Next Article in Journal
Supercritical Fluid Applications in the Design of Novel Antimicrobial Materials
Next Article in Special Issue
Synthesis of Medium-Sized Heterocycles by Transition-Metal-Catalyzed Intramolecular Cyclization
Previous Article in Journal
Adsorption of Sulfamethazine Drug onto the Modified Derivatives of Carbon Nanotubes at Different pH
Previous Article in Special Issue
Facile Synthesis of NH-Free 5-(Hetero)Aryl-Pyrrole-2-Carboxylates by Catalytic C–H Borylation and Suzuki Coupling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Strides in the Transition Metal-Free Cross-Coupling of Haloacetylenes with Electron-Rich Heterocycles in Solid Media

by
Lyubov’ N. Sobenina
and
Boris A. Trofimov
*
A.E. Favorsky Irkutsk Institute of Chemistry, Siberian Branch, Russian Academy of Sciences, 1 Favorsky Str., 664033 Irkutsk, Russia
*
Author to whom correspondence should be addressed.
Submission received: 25 April 2020 / Revised: 20 May 2020 / Accepted: 25 May 2020 / Published: 27 May 2020
(This article belongs to the Special Issue Advances in Cross-Coupling Reactions)

Abstract

:
The publications covering new, transition metal-free cross-coupling reactions of pyrroles with electrophilic haloacetylenes in solid medium of metal oxides and salts to regioselectively afford 2-ethynylpyrroles are discussed. The reactions proceed at room temperature without catalyst and base under solvent-free conditions. These ethynylation reactions seem to be particularly important, since the common Sonogashira coupling does not allow ethynylpyrroles with strong electron-withdrawing substituents at the acetylenic fragments to be synthesized. The results on the behavior of furans, thiophenes, and pyrazoles under the conditions of these reactions are also provided. The reactivity and structural peculiarities of nucleophilic addition to the activated acetylene moiety of the novel C-ethynylpyrroles are considered.

1. Introduction

Functionalized five-membered aromatic heterocycles represent a frequent structural motif of bioactive natural products and pharmaceuticals [1,2,3,4,5,6,7,8,9,10,11,12]. Among them, of particular interest are the compounds bearing acetylenic moieties [13,14,15]. The combination of the electron-rich, five-membered aromatic heterocyclic nucleus with highly reactive carbon-carbon triple bond in one molecule allows using these compounds for the targeted synthesis of various complex heterocyclic systems. Commonly, ethynylation of heterocycles is implemented via Sonogashira reaction employing the halogenated heterocycles and terminal alkynes [16,17,18,19]. In 2004, as a complementation to the existing cross-coupling protocols, the direct palladium- and copper-free ethynylation of the pyrrole ring with haloacetylenes in the Al2O3 medium (room temperature, no solvent) was discovered [20]. Later, haloacetylenes were involved in ethynylation of diverse heterocycles using palladium [21,22], nickel [23,24], copper [25], or gold [26] catalysts. In parallel, the Al2O3-mediated ethynylation of pyrroles and indoles kept being steadily developed [27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43]. A number of pyrroles with alkyl, cycloalkyl, aryl, and hetaryl substituents were successfully ethynylated with acylhaloacetylenes and halopropynoates within a framework of this cross-coupling procedure. It was shown that some other metal oxides (MgO, CaO, BaO) [33] and salts (K2CO3) [35] can also be beneficially applied instead of Al2O3. On the basis of the experimental facts, it was concluded that this cross-coupling includes the nucleophilic attack of pyrroles at the electron-deficient triple bond of haloacetylenes followed by the elimination of hydrogen halide from the zwitterionic intermediates.
As far as the relationship between the reactivity of the heterocycle and the solid salt used is concerned, a broad screening of various metal oxides and salts as mediators for the cross-coupling has shown that some of them are rather active (i.e., BaO). However, due to availability and convenience of the work-up of the reaction mixtures, Al2O3 and K2CO3 were taken as the agents of choice. A selection of specific metal oxides (Al2O3 or K2CO3) for a particular reaction is determined experimentally because the results significantly depend on the structure of both the pyrrole and haloacetylene employed.
This methodology was already partially documented in recent reviews [44,45,46,47,48,49,50,51,52]. This survey covers the recent publications (since 2014) concerning this reaction and the related chemistry which have not been yet summarized in a review.

2. Cross-Coupling of Haloacetylenes with Electron-Rich Heterocycles

2.1. Cross-Coupling of Haloacetylenes with Pyrroles

2.1.1. Cross-Coupling of Bromo- and Iodopropiolaldehydes with Pyrroles

A series of substituted pyrroles 1 were ethynylated by iodopropiolaldehyde in solid K2CO3 (a 10-fold excess) under mild conditions without solvent to afford highly reactive functionalized pyrrole compounds, 3-(pyrrol-2-yl)propiolaldehydes 2, in up to 40% yield (Scheme 1) [53]. The reagents were ground intensively for 5–10 min and allowed to stand at room temperature for 4 h. Iodopropioaldehyde was more preferable over explosive bromopropiolaldehyde.
In this case, the use of Al2O3 proved to be inappropriate, since the above ethynylation, albeit accelerated (the reaction time was 1 h), proceeded non-selectively to form, along with the target 3-(pyrrol-2-yl)propiolaldehydes 2, 3-bis(pyrrol-2-yl)acrylaldehydes 3, with the molar ratio being ~1:1 (Scheme 2).
It was found that 2-phenylpyrrole (1, R1 = H, R2 = Ph, R3 = H) gave the lowest yield (25%) of the ethynylated product that is likely associated with side reactions of the NH-function, e.g., nucleophilic addition across the triple bond or condensation with the aldehyde moiety.
The mechanism of the cross-coupling involves the single-electron transfer (SET) from pyrrole to iodopropiolaldehyde to generate the radical-ion pair A and/or the formation of the zwitterion B followed by elimination of hydrogen iodide (Scheme 3). Apparently, the role of K2CO3 is to stabilize the intermediate ion pairs by dipole-dipole interaction inside the ionic crystalline lattice of the medium, thus somewhat resembling ionic liquids.
The generation of radical-ions during this process was evidenced from ESR signals observed in the reaction of 1-vinyl-2-phenyl-3-amylpyrrole (1, R1 = CH=CH2, R2 = Ph, R3 = C5H11) with iodopropiolaldehyde in solid K2CO3.

2.1.2. Cross-Coupling of Acylbromoacetylenes with Pyrroles

With 2-(Furan-2-yl)- and 2-(2-Thiophen-2-yl)pyrroles

The reaction of 2-(furan-2-yl)- (4) and 2-(thiophen-2-yl)pyrroles 5 with acylbromoacetylenes 6ac was carried out according to the similar procedure: the reactants (1:1 molar ratio) were ground with a 10-fold excess of Al2O3 at room temperature for 1 h [54]. The major direction of this ethynylation of 2-(furan-2-yl)pyrroles 4 was the formation of 2-acylethynyl-5-(furan-2-yl)pyrroles 7 (Scheme 4), while the alternative 2-acylethynyl-5-(pyrrol-2-yl)furans 8 were minor products (7:8 = ~5–7:1). This result was key to understanding the ethynylation of five-membered aromatic heterocycles with haloacetylenes. In fact, this was the first observation of a relative reactivity of the furan ring in this reaction.
Double ethynylation, i.e., ethynylation of each ring, was not observed in any cases. In other words, the reaction occurs either with the pyrrole or furan ring. This points to a strong deactivating effect of the acyl substituent that is transmitted from one ring to another through the system of ten bonds involving conjugated one triple, four double, and five ordinary bonds. The ratio of products 7:8 = ~5–7:1 can be considered as an approximate measure of relative reactivity of the pyrrole and the furan ring towards the acylhaloacetylenes. The reaction of pyrroles with electrophilic acetylenes is commonly regarded as a nucleophilic addition of electron-rich pyrrole moiety (often as the pyrrolate anion) to the electron-deficient triple bond which occurs as N- and C-vinylation [55]. As mentioned above (see Section 2.1.1.) this reaction is likely initiated by the single-electron transfer to generate the radical-ion pairs as key intermediates, further forming C-C covalent bond with a final elimination of hydrogen halide [33]. Such a mechanism and the experimental isomer ratios are in agreement with a lower ionization potential of the pyrrole ring (8.09 eV) compared to that of furan ring (8.69 eV) [56].
In accordance with this rationale, 2-(thiophen-2-yl)pyrroles 5 reacted with acylbromoacetylenes 6ac under the above conditions to give only products of the pyrrole ring ethynylation, ethynylpyrroles 9 (Scheme 5) that also agrees with a higher ionization potentials of the thiophene ring (8.72 eV) [56].
In cases of NH-pyrroles (4, 5, R1 = H), 3-bromo-1-(pyrrol-2-yl)prop-2-en-1-ones 10 were isolated as the E-isomers stabilized by a strong intramolecular hydrogen bond between NH-proton and oxygen atom of the carbonyl group (Scheme 6).
The similar propenones were not observed among the products of ethynylation of N-vinylpyrroles (4, 5, R1 = CH=CH2) because they are not able to form the above stabilizing intramolecular hydrogen bonding.

With Dipyrromethanes

The solid-phase (Al2O3) ethynylation of dipyrromethane 11 with acylbromoacetylenes 6ac afforded 5-acylethynyldipyrromethanes 12 in 38–53% yields (Scheme 7) [57]. In contrast to the ethynylation of pyrrole giving 2-acylethynylpyrroles in the yield of 55–70% for 1 h [20], the cross-coupling of dipyrromethane 11 with acylbromoacetylenes 6ac required a much longer time and portion-wise addition of acetylene 6ac to the reaction mixture.
The low reaction rate in this case is likely resulted from the strong electron-withdrawing effect of the CF3-group, deactivating the pyrrole ring that acts as a nucleophile.
A general synthesis of such non-symmetrical dipyrromethanes was previously developed [58] by the condensation of trifluoropyrrolylethanols with pyrrole.
In the solid K2CO3, effective in the ethynylation of pyrroles with haloacetylenes [35], the above reaction did not take place at all.
In the solid alumina (room temperature, 96 h), dipyrromethane 13 reacted with benzoylbromoacetylene 6a to give insignificant amounts of products. From the reaction mixture, apart from the target dipyrromethane 14, 5-(1-bromo-2-benzoylethenyl)dipyrromethane 15, and the double ethynylation product, (dibenzoylethynyl)dipyrromethane 16, were isolated in low yields (Scheme 8). The formation of dipyrromethane 16 was the first example of ethynylation of the thiophene ring by the reaction studied.
To increase the nucleophilicity of the pyrrole ring, trimethylsilyl group was introduced to nitrogen atom of the pyrrole ring (Scheme 9). The ethynylation (K2CO3, room temperature, 168 h) of a mixture of dipyrromethanes 17 and 18 with acylbromoacetylenes 6ac gave acylethynyldipyrromethanes 14, 19 in 39–44% yields (Scheme 9). Thus, the yields of ethynylated product were increased due to the introduction of trimethylsilyl groups in the pyrrole ring to enhance their nucleophilicity.

