Next Article in Journal
Sustainable and Selective Extraction of Lipids and Bioactive Compounds from Microalgae
Next Article in Special Issue
Revisiting the Structure and Chemistry of 3(5)-Substituted Pyrazoles
Previous Article in Journal
Discovery and Characterisation of Dual Inhibitors of Tryptophan 2,3-Dioxygenase (TDO2) and Indoleamine 2,3-Dioxygenase 1 (IDO1) Using Virtual Screening
Previous Article in Special Issue
Dialkylation of Indoles with Trichloroacetimidates to Access 3,3-Disubstituted Indolenines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Current Advances in the Synthesis of Valuable Dipyrromethane Scaffolds: Classic and New Methods

by
Bruno F. O. Nascimento
,
Susana M. M. Lopes
,
Marta Pineiro
and
Teresa M. V. D. Pinho e Melo
*
CQC and Department of Chemistry, University of Coimbra, Rua Larga, 3004-535 Coimbra, Portugal
*
Author to whom correspondence should be addressed.
Submission received: 8 November 2019 / Revised: 21 November 2019 / Accepted: 22 November 2019 / Published: 28 November 2019
(This article belongs to the Special Issue Development of New Methods of Synthesis of Heterocycles)

Abstract

:
This review presents the most recent developments on the synthesis of dipyrromethanes, covering classical synthetic strategies, using acid catalyzed condensation of pyrroles and aldehydes or ketones, and recent breakthroughs which allow the synthesis of these type of heterocycles with new substitution patterns.

Graphical Abstract

1. Introduction

Dipyrromethanes are well known synthetic scaffolds for the synthesis of macrocycles and dipyrromethene metal complexes. Dipyrromethanes occupy a central place in porphyrin chemistry. The dipyrromethane structures employed in the synthesis of naturally occurring porphyrins typically bear substituents at the β-positions and lack any substituent at the meso-position. However, the dipyrromethanes with substituents at the meso-position have come to play a valuable role in the preparation of synthetic porphyrins [1,2], calixpyrroles [3], chlorins [4], corroles [5], and expanded porphyrins, namely saphyrins and smaragdyrins [6,7] (Figure 1).
The most representative example of dipyrromethene metal complexes are 4,4-difluoro-4-bora-3a,4a-diaza-s-indacenes also known as BODIPYs, which have been successfully used as fluorescent probes in diverse applications [8,9,10]. Other metal complexes have also attracted researchers’ attention, for example, recently, aluminium complexes have been used as catalysts for polymerization reactions [11,12], iron complexes [13] have been used as catalysts for C-H bond amination [14,15], and ruthenium complexes of dipyrromethenes [16,17] were synthetized as precursors of bis(2,2′-bipyridyl)(dipyrrinato)ruthenium(II) complexes.
Beyond its use as synthetic scaffolds, the dipyrromethane framework used as ligand for the synthesis of organometallic complexes has attracted the interest of several research groups, mostly due to the ease of synthetic accessibility and versatility of substitution of this moiety. The electronic properties of this ligand can be modified by substitution at the beta-carbons and at the meso positions, while the steric properties can be tuned by substitution at the alpha-carbons. Zirconium complexes have been applied in olefin hydroamination [18] and ruthenium, rhodium and iridium complexes used as hydrogen transfer catalysts under aqueous and aerobic conditions [19]. The synthesis of dipyrromethene complexes has also been achieved with Mn, Co, Zn, Ni [20,21,22] or Sn [23]. Recently, a dipyrromethane-based diphosphane–germylene was synthetized and used as precursor of tetrahedral Cu(I) and T-shaped Ag(I) and Au(I) flexible pyrrole-derived Phosphorous-Germanium-Phosphorous (PGeP) germylene pincer complexes [24].
Anion recognition is an area of growing interest due to its important role in a wide range of environmental, clinical, chemical and biological applications. Interestingly, the acidic NH protons of dipyrromethanes can be used as anion sensors with good binding activity and selectivity [25,26,27,28]. Furthermore, polymers based on dipyrromethanes were developed for the molecular recognition of two homoserine lactone derivatives involved in bacterial quorum sensing [29].
Herein, bibliographic coverage of the developments on the synthesis of dipyrromethanes since the last reviews in this area [9,30,31,32] is provided (2014–2019). The synthetic strategies have been organized in two main approaches: classical synthetic strategies based on the first report on the synthesis of meso-substituted dipyrromethanes, disclosed in 1974 [33], using acid catalyzed condensation of pyrrole and aldehydes; and recent breakthroughs in dipyrromethane chemistry which allow the synthesis of dipyrromethanes with new substitution patterns.

2. Classic Synthetic Strategies

2.1. Hydrochloric Acid-Catalyzed Dipyrromethane Synthesis

Receptor molecules grounded on guanidinium- and pyrrole-containing binding sites 3 were developed by Kataev and colleagues with the objective of selective recognizing orthophosphate anions in aqueous media (Scheme 1) [27].
Studies demonstrated that the pyrrole-containing binding site was of pronounced influence on the selectivity and that dipyrromethane core structure 2, prepared from the HCl-catalyzed reaction of 4-heptanone 1 and pyrrole in boiling water in 13% isolated yield, demonstrated the highest selectivity for orthophosphate over other inorganic anions. A novel and readily available dipyrromethane-based dual receptor 6 serving as colorimetric sensor for both F- and Cu2+ ions was recently designed and prepared by Pandey and co-workers (Scheme 1) [28]. Treatment of pyrrole and acetophenone 4 in the presence of catalytic HCl in water was key in the formation of meso-methyl-meso-phenyl-dipyrromethane 5 in 75% yield.
Balci et al. established a regioselective method for the preparation of dipyrrolo-diazepine derivatives [34]. This firstly involved the classic room temperature HCl-promoted synthesis of dipyrromethanes 7 (starting from excess pyrrole and suitable aldehydes), followed by reaction of propargyl bromide 8 in the presence of sodium hydride to append an alkyne functionality to the nitrogen atom at one of the pyrrole units. A final seven-exo-dig cyclization, between the alkyne group and the N-deprotonated pyrrole moiety, followed by prototropy produced the target compounds 10 in generally good overall yields (Scheme 2).

2.2. Acetic/propionic Acid-Catalyzed Dipyromethane Synthesis

Interesting work carried out by the research group of Swavey allowed the preparation of two new dipyrromethane bridging ligands [17], as well as their corresponding dimetallic ruthenium(II) [17] and osmium(II) [35] coordination complexes. Reaction of phenanthrolinepyrrole 11 (php) with the selected aryl aldehyde in acetic acid (AcOH) produced dipyrromethanes 12, comprising two php moieties linked by a meso-aryl group (Scheme 3). The use of benzaldehydes substituted with electron donating (vanillin) and electron withdrawing (cyano) groups did not greatly affect the efficiency of the reaction with php. However, the use of sterically hindered aldehydes, e.g., mesitylbenzaldehyde, 3,4,5-trimethoxybenzaldehyde and pentafluorobenzaldehyde, was completely unsuccessful. Coordination with Ru(II) or Os(II) bis(bipyridyl) chloride in refluxing ethanol, followed by saturation with aqueous ammonium hexafluorophosphate, created the novel dimetallic complexes 13 and 14 in good isolated yields (Scheme 3).
Access to C2v symmetric β-substituted porphyrins, e.g., protoporphyrin III 21, using the promptly accessible Knorr’s pyrrole 15, which is crucial in the preparation of the required dipyrromethane building blocks [36], was envisaged by Neya and colleagues in 2016 [37]. Two Knorr’s pyrrole units were coupled into symmetric dipyrromethane 16 in propionic acid (PrOH). Its 3,3′-dibenzyl groups were removed via hydrogenolysis affording the corresponding carboxilic acid substituents, which were removed by iodination giving dipyrromethane 17. This was further reduced to dipyrromethane 18 under a Pd/C-catalyzed hydrogen atmosphere. Reaction of 18 with acetyl chloride in the presence of aluminum chloride led to the formation of 3,3′-diacetyldipyrromethane dimethylester 19, which was then hydrolyzed using aqueous sodium hydroxide into the corresponding dipyrromethene-5,5′-dicarboxylic acid. Its terminal carboxylic residues were subsequently eliminated through iodinative decarboxylation, rendering 5,5′-diiododipyrromethane 20 in 10.2% overall yield after eight reaction steps (Scheme 4).
The multistep synthesis of 1,4,5,8-tetraethyl-2,3,6,7-tetravinylporphyrin 26 was reported by the same research team one year later, this time using closely related Knorr’s pyrrole analogue 22 as starting material [38]. The same experimental protocol was selected in order to produce 5,5′-diiododipyrromethane 23, which was then subjected to reduction to afford dipyrromethane 24, followed by formylation to dipyrromethane 25 (Scheme 5). These two new dipyrromethane derivatives, 24 and 25, were key in the subsequent preparation of the target symmetric porphyrin 26.

2.3. p-Toluenesulphonic Acid-Catalyzed Dipyrromethane Synthesis

A solution of furan-2-carboxaldehyde 27 and ethyl 2-cyano-3-(1H-pyrrol-2-yl)-acrylate 28 in dichloromethane was refluxed for 8 h in the presence of a catalytic amount of p-toluenesulphonic acid (p-TSA) to afford 1,9-bis(2-cyano-2-ethoxycarbonylvinyl)-5-(2-furanyl)-dipyrromethane 30 in 32% yield (Scheme 6). This new dipyrromethane was extensively characterized through experimental spectroscopic measurements and theoretical quantum chemical calculations by Singh and co-workers [39]. The same authors later described the preparation of some novel dipyrromethane-hydrazone derivatives 31, by condensing previously synthesized 2-[(4-isonicotinoyl)-hydrazonomethyl]-1H-pyrrole 29 with suitable aldehydes, also under classic p-TSA catalyzed reactional conditions, high yields being attained [40]. These were screened for antitubercular activity against Mycobacterium tuberculosis H37Rv strains, interesting minimum inhibitory concentration (MIC) values being found (Scheme 6).

