Lipeeka Routa,
Aniket Kumara,
Rajendra S. Dhakab and
Priyabrat Dash*a
aDepartment of Chemistry, National Institute of Technology, Rourkela, Orissa 769008, India. E-mail: dashp@nitrkl.ac.in
bNovel Materials and Interface Physics Laboratory, Department of Physics, Indian Institute of Technology Delhi, HauzKhas, New Delhi-110016, India
First published on 4th May 2016
Bimetallic nanoparticles, particularly those based on copper, have recently attracted a great deal of attention for the development of low cost and highly active catalysts due to the synergistic interaction between individual metal components. In this work, bimetallic Ag–Cu alloy nanoparticles were explored as a highly active and reusable catalyst for the enamination of 1,3-dicarbonyls using diverse amines. The nanocatalysts were intensively characterized by ultraviolet-visible (UV-Vis) spectroscopy, X-ray diffraction (XRD), high-resolution transmission electron microscopy-energy-dispersive spectroscopy (HRTEM-EDS) and valence band and core level X-ray photoelectron spectroscopy (XPS) to study the effect of the bimetallic structure and composition. In comparison to monometallic Ag and Cu nanoparticles, the alloyed Ag–Cu nanoparticles showed a high catalytic performance and the resultant catalytic activity was dependant on the Ag to Cu ratio. This enhanced catalytic activity should be related to the electronic interaction between Ag and Cu nanoparticles formed due to the intimate contact between them. Our study may serve as a foundation for designing efficient alloyed nanocatalysts for fine chemical synthesis via enamination reactions.
The enamination of 1,3-dicarbonyl compounds to form β-enamino ketones and esters synthesized from low-cost raw materials with a very stable structural pattern is considered to be a very important process as these are valuable precursors in organic synthesis.25,26 These compounds are utilized as significant intermediates for the preparation of various heterocyclic compounds.27,28 They are also known to have medicinal applications as anticonvulsants29,30 and anti-inflammatory31,32 and antitumor agents.33 Because of their extensive variety of action and potency, a straightforward and high yielding one-pot approach for the synthesis of β-enaminones and β-enaminoesters is highly desirable. The environmentally benign protocol used for the synthesis of β-enaminones is found to be a time-consuming method.34 Several improved procedures have been reported for the reaction between dicarbonyl compounds and amines using catalysts such as metal triflates by microwave and ultrasound irradiation.35–38 Other synthetic processes to yield β-enaminones and β-enaminoesters includes the cyclization of amino acids, the acylation of amines and condensation reactions.39 The synthesis of some fine chemicals has been reported using different catalysts such as Yb(OTf)3,40 perchlorates,41 [Hbin]Tf],42 HClO4–SiO2,43 montmorillonite K10,44 silica gel,45 natural clays,46 silica chloride,47 InBr348 and CoCl2.49 All the methods discussed so far have their own drawbacks and limitations such as moisture sensitivity and lengthy workups;40 the requirement of harsh conditions and the use of harmful reagents;41 the use of homogenous48 or non-recyclable catalysts;49 and longer reaction times and low yields.49 Hence, there is sufficient scope for the development of a heterogeneous and reusable catalytic system that is able to catalyse the synthesis of β-enaminones and β-enaminoesters at milder operating conditions.
In recent years, transition metal NPs have been used for various organic transformations due to their high surface area and availability of numerous co-ordination sites.2,50 However, these catalysts still suffer from complicated synthesis protocols and low efficiency and are prone to oxidation. Because of the presence of the synergistic effect in bimetallic catalysts and the vast opportunities for engineering the particle size, shape and composition, it is expected that bimetallic catalysts will have a high potential for applications in fine chemical synthesis. Moreover, bimetallic NPs may also provide better stability and functionality at a lower cost.51,52 In this regard, the design of bimetallic NPs involving silver and copper will provide a stable catalyst where silver will help in mitigating the oxidation of copper. In addition, the activity can be controlled by simply changing the Ag to Cu ratio. To the best of our knowledge, the use of bimetallic NPs as an efficient catalyst for the synthesis of fine chemicals has not been demonstrated so far.