With Tetrahydropyrrolo [3,2-c]pyridines

The cross-coupling of pyrrolo[3,2-c]pyridines 20 with acylbromoacetylenes 6a,b in solid K2CO3 was strictly chemo- and regioselective: exclusively propynones 21 were isolated (Scheme 10) [59].
In this case, the use of K2CO3 appeared to be essential, since it allowed the released HBr to be effectively fixed. This prevented the salt formation with the NH-function of the tetrahydropyridine moiety.
Indeed, when Al2O3 (instead of K2CO3) served as an active medium, the reaction of pyrrole 20 (R1 = C6H13) with benzoylbromoacetylene 6a afforded salt of propynone, hydrobromide 22 (Scheme 11). Upon treatment of the aqueous solution of salt 22 with NH4OH propynone 21 was obtained in 61% yield.

With Pyrrole-2-carbaldehydes

Pyrrole-2-carbaldehydes 23 proved to be inactive under usual conditions of the cross-coupling of pyrroles with acylhaloacetylenes in alumina medium (room temperature, 1 h). The reason is likely strong electron-withdrawing effect of the aldehyde group which decreases the pyrrole ring nucleophilicity. This fundamental hurdle was overcome by the acetal protection of the aldehyde function thereby decreasing its electron-withdrawing power [60,61]. The acetals 24 were treated with acylbromoacetylenes 6ac in the alumina medium (room temperature, 6 h) to obtain the expected ethynylated acetals 25. After the deprotection (aqueous acetone, HCl, room temperature, 1 h), the target ethynylated pyrrole-2-carbaldehydes 26 were isolated in 75–89% yields (Scheme 12).

2.1.3. Cross-Coupling of Bromotrifluoroacetylacetylene with Pyrroles

Pyrroles 27, when reacted with bromotrifluoroacetylacetylene 28 in the solid Al2O3 (room temperature, 2 h), gave only 4-bromo-1,1,1-trifluoro-4-(pyrrol-2-yl)but-3-en-2-ones 29 in 12–21% yields [62,63], while the cross-coupling of acetylbromoacetylene 30 with the same pyrroles under the same conditions afforded the expected acetylethynylpyrroles 31 (Scheme 13) [62].
Interestingly, N-vinylpyrroles 32 underwent normal cross-coupling with bromotrifluoroacetylacetylene 28 (Al2O3, rt, 2 h) to deliver ethynylpyrroles 33 in 42–58% yields (Scheme 14).
This implies that the cause of abnormal reaction (Scheme 13) is the interaction between NH and trifluoroacetyl groups that stabilizes 4-bromo-1,1,1-trifluoro-4-(pyrrol-2-yl)but-3-en-2-ones 29 in their E-configuration.
This is evidenced from the extraordinary downfield shift of the NH group proton signal (13–14 ppm) in the 1H NMR spectra of pyrroles 29.
The intramolecular H-O-bonding of such a type is likely realized already in the E-form of the intermediate zwitterion A (Scheme 15). This hydrogen bonding prevents the EZ isomerization and hence elimination of HBr, which usually occurs as a trans-process. Notably, in most cases of ethynylation of pyrroles under similar conditions [20], bromopyrrolylethenylketones of the type 29 are formed just as minor contaminants, if any (0–10% yields), that may also be a result of easier elimination of hydrogen halides (HBr in this case) from their Z-configuration. As the elimination of HBr does not occur at a stage of the zwitterion A formation, the proton in the 2 position of the pyrrole ring is transferred to the carbanionic center. This should be facilitated by a strong electron-withdrawing effect of trifluoroacetyl substituent. Consequently, the target product 29 is formed stereoselectively (as the E-isomer).
In the reaction of N-vinylpyrroles 32 with the bromotrifluoroacetylacetylene 28, the formation of an intramolecular hydrogen bond in the products is impossible (Scheme 16). Moreover, the formation of the E-isomer of 4-bromo-1,1,1-trifluoro-4-(pyrrol-2-yl)but-3-en-2-ones B would be sterically hindered (due to the repulsion between N-vinyl group and trifluoroacetyl substituent).
Probably, the effect of steric strain destabilizes the E-form at the stage of formation of the intermediate zwitterion B (Scheme 16), for which the Z-form turns out to be energetically favorable. At the final stage, the zwitter-ion B is transformed to trifluoroacetylethynylpyrroles 33 via elimination of the bromine anion accompanied by releasing of proton from the position 2 of the pyrrole ring (Scheme 16).
Pyrrole 33a, after 7 days contact with Al2O3, lost the trifluoroacetyl group to give 2-ethynylpyrrole 34 in 24% yield (Scheme 17). The partial detrifluoroacylation of pyrroles 33 also occurred during their passing through Al2O3-packed chromatographic column.

2.1.4. Cross-Coupling of Chloroethynylphosponates with Pyrroles

Pyrroles 35 were cross-coupled with chloroethynylphosponates 36 in solid alumina (room temperature, 24–48 h) to give 2-(pyrrol-2-yl)ethynylphosphonates 37 in 40–58% yields (Scheme 18) [64].
In the absence of Al2O3 (both in a solvent and under solvent-free conditions), the ethynylation did not take place. At room temperature, the complete conversion of the reactants was reached after 24 h. The exception was N-vinyl-4,5,6,7-tetrahydroindole 35a (R1 = H; R2-R3 = (CH2)4), the ethynylation of which lasted twice as long (48 h).
As minor products (2–10%), 2,2-bis(pyrrol-2-yl)vinylphosphonates 38 were detected in the reaction (Figure 1). Also, in the case of 1-vinyl-4,5,6,7-tetrahydroindole [35, R1 = CH=CH2; R2-R3 = (CH2)4], dialkyl 2,2-dichlorovinylphosphonates 39 (2–9%) were present in the reaction mixture (Figure 1).
The formation of the product 39 required [64] a longer reaction time (48 vs. 24 h) that allowed hydrogen chloride to be competitively added to the starting chloroethynylphosphonates according to [65]. A longer reaction rate is also due to the electron-withdrawing effect of the vinyl group, which reduces nucleophilicity of the pyrrole moiety.
In the solid K2CO3 medium (other conditions being the same), the cross-coupling of pyrroles with chloroethynylphosphonates produced only pyrrolylethynylphosphonates 37 in 38–43% yields.
It is suggested [64] that the reaction mechanism in this case represents the direct nucleophilic substitution of chlorine atom by the pyrrole moiety. This is supported by known data [66,67] that the reactions of chloroethynylphosphonates with nucleophiles including the neutral ones proceed mainly as a nucleophilic substitution of chlorine atom at the Csp carbon.

2.1.5. Cross-Coupling of Halopolyynes with Pyrroles

The above transition metal-free solid-phase mediated cross-coupling of haloacetylenes with pyrroles turns out to be efficient also for halopolyynes (di-, tri-, and tetrapolyynes) [68,69], allowing the pyrroles functionalized with polyynes chains to be synthesized. Such rare, highly reactive pyrrole compounds represent exclusively promising building blocks and precursors for the design of biologically and technically valuable heterocyclic molecules of exceptional complexity and structural diversity, including porphyrinoids with the polyyne substituents [70], modified bilirubins [71], and various ensembles of pyrroles with furans [72,73], thiophenes [72,73], pyrroles [72,73], naphthalenes [73], and other cyclic counterparts [73].
Thus, ester end-capped 1-halobutadiynes were successfully cross-coupled with pyrroles 40 in the solid K2CO3 to afford the expected butadiynyl-substituted pyrroles 41 in 43–80% yields (Scheme 19) [68].
The scope of reactions covers 2-phenylpyrrole, NH-4,5,6,7-tetrahydroindole, N-substituted 4,5,6,7-tetrahydroindoles, and chloro-, bromo-, and iodobutadiynes. The most suitable butadiynyl agents proved to be 1-bromobutadiynes.
The reaction rate depends on the pyrrole structure, with tetrahydroindole derivatives being the most reactive. For them, the cross-coupling with one equivalent of various halobutadiynes did not exceed 5 h, whereas for 2-phenylpyrrole, to reach 46–52% yield of the target product it required 2 equivalents of halobutadiynes and much longer reaction time (24 h).
This study was further extended over the longer chain aryl-capped 1-halopolyynes (up to tetraynes) [69]. As pyrrole substrates, 4,5,6,7-tetrahydroindole and its N-substituted derivatives were employed.
For the interaction of N-methyl-4,5,6,7-tetrahydroindole with 1-bromo-2-(4-cyanophenyl)acetylene 42a, 1-bromo-2-(4-cyanophenyl)butadiyne 42b, 1-bromo-2-(4-cyanophenyl)hexatriyne 42c, the expected cross-coupling was observed only for triyne 42c (K2CO3, room temperature, 3 h, 82% yield of hexatriynyl substituted N-methyl-4,5,6,7-tetrahydroindole 43c), while with acetylene 42a no target product was detected (1H-NMR), and in the case of diyne 42b a slow reaction (several days) took place (Scheme 20).
Since longer bromopolyynes were less stable than the corresponding iodine derivatives, 1-bromotetra- and -hexatriynes were used for the synthesis of tetradiynyl- and hexatriynyl-substituted tetrahydroindoles (Scheme 21), while 1-iodotetraynes were employed to produce octatetraynyltetrahydroindoles (Scheme 22).
Like in the work of Trofimov B.A. [33] (see also Section 2.1.1.), it is assumed that the reaction mechanism involves radical-ion pairs generated by the SET process (Scheme 23). According to experimental results longer polyyne chains secure a better stabilization of radical-ion pairs that provide higher yields of polyynyl substituted tetrahydroindoles and a shorter reaction time.

