholar] [CrossRef] [PubMed]
  182. Reddy, R.P.; Davies, H.M.L. Dirhodium Tetracarboxylates Derived from Adamantylglycine as Chiral Catalysts for Enantioselective C−H Aminations. Org. Lett. 2006, 8, 5013–5016. [Google Scholar] [CrossRef] [PubMed]
  183. Liang, C.; Robert-Peillard, F.; Fruit, C.; Müller, P.; Dodd, R.H.; Dauban, P. Efficient Diastereoselective Intermolecular Rhodium-Catalyzed C-H Amination. Angew. Chem. Int. Ed. 2006, 45, 4641–4644. [Google Scholar] [CrossRef] [PubMed]
  184. Leung, S.K.-Y.; Tsui, W.-M.; Huang, J.-S.; Che, C.-M.; Liang, J.-L.; Zhu, N. Imido Transfer from Bis(Imido)Ruthenium(VI) Porphyrins to Hydrocarbons:  Effect of Imido Substituents, C−H Bond Dissociation Energies, and RuVI/V Reduction Potentials. J. Am. Chem. Soc. 2005, 127, 16629–16640. [Google Scholar] [CrossRef]
  185. Yamawaki, M.; Tsutsui, H.; Kitagaki, S.; Anada, M.; Hashimoto, S. Dirhodium(II) Tetrakis[N-Tetrachlorophthaloyl-(S)-Tert-Leucinate]: A New Chiral Rh(II) Catalyst for Enantioselective Amidation of C-H Bonds. Tetrahedron Lett. 2002, 43, 9561–9564. [Google Scholar] [CrossRef]
  186. Kohmura, Y.; Katsuki, T. Mn(Salen)-Catalyzed Enantioselective C-H Amination. Tetrahedron Lett. 2001, 42, 3339–3342. [Google Scholar] [CrossRef]
  187. Albone, D.P.; Aujla, P.S.; Challenger, S.; Derrick, A.M. A Simple Copper Catalyst for Both Aziridination of Alkenes and Amination of Activated Hydrocarbons with Chloramine-T Trihydrate. J. Org. Chem. 1998, 63, 9569–9571. [Google Scholar] [CrossRef]
  188. Powell, D.A.; Fan, H. Copper-Catalyzed Amination of Primary Benzylic C−H Bonds with Primary and Secondary Sulfonamides. J. Org. Chem. 2010, 75, 2726–2729. [Google Scholar] [CrossRef]
  189. Andrus, M.B.; Chen, X. Catalytic Enantioselective Allylic Oxidation of Olefins with Copper(I) Catalysts and New Perester Oxidants. Tetrahedron 1997, 53, 16229–16240. [Google Scholar] [CrossRef]
  190. Kadish, K.M.; Smith, K.M.; Guilard, R. Handbook of Porphyrin Science; World Scientific Publishing: Singapore, 2010. [Google Scholar]
  191. Shy, H.; Mackin, P.; Orvieto, A.S.; Gharbharan, D.; Peterson, G.R.; Bampos, N.; Hamilton, T.D. The Two-Step Mechanochemical Synthesis of Porphyrins. Faraday Discuss. 2014, 170, 59–69. [Google Scholar] [CrossRef] [PubMed]
  192. Fox, S.; Boyle, R.W. Synthetic Routes to Porphyrins Bearing Fused Rings. Tetrahedron 2006, 62, 10039–10054. [Google Scholar] [CrossRef]
  193. Akhigbe, J.; Luciano, M.; Zeller, M.; Brückner, C. Mono- and Bisquinoline-Annulated Porphyrins from Porphyrin β,Β′-Dione Oximes. J. Org. Chem. 2015, 80, 499–511. [Google Scholar] [CrossRef] [PubMed]
  194. Götz, D.C.G.; Gehrold, A.C.; Dorazio, S.J.; Daddario, P.; Samankumara, L.; Bringmann, G.; Brückner, C.; Bruhn, T. Indaphyrins and Indachlorins: Optical and Chiroptical Properties of a Family of Helimeric Porphyrinoids. Eur. J. Org. Chem. 2015, 2015, 3913–3922. [Google Scholar] [CrossRef]
  195. Aguiar, A.; Leite, A.; Silva, A.M.N.; Tomé, A.C.; Cunha-Silva, L.; de Castro, B.; Rangel, M.; Silva, A.M.G. Isoxazolidine-Fused Meso-Tetraarylchlorins as Key Tools for the Synthesis of Mono- and Bis-Annulated Chlorins. Org. Biomol. Chem. 2015, 13, 7131–7135. [Google Scholar] [CrossRef]
  196. Brückner, C. The Breaking and Mending of Meso-Tetraarylporphyrins: Transmuting the Pyrrolic Building Blocks. Acc. Chem. Res. 2016, 49, 1080–1092. [Google Scholar] [CrossRef]
  197. Jaquinod, L.; Gros, C.; Olmstead, M.M.; Antolovich, M.; Smith, K.M. First Syntheses of Fused Pyrroloporphyrins. Chem. Commun. 1996, 12, 1475–1476. [Google Scholar] [CrossRef]
  198. Jaquinod, L.; Gros, C.; Khoury, R.G.; Smith, K.M. A Convenient Synthesis of Functionalized Tetraphenylchlorins. Chem. Commun. 1996, 22, 2581–2582. [Google Scholar] [CrossRef]
  199. Shea, K.M.; Jaquinod, L. Dodecasubstituted Metallochlorins (Metallodihydroporphyrins). Chem. Commun. 1998, 7, 759–760. [Google Scholar] [CrossRef]
  200. Gros, C.P.; Jaquinod, L.; Khoury, R.G.; Olmstead, M.M.; Smith, K.M. Approaches to β-Fused Porphyrinoporphyrins: Pyrrolo- and Dipyrromethaneporphyrins. J. Porphyr. Phthalocyanines 1997, 01, 201–212. [Google Scholar] [CrossRef]
  201. Shea, K.M.; Jaquinod, L.; Smith, K.M. Dihydroporphyrin Synthesis:  New Methodology. J. Org. Chem. 1998, 63, 7013–7021. [Google Scholar] [CrossRef] [PubMed]
  202. Shelnutt, J.A.; Song, X.-Z.; Ma, J.-G.; Jia, S.-L.; Jentzen, W.; Medforth, C.J. Nonplanar Porphyrins and Their Significance in Proteins. Chem. Soc. Rev. 1998, 27, 31–42. [Google Scholar] [CrossRef]
  203. Mandon, D.; Ochenbein, P.; Fischer, J.; Weiss, R.; Jayaraj, K.; Austin, R.N.; Gold, A.; White, P.S.; Brigaud, O. .beta.-Halogenated-Pyrrole Porphyrins. Molecular Structures of 2,3,7,8,12,13,17,18-Octabromo-5,10,15,20-Tetramesitylporphyrin, Nickel(II) 2,3,7,8,12,13,17,18-Octabromo-5,10,15,20-Tetramesitylporphyrin, and Nickel(II) 2,3,7,8,12,13,17,18-Octabromo-5,10,15,20-Tetrakis(Pentafluorophenyl)Porphyrin. Inorg. Chem. 1992, 31, 2044–2049. [Google Scholar] [CrossRef]
  204. Hodge, J.A.; Hill, M.G.; Gray, H.B. Electrochemistry of Nonplanar Zinc(II) Tetrakis(Pentafluorophenyl)Porphyrins. Inorg. Chem. 1995, 34, 809–812. [Google Scholar] [CrossRef]
  205. Kojima, T.; Nakanishi, T.; Harada, R.; Ohkubo, K.; Yamauchi, S.; Fukuzumi, S. Selective Inclusion of Electron-Donating Molecules into Porphyrin Nanochannels Derived from the Self-Assembly of Saddle-Distorted, Protonated Porphyrins and Photoinduced Electron Transfer from Guest Molecules to Porphyrin Dications. Chem. Eur. J. 2007, 13, 8714–8725. [Google Scholar] [CrossRef] [PubMed]
  206. Chaudhri, N.; Grover, N.; Sankar, M. Versatile Synthetic Route for β-Functionalized Chlorins and Porphyrins by Varying the Size of Michael Donors: Syntheses, Photophysical, and Electrochemical Redox Properties. Inorg. Chem. 2017, 56, 11532–11545. [Google Scholar] [CrossRef]
  207. Lindsey, J.S. De Novo Synthesis of Gem-Dialkyl Chlorophyll Analogues for Probing and Emulating Our Green World. Chem. Rev. 2015, 115, 6534–6620. [Google Scholar] [CrossRef]
  208. Taniguchi, M.; Lindsey, J.S. Synthetic Chlorins, Possible Surrogates for Chlorophylls, Prepared by Derivatization of Porphyrins. Chem. Rev. 2017, 117, 344–535. [Google Scholar] [CrossRef]
  209. Samankumara, L.P.; Dorazio, S.J.; Akhigbe, J.; Li, R.; Nimthong-Roldán, A.; Zeller, M.; Brückner, C. Indachlorins: Nonplanar Indanone-Annulated Chlorin Analogues with Panchromatic Absorption Spectra between 300 and 900 Nm. Chem. Eur. J. 2015, 21, 11118–11128. [Google Scholar] [CrossRef]
  210. McCarthy, J.R.; Hyland, M.A.; Brückner, C. Synthesis of Indaphyrins: Meso-Tetraarylsecochlorin-Based Porphyrinoids Containing Direct o-Phenyl-to-β-Linkages. Org. Biomol. Chem. 2004, 2, 1484–1491. [Google Scholar] [CrossRef]
  211. Nishiyabu, R.; Anzenbacher, P. Sensing of Antipyretic Carboxylates by Simple Chromogenic Calix[4]Pyrroles. J. Am. Chem. Soc. 2005, 127, 8270–8271. [Google Scholar] [CrossRef] [PubMed]
  212. Nishiyabu, R.; Anzenbacher, P. 1,3-Indane-Based Chromogenic Calixpyrroles with Push−Pull Chromophores:  Synthesis and Anion Sensing. Org. Lett. 2006, 8, 359–362. [Google Scholar] [CrossRef] [PubMed]
  213. Kim, S.-H.; Hong, S.-J.; Yoo, J.; Kim, S.K.; Sessler, J.L.; Lee, C.-H. Strapped Calix[4]Pyrroles Bearing a 1,3-Indanedione at a β-Pyrrolic Position: Chemodosimeters for the Cyanide Anion. Org. Lett. 2009, 11, 3626–3629. [Google Scholar] [CrossRef] [PubMed]
  214. Chaudhri, N.; Grover, N.; Sankar, M. Nickel-Induced Skeletal Rearrangement of Free Base Trans-Chlorins into Monofused NiII-Porphyrins: Synthesis, Structural, Spectral, and Electrochemical Redox Properties. Inorg. Chem. 2018, 57, 11349–11360. [Google Scholar] [CrossRef] [PubMed]
  215. Mchardy, N.; Hudson, A.T.; Morgan, D.W.T.; Rae, D.G.; Dolan, T.T. Activity of 10 Naphthoquinones, Including Parvaquone (993C) and Menoctone, in Cattle Artificially Infected with Theileria Parva. Res. Vet. Sci. 1983, 35, 347–352. [Google Scholar] [CrossRef]
  216. Patil, P.C.; Akamanchi, K.G. Simple and Effective Route for Synthesis of Parvaquone, an Antiprotozoal Drug. RSC Adv. 2014, 4, 58214–58216. [Google Scholar] [CrossRef]
  217. Britton, H.; Catterick, D.; Dwyer, A.N.; Gordon, A.H.; Leach, S.G.; McCormick, C.; Mountain, C.E.; Simpson, A.; Stevens, D.R.; Urquhart, M.W.J.; et al. Discovery and Development of an Efficient Process to Atovaquone. Org. Process Res. Dev. 2012, 16, 1607–1617. [Google Scholar] [CrossRef]
  218. Holt, G.; Wall, D.K. Some Reactions of 2-Diazoindane-1,3-Dione. J. Chem. Soc. C Org. 1966, 857–858. [Google Scholar] [CrossRef]
  219. Spangler, R.J.; Kim, J.H.; Cava, M.P. Pyrolytic and Photochemical Wolff Rearrangement of Diazoindanones. Synthesis of 2-Carboalkoxybenzocyclobutenones. J. Org. Chem. 1977, 42, 1697–1703. [Google Scholar] [CrossRef]
  220. Regitz, M.; Heck, G. Synthesen Und Einige Umsetzungen Des 2-Diazo- Und Des 2-Hydroxy-Indandions-(1.3). Chem. Ber. 1964, 97, 1482–1501. [Google Scholar] [CrossRef]
  221. Rosenfeld, M.J.; Shankar, B.K.R.; Shechter, H. Rhodium(II) Acetate-Catalyzed Reactions of 2-Diazo-1,3-Indandione and 2-Diazo-1-Indanone with Various Substrates. J. Org. Chem. 1988, 53, 2699–2705. [Google Scholar] [CrossRef]
  222. Alloum, A.B.; Villemin, D. Potassium Fluoride on Alumina: An Easy Preparation of Diazocarbonyl Compounds. Synth. Commun. 1989, 19, 2567–2571. [Google Scholar] [CrossRef]
  223. Zhang, J.; Chen, W.; Huang, D.; Zeng, X.; Wang, X.; Hu, Y. Tandem Synthesis of α-Diazoketones from 1,3-Diketones. J. Org. Chem. 2017, 82, 9171–9174. [Google Scholar] [CrossRef] [PubMed]
  224. Chiang, Y.; Kresge, A.J.; Zhu, Y. Kinetics and Mechanism of the Base–Catalyzed Cleavage of 2-Diazo-1,3-Indanedione and the Acid–Catalyzed Decomposition of Its Hydrolysis Product, 2-(Diazoacetyl)Benzoic Acid. ARKIVOC 2001, 2001, 108–115. [Google Scholar] [CrossRef]
  225. Reichel, L.; Hampel, W. Chemie Und Biochemie Der Pflanzenstoffe, XXI1) Über Das 1.4-Dioxo-Isochroman. Justus Liebigs Ann. Chem. 1968, 712, 152–154. [Google Scholar] [CrossRef]
  226. Asif, M. Mini Review on Important Biological Properties of Benzofuran Derivatives. J. Anal. Pharm. Res. 2016, 3, 11–12. [Google Scholar] [CrossRef] [Green Version]
  227. Miao, Y.; Hu, Y.; Yang, J.; Liu, T.; Sun, J.; Wang, X. Natural Source, Bioactivity and Synthesis of Benzofuran Derivatives. RSC Adv. 2019, 9, 27510–27540. [Google Scholar] [CrossRef]
  228. Sun, W.; Sarma, J.S.M.; Singh, B.N. Electrophysiological Effects of Dronedarone (SR33589), a Noniodinated Benzofuran Derivative, in the Rabbit Heart. Circulation 1999, 100, 2276–2281. [Google Scholar] [CrossRef]
  229. Oter, O.; Ertekin, K.; Kirilmis, C.; Koca, M.; Ahmedzade, M. Characterization of a Newly Synthesized Fluorescent Benzofuran Derivative and Usage as a Selective Fiber Optic Sensor for Fe(III). Sens. Actuators B Chem. 2007, 122, 450–456. [Google Scholar] [CrossRef]
  230. Karatas, F.; Koca, M.; Kara, H.; Servi, S. Synthesis and Oxidant Properties of Novel (5-Bromobenzofuran-2-Yl)(3-Methyl-3-Mesitylcyclobutyl)Ketonethiosemicarbazone. Eur. J. Med. Chem. 2006, 41, 664–669. [Google Scholar] [CrossRef]
  231. Habermann, J.; Ley, S.V.; Smits, R. Three-Step Synthesis of an Array of Substituted Benzofurans Using Polymer-Supported Reagents. J. Chem. Soc. Perkin 1 1999, 17, 2421–2423. [Google Scholar] [CrossRef]
  232. Ameri, M.; Asghari, A.; Amoozadeh, A.; Bakherad, M.; Nematollahi, D. An Efficient, Simple, Non-Catalytic Electrosynthesis of New Polycyclic Benzofuran Derivatives. Tetrahedron Lett. 2015, 56, 2141–2144. [Google Scholar] [CrossRef]
  233. Allais, C.; Grassot, J.-M.; Rodriguez, J.; Constantieux, T. Metal-Free Multicomponent Syntheses of Pyridines. Chem. Rev. 2014, 114, 10829–10868. [Google Scholar] [CrossRef]
  234. Damavandi, S.; Sandaroos, R. Solvent-Free One Pot Synthesis of Indenoquinolinones Catalyzed by Iron(III) Triflate. Heterocycl. Commun. 2011, 17, 121–124. [Google Scholar] [CrossRef]
  235. Maleki, A.; Nooraie Yeganeh, N. Facile One-Pot Synthesis of a Series of 7-Aryl-8H-Benzo[h]Indeno[1,2-b]Quinoline-8-One Derivatives Catalyzed by Cellulose-Based Magnetic Nanocomposite. Appl. Organomet. Chem. 2017, 31, e3814. [Google Scholar] [CrossRef]
  236. Safajoo, N.; Mirjalili, B.B.F.; Bamoniri, A. A Facile and Clean Synthesis of Indenopyrido[2,3-d]Pyrimidines in the Presence of Fe3O4@NCs/Cu(II) as Bio-Based Magnetic Nano-Catalyst. Polycycl. Aromat. Compd. 2021, 41, 1241–1248. [Google Scholar] [CrossRef]
  237. Mamaghani, M.; Shirini, F.; Bassereh, E.; Hossein Nia, R. 1,2-Dimethyl-N-Butanesulfonic Acid Imidazolium Hydrogen Sulfate as Efficient Ionic Liquid Catalyst in the Synthesis of Indeno Fused Pyrido[2,3-d]Pyrimidines. J. Saudi Chem. Soc. 2016, 20, 570–576. [Google Scholar] [CrossRef]
  238. Polo, E.; Arce-Parada, V.; López-Cortés, X.A.; Sánchez-Márquez, J.; Morales-Bayuelo, A.; Forero-Doria, O.; Gutiérrez, M. Synthesis of Pyrazolo-Fused 4-Azafluorenones in an Ionic Liquid. Mechanistic Insights by Joint Studies Using DFT Analysis and Mass Spectrometry. Catalysts 2019, 9, 820. [Google Scholar] [CrossRef]
  239. Sheldon, R. Catalytic Reactions in Ionic Liquids. Chem. Commun. 2001, 23, 2399–2407. [Google Scholar] [CrossRef]
  240. Olivier-Bourbigou, H.; Magna, L. Ionic Liquids: Perspectives for Organic and Catalytic Reactions. J. Mol. Catal. Chem. 2002, 182–183, 419–437. [Google Scholar] [CrossRef]
  241. Wasserscheid, P.; Keim, W. Ionic Liquids—New “Solutions” for Transition Metal Catalysis. Angew. Chem. Int. Ed. 2000, 39, 3772–3789. [Google Scholar] [CrossRef]
  242. Gu, Y. Multicomponent Reactions in Unconventional Solvents: State of the Art. Green Chem. 2012, 14, 2091–2128. [Google Scholar] [CrossRef]
  243. Ahmed, K.; Dubey, B.; Nadeem, S.; Shrivastava, B.; Sharma, P. P-TSA-Catalyzed One-Pot Synthesis and Docking Studies of Some 5H-Indeno[1,2-b]Quinoline-9,11(6H,10H)-Dione Derivatives as Anticonvulsant Agents. Chin. Chem. Lett. 2016, 27, 721–725. [Google Scholar] [CrossRef]
  244. Verma, G.K.; Raghuvanshi, K.; Kumar, R.; Singh, M.S. An Efficient One-Pot Three-Component Synthesis of Functionalized Pyrimido[4,5-b]Quinolines and Indeno Fused Pyrido[2,3-d]Pyrimidines in Water. Tetrahedron Lett. 2012, 53, 399–402. [Google Scholar] [CrossRef]
  245. Bhaskaruni, S.V.H.S.; Maddila, S.; Van Zyl, W.E.; Jonnalagadda, S.B. Ag2O on ZrO2 as a Recyclable Catalyst for Multicomponent Synthesis of Indenopyrimidine Derivatives. Molecules 2018, 23, 1648. [Google Scholar] [CrossRef]
  246. Chupakhin, E.; Babich, O.; Prosekov, A.; Asyakina, L.; Krasavin, M. Spirocyclic Motifs in Natural Products. Molecules 2019, 24, 4165. [Google Scholar] [CrossRef] [Green Version]
  247. Donthi, R.; Reddy, V.R.; Reddy, S.N.; Chandra, R. Base Catalysed Diastereoselective Tamura Cycloaddition of Vinylidene Indanediones. Tetrahedron Lett. 2019, 60, 1–4. [Google Scholar] [CrossRef]
  248. Andrew Evans, P.; Brandt, T.A. Palladium Catalyzed Cross-Coupling Acylation Approach to the Antitumor Antibiotic Fredericamycin A. Tetrahedron Lett. 1996, 37, 1367–1370. [Google Scholar] [CrossRef]
  249. Pizzirani, D.; Roberti, M.; Recanatini, M. Domino Knoevenagel/Diels–Alder Sequence Coupled to Suzuki Reaction: A Valuable Synthetic Platform for Chemical Biology. Tetrahedron Lett. 2007, 48, 7120–7124. [Google Scholar] [CrossRef]
  250. Coldham, I.; Hufton, R. Intramolecular Dipolar Cycloaddition Reactions of Azomethine Ylides. Chem. Rev. 2005, 105, 2765–2810. [Google Scholar] [CrossRef]
