Graphitic Carbon Nitride Nanosheets Covalently Functionalized with

Jul 23, 2019 - (42−45) Because of the remarkable aforementioned properties of ..... were performed using a Thermo Finnigan Flash EA 1112 Element ...
0 downloads 0 Views 5MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2019, 4, 12544−12554

http://pubs.acs.org/journal/acsodf

Graphitic Carbon Nitride Nanosheets Covalently Functionalized with Biocompatible Vitamin B1: Synthesis, Characterization, and Its Superior Performance for Synthesis of Quinoxalines Afsaneh Rashidizadeh, Hossein Ghafuri,* Hamid Reza Esmaili Zand, and Nahal Goodarzi Catalysts and Organic Synthesis Research Laboratory, Department of Chemistry, Iran University of Science and Technology, Tehran 16846-13114, Iran Downloaded via 81.22.47.234 on July 25, 2019 at 00:24:46 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: The physical properties of two-dimensional nanosheet materials make them promising candidates as active materials in the areas of photoelectronics, fuel cells, sensors, water splitting, solar energy conversion, CO2 reduction, and heterogeneous catalysis. Among two-dimensional nanosheet materials, graphitic carbon nitride due to its electronic structure and high chemical and thermal stability possesses unique properties. Covalent functionalization of graphitic carbon nitride could be the key step in modifying its ability and significantly improving its properties. To this purpose, a novel strategy for the covalent functionalization of g-C3N4 nanosheets (CN) with vitamin B1 (VB1) by using 1,3dibromopropane as a covalent linker for the first time is demonstrated. The obtained CN-Pr-VB1 exhibits increased thermal stability compared to the VB1 which is important in the practice application and can be easily dispersed in common organic solvents. The efficacy of the CN-Pr-VB1 as a heterogeneous organocatalyst was evaluated in the quinoxaline synthesis under solvent-free conditions and afforded good isolated yield with high purity. Moreover, the prepared catalyst could be facilely recycled and reused for seven consecutive cycles without a noticeable decrease in the catalytic activity. Extensive characterization confirmed the stability of morphology and chemical structure after recyclability of the CN-Pr-VB1.



and stability.11 To date, the development of two-dimensional (2D) nanosheet materials with molecular thickness has attracted worldwide attention owing to fascinating physicochemical properties.12 Among various 2D layered systems, graphitic carbon nitride (g-C3N4) as an important class of metal-free conjugated polymers with weak van der Waals forces between layers possesses unique electronic band structure, high chemical and thermal stability, large surface area, and environmentally friendly property and nontoxicity.13,14 Therefore, g-C3N4 sheets has been investigated in the areas of photoelectronics, fuel cells, sensors, water splitting, solar energy conversion, CO2 reduction, and heterogeneous catalysis.15−25 Covalent functionalization of g-C3N4 sheets has seldom been used in catalytic applications.26−30 Thus, the utilization of g-C3N4 sheets as a potential support material is still worthy and desirable to explore new heterogeneous catalysts. On the basis of the above-mentioned properties of VB1, gC3N4 sheets, and continuing our efforts toward the design of greener and efficient heterogeneous catalysts for the synthesis

INTRODUCTION Nowadays with the growing severity of energy crisis and environmental pollution, the development of a green sustainable catalytic process is considered as one of the promising methods for a resolution of these problems.1 For this purpose, organocatalysts and biocatalysts based on their advantages such as nontoxicity, ready availability, low cost, stability, and biodegradability have been the subject of considerable interest in chemical syntheses and transformations.1−5 Since the pioneering work reported in 1960, vitamin B1 (VB1) has attracted much attention as a catalyst and an important coenzyme in biochemical transformation of α-keto acids into acetyl CoA in polyketide synthesis and formation of acyloins.6 Moreover, VB1 that consists of multifunctional groups such as hydroxyl and quaternary ammonium salt has already demonstrated its effectiveness to catalyze organic reactions such as Knoevenagel condensation, Michael addition, cyclization, and creating carbon−carbon bonds and carbon− heteroatom bonds.7−10 The focus during the past decades has been on covalent and noncovalent functionalization of solid supports with active biomolecules and organic molecules, which is an effective strategy to overcome the problems of using homogeneous catalysts such as difficult recovery from the reaction mixture and also to improve catalytic performance © 2019 American Chemical Society

Received: June 4, 2019 Accepted: July 11, 2019 Published: July 23, 2019 12544

DOI: 10.1021/acsomega.9b01635 ACS Omega 2019, 4, 12544−12554

ACS Omega

Article

of organic compounds,31−36 we report herein the first successful covalent functionalization of g-C3N4 sheets with VB1 through a facile and innovative strategy. VB1 covalently functionalized g-C3N4 sheets exhibit high thermal stability and much improved catalytic performance compared to VB1 and gC3N4 sheets in organic reactions. Quinoxaline and their derivatives are a valuable and wellestablished class of nitrogen-bearing heterocycles that have found prominence in the area of dyes, pharmaceutical, and agricultural chemistry.37−39 Some commercial antibiotics contain a quinoxaline nucleus as a fundamental scaffold like Echinomycin, Levomycin, and Actinoleutin which are active against Gram-positive bacteria and represent antitumor activity.40,41 Moreover, they have diverse applications in several fields such as significant intermediates in organic synthesis, efficient electron luminescent materials, organic semiconductors, and DNA cleaving agent.42−45 Because of the remarkable aforementioned properties of quinoxaline derivatives, many protocols have been developed for the preparation of these compounds. The conventional synthesis of quinoxalines involves the reaction between aryl 1,2-diamines and 1,2dicarbonyl compounds in refluxing ethanolic medium or acetic acid.46 The use of heterogeneous catalysts has become an essential inspiration for chemists to develop sustainable methods for the synthesis of organic compounds. In view of this, a variety of catalysts such as heterogeneous carbon-based materials, glycerol, CAN, I2, polyaniline sulfate, Montmorillonite K-10, oxalic acid, VB1, and Zn[(L)proline] have been employed to carry out this condensation.47−55 Although some of these reported methods suffer from a number of demerits including prolonged reaction times, the use of toxic organic solvents, a low yield of the product, the use of costly catalysts, and harsh reaction conditions. Hence, a great deal of interest has been focused on the development of environmentally, efficient, and recyclable benign catalytic methods for the synthesis of quinoxaline derivatives. In this work, we demonstrate the application of CN-Pr-VB1 as a heterogeneous organocatalyst for the preparation of quinoxalines via condensation of 1,2-diamines with 1,2-dicarbonyl compounds under solvent-free condition in excellent yields (Scheme 1).

