Next Article in Journal
Design of Highly Efficient Nickel-Cobalt-Manganese-Molybdenum (NCMM) Nano-Catalysts Supported on Activated Carbon for Desulfurization Process
Next Article in Special Issue
Recent Approaches in Transition Metal-Catalyzed Chalcogenative Heteroannulation of Alkenes and Alkynes
Previous Article in Journal
Copper-Catalyzed Reaction of N-Monosubstituted Hydrazones with CBr4: Unexpected Fragmentation and Mechanistic Study
Previous Article in Special Issue
Recent Advances for the Synthesis of Dihydroquinolin-2(1H)-ones via Catalytic Annulation of α,β-Unsaturated N-Arylamides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Recyclable Ni-Containing Coordination Polymer as an Efficient Catalyst for the Synthesis of Oxindole and Quinoline Derivatives through the Borrowing Hydrogen Strategy

1
The Key Laboratory of Synthetic and Biological Colloids, Ministry of Education, School of Chemical and Material Engineering, Jiangnan University, Wuxi 214122, China
2
School of Chemical and Environmental Engineering, Hubei Minzu University, Enshi 445000, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2023, 13(8), 1195; https://doi.org/10.3390/catal13081195
Submission received: 3 July 2023 / Revised: 27 July 2023 / Accepted: 4 August 2023 / Published: 9 August 2023
(This article belongs to the Special Issue Catalytic Annulation Reactions)

Abstract

:
A coordination polymer Ni-CIA has been designed, synthesized, and characterized via Fourier transform infrared spectroscopy (FT-IR), powder X-ray diffraction (XRD), scanning electron microscopy (SEM), X-ray photoelectron spectrometry (XPS), energy-dispersive X-ray spectroscopy (EDS), and thermal gravimetric analysis (TGA). The resulting polymer reveals a slice layer stacked combination structure and good stability. Importantly, this polymer exhibited high catalytic activity for the synthesis of oxindole and quinoline derivatives via borrowing hydrogen and dehydrogenation reactions. Meanwhile, several control reactions were carried out to investigate the possible mechanisms of this transformation. Finally, the results of the recycling experiment indicate that the coordination polymer Ni-CIA could be recovered and reused at least five times.

Graphical Abstract

1. Introduction

Oxindole derivatives are privileged heterocyclic cores, constituting an important skeleton in several compounds of pharmaceutical interest with significant biological activities (Scheme 1) [1,2,3,4,5,6,7,8]. The conventional approaches for the alkylation of oxindoles employed toxic alkyl halides, suffered from harsh reaction conditions, and exhibited poor regioselectivity alongside the generation of undesired side products [9,10]. Therefore, the design and synthesis of a new catalytic system for the alkylation of oxindole derivatives has become a research hotspot in the field of organic synthesis [11,12,13,14,15,16].
The borrowing hydrogen (BH) reaction, also known as the hydrogen autotransfer reaction, employs inexpensive alcohols as alkylation agents, generating water as the sole byproduct [17,18,19,20,21,22,23,24]. Thus, it is an atom-efficient and much greener process for the alkylation of oxindole and indole derivatives via borrowing hydrogen reactions [25,26,27,28,29,30,31]. In 2002, Simig and co-workers first reported that Raney nickel catalyzed the synthesis of 3-alkyloxindoles by means of the dehydrogenative coupling of 2-oxindole and alcohol in moderate yields [32]. Later, the alkylation of oxindole with homogeneous Ru, Ir, Fe, Mn catalysts were also reported [33,34,35,36,37]. In 2018, we developed homogeneous pyridyl triazole-ligated ruthenium(II) complexes which could realize the selective alkylation of 2-oxindole with a variety of primary alcohols via the hydrogen autotransfer process [38]. However, homogeneous catalysts are known to have some disadvantages, such as difficulty in separation from the reaction system and their poor recyclability, which limits their further application [39,40,41,42,43].
Coordination polymers (CPs) have some superior properties, such as their unique channels, good stability, and effective control of their structure and functionality [44,45,46]. CPs have bright future applications in many fields, such as gas adsorption and separation, catalysis, and functional materials [47,48,49]. Organic ligands play a significant role in the design and synthesis of CPs. Recently, we reported a number of nitrogen-containing polycarboxylic acid ligands, which bind with several metal centers into chains to form CPs. Meanwhile, these CPs could be applied to some borrowing hydrogen and dehydrogenation reactions successfully [50,51,52,53]. Herein, an indolyl dicarboxylic acid ligand was used to synthesize a new Ni-containing coordination polymer, Ni-CIA (Ni-indolyl dicarboxylic acid coordination polymer), which showed excellent catalytic activity in the synthesis of oxindole and quinoline derivatives through the borrowing hydrogen strategy.

2. Results and Discussion

2.1. Characterization of Ni-CIA

After synthesis, Ni-CIA was characterized using Fourier transform infrared spectroscopy (FT-IR), powder X-ray diffraction (XRD), scanning electron microscopy (SEM), X-ray photoelectron spectrometry (XPS), energy-dispersive X-ray spectroscopy (EDS), and thermal gravimetric analysis (TGA).
The FT-IR spectra of the H2CIA (indolyl dicarboxylic acid ligand) and Ni-CIA are shown in Figure 1a. In the spectra of H2CIA and Ni-CIA, the characteristic absorption band at 2800~3500 cm−1 is attributed to O–H skeleton stretching vibrations of the hydroxyl groups. The strong absorption bands of H2CIA in the range of 1690~1247 cm−1 were attributed to the asymmetric and symmetric vibrations of the carboxylate groups, respectively. For Ni-CIA, the two corresponding peaks at 1690 and 1247 cm−1 slightly shifted to lower frequencies (1570 and 1390 cm−1, respectively), which indicates that the carboxylate groups in Ni-CIA have been coordinated to the Ni2+ cation [54]. The crystal structure of Ni-CIA was determined via XRD analysis (Figure 1b). The XRD spectrum of Ni-CIA showed various sharp peaks, indicating the polycrystalline nature. Moreover, the main diffraction peaks of the Ni-CIA are observed at 2θ = 15.2°, 26.6°, 30.8°, 34.9°, and 40.8°, corresponding to the crystal faces of (100), (001), (110), (200), and (111), respectively.
To determine the morphology of Ni-CIA, scanning electron microscopy (SEM) was carried out (Figure 2). The result indicated that Ni-CIA has a slice layer stacked combination structure, which could create a favorable environment for rapid ion diffusion and accelerate catalytic mass transfer during the catalytic process.
Furthermore, XPS analysis has been performed to obtain detailed insights into the surface electronic valence states of Ni-CIA. As revealed in Figure 3a, the C 1s spectrum can be deconvoluted into five peaks [55], which were associated with C−C/C=C (284.8 eV), C−N (286.1 eV), C−O (287.0 eV), C=O (288.4 eV), and O–C=O (290.8 eV). In the N 1s spectrum of Ni-CIA (Figure 3b), three peaks with binding energies at 398.6 eV, 399.5 eV, and 400.1 eV are assigned to pyridinic-N, coordinated-N, and pyrrolic-N, respectively [56]. In Figure 3c, the high-resolution spectrum of O 1s is demonstrated. The peak is fitted into three parts at ~531.1 eV, ~532.5 eV, and ~533.3 eV, which can be accredited to Ni–O, O–H, and O–C=O, respectively [57]. Two peaks for Ni 2p (Figure 3d) are detected at binding energies 873.1 and 855.1 eV, corresponding to the Ni 2p1/2 and Ni 2p3/2, respectively. It appears that these values are consistent with the XPS data on Ni2+ [58]. Generally, the binding energy of coordinated-N and Ni–O were indicative of the nitrogen in H2CIA, and oxygen in C–OH participated in the coordination during the Ni-CIA formation.
Further, the elemental compositional distribution determined from EDS analysis confirms the presence of C, N, O, and Ni (Figure 4a). TGA characterization was conducted and analyzed, and the corresponding results are presented in Figure 4b. For pristine Ni-CIA, 11.29% of the weight loss at ~250 °C was associated with the removal of water and decomposition of non-coordinated H2CIA ligands. Subsequently, the coordination polymer Ni-CIA began to disintegrate, losing 49.8% of mass. The weight of Ni-CIA remains basically stable after 400 °C. Generally, the coordination polymer Ni-CIA will not cause thermal decomposition when catalyzing reactions below 200 °C.

