Abstract

Nickel peroxide nanoparticles (NPNPs) were prepared and characterized using various techniques including transmission electron microscope (TEM), scan electron microscope (SEM), energy dispersive spectrometer (EDS), X-ray diffraction (XRD), and FTIR spectra. The aqueous basic catalytic system NiSO4·6H2O/NaOCl/NaOH (pH = 14) was investigated for the catalytic dehydrogenation of benzylamine and parasubstituents to their corresponding nitriles at room temperature. The obtained results confirmed the formation of NiO2 nanocrystalline particles with a size of 20 nm. Benzylamine with electron-donating groups showed higher yields of nitriles compared to electron-withdrawing groups. The mechanism involved in the in situ generated NiO2 nanoparticles dehydrogenating benzylamine to benzonitrile, with the produced NiO converting back to NiO2 nanoparticles through the excess of NaOCl.

1. Introduction

The green synthesis of organic compounds is of great importance for the protection of the environment from the dangerous climate changes [1]. Nitriles are very important functional group because of its presence in many pharmaceutical materials [2, 3]. Many drugs contain nitrile groups in their structures. Nitriles can enhance the biological activity of drugs by improving their stability, lipophilicity, and metabolic properties. For example, nitrile-containing drugs are used to treat hypertension, cancer, and diabetes [4]. Different strategies have been developed for the synthesis of nitriles such as the Sandmeyer reaction [5], Rosenmund-von Braun reaction [6], dehydration of amides and aldoximes [7], nucleophilic substitution of alkyl and aryl halides [8], and oxidation of amines [9]. However, these methods often involve toxic reagents, harsh conditions, and produce chemical waste. Therefore, there is a strong need for an ecofriendly protocol. A number of transition metals such as ruthenium [10, 11], copper [12, 13], and nickel [14, 15] have been used for the catalytic hydrogenation of benzylamine to nitrile. Griffith and coworkers reported the catalytic dehydrogenation of benzyl amine by trans-[RuO3(OH)2]2−/S2O82− in aqueous alkaline medium [10]. Taube and coworkers reported the catalytic dehydrogenation of coordinated benzylamine to coordinated benzonitrile [11]. The simple copper-salt catalyst was used in the selective aerobic oxidation of amines to nitriles [12]. A number of aliphatic and aromatic amines were oxidized to nitriles by the CuCl2/O2 catalytic system in toluene as a solvent and at 80°C [13]. Nickel peroxide was used for the oxidation of many organic compounds [14]. Some alkylamines were dehydrogenated to their corresponding nitriles by NiSO4/K2S2O8 in good yields [15]. Pd@CuO, used as an anodic electrode, resulted in the electro-oxidative coupling of benzyl alcohol and ammonia, producing 83.2% of benzonitrile. This method was also successful in converting various primary alcohols to nitriles [16]. Xiao et al. [17] developed an efficient chemoenzymatic strategy to prepare nitriles from benzylamine by combining selective oxidation and dehydration.

Metal nanoparticles are of great interest to researchers due to their ease of synthesis, unique functional groups, and exceptional properties. They are significantly smaller than conventional materials and exhibit enhanced catalytic activity and increased thermal stability [1821]. Several methods are used for metal nanoparticles preparation, e.g., sol-gel, hydrothermal, chemical vapor deposition, laser vaporization, plane pyrolysis, precipitation, and biosynthesis [2123].

Nickel peroxide (NiO2) nanoparticles have been investigated for their potential use in various applications, including catalysis, energy storage, and biomedical applications. They exhibit high stability, high thermal conductivity, and excellent electrochemical properties [24]. The synthesis of nickel oxide nanoparticles can be achieved using various methods, including sol-gel, hydrothermal, and thermal decomposition techniques [25, 26]. Sodium hypochlorite (bleaching house) is cheap, available, and environmentally acceptable chemical [27]. The synthetic applications of NaOCl in organic synthesis have been reported as a part of researchers’ interest in the use of transition metals as a catalysts in organic synthesis [2831]. NiO2 serves as both a stoichiometric oxidant and a catalyst for the oxidation of benzyl alcohol. This implies that simultaneous stoichiometric and catalytic reactions can potentially take place in different reactions [31, 32].

According to Ji et al. [32], nanosized NiO2 powder can be readily obtained by oxidizing Ni(II) salt with a hypochlorite solution, followed by a wet chemical method and calcination.

