Next Article in Journal
Advancement of Metabolic Engineering Assisted by Synthetic Biology
Previous Article in Journal
The Mechanism of Adsorption, Diffusion, and Photocatalytic Reaction of Organic Molecules on TiO2 Revealed by Means of On-Site Scanning Tunneling Microscopy Observations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Functionalized Ordered Mesoporous Silicas (MCM-41): Synthesis and Applications in Catalysis

by
Gabriel Martínez-Edo
,
Alba Balmori
,
Iris Pontón
,
Andrea Martí del Rio
and
David Sánchez-García
*
Grup d’Enginyeria de Materials (GEMAT), Institut Químic de Sarrià, Universitat Ramon Llull, Via Augusta, 390, 08017 Barcelona, Spain
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2018, 8(12), 617; https://doi.org/10.3390/catal8120617
Submission received: 9 November 2018 / Revised: 30 November 2018 / Accepted: 1 December 2018 / Published: 4 December 2018
(This article belongs to the Section Nanostructured Catalysts)

Abstract

:
Mesoporous silica sieves are among the most studied nano-objects due to their stable pore structure and easy preparation. In particular, MCM-41 have attracted increasing research attention due to their chemical versatility. This review focuses on the synthesis and regioselective functionalization of MCM-41 to prepare catalytic systems. The topics covered are: mono and di-functionalized MCM-41 as basic and acid catalysts, catalysts based on metallic complexes and heteropolyacids supported onto MCM-41, metallic nanoparticles embed onto functionalized MCM-41 and magnetic MCM-41 for catalytic purposes.

1. Introduction

In recent years, the use of molecular sieves in both research and industry has increased exponentially. In particular, in the field of catalysis, special attention has been paid to mesoporous molecular sieves (pore diameter between 2–50 nm). These materials are widely used as adsorbents, catalysts, and catalyst supports due to its high surface areas, large pore volumes, and well-defined pore structure [1,2].
In this context, mesoporous silicas are regarded as very versatile materials, given the possibility to be easily derivatized with a wide variety of functional groups such as acids, amines, metallic nanoparticles, and organometallic complexes through co-condensation and/or post-synthesis grafting methods [2].
Among all the types of MSN reported (SBA-n (Santa Barbara amorphous silica), MSU-n (Michigan State University silica), KIT-1 (Korean Institute of Technology), IBN (Institute of Bioengineering and Nanotechnology) and FDU-n (Fudan University)), the family of MCM-n can be regarded as one of the most studied. Because of their unique properties, MCM-41 have found uses in many biological applications such as drug delivery or cell imaging, cell labeling, and catalysis for catalyst adsorption and enzyme immobilization [3].
The aim of this review is to offer an overview of the most common applications of functionalized MCM-41 in catalysis with a special emphasis in the methodology used to prepare the catalytic system. In the context of the present paper, the concept of functionalized MCM-41 encompasses organic functionalized groups, covalently attached to the silica matrix or forming ionic pairs, and the inclusion of other elements such as Al, B, and Zr into the silica framework or magnetic nanoparticles. Each application will be illustrated with recent examples from the literature, in particular from the period 2015–2018.
In detail, in Section 3.1 simple systems consisting in MCM-41 sieves containing acidic and basic moieties and mixed acid/basic functionalities will be discussed. In Section 3.2, the preparation of metallic complexes anchored onto the nanoparticle will be described. Section 4 is devoted to review the application of solid supported heteropolyacids endowed with improved thermal stability. The following section highlights the application of functionalized MCM41 to prepare supported metallic nanoparticles onto the silica framework (5). Finally, magnetic MCM-41 sieves, which allow a better separation and reuse of solid catalyst by simple application of an external magnetic field.

2. General Synthesis of Mesoporous Silicas (MCM-41) and Applications in Catalysis

The preparation of MCM-41 was first reported by scientists of the Mobil Company in 1992 in the context of a project of the company to find materials with pores larger than that presented by zeolites. Hence the name of the resulting materials, MCM, which stands for Mobil Composition of Matter.
The synthetic methodology is based on the condensation of silica precursors, typically sodium silicate or tetraethylorthosilicate, in the presence of cationic surfactants under basic conditions [4]. In fact, this procedure is an adaptation of the method described by Stöber in 1968 for obtaining silica nanoparticles. In particular, the standard preparation consists of mixing a silicate precursor, usually tetraethylorthosilicate (TEOS), with a cationic surfactant, such as a cetyltrimethylammonium bromide (CTAB), at a temperature between 30 and 60 °C and pH = 11. The nanoparticles are formed by the sol-gel process catalyzed in basic medium [1,5].
In the first stage of the process, the hydrolysis of the alkoxide takes place. Then, the silanol groups polymerize by condensation, forming three-dimensional structures linked by siloxane bonds (Si-O-Si). In the following step, the presence of cylindrical micelles formed by the surfactant is critical, since they act as a template and will give rise to the formation of the pores. The cationic surfactant attracts the negative charges of the silica species, which are concentrated around the micelles forming a tubular silica structure. The nanoparticle increases in size until the negative charge, introduced by the silica species, is so high that it stops growing. It should be noted that the size, hexagonal shape and regularity of the particles depend on various variables such as temperature, rate of addition, agitation and the amount of catalyst used with respect to that of TEOS [1,5].
Finally, the surfactant should be removed from the inside of the pores. To do so, three methods have been recommended: reflux in acidified alcohol with hydrochloric acid, treatment with ammonium nitrate or by calcination. These treatments allow the rupture of the electrostatic interaction that exists between the groups of the cationic surfactant head and the anionic silicates, which facilitates the elimination of the surfactant in the mesopores and the final formation of the particles (Figure 1) [1].
A distinctive characteristic of MCM-41 is the presence of two defined surfaces: an internal one (pores) and an external one (Figure 2). The regioselective modification of the nanoparticles with functional groups makes MCM-41 highly versatile and enables them to perform specialized tasks, such as cooperative catalysis [4]. Two methodologies allow the functionalization of MCM-41: grafting and co-condensation. The so-called grafting is a post-synthetic method that allows the modification of the surface of the nanoparticle by incorporating functional groups by silanization. In this process, silanol groups (Si-OH) at the surface act as anchoring points when treated with a functionalized siloxane bearing the desired group (e.g., APTMS (3-aminopropyltrimethoxysilane)). Silanization proceeds either through free groups of (≡Si-OH) or silanol geminal groups (=Si(OH)2). Importantly, the original structure of the mesoporous support is maintained after grafting (Figure 2). The co-condensation method is fundamentally based on the sol-gel process. The difference lies in the simultaneous addition of a functionalized siloxane with the silica precursor (TEOS). In comparison with the post-graft method in which the distribution of the functional groups tends to be heterogeneous, the co-condensation method is able to homogeneously distribute organic groups on the inner surface of the pores. Another advantage of the co-condensation method over the grafting method is that the morphology of the molecular sieve can be more easily controlled. In any case, each of the two functionalization methods has certain advantages. Direct functionalization of the surface in a single step of synthesis is easily achieved by grafting. However, if it is desired to obtain a homogenous and uniform surface, the co-condensation method is probably more practical.

3. Superficially Functionalized MCM-41

3.1. MCM-41 for Basic and Acid Catalyst

As stated before, the surface of MCM-41 can be relatively easily derivatized, allowing the introduction onto the silica of a wide variety of functional group such amines, thiols or carboxylic acids. The introduction of such functionalities opens the door to the design of tailored catalysts to perform specific tasks [6]. Arguably, in catalysis the most common functional groups introduced onto MCM-41 are amines and sulfonic acids.

3.1.1. Amino Functionalization: NH2-MCM-41

In 2013, Mirza-Aghayan et al. [7] reported the use of NH2-MCM-41 for the synthesis of 4H-chromene and 4H-benzo[b]pyrans as basic catalyst (Figure 3). The interest of these heterocycles lies on their pharmacological activity, mainly as anticoagulant and anticancer drugs. The NH2-MCM-41 were prepared using the grafting methodology. Briefly, MCM-41 were refluxed in toluene in the presence of 3-aminopropyltriethoxysilane (APTES). The three-components reaction (aldehyde, malononitrile and a 5,5-dimethyl-1,3-cyclohexanedione or 1-naphtol) afforded the 2-amino-tetrahydro-4H-chromene derivatives with good to excellent yields (ca 69–90%) using only 10 mol.% of the catalyst, whereas 2-amino-4H-benzo[h]chromene derivatives required a 20 mol.% (69–93%).
The following year, Nale et al. [8] reported on the synthesis of quinazoline-2,4(1H,3H)-dione derivatives from 2-aminobenzonitriles and carbon dioxide under high pressures conditions (3.5 MPa) in aqueous medium using amine functionalized MCM-41 as catalyst. The authors claim that the synthetic methodology can be applied for the synthesis of several biologically active derivatives such as prazosin, bunazosin and doxazosin. According to the report, after completion of the reaction, NH2-MCM-41 could be reused for five consecutive recycles without any significant loss in its catalytic activity (Figure 4).
The same group [9] investigated the application of NH2-MCM-41 catalysts for the reaction of carbon dioxide and aziridines yielding 5-aryl-2-oxazolidinones. Oxazolidinones are very versatile heterocycles used in organic synthesis as chiral auxiliaries. Furthermore, some derivatives are key intermediates compounds in the preparation of drugs such as linezolid (Zyvox), posizolid and torezolid. The catalyst was prepared following the co-condensation method by reaction of TEOS with 3-[2-(2-aminoethylamino)ethylamino]propyltrimethoxysilane (AEPTMS) (Figure 5).
5-aryl-2-oxazolidinones were prepared in good yields (65–99%) and excellent regioselectivity under mild and solvent free conditions in the presence of a 14.6 wt.% of amine functionalized MCM-41. The authors claim that the developed catalyst can be easily separated by filtration. Then, after a washing process with water and a posterior drying, the nanocatalysts could be reused for successive reactions up to five consecutive recycles at least, without apparent loss of activity. A tentative mechanism of the reaction is depicted in Figure 6.
Treatment of calcined MCM-41 (Figure 7) with 3-trimethoxysilylpropylethylenediamine, allowed Choudary et al. [10] the preparation of diamino functionalised MCM-41 sieves. The resulting molecular sieves were studied as catalyst for the aldol and Knoevenagel condensation reactions under very mild conditions. The Knoevenagel condensation renders the product with excellent yield, between 75 and 100%. Whereas the aldol condensation gave lower yields, ranging between 68 and 100%.
Another example of NH2-MSN-41 for the catalysis of the Knoevenagel condensation was reported in 2015 by Zhao et al. [11]. The reaction was conducted in water and under microwave irradiation. The results showed comparable catalytic efficiencies than those obtained with the homogeneous propylamine base-catalyst (99%). Obviously, the advantage of the heterogeneous system is that it can be easily reused, in this case up to eight times. The same year, the group of Röβner et al. [12] prepared MCM-41 functionalized with quaternary amino groups by grafting pristine MCM-41 with N-(trimethoxysilyl)propyl)-N,N,N-trimethylammonium chloride (TMTMAC) in toluene at reflux. The catalytic features of the system were studied using the Knoevenagel condensation of benzaldehyde and malononitrile as model reaction. The authors highlighted the importance of the homogeneity of the distribution of the quaternary amino group to obtain good catalytic activities.
Aminated MCM-41 are interesting catalysts from an industrial perspective given the stability and intrinsic reusability of these particles. Recently, Trisunaryanti et al. [13] reported the use of NH2-MCM-41 in the transesterification of vegetable oils to prepare biodiesel fuel from waste cooking oil. In the same work, the reaction conditions were optimized. It was concluded that applying a 4 wt.% catalyst/waste cook at a methanol/waste cook mole ratio of 15 produced a 49.98% of the desired product.

3.1.2. Acid-Derivatized MCM-41

The need for green processes have stimulated the application of solid supported acids, which minimize production of acidic waste and toxic effluents from mineral acid. In this regard, the silanol groups present on the surface of MCM-41 provide a convenient anchoring point for the inorganic sulfonic acid functionality (-SO3H), which is arguably the most studied acid group [14] (Figure 8).
Sulfonic acid functionalized MCM-41 are used as catalyst for the synthesis of a wide variety of products. A survey of applications of such catalyst have been reviewed elsewhere [14]. Recent examples are the synthesis of benzal-1,1-diacetate [15], a solvent-free synthesis of 2H-indazolo [2,1-b]phthalazine-1,6,11-trione derivatives [16]; multicomponent synthesis of 1H-pyrazolo[3,4-b]pyridines and spiro-pyrazolo-[3,4-b]pyridines [17]. Another important field of application of sulfonic functionalized MCM-41 is the development of “green processes”, for instance, in the production of biodiesel [18,19]. In this regard a recent work, [20] reported a “green synthesis” of 2-amino-6-(hydroxymethyl)-8-oxo-4-aryl-4,8-dihydropyrano[3,2-b]pyrans via a one-pot, three-component reaction of 5-hydroxy-2-(hydroxymethyl)-4H-pyran-4-one aldehydes and malononitrile in aqueous media (Figure 9).
Apart from the treatment of chlorosulfonic acid, SO3H-MCM-41 can be easily synthesized by anchoring 3-((3-(trimethoxysilyl)propyl)thio)propane-1-sulfonic acid onto MCM-41-type by means of the grafting methodology. The performance of these acidic sieves as catalyst was evaluated using the multi-component synthesis of 1H-pyrazolo-[3,4-b]pyridines and spiro-pyrazolo-[3,4-b]pyridines. Typical yields of these syntheses ranged from 84 to 96%. The resulting catalyst could be recycled and reused with negligible loss in activity over seven cycles (Figure 10) [17].
Alternatively, SO3H-MCM-41 can be obtained in two synthetic steps by oxidation of HS-MCM-41 by using 30 wt.% H2O2. The starting nanoparticles (HS-MCM-41) were prepared by co-condensation of TEOS and 3-mercaptopropyltriethoxysilane (MPTES). These acidic sieves were investigated as efficient catalysts for the transformation of xylose into furfural. Interestingly, the substitution of MPTES by 3-mercaptopropyl(methyl)triethoxysilane (MPTES) afforded supports with an improved xylose conversion. This result was rationalized in terms of increase of the hydrophobicity imparted by the methyl group of MPTES. The optimal temperature and reaction time for such process was found to be 155 °C and 2 h, respectively. Furthermore, the authors suggested that the furfural selectivity (>93%) was influenced by the pore architecture (Figure 11) [18].
A similar catalyst was introduced by Samart et al. [19] in 2016 for the production of biodiesel from Hevea brasiliensis oil. The authors investigated the effect on the reaction of four parameters to optimize the process: catalyst loading, reaction time, temperature and MPMDS molar composition. The optimal conditions were found to be: 5.06 wt.% catalyst loading, 120 min, 153 °C, and 0.266 of MPMDS molar composition. Using such conditions, the corresponding fatty acid methyl ester was prepared in 95.5% yield. The performance of the catalyst was benchmarked against H2SO4 and proven to be more effective. Additionally, the catalyst could be reused up to four cycles under the optimum reaction conditions without significant loss of product yield (Figure 12).
More recently, in 2017 Appaturi et al. [21] prepared DL-Alanine functionalized MCM-41 catalyst for the Knoevenagel reaction. The condensation products were obtained in excellent yields (87–96%). The catalyst was prepared by nucleophilic substitution of chloro-functionalized MCM-41 with the aminoacid (Figure 13). This methodology was applied for the synthesis of 3-(2-furylmethylene)-2,4-pentanedione in 92% yield.
In 2014 Vrbkvá et al. [22] studied the aldol condensation of 4-isopropylbenzaldehyde (iPB) and propanal (PA) to yield forcyclamaneldehyde (2-methyl-3-(4-isopropylfenyl)prop)-2-enal, FCA) under acidic and basic conditions using functionalized MCM-41 as catalysts. To this end, three types of functionalized MCM-41 sieves were prepared by post-grafting of pristine MCM-41 with 3-mercaptopropyl)trieth-oxysilane (3-cyanopropyl)trichlorosilane and N-[3-(trimethoxysilyl)propyl)]ethylene-diamine. The synthetic procedure is outlined in Figure 14.
The best results, in terms of conversion and selectivity, were obtained using SO3H-MCM-41. Interestingly, the SO3H-MCM-41 catalyst only yielded 4-isopropylbenzoic acid. While the basic catalyst shown almost no activity. The authors suggest that the good performance of SO3H-MCM-41 was due to the synergic effect of acid –SO3H groups and acid –OH groups of MCM-41 (cf. 3.1.3). The optimization study of the reaction concluded that the optimal reaction conditions were found to be 50 wt.% of SO3H-MCM-41, 100 °C and toluene as solvent. These conditions lead to a 69% conversion of 4-isopropylbenzadehyde after 24 h and 65% selectivity to the final product.