2.2. Catalytic Activity

Initially, the alkylation of 2-oxindole 1a with benzyl alcohol 2a was chosen as a model reaction. Meanwhile, the reaction conditions were optimized as outlined in Table 1. We found that no alkylation of the desired product 3a was observed in the absence of catalyst Ni-CIA, indicating that the Ni-CIA catalyst played a key role in this transformation (Table 1, entry 1). Various inorganic bases, including K2CO3, Cs2CO3, KOtBu, KOH, and K3PO4, were evaluated (Table 1, entries 3–7, respectively). Among the various inorganic bases used, the best results were obtained with KOtBu, which could afford the alkylation product in 84% isolated yield.
Next, we checked the reactivity of the reaction, as polar solvents such as H2O and N,N-dimethylformamide (DMF) all led to poor yields. We then moved to non-polar solvents like xylene and 1,4-dioxane, and a slight decrease in alkylation product 3a yield was observed (Table 1, entries 8–11, respectively). Additionally, we attempted to adjust the reaction time; however, the yield of the desired product decreased (Table 1, entries 12–13, respectively). Subsequently, the reaction was carried out under a N2 atmosphere, affording the target product 3a in excellent yield (88% yield) (Table 1, entry 14). Finally, the O2 atmosphere was screened, though with a greatly reduced yield of 43% (Table 1, entry 15).
With the optimized reaction conditions in hand, the substrate scope of benzyl alcohol was then investigated. We were pleasantly surprised to find that 2-oxindole 1a reacted with benzyl alcohol derivatives 2 smoothly to afford the desired product 3 in moderate to excellent yields. However, the electronic effect of the substituents on the benzene ring had slight distinctions. When a variety of electron-withdrawing groups (such as -F, -Cl, -Br, and -CF3) were present at the para-position, the corresponding products were obtained in good to excellent yields (Scheme 2, entries 3d3g). In contrast, when a variety of electron-donating groups (such as -Me, -OMe, and -tBu) were attached to the para-position (Scheme 2, entries 3b, 3c, and 3h), the desired product was obtained in moderate yields. In comparison, slightly lower yields were observed at the meta-, ortho-position of the benzene ring of benzyl alcohols substituents (3i3l), which are probably due to steric hindrance. To our delight, even if the electronic effect was enhanced through adding substituents (-CF3) on the benzene ring, the yield of the desired product was not reduced (3m). Subsequently, several heteroaryl primary alcohols, including 2-naphthalenemethanol, 2-thiophenemethanol, and piperonyl alcohol, were also investigated (3n3p). To our delight, they were still compatible, and the desired products were obtained in good yields. The scope was further extended to linear aliphatic secondary alcohols. Interestingly, aliphatic alcohols were also effective for the alkylation of 2-oxindole to offer the corresponding products 3q3s in moderate yields. Meanwhile, alicyclic alcohol (cyclopropane methanol 2t) was also effective at yielding the corresponding products 3-(cyclopropyl methyl)indolin-2-one 3t in good yields.
Furthermore, we were eager to explore more synthetic utilities that the Ni-CIA catalyst has to offer, since the quinoline framework is a crucial class of heterocyclic scaffolds, which could be found in many natural products, drugs, and advanced functional materials [59,60,61]. Therefore, exploring new methods for synthesizing quinoline derivatives has become a research hotspot in the field of organic synthesis [62,63,64,65,66,67,68]. To our delight, using Ni-CIA as a catalyst, we have achieved an effective method to synthesize quinoline derivatives via borrowing hydrogen transformation. Meanwhile, we further explored the substrate scope of this process (Scheme 3). A series of acetophenones bearing different groups were investigated. Almost all the acetophenones reacted with o-aminobenzyl alcohol smoothly to produce quinoline derivatives with good to excellent yields. We can find that when using acetophenone 5 with electron-donating substituents, higher yields could be achieved than their counterparts with electron-withdrawing substituents (Scheme 3, entries 6a6f). During the transformation, compared to the para-position of the benzene ring of acetophenone substituents, a slight decrease in the yield of acetophenone substituents in the meta- and ortho- positions of the benzene ring was noticed during this transformation (Scheme 3, entries 6g6j). Heteroaromatic alcohols with a pyridinyl group also afforded the corresponding indoles in 72% yield (Scheme 3, entry 6m). Interestingly, we observed that alkyl ketones could be well tolerated in the reaction of o-aminobenzyl alcohol with ketone, affording moderate yields (Scheme 3, entries 6n, 6o).

