Kheireddine El-Boubbou*abc,
O. M. Lemined and
Daniel Jaqueb
aKing Saud bin Abdulaziz University for Health Sciences (KSAU-HS), King Abdullah International Medical Research Center (KAIMRC), King Abdulaziz Medical City, National Guard Health Affairs, Riyadh 11426, Saudi Arabia. E-mail: elboubboukh@ngha.med.sa
bNanomaterials for Bioimaging Group (nanoBIG), Facultad de Ciencias, Departamento de Física de Materiales, Universidad Autónoma de Madrid (UAM), Madrid 28049, Spain
cDepartment of Chemistry, College of Science, University of Bahrain, Sakhir 32038, Kingdom of Bahrain
dDepartment of Physics, College of Sciences, Imam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh 11623, Saudi Arabia
First published on 15th December 2022
The development of highly efficient, rapid, and recyclable nanocatalysts for effective elimination of toxic environmental contaminants remains a high priority in various industrial applications. Herein, we report the preparation of hybrid mesoporous gold–iron oxide nanoparticles (Au–IO NPs) via the nanocasting “inverse hard-templated replication” approach. Dispersed Au NPs were anchored on amine-functionalized iron oxide incorporated APMS (IO@APMS-amine), followed by etching of the silica template to afford hybrid mesoporous Au–IO NPs. The obtained nanoconstructs were fully characterized using electron microscopy, N2 physisorption, and various spectroscopic techniques. Owing to their magnetic properties, high surface areas, large pore volumes, and mesoporous nature (SBET = 124 m2 g−1, Vpore = 0.33 cm3 g−1, and dpore = 4.5 nm), the resulting Au–IO mesostructures were employed for catalytic reduction of nitroarenes (i.e. nitrophenol and nitroaniline), two of the most common toxic organic pollutants. It was found that these Au–IO NPs act as highly efficient nanocatalysts showing exceptional stabilities (>3 months), enhanced catalytic efficiencies in very short times (∼100% conversions within only 25–60 s), and excellent recyclabilities (up to 8 cycles). The kinetic pseudo-first-order apparent reaction rate constants (kapp) were calculated to be equal to 8.8 × 10−3 and 23.5 × 10−3 s−1 for 2-nitrophenol and 2-nitroaniline reduction, respectively. To our knowledge, this is considered one of the best and fastest Au-based nanocatalysts reported for the catalytic reduction of nitroarenes, promoted mainly by the synergistic cooperation of their high surface area, large pore volume, mesoporous nature, and enhanced Au-NP dispersions. The unique mesoporous hybrid Au–IO nanoconstructs synthesized here make them novel, stable, and approachable nanocatalyst platform for various catalytic industrial processes.
Several methods have been utilized to synthesize porous noble metal–iron oxide hybrid materials, including chemical reduction of metal ion salts in a specific polymeric matrix, porous support matrix, self-assembly, surfactant-assisted processes, and dispersion or blending of metal NPs with iron oxide.30,31 Nonetheless, the fabrication of well-dispersed and stable Au-NPs with controllable sizes and shapes on porous metal matrices having high surface area and large pore volumes is quite challenging. While the mesopores dictate the size and morphology of the loaded Au-NPs as they are formed in the confined matrix, the high surface area and large pore volume aid in the formation of homogeneously dispersed noble metal NPs on the surface of the support. Consequently, urgent development of robust and reliable procedures to synthesize efficient oxide-supported Au-based nanocatalysts are constantly needed, particularly for chemical and environmental industries worldwide. Herein, we demonstrate, for the first time, the fabrication of hybrid mesoporous Au–IO nanostructures, anchored with highly dispersed Au-NPs (sizes ∼ 15 nm) based on our previously reported “inverse hard-templated replication” of an amino-functionalized IO-incorporated acid-prepared mesoporous silica (IO@APMS-amine) support, which acts as metal stabilizing and size-controlling agent. When tested for the catalytic reduction of nitro aromatic compounds (i.e. nitrophenol and nitroaniline), the as-prepared mesoporous Au–IO NPs were found to be highly efficient with very fast conversion rates, enhanced catalytic efficiencies, and excellent stabilities and recyclabilities. To the best of our knowledge, the mesoporous Au-based IO-supported nanocatalyst developed here is among the most efficient and fastest catalysts reported for the catalytic reduction of nitroarenes. Thanks to the synergistic cooperation of their high surface area, large pore volume, mesoporous nature, and dispersed Au-NPs anchored on the nanocasted support.