2.4. Trifluoroacetic Acid-Catalyzed Dipyrromethane Synthesis

Aiming to synthesize a 5,10-diacyltripyrrane, an essential intermediary synthon for the preparation of a 5,10-diacylcalix[4]pyrrole, Mahanta and Panda were able to unexpectedly isolate acyldipyrromethane 32 with a yield up to 31% (Figure 2), following the trifluoroacetic acid (TFA)-catalyzed reaction of 2,3-butanedione and excess pyrrole [41]. Yildiz et al. have described the synthesis, characterization, crystal structure and theoretical calculations of two new meso-borondipyrromethene (BODIPY) incorporating a phthalonitrile moiety [42,43]. Crucial for their detailed report was the preparation of meso-phenoxyphthalonitrile dipyrromethanes 33 and 34 (Figure 2), which were attained in 53% and 61% yields, respectively, after typical room temperature condensation of suitable and previously obtained aldehydes with excess pyrrole in the presence of TFA. A very similar experimental setup was applied to the synthesis of unsubstituted dipyrromethane [44], meso-2-pyrrolyldipyrromethane 7f (see Scheme 2) in 90% yield [45], as well as to the preparation of dipyrromethane 35, in 43% yield [46]. Three meso-(trimethylsilyl)phenyl-dipyrromethane structures 36 were obtained in good yields, ranging from 60% to 66% (Figure 2), after solventless condensation catalyzed by TFA of the corresponding silylated aldehydes with pyrrole [47]. Extensive work on the synthesis of novel pentafluorosulfanyl-substituted A4-type porphyrins (and their respective Zn(II) and Pd(II) complexes), A3-, AB2- and A2B-type corroles, trans-A2B2-type porphyrins and BODIPYs has been reported [48]. Regarding the three latter molecular targets, meta-SF5-phenyl-substituted dipyrromethanes 37ac, prepared in high yields (62%–80%) under standard TFA-catalyzed reaction conditions using appropriate pentafluorosulfanyl-bearing aryl aldehydes and surplus pyrrole, were the strategic intermediates (Figure 2). More recently, the same authors reported the efficient preparation, under similar standard conditions, of useful meso-aryldipyrromethane scaffolds 37d [49] and 37e [50] (in 92% and 87% yields, correspondingly), which were further functionalized and/or used as building blocks for the formation of other interesting BODIPY or porphyrinoid molecular species.
Bis-dipyrromethanes 38ad and 39ad, as well as tris-dipyrromethane 38e and 39e, were prepared from commercially accessible starting materials by Sessler’s research team [51,52], following the strategy summarized in Scheme 7. Their anion binding properties were then evaluated in both organic media and in the solid state, compounds 39 displaying a good affinity for dihydrogenphosphate and pyrophosphate anions (as tetrabutylammonium salts) in chloroform solutions, acting as conformationally switchable receptors.
Acyclic and macrocyclic dipyrromethanes were synthesized by Love et al. by applying common Schiff-base condensation chemistry to meso-pentafluorophenyldipyrromethane dialdehyde 40, diiminodipyrromethane derivative 41 being obtained in 74% yield (Scheme 8) [53]. Bridged macrocyclic dipyrromethanes 43 were prepared by condensation of 40 with either ortho-phenylene 42a or 1,5-anthracene-diamine 42b, again using TFA as catalyst. After neutralization with triethylamine, the target molecules precipitated neatly from the reaction medium (Scheme 8).
Novel bis-dipyrromethane derivatives 45 were synthesized in moderate yields by reacting previously prepared dialdehydes 44 with surplus pyrrole in the presence of TFA as catalyst (Scheme 9). These bis-dipyrromethanes 45, as well as 38c (see Scheme 7) and dipyrromethane 35 (see Figure 2), undergo electropolymerization on the electrode surface occurring upon multiple oxidation cycles [46]. The authors uncovered that quicker electropolymerization rates arise when the monomeric species contains more than one dipyrromethane component, and that the resulting polymers exhibit greater stability, while also showing low roughness and a very uniform and homogenous morphology.
β-Formyl-tetrapyrrolic macrocycles have been used for the synthesis of dipyrromethene derivatives through condensation with pyrroles (Scheme 10). Temelli and Kalkan recently described the synthesis of meso-porphyrinyldipyrromethane 47 in high yield, by reacting 2-formyl-meso-tetraphenylporphyrin 46 prepared beforehand and excess pyrrole under classic TFA catalysis conditions. The unforeseen construction of β/meso-directly connected diporphyrinic molecules in reactions of β-formylated porphyrins with pyrrole under ordinary Adler-Longo conditions was also established, early mechanistic studies showing that dipyrromethane 47-type intermediates are fundamental in the process [54]. On the other hand, Galium(III) corrole-BODIPY hybrid 49 was obtained from the TFA-catalyzed reaction of formyl-corrole 48 with 2,4-dimethylpyrrole followed by oxidation and complexation with boron trifluoride [55].
New meso/meso-straightly linked Ni(II) porphyrin hybrids were designed and prepared by Kong and co-workers, the metalloporphyrinic units being connected with dipyrromethene, bis-dipyrromethene or thiacorrole units [56]. Instrumental in the synthetic strategy of the authors was meso-dipyrromethane-susbtituted Ni(II) porphyrin 51, which was effortlessly obtained in good yield via TFA-catalyzed room temperature condensation of Ni(II) 10,20-di(3,5-di-t-butylphenyl)-5-formyl porphyrin 50 and pyrrole (Scheme 11).
The synthesis and characterization (structural, spectral and electrochemical) of Pd(II), Re(I) and Ru(II) 3-pyrrolyl-BODIPY/dipyrromethenes complexes 55ac was reported by Ravikanth’s research team in 2014 [57]. The pertinent and necessary dipyrromethane-substituted 3-pyrrolyl BODIPY intermediate 53a was prepared in 85% yield by mixing a dichloromethane solution of formylated 3-pyrrolyl BODIPY 52a, produced and isolated beforehand, and excess pyrrole under TFA-catalyzed and inert atmosphere conditions (Scheme 12). Related α-dipyrromethanyl 3-pyrrolyl BODIPY 53b was synthesized in a similar fashion by the same research group a few years later (Scheme 12) [58]. This was further transformed into 3-pyrrolyl BODIPY/BODIPY dimer 54, comprising an ethynyl functionality at the meso-aryl location, which was then coupled to selected monomeric BODIPYs in order to create the authors’ target near-infrared emitting BODIPY oligomers.

2.5. Boron Trifluoride Diethyl Etherate-Catalyzed Dipyrromethane Synthesis

New meso-phenothiazinyldipyrromethanes 56, having alkyl groups of growing bulkiness linked to the heterocyclic nitrogen of the phenothiazine core were synthesized in high yields (81%–89%) by condensing proper N-alkyl-phenothiazin-3-carbaldehydes with pyrrole at room temperature, in the dark, and in the presence of boron trifluoride diethyl etherate (BF3.Et2O) as catalyst (Figure 3) [59]. A BODIPY dye [59], a trans-A2B2-type porphyrin [59], and some Sn(IV) coordination complexes [60] bearing an N-methyl-phenothiazinyl motif were later prepared utilizing dipyrromethane 56a as the building block. Thilagar and colleagues conveyed the simple preparation, under typical BF3.Et2O-catalyzed reaction conditions of four novel triarylborane-dipyrromethane derivatives 57ad (Figure 3) that encompassed dual receptor sites (Lewis acidic boron and hydrogen bond donor NH) and displayed a discriminating fluorogenic response towards the fluoride anion in dichloromethane solution [61]. Broadly acknowledged meso-substituted dipyrromethanes 7c (Scheme 2) and 58 were recently obtained by Bagherzadeh and co-workers via the dropwise addition of pyrrole to a dilute aqueous solution of the aryl aldehydes, using boron trifluoride diethyl etherate as catalyst (Figure 3) [62]. This methodology was adapted from an earlier report that used aqueous HCl at 90 °C [63], tripyrromethane and other oligomers being obtained as byproducts when large and sterically hindered aryl aldehydes are employed. The reaction occurred smoothly in mild conditions, 30–70 min at 70–80 °C under an argon atmosphere, moderate to high yields (60%–85%) being obtained, even when employing bulky electron donating (mesityl) or electron withdrawing (2,6-dichlorophenyl) aldehydes, and no decomposition or scrambling being noticed.
Sessler and co-workers synthesized pyrene-bridged bis-dipyrromethane 60 in 41% yield by reacting previously prepared pyrene dialdehyde 59 with excess pyrrole in ethanol at room temperature for three days and using BF3.Et2O as catalyst (Scheme 13) [64]. After formylation under standard reaction conditions, pyrene-linked tetraformylated bis-dipyrromethane 61 was obtained in moderate yield. Anion recognition studies revealed that the latter performs as a selective fluorescent probe in chloroform solution for dihydrogen phosphate over other tested anions.
Ravikanth’s research group reported the preparation and characterization (structural, photophysical and electrochemical) of β-meso covalently linked azaBODIPY/BODIPY dyad 64 [65] and Pd(II) azaBODIPY/dipyrromethene complex 65 [66,67]. The unconditionally required dipyrromethane-substituted azaBODIPY intermediary 63 was synthesized in reasonable yield by stirring a dichloromethane solution of 2-formyl azaBODIPY 62, and surplus pyrrole, at room temperature, under boron trifluoride diethyl etherate catalysis and inert atmosphere conditions (Scheme 14). A similar synthetic strategy, condensation of 3-formyl-BODIPY with pyrrole catalyzed by BF3-OEt2 was used for the synthesis of BODIPY/BODIPY dimers [68].