and 2θ = 13.1° for (100) reflection can be observed. After exfoliation, the intensity of (002) diffraction is weakened significantly and its related peak is shifted from 27.4° to 27.8°. This change is related to shrinking of the interplanar stacking distance while the typical reflection peak of the bulk g-C3N4 at 2θ = 13.1° disappeared. The FT-IR spectroscopy of the g-C3N4 nanosheet and bulk g-C3N4 is illustrated in Figure 2. As can be seen, the characterization peaks of bulk g-C3N4 were similar to the gC3N4 nanosheet. A strong and broad peak of the N−H group (−NH2 or NH groups) appeared at 3000−3500 cm−1, stretching vibration peaks of CN were observed at 1614 and 1550 cm−1, stretching peaks of the C−N heterocycle were at around 1406, 1319, and 1234 cm−1, and a sharp peak at 808 cm−1 is due to the breathing vibration of tri-s-triazine units. Then, the g-C3N4 nanosheet was modified with 1,3dibromopropane as a covalent linker for the first time. The chemical structure of the new synthesized CN-Pr-Br was confirmed by FT-IR and energy-dispersive X-ray spectroscopy (EDS) (Figure 3a,b). From FT-IR spectra, main adsorption bands of g-C3N4 nanosheets can be observed in CN-Pr-Br, suggesting that the basic g-C3N4 chemical structure is largely retained. Energy-dispersive X-ray spectroscopy (EDS) displays the successful incorporation of the Br atoms within the framework. Intensive efforts have been devoted to evaluate the best conditions for the loading of VB1 on CN sheets, and various amounts of 1,3-dibromopropane as a linker were investigated. In this study, two different amounts of 1,3-dibromopropane, 1.01 and 2.02 mL, were studied under same conditions to investigate the immobilization of VB1 on CN sheets. CHNS elemental analysis of CN-Pr-VB1 in the presence of 1.0 g of CN sheets, 1.01 mL of 1,3-dibromopropane, and 1 mmol of VB1 shows 30.82% (C), 3.97% (H), 45.97% (N), and 0.12% (S) percentage of contents. In the other study, in the presence of 1.0 g of CN sheets, 2.02 mL of 1,3-dibromopropane, and 1 mmol of VB1, the contents of C, H, N, and S surprisingly were changed to 35.20% (C), 3.76% (H), 26.75% (N), and 2.97% (S). According to these results, it is concluded that about 31% vitamin (VB1) molecules have grafted onto the CN sheets. To explain this difference, it can be described that in the presence of 2.02 mL of linker and by participation of nitrogen atoms with sp 2-hybridization in the g-C3N 4 framework, 1,3dibromopropane can react with N, −NH and −NH2 groups; thus, the ratio of reacted functional groups increased and more of VB1 can be attached to g-C3N4-Pr with covalent bonds. The CHNS elemental analysis is in accordance with the EDS spectra. As shown in Figure 4, the presence of C, N, O, Br, S, and Cl elements in the prepared CN-Pr-VB1 structure are proved. Therefore, it is concluded that vitamin (VB1) molecules are grafted onto the CN sheets via the nucleophilic reaction between the hydroxyl group of VB1 and Br atoms of CN-Pr-Br. In addition, also with the aid of thermogravimetric analysis/differential thermal analysis (TGA/DTA), the existence of organic moieties on the CN sheets has been proved. XRD data depicted the retention of the crystalline structure of g-C3N4 without any phase change in the structure of CN-PrVB1 during the functionalization procedure. The main characteristic peak of CN (g-C3N4 nanosheet) did not change after anchoring of VB1 (Figure 5). Further evidence of covalent functionalization of CN sheets with VB1 molecules was also reflected in TGA/DTA (Figure

Scheme 1. Synthesis of Quinoxaline Derivatives Catalyzed by CN-Pr-VB1



RESULTS AND DISCUSSION Catalyst Characterizations. CN-Pr-VB1 was prepared by a facile three-step process as shown in Scheme 2. At the initial step, two-dimensional (2D) g-C3N4 nanosheets were synthesized through the liquid exfoliation followed by sonication treatment. The crystal and chemical structure of g-C3N4 nanosheets were confirmed by X-ray diffraction (XRD) analysis and Fourier transform infrared (FT-IR) spectroscopy. The XRD analysis indicated that the layered g-C3N4 was successfully exfoliated into 2D nanosheets. As verified in the XRD analysis of bulk g-C3N4 (Figure 1), two diffraction peaks which correspond to the interplanar stacking of the conjugated aromatic systems and the in-plane structural packing motif of repeated tri-s-triazine units in 2θ = 27.4° for (002) reflection 12545

DOI: 10.1021/acsomega.9b01635 ACS Omega 2019, 4, 12544−12554

ACS Omega

Article

Scheme 2. Preparation of CN-Pr-VB1

Figure 1. XRD patterns of bulk g-C3N4 and g-C3N4 nanosheets.

Figure 3. (a) FT-IR spectra of g-C3N4 nanosheet (CN) and CN-PrBr, (b) EDS analysis of CN-Pr-Br. Figure 2. FT-IR spectra of bulk g-C3N4 and g-C3N4 nanosheets.

second step in the range of 200−400 °C indicates the presence of organic moieties for the CN-Pr-VB1, which correspond to the loss of the Pr-VB1 to the surface of the g-C3N4 nanosheets (CN). Therefore, it is concluded that the amount of Pr-VB1 grafted on the surface of the g-C3N4 nanosheets (CN) was about 30% that estimated from the percentage of weight loss from the TGA profile of CN-Pr-VB1 and CN. As shown in Figure 7, the FT-IR spectra reveal that the CNPr-VB1 still keeps the original C−N network of g-C3N4 nanosheets and the same chemical structure of VB1.