2.3. Mechanism Explorations

To gain insight into the mechanism of the alkylation of 2-oxindole with alcohols, several catalyst investigative experiments were conducted (Table 2). Firstly, four different commercially available simple nickel salts were examined. The control experiments were carried out with KOtBu as a base, dissolved in toluene for 6 h (Table 2, entries 1–4). It was found that almost no reaction occurred when simple nickel salts were employed. Meanwhile, a low yield was observed when Ni-complexes were used as catalysts (Table 2, entries 5–6). These results reveal that the Ni-CIA catalyst has played a crucial role in the transformation.
In order to better understand the mechanism of this transformation and to clarify the effect of different substituents on reactivity, several control experiments were subsequently carried out. When benzyl alcohol containing electron-donating or electron-withdrawing groups (2c vs. 2g) reacted with 2-oxindole (1a) for 2 h under optimal conditions, the experiment results indicated that the molar ratio of two products (3c vs. 3g) was found to be 1:2.4, indicating that the electron-deficient alcohol was preferentially consumed. Meanwhile, it is concluded that benzyl alcohol with an electron-withdrawing group was much more reactive than that with an electron-donating group for this reaction (Scheme 4).
Further mechanism exploration was also executed using two useful types of radical scavenger, 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO, 1.0 equiv.) and 2,6-ditert-butyl-4-methylphenol (BHT, 1.0 equiv.), to test and verify the possibility of the radical process in this reaction (Scheme 5a). The experimental results showed that the reaction of 2-oxindole 1a and benzyl alcohol 2a produced the desired products 3a in 76% and 84% yields, respectively, and ruled out the possibility of a radical pathway during this Ni-CIA-catalyzed reaction.
In order to find out the right reaction pathway, we next carried out several control experiments to capture possible intermediates (Scheme 5b). When only benzyl alcohol 2a was used as the reactant, the intermediate benzaldehyde 2a’ was formed with a yield of 64% under optimal conditions, which indicated that the Ni-CIA catalyst was essential for the oxidation of benzyl alcohol 2a (i). Furthermore, the reaction of benzaldehyde 2a’ and 2-oxindole 1a afforded an α,β-unsaturated amide 1a’ with 92% yield (ii). At the same time, the experimental results indicate that α,β-unsaturated amide 1a’ and benzyl alcohol 2a could not react to form desired product 3a without the participation of the Ni-CIA catalyst, implying that the Ni-CIA catalyst was necessary for this transformation (iii).
Subsequently, to understand the mechanism of 6a formation, a series of control experiments were performed (Scheme 6). Initially, in the presence of BTH and TEMPO, this transformation proceeded well, and the yield of 6a was only slightly decreased. This rules out the involvement of an organic radical intermediate during this reaction, as it is widely accepted that the capture of key intermediate is an efficient and reliable way to explain what happened in this reaction process. The reaction of only o-aminobenzyl alcohol 4a produced 2-aminobenzaldehyde 4a’ in 58% isolated yield under the standard conditions (i). The reaction of 2-aminobenzaldehyde 4a’ and acetophenone 5a afforded the 2-phenylquinoline 6a in 94% yield in the absence of catalyst Ni-CIA, implying that the carbonyl compound (2-aminobenzaldehyde 4a’) is the intermediate, and Cs2CO3 played a significant role in the deprotonation of alcohols and condensation (ii). Notably, the formation of 2-phenylquinoline 6a was not observed from the coupling of o-aminobenzyl alcohol 4a with acetophenone 5a in the absence of the catalyst (iii). The above control experiment results together confirmed that the catalyst Ni-CIA facilitated the dehydrogenation of the o-aminobenzyl alcohol 4a and that 2-aminobenzaldehyde 4a’ was the intermediate in this process.
Based on our experimental research and previous literature, a possible mechanism is proposed in Scheme 7 [33,34,35,36,37]. First, the catalyst Ni-CIA undergoes an ion exchange with the base KOtBu to produce the active species A, which then participates in the catalytic cycle. Subsequently, A reacts with benzyl alcohol 2a, which results in the formation of alkoxy intermediate B. The intermediate complex B undergoes β-hydride elimination of alkoxide to generate the dehydrogenated product benzaldehyde 2a’ with the formation of [Ni-H] species. After that, the benzaldehyde 2a’ reacted with 2-oxindole 1a to form an α,β-unsaturated amide 1a’ intermediate. Finally, α,β-unsaturated amide 1a’ was reduced by the [Ni-H] species. Meanwhile, A was regenerated to complete the catalytic cycle.
Similarly, a plausible mechanism for the Ni-catalyzed synthesis of quinoline derivatives via the dehydrogenative coupling of o-aminobenzyl alcohol 4a with acetophenone 5a is shown in Scheme 8. Firstly, the deprotonation of 4a assisted by the Cs2CO3 produces the active species C, which then participate in this catalytic cycle. Next, C reacted with Ni-CIA to afford the alkoxy intermediate D. The intermediate D undergoes β-hydride elimination of alkoxide to produce the dehydrogenated product 2-aminobenzaldehyde 4a’ with the formation of [Ni-H] species. Subsequently, 2-aminobenzaldehyde 4a’ undergoes base-catalyzed cross-aldol condensation reaction with acetophenone 5a to form an α,β-unsaturated ketone. Finally, the α,β-unsaturated ketone reacts via intramolecular cyclization to produce the product 6a.

2.4. Reusability of the Catalyst

Finally, the durability of the catalyst Ni-CIA was assessed through recycling experiments. After the reaction of 2-oxindole 1a with benzyl alcohol 2a, the heterogeneous catalyst Ni-CIA was conveniently separated from the reaction mixture via centrifugation, followed by washing with methanol, deionized water, and ethanol, respectively, and dried at 80 °C for 12 h. Then, it was reused in the following run without any pre-treatment, as shown in Scheme 9. To our delight, the catalyst Ni-CIA provided a high yield of 3a even after the fifth recycling experiment. Meanwhile, after the reaction cycle, the leaching of Ni species was measured via ICP-MS analysis. It was shown that the leaching of Ni species in this reaction solution was only 5.2 μg/L, which disclosed that the heterogeneous catalyst Ni-CIA had good stability. After the recycling experiments, the Ni-CIA catalyst was collected and characterized via scanning electron microscopy (SEM) again after being washed three times with methanol, deionized water, and ethanol and dried at 80 °C (Figure S1). From the figure, no significant change in the morphology of the catalyst Ni-CIA is seen, retaining its slice layer stacked combination structure.

3. Materials and Methods

3.1. Catalyst Characterization

All reactions that require air and/or moisture-sensitive conditions were carried out under a nitrogen atmosphere using oven/flame-dried glassware and standard syringe/septa techniques. Unless otherwise noted, all commercial reagents and solvents were obtained from the commercial provider and used without further purification. FT-IR spectra (Thermo Fisher Co., Waltham, MA, USA) were recorded on a Nicolet 360 FT-IR instrument (KBr discs) in the 4000–525 cm−1 regions. SEM and EDS (Hitachi, Tokyo, Japan) was performed on a HITACHI S-4800 field-emission scanning electron microscope. TG analysis was carried out using an STA409 (Mettler-Toli, Greifensee, Switzerland) instrument in N2 at a heating rate of 10 °C/min from 50 to 800 °C. XRD patterns were collected on a Bruker D8 (Bremen, Germany) Advance powder diffractometer, using a Ni-filtered Cu/Kα radiation source at 40 kV and 20 mA, from 5° to 90° with a scan rate of 4°/min. XPS data were recorded with an electron energy analyzer (ESCALAB 250Xi, Thermo Fisher Co., Waltham, MA, USA). Flash column chromatography was performed on 230–430 mesh silica gel. Analytical thin layer chromatography was performed with precoated glass baked plates (250 μ) and visualized via fluorescence and charring after treatment with potassium permanganate stain. 1H NMR and 13C NMR spectra were obtained on a Bruker Advance III HD (Zurich, Switzerland) 400 MHz spectrometer and referenced to CDCl3 (7.26 ppm for 1H, and 77.1 ppm for 13C) or DMSO-d6 (2.50 ppm for 1H, and 39.5 ppm for 13C) with tetramethylsilane as the internal standard (0 ppm).

