Introduction

Natural materials and the supported transition metals are great candidates for catalytic applications due to their availability, relatively low-cost, bio-degradability, and bio-compatibility1,2,3,4,5,6,7,8,9. Boehmite is one of the attractive natural mineral materials, which was recently employed as support in fabricating organometallic catalysts1,10,11. In fact, boehmite is aluminum oxide hydroxide (γ-AlOOH), which is the most stable alumina phase in nature after gibbsite12,13,14,15. Moreover, boehmite turns into γ-Al2O3, δ-Al2O3, θ-Al2O3, and α-Al2O3 at temperatures of 450 °C, 900 °C, 1000 °C, and 1200 °C, respectively16.

Nowadays, it is accepted that boehmite consists of double layers of octahedron structure from oxygen and a central aluminum atom with a cubic orthorhombic unit cell in which aluminum is surrounded by six oxygen atoms17,18,19. The two layers are connected via hydrogen bonds20. Boehmite can be synthesized using various methods such as hydrolysis of aluminum salts19, solid-state decomposition of gibbsite21, precipitation in an aqueous solution from aluminum salt solutions22, sol–gel procedures23, hydrothermal procedures24 and solvothermal procedures12. Most of the boehmite synthetic methods customize the morphology, shape and surface characteristics (pore volume, specific surface area, and pore structure) and also physical and chemical properties. However, boehmite nanomaterials have been rarely used as an ideal support to stabilize various catalysts due to their unique properties such as excellent surface area (> 300 m2/g), nanometer size of crystallite, non-toxicity, easy availability, and mechanical and thermal stability (up to 450 °C)18,25,26. In this sense, it is worth mentioning that boehmite nanomaterials can be employed as a support to stabilize several transition metals catalysts such as palladium27, nickel28, platinum29, molybdenum30,31, vanadium31, cobalt32, copper33, manganese34, iron35, rhodium36, ruthenium37 and zirconium28. Boehmite nanomaterials have also attracted attention in coatings38, absorbents39, flame retardant40, ceramics41, optical materials42, vaccine adjuvants43, cosmetic products32 and in synthesizing alumina as starting materials44. However, boehmite nanoparticles have some specific disadvantages (e.g. impurities of nitrate ions), which can affect the surface property and pore crystalline structure. On the other hand, as mentioned above, boehmite nanoparticles may be converted to Al2O3 at high temperatures; but, as the organic reactions take place at temperatures lower than the boehmite phase change, this cannot effect on the boehmite application in catalysis.

Samarium has rarely been reported as a catalyst for the synthesis of organic compounds. Regarding the fact that selective and reusable catalysts are the main principle of green chemistry, herein a new complex of samarium was investigated using 1,3-bis(pyridin-3-ylmethyl)thiourea on boehmite nanoparticle (Sm-bis(PYT)@boehmite) as a stable, practical, and recyclable nanocatalyst in the homoselective synthesis of 5-substituted 1H-tetrazole derivatives in PEG-400 as a green solvent. Due to the act that one of the principles of green chemistry is to use reusable, cheap and sustainable catalysts, different heterogeneous supported catalysts based on mesoporous materials45,46,47,48, MOF structures49, boehmite50,51, carbon materials52, polymers53, magnetic particles54,55,56,57,58 etc. have been reported as catalysts. Moreover, another principle of green chemistry is to apply safe solvents (such as PEG) and safe reagents.

Besides, tetrazole derivatives which are an important group of organo-heterocyclic compounds can be used in various fields such as coordination chemistry, drugs, synthetic organic chemistry, medicinal chemistry as surrogates for carboxylic acids, catalysis technology, the photographic industry, and organometallic chemistry as effective stabilizers of metallopeptidase structures59,60,61,62,63,64,65,66,67,68,69,70. Furthermore, tetrazole derivatives have some specific biological properties, i.e. analgesics, herbicides, anti-proliferative, anti-inflammatory, antimicrobial, anti-HIV, and anticancer properties18,71,72,73,74,75,76,77,78.