6). Small amounts of adsorbed water are lost in the initial heating step below 200 °C. The DTA curve of CN-Pr-VB1 exhibited an obvious weight loss within the temperature range of 200−400 °C, which indicates the existence of organic moieties on the CN sheets. In the last step, CN sheets start to decompose at >500 °C. Thus, the CN-Pr-VB1 as a catalyst retains its stability up to 200 °C in organic reactions. TGA of the CN-Pr-VB1 were also used to determine the content of organic functional groups (Pr-VB1) of the CN-Pr-VB1. The 12546

DOI: 10.1021/acsomega.9b01635 ACS Omega 2019, 4, 12544−12554

ACS Omega

Article

1, entries 4−7), whereas in EtOH, H2O, and H2O/EtOH, the product was obtained in moderate yields 60−70% (Table 1, entries 8−13). Conducting the model reaction under solventfree condition enhanced significantly the yield of the desired product in a shorter time period (Table 1, entry 14). In addition to this, we performed the model reaction at varying temperatures (Table 1, entries 14−17). It was found that increasing the temperature, from 80 to 100 °C, improves the yield of the product. However, another increase in the reaction temperature to 120 °C had no impact on the product yield. Further to get the optimized amount of catalyst, the model reaction was carried out at 10, 15, 20, and 25 mg of CN-Pr-VB1 at 100 °C under solvent-free conditions (Table 1, entries 17− 20). Therefore, 15 mg of the catalyst is sufficient to achieve the higher yields. The efficiency of CN-Pr-VB1 was also compared with CN (g-C3N4 nanosheets) and VB1 (Table 1, entries 21, 22). These observations reveal that the high catalytic activity of CN-Pr-VB1 may be due to the synergetic effect of CN (g-C3N4 nanosheets) and VB1. Encouraged by the established optimized reaction conditions, we explored/expanded the library construction and scope of the reaction involving various 1,2-diamines and 1,2diketones (Table 2). The results indicate that substrates possessing both electron-donating as well as electron-withdrawing groups afforded good-to-excellent yields of the desired products within a very short period of time. A plausible mechanism for the formation of quinoxalines is presented in Scheme 4. CN (g-C3N4 nanosheet) has a πconjugated structure with surface termination as defects and nitrogen atoms for electron localization or anchoring inorganic/organic functional motifs as the active sites. Because of the incorporation of amine groups in the carbon architecture of CN, the free amino groups distributed on the surface of CNPr-VB1 and also −NH proton of the VB1, and the acidic CH proton (the carbenoid proton) of VB1 activate the carbonyl groups of the 1,2-diketone through hydrogen bonding for nucleophilic attack of 1,2-diamine to afford the corresponding intermediate. Then, dehydration and elimination of a proton were done to afford quinoxaline as the final product. In view of the ion-pair structure of the catalyst (CN-PrVB1), we can postulate an important contribution of the catalyst to the TS stabilization. As a matter of fact, the CN-PrVB1 provides dipolar environment, leading to stabilize the charge distribution in the TS through specific interactions including H-bonding and dipole charge interactions. These observations reveal that the high catalytic activity of CN-PrVB1 may be due to the synergetic effect of CN (g-C3N4 nanosheets) and VB1. In light of the significant features of heterogeneous catalysts including recovery, reusability, and stability, we evaluated the recyclability of the CN-Pr-VB1 for the condensation reaction among o-phenylenediamine and benzil under optimum reaction conditions. The used CN-Pr-VB1 catalyst was separated from the reaction mixture by simple filtration, washed thoroughly with ethanol and water, and dried at 60 °C. The recovered CN-Pr-VB1 catalyst can be reused over six runs without notable descent of catalytic activity (Figure 9). The FT-IR spectra of the CN-Pr-VB1 after six cycles (Figure 10) presented that the structure of the catalyst does not change during the course of reaction. In order to further investigate the catalytic activity of CN-PrVB1, a comparison with other catalysts reported earlier in the literature for this reaction was done and the results are

Figure 4. EDS analysis of CN-Pr-VB1.

Figure 5. XRD patterns of g-C3N4 nanosheets and CN-Pr-VB1.

The morphology and microstructure of the synthesized CNPr-VB1 was characterized via field emission scanning electron microscopy (Figure 8). The FESEM images obtained indicated that the layered structure of CN sheets (g-C3N4 nanosheets) changes during functionalization of CN sheets with VB1, suggesting that the surface of CN sheets were immobilized with VB1. To understand the colloidal stability of the CN-Pr-VB1 sample compared to CN, zeta potential measurements of aqueous solution of CN (g-C3N4 nanosheet) and CN-Pr-VB1 were tested (Figure S1). The zeta potential results indicate that the CN sheet suspension carries a negative charge of −2.56 mV. After covalent functionalization of CN (g-C3N4 sheets) with VB1, the zeta potential increased from −2.56 mV (CN) to −5.76 mV (CN-Pr-VB1) and possessed more negative polarity on its surface in the solution attributed to the existence of the ion pairs (Cl− and Br−) in the structure of CN-Pr-VB1 (Figure S2). Therefore, the CN-Pr-VB1 demonstrates good dispersibility in water. Catalytic Properties. The catalytic potential of CN-PrVB1 was explored for the preparation of quinoxaline derivatives via condensation between aryl 1,2-diamines and 1,2-dicarbonyl compounds. To obtain the optimum reaction conditions regarding the solvent, reaction temperature, and the amount of catalyst, the reaction of o-phenylenediamine and benzil was chosen as the model reaction (Scheme 3). In the absence of the catalyst, no desired product was observed after the prolonged reaction time of 24 h (Table 1, entries 1−3). The progress of the reaction was investigated by using CN-Pr-VB1 as the catalyst in different solvents. In CH3CN and MeOH, the reaction proceeded slowly with low yield of the target product (Table 12547

DOI: 10.1021/acsomega.9b01635 ACS Omega 2019, 4, 12544−12554

ACS Omega

Article

Figure 6. Thermogravimetric and differential thermogravimetric of (a) g-C3N4 nanosheets (CN) and (b) CN-Pr-VB1.

Figure 7. FT-IR spectra of the g-C3N4 nanosheet (CN), VB1, CN-PrBr, and CN-Pr-VB1.

summarized in Table 3. According to these data, our protocol has several advantages over the reported methods in terms of the short reaction time, superior efficiency, mild and environmentally benign conditions, high yield, and recyclability of the catalyst.



CONCLUSIONS Chemical modification of g-C3N4 can enhance its properties as an advantageous 2D nanosheet material such as developing solubility and stability. To enhance g-C3N4’s properties, a new and straightforward synthetic strategy for the covalent functionalization of the g-C3N4 framework was developed through 1,3-dibromopropane as a linker for the first time. The synthesized heterogeneous catalyst (CN-Pr-VB1) was characterized by a range of techniques to confirm the covalent bond

Figure 8. FESEM images of (a,b) CN (g-C3N4 nanosheets) and (c,d) CN-Pr-VB1.