3.1.3. Difunctionalized MCM-41

Biological reactions are commonly mediated by synergistic catalytic processes. The extraordinary efficiency and selectivity of these transformations are greatly dependant on the positions and relative distances of the active sites. In order to mimic these natural systems and to achieve such level of complexity, a rigid support is needed to arrange the proper functional groups with a specific geometry. In this context and due to the chemical versatility and particular geometry, MCM-41 emerge as an ideal building block for the design of bifunctional catalysts for tandem reactions.
A first example was presented in 2007 by Asefa et al. [23] who reported the application of amino-functionalized MCM-41 for the Henry synthesis (nitroaldol). These researchers observed differences in the conversion of the Henry reaction depending on the relative amount of amino groups and silanols and their spatial distribution. Surprisingly, MCM-41 functionalized with a lower number of amino groups displayed a superior catalytic activity. This result was rationalized in terms of increase of cooperative sites, since amino groups and silanols are found spatially spaced (Figure 15). Such catalysts were prepared either by reacting excess amounts of APTMS in ethanol or by postgrafting smaller amounts of the aminoorganosilane in toluene for a short reaction time.
In 2016 the Henry reaction catalyzed with MCM-41 was revisited. Collier et al. [24] studied the aldol and nitroaldol reaction mechanisms using a Hammett analysis to determine the effects on the catalytic activity of electron-donating and electron-withdrawing groups on para-substituted benzaldehydes. Furthermore, the cooperative effect of the silanol groups was assessed by reacting these groups with hexamethyldisilazane (HMDS). As a result of such capping, both reactions required more elevated temperatures to achieve mesurable conversion and reduced the observed TOF values in comparison to catalysts with free silanols groups (Figure 16).
Mukhopadhyay et al. [25] designed a functionalized MCM-41 for the preparation of substituted 2-arylbenzothiazoles by reaction of o-aminothiophenol disulfides and aldehydes through the cleavage of the S-S bond (Figure 17). The catalyst was prepared in two synthetic steps. First, MCM-41 and 3-chloropropyltrimethoxysilane (CTMS) were refluxed for 24 h in acetonitrile. Then, the resulting chloro functionalized MCM-41 were treated with 2-(piperazin-1-yl)pyrimidine) to afford the catalyst. The mechanism of the reaction and role of the catalyst were carefully studied. According to the authors, the acidic character of MCM-41 assists in the formation of an imine between the aniline and benzaldehyde and the cleavage of the disulphide bond. The 2-(piperazin-1-yl)pyrimidine moiety would be responsible for the last dehydration step, which affords the aromatic thiazole ring. The reaction for the synthesis of the heterocycles was conducted in water under a steady flow of oxygen. Typical yields of the benzothiazoles were in the range of 85 up to 98%.
The aforementioned reports base the cooperativity in the catalytic activity on the coexistence of a silanol with an additional group (amine, SO3H). In 2011, Huang et al. [26] reported the preparation of bifunctional MCM-41 decorated with site-separated amino and sulfonic groups. More specifically, the sulfonic moieties were introduced by co-condensation of TEOS and STMOS, while the amino groups were grafted selectively onto the external surface of the nanoparticle. This regioselectivity was achieved by reacting APTMOS and the MCM-41 with the template (CTAB) still inside the cavities, therefore the grafting only affected the external surface of the silica matrix. The detailed synthetic route is depicted in Figure 18. This bifunctional catalyst was applied for the one-pot, two-steps reaction of acetals to render α, β-unsaturated nitro compounds.
A different approach to prepare bifunctional catalyst was proposed the following year by Studer et al. [27]. Instead of taking advantage of the mesoporous structure, this novel methodology relies on the functionalization of the MCM-41 by means of orthogonal click reactions namely thiol–ene reaction, Cu-catalyzed 1,3-dipolar alkyne/azide cycloaddition, and a radical nitroxide exchange reaction. The synthetic pathway is straightforward and consists of the co-condensation of two orthogonally functionalized tetraethoxysilanes. As an example, in Figure 19 the functionalization of MCM-41 with sulfonic and amino groups is outlined. In short, TEOS 1, azide-terminated triethoxyalkylsilane 2 and 7-octenyltrimethoxysilane 6 were mixed under basic conditions to give azido-alkene bifunctionalized MCM-41 (10). In turn, azides reacted with terminal alkynes (4) in the presence of Cu(SO4)2 furnishing triazoles, which allowed the introduction of a sulfonic acid moiety. Moreover, olefins underwent addition of sulphides in the presence of 2,2′-Azobis(2-methylpropionitrile) (AIBN). This reaction led the introduction of amines by treatment of the MCM-41 11 with 8 and AIBN.
In 2013, the same group prepared and tested bifunctional mesoporous silica sieves bearing a Pd-complex and basic sites in order to introduce a cooperative active catalyst in the allylation of ethyl acetoacetate (Tsuji-Trost reaction) [28]. Functionalization of the MCM-41 was effected by post-modification using click chemistry with orthogonal siloxane as has been described before. The structure of some of the systems prepared are shown in Figure 20. The selectivity of mono versus double alkylation was achieved by control of reaction temperature (RT, 30, or 70 °C) and the nature of the catalyst. With systems 17 and 18 the reaction proceeded smoothly and the monoallylated product was obtained almost in quantitative yield. Whilst catalyst 19 provided a 1:2 mixture of mono- (33%) and bisallylated product (66%).
Tsai et al. [29] prepared bifunctional MSN-41 catalysts intended for esterification reactions, containing a Brønsted acid site of diarylammonium triflate (DAT) and a pentafluorophenyl propyl (PFP) group (Figure 21). The high reactivity displayed by this system was attributed to the formation of a surface-bound hydrophobic layer of PFP molecules, which facilitates the extrusion of one of the reaction products (water) from the mesopores by suppressing water adsorption onto the surface, thereby shifting the reaction equilibrium to completion.

3.2. Metallic Complexes

The preparation of MCM-41 decorated with metallic complexes for catalysis has been a very active field of study in recent years. Two strategies are available to build these catalysts: grafting of properly functionalized complexes to the MCM-41 and multi-step functionalization of pristine MCM-41. The later methodology often entails the introduction of the amino functionality. Thus, complexes can be tethered to the MCM-41 by amine or imine linkages.
In 2014, Jabbari et al. [30], reported the encapsulation of salen copper(II) complexes. The catalyst was prepared by condensation of 5-bromo salicylaldehyde with NH2-MCM-41 and the subsequent treatment of these sieves with Cu(NO3)3. This nanocatalyst was applied for the oxidation of sulphides into sulfoxides using urea hydrogen peroxide (UPH) as oxidant. The oxidation reaction was carried out in ethanol at room temperature. The selectivity of the transformation was excellent and typical isolated yield of the products was in the range of 91 to 98% (Figure 22).
Ghadermazi and Ghorbani-Choghamarani et al. [31] reported the preparation of Cu(II) immobilized on Serine@MCM-41. The catalytic system was prepared by refluxing MCM-41 with serine in distilled water for 48 h. Subsequently, this Serine@MCM-41 support was treated with Cu(NO2)3 at reflux in ethanol to afford Serine Cu(II)@MCM-41 (Figure 23). The resulting nanomaterials were studied as a green and efficient catalytic system for the synthesis of four types of compounds: polyhydroquinolins, 2,3-dihydroquinazolin-4(1H)-ones, sulphides, and sulfoxide derivatives. Noteworthy, the researchers studied carefully the copper leaching. Using ICP-OEIS, it was concluded that the amount of copper in the fresh catalyst and the used catalyst after 11 times recycling was 0.51 and 0.45 mmol g−1, respectively, what is considered an excellent performance for such kind of catalysts.
Hajipour et al. [32] prepared Cu(II) Et-S@MCM-41 (Catalyst A, Figure 24) in two synthetic steps: treatment of NH2-MCM-41 with two equivalents of 2-chloroethyl phenyl sulfide in acetonitrile at reflux and, then, complexation with Cu (II) by exposure of the functionalized MCM-41 to Cu(OAc)2·2H2O in ethanol. The catalyst was tested for the preparation of aryl benzyl sulphides using a one-pot thioetherification reaction of aryl halides with benzyl bromide and thiourea in water in the presence of K2CO3 at 100 °C under aerobic conditions. As anticipated, the reaction of aryl chlorides presented lower yields in comparison with aryl bromides, but yields were significantly increased at longer reaction times. When electron-poor aryl bromides and chlorides were used, benzyl aryl sulphides products were obtained in higher yields compared to electron-rich aryl bromides and chlorides.
In 2016, Cai et al. [33] reported the immobilization of 1,10-phenanthroline-copper (I) complex. This catalytic system was designed for the cross-coupling reaction of aryl iodides with aliphatic alcohols. The reaction of iodobenzenes with aliphatic and benzylic alcohols in the presence of 2 equivalents of Cs2CO3 as base provided the corresponding ethers in yields up to 94%. The authors examined the stability of the catalyst by using ICP analysis. The results confirmed that after eight consecutive runs, only 1.3% of copper had been lost from the support. This analysis is in agreement with an additional experiment carried out to confirm that the reaction is actually catalyzed by the heterogeneous catalyst. Briefly, after conversion of 50%, the catalyst was removed and the solution was allowed to react. After the removal of the catalyst, no further conversion was detected. The use of recyclable copper catalyst along with mild and environment-friendly reaction conditions make this methodology a powerful complement to the traditional C-O coupling reactions. The preparation of the catalyst is delineated in Figure 25.
Liu et al. [34] immobilized the complex shown in Figure 26 by electrostatic interaction. The authors reported that the catalyst was capable to effect an efficient conversion of methane into methanol without over-oxidation under ambient conditions. The efficacy of the system to mediate the catalytic conversion of these light alkanes in natural gas into liquid oxidized products was also assessed. The catalyst was prepared by ion exchange via electrostatic attraction with MCM-41 modified by grafting with anionic 3-(trihydroxysilyl)-propylmethyl-phosphonate (TP).
MCM-41-supported N2O-type Schiff base complexes were prepared by condensation of 4-methyl-2,6-diformylphenol (DFP) with APTMS modified molecular sieves. Subsequent treatment with the appropriate Cu(II) and Ni(II) salts afforded active nanocatalysts useful for esterification, Diels-Alder cycloaddition and aldol condensation. The results were moderate to excellent regardless the metal of the complexes. Typical yields ranged from 70 up to 86%. Furthermore, the complexes were thermally stable at the reaction conditions and no leaching of the complexes was observed [35] (Figure 27).
A different approach to prepare immobilized Schiff base complexes was presented by Rangappan et al. [36] in 2016. This group designed a synthetic pathway based on the preparation of a siloxane bearing a preformed Cu(II) complex. Subsequently, this compound was anchored onto the MCM-41 by grafting (Figure 28). The as-synthetized Cu-Schiff base MCM-41 catalyst was tested for Ullmann-type coupling reaction of the aryl halides with aryl halides, phenols, amines and N-heterocyclic amines. The yields of the coupling were excellent using DMSO as solvent. As expected, iodo-aromatics provided the best yields, ranging from 85 to 90%. The catalyst was recovered and could be reused by simple filtration.
Palladium chemistry plays a pivotal role in modern organic synthesis. Palladium-based catalysts are employed in a significant number of modern synthetic transformations such as Tsuji-Trost allylation, hydrogenations, Stille, Heck, Suzuki, and Sonogashira cross-coupling reactions [37]. However, the use of homogenous Pd(0) complexes usually require a purification step for the catalyst removal. Therefore, the immobilization of palladium can simplify the synthetic procedures. In this sense, encapsulation in MCM-41 turns out to be advantageous [38,39].
Tsai et al. [40,41] studied a highly-efficient and recyclable MCM-41 anchored palladium bipyridyl complex-catalyzed (Figure 29). The MCM-41 were prepared by grafting of the palladium bipyridyl complex onto pristine MCM-41. This catalyst was applied for the Sonogashira coupling of heteroaryl halides and acyl chlorides with terminal alkynes. Yields were high in both cases, ranging from 58 to 96%. According to the authors, satisfactory results can be achieved with catalyst loadings as low as 0.002 mol.% Pd.
Dibenzo-18-crown-6 functionalized MCM-41 were prepared by Yousefi et al. [42] by reacting aminopropyl-grafted MCM-41 with 4′,4′′-diformyl-dibenzo-18-crown-6. The catalyst was very efficient for nucleophilic ring opening of oxirans in water with cyanide and azide. The reaction was carried out in water, and a variety of azidohydrins and cyanohydrins was obtained with yields between 75 and 93%. It is worth to mention that the nucleophilicity of the anion could be tuned by using the proper cation. Thus, the use of potassium salts reduced the time of reaction up to 40%. This result was rationalized in terms of better complexation of the K+ by the crown-ether due to the size of the cavity. Remarkably, no evidence for diols formation as reaction by-product was observed (Figure 30).
Nikoorazm et al. [43] anchored a palladium S-benzylisothiourea complex on functionalized MCM-41 as an efficient and reusable catalyst. This MCM-41 catalyst was used for the synthesis of 5-substituted 1H–tetrazoles. This synthesis is predicated on the [2 + 3] cycloaddition reaction of various organic nitriles with sodium azide in poly(ethylene glycol) as green solvent as a palladium complex (Figure 31).
Fadhli-Fraile et al. [44] also used Ti/MCM-41, titanium grafted on the silica surface through the formation of Si-O-Ti bonds, for use in the epoxidation of styrene with tert-butyl hydroperoxide (THP). The catalyst was synthetized by pre-treatment of the surface of the scaffold with Ti(Oi-Pr)4 as titanium precursor. Then, diethyl l-tartrate or diisopropyl l-tartrate (0.4 mmol) was added to the MCM-41. The resulting suspension was refluxed in toluene for 12 h to yield the desired catalyst (Figure 32). According to the authors, Ti/MCM-41 is an active catalyst for the epoxidation of styrene with THP. However, its efficiency was not optimal, probably due to the subsequent epoxide rearrangement to phenyl acetaldehyde.
Zhou and He et al. [45] anchored RhCl(CO)(PPh3)2 to diphenylphosphinopropyl modified (–PrPPh2) MCM-41 with the aim to prepare a catalyst for the hydroformylation 1-octene. The catalytic sieves were prepared by grafting of MCM-41 with CPTMS. In the following step, diphenylphosphine was introduced by nucleophilic substitution. Finally, the as-synthetized material was exposed to PhCl(CO)(PPh3)2 to yield the heterogeneous catalyst. The system exhibited moderate conversions (about 84%) with loading of 1.5 wt.% in Rh. The synthetic pathway for the preparation of the functionalized particles is delineated in Figure 33.
Havasi et al. [46] used a nickel complex anchored onto MCM-41 for the synthesis of 2,3-dihydroquinazolin-4(1H)-ones, which are an important class of fused heterocycles with a range of pharmacological and biological activities, as antibacterial, antitumor, analgesic, antibiotic and antidepressant agents. These heterocycles were synthetized by cyclocondensation of 2-aminobenzamide with aromatic aldehydes under mild reaction condition in polyethylene glycol molecular weight 400 (PEG-400) as solvent (90–98% yield). The catalyst was prepared in two synthetic steps. First, chloro-substituted MCM-41 was obtained by grafting using CPTMS. Then, the particles were treated with dithizone in dimethylformamide at 110 °C for 30 h. Finally, the later material was refluxed in ethanol in the presence of Ni(NO3)2.
Wang et al. [47] synthesized Zr-Schiff base/MCM-41 using the methodology delineated in (Figure 34). This catalyst was applied for the conversion of fructose into 5-hydroxymethylfurfural (HMF) The Zr/MCM-41 shown a superior activity than pure MCM-41 because of the Lewis acidity of the metal. The authors studied the effects of reaction variables such as reaction time and temperature using DMSO as solvent. The best result was obtained at 140 °C, after 4 h of reaction. The optimized yield was 92%. In addition, other sugars were also used as substrates for the synthesis of HMF. Unfortunately, the results were not as satisfactory as the observed ones for fructose.
Molibdenum and Tungsten complexes [MX2(CO)3(L)2] (M = Mo, W; X = I, Br; L = 2- amino-1,3,4-thiadiazole) were immobilized in MCM-41 and their catalytic activity was assessed. As delineated in Figure 35, 2-amino-1,3,4-thiadiazole was attached onto the MCM-41 by refluxing the chlorinated MCM-41 with the ligand in benzene. The required starting material was prepared by grafting of pristine MCM-41 with 3-chloropropyltriethoxysilane. Afterwards, the required complexes were prepared by exposure of the functionalized MCM-41 to the corresponding metallic salt in presence of a base. The supported complexes were tested as catalysts for the epoxidation of several olefins in the presence of tert-butyl hydroperoxide (TBHP). The conversions and TOF significantly depend on the complex and the substrate. However, the authors conclude that, in general, the immobilization of the catalysts improved their performance [48].
Mikoorazm et al. [49] prepared Co(II), Fe(III), Ni (II) and Zn(II) Schiff base complexes anchored on the internal surface of MCM-41 pores. The synthesis of these catalysts required aminopropyl functionalized MCM-41 as starting material. These mesoporous molecular sieves were reacted with 3,4-dihydroxybenzaldehyde under ethanol reflux. Subsequently, the resulting material was treated with the corresponding metallic salt to furnish the aforementioned catalysts. These catalysts exhibited good selectivity and catalytic activity in the oxidation of sulfides to sulfoxides and oxidative coupling of thiols to their corresponding disulfides using urea hydrogen peroxide (UHP) as oxidant at room temperature. The authors highlighted the absence of over-oxidation of sulfides and thiols to sulfoxides and sulfones, respectively [50] (Figure 36). A similar system was prepared by Rangappan et al. [51] by condensation of 3-aminopropyltrimethoxysilane with o-hydroxyacetophenone and exposure to Cu(CH3COO)2·H2O to form a complex, which mimics the “CuN2O2” coordination site that resembles the galactose oxidase enzyme.
In 2015, the group of Bania et al. [52] reported the preparation of immobilized Mn(III)-chiral Schiff base on MCM-41 intended for the enantioselective oxidative dimerization of 2-naphtol to R/S BINOL (1,1’-Bi-2-naphthol) in the presence of oxygen. The catalyst was obtained by reaction of the Mn(III) complex with pyridine functionalized MCM-41 as depicted in Figure 37. In turn, the pyridine moiety was introduced by imine formation between an amino functionalized MCM-41 and 4-pyridinecarboxaldehyde. In the presence of a substoichiometric amount of the nanocatalyst, as low as 20 mg, binaphthols were obtained in nearly quantitative yields with high enantioselectivity of up to 91% ee. These results are comparable with those found for the homogeneous catalyst.
Porphyrins are among the most studied ligands in catalysis for the stabilization of high-valence cations. In 2015, Kong et al. [53] communicated the preparation of [5,10,15,20-tetrakis-(pentafluorophenyl)porphyrin] manganese immobilized on amino-functionalized MCM-41 (Figure 38). The porphyrin was axially coordinated to the catalyst by means of complexation of the amino group. The catalyst was applied for the hydroxylation of naphthalene in the presence of m-chloroperbenzoic acid as the co-oxidant. The system showed an average conversion of up to 31% and good selectivity 1-naphtol/2-naftol. According to the authors, this low conversion, in comparison with the homogeneous catalyst, can be explained by the axial coordination between manganese cation and the amino group. The catalyst could be reused several times (>5) without no apparent loss of activity.
Quijada et al. [54] used MCM-41 impregnated with a zirconium catalyst (Me2Si(Ind)2ZrCl2) for the preparation of nanocomposites through the in situ polymerization of propylene. Three different methodologies for impregnation of the zirconocene were evaluated (Figure 39). Two of the methodologies pre-treat the MCM-41 support with co-catalyst methylaluminoxane (MAO) but differ in drying temperature and catalyst concentration. The third methodology consist of the pre-activation of the catalyst with MAO, followed by impregnation onto the MCM-41. The later procedure achieved about 50% higher activities compared to the other pre-activated metallocene/MAO catalysts. The smaller impregnation time and the significantly lower Zr loads are other advantages of this protocol.

