Next Article in Journal
Synthesis, Spectroscopy, Electrochemistry and DFT of Electron-Rich Ferrocenylsubphthalocyanines
Next Article in Special Issue
Site-Selective Modification of a Porpholactone—Selective Synthesis of 12,13- and 17,18-Dihydroporpholactones
Previous Article in Journal
Synthesis, Molecular Recognition Study and Liquid Membrane-Based Applications of Highly Lipophilic Enantiopure Acridino-Crown Ethers
Previous Article in Special Issue
Rational Design, Synthesis, Characterization and Evaluation of Iodinated 4,4′-Bipyridines as New Transthyretin Fibrillogenesis Inhibitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalyst-Solvent System for PASE Approach to Hydroxyquinolinone-Substituted Chromeno[2,3-b]pyridines Its Quantum Chemical Study and Investigation of Reaction Mechanism

by
Fedor V. Ryzhkov
1,
Yuliya E. Ryzhkova
1,
Michail N. Elinson
1,*,
Stepan V. Vorobyev
2,
Artem N. Fakhrutdinov
1,
Anatoly N. Vereshchagin
1 and
Mikhail P. Egorov
1
1
N. D. Zelinsky Institute of Organic Chemistry Russian Academy of Sciences, Leninsky pr. 47, 119991 Moscow, Russia
2
Department of Organic Chemistry and Petroleum Chemistry, Gubkin Russian State University of Oil and Gas, 65 Leninsky Prospect, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Submission received: 4 May 2020 / Revised: 26 May 2020 / Accepted: 29 May 2020 / Published: 31 May 2020

Abstract

:
The Pot, Atom, and Step Economy (PASE) approach is based on the Pot economy principle and unites it with the Atom and Step Economy strategies; it ensures high efficiency, simplicity and low waste formation. The PASE approach is widely used in multicomponent chemistry. This approach was adopted for the synthesis of previously unknown hydroxyquinolinone substituted chromeno[2,3-b]pyridines via reaction of salicylaldehydes, malononitrile dimer and hydroxyquinolinone. It was shown that an ethanol-pyridine combination is more beneficial than other inorganic or organic catalysts. Quantum chemical studies showed that chromeno[2,3-b]pyridines has potential for corrosion inhibition. Real time 1H NMR monitoring was used for the investigation of reaction mechanism and 2-((2H-chromen-3-yl)methylene)malononitrile was defined as a key intermediate in the reaction.

Graphical Abstract

1. Introduction

Multicomponent reactions (MCRs) employ three or more reactants to obtain heterocycles containing structures of all starting materials in a one-pot process under fixed reaction conditions [1,2,3,4]. It provides powerful productivity to satisfy modern green chemistry requirements, but new synthetic strategies with robust efficiency are still demanded. In this connection, the PASE approach has recently emerged [5,6], it is based on the Pot economy principle and unites it with the Atom and Step Economy strategies (PASE), thereby ensuring high efficiency, simplicity and low waste formation [7,8,9,10]. Nowadays, it is emerging as a fast-paced research front of organic chemistry [11,12,13,14,15,16].
Corrosion is a gradual destruction of refined metal by means of reactions with environment. Today it causes heavy losses to the economy [17,18,19]. Due to electronic configuration of heterocyclic compounds, they are capable of corrosion inhibition or coating metals. The development of PASE approaches to anti-corrosion heterocycles is beneficial for protection of metals.
Chromeno [2,3-b]pyridines are heterocycles with special and useful electronic configuration. Thus, they showed gastric antisecretory activity [20], inhibit mitogen-activated protein kinase 2 and suppress expression of tumor necrosis factor alpha (TNFα) in U937 cells [21].
Similar compounds were earlier synthesized via microwave irradiation of aldehydes, malononitrile dimer and kojic acid in EtOH [22]. The synthesis was carried out under reflux conditions in the presence of Et3N as a catalyst (73–90% yields). Bis-chromeno[2,3-b]pyridine derivatives were obtained via reaction of bis-aldehydes, malononitrile dimer and dimedone under reflux conditions. The reaction was carried out for 5 h in EtOH with a large amount of piperidine [23]. Phenyl substituted derivatives were obtained by reaction of benzaldehydes, malononitrile dimer and naphthols (12%–62% yields). The reaction was carried out under reflux conditions in H2O-EtOH with an equivalent amount of Et2NH [24] and in solvent-free conditions with guanidine hydrochloride catalysis (100 °C, 2 h) [25]. Solvent-free conditions were also used for the transformation of benzaldehydes, malononitrile dimer and coumarin [26]. Carbazole and indole derivatives were synthesized (67–85%) under reflux conditions and microwave irradiation: by reaction of aldehydes, malononitrile dimer and hydroxycarbazole or hydroxyindole in anhydrous EtOH with EtONa [27].
We have already described multicomponent syntheses of chromeno[2,3-b]pyridines [28,29,30,31,32,33,34]. However, there are no reports on a PASE synthesis of hydroxyquinolinone substituted chromeno[2,3-b]pyridines. Therefore, our attention was devoted toward reactions of salicylaldehydes, malononitrile dimer and hydroxyquinolinone. We were interested in developing the PASE method, clarifying reaction mechanism and defining the actual intermediates of the reaction. We were also prompted to carry out quantum chemical studies and estimate chromeno[2,3-b]pyridine potential for corrosion inhibition.