12548

DOI: 10.1021/acsomega.9b01635 ACS Omega 2019, 4, 12544−12554

ACS Omega

Article

Scheme 3. Model Reaction for the Optimization Reaction Conditions

Table 1. Optimization of the Reaction Conditions for the Synthesis of Quinoxalinesa entry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22

solvent

catalyst (mg)

temperature (°C)

time (min)

yieldb (%)

CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (15) CN-Pr-VB1 (10) CN-Pr-VB1 (20) CN-Pr-VB1 (25) CN (g-C3N4 nanosheets) (15) VB1 (2 mol %)

rt. reflux 100 rt. reflux rt. reflux rt. reflux rt. reflux rt. reflux 80 100 110 120 100 100 100 100 100

24 h 24 h 24 h 70 70 70 70 30 30 30 30 30 30 3 3 3 3 3 3 3 3 3

trace trace trace 35 40 40 45 65 70 60 68 73 75 85 97 97 97 90 97 97 45 60

EtOH EtOH CH3CN CH3CN MeOH MeOH EtOH EtOH H2O H2O EtOH/H2O EtOH/H2O

a

Reaction condition: o-phenylenediamine (1.0 mmol), benzil (1.0 mmol), solvent (3.0 mL). bIsolated yield.

using the dynamic light-scattering analysis using the Zeta sizer Nano ZS (Malvern Instruments, England). The suspensions of each sample in aqueous solution media (0.1 mg·mL−1) were prepared through sonication at room temperature. Fourier transform infrared (FT-IR) spectra were obtained on a Shimadzu 800 IR 100 FT-IR spectrometer with a standard KBr disk technique. Melting points were determined on an Electrothermal 9100 apparatus without correction. 1H NMR and 13C NMR spectra were recorded on a Burker DRX-500 Avance spectrometer (500 and 125 MHz) in DMSO-d6 using tetramethylsilane as the internal standard. Preparation of Bulk g-C3N4. Bulk g-C3N4 powder was synthesized according to the previous reported article.16 The melamine was heated at 550 °C in a furnace at a ramp of 2.5 °C min−1 in static air for 4 h, and the resulted yellow solid was grinded into powder in a mortar. Preparation of g-C3N4 Nanosheets. g-C3N4 nanosheets (CN) were prepared according to the previous reported method.17 First, bulk g-C3N4 (1.0 g) was treated in the mixture of 20.0 mL of H2SO4 and 20.0 mL of HNO3 at room temperature for 2 h. The mixture was diluted with 1.0 L of deionized H2O, and the obtained precipitate was filtered and washed several times with deionized water (filtered undispersed g-C3N4) and dried at 60 °C. Then, treated bulk g-C3N4 (100.0 mg) was dispersed in 100.0 mL of water/isopropanol (1:1) and sonicated for 6 h. Finally, to remove the residual unexfoliated g-C3N4 nanoparticles, the formed suspension was centrifuged (5000 rpm).

formation of VB1 with CN sheets. This novel catalyst displayed high thermal stability compared to the VB1. Additionally, the CN-Pr-VB1 was found to be an efficient and biocompatible heterogeneous organocatalyst for the synthesis of quinoxalines. This novel protocol offers several advantages like easy preparation of the catalyst from inexpensive and benign precursors, simple operational procedure, short reaction time period, excellent yield of products, and reusability and stability of the catalyst over several reaction cycles. It can be expected that the g-C3N4 nanosheets are a promising candidate for the catalyst support in organic synthesis and transformation.



EXPERIMENTAL SECTION Chemicals and Instruments. All reagents and solvents were purchased from Merck or Fluka chemical companies and were used without further purification. XRD analyses were obtained on a PANalytical X’Pert-PRO MPD diffractometer with Cu Kα (λ = 1.5406 Å) irradiation in the 2θ = 10°−80° with a 2θ step size of 0.02°. FESEM were acquired using MIRA3 TESCAN-XMU. Energy-dispersive X-ray spectroscopy (EDS) was used to investigate the composition and structure of the samples and was recorded on Oxford Instrument (England). C, H, N, and S elemental analyses for CN-Pr-VB1 were performed using a Thermo Finnigan Flash EA 1112 Element Analyzer. Thermogravimetric/differential thermal analyses (TGA/DTG) were conducted using an STA504 analyzer with a heating rate of 10 °C/min from 20 to 800 °C. Zeta potential measurements of the samples were obtained 12549

DOI: 10.1021/acsomega.9b01635 ACS Omega 2019, 4, 12544−12554

ACS Omega

Article

Table 2. CN-Pr-VB1-Catalyzed Synthesis of Quinoxalines via the Condensation of 1,2-Diamines with 1,2-Diketones under Solvent-Free Conditiona56−67

Reaction conditions: 1.0 mmol 1,2-diamine, 1.0 mmol 1,2-diketone, and 15.0 mg catalyst under solvent-free condition at 100 °C. bIsolated yields.

a

Synthesis of CN-Pr-Br. The resulting g-C3N4 nanosheets (CN) (1.0 g) were dispersed in 25.0 mL of dry toluene and then, 1,3-dibromopropane (2.02 mL) and NaI (1.0 mmol) were added to the dispersed solution, and the reaction mixture was refluxed overnight under a nitrogen atmosphere. The

resulted mixture was cooled to room temperature and collected by centrifugation, washed with ethyl acetate, and dried at room temperature. Synthesis of CN-Pr-VB1. As-prepared CN-Pr-Br (1.0 g) was dispersed in toluene (25.0 mL) by ultrasonication for 30 12550

DOI: 10.1021/acsomega.9b01635 ACS Omega 2019, 4, 12544−12554

ACS Omega

Article

Scheme 4. Plausible Mechanism for the Synthesis of Quinoxalines Catalyzed by CN-Pr-VB1

Table 3. Comparison between the Catalytic Efficiency of the CN-Pr-VB1 with the Reported Catalysts for the Condensation of o-Phenylenediamine and Benzil entry

reaction conditions

1 2

DMSO, I2 (10 mol %), rt. nano-ZrO2 (5 mol %), dichloroethane, 25 °C ZnCl2 (4 mol %), EtOH/H2O (3/1, v/v), rt. glycerol/H2O, 90 °C 17% ZrO2 4% Ga2O3/MCM-41, acetonitrile, rt. polyaniline sulfate, dichloromethane, rt. oxalic acid (20 mol %), EtOH/H2O (1/1), rt. Zn[(L)proline] (10 mol %), HOAc, rt. graphite (2 mmol), EtOH, rt. montmorillonite K-10 (10 mol %), H2O rt. CN-Pr-VB1 (15.0 mg), solvent-free, 100 °C

3 4 5

Figure 9. Reusability of the CN-Pr-VB1 in the synthesis of quinoxalines.