In this study, NiO2 nanoparticles will be prepared at room temperature without using any capping agent compared to recently reported [26, 33, 34]. In addition, the optimum conditions and the mechanism of the catalytic dehydrogenation of benzylamine and some parasubstituents to their corresponding nitriles by the catalytic system, NiSO4·6H2O/NaOCl, in an aqueous basic medium were investigated. Furthermore, the emphasis is placed on investigating several factors to optimize the reaction conditions.

2. Materials and Methods

2.1. Materials

NiSO4·6H2O, benzylamine (BzNH2), and p-substituted benzylamine R-BzNH2 (R = CH3, OCH3, OH, NH2, CHO, CN, NO2, and F3C) were used without purification and purchased from Sigma-Aldrich (Germany, Berlin) and NaOCl from local market.

2.2. NiO2 Nanoparticles Preparation

NiO2 nanoparticles were synthesized following a procedure described in the literature [31] but with a substitution of NaOCl and NaOH with K2S2O8 and KOH. In a typical method, 10 mL of 1.0 M NaOH was added to a solution containing NiSO4·6H2O (0.052 g, 0.2 mM) in 5 mL of H2O. Subsequently, commercial NaOCl (10 mL, 14 mM, 5.25%, 0.7 M) was added dropwise with stirring, resulting in the formation of fine black particles. The precipitate was then filtered, washed with distilled water to remove NaCl, air-dried, and finally packed:

2.3. Characterization

The phase identification (X-ray diffraction) was characterized using a JSX-60PA/Jeol diffractometer (Japan) equipped with a Ni-filtered CuKα radiation (λ = 1.5418 Å). Transmission electron microscopy (TEM) was carried out by using a JEM 100 CXII transmission electron microscope operating at 80 kV. The scanning electron microscopy (SEM) was done using FEGSEM, Thermo Scientific Quattro S. Energy dispersive spectrometer (EDS) measurements were carried out using LEO 1455 VP. Infrared (IR) spectra were recorded on the Alpha Bruker FTIR spectrophotometer (model no. 200695, Berlin, Germany). 1H NMR spectra were recorded on a Bruker DPX (600 MHz, London, UK). Melting points were determined in open capillaries on an electrically heated metal block.

2.4. Catalytic Dehydrogenation of Benzylamine

In a typical experiment, in a 50 mL Schlenk flask, 10 mL of NaOH (1.0 M) was added to the solution of NiSO4·6H2O (0.052 g, 0.2 mM) in 5 mL of H2O and then the commercial NaOCl (10 ml, 14 mM, 5.25%, 0.7 M) was added dropwise with stirring, and a fine black particle was formed. Benzylamine (1.1 g, 10 mM) was added to the reaction mixture with stirring for two hours and then NaHSO3 (10 mL, 10%) was added to quench the excess of NaOCl. The reaction mixture was extracted with CH2Cl2, and the extracts were dried over Na2SO4 anhydrous. The product was isolated and characterized by comparing its spectroscopic data with an authentic sample where appropriate [35] as follows:Benzonitrile: b.p. 188–190°C; 1H NMR (400 MHz, CDCl3) δ: 7.44 (d, 2H, 2 × CH), 7.51 (d, 2H, 2 × CH), 7.54 (t, 1H, CH); IR (KBr, , cm−1): 3067, 2229 cm−1 [35].p-Methylbenzonitrile: m.p. 25–27°C; 1H NMR (400 MHz, CDCl3) δ: 7.23 (2H, d, J = 7.4 Hz, H-Ar), 7.47 (2H, d, J = 7.4 Hz, H-Ar). IR (KBr, , cm−1): 3055, 2224 cm−1 [35].p-Methoxybenzonitrile: m.p. 56–57°C; 1H NMR (400 MHz, CDCl3) δ: 3.881 (s, 3H CH3O), 6.977 (d, 2H, ArH), 7.599 (d, 2H, ArH). IR (KBr, , cm−1): 3060, 2232 cm−1 [35].p-Hydroxybenzonitrile: m.p. 109–110°C; 1H NMR (400 MHz, CDCl3) δ: 6.014 (s, 1H, OH), 6.954 (d, 2H, ArH), 7.558 (d, 2H, ArH). IR (KBr, , cm−1): 3062, 2229 cm−1 [35].

2.5. Recycling of Nickel Oxide

The produced NiO was washed by using acetone and water and then stirred with tenfold excess of NaOCl solution for 30 min. The recovered NiO2 was used similarly as above (Section 3.2) for another dehydrogenation reaction.

