Next Article in Journal
Gene-Regulatory Activity of α-Tocopherol
Next Article in Special Issue
Enhanced Reactivity of [Hydroxy(tosyloxy)iodo]benzene in Fluoroalcohol Media. Efficient Direct Synthesis of Thienyl(aryl)iodonium Salts
Previous Article in Journal
The Molecular Mechanism of Action of Artemisinin—The Debate Continues
Previous Article in Special Issue
Heterocyclic Analogs of Thioflavones: Synthesis and NMR Spectroscopic Investigations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Reactivity and Synthetic Applications of 4,5-Dicyanopyridazine: An Overview †

Laboratorio di Progettazione, Sintesi e Studio di Eterocicli Biologicamente Attivi (HeteroBioLab), Dipartimento di Chimica, Università di Firenze, Polo Scientifico e Tecnologico, Via della Lastruccia 3/13, I-50019 Sesto Fiorentino, Italy
*
Author to whom correspondence should be addressed.
Dedicated to Prof. Rodolfo Nesi on the occasion of his 70th anniversary.
Molecules 2010, 15(3), 1722-1745; https://doi.org/10.3390/molecules15031722
Submission received: 28 December 2009 / Revised: 12 February 2010 / Accepted: 5 March 2010 / Published: 12 March 2010
(This article belongs to the Special Issue Advances in Heterocyclic Chemistry)

Abstract

:
Despite the poor reputation of electron-deficient pyridazines in intermolecular Hetero Diels-Alder (HDA) reactions, 4,5-dicyanopyridazine (DCP) showed a surprising reactivity as a heterocyclic azadiene in inverse electron-demand HDA processes with different dienophiles. The use of alkenes, alkynes and enamines as 2π electron counterparts afforded dicyanocyclohexa-1,3-dienes and substituted phthalonitriles, respectively, while the use of suitable bis-dienophiles provides a general strategy for the one-pot synthesis of polycyclic carbo- and hetero-cage systems through pericyclic three-step homodomino processes. HDA reactions with heterocyclic dienophiles allowed direct benzoannelation: in particular, pyrrole and indole derivatives were converted to dicyano-indoles and -carbazoles. In addition an unprecedented reactivity of DCP as a very reactive heterocyclic electrophile at the C-4 carbon was also evidenced: by changing the experimental conditions, cyanopyrrolyl- and cyanoindolyl-pyridazines were obtained through reactions of pyrrole and indole systems as carbon nucleophiles in formal SNAr2 processes where a CN group of DCP acts as leaving group. Thus, careful control of the reaction conditions allows exploitation of both pathways for the synthesis of different classes of heterocyclic derivatives.

1. Introduction

Pyridazine and its benzo derivatives have been known since the nineteenth century, but interest in these compounds was quite limited compared to other nitrogen heterocycles, like for instance the pyrimidines, because only a few examples of naturally occurring 1,2-diazines have been reported. Anyway, in the last decades, the biological activity exhibited by many synthetic derivatives has stimulated intensive research on this heterocyclic system [1,2,3].
This review article concerns a very specific domain in particular, being focused on the reactivity and synthetic applications of 4,5-dicyanopyridazine (DCP, 1), a pyridazine derivative first synthesized in 1968 but essentially only exploited in organic synthesis in the last fifteen years. According with the most significant aspects of reactivity of DCP, namely as heterocyclic electron-poor azadiene in inverse electron-demand Hetero Diels-Alder (HDA) reactions and heterocyclic electrophile in formal nucleophilic aromatic substitutions SNAr2, this review is organized into four parts concerning the synthesis of DCP (Section 2), Hetero Diels-Alder reactions (Section 3), nucleophilic aromatic substitutions (Section 4), and synthetic applications (Section 5).

2. Syntheses of 4,5-Dicyanopyridazine (DCP)

The synthesis of 4,5-dicyanopyridazine (1) was first accomplished in 1968 by Di Stefano and Castle [4] by dehydration of pyridazine-4,5-dicarboxamide, obtained from pyridazine-4,5-dicarboxylic acid, through the corresponding dimethylpyridazine-4,5-dicarboxylate (Scheme 1, condition a). In 1985 Heinisch and Lötsch [5] reported a facile preparation of diethyl pyridazine-4,5-dicarboxylate in one step from commercially available pyridazine (Scheme 1, condition b), by radical bis-ethoxycarbonylation according to the Minisci procedure [6,7].

3. Hetero Diels-Alder (HDA) Reactions

Aromatic pyridazine derivatives can participate as 4π-electron components in concerted cycloadditions [1,2,3]. In particular, N-oxides, N-imides, and N-ylides behave as 1,3-dipoles with different dipolarophiles and, starting from the pioneering works of Neunhoeffer [8,9,10] and Jojima [11], suitably substituted pyridazines have been exploited as azadienes in inverse electron-demand HDA reactions with electron-rich or strained dienophiles or in intramolecular processes [12,13,14,15,16,17]. In this context, while the great synthetic value of 1,2,4-triazines and 1,2,4,5-tetrazines as excellent azadienes in cycloaddition processes with a variety of dienophiles has been recognized over the past decades [12,13,14,15,16,17,18,19,20,21,22,23], little attention has been devoted to 1,2-diazines, probably due to some discouraging reports suggesting their inertness towards unactivated alkynes and alkenes [15,16]. On the other hand, against all expectations, 4,5-dicyanopyridazine exhibited a surprising behaviour as a heteroaromatic azadiene in inverse electron-demand HDA reactions with different dienophiles.