6 7 8 9 10 11 a

time (min)

yielda (%)

35 45

95 89

68 69

300

80

70

240 120

90 97

47 71

20

95

72

10

93

52

10

95

53

60 150

92 100

54 51

3

97

refs

this work

Isolated yield.

with H2O and EtOH to remove the unreacted substrate to obtain CN-Pr-VB1 and then dried at 50 °C overnight. General Procedure for the Synthesis of Quinoxaline Derivatives. To a round-bottom flask containing 1,2-diamine (1.0 mmol) and 1,2-diketone (1.0 mmol) was added CN-PrVB1 as the catalyst (15.0 mg) and stirred at 100 °C in an oil bath, some minutes later the reagents were melted and after appropriate reaction time (Table 2), and the solid product was achieved. The completion of the reaction was indicated by thin-layer chromatography. After completion of the reaction, EtOAC (10 mL) was added and the catalyst was filtered off; then, the solvent was evaporated under vacuum. The obtained

Figure 10. FT-IR spectra of fresh and used CN-Pr-VB1.

min; then, thiamine hydrochloride (1.0 mmol), K2CO3 (1.0 mmol), and NaI (1.0 mmol) were added to this mixture. The reaction mixture was heated at reflux temperature for 48 h under a nitrogen atmosphere. After being cooled to room temperature, the filtration product was washed several times 12551

DOI: 10.1021/acsomega.9b01635 ACS Omega 2019, 4, 12544−12554

ACS Omega

Article

Notes

crude product was purified by recrystallization from ethanol. The products were identified by 1H NMR, 13C NMR, IR spectroscopy, and melting points. The recovered catalyst was rinsed with ethanol and reused in other fresh reactions without a significant loss of activity for another six times. Selected Spectral Data for Product. 6-Nitro-2,3diphenylquinoxaline. mp: 185−188 °C; FT-IR (KBr) νmax (cm−1): 3055, 2922, 1615, 1515, 1500, 1469, 1446, 1340; 1H NMR (500 MHz, CDCl3): δ (ppm): 7.35−7.44 (m, 6H), 7.6 (t, 4H, J = 6.5 Hz), 8.33 (d, 1H, J = 9.1 Hz), 8.56 (dd, 1H, J = 9.1, 1.4 Hz), 9.10 (d, 1H, J = 1.3 Hz). 2,3-Bis(4-methoxy-phenyl)-6-methylquinoxaline. mp: 124−126 °C; FT-IR (KBr) νmax (cm−1): 3050, 2900, 1655, 1604, 1511, 1343, 1246, 1027; 1H NMR (500 MHz, CDCl3): δ (ppm): 2.62 (s, 3H), 3.85 (s, 6H), 6.90 (d, 4H, J = 9 Hz), 7.50 (dd, 4H, J = 2, 8.75 Hz), 7.58 (dd, 1H, J = 2, 8.5 Hz), 7.92 (s, 1H), 8.03 (d, 1H, J = 8.5 Hz); 13C NMR (125 MHz, CDCl3): δ (ppm): 20.8, 54.2, 112.7, 126.8, 127.5, 130.2, 130.8, 138.5, 138.9, 140.1, 151.1, 151.8, 159.0. Acenaphtho[1,2-b]quinoxaline. mp: 229−232 °C; FT-IR (KBr) νmax (cm−1): 3047, 2312, 1425, 1298, 1093, 754; 1H NMR (500 MHz, CDCl3): δ (ppm): 7.74−7.81 (m, 4H), 8.05 (d, 2H, J = 8 Hz), 8.18−8.20 (m, 2H), 8.37 (d, 2H, J = 7 Hz). 2,3-Diphenylquinoxaline. mp: 116−118 °C; FT-IR (KBr) νmax (cm−1): 3056, 1548, 1342, 767, 730, 696; 1H NMR (DMSO-d6, 500 MHz): δ (ppm): 8.15 (dd, J = 3.4, 6.4 Hz, 2H), 7.88 (dd, J = 3.4, 6.4 Hz, 2H), 7.48 (d, 4H), 7.39 (m, 6H). 2,3-Bis(4-methoxyphenyl)quinoxaline. mp: 147−149 °C; FT-IR (KBr) νmax (cm−1): 3004, 2935, 1604, 1510, 1344, 1026, 833; 1H NMR (DMSO-d6, 500 MHz): δ (ppm): 8.10 (dd, J = 3.5, 6.25 Hz, 2H), 7.82 (dd, J = 3.5, 6.3 Hz, 2H), 7.45 (d, J = 8.75 Hz, 4H), 6.89 (d, J = 8.7 Hz, 4H), 3.78 (s, 6H). 6-Methyl-2,3-diphenylquinoxaline. mp: 110−113 °C; FTIR (KBr) νmax (cm−1): 3063, 1660, 1592, 1210, 874, 719, 640; 1 H NMR (DMSO-d6, 500 MHz): δ (ppm): 8.05 (d, J = 8.5 Hz, 1H), 7.94 (s, 1H), 7.74 (dd, J = 2.1, 8.45 Hz, 1H), 7.45 (m, 4H), 7.35 (m, 6H), 2.5 (s, 3H). 5,6-Diphenylpyrazine-2,3-dicarbonitrile. mp: 210−212 °C; FT-IR (KBr) νmax (cm−1): 3060, 1596, 1512, 1375, 1228, 771, 698; 1H NMR (DMSO-d6, 500 MHz): δ (ppm): 7.40−7.43, 7.48−7.50 and (m, Ph, 10 H). 9-Methylacenaphtho[1,2-b]quinoxaline. mp: 233−235 °C; FT-IR (KBr) νmax (cm−1): 3047, 2360, 1691, 1512, 1425, 1296, 1095, 912, 825, 773; 1H NMR (DMSO-d6, 500 MHz): δ 2.59 (s, 3H), 7.68 (d, J = 8 Hz, 1H), 7.83 (td, J = 7.4 Hz, 2H), 7.99 (s, 1H), 8.10 (d, J = 8.35 1H), 8.30 (q, J = 4.1 Hz, 2H), 8.39 (t, J = 6.65 Hz, 2H).



The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors gratefully acknowledge the partial support from the Research Council of the Iran University of Science and Technology.