3. Results and Discussion

3.1. Characterization of NiO2 Nanoparticles
3.1.1. XRD Analysis

The X-ray diffraction pattern in Figure 1 displays the characteristic pattern of α-type NiO2 nanoparticles formed with the assistance of NaOCl synthesis. The diffraction peaks observed at angles of 17.68°, 25.47°, 36.8°, and 61.48° corresponded to crystal planes (003), (006), (101), and (110), respectively. These peaks closely matched those found in JCPDS card no. 38-0715 [36]. The broadening of these peaks can be attributed to the formation of nanoparticles with a nanostructure, resulting in a smaller particle size [37]. Under the current experimental conditions, the broadening observed in certain diffraction peaks of NiO2 nanoparticles XRD patterns could be attributed to the existence of minuscule particle-like structures. Furthermore, the XRD pattern exhibited “saw-tooth” reflections that are typical of two-dimensional turbostratic phases with layers that lack orientation [38]. The Debye–Scherer’s equation was utilized to determine the average crystallite size of NPNP as follows [39]:where D is the average crystallite size, λ = 1.54056 Å is the wavelength of CuKα radiation, β is the full width at half-maximum (FWHM) intensity of the peak in radian, θ is Bragg’s diffraction angle, and K is a constant usually equal to 0.9. Table 1 exposes that all XRD data of the sample and the estimated crystallite size (by using the Scherrer relation) found to vary between 11 nm and 29 nm (average 20 nm) for various identified diffraction peaks confirm the formation of the nanocrystalline structure which is in accordance with the TEM analyses.

3.1.2. SEM, TEM, and EDS

The SEM micrographs in Figure 2(a) exhibit the synthesized NiO2 nanoparticles, indicating that the particles are spherical in shape and have formed nanoclusters due to the accumulation process. The nanosized crystallites may experience agglomeration because of their small size, which results in a large surface energy. This causes the nanocrystals to aggregate during crystal growth, thereby reducing their surface energy. The SEM images reveal that the average size of the observed NPNP is 24 nm. The evaluation of crystallinity was performed by comparing the crystallite size determined through SEM analysis. The resulting crystallinity index is presented in the following:

The crystallinity index, denoted as Icry, was determined using the following parameters: Dp = 24 nm for the crystallite size (obtained from SEM image) and Dcry = 20 nm for the crystallite size (calculated from the Debye–Scherrer equation). Since the value of Icry = 1.2 is higher than one, it is inferred that the crystallite size corresponds to a polycrystalline structure [40].

The transmission electron microscopy (TEM) analysis of the NiO2 nanoparticles, as illustrated in Figure 2(b), revealed their uniform size characterized by spherical shapes and smooth surfaces, exhibiting even distribution. Nevertheless, the small size and high surface energy of some particles led to their aggregation into secondary particles [41]. In addition, the crystals consisted of mainly spherical particles with the size of 15–20 nm from the TEM observation, which is in good agreement with the result from the XRD patterns.

The EDS elemental analysis in Figure 2(c) confirms that the NiO2 nanoparticles consist solely of Ni and O elements. However, the results obtained from EDS analysis (Table 2) reveal some deviations between theoretical and experimental values, which can be attributed to the presence of water attached to the NPNP [42].

3.1.3. FTIR Spectrum NiO2 Nanoparticles

The FTIR spectrum of the NiO2 nanoparticles sample exhibited typical features of NPNP. The observed spectrum exhibited a significant and wide peak at 3390 cm–1, indicating the stretching vibration of O-H bonds in interlayer water molecules and hydrogen-bonded OH groups. However, the distinct peak associated with the stretching mode of independent Ni-OH groups was not present due to the hydrogen bonding occurring between hydrogen atoms and intercalated anions or water molecules within the layers. In addition, the peak at 1622 cm–1 was assigned to the bending vibration of water. The bands due to carbonate ion are too weak to be identified, suggesting a low carbonate ion content. The band at 1370 cm–1 was characteristic of interlayer SO42− stretching vibration resulted as a contaminant during the preparation process; the vibration at approximately 660 cm–1 was associated with Ni-O-H bonds, whereas a minor and faint vibration centered around 570 cm–1 was attributed to Ni-O stretching [4345]. The FTIR spectrum of NiO2 nanoparticles in Figure 3 displays significant absorption bands. The absorption band in the 500–700 cm⁻1 range indicates nanocrystals, and the small sample size caused the IR band associated with Ni-O stretching vibration to shift towards the blue region. This shift is due to quantum size effect and spherical nanostructures of NPNP [41].