Herein, we report the synthesis of bimetallic Ag–Cu NPs via a low-temperature simple co-reduction method. The composition of the Ag–Cu NPs was tuned by changing the ratio of the metal precursors (3:
1, 1
:
1 and 1
:
3). The formation of the bimetallic Ag–Cu NPs was examined using energy dispersive spectroscopy (EDS) measurements, which helps to determine the distribution of the Ag and Cu components in the bimetallic structure. More importantly, the change in the electronic properties of the bimetallic NPs due to alloying has been studied using core and valence band (VB) X-ray photoelectron spectroscopy (XPS) analysis. Later on, these bimetallic nanoparticles were used as efficient catalysts for the one pot synthesis of β-enaminones and β-enaminoesters. Compared with monometallic NPs, the bimetallic Ag1–Cu3 NPs exhibited a much enhanced activity, providing a high yield of the desired product in less time. In this study, various amine precursors, such as cyclic and branched amines with bulkier groups, were employed for the synthesis of β-enaminones and β-enaminoesters. The catalyst was found to be recyclable up to 4 cycles without any significant loss of activity or yield of desired product. We envision that the current strategy will provide useful clues for the design of novel bimetallic NPs and bimetallic NP-based systems as potential catalysts for other fine chemical syntheses.
![]() | ||
Fig. 1 UV-Vis spectrum of (a) Ag, (b) Cu, (c) CuSO4·5H2O, (d) Ag1Cu1, (e) Ag1–Cu3 and (f) Ag3–Cu1 bimetallic NPs. |
Fig. 2 shows the XRD results of the synthesized monometallic Ag, Cu and bimetallic Ag–Cu NPs. The XRD analysis was conducted with a fully dried powder of the NPs. The prominent peak in the XRD pattern indicates that the samples obtained were of high crystallinity. Major peaks at 2θ values of 38.5°, 44.5°, 64.7° and 77.7° for the Ag NPs correspond to the lattice planes of (111), (200), (220) and (311) of the metallic Ag (JCPDS card no. 03-0931), respectively. In the case of the copper NPs, diffraction peaks formed at the 2θ values of 43.5°, 50.7° and 74.4° correspond to the lattice planes of (111), (200) and (220) from metallic Cu (JCPDS card no. 02-1225). In bimetallic Ag–Cu NPs, all the above Ag and Cu diffraction peaks were observed, suggesting that the bimetallic NPs consist both of Ag and Cu phases. It is important to note that the typical oxide peaks at 61.7° due to CuO and 37.5° due to Cu2O phase are both absent in the bimetallic spectra, demonstrating the formation of an oxide free Ag–Cu bimetallic system.57 From the major diffraction peak, the lattice parameters were obtained and are shown in Table 1. The lattice parameters were found to be in agreement with the selected area electron diffraction (SAED) results as described in the later part of the discussion. Taking the (111) reflection of the XRD spectra, the average crystallite size was calculated using the Debye–Scherrer formula.58,59 The crystallite size of the Ag1–Cu3 bimetallic NPs was found to be around 10 nm, which nearly matches with the particle size obtained from the TEM analysis discussed later.
Peak position (2θ) | Lattice parameter (Å) | |||
---|---|---|---|---|
[111] | [200] | d111 | d200 | |
Ag | 38 | 44.1 | 2.32 | 1.98 |
Cu | 43.1 | 49.5 | 2.08 | 1.76 |
The morphology and structural behaviour of the Ag1–Cu3 catalyst (chosen because of its higher catalytic activity, as described in the catalysis section) were analysed by TEM. Fig. 3a shows the representative TEM image, which shows that the product consists of uniform spherical NPs. The histogram (inset) shows an average size of 7.5 ± 1 nm for the NPs. The crystalline nature of the Ag1–Cu3 NPs was confirmed by SAED analysis which depicts a ring like structure (Fig. 3b). The (111), (200), (220) and (311) rings are indexed to the face centred cubic (fcc) structure of Ag and the (111), (200) and (220) rings are attributed to the fcc structure of Cu.60 These patterns indicated that the nanoparticle is polycrystalline. For further demonstration of the alloy nanostructure, HRTEM and EDS analysis was carried out. The HRTEM image of the Ag1–Cu3 nanoparticle shown in Fig. 3c can be resolved into two groups of lattice fringes. The lattice distance of 0.23 nm corresponds to the lattice fringe distance of the (111) plane of Ag, while the lattice distance of 0.20 nm corresponds to the (111) plane of Cu. The EDS analysis of the Ag1–Cu3 NPs shows the presence of Ag and Cu with a concentration of 25.8 and 74.2%, respectively (Fig. 3d). All the above results confirm the successful synthesis of the bimetallic Ag1–Cu3 NPs.
