Binoyargha Dama,
Mithu Sahab,
Ramen Jamatiaa and
Amarta Kumar Pal*a
aDepartment of Chemistry, Centre for Advanced Studies, North-Eastern Hill University, Shillong-793022, India. E-mail: amartya_pal22@yahoo.com
bState Key Laboratory of Physical Chemistry for Solid Surfaces, College of Chemistry and Chemical Engineering, Xiamen University, China-361005
First published on 25th May 2016
Nano-organocatalyzed one-pot four-component reactions for the synthesis of phthalazine-trione/dione derivatives have been devised for the first time from easily accessible starting materials under solvent-free conditions. This methodology showed very good substrate scope and high degree of tolerance for a variety of aldehydes (including aliphatic and heteroaromatic aldehydes) and active methylene compounds. Moreover, the catalyst can be easily separated from the reaction mixture because of its highly paramagnetic nature, by using an external magnet, and can be reused in five more consecutive runs without much decrease in catalytic activities. Other significant advantages of this method are shorter reaction time, good yield, simple work-up procedure, easy catalyst handling etc.
Heterocyclic compounds possessing an aza group are widely distributed in nature and of a large number of N-containing compounds, those which possess hydrazine moiety at the fusion site are of immense importance because of their pharmacological, clinical and agrochemical applications.11,12 Many of these phthalazine derivatives possess anti-fungal,13 anti-cancer,14 anti-tumor15 and many other pharmaceutical activities.16 Because of these biological and medicinal activities, many researchers developed powerful and promising methods for synthesizing phthalazene molecules by using various types of catalysts like inorganic–organic hybrid materials Al-SBA-15-TPI/H6P2W18O62,17 TBBDA,18 Cs2CO3@magnetic nanoparticles,19 SO3H-FMSM,20 SBA@BiPy2+2Cl−,21 [TMG][Ac],22 RH@[SiPrDABCO@BuSO3H]HSO4,23 CoFe2O4–CS–SO3H (ref. 24) etc. in sequential multi-component reactions (MCRs). These reported methods possess many advantages, but at the same time have many disadvantages too. Previously our group also reported Ni(0) NP-catalyzed synthesis of phthalazine derivatives25 but major difficulties were agglomeration of Ni NPs and inability to form phthalazine molecules with aliphatic and heteroaromatic aldehydes. To address these problems, we applied a surface-stabilized magnetic nano-organocatalyst. In this way we prevented the chance of agglomeration and increased the catalytic activity of nanocomposite. From a literature study we have chosen glutathione as a stabilizer as well as organocatalyst, because it binds to NPs by using –SH functionality while –NH2 and –COOH groups are free for catalysis.10 Furthermore, for easy separation we have selected magnetic NPs (Fe3O4) which avoids the involvement of cumbersome procedures like centrifugation and filtration. Moreover, application of nano-FGT (nano-ferrite-supported glutathione) also made the methodology more efficient as compared to the one reported, since the reaction went smoothly with a diverse range of active methylene compounds and aldehydes including heteroaromatic and aliphatic aldehydes.
So, the chief purpose behind this presented protocol is to highlight the synergetic effects of combined nano-organocatalyst operated in an enhanced synthetic procedure, MCR and solvent-free reaction conditions for development of a new eco-compatible methodology. Therefore, we have chosen nano-FGT as a proficient, harmless, mild, magnetically recyclable, highly stable, powerful solid nano-organocatalyst and have applied it for the synthesis of phthalazine derivatives in four-component MCR under solvent-free conditions.
Formation of nano-FGT was confirmed by FT-IR, TEM, SEM and EDX analyses. Its thermal stability was analyzed by TGA. Functional group identification was done by the FT-IR technique. The FT-IR spectrum of nano-FGT (Fig. SI 1†) showed broad peaks at 3349 and 3166 cm−1 which indicated the presence of –OH and –NH2 groups. The peaks at 1643 and 1634 cm−1 correspond to the carbonyl stretching of acid and amide groups of glutathione respectively. The peak at 2926 cm−1 is due to C–H stretching. The absorption band at 599 cm−1 corresponds to Fe–O stretching of Fe3O4 NPs and absence of S–H stretching band at around 2525 cm−1 clearly indicated that glutathione is successfully anchored onto the surface of Fe3O4 NPs via thiol group.27 Morphology and particle size of nano-FGT were determined from TEM images. TEM images showed the presence of a dark spherical NP core of size 10–20 nm uniformly coated with a layer of glutathione [Fig. 1a and b]. SAED (selected area electron diffraction) of freshly prepared nano-FGT showed spotty diffraction, thereby proving its crystalline character [Fig. 1c]. EDX analysis was carried out to determine the elemental composition of nano-FGT. This confirmed the presence of Fe, O, S, N and C (Fig. 2). SEM images were obtained with the help of a JSM-6360 (JEOL) scanning electron microscope. The SEM image also confirmed the spherical morphology of nano-FGT (Fig. 3).