  251. Kathiravan, S.; Raghunathan, R. Expedient Synthesis of Novel Ferrocenyl Spiropyrrolidines Through 1,3-Dipolar Cycloaddition Reaction. J. Heterocycl. Chem. 2014, 51, 906–910. [Google Scholar] [CrossRef]
  252. Huang, Y.; Sun, J.; Yan, C.-G. Generation of New 1,3-Dipolar Azomethine Ylide via Reaction of Ethyl Glycinate with Dialkyl But-2-Ynedioate and Tandem 1,3-Dipolar Cycloaddition Reaction. ChemistrySelect 2017, 2, 10496–10500. [Google Scholar] [CrossRef]
  253. Qiu, G.; Kuang, Y.; Wu, J. N-Imide Ylide-Based Reactions: C-H Functionalization, Nucleophilic Addition and Cycloaddition. Adv. Synth. Catal. 2014, 356, 3483–3504. [Google Scholar] [CrossRef]
  254. Duan, J.; Cheng, J.; Cheng, Y.; Li, P. Synthesis of Dinitrogen-Fused Spirocyclic Heterocycles via Organocatalytic 1,3-Dipolar Cycloaddition of 2-Arylidene-1,3-Indandiones and an Azomethine Imine. Asian J. Org. Chem. 2016, 5, 477–480. [Google Scholar] [CrossRef]
  255. Yu, J.-K.; Chien, H.-W.; Lin, Y.-J.; Karanam, P.; Chen, Y.-H.; Lin, W. Diversity-Oriented Synthesis of Chromenopyrrolidines from Azomethine Ylides and 2-Hydroxybenzylidene Indandiones via Base-Controlled Regiodivergent (3+2) Cycloaddition. Chem. Commun. 2018, 54, 9921–9924. [Google Scholar] [CrossRef]
  256. Winter, M.; Faust, K.; Himmelsbach, M.; Waser, M. Synthesis of α-CF3-Proline Derivatives by Means of a Formal (3 + 2)-Cyclisation between Trifluoropyruvate Imines and Michael Acceptors. Org. Biomol. Chem. 2019, 17, 5731–5735. [Google Scholar] [CrossRef] [Green Version]
  257. Borges, F.; Roleira, F.; Milhazes, N.; Santana, L.; Uriarte, E. Simple Coumarins and Analogues in Medicinal Chemistry: Occurrence, Synthesis and Biological Activity. Curr. Med. Chem. 2005, 12, 887–916. [Google Scholar] [CrossRef]
  258. Chen, Y.-R.; Ganapuram, M.R.; Hsieh, K.-H.; Chen, K.-H.; Karanam, P.; Vagh, S.S.; Liou, Y.-C.; Lin, W. 3-Homoacyl Coumarin: An All Carbon 1,3-Dipole for Enantioselective Concerted (3+2) Cycloaddition. Chem. Commun. 2018, 54, 12702–12705. [Google Scholar] [CrossRef]
  259. Hu, F.; Wei, Y.; Shi, M. Enantioselective Synthesis of Spirocyclic Cyclopentenes: Asymmetric [3+2] Annulation of 2-Arylideneindane-1,3-Diones with MBH Carbonates Derivatives Catalyzed by Multifunctional Thiourea–Phosphines. Tetrahedron 2012, 68, 7911–7919. [Google Scholar] [CrossRef]
  260. Vetica, F.; Bailey, S.J.; Kumar, M.; Mahajan, S.; von Essen, C.; Rissanen, K.; Enders, D. Palladium-Catalyzed [3+2] Cycloaddition of Vinylaziridine and Indane-1,3-Diones: Diastereo- and Enantioselective Access to Spiro-Pyrrolidines. Synthesis 2020, 52, 2038–2044. [Google Scholar] [CrossRef]
  261. Zhang, H.; Gao, X.; Jiang, F.; Shi, W.; Wang, W.; Wu, Y.; Zhang, C.; Shi, X.; Guo, H. Palladium-Catalyzed Asymmetric [3+2] Cycloaddition of Vinylethylene Carbonates with 2-Arylidene-1,3-Indandiones: Synthesis of Tetrahydrofuran-Fused Spirocyclic 1,3-Indandiones. Eur. J. Org. Chem. 2020, 2020, 4801–4804. [Google Scholar] [CrossRef]
  262. Wei, F.; Ren, C.-L.; Wang, D.; Liu, L. Highly Enantioselective [3+2] Cycloaddition of Vinylcyclopropane with Nitroalkenes Catalyzed by Palladium(0) with a Chiral Bis(Tert-Amine) Ligand. Chem. Eur. J. 2015, 21, 2335–2338. [Google Scholar] [CrossRef] [PubMed]
  263. Ling, J.; Laugeois, M.; Ratovelomanana-Vidal, V.; Vitale, M.R. Palladium(0)-Catalyzed Diastereoselective (3+2) Cycloadditions of Vinylcyclopropanes with Sulfonyl-Activated Imines. Synlett 2018, 29, 2288–2292. [Google Scholar]
  264. Laugeois, M.; Ponra, S.; Ratovelomanana-Vidal, V.; Michelet, V.; Vitale, M.R. Asymmetric Preparation of Polysubstituted Cyclopentanes by Synergistic Pd(0)/Amine Catalyzed Formal [3+2] Cycloadditions of Vinyl Cyclopropanes with Enals. Chem. Commun. 2016, 52, 5332–5335. [Google Scholar] [CrossRef] [PubMed]
  265. Mei, L.-Y.; Tang, X.-Y.; Shi, M. One-Pot Tandem Diastereoselective and Enantioselective Synthesis of Functionalized Oxindole-Fused Spiropyrazolidine Frameworks. Chem. Eur. J. 2014, 20, 13136–13142. [Google Scholar] [CrossRef]
  266. Yuan, Z.; Wei, W.; Lin, A.; Yao, H. Bifunctional Organo/Metal Cooperatively Catalyzed [3 + 2] Annulation of Para-Quinone Methides with Vinylcyclopropanes: Approach to Spiro[4.5]Deca-6,9-Diene-8-Ones. Org. Lett. 2016, 18, 3370–3373. [Google Scholar] [CrossRef]
  267. Corti, V.; Marcantonio, E.; Mamone, M.; Giungi, A.; Fochi, M.; Bernardi, L. Synergistic Palladium-Phosphoric Acid Catalysis in (3 + 2) Cycloaddition Reactions between Vinylcyclopropanes and Imines. Catalysts 2020, 10, 150. [Google Scholar] [CrossRef] [Green Version]
  268. Duan, J.; Cheng, Y.; Li, R.; Li, P. Synthesis of Spiro[Indane-1,3-Dione-1-Pyrrolines] via Copper-Catalyzed Heteroannulation of Ketoxime Acetates with 2-Arylideneindane-1,3-Diones. Org. Chem. Front. 2016, 3, 1614–1618. [Google Scholar] [CrossRef]
  269. Bdiri, B.; Zhou, Z.-M. Novel Asymmetric Synthesis of Spiroindene-1,3dione-Pyrrolidines via CoII/Amino Acids Complex Catalysed Asymmetric 1,3-Dipolar Cycloaddition of Azomethine Ylides and 2-Arylidenindane-1,3-Diones. Tetrahedron Lett. 2017, 58, 4600–4608. [Google Scholar] [CrossRef]
  270. Jiang, Y.-H.; Sun, J.; Sun, Q.; Yan, C.-G. Construction of Spiro[Indene-2,1′-Pyrrolo[2,1-a]Isoquinoline]s through a Visible-Light-Catalyzed Oxidative [3+2] Cycloaddition Reaction. Asian J. Org. Chem. 2017, 6, 862–866. [Google Scholar] [CrossRef]
  271. Domínguez, G.; Pérez-Castells, J. Recent Advances in [2+2+2] Cycloaddition Reactions. Chem. Soc. Rev. 2011, 40, 3430–3444. [Google Scholar] [CrossRef] [PubMed]
  272. Kotha, S.; Sreevani, G. Molybdenum Hexacarbonyl: Air Stable Catalyst for Microwave Assisted Intermolecular [2+2+2] Co-Trimerization Involving Propargyl Halides. Tetrahedron Lett. 2015, 56, 5903–5908. [Google Scholar] [CrossRef]
  273. Tran, C.; Haddad, M.; Ratovelomanana-Vidal, V. Ruthenium-Catalyzed [2+2+2] Cycloaddition of α,ω-Diynes and Selenocyanates: An Entry to Selenopyridine Derivatives. Synthesis 2019, 51, 2532–2541. [Google Scholar] [CrossRef]
  274. Ye, F.; Boukattaya, F.; Haddad, M.; Ratovelomanana-Vidal, V.; Michelet, V. Synthesis of 2-Aminopyridines via Ruthenium-Catalyzed [2+2+2] Cycloaddition of 1,6- and 1,7-Diynes with Cyanamides: Scope and Limitations. New J. Chem. 2018, 42, 3222–3235. [Google Scholar] [CrossRef]
  275. Ye, F.; Haddad, M.; Michelet, V.; Ratovelomanana-Vidal, V. Solvent-Free Ruthenium Trichloride-Mediated [2 + 2 + 2] Cycloaddition of α,ω-Diynes and Cyanamides: A Convenient Access to 2-Aminopyridines. Org. Chem. Front. 2017, 4, 1063–1068. [Google Scholar] [CrossRef]
  276. Tran, C.; Haddad, M.; Ratovelomanana-Vidal, V. Ruthenium-Catalyzed, Microwave-Mediated [2+2+2] Cycloaddition: A Useful Combination for the Synthesis of 2-Aminopyridines. Synlett 2019, 30, 1891–1894. [Google Scholar] [CrossRef]
  277. Ye, F.; Haddad, M.; Ratovelomanana-Vidal, V.; Michelet, V. Ruthenium-Catalyzed [2 + 2 + 2] Cycloaddition Reaction Forming 2-Aminopyridine Derivatives from α,ω-Diynes and Cyanamides. Org. Lett. 2017, 19, 1104–1107. [Google Scholar] [CrossRef]
  278. Li, Q.; Wang, Y.; Li, B.; Wang, B. Cp*Co(III)-Catalyzed Regioselective Synthesis of Cyclopenta[b]Carbazoles via Dual C(Sp2)–H Functionalization of 1-(Pyridin-2-Yl)-Indoles with Diynes. Org. Lett. 2018, 20, 7884–7887. [Google Scholar] [CrossRef]
  279. Wang, Y.; Li, B.; Wang, B. RhIII-Catalyzed Synthesis of Cyclopenta[b]Carbazoles via Cascade C-H/C-C Bond Cleavage and Cyclization Reactions: Using Amide as a Traceless Directing Group. Org. Lett. 2020, 22, 83–87. [Google Scholar] [CrossRef]
  280. Avarvari, N.; Le Floch, P.; Ricard, L.; Mathey, F. 1,3,2-Diazaphosphinines and -Diazaarsinines as Precursors for Polyfunctional Phosphinines and Arsinines. Organometallics 1997, 16, 4089–4098. [Google Scholar] [CrossRef]
  281. Weemers, J.J.M.; van der Graaff, W.N.P.; Pidko, E.A.; Lutz, M.; Müller, C. Bulky Phosphinines: From a Molecular Design to an Application in Homogeneous Catalysis. Chem. Eur. J. 2013, 19, 8991–9004. [Google Scholar] [CrossRef] [PubMed]
  282. Nakajima, K.; Takata, S.; Sakata, K.; Nishibayashi, Y. Synthesis of Phosphabenzenes by an Iron-Catalyzed [2+2+2] Cycloaddition Reaction of Diynes with Phosphaalkynes. Angew. Chem. Int. Ed. 2015, 54, 7597–7601. [Google Scholar] [CrossRef] [PubMed]
  283. Kumar, P.; Zhang, K.; Louie, J. An Expeditious Route to Eight-Membered Heterocycles By Nickel-Catalyzed Cycloaddition: Low-Temperature C-C Bond Cleavage. Angew. Chem. Int. Ed. 2012, 51, 8602–8606. [Google Scholar] [CrossRef] [PubMed]
  284. Wang, J.; Wang, M.; Xiang, J.; Xi, H.; Wu, A. Generation of O-Quinodimethanes (o-QDMs) from Benzo[c]Oxepines and the Synthetic Application for Polysubstituted Tetrahydronaphthalenes. Tetrahedron 2015, 71, 7687–7694. [Google Scholar] [CrossRef]
  285. Yang, Y.; Jiang, Y.; Du, W.; Chen, Y.-C. Asymmetric Cross [10+2] Cycloadditions of 2-Alkylidene-1-Indanones and Activated Alkenes under Phase-Transfer Catalysis. Chem. Eur. J. 2020, 26, 1754–1758. [Google Scholar] [CrossRef]
  286. Ramachary, D.B.; Anebouselvy, K.; Chowdari, N.S.; Barbas, C.F. Direct Organocatalytic Asymmetric Heterodomino Reactions:  The Knoevenagel/Diels−Alder/Epimerization Sequence for the Highly Diastereoselective Synthesis of Symmetrical and Nonsymmetrical Synthons of Benzoannelated Centropolyquinanes. J. Org. Chem. 2004, 69, 5838–5849. [Google Scholar] [CrossRef]
  287. Duan, J.; Cheng, J.; Li, B.; Qi, F.; Li, P. Enantioselective Synthesis of Spiro[1,3-indanedione–Tetrahydrothiophene]s by Organocatalytic Sulfa-Michael/Michael Domino Reaction. Eur. J. Org. Chem. 2015, 2015, 6130–6134. [Google Scholar] [CrossRef]
  288. Mahajan, S.; Chauhan, P.; Blümel, M.; Puttreddy, R.; Rissanen, K.; Raabe, G.; Enders, D. Asymmetric Synthesis of Spiro Tetrahydrothiophene-Indan-1,3-Diones via a Squaramide-Catalyzed Sulfa-Michael/Aldol Domino Reaction. Synthesis 2016, 48, 1131–1138. [Google Scholar] [CrossRef]
  289. Yazdani, H.; Bazgir, A. Lewis Acid Catalyzed Regio- and Diastereoselective Synthesis of Spiroisoxazolines via One-Pot Sequential Knoevenagel Condensation/1,3-Dipolar Cycloaddition Reaction. Synthesis 2019, 51, 1669–1679. [Google Scholar]
  290. Aitha, A.; Yennam, S.; Behera, M.; Anireddy, J.S. Synthesis of Spiroindene-1,3-Dione Isothiazolines via a Cascade Michael/1,3-Dipolar Cycloaddition Reaction of 1,3,4-Oxathiazol-2-One and 2-Arylidene-1,3-Indandiones. Tetrahedron Lett. 2017, 58, 578–581. [Google Scholar] [CrossRef]
  291. Liang, L.; Li, E.; Xie, P.; Huang, Y. Phosphine-Initiated Domino Reaction: A Convenient Method for the Preparation of Spirocyclopentanones. Chem. Asian J. 2014, 9, 1270–1273. [Google Scholar] [CrossRef] [PubMed]
  292. Mao, L.; Li, Y.; Yang, S. Silver-Catalyzed Cascade Radical Cyclization for Stereoselective Synthesis of Exocyclic Phosphine Oxides. Chin. J. Chem. 2017, 35, 316–322. [Google Scholar] [CrossRef]
  293. Shi, W.; Mao, B.; Xu, J.; Wang, Q.; Wang, W.; Wu, Y.; Li, X.; Guo, H. Phosphine-Catalyzed Cascade Michael Addition/[4+2] Cycloaddition Reaction of Allenoates and 2-Arylidene-1,3-Indanediones. Org. Lett. 2020, 22, 2675–2680. [Google Scholar] [CrossRef] [PubMed]
  294. Champetter, P.; Castillo-Aguilera, O.; Taillier, C.; Brière, J.-F.; Dalla, V.; Oudeyer, S.; Comesse, S. N-Alkoxyacrylamides in Domino Reactions: Catalytic and Stereoselective Access to δ-Lactams. Eur. J. Org. Chem. 2019, 2019, 7703–7710. [Google Scholar] [CrossRef]
  295. Yang, S.-M.; Tsai, Y.-L.; Reddy, G.M.; Möhlmann, L.; Lin, W. Chemo- and Diastereoselective Michael–Michael-Acetalization Cascade for the Synthesis of 1,3-Indandione-Fused Spiro[4.5]Decan Scaffolds. J. Org. Chem. 2017, 82, 9182–9190. [Google Scholar] [CrossRef]
  296. Ren, W.; Wang, X.-Y.; Li, J.-J.; Tian, M.; Liu, J.; Ouyang, L.; Wang, J.-H. Efficient Construction of Biologically Important Functionalized Polycyclic Spiro-Fused Carbocyclicoxindoles via an Asymmetric Organocatalytic Quadruple-Cascade Reaction. RSC Adv. 2017, 7, 1863–1868. [Google Scholar] [CrossRef]
  297. Banothu, J.; Basavoju, S.; Bavantula, R. Pyridinium Ylide Assisted Highly Stereoselective One-Pot Synthesis of Trans-2-(4-Chlorobenzoyl)-3-Aryl-Spiro[Cyclopropane-1,2′-Inden]-1′,3′-Diones and Their Antimicrobial and Nematicidal Activities. J. Heterocycl. Chem. 2015, 52, 853–860. [Google Scholar] [CrossRef]
  298. Nassiri, M.; Milani, F.J.; Hassankhani, A. Synthesis of Spiro Pyrrolobenzothiazole Derivatives via a Three-Component Reaction. J. Heterocycl. Chem. 2015, 52, 1162–1166. [Google Scholar] [CrossRef]
  299. Manjappa, K.B.; Peng, Y.-T.; Jhang, W.-F.; Yang, D.-Y. Microwave-Promoted, Catalyst-Free, Multi-Component Reaction of Proline, Aldehyde, 1,3-Diketone: One Pot Synthesis of Pyrrolizidines and Pyrrolizinones. Tetrahedron 2016, 72, 853–861. [Google Scholar] [CrossRef]
  300. Rajeswari, M.; Sindhu, J.; Singh, H.; Khurana, J.M. An Efficient, Green Synthesis of Novel Regioselective and Stereoselective Indan-1,3-Dione Grafted Spirooxindolopyrrolizidine Linked 1,2,3-Triazoles via a One-Pot Five-Component Condensation Using PEG-400. RSC Adv. 2015, 5, 39686–39691. [Google Scholar] [CrossRef]
  301. Firouzi-Haji, R.; Maleki, A. L-Proline-Functionalized Fe3O4 Nanoparticles as an Efficient Nanomagnetic Organocatalyst for Highly Stereoselective One-Pot Two-Step Tandem Synthesis of Substituted Cyclopropanes. ChemistrySelect 2019, 4, 853–857. [Google Scholar] [CrossRef]
  302. Zhou, H.-Y.; Han, Y.; Chen, C.-F. PH-Controlled Motions in Mechanically Interlocked Molecules. Mater. Chem. Front. 2020, 4, 12–28. [Google Scholar] [CrossRef]
  303. Kitson, P.J.; Parenty, A.D.C.; Richmond, C.J.; Long, D.-L.; Cronin, L. A New C-C Bond Forming Annulation Reaction Leading to PH Switchable Heterocycles. Chem. Commun. 2009, 27, 4067–4069. [Google Scholar] [CrossRef]
  304. Qi, J.; Zheng, J.; Cui, S. Facile Synthesis of Carbo- and Heterocycles via Fe(Iii)-Catalyzed Alkene Hydrofunctionalization. Org. Chem. Front. 2018, 5, 222–225. [Google Scholar] [CrossRef]
  305. Duan, J.; Mao, Y.; Zhang, L.; Zhu, N.; Fang, Z.; Guo, K. Copper-Catalyzed [3+2] Annulation of 2-Arylidene-1,3-Indandiones with N-Acetyl Enamides for the Synthesis of Spiropyrrolines. Adv. Synth. Catal. 2020, 362, 695–699. [Google Scholar] [CrossRef]
  306. Zhao, B.; Liang, H.-W.; Yang, J.; Yang, Z.; Wei, Y. Copper-Catalyzed Intermolecular Cyclization between Oximes and Alkenes: A Facile Access to Spiropyrrolines. ACS Catal. 2017, 7, 5612–5617. [Google Scholar] [CrossRef]
  307. Gupta, A.K.; Vaishanv, N.K.; Kant, R.; Mohanan, K. Rapid and Selective Synthesis of Spiropyrazolines and Pyrazolylphthalides Employing Seyferth–Gilbert Reagent. Org. Biomol. Chem. 2017, 15, 6411–6415. [Google Scholar] [CrossRef]
  308. Das, S. Annulations Involving 2-Arylidene-1,3-Indanediones: Stereoselective Synthesis of Spiro- and Fused Scaffolds. New J. Chem. 2020, 44, 17148–17176. [Google Scholar] [CrossRef]
  309. Xiao, P.; Zhang, J.; Dumur, F.; Tehfe, M.A.; Morlet-Savary, F.; Graff, B.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. Visible Light Sensitive Photoinitiating Systems: Recent Progress in Cationic and Radical Photopolymerization Reactions under Soft Conditions. Prog. Polym. Sci. 2015, 41, 32–66. [Google Scholar] [CrossRef]
  310. Al Mousawi, A.; Garra, P.; Schmitt, M.; Toufaily, J.; Hamieh, T.; Graff, B.; Fouassier, J.P.; Dumur, F.; Lalevée, J. 3-Hydroxyflavone and N-Phenylglycine in High Performance Photoinitiating Systems for 3D Printing and Photocomposites Synthesis. Macromolecules 2018, 51, 4633–4641. [Google Scholar] [CrossRef]