Results and discussion

In the first step, modified boehmite nanoparticles through (3-chloropropyl)triethoxysilane were synthesized based on the new procedure60. Subsequently, a new complex of samarium was fabricated on their surface (Sm-bis(PYT)@boehmite) (Fig. 1). Sm-bis(PYT)@boehmite was employed as a practical, reusable, and homoselective nanocatalyst used to synthesize tetrazoles. This is the first report proposing the immobilization of 1,3-bis(pyridin-3-ylmethyl)thiourea on boehmite nanoparticles. Moreover, this is the first report in which the samarium complex of 1,3-bis(pyridin-3-ylmethyl)thiourea was used as a catalyst in the synthesis of organic reactions. Therefore, it can be said that this catalyst is an innovation in organic reactions. This nanocatalyst was characterized using N2 adsorption–desorption isotherms, TGA, EDS, WDX, SEM, FT-IR, ICP-MS, DLS and XRD techniques.

Figure 1
figure 1

Synthesis of Sm-bis(PYT)@boehmite.

FESEM-TESCAN MIRA III Scanning Electron Microscope device from Czechia country was employed to study the shape, morphology, and diameters of Sm-bis(PYT)@boehmite. The SEM images are shown in Fig. 2. As shown, the particles of Sm-bis(PYT)@boehmite are formed in uniform spherical shapes and relatively homogeneous diameters of less than 70 nm. Dynamic light scattering (DLS) of Sm-bis(PYT)@boehmite is shown in Fig. 3. Based on DLS analysis, the diameter of Sm-bis(PYT)@boehmite was found to be 129.55 nm, which is greater than the SEM analysis results due to the agglomeration and solvation of the catalyst particles in water79.

Figure 2
figure 2

SEM images of Sm-bis(PYT)@boehmite.

Figure 3
figure 3

DLS analysis of Sm-bis(PYT)@boehmite.

Regarding the elemental composition of Sm-bis(PYT)@boehmite, the energy-dispersive X-ray spectroscopy (EDS) analysis was used—indicating that Sm-bis(PYT)@boehmite is organized from Al, O, Si, C, N, S, and Sm elements. As shown in Fig. 4, the intensity of aluminum and oxygen peaks is higher than other elements, which form the skeleton of boehmite nanoparticles. Furthermore, the presence of Si, C, N, S, and Sm elements indicates the successful stabilization of the samarium complex on the surface of boehmite nanoparticles. In addition, wavelength dispersive X-ray spectroscopy (WDX) analysis (Fig. 5) illustrates the homogeneous distribution of Al, Si, O, C, N, S, and Sm in the structure of Sm-bis(PYT)@boehmite. Moreover, the exact amount of samarium metal was calculated using ICP-MS analysis. The exact content of samarium—in the structure of Sm-bis(PYT)@boehmite—was obtained as 0.386 mmol/g.

Figure 4
figure 4

EDS diagram of Sm-bis(PYT)@boehmite.

Figure 5
figure 5

Elemental mapping of (a) Al, (b) Si, (c) O, (d) C, (e) S, (f) N, (g) Sm and (h) combine of all elements for Sm-bis(PYT)@boehmite.

TGA analysis can be employed to determine the content of inorganic materials and organic compounds in an organic–inorganic hybrid sample and to calculate thermal stability. Therefore, TGA analysis of Sm-bis (PYT)@boehmite was recorded from 25 to 800 °C within an increasing temperature rate of 10 °C/min under air atmosphere (Fig. 6). In this analysis, a small weight loss (7% of weight) up to 150 °C occurred, corresponding to the evaporation of solvents80,81. As shown, no weight loss occurred at up to 300 °C except evaporation of solvents, indicating excellent thermal stability of Sm-bis(PYT)@boehmite. Therefore, Sm-bis(PYT)@boehmite can be used as a catalyst under hard conditions in a wide range of organic reactions. TGA analysis of Sm-bis(PYT)@boehmite illustrated a considerable mass loss (27% of weight) between 300 and 600 °C due to the decomposition of immobilized organic layers on the surface of modified boehmite nanoparticles60.

Figure 6
figure 6

TGA diagram of Sm-bis(PYT)@boehmite.