3.2. Catalytic Dehydrogenation of Benzyl Amine

Yamazaki [46] reported the stoichiometric dehydrogenation of benzylamine to benzonitrile by NaOCl in ethanol; this encouraged us to investigate the catalytic dehydrogenation of benzylamine and some parasubstituents containing electron-donating and electron-withdrawing groups by the catalytic system NiSO4·6H2O/NaOCl/NaOH (pH = 14) to their corresponding nitriles, as shown in Scheme 1.

Benzylamine was selected as a model substrate for the optimization of the reaction conditions. The results are given in Table 3.

Benzylamine was dehydrogenated by the in situ-generated NiO2 nanoparticles to benzonitrile and water according to the following equation:

A set of experiments were carried out by the catalytic system NiSO4·6H2O (0.2 mM)/NaOCl (20 mL, 5.25%)/NaOH (10 mL, 1.0 M) (entry 1, Table 3).

It was found that benzylamine (10 mM) was smoothly dehydrogenated to benzonitrile in 95% yield within two hours.

On the other hand, a control experiment was performed in the absence of NiSO4·6H2O; the yield of benzonitrile was 20% (entry 2, Table 3). This result is in agreement with the reported stoichiometric dehydrogenation of benzylamine [46].

The effect of the amount of NiSO4·6H2O on the yield of benzonotrile was studied by performing two reactions; it was found that the yield of benzonitrile was not improved. That is because the experiments with 0.1 mM and 0.3 mM of NiSO4·6H2O gave 51% and 94% of benzonitrile, respectively (entries 3 and 4, Table 3).

The effect of the amount of the co-oxidant was investigated by conducting one experiment with 5 mL of NaOCl (5.25%) and another one with 10 mL of NaOCl (5.25%). The yields of benzonitrile were 50% and 60%, respectively (entries 5 and 6, Table 3); these results were probably because the amount of the produced NiO2 was not enough to dehydrogenate the whole amount of benzylamine and this consequently led to the formation of some other biproducts such as benzaldehyde and N-benzylidenebenzylamine.

To study the effect of the reaction time, two experiments were conducted under the same conditions of entry 1 but with different reaction times, 0.5 h and 1 h, and it was found that the yields of benzonitrile were 35% and 44%, respectively (entries 7 and 8, Table 3). These low yields were probably because these reaction times did not allow the catalytic dehydrogenation to go to completion in addition to the formation of some other side products like N-benzylidenebenzylamine.

3.2.1. The Produced NiO

The obtained NiO (equation (1)) powder was filtered and washed with deionized water and recycled by treatment with NaOCl and NaOH to produce NiO2 nanoparticles that were used for the dehydrogenation of 10 mM of benzylamine. The yield obtained was 80% of benzonitrile (entries 9, Table 3); this recycling process was repeated two times and gave 60% and 55% of benzonitrile, respectively (entry 10 and 11, Table 3). The observed low yields in the recycling process can be attributed to the catalyst’s loss during the reaction workup, as well as the loss of its active sites.

3.2.2. Reaction Conditions

All reactions were conducted at room temperature using 20 ml of NaOCl (5.25%), 10 mL of NaOH (1.0 M), 0.2 mM NiSO4·6H2O, and 10 mM of benzylamine. Entries 2, 3, and 4 are as follows: the amounts of NiSO4·6H2O were 0.1, 0.3, and 0 mM, respectively. Entries 5 and 6 are as follows: the amounts of NaOCl (5%) were 5 mL and 10 mL, respectively. Entries 7 and 8 are as follows: the reaction times were 0.5 h and 1 h, respectively. Entries 9, 10, and 11 are the recycling of the produced NiO powder after the workup. Y = yield (%) = number of moles of produced nitrile × 100/number of moles of benzylamine; TO = turn over = number of moles of product/number of moles of catalyst; TOF (h−1) = turn over frequency = number of moles of product/number of moles of catalyst per hour.

The effect of the parasubstituents on the yield of the nitrile was studied via the dehydrogenation of four substituents containing electron-donating groups (p-CH3, p-CH3O, OH, and NH2) and four substituents containing electron-withdrawing groups (p-CHO, p-CN, p-NO2, and p-CF3).

It was noticed that the yields that were found with the electron-withdrawing group (entries 16, 17, 18, and 19, Table 4) were lower than those with the electron-donating substituents (entries 12, 13, 14, and 15, Table 4).