3.3. MCM-41 Functionalized with Metallic Complexes for Enzyme Entrapment

Mesoporous silica materials have been shown to be efficient supports for immobilization of enzymes. These heterogeneous catalysts have been explored extensively in organic catalysis because they can provide important advantages over the free enzyme, such as improved enzyme recovery and reutilization, easy continuous operation, higher stability of the immobilized enzyme in terms of temperature and pH. [55,56]. All of these materials exhibit very similar chemical properties, but hexagonally ordered pore size of MCM-41 are smaller (20–40 Å) than other mesoporous silica materials, such as SBA-15 (60–150 Å) and MCF (200–420 Å), restricting the enzyme loading on the pore system [57,58].
However, enzymes such as cytochrome c, lysozyme, lipase, and albumin have been successfully immobilized onto the surface of functionalized ordered mesoporous materials MCM-41. A pioneering study was reported by Diaz and Balkus in 1996 [59]. In this regard, a simple strategy for the conjugation of MCM-41 with enzimes is based on the formation of metallic chelates between the two moieties.
Faramarzi-Khoobi et al. [60] took advantage of this strategy for the immobilization of a lipase. To do so, polyethylenimie (PEI) coated MCM-41 was prepared by treatment of MCM-41 with [3-(2,3-epoxypropoxy)propyl] (EPO) in toluene at 80 °C. The, the resulting solid was exposed to different solutions of divalent metal ions (Co2+, Cu2+, and Pd2+). Finally, the surface of the molecular sieve was coated with the lipase through the formation of a chelate. The ions facilitate the adsorption thanks to the interaction of the thiol group of cysteine (Cys), indole group of tryptophan (Trp), and imidazole ring of histidine (His) present on lipase surface. The catalytic activity of the immobilized lipase was studied in the esterification reaction of ethanol with valeric acid. The resulting ester, ethyl valerate (green apple flavor), was obtained in good yield (60%).

4. MCM-41 as Support for Immobilization Heteropolyacids

The use of heteropolyacids (HPAs) or their more often used equivalent polyoxometalates (POMs) in acid catalyzed reactions has attracted increasing attention in the last years especially for their application in processes requiring high acidic conditions [61]. Chemically, HPAs possess both redox potential and Brønsted acidity, structural flexibility and can stabilize cationic organic intermediates. Moreover, these reagents are used in several industrial processes since they are relatively stable, strongly acidic and many of the environmental pollution and corrosion problems of the traditional technologies are avoided.
Arguably, the most important and common HPAs for catalysis are of Keggin structure, which comprise heteropolyanions of the formula [XM12O40]n−, where X is the heteroatom (P, Si, etc.) and M the addendum atom (Mo, W, V etc.), because of the higher thermal stability and easier preparation compared to others species [62]. However, the use of HPAs has encountered some difficulties, which have reduced their application in some organic transformations. The main drawbacks of this class of catalysts include high solubility in polar solvents, weak thermal stability, low specific surface area and difficulty of recovery [63,64].
To circumvent these issues, the immobilization of heteropolyacids on inorganic supports has been suggested. In this regard, MCM-41 are especially promising because they can improve their catalytic activity thanks to their high specific surface area and pore volume [65]. Furthermore, the regular pore structure and tuneable pore size enables rapid diffusion of reactants and products through the pores to minimize consecutive reactions. Usually, tungsten HPAs, in particular 12-tungstophosphoric acid (H3PW12O40, HPW, TPA), are the catalysts of choice considering their stronger Brønsted acidity, lower oxidation potential and higher thermal stability [66,67,68,69,70,71].
In general, HPAs have been introduced inside the pore structure of MCM-41 by common impregnation techniques involving passive diffusion of poly-oxometalate anions into the pores of mesoporous silica [72,73,74]. Unfortunately, the HPA/MCM-41 impregnated system commonly lacks of stability in polar medium because of species are not anchored strongly onto the surface. In an attempt to diminish the leaching of HPAs, functionalization of MCM-41 with an alkyl amine or other functional groups bearing positive charges on the surface for supporting heteropolyacids has been studied extensively.
In 2006, Valente et al. [75] reported the use 12-tungstophosphoric acid (HPW) supported on different MCM-41 hybrid composites for the conversion of xylose to furfural in liquid phase and they compare the catalytic performance between them and the unsupported HPW catalyst. The heteropolyacid was immobilized in medium-pore (MP) and large-pore (LP) MTS silicas by incipient wetness impregnation, giving materials with HPW loadings in the range of 15–34 wt.%. Besides, HPW was also heterogenized in aminopropyl functionalized MP and LP silicas at loadings of 15.3 and 30.6 wt.%, respectively. These catalysts exhibited higher activity than homogeneous PW, and were comparable with H2SO4 under similar reaction conditions, achieving a yield of 58%. The study showed that the catalytic performance of the HPW-supported catalysts depends on the interplay of several variables, such as the preparation method, type of support, HPW loading, temperature (140–160 °C) and type of solvent (toluene/water or DMSO). According to the characterization data, the authors concluded that catalytic results were at least be partly related with the presence of stronger host–guest interactions in these materials, which lead to better dispersions of the polyoxotungstate anions.
HPW anchored on amino-functionalized MCM-41 (Figure 40) was also used for the oxidesulfurization of dibenzothiophene (DBT) by Zhu et al. [76]. In an attempt to produce zero sulfur fuel and reduce the SO2 emission, reaction conditions including HPW loading, reaction time, catalyst dosage, molar ratio of H2O2 to sulfur (O/S) and reaction temperature were investigated. The results showed excellent catalytic activity in the ultra-deep desulfurization of fuel oils with a conversion of DBT up to 100% when the reaction was conducted using 20% HPW–NH2–MCM-41 at 333 K. Noteworthy, the catalyst could be recovered by simple filtration retaining its high catalytic activity up to five runs. The authors concluded that the good recoverability of HPW–NH2–MCM-41 catalyst was attributed to HPW species are chemically immobilized on the surface of modified MCM-41.
Immobilization of phosphotungstic acid with Wells–Dawson structure (H6P2W18O62, D-HPW) on different amino siloxane functionalized MCM-41 was studied by Ding and Han et al. [77]. The amino siloxanes (3-aminopropyltrimethoxysilane (KH-540), N-(2-aminoethyl)-3 aminopropyltrimethoxysilane (KH-792) and 3-[2-(2-aminoethylamino)ethylamino]-propyl-trimethoxysilane (NQ-62)) were chosen as the grafting agents. Firstly, MCM-41 was functionalized with the amino siloxanes and, afterwards, the polyoxometalate was introduced by wet impregnation method. The catalytic activities of the three catalysts were compared with pure MCM-41 for the gas phase dehydration of glycerol to acrolein. The best catalytic performance was achieved using D-HPW/NQ-62-MCM-41. The strong interaction between amino siloxane and D-HPW generated strongly distorted heteropolyanions on the surface of the host materials. The authors suggested that this highly deformed structure probably rendered H+ more active and resulted in an increase of Brønsted sites and a decrease in the amount of Lewis sites. Accordingly, glycerol conversion and acrolein selectivity was improved with the increase in the number of amine groups, D-HPW/MCM-41 < D-HPW/KH-540-MCM-41 < D-HPW/KH-792-MCM-41 < D-HPW/NQ-62-MCM-41. Modification with amino siloxane also improved the catalyst stability, decreasing the leaching ratio and the amount of coke formation. The amine groups not only interacted with the heteropolyanions slowing the leaching but also behaved as basic sites and decrease the total acidity, which was beneficial to suppress the coke formation. In order to study the recovery the D-HPW/NQ-62-MCM-41 catalyst, burn the soft coke, and avoid the decomposition of NQ-62 and D-HPW, thermogravimetric analysis was carried out. The results established that the treatment of D-HPW/NQ-62-MCM-41 catalyst at 350 °C under an oxygen flow for 3 h enables the reusability of the catalyst with a slight decrease in glycerol conversion and acrolein selectivity.
The catalytic activity of HPAs can be improved by partial substitution of redox metal ions in Keggin structure, among which the introduction of vanadium has been demonstrated to be beneficial for redox catalysis. On this subject, amino-functionalized MCM-41 has been used to strongly immobilize through hydrogen interactions a new series of vanadium-substituted phosphotungstates (Figure 41) [78]. Modification of the mesoporous materials has been accomplished using APTES as the grafting agent. Subsequently, HPW (H3[PW12O40]) and different vanadium substituted polyoxometalates (H4[PVW11O40], H5[PV2W10O40] and H6[PV3W9O40]) have been introduced inside the pores by wet impregnation. The catalytic performance of these hybrid materials has been compared in the selective oxidation of benzyl alcohol to benzaldehyde. The results have established that the catalytic activity to be as follows: HPW < H4[PVW11O40] < H6[PV3W9O40] < H5[PV2W10O40], reaching 97% of conversion and 99% of selectivity in the oxidation of benzyl alcohol when H5[PV2W10O40] is employed. Using optimum reaction conditions; 30% aqueous H2O2 as oxidant, 0.05 g of catalyst, 5 mL of toluene as solvent, 4 mmol alcohol, H2O2/benzyl alcohol ratio = 5:1 at 80 °C for 8 h, the generality of the reaction has been studied. The study has verified that conversion depends on the alcohol nature: aromatic alcohols (82–100%) > cyclohexanol (~80%) > aliphatic alcohol chains (15–27%). Noteworthy, the catalyst can be recovered and reused without appreciable loss of activity. Then, it can be concluded that the hybrid system is a highly efficient recyclable catalyst for the selective oxidation of aromatic alcohols to the corresponding aldehydes.
In some applications, the presence of the support is critical in order to achieve high reaction selectivity. An example of such effect was reported by Popa et al. [79]. In 2017, these researchers carried out the preparation of novel catalysts based on molybdophosphoric acid H3PMo12O40 (HPM) and molybdovanadophosphoric acid H4PVMo11O40 (HPVM) Keggin-type HPAs by supporting on amino-functionalized MCM-41 and SBA-15 molecular sieves. The authors conducted thermal analysis to determine the effect of the support on heteropolyacid stability. An important mass loss above 543–553 K was observed, probably due to decomposition of amino propyl functional groups. The catalytic activity of HPA/MCM-41-NH2 and HPA/SBA-15-NH2 was tested for vapor-phase ethanol dehydration reaction, between 463 and 543 K of temperature range in nitrogen and air atmospheres. The results displayed an enhanced oxidation catalytic activity (formation of acetaldehyde) and a lower acid catalytic activity (formation of ethylene and diethyl ether) compared to the unsupported catalyst. At 463 K, a good selectivity was achieved albeit yields were low. However, between 483 and 503 K, acetaldehyde was obtained in high selectivity and at good values of conversion.
Ren et al. [80] reported the preparation of di-vacant Keggin based lacunary silicious polyoxometalate [γ-SiW10(H2O)2O34]4− supported on MCM-41 (physisorbed) and amino-functionalized MCM-41 (chemisorbed). Both materials were employed to investigate their activity and reusability for the oxygenation of organic sulphides using H2O2. To this end, MCM-41 was functionalized with APTES and then the amino moieties were protonated with nitric acid to anchor the polyoxometalate (Figure 42). The recyclability of both physisorbed and chemisorbed catalysts were tested, and the latter was clearly advantageous in maintaining full reactivity throughout recycling. The decrease in activity for the physisorbed catalyst was likely caused by the loss of catalyst back into solution. The reactivity of chemisorb catalyst was examined with several other substrates under similar conditions, revealing its utility in the removal of refractory sulfides via oxidative desulfurization. The study revealed that the selectivity to convert sulfide into the sulfoxide entirely before the appearance of sulfone was 90% using methyl phenyl sulfide and 75% using diphenylsulfide as substrates.
In addition to amino-propyl grafting, other aminated modifiers have been examined. In 2015, Tayebee et al. [81] designed a novel pyridine-functionalized mesoporous MCM-41 material (TPI-MCM-41). The MCM-41 support was functionalized via co-condensation of hydroxyl groups of the support and ethoxy groups of the TPI (N-[3-(triethoxysilyl)propyl]isonicotinamide) linker (Figure 43). The heteropolyacid (H4SiW12O40) was anchored to TPI-MCM-41 through electrostatic interaction in a second step. The authors fully characterized the catalysts by using various analyses such as XRD, SEM, EDX, UV-Vis, DTA-TGA, DLS, and FT-IR spectroscopy. The obtained catalyst exhibited a good dispersion of the heteropolyanions on the solid support surface and its structure was preserved after TPI immobilization. The catalytic performance of the catalyst was assessed for the one-pot synthesis of 1-amidoalkyl-2-naphthols by condensation of β-naphthol and benzaldehyde with benzamide under solvent-free conditions. The authors compared the catalytic activity of the silicotungstic Keggin-type heteropolyacid (H4SiW12O40) with several heteropolyacids (H3PMo12O40 (Keggin), H3PW12O40 (Keggin), H6P2W18O62 (Wells–Dawson), H6P2Mo18O62 (Wells–Dawson), H4SiW9Mo2O40 (Si-substituted mixed metal lacunary Keggin) and H4SiW9Mo3O40 (Si-substituted mixed metal Keggin)). The novel system showed the best catalytic activity (95% of yield) when compared with these heteropolyacids (75–80% yield). Moreover, the catalytic performance of the catalyst was compared with some mesoporous and heteropolyacid systems reported previously in the literature (H6P2W18O62/TPI-CNT, H5PW10V2O40/TPI-SBA-15, H5PW10V2O40/Pip-SBA-15, H6P2W18O62/TPI-(Al)SBA-15). The results revealed the superior catalytic activity of the new TPI-MCM-41 support (94% of yield) in front of the SBA-15 systems (yields ranging between 33–55%). In addition, the inorganic–organic hybrid catalyst was recovered and reused for seven reaction runs with only slight loss of the catalytic activity.
An alternative strategy to immobilize HPAs in MCM-41 is predicated on (POM)-based metal–organic framework (MOF). MOF are crystalline porous coordination polymers composed of inorganic clusters bridged by organic ligands which possess a 3D ordered porous structure [82]. POM-based MOF materials normally contain POM units and thus, generate MOF materials with open network structures [83]. POMs can be either directly part of the frameworks of MOFs or can be encapsulated within the cavities of the MOFs networks.
Following this approach, a novel catalyst containing (POM)-based metal–organic framework (MOF) in the pore of MCM-41 (POM@MOF-199@MCM-41) was designed by Zhao et al. [84]. In detail, MOF-199 is a kind of micropore MOF with Cu(II) dimers linked into paddlewheels by 1,3,5-benzenetricarboxylic acid (BTC), owning two types of cages of cuboctahedral symmetry (Figure 44). The preparation of the hybrid material POM@MOF-199 was carried out by a one-pot POM template self-construction of MOF-199 in the pore of MCM-41 via the hydrothermal method. The catalyst was synthetized mixing Cu(NO3)2·3H2O, PMo12−xWxO40 and MCM-41 in a water solution. Afterwards, (CH3)4NOH·5H2O (TMA) and BTC were added with stirring. Finally, the mixture was heated at 180 °C for 16 h in an autoclave. The hybrid POM/MOF material was exploited in the direct oxidative desulfurization of fuel oil, using dibenzothiophene (DBT) as model, in the presence of O2 as oxidant. In order to prove the synergistic effect of the hybrid system, the oxidative desulfurization rate of POM@MOF-199 and POM@MCM-41 was analyzed. As expected, POM@MOF-199@MCM-41 exhibited the highest desulfurization reactivity (98.5% DBT conversion) and the highest level of reusability. Additionally, a series of PMo12−xWxO40/MCM-41 ratios (20%, 40%, 60% and 80%) and different phosphomolybdenum polyoxometalates were explored. The results showed that conversion of DBT increased with the W loaded amount. Among the supported heteropolyacid catalysts, the PMo6W6O40@MOF-199@MCM-41 showed superior performance, which could be enhanced with an increment of POM content in the MOF/MCM-41 structure.
One year later, the same group reported the preparation of Cs+ ion modified POM hybrid MOF in MCM-41 (Cs-POM@MOF-199@MCM-41) for the DBT oxidative desulfurization in the presence of oxygen [85]. Two different synthetic methods, the direct method and the substitution method, were used to construct the Cs-POM. The heteropolyanions were encapsulated into the pores of MCM-41 with MOF using the protocol previously reported. The results established that the introduction of Cs+ ion modifies the Keggin heteropolyacid catalytic performance achieving a DBT desulfurization improvement comparing with the previous POM@MOF-199@MCM-41 system. Regarding Cs-POM@MOF-199@MCM-41, better results were obtained using Cs-POM synthetized by substitution method. The relative purer active constituent of Cs+ modified heteropolyacid increase Lewis acidity and promoted DBT conversion (up a 97.8%), while using the system obtained by direct method the conversion was only 76.1% in 150min. Under optimal conditions of Cs+ = 2 molar ratio (replacing the H+ with Cs+ cations in H3PMo6W6O40), 2.0 g/L of the catalyst dosage, 80 °C as reaction temperature and 500 rpm the DBT conversion rate achieved was 99.6% in only 180 min. Moreover, the catalyst was recyclable and could be recovered and reused up to 10 times with a small decrease in DBT conversion (from 99.6% to 98.5%).
In another attempt to optimise the desulfurization processes, the same authors used the broad experience in MOF-POM hybrid materials to introduce a family of heteropolyacid-type surfactants [(CH3)2(CnH2n+1)N(CH2)]2[POM] (SRL-POM), n = 10, 12, 14, and 16) in the pore of MCM-41 (SRL-POM@MOF-199@MCM-41) [86]. These surfactant Keggin-type POMs are amphiphilic quaternary ammonium micelles of POMs, with a surfactant surface and a POM core. Therefore, they can act as both the surfactant and the catalyst helping in the hydrophobic distribution of the catalyst in oil to favor the mass transfer and the reaction rate. Optimizing the reaction conditions for the catalytic oxidation of dibenzothiophene (DBT), 100% selectivity was obtained. The reusability results showed that the conversion of DBT decreased only from 100% to 99% even after 14 recycles.