3. Reaction of Acylhaloacetylenes with Furans

A logical development of ethynylation of pyrroles with haloacetylenes [20] was the translation of this methodology to the furan compounds. In this line, on the example of menthofuran (3,6-dimethyl-4,5,6,7-tetrahydrobenzofuran 44), first synthetically appropriate results on the transition metal-free cross-coupling of the furan ring with haloacetylenes 6af initiated by their grinding with solid Al2O3 (room temperature, 1–72 h) were attained [74].
As it was found by Trofimov B.A. [74], after 1 h the reaction of menthofuran 44 with benzoylbromoacetylene 6a in solid Al2O3 resulted in the formation of ethynylfuran 45 along with the pair of diastereomeric cycloadducts of oxanorbornadiene structure 46 in 44:56 ratio (Scheme 24). The reaction is regioselective: the bromine atom is neighboring the position 2 of the furan ring exclusively.
Upon standing the reaction mixture for 72 h, the content of ethynylfuran 45 was increased up to 88%, while amount of cycloadduct 46 was reduced. These results indicate that cycloadduct 46 converts to ethynylfuran 45, i.e., the cycloadduct 46 is a kinetic intermediate of the ethynylation (this transformation is accompanied by elimination of hydrogen bromide).
The reaction of menthofuran 44 with chloro- and iodobenzoylacetylenes proceeded analogously leading for 1 h to a mixture of ethynylfuran 45 and cycloadduct 46, the latter disappearing completely after 72 h.
Thus, in contrast to cross-coupling of pyrroles with acylhaloacetylenes under similar conditions, ethynylation of the furan ring with acylhaloacetylenes occurred through [4+2]-cycloaddition followed by the elimination of HX during the ring-opening of the cycloadducts.
This reaction was proved to be applicable to bromoacetylenes with formyl (6d), acetyl (6e), furoyl (6b), thenoyl (6c), and ethoxy (6f) groups at the triple bond, which reacted with menthofuran 44 in the solid Al2O3 to afford the acetylenic derivatives 45af in 40–88% yields (Scheme 25).
Following the experimental results, the oxanorbornadiene intermediates such as 46, reversibly generated on the first reaction step, are transformed to the ethynyl derivatives of menthofuran 45 via a zwitterion with the positive charge distributed over the whole furan ring. The latter eliminates hydrogen bromide in the concerted process (hydrogen is released from the position 2 of the furan moiety, Scheme 26).
An experimental evidence for the proposed mechanism is the observation that cycloadducts 46 are gradually transformed to ethynylated products 45 in the solid Al2O3.

4. Reaction of Acylhaloacetylenes with Pyrazoles

The reaction of benzoylbromoacetylene 6a with pyrazole under conditions similar for the ethynylation of pyrroles (10-fold excess of Al2O3, room temperature, the molar ratio 1:1, 24 h), instead of the expected 2-benzoylethynylpyrazole 47, led to dipyrazolylenone 48a in 18% isolated yield (Scheme 27) [75]. The yield of enone 48a increased to 32%, when 2 equivalents of pyrazole was taken and reached 43% for the reaction with a 3-molar excess of the starting heterocycle.
The reaction proceeded via the intermediate (Z)-2-bromo-2-(pyrazol-1-yl)enone 49a and was accompanied by the formation of 2,2-dibromoenone 50a (Scheme 27).
Modest yields (22–35%) of dipyrazolylenones 48ac were observed using a two-fold molar excess of pyrazole relative to acylbromoacetylenes 6ac, with the yields of bromopyrazolylenones 49ac and dibromoenones 50ac being 10–18% and 8–14%, respectively (Scheme 27).
Surprisingly, no traces of ethynylpyrazoles 47 were detectable in the reaction mixture, implying that dipyrazolylenones 48a–c are not adducts of the reaction of pyrazole with the intermediate ethynylated pyrazoles 47.
3,5-Dimethylpyrazole reacted with acylbromoacetylenes 6ac,e in a 2:1 molar ratio to form dipyrazolylenones 51ac,e in 42–55% yields (Scheme 28). Bromopyrazolylenone of the type 49 in this case, was not discernible in the reaction mixture.
On the basis of the results obtained and previous mechanistic rationalizations concerning the reactions of pyrroles with haloacetylenes [20], it may be suggested that the synthesis of dipyrazolylenones 48 is triggered by the nucleophilic addition of pyrazole to the triple bond of acylbromoacetylenes 6ac to form the intermediate zwitterion (Scheme 29), which converts via proton transfer from the pyrazole moiety to its carbanionic center to give isolable intermediate 49. Subsequent nucleophilic substitution of the bromine atom by a second molecule of pyrazole affords dipyrazolylenone 48.
Unlike the ethynylation of pyrroles, where the initial zwitterion releases a halogen anion to restore the triple bond, for pyrazole, rapid intramolecular neutralization of the carbanionic site of the intermediate zwitterion occurs, which precludes formation of the ethynyl derivatives. Such a change of the reaction mechanism is likely due to the higher acidity of pyrazoles compared with pyrroles (pka of pyrazole is 14.2 whereas pka of pyrrole is 17.5).

5. Selected Reactions of Acylethynylpyrroles and Their Analogs

To demonstrate the possibilities of the cross-coupling developed for the construction of important functionalized heterocyclic systems, some selected synthetically attractive reactions of acylethynylpyrroles and their analogs are considered below.

5.1. Cyclizations with Propargylamine

5.1.1. Synthesis of Pyrrolo[1,2-a]pyrazines

Acylethynylpyrroles 52 were used for the synthesis of pyrrolo[1,2-a]pyrazines 53a,b according to the strategy which includes the following steps: (i) the non-catalyzed chemo- and regioselective nucleophilic addition of propargylamine to the triple bond of acylethynylpyrroles 52 to afford N-propargyl(pyrrolyl)aminoenones 54 and (ii) base-catalysed intramolecular cyclization of N-propargyl(pyrrolyl)aminoenones 54 to pyrrolo[1,2-a]pyrazines 53a,b (Scheme 30) [76].
Nucleophilic addition of propargylamine to the triple bond of acylethynylpyrroles 52 was carried out under reflux of reactants (52: propargylamine ratio being 1:2) in methanol for 5 h to deliver N-propargyl(pyrrolyl)aminoenones 54 (Scheme 30). The latter were formed as a mixture of E/Z isomers stabilized by intramolecular H-bonds between carbonyl group and NH-function of the amino moiety (the Z-isomer) or NH-function of the pyrrole ring (the E-isomer) with predominance of the Z-isomer.
The electronic nature of the substituents attached to the pyrrole ring determines the isomers ratio. Thus, for aminoenone 54 with unsubstituted pyrrole ring, the Z/E ratio is ~9:1. When a donor cyclohexane moiety is attached to the pyrrole ring [R1-R2 = (CH2)4], this ratio becomes 15:1, probably owing to a lower NH-acidity of the pyrrole counterpart and hence a weaker stabilization of the E-isomer by the intramolecular H-bonding. Consequently, for pyrroles with electron-withdrawing aryl substituents, having more acidic pyrrole NH-proton, the content of the E-isomer increases, Z/E ratio being ~4:1.
The cyclization of N-propargyl(pyrrolyl)aminoenones 54 was implemented by heating (60 °C, 15–30 min ) in the system Cs2CO3/DMSO to afford pyrazines 53a with exocyclic double bond and their thermodynamically more stable endocyclic isomers 53b. Pyrrolopyrazines 53b with the endocyclic double bond were formed selectively only from aminoenones 54 with unsubstituted pyrrole ring or with tetrahydroindole derivatives. In the case of enaminones with phenyl or fluorophenyl substituents, the major products were pyrrolopyrazines having the exocyclic double bond 53a (their content in the reaction mixture was spanned 70–90%), while pyrrolopyrazines 53b were minor products. The total yield of both isomers remained almost quantitative (90–96%).
Later, pyrrolopyrazines 53a,b were obtained (90–95% yields) via a one-pot procedure from ethynylpyrroles 52 and propargylamine when the reactants were heated (60–65 °C) in DMSO [77].

5.1.2. Synthesis of Pyrrolyl Pyridines

The one-pot reaction of N-substituted acylethynylpyrroles 55 with propargylamine in the presence of CuI selectively afforded 2-(pyrrol-2-yl)-3-acylpyridines 56 (Scheme 31) [78].
Catalyst-free heating of the reactants led to N-propargyl(pyrrolyl)aminoenones 57 which, upon keeping with CuI (equimolar amount) for 2.5 h at the same temperature, underwent the dihydrogenative ring closure to give pyrrolyl pyridines 56 (Scheme 31).
The duration of non-catalytic step strongly depended on the pyrrole structure: the acceptor substituents in the pyrrole ring facilitated the reaction (the reaction time was 6 h), while the donor ones slowed down the process (the reaction time was 16 h). A peculiar feature of this dehydrogenative cyclization is that the intermediate dihydropyridines 58 were aromatized rapidly (they are not usually detectable in the reaction mixture). Only in the case of acylethynyltetrahydroindole dihydropyridine 58 was isolated in 4% yield. Notably, the catalytic ring closure was almost insensitive to the structure of the initial acylethynylpyrroles 55 (the reaction time was about 2.5 h for all the cases).
A less predictable step of the synthesis is the intramolecular nucleophilic addition of the CH-bond adjacent to carbonyl group across the acetylenic moiety (Scheme 32). This CH-bond can be deprotonated under the action of amino group, either intramolecularly (autodeprotonation) to generate intermediate A or intermolecularly. Upon the complexing of Cu+ cation with the triple bond, the latter should be polarized to increase sensitivity towards the nucleophilic attack. This attack is completed by the addition of the carbanionic site to the terminal acetylenic atom to give the intermediate dihydropyridine 58.
The MS spectra of the reaction mixtures showed that the oxidation of intermediate 58 did not take place under the action of DMSO (no Me2S was detected). The air oxygen also did not participate in this process: the same results were obtained both under argon blanket and on air. Therefore, the Cu+ cation was considered [78] as a likely oxidant.
Latter [79] from the reaction of NH-acylethynylpyrroles 59 with propargylamine in the presence of CuX (X = Cl, Br, I), 3-acyl-2-(pyrrol-2-yl)-5-halopyridines 60 were unexpectedly isolated in 4–14% yields along with 3-acyl-2-(pyrrol-2-yl)pyridines 61 (28–61% yields) (Scheme 33). Evidently, the cause of this difference compared to the previous cyclization [78] was the NH-functionality of the starting acylethynylpyrroles.
Under the above conditions, pyrrolyl pyridines 61 were not halogenated with CuX, thus indicating that construction of the halogenated pyridine ring occurred before its closure. It is supposed (Scheme 34) [79] that hydrogen halides, reversibly generated by the interaction of the NH pyrrole moiety of the intermediate N-propargyl(pyrrolyl)aminoenone 57 with CuX, add to the triple bond activated by π-complexing with other CuX molecules to give haloallyl intermediate B (Scheme 34). Afterwards, the intramolecular addition of the CH bond to the allyl moiety takes place to form the intermediate 5-halotetrahydropyridyl intermediate C. Aromatization of the latter is finalized via the reaction with CuX and further oxidation by Cu+ cations as previously described for a similar process [78].