(1) Savateev, A.; Antonietti, M. Heterogeneous Organocatalysis for Photoredox Chemistry. ACS Catal. 2018, 8, 9790−9808. (2) Nestl, B. M.; Hammer, S. C.; Nebel, B. A.; Hauer, B. New generation of biocatalysts for organic synthesis. Angew. Chem., Int. Ed. 2014, 53, 3070−3095. (3) Maleki, A.; Rahimi, J.; Demchuk, O. M.; Wilczewska, A. Z.; Jasiński, R. Green in water sonochemical synthesis of tetrazolopyrimidine derivatives by a novel core-shell magnetic nanostructure catalyst. Ultrason. Sonochem. 2018, 43, 262−271. (4) Dekamin, M. G.; Karimi, Z.; Latifidoost, Z.; Ilkhanizadeh, S.; Daemi, H.; Naimi-Jamal, M. R.; Barikani, M. Alginic acid: A mild and renewable bifunctional heterogeneous biopolymeric organocatalyst for efficient and facile synthesis of polyhydroquinolines. Int. J. Biol. Macromol. 2018, 108, 1273−1280. (5) Maleki, A.; Firouzi-Haji, R.; Hajizadeh, Z. Magnetic guanidinylated chitosan nanobiocomposite: A green catalyst for the synthesis of 1, 4-dihydropyridines. Int. J. Biol. Macromol. 2018, 116, 320−326. (6) Breslow, R. On the mechanism of thiamine action. IV. 1 Evidence from studies on model systems. J. Am. Chem. Soc. 1958, 80, 3719−3726. (7) Liu, J.; Lei, M.; Hu, L. Thiamine hydrochloride (VB1): an efficient promoter for the one-pot synthesis of benzo [4, 5] imidazo [1, 2-a] pyrimidine and [1, 2, 4] triazolo [1, 5-a] pyrimidine derivatives in water medium. Green Chem. 2012, 14, 840−846. (8) Chen, Y.; Zhang, X. P. Vitamin B12 derivatives as natural asymmetric catalysts: enantioselective cyclopropanation of alkenes. J. Org. Chem. 2004, 69, 2431−2435. (9) Lei, M.; Ma, L.; Hu, L. Highly chemoselective condensation of β-naphthol, aldehyde, and urea catalyzed by thiamine hydrochloride. Synth. Commun. 2011, 41, 3424−3432. (10) Yi, Z.; Lan, D.; Wang, Y.; Chen, L.; Au, C.; Yin, S. Green and efficient cycloaddition of CO2 toward epoxides over thiamine derivatives/GO aerogels under mild and solvent-free conditions. Sci. China Chem. 2017, 60, 990−996. (11) Lam, E.; Luong, J. H. T. Carbon materials as catalyst supports and catalysts in the transformation of biomass to fuels and chemicals. ACS Catal. 2014, 4, 3393−3410. (12) Dubertret, B.; Heine, T.; Terrones, M. The Rise of TwoDimensional Materials; ACS Publications, 2015. (13) Ong, W.-J.; Tan, L.-L.; Ng, Y. H.; Yong, S.-T.; Chai, S.-P. Graphitic carbon nitride (g-C3N4)-based photocatalysts for artificial photosynthesis and environmental remediation: are we a step closer to achieving sustainability? Chem. Rev. 2016, 116, 7159−7329. (14) Zhang, Y.; Zhao, H.; Hu, Z.; Chen, H.; Zhang, X.; Huang, Q.; Wo, Q.; Zhang, S. Protic salts of high nitrogen content as versatile precursors for graphitic carbon nitride: anion effect on the structure, properties, and photocatalytic activity. ChemPlusChem 2015, 80, 1139−1147. (15) Zhuang, X.; Mai, Y.; Wu, D.; Zhang, F.; Feng, X. Twodimensional soft nanomaterials: a fascinating world of materials. Adv. Mater. 2015, 27, 403−427. (16) Zheng, Y.; Lin, L.; Ye, X.; Guo, F.; Wang, X. Helical graphitic carbon nitrides with photocatalytic and optical activities. Angew. Chem., Int. Ed. 2014, 53, 11926−11930. (17) Li, C.-Z.; Wang, Z.-B.; Sui, X.-L.; Zhang, L.-M.; Gu, D.-M. Ultrathin graphitic carbon nitride nanosheets and graphene composite material as high-performance PtRu catalyst support for methanol electro-oxidation. Carbon 2015, 93, 105−115.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b01635. List of materials and characterization analysis (PDF)



REFERENCES

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Fax: 982177491204. Phone: 982173228312. ORCID