2. Results and Discussion

2.1. PASE Approach

Initially, to examine the reaction of salicylaldehyde 1a, malononitrile dimer 2 and hydroxyquinolinone 3, we have carried out multicomponent synthesis of chromeno[2,3-b]pyridine 4a under solvent-free conditions (Scheme 1, Table 1). Previously, AcONa and KF have successfully catalyzed similar processes under solvent-free conditions [32]. AcONa showed better results (Entry 3) than KF (Entry 2) and provided twice the yield as the reaction without catalyst (Entry 1). ‘On-water’ [33,34,35] reactions showed almost the same results as solvent-free experiments (Entries 4,5).
Further, we have examined the multicomponent reaction in alcohol media. The reaction in EtOH without catalyst resulted in assembling of chromeno[2,3-b]pyridine 4a in 19% yield (Entry 6), in the presence of AcONa it was formed in 35% yield (Entry 7) whereas organocatalysis by Et3N and morpholine resulted in 63% and 57% yields (Entries 8 and 9).
Pyridine (Py) has been already employed as a solvent and catalyst for similar processes [36]. Refluxing of salicylaldehyde 1a, malononitrile dimer 2 and hydroxyquinone 3 in Py for 2 h resulted in 61% yield (Entry 10). Py is a good solvent for structures of this type, however, a small amount of chromeno[2,3-b]pyridine 4a always remains dissolved in Py.
Then, reaction in a mixture of EtOH and Py was carried out. In the volumetric range of EtOH:Py (2-4:1), the best yield was in EtOH-Py (3:1) and chromeno[2,3-b]pyridine 4a was isolated in excellent 95% yield (Entry 12).
Figure 1 shows that this organic catalyst is more effective than inorganic KF or AcONa catalysts and the ‘EtOH-Py’ system is more beneficial than ‘solvent-free’ and ‘on-water’ approaches to chromeno[2,3-b]pyridine 4a. Py in EtOH keeps the intermediates dissolved and ensures optimal basicity [37] supporting the reaction.
When the reaction in EtOH:Py mixture was finished, the final compound was directly crystallized in pure form after cooling. Under the optimal conditions (Entry12: refluxing for 2 h in 4 mL of EtOH-Py (3:1) mixture) multicomponent reactions of salicylaldehydes 1ai, malononitrile dimer 2 and hydroxyquinolinone 3 were carried out. Chromeno[2,3-b]pyridines 4ai were obtained in 59–98% yields (Table 2). In general, substitution reduced yields of chromeno[2,3-b]pyridines 4bi. Presumably, electronic effects were taking place: electron-donating methyl, methoxy and ethoxy groups tended to support higher yields than bromo and nitro groups. The yield of chlorine substituted compound was bigger than in the case of bromo substituted compound due to a mesomeric effect.
Considering the mechanism of this multicomponent reaction, several path-ways for the reaction were possible. Salicylaldehyde 1 may have undergone condensation either with malononitrile dimer 2 or hydroxyquinolinone 3 (Scheme 2). Hence, four main pathways of the reaction could be discussed. Path 1 started from Knoevenagel condensation of salicylaldehyde 1 and malononitrile dimer 2, then cyclization into intermediate 5 proceeded. The cyclization was followed by Michael addition of hydroxyquinolinone 3 and another cyclization. Path 2 [38,39] started from Knoevenagel condensation of salicylaldehyde 1 and malononitrile dimer 2. Further Michael addition of hydroxyquinolinone 3 to intermediate 7 took place, and subsequent double cyclization of the malononitrile fragment formed chromeno[2,3-b]pyridine 4. Path 3 began with Knoevenagel condensation of salicylaldehyde 1 and hydroxyquinolinone 3. Formation of chromeno[2,3-b]pyridine 4 proceeded via intermediate 8. In Path 4, salicylaldehyde 1 undergoes condensation with malononitrile dimer 2, and then double cyclization took place to form intermediate 10. Further, addition of hydroxyquinolinone 3 formed chromeno[2,3-b]pyridine 4.
In order to gain insight into the reaction mechanism, we have performed 1H NMR monitoring, to obtain constituent data on the reaction. To reduce the influence of sample preparation, the transformation of starting materials into chromeno[2,3-b]pyridine 4a was carried out and monitored directly into a spectrometer without catalyst to slow down the reaction.
During the NMR study, six major components were recorded: salicylaldehyde 1a, malononitrile dimer 2 and hydroxyquinolinone 3, intermediate 5, intermediate 7 and chromeno[2,3-b]pyridine 4a. A representative 1H NMR spectrum with assignment of peaks showed on Figure 2.
The starting materials are well known (we estimated the presence of salicylaldehyde 1a by the singlet at 10.25 ppm, malononitrile dimer 2 by the singlet at 3.85 ppm and hydroxyquinolinone by the singlet at 5.78 ppm); compound 4a is described in this manuscript and was estimated by the signal at 5.61 ppm; intermediate 5 was described previously (7.75 ppm) [40]; intermediate 7 was estimated by the signals at 8.04–8.11 ppm [41].
Figure 3 presents a series of 1H NMR spectra for a conversion of starting materials into the final compound 4a in DMSO-d6. It shows the disappearance of malononitrile dimer 2 (in 20 min to 2 h from the beginning of experiment) and the simultaneous formation of intermediate 7 and intermediate 5. Intermediate 7 formed by Knoevenagel condensation and it was followed by cyclization into intermediate 5 (Scheme 2). Thus, intermediate 7 was forming and rapidly converting into intermediate 5 (1–3 h of experiment). Eventually the signals of intermediate 7 disappeared.
As shown in Figure 3, when intermediate 7 disappeared (Figure 3), hydroxyquinolinone 3 signals disappeared and chromeno[2,3-b]pyridine 4a signals appeared. Thus, hydroxyquinolinone 3 took part in Michael addition to the electron-deficient intermediate 5 to form final compound 4a.
Based on these data and taking into consideration earlier published results [42,43], we suggest that the first stage was a rapid formation of intermediate 7 with expulsion of a hydroxide anion [37]. This hydroxide anion instantly catalyzed a rapid cyclization of intermediate 7 into intermediate 5. Then, subsequent Michael addition and cyclization formed the chromeno[2,3-b]pyridine 4a.
To prove this conclusion, the intermediate 5 was isolated in a one-step reaction. Further addition of hydroxyquinolinone 3 to the obtained intermediate 5 in a distinct reactor afforded chromeno[2,3-b]pyridine 4a. The reaction conditions were the same as developed for multicomponent process and the yield of chromeno[2,3-b]pyridine 4a was 95% again. More than that, the two-component reaction of salicylaldehyde 1a and hydroxyquinolinone 3 in a separate reactor did not form intermediate 9. This experimental data and 1NMR monitoring data were consistent with each other. Thus, the results from both intermediate synthesis and 1NMR monitoring turned out to be consistent with each other.
Moreover, both 1NMR monitoring and experimental data (Table 1 and Table 2) revealed no signals of intermediate 9 [44]: 7.11–7.16 (m, 1H), 7.36 (s, 1H), 7.54–7.58 (m, 1H), 7.67–7.70 (d, 1H, J = 8.4 Hz), 7.93–7.99 (m, 4H), 8.45–8.48 (d, 1H, J = 8.4 Hz), 10.22 (s, 1H), 11.19 (s, 1H) ppm.
To sum up, we postulated Path 1 for the assembling of chromeno[2,3-b]pyridine 4a in this multicomponent process. The proposed mechanism for the developed reaction is outlined in Scheme 3.

2.2. Quantum Chemical Investigation

2.2.1. Quantum Chemistry Approach

One of the most well-known corrosion inhibitors are heterocycles, such as azoles [45,46], pyridines [47], several bioorganic compounds [48] and chromenes [49]. The mechanism of their action is the adsorption on the metal’s surface (whatever metal is used–copper, steel, brass, etc.). Generally, the greater the number of electron-donating groups in molecule, the higher the probability of adsorption and inhibition activity. As it was shown previously, the inhibitory activity of organic compounds can be predicted properly by quantum-chemical methods [50,51,52]. This reduces the time of the study and hazardous reactant wasting by selecting the hit-compounds. They must have appropriate electronic properties, which can be estimated by quantum chemistry calculations, to interact with metal’s surface.
To estimate electronic properties, the next method can be used [53]: according to Koopman’s theorem, the energy of the highest occupied orbital (EHOMO) is equal to the ionization potential (I) having the opposite charge, the energy of the lowest unoccupied orbital (ELUMO) is related to the electron affinity (A).
Considering the quantum chemical parameters of the inhibitor, the higher the energy of the highest occupied orbital (HOMO), the higher the donation of the inhibitor to metal’s vacant d-orbital; the lower by the energy of the lowest unoccupied orbital (LUMO), the greater is the electron acception from the metal to inhibitor. The electron release is characterized by energy difference (ΔE(L-H)). The electron release is easier when the ΔE(L-H) is lower, and then the adsorption of inhibitor is stronger [54]. The electronic density distribution is an important parameter as well, and it can be estimated by a frontier orbital localization study and partial atomic charge distribution [55].
In addition to HOMO and LUMO, several significant parameters were calculated as well. Electronegativity shows the ability of the molecule to attract electrons towards itself and therefore a molecule with the lowest value tends to has the highest ability to donate electrons.
The global chemical hardness is related to the resistance of a molecule to charge transfer [56]. The global chemical softness (σ) is inversely proportional from the chemical hardness. The higher softness, the better adsorption [57]. Chemical hardness (η) as well as electronegativity (χ) can be calculated from ionization potential and electron affinity, and then chemical softness can be calculated from chemical hardness. Since the molecule of inhibitor must donate its electron to vacant metal orbitals, its nucleophilicity should be high, opposite to electrophilicity and hence to the global electrophilicity index (ω) [58].
All quantum chemical calculations were performed using the Gaussian09 program package [59]. The structures 4ai were optimized and the required parameters were calculated using the density functional theory method B3LYP [60] with the 6-311G(d,p) basic set. This method is recommended for calculation of frontier orbitals energies and related values, such as electronegativity, polarizability etc. [61]. Our study was carried out for molecules in gas phase as well as for solvated ones in water (PCM model) [62].
The results of quantum chemical calculations for gas phase, solvated forms and protonated forms are presented in Table 3, frontier orbitals of several studied compounds are shown in Table 4.

2.2.2. Gas Phase Calculations

During quantum chemical calculations, the structures of the studied compounds were optimized as shown in Figure 4 for 4a. The chromeno[2,3-b]pyridine fragment was not planar and hydroxyquinolinone ring was localized nearly perpendicular to chromeno[2,3-b]pyridine ring. Calculated key parameters of the studied compounds 4ai are shown in Table 3.
As mentioned above, corrosion inhibition depends on the energies and distributions of frontier orbitals. For most cases in considered compounds, the HOMOs are localized on the benzene ring atoms of chromeno[2,3-b]pyridines, making them sites for interaction with metal cations that formed by dissolving the metal in acid. The LUMOs are localized in the hydroxyquinolinone ring of the substituent. Such a type of distribution of the frontier orbitals is represented in compounds 4a, 4b, 4d and 4i. The frontier orbitals of several compounds are shown in Table 4.
Introducing of a strong electron-withdrawing substituent, such as a nitro-group, tended to change the orbital distribution. The HOMO became localized on the atoms of the pyridine ring, and the LUMO on the nitro-group of the substituent (4h, Table 4). A similar distribution was observed for the compounds 4e, 4f and 4g: the HOMOs were localized in the pyridine ring, while the LUMOs were in the hydroxyquinolinone ring (Table 4). This was caused by the withdrawing effect of the halogen atom. It is noteworthy that compound 4c had the same distribution of the orbitals, which was probably related to poor overlapping of oxygen orbital of methoxy-group with ones of benzene ring.
The energy difference is another important parameter. The lower the energy difference of the molecule, the higher the reaction ability (due to easier electron release). For the studied compounds these values were comparable, but 4b, 4e and 4i had the lowest ones. The global electronegativity χ values for 4b, 4e and 4i were almost twice as low as values calculated for similar tricyclic cationic inhibitors (<7.2 is accepted, [53]). The global electrophilicity index ω was low as well. It was almost five times lower than the values calculated for the studied inhibitors (<19.3 is accepted, [53]).