5.2. Synthesis of Pyrrolizines via Three-Component Cyclization with Benzylamine and Acylacetylenes

On the platform of acylethynylpyrroles 52, a new general strategy for the synthesis of functionalized pyrrolizines was developed [80]. It consisted of the two steps: (i) the base-catalyzed addition of a benzylamine to 2-acylethynylpyrroles 52 to give pyrrolylaminoenones 62; (ii) non-catalyzed addition of N-benzyl(pyrrolyl)aminoenones 62 to the triple bond of acylacetylenes 63 followed by the intramolecular cyclization of the intermediate pentadiendiones 64 thus formed to 1-benzylamino-2-acyl-3-methylenoacylpyrrolizines 65 (Scheme 35).
The nucleophilic addition of benzylamine to the triple bond of 2-acylethynylpyrroles 52 was realized in the presence K3PO4/DMSO catalytic system to smoothly deliver N-benzyl(pyrrolyl)aminoenones 62 in up to 97% yield (Scheme 35). The latter were formed as a mixture of the E/Z isomers, the E-isomer being obviously stabilized by intramolecular H-bonds between the carbonyl group and NH-function of the pyrrole ring. As in the case of the addition of propargylamine to acylethynylpyrrole (see Section 5.1), the structure of the substituents of the pyrrole ring strongly influences the isomer ratio of the adducts: the donor substituents increase the content of the Z-isomers.
Further, the aminoenones 62 chemo- and regioselectively reacted with acylacetylenes 63 to afford the intermediate pentadiendiones 64, which then cyclized to 1-benzylamino-2-acyl-3-methylenoacylpyrrolyzines 65 in up to 80% yield (Scheme 35).

5.3. Reactions with Ethylenediamine

The reaction of 2-benzoylethynylpyrroles 55a,b with ethylenediamine was realized upon reflux of their equimolar mixture in dioxane (40 h) [81]. Expectedly, first the addition of diamine gave monoadduct 66a,b, which, in the case of acylethynylpyrrole 55a, underwent intramolecular cyclization/fragmentation to afford tetrahydroindolyl imidazoline 67a and acetophenone (Scheme 36).
In this reaction, in the case of acylethynylpyrrole 55b, the formation of dihydrodiazepine 68 takes place. This is a result of the intramolecular cyclization of monoadduct 66b with the participation of the carbonyl group followed by dehydration (Scheme 37).

5.4. Cyclization with Hydrazine: Synthesis of Pyrrolyl Pyrazoles

The building up of the pyrazole ring over acetylenic moiety of pyrrolopyridine propynones 21 via its ring closure with hydrazine gave 4,5,6,7-tetrahydropyrrolo[3,2-c]pyridine-pyrazole ensembles 69 in 92–98% yields (Scheme 38) [59].
According to the above procedure, a new extended dipyrromethane system conjugated with pyrazole cycle 70 was obtained in almost quantitative yield (Scheme 39) [57].

5.5. Cyclization with Hydroxylamine: Synthesis of Pyrrolyl Isoxazoles

Acylethynyltetrahydroindoles 71 readily cyclized with hydroxylamine to give regioselectively either 3-(4,5,6,7-tetrahydroindol-2-yl)-4,5-dihydroisoxazol-5-ols 72 or 5-(4,5,6,7-tetrahydroindol-2-yl)isoxazoles 73 (Scheme 40) [82]. The cyclization can be easily switched from the direction leading exclusively to isoxazoles 72 to the formation of isoxazoles 73 by simple changing of the proton concentration in the reaction mixture. When the reaction was carried out in the presence of acetic acid (NH2OH·HCl/NaOAc, 1:1 system), only isoxazoles 72 were formed, whereas under neutral or basic conditions (NH2OH·HCl/NaOH (1:1 or 1:1.5 system), the cyclization took another pathway to produce preferably (94s–97% or entirely) isoxazoles 73.
Apparently, in the presence of acetic acid, the attack of the NH2OH nucleophile at the β-acetylenic carbon of tetrahydroindoles 71 is electrophilically assisted by the simultaneous protonation of the carbonyl group (and finally 1,4-addition takes place to deliver isoxazoles 72), as shown in Scheme 41.
In the presence of the NH2OH·HCl/NaOH system, which is unable to exert the electrophilic assistance, the common oximation of the carbonyl group prevailed.
Moreover, 4,5-dihydroisoxazol-5-ols 72 underwent easy aromatization when refluxing (benzene, 1 h) in the presence of TsOH·H2O to isoxazoles 74 in 73–91% yields (Scheme 42) [82].
On the basis of the above cycloaddition, two approaches to the synthesis of meso-CF3 substituted dipyrromethanes 75–77 bearing isoxazole moieties were developed [83].
The key stages of these approaches are the cycloaddition of hydroxylamine to the triple bond of ethynyldipyrromethanes 12a, 78 (Scheme 43 and Scheme 44), or the synthesis of pyrrolyl isoxazoles 75, 77 from ethynylpyrrole 79 (accessible from pyrrole and benzoylbromoacetylene), and its further condensation with 2,2,2-trifluoro-1-(pyrrol-2-yl)-1-ethanols 80, as described in Section 2.1.2. (Scheme 45).

5.6. Cyclization with Methylene Active Esters: Synthesis of Pyrrolyl Pyrones

The [4+2]-cycloaddition between 2-acylethynylpyrroles 83 and methylene active esters (Scheme 46), offering a short-cut to pyrrolyl pyrones 84 in good to high yields, was described [84].
The reaction was carried out in acetonitrile in the presence of 1.5 molar excess of KOH. As methylene active esters, diethylmalonate, ethyl acetoacetate and ethyl cyanoacetate were used.
The cyclization is triggered by the proton abstraction from the active CH2 group of methylene active esters followed by the nucleophilic attack of the carbanion A, thus generated at the triple bond of acylethynylpyrroles 83 to afford intermediate B. The subsequent intramolecular nucleophilic substitution of the ethoxy group in the ester function by the oxygen-centered anion (the resonance form of the intermediate B) furnishes the target products (Scheme 47).

5.7. Unprecedented Four-Proton Migration in Acylethynylmenthofurans: “A Proton Pump”

When benzoylethynylmenthofuran 45a was heated at reflux in CHCl3 in the presence of HBr, the formation of benzoylethylbenzofuran 85a in 95% yield was observed (Scheme 48) [85]. Thus, the transfer of four hydrogen atoms from the cyclohexane ring to the triple bond took place.
This rearrangement was found to be general for other acylethynyl derivatives (furoyl, thenoyl, alkoxycarbonyl) of menthofuran to give their acylethylbenzofuran derivatives in the yield of 44%, 48%, and 24% respectively (Scheme 48).
Basing on these experimental results, it can be postulated that the rearrangement starts with protonation of acylethynyltetahydrobenzofuran moiety with HBr to give carbocation A, which in its more stable mesomeric form B abstracts a hydride-ion from the adjacent position (C-7) with positive charge transfer to form carbocation C. Then, two hydride shifts in the cyclohexane ring transform carbocation C into carbocation D with the positive charge at C-5. Proton abstraction from the C-4 position of this carbocation leads to the cyclohexene moiety and regenerates HBr. Simultaneously, after two 1,3-hydrogen shifts in the furan counterpart, it is transformed into vinyl intermediate E. Next, protonation of the double bond with HBr results in the formation of carbocation F which in its stable endocyclic form accepts the hydride ion from the cyclohexene ring to give cyclohexene carbocation G. The release of a proton from the latter gives the cyclohexadiene ring and HBr. Two 1,3-hydrogen shifts in the furan moiety completes the four-hydrogen transfer to the side chain giving 3,6-dimethylbenzofuran 87 with a saturated side chain, i.e., an exhaustively hydrogenated acetylene moiety (Scheme 49).
The driving force of this spectacular “hydrogen pump” is the energy gain due to the formation of the aromatic benzofuran system.

6. Concluding Remarks and Outlook

This review evidences that the cross-coupling reactions between electrophilic haloacetylenes and electron-rich heterocycles assisted by Al2O3 or K2CO3 or similar solid oxides and salts continue to be expanded, occupying more and more areas of heterocyclic chemistry. These endeavors are stimulated by such competitive beneficial features of this methodology as transition metal-free, no-solvent, mild conditions, availability of the starting materials, very simple synthetic operations, and possibility to introduce acetylenic substituents with electron-withdrawing groups into a heterocyclic core. Now, these reactions pave a short way to previously inaccessible or unknown, highly reactive heterocyclic building blocks and precursors to create novel heterocyclic systems of greater diversity and complexity.

Funding

This work was supported by the Ministry of Science and Higher Education of the Russian Federation (topic № AAAA-A16-116112510005-7). APC was sponsored by MDPI.

Acknowledgments

Authors acknowledge Baikal Analytical Center for collective use SB RAS for the equipment.

Conflicts of Interest

There are no conflicts to declare.