Though the results from XRD, TEM, SAED and HRTEM are consistent, they are not sufficient to fully confirm the bimetallic alloyed structure and distribution of the Ag and Cu NPs. In order to obtain conclusive confirmation of the alloy structure of the Ag1–Cu3 NPs, HAADF-STEM, EDS elemental mapping patterns and EDS line scanning profiles were performed and are shown in Fig. 4. The HAADF-STEM image (Fig. 4a) shows a clear luminance between Ag and Cu, suggesting an alloy structure of the as-prepared Ag1Cu3 NPs rather than a core–shell structure of Ag and Cu.61–63 EDS elemental mapping patterns further revealed that Ag and Cu elements are uniformly distributed throughout the whole bimetallic Ag1Cu3 nanostructure, indicating the formation of an Ag–Cu alloy without phase segregation, as shown in Fig. 4b (overlap, Ag–Cu), 4c (red, Ag) and 4d (green, Cu). The results demonstrated that Ag and Cu are well overlapped in the entire part. The line scan along the direction derived in Fig. 4e demonstrated that Ag and Cu are mixed well in the NPs.64–67 The EDS analysis indicates an average composition of approximately 1:
3 of Ag and Cu, which is in agreement with the molar percentage of the respective slats taken during the synthesis. This further shows the high compositional uniformity of the bimetallic NPs.
![]() | ||
Fig. 4 (a–d) HAADF-STEM-EDS mapping of the Ag1–Cu3 NPs and (e) EDS line scan of the Ag1–Cu3 bimetallic NPs. |
We performed XPS measurements to determine the chemical composition and oxidation state of the Ag1–Cu3 NPs. A survey scan, as shown in Fig. 5a, revealed the presence of elemental Ag, Cu and O in the sample. In Fig. 5b and c, we show the core-level spectra of Ag 3d and Cu 2p, respectively. The binding energy (BE) values of the Ag 3d5/2 and Ag 3d3/2 core levels for the Ag1–Cu3 NPs appear at 367.1 and 373.3 eV, respectively, which are associated with pure Ag NPs. The absence of peaks at 367.3 and 367.6 eV suggest that there is no significant formation of AgO and Ag2O species, respectively, in our nanoparticle sample.68–70 In the Cu 2p core-level spectrum, we observed two strong peaks at 931.0 eV and 950.7 eV, which are associated with Cu 2p3/2 and Cu 2p1/2, respectively. These values are found to be consistent with those reported for Cu(0).71 Both the Ag and Cu data suggested the metallic nature of Ag and Cu in the Ag1–Cu3 bimetallic NPs, which is consistent with the XRD and TEM analysis. Though the Cu 2p spectrum demonstrated that most of the Cu exists in the form of the metallic Cu(0) (931.0 eV and 950.7 eV) state, the presence of a small amount of Cu(II) (934.1 eV and 954.5 eV) can also be found. The presence of Cu(II) was further confirmed by a satellite peak at 942.0 eV. This can be attributed to the oxidation of surface Cu atoms in air.72–74 It is interesting to find that while XRD does not show any evidence of the CuO phase, XPS analysis demonstrates the surface presence of Cu2+ ions, which suggests that CuO is present only on the surface.75 Similar characteristic XPS spectra containing shakeup satellite peaks were also reported in the literature, in which an excess of copper was used in relation to other species.22 Moreover, by fitting the spin–orbit splitting peaks and taking the area under the curve, we calculated the surface composition of Ag:
Cu, which was found to be 1
:
2.8 (Table 2), very close to the original value of 1
:
3. This further suggests the oxidation stability of the Ag–Cu alloy NPs, along with the presence of significantly smaller amounts of CuO, possibly formed due to the chemisorption of small amounts of O2 on the nanoparticle surface. Due to this, the core-level peaks of Ag 3d5/2 and Cu 2p3/2 were found to be asymmetric towards the higher BE side. In addition, it can be observed that the BE of the Ag 3d5/2 and Cu 2p3/2 core level peak position in the bimetallic Ag1–Cu3 NPs shifted to lower energies (∼1.3 for Ag and ∼1.4 for Cu) in comparison to their bulk values (368.3 eV for Ag and 932.4 eV for Cu). Abrikosov et al. observed the shift in the core-level peaks when alloy formation occurs, which was explained in terms of the intra-atomic charge re-distribution caused by valence electron hybridization.76 Therefore, the observed shift in the Ag 3d5/2 and Cu 2p3/2 core-levels suggests the alloy formation of the Ag1–Cu3 NPs. Also, these peak shifts in the XPS spectra further suggest the possible electronic interaction (synergistic effect) between Ag and Cu in the Ag1–Cu3 bimetallic nanoparticle. This is because, in comparison to Cu, Ag has a higher redox potential, which leads to lower electron densities in the Cu atom in the bimetallic Ag1–Cu3 nanoparticle.