Fig. 1 Schematic illustration for the preparation of hybrid mesoporous gold iron oxide nanoparticles (Au–IO NPs) via nano-casting reverse methodology using APMS as the silica template. |
The prepared mesoporous Au–IO NPs were fully characterized by electron microscopy, dynamic light scattering (DLS), Fourier transform infrared (FTIR), X-ray diffraction (XRD), and vibrating-sample magnetometer (VSM). First, the structure, morphology, and composition of mesoporous IO–Au NPs were examined by transmission electron microscopy (TEM) and scanning electron microscopy energy dispersive X-ray (SEM-EDX). Fig. 2a–c and S1† depict TEM images of ∼1.5 μm IO@APMS clearly showing the formation of ultrasmall sized iron oxide nanoclusters (∼1–5 nm) within and surrounding the APMS silica template (Fig. 2a). Once Au NPs were incorporated, a regular arrangement of additional well-dispersed black colored spherical dots (∼15–20 nm) are evident along the surface of the APMS particles (Fig. 2b). No independent agglomeration of the Au NPs was observed along the TEM grid, revealing that the amine moieties controllably anchored the formed Au NPs. Finally, spheroid-like Au–IO nanoparticulate mesostructures are remarkably evident, with mesochannels clearly seen, proving the successful replication of the original mesoporous template. The obtained structures possess mesopores of nanoclustered IO (gray) and the heavier Au (black) NPs depicted in yellow circles (Fig. 2c). SEM-EDX analyses were also conducted to determine the elemental composition of the different prepared samples (Fig. 2d–f). SEM images clearly show the IO- and Au-dotted architectures of both IO@APMS and Au–IO@APMS, along with the formation of smaller-sized mesoporous Au–IO NPs, in agreement with TEM. EDX performed on selected SEM areas confirmed the presence of Fe, O, and Si elements in IO@APMS; Fe, O, Si, and Au in Au–IO@APMS; and only Fe, O, and Au in the final replicated mesoporous Au–IO material. Elemental analysis revealed that the mesoporous support can be incorporated with as high as 14.6 wt% of Au (Table S1†), which is much higher than the loading achieved in previous mesoporous IO-based and nonporous commercial IO supports.27
Next, the porous nature of the samples was examined by nitrogen (N2) adsorption/desorption isotherms (Fig. 3). The isotherms clearly demonstrated type IV curves, consistent with reported porosities for similar materials, where the Brunauer–Emmett–Teller (BET) surface area (SBET), Barrett–Joyner–Halenda (BJH) pore volume (Vpore), and BJH average pore size diameter (dpore) were calculated from the obtained desorption graphs. As expected, iron impregnation indicated reduction in the specific surface area and total pore volume of the original APMS-amine (SBET = 534 m2 g−1 and Vpore = 0.96 cm3 g−1) affording IO@APMS with SBET = 265 m2 g−1, Vpore = 0.73 cm3 g−1, and dpore = 9.5 nm (Fig. 3a). Once incorporated with Au NPs, SBET was found to increase to 360 m2 g−1 and Vpore = 1.0 cm3 g−1 for IO–Au@APMS. This result could be mainly due to the formation of Au NPs on the outer surface via the amine moieties coating the APMS, indications of both IO and Au NPs incorporating and coating the surfaces of the silica walls, respectively. Importantly, following the loading of Au NPs, the mesoporous structure of the sample is still well-retained with dpore = 9.0 nm. Notably, N2 isotherms showed that the final etched mesoporous Au–IO NP sample exhibit relatively high SBET = 124 m2 g−1, Vpore = 0.33 cm3 g−1, and major average dpore = 4.5 nm (Fig. 3b). It is worth pinpointing that this procedure was repeated more than once using different APMS samples with various surface areas and pore volumes obtaining mesoporous Au–IO NPs with SBET of 95–125 m2 g−1 and Vpore of 0.15–0.35 cm3 g−1. All the N2 physisorption data are summarized in Table 1.