References

  1. Mishra, R.; Jha, K.K.; Kumar, S.; Tomer, I. Synthesis, properties and biological activity of thiophene: A review. Der. Pharm. Chem. 2011, 3, 38–54. [Google Scholar]
  2. Chaudhary, A.; Jha, K.K.; Kumar, S. Biological diversity of thiophene: A review. J. Adv. Sci. Res. 2012, 3, 3–10. [Google Scholar]
  3. Kharb, R.; Birla, S.; Sharma, A.K. Recent updates on antimicrobial potential of novel furan derivatives. Int. J. Pharm. Phytopharmacol. Res. 2014, 3, 451–459. [Google Scholar]
  4. Bhardwaj, V.; Gumber, D.; Abbot, V.; Dhiman, S.; Sharma, P. Pyrrole: A resourceful small molecule in key medicinal hetero-aromatics. RSC Adv. 2015, 5, 15233–15266. [Google Scholar] [CrossRef]
  5. Konstantinidou, M.; Gkermani, A.; Hadjipavlou-Litina, D. Synthesis and pharmacochemistry of new pleiotropic pyrrolyl derivatives. Molecules 2015, 20, 16354–16374. [Google Scholar] [CrossRef]
  6. Mishra, R.; Sharma, P.K. A review on synthesis and medicinal importance of thiophene. Int. J. Eng. Al. Sci. 2015, 1, 46–59. [Google Scholar]
  7. Heravi, M.M.; Zadsirjan, V. Recent advances in the synthesis of biologically active compounds containing benzo[b]furans as a framework. Curr. Org. Syn. 2016, 13, 780–833. [Google Scholar] [CrossRef]
  8. Gholap, S.S. Pyrrole: An emerging scaffold for construction of valuable therapeutic agents. Eur. J. Med. Chem. 2016, 110, 13–31. [Google Scholar] [CrossRef]
  9. Yet, L. Privileged Structures in Drug Discovery: Medicinal Chemistry and Synthesis, 1st ed.; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2018. [Google Scholar]
  10. Shah, R.; Verma, P.K. Therapeutic importance of synthetic thiophene. Chem. Centr. J. 2018, 12, 137–158. [Google Scholar] [CrossRef] [Green Version]
  11. Ahmad, S.; Alam, O.; Naim, M.J.; Shaquiquzzaman, M.; Alam, M.M.; Iqbal, M. Pyrrole: An insight into recent pharmacological advances with structure activity relationship. Eur. J. Med. Chem. 2018, 157, 527–561. [Google Scholar] [CrossRef]
  12. Chiurchiù, E.; Gabrielli, S.; Ballini, R.; Palmieri, A. A new valuable synthesis of polyfunctionalized furans starting from β-nitroenones and active methylene compounds. Molecules 2019, 24, 4575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Zeni, G.; Lüdtke, D.S.; Nogueira, C.W.; Panatieri, R.B.; Braga, A.L.; Silveira, C.C.; Stefani, H.A. New acetylenic furan derivatives: Synthesis and anti-inflammatory activity. Tetrahedron Lett. 2001, 42, 8927–8930. [Google Scholar] [CrossRef]
  14. Acetylene Chemistry: Chemistry, Biology and Material Science; Diederich, F.; Stang, P.J.; Tykwinski, R.R. (Eds.) Wiley-VCH: Weinheim, Germany, 2005. [Google Scholar]
  15. Li, Y.; Waser, J. Platinum-catalyzed domino reaction with benziodoxole reagents for accessing benzene-alkynylated indoles. Angew. Chem. Int. Ed. 2015, 54, 5438–5442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Sonogashira, K. Comprehensive Organic Synthesis; Trost, B.M., Fleming, I., Pattenden, G., Eds.; Pergamon Press: Oxford, UK, 1991; pp. 521–561. [Google Scholar]
  17. Chinchilla, R.; Nájera, C. The Sonogashira reaction:  A booming methodology in synthetic organic chemistry. Chem. Rev. 2007, 107, 874–922. [Google Scholar] [CrossRef]
  18. Chinchilla, R.; Njera, C. Recent advances in Sonogashira reactions. Chem. Soc. Rev. 2011, 40, 5084–5121. [Google Scholar] [CrossRef]
  19. Heravi, M.; Dehghani, M.; Zadsirjan, V.; Ghanbarian, M. Alkynes as privileged synthons in selected organic name reactions. Curr. Org. Syn. 2019, 16, 205–243. [Google Scholar] [CrossRef]
  20. Trofimov, B.A.; Stepanova, Z.V.; Sobenina, L.N.; Mikhaleva, A.I.; Ushakov, I.A. Ethynylation of pyrroles with 1-acyl-2-bromoacetylenes on alumina: A formal-inverse Sonogashira coupling. Tetrahedron Lett. 2004, 34, 6513–6516. [Google Scholar] [CrossRef]
  21. Seregin, I.V.; Ryabova, V.; Gevorgyan, V. Direct palladium-catalyzed alkynylation of N-fused heterocycles. J. Am. Chem. Soc. 2007, 129, 7742–7743. [Google Scholar] [CrossRef] [Green Version]
  22. Gu, Y.; Wang, X. Direct palladium-catalyzed C-3 alkynylation of indoles. Tetrahedron Lett. 2009, 50, 763–766. [Google Scholar] [CrossRef]
  23. Matsuyama, N.; Hirano, K.; Satoh, T.; Miura, M. Nickel-catalyzed direct arylation of azoles with aryl bromides. Org. Lett. 2009, 11, 4156–4159. [Google Scholar] [CrossRef]
  24. Matsuyama, N.; Kitahara, M.; Hirano, K.; Satoh, T.; Miura, M. Nickel- and copper- direct alkynylation of azoles and polyfluoroarenes with terminal alkynes under O2 or atmospheric conditions. Org. Lett. 2010, 12, 2358–2361. [Google Scholar] [CrossRef] [PubMed]
  25. Besselievre, F.; Piguel, S. Copper as a powerful catalyst in the direct alkynylation of azoles. Angew. Chem. Int. Ed. 2009, 48, 9553–9556. [Google Scholar] [CrossRef] [PubMed]
  26. Brand, J.P.; Charpentier, J.; Waser, J. Direct alkynylation of indole and pyrrole heterocycles. Angew. Chem. Int. Ed. 2009, 48, 9346–9349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Stepanova, Z.V.; Sobenina, L.N.; Mikhaleva, A.I.; Ushakov, I.A.; Chipanina, N.N.; Elokhina, V.N.; Voronov, V.K.; Trofimov, B.A. Silica-assisted reactions of pyrroles with 1-acyl-2-bromoacetylenes. Synthesis 2004, 2736–2742. [Google Scholar] [CrossRef]
  28. Trofimov, B.A.; Sobenina, L.N.; Stepanova, Z.V.; Demenev, A.P.; Mikhaleva, A.I.; Ushakov, I.A.; Vakul’skaya, T.I.; Petrova, O.V. Synthesis of 2-benzoylethynylpyrroles by cross-coupling of 2-arylpyrroles with 1-benzoyl-2-bromoacetylene over aluminium oxide. Russ. J. Org. Chem. 2006, 42, 1348–1355. [Google Scholar] [CrossRef]
  29. Trofimov, B.A.; Stepanova, Z.V.; Sobenina, L.N.; Mikhaleva, A.I.; Sinegovskaya, L.M.; Potekhin, K.A.; Fedyanin, I.V. 2-(2-Benzoylethynyl)-5-phenylpyrrole: Fixation of cis- and trans-rotamers in a crystal state. Mendeleev Commun. 2005, 15, 229–232. [Google Scholar] [CrossRef]
  30. Sobenina, L.N.; Demenev, A.P.; Mikhaleva, A.I.; Ushakov, I.A.; Vasil’tsov, A.M.; Ivanov, A.V.; Trofimov, B.A. Ethynylation of indoles with 1-benzoyl-2-bromoacetylene on Al2O3. Tetrahedron Lett. 2006, 47, 7139–7141. [Google Scholar] [CrossRef]
  31. Trofimov, B.A.; Sobenina, L.N.; Stepanova, Z.V.; Ushakov, I.A.; Petrova, O.V.; Tarasova, O.A.; Volkova, K.A.; Mikhaleva, A.I. Regioselective cross-coupling of 1-vinylpyrroles with acylbromoacetylenes on Al2O3: A synthesis of 1-vinyl-2-(2-acylethynyl)pyrroles. Synthesis 2007, 39, 447–451. [Google Scholar] [CrossRef]
  32. Trofimov, B.A.; Sobenina, L.N.; Demenev, A.P.; Stepanova, Z.V.; Petrova, O.V.; Ushakov, I.A.; Mikhaleva, A.I. A palladium- and copper-free cross-coupling of ethyl 3-halo-2-propynoates with 4,5,6,7-tetrahydroindoles on alumina. Tetrahedron Lett. 2007, 48, 4661–4664. [Google Scholar] [CrossRef]
  33. Trofimov, B.A.; Sobenina, L.N.; Stepanova, Z.V.; Vakul’skaya, T.I.; Kazheva, O.N.; Aleksandrov, G.G.; Dyachenko, O.A.; Mikhaleva, A.I. Reactions of 2-phenylpyrrole with bromobenzoylacetylene on metal oxides active surfaces. Tetrahedron 2008, 64, 5541–5544. [Google Scholar] [CrossRef]
  34. Petrova, O.V.; Sobenina, L.N.; Ushakov, I.A.; Mikhaleva, A.I. Reaction of indoles with ethyl bromopropynoate over Al2O3 surface. Russ. J. Org. Chem. 2008, 44, 1512–1516. [Google Scholar] [CrossRef]
  35. Trofimov, B.A.; Sobenina, L.N.; Stepanova, Z.V.; Petrova, O.V. Chemo- and regioselective ethynylation of 4,5,6,7-tetrahydroindoles with ethyl 3-halo-2-propynoates. Tetrahedron Lett. 2008, 49, 3946–3949. [Google Scholar] [CrossRef]
  36. Sobenina, L.N.; Tomilin, D.N.; Petrova, O.V.; Gulia, N.; Osowska, K.; Szafert, S.; Mikhaleva, A.I.; Trofimov, B.A. Cross-coupling of 4,5,6,7-tetrahydroindole with functionalized haloacetylenes on active surfaces of metal oxides and salts. Russ. J. Org. Chem. 2010, 46, 1373–1377. [Google Scholar] [CrossRef]