5. MCM-41 for the Incorporation of Metal Nanoparticles (MNP)

The incorporation of metal nanoparticles in the pores of MCM-41 produce materials with interesting catalytic capabilities [87]. The effect of the mesoporous support is twofold: the aggregation of metal nanoparticles is prevented and its dispersion is enhanced. Several methods have been reported in the literature for the preparation of MCM-41 supported metal nanoparticles, which can be categorized into two groups [88]. The first one, the so-called two-step procedure, relies on the embedding of the MNP through simple physical entrapment into the pristine MCM-41. The main disadvantage of this procedure is the possible leaching of the metal species leading to the formation of large particles because of the weak interaction between the catalyst and the silica framework [89]. The second one requires the preparation of functionalized MCM-41 with organic ligands containing one or more heteroatoms. The role of these ligands is to interact by electrostatic or non-covalent forces with the MNP avoiding migration.
A wide variety of catalytic systems using MNP supported on MCM-41 have been reported. Most of these nanomaterials have been proposed as catalysts for applications in organic synthesis. A clear example is the use of palladium nanoparticles for the catalysis of organometallic cross-coupling reactions [88,90]. Another metal that has lately received attention is gold, which is highly selective in hydrogenation and oxidations and plays a major role in A3 coupling reactions, which involves multicomponent reactants [91]. Nickel and titanium nanoparticles have found also many applications as catalysts for hydrogenation of aromatics and epoxidations reactions [92]. Finally, ruthenium is employed for industrial processes such as the synthesis of paraffins, methanation of CO, or hydrogenations [93].

5.1. Palladium Nanoparticles Supported on MCM-41

The Suzuki reaction is one of the most versatile palladium-catalyzed cross-coupling reactions. This transformation requires as coupling partners a boronic acid and a halogenated derivative and is usually catalyzed by Pd0 complexes.
In order to simplify the workup of the reaction, several groups have reported the encapsulation of Pd nanoparticles inside the MCM-41 channels. An example of this kind of system was reported by Yamashita et al. [90]. Interestingly, the authors proved that the performance of the catalyst can be enhanced if the template (CTAB) was inside the mesoporous sieves due to the basic character of the MCM-41. It was hypothesized, that more active Pd nanoparticles were generated when the template was kept within the silica framework in comparison with the removed ones. Such assumption was proven for the catalytic activity of the Suzuki reaction using Pd/MCM-41 (CTAB) and Pd/MCM-41 (non-template). The results showed a catalytic activity of 100% when the template was kept within the silica framework, whereas only a 50% of conversion was achieved when the template had been previously removed.
In 2015, Wan et al. [88] proposed a similar methodology. This group developed a Pd/MCM-41 catalyst by using (NH4)2PdCl4 as Pd precursor. The authors postulated that surfactant micelles, which were used as a scaffold for the MCM-41 formation, helped the grafting of Pd species. As a result, highly dispersed Pd nanoparticles could be immobilized inside the channels of mesoporous SiO2 through a one-pot reaction (Figure 45), whereas merely non-encapsulation was achieved when the CTAB was previously removed. Accordingly, these Pd nanoparticles were well protected by the channels avoiding any leakage, showing a high catalytic activity (96% of conversion for iodobenzene to biphenyl) and notable cyclability.
Another relevant palladium-catalyzed reaction is the Heck coupling between olefins and organohalides. A catalyst intended for this coupling was presented in 2017 Azaroon et al. [94]. In this catalyst, the Pd nanoparticles were immobilized by interaction with the macrocyclic ligand dibenzo-18-crown-6-ether (cf. 0), which was previously attached to the MCM-41 by means of amine groups (Figure 46). The incorporation of the nanoparticles was achieved thanks to the use of palladium acetate as a source of the metal, which was reduced to Pd0 in ethanol and stabilized within the cavity of dibenzo-18-crown-6-ether. Therefore, Pd aggregation was prevented and a high degree of metal dispersion was attained. The synthesized catalyst was successfully used for the coupling of iodobenzene and styrene. The best reaction conditions (92–98%) were found to be using water as solvent at 90 °C and in the presence of K2CO3 as base.
The presence of doping atoms within the silica framework makes a great impact in the catalytic performance of the catalyst. Bearing in mind this concept, Stepnicka et al. [95] reported the preparation of doped MCM-41, with aluminium, caesium and potassium cations as doping agents. Pd salts were incorporated within the mesoporous through a wetness method and posterior reduction of the salt with tetrabutylammonium acetate. The performance of this catalyst was assessed using the cross-coupling reaction of bromobenzene and butyl acrylate as model reaction. The authors reported that Pd@Cs+-(Al)MCM-41 displayed a superior catalytic activity compared to related catalysts in terms of turnover number and turnover frequency [96].
Kawanami et al. [97] investigated the effect of boron as a dopant to enhance the dispersion of Pd nanoparticles. The supported catalyst was synthesized by a hydrothermal method. Briefly, CTAB, H3BO3, and a Pd salt were added sequentially to a sodium hydroxide solution. The analysis of the nanomaterial revealed that Pd particle size was dependant on the Si/B ratio, the higher the content of B the higher size of the Pd nanoparticle was. The catalytic activity of such materials was tested for the hydrogenation of nitrobenzene to aniline in supercritical CO2. The catalyst was compared with similar systems containing Al and Ga in their structure, instead of B, obtaining a faster reaction when boron was found within the silica framework [98].
By applying the same concept, Yan et al. [99] developed a series of Pd nanoparticles supported on MCM-41 for the production of biofuel γ-valerolactone (GVL) and valeric acid (VA) from biomass-derived levulinic acid (LA). The best catalytic conditions for LA hydrogenation was 5 wt.% of Pd-MCM-41, obtaining a 96.3% of GVL production at 240 °C for 10 h. In 2018, the same group [100] showed that LA can be selectively transformed into GVL and VA through Pd nanoparticles supported onto MCM-41 doped with Al (Al/MCM-41). The experiments performed with Pd/AlMCM-41 showed that the best catalytic transformation (88.5% of GVL) was achieved with 5 wt.% Pd/AlMCM-41 within 10 h at 240 °C under 60 bar. Furthermore, the catalyst could be run several times without a significant loss of its catalytic activity.
A critical step in the preparation of the Pd nanoparticles is the reduction of Pd salts. This transformation can be achieved by many reductants, including sodium borohydride, hydrazine or a flow of H2 (5%) [99]. Alternatively, grafted Pd nanoparticles can be produced by using hydride-functionalized MCM-41. The hydride was introduced to the silica surface through grafting procedure. Because of the strong reductive character of silicon hydride, Pd salts were immediately reduced to Pd(0) leading to the formation of small metal nanoparticles. Interestingly, conventional reducing agents produced metal aggregation, hence catalytic activity is clearly diminished. TEM and BET observations concluded that Pd nanoparticles were preferentially located in the outer part of the MCM-41, which is consistent with the presence of hydride groups. The obtained system was used as a catalyst for methane oxidation, obtaining better activity than Pd-SBA. Hydride groups in the latter catalytic system were preferably located in the inner part of the particle due to presence of large pores, hence reactants were less capable to reach the active sites [101].