2.2.3. Solvated Form Calculations

Several parameters changed by taking into consideration the solvation processes. Thus, the frontier orbitals in compounds 4a, 4b, 4d, 4f and 4i were located in the same manner as in non-solvated ones, but for 4c, 4e, 4g and 4h shifting of the HOMOs to the carbonyl atoms of hydroxyquinolinone ring is observed, while the LUMOs remained in the same place (Table 4).
The analysis of the basic parameters (Table 3) shows that assuming solvation led to changing of some values mostly for compound 4h; we can expect the highest anticorrosive effect from 4b, 4c and 4d due to their low values of general electrophilicity ω, which can be predicted because of electron-donating substituents. The values of electronegativity χ for 4b, 4c and 4d remained low, almost the same as was calculated for the gas phase. It is noteworthy that the value of the global softness σ for these compounds tended to decrease while taking to consideration the solvation effects.

2.2.4. Protonated form Calculations

In the case of a corrosion inhibition study in high acidic media, we must take in consideration the possibility of protonation of our compounds because of the basicity of some functional groups.
As it was shown previously [63], the main center of basicity in aminopyridines is the nitrogen atom in heterocyclic ring. We decided to confirm this statement by calculating the protonation energy for the compound 4a (Scheme 4).
Proton affinity (PA) is Equation:
PA = EPyH + EH2OEPyEH3O+
where Ei–the total energy of the corresponding compound
We found that for the compound 4a protonation took place on the pyridine ring nitrogen atom, while it was preferable by 15.1 kcal/mol in comparison to the α-amino-group protonation and by 14.2 kcal/mol to the γ-amino-group.
The calculation results confirmed that the protonation took place first on the pyridine ring, so we have not carried out such calculations for the other compounds in order to reduce computation time, and optimized their geometry assuming protonation of the first pyridine position.
The calculation for the protonated molecules shows the great changing of electron density distribution in them. As the nitrogen atom in chromeno[2,3-b]pyridine cycle became positively charged and revealed high electron-withdrawing effect, the electron density in the pyridine ring decreased. This process was also strengthened by the influence of the cyano-group in this ring. In all nine compounds the LUMOs were localized on the atoms of pyridine ring, while the HOMOs were localized either on the benzene ring atoms (compounds 4a, 4b, 4d, 4i, Table 4) or on the carbonyl group of the hydroxyquinolinone ring (compounds 4c, 4e, 4f, 4g, 4h, Table 4).
The key parameters remained low for compounds 4ai even in the protonated state. Global electrophilicity ω was extremely low (<60 is acceptable, [53]), the softness σ and electronegativity χ were almost twice as low as the values for known cationic dies (<1 and <10 are acceptable, respectively [53]).
The analysis of the key parameters for the protonated molecules of 4ai shows that the highest anticorrosion effect was possessed by the compounds 4b, 4c and 4d, the lowest by 4f and 4h. However, despite of the presence of an electron-withdrawing substituent and the high value of global electrophilicity ω, compound 4h had the highest value of chemical softness σ, which indicates that this compound can readily react with forming Fe2+ ions (soft acid with soft base).

3. Materials and Methods

3.1. Synthesis

3.1.1. General Information

All melting points were measured with a Gallenkamp melting point apparatus (London, UK). 1H and 13C NMR spectra were recorded with Bruker AM−300 spectrometer (Billerica, MA, USA) at ambient temperature. Chemical shifts values were relative to Me4Si. IR spectra were registered with a Bruker ALPHA-T FT-IR spectrometer (Billerica, MA, USA) in KBr pellets. Mass spectra (EI, 70 eV) were obtained directly with a Finnigan MAT INCOS 50 spectrometer (Bremen, Germany). High-resolution mass spectra (HRMS) were measured on a Bruker micrOTOF II instrument (Billerica, MA, USA) using electrospray ionization (ESI).