3.2. Chemicals

The following chemicals were used as received: methyl indole-5-carboxylate (>98%, Adamas, Shanghai, China), methyl 2-bromoacetate (>98%, Adamas), anhydrous potassium carbonate (K2CO3, >99%, Sinopharm, Beijing, China), anhydrous magnesium sulfate (MgSO4, >99%, Sinopharm), acetonitrile (MeCN, 99.9%, Sigma, Shanghai, China), LiOH·H2O (>95%, Sinopharm, Beijing, China), Tetrahydrofuran (THF, 99.9%, Sigma, Shanghai, China), NiCl2.6H2O (Adamas, Shanghai, China), N,N-dimethylformamide (DMF, 99.9%, Sigma, Shanghai, China), ethanol (99%, Sigma, Shanghai, China), KOtBu (>98%, Sinopharm, Beijing, China), 2-oxindole derivative (>98%, Adamas, Shanghai, China), benzyl alcohol derivatives (>99%, Adamas, Shanghai, China), o-aminobenzyl alcohol (99%, Sigma, Shanghai, China), acetophenone derivatives (>99%, Adamas, Shanghai, China). Unless otherwise noted, all commercial reagents and solvents were obtained from the commercial provider and used without further purification.

3.3. Catalyst Preparation

To a 500 mL round-bottom bottle, we successively added methyl indole-5-carboxylate (3.504 g, 20 mmol), methyl 2-bromoacetate (7.649 g, 50 mmol), anhydrous potassium carbonate (11.057 g, 80 mmol), and MeCN (100 mL). The reaction mixture was refluxed for 20 h and monitored via thin layer chromatography (TLC) until the complete disappearance of methyl indole-5-carboxylate was confirmed. Then, the mixture was extracted with water and ethyl acetate. The combined organic phases were washed and dried over anhydrous MgSO4. The solvent was removed using a rotary evaporator under reduced pressure, and the crude product was purified via column chromatography on silica gel to afford methyl 1-(2-methoxy-2-oxoethyl)-1H-indole-5-carboxylate in 88% yield as a pale yellow solid. 1H NMR (400 MHz, CDCl3) δ 8.43 (s, 1H), 7.96 (d, J = 10.3 Hz, 1H), 7.28 (d, J = 6.8 Hz, 1H), 7.17 (d, J = 3.3 Hz, 1H), 6.69 (d, J = 4.0 Hz, 1H), 4.91 (s, 2H), 3.95 (s, 3H), 3.78 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 168.58, 168.04, 138.98, 129.83, 128.22, 124.13, 123.56, 122.16, 108.64, 104.08, 52.65, 51.85, 47.79.
To a mixed solution of THF–water (v/v = 1:1) (200 mL) and LiOH·H2O (40 mL), we successively added 1-(2-methoxy-2-oxoethyl)-1H-indole-5-carboxylate (20 mmol) in THF/water (v/v = 1:1) (200 mL) at room temperature and at 800 rpm. After stirring for 4 h, the reaction system was acidified to pH = 2 through adding an aqueous solution of 1M HCl. The product, 1-(carboxymethyl)-1H-indole-5-carboxylic acid (H2CICA), was washed with water, then evaporated to dry, obtaning a pale yellow solid in 95% yield (Scheme 10). 1H NMR (400 MHz, DMSO-d6) δ 12.58 (s, 2H), 8.04 (m, 2H), 7.48 (d, J = 7.0 Hz, 1H), 7.22 (m, 2H), 5.13 (s, 2H). 13C NMR (101 MHz, DMSO-d6) δ 170.32, 166.05, 137.45, 136.82, 126.74, 122.81, 121.81, 121.19, 111.14, 107.46, 47.80.
A mixture of NiCl2.6H2O (0.237 g, 1.0 mmol) and H2CIA (0.219 g, 1.0 mmol) was dissolved in a 5:2:1 (v/v/v) DMF/ethanol/water mixture of (8 mL) in a 15 mL Teflon reactor. The mixture was heated in an oven at 110 °C for 48 h and then cooled to room temperature. The obtained mixture was filtered, successively washed with DMF/ethanol/water and ethanol twice, and finally heated at 80 °C for 5 h to produce a green powder (Ni-CIA).

3.4. Alkylation of 2-Oxindole with Alcohols

A mixture of Ni-CIA (1.0 mol%), 1 (0.5 mmol), 2 (1.0 mmol), KOtBu (2.0 equiv.), and toluene (2.0 mL) was stirred in a 20 mL Schlenk tube under N2 at 120 °C for 6 h. The reaction mixture was cooled down to room temperature, then removed via filtration to obtain the residue. The residue was then directly purified via column chromatography with petroleum ether/ethyl acetate as eluent to produce the desired product.

3.5. Reaction of Synthesis of Quinoline Derivatives

A mixture of Ni-CIA (1.0 mol%), 4 (0.5 mmol), 5 (0.6 mmol), Cs2CO3 (1.0 equiv.) and toluene (2.0 mL) was stirred in a 20 mL Schlenk tube under 100 °C for 8 h. The reaction mixture was cooled down to room temperature, then removed via filtration to produce the residue. The residue was then directly purified via column chromatography with petroleum ether/ethyl acetate as eluent to yield the desired product.

4. Conclusions

In summary, we herein report a simple, effective, and reusable heterogeneous coordination polymer Ni-CIA catalyst for promoting the synthesis of oxindole and quinoline derivatives through the borrowing hydrogen strategy. The slice layer stacked combination structure of the catalyst can effectively promote ion diffusion and accelerate catalytic mass transfer during the catalytic reaction. Moreover, various primary alcohol and acetophenone substrates were tested under optimized reaction conditions to form the desired compounds in moderate to excellent yields. After that, mechanism explorations including control experiments were conducted to clarify the effect of different substituents, and cooperative effects were investigated to clarify these transformations. The above experimental results confirmed that catalysis with Ni-CIA facilitates the dehydrogenation of the benzyl alcohol 2a and o-aminobenzyl alcohol 4a, revealing that the reaction followed the hydrogen-borrowing mechanism; thus, we proposed possible reaction mechanisms for the two types of reactions. Finally, the recycling experiments showed that the Ni-CIA catalyst had fine stability. In particular, it was shown that no significant reduction in activity and no significant changes in surface morphology were observed for Ni-CIA after five runs. This catalytic system is shown as a much greener method for the synthesis of oxindole and quinoline derivatives, in accordance with the principles of sustainable development and atom economy, making it highly promising for organic catalysis.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13081195/s1, Figure S1: (a) and (b) SEM image of Ni-CIA after being recycled five times, Analytical data of the obtained compounds. References [69,70] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, J.L.; methodology, J.L. and J.T.; validation, J.L., J.T. and L.W.; formal analysis, J.L. and J.T.; investigation, J.T. and L.W.; resources, D.W. and Z.-C.D.; writing—original draft preparation, J.L.; writing—review and editing, D.W.; visualization, D.W. and Z.-C.D.; supervision, D.W. and Z.-C.D.; project administration, D.W.; funding acquisition, D.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 21861039).