The X-ray diffraction (XRD) pattern is shown in Fig. 7. The XRD pattern exhibited several peaks of 2θ position at 14.01° (0 2 0), 27.93° (1 2 0), 38.19° (0 3 1), 45.89° (1 3 1), 49.64° (0 5 1), 51.81° (2 0 0), 55.24° (1 5 1), 58.66° (0 8 0), 64.34° (2 3 1), 66.98° (0 0 2), 68.49° (1 7 1), and 72.30° (2 5 1), which can be indexed to the cubic orthorhombic unit cell of boehmite according to X-ray diffraction pattern (Joint Committee on Powder Diffraction Standards (JCPDS)-No. 00-049-0133 and JCPDS-No. 01-074-1895)82,83,84,85,86. Moreover, a list of the identified patterns from XRD results of boehmite NPs (AlO(OH)) is summarized in Table 1.

Figure 7
figure 7

Original XRD pattern of boehmite NPs (a), Observed peaks list from normal XRD pattern of boehmite NPs (b), standard pattern code 00-049-0133 of boehmite NPs (c), and standard pattern code 01-074-1895 of boehmite NPs (d).

Table 1 Identified patterns list from XRD results of boehmite NPs (Aluminum Oxide Hydroxide).

X-ray diffraction (XRD) pattern of Sm-bis(PYT)@boehmite is presented in Fig. 8. The XRD pattern of Sm-bis(PYT)@boehmite shows several peaks of 2θ position at 14.49° (0 2 0), 28.29° (1 2 0), 38.84° (0 3 1), 46.09° (1 3 1), 49.69° (0 5 1), 52.09° (2 0 0), 56.34° (1 5 1), 58.99° (0 8 0), 64.89° (2 3 1), 66.19° (0 0 2), 67.99° (1 7 1), and 71.99° (2 5 1). These peaks confirm that boehmite nanoparticles are stable as the orthorhombic unit cell (according to XRD pattern codes JCPDS-No. 00-049-0133 and JCPDS-No. 01-074-1895)12,60 after modification or stabilization of Sm-complex. Moreover, several peaks of 2θ value at 19.09°, 21.59°, 27.59°, 28.94°, 30.84°, 32.04°, 39.44°, 45.24°, 47.64°, and 55.54° positions are related to the samarium on boehmite nanoparticles87. In addition, a broad peak of 2θ from 16° to 25° is related to the amorphous SiO288.

Figure 8
figure 8

Normal XRD pattern of Sm-bis(PYT)@boehmite.

The N2 adsorption–desorption isotherms of Sm-bis(PYT)@boehmite are shown in Fig. 9. Moreover, the results of the BET analysis are summarized in Table 2. Based on Brunauer–Emmett–Teller (BET), the surface area of Sm-bis(PYT)@boehmite is 5.30 m2/g, which is lower than the unmodified boehmite nanoparticles (which reported about 128.8 m2/g19). In this sense, it is worth mentioning that, the pore volume of Sm-bis(PYT)@boehmite was calculated at 0.012 cm3/g, which is lower than the unmodified boehmite nanoparticles (reported at about 0.22 cm3/g19). As observed, BET surface area and pore volume of Sm-bis(PYT)@boehmite are lower than the unmodified boehmite nanoparticles due to the grafting of Sm-complex on boehmite nanoparticles.

Figure 9
figure 9

N2 adsorption–desorption isotherms of Sm-bis(PYT)@boehmite.

Table 2 Textural properties of boehmite nanoparticles and Sm-bis(PYT)@boehmite.

The FT-IR spectra of boehmite, boehmite@CPTMS, bis(PYT)@boehmite and Sm-bis(PYT)@boehmite are shown in Fig. 10. Three peaks at 484, 623, and 743 cm−1 can be related to the Al–O vibrations in the structure of boehmite89. The stretching vibration of Si–O–Si at 1077 cm−147,90 indicates successful surface modification of boehmite nanoparticles. Also, the stretching vibration of C–H bonds can be observed at around 3000 cm−1 in the FT-IR spectra90,91—which not observed in the FT-IR spectrum of boehmite nanoparticles—indicating successful surface modification of boehmite nanoparticles. Moreover, two bands at 1559 cm−1 and 1429 cm−1 correspond to the aromatic C=C bonds73,47. The stretching vibration of C=N is observed at 1633 cm−173.