Sample | SBET (m2 g−1) | Vpore (cm3 g−1) | dpore (nm) |
---|---|---|---|
APMS-amine | 534 | 0.99 | 5.9 |
IO@APMS | 265 | 0.73 | 9.5 |
Au–IO@APMS | 360 | 1.00 | 9.0 |
Mesoporous Au–IO NPs | 124 | 0.15 | 4.5 |
FTIR analysis was then extended to further corroborate the structures of IO@APMS-amine, Au–IO@APMS, and mesoporous Au–IO NP samples (Fig. 4a). In all the samples, the presence of the Fe–O stretching bands at ∼456 to 585 cm−1 confirm the presence of iron oxide. The broad peaks observed at ∼3400 cm−1 depict the O–H and N–H stretching vibration of hydroxyl and amine groups, while the distinctive peaks at 2850 and 2920 cm−1 show the stretching modes of C–H and –CH2– of methyl and methylene groups. In the case of mesoporous Au–IO material, a weaker peak intensity in the 3400 cm−1 region was observed, signifying the interaction of Au NPs with the amine surface moieties. Moreover, the slight decrease of peak at 733 cm−1 ascribed to the Si–O–Si symmetric stretching and Si–O–Fe moiety with Au loading suggests a strong interaction of AuNPs with silica. Importantly, the IR spectrum showed the disappearance of the bands at 1080 cm−1 (corresponding to Si–O–Si) and 960 cm−1 (Si–OH and Si–Oδ−) stretching vibrations and at ∼3080 cm−1 (–NH2 stretching vibrations of the primary amine), indicating the successful incorporation of Au NPs into IO@APMS via the involvement of –NH2 groups. Additionally, the UV-vis absorption spectra of three different samples were also recorded (Fig. 4b). The results confirm the featureless absorption of IO@APMS, with the successful formation of Au NPs displaying a characteristic absorption band centered ∼550 nm due to the Au surface plasmon band (SPB), indicating the presence of well-dispersed nano-sized Au particles. It is well known that with the increase of Au NP sizes and particle agglomeration, the SPB peak shifts to longer wavelength with higher absorbance intensity (∼700 nm). The peak becomes broader and broader due to the coupling of the individual Au NPs when aggregated.42 Thus, the sharp SPB peaks of Au-incorporated samples obtained here indicates a narrow size distribution of well-dispersed Au NPs, corroborating the obtained electron microscopy results. Perimeter sites.
We then assessed the average hydrodynamic sizes (DH) and zeta potentials (ζ) of the different mesoporous constructs in their aqueous dispersions (Fig. 5 and S2†). While the DH sizes for IO@APMS and Au–IO@APMS were found to be similar measuring ∼1500 nm, Au–IO NPs possessed much smaller sizes with average DH ∼ 750 nm (Fig. 5a). This is expected and agrees well with our previous results, mainly due to effective etching of the original mesoporous template. The relatively sharp peaks obtained pinpoint the uniformity and dispersity of the different as-synthesized samples (PDI ∼ 0.5). The surface charges of the different samples were then computed using zeta potentials (ζ) analysis. The results showed average zeta potentials ξ = +9.40 ± 0.56 mV, −16.8 ± 0.50 mV, and −26.4 ± 0.98 mV for IO@APMS-amine, Au–IO@APMS, and mesoporous Au–IO samples, respectively (Fig. 5b). The results indicate a strong interaction of Au NPs with the support enabled by the amine surface modification where positively charged amine-functionalized IO@APMS-amines (ξ-potential = + 9.40 mV) were able to electrostatically attract negatively charged Au NPs (ξ-potential of only Au-NPs = −29.2 mV). This clearly confirm the appropriate anchoring of AuNPs on the aminated surface of IO@APMS as well as their high and good dispersion stability due to electrostatic repulsive forces. It is anticipated that the surface of the IO@APMS microspheres consists mainly of –NH3+ and unmodified Si–OH/Fe–OH groups, where the negatively charged tetrachloroaurate (AuCl4−) ions interact with the ammonium groups through electrostatic interactions resulting in Au dispersed deposition over the whole APMS surface. Both the presence of a mesoporous structure in the support along with a mixed surface phase of Si and Fe oxides, seem to be key parameters, providing improved Au dispersion and higher contact surface area between the confined AuNPs within and/or at the exterior surface of the porous nanostructured support, in accordance with earlier observation.29
To determine the magnetic characteristics of the Au–IO samples, field-dependent magnetization (M–H)20–22 of the as-prepared mesostructures at 298 K were recorded (Fig. S3†). The saturation magnetization (Ms) obtained for IO@APMS-amine, Au–IO@APMS, and mesoporous Au–IO NP samples were found to be equal to 1.85 emu g−1, 1.09 emu g−1 and 1.40 emu g−1, respectively (these values are expected to be higher when calculated per gram IO). It is evident that the all-magnetization curves obtained have almost negligible retentivity and coercivity, signifying the superparamagnetic nature of the mesoconstructs. This is advantageous for catalytic reactions, as it provides ease of separation from the reaction medium, enabling regeneration of the catalyst. To further elucidate which phase of magnetic iron oxide we have, XRD of mesoporous Au–IO NPs was conducted. XRD data revealed that the produced iron oxide phase is γ-Fe2O3 (maghemite), with observed