Acknowledgments

We gratefully acknowledge financial support of this work by the National Natural Science Foundation of China (21861039) and Central Laboratory, School of Chemical and Material Engineering, Jiangnan University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jiang, T.; Kuhen, K.L.; Wolff, K.; Yin, H.; Bieza, K.; Caldwell, J.; Bursulaya, B.; Wu, T.Y.; He, Y. Design, synthesis and biological evaluations of novel oxindoles as HIV-1 non-nucleoside reverse transcriptase inhibitors. Part I. Bioorg. Med. Chem. Lett. 2006, 16, 2105–2108. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, H.; Chen, M.; Wang, L. Selective Synthesis and Utility of One Tripyrrolic Compound and Its Intermediates. Chem. Pharm. Bull. 2007, 55, 1439–1441. [Google Scholar] [CrossRef] [Green Version]
  3. Reisman, S.E.; Ready, J.M.; Weiss, M.M.; Hasuoka, A.; Hirata, M.; Tamaki, K.; Ovaska, T.V.; Smith, C.J.; Wood, J.L. Evolution of a Synthetic Strategy:  Total Synthesis of (±)-Welwitindolinone A Isonitrile. J. Am. Chem. Soc. 2008, 130, 2087–2100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Rodrigues, T.; Reker, D.; Schneider, P.; Schneider, G. Counting on natural products for drug design. Nat. Chem. 2016, 8, 531–541. [Google Scholar] [CrossRef]
  5. Huang, L.; Yang, L.; Wan, J.P.; Zhou, L.; Liu, Y.; Hao, G. Metal-free three-component assemblies of anilines, alpha-keto acids and alkyl lactates for quinoline synthesis and their anti-inflammatory activity. Org. Biomol. Chem. 2022, 20, 4385–4390. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, H.; Zeng, X.-Q.; Zuo, H.-W.; Zhang, N.; Huang, Z.-Y. Design, Synthesis and Seed Germination Inhibition Activity of Quinoline-6-sulfonamide Compounds. Chin. J. Org. Chem. 2022, 42, 2947–2953. [Google Scholar] [CrossRef]
  7. Moutaouakil, M.; Tighadouini, S.; Almarhoon, Z.M.; Al-Zaben, M.I.; Bacha, A.B.; Masand, V.H.; Jamaleddine, J.; Saddik, R. Quinoline Derivatives with Different Functional Groups: Evaluation of Their Catecholase Activity. Catalysts 2022, 12, 1468. [Google Scholar] [CrossRef]
  8. Hu, C.; Lu, L.; Wan, J.P.; Wen, C. The Pharmacological Mechanisms and Therapeutic Activities of Hydroxychloroquine in Rheumatic and Related Diseases. Curr. Med. Chem. 2017, 24, 2241–2249. [Google Scholar] [CrossRef]
  9. Trost, B.M.; Zhang, Y. Molybdenum-Catalyzed Asymmetric Allylation of 3-Alkyloxindoles: Application to the Formal Total Synthesis of (−)-Physostigmine. J. Am. Chem. Soc. 2006, 128, 4590–4591. [Google Scholar] [CrossRef]
  10. Kende, A.S.; Hodges, J.C. Regioselective C-3 Alkylations of Oxindole Dianion. Synth. Commun. 2007, 12, 1–10. [Google Scholar] [CrossRef]
  11. Chaudhari, M.B.; Sutar, Y.; Malpathak, S.; Hazra, A.; Gnanaprakasam, B. Transition-Metal-Free C-H Hydroxylation of Carbonyl Compounds. Org. Lett. 2017, 19, 3628–3631. [Google Scholar] [CrossRef] [PubMed]
  12. Chakraborty, S.; Daw, P.; David, Y.B.; Milstein, D. Manganese-Catalyzed alpha-Alkylation of Ketones, Esters, and Amides Using Alcohols. ACS Catal. 2018, 8, 10300–10305. [Google Scholar] [CrossRef]
  13. Duan, Y.T.; Wang, Z.X. Transition-Metal-Free α-C-Alkylation of Oxindoles with Alcohols. Asian J. Org. Chem. 2023, 12, e202300015. [Google Scholar] [CrossRef]
  14. Luo, X.; Tian, A.; Pei, M.; Yan, J.; Liu, X.; Wang, L. Highly Stable Univalent Copper of a Cu@Al/SBA-15 Nanocomposite Catalyzes the Synthesis of Fluorescent Aminobenzotriazoles Derivatives. Chem. Eur. J. 2022, 28, e202103361. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, D.; Xia, X.; Hu, W.; Yang, Y.; Li, J. Catalytic Synthesis of Pyrazine and Ketone Derivatives by Unsymmetrical Triazolyl-Naphthyridinyl-Pyridine Copper. Chin. J. Org. Chem. 2022, 42, 190–199. [Google Scholar] [CrossRef]
  16. Pena-Lopez, M.; Piehl, P.; Elangovan, S.; Neumann, H.; Beller, M. Manganese-Catalyzed Hydrogen-Autotransfer C-C Bond Formation: Alpha-Alkylation of Ketones with Primary Alcohols. Angew. Chem. Int. Ed. Engl. 2016, 55, 14967–14971. [Google Scholar] [CrossRef]
  17. Li, J.; Mao, A.; Yao, W.; Zhu, H.; Wang, D. Iridium supported on porous polypyridine-oxadiazole as high-activity and recyclable catalyst for the borrowing hydrogen reaction. Green Chem. 2022, 24, 2602–2612. [Google Scholar] [CrossRef]
  18. Wang, D.; Ge, B.; Zhang, B.; Liu, H.; Li, J. Synthesis of Supported Indazolyl-Pyridyl-Quinoline Iridium Catalyst and Its Application to N-Alkylation of 2-Aminobenzothiazoles. Chin. J. Org. Chem. 2022, 42, 619–630. [Google Scholar] [CrossRef]
  19. Banik, A.; Datta, P.; Mandal, S.K. C-Alkylation by Phenalenyl-Based Molecule via a Borrowing Hydrogen Pathway. Org. Lett. 2023, 25, 1305–1309. [Google Scholar] [CrossRef]
  20. Chen, X.; Liu, D.; Yang, C.; Shi, L.; Li, F. Hexaazatrinaphthalene-Based Covalent Triazine Framework-Supported Rhodium(III) Complex: A Recyclable Heterogeneous Catalyst for the Reductive Amination of Ketones to Primary Amines. Inorg. Chem. 2023, 62, 9360–9368. [Google Scholar] [CrossRef]
  21. Liu, Y.; Diao, H.; Hong, G.; Edward, J.; Zhang, T.; Yang, G.; Yang, B.M.; Zhao, Y. Iridium-Catalyzed Enantioconvergent Borrowing Hydrogen Annulation of Racemic 1,4-Diols with Amines. J. Am. Chem. Soc. 2023, 145, 5007–5016. [Google Scholar] [CrossRef]
  22. Ng, T.W.; Tao, R.; See, W.W.L.; Poh, S.B.; Zhao, Y. Economical Access to Diverse Enantiopure Tetrahydropyridines and Piperidines Enabled by Catalytic Borrowing Hydrogen. Angew. Chem. Int. Ed. 2023, 62, e202212528. [Google Scholar] [CrossRef] [PubMed]