  311. Tehfe, M.-A.; Dumur, F.; Xiao, P.; Delgove, M.; Graff, B.; Fouassier, J.-P.; Gigmes, D.; Lalevée, J. Chalcone Derivatives as Highly Versatile Photoinitiators for Radical, Cationic, Thiol–Ene and IPN Polymerization Reactions upon Exposure to Visible Light. Polym. Chem. 2014, 5, 382–390. [Google Scholar] [CrossRef]
  312. Zivic, N.; Zhang, J.; Bardelang, D.; Dumur, F.; Xiao, P.; Jet, T.; Versace, D.-L.; Dietlin, C.; Morlet-Savary, F.; Graff, B.; et al. Novel Naphthalimide–Amine Based Photoinitiators Operating under Violet and Blue LEDs and Usable for Various Polymerization Reactions and Synthesis of Hydrogels. Polym. Chem. 2015, 7, 418–429. [Google Scholar] [CrossRef]
  313. Tehfe, M.-A.; Dumur, F.; Contal, E.; Graff, B.; Morlet-Savary, F.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. New Insights into Radical and Cationic Polymerizations upon Visible Light Exposure: Role of Novel Photoinitiator Systems Based on the Pyrene Chromophore. Polym. Chem. 2013, 4, 1625–1634. [Google Scholar] [CrossRef]
  314. Zhang, J.; Zivic, N.; Dumur, F.; Xiao, P.; Graff, B.; Fouassier, J.P.; Gigmes, D.; Lalevée, J. N-[2-(Dimethylamino)Ethyl]-1,8-Naphthalimide Derivatives as Photoinitiators under LEDs. Polym. Chem. 2018, 9, 994–1003. [Google Scholar] [CrossRef]
  315. Tehfe, M.-A.; Dumur, F.; Contal, E.; Graff, B.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. Novel Highly Efficient Organophotocatalysts: Truxene–Acridine-1,8-Diones as Photoinitiators of Polymerization. Macromol. Chem. Phys. 2013, 214, 2189–2201. [Google Scholar] [CrossRef]
  316. Dumur, F. Recent Advances on Visible Light Metal-Based Photocatalysts for Polymerization under Low Light Intensity. Catalysts 2019, 9, 736. [Google Scholar] [CrossRef]
  317. Bonardi, A.-H.; Dumur, F.; Noirbent, G.; Lalevée, J.; Gigmes, D. Organometallic vs Organic Photoredox Catalysts for Photocuring Reactions in the Visible Region. Beilstein J. Org. Chem. 2018, 14, 3025–3046. [Google Scholar] [CrossRef]
  318. Zivic, N.; Bouzrati-Zerelli, M.; Kermagoret, A.; Dumur, F.; Fouassier, J.-P.; Gigmes, D.; Lalevée, J. Photocatalysts in Polymerization Reactions. ChemCatChem 2016, 8, 1617–1631. [Google Scholar] [CrossRef]
  319. Lalevée, J.; Telitel, S.; Xiao, P.; Lepeltier, M.; Dumur, F.; Morlet-Savary, F.; Gigmes, D.; Fouassier, J.-P. Metal and Metal-Free Photocatalysts: Mechanistic Approach and Application as Photoinitiators of Photopolymerization. Beilstein J. Org. Chem. 2014, 10, 863–876. [Google Scholar] [CrossRef]
  320. Sun, K.; Pigot, C.; Zhang, Y.; Borjigin, T.; Morlet-Savary, F.; Graff, B.; Nechab, M.; Xiao, P.; Dumur, F.; Lalevée, J. Sunlight Induced Polymerization Photoinitiated by Novel Push–Pull Dyes: Indane-1,3-Dione, 1H-Cyclopenta[b]Naphthalene-1,3(2H)-Dione and 4-Dimethoxyphenyl-1-Allylidene Derivatives. Macromol. Chem. Phys. 2022, 223, 2100439. [Google Scholar] [CrossRef]
  321. Sun, K.; Liu, S.; Chen, H.; Morlet-Savary, F.; Graff, B.; Pigot, C.; Nechab, M.; Xiao, P.; Dumur, F.; Lalevée, J. N-Ethyl Carbazole-1-Allylidene-Based Push-Pull Dyes as Efficient Light Harvesting Photoinitiators for Sunlight Induced Polymerization. Eur. Polym. J. 2021, 147, 110331. [Google Scholar] [CrossRef]
  322. Sun, K.; Liu, S.; Pigot, C.; Brunel, D.; Graff, B.; Nechab, M.; Gigmes, D.; Morlet-Savary, F.; Zhang, Y.; Xiao, P.; et al. Novel Push–Pull Dyes Derived from 1H-Cyclopenta[b]Naphthalene-1,3(2H)-Dione as Versatile Photoinitiators for Photopolymerization and Their Related Applications: 3D Printing and Fabrication of Photocomposites. Catalysts 2020, 10, 1196. [Google Scholar] [CrossRef]
  323. Pigot, C.; Noirbent, G.; Brunel, D.; Dumur, F. Recent Advances on Push–Pull Organic Dyes as Visible Light Photoinitiators of Polymerization. Eur. Polym. J. 2020, 133, 109797. [Google Scholar] [CrossRef]
  324. Mokbel, H.; Toufaily, J.; Hamieh, T.; Dumur, F.; Campolo, D.; Gigmes, D.; Fouassier, J.P.; Ortyl, J.; Lalevée, J. Specific Cationic Photoinitiators for near UV and Visible LEDs: Iodonium versus Ferrocenium Structures. J. Appl. Polym. Sci. 2015, 132, 42759. [Google Scholar] [CrossRef]
  325. Dumur, F.; Nasr, G.; Wantz, G.; Mayer, C.R.; Dumas, E.; Guerlin, A.; Miomandre, F.; Clavier, G.; Bertin, D.; Gigmes, D. Cationic Iridium Complex for the Design of Soft Salt-Based Phosphorescent OLEDs and Color-Tunable Light-Emitting Electrochemical Cells. Org. Electron. 2011, 12, 1683–1694. [Google Scholar] [CrossRef]
  326. Zhang, J.; Zivic, N.; Dumur, F.; Xiao, P.; Graff, B.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. A Benzophenone-Naphthalimide Derivative as Versatile Photoinitiator of Polymerization under near UV and Visible Lights. J. Polym. Sci. Part Polym. Chem. 2015, 53, 445–451. [Google Scholar] [CrossRef]
  327. Xiao, P.; Frigoli, M.; Dumur, F.; Graff, B.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. Julolidine or Fluorenone Based Push–Pull Dyes for Polymerization upon Soft Polychromatic Visible Light or Green Light. Macromolecules 2014, 47, 106–112. [Google Scholar] [CrossRef]
  328. Tehfe, M.-A.; Dumur, F.; Graff, B.; Morlet-Savary, F.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. Push–Pull (Thio)Barbituric Acid Derivatives in Dye Photosensitized Radical and Cationic Polymerization Reactions under 457/473 Nm Laser Beams or Blue LEDs. Polym. Chem. 2013, 4, 3866–3875. [Google Scholar] [CrossRef]
  329. Dumur, F.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. Organic Electronics: An El Dorado in the Quest of New Photocatalysts for Polymerization Reactions. Acc. Chem. Res. 2016, 49, 1980–1989. [Google Scholar] [CrossRef]
  330. Mokbel, H.; Dumur, F.; Telitel, S.; Vidal, L.; Xiao, P.; Versace, D.-L.; Tehfe, M.-A.; Morlet-Savary, F.; Graff, B.; Fouassier, J.-P.; et al. Photoinitiating Systems of Polymerization and in Situ Incorporation of Metal Nanoparticles into Polymer Matrices upon Exposure to Visible Light: Push–Pull Malonate and Malononitrile Based Dyes. Polym. Chem. 2013, 4, 5679–5687. [Google Scholar] [CrossRef]
  331. Tehfe, M.-A.; Dumur, F.; Graff, B.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. Blue-to-Red Light Sensitive Push–Pull Structured Photoinitiators: Indanedione Derivatives for Radical and Cationic Photopolymerization Reactions. Macromolecules 2013, 46, 3332–3341. [Google Scholar] [CrossRef]
  332. Xiao, P.; Dumur, F.; Graff, B.; Morlet-Savary, F.; Vidal, L.; Gigmes, D.; Fouassier, J.P.; Lalevée, J. Structural Effects in the Indanedione Skeleton for the Design of Low Intensity 300–500 Nm Light Sensitive Initiators. Macromolecules 2014, 47, 26–34. [Google Scholar] [CrossRef]
  333. Tehfe, M.-A.; Dumur, F.; Graff, B.; Morlet-Savary, F.; Fouassier, J.-P.; Gigmes, D.; Lalevée, J. New Push–Pull Dyes Derived from Michler’s Ketone For Polymerization Reactions Upon Visible Lights. Macromolecules 2013, 46, 3761–3770. [Google Scholar] [CrossRef]
  334. Telitel, S.; Dumur, F.; Kavalli, T.; Graff, B.; Morlet-Savary, F.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. The 1,3-Bis(Dicyanomethylidene)Indane Skeleton as a (Photo) Initiator in Thermal Ring Opening Polymerization at RT and Radical or Cationic Photopolymerization. RSC Adv. 2014, 4, 15930–15936. [Google Scholar] [CrossRef]
  335. Sun, K.; Pigot, C.; Chen, H.; Nechab, M.; Gigmes, D.; Morlet-Savary, F.; Graff, B.; Liu, S.; Xiao, P.; Dumur, F.; et al. Free Radical Photopolymerization and 3D Printing Using Newly Developed Dyes: Indane-1,3-Dione and 1H-Cyclopentanaphthalene-1,3-Dione Derivatives as Photoinitiators in Three-Component Systems. Catalysts 2020, 10, 463. [Google Scholar] [CrossRef]
  336. Bakshiev, N.G. Universal Intermolecular Interactions and Their Effect on the Position of the Electronic Spectra of Molecules in Two Component Solutions. Opt. Spektrosk. 1964, 16, 821–832. [Google Scholar]
  337. Kawski, A. Der Wellenzahl von Elecktronenbanden Lumineszierenden Molecule. Acta Phys Pol. 1966, 29, 507–518. [Google Scholar]
  338. Lippert, E. Dipolmoment Und Elektronenstruktur von Angeregten Molekülen. Z. Für Naturforschung A 1955, 10, 541–545. [Google Scholar] [CrossRef]
  339. McRae, E.G. Theory of Solvent Effects on Molecular Electronic Spectra. Frequency Shifts. J. Phys. Chem. 1957, 61, 562–572. [Google Scholar] [CrossRef]
  340. Suppan, P. Solvent Effects on the Energy of Electronic Transitions: Experimental Observations and Applications to Structural Problems of Excited Molecules. J. Chem. Soc. Inorg. Phys. Theor. 1968, 3125–3133. [Google Scholar] [CrossRef]
  341. Steyrer, B.; Neubauer, P.; Liska, R.; Stampfl, J. Visible Light Photoinitiator for 3D-Printing of Tough Methacrylate Resins. Materials 2017, 10, 1445. [Google Scholar] [CrossRef] [PubMed]
  342. Quan, H.; Zhang, T.; Xu, H.; Luo, S.; Nie, J.; Zhu, X. Photo-Curing 3D Printing Technique and Its Challenges. Bioact. Mater. 2020, 5, 110–115. [Google Scholar] [CrossRef] [PubMed]
  343. Lai, H.; Zhu, D.; Xiao, P. Yellow Triazine as an Efficient Photoinitiator for Polymerization and 3D Printing under LEDs. Macromol. Chem. Phys. 2019, 220, 1900315. [Google Scholar] [CrossRef]
  344. Mitterbauer, M.; Knaack, P.; Naumov, S.; Markovic, M.; Ovsianikov, A.; Moszner, N.; Liska, R. Acylstannanes: Cleavable and Highly Reactive Photoinitiators for Radical Photopolymerization at Wavelengths above 500 Nm with Excellent Photobleaching Behavior. Angew. Chem. Int. Ed. 2018, 57, 12146–12150. [Google Scholar] [CrossRef]
  345. Bouzrati-Zerelli, M.; Zivic, N.; Dumur, F.; Gigmes, D.; Graff, B.; Fouassier, J.P.; Lalevée, J. New Violet to Yellow Light Sensitive Diketo Pyrrolo–Pyrrole Photoinitiators: High Performance Systems with Unusual Bleaching Properties and Solubility in Water. Polym. Chem. 2017, 8, 2028–2040. [Google Scholar] [CrossRef]
  346. Li, J.; Hao, Y.; Zhong, M.; Tang, L.; Nie, J.; Zhu, X. Synthesis of Furan Derivative as LED Light Photoinitiator: One-Pot, Low Usage, Photobleaching for Light Color 3D Printing. Dyes Pigments 2019, 165, 467–473. [Google Scholar] [CrossRef]
  347. Dumur, F. Zinc Complexes in OLEDs: An Overview. Synth. Met. 2014, 195, 241–251. [Google Scholar] [CrossRef]
  348. Dumur, F.; Bertin, D.; Gigmes, D. Iridium (III) Complexes as Promising Emitters for Solid–State Light–Emitting Electrochemical Cells (LECs). Int. J. Nanotechnol. 2012, 9, 377–395. [Google Scholar] [CrossRef]
  349. Terenziani, F.; Mongin, O.; Katan, C.; Bhatthula, B.K.G.; Blanchard-Desce, M. Effects of Dipolar Interactions on Linear and Nonlinear Optical Properties of Multichromophore Assemblies: A Case Study. Chem. Eur. J. 2006, 12, 3089–3102. [Google Scholar] [CrossRef]
  350. Lou, A.J.-T.; Marks, T.J. A Twist on Nonlinear Optics: Understanding the Unique Response of π-Twisted Chromophores. Acc. Chem. Res. 2019, 52, 1428–1438. [Google Scholar] [CrossRef]
  351. Rutkis, M.; Tokmakovs, A.; Jecs, E.; Kreicberga, J.; Kampars, V.; Kokars, V. Indanedione Based Binary Chromophore Supramolecular Systems as a NLO Active Polymer Composites. Opt. Prop. Funct. Mater. Nanomater. 2010, 32, 796–802. [Google Scholar] [CrossRef]
  352. Acharya, S.; Krief, P.; Khodorkovsky, V.; Kotler, Z.; Berkovic, G.; Klug, J.T.; Efrima, S. Studies of Langmuir and Langmuir–Blodgett Films of NLO-Active Amphiphilic 1,3-Indanedione Derivatives. New J. Chem. 2005, 29, 1049–1057. [Google Scholar] [CrossRef]
  353. Schwartz, H.; Mazor, R.; Khodorkovsky, V.; Shapiro, L.; Klug, J.T.; Kovalev, E.; Meshulam, G.; Berkovic, G.; Kotler, Z.; Efrima, S. Langmuir and Langmuir−Blodgett Films of NLO Active 2-(p-N-Alkyl-N-Methylamino)Benzylidene-1,3-Indandioneπ/A Curves, UV−Vis Spectra, and SHG Behavior. J. Phys. Chem. B 2001, 105, 5914–5921. [Google Scholar] [CrossRef]
  354. Doddamani, R.V.; Tasaganva, R.G.; Inamdar, S.R.; Kariduraganavar, M.Y. Synthesis of Chromophores and Polyimides with a Green Chemistry Approach for Second-Order Nonlinear Optical Applications. Polym. Adv. Technol. 2018, 29, 2091–2102. [Google Scholar] [CrossRef]
  355. Wu, D.; Sedgwick, A.C.; Gunnlaugsson, T.; Akkaya, E.U.; Yoon, J.; James, T.D. Fluorescent Chemosensors: The Past, Present and Future. Chem. Soc. Rev. 2017, 46, 7105–7123. [Google Scholar] [CrossRef]
  356. Wu, X.; Xu, B.; Tong, H.; Wang, L. Highly Selective and Sensitive Detection of Cyanide by a Reaction-Based Conjugated Polymer Chemosensor. Macromolecules 2011, 44, 4241–4248. [Google Scholar] [CrossRef]
  357. Tomal, W.; Ortyl, J. Water-Soluble Photoinitiators in Biomedical Applications. Polymers 2020, 12, 1073. [Google Scholar] [CrossRef]
  358. Hu, J.-W.; Lin, W.-C.; Hsiao, S.-Y.; Wu, Y.-H.; Chen, H.-W.; Chen, K.-Y. An Indanedione-Based Chemodosimeter for Selective Naked-Eye and Fluorogenic Detection of Cyanide. Sens. Actuators B Chem. 2016, 233, 510–519. [Google Scholar] [CrossRef]
  359. Wang, L.; Du, J.; Cao, D. A Colorimetric and Fluorescent Probe Containing Diketopyrrolopyrrole and 1,3-Indanedione for Cyanide Detection Based on Exciplex Signaling Mechanism. Sens. Actuators B Chem. 2014, 198, 455–461. [Google Scholar] [CrossRef]
  360. Facchetti, A. π-Conjugated Polymers for Organic Electronics and Photovoltaic Cell Applications. Chem. Mater. 2011, 23, 733–758. [Google Scholar] [CrossRef]
  361. Bian, L.; Zhu, E.; Tang, J.; Tang, W.; Zhang, F. Recent Progress in the Design of Narrow Bandgap Conjugated Polymers for High-Efficiency Organic Solar Cells. Top. Issue Conduct. Polym. 2012, 37, 1292–1331. [Google Scholar] [CrossRef]
  362. Zhou, H.; Yang, L.; You, W. Rational Design of High Performance Conjugated Polymers for Organic Solar Cells. Macromolecules 2012, 45, 607–632. [Google Scholar] [CrossRef]
  363. Bloking, J.T.; Han, X.; Higgs, A.T.; Kastrop, J.P.; Pandey, L.; Norton, J.E.; Risko, C.; Chen, C.E.; Brédas, J.-L.; McGehee, M.D.; et al. Solution-Processed Organic Solar Cells with Power Conversion Efficiencies of 2.5% Using Benzothiadiazole/Imide-Based Acceptors. Chem. Mater. 2011, 23, 5484–5490. [Google Scholar] [CrossRef]
  364. Lin, Y.; Li, Y.; Zhan, X. A Solution-Processable Electron Acceptor Based on Dibenzosilole and Diketopyrrolopyrrole for Organic Solar Cells. Adv. Energy Mater. 2013, 3, 724–728. [Google Scholar] [CrossRef]
  365. Patil, H.; Gupta, A.; Bilic, A.; Bhosale, S.V.; Bhosale, S.V. A Solution-Processable Electron Acceptor Based on Diketopyrrolopyrrole and Naphthalenediimide Motifs for Organic Solar Cells. Tetrahedron Lett. 2014, 55, 4430–4432. [Google Scholar] [CrossRef]
  366. Lin, Y.; Wang, Y.; Wang, J.; Hou, J.; Li, Y.; Zhu, D.; Zhan, X. A Star-Shaped Perylene Diimide Electron Acceptor for High-Performance Organic Solar Cells. Adv. Mater. 2014, 26, 5137–5142. [Google Scholar] [CrossRef]
  367. Cheng, P.; Ye, L.; Zhao, X.; Hou, J.; Li, Y.; Zhan, X. Binary Additives Synergistically Boost the Efficiency of All-Polymer Solar Cells up to 3.45%. Energy Environ. Sci. 2014, 7, 1351–1356. [Google Scholar] [CrossRef]
  368. Patil, H.; Zu, W.X.; Gupta, A.; Chellappan, V.; Bilic, A.; Sonar, P.; Rananaware, A.; Bhosale, S.V.; Bhosale, S.V. A Non-Fullerene Electron Acceptor Based on Fluorene and Diketopyrrolopyrrole Building Blocks for Solution-Processable Organic Solar Cells with an Impressive Open-Circuit Voltage. Phys. Chem. Chem. Phys. 2014, 16, 23837–23842. [Google Scholar] [CrossRef]
  369. Raynor, A.M.; Gupta, A.; Patil, H.; Bilic, A.; Bhosale, S.V. A Diketopyrrolopyrrole and Benzothiadiazole Based Small Molecule Electron Acceptor: Design, Synthesis, Characterization and Photovoltaic Properties. RSC Adv. 2014, 4, 57635–57638. [Google Scholar] [CrossRef]