![]() | ||
Fig. 5 (a) XPS survey spectra of the Ag1–Cu3 bimetallic NPs, (b) Ag 3d core level XPS spectra of the Ag1–Cu3 bimetallic NPs, (c) Cu 2p core level XPS spectra of the Ag1–Cu3 bimetallic NPs. |
Sample | Ag | Cu | Ag![]() ![]() |
---|---|---|---|
Ag![]() ![]() ![]() ![]() |
26.4 | 73.6 | 1![]() ![]() |
Besides the shift in the core-level peak positions in the metal NPs, the changes in the VB spectra are crucial to understand the formation of alloy NPs.77 Therefore, we have recorded the XPS VB spectra of the Ag1–Cu3 nanoparticle (as shown in Fig. 6), which shows a broad peak centred around 3 eV below the Fermi level. The characteristic features of the Ag1–Cu3 VB can be seen as: a broad peak at ∼2.5 eV and a shoulder ∼4.5 eV. To understand the contribution from Ag and Cu, the VB spectrum is deconvoluted with two peaks at around 2 eV and 4.5 eV, which are attributed to the Cu 3d and Ag 4d, respectively. These values are relatively in agreement with the alloy NPs reported in the literature, further demonstrating the formation of the alloy Ag1–Cu3 nanoparticle.2 Moreover, there is a shift in the VB peak position of 1.5 and 0.5 for Cu 3d and Ag 3d, respectively, as compared to their bulk values (Cu = 3.5 and Ag = 5 eV).78,79 This shift of VB energy indicates the charge separation behaviour of Ag and Cu and suggests the possibility of an electronic interaction between Ag and Cu in the nanoparticle alloy formation.78,80,81
![]() | ||
Scheme 1 One pot synthesis of β-enaminones and β-enaminoesters using a Ag1–Cu3 bimetallic nanoparticle catalyst. |
It was observed that in comparison to monometallic NPs, bimetallic NPs of different ratios showed the best performance for the catalytic reaction, providing higher yields of the desired product (Table 3). Fig. 7 shows the catalytic activity of the bimetallic series of Ag–Cu NPs for the model reaction; bimetallic NPs that were rich in Cu (Ag1–Cu3) showed the highest activity. Such electronic enhancement effects in bimetallic nanoparticle catalysts have been previously documented in various Ag–Cu bimetallic catalysts and other bimetallic systems (Ag–Pd and Au–Pd).82–85 Therefore, the Ag1–Cu3 nanocatalyst was chosen as the preferred catalyst for optimizing the reaction conditions. As shown in Table 3, entry 1, for the control experiment (without metallic NPs), only trace products were formed, confirming that almost no condensation takes place without metal NPs.