  23. Hamid, M.H.S.A.; Slatford, P.A.; Williams, J.M.J. Borrowing Hydrogen in the Activation of Alcohols. Adv. Synth. Catal. 2007, 349, 1555–1575. [Google Scholar] [CrossRef]
  24. Nakayama, T.; Harada, S.; Kikkawa, S.; Hikawa, H.; Azumaya, I. Palladium-Catalyzed Dehydrogenative Synthesis of Imidazoquinolines in Water. Asian J. Org. Chem. 2022, 11, e202200510. [Google Scholar] [CrossRef]
  25. Putra, A.E.; Takigawa, K.; Tanaka, H.; Ito, Y.; Oe, Y.; Ohta, T. Transition-Metal-Catalyzed Regioselective Alkylation of Indoles with Alcohols. Eur. J. Org. Chem. 2013, 2013, 6344–6354. [Google Scholar] [CrossRef]
  26. Yang, Q.; Wang, Q.; Yu, Z. Substitution of alcohols by N-nucleophiles via transition metal-catalyzed dehydrogenation. Chem. Soc. Rev. 2015, 44, 2305–2329. [Google Scholar] [CrossRef]
  27. Huang, F.; Liu, Z.; Yu, Z. C-Alkylation of Ketones and Related Compounds by Alcohols: Transition-Metal-Catalyzed Dehydrogenation. Angew. Chem. Int. Ed. 2016, 55, 862–875. [Google Scholar] [CrossRef]
  28. Corma, A.; Navas, J.; Sabater, M.J. Advances in One-Pot Synthesis through Borrowing Hydrogen Catalysis. Chem. Rev. 2018, 118, 1410–1459. [Google Scholar] [CrossRef]
  29. Alig, L.; Fritz, M.; Schneider, S. First-Row Transition Metal (De)Hydrogenation Catalysis Based on Functional Pincer Ligands. Chem. Rev. 2019, 119, 2681–2751. [Google Scholar] [CrossRef]
  30. Irrgang, T.; Kempe, R. 3d-Metal Catalyzed N- and C-Alkylation Reactions via Borrowing Hydrogen or Hydrogen Autotransfer. Chem. Rev. 2019, 119, 2524–2549. [Google Scholar] [CrossRef]
  31. Reed-Berendt, B.G.; Latham, D.E.; Dambatta, M.B.; Morrill, L.C. Borrowing Hydrogen for Organic Synthesis. ACS Cent. Sci. 2021, 7, 570–585. [Google Scholar] [CrossRef]
  32. Volk, B.; Mezei, T.; Simig, G. Raney Nickel-Induced 3-Alkylation of Oxindole with Alcohols and Diols. Synthesis 2002, 2002, 595–597. [Google Scholar] [CrossRef]
  33. Grigg, R.; Whitney, S.; Sridharan, V.; Keep, A.; Derrick, A. Iridium catalysed C-3 alkylation of oxindole with alcohols under solvent free thermal or microwave conditions. Tetrahedron 2009, 65, 4375–4383. [Google Scholar] [CrossRef]
  34. Jensen, T.; Madsen, R. Ruthenium-Catalyzed Alkylation of Oxindole with Alcohols. J. Org. Chem. 2009, 74, 3990–3992. [Google Scholar] [CrossRef]
  35. Bisht, G.S.; Chaudhari, M.B.; Gupte, V.S.; Gnanaprakasam, B. Ru-NHC Catalyzed Domino Reaction of Carbonyl Compounds and Alcohols: A Short Synthesis of Donaxaridine. ACS Omega 2017, 2, 8234–8252. [Google Scholar] [CrossRef] [Green Version]
  36. Dambatta, M.B.; Polidano, K.; Northey, A.D.; Williams, J.M.J.; Morrill, L.C. Iron-Catalyzed Borrowing Hydrogen C-Alkylation of Oxindoles with Alcohols. ChemSusChem 2019, 12, 2345–2349. [Google Scholar] [CrossRef] [Green Version]
  37. Saini, P.; Dolui, P.; Nair, A.; Verma, A.; Elias, A.J. A Bench-stable 8-Aminoquinoline Derived Phosphine-free Manganese (I)-Catalyst for Environmentally Benign C(alpha)-Alkylation of Oxindoles with Secondary and Primary Alcohols. Chem. Asian J. 2023, 18, e202201148. [Google Scholar] [CrossRef]
  38. Wu, Q.; Pan, L.; Du, G.; Zhang, C.; Wang, D. Preparation of pyridyltriazole ruthenium complexes as effective catalysts for the selective alkylation and one-pot C–H hydroxylation of 2-oxindole with alcohols and mechanism exploration. Org. Chem. Front. 2018, 5, 2668–2675. [Google Scholar] [CrossRef]
  39. Liu, G.; Huang, T.; Zhang, Y.; Liang, X.; Li, Y.; Li, H. C-3 alkylation of oxindole with alcohols catalyzed by an indene-functionalized mesoporous iridium catalyst. Catal. Commun. 2011, 12, 655–659. [Google Scholar] [CrossRef]
  40. Chaudhari, C.; Siddiki, S.M.A.H.; Kon, K.; Tomita, A.; Tai, Y.; Shimizu, K. C-3 alkylation of oxindole with alcohols by Pt/CeO2 catalyst in additive-free conditions. Catal. Sci. Technol. 2014, 4, 1064–1069. [Google Scholar] [CrossRef]
  41. Putra, A.E.; Oe, Y.; Ohta, T. Pd/C-Catalyzed Alkylation of Heterocyclic Nucleophiles with Alcohols through the “Borrowing Hydrogen” Process. Eur. J. Org. Chem. 2015, 2015, 7799–7805. [Google Scholar] [CrossRef]
  42. Fujita, S.; Imagawa, K.; Yamaguchi, S.; Yamasaki, J.; Yamazoe, S.; Mizugaki, T.; Mitsudome, T. A nickel phosphide nanoalloy catalyst for the C-3 alkylation of oxindoles with alcohols. Sci. Rep. 2021, 11, 10673. [Google Scholar] [CrossRef] [PubMed]
  43. Li, J.; Yu, M.; Duan, Z.-C.; Zhu, H.; Yao, W.; Wang, D. Porous cross-linked polymer copper and iridium catalyzed the synthesis of quinoxalines and functionalized ketones under solvent-free conditions. Mater. Chem. Front. 2021, 5, 7861–7872. [Google Scholar] [CrossRef]
  44. Zhao, S.; Wang, Y.; Dong, J.; He, C.-T.; Yin, H.; An, P.; Zhao, K.; Zhang, X.; Gao, C.; Zhang, L.; et al. Ultrathin metal–organic framework nanosheets for electrocatalytic oxygen evolution. Nat. Energy 2016, 1, 16184. [Google Scholar] [CrossRef]
  45. Stoica, A.-C.; Damoc, M.; Zaltariov, M.-F.; Racles, C.; Cazacu, M. Two-dimensional coordination polymers containing permethylated motifs-promising candidates for 2D emerging materials. Structural, behavioral and functional particularities. React. Funct. Polym. 2021, 168, 105039. [Google Scholar] [CrossRef]