5.2. Gold Nanoparticles Supported on MCM-41

Gold is a usual metal catalyst for the A3 multicomponent coupling of aldehydes with alkynes and amines, which leads to the formation of propargyl-amines. Typically, cationic gold species present the best catalytic activity. Unfortunately, these catalysts are quickly reduced to non-active Au species during the activation of alkynes/alkenes.
With the aim to overcome this drawback, Feiz and Bazgir et al. [91] prepared gold nanoparticles supported on thiol functionalized MCM-41 (MCM-SH). The introduction of such functional group was achieved by sequential treatment of MCM-41 with thionyl chloride and 2-mercaptoethanol. Finally, an Au salt was dispersed within the silica matrix and reduced to metallic Au using NaBH4 (Figure 47). Gold nanoparticles were immobilized within the silica matrix due to the electrostatic/covalent bond between the metal and the thiol group.
The optimal catalytic activity of these MCM-41 in the catalysis of the A3 coupling reaction was achieved using a loading of 2 mol.% Au, a temperature of at 80 °C and water as a solvent. The catalyst could be recovered and reused 5 cycles without a significant loss of activity. The proposed reaction mechanism starts with the alkyne absorption on a gold atom at the support interface, followed by its deprotonation. This intermediate reacted with an iminium cation, formed from the condensation of the aldehyde and the amine, affording the corresponding propargylamine (Figure 48).
Although it is well established that -SH groups improves the dispersion of the MNP onto the surface of the MCM-1, Yang et al. [102] demonstrated that amines display a similar effect thanks to the formation of a partial covalent bond. Au catalyst prepared using this methodology have showed highly stable and reusable capability while recycling the same sample.
Accordingly, Au nanoparticles deposited on functionalized NH2-MCM-41 have been studied as catalysis of different types of reactions. In 2018, Qu et al. [103] studied the oxidation of HCHO in Au/NH2-MCM-41. The authors reported that formaldehyde activation merely depended on the oxidation state of Au and its particle size. The results established that the best catalytic activity was obtained with Au nanoparticles of 2–4 nm, which were highly dispersed on the silica framework thanks to the amine groups. Moreover, amine groups partially inhibited the reduction of Auδ+ (electron-deficient) to Au0. The coexistence of both species along with the optimal size of the MNP (2–4 nm) provided the highest HCHO oxidation at 55 °C, whereas non-aminated MCM-41 required 200 °C due to the presence of gold aggregations.
Yang et al. [102] compared the catalytic performance of gold nanoparticles supported on aminated MCM-41 and SBA for the aerobic oxidation of 1-phenylethanol and some para-derivates. The best catalytic results were observed in the case of NH2-MCM-41 with a 2.4 wt.% of Au and a size of 1–3 nm. Interestingly, molecular sieves with a higher gold weight ratio (10.3%) produced metal nanoparticle aggregates of about 5–15 nm. Again, the formation of these aggregates reduced the oxidation activity. The authors concluded that the degree dispersion of Au is of paramount importance for the reaction yield.
Besides, Finashina et al. [104] studied the intramolecular hydroamination of 2-(2-phenylethynyl)aniline to form 2-phenylindole. As mentioned above, the presence of amine groups within the silica matrix prevents Au agglomeration leading to small metal nanoparticles (3 nm), while Au aggregates were formed in the non-aminated MCM-41. A yield of 22% was obtained for this catalyst system, whereas a 72% was achieved by the -NH2 functionalized one.
A different approach was presented in 2001 by Satry et al. [105], to obtain gold nanoparticles in amine functionalized MCM-41. The methodology relies on spontaneous reduction of chloroaurate ions within the silica matrix (Figure 49). It was hypothesised that hydroxy groups spontaneously reduce chloroaurate ions on the inner surface of the silica support. Thus, the gold nanoparticles are formed within the pores due to the interaction between the amine and the hydroxy functionality. The as-synthesized catalyst showed a good catalytic activity for the hydrogenation of a broad substrate scope. For instance, styrene was hydrogenated to ethylbenzene in a 30% of conversion and a 100% of selectivity [106].
Alternatively, bimetallic nanoparticles with Au-Pt were synthesized by simultaneous reduction of HAuCl4 and HPtCl6 by NaBH4. These metals were deposited into an amino functionalized MCM-41. The catalytic activity for the hydrogenation of aromatic nitro compounds of Au-Pt/MCN-41 was found to be superior to monometallic Au nanoparticles [107] (Figure 50).
Interestingly, Ratnasamy et al. [108] used the same catalyst for the selective oxidation of CO in the presence of an excess amount of H2. The authors reported that the best Au:Pt ratio was 5:95 decreasing the amount of CO to 0.3% at moderate temperatures (80 °C), whereas almost non-conversion was found for monometallic or bimetallic with higher rates of Au. Moreover, its activity was not even increased at higher temperatures (150–200 °C). It is worth nothing that at such conditions, the oxidation of H2 was enhanced, while the oxidation activity of CO decreased. This bimetallic catalyst was used for 16 h with no apparent deactivation.
On the other hand, Mou et al. [109] studied the CO oxidation using as catalyst Au and Ag. Catalyst introduction was successfully achieved through the dispersion of AuCl4 and AgNO3 followed by a posterior reduction with a H2 flow. Besides, amine functionalization of MCM-41 enabled to control the size of the bimetallic nanoparticles from 4–6 nm to 20 nm [110,111]. Thereby, its catalytic activity was enhanced achieving a reaction rate of 1.097 mol g Au−1·h−1 with a ratio of Au:Ag = 8:1 at 30 °C. Surprisingly, the catalyst was still active working at 79 °C after one year of storage at ambient conditions.
C. Torres et al. [112] developed a series of Au/MCM-41 catalyst modified with Ti, instead of -NH2 or -SH groups, in order to achieve a good dispersion and stabilization of gold nanoparticles within the framework. The catalytic test for hydrogenation of nitro compounds showed that modified Au-MCM-41 with Ti gave a better conversion rate than the unmodified catalyst. The improvement of the catalytic activity was ascribed to the effect of Ti on the gold nucleation and growth as the structural characterization pointed out.
Similarly, Yamashita et al. [113] used doped silica framework with Ti (Ti/Si = 0.01) for the preparation of H2O2. Again, the presence of Ti enhances metal dispersion within the MCM-41. Besides, the photocatalytic character of such dopant allows the metal nanoparticle formation through the photo assisted deposition method (Figure 51). Palladium nanoparticles were successfully deposited onto Ti/MCM-41 and its catalytic activity was tested in the production of H2O2 from H2 and O2 and compared to the typical impregnation method obtaining a concentration of 50 and 30 mM of H2O2, respectively. Notably, bimetallic Au-Pd were also deposited within a Ti modified MCM-41 through the photocatalytic deposition method achieving a higher H2O2 formation than the Pd monometallic catalyst.
Clearly, the presence of doping atoms into the silica matrix imparts a strong effect on the performance of the catalyst. Another example was given by Pescarmona et al. [114] who studied the incorporation of gold nanoparticles within Sn modified MCM-41 through impregnation method and posterior calcination. Au catalyst and Sn-MCM-41 possess two independent important roles for the selective transformation of glycerol. The first one promotes the oxidation of glycerol to trioses, whereas the solid acid Sn-MCM-41 catalyzes the rearrangement of the intermediate trioses to methyl lactate. Au/Sn-MCM-41 showed better catalytic performance than a benchmark solid acid catalyst (zeolite). The authors postulated that the best catalytic activity for the glycerol conversion was achieved with a particle size of 3 nm.

5.3. Nickel Nanoparticles Supported on MCM-41

MCM-41 supported nickel catalysts have attracted wide attention for their extensive applications in industrial processes such as hydrogenation of aromatics, production of synthesis gas, methanation of CO2, and syngas methanation [92].
Immobilization of nickel nanoparticles inside of MCM-41 (Figure 52) was achieved by using H2N-MCM-41 by Zhu et al. [92]. As stated before, amino functionalization was reported to inhibit the aggregation of MNP on the mesoporous enhancing their dispersion. The resulting catalyst was applied for the methanation of synthetic gas. The stability of the system was carefully studied since the formation of [Ni(CO)4] during the catalytic process, which deactivates the catalyst. Gratifyingly, such deactivation was not observed when using amino-functionalized MCM-41 as support, because of the low formation of the later complex. Furthermore, the presence of amine groups in the catalysts avoided metal sintering, increasing its thermal stability and hence the catalytic activity.
A different approach to avoid the aggregation and migration of metal nanoparticles to the outer part of the silica framework is predicated on the incorporation of a doping agent within the silica matrix [95].
In this regard, Williams et al. [115] employed aluminium salts in order to create acidic sites within the silica framework of the MCM-41. The resulting nanomaterial was studied for the biomass gasification of wood sawdust to produce H2. The catalyst was stable and no sintering was observed after the gasification at 800 °C. The same system was studied for the oligomerization of ethylene and compared to two typical acidic porous aluminosilicates: nanocrystalline β-zeolite and silica-doped alumina Siralox-30 by Martínez et al. [116]. The authors concluded that the production of oligomers with more than 5 carbon-chain was maximal, with a value of 12 mmol/(Kgcatalyst · s), for the case of Ni/Al-MCM-41.
In 2018, Abu-Darieh et al. [117] introduced gallium as a doping agent to develop a potential Ni/MCM-41 catalyst for methane reforming. The authors reported that Ga affected the surface area, decreasing the number of basic sites and then increasing the catalytic activity. The 2% gallium loading showed the highest methane conversion of 88%. In the same year and with the same aim, Wang et al. [118] prepared an analogous system but replacing Ga with CeO2. The introduction of CeO2 improved the stability and the reduction of nickel species inside the mesoporous support, resulting in a good stability and better catalytic activity.
Pirouzmand et al. [119] prepared Ni/MCM-41 by direct synthesis and wet impregnation methods with the aim of study the template effect in the CO2 hydration catalysis. The authors employed a calcination method to remove the template from Ni/MCM-41 and compared its catalytic activity to Ni/MCM-41 containing the template inside the pores. As a result, the template-containing catalyst displayed a higher catalytic activity, probably due to the presence of strong basic sites, which lead to a major grafting of metal nanoparticles within the silica matrix. Zheng et al. [120], studied the same effect in a catalysis for the hydrodechlorination of 1,2-dicholoethane. By contrast to the later report, these authors reported that the strong electrostatic repulsive interactions between Ni (II) and CTAB lead to aggregation of Ni nanoparticles upon calcination resulting in a low dispersion of the nickel source (Ni(NO3)2) into the mesopores of MCM-41.

5.4. Ruthenium Nanoparticles Supported on MCM-41

Supported ruthenium nanoparticles have emerged as versatile catalysts in industrial processes such as hydrogenation and oxidation reactions. Carrillo et al. [121] reported the use of an amino functionalized MCM-41 catalyst which enhance the dispersion of Ru nanoparticles inside the support by covalent interaction between the metal and the ligand. Therefore, the leaching of metal nanoparticles was precluded. The resulting materials showed high catalytic activity a tandem oxidation-Wittig olefination of different benzyl alcohols (Figure 53).
In 2016, Roy et al. [122] developed a Ru nanoparticles supported on a CeO2 doped MCM-41 for the treatment of phenol-containing wastewater. The use of metal catalyst for the removal of such pollutants represent a more environmentally friendly alternative, in contrast of the most commonly used methods such as incineration and biological treatments. The main drawbacks of these methods are the generation of noxious compounds and highly toxic effluents. On the other hand, the use of unsupported metal catalyst is not recommended because of their low reusability. For this reason, Ru nanoparticles on a silica support have been proposed to overcome this issue. Moreover, doping of the MCM-41 framework with 14 wt.% CeO2 enhanced the dispersion of the metal catalyst (93%) and inhibited its aggregation. This catalytic system was employed in the wet air oxidation of phenol achieving up to 99% of conversion and could be used up to 15 days without any leaching of the catalyst.
Another example of Ru nanoparticles supported on MCM-41was reported by Liu et al. [93]. This group synthesised a series of zirconia-modified MCM-41 supported ruthenium-lanthanum catalysts by double solvent impregnation method. The purpose of the synthesized catalyst was the hydrogenation of benzene to cyclohexene, which is a valuable intermediate for the synthesis of nylon-6 and nylon-66. The critical issue in this procedure is to avoid the over-oxidation of cyclohexene to cyclohexane. Thereby, it is relevant to promote the desorption of the previously hydrogenated cyclohexene from the catalytical system. To this end, the modification of MCM-41 with zirconia creates a thin layer of water that improves the hydrophilicity of the silicate matrix promoting the desorption of the synthesized cyclohexene. The authors have postulated that tetragonal zirconia-modified MCM-41 (Figure 54) showed higher hydrophilicity and more abundant surface hydroxyl groups than MCM-41 modified by monoclinic and amorphous zirconia. Moreover, tetragonal zirconia MCM-41 supporting ruthenium showed a better performance in the hydrogenation of benzene to cyclohexene (55.8%). It is noteworthy, that cyclohexene selectivity (82.1%) was enhanced due to the hydrophilicity introduced by tetragonal zirconia.

5.5. Titanium Nanoparticles Supported on MCM-41

TiO2 is a widely used photo-catalyst due to its lower cost, nontoxic nature and chemically inertness. Unfortunately, during the preparation process, the oxide can easily agglomerate what leads to a material with an unsatisfactory catalytic performance. To optimize the performance and inhibit the aggregation of the catalyst, the use of MCM-41 as support have been recommended.
In 2015, H. Entezari et al. [123] studied the encapsulation of TiO2 within the silica framework for the removal of sulphur compounds of liquid hydrocarbon fuels. The authors reported that the catalytic activity of supported TiO2 was higher than the unsupported ones due to long-lived photo induced charge separation, large pore size and high specific surface of MCM-41, which can adsorb some bulky species. However, TiO2 exhibits low catalytic activity under visible light conditions because it has a large band gap energy of 3.2 eV corresponding to a threshold wavelength of 388 nm. In order to overcome such issue, the doping of the silica framework with non-metallic elements such as carbon, sulphur and nitrogen have become an interesting strategy for improving the visible light activity of the catalyst. Among them, the authors selected carbon as doping agent of MCM-41 to enhance visible light activity of TiO2/MCM-41. Thereby, carbon modified TiO2/MCM-41 catalyst was used for the oxidative desulfurization of dibenzothiophene under visible light and mild conditions within 5 h achieving a photocatalytic degradation of 95.6% for DBT with Si/Ti = 5 ratio. Moreover, the catalyst system could be reused at least two times without a significant loss of catalytic activity.
In 2017, Yamashita et al. [124], prepared TiO2/MCM-41 catalyst with alkali metal cations embebed (Li+, Na+, K+, Rb+, Cs+) within its pores. In this regard, a previous post-modification of the MCM-41 with carboxylic acid groups was needed. Therefore, alkali cations were successfully anchored onto the silica matrix by an ion-exchange process. This catalytic system was assessed for the photocatalytic degradation of phenol in water under UV light irradiation (Figure 55). It was concluded that the strong interaction between the aromatic ring and the alkali metal cations, was responsible for the high catalytic activity observed.
Pescarmona et al. [125] studied the functionalization of Ti/MCM-41 with methyl groups in order to increase the hydrophobicity of the support, which minimize the presence of water within the pores and stabilize the MNP. Besides, the diffusion of non-polar alkenes through the pores is favored. Ti/MCM-41 methyl functionalized was satisfactory synthesized by hydrothermal method and was tested for the epoxidation of cyclohexene obtaining a higher TOF in comparison with the unmodified Ti/MCM-41. On the other hand, a complete regeneration of the catalyst through simple washing was not achieved. It was reported that a gradual deposition of the product on the catalyst inhibits the conversion of cyclohexene.

6. Magnetic Core Shell Nanoparticles

As stated before, MCM-41 are endowed with ideal features as supports in catalysis because of the inherent stability of SiO2 and the ordered mesoporous structure, which allow a high loading of the catalyst [126]. However, the inefficiency of separation methods such as centrifugation and filtration, hamper the sustainability and economy of their use in bulk organic transformation catalysis. To overcome this drawback, magnetic nanoparticles (MNPs) have emerged as a powerful solution to optimize the operations of separation and reuse of solid catalysts by the simple application of an external magnetic field [127,128,129].
Magnetic MCM-41 consist of magnetic particles located in the core or within/onto the silica framework. The most common magnetic materials used are Fe3O4 (magnetite), γ-Fe2O3 (maghemite), α-Fe2O3 (hematite), MnFe2O4, and CoFe2O4 since these compounds are easy to synthesize with size-monodisperse particles and with high saturation magnetization. Among these, monometallic iron particles are arguably the most studied systems [130,131].