  37. Sobenina, L.N.; Tomilin, D.N.; Petrova, O.V.; Ushakov, I.A.; Mikhaleva, A.I.; Trofimov, B.A. Hydroamination of 2-ethynyl-4,5,6,7-tetrahydroindoles: Towards 2-substituted aminoderivatives of indole. Synthesis 2010, 42, 2468–2474. [Google Scholar]
  38. Trofimov, B.A.; Sobenina, L.N.; Stepanova, Z.V.; Ushakov, I.A.; Mikhaleva, A.I.; Tomilin, D.N.; Kazheva, O.N.; Alexandrov, G.G.; Chekhlov, A.N.; Dyachenko, O.A. Peculiar rearrangement of the [2+2]-cycloadducts of DDQ and 2-ethynylpyrroles. Tetrahedron Lett. 2010, 51, 5028–5031. [Google Scholar] [CrossRef]
  39. Trofimov, B.A.; Sobenina, L.N.; Stepanova, Z.V.; Ushakov, I.A.; Sinegovskaya, L.M.; Vakul’skaya, T.I.; Mikhaleva, A.I. Facile [2+2]-cycloaddition of DDQ to acetylenic moiety: Synthesis of pyrrole-(indole)bicyclo[4.2.0]octadiene ensembles from C-ethynylpyrroles or –indoles. Synthesis 2010, 42, 470–477. [Google Scholar] [CrossRef]
  40. Sobenina, L.N.; Stepanova, Z.V.; Ushakov, I.A.; Mikhaleva, A.I.; Tomilin, D.N.; Kazheva, O.N.; Alexandrov, G.G.; Dyachenko, O.A.; Trofimov, B.A. From 4,5,6,7-tetrahydroindole to functionalized furan-2-one-4,5,6,7-tetrahydroindole-cyclobutene sequence in two steps. Tetrahedron 2011, 67, 4832–4837. [Google Scholar] [CrossRef]
  41. Sobenina, L.N.; Tomilin, D.N.; Ushakov, I.A.; Mikhaleva, A.I.; Trofimov, B.A. Transition-metal-free stereoselective and regioselective hydroamination of 2-benzoylethynyl-4,5,6,7-tetrahydroindoles with amino acids. Synthesis 2012, 44, 2084–2090. [Google Scholar] [CrossRef]
  42. Sobenina, L.N.; Tomilin, D.N.; Ushakov, I.A.; Mikhaleva, A.I.; Ma, J.S.; Yang, G.; Trofimov, B.A. The “Click-Chemistry” with 2-ethynyl-4,5,6,7-tetrahydroindoles: Towards a functionalized tetrahydroindole-triazole ensembles. Synthesis 2013, 45, 678–682. [Google Scholar] [CrossRef]
  43. Sobenina, L.N.; Stepanova, Z.V.; Petrova, O.V.; Ma, J.S.; Yang, G.; Tatarinova, A.A.; Mikhaleva, A.I.; Trofimov, B.A. Synthesis of 3-[5-(biphenyl-4-yl)pyrrol-2-yl]-1-phenylprop-2-yn-1-ones by palladium-free cross-coupling between pyrroles and haloalkynes on aluminum oxide. Russ. Chem. Bull. Int. Ed. 2013, 62, 88–92, [Izv. Akad. Nauk Ser. Khim. 2013, 88–92]. [Google Scholar] [CrossRef]
  44. Banwell, M.G.; Goodwin, T.E.; Ng, S.; Smith, J.A.; Wong, D.J. Palladium-catalysed cross-coupling and reactions involving pyrroles. Eur. J. Org. Chem. 2006, 3043–3060. [Google Scholar] [CrossRef]
  45. Trofimov, B.A.; Nedolya, N.A. Comprehensive Heterocyclic Chemistry III; Katritzky, A.R., Ramsden, C.A., Scriven, E.F.V., Taylor, R.J.K., Eds.; Elsevier: Amsterdam, The Netherlands, 2008; Volume 3, p. 45. [Google Scholar]
  46. Trofimov, B.A.; Sobenina, L.N. Targets in Heterocyclic Chemistry; Attanasi, O.A., Spinelli, D., Eds.; Societa Chimica Italiana: Rome, Italy, 2009; pp. 92–119. [Google Scholar]
  47. Tanaka, K.; Kaupp, G. Solvent-Free Organic Synthesis; WILEY-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2009; 468p. [Google Scholar]
  48. Trofimov, B.A.; Mikhaleva, A.I.; Schmidt, E.Y.; Sobenina, L.N. Pyrroles and N-vinylpyrroles from ketones and acetylenes: Recent strides. Adv. Heterocycl. Chem. 2010, 99, 209–254. [Google Scholar]
  49. Dudnik, A.S.; Gevorgyan, V. Formal inverse Sonogashira reaction: Direct alkynylation of arenes and heterocycles with alkynyl halides. Angew. Chem. Int. Ed. 2010, 49, 2096–2098. [Google Scholar] [CrossRef] [PubMed]
  50. Messaoudi, S.; Brion, J.-D.; Alami, M. Transition-metal-catalyzed direct C–H alkenylation, alkynylation, benzylation, and alkylation of (hetero)arenes. Eur. J. Org. Chem. 2010, 6495–6516. [Google Scholar] [CrossRef]
  51. Trofimov, B.A.; Mikhaleva, A.I.; Schmidt, E.Y.; Sobenina, L.N. Chemistry of Pyrroles; CRC Press Inc: Boca-Raton, LA, USA, 2014. [Google Scholar]
  52. Sobenina, L.N.; Tomilin, D.N.; Trofimov, B.A. C-Ethynylpyrroles: Synthesis and reactivity. Russ. Chem. Rev. 2014, 83, 475–501. [Google Scholar] [CrossRef]
  53. Gotsko, M.D.; Sobenina, L.N.; Tomilin, D.N.; Markova, M.V.; Ushakov, I.A.; Vakul’skaya, T.I.; Khutsishvili, S.S.; Trofimov, B.A. Pyrrole acetylenecarbaldehydes: An entry to a novel class of functionalized pyrroles. Tetrahedron Lett. 2016, 57, 4961–4964. [Google Scholar] [CrossRef]
  54. Sobenina, L.N.; Petrova, O.V.; Tomilin, D.N.; Gotsko, M.D.; Ushakov, I.A.; Klyba, L.V.; Mikhaleva, A.I.; Trofimov, B.A. Ethynylation of 2-(furan-2-yl)- and 2-(thiophen-2-yl)pyrroles with acylbromoacetylenes in the Al2O3 medium: Relative reactivity of heterocycles. Tetrahedron 2014, 70, 9506–9511. [Google Scholar] [CrossRef]
  55. Trofimov, B.A.; Sobenina, L.N.; Mikhaleva, A.I.; Ushakov, I.A.; Vakul’skaya, T.I.; Stepanova, Z.V.; Toryashinova, D.-S.D.; Mal’kina, A.G.; Elokhina, V.N. N- and C-Vinylation of pyrroles with disubstituted activated acetylenes. Synthesis 2003, 35, 1272–1279. [Google Scholar] [CrossRef]
  56. Trofimov, A.B.; Zaitseva, I.L.; Moskovskaya, T.E.; Vitkovskaya, N.M. Theoretical investigation of photoelectron spectra of furan, pyrrole, thiophene, and selenole. Chem. Heterocycl. Compd. 2008, 44, 1101–1112. [Google Scholar] [CrossRef]
  57. Tomilin, D.N.; Petrushenko, K.B.; Sobenina, L.N.; Gotsko, M.D.; Ushakov, I.A.; Skitnevskaya, A.D.; Trofimov, A.B. Synthesis and optical properties of meso-CF3-BODIPY with acylethynyl substituents in the position 3 of the indacene core. Asian J. Org. Chem. 2016, 5, 1288–1294. [Google Scholar] [CrossRef]
  58. Sobenina, L.N.; Vasil’tsov, A.M.; Petrova, O.V.; Petrushenko, K.B.; Ushakov, I.A.; Clavier, G.; Meallet-Renault, R.; Mikhaleva, A.I.; Trofimov, B.A. A general route to symmetric and asymmetric meso-CF3-3(5)-aryl(hetaryl)- and 3,5-diaryl(dihetaryl)-BODIPY dyes. Org. Lett. 2011, 13, 2524–2527. [Google Scholar] [CrossRef] [PubMed]
  59. Sagitova, E.F.; Tomilin, D.N.; Petrova, O.V.; Budaev, A.B.; Sobenina, L.N.; Trofimov, B.A.; Yang, G.Q.; Hu, R. Acetylene-based short-cut from oxime of 2,2,6,6-tetramethylpiperidine-4-one to 4,4,6,6-tetramethyl-4,5,6,7-tetrahydropyrrolo[3,2-c]pyridine-pyrazole ensembles. Mendeleev Commun. 2019, 29, 658–660. [Google Scholar] [CrossRef]
  60. Sobenina, L.N.; Markova, M.V.; Tomilin, D.N.; Tret’yakov, E.V.; Ovcharenko, V.I.; Mikhaleva, A.I.; Trofimov, B.A. First example of the synthesis of pyrrolecarbaldehyde with electron-deficient acetylene substituents. Russ. J. Org. Chem. 2013, 49, 1241–1243. [Google Scholar] [CrossRef]
  61. Tomilin, D.N.; Sobenina, L.N.; Markova, M.V.; Gotsko, M.D.; Ushakov, I.A.; Smirnov, V.I.; Vaschenko, A.V.; Mikhaleva, A.I.; Trofimov, B.A. Expediant strategy for the synthesis of 5-acylethynylpyrrole-2-carbaldehydes. Synth. Commun. 2015, 45, 1652–1661. [Google Scholar] [CrossRef]
  62. Tomilin, D.N.; Soshnikov, D.Y.; Trofimov, A.B.; Gotsko, M.D.; Sobenina, L.N.; Ushakov, I.A.; Afonin, A.V.; Koldobsky, A.B.; Vitkovskaya, N.M.; Trofimov, B.A. A peculiarity in the Al2O3-mediated cross-coupling of pyrroles with bromotrifluoroacetylacetylene: A quantum-chemical insight. Mendeleev Commun. 2016, 26, 480–482. [Google Scholar] [CrossRef]
  63. Tomilin, D.N.; Gotsko, M.D.; Sobenina, L.N.; Ushakov, I.A.; Afonin, A.V.; Soshnikov, D.Y.; Trofimov, A.B.; Koldobsky, A.B.; Trofimov, B.A. N-Vinyl-2-(trifluoroacetylethynyl)pyrroles and E-2-(1-bromo-2-trifluoroacetylethenyl)pyrroles: Cross-coupling vs. addition during C-H-functionalization of pyrroles with bromotrifluoroacetylacetylene in solid Al2O3 medium. H-bonding control. J. Fluor. Chem. 2016, 186, 1–6. [Google Scholar] [CrossRef]
  64. Gotsko, M.D.; Sobenina, L.N.; Tomilin, D.N.; Ushakov, I.A.; Dogadina, A.V.; Trofimov, B.A. Topochemical mechanoactivated phosphonylethynylation of pyrroles with chloroethynylphosphonates in solid Al2O3 or K2CO3 media. Tetrahedron Lett. 2015, 57, 4961–4964. [Google Scholar] [CrossRef]