  370. Lin, Y.; Cheng, P.; Li, Y.; Zhan, X. A 3D Star-Shaped Non-Fullerene Acceptor for Solution-Processed Organic Solar Cells with a High Open-Circuit Voltage of 1.18 V. Chem. Commun. 2012, 48, 4773–4775. [Google Scholar] [CrossRef]
  371. Zhang, X.; Lu, Z.; Ye, L.; Zhan, C.; Hou, J.; Zhang, S.; Jiang, B.; Zhao, Y.; Huang, J.; Zhang, S.; et al. A Potential Perylene Diimide Dimer-Based Acceptor Material for Highly Efficient Solution-Processed Non-Fullerene Organic Solar Cells with 4.03% Efficiency. Adv. Mater. 2013, 25, 5791–5797. [Google Scholar] [CrossRef] [PubMed]
  372. Chen, W.; Yang, X.; Long, G.; Wan, X.; Chen, Y.; Zhang, Q. A Perylene Diimide (PDI)-Based Small Molecule with Tetrahedral Configuration as a Non-Fullerene Acceptor for Organic Solar Cells. J. Mater. Chem. C 2015, 3, 4698–4705. [Google Scholar] [CrossRef]
  373. Chen, W.; Salim, T.; Fan, H.; James, L.; Lam, Y.M.; Zhang, Q. Quinoxaline-Functionalized C60 Derivatives as Electron Acceptors in Organic Solar Cells. RSC Adv. 2014, 4, 25291–25301. [Google Scholar] [CrossRef]
  374. Zhang, Q.; Xiao, J.; Yin, Z.; Duong, H.M.; Qiao, F.; Boey, F.; Hu, X.; Zhang, H.; Wudl, F. Synthesis, Characterization, and Physical Properties of a Conjugated Heteroacene: 2-Methyl-1,4,6,7,8,9-Hexaphenylbenz(g)Isoquinolin-3(2H)-One (BIQ). Chem. Asian J. 2011, 6, 856–862. [Google Scholar] [CrossRef]
  375. Lin, Y.; Wang, J.; Dai, S.; Li, Y.; Zhu, D.; Zhan, X. A Twisted Dimeric Perylene Diimide Electron Acceptor for Efficient Organic Solar Cells. Adv. Energy Mater. 2014, 4, 1400420. [Google Scholar] [CrossRef]
  376. Bai, H.; Wang, Y.; Cheng, P.; Wang, J.; Wu, Y.; Hou, J.; Zhan, X. An Electron Acceptor Based on Indacenodithiophene and 1,1-Dicyanomethylene-3-Indanone for Fullerene-Free Organic Solar Cells. J. Mater. Chem. A 2015, 3, 1910–1914. [Google Scholar] [CrossRef]
  377. Zhan, X.; Tan, Z.; Domercq, B.; An, Z.; Zhang, X.; Barlow, S.; Li, Y.; Zhu, D.; Kippelen, B.; Marder, S.R. A High-Mobility Electron-Transport Polymer with Broad Absorption and Its Use in Field-Effect Transistors and All-Polymer Solar Cells. J. Am. Chem. Soc. 2007, 129, 7246–7247. [Google Scholar] [CrossRef]
  378. Holliday, S.; Ashraf, R.S.; Nielsen, C.B.; Kirkus, M.; Röhr, J.A.; Tan, C.-H.; Collado-Fregoso, E.; Knall, A.-C.; Durrant, J.R.; Nelson, J.; et al. A Rhodanine Flanked Nonfullerene Acceptor for Solution-Processed Organic Photovoltaics. J. Am. Chem. Soc. 2015, 137, 898–904. [Google Scholar] [CrossRef]
  379. Patil, H.; Gupta, A.; Alford, B.; Ma, D.; Privér, S.H.; Bilic, A.; Sonar, P.; Bhosale, S.V. Conjoint Use of Dibenzosilole and Indan-1,3-Dione Functionalities to Prepare an Efficient Non-Fullerene Acceptor for Solution-Processable Bulk-Heterojunction Solar Cells. Asian J. Org. Chem. 2015, 4, 1096–1102. [Google Scholar] [CrossRef]
  380. Srivani, D.; Gupta, A.; Bhosale, S.V.; Puyad, A.L.; Xiang, W.; Li, J.; Evans, R.A.; Bhosale, S.V. Non-Fullerene Acceptors Based on Central Naphthalene Diimide Flanked by Rhodanine or 1,3-Indanedione. Chem. Commun. 2017, 53, 7080–7083. [Google Scholar] [CrossRef]
  381. Chen, X.; Feng, H.; Lin, Z.; Jiang, Z.; He, T.; Yin, S.; Wan, X.; Chen, Y.; Zhang, Q.; Qiu, H. Impact of End-Capped Groups on the Properties of Dithienosilole-Based Small Molecules for Solution-Processed Organic Solar Cells. Dyes Pigments 2017, 147, 183–189. [Google Scholar] [CrossRef]
  382. Raynor, A.M.; Gupta, A.; Plummer, C.M.; Jackson, S.L.; Bilic, A.; Patil, H.; Sonar, P.; Bhosale, S.V. Significant Improvement of Optoelectronic and Photovoltaic Properties by Incorporating Thiophene in a Solution-Processable D–A–D Modular Chromophore. Molecules 2015, 20, 21787–21801. [Google Scholar] [CrossRef] [PubMed]
  383. Liu, Y.; Sun, Y.; Li, M.; Feng, H.; Ni, W.; Zhang, H.; Wan, X.; Chen, Y. Efficient Carbazole-Based Small-Molecule Organic Solar Cells with an Improved Fill Factor. RSC Adv. 2018, 8, 4867–4871. [Google Scholar] [CrossRef] [PubMed]
  384. Mishra, A.; Rana, T.; Looser, A.; Stolte, M.; Würthner, F.; Bäuerle, P.; Sharma, G.D. High Performance A–D–A Oligothiophene-Based Organic Solar Cells Employing Two-Step Annealing and Solution-Processable Copper Thiocyanate (CuSCN) as an Interfacial Hole Transporting Layer. J. Mater. Chem. A 2016, 4, 17344–17353. [Google Scholar] [CrossRef]
  385. Qi, X.; Lo, Y.-C.; Zhao, Y.; Xuan, L.; Ting, H.-C.; Wong, K.-T.; Rahaman, M.; Chen, Z.; Xiao, L.; Qu, B. Two Novel Small Molecule Donors and the Applications in Bulk-Heterojunction Solar Cells. Front. Chem. 2018, 6, 260. [Google Scholar] [CrossRef]
  386. Adhikari, T.; Solanke, P.; Pathak, D.; Wagner, T.; Bureš, F.; Reed, T.; Nunzi, J.-M. T-Shaped Indan-1,3-Dione Derivatives as Promising Electron Donors for Bulk Heterojunction Small Molecule Solar Cell. Opt. Mater. 2017, 69, 312–317. [Google Scholar] [CrossRef]
  387. Walker, T.K.; Suthers, A.J.; Roe, L.L.; Shaw, H. LXIX.—Syntheses of Antiseptic Derivatives of Indan-1: 3-Dione. Part II. Interaction of Alkylmalonyl Chlorides with p-Tolyl Methyl Ether. J. Chem. Soc. Resumed 1931, 514–520. [Google Scholar] [CrossRef]
  388. Robinson, F.A.; Suthers, A.J.; Walker, T.K. Some New Antiseptics Related to Indan-I:3-Dione. Biochem. J. 1932, 26, 1890–1901. [Google Scholar] [CrossRef]
  389. Wen, A.; Delaquis, P.; Stanich, K.; Toivonen, P. Antilisterial Activity of Selected Phenolic Acids. Food Microbiol. 2003, 20, 305–311. [Google Scholar] [CrossRef]
  390. Hudzicki, J. Kirby-Bauer Disk Diffusion Susceptibility Test Protocol; American Society for Microbiology: Washington, DC, USA, 2009. [Google Scholar]
  391. Fairbrother, R.W. A Text-Book of Medical Bacteriology; William Heinemann: London, UK, 1937. [Google Scholar]
  392. Mohil, R.; Kumar, D.; Mor, S. Synthesis and Antimicrobial Activity of Some 1,3-Disubstituted Indeno[1,2-c]Pyrazoles. J. Heterocycl. Chem. 2014, 51, 203–211. [Google Scholar] [CrossRef]
  393. Mor, S.; Mohil, R.; Nagoria, S.; Kumar, A.; Lal, K.; Kumar, D.; Singh, V. Regioselective Synthesis, Antimicrobial Evaluation and QSAR Studies of Some 3-Aryl-1-Heteroarylindeno[1,2-c]Pyrazol-4(1H)-Ones. J. Heterocycl. Chem. 2017, 54, 1327–1341. [Google Scholar] [CrossRef]
  394. Alsharif, M.A.; Mukhtar, S.; Asiri, A.M.; Khan, S.A. One Pot Synthesis, Physicochemical and Photophysical Investigation of Biologically Active Pyridine-3-Carboxylate (ECPC) as Probe to Determine CMC of Surfactants in Organized Media. Colloids Surf. Physicochem. Eng. Asp. 2018, 543, 38–45. [Google Scholar] [CrossRef]
  395. Pandian, R.; Naushad, E.; Vijayakumar, V.; Peters, G.H.; Mondikalipudur Nanjappagounder, P. Synthesis and Crystal Structures of 2-Methyl-4-Aryl-5-Oxo-5H-Indeno [1,2-b] Pyridine Carboxylate Derivatives. Chem. Cent. J. 2014, 8, 34. [Google Scholar] [CrossRef]
  396. Gupta, R.; Chaudhary, R.P. Efficient Ionic Liquid-Catalysed Synthesis and Antimicrobial Studies of 4,6-Diaryl- and 4,5-Fused Pyrimidine-2-Thiones. J. Chem. Res. 2012, 36, 718–721. [Google Scholar] [CrossRef]
  397. Meena, K.; Kumari, S.; Khurana, J.M.; Malik, A.; Sharma, C.; Panwar, H. One Pot Three Component Synthesis of Spiro [Indolo-3,10′-Indeno[1,2-b] Quinolin]-2,4,11′-Triones as a New Class of Antifungal and Antimicrobial Agents. Chin. Chem. Lett. 2017, 28, 136–142. [Google Scholar] [CrossRef]
  398. Khanna, G.; Aggarwal, K.; Khurana, J. An Efficient and Confluent Approach for the Synthesis of Novel 3,4-Dihydro-2H-Naphtho[2,3-e][1,3]Oxazine-5,10-Dione Derivatives by a Three Component Reaction in Ionic Liquid. ChemInform 2015, 46, 46448–46454. [Google Scholar] [CrossRef]
  399. Saluja, P.; Khurana, J.M.; Sharma, C.; Aneja, K.R. An Efficient and Convenient Approach for the Synthesis of Novel Pyrazolo[1,2-a]Triazole-Triones and Evaluation of Their Antimicrobial Activities. Aust. J. Chem. 2014, 67, 867–874. [Google Scholar] [CrossRef]
  400. Singh, H.; Sindhu, J.; Khurana, J.M.; Sharma, C.; Aneja, K.R. Ultrasound Promoted One Pot Synthesis of Novel Fluorescent Triazolyl Spirocyclic Oxindoles Using DBU Based Task Specific Ionic Liquids and Their Antimicrobial Activity. Eur. J. Med. Chem. 2014, 77, 145–154. [Google Scholar] [CrossRef]
  401. Rajeswari, M.; Saluja, P.; Khurana, J.M. A Facile and Green Approach for the Synthesis of Spiro[Naphthalene-2,5′-Pyrimidine]-4-Carbonitrile via a One-Pot Three-Component Condensation Reaction Using DBU as a Catalyst. RSC Adv. 2016, 6, 1307–1312. [Google Scholar] [CrossRef]
  402. Avendano, C.; Menendez, J.C. Medicinal Chemistry of Anticancer Drugs; Elsevier Science: Amsterdam, The Netherlands, 1995; ISBN 978-0-444-62667-7. [Google Scholar]
  403. Kumar, D.; Jain, S.K. A Comprehensive Review of N-Heterocycles as Cytotoxic Agents. Curr. Med. Chem. 2016, 23, 4338–4394. [Google Scholar] [CrossRef]
  404. Lang, D.K.; Kaur, R.; Arora, R.; Saini, B.; Arora, S. Nitrogen-Containing Heterocycles as Anticancer Agents: An Overview. Anticancer Agents Med. Chem. 2020, 20, 2150–2168. [Google Scholar] [CrossRef] [PubMed]
  405. Nagarajan, M.; Morrell, A.; Ioanoviciu, A.; Antony, S.; Kohlhagen, G.; Agama, K.; Hollingshead, M.; Pommier, Y.; Cushman, M. Synthesis and Evaluation of Indenoisoquinoline Topoisomerase I Inhibitors Substituted with Nitrogen Heterocycles. J. Med. Chem. 2006, 49, 6283–6289. [Google Scholar] [CrossRef] [PubMed]
  406. Prachayasittikul, S.; Manam, P.; Chinworrungsee, M.; Isarankura-Na-Ayudhya, C.; Ruchirawat, S.; Prachayasittikul, V. Bioactive Azafluorenone Alkaloids from Polyalthia Debilis (Pierre) Finet & Gagnep. Molecules 2009, 14, 4414–4424. [Google Scholar] [CrossRef] [PubMed]
  407. Manpadi, M.; Uglinskii, P.Y.; Rastogi, S.K.; Cotter, K.M.; Wong, Y.-S.C.; Anderson, L.A.; Ortega, A.J.; Van slambrouck, S.; Steelant, W.F.A.; Rogelj, S.; et al. Three-Component Synthesis and Anticancer Evaluation of Polycyclic Indenopyridines Lead to the Discovery of a Novel Indenoheterocycle with Potent Apoptosis Inducing Properties. Org. Biomol. Chem. 2007, 5, 3865–3872. [Google Scholar] [CrossRef] [PubMed]
  408. Vermes, I.; Haanen, C.; Steffens-Nakken, H.; Reutellingsperger, C. A Novel Assay for Apoptosis Flow Cytometric Detection of Phosphatidylserine Expression on Early Apoptotic Cells Using Fluorescein Labelled Annexin V. J. Immunol. Methods 1995, 184, 39–51. [Google Scholar] [CrossRef]
  409. Evdokimov, N.M.; Van slambrouck, S.; Heffeter, P.; Tu, L.; Le Calvé, B.; Lamoral-Theys, D.; Hooten, C.J.; Uglinskii, P.Y.; Rogelj, S.; Kiss, R.; et al. Structural Simplification of Bioactive Natural Products with Multicomponent Synthesis. 3. Fused Uracil-Containing Heterocycles as Novel Topoisomerase-Targeting Agents. J. Med. Chem. 2011, 54, 2012–2021. [Google Scholar] [CrossRef]
  410. Abdel-Aziz, A.A.-M.; El-Azab, A.S.; Ceruso, M.; Supuran, C.T. Carbonic Anhydrase Inhibitory Activity of Sulfonamides and Carboxylic Acids Incorporating Cyclic Imide Scaffolds. Bioorg. Med. Chem. Lett. 2014, 24, 5185–5189. [Google Scholar] [CrossRef]
  411. Ghorab, M.M.; Al-Said, M.S. Anticancer Activity of Novel Indenopyridine Derivatives. Arch. Pharm. Res. 2012, 35, 987–994. [Google Scholar] [CrossRef]
  412. Arnold, L.D.; Xu, Y.; Barlozzari, T.; Rafferty, P.; Hockley, M.; Turner, A. Indeno[1,2-C]Pyrazole Derivatives for Inhibiting Tyrosine Kinase Activity. U.S. Patent 6,534,655, 18 March 2003. [Google Scholar]
  413. Nugiel, D.A.; Vidwans, A.; Etzkorn, A.-M.; Rossi, K.A.; Benfield, P.A.; Burton, C.R.; Cox, S.; Doleniak, D.; Seitz, S.P. Synthesis and Evaluation of Indenopyrazoles as Cyclin-Dependent Kinase Inhibitors. 2. Probing the Indeno Ring Substituent Pattern. J. Med. Chem. 2002, 45, 5224–5232. [Google Scholar] [CrossRef]
  414. Yue, E.W.; Higley, C.A.; DiMeo, S.V.; Carini, D.J.; Nugiel, D.A.; Benware, C.; Benfield, P.A.; Burton, C.R.; Cox, S.; Grafstrom, R.H.; et al. Synthesis and Evaluation of Indenopyrazoles as Cyclin-Dependent Kinase Inhibitors. 3. Structure Activity Relationships at C3. J. Med. Chem. 2002, 45, 5233–5248. [Google Scholar] [CrossRef]
  415. Nugiel, D.A.; Carini, D.J.; Yue, E.W.; Dimeo, S.V. 5-Aminoindeno(1,2-c)Pyrazol-4-Ones as Anti-Cancer and Anti-Proliferative Agents. PCT/US1999/008616, 28 October 1999. [Google Scholar]
  416. Matsuo, T.; Nishida, S.; Takagi, T. Anticancer Agent. JPS60130521A, 12 July 1985. [Google Scholar]
  417. Heck, A.J.R.; Chandler, D.W. Imaging Techniques for the Study of Chemical Reaction Dynamics. Annu. Rev. Phys. Chem. 1995, 46, 335–372. [Google Scholar] [CrossRef] [PubMed]
  418. Qian, J.; Tang, B.Z. AIE Luminogens for Bioimaging and Theranostics: From Organelles to Animals. Chem 2017, 3, 56–91. [Google Scholar] [CrossRef]
  419. Dai, Y.; He, F.; Ji, H.; Zhao, X.; Misal, S.; Qi, Z. Dual-Functional NIR AIEgens for High-Fidelity Imaging of Lysosomes in Cells and Photodynamic Therapy. ACS Sens. 2020, 5, 225–233. [Google Scholar] [CrossRef]
  420. Zhang, H.; Xu, L.; Chen, W.; Huang, J.; Huang, C.; Sheng, J.; Song, X. A Lysosome-Targetable Fluorescent Probe for Simultaneously Sensing Cys/Hcy, GSH, and H2S from Different Signal Patterns. ACS Sens. 2018, 3, 2513–2517. [Google Scholar] [CrossRef] [PubMed]
  421. Gao, M.; Su, H.; Li, S.; Lin, Y.; Ling, X.; Qin, A.; Tang, B.Z. An Easily Accessible Aggregation-Induced Emission Probe for Lipid Droplet-Specific Imaging and Movement Tracking. Chem. Commun. 2017, 53, 921–924. [Google Scholar] [CrossRef] [PubMed]
  422. Daemen, S.; van Zandvoort, M.A.M.J.; Parekh, S.H.; Hesselink, M.K.C. Microscopy Tools for the Investigation of Intracellular Lipid Storage and Dynamics. Mol. Metab. 2016, 5, 153–163. [Google Scholar] [CrossRef]
  423. Zhai, J.; Zhang, Y.; Yang, C.; Xu, Y.; Qin, Y. A Long Wavelength Hydrophobic Probe for Intracellular Lipid Droplets. Analyst 2014, 139, 52–54. [Google Scholar] [CrossRef]
  424. Kim, E.; Lee, S.; Park, S.B. A Seoul-Fluor-Based Bioprobe for Lipid Droplets and Its Application in Image-Based High Throughput Screening. Chem. Commun. 2012, 48, 2331–2333. [Google Scholar] [CrossRef]
  425. Lee, J.H.; So, J.-H.; Jeon, J.H.; Choi, E.B.; Lee, Y.-R.; Chang, Y.-T.; Kim, C.-H.; Bae, M.A.; Ahn, J.H. Synthesis of a New Fluorescent Small Molecule Probe and Its Use for in Vivolipid Imaging. Chem. Commun. 2011, 47, 7500–7502. [Google Scholar] [CrossRef]
  426. Sharma, A.; Umar, S.; Kar, P.; Singh, K.; Sachdev, M.; Goel, A. A New Type of Biocompatible Fluorescent Probe AFN for Fixed and Live Cell Imaging of Intracellular Lipid Droplets. Analyst 2016, 141, 137–143. [Google Scholar] [CrossRef]
  427. Neef, A.B.; Schultz, C. Selective Fluorescence Labeling of Lipids in Living Cells. Angew. Chem. Int. Ed. 2009, 48, 1498–1500. [Google Scholar] [CrossRef] [PubMed]
  428. Eggert, D.; Rösch, K.; Reimer, R.; Herker, E. Visualization and Analysis of Hepatitis C Virus Structural Proteins at Lipid Droplets by Super-Resolution Microscopy. PLoS ONE 2014, 9, e102511. [Google Scholar] [CrossRef] [PubMed]
  429. Greenspan, P.; Mayer, E.P.; Fowler, S.D. Nile Red: A Selective Fluorescent Stain for Intracellular Lipid Droplets. J. Cell Biol. 1985, 100, 965–973. [Google Scholar] [CrossRef]
  430. Spandl, J.; White, D.J.; Peychl, J.; Thiele, C. Live Cell Multicolor Imaging of Lipid Droplets with a New Dye, LD540. Traffic 2009, 10, 1579–1584. [Google Scholar] [CrossRef] [PubMed]
  431. Mondal, A.; Naskar, B.; Goswami, S.; Prodhan, C.; Chaudhuri, K.; Mukhopadhyay, C. A Quick Accelerating Microwave-Assisted Sustainable Technique: Permutated Spiro-Casing for Imaging Experiment. Mol. Divers. 2020, 24, 93–106. [Google Scholar] [CrossRef]