diffraction patterns in excellent agreement with previously reported data for maghemite (JCPDS #039-1346).32 Noticeably, the diffraction peaks at 2θ = 38°, 44°, 64°, and 77° indexed to (111), (200), (220), and (311) planes of crystalline Au phases (JCPDS #96-901-3039) are well observed (Fig. S4†). The weight percentage ratio of maghemite to Au phase was found to be 80% γ-Fe2O3: 20% Au. This is in accordance with the EDX results and affirms the very high loading amounts of Au NPs onto the support mainly due to the remarkably high surface area (SBET = 260–360 m2 g−1), large pore volume (Vpore = 0.70–1 cm3 g−1), and appropriate pore diameter (dpore = 9 nm) of the mesostructured support. Moreover, the greater number of surface defects (i.e. steps, edges and kinks) present on mesoporous IO@APMS might be another important factor for the high Au dispersion (Au NPs cannot be easily adsorbed on a flat metal oxide surface nor on nonporous silica-type materials).27
It is well established that at the nano-regime, Au and Au-hybrid NPs are particularly useful for the as-mentioned catalytic conversions. However, bare AuNPs ae usually not that stable, can easily agglomerate through catalytic processes, and are difficult to separate from reaction medium, resulting in clear reduction in catalytic activities. In contrast, the large surface area and pore volume exhibited by the as-prepared Au–IO mesostructures are expected to be beneficial for accommodating substrates, while also providing enhanced diffusion rates of reactant molecules during the catalytic reaction, significantly improving the overall catalytic performance. Furthermore, their intrinsic magnetism aids in their fast separation from the reaction medium by applying external magnet. First, the catalytic performances of the as-synthesized hybrid mesoporous nanocatalysts IO@APMS, Au–IO@APMS, and Au–IO NPs towards 2-nitrophenol reduction to their non-toxic counterparts in aqueous media is shown in Fig. 7. Time-dependent UV-vis absorption was monitored throughout the reductive catalytic activity of 2-nitrophenol to 2-aminophenol in the presence of NaBH4. As shown in Fig. 7a, the absorption peak at 350 nm corresponds to pure 2-nitrophenol aqueous solution. Upon the addition of freshly prepared NaBH4, the absorption peak shifts to 417 nm, with a color change from pale to dark yellow, ensuring the formation of nitrophenolate solution. Without the addition of the Au catalyst, the reduction reaction cannot proceed even in the presence of NaBH4. Remarkably, when only 1.5 mg of the mesoporous Au–IO NP catalyst was added into the reaction mixture, the reaction rapidly progressed to completion in a very short reaction time of only 60 seconds (s) (Fig. 7b). As can be seen, the absorption peak of 2-nitrophenolate anion dramatically decreased, with a simultaneous new and weak band appearing at 297 nm, confirming the formation of a transparent and colorless non-toxic 2-aminophenol solution. When same amount of Au–IO@APMS was used, the reaction rate was found to be slightly slower with completion of the reaction within 90 s (Fig. 7c). Importantly, same concentration of IO@APMS sample with no incorporated Au exhibited no reduction of 2-nitrophenol even after 24 h of reaction time. When compared to only Au-NPs produced in aqueous basic NaBH4 solution, without the incorporation of IO@APMS support, the catalytic efficiencies were much less taking ∼4 minutes to complete the conversions (Fig. S5†). It is to be noted that when using higher concentrations of mesoporous Au–IO NP or Au–IO@APMS (2 mg mL−1), the reduction reaction was completed extremely fast within only few seconds. All these results strongly confirm that the presence of dispersed Au, but not IO, in addition to its confinement onto and within the mesoporous nanocatalyst support (large surface area and pore volume) are directly responsible for the fast catalytic reduction. Mechanistically, it is believed that upon the addition of Au NPs, the electron donor (BH4−) and electron acceptor (4-nitrophenolate) are both adsorbed on the NP surface where the catalytic reduction of nitrophenolate by hydride ions (H−) inducts via Au–hydride complex.45 Au–H and/or Au–BH4 bond formation with the AuNPs has been suggested.38 Then, interfacial electron transfer occurs from hydrides to nitrobenzene where reduction to corresponding aminobenzene involves two fast intermediate steps via nitrosobenzene and phenylhydroxylamine. Finally, desorption of the product takes place from the nanocatalyst surface to make it free for another cycle. It is to be noted that in the absence of the Au catalyst and the presence of the reducing agent NaBH4, the formation of the 2-nitrophenolate ion is observed, but no further conversion to 2-aminophenol occurs with 417 nm peak remaining unaltered for a long period of time. Likewise, in the absence of NaBH4 and the presence of the Au catalyst, the Au-incorporated IO mesostructures showed no catalytic activities. Thus, both the Au and the hydride species (Au–hydride complex) are crucial to promote the reduction of 2-nitrophenol.46 A plausible mechanism for the reduction of nitroarenes using the Au–IO nanocatalyst is depicted in Fig. 8.
Fig. 8 A plausible mechanism for catalytic reduction of nitroarenes using mesoporous Au–IO nanocatalysts. |
The catalytic conversion kinetic rates for reduction of 2-nitrophenol detected using UV-vis were also investigated (Fig. 9a). Since the reducing agent NaBH4 is in high excess relative to 2-nitrophenol, the reaction is expected to be independent of NaBH4 concentration and follow pseudo-first-order kinetics,47 leading to the determination of the rate constant kapp according to the following equation:
ln(Ct/C0) = −kappt |
Next, we also studied the use of the hybrid mesoporous Au–IO nanocatalysts for catalytic reduction of 2-nitroaniline monitored by UV-vis spectroscopy (Fig. 10). Although the chemical reduction of nitroaniline,43 a highly toxic pollutant, into the less harmful counterparts is highly needed, its conversion and removal using noble metal-based nanocatalytic systems is studied to much less extent. The 2-nitroaniline orange colored aqueous solution presented two distinguished absorption peaks at 290 and 415 nm (Fig. 10a). In the presence of NaBH4 solution, the intensity of the aforementioned peaks gradually decreased upon addition of mesoporous Au nanocatalysts to 2-nitroaniline/NaBH4 aqueous solution. Similar to 2-nitrophenol reduction reaction, upon adding only NaBH4, no catalytic reduction of 2-nitroaniline to 2-aminoaniline is observed. However, addition of 1.5 mg of mesoporous Au–IO NPs or Au–IO@APMS afforded the non-toxic 2-aminoaniline very fast in 25 and 40 s, respectively (Fig. S6†). The catalytic conversion rates were found to be equal to with kapp = 15.9 × 10−3 s−1 and 23.5 × 10−3 s−1 for the mesoporous Au–IO nanocomposites (Fig. 10b and c). Table 2 shows the computed kinetic catalytic reaction rates of the different mesoporous Au–IO nanocatalysts for reduction of 2-nitrophenol and 2-nitroaniline. Those Au-incorporated mesostructured nanocatalysts were found to have very fast and efficient reduction performance compared to other nanocatalysts reported in the literature, where the hybrid mesoporous Au–IO NPs retaining the fastest catalytic activities. However, extra care should be taken when comparing the efficiencies of different catalysts, as many factors such as nanocatalyst concentrations, amount of the substrate, and overall reaction conditions might be variables. For better quantitative comparison with other systems, the catalytic activity parameter kapp/M (ratio of kapp to the total mass of the catalyst) is typically presented. For instance, the reaction rate constant per unit mass for Au–IO NP towards 2-nitrophenol reduction is calculated to be equal to 17.6 × 10−3 s−1. Table 3 summarizes the catalytic activities of our mesoporous Au–IO NPs compared to other transition metal-based nanoparticulate catalytic platforms used for the reduction of nitroarenes. As depicted, the catalytic activities of the mesoporous Au–IO nanocatalyst developed in this work is considerably greater than those reported in previous works. This affirms the excellent catalytic behavior of our Au-based mesoconstructs due to synergistic factors mainly the dispersed small-sized Au NPs, mesostructures, high surface area, and large pore volumes, facilitating optimum access to Au perimeter active sites, as well as faster transport of the substrates towards the reduction. These interesting results clearly demonstrate that very low concentrations of Au–IO nanocatalysts (only 0.5–1.5 mg) are needed to convert toxic nitroarene solutions to non-toxic counterparts in very short times (only few seconds). This strongly suggests the potential use of the as-prepared hybrid mesoporous Au–IO based nanocatalysts for industrial applications and other catalytic conversion systems.