The prepared nano-FGT was also characterized by the powder XRD technique. The XRD diffraction pattern (Fig. 4) shows characteristic 2θ peaks at around 30.33°, 35.61°, 43.19°, 57.28°, 62.83°, 74.30° which are in good agreement with those reported in the literature.4
In order to gain information regarding thermal stability of nano-FGT, we performed TGA (thermogravimetric analysis). There is a weight loss at around 53 °C in the thermogram which corresponds to the degradation of solvent molecules trapped in the catalyst. There is another weight loss at 185 °C which is due to the degradation of glutathione molecules of nano-FGT. These observations prove that the catalyst is stable below 180 °C and therefore can be easily applied in reactions under that temperature (Fig. 5).
ICP-OES analysis was performed in order to determine the actual amount of Fe in the nano-ferrite heterogeneous catalyst. It was found to be 63.23%.
After successful surface modification and characterization of nano-FGT, we explored its applicability for the synthesis of phthalazine dione and trione derivatives by four-component condensation reaction of phthalic anhydride (1), hydrazinium hydroxide (2), active methylene compounds (3, 4 or 5) and aldehydes (6a–q) (Scheme 2). Initially a mixture of phthalic anhydride (1) (1 mmol), hydrazinium hydroxide (2) (1.2 mmol), dimedone (3) (1 mmol) and 4-chlorobenzaldehyde (6a) (1 mmol) was taken as a pilot reaction. Following which the reaction was stirred at room temperature without any catalyst. It was found that without catalyst the reaction failed to form the desired product even after 24 h of constant stirring. Only starting materials were evidenced using TLC. Then another new reaction was set up with addition of nano-FGT which was stirred according to known procedures.28 It is important to mention here that since the reaction is solvent free, reactants and magnetic nano-organocatalyst (nano-FGT) were mixed properly by using a glass rod into which a magnetic bit was added in order to maintain proper mixing during the reaction period, which in turn increases the contact between the surface of the catalyst and reactant molecules. Firstly, the reaction was carried out at room temperature, and to our delight it was found that after 12 h of continuous stirring the desired product was formed but the conversion was very poor, most of the starting materials remaining unreacted. This observation prompted us to apply thermal energy to the reaction. A number of reactions were set up in a pre-heated oil bath using a temperature-controlled magnetic stirrer at different temperatures ranging from 40 to 120 °C, the best result being obtained at 80 °C. However, no desired product was formed without addition of catalyst even at 80 °C after 24 h of stirring. All these observations proved the efficient role of catalyst in carrying out the reaction. The structure of the compound was identified by elemental and spectral analysis. The IR spectrum of product compound 15a showed absorption bands at 1453 and 1109 cm−1 which are due to C–N and N–N band stretching, and the band at 1666 cm−1 corresponds to carbonyl group of dimedone. In the 1H NMR spectrum of compound 15a, eight aromatic protons appear at δ 8.29–7.24. The methine proton was observed as a singlet at δ 6.34. The four methylene protons of the dimedone residue appeared as an AB system at δ 3.35–3.13 and singlet at δ 2.26. The two methyl groups were observed as two singlets at δ 1.137 and δ 1.132. Next, the reaction conditions were optimized and, for this, various reaction parameters were examined.
Firstly we focused on the optimization of the amount of catalyst required to catalyze the reaction. It was discovered that 10 mg (6.32 mg of Fe, 11.31 mol%) of the catalyst was sufficient to furnish maximum yield within a very short period of time. Application of higher amount of catalyst, viz. 12 mg, showed no increase in the yield of desired product, neither did it result in decreased reaction time. On the other hand, reduction in the amount of catalyst below 10 mg produced lower yield within the same amount of time. The optimization plot of the reaction time and the amount of catalyst is shown in Fig. 6.