3.1. Reactions of DCP with Dienes: A Direct Access to Carbo- and Hetero-Cage Systems

On the basis of a long-term study concerning Diels-Alder (DA) reactions of nitrogen heterocycles, the strongly electron-deficient DCP, structurally resembling the highly dienophilic tetracyanoethylene, was first chosen to test the almost completely disregarded applications of aromatic pyridazines as 2π-electron components in [2+4] cycloadditions. In this context, the reaction of DCP (1) with 2,3-dimethylbuta-1,3-diene (DMB) as 4π-electron counterpart was investigated (Scheme 2) [24]. By heating 1 with an excess of DMB in CHCl3 at 110 °C, in a sealed tube, only minor amounts of compounds 3 and 4 resulting from the expected [2+4] process were isolated, while the main reaction product was the tricyclic cage compound 2, obtained in 64% yield.
The non-straightforward formation of 2 can be rationalized on the basis of a three-step pericyclic homodomino process [25] involving DCP as masked bis-diene [26]. A first inverse electron-demand [4+2] intermolecular HDA reaction of the electron-poor azadiene moiety of 1 on a double bond of DMB, acting as dienophile, affords the primary cycloadduct 5 able to evolve via retro-DA with nitrogen extrusion into the vinyl substituted cyclohexadiene intermediate 6, then converted into the final product by an entropically assisted intramolecular Diels-Alder (IMDA) cycloaddition (Scheme 3).
The above reaction was also performed with 4,5-dicyano-3,6-dimethylpyridazine [27]. Anyway, likely due to less favorable steric and electronic effects, the corresponding tetramethyl tricyclic derivative was isolated only operating at higher temperature and in low yield (26%) [24].
Analogous domino processes of DCP with different bis-dienophiles have been exploited for direct access to polyfunctionalized carbo- and hetero-cage compounds, often isolated in high yields as sole reaction products (Table 1, entries 1–8) [28,29]. In particular, the almost quantitative formation of the tetracyclic cage system obtained by treatment with cycloocta-1,5-diene (COD, Table 1, entry 1) allowed evaluation of the higher reactivity of DCP as an azadiene in inverse electron-demand HDA reactions with respect to other electron-poor 1,2-diazines: in fact, a previous study concerning different pyridazine derivatives reported that only the highly activated tetramethyl pyridazine-3,4,5,6-tetracarboxylate was able to react with COD, in more drastic conditions (150 °C), affording the corresponding cage system in only 19% yield [17].
Although FMO calculations evidenced for DCP a LUMO of lower energy (0.067 eV, Figure 1b) with respect to the corresponding orbital of the tetraester (0.966 eV, Figure 1a), this orbital can’t participate to the HDA process due to unfavourable symmetry showing two nodes in the C-3 and C-6 positions. For this reason, DCP would be able to react with the LUMO+1 orbital, at higher energy (1.636 eV, Figure 1c) with respect to the tetraester LUMO (having a correct symmetry), and likely steric factors can be mainly invoked to justify the higher reactivity observed for the dinitrile.
A different reaction course was observed with cyclohexa-1,4-diene: a hydrogen transfer from the diene to DCP, followed by HCN elimination, is now preferred over the HDA process leading to 4-cyanopyridazine as the main reaction product (40%) [28]. Higher functionalized bis-dienophiles gave the desired cage compounds after longer reaction times and in lower yields, likely due to unfavorable steric factors associated with competitive reactions of the cyclohexadiene intermediates (Table 1, entries 9–11). Moreover, reacting with allyl acrylate, a preferential aromatization of the cyclohexadiene derivative led exclusively to the corresponding phthalonitrile (Table 1, entry 12). On the other hand, a domino reaction of DCP and diallyldimethylsilane was unsuccessful and the cage system was synthesized through a two-step process involving formation of the cyclohexadiene intermediate under milder reaction conditions (70 °C, 6 days, 66%) and subsequent IMDA reaction (110 °C, 9 days, 32%), likely hampered by the encumbering methyl groups (Table 1, entry 13). Similar steric effects caused a significant decreasing of reactivity in the reaction with myrcene, characterized by a total sitoselectivity, and with the enantiopure terpenes (R)-(–)-β-citronellene and (R)-(–)-linalool affording the diastereomeric tricycloundecene systems (Table 1, entries 14–16). With longer bis-dienophiles as octa-1,7-diene, nona-1,8-diene, and diallyl carbonate, the IMDA process was totally inhibited, probably due to unfavorable entropic effects and repulsive non-bonding interactions, and only cyclohexadiene derivatives were isolated (Table 1, entries 17–19).

3.2. Reactions of DCP with Alkenes

4,5-Dicyanopyridazine shows an exceptional reactivity towards different 2π-electron counterparts [30]. By treatment with an excess of the activated ethyl vinyl ether in chloroform at 50 °C: the intermediate 7, even under these mild conditions, escaped from isolation and it was converted into benzonitrile 8, by loss of EtOH, and mainly into the endo/syn and endo/anti adducts 9a,b and 10a,b, isolated as 1:1 and 4:1 regioisomeric mixtures in 65 and 15% yields, respectively (Scheme 4) [31].
The reaction with the strained 1,2,3-triphenylcyclopropene in chloroform at 110 °C afforded the cycloheptatriene 12 in 77% yield through a [4+2] HDA cycloaddition followed by ring-enlargement via electrocyclic ring-opening of the cyclohexadiene moiety (Scheme 5). Nevertheless, DCP prefers to react as nucleophile with the strongly electrophilic diphenylcyclopropenone leading to bicyclic pyridazine derivatives 13 and 14 [31]. HDA reactions with different cycloalkenes allowed to prepare bicyclic dicyanocyclohexadienes 15a-c (Scheme 5). With cyclopentene and cyclohexene, the yields of compounds 15a,b were lowered due to the competing formation of bis-adducts and aromatization products [31].
Treatment of DCP with an excess of styrene at 110 °C in the presence of 10% Pd/C, to favor the aromatization of the cyclohexadiene intermediate 16, allowed to isolate 3,4-dicyanobiphenyl 17 in 38% yield. Compound 16 was the predominant product operating at 70 °C with a stoichiometric amount of reagent but it escaped any efforts of isolation [31]. The dinitrile 1 was also able to react at 110 °C with electron-poor alkenes, as methyl acrylate and methyl methacrylate, leading to the aromatic ester 19 and the diene 18b in 58 and 64% yields, respectively (Scheme 6) [31].

3.3. Reactions of DCP with Alkynes and Enamines

Phthalonitriles, key building blocks for the synthesis of phthalocyanines whose importance is associated to manifold applications in different fields of material science (see Section 5), are generally obtained by multistage elaboration of aromatic precursors. On the other hand, a novel complementary direct approach to variously substituted phthalonitriles was achieved through HDA cycloadditions of DCP with alkynes and enamines (Table 2). [32]. The aromatic products 22, isolated in satisfactory yields, come from the primary adducts 20 and 21 by nitrogen extrusion or sequential elimination of nitrogen and amine, respectively (Scheme 7).
While DCP was able to react at room temperature with the electron-rich ynamine affording the corresponding phthalonitrile derivative in 85% yield, silylacetylenes required more drastic conditions (110 °C) but allowed the isolation of the reaction products in almost quantitative yields (Table 2, entries 1–3). Reactions of scarcely activated or hindered alkynes were performed at 150 °C in xylene, for longer reaction times, leading to substituted phthalonitriles in lower yields (42–72%) (Table 2, entries 4a, 5a, 6, 7a, 8a). With the exception of highly hindered and/or less activated enamines that showed poor dienophilic properties towards 1 (Table 2, entries 5b and 8b), replacement of alkynes with the corresponding enamines allowed to achieve satisfactory yields (Table 2, entries 4b and 7b). Good results were obtained with cyclic enamines, affording condensed phthalonitriles (Table 2, entries 9a, 10a, 11a) [32]. When the reactions of DCP with cyclic enamines and (E)-4-(pent-2-en-3-yl)morpholine were performed in dichloromethane at room temperature for 24 hours, morpholinodicyano-cyclohexadiene derivatives 24 were obtained in high yields through a [1,5] sigmatropic rearrangement of intermediates 23 (Scheme 8). Operating with 4-(1-cyclopenten-1-yl) morpholine, it was possible to characterize the corresponding intermediate 23 via 1H and 13C NMR analyses.
This observation allowed an improved one-pot synthesis of the corresponding phthalonitriles 22: in fact, by operating first at room temperature in dichloromethane for 24 hours and then, once formed compounds 24, by addition of acetic acid and heating at 50 °C for the same time, to favor the amine elimination step, the desired products were isolated in almost quantitative yields (Table 2, entries 7c, 9b, 10b, 11b) [33].