Figure 10
figure 10

FT-IR spectra of (a) boehmite, (b) boehmite@CPTMS, (c) bis(PYT)@boehmite and (d) Sm-bis(PYT)@boehmite.

Catalytic study of Sm-bis(PYT)@boehmite

The catalytic application of Sm-bis(PYT)@boehmite is illustrated in the synthesis of tetrazole derivatives (Fig. 11). The best reaction conditions were found in the [3 + 2] cycloaddition of sodium azide (NaN3) and benzonitrile as the model reaction (Table 3). The model reaction cannot take place in the absence of Sm-bis(PYT)@boehmite catalyst (Table 3, entry 1). In this sense, the existence of Sm-bis(PYT)@boehmite is required to synthesize 5-substituted 1H-tetrazole derivatives. As accepted, the reaction proceeds in presence of the catalyst and, then, proceeds faster while increasing the amount of the catalyst. As shown, by increasing the amount of the catalyst up to 50 mg (containing 1.93 mol% of samarium), the model reaction can be completed within an acceptable time (Table 3, entry 3). As shown, when the amount of Sm-bis(PYT)@boehmite increased from 40 to 50 g, the turnover number (TOF) value improved more than twice. Among the several tested solvents (such as PEG-400, dimethyl sulfoxide (DMSO) and H2O), PEG-400 provided the best results in terms of TOF, reaction time, and isolated yield of the product (Table 3, entry 3). Furthermore, the effects of temperature and NaN3/benzonitrile ratio were checked on the model reaction and, accordingly, the best results were obtained at 120 °C with 1.4 mmol of NaN3 per 1 mmol of benzonitrile. As shown in Table 3, the best results (including TOF number, turnover number (TON), reaction time, and yield of the product) were obtained in presence of 50 mg of the catalyst in PEG-400 at 120 °C.

Figure 11
figure 11

The general method used to synthesize 5-substituted 1H-tetrazoles in presence of Sm-bis(PYT)@boehmite nanocatalyst.

Table 3 Definition of the best reaction conditions for synthesizing 5-substituted 1H-tetrazoles in presence of Sm-bis(PYT)@boehmite.

To show the effect of the catalytic activity of the supported samarium, the [3 + 2] cycloaddition of benzonitrile with NaN3 under optimized conditions was studied in the presence of bare boehmite nanoparticles, boehmite@CPTMS, bis(PYT)@boehmite, and samarium(III) acetate salt. The obtained results were compared with the same reaction in the presence of Sm-bis(PYT)@boehmite (Table 4). As shown in Table 4, appropriate results can not be obtained for the synthesis of 5-phenyl-1H-tetrazole in the presence of unfunctionalized boehmite nanoparticles, boehmite@CPTMS, and bis(PYT)@boehmite, while 5-phenyl-1H-tetrazole was synthesized in the presence of Sm-bis(PYT)@boehmite in 90 min with a yield of 96%.

Table 4 Synthesis of 5-phenyl-1H-tetrazole in the presence of boehmite nanoparticles, boehmite@CPTMS, bis(PYT)@boehmite, samarium(III) acetate salt and Sm-bis(PYT)@boehmite.

The scope of catalytic application of Sm-bis(PYT)@boehmite catalyst was extended in the [3 + 2] cycloaddition of NaN3 and other benzonitrile derivatives (Table 5). In this regard, benzonitriles having electron-withdrawing or electron-donating groups (on para- meta- or ortho-position of aromatic ring) were investigated, and all corresponding tetrazoles were synthesized in good yields.

Table 5 Synthesis of 5-substituted 1H-tetrazole derivatives catalyzed by Sm-bis(PYT)@boehmite.

TON and TOF values are two valuable factors to evaluate the efficiency and practicality of catalysts. As shown in Table 5, all tetrazole products can be obtained with good TON and TOF numbers in the presence of Sm-bis(PYT)@boehmite catalyst. Therefore, one of the most important innovations of this work is the good TON and TOF values of the obtained products in the presence of Sm-bis(PYT)@boehmite catalyst.