  432. Chaudhuri, K.R.; Schapira, A.H. Non-Motor Symptoms of Parkinson’s Disease: Dopaminergic Pathophysiology and Treatment. Lancet Neurol. 2009, 8, 464–474. [Google Scholar] [CrossRef]
  433. Weintraub, D.; Comella, C.L.; Horn, S. Parkinson’s Disease—Part 2: Treatment of Motor Symptoms. Am. J. Manag. Care 2008, 14, S49–S58. [Google Scholar]
  434. Shook, B.C.; Rassnick, S.; Osborne, M.C.; Davis, S.; Westover, L.; Boulet, J.; Hall, D.; Rupert, K.C.; Heintzelman, G.R.; Hansen, K.; et al. In Vivo Characterization of a Dual Adenosine A2A/A1 Receptor Antagonist in Animal Models of Parkinson’s Disease. J. Med. Chem. 2010, 53, 8104–8115. [Google Scholar] [CrossRef]
  435. Shook, B.C.; Rassnick, S.; Hall, D.; Rupert, K.C.; Heintzelman, G.R.; Hansen, K.; Chakravarty, D.; Bullington, J.L.; Scannevin, R.H.; Magliaro, B.; et al. Methylene Amine Substituted Arylindenopyrimidines as Potent Adenosine A2A/A1 Antagonists. Bioorg. Med. Chem. Lett. 2010, 20, 2864–2867. [Google Scholar] [CrossRef]
  436. Atack, J.R.; Shook, B.C.; Rassnick, S.; Jackson, P.F.; Rhodes, K.; Drinkenburg, W.H.; Ahnaou, A.; te Riele, P.; Langlois, X.; Hrupka, B.; et al. JNJ-40255293, a Novel Adenosine A2A/A1 Antagonist with Efficacy in Preclinical Models of Parkinson’s Disease. ACS Chem. Neurosci. 2014, 5, 1005–1019. [Google Scholar] [CrossRef]
  437. Lim, H.-K.; Chen, J.; Sensenhauser, C.; Cook, K.; Preston, R.; Thomas, T.; Shook, B.; Jackson, P.F.; Rassnick, S.; Rhodes, K.; et al. Overcoming the Genotoxicity of a Pyrrolidine Substituted Arylindenopyrimidine As a Potent Dual Adenosine A2A/A1 Antagonist by Minimizing Bioactivation to an Iminium Ion Reactive Intermediate. Chem. Res. Toxicol. 2011, 24, 1012–1030. [Google Scholar] [CrossRef] [PubMed]
  438. Heintzelman, G.R.; Bullington, J.L.; Rupert, K.C. Arylindenopyridines and Arylindenopyrimidines and Related Therapeutic and Prophylactic Methods. U.S. 20050267138, 1 December 2005. [Google Scholar]
  439. Gijsen, H.J.M.; Berthelot, D.; De Cleyn, M.A.J.; Geuens, I.; Brône, B.; Mercken, M. Tricyclic 3,4-Dihydropyrimidine-2-Thione Derivatives as Potent TRPA1 Antagonists. Bioorg. Med. Chem. Lett. 2012, 22, 797–800. [Google Scholar] [CrossRef]
  440. Edwards, P.J.; Sturino, C. Managing the Liabilities Arising from Structural Alerts: A Safe Philosophy for Medicinal Chemists. Curr. Med. Chem. 2011, 18, 3116–3135. [Google Scholar] [CrossRef] [PubMed]
  441. Fisher, R.S.; Boas, W.V.E.; Blume, W.; Elger, C.; Genton, P.; Lee, P.; Engel, J., Jr. Epileptic Seizures and Epilepsy: Definitions Proposed by the International League Against Epilepsy (ILAE) and the International Bureau for Epilepsy (IBE). Epilepsia 2005, 46, 470–472. [Google Scholar] [CrossRef]
  442. Picot, M.-C.; Baldy-Moulinier, M.; Daurès, J.-P.; Dujols, P.; Crespel, A. The Prevalence of Epilepsy and Pharmacoresistant Epilepsy in Adults: A Population-Based Study in a Western European Country. Epilepsia 2008, 49, 1230–1238. [Google Scholar] [CrossRef]
  443. Rashid, U.; Ansari, F.L. Chapter 2—Challenges in Designing Therapeutic Agents for Treating Alzheimer’s Disease-from Serendipity to Rationality. In Drug Design and Discovery in Alzheimer’s Disease; Atta-ur-Rahman, Choudhary, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2014; pp. 40–141. ISBN 978-0-12-803959-5. [Google Scholar]
  444. LaFerla, F.M.; Green, K.N.; Oddo, S. Intracellular Amyloid-β in Alzheimer’s Disease. Nat. Rev. Neurosci. 2007, 8, 499–509. [Google Scholar] [CrossRef]
  445. Tanoli, S.T.; Ramzan, M.; Hassan, A.; Sadiq, A.; Jan, M.S.; Khan, F.A.; Ullah, F.; Ahmad, H.; Bibi, M.; Mahmood, T.; et al. Design, Synthesis and Bioevaluation of Tricyclic Fused Ring System as Dual Binding Site Acetylcholinesterase Inhibitors. Bioorganic Chem. 2019, 83, 336–347. [Google Scholar] [CrossRef]
  446. Campagna, F.; Palluotto, F.; Carotti, A.; Maciocco, E. Synthesis, Central and Peripheral Benzodiazepine Receptor Affinity of Pyrazole and Pyrazole-Containing Polycyclic Derivatives. Il Farm. 2004, 59, 849–856. [Google Scholar] [CrossRef]
  447. Catto, M.; Aliano, R.; Carotti, A.; Cellamare, S.; Palluotto, F.; Purgatorio, R.; De Stradis, A.; Campagna, F. Design, Synthesis and Biological Evaluation of Indane-2-Arylhydrazinylmethylene-1,3-Diones and Indol-2-Aryldiazenylmethylene-3-Ones as β-Amyloid Aggregation Inhibitors. Eur. J. Med. Chem. 2010, 45, 1359–1366. [Google Scholar] [CrossRef]
  448. Shapiro, S.L.; Geiger, K.; Freedman, L. Indandione Anticoagulants. J. Org. Chem. 1960, 25, 1860–1865. [Google Scholar] [CrossRef]
  449. Murphy, M.J. Chapter 46—Anticoagulant Rodenticides. In Veterinary Toxicology, 3rd ed.; Gupta, R.C., Ed.; Academic Press: Cambridge, MA, USA, 2018; pp. 583–612. ISBN 978-0-12-811410-0. [Google Scholar]
  450. Daveluy, A.; Milpied, B.; Barbaud, A.; Lebrun-Vignes, B.; Gouraud, A.; Laroche, M.-L.; Ciobanu, E.; Bégaud, B.; Moore, N.; Miremont-Salamé, G.; et al. Fluindione and Drug Reaction with Eosinophilia and Systemic Symptoms: An Unrecognised Adverse Effect? Eur. J. Clin. Pharmacol. 2012, 68, 101–105. [Google Scholar] [CrossRef]
  451. Wehling, M.; Collins, R.; Gil, V.M.; Hanon, O.; Hardt, R.; Hoffmeister, M.; Monteiro, P.; Quinn, T.J.; Ropers, D.; Sergi, G.; et al. Appropriateness of Oral Anticoagulants for the Long-Term Treatment of Atrial Fibrillation in Older People: Results of an Evidence-Based Review and International Consensus Validation Process (OAC-FORTA 2016). Drugs Aging 2017, 34, 499–507. [Google Scholar] [CrossRef] [PubMed]
  452. Ma, S.; Long, D.; Chen, P.; Shi, H.; Li, H.; Fang, R.; Wang, X.; Xie, X.; She, X. Synthesis of 2,3-Disubstituted Indoles via a Tandem Reaction. Org. Chem. Front. 2020, 7, 2689–2695. [Google Scholar] [CrossRef]
  453. Murdock, K.C. Triacylhalomethanes: 2-Halo-2-Acyl-1,3-Indandiones. J. Org. Chem. 1959, 24, 845–849. [Google Scholar] [CrossRef]
  454. Larsen, B.J.; Rosano, R.J.; Ford-Hutchinson, T.A.; Reitz, A.B.; Wrobel, J.E. A Method for C2 Acylation of 1,3-Indandiones. Tetrahedron 2018, 74, 2762–2768. [Google Scholar] [CrossRef]
  455. Regnery, J.; Friesen, A.; Geduhn, A.; Göckener, B.; Kotthoff, M.; Parrhysius, P.; Petersohn, E.; Reifferscheid, G.; Schmolz, E.; Schulz, R.S.; et al. Rating the Risks of Anticoagulant Rodenticides in the Aquatic Environment: A Review. Environ. Chem. Lett. 2019, 17, 215–240. [Google Scholar] [CrossRef]
  456. Bates, N. Anticoagulant Rodenticide Toxicosis. Companion Anim. 2016, 21, 466–471. [Google Scholar] [CrossRef]
Scheme 1. Synthetic routes to indane-1,3-dione 4.
Scheme 1. Synthetic routes to indane-1,3-dione 4.
Molecules 27 05976 sch001
Scheme 2. Synthetic routes to substituted indane-1,3-diones 19–21 and 25, 26.
Scheme 2. Synthetic routes to substituted indane-1,3-diones 19–21 and 25, 26.
Molecules 27 05976 sch002
Scheme 3. Synthetic routes to 28, 31, 31′, 33, 33′, 35, 35′.
Scheme 3. Synthetic routes to 28, 31, 31′, 33, 33′, 35, 35′.
Molecules 27 05976 sch003
Scheme 4. Synthetic routes to 37a–37d, 39, 41, 43, 45, 47 and 49.
Scheme 4. Synthetic routes to 37a–37d, 39, 41, 43, 45, 47 and 49.
Molecules 27 05976 sch004
Scheme 5. Synthesis of bindone 50.
Scheme 5. Synthesis of bindone 50.
Molecules 27 05976 sch005
Scheme 6. Synthetic route to bis-thiazoles and bis-thiazolidinone starting from 4.
Scheme 6. Synthetic route to bis-thiazoles and bis-thiazolidinone starting from 4.
Molecules 27 05976 sch006
Scheme 7. Synthetic routes to 68, 72 and 75.
Scheme 7. Synthetic routes to 68, 72 and 75.
Molecules 27 05976 sch007
Scheme 8. Examples of halogenated indane-1,3-diones 76–82, 84 reported in the literature.
Scheme 8. Examples of halogenated indane-1,3-diones 76–82, 84 reported in the literature.
Molecules 27 05976 sch008
Scheme 9. Synthesis of nitro-substituted indane-1,3-diones 19 and 21.
Scheme 9. Synthesis of nitro-substituted indane-1,3-diones 19 and 21.
Molecules 27 05976 sch009
Scheme 10. Ethoxy-carbonyl compound 32.
Scheme 10. Ethoxy-carbonyl compound 32.
Molecules 27 05976 sch010
Scheme 11. Synthesis of push–pull dyes 95–114.
Scheme 11. Synthesis of push–pull dyes 95–114.
Molecules 27 05976 sch011
Scheme 12. Mechanism resulting in the synthesis of 3-(dialkylamino)-1,2-dihydro-9-oxo-9H-indeno [2,1-c]pyridine-4-carbonitrile derivatives 115.
Scheme 12. Mechanism resulting in the synthesis of 3-(dialkylamino)-1,2-dihydro-9-oxo-9H-indeno [2,1-c]pyridine-4-carbonitrile derivatives 115.
Molecules 27 05976 sch012
Scheme 13. Mechanism supporting the formation of 117 starting from 116.
Scheme 13. Mechanism supporting the formation of 117 starting from 116.
Molecules 27 05976 sch013
Figure 1. Crystal structure of 117. Reproduced with permission from Ref. [57].
Figure 1. Crystal structure of 117. Reproduced with permission from Ref. [57].
Molecules 27 05976 g001
Scheme 14. Synthetic routes to 119 starting from 29.
Scheme 14. Synthetic routes to 119 starting from 29.
Molecules 27 05976 sch014
Scheme 15. Knoevenagel reactions performed in classical ethanol/piperidine conditions.
Scheme 15. Knoevenagel reactions performed in classical ethanol/piperidine conditions.
Molecules 27 05976 sch015
Scheme 16. Synthetic route to push–pull dye 127.
Scheme 16. Synthetic route to push–pull dye 127.
Molecules 27 05976 sch016
Scheme 17. Synthetic route to ninhydrin 138.
Scheme 17. Synthetic route to ninhydrin 138.
Molecules 27 05976 sch017
Scheme 18. Different routes for halogenation of indane-1,3-dione 4 at the methylene position.
Scheme 18. Different routes for halogenation of indane-1,3-dione 4 at the methylene position.
Molecules 27 05976 sch018
Scheme 19. Fluorination reactions of indane-1,3-dione 4.
Scheme 19. Fluorination reactions of indane-1,3-dione 4.
Molecules 27 05976 sch019
Scheme 20. Mechanism involved in the fluorination of 4.
Scheme 20. Mechanism involved in the fluorination of 4.
Molecules 27 05976 sch020
Scheme 21. Mechanism supporting the coexistence of a radical and an electrophilic bromination pathway for the bromination of 4.
Scheme 21. Mechanism supporting the coexistence of a radical and an electrophilic bromination pathway for the bromination of 4.
Molecules 27 05976 sch021
Scheme 22. Synthesis of 150 using NBS as the bromination agent.
Scheme 22. Synthesis of 150 using NBS as the bromination agent.
Molecules 27 05976 sch022
Scheme 23. Synthetic route to 153.
Scheme 23. Synthetic route to 153.
Molecules 27 05976 sch023
Scheme 24. Synthetic route to 154.
Scheme 24. Synthetic route to 154.
Molecules 27 05976 sch024
Scheme 25. Synthetic routes to cyclophanes 158a and 158b.
Scheme 25. Synthetic routes to cyclophanes 158a and 158b.
Molecules 27 05976 sch025
Scheme 26. Synthetic routes to the crown ether derivative of 1,3-indandione 159.
Scheme 26. Synthetic routes to the crown ether derivative of 1,3-indandione 159.
Molecules 27 05976 sch026
Scheme 27. Synthetic route to 162.
Scheme 27. Synthetic route to 162.
Molecules 27 05976 sch027
Scheme 28. Synthetic routes to tetracycline heterocyclic analogues.
Scheme 28. Synthetic routes to tetracycline heterocyclic analogues.
Molecules 27 05976 sch028
Scheme 29. Synthetic routes to 172.
Scheme 29. Synthetic routes to 172.
Molecules 27 05976 sch029
Scheme 30. Synthetic Routes to 177, 179, 181 and 182.
Scheme 30. Synthetic Routes to 177, 179, 181 and 182.
Molecules 27 05976 sch030
Scheme 31. Synthetic routes to 185–192.
Scheme 31. Synthetic routes to 185–192.
Molecules 27 05976 sch031
Scheme 32. Synthetic route to 196.
Scheme 32. Synthetic route to 196.
Molecules 27 05976 sch032
Scheme 33. Synthetic routes to porphyrins 199 and 200.
Scheme 33. Synthetic routes to porphyrins 199 and 200.
Molecules 27 05976 sch033
Scheme 34. Synthetic routes to various (metallo)porphyrins.
Scheme 34. Synthetic routes to various (metallo)porphyrins.
Molecules 27 05976 sch034
Scheme 35. Synthetic route to 208.
Scheme 35. Synthetic route to 208.
Molecules 27 05976 sch035
Scheme 36. Mechanism involved in the synthesis of 212.
Scheme 36. Mechanism involved in the synthesis of 212.
Molecules 27 05976 sch036
Scheme 37. Synthetic routes to 212 and 213.
Scheme 37. Synthetic routes to 212 and 213.
Molecules 27 05976 sch037
Scheme 38. Chemical structures of the indeno-fused structures 216–224 treated below.
Scheme 38. Chemical structures of the indeno-fused structures 216–224 treated below.
Molecules 27 05976 sch038
Scheme 39. Synthetic route to 217b.
Scheme 39. Synthetic route to 217b.
Molecules 27 05976 sch039
Scheme 40. Mechanism involved in the synthesis of indeno-fused structures 218.
Scheme 40. Mechanism involved in the synthesis of indeno-fused structures 218.
Molecules 27 05976 sch040
Scheme 41. General mechanism of indeno-fused structures synthetized by MCR.
Scheme 41. General mechanism of indeno-fused structures synthetized by MCR.
Molecules 27 05976 sch041
Scheme 42. The different synthetic routes to spiroindanediones discussed in this part.
Scheme 42. The different synthetic routes to spiroindanediones discussed in this part.
Molecules 27 05976 sch042
Scheme 43. Synthesis of the ferrocenecarbocyaldehyde adduct 230.
Scheme 43. Synthesis of the ferrocenecarbocyaldehyde adduct 230.
Molecules 27 05976 sch043
Scheme 44. Synthetic route to 235.
Scheme 44. Synthetic route to 235.
Molecules 27 05976 sch044
Scheme 45. Synthetic routes to 236, 238, 240 and 242.
Scheme 45. Synthetic routes to 236, 238, 240 and 242.
Molecules 27 05976 sch045
Scheme 46. Synthetic route to azomethine ylide 245.
Scheme 46. Synthetic route to azomethine ylide 245.
Molecules 27 05976 sch046
Scheme 47. Synthetic routes to dihydro-spiro[indene-2,3′-pyrrolidines] 247.
Scheme 47. Synthetic routes to dihydro-spiro[indene-2,3′-pyrrolidines] 247.
Molecules 27 05976 sch047
Scheme 48. Mechanism supporting the formation of a unique diastereoisomer.
Scheme 48. Mechanism supporting the formation of a unique diastereoisomer.
Molecules 27 05976 sch048
Scheme 49. Synthetic route to 250.
Scheme 49. Synthetic route to 250.
Molecules 27 05976 sch049
Scheme 50. Synthesis of azomethine imine.
Scheme 50. Synthesis of azomethine imine.
Molecules 27 05976 sch050
Scheme 51. Synthesis of 256.
Scheme 51. Synthesis of 256.
Molecules 27 05976 sch051
Scheme 52. Mechanism supporting the synthesis of 256. Reproduced with permission from Ref. [254].
Scheme 52. Mechanism supporting the synthesis of 256. Reproduced with permission from Ref. [254].
Molecules 27 05976 sch052
Scheme 53. Synthetic route to chromeno [3,4-b]pyrrolidine 259.
Scheme 53. Synthetic route to chromeno [3,4-b]pyrrolidine 259.
Molecules 27 05976 sch053
Scheme 54. Synthesis of chromeno [3,4-c]pyrrolidine 262 while using TMG as the base.
Scheme 54. Synthesis of chromeno [3,4-c]pyrrolidine 262 while using TMG as the base.
Molecules 27 05976 sch054
Scheme 55. Synthesis of 259.
Scheme 55. Synthesis of 259.
Molecules 27 05976 sch055
Scheme 56. Synthesis of 262.
Scheme 56. Synthesis of 262.
Molecules 27 05976 sch056
Scheme 57. Synthetic route to a fluorinated dipolarophile 264 further used for cycloaddition reactions.
Scheme 57. Synthetic route to a fluorinated dipolarophile 264 further used for cycloaddition reactions.
Molecules 27 05976 sch057
Scheme 58. Cycloaddition reaction using TEBAB 266 as the catalyst.
Scheme 58. Cycloaddition reaction using TEBAB 266 as the catalyst.
Molecules 27 05976 sch058
Scheme 59. Synthesis of 269.
Scheme 59. Synthesis of 269.
Molecules 27 05976 sch059
Scheme 60. Synthesis of 271.
Scheme 60. Synthesis of 271.