  65. Garibina, B.A.; Dogadina, A.V.; Zakharov, V.I.; Ionin, B.I.; Petrov, A.A. Phosphorus containing ynamines. Zh. Obshch. Khim. 1979, 71, 1964–1973. [Google Scholar]
  66. Petrov, A.A.; Dogadina, A.V.; Ionin, B.I.; Garibina, B.A.; Leonov, A.A. The Arbuzov rearrangement with participation of halogenoacetylenes as a method of synthesis of ethynylphosphonates and other organophosphorus compounds. Russ. Chem. Rev. 1983, 52, 1030–1035. [Google Scholar] [CrossRef]
  67. Dogadina, A.V.; Svintsitskaya, N.I. New reactions of chloroethynylphosphonates. Russ. J. Gen. Chem. 2015, 85, 351–358. [Google Scholar] [CrossRef]
  68. Tomilin, D.N.; Pigulski, B.; Gulia, N.; Arendt, A.; Sobenina, L.N.; Mikhaleva, A.I.; Szafert, S.; Trofimov, B.A. Direct synthesis of butadiynyl-substituted pyrroles under solvent and transition metal-free conditions. RSC Adv. 2015, 5, 73241–73248. [Google Scholar] [CrossRef]
  69. Pigulski, B.; Arendt, A.; Tomilin, D.N.; Sobenina, L.N.; Trofimov, B.A.; Szafert, S. Transition-metal free mechanochemical approach to polyyne substituted pyrroles. J. Org. Chem. 2016, 81, 9188–9198. [Google Scholar] [CrossRef] [PubMed]
  70. Cho, D.H.; Joon, H.L.; Kim, B. An improved synthesis of 1,4-bis(3,4-dimethyl-5-formyl-2-pyrryl)butadiyne and 1,2-bis(3,4-dimethyl-5-formyl-2-pyrryl)ethyne. J. Org. Chem. 1999, 64, 8048–8050. [Google Scholar] [CrossRef]
  71. McDonagh, A.F.; Lightner, D.A. Influence of conformation and intramolecular hydrogen bonding on the acyl glucuronidation and biliary excretion of acetylenic bis-dipyrrinones related to bilirubin. J. Med. Chem. 2007, 50, 480–488. [Google Scholar] [CrossRef] [PubMed]
  72. Zheng, Q.; Hua, R.; Jiang, J.; Zhang, L. A general approach to arylated furans, pyrroles, and thiophenes. Tetrahedron 2014, 70, 8252–8256. [Google Scholar] [CrossRef]
  73. Nizami, T.A.; Hua, R. Cycloaddition of 1,3-butadiynes: Efficient synthesis of carbo- and heterocycles. Molecules 2014, 19, 13788–13802. [Google Scholar] [CrossRef]
  74. Sobenina, L.N.; Tomilin, D.N.; Gotsko, M.D.; Ushakov, I.A.; Trofimov, B.A. Transition metal-free cross-coupling of furan ring with haloacetylenes. Tetrahedron 2018, 74, 1565–1570. [Google Scholar] [CrossRef]
  75. Gotsko, M.D.; Sobenina, L.N.; Vashchenko, A.V.; Trofimov, B.A. Synthesis of 2,2-(pyrazol-1-yl)enones via 2:1 coupling of pyrazoles and acylbromoacetylenes in solid alumina. Tetrahedron Lett. 2018, 59, 4231–4235. [Google Scholar] [CrossRef]
  76. Sobenina, L.N.; Sagitova, E.F.; Ushakov, I.A.; Trofimov, B.A. Transition-metal-free synthesis of pyrrolo[1,2-a]pyrazines via intramolecular cyclization of N-propargyl(pyrrolyl)enaminones. Synthesis 2017, 49, 4065–4081. [Google Scholar] [CrossRef] [Green Version]
  77. Sagitova, E.F.; Sobenina, L.N.; Trofimov, B.A. From acylethynylpyrroles to pyrrolo[1,2-a]pyrazines in one step. Russ. J. Org. Chem. 2020, 56, 225–233. [Google Scholar] [CrossRef]
  78. Sobenina, L.N.; Sagitova, E.F.; Markova, M.V.; Ushakov, I.A.; Ivanov, A.V.; Trofimov, B.A. Acylethynylpyrroles as a platform for the one-pot access to 2-(pyrrol-2-yl)-3-acylpyridines via the dihydrogenative annelation with propargylamine. Tetrahedron Lett. 2018, 59, 4047–4949. [Google Scholar] [CrossRef]
  79. Sagitova, E.F.; Sobenina, L.N.; Tomilin, D.N.; Markova, M.V.; Ushakov, I.A.; Trofimov, B.A. Formation of 2-(3-acyl-5-halopyridin-2-yl)pyrroles during annulation/aromatization of NH-2-acylethynylpyrroles with propargylamine in the presence of copper(I) halide. Mendeleev Commun. 2019, 29, 252–253. [Google Scholar] [CrossRef]
  80. Sobenina, L.N.; Tomilin, D.N.; Sagitova, E.F.; Ushakov, I.A.; Trofimov, B.A. Metal-free, atom- and step-economic synthesis of aminoketopyrrolizines from benzylamine, acylethynylpyrroles and acylacetylenes. Org. Lett. 2017, 19, 1586–1589. [Google Scholar] [CrossRef] [PubMed]
  81. Vasilevsky, S.F.; Davydova, M.P.; Tomilin, D.N.; Sobenina, L.N.; Mamatuyk, V.I.; Pleshkova, N.V. Peculiarities of the cascade cleavage of the polarized C-C-fragment in α-ketoacetylenes on reaction with ethylene diamine. ARKIVOC 2014, 5, 132–144. [Google Scholar]
  82. Sobenina, L.N.; Tomilin, D.N.; Gotsko, M.D.; Ushakov, I.A.; Mikhaleva, A.I.; Trofimov, B.A. From 4,5,6,7-tetrahydroindoles to 3- or 5-(4,5,6,7-tetrahydroindol-2-yl)isoxazoles in two steps: A regioselective switch between 3- and 5-isomers. Tetrahedron 2014, 70, 5168–5174. [Google Scholar] [CrossRef]
  83. 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 Pigments 2018, 152, 14–18. [Google Scholar] [CrossRef]
  84. Gotsko, M.D.; Saliy, I.V.; Sobenina, L.N.; Ushakov, I.A.; Trofimov, B.A. From acylethynylpyrroles to pyrrole-pyrone ensembles in a one step. Tetrahedron Lett. 2019, 60, 151126. [Google Scholar] [CrossRef]
  85. Tomilin, D.N.; Gotsko, M.D.; Sobenina, L.N.; Ushakov, I.A.; Trofimov, B.A. A hydrogen pump: Transfer of four hydrogens from a cyclohexane ring to a triple bond during a menthofuran/bromoacetylene adduct rearrangement. Tetrahedron Lett. 2019, 60, 1864–1867. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are available from the authors.
Scheme 1. Synthesis of 3-(pyrrol-2-yl)propiolaldehydes 2.
Scheme 1. Synthesis of 3-(pyrrol-2-yl)propiolaldehydes 2.
Molecules 25 02490 sch001
Scheme 2. Reaction of pyrroles 1 with iodopropiolaldehyde in solid Al2O3.
Scheme 2. Reaction of pyrroles 1 with iodopropiolaldehyde in solid Al2O3.
Molecules 25 02490 sch002
Scheme 3. Reaction of pyrroles1 with iodopropiolaldehyde in solid Al2O3.
Scheme 3. Reaction of pyrroles1 with iodopropiolaldehyde in solid Al2O3.
Molecules 25 02490 sch003
Scheme 4. Reaction of 2-(furan-2-yl)pyrroles 4 with acylbromoacetylenes 6ac in the solid Al2O3.
Scheme 4. Reaction of 2-(furan-2-yl)pyrroles 4 with acylbromoacetylenes 6ac in the solid Al2O3.
Molecules 25 02490 sch004
Scheme 5. Reaction of 2-(thiophen-2-yl)pyrroles 5 with acylbromoacetylenes 6ac in the solid Al2O3.
Scheme 5. Reaction of 2-(thiophen-2-yl)pyrroles 5 with acylbromoacetylenes 6ac in the solid Al2O3.
Molecules 25 02490 sch005
Scheme 6. Formation of (E)-3-bromo-1-(pyrrol-2-yl)prop-2-en-1-ones 10.
Scheme 6. Formation of (E)-3-bromo-1-(pyrrol-2-yl)prop-2-en-1-ones 10.
Molecules 25 02490 sch006
Scheme 7. Ethynylation of the dipyrromethane 11 in the solid Al2O3.
Scheme 7. Ethynylation of the dipyrromethane 11 in the solid Al2O3.
Molecules 25 02490 sch007
Scheme 8. Ethynylation of the dipyrromethane 13 in the solid Al2O3.
Scheme 8. Ethynylation of the dipyrromethane 13 in the solid Al2O3.
Molecules 25 02490 sch008
Scheme 9. Synthesis and ethynylation of dipyrromethanes 17, 18.
Scheme 9. Synthesis and ethynylation of dipyrromethanes 17, 18.
Molecules 25 02490 sch009
Scheme 10. Cross-coupling of tetrahydropyrrolo[3,2-c]pyridines 20 with acylbromoacetylenes 6a,b in the solid K2CO3.
Scheme 10. Cross-coupling of tetrahydropyrrolo[3,2-c]pyridines 20 with acylbromoacetylenes 6a,b in the solid K2CO3.
Molecules 25 02490 sch010
Scheme 11. Cross-coupling of pyrrolo[3.2-c]pyridine 20 with benzoylbromoacetylene 6a in the solid Al2O3.
Scheme 11. Cross-coupling of pyrrolo[3.2-c]pyridine 20 with benzoylbromoacetylene 6a in the solid Al2O3.
Molecules 25 02490 sch011
Scheme 12. Synthesis of pyrrole-2-carbaldehydes with the electron-deficient acetylenic substituents.
Scheme 12. Synthesis of pyrrole-2-carbaldehydes with the electron-deficient acetylenic substituents.
Molecules 25 02490 sch012
Scheme 13. The cross-coupling of NH-pyrroles 27 with acetylbromoacetylenes 28 and 30 in the solid Al2O3.
Scheme 13. The cross-coupling of NH-pyrroles 27 with acetylbromoacetylenes 28 and 30 in the solid Al2O3.
Molecules 25 02490 sch013
Scheme 14. The cross-coupling of N-vinylpyrroles 32 with bromotrifluoroacetylacetylene 28 in the solid Al2O3.
Scheme 14. The cross-coupling of N-vinylpyrroles 32 with bromotrifluoroacetylacetylene 28 in the solid Al2O3.