Entry | Catalyst | Solvent | Temp. (°C) | Catalyst loading (mg) | Yieldb (%) |
---|---|---|---|---|---|
a Reaction conditions: acetyl acetone (1 mmol), aniline (1 mmol), catalyst loading (20 mg), methanol (5 mL), temperature (60 °C) and time (1 h 45 min).b Isolated and unoptimized yield. | |||||
1 | Without catalyst | Methanol | 60 | 20 | Traces |
2 | Ag NPS | Methanol | 60 | 20 | 30 |
3 | Cu NPs | Methanol | 60 | 20 | 50 |
4 | Ag3Cu1 | Methanol | 60 | 20 | 85 |
5 | Ag1Cu3 | Methanol | 60 | 20 | 95 |
6 | Ag1Cu1 | Methanol | 60 | 20 | 92 |
![]() | ||
Fig. 7 Catalytic activities of the monometallic and bimetallic NPs over a series of catalytic reactions. |
Further, the influence of various reaction parameters, such as solvent, temperature and catalyst loading, was studied by using the model reaction. In order to verify the effect of different solvent media on the condensation of the model reaction, solvents with different polarity, such as acetonitrile, tetrahydrofuran (THF), dichloromethane (DCM), hexane, ethanol and toluene, were used to optimize the reaction conditions (Table 4). It was noticed that for a nonpolar solvent, such as hexane, the yield of the product is much lower. The yield of the product was found to be higher when polar solvents such as methanol, ethanol, and acetonitrile were used. Among the various solvents tested, the highest yields of β-enaminones and β-enaminoesters were obtained in methanol solvent, which can be attributed to the interaction of the strong hydrogen bond in methanol that stabilises the reaction intermediates and increases the rate of reaction.86 Therefore, methanol was used to further study the catalytic reaction. The reaction was also performed under solvent free conditions to investigate the influence of the solvent parameters, but only traces of the product were found (Table 4, entry 8). These observations indicated that the solvent plays an important role in the condensation reaction, in which bimetallic Ag1–Cu3 NPs promote the mass diffusion and transport of the reactants.69 Later on, the amount of catalyst loading was varied from 5 mg to 25 mg. It was observed that the yield of the reaction involving the condensation of 1 mmol of acetylacetone and 1 mmol of aniline increases up to a catalyst dose of 20 mg. Further increasing the catalyst dose only marginally affects the yield of the product (Table 5, entries 1–5). The best result with respect to yield was obtained for 20 mg of Ag1–Cu3 catalyst (Table 5, entry 4). Pure Ag and Cu NPs showed very low activity (a yield of 30% and 50%, respectively) with the same catalyst loading (Table 5, entry 6–7). Hence, 20 mg of catalyst was utilised for further catalytic studies. Finally, the influence of temperature on the activity of the Ag1–Cu3 catalyst for the model reaction was studied varying the temperature from room temperature to reflux conditions. As shown in Fig. 8, it can be observed that upon increasing the temperature, the yield of the desired product initially increased. Upon further increasing of the temperature, i.e. at reflux conditions, the yield of the product significantly decreased. Based on these findings, the optimum reaction conditions were found to be, catalyst: Ag1–Cu3 bimetallic NPs, solvent: methanol, catalyst loading: 20 mg, and temperature: 60 °C.
Entry | Solvent | Catalyst | Catalyst loading (mg) | Temperature | Yieldb (%) |
---|---|---|---|---|---|
a Reaction conditions: acetyl acetone (1 mmol), aniline (1 mmol), methanol (5 mL), time (1 h 45 min) and temperature (60 °C).b Isolated and unoptimized yield. | |||||
1 | Methanol | Ag1–Cu3 NPs | 5 | 60 °C | 58 |
2 | Methanol | Ag1–Cu3 NPs | 10 | 60 °C | 72 |
3 | Methanol | Ag1–Cu3 NPs | 15 | 60 °C | 88 |
4 | Methanol | Ag1–Cu3 NPs | 20 | 60 °C | 95 |
5 | Methanol | Ag1–Cu3 NPs | 25 | 60 °C | 95.5 |
6 | Methanol | Ag NPs | 20 | 60 °C | 30 |
7 | Methanol | Cu NPs | 20 | 60 °C | 50 |
![]() | ||
Fig. 8 Effect of temperature on the percentage yield of enaminoesters synthesized by the one pot condensation of ethylacetoacetate and methyl amine. |
After optimizing the reaction conditions, we further investigated the scope of the optimized protocol using different dicarbonyl compounds and different substituted aliphatic and aromatic amines. Excellent yields and a high purity of products were obtained in all cases (Table 6, entries 1–40). All the condensation reactions were completed within 45–180 min at 60 °C without formation of any side products. Acetylacetone was found to give a good yield of 95% with aniline (Table 6, entry 14). Aliphatic and alicyclic amines, such as methyl amine, ethyl amine, propyl amine, butyl amine, and morpholine, were found to give an excellent yield of the condensed product with acetylacetone using Ag1–Cu3 bimetallic NPs as the catalyst (Table 6, entries 13–16, and 18). It was found that the reaction with aliphatic amine proceeded smoothly in a short period of time as compared to aromatic amines, which can be attributed to the higher nucleophilicity of aliphatic amines than aromatic amines.87 Later on, a variety of amines possessing a wide range of functional groups were selected for further studies. It was observed that the amines containing electron donating groups provided better yields in less time as compared to the amines containing electron withdrawing groups.88 For example, anisidine and p-nitroaniline, containing the electron withdrawing groups OCH3− and NO2−, respectively, have a strong deactivating effect, resulting in lower yields (Table 6, entries 20–21).