3.1.2. Synthesis of 4ai

Salicylaldehyde 1ai (1 mmol), 2-aminoprop-1-ene-1,1,3-tricarbonitrile 2 (0.13 g, 1 mmol) and 4-hydroxyquinolin-2(1H)-one 3 (0.16 g, 1 mmol) were refluxed in 4 mL of ethanol–pyridine (3:1) mixture for 2 h. After the reaction was completed, the solid was filtered, washed with well-chilled methanol (3 × 2 mL) and dried to isolate pure substituted 2,4-diamino-5-(4-hydroxy-2-oxo-1,2-dihydroquinolin-3-yl)-5H-chromeno[2,3-b]pyridine-3-carbonitriles 4ai.
2,4-Diamino-5-(4-hydroxy-2-oxo-1,2-dihydroquinolin-3-yl)-5H-chromeno[2,3-b]pyri-dine-3-carbonitrile 4a, (white solid, 0.377g, 95%), mp > 350 °C (from Py-EtOH), FTIR (KBr) cm−1: 3391, 3184, 2993, 2846, 2201, 1642, 1606, 1399, 1237, 751. 1H-NMR (400 MHz, DMSO-d6) δ 5.58 (s, 1H, CH), 6.31 (s, 2H, NH2), 6.44 (s, 2H, NH2), 6.90–7.05 (m, 3H, Ar), 7.06–7.22 (m, 2H, Ar), 7.28–7.41 (m, 2H, Ar), 7.86 (d, J = 7.8 Hz, 1H, Ar), 10.77 (br s, 1H, OH), 11.86 (s, 1H, NH) ppm. 13C-NMR (100 MHz, DMSO-d6) δ 28.68, 70.32, 88.97, 115.24, 115.41 (2C), 115.61, 116.54, 121.60, 122.16, 123.39, 127.42 (2C), 128.68, 130.76, 137.49, 151.27, 156.72, 159.19, 159.35, 160.19, 164.18. MS (EI, 70 eV) m/z (%): 397 (M+, 17), 376 (11), 304 (7), 252 (6), 237 (100), 171 (12), 161 (26), 119 (12), 79 (86), 52 (55). HRMS-ESI: [M + H]+, calcd for C22H16N5O3 398.1253, found 398.1252.
2,4-Diamino-5-(4-hydroxy-2-oxo-1,2-dihydroquinolin-3-yl)-7-methyl-5H-chromeno[2,3-b]pyridine-3-carbonitrile 4b, (white solid, 0.358g, 87%), mp > 350 °C (from Py-EtOH), FTIR (KBr) cm−1: 3419, 3361, 2881, 2845, 2203, 1634, 1580, 1402, 1220, 752. 1H-NMR (400 MHz, DMSO-d6) δ 2.14 (s, 3H, CH3), 5.53 (s, 1H, CH), 6.28 (s, 2H, NH2), 6.41 (s, 2H, NH2), 6.78 (s, 1H, Ar), 6.93 (dd, 3J = 16.6 Hz, 4J = 8.0 Hz, 2H, Ar), 7.11 (t, J = 7.4Hz, 1H, Ar), 7.33 (d, J = 7.8 Hz, 1H, Ar), 7.48 (t, J = 7.3 Hz, 1H, Ar), 7.86 (d, J = 7,8 Hz, 1H, Ar), 10.74 (s, 1H, OH), 11.84 (s, 1H, NH) ppm. 13C-NMR (100 MHz, DMSO-d6) δ 20.11, 28.64, 70.20, 89.00, 115.18 (2C), 115.61, 116.58, 121.48, 122.14, 123.15, 127.97 (2C), 128.34, 130.73, 132.22, 137.46, 149.20, 156.68, 159.15,159.49 160.09, 164.18. MS (EI, 70 eV) m/z (%): 411 (M+, 26), 390 (7), 251 (100), 223 (14), 185 (28), 161 (61), 119 (43), 92 (30), 77 (35), 42 (15). HRMS-ESI: [M + H]+, calcd for C23H18N5O3 412.1410, found 412.1402.
2,4-Diamino-5-(4-hydroxy-2-oxo-1,2-dihydroquinolin-3-yl)-8-methoxy-5H-chromeno[2,3-b]pyridine-3-carbonitrile 4c, (white solid, 0.355g, 83%), mp > 350°C (from Py-EtOH), FTIR (KBr) cm−1: 3428, 3381, 2887, 2836, 2199, 1632, 1607, 1569, 1402, 1201, 761. 1H-NMR (400 MHz, DMSO-d6) δ 3.72 (s, 3H, OMe), 5.50 (s, 1H, CH), 6.29 (s, 2H, NH2), 6.43 (s, 2H, NH2), 6.51–6.65 (m, 2H, Ar), 6.88 (d, J = 7.9 Hz, 1H, Ar), 7.11 (t, J = 7.2 Hz, 1H, Ar), 7.33 (d, J = 7.9 Hz, 1H, Ar), 7.48 (t, J = 7.2 Hz, 1H, Ar), 7.86 (d, J = 7.3 Hz, 1H, Ar), 10.71 (br s, 1H, OH), 11.83 (s, 1H, NH) ppm. 13C-NMR (100 MHz, DMSO-d6) δ 28.08, 30.63, 55.21, 70.29, 89.12, 100.60 (2C), 109.69, 115.27, 115.55 (2C), 116.50, 121.43, 122.11, 128.75, 130.64, 137.40, 151.93, 156.68, 158.53, 159.10, 159.97, 164.13. MS (EI, 70 eV) m/z (%): 407 (18), 267 (100), 252 (11), 237 (9), 224 (17), 201 (4), 161 (60), 119 (43), 92 (33), 15 (63). HRMS-ESI: [M + H]+, calcd for C23H18N5O4 428.1359, found 428.1351.
2,4-Diamino-9-ethoxy-5-(4-hydroxy-2-oxo-1,2-dihydroquinolin-3-yl)-5H-chromeno[2,3-b]pyridine-3-carbonitrile 4d, (white solid, 0.433 g, 98%), mp > 350 °C (from Py-EtOH), FTIR (KBr) cm−1: 3409, 3372, 2979, 2881, 2202, 1631, 1568, 1483, 1222, 751. 1H-NMR (400 MHz, DMSO-d6) δ 1.39 (t, J = 6.8 Hz, 3H, OEt), 4.04 (m, 2H, OEt), 5.56 (s, 1H, CH), 6.35 (s, 2H, NH2), 6.41 (s, 2H, NH2), 6.52 (d, J = 6.4 Hz, 1H, Ar), 6.78–6.90 (m, 2H, Ar), 7.10 (t, J = 7.6 Hz, 1H, Ar), 7.33 (d, J = 8.1 Hz, 1H, Ar), 7.48 (t, J = 7.6 Hz, 1H, Ar), 7.88 (d, J = 8.1 Hz, 1H, Ar), 10.80 (s, 1H, OH), 11.83 (s, 1H, NH) ppm. 13C-NMR (100 MHz, DMSO-d6) δ 14.81, 28.75, 63.62, 70.22, 88.77, 110.79, 115.22, 115.57, 116.55, 119.49 (2C), 121.47, 122.10, 122.90, 124.00, 130.71, 137.43, 140.78, 145.90, 156.64, 159.18 (2C), 160.21, 164.12. MS (EI, 70 eV) m/z (%): 441 (M+, 19), 412 (18), 392 (26), 281 (74), 253 (77), 187 (14), 161 (100), 119 (57), 92 (42), 29 (95). HRMS-ESI: [M + H]+, calcd for C24H19N5O4 442.1510, found 442.1503.
2,4-Diamino-7-bromo-5-(4-hydroxy-2-oxo-1,2-dihydroquinolin-3-yl)-9-methoxy-5H-chromeno[2,3-b]pyridine-3-carbonitrile 4e, (white solid, 0.344 g, 68%), mp > 350 °C (from Py-EtOH), FTIR (KBr) cm−1: 3445, 3387, 2204, 1639, 1600, 1567, 1398, 1224, 1013, 768. 1H-NMR (400 MHz, DMSO-d6) δ 3.84 (s, 3H, OMe), 5.53 (s, 1H, CH), 6.33 (s, 2H, NH2), 6.42 (br s, 2H, NH2), 6.64 (s, 1H, Ar), 7.04 (s, 1H, Ar), 7.12 (t, J = 7.3 Hz, 1H, Ar), 7.34 (d, J = 8.1 Hz, 1H, Ar), 7.50 (t, J = 7.3 Hz, 1H, Ar), 7.88 (d, J = 7.2 Hz, 1H, Ar), 10.90 (s, 1H, OH), 11.86 (s, 1H, NH) ppm. 13C-NMR (100 MHz, DMSO-d6) δ 28.64, 56.06, 70.39, 88.34, 113.03, 113.23, 114.08, 115.07, 115.62, 116.37, 121.62, 122.20, 123.27, 125.73, 130.88, 131.78, 137.51, 137.71, 147.67, 156.58, 156.74, 159.13, 163.94. MS (EI, 70 eV) m/z (%): 507 (M+, 3), 346 (51), 331 (8), 305 (51), 252 (14), 204 (45), 161 (100), 119 (59), 92 (39), 15 (28). HRMS-ESI: [M + H]+, calcd for C23H17BrN5O4 506.0464 [79Br], 508.0443 [81Br], found 506.0458 [79Br], 508.0434 [81Br].
2,4-Diamino-7-chloro-5-(4-hydroxy-2-oxo-1,2-dihydroquinolin-3-yl)-5H-chrome-no[2,3-b]pyridine-3-carbonitrile 4f, (white solid, 0.393 g, 91%), mp > 350°C (from Py-EtOH), FTIR (KBr) cm−1: 3393, 3217, 2975, 2894, 2203, 1633, 1607, 1404, 1259, 757. 1H-NMR (400 MHz, DMSO-d6) δ 5.56 (s, 1H, CH), 6.35 (s, 2H, NH2), 6.47 (br s, 2H, NH2), 6.96 (s, 1H, Ar), 7.06 (d, J = 8.7 Hz, 1H, Ar), 7.13 (t, J = 7.6 Hz, 1H, Ar), 7.22 (d, J = 8.6 Hz, 1H, Ar), 7.34 (d, J = 7.9 Hz, 1H, Ar), 7.50 (t, J = 7.6 Hz, 1H, Ar), 7.88 (d, J = 6.0 Hz, 1H, Ar), 10.89 (br s, 1H, OH), 11.88 (s, 1H, NH) ppm. 13C-NMR (100 MHz, DMSO-d6) δ 28.73, 70.43, 88.31, 115.15, 115.37, 115.66, 116.39, 117.27 (2C), 121.51, 122.27, 125.62, 126.68, 127.36, 127.48, 130.90, 137.60, 150.23, 156.65, 156.80, 159.20, 163.98. MS (EI, 70 eV) m/z (%): 431 (M+, 17), 409 (9), 271 (100), 243 (10), 205 (21), 161 (79), 119 (73), 92 (92), 77 (54), 28 (52). HRMS-ESI: [M + H]+, calcd for C22H15ClN5O3 432.0863 [35Cl], 434.0834 [37Cl], found 432.0856 [35Cl], 434.0824 [37Cl].
2,4-Diamino-7-bromo-5-(4-hydroxy-2-oxo-1,2-dihydroquinolin-3-yl)-5H-chrome-no[2,3-b]pyridine-3-carbonitrile 4g, (white solid, 0.324 g, 68%), mp > 231–233 °C (from Py-EtOH), FTIR (KBr) cm−1: 3343, 3184, 2993, 2846, 2201, 1642, 1609, 1398, 1258, 775. 1H-NMR (400 MHz, DMSO-d6) δ 5.56 (s, 1H, CH), 6.35 (s, 2H, NH2), 6.46 (br s, 2H, NH2), 7.01 (d, J = 8.5 Hz, 1H, Ar), 7.08 (s, 1H, Ar), 7.13 (t, J = 7.5 Hz, 1H, Ar), 7.35 (d, J = 7.9 Hz, 2H Ar), 7.50 (t, J = 7.5 Hz, 1H, Ar), 7.88 (d, J = 6.0 Hz, 1H Ar), 10.89 (s, 1H, OH), 11.87 (s, 1H, NH) ppm. 13C-NMR (100 MHz, DMSO-d6) δ 28.61, 70.43, 88.34, 114.47, 115.07, 115.65 (2C), 116.36, 117.72, 121.52, 122.27, 126.09, 130.23, 130.34, 130.90, 137.56, 150.67, 156.67, 158.92, 159.20, 160.32, 163.90. MS (EI, 70 eV) m/z (%): 316 ([M-C9H6NO2]+, 81Br, 2), 314 ([M-C9H6NO2]+, 79Br, 2), 236 (2), 202 (1), 171 (1), 161 (13), 119 (9), 92 (5), 78 (73), 63 (100), 15 (31). HRMS-ESI: [M + H]+, calcd for C22H14BrN5O3 476.0353 [79Br], 478.0333 [81Br], found 476.0324 [79Br], 478.0323 [81Br].
2,4-Diamino-5-(4-hydroxy-2-oxo-1,2-dihydroquinolin-3-yl)-7-nitro-5H-chrome-no[2,3-b]pyridine-3-carbonitrile 4h, (yellow solid, 0.261 g, 59%), mp > 350 °C (from Py-EtOH), FTIR (KBr) cm−1: 3357, 3184, 2199, 1661, 1627, 1336, 1242, 1026, 828, 752. 1H-NMR (400 MHz, DMSO-d6) δ 5.24 (s, 1H, CH), 6.53 (br s, 2H, NH2), 6.56 (s, 2H, NH2), 7.28 (t, J = 7.6 Hz, 1H, Ar), 7.34 (d, J = 8.3 Hz, 1H Ar), 7.42 (d, J = 8.3 Hz, 1H Ar), 7.54 (t, J = 7.6 Hz, 1H, Ar), 7.87–7.94 (m, 1H, Ar), 8.02 (d, J = 8.3 Hz, 1H Ar), 8.08 (dd, 3J = 7.0 Hz, 4J = 2.6 Hz, 1H, Ar), 9.57 (s, 1H, OH), 11.62 (s, 1H, NH) ppm. 13C-NMR (100 MHz, DMSO-d6) δ 28.45, 98.42, 107.18, 112.75, 115.20 (2C), 116.43, 116.77, 121.76 (2C), 123.02, 124.14, 126.22, 130.52, 137.22, 143.17, 152.75, 153.74, 154.75, 155.08, 159.76, 161.57. MS (EI, 70 eV) m/z (%): 294 (11), 247 (10), 236 (3), 190 (3), 161 (20), 150 (100), 122 (65), 78 (16), 63 (31), 18 (32). HRMS-ESI: [M + H]+, calcd for C22H15N6O5 443.1104, found 443.1098.
9,11-Diamino-12-(4-hydroxy-2-oxo-1,2-dihydroquinolin-3-yl)-12H-benzo[5,6]chromeno[2,3-b]pyridine-10-carbonitrile 4i, (white solid, 0.389 g, 87%), mp > 350 °C (from Py-EtOH), FTIR (KBr) cm−1: 3454, 3400, 2875, 2835, 2203, 1634, 1607, 1409, 1239, 755. 1H-NMR (400 MHz, DMSO-d6) δ 6.06 (s, 1H, CH), 6.35 (s, 2H, NH2), 6.69 (s, 2H, NH2), 7.07 (t, J = 7.6 Hz, 1H, Ar), 7.25–7.51 (m, 5H, Ar), 7.76–7.93 (m, 3H, Ar), 8.02 (d, J = 8.4 Hz, 1H Ar), 10.82 (s, 1H, OH), 11.95 (s, 1H, NH) ppm. 13C-NMR (100 MHz, DMSO-d6) δ 26.76, 70.47, 89.08, 114.46, 115.21, 115.58, 116.51, 117.06 (2C), 121.56, 122.13, 122.68, 124.08, 126.83, 128.30, 128.44 (2C), 130.18, 130.78, 131.32, 137.26, 149.24, 156.86, 159.17, 160.96, 163.90. MS (EI, 70 eV) m/z (%): 429 ([M-H2O]+, 5), 410 (14), 364 (19), 287 (100), 258 (6), 221 (13), 161 (35), 144 (38), 92 (12), 63 (13). HRMS-ESI: [M + H]+, calcd for C26H18N5O3 448.1410, found 448.1401.