2.6. Indium(III) Chloride-Catalyzed Dipyrromethane Synthesis

meso-Substituted dipyrromethanes 66a,b were prepared by simple solvent-free condensation of the appropriate aryl aldehydes with excess pyrrole, under a saturated argon atmosphere and indium chloride (InCl3)-catalyzed conditions, (Figure 4) moderate yields being attained [69]. Lindsey’s research team designed and synthesized a series of interesting trans-AB-kind porphyrins and metalloporphyrins comprising one water-solubilization moiety and one bioconjugatable functionality [70]. Key for their strategy was the previous preparation of suitable dipyrromethane building blocks, including novel compound 66c, which was obtained in 33% yield using the same catalytic conditions (Figure 4). A related setting was also applied for the synthesis of meso-nonyldipyrromethane 67 in 78% yield and meso-methoxycarbonyldipyrromethane 68 in 52% yield [71]. In addition to the latter, the synthetic process also provided regioisomer 69 and cyclic derivative 70, a by-product resulting from intramolecular aminolysis of ester 68, in 32% and 10% isolated yields, respectively (Figure 4). Moreover, the same authors prepared dipyrromethane derivatives 71 and 72 in moderate to reasonable yields, 47%–62% by simply mixing the adequate aldehydes or acetals with excess pyrrole in an inert atmosphere using indium chloride as catalyst. Dipyrromethanes 67–69 and 71–72 were later crucial for the preparation of several trans-AB-type porphyrins and metalloporphyrins (Figure 4) [71].
Aiming to combine porphyrin, BODIPY and triptycene chemistry, Senge et al. recently presented meso-triptycenyldipyrromethane synthon 74, synthesized in 60% yield from the condensation of 2-formyltriptycene 73 and surplus pyrrole under standard InCl3-catalyzed conditions (Scheme 15) [72]. The extreme utility of dipyrromethane 74 was noticeably demonstrated by its application in the preparation of triptycene-substituted BODIPY 75a, triptycene-substituted 3-pyrrolyl-BODIPY 75b, trans-A2B2 triptycenylporphyrins 76a,b and A3B-type triptycenylporphyrins 76c,d.
Following a similar synthetic approach, the same authors also devised the synthesis of a more complex tris-dipyrromethane-substituted triptycene 78 in 37% yield [72], starting from 2,6,14-tri(4-formylphenyl)triptycene derivative 77, which was prepared and isolated beforehand (Scheme 16). Tris-dipyrromethane 78 was later successfully utilized as a valuable intermediate in the synthesis of tris-BODIPY-substituted triptycene 79.

2.7. Other Strategies

The synthetic route that Trofimov’s research group developed in order to obtain new meso-trifluoromethyldipyrromethanes 85 and 86 using 2-aminophenyl-1H-pyrroles 80 as starting material is depicted in Scheme 17 [73]. Briefly, the protection of the amino functionality of pyrroles 80 with acetic anhydride, followed by a reaction with trifluoroacetic anhydride (TFAA), rendered 2-trifluoroacetylpyrroles 82 in high yields. Sodium borohydride-promoted reduction of pyrroles 82 and ensuing reaction of pyrrole carbinols 83 with 2-phenylpyrrole 84, in the presence of dehydration agent phosphorous pentoxide (P2O5), gave dipyrromethanes 85 in good yields. Finally, conversion of the acetamide substituents into amino groups in refluxing acidic media originated dipyrromethanes 86 (Scheme 17). Both dipyrromethane derivatives 85 and 86 were later used in the preparation of their corresponding meso-CF3-BODIPY dyes, a big influence of the coexistence of the strong electron donor NH2 and electron acceptor CF3 groups, along with the location of the amine at the aryl ring, being determined on the optical properties of chromophores 86 [73].
The preparation of novel meso-trifluoromethyldipyrromethanes 92 and 93, comprising isoxazole substituents in their molecular structure, starting from ethynylpyrrole 87 was also recently described by the same authors (Scheme 18) [74]. Initial cyclization of 87 with hydroxylamine hydrochloride and subsequent condensation of the obtained isoxazoles 89 or 90 with previously prepared pyrrole carbinols 91 in the presence of dehydration agent P2O5 leads to the desired and unreported dipyrromethane derivatives 92 and 93 (Scheme 18). The latter were again further explored in the synthesis of their respective meso-CF3-BODIPY dyes, some photophysical studies and quantum chemical calculations having been carried out [74].
Aiming to synthesize the illusive and highly sought 2,3,7,8,12,13,17,18-octafluoroporphyrin with a reasonable yield, Chang and colleagues choose the approach summarized in Scheme 19 [75], ensuing an older account by Clezy and Smythe [76]. In brief, tetrafluorinated dipyrromethane 97 was obtained in three steps starting from 3,4-difluoropyrrole 94. Reaction with thiophosgene under an inert atmosphere rendered dipyrrothioketone 95 in high yield. Subsequent hydrogen peroxide-promoted oxidation to dipyrroketone 96, followed by sodium borohydride-mediated reduction, originated dipyrromethane derivative 97. Having this valuable scaffold in hands, the authors were thus able to prepare trans-A2B2-type porphyrins 98, as well as β-octafluoroporphyrins 99.

3. Novel Synthetic Strategies

3.1. Dipyrromethane Synthesis from Aldehydes and Pyrroles

Despite the wide variety of standard available methods, there is still an open door for the development of new synthetic approaches for dipyrromethanes. For instance, there is a growing interest in the use of catalysts that can be easily removed from the reaction medium and reused, while also having an economically and ecologically friendly access. Konar and co-workers described the synthesis of a wide range of meso-thienyldipyrromethanes 101 using an amine functionalized MOF (Metal-Organic Framework) for the controlled release of the catalyst, iodine (Scheme 20) [77]. Dipyrromethanes 101 were obtained in high yields (50%–69%) with the ratio aldehyde/pyrrole (1:5) in the presence of 20 mol% of NH2-MOF(I2), without organic solvents and under mild conditions. The catalyst was reused for three cycles, although with a slight yield decrease. After immersion in an iodine solution, the catalytic performance was comparable with the freshly prepared NH2-MOF(I2). The authors tested other catalysts, such as unfunctionalized MOFs (H-MOF(I2)), conventional TFA and molecular iodine; however, dipyrromethanes 101 were obtained with lower yields.
Copper nanoparticles (CuNPs) [78] and celite-supported glycine (glycine@celite) [79] have been explored as catalysts in the synthesis of dipyrromethanes because they are easily removed from the reaction medium and can be reutilized (Scheme 21). The reaction of aromatic aldehydes with pyrrole (10 equiv) under solvent-free conditions using CuNPs as catalyst gave dipyrromethanes 7c,e, 58d and 103 in high yields (65%–77%). In this catalytic system, the pyrrole nucleophilicity was increased by the adsorption over CuNPs and reacted with the electrophilic carbon of the aldehyde to give an intermediate, which reacted with another molecule of pyrrole adsorbed on CuNPs, giving dipyrromethanes 7c,e, 58d, 103. The catalyst was recovered and reused in four cycles with a similar catalytic activity without any deterioration. Dipyrromethane 104d was also synthesized using the CuNPs as catalyst in 75% yield (Scheme 21) [78]. Glycine@celite was a milder acid catalyst system whose efficacy was proven in the synthesis of dipyrromethanes 104. The reaction of aromatic aldehydes with 2,4-dimethylpyrrole (3 equiv) in the presence of 10 mol% of the catalyst in dichloromethane for 30 min gave the corresponding dipyrromethanes 104 in very good yields (Scheme 21). The authors described the reuse of the catalyst for 5 cycles without any deterioration and with similar catalytic activity [79].
Dipyrromethane 104e underwent oxidation and chelation by copper producing a red bis(dipyrrinato)copper(II) complex 105 (Scheme 22). This characteristic makes it an efficient naked eye colorimetric chemosensor for copper ions, even in the presence of several other metal ions [79,80].
Dipyrromethanes 7c, 103b and 106 were synthesized in good yields by the iodine-catalyzed double Friedel-Crafts reaction, using toluene or water as solvent (Scheme 23) [81]. The reaction of pyrrole derivatives with aldehydes, (2:1) ratio, in presence of 10 mol% of molecular iodine in water gave dipyrromethanes in better yields (60%–87%) than the reaction carried out in toluene (42%–62%).
Sirion and co-workers described the synthesis of dipyrromethanes through the reaction of aldehydes with pyrrole catalyzed by SO3H-functionalized ionic liquids (SO3H-ILs) in aqueous media [82]. The authors tested a few SO3H-ILs containing imidazolium or pyridinium cations and different anions, and found that [bsmim][HSO4] (i.e., 1-butylsulfonic-3-methylimidazolium hydrogen sulfate) was the ideal catalyst for the synthesis of dipyrromethanes (Scheme 24). Dipyrromethanes 107 were obtained in moderate to good yields from the reaction of pyrrole with aliphatic or aromatic aldehydes, in the presence of 10 mol% of catalyst in water under mild conditions. The described method comprised a large variety of aromatic aldehydes with both electron withdrawing and electron donating substituents, heteroaromatic aldehydes as well as alkyl aldehydes. Moreover, the catalyst was easily removed from the reaction media trough a simple extraction and recycling [82]. Later, the synthesis of meso-aryldipyrromethanes using 1-propylsulfonic-3-methylimidazolium trifluoromethylacetate as a catalyst was described, however, organic solvents were used [83].
The research group of Majee explored the use of imidazolium zwitterionic molten salt as organocatalyst in the synthesis of meso-substituted dipyrromethanes under solvent-free conditions (Scheme 25) [84]. This catalytic system acts as an electrophilic activator of the aldehyde by the hydrogen bond with imidazolium C-2 hydrogen, emphasizing the importance of this cationic moiety. The reaction of pyrrole or N-methylpyrrole with aromatic or aliphatic aldehydes (ratio 2:1), in the presence of 10 mol% of the catalyst at room temperature and without solvent, gave access to a wide range of dipyrromethanes in very high yields. Aromatic aldehydes containing electron withdrawing or donating groups react with pyrrole or N-methylpyrrole to give dipyrromethanes 108a in yields from 72% to 87%. Dipyrromethanes with a naphthyl 108b, pyrrole or indole 108c, and propyl 108d substituents were also synthesized in high yields, using the same imidazolium zwitterionic molten salt as catalyst. The methodology developed was seen by the authors as being a green synthetic protocol, because it was metal- and solvent-free, and environmentally friendly with a good atom economy [84].
meso-Acetyldipyrromethane 110 was synthesized in 70% yield by the reaction of methylglyoxal 109 with pyrrole, in a 1:2.5 ratio, using boric acid as catalyst in aqueous media (Scheme 26) [85]. In addition to dipyrromethane 110, other dipyrromethanes were synthesized using aromatic aldehydes and similar reaction conditions. Boric acid is weakly acidic and reacts with water decreasing the pH of the aqueous layer; given that the reaction of pyrrole with aldehydes occurs in the interface with the organic layer, the formation of side products is thus prevented.