3.4. Reactions of DCP with Heterocyclic Dienophiles

Although indole and carbazole derivatives, intensively studied for their pharmacological properties, are mainly synthesized through cyclization of suitably substituted aromatic carbocycles, alternative routes based on cycloaddition processes have also been reported [34,35,36,37,38]. In particular, in the last fifteen years electron-rich heterocycles with latent enamine functionalities, such as pyrrole and indole derivatives, have been widely exploited as 2π-electron components in inter- and intramolecular inverse electron-demand HDA reactions, mainly with 1,2,4,5-tetrazines and 1,2,4-triazines [16,20,21] for a facile access to complex indole alkaloids. Although in this context, apart from a few exceptions, 1,2-diazines played a minor role, reactions of dinitrile 1 with pyrrole and indole dienophiles were studied. HDA reactions of DCP with N-methylpyrrole in xylene at 150 °C afforded dicyanoindole 27 in only 17% yield, by spontaneous aromatization of intermediate 26, coming from the primary adduct 25. Attempts to improve this result by changing the reaction conditions were unsuccessful and operating in the presence of catalyst, as CuI or 10% Pd/C to favor the final aromatization step, compounds 28 and 29, coming from competitive SNAr2 reactions of the pyrrole derivative on the strongly electrophilic C-4 carbon of DCP (see Section 4), were obtained as main reaction products (Scheme 9) [39].
Better results were achieved with the more dienophilic indole and N-methylindole, leading to the carbazole derivatives 30a,b in 57 and 62% yields, respectively. Anyway, in both cases, minor amounts of substitution products 31a,b were obtained. Moreover, while 3-acetoxy and 3-ethylthioindole were completely inert towards DCP, 2-ethylthioindole easily reacted with 1 but only compound 31c was isolated in 92% yield (Scheme 10) [39].
The synthesis of indole derivatives was improved by using the non-aromatic 2-methylthiopyrroline as a strongly activated dienophile (Scheme 11).
Although at low temperature (-78 °C) the substitution pathway leading to compound 32 as predominant product is strongly preferred, by operating at 0 °C in THF for 30 minutes the bicyclic systems 34 and 35 were isolated in 50 and 23% yields, respectively [40]. The formation of 34 and 35 can be rationalized through two domino processes involving a [4+2] HDA reaction of 1 with loss of nitrogen affording the intermediate 33 and, respectively, a [1,5] sigmatropic rearrangement, as previously observed with enamines [33], or methanethiol elimination. Moreover, compound 34 is able to evolve into 35 and 36 by elimination of MeSH or H2. Then, this alternative methodology, now involving multi-step processes, allows to synthesize 27 in 70% overall yield, as well as the thiomethyl derivative 37, through aromatization of 35 and 36, respectively.
On the other hand, HDA reactions with heterocyclic enol ethers as dienophiles resulted in the formation of bis-adducts, as previously observed with ethyl vinyl ether (see Section 3.2). In fact, 2,3-dihydrofuran afforded the tetracyclic derivatives 39a,b, as a 3:1 regioisomeric mixture, through endo/anti cycloaddition of the excess of dienophile on the cyclohexadiene intermediate 38 (n = 0), together with minor amounts of 40a,b (1.5:1 mixture) arising from endo/syn or exo/anti approaches. The phthalonitrile 41 was the main product from the reaction of DCP and 3,4-dihydro-2H-pyran, obtained by ring-opening of the corresponding intermediate 38 (n = 1)and trapping with the excess of reagent (Scheme 12) [40].

4. Nucleophilic Aromatic Substitutions on DCP

The nucleophilic substitution of the cyano group in cyano heterocycles is quite unusual and only a few examples have been reported concerning 1,2,4-triazinecarbonitriles [41,42], quinazoline-[42] and quinazolinone-carbonitriles [43], pyrazine-2,3-dicarbonitriles [44], and pyridine-[45] and pyridinium-carbonitriles [46]. In the pyridazine series, the formal replacement of the CN group has been observed in the treatment of 4-cyano-3(2H)-pyridazinone or tetrazolo[1,5-b]pyridazine-8-carbonitrile with phenylmagnesium chloride [47], as well as in the reactions of 1-phthalazine- and 4-cinnoline-carbonitrile with Grignard reagents [48] and cinnoline-3,4-dicarbonitrile with amines [49]. Moreover, as reported in Scheme 9 and Scheme 10, the study of HDA reactions of DCP with pyrrole and indole systems allowed to isolate the corresponding substitution products 28, 29, and 31 [39]. On this ground, and with the aim to elucidate the behavior of the dinitrile 1 as heterocyclic electrophile in nucleophilic aromatic substitution, the above reactions were repeated in different experimental conditions. In particular, operating at room temperature (or at 50 °C) in glacial acetic acid as solvent, without any catalysts, indole and N-methylindole led to indolylpyridazines 31a,b in 91 and 99% yields, respectively, while N-methylpyrrole gave the regioisomeric pyrrolylpyridazines 28 and 29 in 68 and 13% yields and the more reactive pyrrole afforded, after only 4 days at room temperature, derivative 42 in 91% yield, together with minor amount of dihydropyridazine 43 (8%) (Scheme 13) [50].
These results clearly show the importance of the acidity in the substitution pathway, likely involving the LUMO of protonated DCP at lower energy (-5.883 eV) with respect to the corresponding orbital of 1 (0.067 eV). Moreover, the acidic medium, through protonation of the pyridazine nitrogen, seems to play a crucial role in the HDA/SNAr2 competition, in favor of the latter process.
The isolation of the 1,4-adduct 43, from reaction with pyrrole, led to a mechanistic investigation of the substitution processes. Then, performing the above reactions in the same conditions but for shorter reaction times, dihydropyridazine derivatives were isolated in 52–88% yields. These intermediates were easily converted into the final products by simple stirring in AcOH at room temperature, through spontaneous HCN elimination. These data clearly show that the nucleophilic substitution products are actually the fruit of a two-step process, involving nucleophilic addition and elimination reactions (Scheme 14). To assess applications and limits of this new reactivity of DCP, reactions with different pyrrole and indole nucleophiles in acetic acid were studied.
While DCP was absolutely inert towards electron-deficient-pyrrole and indole derivatives, 2-methylindole gave compound 44 in 94% yield by heating at 50 °C for 24 hours and the sterically hindered 2-phenyl derivative, although required longer reaction times and higher temperature, led to 45 in 88% yield. Reaction of 1 with the bulky N-phenylpyrrole was achieved in even more drastic conditions affording the pyrrolylpyridazine 46 in only 23% yield while the electron-rich N-(dimethylamino)pyrrole allowed to isolate 47 in 78% yield (Scheme 15) [50]. DCP was also able to react with 1-(2-pyridyl)-2-propen-1-ol, acting as C-3 carbon nucleophile, to give the polyfunctionalyzed derivative 48 by replacement of a CN group (Scheme 16) [51].
Moreover, on the basis of the reactivity observed with carbon nucleophiles, the behavior of DCP towards other nucleophilic species was also studied. In particular, some preliminary results concerning reactions with amines evidenced a facile access to aminocyanopyridazines 49, promising for further synthetic elaborations (Scheme 16) [unpublished results].