Hossein Ghafuri: 0000-0001-6778-985X 12552

DOI: 10.1021/acsomega.9b01635 ACS Omega 2019, 4, 12544−12554

ACS Omega

Article

heterogeneous solid acid catalyst for Biginelli reaction. Microporous Mesoporous Mater. 2019, 274, 83−93. (37) Seitz, L. E.; Suling, W. J.; Reynolds, R. C. Synthesis and antimycobacterial activity of pyrazine and quinoxaline derivatives. J. Med. Chem. 2002, 45, 5604−5606. (38) Sonawane, N. D.; Rangnekar, D. W. Synthesis and application of 2-styryl-6, 7-dichlorothiazolo [4, 5-b]-quinoxaline based fluorescent dyes: Part 3. J. Heterocycl. Chem. 2002, 39, 303−308. (39) Knowles, C. O. Chemistry and toxicology of quinoxaline, organotin, organofluorine, and formamidine acaricides. Environ. Health Perspect. 1976, 14, 93−102. (40) Bailly, C.; Echepare, S.; Gago, F.; Waring, M. J. Recognition elements that determine affinity and sequence-specific binding to DNA of 2QN, a biosynthetic bis-quinoline analogue of echinomycin. Anti-Cancer Drug Des. 1999, 14, 291−303. (41) Dell, A.; Williams, D. H.; Morris, H. R.; Smith, G. A.; Feeney, J.; Roberts, G. C. K. Structure revision of the antibiotic echinomycin. J. Am. Chem. Soc. 1975, 97, 2497−2502. (42) O’Brien, D.; Weaver, M.; Lidzey, D.; Bradley, D. Use of poly (phenyl quinoxaline) as an electron transport material in polymer light-emitting diodes. Appl. Phys. Lett. 1996, 69, 881−883. (43) Thomas, K. R. J.; Velusamy, M.; Lin, J. T.; Chuen, C.-H.; Tao, Y.-T. Chromophore-labeled quinoxaline derivatives as efficient electroluminescent materials. Chem. Mater. 2005, 17, 1860−1866. (44) Dailey, S.; Feast, W. J.; Peace, R. J.; Sage, I. C.; Till, S.; Wood, E. L. Synthesis and device characterisation of side-chain polymer electron transport materials for organic semiconductor applications. J. Mater. Chem. 2001, 11, 2238−2243. (45) Toshima, K.; Takano, R.; Ozawa, T.; Matsumura, S. Molecular design and evaluation of quinoxaline-carbohydrate hybrids as novel and efficient photo-induced GG-selective DNA cleaving agents. Chem. Commun. 2002, 3, 212−213. (46) Brown, D. In Chemistry of Heterocyclic Compounds, Quinoxalines Supplements II; Taylor, EC., Wipf, P., Eds.; John Wiley and Sons: New Jersey, 2004. (47) Bachhav, H. M.; Bhagat, S. B.; Telvekar, V. N. Efficient protocol for the synthesis of quinoxaline, benzoxazole and benzimidazole derivatives using glycerol as green solvent. Tetrahedron Lett. 2011, 52, 5697−5701. (48) More, S. V.; Sastry, M. N. V.; Yao, C.-F. Cerium (IV) ammonium nitrate (CAN) as a catalyst in tap water: A simple, proficient and green approach for the synthesis of quinoxalines. Green Chem. 2006, 8, 91−95. (49) More, S. V.; Sastry, M. N. V.; Wang, C.-C.; Yao, C.-F. Molecular iodine: a powerful catalyst for the easy and efficient synthesis of quinoxalines. Tetrahedron Lett. 2005, 46, 6345−6348. (50) Srinivas, C.; Kumar, C. N. S. S. P.; Rao, V. J.; Palaniappan, S. Green approach for the synthesis of quinoxaline derivatives in water medium using reusable polyaniline-sulfate salt catalyst and sodium laurylsulfate. Catal. Lett. 2008, 121, 291−296. (51) Huang, T.-k.; Wang, R.; Shi, L.; Lu, X.-x. Montmorillonite K10: an efficient and reusable catalyst for the synthesis of quinoxaline derivatives in water. Catal. Commun. 2008, 9, 1143−1147. (52) Hasaninejad, A.; Zare, A.; Mohammadizadeh, M. R.; Shekouhya, M. Oxalic acid as an efficient, cheap, and reusable catalyst for the preparation of quinoxalines via condensation of 1, 2diamines with α-diketones at room temperature. ARKIVOC 2008, 13, 28−35. (53) Heravi, M. M.; Tehrani, M. H.; Bakhtiari, K.; Oskooie, H. A. Zn [(L) proline]: A powerful catalyst for the very fast synthesis of quinoxaline derivatives at room temperature. Catal. Commun. 2007, 8, 1341−1344. (54) Kadam, H. K.; Khan, S.; Kunkalkar, R. A.; Tilve, S. G. Graphite catalyzed green synthesis of quinoxalines. Tetrahedron Lett. 2013, 54, 1003−1007. (55) Pawar, O. B.; Chavan, F. R.; Suryawanshi, V. S.; Shinde, V. S.; Shinde, N. D. Thiamine hydrochloride: An efficient catalyst for onepot synthesis of quinoxaline derivatives at ambient temperature. J. Chem. Sci. 2013, 125, 159−163.

(18) Gong, Y.; Zhang, P.; Xu, X.; Li, Y.; Li, H.; Wang, Y. A novel catalyst Pd@ ompg-C3N4 for highly chemoselective hydrogenation of quinoline under mild conditions. J. Catal. 2013, 297, 272−280. (19) Huang, Y.; Wang, Y.; Bi, Y.; Jin, J.; Ehsan, M. F.; Fu, M.; He, T. Preparation of 2D hydroxyl-rich carbon nitride nanosheets for photocatalytic reduction of CO2. RSC Adv. 2015, 5, 33254−33261. (20) Vattikuti, S. V. P.; Reddy, P. A. K.; Shim, J.; Byon, C. VisibleLight-Driven Photocatalytic Activity of SnO2−ZnO Quantum Dots Anchored on g-C3N4 Nanosheets for Photocatalytic Pollutant Degradation and H2 Production. ACS Omega 2018, 3, 7587−7602. (21) Zhong, Y.; Chen, W.; Yu, S.; Xie, Z.; Wei, S.; Zhou, Y. CdSe Quantum Dots/g-C3N4 Heterostructure for Efficient H2 Production under Visible Light Irradiation. ACS Omega 2018, 3, 17762−17769. (22) Bahuguna, A.; Choudhary, P.; Chhabra, T.; Krishnan, V. Ammonia-Doped Polyaniline−Graphitic Carbon Nitride Nanocomposite as a Heterogeneous Green Catalyst for Synthesis of IndoleSubstituted 4 H-Chromenes. ACS Omega 2018, 3, 12163−12178. (23) Gong, Y.; Li, M.; Li, H.; Wang, Y. Graphitic carbon nitride polymers: promising catalysts or catalyst supports for heterogeneous oxidation and hydrogenation. Green Chem. 2015, 17, 715−736. (24) Barrio, J.; Shalom, M. Rational design of carbon nitride materials by supramolecular preorganization of monomers. ChemCatChem 2018, 10, 5573−5586. (25) Shahabi, S. S.; Azizi, N.; Vatanpour, V. Synthesis and characterization of novel g-C3N4 modified thin film nanocomposite reverse osmosis membranes to enhance desalination performance and fouling resistance. Sep. Purif. Technol. 2019, 215, 430−440. (26) Fu, Y.; Zhu, J.; Hu, C.; Wu, X.; Wang, X. Covalently coupled hybrid of graphitic carbon nitride with reduced graphene oxide as a superior performance lithium-ion battery anode. Nanoscale 2014, 6, 12555−12564. (27) Wang, Y.; Hai, X.; E, S.; Chen, M.; Yang, T.; Wang, J. Boronic acid functionalized g-C3N4 nanosheets for ultrasensitive and selective sensing of glycoprotein in the physiological environment. Nanoscale 2018, 10, 4913−4920. (28) Sun, J.; Phatake, R.; Azoulay, A.; Peng, G.; Han, C.; Barrio, J.; Xu, J.; Wang, X.; Shalom, M. Covalent Functionalization of Carbon Nitride Frameworks through Cross-Coupling Reactions. Chem. Eur J. 2018, 24, 14921−14927. (29) Chauhan, D. K.; Jain, S.; Battula, V. R.; Kailasam, K. Organic motif’s functionalization via covalent linkage in carbon nitride: An exemplification in photocatalysis. Carbon 2019, 152, 40−58. (30) Kumru, B.; Barrio, J.; Zhang, J.; Antonietti, M.; Shalom, M.; Schmidt, B. V. K. J. Robust Carbon Nitride based Thermoset Coatings for Surface Modification and Photochemistry. ACS Appl. Mater. Interfaces 2019, 11, 9462−9469. (31) Ghafuri, H.; Rashidizadeh, A.; Esmaili Zand, H. R. Highly efficient solvent free synthesis of α-aminophosphonates catalyzed by recyclable nano-magnetic sulfated zirconia (Fe3O4@ ZrO2/SO42−). RSC Adv. 2016, 6, 16046−16054. (32) Ghafuri, H.; Talebi, M. Water-soluble phosphated graphene: preparation, characterization, catalytic reactivity, and adsorption property. Ind. Eng. Chem. Res. 2016, 55, 2970−2982. (33) Ghafuri, H.; Joorabchi, N.; Emami, A.; Esmaili Zand, H. R. Covalent Modification of Graphene Oxide with Vitamin B1: Preparation, Characterization, and Catalytic Reactivity for Synthesis of Benzimidazole Derivatives. Ind. Eng. Chem. Res. 2017, 56, 6462− 6467. (34) Zand, H. R. E.; Ghafuri, H.; Ghanbari, N. Lewis Acidic Ionic Liquid Immobilized on Graphene Oxide: Synthesis, Characterization, and Application in the Ultrasound-Assisted Condensation Reactions. ChemistrySelect 2018, 3, 8229−8237. (35) Emami, A.; Ghafuri, H. Synergetic catalytic effect of rGO, Pd, Fe3O4 and PPy as a magnetically separable and recyclable nanocomposite for coupling reactions in green media. Appl. Organomet. Chem. 2018, 32, No. e4439. (36) Fard, M. A. D.; Ghafuri, H.; Rashidizadeh, A. Sulfonated highly ordered mesoporous graphitic carbon nitride as a super active 12553