As shown in Fig. 12, Sm-bis(PYT)@boehmite catalyst shows a good homoselectivity in the synthesis of tetrazoles through [3 + 2] cycloaddition of sodium azide and benzonitrile derivatives, which has two similar cyano groups in their structure such as phthalonitrile (Table 5, entry 4). As shown in Table 5 (entry 4) and Fig. 12, this strategy provided only monoaddition, which indicates an excellent homoselectivity of this catalyst.

Figure 12
figure 12

Homoselectivity of Sm-bis(PYT)@boehmite in the synthesis of 5-substituted 1H-tetrazoles from [3 + 2] cycloaddition of NaN3 with dicyano substituted derivatives.

Based on authentically reported strategies for the synthesis of tetrazoles that are catalyzed by transition metal catalysts19,95, a catalytic cycle mechanism for the production of tetrazoles in the presence of Sm-bis(PYT)@boehmite catalyst is offered in Fig. 13.

Figure 13
figure 13

An expected mechanism for synthesizing 5-substituted 1H-tetrazoles in the presence of Sm-bis(PYT)@boehmite.

Comparison of the catalyst

The advantages of Sm-bis(PYT)@boehmite catalyst over previous catalysts was investigated in the synthesis of 5-phenyl-1H-tetrazole through [3 + 2] cycloaddition of benzonitrile with NaN3 in the presence of Sm-bis(PYT)@boehmite and previous catalysts (Table 6). As shown, Sm-bis(PYT)@boehmite catalyst affords 96% of the product in 90 min, which is better than the previous catalysts in terms of time and yields. Moreover, some of previous catalysts have several limitations or drawbacks, such as long reaction times, non or difficult separation of the catalysts and utilizing hazard solvents. Significantly, herein the synthesis of tetrazoles was introduced in a green solvent such as PEG, having a short reaction time and acceptable yield in presence of reusable Sm-bis(PYT)@boehmite catalyst.

Table 6 Comparison results of Sm-bis(PYT)@boehmite with other catalysts for the synthesis of 5-phenyl-1H-tetrazole.

Recycling ability and leaching study of the catalyst

As mentioned, TGA analysis indicates that Sm-bis(PYT)@boehmite catalyst is stable at up to 300 °C. Therefore, the reusability of Sm-bis(PYT)@boehmite catalyst was investigated in the synthesis of 5-phenyl-1H-tetrazole through [3 + 2] cycloaddition of benzonitrile and NaN3. In this regard, at the end of the reaction, the reaction mixture was diluted and, then, the catalyst was recovered by filtration and reused in the next run. As shown in Fig. 14, Sm-bis(PYT)@boehmite catalyst can be recovered and reused for up to 6 runs without further activation.

Figure 14
figure 14

The recoverability and reusability of Sm-bis(PYT)@boehmite nanocatalyst in the synthesis of 5-phenyl-1H-tetrazole.

The samarium leaching from Sm-bis(PYT)@boehmite catalyst in the synthesis of tetrazoles was studied applying hot filtration test and ICP-MS analysis. Thus, the [3 + 2] cycloaddition of benzonitrile and NaN3 in the presence of Sm-bis(PYT)@boehmite catalyst was repeated and, accordingly, the catalyst was removed after 0.75 h. The reaction mixture was allowed to continue without catalyst presence for up to 1.5 h. The reaction did not proceed after catalyst removing, signifying that samarium was not leached under reaction conditions.

In addition, this reaction was repeated and, at the end of the reaction, the catalyst was removed by simple filtration. Afterwards, the exact amount of the leached samarium in the reaction solution was calculated using ICP-MS analysis. No significant amount of the leached samarium was detected in the reaction solution.

The recovered Sm-bis(PYT)@boehmite nanocatalyst was characterized applying FT-IR, XRD, SEM, EDS, WDX and ICP-MS techniques.

The XRD pattern of the recovered Sm-bis(PYT)@boehmite nanocatalyst is illustrated in Fig. 15. A good agreement between fresh and recovered catalysts can be seen. No change was observed in the XRD pattern after reusing. These results verify the good stability of Sm-bis(PYT)@boehmite after reusing in the synthesis of tetrazoles.