Molecules 27 05976 sch060
Scheme 61. Synthesis of 274.
Scheme 61. Synthesis of 274.
Molecules 27 05976 sch061
Scheme 62. Mechanism of formation of the coumarin-indanedione cycloadducts. Reproduced with permission from Ref. [258].
Scheme 62. Mechanism of formation of the coumarin-indanedione cycloadducts. Reproduced with permission from Ref. [258].
Molecules 27 05976 sch062
Scheme 63. Synthesis of 277.
Scheme 63. Synthesis of 277.
Molecules 27 05976 sch063
Scheme 64. Mechanism involved in the cycloaddition reaction with Morita–Baylis–Hillman carbonates 275. Reproduced with permission from Ref. [259].
Scheme 64. Mechanism involved in the cycloaddition reaction with Morita–Baylis–Hillman carbonates 275. Reproduced with permission from Ref. [259].
Molecules 27 05976 sch064
Scheme 65. Synthesis of 281.
Scheme 65. Synthesis of 281.
Molecules 27 05976 sch065
Scheme 66. Mechanism supporting the formation of only one diastereoisomer during the Pd-catalyzed reaction.
Scheme 66. Mechanism supporting the formation of only one diastereoisomer during the Pd-catalyzed reaction.
Molecules 27 05976 sch066
Scheme 67. Synthesis of 285.
Scheme 67. Synthesis of 285.
Molecules 27 05976 sch067
Scheme 68. Synthetic route to spirovinylcyclopropaneindanedione (VCP) 287.
Scheme 68. Synthetic route to spirovinylcyclopropaneindanedione (VCP) 287.
Molecules 27 05976 sch068
Scheme 69. Synthetic route to a five-membered spiroindanedione 289.
Scheme 69. Synthetic route to a five-membered spiroindanedione 289.
Molecules 27 05976 sch069
Scheme 70. Cycloaddition reaction with VCP 287 and cinnamaldehyde 292.
Scheme 70. Cycloaddition reaction with VCP 287 and cinnamaldehyde 292.
Molecules 27 05976 sch070
Scheme 71. Cycloaddition reactions with aryl and naphthyl derivatives 295 and 296.
Scheme 71. Cycloaddition reactions with aryl and naphthyl derivatives 295 and 296.
Molecules 27 05976 sch071
Scheme 72. Mechanism of cyclization determined for cycloaddition reactions occurring with nitroalkenes. Reproduced with permission from Ref. [262].
Scheme 72. Mechanism of cyclization determined for cycloaddition reactions occurring with nitroalkenes. Reproduced with permission from Ref. [262].
Molecules 27 05976 sch072
Scheme 73. Synthesis of 299.
Scheme 73. Synthesis of 299.
Molecules 27 05976 sch073
Scheme 74. Synthesis of 302.
Scheme 74. Synthesis of 302.
Molecules 27 05976 sch074
Scheme 75. Mechanism involved in the synthesis of the oxindole-fused spiropyrazolidine. Reproduced with permission from Ref. [265].
Scheme 75. Mechanism involved in the synthesis of the oxindole-fused spiropyrazolidine. Reproduced with permission from Ref. [265].
Molecules 27 05976 sch075
Scheme 76. Annulation reaction between para-quinone methide 307 and 2-vinylspiro[cyclopropane-1,2′-indene]-1′,3′-dione 289.
Scheme 76. Annulation reaction between para-quinone methide 307 and 2-vinylspiro[cyclopropane-1,2′-indene]-1′,3′-dione 289.
Molecules 27 05976 sch076
Scheme 77. Synthesis of 308.
Scheme 77. Synthesis of 308.
Molecules 27 05976 sch077
Scheme 78. Synthesis of 310.
Scheme 78. Synthesis of 310.
Molecules 27 05976 sch078
Scheme 79. Mechanism proposed to support the formation of 2-arylideneindane-1,3-diones. Reproduced with permission from Ref. [268].
Scheme 79. Mechanism proposed to support the formation of 2-arylideneindane-1,3-diones. Reproduced with permission from Ref. [268].
Molecules 27 05976 sch079
Scheme 80. Synthesis of 315.
Scheme 80. Synthesis of 315.
Molecules 27 05976 sch080
Scheme 81. Mechanism supporting the formation of a unique diastereoisomer during the cycloaddition of azomethine ylides 314 and arylidene-indane-1,3-diones 255. Reproduced with permission from Ref. [269].
Scheme 81. Mechanism supporting the formation of a unique diastereoisomer during the cycloaddition of azomethine ylides 314 and arylidene-indane-1,3-diones 255. Reproduced with permission from Ref. [269].
Molecules 27 05976 sch081
Scheme 82. Synthesis of 317.
Scheme 82. Synthesis of 317.
Molecules 27 05976 sch082
Scheme 83. The two plausible mechanisms supporting the cycloaddition reaction or the epoxidation reaction. Reproduced with permission from Ref. [270].
Scheme 83. The two plausible mechanisms supporting the cycloaddition reaction or the epoxidation reaction. Reproduced with permission from Ref. [270].
Molecules 27 05976 sch083
Scheme 84. Synthesis of 318.
Scheme 84. Synthesis of 318.
Molecules 27 05976 sch084
Scheme 85. Synthetic route to 320.
Scheme 85. Synthetic route to 320.
Molecules 27 05976 sch085
Scheme 86. Cyclotrimerization reaction of 319 and 320.
Scheme 86. Cyclotrimerization reaction of 319 and 320.
Molecules 27 05976 sch086
Scheme 87. Synthesis of 325.
Scheme 87. Synthesis of 325.
Molecules 27 05976 sch087
Scheme 88. Synthetic route to 327.
Scheme 88. Synthetic route to 327.
Molecules 27 05976 sch088
Scheme 89. Synthetic route to 329 and 331.
Scheme 89. Synthetic route to 329 and 331.
Molecules 27 05976 sch089
Scheme 90. Synthetic route to 329 and 334.
Scheme 90. Synthetic route to 329 and 334.
Molecules 27 05976 sch090
Scheme 91. Synthesis of 335.
Scheme 91. Synthesis of 335.
Molecules 27 05976 sch091
Scheme 92. Synthesis of compounds 338 and 339.
Scheme 92. Synthesis of compounds 338 and 339.
Molecules 27 05976 sch092
Scheme 93. Cobalt-catalyzed cycloaddition reactions.
Scheme 93. Cobalt-catalyzed cycloaddition reactions.
Molecules 27 05976 sch093
Scheme 94. Rhodium-catalyzed cycloaddition reactions.
Scheme 94. Rhodium-catalyzed cycloaddition reactions.
Molecules 27 05976 sch094
Scheme 95. Mechanism involved in the Co and Rh-catalyzed cyclization reaction.
Scheme 95. Mechanism involved in the Co and Rh-catalyzed cyclization reaction.
Molecules 27 05976 sch095
Scheme 96. Synthetic route to 346.
Scheme 96. Synthetic route to 346.
Molecules 27 05976 sch096
Scheme 97. Synthesis of 349.
Scheme 97. Synthesis of 349.
Molecules 27 05976 sch097
Scheme 98. Ni-catalyzed cycloaddition reaction furnishing 352.
Scheme 98. Ni-catalyzed cycloaddition reaction furnishing 352.
Molecules 27 05976 sch098
Scheme 99. Base-catalyzed [4+1] cycloaddition.
Scheme 99. Base-catalyzed [4+1] cycloaddition.
Molecules 27 05976 sch099
Scheme 100. Rh-catalyzed cycloaddition reactions.
Scheme 100. Rh-catalyzed cycloaddition reactions.
Molecules 27 05976 sch100
Scheme 101. [4+4] Cycloaddition of indanone containing benzo[c]oxepines providing dibenzocycloooctadiene derivatives 359, 360, 363 and 364.
Scheme 101. [4+4] Cycloaddition of indanone containing benzo[c]oxepines providing dibenzocycloooctadiene derivatives 359, 360, 363 and 364.
Molecules 27 05976 sch101
Scheme 102. Examples of asymmetric cross [10+2] cycloadditions producing 367 starting from 365 and 249.
Scheme 102. Examples of asymmetric cross [10+2] cycloadditions producing 367 starting from 365 and 249.
Molecules 27 05976 sch102
Scheme 103. Synthesis of 370.
Scheme 103. Synthesis of 370.
Molecules 27 05976 sch103
Scheme 104. Mechanism of domino Knoevenagel/Diels–Alder/Epimerization sequence providing 370.
Scheme 104. Mechanism of domino Knoevenagel/Diels–Alder/Epimerization sequence providing 370.
Molecules 27 05976 sch104
Scheme 105. Synthesis of 373.
Scheme 105. Synthesis of 373.
Molecules 27 05976 sch105
Figure 2. Crystal structure of a product used to determine the reaction mechanism. Reproduced with permission of Duan et al. [287].
Figure 2. Crystal structure of a product used to determine the reaction mechanism. Reproduced with permission of Duan et al. [287].
Molecules 27 05976 g002
Scheme 106. Synthesis of 376.
Scheme 106. Synthesis of 376.
Molecules 27 05976 sch106
Scheme 107. Synthesis of 378.
Scheme 107. Synthesis of 378.
Molecules 27 05976 sch107
Scheme 108. Synthesis of 380.
Scheme 108. Synthesis of 380.
Molecules 27 05976 sch108
Scheme 109. Domino reactions carried out with (benzo)thiazoles 382 and 383.
Scheme 109. Domino reactions carried out with (benzo)thiazoles 382 and 383.
Molecules 27 05976 sch109
Scheme 110. Domino reactions involving a Michael addition followed by a 1,3 dipolar cycloaddition of 2-arylidene-1,3-indanediones 255 and 5-aryl-1,3,4-oxathiazol-2-ones 388.
Scheme 110. Domino reactions involving a Michael addition followed by a 1,3 dipolar cycloaddition of 2-arylidene-1,3-indanediones 255 and 5-aryl-1,3,4-oxathiazol-2-ones 388.
Molecules 27 05976 sch110
Scheme 111. Mechanism supporting the formation of the previous compound 395. Reproduced with permission from Ref. [290].
Scheme 111. Mechanism supporting the formation of the previous compound 395. Reproduced with permission from Ref. [290].
Molecules 27 05976 sch111
Scheme 112. Synthesis of 392.
Scheme 112. Synthesis of 392.
Molecules 27 05976 sch112
Scheme 113. Synthetic route to 395.
Scheme 113. Synthetic route to 395.
Molecules 27 05976 sch113
Scheme 114. Mechanism of the domino reaction between ynones 397 and 2-arylidene-indane-1,3-diones 255. Reproduced with permission from Ref. [291].
Scheme 114. Mechanism of the domino reaction between ynones 397 and 2-arylidene-indane-1,3-diones 255. Reproduced with permission from Ref. [291].
Molecules 27 05976 sch114
Scheme 115. Synthesis of spiro-compounds 403 by domino reaction involving a silver-based catalyst.
Scheme 115. Synthesis of spiro-compounds 403 by domino reaction involving a silver-based catalyst.
Molecules 27 05976 sch115
Scheme 116. Mechanism of the domino reaction involving a silver-based catalyst. Reproduced with permission from Ref. [292].
Scheme 116. Mechanism of the domino reaction involving a silver-based catalyst. Reproduced with permission from Ref. [292].
Molecules 27 05976 sch116
Scheme 117. Synthetic access to 404.
Scheme 117. Synthetic access to 404.
Molecules 27 05976 sch117
Scheme 118. Synthesis of 5′-hydroxy-6′-methyl-1′,3′-dihydro-2,2′-spirobi[indene]-1,3-dione 406.
Scheme 118. Synthesis of 5′-hydroxy-6′-methyl-1′,3′-dihydro-2,2′-spirobi[indene]-1,3-dione 406.
Molecules 27 05976 sch118
Scheme 119. Synthesis of 408.
Scheme 119. Synthesis of 408.
Molecules 27 05976 sch119
Scheme 120. Mechanism of the cascade Michael addition/cycloaddition reaction between 2-arylidene-indane-1,3-diones 255 and allenoates 407. Reproduced with permission from Ref. [293].
Scheme 120. Mechanism of the cascade Michael addition/cycloaddition reaction between 2-arylidene-indane-1,3-diones 255 and allenoates 407. Reproduced with permission from Ref. [293].
Molecules 27 05976 sch120
Scheme 121. Synthesis of 411.
Scheme 121. Synthesis of 411.
Molecules 27 05976 sch121
Scheme 122. Mechanism of the domino reaction between 2-arylidene-indane-1,3-diones 255 and N-alkoxyacrylamides 410 in the presence of a base. Reproduced with permission from Ref. [294].
Scheme 122. Mechanism of the domino reaction between 2-arylidene-indane-1,3-diones 255 and N-alkoxyacrylamides 410 in the presence of a base. Reproduced with permission from Ref. [294].
Molecules 27 05976 sch122
Scheme 123. Synthesis of 417.
Scheme 123. Synthesis of 417.
Molecules 27 05976 sch123
Scheme 124. Mechanism involved in the cascade double Michael addition/acetalization reactions. Reproduced with permission from Ref. [295].
Scheme 124. Mechanism involved in the cascade double Michael addition/acetalization reactions. Reproduced with permission from Ref. [295].
Molecules 27 05976 sch124
Scheme 125. Product 423 obtained in a quadruple cascade reaction.
Scheme 125. Product 423 obtained in a quadruple cascade reaction.
Molecules 27 05976 sch125
Scheme 126. Mechanism involved in the quadruple cascade reaction.
Scheme 126. Mechanism involved in the quadruple cascade reaction.
Molecules 27 05976 sch126
Scheme 127. Mechanism involved in the synthesis of spiro-indane-1,3-diones 428.
Scheme 127. Mechanism involved in the synthesis of spiro-indane-1,3-diones 428.
Molecules 27 05976 sch127
Scheme 128. Synthesis of 428.
Scheme 128. Synthesis of 428.
Molecules 27 05976 sch128
Scheme 129. Examples of compounds 437 and 438 obtained during the MRC of indane-1,3-dione 4, dimethyl but-2-ynedioate 434 and various substituted benzothiazoles 435 or 436.
Scheme 129. Examples of compounds 437 and 438 obtained during the MRC of indane-1,3-dione 4, dimethyl but-2-ynedioate 434 and various substituted benzothiazoles 435 or 436.
Molecules 27 05976 sch129
Scheme 130. Mechanism of MRC between the MRC between indane-1,3-dione 4, dimethyl but-2-ynedioate 434 and benzothiazoles 437 or 438.
Scheme 130. Mechanism of MRC between the MRC between indane-1,3-dione 4, dimethyl but-2-ynedioate 434 and benzothiazoles 437 or 438.
Molecules 27 05976 sch130
Scheme 131. Synthesis of spiro-N-fused indane-1,3-diones 441 and 442.
Scheme 131. Synthesis of spiro-N-fused indane-1,3-diones 441 and 442.
Molecules 27 05976 sch131
Scheme 132. Synthesis of 441 and 442.
Scheme 132. Synthesis of 441 and 442.
Molecules 27 05976 sch132
Scheme 133. Synthesis of 445.
Scheme 133. Synthesis of 445.
Molecules 27 05976 sch133
Scheme 134. Synthesis of 448 and 450.
Scheme 134. Synthesis of 448 and 450.
Molecules 27 05976 sch134
Scheme 135. Synthesis of 452.
Scheme 135. Synthesis of 452.
Molecules 27 05976 sch135
Scheme 136. Mechanism involved in the MRC reaction using Fe-particles as catalysts. Reproduced with permission from Ref. [301].
Scheme 136. Mechanism involved in the MRC reaction using Fe-particles as catalysts. Reproduced with permission from Ref. [301].
Molecules 27 05976 sch136
Scheme 137. Example of pH-switchable compound 454.
Scheme 137. Example of pH-switchable compound 454.
Molecules 27 05976 sch137
Scheme 138. Various products obtained by reacting indane-1,3-dione 4 in the presence of base.
Scheme 138. Various products obtained by reacting indane-1,3-dione 4 in the presence of base.
Molecules 27 05976 sch138
Scheme 139. Synthetic routes to 50 and 458.
Scheme 139. Synthetic routes to 50 and 458.
Molecules 27 05976 sch139
Scheme 140. Synthetic route to 455.
Scheme 140. Synthetic route to 455.
Molecules 27 05976 sch140
Scheme 141. Synthetic route to 457.
Scheme 141. Synthetic route to 457.
Molecules 27 05976 sch141
Scheme 142. Synthesis of 463.
Scheme 142. Synthesis of 463.
Molecules 27 05976 sch142
Scheme 143. Mechanism occurring during alkene hydrofunctionnalizations of 2-arylidene-indane-1,3-diones 255 providing 463. Reproduced with permission from Ref. [304].
Scheme 143. Mechanism occurring during alkene hydrofunctionnalizations of 2-arylidene-indane-1,3-diones 255 providing 463. Reproduced with permission from Ref. [304].
Molecules 27 05976 sch143
Scheme 144. Synthesis of 465.
Scheme 144. Synthesis of 465.
Molecules 27 05976 sch144
Scheme 145. Mechanism involved in the copper-catalyzed synthesis of spiro compounds. Reproduced with permission from Ref. [305].
Scheme 145. Mechanism involved in the copper-catalyzed synthesis of spiro compounds. Reproduced with permission from Ref. [305].
Molecules 27 05976 sch145
Scheme 146. Synthesis of 467.
Scheme 146. Synthesis of 467.
Molecules 27 05976 sch146
Scheme 147. Mechanism occurring in the copper catalyzed cycloaddition reaction of 2-arylene-indane-1,3-dione 255 with various oximes 466. Reproduced with permission from Ref. [306].
Scheme 147. Mechanism occurring in the copper catalyzed cycloaddition reaction of 2-arylene-indane-1,3-dione 255 with various oximes 466. Reproduced with permission from Ref. [306].
Molecules 27 05976 sch147
Scheme 148. The different products obtained during the reaction of 2-arylidene-indane-1,3-diones 255 and the Seyferth–Gilbert reagent 468.
Scheme 148. The different products obtained during the reaction of 2-arylidene-indane-1,3-diones 255 and the Seyferth–Gilbert reagent 468.
Molecules 27 05976 sch148
Scheme 149. Synthesis of 469.
Scheme 149. Synthesis of 469.
Molecules 27 05976 sch149
Scheme 150. Chemical structures of indane-1,3-dione-based push–pull dyes 470–497.
Scheme 150. Chemical structures of indane-1,3-dione-based push–pull dyes 470–497.
Molecules 27 05976 sch150
Scheme 151. Chemical structures of dyes 498–506.
Scheme 151. Chemical structures of dyes 498–506.
Molecules 27 05976 sch151
Scheme 152. Reaction occurring with cyanide anions.
Scheme 152. Reaction occurring with cyanide anions.
Molecules 27 05976 sch152
Scheme 153. Reaction occurring with cyanide anions.
Scheme 153. Reaction occurring with cyanide anions.
Molecules 27 05976 sch153
Figure 3. Cyanide ion detection with 509: (a) optically with the naked eye; (b) by fluorescence changes. Reproduced with permission from Wang et al. [359].
Figure 3. Cyanide ion detection with 509: (a) optically with the naked eye; (b) by fluorescence changes. Reproduced with permission from Wang et al. [359].
Molecules 27 05976 g003
Scheme 154. Chemical structures of 510–518.
Scheme 154. Chemical structures of 510–518.
Molecules 27 05976 sch154
Scheme 155. Synthesis of 4-hydroxyindan-1,3-diones 521 studied in the 1930s by Robinson et al. and Walker et al., first indanediones known for their antiseptic activities [388].
Scheme 155. Synthesis of 4-hydroxyindan-1,3-diones 521 studied in the 1930s by Robinson et al. and Walker et al., first indanediones known for their antiseptic activities [388].
Molecules 27 05976 sch155
Scheme 156. Examples of hydroxyindanediones 522–524 with important antibacterial activities against Bacterium typhosum.
Scheme 156. Examples of hydroxyindanediones 522–524 with important antibacterial activities against Bacterium typhosum.
Molecules 27 05976 sch156
Scheme 157. The two families of indeno [1,2-c]pyrazoles examined for their antimicrobial activities.
Scheme 157. The two families of indeno [1,2-c]pyrazoles examined for their antimicrobial activities.
Molecules 27 05976 sch157
Scheme 158. Synthetic routes to 3-aryl-1-heteroarylindeno [1,2-c]pyrazol-4(1H)-ones 532.
Scheme 158. Synthetic routes to 3-aryl-1-heteroarylindeno [1,2-c]pyrazol-4(1H)-ones 532.
Molecules 27 05976 sch158
Scheme 159. Synthesis of ethyl 4-(9-ethyl-9H-carbazol-3-yl)-2-methyl-5-oxo-4,5-dihydro-1H-indeno [1,2-b]pyridine-3-carboxylate (ECPC) 536.