  46. Farid, S.; Mao, Q.; Ren, S.; Hao, C.; Dong, X. Promoting the Oxygen Evolution Reaction via Morphological Manipulation of a Lamellar Nanorod-Assembled Ni(II)-Pyrazolate Superstructure. ACS Appl. Mater. Interfaces 2022, 14, 47775–47787. [Google Scholar] [CrossRef]
  47. Duan, J.; Jin, W.; Kitagawa, S. Water-resistant porous coordination polymers for gas separation. Coord. Chem. Rev. 2017, 332, 48–74. [Google Scholar] [CrossRef] [Green Version]
  48. Griffin, S.E.; Domecus, G.P.; Flores, C.E.; Sikma, R.E.; Benz, L.; Cohen, S.M. A Coordination Polymer of Vaska’s Complex as a Heterogeneous Catalyst for the Reductive Formation of Enamines from Amides. Angew. Chem. Int. Ed. 2023, 62, e202301611. [Google Scholar] [CrossRef]
  49. Gui, S.; Tang, W.; Huang, Z.; Wang, X.; Gui, S.; Gao, X.; Xiao, D.; Tao, L.; Jiang, Z.; Wang, X. Ultrasmall Coordination Polymer Nanodots Fe-Quer Nanozymes for Preventing and Delaying the Development and Progression of Diabetic Retinopathy. Adv. Electron. Mater. 2023, 300261. [Google Scholar] [CrossRef]
  50. Hu, X.; Zhu, H.; Sang, X.; Wang, D. Design and Synthesis of Zirconium-Containing Coordination Polymer Based on Unsymmetric Indolyl Dicarboxylic Acid and Catalytic Application on Borrowing Hydrogen Reaction. Adv. Synth. Catal. 2018, 360, 4293–4300. [Google Scholar] [CrossRef]
  51. Sang, X.; Hu, X.; Tao, R.; Zhang, Y.; Zhu, H.; Wang, D. A Zirconium Indazole Carboxylate Coordination Polymer as an Efficient Catalyst for Dehydrogenation-Cyclization and Oxidative Coupling Reactions. ChemPlusChem 2020, 85, 123–129. [Google Scholar] [CrossRef] [Green Version]
  52. Tao, R.; Yang, Y.; Zhu, H.; Hu, X.; Wang, D. Ligand-tuned cobalt-containing coordination polymers and applications in water. Green Chem. 2020, 22, 8452–8461. [Google Scholar] [CrossRef]
  53. Chang, S.; Liu, H.; Shi, G.; Xia, X.-F.; Wang, D.; Duan, Z.-C. Copper–cobalt coordination polymers and catalytic applications on borrowing hydrogen reactions. New J. Chem. 2022, 46, 15929–15936. [Google Scholar] [CrossRef]
  54. Han, M.-R.; Kong, X.-J.; Feng, S.-S.; Lu, L.-P. Synthesis, Structure and Magnetic Studies of a Ni(II) Coordination Polymer Constructed from Asymmetrical Tetracarboxylic Acid and N-Donor Ligand. J. Clust. Sci. 2021, 33, 809–814. [Google Scholar] [CrossRef]
  55. Xu, J.; He, S.; Zhang, H.; Huang, J.; Lin, H.; Wang, X.; Long, J. Layered metal–organic framework/graphene nanoarchitectures for organic photosynthesis under visible light. J. Mater. Chem. A 2015, 3, 24261–24271. [Google Scholar] [CrossRef]
  56. Osadchii, D.Y.; Olivos-Suarez, A.I.; Bavykina, A.V.; Gascon, J. Revisiting Nitrogen Species in Covalent Triazine Frameworks. Langmuir 2017, 33, 14278–14285. [Google Scholar] [CrossRef] [Green Version]
  57. Jia, X.; Zhang, B.; Chen, C.; Fu, X.; Huang, Q. Immobilization of chitosan grafted carboxylic Zr-MOF to porous starch for sulfanilamide adsorption. Carbohydr. Polym. 2021, 253, 117305. [Google Scholar] [CrossRef]
  58. Wu, H.; Li, S.; Liu, Y.; Liu, M.; Shi, Y. Metal–Organic Framework-Derived Hollow Carbon Nanosphere@Ni/Reduced Graphene Oxide Composites for Supercapacitor Electrodes with Enhanced Performance. ACS Appl. Nano Mater. 2023, 6, 1582–1591. [Google Scholar] [CrossRef]
  59. Shang, X.F.; Morris-Natschke, S.L.; Yang, G.Z.; Liu, Y.Q.; Guo, X.; Xu, X.S.; Goto, M.; Li, J.C.; Zhang, J.Y.; Lee, K.H. Biologically active quinoline and quinazoline alkaloids part II. Med. Res. Rev. 2018, 38, 1614–1660. [Google Scholar] [CrossRef]
  60. Juvale, K.; Gallus, J.; Wiese, M. Investigation of quinazolines as inhibitors of breast cancer resistance protein (ABCG2). Bioorg. Med. Chem. 2013, 21, 7858–7873. [Google Scholar] [CrossRef]
  61. Michael, J.P. Quinoline, quinazoline and acridone alkaloids. Nat. Prod. Rep. 2002, 19, 742–760. [Google Scholar] [CrossRef] [PubMed]
  62. Chen, Z.; Song, J.; Peng, X.; Xi, S.; Liu, J.; Zhou, W.; Li, R.; Ge, R.; Liu, C.; Xu, H.; et al. Iron Single Atom Catalyzed Quinoline Synthesis. Adv. Mater. 2021, 33, e2101382. [Google Scholar] [CrossRef] [PubMed]
  63. Xu, X.; Ai, Y.; Wang, R.; Liu, L.; Yang, J.; Li, F. Ruthenium-catalyzed acceptorless dehydrogenative coupling of o-aminobenzyl alcohols with ketones to quinolines in the presence of carbonate salt. J. Catal. 2021, 395, 340–349. [Google Scholar] [CrossRef]
  64. Yu, K.; Chen, Q.; Liu, W. Iron-catalysed quinoline synthesis via acceptorless dehydrogenative coupling. Org. Chem. Front. 2022, 9, 6573–6578. [Google Scholar] [CrossRef]
  65. Hao, Z.; Zhou, X.; Ma, Z.; Zhang, C.; Han, Z.; Lin, J.; Lu, G.L. Dehydrogenative Synthesis of Quinolines and Quinazolines via Ligand-Free Cobalt-Catalyzed Cyclization of 2-Aminoaryl Alcohols with Ketones or Nitriles. J. Org. Chem. 2022, 87, 12596–12607. [Google Scholar] [CrossRef]
  66. Chen, X.; Ai, Y.; Liu, D.; Liu, P.; Xu, X.; Yang, J.; Li, F. A recyclable covalent triazine framework-supported iridium(iii) terpyridine complex for the acceptorless dehydrogenative coupling of o-aminobenzyl alcohols with ketones to form quinolines. Mater. Chem. Front. 2022, 6, 1228–1235. [Google Scholar] [CrossRef]
  67. Yang, L.; Wan, J.-P. Ethyl lactate-involved three-component dehydrogenative reactions: Biomass feedstock in diversity-oriented quinoline synthesis. Green Chem. 2020, 22, 3074–3078. [Google Scholar] [CrossRef]
  68. Li, R.; Jiang, S.; Zheng, H.; Lei, H.; Huang, Z.; Chen, S.; Deng, G.-J. Iron-catalyzed indolo[2,3-c]quinoline synthesis from nitroarenes and benzylic alcohols/aldehydes promoted by elemental sulfur. Green Synth. Catal. 2022, 3, 95–101. [Google Scholar] [CrossRef]