Figure 15
figure 15

The powder XRD pattern of recovered Sm-bis(PYT)@boehmite.

The FE-SEM images of the recovered Sm-bis(PYT)@boehmite nanocatalyst are recorded using an electron microscope (FE-SEM ZEISS Sigma 300) model SIGMA VP from the Zeiss Company of German (Fig. 16). As shown, no significant change was observed in FESEM images of this catalyst after recovering in terms of particles size or shape. The SEM images indicate that the particles of the recovered Sm-bis(PYT)@boehmite have uniform spherical shapes and quite homogeneous diameter less than 70 nm, as compared to the catalyst before recovering.

Figure 16
figure 16

FESEM images of recovered Sm-bis(PYT)@boehmite.

Furthermore, the electron microscope (FE-SEM ZEISS Sigma 300) model SIGMA VP from the Zeiss Company of German was employed for the elemental content of the recovered Sm-bis(PYT)@boehmite using EDS analysis. EDS diagram of the recovered Sm-bis(PYT)@boehmite is shown in Fig. 17. The EDS analysis indicated that the presence of all elements (e.g. Al, O, Si, C, N, S, and Sm elements) in the structure of the recovered Sm-bis(PYT)@boehmite shows a good agreement with the fresh catalyst.

Figure 17
figure 17

EDS diagram of recovered Sm-bis(PYT)@boehmite.

In addition, WDX analysis (Fig. 18) of the recovered Sm-bis(PYT)@boehmite illustrates a homogeneous distribution of Al, Si, O, C, N, S, and Sm in the structure of the recovered Sm-bis(PYT)@boehmite, as compared to this catalyst before recovering.

Figure 18
figure 18

Elemental mapping of (a) Al, (b) Si, (c) O, (d) C, (e) S, (f) N and (g) Sm for recovered Sm-bis(PYT)@boehmite.

Also, the FT-IR spectrum of the recovered Sm-bis(PYT)@boehmite was compared with the FT-IR spectrum of fresh Sm-bis(PYT)@boehmite. Excellent agreement was observed in the FT-IR spectra of the recovered and the fresh catalyst in terms of the position and shape of the stretching vibrations (Fig. 19). Therefore, Sm-bis(PYT)@boehmite is stable under reaction conditions during the synthesis of tetrazole derivatives.

Figure 19
figure 19

FT-IR spectra of (a) Sm-bis(PYT)@boehmite and (b) recovered Sm-bis(PYT)@boehmite.

FT-IR, XRD, SEM, EDS, WDX and ICP-MS analysis of the recovered Sm-bis(PYT)@boehmite proved the good stability of this catalyst after being reused in the synthesis of tetrazoles. Therefore, this catalyst is stable under reaction conditions of the tetrazoles synthesis. Therefore, it can be recovered and reused for several runs without any change in its structure.

Methods

Synthesis of 1,3-bis(pyridin-3-ylmethyl)thiourea ligand (3)

In a round-bottomed flask, 10 mmol of 3-(aminomethyl)pyridine (1) was added to 5 mmol of carbon disulfide (2) in water and, then, stirred at room temperature for 7 h. The reaction progress was consecutively monitored using thin-layer chromatography (TLC, ethyl acetate: n-hexane, 1:2). Since this reaction is exothermic, the temperature increases during the reaction and, therefore, the heat is sufficient to release H2S (4, confirmed by smell and lead acetate paper, which turns black). After completion of the reaction, the water-insoluble product (3) was filtered and, then, recrystallized from hot water and ethanol (1: 1 v/v).

Synthesis of catalyst

In order to prepare boehmite, 50 mL of an aqueous solution of sodium hydroxide (6.490 g) was added to 30 mL of an aqueous solution of aluminum nitrate (20 g) as drop-by-drop under vigorous stirring. The resulting milky mixture was transferred to an ultrasonic bath for 3 h at room temperature. The resulting boehmite was filtered and, then, washed with distilled water. Subsequently, the obtained boehmite was kept in the oven at 220 °C for 4 h60.