3.1.3. Isolation of the Intermediate 5

Salicylaldehyde 1a (0.12 g, 1 mmol) and 2-aminoprop-1-ene-1,1,3-tricarbonitrile 2 (0.13 g, 1 mmol) were stirred at room temperature in 4 mL of ethanol–pyridine (3:1) mixture for 2 h. After the reaction was completed, the solid was filtered, washed with well-chilled methanol (3 × 2 mL) and dried to isolate pure 2-(amino-(2-imino-2H-chromen-3-yl)methylene)malononitrile 5.
2-(Amino-(2-imino-2H-chromen-3-yl)methylene)malononitrile 5, (Yellow solid, 0.208 g, yield 88%), m.p. 272–273 °C (decomp.) (lit [40] m.p. 271–272 °C (decomp.)). 1H-NMR (300 MHz, DMSO-d6) δ 7.11–7.31 (m, 2H, 2CH Ar), 7.45–7.63 (m, 2H, 2CH Ar), 7.75 (s, 1H, CH), 8.65 (s, 1H, NH), 8.93 (br.s, 1H, NHH), 8.95 (br.s, 1H, NHH) ppm.

3.1.4. Synthesis of Chromeno[2,3-b]pyridine 4a from 2-(amino-(2-imino-2H-chromen-3-yl)methylene)malononitrile 5 and 4-hydroxyquinolin-2(1H)-one 3

2-(Amino-(2-imino-2H-chromen-3-yl)methylene)malononitrile 5 (0.24 g, 1 mmol) and 4-hydroxyquinolin-2(1H)-one (0.16 g, 1 mmol) 3 were refluxed in 4 mL of ethanol–pyridine (3:1) mixture for 2 h. After the reaction was completed, the solid was filtered, washed with well-chilled methanol (3 × 2 mL) and dried to isolate pure chromeno[2,3-b]pyridine 4a.
2,4-Diamino-5-(4-hydroxy-2-oxo-1,2-dihydroquinolin-3-yl)-5H-chromeno[2,3-b]pyridine-3-car-bonitrile 3a, (White solid, 0.377g, 95%), mp > 350°C (from Py-EtOH), 1H-NMR (400 MHz, DMSO-d6) δ 5.58 (s, 1H, CH), 6.31 (s, 2H, NH2), 6.44 (s, 2H, NH2), 6.90–7.05 (m, 3H, Ar), 7.06–7.22 (m, 2H, Ar), 7.28–7.41 (m, 2H, Ar), 7.86 (d, J = 7.8 Hz, 1H, Ar), 10.77 (br s, 1H, OH), 11.86 (s, 1H, NH) ppm.

4. Conclusions

In summary, the PASE transformation of salicylaldehydes, malononitrile dimer and hydroxyquinolinone into previously unknown hydroxyquinolinone substituted chromeno[2,3-b]pyridines has been found. The proposed EtOH-Py catalyst-solvent system provides an efficient rout to hydroxyquinolinone substituted chromeno[2,3-b]pyridines.
The developed approach is facile, it is easy to isolate final compounds directly from reaction mixture and the yields of final compounds are 59%–98%.
During investigation of reaction mechanism using a real-time 1H NMR monitoring, 2-((2H-chromen-3-yl)methylene)malononitrile was determined and proven as the key intermediate in assembling of chromeno[2,3-b]pyridines.
Quantum chemistry calculations showed that hydroxyquinolinone-substituted chromeno[2,3-b]pyridines bearing electron-donating groups and hydroxynaphthaldehyde derivatives are prospective compounds for corrosion inhibition. In particular, methyl, methoxy, ethoxy substituted 5-(1,2-dihydroquinolin-3-yl)-5H-chromeno[2,3-b]pyridines are the most effective among those examined according to calculation.

Supplementary Materials

1H and 13C Spectra of synthesized compounds 4ai and intermediate 5 (Figure S1–S22), NMR monitoring spectra (Figure S23–S42), results of quantum chemical calculations with extended decimal point (Table S1) and figures of frontier orbitals (Figure S44–S55) are available online.

Author Contributions

Conceptualization: F.V.R. and Y.E.R.; Methodology: F.V.R., Y.E.R. and M.N.E.; writing—original draft preparation: F.V.R., Y.E.R., and S.V.V.; supervision: M.N.E. and M.P.E.; funding acquisition: M.N.E. and M.P.E.; performed the synthesis: Y.E.R.; quantum chemistry simulation: S.V.V.; performed the NMR monitoring: A.N.F.; supervised, wrote and edited the manuscript: F.V.R. and M.N.E; visualization: F.V.R. and A.N.V.; Resources: A.N.V. All authors contributed to writing the manuscript, editing, and reviewing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial support of the Russian Foundation for Basic Research (Project 18-03-00212).

Conflicts of Interest

There are no conflicts to declare.