3.2. Dipyrromethane Synthesis via Alternative Methods

Pinho e Melo and co-workers developed an on-water one-pot synthetic approach to meso-substituted dipyrromethanes via hetero-Diels-Alder reaction (or conjugated addition) of nitrosoalkenes and azoalkenes with pyrrole (Scheme 27) [86,87]. Dehydrohalogenation of α,α-dihalooximes or α,α-dihalohydrazones, in the presence of base, produces transient nitrosoalkenes or azoalkenes I. These reactive species react with pyrrole to give pyrroles II functionalized at C-2 with a side chain, which undergo dehydrohalogenation to form the second transient nitrosoalkenes or azoalkenes III. The reaction with another molecule of pyrrole gives the dipyrromethanes 112 or 113, in moderate to high yields (21%–82%). The formation of dipyrromethanes 112 and 113 are accelerated and more efficient using water as solvents allowing easier purification procedures than the reaction performed in dichloromethane or in the absence of solvent. Dipyrromethanes synthesized by this approach have the unique feature of being meso-substituted with oxime and hydrazone moieties [86]. The same research group described the functionalization of dipyrromethanes at positions 1 and/or 9 through hetero-Diels-Alder reaction or conjugated addition of nitrosoalkenes and azoalkenes [88,89,90].
The one-pot synthesis of ortho-hydroxymethyl 8-C-aryl BODIPY derivatives 117 was achieved through the key intermediate dipyrromethanes 116 (Scheme 28) [91]. Ethyl phthalidinium salts 115, obtained by O-ethylation of phathalides 114 using Meerweins reagent, reacted with pyrrole to form intermediate ketal I. Elimination of the ethoxy group from intermediate I gave the oxonium ion II, which reacted with a second pyrrole unit to produce dipyrromethane 116. Reaction of dipyrromethanes 116 with BF3.OEt2 gave the BODIPY derivatives 117 in moderate yields (26%–45%). The masked 5-alkoxy-5-phenyldipyrromethane 116 with R1 = H, was isolated and treated with BF3.OEt2 in order to confirm that this is a key intermediate in the synthesis of the corresponding borondipyrromethenes 117.
Borbas and Xiong developed a strategy to synthesize unsymmetrical dipyrromethanes 120 through the Mannich reaction between pyrroles and Eschenmoser’s salt (Scheme 29) [92]. Initially, pyrroles 118 reacted with Eschenmoser’s salt to give the Mannich product 119, which undergo substitution of the N,N-dimethylamino group under microwave irradiation using pyrrole as reactant and solvent. This method encompasses acid-sensitive and formyl groups and does not require the use of acid to activate the pyrrole unit.
meso- and α-Unsubstituted dipyrromethanes 124 were formed by the decarboxylation of 1,9-diethoxycarbonyldiyrromethanes 123 with KOH in ethylene glycol (Scheme 30) [93]. Bromination of the α-methyl group of pyrrole 121, followed by nucleophilic substitution generated α-acethoxymethyl pyrroles 122, which underwent self-condensation in the presence of HCl to give meso-unsubstituted dipyrromethanes 123 in good yields. Dipyrromethanes 124 are key intermediates in the synthesis of porphyrins 125, that are meso-unsubstituted and β-substituted.
Thompson and co-workers developed a methodology to generate dipyrromethanes through the microwave-assisted reduction of F-BODIPYs and dipyrromethenes (Scheme 31) [94]. meso-Aryl BODIPYs 126 or dipyrromethenes 127 are reduced to the corresponding dipyrromethanes 128 in ethylene glycol and an excess of sodium methoxide under microwave irradiation at 215 °C for 10 minutes. This methodology is useful when BODIPYs or dipyrromethenes are formed in one-pot procedures and it is necessary to regenerate the dipyrromethane.
Neo-confused porphyrins 133 have been synthesized from the reaction of neo-confused dipyrromethanes 131 with dipyrromethane 132 (Scheme 32) [95]. The treatment of pyrrole-3-carboxaldehyde 129 with NaH in DMF at room temperature, followed by addition of methyl 4-formylpyrrole-2-carboxylate (130) gave the corresponding neo-confused dipyrromethanes 131 in good yields (45%–75%).

4. Conclusions

Given the continual relevance of the dipyrromethane scaffold, either by its own merits and applications or because of its extreme usefulness as a synthetic intermediate for other high value molecules, e.g., calix[4]pyrroles, (hydro)porphyrins, expanded porphyrins, corroles, BODIPY dyes, and metal coordination compounds, classical synthetic methods employing effective tried-and-tested catalysts still find their place on laboratory benches across the world. Nonetheless, as can be realized from this literature review covering the past six years, it is highly expected that organic and medicinal chemists, as well as material scientists, will keep pursuing innovative technologies and/or novel synthetic approaches with the aim of obtaining original, interesting, and much needed dipyrromethane structures.

Author Contributions

Literature review and original draft preparation, B.F.O.N. and S.M.M.L.; final draft review and editing, M.P. and T.M.V.D.P.eM.