Encouraged by the success discussed above, we investigated the effect of various solvents on the pilot reaction at 80 °C using 10 mg nano-FGT as a catalyst. In the presence of solvents like water, ethanol, acetonitrile and THF, bis-adduct 15a′, 2,2′-(4-chlorophenylmethylene)-bis(3 hydroxy-5,5-dimethylcyclohex-2-enone) (see ESI†), was the major product instead of phthalazine derivative 15a. But under solvent-free conditions, the reaction proceeded very smoothly and afforded phthalazine derivative 15a as major product within a very short period of time (Fig. 7).
Now in order to determine whether the catalytic activity of nano-FGT in the above reaction was the best, we compared its catalytic activity with that of other homogeneous and heterogeneous catalysts (Scheme 3). The cases where FeCl3, Fe2SO4 and Fe3O4 NPs were used as catalysts led to very much less conversion thereby leading to a poor yield of the desired product (Table 1). The fourth case where glutathione was used as the catalyst showed very good improvement in the result but the time required for the completion of the reaction was high, and in this case glutathione, being soluble in water, could not be recycled back. Application of nano-FGT however solved this problem of recyclability, and since it is attached to NPs which have higher surface/volume ratios, it shows better catalytic activity,9 which in turn reduces the time and increases the product yield.
No. | Catalystb | Time | Yieldc (%) |
---|---|---|---|
a Reaction conditions: phthalic anhydride (1 mmol), hydrazinium hydroxide (1.2 mmol), dimedone (1 mmol) and 4-chlorobenzaldehyde (1 mmol), 80 °C, SFRC.b Amount of catalyst: 10 mg.c Isolated yield. | |||
1 | No catalyst | 24 h | 0 |
2 | FeCl3 | 180 min | 21 |
3 | Fe2SO4 | 180 min | 25 |
4 | Fe3O4 NPs | 180 min | 53 |
5 | Glutathione | 120 min | 78 |
6 | Nano-FGT | 20 min | 97 |
To explore the generality of the reaction, we extended the scope of the present protocol for the synthesis of various phthalazine derivatives (Scheme 4). The results of this study are shown in Table 2. As shown in Table 2, aromatic aldehydes carrying either electron donating or withdrawing substituents worked well, giving excellent yield of products (87–97%) with high purity. The reaction went smoothly with various active methylene compounds like dimedone (3), 1,3-cyclohexanedione (4), ethyl cyanoacetate (5a), and malononitrile (5b). To analyze the generality of our methodology, the said synthesis was also performed with different heteroaromatic and aliphatic aldehydes (15o, 15p and 18a) under similar conditions and we were delighted to see the formation of our desired product in good yields within similar time duration. As per a literature survey, very much less reported procedure for the synthesis of phthalazine derivatives showed good results with aliphatic and heteroaromatic aldehydes, but our methodology provides excellent results with both kinds of aldehydes (Table 2). Reaction was also tried with ethyl acetoacetate as active methylene compound but no desired product formed after 6 h of constant stirring; only starting materials were recovered. All the synthesized products were characterized from their analytical and spectroscopic data. Structure of compound 17a was again confirmed by single-crystal X-ray crystallography (Fig. 8).