The reason behind these observations could be attributed to the activation of the ring by electron-donating groups, which leads to an increase in the dehydrogenation of benzylamine into the corresponding nitrile. Conversely, electron-withdrawing groups deactivate the phenyl ring, causing a slowdown in the catalytic dehydrogenation process.

The reaction exhibited self-indication, as evidenced by the color change of the reaction mixture from green to black upon the addition of NaOCl, indicating the formation of NPNP. Over time, this color gradually disappeared with the formation of NiO. However, this catalytic dehydrogenation reaction is considered to be selective and catalytic (benzylamine is converted mainly to benzonitrile) and inexpensive. In addition, dehydrogenation was performed at room temperature. However, the obtained results were analogous with some recently reported protocols for the catalytic dehydrogenation of benzylamine but it is appeared to be superior, simpler, and more practical than the most previously reported approaches [4749]. Dutta et al. reported the dehydrogenation of benzylamine to benzonitrile by a ruthenium(II) complex bearing a naphthyridine-functionalized pyrazole ligand in 84% yield in toluene and at 70°C [50] (our protocol did use any organic additive or organic solvent). Benzonitrile was obtained from benzylamine in 80% yield after 24 hours in dichloroethane as a solvent and at 110°C (higher temperature and longer reaction time than those applied in this catalytic system) [51]. Recently, the complex [RuCl2(p-cymene)]2 catalyzed the conversion of benzylamine to nitrile in dichlorobenzene and in the presence of Me4NCl at 150°C [52] (here, water was used as an ecofriendly solvent). Benzonitrile was produced from the photocatalytic oxidation of benzylamine by zirconium trisulfide (ZrS3) but the workup is complicated and involved many steps [53]. To conclude, this protocol was highly selective, producing only nitrile as the product, and water as the sole byproduct. Moreover, it was free from the use of any toxic solvent or chemical.

3.3. Mechanism of Catalysis

The time-dependent profile illustrated the mechanism of dehydrogenation of benzylamine to benzonitrile through tracing the amount of benzylamine (% BzNH2) and the produced amount of benzonitrile (% BzCN) with time. As expected, the amount of benzylamine was decreased and the amount of benzonitrile was increased, as shown in Figure 4.

The in situ-generated NiO2 nanoparticles that were produced from the reaction of NiSO4·6H2O with NaOCl and in the presence of NaOH (equation (1)) were attacked by benzylamine forming the unstable intermediate PhCH2NH2NiO2 (equation (5)), which is further dehydrogenated through the hydride abstraction mechanism [54] to liberate the corresponding imine (equation (6)). This imine reacted with one another molecule of NiO2 to form the unstable PhCHNHNiO2 (equation (7)) that underwent dehydrogenation to liberate benzonitrile and water (equation (8)).

The produced NiO is further oxidized by the excess of NaOCl to produce NiO2 as follows:

This catalytic cycle was repeated until the amount of benzylamine was completely consumed, as shown in Scheme 2.

4. Conclusion

NiSO4·6H2O and NaOCl in the presence of aqueous 1.0 molar KOH were used to generate NiO2 nanoparticles. The particles were characterized and found to be spherical in shape with an average size of 20 nm, confirming their nanocrystalline structure. The catalytic system showed good performance in the dehydrogenation of benzylamine and some parasubstituents to benzonitriles. Optimum reaction conditions were determined, and the yields, turnover, and turnover frequency were calculated. The reaction mechanism was analyzed, and it was determined that the in situ-generated NPNP was the active species accountable for the dehydrogenation process. This promising reaction has several advantages as follows: the chemicals are inexpensive, water is used as a solvent at ambient temperature, the yields and turnover are good, and the produced NiO can be recycled for further catalytic reactions.

Data Availability

The datasets generated and analyzed during the current study are available from the corresponding author upon request.

Ethical Approval

All authors have read, understood, and have complied as applicable with the statement on “Ethical responsibilities.”

Conflicts of Interest

The authors declare that they have no conflicts of interest.

Authors’ Contributions

Shoair was the research director and performed objection discussion and article screening. Khalaf performed data interpretation, rote the characterization section and revised the final form. Almalki performed data extraction, methodology, and literature quality evaluation. Shanab and Nasef performed data extraction, methodology, and meta-analysis. Sheta performed study designing and article-draft writing. El-Basiony performed article-draft writing. Nasef performed data extraction, proposed the methodology, and performed meta-analysis. All the authors agreed with the final version of the article and the authors list. All the authors have read and approved the final manuscript.

Acknowledgments

The authors would like to thank the Deanship of Scientific Research, Taif University, Saudi Arabia, for supporting this work.