Funding

This research was supported by the Portuguese Foundation for Science and Technology (FCT), co-funded by the European Regional Development Fund (FEDER) through COMPETE 2020 and through PT 2020/CENTRO 2020 (CENTRO-01-0145-FEDER-000014/MATIS). The Coimbra Chemistry Centre (CQC) was also supported via project UID/QUI/00313/2019.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lindsey, J.S. Synthetic routes to meso-patterned porphyrins. Acc. Chem. Res. 2010, 43, 300–311. [Google Scholar] [CrossRef] [PubMed]
  2. Yedukondalu, M.; Ravikanth, M. Core-modified porphyrin based assemblies. Coord. Chem. Rev. 2011, 255, 547–573. [Google Scholar] [CrossRef]
  3. Gale, P.A.; Anzenbacher, P.; Sessler, J.S. Calixpyrroles II. Coord. Chem. Rev. 2001, 222, 57–102. [Google Scholar]
  4. 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]
  5. Orłowski, R.; Gryko, D.; Gryko, D.T. Synthesis of corroles and their heteroanalogs. Chem. Rev. 2017, 117, 3102–3137. [Google Scholar] [CrossRef]
  6. Pareek, Y.; Ravikanth, M.; Chandrashekar, T.K. Smaragdyrins: Emeralds of expanded porphyrin family. Acc. Chem. Res. 2012, 45, 1801–1816. [Google Scholar] [CrossRef] [PubMed]
  7. Chatterjee, T.; Srinivasan, A.; Ravikanth, M.; Chandrashekar, T.K. Smaragdyrins and sapphyrins analogues. Chem. Rev. 2017, 117, 3329–3376. [Google Scholar] [CrossRef]
  8. Wood, T.E.; Thompson, A. Advances in the chemistry of dipyrrins and their complexes. Chem. Rev. 2007, 107, 1831–1861. [Google Scholar] [CrossRef] [PubMed]
  9. Clarke, R.C.; Hall, M.J. Recent developments in the synthesis of the BODIPY dyes. In Advances in Heterocyclic Chemistry; Eric, F.V., Scriven, C.A.R., Eds.; Elsevier: Amsterdam, the Netherlands, 2019; Volume 128, pp. 181–261. [Google Scholar]
  10. Loudet, A.; Burgess, K. BODIPY® dyes and their derivatives: Syntheses and spectroscopic properties. In Handbook of Porphyrin Science; Kadish, K.M., Smith, K.M., Guilard, R., Eds.; World Scientific: Hackensack, NJ, USA, 2010; Volume 8, pp. 1–164. [Google Scholar]
  11. Gianapoulos, C.G.; Kirschbaum, K.; Mason, M.R. Mono- and bimetallic aluminum alkyl, aryl, and hydride complexes of a bulky dipyrromethene ligand. Organometallics 2014, 33, 4503–4511. [Google Scholar] [CrossRef]
  12. Gianapoulos, C.G.; Kumar, A.; Zhao, Y.; Jia, L.; Kirschbaum, K.; Mason, M.R. Aluminum alkoxide, amide and halide complexes supported by a bulky dipyrromethene ligand: Synthesis, characterization, and preliminary ε-caprolactone polymerization activity. Dalton Trans. 2016, 45, 13787–13797. [Google Scholar] [CrossRef] [PubMed]
  13. Cohen, S.M.; Halper, S.R. Dipyrromethene complexes of iron. Inorg. Chim. Acta 2002, 341, 12–16. [Google Scholar] [CrossRef]
  14. King, E.R.; Betley, T.A. C-H bond amination from a ferrous dipyrromethene complex. Inorg. Chem. 2009, 48, 2361–2363. [Google Scholar] [CrossRef] [PubMed]
  15. King, E.R.; Hennessy, E.T.; Betley, T.A. Catalytic C-H bond amination from high-spin iron imido complexes. J. Am. Chem. Soc. 2011, 133, 4917–4923. [Google Scholar] [CrossRef] [PubMed]
  16. Weng, W.; Parkin, S.; Ozerov, O.V. Double C-H activation results in ruthenium complexes of a neutral PCP ligand with a central carbene moiety. Organometallics 2006, 25, 5345–5354. [Google Scholar] [CrossRef]
  17. Swavey, S.; DeBeer, M.; Li, K. Photoinduced interactions of supramolecular ruthenium(II) complexes with plasmid DNA: Synthesis and spectroscopic, electrochemical, and DNA photocleavage studies. Inorg. Chem. 2015, 54, 3139–3147. [Google Scholar] [CrossRef]
  18. Majumder, S.; Odom, A.L. Group-4 dipyrrolylmethane complexes in intramolecular olefin hydroamination. Organometallics 2008, 27, 1174–1177. [Google Scholar] [CrossRef]
  19. Yadav, M.; Singh, A.K.; Pandey, R.; Pandey, D.S. Synthesis and characterization of complexes imparting N-pyridyl bonded meso-pyridyl substituted dipyrromethanes. J. Organometall. Chem. 2010, 695, 841–849. [Google Scholar] [CrossRef]
  20. Yin, Z.; Yan, Y.; Sun, S.; Wang, W. Syntheses, structures, and properties of Ni(II) complexes with 5,5’-bis(4-halogenphenyl)diazo-dipyrromethane. J. Coord. Chem. 2012, 65, 865–874. [Google Scholar] [CrossRef]
  21. King, E.R.; Betley, T.A. Unusual electronic structure of first row transition metal complexes featuring redox-active dipyrromethane ligands. J. Am. Chem. Soc 2009, 131, 14374–14380. [Google Scholar] [CrossRef]
  22. Reid, S.D.; Wilson, C.; Blake, A.J.; Love, J.B. Tautomerisation and hydrogen-bonding interactions in four-coordinate metal halide and azide complexes of N-donor-extended dipyrromethanes. Dalton Trans. 2010, 39, 418–425. [Google Scholar] [CrossRef]
  23. Tamaru, S.-I.; Yu, L.; Youngblood, W.J.; Muthukumaran, K.; Taniguchi, M.; Lindsey, J.S. A tin-complexation strategy for use with diverse acylation methods in the preparation of 1,9-diacyldipyrromethanes. J. Org. Chem. 2004, 69, 765–777. [Google Scholar] [CrossRef] [PubMed]
  24. Cabeza, J.A.; Fernández, I.; García-Álvarez, P.; Laglera-Gándara, C.J. A dipyrromethane-based diphosphane–germylene as precursor to tetrahedral copper(I) and T-shaped silver(I) and gold(I) PGeP pincer complexes. Dalton Trans. 2019, 48, 13273–13280. [Google Scholar] [CrossRef] [PubMed]
  25. Alešković, M.; Basarić, N.; Mlinarić-Majerski, K.; Molčanov, K.; Kojić-Prodić, B.; Kesharwani, M.K.; Ganguly, B. Anion recognition through hydrogen bonding by adamantane-dipyrromethane receptors. Tetrahedron 2010, 66, 1689–1698. [Google Scholar] [CrossRef]
  26. You, J.-M.; Jeong, H.; Seo, H.; Jeon, S. A new fluoride ion colorimetric sensor based on dipyrrolemethanes. Sens. Actuators B Chem. 2010, 146, 160–164. [Google Scholar] [CrossRef]
  27. Kataev, E.A.; Müller, C.; Kolesnikov, G.V.; Khrustalev, V.N. Guanidinium-based artificial receptors for binding orthophosphate in aqueous solution. Eur. J. Org. Chem. 2014, 2014, 2747–2753. [Google Scholar] [CrossRef]
  28. Muwal, P.K.; Nayal, A.; Jaiswal, M.K.; Pandey, P.S. A dipyrromethane based receptor as a dual colorimetric sensor for F and Cu2+ ions. Tetrahedron Lett. 2018, 59, 29–32. [Google Scholar] [CrossRef]
  29. Susmel, S.; Comuzzi, C. Selectivity and efficiency of conductive molecularly imprinted polymer (c-MIP) based on 5-phenyl-dipyrromethane and 5-phenol-dipyrromethane for quorum sensing precursors detection. Chemosensors 2017, 5, 5. [Google Scholar] [CrossRef]
  30. Pereira, N.A.M.; Pinho e Melo, T.M.V.D. Recent developments in the synthesis of dipyrromethanes. A review. Org. Prep.Proc. Int. 2014, 46, 183–213. [Google Scholar] [CrossRef]
  31. Gryko, D.T.; Gryko, D.; Lee, C.-H. 5-Substituted dipyrranes: Synthesis and reactivity. Chem. Soc. Rev. 2012, 41, 3780–3789. [Google Scholar] [CrossRef]
  32. Boens, N.; Leen, V.; Dehaen, W. Fluorescent indicators based on BODIPY. Chem. Soc. Rev. 2012, 41, 1130–1172. [Google Scholar] [CrossRef]
  33. Nagarkatti, J.P.; Ashley, K.R. Synthesis of pyridyl meso substituted dipyrrylmethanes. Synthesis 1974, 186–187. [Google Scholar] [CrossRef]
  34. Baskın, D.; Çetinkaya, Y.; Balci, M. Synthesis of dipyrrolo-diazepine derivatives via intramolecular alkyne cyclization. Tetrahedron 2018, 74, 4062–4070. [Google Scholar] [CrossRef]
  35. Swavey, S.; Li, K. A dimetallic osmium(II) complex as a potential phototherapeutic agent: Binding and photocleavage studies with plasmid DNA. Eur. J. Inorg. Chem. 2015, 2015, 5551–5555. [Google Scholar] [CrossRef]
  36. Vicente, M.G.H.; Smith, K.M. Syntheses and functionalizations of porphyrin macrocycles. Curr. Org. Synth. 2014, 11, 3–28. [Google Scholar] [CrossRef] [Green Version]
  37. Neya, S.; Yoneda, T.; Hoshino, T.; Kawaguchi, A.T.; Suzuki, M. Synthesis of type III isomers of diacetyldeutero-, hemato-, and protoporphyrins with the use of Knorr’s pyrrole. Tetrahedron 2016, 72, 4022–4026. [Google Scholar] [CrossRef]