Scheme 159. Synthesis of ethyl 4-(9-ethyl-9H-carbazol-3-yl)-2-methyl-5-oxo-4,5-dihydro-1H-indeno [1,2-b]pyridine-3-carboxylate (ECPC) 536.
Molecules 27 05976 sch159
Scheme 160. Synthesis of pyrimidine-2-thiones 538 starting from indane-1,3-dione 4 along with their corresponding antibacterial and antifungal activities.
Scheme 160. Synthesis of pyrimidine-2-thiones 538 starting from indane-1,3-dione 4 along with their corresponding antibacterial and antifungal activities.
Molecules 27 05976 sch160
Scheme 161. Strategy used for the synthesis of spiro[indolo-3,10′-indeno[1,2-b]quinolin]-2,4,11′-triones 540-(IVa-IVv) along with their graphical representations of the diameter of growth of inhibition (mm) against bacteria strains. Reproduced with permission from Ref. [397].
Scheme 161. Strategy used for the synthesis of spiro[indolo-3,10′-indeno[1,2-b]quinolin]-2,4,11′-triones 540-(IVa-IVv) along with their graphical representations of the diameter of growth of inhibition (mm) against bacteria strains. Reproduced with permission from Ref. [397].
Molecules 27 05976 sch161
Scheme 162. MCR leading to two different structures of indeno-heterocycles 541. Apoptosis properties were evaluated at 25 µM.
Scheme 162. MCR leading to two different structures of indeno-heterocycles 541. Apoptosis properties were evaluated at 25 µM.
Molecules 27 05976 sch162
Scheme 163. Camptothecin-inspired pentacycle-based indeno-heterocycles 544 and 545.
Scheme 163. Camptothecin-inspired pentacycle-based indeno-heterocycles 544 and 545.
Molecules 27 05976 sch163
Scheme 164. Synthetic routes to indenopyridine derivatives 549 examined in Ghorab’s study.
Scheme 164. Synthetic routes to indenopyridine derivatives 549 examined in Ghorab’s study.
Molecules 27 05976 sch164
Scheme 165. Synthetic routes to indenopyrazoles and the best candidate 553-k.
Scheme 165. Synthetic routes to indenopyrazoles and the best candidate 553-k.
Molecules 27 05976 sch165
Scheme 166. “Push–pull” effect in IND-TPA with TICT schematisation used in this study.
Scheme 166. “Push–pull” effect in IND-TPA with TICT schematisation used in this study.
Molecules 27 05976 sch166
Figure 4. (A) Generation of ROS in PBS buffer. (B) Cell viability of HeLa living cells, stained with MPAT, upon irradiation with a green light for 10 min. Reproduced with permission from Ref. [419].
Figure 4. (A) Generation of ROS in PBS buffer. (B) Cell viability of HeLa living cells, stained with MPAT, upon irradiation with a green light for 10 min. Reproduced with permission from Ref. [419].
Molecules 27 05976 g004
Scheme 167. Chemical structure of 557 and the different advantages of this AIE dye.
Scheme 167. Chemical structure of 557 and the different advantages of this AIE dye.
Molecules 27 05976 sch167
Figure 5. CLSM images of HCC827 (AE) and A549 (FJ) cells after incubation with 557 (5 mM) and BODIPY493/503 (100 nM) at 37 °C for 15 min. (A,F) Bright-field images. (B,G) Fluorescence image from 557 and from BODIPY493/503. (D,I) The merged images. (E,J) The intensity profile of ROI lines. Scale bar = 20 mm. Reproduced with permission of Gao M, Su H, Lin Y, Ling X, Li S, Qin A, and Zhong Tang B. Photoactivatable aggregation-induced emission probes for lipid droplets-specific live cell imaging. Reproduced with permission from Ref. [421].
Figure 5. CLSM images of HCC827 (AE) and A549 (FJ) cells after incubation with 557 (5 mM) and BODIPY493/503 (100 nM) at 37 °C for 15 min. (A,F) Bright-field images. (B,G) Fluorescence image from 557 and from BODIPY493/503. (D,I) The merged images. (E,J) The intensity profile of ROI lines. Scale bar = 20 mm. Reproduced with permission of Gao M, Su H, Lin Y, Ling X, Li S, Qin A, and Zhong Tang B. Photoactivatable aggregation-induced emission probes for lipid droplets-specific live cell imaging. Reproduced with permission from Ref. [421].
Molecules 27 05976 g005
Figure 6. I/I0 (%) of fluorescence intensity of HCC827 cells colored with 557 (5 µM) with increasing time of irradiation at 514 nm with 7% laser power. Inset: Fluorescence images of HCC827 cells with increasing time of irradiation. Reproduced with permission from Gao M, Su H, Lin Y, Ling X, Li S, Qin A, Zhong Tang B. Photoactivatable aggregation-induced emission probes for lipid droplets-specific live cell imaging. Reproduced with permission from Ref. [421].
Figure 6. I/I0 (%) of fluorescence intensity of HCC827 cells colored with 557 (5 µM) with increasing time of irradiation at 514 nm with 7% laser power. Inset: Fluorescence images of HCC827 cells with increasing time of irradiation. Reproduced with permission from Gao M, Su H, Lin Y, Ling X, Li S, Qin A, Zhong Tang B. Photoactivatable aggregation-induced emission probes for lipid droplets-specific live cell imaging. Reproduced with permission from Ref. [421].
Molecules 27 05976 g006
Scheme 168. Series of molecules 561–564 synthetized by MWAMCR.
Scheme 168. Series of molecules 561–564 synthetized by MWAMCR.
Molecules 27 05976 sch168
Scheme 169. Plausible mechanism of the above molecules in the presence of Zn2+ cations.
Scheme 169. Plausible mechanism of the above molecules in the presence of Zn2+ cations.
Molecules 27 05976 sch169
Scheme 170. Synthetic route to arylindenopyrimidines 565.
Scheme 170. Synthetic route to arylindenopyrimidines 565.
Molecules 27 05976 sch170
Scheme 171. Summary of the synthetic strategy developed to access compounds 577 and 584 substituted at the 8- and 9-positions.
Scheme 171. Summary of the synthetic strategy developed to access compounds 577 and 584 substituted at the 8- and 9-positions.
Molecules 27 05976 sch171aMolecules 27 05976 sch171b
Scheme 172. Comparisons between six compounds substituted at the 8- and the 9-positions for their in vitro and in vivo activities. In vitro activity for A2a and A1 functional assays and in vivo results for mouse catalepsy at 10 mg/kg, po.
Scheme 172. Comparisons between six compounds substituted at the 8- and the 9-positions for their in vitro and in vivo activities. In vitro activity for A2a and A1 functional assays and in vivo results for mouse catalepsy at 10 mg/kg, po.
Molecules 27 05976 sch172
Scheme 173. Chemical structure of JNJ-40255293 (585).
Scheme 173. Chemical structure of JNJ-40255293 (585).
Molecules 27 05976 sch173
Scheme 174. Strategy employed for the synthesis of arylindenopyridines 589 and 593 in the patents.
Scheme 174. Strategy employed for the synthesis of arylindenopyridines 589 and 593 in the patents.
Molecules 27 05976 sch174
Scheme 175. Synthesis of tricyclic 3,4-dihydropyrimidine derivatives 595 via Biginelli reaction along with the most promising compounds. a human TRPA1 antagonism b rat TRPA1 antagonism.
Scheme 175. Synthesis of tricyclic 3,4-dihydropyrimidine derivatives 595 via Biginelli reaction along with the most promising compounds. a human TRPA1 antagonism b rat TRPA1 antagonism.
Molecules 27 05976 sch175
Scheme 176. MCR for the synthesis of 222 with three examples having decent anticonvulsant activity. a, Values represent means SEM (n = 3). b, Rotarod toxicity (number of animals exhibiting toxicity/number of animals tested).
Scheme 176. MCR for the synthesis of 222 with three examples having decent anticonvulsant activity. a, Values represent means SEM (n = 3). b, Rotarod toxicity (number of animals exhibiting toxicity/number of animals tested).
Molecules 27 05976 sch176
Scheme 177. Synthetic route to dihydropyrimidine.
Scheme 177. Synthetic route to dihydropyrimidine.
Molecules 27 05976 sch177
Figure 7. Close-up depiction of the lowest-energy three-dimensional (3-D) docking poses of 600 into the binding site of Torpedo californica acetylcholinesterase TcAChE. Reproduced with permission from Ref. [445].
Figure 7. Close-up depiction of the lowest-energy three-dimensional (3-D) docking poses of 600 into the binding site of Torpedo californica acetylcholinesterase TcAChE. Reproduced with permission from Ref. [445].
Molecules 27 05976 g007
Scheme 178. Chemical structures of indane-2-arylhydrazinylmethylene-1,3-diones 605a–f and indol-2-aryldiazenylmethylene-3-ones 608a–m.
Scheme 178. Chemical structures of indane-2-arylhydrazinylmethylene-1,3-diones 605a–f and indol-2-aryldiazenylmethylene-3-ones 608a–m.
Molecules 27 05976 sch178
Scheme 179. Chemical structures of various indane-1,3-dione derivatives 609–611 with anticoagulant properties.
Scheme 179. Chemical structures of various indane-1,3-dione derivatives 609–611 with anticoagulant properties.
Molecules 27 05976 sch179
Scheme 180. The different synthetic routes to acylindane-1,3-diones 613.
Scheme 180. The different synthetic routes to acylindane-1,3-diones 613.
Molecules 27 05976 sch180
Scheme 181. New synthetic route developed by Larsen et al. to access to acylindanediones 616.
Scheme 181. New synthetic route developed by Larsen et al. to access to acylindanediones 616.
Molecules 27 05976 sch181
Table 1. Reaction yields obtained for the synthesis of 95–114.
Table 1. Reaction yields obtained for the synthesis of 95–114.
compounds9596979899100101102103104
reaction yields88848874948992858488
compounds105106107108109110111112113114
reaction yields74857582928781788985
Table 2. Examples of cycloaddition reactions performed with various arylidene indane-1,3-diones 255 and azomethine imines 254.
Table 2. Examples of cycloaddition reactions performed with various arylidene indane-1,3-diones 255 and azomethine imines 254.
Ar1Ar2Yield adr b
C6H5C6H598>20:1
4-FC6H4C6H575>20:1
4-ClC6H4C6H585>20:1
4-BrC6H4C6H584>20:1
4-CF3C6H4C6H5924:1
4-MeC6H4C6H583>20:1
3-BrC6H4C6H565>20:1
3-NO2C4H4C6H571>20:1
3-MeC6H4C6H593>20:1
3-MeOC6H4C6H5827:1
2-C4H3SC6H570>20:1
a Yield of isolated product; b Determined by 1H NMR spectroscopy on the crude mixture.
Table 3. Examples of chromeno [3,4-b]pyrrolidines 259 obtained using DMAP as the base.
Table 3. Examples of chromeno [3,4-b]pyrrolidines 259 obtained using DMAP as the base.
R1R2t (h)Yield (%) a
5-BrH1869
HH1862
5-ClH1270
5-NO2H3058
5-OMeH1859
4-OMeH2471
3-OMeH1879
5-FH1146 b
5-Br5-Cl1875
5-Br5-Br1272
5-Br5-NO21863
5-Br5-OMe2472
5-Br4-OMe1861
5-Br3-OMe4848
a Isolated yield; b 17% of the other product was obtained.
Table 4. Examples of chromeno [3,4-c]pyrrolidines 262 obtained using TMG as the base.
Table 4. Examples of chromeno [3,4-c]pyrrolidines 262 obtained using TMG as the base.
R1R2t (h)Yield (%) a
5-BrH479
HH565
5-ClH471
5-NO2H450
5-OMeH2426
4-OMeH1510
3-OMeH1547
5-FH553
5-Br5-Cl675
5-Br5-Br479
5-Br5-NO24856
5-Br5-OMe570
5-Br4-OMe876
5-Br3-OMe469
a Isolated yield.
Table 5. Reaction yields obtained during the cycloaddition reactions using TEBAB 266 as the catalyst.
Table 5. Reaction yields obtained during the cycloaddition reactions using TEBAB 266 as the catalyst.
R1Ar1Ar2Yield (%) adr b
HC6H5C6H598100:0:0:0
H4-BrC6H4C6H570100:0:0:0
H4-OMeC6H4C6H53296:4:0:0
H4-NO2C6H4C6H573100:0:0:0
H4-ClC6H4C6H568100:0:0:0
H4-PhC6H4C6H584100:0:0:0
HC10H7C6H572100:0:0:0
HC4H3OC6H561100:0:0:0
HC4H3SC6H556100:0:0:0
FC6H5C6H54050:50:0:0
NO2C6H5C6H54660:40:0:0
HC6H54-ClC6H469100:0:0:0
HC6H54-NO2C6H474100:0:0:0
HC6H53,5-(CF3)2C6H331100:0:0:0
a Isolated yield; b determined by 1H and 19F NMR.
Table 6. Reaction yields obtained during the formation of spiro-indanedione derivatives 271.
Table 6. Reaction yields obtained during the formation of spiro-indanedione derivatives 271.
RYield (%) a
C6H594
4-FC6H493
2-ClC6H488
4-ClC6H489
2-BrC6H491
4-BrC6H492
4-CNC6H490
4-NO2C6H491
4-MeC6H495
4-OMeC6H496
3,4,5-Me3C6H291
2-C10H792
2-C4H3O88
2-C4H3S91
C6H1365
a Reaction yield determined after purification and column chromatography.
Table 7. Reaction yields obtained during the synthesis of coumarin-indanedione derivatives 274.
Table 7. Reaction yields obtained during the synthesis of coumarin-indanedione derivatives 274.
R1R2R3t (d)Yield (%)ee a
C6H5HC6H518996
4-NO2C6H4HC6H518492
4-CNC6H4HC6H529094
4-ClC6H4HC6H528894
4-BrC6H4HC6H518593
4-MeC6H4HC6H539293
4-MeOC6H4HC6H548491
4-HOC6H4HC6H577069
2-MeOC6H4HC6H579388
2-HOC6H4HC6H554225
2-BrC6H4HC6H528891
C4H3OHC6H547389
C4H3SHC6H557688
C5H4NHC6H518277
C6H5H4-ClC6H429095
C6H5HCH338593
C6H54-ClC6H519291
C6H54-BrC6H528295
C6H54-MeOC6H518692
C6H52,4-Cl2C6H528790
C6H52-MeOC6H529593
a Enantiomeric excess determined by HPLC analyses on a chiral stationary phase.
Table 8. Reaction yields obtained during asymmetric cycloadditions between 2-arylidene-indane-1,3-diones 255 and Morita–Baylis–Hillman carbonates 275.
Table 8. Reaction yields obtained during asymmetric cycloadditions between 2-arylidene-indane-1,3-diones 255 and Morita–Baylis–Hillman carbonates 275.
R1R2Yield (%) aee (%) b
4-NO2C6H44-Cl5288
4-NO2C6H44-Br5192
4-NO2C6H44-CN6591
4-NO2C6H44-CF35091
4-NO2C6H42-Cl6897
4-NO2C6H42-Br5097
4-NO2C6H43-NO27587
4-NO2C6H43-Cl6091
4-NO2C6H43-Br5190
4-NO2C6H42,4-Cl26296
4-NO2C6H42,3-Cl26798
4-NO2C6H43,4-Cl25492
4-NO2C6H42,6-Cl23092
4-NO2C6H44-F, 3-Br5090
4-NO2C6H44-CH3Trace-
C6H54-NO275 c66
3-BrC6H44-NO264 c67
4-C6H44-NO275 c65
C6H114-NO268 c67
a Isolated yield; b determined by Chiral HPLC; c the reaction was carried out within 48 h.
Table 9. Reaction yields obtained during the [3+2] cycloaddition of vinylaziridine 280 and indane-1,3-dione derivatives 255.
Table 9. Reaction yields obtained during the [3+2] cycloaddition of vinylaziridine 280 and indane-1,3-dione derivatives 255.
RYield (%)erdr
C6H59792:83:1
4-MeC6H49875:253:1
4-BrC6H49077:233:1
4-NO2C6H49880:203:1
4-CF3C6H49387:133:1
C10H79984:1610:1
3-ClC6H49072:283:1
3-MeOC6H49083:173:1
C5H4N8175:253:1
C4H3O5997:34:1
CH2C6H59174:263:1
Table 10. Reaction yields obtained during the cyclization reaction with vinylethylene carbonate 282 and arylidene-indane-1,3-diones 255.
Table 10. Reaction yields obtained during the cyclization reaction with vinylethylene carbonate 282 and arylidene-indane-1,3-diones 255.
R1R2Yield (%) adr bee c
C6H5C6H59953:4799.98
3-MeC6H4C6H59952:4899.95
4-MeC6H4C6H59751:4996.95
2-MeOC6H4C6H58254:4699.95
3-MeOC6H4C6H58453:4799.98
4-MeOC6H4C6H59354:4696.89
2-FC6H4C6H59653:4793.92
3-FC6H4C6H58656:4497.94
4-FC6H4C6H59652:4894.92
3-ClC6H4C6H59452:4893.95
4-ClC6H4C6H57651:4996.95
3-BrC6H4C6H57752:4893.89
4-BrC6H4C6H56850:5095.95
4-PhC6H4C6H57153:4793.95
C6H53-MeC6H49456:4496.97
C6H52-MeOC6H47364:3684.88
C6H53-MeOC6H49962:3896.96
C6H54-MeOC6H49359:4198.97
C6H53,4-(MeO)2C6H47981:1999.95
C6H52-FC6H48968:3291.90
C6H53-FC6H49168:3296.90
C6H54-FC6H48953:4794.92
C6H53-CF3C6H49073:2793.91
C6H54-CF3C6H49959:4197.81
C6H52-C4H3O9763:3799.99
C6H54-Br-2-C4H2S9976:2499.96
a Isolated yield; b dr was determined by 1H NMR analysis of the product; c determined by chiral HPLC analysis.
Table 11. Reaction yields obtained with various nitroalkene derivatives.
Table 11. Reaction yields obtained with various nitroalkene derivatives.
R.Yield (%) adr bee c
C6H5805:197:95
4-FC6H4875:196:95
4-BrC6H4844.4:192:80
2-CF3C6H47514:199:55
2-FC6H4826.4:196:89
4-MeOC6H4925.5:198:90
2-F-6-ClC6H3855.3:197:90
2-MeOC6H4902.6:197:97
4-MeC6H4815:196:88
2-BrC6H4794.1:192:84
4-ClC6H4905.3:192:68
1-C10H7832.3:197:97
2-C10H7815.7:198:80
C4H3S806:197:87
C4H3O895.7:196:78
C3H4861.7:199:99
C6H11881.2:199:99
a Isolated yield of diastereoisomer; b the diastereoisomeric ratios were determined by 1H NMR spectroscopy; c the enantiomeric excess values were determined by chiral HPLC analyses.
Table 12. Reaction yields obtained during the synthesis of oxindole-fused spiropyrazolidine compounds.
Table 12. Reaction yields obtained during the synthesis of oxindole-fused spiropyrazolidine compounds.
R1R2Yield % aee % b
HCH2C6H59578
5-FCH2C6H59964
5-ClCH2C6H59278
5-ICH2C6H59869
5-MeCH2C6H59982
6-MeCH2C6H57859
6-MeOCH2C6H59677
6-ClCH2C6H59665
7-CF3CH2C6H55268
5,7-Me2CH2C6H58048
5,7-Cl2CH2C6H53975
HCH2OMe6577
HMe8184
a Isolated yield; b the enantiomeric excess was determined by chiral HPLC analysis on chiral cel IB-3.
Table 13. Reaction yields obtained during cycloaddition reactions using diphenylphosphoric acid 306 and a Pd (0) catalyst.
Table 13. Reaction yields obtained during cycloaddition reactions using diphenylphosphoric acid 306 and a Pd (0) catalyst.
R1R2Yield (%) adr b
C10H7C6H5753:1
3,4-(OMe)2C6H4C6H5931:4.2
4-BrC6H4C6H5803.5:1
3-ClC6H4C6H58610.0:1
C4H9C6H5849.5:1
C6H54-NO2C6H4814.8:1
C6H54-OMeC6H4771:1.2
C6H53-MeC6H4791.6:1
C6H54-ClC6H4CH2501:9.2
a Yields determined after purification by chromatography on silica gel; b diastereomeric ratio determined after purification by chromatography.
Table 14. [3+2] Cycloadditions of 2-arylideneindane-1,3-diones 255 with ketoxime acetates 309 using a copper catalyst.
Table 14. [3+2] Cycloadditions of 2-arylideneindane-1,3-diones 255 with ketoxime acetates 309 using a copper catalyst.