DOI: 10.1021/acsomega.9b01635 ACS Omega 2019, 4, 12544−12554

ACS Omega

Article

(56) Harsha, K. B.; Rangappa, K. S. One-step approach for the synthesis of functionalized quinoxalines mediated by T3P®−DMSO or T3P® via a tandem oxidation−condensation or condensation reaction. RSC Adv. 2016, 6, 57154−57162. (57) Jafarpour, M.; Rezaeifard, A.; Ghahramaninezhad, M.; Tabibi, T. Reusable α-MoO3 nanobelts catalyzes the green and heterogeneous condensation of 1, 2-diamines with carbonyl compounds. New J. Chem. 2013, 37, 2087−2095. (58) Fathi, S.; Sardarian, A. R. Nitrilotris (Methylenephosphonic Acid) as a New Highly Efficient and Recyclable BrØNested Acid Catalyst for the Synthesis Of Quinoxaline Derivatives under Mild and Green Conditions. Phosphorus, Sulfur, Silicon Relat. Elem. 2015, 190, 1471−1478. (59) Nandi, G. C.; Samai, S.; Kumar, R.; Singh, M. S. Silica-gel− catalyzed efficient synthesis of quinoxaline derivatives under solventfree conditions. Synth. Commun. 2011, 41, 417−425. (60) Jeena, V.; Sithebe, S.; Robinson, R. S. Copper-catalyzed synthesis of valuable heterocyclic compounds using a tandem oxidation process approach. Synth. Commun. 2015, 45, 1484−1491. (61) Shamsi-Sani, M.; Shirini, F.; Abedini, M.; Seddighi, M. Synthesis of benzimidazole and quinoxaline derivatives using reusable sulfonated rice husk ash (RHA-SO3H) as a green and efficient solid acid catalyst. Res. Chem. Intermed. 2016, 42, 1091−1099. (62) Daragahi, S. A. H.; Mohebat, R.; Mosslemin, M. H. Green and Eco-Friendly Synthesis of Quinoxalines by Brönsted Acidic Ionic Liquid Supported on Nano-SiO2 under Solvent-Free Conditions. Org. Prep. Proced. Int. 2018, 50, 301−313. (63) Mohammadi, K.; Hasaninejad, A.; Niad, M.; Najmi, M. Application of metalloporphyrins as new catalysts for the efficient, mild and regioselective synthesis of quinoxaline derivatives. J. Porphyrins Phthalocyanines 2010, 14, 1052−1058. (64) Shaabani, A.; Maleki, A. Green and efficient synthesis of quinoxaline derivatives via ceric ammonium nitrate promoted and in situ aerobic oxidation of α-hydroxy ketones and α-keto oximes in aqueous media. Chem. Pharm. Bull. 2008, 56, 79−81. (65) Karami, B.; Rooydel, R.; Saeed, K. A rapid synthesis of some 1, 4-aryldiazines by the use of lithium chloride as an effective catalyst. Acta Chim Slov 2012, 59, 183−188. (66) Hou, J.-T.; Liu, Y. H.; Zhang, Z. H. NbCl5 as an efficient catalyst for rapid synthesis of quinoxaline derivatives. J. Heterocycl. Chem. 2010, 47, 703−706. (67) Hasaninejad, A.; Zare, A.; Zolfigol, M. A.; Shekouhy, M. Zirconium tetrakis (dodecyl sulfate)[Zr(DS)4] as an efficient lewis acid−surfactant combined catalyst for the synthesis of quinoxaline derivatives in aqueous media. Synth. Commun. 2009, 39, 569−579. (68) Bhosale, R. S.; Sarda, S. R.; Ardhapure, S. S.; Jadhav, W. N.; Bhusare, S. R.; Pawar, R. P. An efficient protocol for the synthesis of quinoxaline derivatives at room temperature using molecular iodine as the catalyst. Tetrahedron Lett. 2005, 46, 7183−7186. (69) Alinezhad, H.; Tajbakhsh, M.; Salehian, F.; Biparva, P.; Branch, U. Synthesis of Quinoxaline Derivatives Using TiO2 Nanoparticles as an Efficient and Recyclable Catalyst. Bull. Korean Chem. Soc. 2012, 43, 3720−3725. (70) Hasaninejad, A.; Zare, A.; Mohammadizadeh, M. R.; Karami, Z. Synthesis of quinoxaline derivatives via condensation of aryl-1, 2diamines with 1, 2-Diketones using (NH4)6Mo7O24.4H2O as an efficient, mild and reusable catalyst. J. Iran. Chem. Soc. 2009, 6, 153− 158. (71) Ajaikumar, S.; Pandurangan, A. Efficient synthesis of quinoxaline derivatives over ZrO2/MxOy (M= Al, Ga, In and La) mixed metal oxides supported on MCM-41 mesoporous molecular sieves. Appl. Catal., A 2009, 357, 184−192. (72) Srinivas, C.; Kumar, C. N. S. S. P.; Rao, V. J.; Palaniappan, S. Efficient, convenient and reusable polyaniline-sulfate salt catalyst for the synthesis of quinoxaline derivatives. J. Mol. Catal. A: Chem. 2007, 265, 227−230.

12554

DOI: 10.1021/acsomega.9b01635 ACS Omega 2019, 4, 12544−12554