  38. Neya, S.; Yoneda, T.; Omori, H.; Hoshino, T.; Kawaguchi, A.T.; Suzuki, M. Synthesis of 1,4,5,8-tetraethyl-2,3,6,7-tetravinylporphyrin from a Knorr’s pyrrole analogue. Tetrahedron 2017, 73, 6780–6785. [Google Scholar] [CrossRef]
  39. Singh, R.N.; Rawat, P.; Kumar, A.; Kant, P.; Srivastava, A. Spectral analysis, chemical reactivity and first hyperpolarizability evaluation of a novel 1,9–bis(2–cyano–2–ethoxycarbonylvinyl)–5–(2–furyl)–dipyrromethane: Experimental and theoretical approaches. Spectrosc. Lett. 2015, 48, 235–250. [Google Scholar] [CrossRef]
  40. Rawat, P.; Singh, R.N.; Niranjan, P.; Ranjan, A.; Holguín, N.R.F. Evaluation of antituberculosis activity and DFT study on dipyrromethane-derived hydrazone derivatives. J. Mol. Struct. 2017, 1149, 539–548. [Google Scholar] [CrossRef]
  41. Mahanta, S.P.; Panda, P.K. 5,10-Diacylcalix[4]pyrroles: Synthesis and anion binding studies. Org. Biomol. Chem. 2014, 12, 278–285. [Google Scholar] [CrossRef]
  42. Sen, P.; Yildiz, S.Z.; Atalay, Y.; Dege, N.; Demirtas, G. The synthesis, characterization, crystal structure and theoretical calculations of a new meso-BODIPY substituted phthalonitrile. J. Lumin. 2014, 149, 297–305. [Google Scholar] [CrossRef]
  43. Sen, P.; Atmaca, G.Y.; Erdoğmuş, A.; Dege, N.; Genç, H.; Atalay, Y.; Yildiz, S.Z. The synthesis, characterization, crystal structure and photophysical properties of a new meso-BODIPY substituted phthalonitrile. J. Fluoresc. 2015, 25, 1225–1234. [Google Scholar] [CrossRef]
  44. Radzuan, N.H.M.; Malek, N.H.A.; Ngatiman, M.F.; Xin, T.K.; Bakar, M.B.; Hassan, N.I.; Bakar, M.A. Synthesis and X-ray single crystal study of 5-(4,4,5,5–tetramethyl–1,3,2–dioxoborolane)-10,20–diphenylporphyrin. Sains Malays. 2018, 47, 2083–2090. [Google Scholar] [CrossRef]
  45. Umasekhar, B.; Ganapathi, E.; Chatterjee, T.; Ravikanth, M. Synthesis, structure, and spectral, electrochemical and fluoride sensing properties of meso-pyrrolyl boron dipyrromethene. Dalton Trans. 2015, 44, 16516–16527. [Google Scholar] [CrossRef]
  46. Wałęsa-Chorab, M.; Banasz, R.; Kubicki, M.; Patroniak, V. Dipyrromethane functionalized monomers as precursors of electrochromic polymers. Electrochim. Acta 2017, 258, 571–581. [Google Scholar] [CrossRef]
  47. Swamy, C.A.; Pakkirisamy, T. Effect of substituent position on optical properties of boron-dipyrromethane isomers. Inorg. Chim. Acta 2014, 411, 97–101. [Google Scholar]
  48. Golf, H.R.A.; Reissig, H.-U.; Wiehe, A. Synthesis of 5-substituted tetrapyrroles, metalloporphyrins, BODIPYs, and their dipyrrane precursors. J. Org. Chem. 2015, 80, 5133–5143. [Google Scholar] [CrossRef]
  49. Gutsche, C.S.; Hohlfeld, B.F.; Flanagan, K.J.; Senge, M.O.; Kulak, N.; Wiehe, A. Sequential nucleophilic substitution of the α-pyrrole and p-aryl positions of meso-pentafluorophenyl-substituted BODIPYs. Eur. J. Org. Chem. 2017, 2017, 3187–3196. [Google Scholar] [CrossRef]
  50. Hohlfeld, B.F.; Flanagan, K.J.; Kulak, N.; Senge, M.O.; Christmann, M.; Wiehe, A. Synthesis of porphyrinoids, BODIPYs, and (dipyrrinato)ruthenium(II) complexes from prefunctionalized dipyrromethanes. Eur. J. Org. Chem. 2019, 2019, 4020–4033. [Google Scholar] [CrossRef]
  51. Deliomeroglu, M.K.; Lynch, V.M.; Sessler, J.L. Conformationally switchable non-cyclic tetrapyrrole receptors: Synthesis of tetrakis(1H-pyrrole-2-carbaldehyde) derivatives and their anion binding properties. Chem. Commun. 2014, 50, 11863–11866. [Google Scholar] [CrossRef]
  52. Deliomeroglu, M.K.; Lynch, V.M.; Sessler, J.L. Non-cyclic formylated dipyrromethanes as phosphate anion receptors. Chem. Sci. 2016, 7, 3843–3850. [Google Scholar] [CrossRef] [Green Version]
  53. Pankhurst, J.R.; Cadenbach, T.; Betz, D.; Finn, C.; Love, J.B. Towards dipyrrins: Oxidation and metalation of acyclic and macrocyclic Schiff-base dipyrromethanes. Dalton Trans. 2015, 44, 2066–2070. [Google Scholar] [CrossRef] [Green Version]
  54. Temelli, B.; Kalkan, H. Unexpected formation of β, meso-directly linked diporphyrins under Adler–Longo reaction conditions. Synth. Commun. 2018, 48, 2112–2117. [Google Scholar] [CrossRef]
  55. Basumatary, B.; Reddy, R.V.R.; Bhandary, S.; Sanka, J. Gallium(III)corrole–BODIPY hybrid: Novel photophysical properties and first observation of b–f⋯f interactions. Dalton Trans. 2015, 44, 20817–20821. [Google Scholar] [CrossRef]
  56. He, R.B.; Yue, H.; Kong, J.H. Covalent porphyrin hybrids linked with dipyrrin, bidipyrrin or thiacorrole. Molecules 2017, 22, 1400. [Google Scholar]
  57. Lakshmi, V.; Lee, W.-Z.; Ravikanth, M. Synthesis, structure and spectral and electrochemical properties of 3-pyrrolyl BODIPY-metal dipyrrin complexes. Dalton Trans. 2014, 43, 16006–16014. [Google Scholar] [CrossRef]
  58. Sharma, R.; Gobeze, H.B.; D’Souza, F.; Ravikanth, M. Panchromatic light capture and efficient excitation transfer leading to near-ir emission of BODIPY oligomers. ChemPhysChem 2016, 17, 2516–2524. [Google Scholar] [CrossRef]
  59. Brem, B.; Gal, E.; Gaina, L.; Cristea, C.; Silaghi-Dumitrescu, L. Synthesis of novel (phenothiazinyl)dipyrrolylmethanes. Revue Roumaine De Chimie 2014, 59, 947–952. [Google Scholar]
  60. Brem, B.; Gal, E.; Gaina, L.; Lovasz, T.; Molnar, E.A.; Porumb, D.; Cristea, C. Novel 1,9-diacyl-5-(phenothiazinyl)dipyrromethane dialkyltin complexes. Studia Universitatis Babes-Bolyai Chemia 2016, 61, 73–80. [Google Scholar]
  61. Swamy, C.A.; Priyanka, R.N.; Thilagar, P. Triarylborane–dipyrromethane conjugates bearing dual receptor sites: The synthesis and evaluation of the anion binding site preference. Dalton Trans. 2014, 43, 4067–4075. [Google Scholar] [CrossRef]
  62. Bagherzadeh, M.; Jonaghani, M.A.; Amini, M.; Mortazavi-Manesh, A. Synthesis of dipyrromethanes in water and investigation of electronic and steric effects in efficiency of olefin epoxidation by sodium periodate catalyzed by manganese tetraaryl and trans disubstituted porphyrin complexes. J. Porphyr. Phthalocyan. 2019, 23, 671–678. [Google Scholar] [CrossRef]
  63. Sobral, A.l.J.F.N.; Rebanda, N.G.C.L.; da Silva, M.; Lampreia, S.H.; Ramos Silva, M.; Beja, A.M.; Paixão, J.A.; Rocha Gonsalves, A.M.d.A. One-step synthesis of dipyrromethanes in water. Tetrahedron Lett. 2003, 44, 3971–3973. [Google Scholar] [CrossRef] [Green Version]
  64. Guo, C.; Sun, S.; He, Q.; Lynch, V.M.; Sessler, J.L. Pyrene-linked formylated bis(dipyrromethane): A fluorescent probe for dihydrogen phosphate. Org. Lett. 2018, 20, 5414–5417. [Google Scholar] [CrossRef]
  65. Kumar, S.; Gobeze, H.B.; Chatterjee, T.; D’Souza, F.; Ravikanth, M. Directly connected azaBODIPY–BODIPY dyad: Synthesis, crystal structure, and ground- and excited-state interactions. J. Phys. Chem. A 2015, 119, 8338–8348. [Google Scholar] [CrossRef]
  66. Kumar, A.; Kumar, S.; Chatterjee, T.; Ravikanth, M. β-meso covalently linked azaBODIPY-Pd(II) dipyrrin conjugate. ChemistrySelect 2016, 1, 94–100. [Google Scholar] [CrossRef]
  67. Koch, A.; Ravikanth, M. Monofunctionalized 1,3,5,7-tetraarylazaBODIPYs and their application in the synthesis of azaBODIPY based conjugates. J. Org. Chem. 2019, 84, 10775–10784. [Google Scholar] [CrossRef]
  68. Epelde-Elezcano, N.; Palao, E.; Manzano, H.; Prieto-CastaÇeda, A.; Agarrabeitia, A.R.; Tabero, A.; Villanueva, A.; Moya, S.; López-Arbeloa, I.; Martínez-Martínez, V.; et al. Rational design of advanced photosensitizers based on orthogonal BODIPY dimers to finely modulate singlet oxygen generation. Chem. Eur. J. 2017, 23, 4837–4848. [Google Scholar] [CrossRef]
  69. Esipova, T.V.; Vinogradov, S.A. Synthesis of phosphorescent asymmetrically π-extended porphyrins for two-photon applications. J. Org. Chem. 2014, 79, 8812–8825. [Google Scholar] [CrossRef] [Green Version]