Ar1Ar2Yield (%) a
C6H5C6H595
2-BrC6H5C6H592
2-NO2C6H5C6H578
4-FC6H4C6H585
4-CF3C6H4C6H588
4-MeC6H4C6H581
4-MeOC6H4C6H561
3-BrC6H4C6H595
C10H7C6H551
C4H3OC6H556
C4H3SC6H554
C3H7C6H521
C6H52-BrC6H482
C6H54-MeC6H485
C6H54-MeOC6H486
C6H54-NO2C6H486
4-FC6H43-BrC6H493
3-BrC6H44-ClC6H484
4-BrC6H44-MeC6H487
3-NO2C6H44-MeOC6H484
2-ClC6H44-MeOC6H492
C6H5C4H3S77
a Isolated yield.
Table 15. Reaction yields obtained during the cycloaddition of azomethine ylides and arylidene-indane-1,3-diones.
Table 15. Reaction yields obtained during the cycloaddition of azomethine ylides and arylidene-indane-1,3-diones.
Ar1RAr2Yield (%) aee (%) b
C6H5Me4-BrC6H456−87
C6H5Et4-BrC6H460−74
2,4-Cl2C6H3Et4-BrC6H485−70
2-MeC6H4Me4-BrC6H479−73
1-Br-2-C10H6Me4-BrC6H489−81
1-Br-2-C10H6Et3-NO2C6H473−81
2,4-F2C6H3Et3-NO2C6H489−76
3-MeC6H4Me3-NO2C6H475−76
2-MeC6H4Et3-NO2C6H476−72
4-ClC6H5Et3-NO2C6H45564
4-BrC6H4Et3-NO2C6H476−77
2-ClC6H4Et2-NO2C6H459−79
4-NO2C6H4MeC6H590−74
4-ClC6H4Me2-C4H3O65−78
a yield of the isolated product; b determined by Chiral HPLC.
Table 16. Reaction yields obtained during the [3+2] cycloadditions of 2-arylidene-indane-1,3-diones 255 with methyl-2-(3,4-dihydroisoquinolin-2-yl) acetate 316.
Table 16. Reaction yields obtained during the [3+2] cycloadditions of 2-arylidene-indane-1,3-diones 255 with methyl-2-(3,4-dihydroisoquinolin-2-yl) acetate 316.
Ar.RYield (%) a
4-OMeC6H4CO2Me81
4-MeC6H4CO2Me88
C6H5CO2Me92
3-ClC6H4CO2Me92
3-NO2C6H4CO2Me70
4-ClC6H4CO2Me73
4-OMeC6H4CO2Et56
4-MeC6H4CO2Et72
C6H5CO2Et90
3-FC6H4CO2Et68
3-ClC6H4CO2Et70
3-NO2C6H4CO2Et65
4-ClC6H4CO2Et78
4-BrC6H4CO2Et90
4-NO2C6H4CO2Et75
C6H5CO2tBu86
2-ClC6H4CO2tBu55
3-ClC6H4CO2tBu60
3-FC6H4CO2tBu63
3-NO2C6H4CO2tBu54
4-BtC6H4CO2tBu62
4-NO2C6H4CO2tBu63
C6H5CN73
4-MeC6H4CN48
3-ClC6H4CN37
4-ClC6H4CN55
4-BrC6H4CN57
a yield of the isolated product.
Table 17. Reaction yields obtained during the synthesis of various 3′-arylspiro[indene-2,2′-oxirane]-1,3-diones 318.
Table 17. Reaction yields obtained during the synthesis of various 3′-arylspiro[indene-2,2′-oxirane]-1,3-diones 318.
Ar.Yield (%) a
4-MeC6H470
3-OMeC6H468
2-ClC6H492
2-BrC6H481
3-FC6H473
3-ClC6H489
4-ClC6H456
4-BrC6H472
a yield of the isolated product.
Table 18. Reaction yields obtained during the synthesis of spiroindanediones 325.
Table 18. Reaction yields obtained during the synthesis of spiroindanediones 325.
R1R2Yield (%) a
CH2BrCH2Br72
CH2ClH80
CH2ClCH2Cl79
a yield after purification by column chromatography.
Table 19. Reaction yields obtained during cycloaddition reactions of 2-arylidene-indane-1,3-dione 255 and cyclic carbonate 347.
Table 19. Reaction yields obtained during cycloaddition reactions of 2-arylidene-indane-1,3-dione 255 and cyclic carbonate 347.
Rt (h)Yield (%) aee (%) b
C6H5128797
2-FC6H4149998
3-FC6H4486297
4-FC6H4109497
2-ClC6H456298
3-ClC6H4369498
4-ClC6H4249998
2,4-Cl2C6H349098
2-MeC6H459998
3-MeC6H4129996
4- MeC6H4128395
4-EtC6H4126497
2-MeOC6H4119997
3-MeOC6H4149999
4- MeOC6H4149994
1-C10H6209999
2-C4H3S24n.r.-
n.r., no reaction; a yield of isolated product; b determined by HPLC analysis using a chiral stationary phase.
Table 20. Examples of a domino Knoevenagel/Diels–Alder/Epimerization sequence.
Table 20. Examples of a domino Knoevenagel/Diels–Alder/Epimerization sequence.
R1R2Yield (%)dr cis:trans
C6H54-NO2C6H497≥99:1
C6H54-MeOC6H4716:1
C6H54-HOC6H4≥99≥99:1
C6H54-ClC6H4≥99≥99:1
C6H52-NO2C6H480≥99:1
C6H54-CNC6H4≥99≥99:1
C6H54-CO2MeC6H4≥99≥99:1
C6H5C10H7≥99≥99:1
C6H5C4H3O≥9910:1
C6H5C4H3S5713:1
C6H5C4H4N30≥99:1
C6H5C2H2C6H574≥99:1
C6H5C6H593≥99:1
C10H7C10H795≥99:1
C4H3SC4H3S93≥99:1
C4H3OC4H3O60≥99:1
4-MeOC6H44-MeOC6H495≥99:1
C7H5O2C7H5O298≥99:1
4-Me2NC6H44-Me2NC6H490≥99:1
4-HOC6H44-HOC6H495≥99:1
4-ClC6H44-ClC6H498≥99:1
4-NO2C6H44-NO2C6H498≥99:1
4-CNC6H44-CNC6H485≥99:1
4-CO2MeC6H44-CO2MeC6H483≥99:1
Table 21. Domino reaction involving a sulfa Michael/Michael sequence.
Table 21. Domino reaction involving a sulfa Michael/Michael sequence.
ArYield (%) bee (%) cdr cis:trans
C6H5869994.8:5.2
4-FC6H4819995.3:4.7
4-MeOC6H4809895.6:4.4
3-BrC6H4789791.6:8.4
3-NO2C6H4 a859184.4:15.6
3-MeC6H4839895.0:5.0
2-C4H3O789792.1:7.9
a The reaction time was 48 h; b yield of a mixture of two isolated isomers; c determined by HPLC analysis on a chiral stationary phase; ee of the major diastereoisomer.
Table 22. Reaction yields determined during the synthesis of spiro tetrahydrothiophene-indan-1,3-diones 376 starting from 1,4-dithiane-2,5-diol 374.
Table 22. Reaction yields determined during the synthesis of spiro tetrahydrothiophene-indan-1,3-diones 376 starting from 1,4-dithiane-2,5-diol 374.
RTime (h)Yield (%) adr cis:trans bee c
C6H53969:168
C10H73938:165
4-NO2C6H42997.5:156
4-FC6H41969:166
4-BrC6H45971.3:168
2-ClC6H42931.5:172
4-MeC6H42.5989:171
3-MeOC6H43889:160
3,4-(OCH2O)C6H34.5969:174
3-C4H3S24989:172
a Yield of isolated product after column chromatography; b diastereoisomeric ratio (cis to trans) was determined by HPLC of the acylated product; c enantiomeric excess of the major diastereoisomer was determined by chiral HPLC analysis of the acylated product.
Table 23. Domino reactions involving a Knoevenagel condensation/1,3 dipolar cycloaddition sequence.
Table 23. Domino reactions involving a Knoevenagel condensation/1,3 dipolar cycloaddition sequence.
ArYield (%) arr b
4-MeC6H478≥95:5
4-MeOC6H483≥95:5
2-C10H774≥95:5
4-ClC6H479≥95:5
4-BrC6H480≥95:5
a Isolated yield; b determined by 1H NMR analysis of the crude reaction mixture.
Table 24. Domino reactions carried out with benzoimidazoles 379.
Table 24. Domino reactions carried out with benzoimidazoles 379.
ArXYield (%)
4-BrC6H4NH78
4-MeOC6H4NH80
4-ClC6H4S76
4-MeOC6H4S85
Table 25. Reaction yields obtained during domino reaction of ynones 391 and 2-arylidene-indane-1,3-diones 255.
Table 25. Reaction yields obtained during domino reaction of ynones 391 and 2-arylidene-indane-1,3-diones 255.
R1R2R3t (h)Yield (%) a
C6H5HC6H51590
C6H5H2-BrC6H41093
C6H5H3-BrC6H41096
C6H5H4-BrC6H41090
C6H5H2-MeC6H41091
C6H5H3-MeC6H41095
C6H5H4-MeC6H41094
C6H5H4-ClC6H41288
C6H5H4-FC6H41287
C6H5H4-MeOC6H4390 b
C6H5H4-NO2C6H42473
C6H5H2,4-Cl2C6H31262
C6H5H2-C4H3O4887
C6H5H2-C4H3S5384
C6H5H1-C10H72096
C6H5HC2H4C6H43589
C6H5Me4-BrC6H41895
C6H5Di-Me4-BrC6H41897
4-FC6H4H4-BrC6H42183
4-MeC6H4Me4-BrC6H42184
n-BuH4-BrC6H422NR
a Isolated yields; b the reaction was carried out at 70 °C with a 2:1 ratio for ynone:indane-1,3-dione.
Table 26. Spiro compounds 408 obtained by cascade Michael addition/cycloaddition reactions between 2-arylidene-indane-1,3-diones 255 and allenoates 407.
Table 26. Spiro compounds 408 obtained by cascade Michael addition/cycloaddition reactions between 2-arylidene-indane-1,3-diones 255 and allenoates 407.
R1R2R3t (h)Yield (%)
C6H5HC6H51590
C6H5H2-BrC6H41093
C6H5H3-BrC6H41096
C6H5H4-BrC6H41090
C6H5H2-MeC6H41091
C6H5H3-MeC6H41095
C6H5H4-MeC6H41094
C6H5H4-ClC6H41288
C6H5H4-FC6H41287
C6H5H4-MeOC6H4390
C6H5H4-NO2C6H42473
C6H5H2,4-Cl2C6H31262
C6H5H2-C4H3O4887
C6H5H2-C4H3S5384
C6H5H1-C10H72096
C6H5HC2H4C6H43589
C6H5Me4-BrC6H41895
C6H5Di-Me4-BrC6H41897
4-FC6H4H4-BrC6H42183
4-MeC6H4Me4-BrC6H42184
n-BuH4-BrC6H422NR
Table 27. Domino reaction between 2-arylidene-indane-1,3-diones 255 and N-alkoxyacrylamides 410 in the presence of a base.
Table 27. Domino reaction between 2-arylidene-indane-1,3-diones 255 and N-alkoxyacrylamides 410 in the presence of a base.
ArR1R2Yield %dr
C6H5C6H5H854:1
C6H54-FC6H4H814:1
C6H54-OMeC6H4H694:1
C6H54-NO2C6H4H514:1
4-OMeC6H4C6H5H604:1
4-OMeC6H44-FC6H4H484:1
4-OMeC6H44-OMeC6H4H514:1
4-OMeC6H44-NO2C6H4H404:1
C6H5MeH723:2
4-OMeC6H4MeH603:2
C6H5HMe60 a- b
C6H5MeMe69-
a The reaction was carried at 60 °C; b only one diastereoisomer was observed by 1H NMR.
Table 28. Reaction yields obtained during the cascade double Michael addition/acetalization reactions.
Table 28. Reaction yields obtained during the cascade double Michael addition/acetalization reactions.
R1 aR2R3Time (h)Yield % b
-C6H5C6H52482
5-OHC6H5C6H52482
5-OMeC6H5C6H53677
4-OMeC6H5C6H52783
5-BrC6H5C6H52482
-4-OMeC6H4C6H575.572
-4-BrC6H4C6H59672
-4-ClC6H4C6H524.567
-3-ClC6H4C6H52472
-2-ClC6H4C6H52472
-C6H54-BrC6H424.576c
-C6H53-BrC6H424.576
-C6H54-ClC6H42782
-C6H53-ClC6H424.573
-C6H52-ClC6H424.577 c
-C6H54-OMeC6H44884
-C6H5C4H3S2471
-C6H5C3H72458 c
a If nothing, there is 2-OHC6H4; b isolated yield; c in this case, only the intermediate was obtained in this yield.
Table 29. Reaction yields obtained during the synthesis of spiro-indane-1,3-diones 428.
Table 29. Reaction yields obtained during the synthesis of spiro-indane-1,3-diones 428.
Rt (h)Yield (%) a
C6H51067
2-ClC6H41069
4-ClC6H4872
4-FC6H4964
3-NO2C6H41062
4-MeC6H41052
4-MeOC6H4962
3,5-(MeO)2C6H3958
2-HOC6H41063
4-HOC6H4955
2-HO-3-MeOC6H4857
3-MeO-4-HOC6H4860
4-NMe2C6H41052
a Isolated yield.
Table 30. Scope of application of the microwave-assisted reaction.
Table 30. Scope of application of the microwave-assisted reaction.
R1R2Yield Product 1 (%) Yield Product 2 (%)
C6H5C6H53031
4-NMe2C6H44-NMe2C6H42832
4-OMeC6H44-OMeC6H43236
4-NO2C6H44-NO2C6H42022
4-C6H5C6H44-C6H5C6H43228
C4H3SC4H3S4241
C4H3OC4H3O3636
3-NO2C4H2O3-NO2C4H2O3931
3-MeC4H2O3-MeC4H2O3033
Table 31. Reaction yields obtained during the synthesis of triazole-containing spiro-indane-1,3-diones 445.
Table 31. Reaction yields obtained during the synthesis of triazole-containing spiro-indane-1,3-diones 445.
R1R2Time (min)Yield (%)
4-ClC6H44-FC6H44585
4-ClC6H44-MeC6H45080
4-BrC6H44-FC6H45083
4-FC6H44-MeC6H44082
4-BrC6H44-MeC6H45081
4-FC6H44-FC6H44584
4-ClC6H44-NO2C6H45082
4-BrC6H44-NO2C6H44586
4-MeC6H44-NO2C6H44079
4-NO2C6H44-OMeC6H44584
4-FC6H44-NO2C6H44082
4-(CF3)C6H44-NO2C6H44087
4-(CF3)C6H47-ClC9H5N5080
4-(CF3)C6H44-FC6H44585
4-MeC6H47-ClC9H5N5074
C4H3OC4H95080
(CH2O2)C6H34-FC6H45086
C4H97-ClC9H5N6078
C4H94-(Ome)C6H46574
Table 32. MRC used for the design of spiro-compounds 450 with biological properties.
Table 32. MRC used for the design of spiro-compounds 450 with biological properties.
ArYield (%) a
4-CH2OHC6H481
4-CHOC6H480
4-OMeC6H493
3,5-(OMe)2C6H440
4-CF3C6H481
4-OHC6H470
4-COOHC6H454
a Yield of the purified product after chromatography column.
Table 33. MRC reactions involving Fe-based nanoparticles.
Table 33. MRC reactions involving Fe-based nanoparticles.
R1R2Yield (%) a
--84
4-Me-80
4-OMe-73
2-OMe-70
3-OMe-78
4-Cl-92
4-Br-90
3-OMe4-OMe83
4-Br4-OMe86
4-Br4-Br91
a Isolated yield.
Table 34. Reaction yields obtained during alkene hydrofunctionnalizations of 2-arylidene-indane-1,3-diones 255.
Table 34. Reaction yields obtained during alkene hydrofunctionnalizations of 2-arylidene-indane-1,3-diones 255.
RMethylenYield (%) a
C6H5Yes182
4-BrC6H4Yes155
4-CF3C6H4Yes159
4-MeC6H4Yes153
4-NMe2C6H4Yes145
4-ClC6H4Yes142
4-CO2MeC6H4Yes169
4-OMeC6H4Yes154
2-OMeC6H4Yes172
3-OMeC6H4Yes154
3-CNC6H4Yes148
2,4,5-C6H2Yes150
C4H3OYes158
C4H3SYes148
C10H7Yes174
3-CNC6H4No163 b
C4H3SNo146 b
4-OMeC6H4No258
C6H5Yes250
a Yield refers to isolated product; b dr was determined by 1H NMR.
Table 35. Copper-catalyzed synthesis of spiro-compounds.
Table 35. Copper-catalyzed synthesis of spiro-compounds.
R1R2Yield (%) a
4-FC6H4C6H590
4-BrC6H4C6H561
4-CF3C6H4C6H584
4-MeC6H4C6H573
4-OMeC6H4C6H582
3-MeC6H4C6H581
3-OMeC6H4C6H566
3-ClC6H4C6H569
2-BrC6H4C6H585
C10H7C6H571
C4H3OC6H564
C4H3SC6H566
C6H54-FC6H474
C6H54-ClC6H487
C6H54-BrC6H491
C6H54-NO2C6H475
C6H54-MeC6H476
C6H54-OMeC6H460
C6H53-ClC6H489
C6H53-MeC6H478
C6H53-OMeC6H474
C6H52-BrC6H468
C6H52-MeC6H470
4-BrC6H44-MeC6H458
4-FC6H43-BrC6H462
3-BrC6H44-ClC6H471
C4H3O4-NO2C6H471
C6H5CO2MeNr
C6H5C6H11Nr
a Isolated yield; Nr, no reaction.
Table 36. Copper-catalyzed cycloaddition reaction of 2-arylene-indane-1,3-dione 255 with various oximes 466.
Table 36. Copper-catalyzed cycloaddition reaction of 2-arylene-indane-1,3-dione 255 with various oximes 466.
R1R2R3R4Yield (%) a
C6H5C6H5HH74
C6H54-MeC6H4HH82
C6H54-OMeC6H4HH80
C6H54-IC6H4HH64
C6H54-ClC6H4HH83
C6H54-CF3C6H4HH81
C6H54-NO2C6H4HH66
C6H54-CNC6H4HH78
C6H54-MeSC6H4HH81
C6H53-BrC6H4HH85
C6H53-MeC6H4HH89
C6H52-MeC6H4HH78
C6H5C10H7HH74
C6H5C4H9HH79
C6H5C6H5CH3CH350
C6H5C6H5C2H5H65 a
C6H5C3H5HH78
C6H5CO2EtHH30
C6H5C10H20H32 a
C6H5C8H8H40 a
4-OMeC6H4C2H5CH3H51 a
C6H5C5H4NHH51
C6H5C4H3OHH86
C6H5C4H3SHH56
C6H5C9H8NHH40
C6H5C5H6NHH96
C6H5C4H5N2HH77
C6H5C6H4C3H2SNHH70
4-BrC6H4C6H3(O2CH2)HH72
2-MeC6H4C8H5SHH52
a > 19:1 dr value was determined by 1H NMR spectroscopy.
Table 37. Reaction yields obtained during the reaction of of 2-arylidene-indane-1,3-diones 255 and the Seyferth–Gilbert reagent 457 in the presence of CsF.
Table 37. Reaction yields obtained during the reaction of of 2-arylidene-indane-1,3-diones 255 and the Seyferth–Gilbert reagent 457 in the presence of CsF.
RYield (%) a
C6H587
4-OMeC6H475
4-iPrC6H492
4-EtC6H488
4-MeC6H488
4-PhC6H490
4-BrC6H485
4-ClC6H477
3-ClC6H493
4-OMeC6H486
2-MeC6H488
3,4-(OMe)2C6H365
3,4,5-(OMe)3C6H280
C4H4N80
C4H3S87
C10H789
Fc87
a Isolated yield after chromatography by silica gel.
Table 38. Antibacterial and antifungal activities of 538 derivatives.
Table 38. Antibacterial and antifungal activities of 538 derivatives.
EntryZone of Inhibition/mm
Gram-PositiveGram-NegativeFungi
B. subtilisS. aureusP. aeruginosaE. coliA. nigerC. albicansA. fumigatus
538a9106881011
538b15141011121417
538c21241214191719
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pigot, C.; Brunel, D.; Dumur, F. Indane-1,3-Dione: From Synthetic Strategies to Applications. Molecules 2022, 27, 5976. https://doi.org/10.3390/molecules27185976

AMA Style

Pigot C, Brunel D, Dumur F. Indane-1,3-Dione: From Synthetic Strategies to Applications. Molecules. 2022; 27(18):5976. https://doi.org/10.3390/molecules27185976

Chicago/Turabian Style

Pigot, Corentin, Damien Brunel, and Frédéric Dumur. 2022. "Indane-1,3-Dione: From Synthetic Strategies to Applications" Molecules 27, no. 18: 5976. https://doi.org/10.3390/molecules27185976

Article Metrics

Back to TopTop