Entry | R | Amine (R1) | Product | Time | Yieldb (%) |
---|---|---|---|---|---|
a Reaction conditions: dicarbonyl compound (1 mmol), amine (1 mmol), Ag1–Cu3 NPs (20 mg), methanol (5 mL) and temperature (60 °C).b Isolated and unoptimized yield.c In this reaction, the catalyst was recycled for four consecutive cycles and showed the same activity without any significant loss of yield. Amine (R1) – entry (1–3, 13–16, 26–28, 38) alkyl amine, entry (4, 7–8, 17, 20–21, 29, 32–33, 39) aryl amine. | |||||
1 | OC2H5 | CH3 | ![]() |
1 h | 91 |
2 | OC2H5 | C3H7 | ![]() |
2 h | 90 |
3 | OC2H5 | C4H9 | ![]() |
1 h 10 min | 88 |
4 | OC2H5 | C6H5 | ![]() |
2 h 30 min | 93 |
5 | OC2H5 | ![]() |
![]() |
1 h 45 min | 85 |
6 | OC2H5 | ![]() |
![]() |
1 h 30 min | 88 |
7 | OC2H5 | PhNO2 | ![]() |
2 h 15 min | 75 |
8 | OC2H5 | C7H9NO | ![]() |
2 h | 82 |
9 | OC2H5 | ![]() |
![]() |
1 h 45 min | 91 |
10 | OC2H5 | ![]() |
![]() |
1 h | 92 |
11 | OC2H5 | ![]() |
![]() |
6 h | 72 |
12 | OC2H5 | ![]() |
![]() |
2 h | 83 |
13 | CH3 | CH3 | ![]() |
45 min | 93 |
14 | CH3 | C2H5 | ![]() |
45 min | 92 |
15 | CH3 | C3H7 | ![]() |
1 h 15 min | 92 |
16 | CH3 | C4H9 | ![]() |
45 min | 91 |
17 | CH3 | C6H5 | ![]() |
1 h 45 min | 95c |
18 | CH3 | ![]() |
![]() |
1 h | 85 |
19 | CH3 | ![]() |
![]() |
50 min | 90 |
20 | CH3 | PhNO2 | ![]() |
2 h 45 min | 82 |
21 | CH3 | C7H9NO | ![]() |
2 h 30 min | 85 |
22 | CH3 | ![]() |
![]() |
2 h | 90 |
23 | CH3 | ![]() |
![]() |
45 min | 95 |
24 | CH3 | ![]() |
![]() |
5 h | 75 |
25 | CH3 | ![]() |
![]() |
1 h 30 min | 85 |
26 | OCH3 | C2H5 | ![]() |
2 h 30 min | 92 |
27 | OCH3 | C3H7 | ![]() |
2 h 15 min | 90 |
28 | OCH3 | C4H9 | ![]() |
45 min | 91 |
29 | OCH3 | C6H5 | ![]() |
1 h | 90 |
30 | OCH3 | ![]() |
![]() |
1 h 30 min | 88 |
31 | OCH3 | ![]() |
![]() |
1 h 45 min | 91 |
32 | OCH3 | PhNO2 | ![]() |
3 h | 80 |
33 | OCH3 | C7H9NO | ![]() |
2 h 30 min | 87 |
34 | OCH3 | ![]() |
![]() |
1 h 45 min | 91 |
35 | OCH3 | ![]() |
![]() |
1 h 30 min | 90 |
36 | OCH3 | ![]() |
![]() |
6 h 30 min | 70 |
37 | OCH3 | ![]() |
![]() |
3 h | 80 |
38 | Ph | C4H9 | ![]() |
2 h | 80 |
39 | Ph | C6H5 | ![]() |
3 h | 85 |
40 | Ph | ![]() |
![]() |
2 h 15 min | 83 |
In addition, other bulkier and cyclic amines were tested in the condensation reaction. It was observed that the condensation of dicarbonyl compounds with 2,5-dimethyl aniline produce a very low yield of product (75%) even when the reaction was carried out for a longer reaction time. This lower yield of β-enaminones by the condensation reaction can be attributed to the steric hindrance between the two adjacent methyl groups (Table 6, entry 11). Amines with cyclic group such as cyclohexyl amine gave good yields in the condensation reaction (Table 6, entry 23). Substituted aromatic amines with electron donating substituents were more reactive and provide a better yield of the corresponding condensed product. The condensation of amines to other dicarbonyl compounds, such as ethyl acetoacetate, methylacetoacetate and benzoyl acetone, were also studied (Table 6, entry 1–12, 26–37 and 38–40). Among all the entries, the minimum yield was observed when benzoyl acetone was used as the β-dicarbonyl compound in the optimized condition. In benzoyl acetone, the presence of the electron withdrawing group (–COC6H5) generates weak acidic protons compared to its ester counterparts. Therefore, a lower yield of product was obtained in the case of benzoyl acetone.