  70. Sahin, T.; Vairaprakash, P.; Borbas, K.E.; Balasubramanian, T.; Lindsey, J.S. Hydrophilic bioconjugatable trans-AB-porphyrins and peptide conjugates. J. Porphyr. Phthalocyan. 2015, 19, 663–678. [Google Scholar] [CrossRef]
  71. Smoleń, S.; Walaszek, D.J.; Karczewski, M.; Martin, E.; Gryko, D. Towards no-free regulation of sGC: Design and synthesis of trans-AB-porphyrins. Isr. J. Chem. 2016, 56, 156–168. [Google Scholar] [CrossRef]
  72. Emandi, G.; Shaker, Y.M.; Flanagan, K.J.; O’Brien, J.M.; Senge, M.O. Merging triptycene, BODIPY and porphyrin chemistry: Synthesis and properties of mono- and trisubstituted triptycene dye arrays. Eur. J. Org. Chem. 2017, 2017, 6680–6692. [Google Scholar] [CrossRef]
  73. Petrushenko, K.B.; Petrushenko, I.K.; Petrova, O.V.; Sobenina, L.N.; Ushakov, I.A.; Trofimov, B.A. Environment-responsive 8-CF3-BODIPY dyes with aniline groups at the 3 position: Synthesis, optical properties and RI-CC2 calculations. Asian J. Org. Chem. 2017, 6, 852–861. [Google Scholar] [CrossRef]
  74. Tomilin, D.N.; Sobenina, L.N.; Petrushenko, K.B.; Ushakov, I.A.; Trofimov, B.A. Design of novel meso-CF3-BODIPY dyes with isoxazole substituents. Dyes Pigm. 2018, 152, 14–18. [Google Scholar] [CrossRef]
  75. Kashi, C.; Wu, C.-C.; Mai, C.-L.; Yeh, C.-Y.; Chang, C.K. Synthesis of octafluoroporphyrin. Angew. Chem. Int. Ed. 2016, 55, 5035–5039. [Google Scholar] [CrossRef]
  76. Clezy, P.S.; Smythe, G.A. The chemistry of pyrrolic compounds. VIII. Dipyrrylthiones. Aust. J. Chem. 1969, 22, 239–249. [Google Scholar] [CrossRef]
  77. Rangaraj, P.; Parshamoni, S.; Konar, S. MOF as a syringe pump for the controlled release of iodine catalyst in the synthesis of meso-thienyl dipyrromethanes. Chem. Commun. 2015, 51, 15526–15529. [Google Scholar] [CrossRef]
  78. Megarajan, S.; Ayaz Ahmed, K.B.; Rajmohan, R.; Vairaprakash, P.; Anbazhagan, V. An easily accessible and recyclable copper nanoparticle catalyst for the solvent-free synthesis of dipyrromethanes and aromatic amines. RSC Adv. 2016, 6, 103065–103071. [Google Scholar] [CrossRef]
  79. Rajmohan, R.; Ayaz Ahmed, K.B.; Sangeetha, S.; Anbazhagan, V.; Vairaprakash, P. C–H oxidation and chelation of a dipyrromethane mediated rapid colorimetric naked-eye Cu(II) chemosensor. Analyst 2017, 142, 3346–3351. [Google Scholar] [CrossRef]
  80. Rajaswathi, K.; Jayanthi, M.; Rajmohan, R.; Anbazhagan, V.; Vairaprakash, P. Simple admixture of 4-nitrobenzaldehyde and 2,4-dimethylpyrrole for efficient colorimetric sensing of copper(II) ions. Spectrochim. Acta A 2019, 212, 308–314. [Google Scholar] [CrossRef]
  81. Jaratjaroonphong, J.; Tuengpanya, S.; Saeeng, R.; Udompong, S.; Srisook, K. Green synthesis and anti-inflammatory studies of a series of 1,1-bis(heteroaryl)alkane derivatives. Eur. J. Med. Chem. 2014, 83, 561–568. [Google Scholar] [CrossRef]
  82. Senapak, W.; Saeeng, R.; Jaratjaroonphong, J.; Kasemsuk, T.; Sirion, U. Green synthesis of dipyrromethanes in aqueous media catalyzed by SO3H-functionalized ionic liquid. Org. Biomol. Chem. 2016, 14, 1302–1310. [Google Scholar] [CrossRef]
  83. Rawat, A.K. Highly efficient synthesis of aryldipyrromethanes in presence of bronsted acidic ionic liquids. Heterocycl. Lett. 2018, 8, 43–47. [Google Scholar]
  84. Chatterjee, R.; Mahato, S.; Santra, S.; Zyryanov, G.V.; Hajra, A.; Majee, A. Imidazolium zwitterionic molten salt: An efficient organocatalyst under neat conditions at room temperature for the synthesis of dipyrromethanes as well as bis(indolyl)methanes. ChemistrySelect 2018, 3, 5843–5847. [Google Scholar] [CrossRef]
  85. Singhal, A.; Singh, S.; Chauhan, S.M.S. Synthesis of dipyrromethanes in aqueous media using boric acid. Arkivoc 2016, vi, 144–151. [Google Scholar] [CrossRef] [Green Version]
  86. Pereira, N.A.M.; Lopes, S.M.M.; Lemos, A.; Pinho e Melo, T.M.V.D. On-water synthesis of dipyrromethanes via bis-hetero-diels-alder reaction of azo- and nitrosoalkenes with pyrrole. Synlett 2014, 25, 423–427. [Google Scholar] [CrossRef]
  87. Lopes, S.M.M.; Cardoso, A.L.; Lemos, A.; Pinho e Melo, T.M.V.D. Recent advances in the chemistry of conjugated nitrosoalkenes and azoalkenes. Chem. Rev. 2018, 118, 11324–11352. [Google Scholar] [CrossRef]
  88. Lopes, S.M.M.; Lemos, A.; Pinho e Melo, T.M.V.D. Reactivity of dipyrromethanes towards azoalkenes: Synthesis of functionalized dipyrromethanes, calix[4]pyrroles, and bilanes. Eur. J. Org. Chem. 2014, 2014, 7039–7048. [Google Scholar] [CrossRef]
  89. Nunes, S.C.C.; Lopes, S.M.M.; Gomes, C.S.B.; Lemos, A.; Pais, A.; Pinho e Melo, T.M.V.D. Reactions of nitrosoalkenes with dipyrromethanes and pyrroles: Insight into the mechanistic pathway. J. Org. Chem. 2014, 79, 10456–10465. [Google Scholar] [CrossRef]
  90. Jorda, R.; Lopes, S.M.M.; Řezníčková, E.; Kryštof, V.; Pinho e Melo, T.M.V.D. Biological evaluation of dipyrromethanes in cancer cell lines: Antiproliferative and pro-apoptotic properties. ChemMedChem 2017, 12, 701–711. [Google Scholar] [CrossRef]
  91. del Río, M.; Lobo, F.; López, J.C.; Oliden, A.; Bañuelos, J.; López-Arbeloa, I.; Garcia-Moreno, I.; Gómez, A.M. One-pot synthesis of rotationally restricted, conjugatable, BODIPY derivatives from phthalides. J. Org. Chem. 2017, 82, 1240–1247. [Google Scholar] [CrossRef]
  92. Xiong, R.; Borbas, K.E. Mild microwave-assisted synthesis of dipyrromethanes and their analogues. Synlett 2015, 26, 484–488. [Google Scholar]
  93. Lyubimova, T.V.; Syrbu, S.A.; Semeikin, A.S. Synthesis of porphyrins from alpha-unsubstituted dipyrromethanes. Macroheterocycles 2016, 9, 59–64. [Google Scholar] [CrossRef] [Green Version]
  94. Melanson, J.A.; Smithen, D.A.; Cameron, T.S.; Thompson, A. Microwave-assisted reduction of F-BODIPYs and dipyrrins to generate dipyrromethanes. Can. J. Chem. 2014, 92, 688–694. [Google Scholar] [CrossRef]
  95. Li, R.; Lammer, A.D.; Ferrence, G.M.; Lash, T.D. Synthesis, structural characterization, aromatic characteristics, and metalation of neo-confused porphyrins, a newly discovered class of porphyrin isomers. J. Org. Chem. 2014, 79, 4078–4093. [Google Scholar] [CrossRef]
Figure 1. Dipyrromethane applications.
Figure 1. Dipyrromethane applications.
Molecules 24 04348 g001
Scheme 1. Synthesis of meso-disubstituted dipyrromethanes 2 and 5 featured in ion receptor dipyrromethanes 3 and 6, respectively.
Scheme 1. Synthesis of meso-disubstituted dipyrromethanes 2 and 5 featured in ion receptor dipyrromethanes 3 and 6, respectively.
Molecules 24 04348 sch001
Scheme 2. Synthesis of dipyrromethanes 7, alkyne-substituted dipyrromethanes 9 and dipyrrolo-diazepine derivatives 10.
Scheme 2. Synthesis of dipyrromethanes 7, alkyne-substituted dipyrromethanes 9 and dipyrrolo-diazepine derivatives 10.
Molecules 24 04348 sch002
Scheme 3. Synthesis of diphenanthrolinepyrromethanes 12 and their corresponding dimetallic Ru(II) 13 and Os(II) 14 coordination complexes.
Scheme 3. Synthesis of diphenanthrolinepyrromethanes 12 and their corresponding dimetallic Ru(II) 13 and Os(II) 14 coordination complexes.
Molecules 24 04348 sch003
Scheme 4. Synthesis of dipyrromethanes 1620 and protoporphyrin III 21 starting from Knorr’s pyrrole 15.
Scheme 4. Synthesis of dipyrromethanes 1620 and protoporphyrin III 21 starting from Knorr’s pyrrole 15.
Molecules 24 04348 sch004
Scheme 5. Synthesis of dipyrromethanes 2325 and 1,4,5,8-tetraethyl-2,3,6,7-tetravinylporphyrin 26 starting from Knorr’s pyrrole analogue 22.
Scheme 5. Synthesis of dipyrromethanes 2325 and 1,4,5,8-tetraethyl-2,3,6,7-tetravinylporphyrin 26 starting from Knorr’s pyrrole analogue 22.
Molecules 24 04348 sch005
Scheme 6. Meso-Dipyrromethanes 30 and 31 synthesized under standard p-TSA-catalyzed conditions.
Scheme 6. Meso-Dipyrromethanes 30 and 31 synthesized under standard p-TSA-catalyzed conditions.
Molecules 24 04348 sch006
Figure 2. Dipyrromethanes 32–37 synthesized under standard TFA-catalyzed conditions.