5. Synthetic applications

5.1. Synthesis of Phthalocyanine Analogues

As previously reported, phthalocyanines have been the subject of intensive study due to their multipurpose applications as, for instance, dyes and pigments, photoconducting reagents, catalysts, deodorants, germicides, in the photodynamic therapy of cancer, and so on [52,53,54].
In this context, dicyanocycloheptatrienes 12 and 50, obtained by HDA reactions of DCP with 1,2,3-triphenyl- and 1,2-diphenyl-cyclopropene, have been exploited for the synthesis of seven-membered carbon-ring-fused phthalocyanine analogue complexes 51a,b, in which the π system changes during dehydrogenation/hydrogenation cycles, according with nonaromatic cycloheptatriene/aromatic tropylium cation interconversion (Scheme 17) [55].
Moreover, crossover Linstead macrocyclization of 12 and bis(dimethylamino)maleonitrile gave rise to an unsymmetrical (A3B) aminoporphyrazine 52 converted by hydride abstraction into a tropylium-fused aminoporphyrazine, labelled as a ‘push-pull’ macrocycle for the presence of both electron-donating and electron-withdrawing groups (Scheme 17) [56].
On the other hand, diazotization of 3,4,5,6-tetraphenylanthranilic acid in the presence of DCP allowed HDA trapping of the aryne intermediate for the synthesis of dicyanonaphthalene 53, converted to the polyphenylated naphthalocyanine 54. The sixteen phenyl substituents, orthogonal to the macrocyclic plane, are responsible for improved solubility and minimal intermolecular aggregation (Scheme 17) [57].

5.2. Synthesis of Enantiopure C2 Symmetric bis(2-Oxazolinyl) Cage Ligands

In the last decades, a great range of ligands containing one, two or three oxazoline rings has been successfully developed and applied in asymmetric catalysis [58]. In particular, the attention towards enantiopure C2 symmetric bis(oxazoline) (Box) ligands has increased enormously and a large number of chiral ligands with a great deal of structural diversity has been synthesized and applied in an impressive variety of metal-catalyzed processes [59]. The synthesis of oxazoline systems can be easily achieved from amino alcohols and nitrile or carboxylic acid derivatives [60] and, in particular, Bolm et al. described a direct reaction of dinitriles with enantiomerically pure β-amino alcohols under zinc chloride catalysis [61]. On these grounds, the cage systems obtained by domino reactions of DCP with dienes, all containing a maleonitrile moiety coming from 1, appeared interesting substrates for the synthesis of new classes of chiral bis(oxazoline) ligands.
Then, the tetracyclic compound 55, prepared almost quantitatively from reaction of 1 with COD (Table 1, entry 1), was allowed to react with enantiopure β-amino alcohols 56a-d in anhydrous chlorobenzene under reflux, in the presence of catalytic amounts of anhydrous ZnCl2. Depending on the experimental conditions, different quantities of enantiomerically pure mono(2-oxazolines) 57a-d and bis(2-oxazoline) (Cage-Box) ligands 58a-d were isolated from the reaction mixtures. Anyway, mono(oxazolines) can be converted into the bis(oxazoline) derivatives by treatment with β-amino alcohols in the previously reported conditions (Scheme 18) [62]. A 1:1 Zn(II) complex 59 was isolated in 99% yield by treatment of 58a with ZnCl2 and its molecular structure determined by single-crystal X-ray diffraction analysis (Figure 2).

6. Conclusions

These results clearly evidence for 4,5-dicyanopyridazine a polyvalent reactivity towards different species. In fact, although electron-deficient pyridazines, and especially symmetric derivatives [63], appeared quite inert in inverse electron-demand HDA reactions, DCP exhibited an exceptional behavior as heterocyclic azadiene able to react not only with electron-rich dienophiles but also with unactivated alkenes and alkynes, as well as electron-poor 2π electron counterparts. In addition to the synthesis of dicyanocyclohexa-1,3-dienes, a general and complementary methodology for a facile access to substituted phthalonitriles through reactions with alkynes and enamines was accomplished. Moreover, pericyclic three-step homodomino processes with bis-dienophiles allowed to realize a general strategy for the one-pot synthesis of carbo- and hetero-cage systems.
Pyrrole and indole derivatives reacted in different ways with DCP. In particular, it appears possible to modulate the reaction course by changing the experimental conditions. By simple heating at 110–150 °C in chloroform or xylene (sealed tube), HDA reactions afforded dicyano-indoles and -carbazoles via direct benzoannelations. On the other hand, operating in acetic acid in milder conditions (25–110 °C), DCP was able to react with the same species as a very reactive heterocyclic electrophile at C-4 carbon in formal SNAr2 processes where a CN group acts as leaving group. Mechanistic investigations evidenced that cyanopyrrolyl- and cyanoindolyl-pyridazines, obtained in satisfactory yields, are the fruit of addition-elimination processes. Synthetic applications of the above reaction products were also observed in different domains.
In conclusion, the exceptional reactivity of DCP compared to other electron-poor pyridazines, mainly evidenced in HDA reactions, could be likely ascribed to the lack of any steric hindrance experienced by the approaching reagents. On this ground, the ‘flat’ 4,5-dicyanopyridazine could certainly be welcome in the fantastic ‘Flatland’ country that in 1882 the London Reverend Edwin A. Abbott described in his book.