6.1. General Synthesis of Magnetic MCM-41

Several methods to synthesize iron-oxide-based nanomaterials have been developed [132]. These methods include micro-emulsion [133,134], thermal decomposition of iron complex in an organic solvent [135], co-precipitation from the Fe2+/Fe3+ mixed salt solution in an alkaline medium [136,137], hydrothermal synthesis [138,139] and the use of successive redox reactions in solvothermal conditions under microwave irradiation [140].
A typical procedure to prepare magnetic particles in the mesoporous material is based on the deposition of magnetic nanoparticles on the external surface of MCM-41 by thermolysis of Fe(NO3)3 [141]. Similar approaches rely on the wet impregnation and chemical vapor deposition. Arruebo et al. [142] used wet impregnation of iron precursor followed by repeated cycles of oxidation and reduction to obtain magnetic MCM-41, while Zhang et al. [143] employed chemical vapor deposition to prepare iron oxide nanoclusters in MCM-41 from volatile Fe(CO)5. An alternate common approach is to graft magnetic nanoparticles on the nanoparticles. Hyeon et al. [144] deposits molybdenum oxide catalyst on the mesoporous silica shell coated on dense silica-encapsulated magnetic iron oxide nanoparticles. Other works reported single-pot synthesis of magnetic mesoporous materials by adding iron (III) ethoxide during the synthesis of mesoporous silica and chemically transform it to superparamagnetic γ-Fe2O3 particles [145,146,147]. Chen et al. [148] reported a simple synthesis method for embedding the magnetic iron oxide nanoparticles inside MCM-41. According to these authors, magnetic mesoporous molecular sieves can be prepared by adding the magnetic colloid to an ammonia solution containing CTAB as a template. The silica precursor, tetraethyl orthosilicate (TEOS) was then added and the mixture was allowed to react at room temperature under well-mixed conditions.
In attempt to improve the preparation of magnetic MCM-41, Iamamoto et al. disclosed a new method which relies on the coating of Fe3O4 nanoparticles with a thin silica layer before preparation of the mesostructured framework [149]. The catalyst preparation procedure is depicted in Figure 56. First, Fe3O4 particles were coated with SiO2 resulting in the formation of the silica-Fe3O4 composites with a nonporous silica layer of 5 nm of thickness (denoted as Fe3O4@nSiO2). Then, a MCM-41-type ordered mesoporous silica structure was grown on the support to yield the Fe3O4@nSiO2@MCM-41 system. At this step, mesitylene could also be incorporated as a structure expanding agent for the production of larger pores.

6.2. Magnetic MCM-41 for Basic and Acid Catalysis

The silanol component permits grafting of a wide variety of functional groups to the molecular sieves, which can be used to develop a basic catalyst. A typical modified organic–inorganic hybrid material that have been applied as effective solid base catalyst is 3-aminopropylated silica (MCM-41–nPrNH2) (cf. 3.1.1).In this regard, magnetic MCM-41 has been functionalized with the amino moiety by the post-modification method from the reaction of MCM-41 with APTES. Fe2O3-MCM-41-nPrNH2 has found to be an efficient catalyst for the synthesis of 2,4-diphenylpyrido[4,3-d]pyrimidines through the two component reaction between (E)-3,5-bis-(benzylidene)-4-piperidones and benzamidine hydrochloride under solvent free conditions by Shadjou and Hasanzadeh [150]. Using 0.04 g α(Fe2O3)–MCM-41–nPrNH2 in solvent free conditions at temperature of 130 °C during 1:20 h the desired product was furnished in 98% yield. A new series of phenylpyrido[4,3-d]pyrimidin-2-amino derivatives have been also synthesized in the presence of amino modified magnetic MCM-41 from the reaction of (E)-3,5-bis(benzylidene)-4-piperidones and guanidine carbonate [151]. The efficiency of the catalysts was due to the synergetic effect of Brønsted basicity of –NH2 and Lewis acidity of the Fe3+ groups. Mild reaction conditions, good to high yields, simple work-up and ease of recovery and reusability of the catalyst, make this MCM-41 system attractive and useful for the formation of these cyclic compounds.
Silica supported sulfonic acid has received considerable attention as an inexpensive, nontoxic and recyclable catalyst for numerous organic transformations [14] (cf. 3.1.2). Among them, magnetite MCM-41 (Fe3O4@MCM-41-OSO3H) has been widely used because of its paramagnetic character combined with its structurally well-ordered channels and high external surface area. In this regard, Khorshidi et al. [152] studied the performance of this sulfonated catalyst for the three component reaction of indoles, aldehydes and thiols to furnish the desired 3-[(aryl)(arylthio)methyl]-1H-indoles. After obtaining the optimum reaction conditions, the scope of the reaction was investigated affording the products in yields ranging from 25 to 80%.
Rostamizadeh et al. [153] reported the preparation of the maghemite (α-Fe2O3)-MCM-41-SO3H for the synthesis of N-aryl-2-amino-1,6-naphthyridine derivatives. The magnetic catalyst was prepared directly through the reaction of chlorosulfonic acid with silica-coated sieves (α-Fe2O3)-MCM-41. Because of the low stability of Fe3O4 nanoparticles in the presence of chlorosulfonic acid, the desired catalyst was not formed and a non-magnetic catalyst was obtained instead (Figure 57). Therefore, (Fe3O4)-MCM-41 was calcined at 450 °C and converted to its stable form (α-Fe2O3)-MCM-41. The catalyst showed high efficiencies due to the synergetic effect of Brønsted acidity of -SO3H and Lewis acidity of the Fe3+ groups.
Alternatively, Shariati and Golshekan et al. [154] studied the decoration of MCM-41 with sulfonic acid in two steps by the oxidation of mesoporous structure with thiol groups on the surface. The co-condensation of Fe3O4 and TEOS with terminal trialkoxy organosilanes incorporating the thiol group ((RO)3Si(CH2)2SH) in the presence of structure-directing agent (CTAB) leads to the formation of Fe3O4@MCM-41 with thiol residues anchored covalently onto the silica network (Fe3O4@MCM-41-SH). The oxidation of thiol groups rendered the final sulfonic acid functionalized system (Fe3O4@MCM-41-SO3H) (Figure 58). The applicability of the nanocomposite was assessed at room temperature for the preparation of 12-aryl-benzo[α]-xanthenonederivatives by the one-pot three-component reaction of various arylaldehydes, β-naphthol and 1,3-cyclohexadione. The study showed the unique advantages of these molecular sieves, such as their high magnetic properties, high density of sulfonic acid functional groups of the nanocomposite surface, simple synthesis of the catalyst in solvent-free conditions and room temperature, excellent yields, shorter reaction time, activity as stereoselective catalyst and easy separation.
Cai et al. [155] reported a new magnetic mesoporous solid acid catalyst containing NiFe2O nanoparticles as the magnetic material (SO3H-MCM-41@NiFe2O4). The system was prepared by impregnation-pyrolysis decoration method (Figure 59) and its catalytic activity was compared to SO3H-MCM-41@Fe2O3 synthetized by one-step sol–gel route. The magnetic mesoporous systems were, afterwards, functionalized with the thiol groups by grafting and then oxidised using H2O2. The catalytic performance of these two catalysts was evaluated using a three-component condensation of aromatic aldehydes, 2-naphthol and benzamides to give the corresponding 1-amidoalkyl-2-naphthols. The NiFe2O4 system exhibits better catalytic activity (92% of yield in 4 h) than SO3H-MCM-41@Fe2O3 nanoparticle analogue (70% of yield in 7 h).

6.3. Magnetic MCM-41 to Embed Ionic Liquids

In the field of catalysis, ionic liquids (ILs) play an important role as solvents and organocatalysts [156]. Unfortunately, their high cost, difficulty in products separation and recyclability hamper their further application at industrial scale. On the other hand, ILs are highly viscous, which often limits mass transfer during catalytic reactions. To overcome these short comings, the immobilization of ILs onto the solid supports has become a green approach for their catalytic applications [157,158].
The use of supported Fe3O4@MCM-41-SO3H in the presence of the Brønsted acidic ionic liquid 1-methylimidazolium hydrogen sulphate [HMim][HSO4] has been evaluated [159]. Three types of MCM-41 mesoporous magnetite catalyst, Fe3O4@MCM-41-(CH2)3-SO3H, Fe3O4@MCM-41-SO3H and Fe3O4@MCM-41-SO3H@[HMIm][HSO4] were compared. The catalyst performance was studied for the preparation of spiro[benzo[5,6]chromeno[2,3-d]pyrimidin-indolines] by using a variety of structurally diverse α or β-naphtols, barbituric acid and isatin derivatives under solvent-free conditions. The reaction test showed a satisfactory catalytic performance when the reaction was carried out in the presence of Fe3O4@MCM-41-SO3H@[HMIm][HSO4], achieving high yields (90–95%) and good reaction rates (15–30 min). Furthermore, it can be easily separated from the reaction medium with a permanent magnet and directly reused without any deactivation even after five rounds of application.
Dual acidic ionic liquid (DAIL) (1-Methyl-3-(4-sulfobutyl) imidazolium hydrogen sulfate) has been also supported on magnetic MCM-41 [160]. The system has been successfully used for one-pot three-component synthesis of some new and known 2-amino-4H-chromenes in aqueous conditions (Figure 60). The catalytic activity of (α-Fe2O3)-MCM-41-DAIL) was compared with reported catalysts and its building components, showing that the use of DAIL, (α-Fe2O3)-MCM-41 or MCM-41 alone as catalyst was not sufficient to run the reaction to completion. It is hypothesized that the novel catalyst was more efficient due to the Brønsted dual acidic property of the IL together with the Lewis acidity of ferric oxide and the high surface area of MCM-41. The study also evaluates the scope and generality of the reaction showing a marked efficiency for all the substrates tested.

6.4. Magnetic MCM-41 Decorated with Metallic Complexes

Encapsulation of metalloporphyrins in porous compounds protects them from harsh reaction condition and also results in catalysts that mimic the features of natural enzymes (cf. 3.2). MCM-41, with its large pore size and a high amount of silanol groups on the surface of its channels, acts as an attractive candidate for encapsulate metalloporphyrins via physical adsorption or covalent linkage [161].
Zanardi et al. [150] reported the immobilization of manganese porphyrin (MnP) on Fe3O4@SiO2@MCM-41 as a catalyst for hydrocarbon oxidation in the presence of iodosylbenzene (PhIO). The authors synthetized the catalyst with two different pore sizes, the Fe3O4@nSiO2@MCM-41-MnP and Fe3O4@nSiO2@MCM-41(E)-MnP (expanded porous structure), using mesitylene as an expanding agent for the production of larger pores. The synthetic method for the preparation of the catalyst is illustrated in Figure 56. Magnetite sieves were first coated with a thin silica layer (Fe3O4@nSiO2) before the preparation of the mesostructured composites. The MCM-41 framework was functionalized with APTES to enable the covalent immobilization of Mn(III)-5,10,15,20-tetrakis(pentafluorophenyl) porphyrin onto the MCM-41 via an aromatic nucleophilic substitution reaction. The oxidization of (Z)-cyclooctene and cyclohexane was performed to assess the catalytic activity of the composites and to evaluate the accessibility of substrate and oxidant molecules to the catalytic sites. The study concluded that Fe3O4@nSiO2@MCM-41(E)-MnP provided higher yields than Fe3O4@nSiO2@MCM-41-MnP (93% vs. 77%, respectively). This result can be attributed to the greater textural mesoporosity volume of Fe3O4@nSiO2@MCM-41(E)-MnP, which increases diffusional processes that facilitate the access to the region of the framework-confined mesoporosity. The recycling studies demonstrated that Fe3O4@nSiO2@MCM-41(E)-MnP catalyst was also more stable than Fe3O4@nSiO2@MCM-41-MnP. Both heterogeneous catalysts were selective for the oxidation of cyclohexane to cyclohexanol and no cyclohexanone formation was observed. The higher selectivity for the alcohol product in comparison to homogeneous MnP indicates a P450-type biomimetic behavior favored by the inorganic support.
MnP was also bonded to MCM-41 using imidazole as axial ligation (Fe3O4@MCM-41-Im@MnPor) [162]. Fe3O4@MCM-41 were functionalized with 3-chloropropyltriethoxysilane and then with imidazole (Figure 61). The catalytic performance was studied in the epoxidation of different olefins varying catalyst amounts, oxygen donors and reaction media. The optimum reaction conditions were reached using 100 mg of catalyst, acetonitrile/water mixture as solvent in the presence of NaIO4 as the oxygen source. The activity and selectivity of the heterogeneous catalyst was comparable to those of its homogeneous analogue (MnPor). The reaction of different cyclic and linear alkenes was also investigated to examine the scope of the method. Using the optimized reaction conditions, the epoxide was afforded with high selectivity and good yields, ranging between 50 and 98%. Moreover, the novel catalyst was compared to a system reported earlier, which contains the manganese porphyrin complex supported on amorphous silica (Fe3O4@SiO2@MnTPPCl). This study provides evidence of the matrix effect on catalyst performance, which leads to the separation of active catalytic sites and thereby increases the catalytic activity.
Rayati et al. [163] immobilize a molybdenum Schiff base complex (dioxomolybdenum(VI) complex) onto magnetite sieves by an hydrogen bonding ((α-Fe2O3)–MCM-41–Mo(O)2L). The results showed an efficient and highly selective oxidation of a wide range of olefins using hydrogen peroxide and tert-butyl hydroperoxide as oxidizing agents. Complete selectivity for epoxide formation was obtained in the case of cyclooctene, 1-octene, indene, and 1-methyl cyclohexene, while the oxidation of styrene substrates led to the formation of the corresponding substituted benzaldehyde.
Nunes and Vaz et al. [164] reported a novel mesoporous nanocomposite material comprising helical chiral channels, with magnetic iron nanoparticles attached to the silica MCM-41 framework, as support for a molybdenum complex. The helical mesoporous magnetic nanocomposite (magMSh) was prepared (Figure 62) using an entropy-driven procedure with achiral cationic surfactant template (CTAB) and a co-surfactant (ammonia) in the presence of silica-coated magnetic MCM-41 that were co-condensed with TEOS. Afterwards, a bipyridine derivative was grafted to the inner silanol surface and used as ligand to coordinate a Mo(II) precursor complex [MoI2(CO)3(CH3CN)2]. The derivatized chiral nanocomposite (magMSh-bpy-Mo) was used as an efficient and reusable catalyst precursor for the asymmetric epoxidation of olefins using tert-butyl hydroperoxide as oxygen donor. The catalysis of cis-cyclooctene, R-(+)-Limonene and trans-hex-2-en-1-ol yielded the epoxide without any formation of by-products and conversions raging between 98 and 100% when toluene was used as solvent at 353 K or 383 K. The authors suggested that the good stereoselectivity arises from the confinement effect imposed by the helical rigid chiral channels. In addition, data obtained from the selective adsorption of enantiopure d- and l-phenylalanine supported the presence of channel chirality and proved the utility of this novel magnetic system for chiral recognition applications.