From all the catalysis data, it can be concluded that bimetallic NPs of different ratios showed more activity in comparison to their monometallic counterparts. For the bimetallic Ag–Cu NPs, the electron density on the surface was higher than that of the monometallic NPs because of electron transfers from Cu to Ag owing to higher redox potential of the later. The core level peak shifts of Ag and Cu in the XPS analysis clearly suggest this electron transfer, resulting in a lower electron density of Cu in the bimetallic Ag–Cu NPs. This synergistic electronic effect behaviour enhances the catalytic activity in the bimetallic Ag–Cu NPs. In addition, the chemisorption strength (BE) of the organic moieties plays an important role in controlling the catalytic activity that depends on the d-band centre of the metal surface. According to the Hammer–Norskov model, the d-band centre of Ag is located at −3.9 eV, whereas that of Cu is located at −2.4 eV.89–91 When the organic moiety (diketones or amines) interacts with the d-band of Ag and Cu, an overlap of the adsorbate state with the metal state happens. In the case of the monometallic Ag and Cu NPs, this overlap results in weak binding, thereby lowering the activity of the catalyst. However, in the case of the bimetallic Ag–Cu NPs, an alloy structure is formed (confirmed by HRTEM-EDS and XPS) due to the close contact between Ag and Cu. This generates a stronger BE and higher activity in the case of the bimetallic Ag–Cu NPs. Therefore, the structural effects and composition effects in the bimetallic NPs play an important role in their enhanced catalytic performance.
Based on the observations discussed above, a plausible Ag1–Cu3 catalysed reaction mechanism for the condensation reaction is proposed in Scheme 2. The condensation product generally forms through an addition–elimination reaction. The Ag1–Cu3 NPs co-ordinate to the carbonyl oxygen of the enol form of acetylacetone (the enol form of acetyl acetone is more stable than the diketone), followed by the addition of aniline.92 This generates a tetrahedral intermediate after passing through a four-membered transition state (II–IV), which further undergoes a water elimination reaction to yield the imine moiety. The final product β-enaminones form after the tautomerization of the imine intermediate. Overall, the NPs play an important role in catalyzing/activating the reaction by properly binding or co-ordinating with the organic substrate, resulting in the lowering of the energy barrier required for all the intermediate steps in the formation of β-enaminones and β-enaminoesters. A similar observation has previously been documented for the binding of Cu NPs on organic substrates.93,94
S. no. | Catalyst | Product | Time | Catalyst amount (mol%) | Catalyst amount (mg) | Yield (%) | Particle size | Ref. |
---|---|---|---|---|---|---|---|---|
1 | Cu NPs | ![]() |
2 h 50 min | 10 | — | 92 | 20 nm | 77 |
2 | Ag NPs | ![]() |
8 h | 10 | — | 90 | 40 nm | 83 |
3 | Ag NPs in hollow magnetic spheres | ![]() |
8 h | — | 31 | 98 | 10 nm | 32 |
4 | Zn(oAc)2·H2O | ![]() |
48 h | 5 | — | 86 | — | 84 |
5 | Ag1–Cu3 NPs | ![]() |
1 h 45 min | 10 | 20 | 95 | 8 nm | This work |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra04569c |
This journal is © The Royal Society of Chemistry 2016 |