References

  1. Dömling, A. Recent developments in isocyanide based multicomponent reactions in applied chemistry. Chem. Rev. 2006, 106, 17–89. [Google Scholar] [CrossRef] [PubMed]
  2. Dömling, A.; Ugi, I. Multicomponent reactions with isocyanides. Angew. Chemie Int. Ed. 2000, 39, 3168–3210. [Google Scholar] [CrossRef]
  3. Nair, V.; Rajesh, C.; Vinod, A.U.; Bindu, S.; Sreekanth, A.R.; Mathen, J.S.; Balagopal, L. Strategies for heterocyclic construction via novel multicomponent reactions based on isocyanides and nucleophilic carbenes. Acc. Chem. Res. 2003, 36, 899–907. [Google Scholar] [CrossRef] [PubMed]
  4. Wan, J.-P.; Liu, Y. Multicomponent reactions promoted by organosilicon reagents. Curr. Org. Chem. 2011, 15, 2758–2773. [Google Scholar] [CrossRef]
  5. Trost, B.M. The atom economy—A search for synthetic efficiency. Science 1991, 254, 1471–1477. [Google Scholar] [CrossRef] [PubMed]
  6. Trost, B.M. Atom economy—A challenge for organic synthesis: Homogeneous catalysis leads the way. Angew. Chemie Int. Ed. English 1995, 34, 259–281. [Google Scholar] [CrossRef]
  7. Clarke, P.A.; Santos, S.; Martin, W.H.C. Combining pot, atom and step economy (PASE) in organic synthesis. Synthesis of tetrahydropyran-4-ones. Green Chem. 2007, 9, 438–440. [Google Scholar] [CrossRef] [Green Version]
  8. Newhouse, T.; Baran, P.S.; Hoffmann, R.W. The economies of synthesis. Chem. Soc. Rev. 2009, 38, 3010–3021. [Google Scholar] [CrossRef]
  9. Hayashi, Y. Pot economy and one-pot synthesis. Chem. Sci. 2016, 7, 866–880. [Google Scholar] [CrossRef] [Green Version]
  10. Khasanov, A.F.; Kopchuk, D.S.; Kim, G.A.; Slepukhin, P.A.; Kovalev, I.S.; Santra, S.; Zyryanov, G.V.; Majee, A.; Chupakhin, O.N.; Charushin, V.N. Pot, Atom, Step Economic (PASE) approach towards (Aza)-2, 2′-Bipyridines: Synthesis and photophysical studies. ChemistrySelect 2018, 3, 340–347. [Google Scholar] [CrossRef]
  11. Zhang, W.; Yi, W.B. Pot, Atom, and Step Economy (PASE) Synthesis; Springer Briefs in Molecular Science; Springer International Publishing: Cham, Switzerland, 2019; ISBN 9783030225964. [Google Scholar]
  12. Zhi, S.; Ma, X.; Zhang, W. Consecutive multicomponent reactions for the synthesis of complex molecules. Org. Biomol. Chem. 2019, 17, 7632–7650. [Google Scholar] [CrossRef] [PubMed]
  13. Moseev, T.D.; Varaksin, M.V.; Gorlov, D.A.; Nikiforov, E.A.; Kopchuk, D.S.; Starnovskaya, E.S.; Khasanov, A.F.; Zyryanov, G.V.; Charushin, V.N.; Chupakhin, O.N. Direct CH/CLi coupling of 1,2,4-triazines with C6F5Li followed by aza-Diels-Alder reaction as a pot, atom, and step economy (PASE) approach towards novel fluorinated 2,2′-bipyridine fluorophores. J. Fluor. Chem. 2019, 224, 89–99. [Google Scholar] [CrossRef]
  14. Chen, X.-B.; Xiong, S.-L.; Xie, Z.-X.; Wang, Y.-C.; Liu, W. Three-Component One-Pot Synthesis of Highly Functionalized Bis-Indole Derivatives. ACS omega 2019, 4, 11832–11837. [Google Scholar] [CrossRef] [PubMed]
  15. Vereshchagin, A.N.; Elinson, M.N.; Anisina, Y.E.; Ryzhkov, F.V.; Novikov, R.A.; Egorov, M.P. PASE Pseudo-Four-Component Synthesis and Docking Studies of New 5-C-Substituted 2,4-Diamino-5H-Chromeno[2,3-b]pyridine-3-Carbonitriles. ChemistrySelect 2017, 2, 4593–4597. [Google Scholar] [CrossRef]
  16. Elinson, M.N.; Ryzhkov, F.V.; Nasybullin, R.F.; Vereshchagin, A.N.; Egorov, M.P. Fast Efficient and General PASE Approach to Medicinally Relevant 4H,5H-Pyrano-[4,3-b]pyran-5-one and 4,6-Dihydro-5H-pyrano-[3,2-c]pyridine-5-one Scaffolds. Helv. Chim. Acta 2016, 99, 724–731. [Google Scholar] [CrossRef]
  17. Cabeza, L.F.; Illa, J.; Roca, J.; Badia, F.; Mehling, H.; Hiebler, S.; Ziegler, F. Immersion corrosion tests on metal-salt hydrate pairs used for latent heat storage in the 32 to 36 °C temperature range. Mater. Corros. 2001, 52, 140–146. [Google Scholar] [CrossRef]
  18. NACE—Economic Impact, Assessment of the Global Cost of Corrosion. Available online: http://impact.nace.org/economic-impact.aspx (accessed on 18 January 2020).
  19. U.S. Report on Corrosion Costs and Preventive Strategies in the United States. Available online: http://www.corrosioncost.com (accessed on 18 January 2020).
  20. Bristol, J.A.; Gold, E.H.; Gross, I.; Lovey, R.G.; Long, J.F. Gastric antisecretory agents. 2. Antisecretory activity of 9-[(aminoalkyl)thio]-9H-xanthenes and 5-[(aminoalkyl)thio]-5H-[1]benzopyrano-[2,3-b]pyridines. J. Med. Chem. 1981, 24, 1010–1013. [Google Scholar] [CrossRef]
  21. Anderson, D.R.; Hegde, S.; Reinhard, E.; Gomez, L.; Vernier, W.F.; Lee, L.; Liu, S.; Sambandam, A.; Snider, P.A.; Masih, L. Aminocyanopyridine inhibitors of mitogen activated protein kinase-activated protein kinase 2 (MK-2). Bioorg. Med. Chem. Lett. 2005, 15, 1587–1590. [Google Scholar] [CrossRef]
  22. Tu, X.-J.; Fan, W.; Hao, W.-J.; Jiang, B.; Tu, S.-J. Three-component bicyclization providing an expedient access to Pyrano[2′,3′:5,6]pyrano[2,3-b]Pyridines and its derivatives. ACS Comb. Sci. 2014, 16, 647–651. [Google Scholar] [CrossRef]
  23. Abdelmoniem, A.M.; Ghozlan, S.A.S.; Abdelmoniem, D.M.; Elwahy, A.H.M.; Abdelhamid, I.A. Facile one-pot, three-component synthesis of novel Bis-Heterocycles incorporating 5H-chromeno[2,3-b]pyridine-3-carbonitrile Derivatives. J. Heterocycl. Chem. 2017, 54, 2844–2849. [Google Scholar] [CrossRef]
  24. Bardasov, I.N.; Alekseeva, A.U.; Mihailov, D.L.; Ershov, O.V.; Grishanov, D.A. Double heteroannulation reactions of 1-naphthol with alkyl-and arylmethylidene derivatives of malononitrile dimer. Tetrahedron Lett. 2015, 56, 1830–1832. [Google Scholar] [CrossRef]
  25. Olyaei, A.; Shahsavari, M.S.; Sadeghpour, M. Organocatalytic approach toward the green one-pot synthesis of novel benzo[f]chromenes and 12H-benzo[5,6]chromeno[2,3-b]pyridines. Res. Chem. Intermed. 2018, 44, 943–956. [Google Scholar] [CrossRef]
  26. Olyaei, A.; Shafie, Z.; Sadeghpour, M. An efficient and one-pot green synthesis of novel 6-oxo-7-aryl-6,7-dihydrochromeno pyrano[2,3-b]pyridine derivatives. Tetrahedron Lett. 2018, 59, 3567–3570. [Google Scholar] [CrossRef]
  27. Zhang, W.; Wang, J.; Mao, J.; Hu, L.; Wu, X.; Guo, C. Three-component domino cyclization of novel carbazole and indole fused pyrano[2,3-c]pyridine derivatives. Tetrahedron Lett. 2016, 57, 1985–1989. [Google Scholar] [CrossRef]
  28. Elinson, M.N.; Vereshchagin, A.N.; Anisina, Y.E.; Egorov, M.P. Efficient Multicomponent Approach to the Medicinally Relevant 5-aryl-chromeno[2,3-b]pyridine Scaffold. Polycycl. Aromat. Compd. 2020, 40, 108–115. [Google Scholar] [CrossRef]
  29. Elinson, M.N.; Vereshchagin, A.N.; Anisina, Y.E.; Krymov, S.K.; Fakhrutdinov, A.N.; Egorov, M.P. Selective multicomponent ‘one-pot’ approach to the new 5-(4-hydroxy-6-methyl-2-oxo-2H-pyran-3-yl)chromeno[2,3-b]pyridine scaffold in pyridine-ethanol catalyst/solvent system. Monatshefte für Chemie Chem. Mon. 2019, 150, 1073–1078. [Google Scholar] [CrossRef]
  30. Elinson, M.; Vereshchagin, A.; Anisina, Y.; Krymov, S.; Fakhrutdinov, A.; Egorov, M. Potassium fluoride catalysed multicomponent approach to medicinally privileged 5-[3-hydroxy-6-(hydroxymethyl)-4H-pyran-2-yl] substituted chromeno[2,3-b]pyridine scaffold. Arkivoc 2019, part ii, 38–49. [Google Scholar] [CrossRef]
  31. Vereshchagin, A.N.; Elinson, M.N.; Anisina, Y.E.; Ryzhkov, F.V.; Goloveshkin, A.S.; Novikov, R.A.; Egorov, M.P. Synthesis, structural, spectroscopic and docking studies of new 5C-substituted 2,4-diamino-5H-chromeno[2,3-b]pyridine-3-carbonitriles. J. Mol. Struct. 2017, 1146, 766–772. [Google Scholar] [CrossRef]
  32. Elinson, M.N.; Ryzhkov, F.V.; Korolev, V.A.; Egorov, M.P. Pot, atom and step-economic (PASE) synthesis of medicinally relevant spiro [oxindole-3, 4′-pyrano [4, 3-b] pyran] scaffold. Heterocycl. Commun. 2016, 22, 11–15. [Google Scholar] [CrossRef]
  33. Elinson, M.N.; Nasybullin, R.F.; Ryzhkov, F.V.; Egorov, M.P. Solvent-free and ‘on-water’ multicomponent assembling of salicylaldehydes, malononitrile and 3-methyl-2-pyrazolin-5-one: A fast and efficient route to the 2-amino-4-(1H-pyrazol-4-yl)-4H-chromene scaffold. Comptes Rendus Chim. 2014, 17, 437–442. [Google Scholar] [CrossRef]
  34. Elinson, M.N.; Nasybullin, R.F.; Ryzhkov, F.V.; Zaimovskaya, T.A.; Nikishin, G.I. Solvent-free and ‘on-water’ multicomponent assembling of aldehydes, 3-methyl-2-pyrazoline-5-one, and malononitrile: Fast and efficient approach to medicinally relevant pyrano[2,3-c]pyrazole scaffold. Monatshefte für Chemie Chem. Mon. 2015, 146, 631–635. [Google Scholar] [CrossRef]
  35. Chanda, A.; Fokin, V. V Organic synthesis “On Water”. Chem. Rev. 2009, 109, 725–748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Vereshchagin, A.N.; Elinson, M.N.; Anisina, Y.E.; Ryzhkov, F.V.; Goloveshkin, A.S.; Bushmarinov, I.S.; Zlotin, S.G.; Egorov, M.P. Pot, atom and step economic (PASE) synthesis of 5-isoxazolyl-5H-chromeno[2,3-b]pyridine scaffold. Mendeleev Commun. 2015, 25, 424–426. [Google Scholar] [CrossRef]
  37. Patai, S.; Israeli, Y. 411. The kinetics and mechanisms of carbonyl–methylene condensations. Part VII. The reaction of malononitrile with aromatic aldehydes in ethanol. J. Chem. Soc. 1960, 2025–2030. [Google Scholar] [CrossRef]
  38. Elinson, M.N.; Vereshchagin, A.N.; Anisina, Y.E.; Goloveshkin, A.S.; Ushakov, I.E.; Egorov, M.P. PASE facile and efficient multicomponent approach to the new type of 5-C-substituted 2,4-diamino-5H-chromeno[2,3-b]pyridine scaffold. Mendeleev Commun. 2018, 28, 372–374. [Google Scholar] [CrossRef]
  39. Elinson, M.N.; Gorbunov, S.V.; Vereshchagin, A.N.; Nasybullin, R.F.; Goloveshkin, A.S.; Bushmarinov, I.S.; Egorov, M.P. Chemical and electrocatalytic cascade cyclization of salicylaldehyde with three molecules of malononitrile: ‘one-pot’ simple and efficient way to the chromeno[2,3-b]pyridine scaffold. Tetrahedron 2014, 70, 8559–8563. [Google Scholar] [CrossRef]
  40. Elinson, M.N.; Vereshchagin, A.N.; Anisina, Y.E.; Fakhrutdinov, A.N.; Goloveshkin, A.S.; Egorov, M.P. Pot-, Atom- and Step-Economic (PASE) Multicomponent approach to the 5-(Dialkylphosphonate)-Substituted 2,4-Diamino-5H-chromeno[2,3-b]pyridine scaffold. Eur. J. Org. Chem. 2019, 2019, 4171–4178. [Google Scholar] [CrossRef]
  41. Amer, A.A. Synthesis of Some New Polyfunctionalized Pyridines. J. Heterocycl. Chem. 2018, 55, 297–301. [Google Scholar] [CrossRef]
  42. Abou Elmaaty, T.; el taweel, M.; Elmougi, S.; Elagamey, A. The condensation of active methylene reagents with salicylaldehyde: Novel synthesis of chromene, azaanthracene, pyrano[3,4-c]chromene and chromeno[3,4-c]pyridine derivatives. J. Heterocycl. Chem. 2004, 41, 655–658. [Google Scholar] [CrossRef]