Figure 2. Dipyrromethanes 32–37 synthesized under standard TFA-catalyzed conditions.
Molecules 24 04348 g002
Scheme 7. Dipyrromethane derivatives 38 synthesized under typical TFA-catalyzed conditions and subsequent formylation to dipyrromethane products 39.
Scheme 7. Dipyrromethane derivatives 38 synthesized under typical TFA-catalyzed conditions and subsequent formylation to dipyrromethane products 39.
Molecules 24 04348 sch007
Scheme 8. Acyclic and macrocyclic dipyrromethane derivatives 41 and 43 synthesized under standard TFA-catalyzed Schiff-base conditions.
Scheme 8. Acyclic and macrocyclic dipyrromethane derivatives 41 and 43 synthesized under standard TFA-catalyzed Schiff-base conditions.
Molecules 24 04348 sch008
Scheme 9. Bis-dipyrromethanes 45 synthesized under normal TFA-catalyzed conditions.
Scheme 9. Bis-dipyrromethanes 45 synthesized under normal TFA-catalyzed conditions.
Molecules 24 04348 sch009
Scheme 10. Synthesis of dipyrromethane derivatives with tetrapyrrolic macrocycle substituents.
Scheme 10. Synthesis of dipyrromethane derivatives with tetrapyrrolic macrocycle substituents.
Molecules 24 04348 sch010
Scheme 11. Synthesis of meso/meso-linked porphyrin-dipyrromethane 51 catalyzed by TFA.
Scheme 11. Synthesis of meso/meso-linked porphyrin-dipyrromethane 51 catalyzed by TFA.
Molecules 24 04348 sch011
Scheme 12. TFA-catalyzed synthesis of α-dipyrromethanyl 3-pyrrolyl BODIPY 53 required for the preparation of Pd(II), Re(I) and Ru(II) 3-pyrrolyl-BODIPY/dipyrromethenes complexes 55a-c, and 3-pyrrolyl BODIPY/BODIPY 54.
Scheme 12. TFA-catalyzed synthesis of α-dipyrromethanyl 3-pyrrolyl BODIPY 53 required for the preparation of Pd(II), Re(I) and Ru(II) 3-pyrrolyl-BODIPY/dipyrromethenes complexes 55a-c, and 3-pyrrolyl BODIPY/BODIPY 54.
Molecules 24 04348 sch012
Figure 3. Dipyrromethanes 56-58 synthesized under standard BF3.Et2O-catalyzed conditions.
Figure 3. Dipyrromethanes 56-58 synthesized under standard BF3.Et2O-catalyzed conditions.
Molecules 24 04348 g003
Scheme 13. Bis-dipyrromethane derivative 60 synthesized under typical BF3.Et2O-catalyzed conditions and subsequent formylation.
Scheme 13. Bis-dipyrromethane derivative 60 synthesized under typical BF3.Et2O-catalyzed conditions and subsequent formylation.
Molecules 24 04348 sch013
Scheme 14. BF3.Et2O-catalyzed synthesis of dipyrromethane-substituted azaBODIPY 63 required for the preparation of azaBODIPY/BODIPY 64 and Pd(II) azaBODIPY/dipyrromethene complex 65.
Scheme 14. BF3.Et2O-catalyzed synthesis of dipyrromethane-substituted azaBODIPY 63 required for the preparation of azaBODIPY/BODIPY 64 and Pd(II) azaBODIPY/dipyrromethene complex 65.
Molecules 24 04348 sch014
Figure 4. Dipyrromethanes 66-72 synthesized under standard InCl3-catalyzed conditions.
Figure 4. Dipyrromethanes 66-72 synthesized under standard InCl3-catalyzed conditions.
Molecules 24 04348 g004
Scheme 15. InCl3-catalyzed synthesis of triptycenyldipyrromethane 74 for the preparation of triptycene-substituted BODIPY, triptycene-substituted 3-pyrrolyl-BODIPY, trans-A2B2 triptycenylporphyrins and A3B triptycenylporphyrins.
Scheme 15. InCl3-catalyzed synthesis of triptycenyldipyrromethane 74 for the preparation of triptycene-substituted BODIPY, triptycene-substituted 3-pyrrolyl-BODIPY, trans-A2B2 triptycenylporphyrins and A3B triptycenylporphyrins.
Molecules 24 04348 sch015
Scheme 16. InCl3-catalyzed synthesis of tris-dipyrromethane-substituted triptycene 78 required for the preparation of tris-BODIPY-substituted triptycene.
Scheme 16. InCl3-catalyzed synthesis of tris-dipyrromethane-substituted triptycene 78 required for the preparation of tris-BODIPY-substituted triptycene.
Molecules 24 04348 sch016
Scheme 17. Synthesis of meso-trifluoromethyldipyrromethanes 85 and 86 from 2-aminophenyl-1H-pyrroles.
Scheme 17. Synthesis of meso-trifluoromethyldipyrromethanes 85 and 86 from 2-aminophenyl-1H-pyrroles.
Molecules 24 04348 sch017
Scheme 18. Synthesis of meso-trifluoromethyldipyrromethanes 92 and 93 from ethynylpyrroles.
Scheme 18. Synthesis of meso-trifluoromethyldipyrromethanes 92 and 93 from ethynylpyrroles.
Molecules 24 04348 sch018
Scheme 19. Synthesis of tetrafluorinated dipyrromethene 97 for the preparation of β-octafluoroporphyrins.
Scheme 19. Synthesis of tetrafluorinated dipyrromethene 97 for the preparation of β-octafluoroporphyrins.
Molecules 24 04348 sch019
Scheme 20. Synthesis of meso-thienyldipyrromethanes.
Scheme 20. Synthesis of meso-thienyldipyrromethanes.
Molecules 24 04348 sch020
Scheme 21. Synthesis of dipyrromethanes 7c,e, 58d, 103 and 104, catalyzed by CuNPs and celite-supported glycine.
Scheme 21. Synthesis of dipyrromethanes 7c,e, 58d, 103 and 104, catalyzed by CuNPs and celite-supported glycine.
Molecules 24 04348 sch021
Scheme 22. Formation of bis(dipyrromethene)copper(II) complex 105.
Scheme 22. Formation of bis(dipyrromethene)copper(II) complex 105.
Molecules 24 04348 sch022
Scheme 23. Synthesis of dipyrromethanes 7c, 103b and 106 catalyzed by iodine.
Scheme 23. Synthesis of dipyrromethanes 7c, 103b and 106 catalyzed by iodine.
Molecules 24 04348 sch023
Scheme 24. Synthesis of dipyrromethanes 7f and 107 catalyzed by a SO3H-functionalized ionic liquid.
Scheme 24. Synthesis of dipyrromethanes 7f and 107 catalyzed by a SO3H-functionalized ionic liquid.
Molecules 24 04348 sch024
Scheme 25. Synthesis of dipyrromethanes 108 catalyzed by an imidazolium zwitterionic molten salt.
Scheme 25. Synthesis of dipyrromethanes 108 catalyzed by an imidazolium zwitterionic molten salt.
Molecules 24 04348 sch025
Scheme 26. Synthesis of meso-acetyldipyrromethane catalyzed by boric acid.
Scheme 26. Synthesis of meso-acetyldipyrromethane catalyzed by boric acid.
Molecules 24 04348 sch026
Scheme 27. Synthesis of dipyrromethanes 112 and 113 based on the chemistry of nitrosoalkenes and azoalkenes.
Scheme 27. Synthesis of dipyrromethanes 112 and 113 based on the chemistry of nitrosoalkenes and azoalkenes.
Molecules 24 04348 sch027
Scheme 28. One-pot synthesis of ortho-hydroxymethyl 8-C-aryl BODIPY derivatives.
Scheme 28. One-pot synthesis of ortho-hydroxymethyl 8-C-aryl BODIPY derivatives.
Molecules 24 04348 sch028
Scheme 29. Synthesis of unsymmetrical dipyrromethanes 120.
Scheme 29. Synthesis of unsymmetrical dipyrromethanes 120.
Molecules 24 04348 sch029
Scheme 30. Synthesis and reactivity of meso- and α-unsubstituted dipyrromethanes.
Scheme 30. Synthesis and reactivity of meso- and α-unsubstituted dipyrromethanes.
Molecules 24 04348 sch030
Scheme 31. Microwave-assisted reduction of F-BODIPYs and dipyrromethenes to dipyrromethanes.
Scheme 31. Microwave-assisted reduction of F-BODIPYs and dipyrromethenes to dipyrromethanes.
Molecules 24 04348 sch031
Scheme 32. Synthesis and reactivity of neo-confused dipyrromethanes 131 and neo-confused porphyrins.
Scheme 32. Synthesis and reactivity of neo-confused dipyrromethanes 131 and neo-confused porphyrins.
Molecules 24 04348 sch032

Share and Cite

MDPI and ACS Style

Nascimento, B.F.O.; Lopes, S.M.M.; Pineiro, M.; Pinho e Melo, T.M.V.D. Current Advances in the Synthesis of Valuable Dipyrromethane Scaffolds: Classic and New Methods. Molecules 2019, 24, 4348. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules24234348

AMA Style

Nascimento BFO, Lopes SMM, Pineiro M, Pinho e Melo TMVD. Current Advances in the Synthesis of Valuable Dipyrromethane Scaffolds: Classic and New Methods. Molecules. 2019; 24(23):4348. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules24234348

Chicago/Turabian Style

Nascimento, Bruno F. O., Susana M. M. Lopes, Marta Pineiro, and Teresa M. V. D. Pinho e Melo. 2019. "Current Advances in the Synthesis of Valuable Dipyrromethane Scaffolds: Classic and New Methods" Molecules 24, no. 23: 4348. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules24234348

Article Metrics

Back to TopTop