6.5. Magnetic MCM-41 to Incorporate Metal Nanoparticles

Homogeneous metal catalysts have been extensively studied in synthetic organic chemistry because of their high selectivity and catalytic activity. Unfortunately, such catalysts are not endowed with high reusability and their use leads metal contamination on the final products (cf. 0). In attempt to solve these shortcomings, a great number of metal-based catalytic species have been immobilized on different heterogeneous supports containing superparamagnetic materials. MCM-41 functionalized with organic moieties can be used as support of metal nanoparticles increasing their dispersion and catalytic performance.
A Pd Schiff base complex immobilized onto the surface of magnetic MCM-41 (Fe3O4@MCM-41@Pd(0)-P2C) has been designed as a novel reusable magnetic catalyst [165]. The catalyst was prepared by modification of Fe3O4@MCM-41@nPr-NH2 using pyrrole-2-carbaldehyde in ethanol and heating under a N2 atmosphere. The metal was introduced into the pores from a solution of Pd(OAc)2in ethanol, to finally reduce Pd(II) using NaBH4. The obtained palladium system was successfully used in Suzuki–Miyaura (88–98% yield) and Heck (80–96% yield) reactions, employing polyethylene glycol (PEG) as a green solvent. The broad substrate scope, simple experimental procedure and catalyst separation, utilization of an inexpensive and readily available catalyst with excellent reusability, non-chromatographic purification of the products, simple recrystallization from diethyl ether and the use of green solvent were the merits of this novel catalyst.
The same research group also reported a Pd(0)-S-propyl-2-aminobenzothioate complex immobilized onto functionalized magnetic MCM-41(Fe3O4@MCM-41@Pd-SPATB) as an efficient and recyclable nano-organometallic catalyst for Suzuki, Stille and Heck reactions [166]. For the preparation of Fe3O4@MCM-41@Pd-SPATB (Figure 63), Fe3O4@MCM-41 was functionalized with (3-mercaptopropyl)trimethoxysilane (MPTMS) and, afterwards, with S-propyl-2-aminobenzothioate, providing the denoted Fe3O4@MCM-41@SPATB. Finally, Pd(0) was introduced by impregnation and reduction. The protocol was carried out also in PEG-400 as green solvent achieving good to excellent yields in short reaction time for all the reactions tested.
Jiang et al. [167] reported the catalytic hydrogenation of aqueous bromate using a catalyst that contains Pd nanoparticles supported on amino functionalized magnetic MCM-41. The heterogeneous system was synthesized using two different methods, deposition-precipitation (DP) and impregnation (IMP). The catalysts were functionalized with different loading amounts of –NH2 to evaluate the impact on Pd nanoparticle dispersion and the catalytic behavior. The study of the liquid phase catalytic hydrogenation of bromate established that DP catalyst exhibited a better activity than IMP as a result of improved Pd dispersion. The authors concluded that it was feasible to improve hydrogenation efficiency (100% of conversion) by applying DP catalyst prepared using 0.25 APTES:magMCM-41 ratio. Noteworthy, a decline in activity was observed with a further increase of NH2 loading. The optimized catalyst could be reused five times without significant loss of activity (<6%).
Taking into account the soft interaction between Au and S and the resulting semi-covalent bond, magnetic MCM-41 have been functionalized with thiol groups to incorporate Au nanoparticles [168]. Firstly, magnetic sieves were embedded in MCM-41 framework and subsequently functionalized with (3-mercaptopropyl)trimethoxysilane (MPTMS) (Figure 64). Then, HAuCl4 was deposited into (α-Fe2O3)-MCM-41-SH and reduced to create fine Au nanoparticles bonded to thiol groups. The optimization of the reaction indicated that the best conditions were achieved using H2O as solvent at 80 °C and solvent-free at 70 °C gave the best yields (98% and 100%, respectively). The hydration of a range of alkynes and aliphatic alkyne was investigated to ascertain the scope of the methodology (90–100% of yield). In order to extend the applicability of this catalyst, the authors also reported its catalytic performance in Ullmann homocoupling of aryl iodides to produce symmetrical biaryls [169]. Short reaction time, high selectivity, no observation of undesired side products and utilization of green and magnetically separable catalyst were the advantages of this procedure. This protocol produces very finely Au nanoparticles distribution on the pore surface resulting in excellent catalytic behavior.
The incorporation of metal on the magnetic catalyst can be also accomplished adding a silicon precursor that contains the metal source. In this sense, it was reported the synthesis of Fe2O3/MCM-41 silicoaluminate magnetic catalyst prepared employing a modified assembling method that introduces clinoplitolite as aluminium natural source [170]. The Fe2O3/MCM-41 catalyst was prepared coating Fe2O3 nanoparticles with a layer of SiO2, and then synthetizing the mesoporous framework by using natural clinoptilolite. The catalyst was employed for hydrothermal liquefaction (HTL) of microalgae and deoxygenation upgrading of derived biocrude and its model compounds under the absence of hydrogen. Chlorella, palmitic acid and methyl palmitate were chosen as reactants. The magnetic Fe2O3/MCM-41 catalyst showed good performance; in the 320–350 °C regime and under subcritical water, palmitic acid conversion was improved by 14–29%. Methyl palmitate conversion was 56% and decarboxylation selectivity to pentadecane was improved to 62% carrying on the deoxygenation at 342 °C.
A strategy to assemble superparamagnetic iron oxide and platinum nanoparticles onto mesoporous MCM-41 silica host with a high degree of spatial precision has been reported by Santamaria and Martinez et al. [171]. The system was prepared by a size-exclusion strategy that allows a highly precise segregation of iron nanoparticles outside the MCM-41 and platinum inside the pores. The direct assembly of magnetic nanoparticles onto the outer surface of silica framework was carried out using nanoparticles with a size larger than the MCM-41 pores and grafting covalently through a peptide-like bonding onto their external surface. Magnetic functionalized MCM-41 with carboxylic groups on the surface (DMSA−FexOy) were obtained via a ligand-exchange reaction of triethylene glycol by meso-2,3-dimercaptosuccinic acid (DMSA). These carboxylic groups were suitable for the later condensation reaction with the MCM-41 amino silica groups to anchor the magnetic nanoparticles through a peptide-like link. This procedure endows a strong superparamagnetic response and overcome the drawbacks associated with the encapsulation of magnetic cores within the mesoporous silica shell (formation of irregular core/shell structures and low magnetization saturation values due to poor magnetic cores encapsulated). Furthermore, this protocol allows to preserve a regular porous structure capable to host other functional guest species.
Secondly, the preformed Pt nanoparticles (1.5 nm) were “electrostatically pumped” within the channels of the magnetic amino-functionalized MCM-41 to confer a catalytic functionality (Figure 65). The Pt nanoparticles were stabilized with tetrakis(hydroxymethyl)phosphonium chloride (THPC) (THPC−Pt NPs) proving a high negative surface charge (−30 mV) and strong adhesion to the positively charged NH3+ mesoporous silica pores (27 mV). A sufficient high repulsion among NPs provided the “pumping” effect and propelled them within the channels. Hence, the penetration depth of the Pt-NPs can be explained as a result of the interplay between the particle−wall electrostatic attraction and the repulsive forces between neighbouring Pt-NPs.
The catalytic performance of this multi assembled nanoplatform has been evaluated in the selective hydrogenation of p-nitrophenol (4-NP) by NaBH4 as a model reaction at room temperature. The reaction has been monitored along the time measuring the 4-NP absorption peak at 400 nm. The results proved the efficiency of the catalyst, reaching a k-value comparable to the best catalyst reported in the literature for 4-NP hydrogenation. This multifunctional hybrid catalyst performed an excellent catalytic activity and could be efficiently reused during 4 cycles with negligible catalyst loss.

6.6. Magnetic MCM-41 to Immobilize Enzymes

Enzyme immobilization onto MCM-41 supports improves thermal and pH stability and enables easier separation, recyclability and reutilization of the enzymes (cf. 3.3). Embedding MCM-41 with magnetic nanoparticles enhances fast separation by applying a magnetic field and allows bulk operation and enzyme reusability.
Jing et al. [172] immobilized porcine pancrease lipase (PPL) onto magnetic MCM-41 and compared three different attachment methods. For the support preparation, Fe3O4 nanoparticles were synthetized using the solvothermal method and then coated with mesoporous silica. The enzyme PPL was immobilized by covalent attachment, physical absorption and cross-linking methods (Figure 66). For the covalent bonding, the requisite support was prepared by grafting of MCM-41 with 3-chloropropyltriethoxysilane (CPS or CPTMS). The as-synthetized support can react covalently with biological macromolecules containing amino groups. The physical adsorption was carried out without MCM-41 surface modification, while for the cross-linking method, the mesoporous nanoparticle was grafted with APTES and treated with the PPL solution using glutaraldehyde as a coupling reagent.
PPL was immobilized onto the three mesoporous magnetic sieves and the as-synthesized catalysts were studied. Enzyme immobilization was determined analysing by Bradford method the PPL supernatant. Optimal concentration of the enzyme, reaction temperature and time of incubation were found to be 0.30 mg·mL−1, 30 °C, and 4 h for all these three materials. The results showed that covalent attachment exhibit the best immobilization efficiency (maximum 96%) and was not sensitive to an increase in temperature or enzyme concentration. The cross-linked PPL nanoparticle also exhibited an enhanced immobilization when compared with the PPL attached by physical absorption. Activity of the different immobilized PPL and non-immobilized one was determined using olive oil as substrate. The activity performance, stability and reusability studies confirm the same trends. Maximum relative activity (up to 96%), high stability and reusability (83% 56 days and 86.7% ten cycles) have been achieved using the covalent bonding method when the enzymatic reaction was carried out at optimum temperature and pH (35 °C and 7.5, respectively). Based on the results obtained, chemically covalent attachment to the enzyme without a cross-linking agent instead of physical adsorption seems to be an appropriate immobilization method.
In this regard, Candida rugosa lipase was bound to Fe3O4@MCM-41 nanocomposite by using glutaraldehyde as a cross-linking specie [173]. To prepare the catalyst, MCM-41 was derivatized with APTES and the lipase was immobilized on the magnetic nanocomposite by cross-linking with glutaraldehyde (Figure 67). The enzymatic catalyst was used for the interesterification of lard and soybean oil mixtures in aim to modify the physicochemical properties of vegetable oils in the food industry because the formation of trans fatty acids is avoided. The immobilized lipase displayed a better catalytic activity towards the interesterification reaction and no trans fatty acids were detected in the interesterified blends. The enzymatic catalyst displayed good magnetic properties and could be facilely separated by applying an external magnetic field and reused for four times without significant loss of catalytic efficiency.
Ates et al. developed magnetic Fe3O4@MCM-41 functionalized with CPTMS for the immobilization of L-asparaginase (L-ASNase) [174]. To this purpose, CPTMS was used as a surface-modifying agent for covalent binding of L-ASNase on the magnetic sieves (Figure 68). The immobilization efficiency onto Fe3O4@MCM-41-Cl was as high as 63%. The catalytic activity, reusability, storage, pH, and thermal stabilities of the immobilized L-ASNase were investigated and compared with the free enzyme. It was found that both L-ASNase forms exhibited the highest activity at pH 8.5, however the immobilized enzyme retained over 85% activity in a broad pH range. The residual activity of soluble L-ASNase at temperatures higher than 60 °C was negligible, while immobilized one retained below 80% activity. It was evidenced that immobilized L-ASNase was more stable towards pH and temperature compared with soluble enzyme. The results in thermal stability showed that at catalytic activity was maintained after 180 min at 55 °C (69.7%) while soluble enzyme was totally inactivated. Furthermore, could be reused for 18 cycles retaining 42.2% of its original activity.
In another attempt to immobilize L-ASNase, the authors functionalized Fe3O4@MCM-41 magnetic sieves employing (3-glycidyloxypropyl) trimethoxysilane (GPTMS) [175] and 3-mercaptopropyltrimethoxysilane (MPTMS) [176] as the grafting agents. The generated epoxy (Figure 69) and thiol functionalities were used to covalently bind the L-ASNase. The enzyme activity at different pH and temperatures were determined measuring the concentration of ammonia produced during catalysis of L-asparagine. The two immobilized L-ASNase had greater activity compared to free enzyme, especially at high pH and temperature values. Regarding GPTMS functionalization, the highest relative activity for the native and immobilized L-ASNase was observed at pH 8.5 and 9.0, respectively. The temperature conditions analysis showed that at pH 8.6 the optimum activity was at 50 °C for soluble enzyme and 55 °C for the immobilized. The authors also studied the thermal stability at 55 °C and found that the immobilization had a great effect in denaturalization. The immobilized L-ASNase maintained more than 92% of its initial activity after incubation for 3 h while the native enzyme was deactivated due to denaturation. Moreover, immobilized L-ASNase showed higher kinetic values, storage stability and an excellent reusability, retaining 93.9% of their initial activity after 5 successive cycles and 65% after 10 cycles.
The properties of thiol-functionalized Fe3O4@MCM-41 magnetic sieves, including pH, temperature, kinetic values, thermal stability, reusability, and storage stability were also investigated and compared to native enzyme. The results showed that pH and temperature stability of immobilized enzyme were also significantly higher than those of free enzyme, obtaining a faster decrease on stability. For instance, the free L-ASNase still retained 73% of its initial activity, whilst the immobilized L-ASNase retained 81% after 3 h incubation. Besides, at 50 °C native enzyme lost 70% of its initial activity and immobilized enzyme still retained 94% after 3h. Noteworthy, the immobilized enzyme kept 63% of its original activity after 16 times of reuse. The study confirmed that functionalized Fe3O4@MCM-41 sieves have high efficiency for L-ASNase immobilization and could be reused several cycles maintaining a good performance.

7. Concluding Remarks

In this review article, the main types of functionalized MCM-41 for the immobilization of catalysts have been presented. The catalytic systems discussed include: mono and di-functionalized MCM-41 as basic and acid catalysts, catalysts based on metallic complexes supported by MCM-41, metallic nanoparticles embed onto functionalized MCM-41 and magnetic MCM-41 for catalytic purposes.
Similarly to other supports, such as SBA-15, MCM-41 are cheap, synthetically versatile, and can be reused. However, an advantage over other systems is that MCM-41 have been extensively studied due to their applications in biological applications. The chemical versatility of MCM-41 relies not only on the easy preparation of the molecular sieves, but on the straightforward regioselective functionalization, both within the inner walls of the pores and the outside surface of the particle. The functionalization of the particles can be accomplished either by grafting or post-functionalization.
Both methodologies of functionalization allow the preparation amino and acid derivatized particles or even mixed functionalizations NH2/SO3H. Many important reactions can benefit of these catalyts, such as aldol condensation, Knoevenagel to name some. Furthermore, the functionalization of the MCM-41 opens the door to the preparation of more sophisticate systems incorporating metallic complexes, which catalyze palladium cross-coupling reactions (e.g., Sonogashira, Tsuji-Trost reactions), cicloadditions, and hydroformylations among others.
Functionalization not only enables the decoration of the MCM-41 with a wide variety of functional groups or complexes but improves the dispersion of metallic nanoparticles absorbed onto the surface of the MCM-41, minimizing the aggregation of the metallic fragments and avoiding their migration. The obtained hybrid catalysts exhibited excellent catalytic activity and reusability, as a result of a better embedment of the metal catalyst, in comparison with traditional physical entrapment methods. Following this strategy metal nanoparticles such as Pd, Au, Ni, Ru, and Ti nanoparticles could be efficiently deposited on MCM-41 support. Besides, the presence of basic or acid sites, by doping the support with metal cations, enhance the dispersion of the metal nanoparticle within the silica matrix.
The incorporation of a magnetic particle in the core of the MCM-41 give rise a novel family of catalytic systems characterized by a high recovery and reusability. These features make magnetic MCM-41 ideal supports for industrial applications or solid phase synthesis.
In our opinion, the present challenges in the field of MCM-41 in catalysis are twofold: the scale up of the synthesis and the design of “click” methodologies for the decoration of the molecular sieves. In order to be implemented in industrial processes, large amount of homogeneous MCM-41 must be produced to satisfy the needs of the demand. Moreover, there is still a need for straightforward methods of functionalization to provide the optimum catalyst avoiding tedious and long synthetic procedures.
To conclude, the available methodologies to prepare and characterize molecular sieves are mature enough to allow specialists in catalysis to improve the behavior of their systems by anchor them onto MCM-41. This review has shown the main results in the field that give support to this statement.