Acknowledgements

We would like to thank Rodolfo Nesi for his decisive contribution to the realization of this research and all the students that with their efforts, enthusiasm, and dedication allowed to perform and develop the synthetic work. Moreover, we thank B. Innocenti and M. Passaponti for technical assistance and Ministry of University and Research (MIUR Rome-Italy) for financial support.

References

  1. Tišler, M.; Stanovnik, B. Pyridazines and their Benzo Derivatives. In Comprehensive Heterocyclic Chemistry; Boulton, A.J., McKillop, A., Eds.; Pergamon: Oxford, UK, 1984; Volume 3, pp. 1–56. [Google Scholar]
  2. Coates, W.J. Pyridazines and their Benzo Derivatives. In Comprehensive Heterocyclic Chemistry II; Boulton, A.J., Ed.; Pergamon: Oxford, UK, 1996; Volume 6, pp. 1–91. [Google Scholar]
  3. Maes, B.U.W.; Lemière, G.L.F. Pyridazines and their Benzo Derivatives. In Comprehensive Heterocyclic Chemistry III; Aitken, A., Ed.; Elsevier: Oxford, UK, 2008; Volume 8, pp. 1–116. [Google Scholar]
  4. Di Stefano, L.; Castle, R.N. The synthesis of pyridazino[4,5-d]pyridazines, pyrazino [2,3-d] pyridazines and a pyrimido [4,5-d] pyridazine. J. Heterocycl. Chem. 1968, 5, 53–59. [Google Scholar] [CrossRef]
  5. Heinisch, G.; Lötsch, G. Pyridazines XXIV. Application of Radicalic ethoxycarbonylation to the synthesis of pyridazine mono- and polycarboxylic acid esters. Tetrahedron 1985, 41, 1199–1205. [Google Scholar] [CrossRef]
  6. Minisci, F. Novel applications of free-radical reactions in preparative organic chemistry. Synthesis 1973, 1–24. [Google Scholar] [CrossRef]
  7. Minisci, F.; Fontana, F.; Vismara, E. Substitutions by nucleophilic free radicals: A new general reaction of heteroaromatic bases. J. Heterocycl. Chem. 1990, 27, 79–96. [Google Scholar] [CrossRef]
  8. Neunhoeffer, H.; Werner, G. Reaktion von pyridazincarbonestern mit 1-diäthylaminopropin. Tetrahedron Lett. 1972, 13, 1517–1518. [Google Scholar] [CrossRef]
  9. Neunhoeffer, H.; Werner, G. Reaktion von pyridazinen mit 1-diäthylaminopropin. Liebigs Ann. Chem. 1973, 437–442. [Google Scholar] [CrossRef]
  10. Neunhoeffer, H.; Werner, G. Reaktion von pyridazinen mit 1-methoxy-N,N-dimethylvinylamin. Liebigs Ann. Chem. 1973, 1955–1962. [Google Scholar]
  11. Jojima, T.; Takeshiba, H.; Konotsune, T. Pyridazines I. Novel intramolecular cycloaddition of 3-chloro-6-(2-allylphenoxy)pyridazines. Chem. Pharm. Bull. 1973, 20, 2191–2203. [Google Scholar] [CrossRef]
  12. Boger, D.L. Diels-Alder reactions of azadienes. Tetrahedron 1983, 39, 2869–2939. [Google Scholar] [CrossRef]
  13. Boger, D.L. Diels-Alder reactions of heterocyclic azadienes: Scope and applications. Chem. Rev. 1986, 86, 781–793. [Google Scholar] [CrossRef]
  14. Boger, D.L.; Weinreb, S.N. Heteroaromatic Azadienes. In Hetero Diels-Alder methodology in organic synthesis; Wasserman, H.H., Ed.; Academic: San Diego, CA, USA, 1987; pp. 300–357. [Google Scholar]
  15. Boger, D.L.; Patel, M. Recent applications of the inverse electron demand Diels-Alder reaction. Progr. Heterocycl. Chem. 1989, 1, 30–64. [Google Scholar]
  16. Benson, S.C.; Gross, J.L.; Snyder, J.K. Indole as a Dienophile in Inverse Electron Demand Diels-Alder Reactions: Reactions with 1,2,4-Triazines and 1,2-Diazines. J. Org. Chem. 1990, 55, 3257–3269. [Google Scholar] [CrossRef]
  17. Lantos, I.; Sheldrake, P.W.; Wells, A.S. Novel cage compounds from inter- and intra-molecular Diels-Alder reactions of heteroaromatic azadienes and methyl coumalate with cyclo-octa-1,5-diene. J. Chem. Soc. Perkin Trans. 1 1990, 1887–1890. [Google Scholar] [CrossRef]
  18. Neunhoeffer, H. 1,2,4-Triazines and their Benzo Derivatives. In Comprehensive Heterocyclic Chemistry; Boulton, A.J., McKillop, A., Eds.; Pergamon: Oxford, UK, 1984; Volume 3, pp. 422–429. [Google Scholar]
  19. Neunhoeffer, H. Tetrazines and Pentazines. In Comprehensive Heterocyclic Chemistry; Boulton, A.J., McKillop, A., Eds.; Pergamon: Oxford, UK, 1984; Volume 3, pp. 550–555. [Google Scholar]
  20. Neunhoeffer, H. 1,2,4-Triazines and their Benzo Derivatives. In Comprehensive Heterocyclic Chemistry II; Boulton, A.J., Ed.; Pergamon: Oxford, UK, 1996; Volume 6, pp. 533–538. [Google Scholar]
  21. Sauer, J. 1,2,4,5-Tetrazine. In Comprehensive Heterocyclic Chemistry II; Boulton, A.J., Ed.; Pergamon: Oxford, UK, 1996; Volume 6, pp. 912–944. [Google Scholar]
  22. Van der Plas, H.C. Cycloalkenopyridines by ring transformations of diazines and triazines. ARKIVOC 2008, (iii), 127–139. [Google Scholar]
  23. Saracoglu, N. Recent advances and applications in 1,2,4,5-tetrazine chemistry. Tetrahedron 2007, 63, 4199–4236. [Google Scholar] [CrossRef]
  24. Nesi, R.; Giomi, D.; Turchi, S. [4+2]- and [2+4]-Cycloaddition processes of 4,5-dicyanopyridazine with 2,3-dimethylbuta-1,3-diene. Tetrahedron 1994, 50, 9189–9194. [Google Scholar] [CrossRef]
  25. Tietze, L.F. Domino reactions in organic synthesis. Chem. Rev. 1996, 96, 115–136. [Google Scholar] [CrossRef] [PubMed]