  43. Sakurai, A.; Midorikawa, H.; Hashimoto, Y. The cyclization of ethyl cyanoacetate and salicylaldehyde or 3-methoxysalicylaldehyde with ketones by means of ammonium acetate. Bull. Chem. Soc. Jpn. 1970, 43, 2925–2933. [Google Scholar] [CrossRef] [Green Version]
  44. Soleiman, H.A.; Koraiem, A.I.M.; Mahmoud, N.Y. Synthesis of new fused hetecrocyclic compounds of benzpyrid-4-one derivatives and their some biological activity. J. Chin. Chem. Soc. 2004, 51, 553–560. [Google Scholar] [CrossRef]
  45. Finšgar, M.; Milošev, I. Inhibition of copper corrosion by 1,2,3-benzotriazole: A review. Corros. Sci. 2010, 52, 2737–2749. [Google Scholar] [CrossRef]
  46. Al-Mayouf, A.M.; Al-Ameery, A.K.; Al-Suhybani, A.A. Inhibition of type 304 stainless steel corrosion in 2 M sulfuric acid by some benzoazoles—Time and temperature effects. Corrosion 2001, 57, 614–620. [Google Scholar] [CrossRef]
  47. Lashgari, M.; Arshadi, M. DFT studies of pyridine corrosion inhibitors in electrical double layer: Solvent, substrate, and electric field effects. Chem. Phys. 2004, 299, 131–137. [Google Scholar] [CrossRef]
  48. Levin, M.; Wiklund, P.; Leygraf, C. Bioorganic compounds as copper corrosion inhibitors in hydrocarbon media. Corros. Sci. 2012, 58, 104–114. [Google Scholar] [CrossRef]
  49. Paul, P.K.; Saraswat, V.; Yadav, M. Chromenes as efficient corrosion inhibitor for mild steel in HCl solution. J. Adhes. Sci. Technol. 2019, 33, 1275–1293. [Google Scholar] [CrossRef]
  50. Khaled, K.; Al-Mobarak, N.A. A predictive model for corrosion inhibition of mild steel by thiophene and its derivatives using artificial neural network. Int. J. Electrochem. Sci. 2012, 7, 1045–1059. [Google Scholar]
  51. Taylor, C.D. Design and prediction of corrosion inhibitors from quantum chemistry. J. Electrochem. Soc. 2015, 162, 340–346. [Google Scholar] [CrossRef]
  52. Taylor, C.D.; Chandra, A.; Vera, J.; Sridhar, N. Design and prediction of corrosion inhibitors from quantum chemistry II. A general framework for prediction of effective oil/water partition coefficients and speciation from quantum chemistry. J. Electrochem. Soc. 2015, 162, 347–353. [Google Scholar] [CrossRef]
  53. Ebenso, E.E.; Kabanda, M.M.; Murulana, L.C.; Singh, A.K.; Shukla, S.K. Electrochemical and quantum chemical investigation of some azine and thiazine dyes as potential corrosion inhibitors for mild steel in hydrochloric acid solution. Ind. Eng. Chem. Res. 2012, 51, 12940–12958. [Google Scholar] [CrossRef]
  54. Amin, M.A.; Khaled, K.F.; Mohsen, Q.; Arida, H.A. A study of the inhibition of iron corrosion in HCl solutions by some amino acids. Corros. Sci. 2010, 52, 1684–1695. [Google Scholar] [CrossRef]
  55. Khaled, K.; Fadl-allah, S.; Hammouti, B. Some benzotriazole derivatives as corrosion inhibitors for copper in acidic medium: Experimental and quantum chemical molecular dynamics approach. Mater. Chem. Phys. 2009, 117, 148–155. [Google Scholar] [CrossRef]
  56. Parr, R.G.; Pearson, R.G. Absolute hardness: Companion parameter to absolute electronegativity. J. Am. Chem. Soc. 1983, 105, 7512–7516. [Google Scholar] [CrossRef]
  57. Zhang, J.; Liu, J.; Yu, W.; Yan, Y.; You, L.; Liu, L. Molecular modeling of the inhibition mechanism of 1-(2-aminoethyl)-2-alkyl-imidazoline. Corros. Sci. 2010, 52, 2059–2065. [Google Scholar] [CrossRef]
  58. Sulaiman, K.O.; Onawole, A.T. Quantum chemical evaluation of the corrosion inhibition of novel aromatic hydrazide derivatives on mild steel in hydrochloric acid. Comput. Theor. Chem. 2016, 1093, 73–80. [Google Scholar] [CrossRef]
  59. Dega-Szafran, Z.; Kania, A.; Nowak-Wydra, B.; Szafran, M. UV, 1H and 13C NMR spectra, and AM1 studies of protonation of aminopyridines. J. Mol. Struct. 1994, 322, 223–232. [Google Scholar] [CrossRef]
  60. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria., G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09; Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  61. Becke, A.D. Becke’s three parameter hybrid method using the LYP correlation functional. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  62. Metri, N.; Sallenave, X.; Plesse, C.; Beouch, L.; Aubert, P.-H.; Goubard, F.; Chevrot, C.; Sini, G. Processable star-shaped molecules with triphenylamine core as hole-transporting materials: Experimental and theoretical approach. J. Phys. Chem. C 2012, 116, 3765–3772. [Google Scholar] [CrossRef]
  63. Cances, E.; Mennucci, B.; Tomasi, J. A new integral equation formalism for the polarizable continuum model: Theoretical background and applications to isotropic and anisotropic dielectrics. J. Chem. Phys. 1997, 107, 3032–3041. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds 4ai are available from the authors.
Scheme 1. Reaction of salicylaldehyde, malononitrile dimer and hydroxyquinolinone.
Scheme 1. Reaction of salicylaldehyde, malononitrile dimer and hydroxyquinolinone.
Molecules 25 02573 sch001
Figure 1. Comparison diagram of approaches to chromeno[2,3-b]pyridine 4a. The mixture of EtOH-Py (3:1) provides the best result.
Figure 1. Comparison diagram of approaches to chromeno[2,3-b]pyridine 4a. The mixture of EtOH-Py (3:1) provides the best result.
Molecules 25 02573 g001
Scheme 2. The key intermediates for the possible pathways to chromeno[2,3-b]pyridine formation.
Scheme 2. The key intermediates for the possible pathways to chromeno[2,3-b]pyridine formation.
Molecules 25 02573 sch002
Figure 2. Representative 1H NMR spectrum of multicomponent reaction in DMSO-d6 at 313 K recorded 2 h after dissolution.
Figure 2. Representative 1H NMR spectrum of multicomponent reaction in DMSO-d6 at 313 K recorded 2 h after dissolution.
Molecules 25 02573 g002
Figure 3. Real-time monitoring of chemical reaction by 1H NMR spectroscopy in DMSO-d6 at 313 K. The NMR spectra of all the synthesized compounds, as well as the NMR spectra of the real-time monitoring are presented in Supplementary Materials (Figures S1–S42).
Figure 3. Real-time monitoring of chemical reaction by 1H NMR spectroscopy in DMSO-d6 at 313 K. The NMR spectra of all the synthesized compounds, as well as the NMR spectra of the real-time monitoring are presented in Supplementary Materials (Figures S1–S42).
Molecules 25 02573 g003
Scheme 3. Mechanism of salicylaldehydes 1, malononitrile dimer 2 and hydroxyquinolinone 3 transformation into chromeno[2,3-b]pyridine 4. Catalytic cycles are simplified.
Scheme 3. Mechanism of salicylaldehydes 1, malononitrile dimer 2 and hydroxyquinolinone 3 transformation into chromeno[2,3-b]pyridine 4. Catalytic cycles are simplified.
Molecules 25 02573 sch003
Figure 4. Optimized structure for compound 4a (gas phase).
Figure 4. Optimized structure for compound 4a (gas phase).
Molecules 25 02573 g004
Scheme 4. Protonation of chromeno[2,3-b]pyridine 4a.
Scheme 4. Protonation of chromeno[2,3-b]pyridine 4a.
Molecules 25 02573 sch004
Table 1. Optimization of reaction conditions 1.
Table 1. Optimization of reaction conditions 1.
Entry SolventCatalystTime (h)Temp (°C)Yield (%)
1Solvent-free-28017
2Solvent-freeKF28025
3Solvent-freeAcONa28037
4H2O-28016
5H2OAcONa28031
6EtOH-27819 2
7EtOHAcONa27835 2
8EtOHEt3N27863 2
9EtOHMorph27857 2
10Py-211661 2
11EtOH-Py (4:1)-27887 2
12EtOH-Py (3:1)-27895 2
13EtOH-Py (2:1)-27892 2
14EtOH-Py (3:1)-17889 2
1 Reaction conditions 1a (1 mmol), 2 (1 mmol), 3 (1 mmol) were heated in 4 mL of solvent or without solvent; with 10 mol% of catalyst or without catalyst. 2 Isolated yield, in other cases NMR data.
Table 2. PASE reaction of salicylaldehydes 1a–i, malononitrile dimer 2 and hydroxyquinolinone 3 1.
Table 2. PASE reaction of salicylaldehydes 1a–i, malononitrile dimer 2 and hydroxyquinolinone 3 1.
Molecules 25 02573 i001
Molecules 25 02573 i002
1 Reaction conditions: 1ai (1 mmol), 2 (1 mmol), 3 (1 mmol) were refluxed in 4 mL of EtOH-Py mixture (3:1). Isolated yields.
Table 3. Results of quantum chemical calculations for studied compounds 4ai.
Table 3. Results of quantum chemical calculations for studied compounds 4ai.
Compound4a4b4c4d4e4f4g4h4i
Gas Phase Calculations
Total energy, a.u.−1347.19−1386.52−1461.75−1501.07−4035.28−1806.81−3920.74−1551.75−1500.87
E(HOMO), eV−5.707−5.649−5.644−5.558−5.695−5.835−5.830−6.065−5.669
E(LUMO), eV−2.040−2.023−1.997−1.982−2.072−2.135−2.131−2.416−2.047
ΔE(L-H), eV3.6673.6263.6473.5763.6233.7003.6993.6493.622
μ, D7.8968.0717.6458.1197.4877.3257.3517.6197.670
χ3.8743.8363.8213.7703.8843.9863.9814.2413.858
η1.8341.8131.8241.7881.8121.851.8491.8251.811
ω4.0924.0584.0033.9754.1644.3014.2864.9294.109
σ0.5450.5520.5480.5590.5520.5410.5410.5480.552
Calculations for Solvated Compounds
Total energy, a.u.−1347.22−1386.55−1461.77−1501.10−4035.31−1806.84−3920.76−1551.78−1500.90
E(HOMO), eV−5.993−5.947−5.940−5.900−5.968−6.037−6.035−6.139−5.934
E(LUMO), eV−1.795−1.791−1.783−1.791−1.822−1.826−1.825−2.726−1.810
ΔE(L-H), eV4.1984.1564.1574.1094.1464.2114.2103.4134.124
μ, D10.26311.7899.88310.87311.1379.4909.4869.60910.135
χ3.8943.8693.8673.8463.8953.9323.9304.4333.872
η2.0992.0782.0792.0552.0732.1062.1051.7072.062
ω3.6123.6023.5963.5993.6593.6713.6695.7563.635
σ0.4760.4810.4810.4870.4820.4750.4750.5860.485
Calculations for protonated forms of studied compounds
Total energy, a.u.−1347.67−1386.99−1462.22−1501.55−4035.76−1807.29−3921.21−1552.22−1501.34
E(HOMO), eV−6.592−6.541−6.374−6.350−6.449−6.623−6.617−6.670−6.321
E(LUMO), eV−1.983−1.971−1.967−1.977−2.026−2.035−2.033−2.927−2.044
ΔE(L-H), eV4.6094.574.4074.3734.4234.5884.5843.7434.277
μ, D9.7049.78511.0998.72213.35612.54313.78416.31911.489
χ4.2884.2564.1714.1644.2634.3294.3254.7994.183
η2.3052.2852.2042.1872.2122.2942.2921.8722.139
ω3.9883.9643.9473.9644.1084.0854.0816.1514.090
σ0.4340.4380.4540.4570.4520.4360.4360.5340.468
Ionization potential [53]: I ≈ −E(HOMO); electron affinity as: A ≈ −E(LUMO); electronegativity: χ ≈ (I + A)/2; chemical hardness: η ≈ (I − A)/2; global electrophilicity index: ω = (χ2)/2η; global chemical softness: σ = 1/η. this table with the extended decimal point is presented in Supplementary Materials (Table S1).
Table 4. Frontier orbitals of several studied compounds.
Table 4. Frontier orbitals of several studied compounds.
CompoundHOMOLUMO
In Gas Phase
4a Molecules 25 02573 i003 Molecules 25 02573 i004
4e Molecules 25 02573 i005 Molecules 25 02573 i006
4h Molecules 25 02573 i007 Molecules 25 02573 i008
Solvated Molecules
4c Molecules 25 02573 i009 Molecules 25 02573 i010
Protonated Molecules
4a Molecules 25 02573 i011 Molecules 25 02573 i012
4h Molecules 25 02573 i013 Molecules 25 02573 i014
Frontier orbitals of these compounds are also presented in Figures S43–S55 of Supplementary Materials.