Molecules 25 02490 sch014
Scheme 15. Proposed mechanism of formation of E-isomers of the hydrogen-bonded compounds 29.
Scheme 15. Proposed mechanism of formation of E-isomers of the hydrogen-bonded compounds 29.
Molecules 25 02490 sch015
Scheme 16. Formation of products 33 from N-vinylpyrroles 32 and bromotrifluoroacetylacetylene 28.
Scheme 16. Formation of products 33 from N-vinylpyrroles 32 and bromotrifluoroacetylacetylene 28.
Molecules 25 02490 sch016
Scheme 17. Detrifluoroacylation of compound 33a after 7 days contact with Al2O3.
Scheme 17. Detrifluoroacylation of compound 33a after 7 days contact with Al2O3.
Molecules 25 02490 sch017
Scheme 18. Synthesis of 2-(pyrrol-2-yl)ethynylphosphonates 37.
Scheme 18. Synthesis of 2-(pyrrol-2-yl)ethynylphosphonates 37.
Molecules 25 02490 sch018
Figure 1. Side products of cross-coupling of pyrroles 35 with chloroethynylphosphonates 36.
Figure 1. Side products of cross-coupling of pyrroles 35 with chloroethynylphosphonates 36.
Molecules 25 02490 g001
Scheme 19. Cross-coupling of electron-deficient halobutadiynes with pyrroles 40.
Scheme 19. Cross-coupling of electron-deficient halobutadiynes with pyrroles 40.
Molecules 25 02490 sch019
Scheme 20. The reaction of polyynes 42ac with N-methyl-4,5,6,7-tetrahydroindole.
Scheme 20. The reaction of polyynes 42ac with N-methyl-4,5,6,7-tetrahydroindole.
Molecules 25 02490 sch020
Scheme 21. Synthesis of tetradiynyl- and hexatriynyl-substituted tetrahydroindoles.
Scheme 21. Synthesis of tetradiynyl- and hexatriynyl-substituted tetrahydroindoles.
Molecules 25 02490 sch021
Scheme 22. Synthesis of octatetraynyl-substituted tetrahydroindoles.
Scheme 22. Synthesis of octatetraynyl-substituted tetrahydroindoles.
Molecules 25 02490 sch022
Scheme 23. Proposed mechanism of long-chain stabilization of a radical intermediate product.
Scheme 23. Proposed mechanism of long-chain stabilization of a radical intermediate product.
Molecules 25 02490 sch023
Scheme 24. Reaction of benzoylbromoacetylene 6a with menthofuran 44 in the solid Al2O3.
Scheme 24. Reaction of benzoylbromoacetylene 6a with menthofuran 44 in the solid Al2O3.
Molecules 25 02490 sch024
Scheme 25. Reaction of haloacetylenes 6af with menthofuran 44 in thesolid Al2O3.
Scheme 25. Reaction of haloacetylenes 6af with menthofuran 44 in thesolid Al2O3.
Molecules 25 02490 sch025
Scheme 26. Possible reaction pathway.
Scheme 26. Possible reaction pathway.
Molecules 25 02490 sch026
Scheme 27. Reaction of pyrazole with acylbromoacetylenes 6ac.
Scheme 27. Reaction of pyrazole with acylbromoacetylenes 6ac.
Molecules 25 02490 sch027
Scheme 28. Reaction of 3,5-dimethylpyrazole with acylbromoacetylenes 6ac, e.
Scheme 28. Reaction of 3,5-dimethylpyrazole with acylbromoacetylenes 6ac, e.
Molecules 25 02490 sch028
Scheme 29. Proposed mechanism of dipyrazolylenones 48 formation.
Scheme 29. Proposed mechanism of dipyrazolylenones 48 formation.
Molecules 25 02490 sch029
Scheme 30. Synthesis of pyrrolo[1,2-a]pyrazines 53a,b from acylethynylpyrroles 52 and propargylamine.
Scheme 30. Synthesis of pyrrolo[1,2-a]pyrazines 53a,b from acylethynylpyrroles 52 and propargylamine.
Molecules 25 02490 sch030
Scheme 31. The formation of pyrrolyl pyridines 56 from acylethynylpyrroles and propargylamine.
Scheme 31. The formation of pyrrolyl pyridines 56 from acylethynylpyrroles and propargylamine.
Molecules 25 02490 sch031
Scheme 32. Formation of pyrrolyl pyridines 56 from aminoenones 57.
Scheme 32. Formation of pyrrolyl pyridines 56 from aminoenones 57.
Molecules 25 02490 sch032
Scheme 33. Synthesis of pyrrolyl pyridines 60 and 61 from NH-2-acylethynylpyrroles and propargylamine.
Scheme 33. Synthesis of pyrrolyl pyridines 60 and 61 from NH-2-acylethynylpyrroles and propargylamine.
Molecules 25 02490 sch033
Scheme 34. Proposed scheme of halopyridines 60 formation.
Scheme 34. Proposed scheme of halopyridines 60 formation.
Molecules 25 02490 sch034
Scheme 35. Synthesis of 1-benzylamino-2-acyl-3-methylenoacylpyrrolizines 65.
Scheme 35. Synthesis of 1-benzylamino-2-acyl-3-methylenoacylpyrrolizines 65.
Molecules 25 02490 sch035
Scheme 36. Reaction of 2-benzoylethynylpyrroles 55a,b with ethylenediamine.
Scheme 36. Reaction of 2-benzoylethynylpyrroles 55a,b with ethylenediamine.
Molecules 25 02490 sch036
Scheme 37. The formation of tetrahydroindolyl dihydrodiazepine 68b.
Scheme 37. The formation of tetrahydroindolyl dihydrodiazepine 68b.
Molecules 25 02490 sch037
Scheme 38. Synthesis of 4,5,6,7-tetrahydropyrrolo[3,2-c]pyridine-pyrazole ensembles 69.
Scheme 38. Synthesis of 4,5,6,7-tetrahydropyrrolo[3,2-c]pyridine-pyrazole ensembles 69.
Molecules 25 02490 sch038
Scheme 39. Synthesis of dipyrromethane-pyrazole ensemble 70.
Scheme 39. Synthesis of dipyrromethane-pyrazole ensemble 70.
Molecules 25 02490 sch039
Scheme 40. Reaction of acylethynyltetrahydroindoles 71 with hydroxylamine.
Scheme 40. Reaction of acylethynyltetrahydroindoles 71 with hydroxylamine.
Molecules 25 02490 sch040
Scheme 41. The formation of 3-(4,5,6,7-tetrahydroindol-2-yl)-4,5-dihydroisoxazol-5-ols 72.
Scheme 41. The formation of 3-(4,5,6,7-tetrahydroindol-2-yl)-4,5-dihydroisoxazol-5-ols 72.
Molecules 25 02490 sch041
Scheme 42. Dehydration of 3-(4,5,6,7-tetrahydroindol-2-yl)-4,5-dihydroisoxazol-5-ols 72.
Scheme 42. Dehydration of 3-(4,5,6,7-tetrahydroindol-2-yl)-4,5-dihydroisoxazol-5-ols 72.
Molecules 25 02490 sch042
Scheme 43. Synthesis of (3-phenylisoxazol-5-yl)dipyrromethanes 75a,b from ethynyldipyrromethanes 12, 78.
Scheme 43. Synthesis of (3-phenylisoxazol-5-yl)dipyrromethanes 75a,b from ethynyldipyrromethanes 12, 78.
Molecules 25 02490 sch043
Scheme 44. Synthesis of (5-phenylisoxazol-3-yl)dipyrromethanes 77a,b from ethynyldipyrromethanes 12, 78.
Scheme 44. Synthesis of (5-phenylisoxazol-3-yl)dipyrromethanes 77a,b from ethynyldipyrromethanes 12, 78.
Molecules 25 02490 sch044
Scheme 45. Alternative synthesis of (3- or 5-phenylisoxazolyl)dipyrromethanes 75 and 77 by condensation of pyrrolylisoxazoles 81 or 82 and with 2,2,2-trifluoro-1-(pyrrol-2-yl)-1-ethanols 80.
Scheme 45. Alternative synthesis of (3- or 5-phenylisoxazolyl)dipyrromethanes 75 and 77 by condensation of pyrrolylisoxazoles 81 or 82 and with 2,2,2-trifluoro-1-(pyrrol-2-yl)-1-ethanols 80.
Molecules 25 02490 sch045
Scheme 46. The synthesis of pyrrolyl pyrones 84.
Scheme 46. The synthesis of pyrrolyl pyrones 84.
Molecules 25 02490 sch046
Scheme 47. Scheme of pyrrolyl pyrones 84 formation.
Scheme 47. Scheme of pyrrolyl pyrones 84 formation.
Molecules 25 02490 sch047
Scheme 48. Rearrangement of acylethynylmenthofurans 45 to acylethylbenzofurans 85.
Scheme 48. Rearrangement of acylethynylmenthofurans 45 to acylethylbenzofurans 85.
Molecules 25 02490 sch048
Scheme 49. Proposed mechanism for the transfer of four hydrogens.
Scheme 49. Proposed mechanism for the transfer of four hydrogens.
Molecules 25 02490 sch049

Share and Cite

MDPI and ACS Style

Sobenina, L.N.; Trofimov, B.A. Recent Strides in the Transition Metal-Free Cross-Coupling of Haloacetylenes with Electron-Rich Heterocycles in Solid Media. Molecules 2020, 25, 2490. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25112490

AMA Style

Sobenina LN, Trofimov BA. Recent Strides in the Transition Metal-Free Cross-Coupling of Haloacetylenes with Electron-Rich Heterocycles in Solid Media. Molecules. 2020; 25(11):2490. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25112490

Chicago/Turabian Style

Sobenina, Lyubov’ N., and Boris A. Trofimov. 2020. "Recent Strides in the Transition Metal-Free Cross-Coupling of Haloacetylenes with Electron-Rich Heterocycles in Solid Media" Molecules 25, no. 11: 2490. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25112490

Article Metrics

Back to TopTop