No. | Aldehyde | Active methylene compound | Product | Time (min) | Yieldb (%) | Mp (°C) |
---|---|---|---|---|---|---|
a Reaction conditions: phthalic anhydride (1 mmol), hydrazinium hydroxide (1.2 mmol), active methylene compounds (1 mmol) and aldehydes (1 mmol), 80 °C, SFRC,10 mg of nano-FGT.b Isolated yield. | ||||||
1 | 4-ClC6H5 (6a) | 3 | 15a | 20 | 97 | 261–262 (ref. 17) |
2 | 4-CNC6H5 (6b) | 3 | 15b | 20 | 96 | 271–273 (ref. 25) |
3 | 4-NO2C6H5 (6c) | 3 | 15c | 20 | 97 | 224–226 (ref. 17) |
4 | 4-BrC6H5 (6d) | 3 | 15d | 25 | 96 | 265–267 (ref. 17) |
5 | 4-FC6H5 (6e) | 3 | 15e | 20 | 95 | 219–220 (ref. 17) |
6 | 3-ClC6H5 (6f) | 3 | 15f | 20 | 92 | 205–207 (ref. 25) |
7 | 2-ClC6H5 (6g) | 3 | 15g | 23 | 91 | 266–268 (ref. 17) |
8 | 3-NO2C6H5 (6h) | 3 | 15h | 24 | 92 | 269–270 (ref. 25) |
9 | 3-BrC6H5 (6i) | 3 | 15i | 21 | 92 | 266–268 (ref. 25) |
10 | 3-FC6H5 (6j) | 3 | 15j | 21 | 90 | 271–273 (ref. 25) |
11 | 2-BrC6H5 (6k) | 3 | 15k | 22 | 91 | 262–265 (ref. 25) |
12 | C6H5 (6l) | 3 | 15l | 25 | 93 | 205–207 (ref. 17) |
13 | 4-MeC6H5 (6m) | 3 | 15m | 25 | 87 | 228–230 (ref. 17) |
14 | 4-MeOC6H5 (6n) | 3 | 15n | 20 | 89 | 220–222 (ref. 25) |
15 | 2-C5NH4 (6o) | 3 | 15o | 30 | 88 | 227–230 (ref. 24) |
16 | C4H8 (6p) | 3 | 15p | 30 | 87 | 133–136 (ref. 22) |
17 | 4-NO2C6H5 (6c) | 4 | 16a | 20 | 94 | 237–239 |
18 | 4-Cl,3-NO2C6H5 (6q) | 5a | 17aa | 20 | 97 | 212–214 |
19 | C4H8 (6p) | 5b | 18a | 25 | 92 | 220–223 |
![]() | ||
Fig. 8 Molecular structure of compound 17a (CCDC 1438715). |
The most probable mechanism of the said conversion is shown in Scheme 5. In the initial step nano-FGT facilitates the nucleophilic attack of hydrazinium hydroxide (2) to phthalic anhydride (1) leading to the formation of phthalhydrazide (7). On the other hand, aldehydes react with nano-FGT forming iminium intermediate (8).9 Then cyclic 1,3-diketones (3 and 4) form 9 which reacts with 8 and affords Knoevenagel product 11 (path a)9 where acyclic active methylene compounds (5a or 5b) generate Knoevenagel product 12 through carbanion formation (path b).29 Phthalhydrazide (7) attacks these intermediates (11 and 12) under Michael fashion leading to the formation of other intermediates 13 and 14 which eventually undergo cyclization followed by dehydration to furnish our desired product.
![]() | ||
Scheme 5 Plausible mechanism for the synthesis of 2H-indazolo[1,2-b]phthalazine-triones and 1H-pyrazolo[1,2-b]phthalazine-diones. |
Following this, the possibility of recovery and reusability of the catalyst was investigated. After completion of the reaction, the catalyst was recovered from the reaction mixture simply by using an external magnet and washed with ethyl acetate and acetone and dried. The recovered catalyst was reused under the same reaction conditions, and it was found that the catalyst could be reused five times without any significant loss of the product yield. The reproducibility of the reaction was then investigated by performing five sets of reactions under the same conditions. Each set was repeated five times and the mean yield was calculated (95.6, 94.8, 94.2, 93.2 and 92.6%) by the standard deviation method; the error bars were also determined (Fig. 9). After carrying out the 5th set of reactions, again ICP-OES analysis was performed and the amount of Fe found was 60.04%. This observation confirmed there is 3.19% leaching of the catalyst. Then, in order to confirm the structure of the reused (after 5 consecutive runs) catalyst, we performed powder XRD (Fig. 10a), SEM (Fig. 10b), TEM (Fig. 10c), FT-IR (Fig. SI 1b†) and TGA (Fig. 10d) analyses. The structure was very similar to that of freshly synthesized catalyst and it also remained well dispersed and retained its size (10–20 nm).
![]() | ||
Fig. 10 (a) Powder XRD pattern, (b) TEM image of nano-FGT at 50 nm, (c) SEM image of nano-FGT at 500 nm and (d) TGA curve of nano-FGT after the 5th run. |
Footnote |
† Electronic supplementary information (ESI) available. CCDC 1438715. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6ra06376d |
This journal is © The Royal Society of Chemistry 2016 |