Share and Cite

MDPI and ACS Style

Ryzhkov, F.V.; Ryzhkova, Y.E.; Elinson, M.N.; Vorobyev, S.V.; Fakhrutdinov, A.N.; Vereshchagin, A.N.; Egorov, M.P. Catalyst-Solvent System for PASE Approach to Hydroxyquinolinone-Substituted Chromeno[2,3-b]pyridines Its Quantum Chemical Study and Investigation of Reaction Mechanism. Molecules 2020, 25, 2573. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25112573

AMA Style

Ryzhkov FV, Ryzhkova YE, Elinson MN, Vorobyev SV, Fakhrutdinov AN, Vereshchagin AN, Egorov MP. Catalyst-Solvent System for PASE Approach to Hydroxyquinolinone-Substituted Chromeno[2,3-b]pyridines Its Quantum Chemical Study and Investigation of Reaction Mechanism. Molecules. 2020; 25(11):2573. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25112573

Chicago/Turabian Style

Ryzhkov, Fedor V., Yuliya E. Ryzhkova, Michail N. Elinson, Stepan V. Vorobyev, Artem N. Fakhrutdinov, Anatoly N. Vereshchagin, and Mikhail P. Egorov. 2020. "Catalyst-Solvent System for PASE Approach to Hydroxyquinolinone-Substituted Chromeno[2,3-b]pyridines Its Quantum Chemical Study and Investigation of Reaction Mechanism" Molecules 25, no. 11: 2573. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25112573

Article Metrics

Back to TopTop