  69. Ghosh, A.; Bera, A.; Banerjee, D. Nickel Catalyzed Alkylation of Oxindoles with Alkyl Alcohols. ChemCatChem 2023, 15, e202201433. [Google Scholar] [CrossRef]
  70. Jana, A.; Kumar, A.; Maji, B. Manganese catalyzed C-alkylation of methyl N-heteroarenes with primary alcohols. Chem. Commun. 2021, 57, 3026–3029. [Google Scholar] [CrossRef]
Scheme 1. Examples of biologically active C3-substituted oxindoles.
Scheme 1. Examples of biologically active C3-substituted oxindoles.
Catalysts 13 01195 sch001
Figure 1. (a) FT−IR spectra of H2CIA and Ni-CIA; (b) XRD pattern of Ni-CIA.
Figure 1. (a) FT−IR spectra of H2CIA and Ni-CIA; (b) XRD pattern of Ni-CIA.
Catalysts 13 01195 g001
Figure 2. SEM image of (a,b) Ni-CICA.
Figure 2. SEM image of (a,b) Ni-CICA.
Catalysts 13 01195 g002
Figure 3. (a) High-resolution C 1s spectrum, (b) high-resolution N 1s spectrum, (c) high-resolution O 1s spectrum, and (d) high-resolution Ni 2p spectrum.
Figure 3. (a) High-resolution C 1s spectrum, (b) high-resolution N 1s spectrum, (c) high-resolution O 1s spectrum, and (d) high-resolution Ni 2p spectrum.
Catalysts 13 01195 g003
Figure 4. (a) EDS pattern of Ni-CIA. (b) TGA pattern of Ni-CIA.
Figure 4. (a) EDS pattern of Ni-CIA. (b) TGA pattern of Ni-CIA.
Catalysts 13 01195 g004
Scheme 2. Alkylation of 2-oxindole with alcohols a,b. a Reaction conditions: 1a (0.5 mmol), 2a (1.0 mmol), Ni-CIA (1.0 mol%), KOtBu (2.0 equiv.), and toluene (2 mL), N2,6 h, 120 °C. b Isolated yields.
Scheme 2. Alkylation of 2-oxindole with alcohols a,b. a Reaction conditions: 1a (0.5 mmol), 2a (1.0 mmol), Ni-CIA (1.0 mol%), KOtBu (2.0 equiv.), and toluene (2 mL), N2,6 h, 120 °C. b Isolated yields.
Catalysts 13 01195 sch002
Scheme 3. Ni-CIA-catalyzed reaction of acetophenones with o-aminobenzyl alcohol a,b. a Reaction conditions: 4a (0.5 mmol), 5 (0.6 mmol), Ni-CIA (1.0 mol%), Cs2CO3 (1.0 equiv.), and toluene (2 mL), 100 °C, 8 h. b Isolated yields.
Scheme 3. Ni-CIA-catalyzed reaction of acetophenones with o-aminobenzyl alcohol a,b. a Reaction conditions: 4a (0.5 mmol), 5 (0.6 mmol), Ni-CIA (1.0 mol%), Cs2CO3 (1.0 equiv.), and toluene (2 mL), 100 °C, 8 h. b Isolated yields.
Catalysts 13 01195 sch003
Scheme 4. The control experiments.
Scheme 4. The control experiments.
Catalysts 13 01195 sch004
Scheme 5. Mechanism exploration experiments.
Scheme 5. Mechanism exploration experiments.
Catalysts 13 01195 sch005
Scheme 6. Mechanism exploration experiments for the synthesis of quinolines.
Scheme 6. Mechanism exploration experiments for the synthesis of quinolines.
Catalysts 13 01195 sch006
Scheme 7. Proposed possible reaction mechanism.
Scheme 7. Proposed possible reaction mechanism.
Catalysts 13 01195 sch007
Scheme 8. Proposed possible reaction mechanism for the synthesis of quinolines.
Scheme 8. Proposed possible reaction mechanism for the synthesis of quinolines.
Catalysts 13 01195 sch008
Scheme 9. Recycling experiments.
Scheme 9. Recycling experiments.
Catalysts 13 01195 sch009
Scheme 10. The synthesis of H2CICA.
Scheme 10. The synthesis of H2CICA.
Catalysts 13 01195 sch010
Table 1. Optimization of reaction conditions for the alkylation of 2-oxindole a.
Table 1. Optimization of reaction conditions for the alkylation of 2-oxindole a.
Catalysts 13 01195 i001
EntryCatalystBaseTime [h]SolventYield [%] b
1-KOtBu6toluene<5
2Ni-CIANEt36toluene17
3Ni-CIAK2CO36toluene34
4Ni-CIACs2CO36toluene65
5Ni-CIAKOtBu6toluene84
6Ni-CIAKOH6toluene72
7Ni-CIAK3PO46toluene56
8Ni-CIAKOtBu6H2O<5
9Ni-CIAKOtBu61,4-dioxane76
10Ni-CIAKOtBu6xylene80
11Ni-CIAKOtBu8DMF12
12Ni-CIAKOtBu2toluene63
13Ni-CIAKOtBu6toluene82
14 cNi-CIAKOtBu6toluene88
15 dNi-CIAKOtBu6toluene43
a Reaction conditions: 1a (0.5 mmol), 2a (1.0 mmol), catalyst (1.0 mol%), base (2.0 equiv.), and solvent (2 mL), 120 °C. b Isolated yields. c N2. d O2.
Table 2. Catalyst investigation experiments a,b.
Table 2. Catalyst investigation experiments a,b.
Catalysts 13 01195 i002
EntryCatalystYield [%] bEntryCatalystYield [%] b
1NiCl2 · 6H2O155NiCl2(bipy)232
2NiBr2 · 3H2O<56NiCl2(phen)245
3Ni(NO3)2 · 6H2O<57Ni-CIA88
4Ni(OAc)2 · 6H2O<58none<5
a Reaction conditions: 1a (0.5 mmol), 2a (1.0 mmol), catalyst (1.0 mol%), KOtBu (2.0 equiv.), and toluene (2 mL), N2, 6 h, 120 °C. b Isolated yields.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, J.; Tang, J.; Wang, L.; Wang, D.; Duan, Z.-C. Recyclable Ni-Containing Coordination Polymer as an Efficient Catalyst for the Synthesis of Oxindole and Quinoline Derivatives through the Borrowing Hydrogen Strategy. Catalysts 2023, 13, 1195. https://doi.org/10.3390/catal13081195

AMA Style

Li J, Tang J, Wang L, Wang D, Duan Z-C. Recyclable Ni-Containing Coordination Polymer as an Efficient Catalyst for the Synthesis of Oxindole and Quinoline Derivatives through the Borrowing Hydrogen Strategy. Catalysts. 2023; 13(8):1195. https://doi.org/10.3390/catal13081195

Chicago/Turabian Style

Li, Jiahao, Jiajie Tang, Likui Wang, Dawei Wang, and Zheng-Chao Duan. 2023. "Recyclable Ni-Containing Coordination Polymer as an Efficient Catalyst for the Synthesis of Oxindole and Quinoline Derivatives through the Borrowing Hydrogen Strategy" Catalysts 13, no. 8: 1195. https://doi.org/10.3390/catal13081195

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop