Xiao Yua,
Atika Muhammada,
Boya Qiua,
Aristarchos Mavridisa,
Min Hu
*a and
Carmine D'Agostino
*ab
aDepartment of Chemical Engineering, The University of Manchester, Oxford Road, Manchester, M13 9PL, UK. E-mail: min.hu@manchester.ac.uk; carmine.dagostino@manchester.ac.uk
bDipartimento di Ingegneria Civile, Chimica, Ambientale e dei Materiali (DICAM), Alma Mater Studiorum-Università di Bologna, Via Terracini, 28, Bologna, 40131, Italy
First published on 22nd April 2025
The heterogenization of aluminium isopropoxide [Al(OiPr)3] on modified magnesium oxide (MgO) supports was investigated to develop efficient catalysts for hydrogen-transfer (H-transfer) reduction reactions. MgO surfaces were functionalized with octadecyltrichlorosilane (OTES) or dibromobutane (DBB) to optimize the surface chemistry of commercial MgO. The OTES-modified MgO exhibited a porous “nest-like” structure with a markedly increased surface area of 84.3 m2 g−1, compared to 2.9 m2 g−1 for the unmodified MgO. In contrast, DBB-modified MgO displayed a “brush-like” morphology attributed to the flexibility of the immobilized carbon chains. The modified heterogenized catalysts demonstrated substantial improvements in activity compared to the unmodified MgO-based systems. Among the heterogenized catalysts, Al–DBB–MgO achieved the highest turnover frequency (TOF), which is attributed to enhanced substrate adsorption and reduced steric hindrance, facilitating efficient reactant access to active sites. The activity of the two modified catalysts after 5 reduction cycles shows no obvious change in terms of the TOF of Al–DBB–MgO, while the TOF of Al–OTES–MgO dropped by around 14%. These findings highlight the critical role of MgO surface modification in enabling effective Al(OiPr)3 immobilization and enhancing catalytic performance for H-transfer reactions, offering a promising strategy for designing advanced heterogenized catalysts for reduction of carbonyl compounds.
Initial studies on catalyst design for H-transfer reductions predominantly explored metal oxides and zeolites, capitalizing on their intrinsic Lewis acidic properties as an extension of homogeneous catalytic systems. For example, Lavalley et al.9 systemically investigated heterogeneous acetone reduction over various metal oxide catalysts, demonstrating that the catalytic activity is strongly correlated with surface acid–base properties. Lewis acidic catalysts, such as chloride-modified Al2O3 (Al2O3–Cl), and basic catalysts, including MgO and ZrO2, exhibited comparable activity. Similarly, H van Bekkumn's group10 reported the reduction of substituted cyclohexanones over zeolite BEA, achieving superior seteroselectivity towards cis-4-tert-butylcyclohexanol. This performance was attributed to Lewis acidic aluminium sites located within the zeolite's micropores. Efforts to enhance zeolite catalytic performance have included incorporating elements like Sn and Zr into the framework to increase Lewis acidity.11,12
Heterogenized homogeneous catalysts represent another significant advancement, achieved by covalently grafting metal alkoxide catalysts onto solid supports. Porous silicas are among the most used supports due to their diverse porous structures.13,14 Such heterogenized catalysts can often retain the high activity and selectivity of their homogeneous counterparts, particularly in the synthesis of α,β-unsaturated alcohols. For example, Uysal and Oksal,15 developed a boron-based heterogeneous catalyst by grafting the boron tri-ethoxide onto mesoporous MCM-41 [B(OEt)3-MCM-41], which exhibited comparable catalytic activity to homogeneous B(OiPr)3 and B(OEt)3 catalysts. This system demonstrated excellent stability and recyclability, maintaining catalytic activity over six cycles without significant performance loss. Similarly, in our previous work,16 we immobilised aluminium isopropoxide [Al(OiPr)3] onto various mesoporous supports, including SiO2, TiO2 and γ-Al2O3, and evaluated their performance in H-transfer reductions. The heterogenized catalysts exhibited excellent activity and high selectivity, closely mirroring the performance of homogeneous Al(OiPr)3. Among these, Al(OiPr)3 grafted onto SiO2 outperformed those on γ-Al2O3 and TiO2, owning to its higher surface area and large pore structure, which facilitated improved Al(OiPr)3 dispersion and enhanced accessibility of the carbonyl substrates to active sites. These findings underscore the importance of carefully selecting and engineering support materials to achieve optimal catalytic performance.
In this study, we explore the impact of MgO surface modification using different functional groups, octyltriethoxysilane, OTES and 1,4-dibromobutane, DBB, to tailor MgO surface chemistry for improved catalytic performance. A comprehensive set of characterization techniques, including FTIR, XRD, SEM, and nitrogen adsorption, were employed to probe the structural and surface chemistry changes induced by the modifications. The catalyst activity of the unmodified and modified supports was evaluated in H-transfer reduction reactions with various substrates including aldehydes and ketones. This work aims to provide insights into the relationship between surface modification and catalytic efficiency, paving the way for the development of advanced heterogenized catalysts with improved performances through relatively simple support modifications.
![]() | (1) |
![]() | (2) |
![]() | (3) |
![]() | ||
Fig. 1 Processes of (a) MgO support modification followed by Al(OiPr)3 immobilization (a); and (b) Al(OiPr)3 immobilization onto MgO directly (b). |
Al (wt%) | SBET (m2 g−1) | |
---|---|---|
MgO | — | 2.9 |
OTES–MgO | — | 84.3 |
DBB–MgO | — | 6.8 |
Al–MgO | 0.25 | 5.1 |
Al–MgO–OTES | 3.72 | 57.4 |
Al–DBB–MgO | 11.42 | 29.5 |
Similarly, the addition of DBB to the MgO surface also increased the surface area, resulting a material with a surface area of 6.8 m2 g−1. The N2 adsorption–desorption isotherm of DBB–MgO (Fig. 2c and Table 1) exhibited a profile similar to that of bare MgO, with the steepest slope at higher relative pressures due to the high difference in volumes of gas adsorbed at the high-pressure points, indicating its microporous nature induced by the presence of the DBB carbon chain layer.19,20 This DBB layer acts as a “brush” on the MgO surface, consisting of extended hydrocarbon chains terminated with hydroxyl groups (–OH). The terminal bromine (Br) of DBB likely undergoes nucleophilic substitution with water (solvent) or surface hydroxyl groups on MgO, forming C–O bond.21 The FTIR spectrum of DBB–MgO (Fig. 2a) showed a characteristic C–O stretching peak at around 1050 cm−1.
Notably, the addition of OTES and DBB did not significantly alter the crystal structure of MgO, as indicated by the XRD patterns of both supports and catalysts (Fig. 2b). The diffraction peaks at 2θ values of 37.2°, 43.2°, 62.5°, 74.9° and 78.8° correspond to the (111), (200), (220), (311), and (222) planes of cubic polycrystalline MgO, consistent with the JCPDS No. 01-087-0652 standard.22 However, a minor Mg(OH)2 phase was detected in the modified MgO supports, as evidenced by additional diffraction peaks (marked with black circles) corresponding to the (101), (102), and (110) planes at 2θ = 38.2°, 51.0°, and 58.9° (JCPDS No. 00-007-0239), respectively.23,24 This phase likely resulted from surface hydrolysis induced by water.25 After immobilization, significant differences in crystal structure were not observed between the supports and the catalysts.
The immobilization of Al(OiPr)3 was achieved through exchange reactions between the surface/terminated hydroxyl groups and the isopropyl linkers of Al(OiPr)3.3,26 The amount of immobilized Al(OiPr)3 was determined by measuring the Al loading via ICP-OES, with results summarized in Table 1. Among the samples, Al–MgO exhibited the lowest Al loading, which can be attributed to its minimal surface area. Furthermore, due to the relatively large size of the Al(OiPr)3 molecule, its grafting onto the MgO surface likely hindered access to adjacent hydroxyl groups, thereby limiting further immobilization.
In contrast, the Al loading increased significantly after the MgO surface was modified with OTES or DBB, with Al–DBB–MgO showing the highest Al loading. Both OTES and DBB contain terminal hydroxyl groups that can facilitate the immobilization of Al(OiPr)3. For Al–OTES–MgO, the Al(OiPr)3 was likely grafted primarily onto the outer layer of OTES, as the densely packed OTES near the MgO surface hindered access to inner hydroxyl groups. Additionally, the lower surface area of MgO limited the amount of OTES grafting, reducing the number of hydroxyl groups available for further Al(OiPr)3 immobilization. The grafting of Al(OiPr)3 may have also occupied some of the cavity space within the OTES framework, contributing to a decrease in surface area.
Conversely, the DBB modification significantly increased the number of the terminal hydroxyl groups. The flexibility of carbon chains of DBB also minimized steric hindrance during Al(OiPr)3 grafting, resulting in the highest Al loading among the samples. The densely grafted Al(OiPr)3 molecules formed a onto DBB–MgO might pack a also contributed to the overall surface area, explaining the increased surface area observed for Al–DBB–MgO.
The surface chemical states of the fresh catalysts were further examined using X-ray photoelectron spectroscopy (XPS), with the Al 2p and O 1s spectra presented in Fig. 3. The O 1s peak in all samples appeared as a single peak at approximately 531.7 eV, corresponding to hydroxyl groups on the MgO surface or terminated hydroxyl groups introduced by the OTES and DBB functional groups. Due to the similar binding energies of these two types of hydroxyl groups, distinguishing them from one another is challenging. The Al 2p peak confirmed the successful immobilization of Al(OiPr)3 molecules onto the support surfacers. The Al–MgO exhibited a single Al 2p doublet peak, indicative of Al–O bond formation between Al(OiPr)3 and the surface hydroxyl groups. In contrast, the modified catalyst, Al–OTES–MgO and Al–DBB–MgO, displayed two doublet Al 2p peaks, suggesting that Al(OiPr)3 molecules were anchored both on the MgO surface (at 74.2–74.6 eV) and on the terminated hydroxyls groups from the OTES or DBB modifiers (at 73.7–74.0 eV), with the area ratio of around 0.25 (80% terminated –OH to 20% surface –OH). These findings support our hypothesis that hydroxyl groups introduced by the modification process serve as active sites for catalyst grafting and play a crucial role in facilitating the subsequent hydrogen transfer reactions.27,28
![]() | ||
Fig. 3 XPS spectra of O 1s and Al 2p in the fresh catalyst (a) Al–MgO; (b) Al–OTES–MgO and (c) Al–DBB–MgO. |
The morphology and distribution of relevant elements were further analysed using SEM and EDX mapping, as shown in Fig. 4. The SEM images revealed an intensive and uniform distribution of carbon on DBB–MgO, whereas only a small amount of silicon was detected on OTES–MgO. This supports the conclusion that limited OTES was grafted onto the MgO surface. As SEM primarily examines surface features, the detected silicon distribution likely originated from the outer layer of the OTES coating. All catalyst samples showed the presence of aluminum, confirming the successful immobilization of Al(OiPr)3.
![]() | ||
Fig. 4 SEM images and EDX elemental maps of catalysts: (a) Al–MgO; (b) Al–OTES–MgO; (c) Al–DBB–MgO. Elemental maps of Al (red), Mg (blue), O (green) and Si or C (turquoise blue) are shown. |
Substrate | Structure | Catalyst | Conversion (%) | Selectivity (%) | TOFa (h−1) |
---|---|---|---|---|---|
a TOF was calculated as the average value (over 3 h). | |||||
Hexanal | ![]() |
Al–DBB–MgO | 46.6 | 92.7 | 0.21 |
Al–OTES–MgO | 22.9 | 87.2 | 0.14 | ||
Al–MgO | 12.8 | >99 | 0.06 | ||
Cyclohexanone | ![]() |
Al–DBB–MgO | 28.1 | >99 | 0.13 |
Al–OTES–MgO | 22.3 | 92.1 | 0.10 | ||
Al–MgO | 5.6 | >99 | 0.03 | ||
Benzaldehyde | ![]() |
Al–DBB–MgO | 23.9 | 84.4 | 0.23 |
Al–OTES–MgO | 13.0 | 87.5 | 0.12 | ||
Al–MgO | 7.8 | 94.6 | 0.07 | ||
Cinnamaldehyde | ![]() |
Al–DBB–MgO | 12.6 | 96.5 | 0.12 |
Al–OTES–MgO | 6.7 | >99 | 0.05 | ||
Al–MgO | 4.6 | >99 | 0.02 |
As previously reported,29 the reaction proceeds via a cyclic six-membered transition state involving direct hydrogen transfer from the H-donor (alcohol) to the H-acceptor (aldehyde or ketone). Among the substrates tested, saturated aldehydes were generally easier to reduce than ketones, particularly liner chain aldehydes. For example, all heterogenized catalysts demonstrated the highest activity in reducing hexanal compared to cyclohexanone. The reduced activity observed for cyclohexanone is likely due to the steric hindrance, which restricts its coordination of the ketones to the Al centre.30 In the case of unsaturated aldehydes, such as cinnamaldehyde, a typical α,β-unsaturated aldehyde with a conjugated aromatic ring and CC bond, exhibited the poorest conversion. This reduced activity can be attributed to the combined influence of steric and electronic effects, resulting in the lowest observed activity.
Catalyst based on the unmodified MgO demonstrated the lowest activity among various substrates, owning to the lowest Al(OiPr)3 grafting. In contrast, the modified catalysts demonstrated significantly improved activity across all the substrates, achieving approximately 2- to 4-fold increase compared to unmodified catalyst. Moreover, the heterogeneous catalysis followed the same substrate-specific trends as the homogeneous process, suggesting a similar reaction mechanism for the classical H-transfer reduction.
When normalised by turnover frequency (TOF, last entry in Table 2), Al–DBB–MgO outperformed the other solid catalysts. This superior performance can be attributed to the flexibility of the DBB layer grafted onto MgO, where rotation of the carbon chains reduces steric limitations, facilitating coordination between the Al canter and the substrates. Additionally, the structural similarity between DBB and certain substrates, such as hexanal, enhances reactant adsorption through stronger interaction.
For Al–OTES–MgO, an activity increase was also observed compared to the unmodified catalyst. The enhanced surface area of Al–OTES–MgO is thought to improve contact between the reactants and active sites. However, the densely packed OTES layer induces significant steric hindrance, which was particularly detrimental for substrates with ring structures. This may explain the comparatively lower activity of catalyst Al–OTES–MgO for such substrates.
To further evaluate the impact of the support modification on catalytic performance, we conducted recyclability tests on the modified catalysts Al–DBB–MgO and Al–OTES–MgO using hexanal reduction as a model reaction. As shown in Fig. 5, the turnover frequency (TOF) of Al–DBB–MgO remained stable after 5 reaction cycles, indicating excellent catalytic stability. While the TOF of Al–OTES–MgO dropped by around 14%, likely due to the leaching of Al(OiPr)3.
![]() | ||
Fig. 5 Recycling performance of modified Al–MgO catalysts (Al–DBB–MgO and Al–OTES–MgO) for the H-transfer reduction of hexanal. |
XPS and XRD analyse were performed on the spent catalysts after the recyclability tests investigate potential structural and surface property changes during the reaction. As depicted in Fig. 6a and b, an additional peak emerged at approximately 530.5 eV with an area of ∼20%, suggesting the formation of the more reduced oxygen species, which can be attributed to Mg(OH)2. The XRD patterns of the spent catalysts, as shown in Fig. 6c, further supported these fundings, showing an increased intensity of the diffraction peaks at 2θ = 38.2°, 51.0°, and 58.9°, corresponding to the (101), (102), and (110) planes of Mg(OH)2. This increment is likely due to accumulated surface hydrolysis accelerated by water during the catalyst recovery and washing. Additionally, oxidation state of Al remained largely unchanged. These observations confirm that the modified catalysts retain their structural integrity upon recycling, with only limited surface changes.
![]() | ||
Fig. 6 XPS analysis of spent Al–OTES–MgO (a) and Al–DBB–MgO (b); XRD patterns of fresh and spent catalyst for comparison (c). |
Among all the solid catalysts tested, Al–DBB–MgO demonstrated the highest catalytic activity, which can be attributed to the flexibility of the DBB carbon chains, facilitating enhanced reactant coordination. In comparison, Al–OTES–MgO showed moderate improvements in activity; however, the steric hindrance from densely packed OTES layers limited its performance, particularly with bulky substrates. Additionally, both Al–DBB–MgO and Al–OTES–MgO demonstrated remarkable stability during the cycling tests with minor structure change. These findings highlight the critical role of optimizing the surface chemistry of support materials to enhance catalytic performance.
This journal is © The Royal Society of Chemistry 2025 |