  26. Winkler, J.D. Tandem Diels-Alder cycloadditions in organic synthesis. Chem. Rev. 1996, 96, 167–176. [Google Scholar] [CrossRef] [PubMed]
  27. Adembri, G.; De Sio, F.; Nesi, R.; Scotton, M. Synthesis and properties of 1,4-dimethylpyridazino[4,5-d] pyridazines. J. Heterocycl. Chem. 1975, 12, 95–99. [Google Scholar] [CrossRef]
  28. Giomi, D.; Nesi, R.; Turchi, S.; Coppini, R. Tandem inverse electron-demand Diels-Alder reactions of 4,5-dicyanopyridazine with nonconjugated dienes: An excellent direct entry into carbo- and hetero-cage systems. J. Org. Chem. 1996, 61, 6028–6030. [Google Scholar] [CrossRef]
  29. Giomi, D.; Nesi, R.; Turchi, S.; Mura, E. Polyfunctionalized cage compounds by pericyclic domino processes of 4,5-dicyanopyridazine with dienes: Applications and limits. J. Org. Chem. 2000, 65, 360–364. [Google Scholar] [CrossRef] [PubMed]
  30. Nesi, R.; Giomi, D.; Turchi, S.; Falai, A. 4,5-Dicyanopyridazine: A 1,2-diazine superheterodiene for [4+2] cycloaddition reactions. J. Chem. Soc. Chem. Commun. 1995, 2201–2202. [Google Scholar] [CrossRef]
  31. Turchi, S.; Giomi, D.; Capaccioli, C.; Nesi, R. Hetero Diels-Alder reactions of 4,5-dicyanopyridazine with alkenes. Tetrahedron 1997, 53, 11711–11720. [Google Scholar] [CrossRef]
  32. Turchi, S.; Nesi, R.; Giomi, D. Reactions of 4,5-dicyanopyridazine with alkynes and enamines: A new straightforward complementary route to 4-mono- and 4,5-disubstituted phthalonitriles. Tetrahedron 1998, 54, 1809–1816. [Google Scholar] [CrossRef]
  33. Nesi, R.; Turchi, S.; Giomi, D.; Corsi, C. Phthalonitriles by hetero Diels-Alder reactions of 4,5-dicyanopyridazine with enamines: Isolation and characterization of unprecedented intermediates. Tetrahedron 1998, 54, 10851–10856. [Google Scholar] [CrossRef]
  34. Sundberg, R.J. Pyrroles and their Benzo Derivatives: (iii) Synthesis and Applications. In Comprehensive Heterocyclic Chemistry; Bird, C.W., Cheeseman, G.W.H., Eds.; Pergamon: Oxford, UK, 1984; Volume 4, pp. 313–376. [Google Scholar]
  35. Sundberg, R.J. Pyrroles and their Benzo Derivatives: Synthesis. In Comprehensive Heterocyclic Chemistry II; Bird, C.W., Ed.; Pergamon: Oxford, UK, 1996; Volume 2, pp. 119–206. [Google Scholar]
  36. Gribble, G.W. Recent developments in indole ring synthesis–methodology and applications. J. Chem. Soc. Perkin Trans. 1 2000, 1045–1075. [Google Scholar] [CrossRef]
  37. Jones, R.A. Pyrroles and their Benzo Derivatives: Reactivity. In Comprehensive Heterocyclic Chemistry; Bird, C.W., Cheeseman, G.W.H., Eds.; Pergamon: Oxford, UK, 1984; Volume 4, pp. 261–268. [Google Scholar]
  38. Black, D.St.C. Pyrroles and their Benzo Derivatives: Reactivity. In Comprehensive Heterocyclic Chemistry II; Bird, C.W., Ed.; Pergamon: Oxford, UK, 1996; Volume 2, pp. 81–88. [Google Scholar]
  39. Giomi, D.; Cecchi, M. Study on direct benzoannelation of pyrrole and indole systems by domino reactions with 4,5-dicyanopyridazine. Tetrahedron 2002, 58, 8067–8071. [Google Scholar] [CrossRef]
  40. Giomi, D.; Cecchi, M. Domino reactions of 4,5-dicyanopyridazine with dihydroheterocycles: Synthetic and mechanistic features. J. Org. Chem. 2003, 68, 3340–3343. [Google Scholar] [CrossRef] [PubMed]
  41. Huang, J.J. Synthesis of fused 1,2,4-triazines: 6- and 7-azapteridine and 6-azapurine ring systems. J. Org. Chem. 1985, 50, 2293–2298. [Google Scholar] [CrossRef]
  42. Ohba, S.; Konno, S.; Yamanaka, H. Studies on as-triazine derivatives. XVI. Reaction of 1,2,4-triazinecarbonitriles with carbanions. Chem. Pharm. Bull. 1991, 39, 486–488. [Google Scholar] [CrossRef]
  43. Ozaki, K.; Yamada, Y.; Oine, T. Studies on 4(1H)-quinazolinones. III. Some derivatizations of 2-ethoxycarbonylalkyl-1-substituted-4(1H)-quinazolinones. Chem. Pharm. Bull. 1983, 31, 2234–2243. [Google Scholar] [CrossRef]
  44. Hirano, H.; Lee, R.; Tada, M. The substitution reaction of pyrazine-2,3-dicarbonitrile derivatives with ammonia, amines, water and alcohols. J. Heterocycl. Chem. 1982, 19, 1409–1413. [Google Scholar] [CrossRef]
  45. Penney, J.M. Synthesis of aminopyridines via an unprecedented nucleophilic aromatic substitution of cyanopyridines. Tetrahedron Lett. 2004, 45, 2667–2669. [Google Scholar] [CrossRef]
  46. Landquist, J.K. Nucleophilic replacements in 4-cyanopyridinium salts. J. Chem. Soc. Perkin Trans. 1 1976, 454–456. [Google Scholar] [CrossRef]
  47. Haider, N.; Heinisch, G.; Moshuber, J. Pyridazines – 61. Unexpected reaction behavior of pyridazinecarbonitrile derivatives towards phenylmagnesium chloride. Tetrahedron 1991, 47, 8573–8578. [Google Scholar] [CrossRef]
  48. Hayashi, E.; Iinuma, M.; Utsunomiya, I.; Iijima, C.; Oishi, E.; Higashino, T. Grignard reactions, of 1-phthalazine-, 2-quinoxaline-, and 4-cinnoline-carbonitrile. Chem. Pharm. Bull. 1977, 25, 579–589. [Google Scholar] [CrossRef]
  49. Ames, D.E.; Bull, D. Preparation of cinnoline-3,4-dicarbonitrile and –dicarboxylic acid. Tetrahedron 1981, 37, 2489–2491. [Google Scholar] [CrossRef]