The modified boehmite nanoparticles were synthesized according to a reported procedure in the literature60. In this regard, 1.5 g of boehmite nanoparticles was dispersed in n-hexane by sonication for up to 0.42 h and, then, 2 mL of (3-choloropropyl)triethoxysilane (CPTMS) was added to the reaction mixture. The obtained reaction mixture was stirred for 24 h under reflux conditions. The modified boehmite nanoparticles were produced through CPTMS (boehmite@CPTMS). The prepared boehmite@CPTMS was filtered, washed with ethanol, and dried at room temperature. In order to immobilize of 1,3-bis(pyridin-3-ylmethyl)thiourea ligand (3) on boehmite@CPTMS (bis(PYT)@boehmite), 1 g of boehmite@CPTMS was refluxed with 3 in toluene for 40 h. Afterwards, the obtained bis(PYT)@boehmite was isolated via simple filtration, washed with DMSO and ethanol and, then, dried at 60 °C. Finally, 1 g of bis(PYT)@boehmite was dispersed in EtOH, and then samarium(III) acetate was added to the mixture and stirred for 24 h under reflux conditions. The obtained catalyst (Sm-bis(PYT)@boehmite) was filtered, washed, and dried at 60 °C.

General method for the synthesis of 5-substituted 1H-tetrazoles catalyzed by Sm-bis(PYT)@boehmite

[3 + 2] cycloaddition of sodium azide (NaN3) with organic nitrile compounds was selected for the synthesis of tetrazole derivatives in the presence of Sm-bis(PYT)@boehmite as a nanocatalyst. In this regard, 1.4 mmol of NaN3 and 1 mmol of nitrile were stirred in the presence of 50 mg of Sm-bis(PYT)@boehmite (1.93 mol% of samarium) in 2 mL of PEG-400 at 120 °C. At the end of the reaction, which was controlled using TLC, the reaction mixture was cooled down to room temperature. The reaction mixture was diluted by water and ethyl acetate and, then, Sm-bis(PYT)@boehmite nanocatalyst was separated by simple filtration. Afterwards, HCl (10 mL, 4 N) was added to the solution and, then, tetrazole products were extracted from ethyl acetate. The organic solvent was evaporated and dried using anhydrous sodium sulfate.

Spectral data

5-(2-Chlorophenyl)-1H-tetrazole 1H NMR (400 MHz, DMSO): δH = 7.81–7.78 (d, J = 12 Hz, 1H), 7.72–7.69 (d, J = 12 Hz, 1H), 7.64–7.60 (t, J = 8 Hz, 1H), 7.57–7.52 (t, J = 8 Hz, 1H) ppm (Fig. S1, Supplementary Material).

5-(4-Nitrophenyl)-1H-tetrazole 1H NMR (400 MHz, DMSO): δH = 16.55 (br, 1H), 8.46–8.43 (d, J = 12 Hz, 2H), 8.31–8.28 (d, J = 12 Hz, 2H) ppm (Fig. S2, Supplementary Material).

Conclusion

In Conclusion, natural and environmentally friendly boehmite nanoparticles were synthesized using sodium hydroxide and aluminum nitrate in an aqueous solution. The surface of boehmite nanoparticles was modified applying CPTMS. Afterwards, a new samarium complex of 1,3-bis(pyridin-3-ylmethyl)thiourea was stabilized on the modified boehmite nanoparticles (Sm-bis(PYT)@boehmite). Moreover, Sm-bis(PYT)@boehmite was used as an efficient, environmentally friendly, and reusable nanocatalyst in the homoselective synthesis of tetrazole derivatives from [2 + 3] cycloaddition reaction of NaN3 and nitriles in PEG-400 solvent. Significantly all tetrazoles were obtained with excellent yields and good TON and TOF values within short reaction times. This nanocatalyst was characterized using several techniques, such as FT-IR, SEM, EDS, WDX, TGA, XRD, DLS, ICP-MS, and BET. This nanocatalyst showed high activity, good selectivity, stability, and recyclability in the synthesis of tetrazole derivatives. Sm-bis(PYT)@boehmite was recovered and reused for six times without notable loss of catalytic efficiency. After being reused, the stability and heterogeneity of the catalyst were investigated using techniques such as FT-IR, XRD, SEM, EDS, WDX and ICP-MS analysis.