Catalyst | Substrate | 100% conv. time (sec) | kapp (s−1) | kapp/M (s−1 g−1) |
---|---|---|---|---|
a Reaction conditions: 2.5 mM of nitroarene solution, 25 mM of freshly prepared NaBH4 (excess), 0.5 mg of catalyst, total volume = 1 mL. | ||||
IO@APMS | 2-Nitrophenol | No reduction | N/A | N/A |
Au–IO@APMS | 2-Nitrophenol | 60 | 5.2 × 10−3 | 10.4 |
Mesoporous Au–IO | 2-Nitrophenol | 45 | 8.8 × 10−3 | 17.6 |
IO@APMS | 2-Nitroaniline | No reduction | N/A | N/A |
Au–IO@APMS | 2-Nitroaniline | 40 | 15.9 × 10−3 | 31.8 |
Mesoporous Au–IO | 2-Nitroaniline | 25 | 23.5 × 10−3 | 47 |
Catalyst | kapp (s−1) x 10−3 | kapp/M (s−1 g−1) | Ref. |
---|---|---|---|
a Reaction conditions: comparison of catalytic activities between mesoporous Au–IO nanocatalyst and other reported Au nanocatalysts for reduction of 4-nitrophenol.b Also compare with other state-of-the-art transition metal-based nanocatalysts reported in Table 3.17,49 GO/Fe3O4@PDA/Au stands for Au-NP incorporated polydopamine (PDA)-shelled graphene oxide (GO)/Fe3O4.c It is to be noted that reduction of 4-nitrophenol is ∼ 3–4 times faster than that of 2-nitrophenol (i.e. 16 min vs. 60 min for Fe3O4@SiO2–NH2–Au catalyst).49 | |||
GO/Fe3O4@PDA/Au | 14.4 | 14.4b | 49 |
Fe3O4@SiO2–NH2–Au | 7.8 | 2.6b | 17 |
Fe3O4@SiO2@Aushell | 6.6 | 1.76 | 48 |
Au@MSNscore–shell | 3 | N/A | 50 |
Au@SiO2yolk–shell | 3.9 | 12 | 51 |
Au@Fe3O4dumbbell | 10.5 | 5.25 | 52 |
Mesoporous Au–IO | 8.8 | 17.6 | This workc |
Finally, the stability and recyclability of the mesoporous Au–IO nanocatalyst were also investigated (Fig. 11). It is extremely significant for industrial applications that the utilized nanocatalysts are stable, ready to be used in harsh conditions, and can be recycled many times without the loss in activity. Different strategies have been reported to achieve these goals varying from core-satellite, yolk–shell nanorattle type structures, to magnetic core–gold shells and Au-NPs embedded in core–shell (magnetite-mesoporous silica) microspheres.48 It is therefore imperative to keep enhancing on catalyst stabilities and recyclabilities. Herein, the nanocatalysts were magnetically recovered, washed with water, and then used for the next reduction cycles. As evident, the catalytic conversions didn't stagnate even after the use of the nanocatalyst for 8 consecutive runs (Fig. 11a). The nanocatalyst did not undergo any substantial change in its activity, with only very slight decrease in reaction times noticed during the 8 cycles. This is probably due to adsorption of product (2-aminophenol) over the catalyst surface. To investigate the structural stability of the reused nanocatalyst after the 8th run, TEM and FTIR analysis were conducted. As depicted, the morphology and structural signals were found to be the same as the initial Au–IO NP sample without changes in functionality (Fig. 11b and c). Remarkably, even after 3 months of storage in the fridge at 4 °C, the catalyst efficiently carried out 100% conversion of 2-nitrophenol and 2-nitroaniline within only few seconds, exactly similar to the freshly prepared nanocatalyst samples. This demonstrates the exceptional stability of the mesoporous Au–IO nanocatalyst and that the dispersed AuNPs are still active, validating its outstanding reusability.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2ra05990h |
This journal is © The Royal Society of Chemistry 2022 |