  50. Cecchi, M.; Micoli, A.; Giomi, D. Nucleophilic aromatic substitutions on 4,5-dicyanopyridazine. Pyrrole and indole systems as carbon nucleophiles. Tetrahedron 2006, 62, 12281–12287. [Google Scholar] [CrossRef]
  51. Giomi, D.; Piacenti, M.; Brandi, A. New reactivity of hydroxyallylpyridyl derivatives as C-3 carbon nucleophiles. Tetrahedron Lett. 2004, 45, 2113–2115. [Google Scholar] [CrossRef]
  52. Leznoff, C.C.; Lever, A.B.P. Phthalocyanines: Properties and Applications; VCH Publishers: New York, NY, USA, 1989; Volumes 1–4. [Google Scholar]
  53. Shirai, H.; Kobayashi, N. Phthalocyanines – Chemistry and Functions; IPC: Tokyo, Japan, 1997. [Google Scholar]
  54. Tanaka, M. Phthalocyanines—High performance pigments and their applications. In High performance pigments, 2nd ed.; Faulkner, E.B., Schwartz, R.J., Eds.; Wiley-VCH: Weinheim, Germany, 2009; pp. 275–291. [Google Scholar]
  55. Kobayashi, N.; Nonomura, T.; Nakai, K. A seven-membered carbon-ring-fused phthalocyanine analogue in which the π system changes during dehydrogenation/hydrogenation cycles. Angew. Chem. Int. Ed. 2001, 40, 1300–1303. [Google Scholar] [CrossRef]
  56. Ruggiero, A.; Fuchter, M.J.; Kokas, O.J.; Negru, M.; White, A.J.P.; Haycock, P.R.; Hoffman, B.M.; Barrett, A.G.M. A ‘push-pull’ tropylium-fused aminoporphyrazine. Tetrahedron 2009, 65, 9690–9693. [Google Scholar] [CrossRef]
  57. Lee, C.H.; Sooksimuang, T.; Mandal, B.K. Synthetic methodology towards novel polyphenylated naphthalocyanines. J. Porphyr. Phthalocyanines 2006, 10, 135–139. [Google Scholar] [CrossRef]
  58. McManus, H.A.; Guiry, P.J. Recent developments in the application of oxazoline-containing ligands in asymmetric catalysis. Chem. Rev. 2004, 104, 4151–4202. [Google Scholar] [CrossRef] [PubMed]
  59. Desimoni, G.; Faita, G.; Jørgensen, K.A. C2-Symmetric chiral bis(oxazoline) ligands in asymmetric catalysis. Chem. Rev. 2006, 106, 3561–3651. [Google Scholar] [CrossRef] [PubMed]
  60. Gant, T.G.; Meyers, A.I. The chemistry of 2-oxazolines (1985-present). Tetrahedron 1994, 50, 2297–2360. [Google Scholar] [CrossRef]
  61. Bolm, C.; Weickhardt, K.; Zehnder, M.; Ranff, T. Synthesis of optically active bis(2-oxazolines): Crystal structure of a 1,2-bis(2-oxazolinyl)benzene zinc chloride complex. Chem. Ber. 1991, 124, 1173–1180. [Google Scholar] [CrossRef]
  62. Cecchi, M.; Faggi, C.; Giomi, D. New enantiopure C2 symmetric bis(2-oxazolinyl)cage (Cage-Box) ligands from 4,5-dicyanopyridazine. Tetrahedron Asymmetry 2005, 16, 3998–4003. [Google Scholar] [CrossRef]
  63. Ho, T.L. Symmetry: A Basis for Synthesis Design; Wiley: New York, NY, USA, 1995; p. 474. [Google Scholar]
Scheme 1. Syntheses of 4,5-Dicyanopyridazine (1).
Scheme 1. Syntheses of 4,5-Dicyanopyridazine (1).
Molecules 15 01722 sch001
Scheme 2. Reaction of DCP with 2,3-dimethylbuta-1,3-diene.
Scheme 2. Reaction of DCP with 2,3-dimethylbuta-1,3-diene.
Molecules 15 01722 sch002
Scheme 3. Mechanistic rationale for the formation of compound 2.
Scheme 3. Mechanistic rationale for the formation of compound 2.
Molecules 15 01722 sch003
Figure 1. Molecular orbitals of tetramethyl pyridazine-3,4,5,6-tetracarboxylate and 4,5-dicyanopyridazine.
Figure 1. Molecular orbitals of tetramethyl pyridazine-3,4,5,6-tetracarboxylate and 4,5-dicyanopyridazine.
Molecules 15 01722 g001
Scheme 4. HDA reaction of 1 with ethyl vinyl ether.
Scheme 4. HDA reaction of 1 with ethyl vinyl ether.
Molecules 15 01722 sch004
Scheme 5. Reactions of 1 with cycloalkenes.
Scheme 5. Reactions of 1 with cycloalkenes.
Molecules 15 01722 sch005
Scheme 6. HDA reactions of 1 with linear alkenes.
Scheme 6. HDA reactions of 1 with linear alkenes.
Molecules 15 01722 sch006
Scheme 7. HDA reactions of 1 with alkynes and enamines.
Scheme 7. HDA reactions of 1 with alkynes and enamines.
Molecules 15 01722 sch007
Scheme 8. Mechanistic study of reactions of 1 with enamines.
Scheme 8. Mechanistic study of reactions of 1 with enamines.
Molecules 15 01722 sch008
Scheme 9. HDA reactions of DCP with N-methylpyrrole.
Scheme 9. HDA reactions of DCP with N-methylpyrrole.
Molecules 15 01722 sch009
Scheme 10. HDA reactions of DCP with indole derivatives.
Scheme 10. HDA reactions of DCP with indole derivatives.
Molecules 15 01722 sch010
Scheme 11. HDA reactions of DCP with 2-methylthiopyrroline.
Scheme 11. HDA reactions of DCP with 2-methylthiopyrroline.
Molecules 15 01722 sch011
Scheme 12. HDA reactions of DCP with dihydroheterocycles.
Scheme 12. HDA reactions of DCP with dihydroheterocycles.
Molecules 15 01722 sch012
Scheme 13. SNAr2 reactions of DCP with pyrrole and indole nucleophiles.
Scheme 13. SNAr2 reactions of DCP with pyrrole and indole nucleophiles.
Molecules 15 01722 sch013
Scheme 14. Two-step mechanism involving nucleophilic addition and elimination reactions.
Scheme 14. Two-step mechanism involving nucleophilic addition and elimination reactions.
Molecules 15 01722 sch014
Scheme 15. SNAr2 reactions of DCP with substituted pyrrole and indole nucleophiles.
Scheme 15. SNAr2 reactions of DCP with substituted pyrrole and indole nucleophiles.
Molecules 15 01722 sch015
Scheme 16. SNAr2 reactions of DCP with 1-(2-pyridyl)-2-propen-1-ol and amines.
Scheme 16. SNAr2 reactions of DCP with 1-(2-pyridyl)-2-propen-1-ol and amines.
Molecules 15 01722 sch016
Scheme 17. Synthesis of phthalocyanine analogues.
Scheme 17. Synthesis of phthalocyanine analogues.
Molecules 15 01722 sch017
Scheme 18. Synthesis of mono- and bis(oxazoline) ligands.
Scheme 18. Synthesis of mono- and bis(oxazoline) ligands.
Molecules 15 01722 sch018
Figure 2. ORTEP drawing of complex 59.
Figure 2. ORTEP drawing of complex 59.
Molecules 15 01722 g002
Table 1. Reactions of DCP with dienes. a
Table 1. Reactions of DCP with dienes. a
EntryReagentTimeProductYield % b [Ref.]
1Molecules 15 01722 i00124 hMolecules 15 01722 i00298 [28]
2Molecules 15 01722 i00348 hMolecules 15 01722 i00471 [28]
3Molecules 15 01722 i00548 hMolecules 15 01722 i00698 [28]
4Molecules 15 01722 i00748 hMolecules 15 01722 i00897 [28]
5Molecules 15 01722 i0094 dMolecules 15 01722 i01086 [28]
6Molecules 15 01722 i01124 hMolecules 15 01722 i01298 [28]
7Molecules 15 01722 i01372 hMolecules 15 01722 i01497 [28]
8Molecules 15 01722 i01524 hMolecules 15 01722 i01699 [29]
9Molecules 15 01722 i0174 dMolecules 15 01722 i01856 [29] c
10Molecules 15 01722 i0199 dMolecules 15 01722 i02034 [29]
11Molecules 15 01722 i0216 dMolecules 15 01722 i02237 [29]
12Molecules 15 01722 i02348 hMolecules 15 01722 i02440 [29]
13dMolecules 15 01722 i02515 dMolecules 15 01722 i02621 [29]d
14Molecules 15 01722 i0274 dMolecules 15 01722 i02824 [29]e
15Molecules 15 01722 i0295 dMolecules 15 01722 i03028 [29]e
16Molecules 15 01722 i0318 dMolecules 15 01722 i03231 [29]e
17Molecules 15 01722 i0334 dMolecules 15 01722 i03452 [29]f
18Molecules 15 01722 i0355 dMolecules 15 01722 i03650 [29]f
19Molecules 15 01722 i0375 dMolecules 15 01722 i03852 [29]f
a Reactions performed in a sealed tube in chloroform at 110 °C with an excess (5 equiv) of reagent. b Isolated yields. c Phthalonitrile, formed by competitive elimination of methacrylic acid from the corresponding cyclohexadiene intermediate, was also recovered in 39% yield. d Reaction performed at 70 °C for 6 days to isolate the cyclohexadiene intermediate then converted into the cage system by heating at 110 °C for 9 days. e Competitive aromatization of the corresponding cyclohexadiene intermediates afforded minor amounts of substituted phthalonitriles. f Reactions performed with 1.1 equiv. of reagent to reduce the formation of by-products.
Table 2. Synthesis of phthalonitriles from reactions of DCP with alkynes and enamines. a
Table 2. Synthesis of phthalonitriles from reactions of DCP with alkynes and enamines. a
EntryReagentConditionsProductYield %b [Ref.]
1Molecules 15 01722 i039CH2Cl2, 25 °C, 19 hMolecules 15 01722 i04085 [32]
2Molecules 15 01722 i041CHCl3, 110 °C, 48 hMolecules 15 01722 i04298 [32]
3Molecules 15 01722 i043CHCl3, 110 °C, 8 dMolecules 15 01722 i04498 [32]
4aMolecules 15 01722 i045xylene, 150 °C, 24 hMolecules 15 01722 i04642 [32]
4bMolecules 15 01722 i047CHCl3, 70 °C, 24 hMolecules 15 01722 i04872 [32]
5aMolecules 15 01722 i049xylene, 150 °C, 5 dMolecules 15 01722 i05055 [32]
5bMolecules 15 01722 i051CHCl3, 70 °C, 24 hMolecules 15 01722 i05229 [32]
6Molecules 15 01722 i053xylene, 150 °C, 10 dMolecules 15 01722 i05455 [32]
7aMolecules 15 01722 i055xylene, 150 °C, 18 dMolecules 15 01722 i05651 [32]
7b
7c
Molecules 15 01722 i057CHCl3, 70 °C, 7 d
CH2Cl2, rt, 24 h +
AcOH, 50 °C, 24 h
Molecules 15 01722 i05873 [32]
95 [33]
8aMolecules 15 01722 i059xylene, 150 °C, 21 dMolecules 15 01722 i060 72 [32]
8bMolecules 15 01722 i061CHCl3, 70 °C, 48 hMolecules 15 01722 i06211 [32]
9a
9b
Molecules 15 01722 i063CHCl3, 70 °C, 24 h
CH2Cl2, rt, 24 h +
AcOH, 50 °C, 24 h
Molecules 15 01722 i06496 [32]
98 [33]
10a
10b
Molecules 15 01722 i065CHCl3, 70 °C, 6 d
CH2Cl2, rt, 24 h +
AcOH, 50 °C, 24 h
Molecules 15 01722 i06682 [32]
99 [33]
11a
11b
Molecules 15 01722 i067CHCl3, 70 °C, 7 d
CH2Cl2, rt, 24 h +
AcOH, 50 °C, 24 h
Molecules 15 01722 i06871 [32]
96 [33]
a Reactions performed in a sealed tube with an excess of reagent (1.1–10 equiv). b Isolated yields.

Share and Cite

MDPI and ACS Style

Alfini, R.; Cecchi, M.; Giomi, D. Reactivity and Synthetic Applications of 4,5-Dicyanopyridazine: An Overview. Molecules 2010, 15, 1722-1745. https://doi.org/10.3390/molecules15031722

AMA Style

Alfini R, Cecchi M, Giomi D. Reactivity and Synthetic Applications of 4,5-Dicyanopyridazine: An Overview. Molecules. 2010; 15(3):1722-1745. https://doi.org/10.3390/molecules15031722

Chicago/Turabian Style

Alfini, Renzo, Marco Cecchi, and Donatella Giomi. 2010. "Reactivity and Synthetic Applications of 4,5-Dicyanopyridazine: An Overview" Molecules 15, no. 3: 1722-1745. https://doi.org/10.3390/molecules15031722

Article Metrics

Back to TopTop