Funding

G.M.-E. and A.M.d.R. are grateful to IQS for financial support. We would like to thank Generalitat de Catalunya for the consolidated Research Group Grant 2017SGR 01559 to GEMAT.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Llinas, M.C.; Sánchez, D. Nanopartículas de sílice: Preparación y aplicaciones en biomedicina. Affinidad 2014, LXXI, 20–31. [Google Scholar]
  2. Huang, Y. Functionalization of mesoporous silica nanoparticles and their applications in organo-, metallic and organometallic catalysis. Ph.D. Thesis, Iowa State University, Ames, IA, USA, 2009. [Google Scholar]
  3. Popat, A.; Hartono, S.B.; Stahr, F.; Liu, J.; Qiao, S.Z.; Qing Lu, G. Mesoporous silica nanoparticles for bioadsorption, enzyme immobilisation, and delivery carriers. Nanoscale 2011, 3, 2801–2818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Sun, X. Mesoporous Silica Nanoparticles for Aplications in Drug Delivery and Catalysis. Ph.D. Thesis, Iowa State University, Ames, IA, USA, 2012. [Google Scholar]
  5. Hoffmann, F.; Cornelius, M.; Morell, J.; Fröba, M. Silica-based mesoporous organic-inorganic hybrid materials. Angew. Chem. Int. Ed. 2006, 45, 3216–3251. [Google Scholar] [CrossRef] [PubMed]
  6. Slowing, I.I.; Vivero-Escoto, J.L.; Trewyn, B.G.; Lin, V.S.Y. Mesoporous silica nanoparticles: Structural design and applications. J. Mater. Chem. 2010, 20, 7924–7937. [Google Scholar] [CrossRef]
  7. Mirza-Aghayan, M.; Nazmdeh, S.; Boukherroub, R.; Rahimifard, M.; Tarlani, A.A.; Abolghasemi-Malakshah, M. Convenient and efficient one-pot method for the synthesis of 2-amino-tetrahydro-4H-chromenes and 2-amino-4H-benzo[h]-chromenes using catalytic amount of amino-functionalized MCM-41 in aqueous media. Synth. Commun. 2013, 43, 1499–1507. [Google Scholar] [CrossRef]
  8. Nale, D.B.; Rana, S.; Parida, K.; Bhanage, B.M. Amine functionalized MCM-41: An efficient heterogeneous recyclable catalyst for the synthesis of quinazoline-2,4(1H,3H)-diones from carbon dioxide and 2-aminobenzonitriles in water. Catal. Sci. Technol. 2014, 4, 1608–1614. [Google Scholar] [CrossRef]
  9. Nale, D.B.; Rana, S.; Parida, K.; Bhanage, B.M. Amine functionalized MCM-41 as a green, efficient, and heterogeneous catalyst for the regioselective synthesis of 5-aryl-2-oxazolidinones, from CO2 and aziridines. Appl. Catal. A Gen. 2014, 469, 340–349. [Google Scholar] [CrossRef]
  10. Choudary, B.M.; Kantam, M.L.; Sreekanth, P.; Bandopadhyay, T.; Figueras, F.; Tuel, A. Knoevenagel and aldol condensations catalyzed by a new diamino-functionalized mesoporous material. J. Mol. Catal. A Chem. 1999, 142, 361–365. [Google Scholar] [CrossRef]
  11. Wu, N.; Li, B.; Liu, J.; Zuo, S.; Zhao, Y. Preparation and catalytic performance of a novel highly dispersed bifunctional catalyst Pt@Fe-MCM-41. RSC Adv. 2016, 6, 13461–13468. [Google Scholar] [CrossRef]
  12. Borodina, E.; Karpov, S.I.; Selemenev, V.F.; Schwieger, W.; Maracke, S.; Fröba, M.; Rößner, F. Surface and texture properties of mesoporous silica materials modified by silicon-organic compounds containing quaternary amino groups for their application in base-catalyzed reactions. Microporous Mesoporous Mater. 2015, 203, 224–231. [Google Scholar] [CrossRef]
  13. Trisunaryanti, W.; Dwi Putri, A.; Lutfiana, A.; Dewi, K. Transesterification of Waste Cooking Oil Using NH2/MCM-41 Base Catalyst: Effect of Methanol/Oil Mole Ratio And Catalyst/Oil Weight Ratio towards Conversion of Ester. Asian J. Chem. 2018, 30, 953–957. [Google Scholar] [CrossRef]
  14. Vekariya, R.H.; Prajapati, N.P.; Patel, H.D. MCM-41-anchored sulfonic acid (MCM-41-SO3H): An efficient heterogeneous catalyst for green organic synthesis. Synth. Commun. 2016, 46, 1713–1734. [Google Scholar] [CrossRef]
  15. Luštická, I.; Vrbková, E.; Vyskočilová, E.; Paterová, I.; Červený, L. Acid functionalized MCM-41 as a catalyst for the synthesis of benzal-1,1-diacetate. React. Kinet. Mech. Catal. 2013, 108, 205–212. [Google Scholar] [CrossRef]
  16. Alinasab Amiri, A.; Javanshir, S.; Dolatkhah, Z.; Dekamin, M.G. SO3H-functionalized mesoporous silica materials as solid acid catalyst for facile and solvent-free synthesis of 2H-indazolo[2,1-b]phthalazine-1,6,11-trione derivatives. New J. Chem. 2015, 39, 9665–9671. [Google Scholar] [CrossRef]
  17. Safaei, S.; Mohammadpoor-Baltork, I.; Khosropour, A.R.; Moghadam, M.; Tangestaninejad, S.; Mirkhani, V. SO3H-functionalized MCM-41 as an efficient catalyst for the combinatorial synthesis of 1H-pyrazolo-[3,4-b]pyridines and spiro-pyrazolo-[3,4-b]pyridines. J. Iran. Chem. Soc. 2017, 14, 1583–1589. [Google Scholar] [CrossRef]
  18. Kaiprommarat, S.; Kongparakul, S.; Reubroycharoen, P.; Guan, G.; Samart, C. Highly efficient sulfonic MCM-41 catalyst for furfural production: Furan-based biofuel agent. Fuel 2016, 174, 189–196. [Google Scholar] [CrossRef]
  19. Karnjanakom, S.; Kongparakul, S.; Chaiya, C.; Reubroycharoen, P.; Guan, G.; Samart, C. Biodiesel production from Hevea brasiliensis oil using SO3H-MCM-41 catalyst. J. Environ. Chem. Eng. 2016, 4, 47–55. [Google Scholar] [CrossRef]
  20. Sarrafi, Y.; Mehrasbi, E.; Mashalchi, S.Z. MCM-41-SO3H: An efficient, reusable, heterogeneous catalyst for the one-pot, three-component synthesis of pyrano[3,2-b]pyrans. In Research on Chemical Intermediates; Springer: Dordrecht, The Netherlands, 2015; pp. 1–13. [Google Scholar]
  21. Appaturi, J.N.; Selvaraj, M.; Abdul Hamid, S.B.; Bin Johan, M.R. Synthesis of 3-(2-furylmethylene)-2,4-pentanedione using DL-Alanine functionalized MCM-41 catalyst via Knoevenagel condensation reaction. Microporous Mesoporous Mater. 2018, 260, 260–269. [Google Scholar] [CrossRef]
  22. Vrbková, E.; Vyskočilová, E.; Červený, L. Functionalized MCM-41 as a catalyst for the aldol condensation of 4-isopropylbenzaldehyde and propanal. React. Kinet. Mech. Catal. 2015, 114, 675–684. [Google Scholar] [CrossRef]
  23. Sharma, K.K.; Asefa, T. Efficient bifunctional nanocatalysts by simple postgrafting of spatially isolated catalytic groups on mesoporous materials. Angew. Chem. Int. Ed. 2007, 46, 2879–2882. [Google Scholar] [CrossRef]
  24. Collier, V.E.; Ellebracht, N.C.; Lindy, G.I.; Moschetta, E.G.; Jones, C.W. Kinetic and Mechanistic Examination of Acid-Base Bifunctional Aminosilica Catalysts in Aldol and Nitroaldol Condensations. ACS Catal. 2016, 6, 460–468. [Google Scholar] [CrossRef]
  25. Ray, S.; Das, P.; Banerjee, B.; Bhaumik, A.; Mukhopadhyay, C. Piperazinylpyrimidine modified MCM-41 for the ecofriendly synthesis of benzothiazoles by the simple cleavage of disulfide in the presence of molecular O2. RSC Adv. 2015, 5, 72745–72754. [Google Scholar] [CrossRef]
  26. Huang, Y.; Xu, S.; Lin, V.S.Y. Bifunctionalized Mesoporous Materials with Site-Separated Brønsted Acids and Bases: Catalyst for a Two-Step Reaction Sequence. Angew. Chem. Int. Ed. Engl. 2011, 50, 661–664. [Google Scholar] [CrossRef] [PubMed]
  27. Dickschat, A.T.; Behrends, F.; Buehner, M.; Ren, J.; Weiss, M.; Eckert, H.; Studer, A. Preparation of Bifunctional Mesoporous Silica Nanoparticles by Orthogonal Click Reactions and Their Application in Cooperative Catalysis. Chem. Eur. J. 2012, 18, 16689–16697. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Dickschat, A.T.; Behrends, F.; Surmiak, S.; Weiß, M.; Eckert, H.; Studer, A. Bifunctional mesoporous silica nanoparticles as cooperative catalysts for the Tsuji-Trost reaction-tuning the reactivity of silica nanoparticles. Chem. Commun. 2013, 49, 2195–2197. [Google Scholar] [CrossRef]
  29. Tsai, C.-H.; Chen, H.-T.; Althaus, S.M.; Mao, K.; Kobayashi, T.; Pruski, M.; Lin, V.S.Y. Rational Catalyst Design: A Multifunctional Mesoporous Silica Catalyst for Shifting the Reaction Equilibrium by Removal of Byproduct. ACS Catal. 2011, 1, 729–732. [Google Scholar] [CrossRef]
  30. Jabbari, A.; Mahdavi, H.; Nikoorazm, M.; Ghorbani-Choghamarani, A. Salen copper(II) complex heterogenized on mesoporous MCM-41 as nanoreactor catalyst for the selective oxidation of sulfides using urea hydrogen peroxide (UHP). Res. Chem. Intermed. 2015, 41, 5649–5663. [Google Scholar] [CrossRef]
  31. Tamoradi, T.; Ghadermazi, M.; Ghorbani-Choghamarani, A. Synthesis of Polyhydroquinoline, 2,3-Dihydroquinazolin-4(1H)-one, Sulfide and Sulfoxide Derivatives Catalyzed by New Copper Complex Supported on MCM-41. Catal. Lett. 2018, 148, 857–872. [Google Scholar] [CrossRef]
  32. Hajipour, A.R.; Hosseini, S.M.; Jajarmi, S. Cu(II)-Et-S@MCM-41: A Green and Cost-Effective Catalytic System for S-arylation of Aryl Halides Using Thiourea and Benzyl Bromide. ChemistrySelect 2017, 2, 2388–2394. [Google Scholar] [CrossRef]
  33. Lin, Y.; Cai, M.; Fang, Z.; Zhao, H. MCM-41-immobilized 1,10-phenanthroline-copper(I) complex: A highly efficient and recyclable catalyst for the coupling of aryl iodides with aliphatic alcohols. RSC Adv. 2016, 6, 85186–85193. [Google Scholar] [CrossRef]
  34. Liu, C.-C.; Janmanchi, D.; Wen, D.-R.; Oung, J.-N.; Mou, C.-Y.; Yu, S.S.F.; Chan, S.I. Catalytic Oxidation of Light Alkanes Mediated at Room Temperature by a Tricopper Cluster Complex Immobilized in Mesoporous Silica Nanoparticles. ACS Sustain. Chem. Eng. 2018, 6, 5431–5440. [Google Scholar] [CrossRef]
  35. Zendehdel, M.; Zamani, F. MCM-41-supported NNO type Schiff base complexes: A highly selective heterogeneous nanocatalyst for esterification, Diels–Alder and aldol condensation. J. Porous Mater. 2017, 24, 1263–1277. [Google Scholar] [CrossRef]
  36. Moorthy, M.; Kannan, B.; Madheswaran, B.; Rangappan, R. Tethering of Cu(II) Schiff base metal complex on mesoporous material MCM-41: Catalyst for Ullmann-type coupling reactions. J. Porous Mater. 2016, 23, 977–986. [Google Scholar] [CrossRef]
  37. Nejat, R.; Mahjoub, A.R.; Hekmatian, Z.; Azadbakht, T. Pd-functionalized MCM-41 nanoporous silica as an efficient and reusable catalyst for promoting organic reactions. RSC Adv. 2015, 5, 16029–16035. [Google Scholar] [CrossRef]
  38. Das, T.; Uyama, H.; Nandi, M. Pronounced effect of pore dimension of silica support on Pd-catalyzed Suzuki coupling reaction under ambient conditions. New J. Chem. 2018, 42, 6416–6426. [Google Scholar] [CrossRef]
  39. Adam, F.; Abbas, S.H. The imprinted MCM-41 palladium(II) salen complex: Its synthesis, characterisation and catalytic activity for the suzuki-miyaura reaction. J. Phys. Sci. 2016, 27, 15–38. [Google Scholar]
  40. Lin, B.-N.; Huang, S.-H.; Wu, W.-Y.; Mou, C.-Y.; Tsai, F.-Y. Sonogashira reaction of aryl and heteroaryl halides with terminal alkynes catalyzed by a highly efficient and recyclable nanosized MCM-41 anchored palladium bipyridyl complex. Molecules 2010, 15, 9157–9173. [Google Scholar] [CrossRef]
  41. Chen, J.-Y.; Lin, T.-C.; Chen, S.-C.; Chen, A.-J.; Mou, C.-Y.; Tsai, F.-Y. Highly-efficient and recyclable nanosized MCM-41 anchored palladium bipyridyl complex-catalyzed coupling of acyl chlorides and terminal alkynes for the formation of ynones. Tetrahedron 2009, 65, 10134–10141. [Google Scholar] [CrossRef]
  42. Yousefi, S.; Kiasat, A.R. MCM-41 bound dibenzo-18-crown-6 ether: A recoverable phase-transfer nano catalyst for smooth and regioselective conversion of oxiranes to β-azidohydrins and β-cyanohydrins in water. RSC Adv. 2015, 5, 92387–92393. [Google Scholar] [CrossRef]
  43. Nikoorazm, M.; Ghorbani-Choghamarani, A.; Ghobadi, M.; Massahi, S. Pd-SBT@MCM-41: As an efficient, stable and recyclable organometallic catalyst for C-C coupling reactions and synthesis of 5-substituted tetrazoles. Appl. Organomet. Chem. 2017, 31. [Google Scholar] [CrossRef]
  44. Fadhli, M.; Khedher, I.; Fraile, J.M. Modified Ti/MCM-41 catalysts for enantioselective epoxidation of styrene. J. Mol. Catal. A Chem. 2016, 420, 282–289. [Google Scholar] [CrossRef]
  45. Zhou, W.; He, D. Anchoring RhCl(CO)(PPh3)2 to -PrPPh2 Modified MCM-41 as Effective Catalyst for 1-Octene Hydroformylation. Catal. Lett. 2009, 127, 437–443. [Google Scholar] [CrossRef]
  46. Havasi, F.; Ghorbani-Choghamarani, A.; Nikpour, F. Synthesis and characterization of nickel complex anchored onto MCM-41 as a novel and reusable nanocatalyst for the efficient synthesis of 2,3-dihydroquinazolin-4(1H)-ones. Microporous Mesoporous Mater. 2016, 224, 26–35. [Google Scholar] [CrossRef]
  47. Xie, Y.; Yuan, W.; Huang, Y.; Wu, C.; Wang, H.; Xia, Y.; Liu, X. Zirconyl schiff base complex-functionalized MCM-41 catalyzes dehydration of fructose into 5-hydroxymethylfurfural in organic solvents. Chin. Chem. Lett. 2018. [Google Scholar] [CrossRef]
  48. Vasconcellos Dias, M.; Saraiva, M.S.; Ferreira, P.; Calhorda, M.J. Catalytic activity of molybdenum(II) complexes in homogeneous and heterogeneous conditions. Organometallics 2015, 34, 1465–1478. [Google Scholar] [CrossRef]
  49. Noori, N.; Nikoorazm, M.; Ghorbani-Choghamarani, A. Synthesis and characterization of Co (II) and Fe (III) Schiff base complexes grafted onto mesoporous MCM-41: A heterogeneous and recyclable nanocatalysts for the selective oxidation of sulfides and oxidative coupling of thiols. Phosphorus Sulfur Silicon Relat. Elem. 2016, 191, 1388–1395. [Google Scholar] [CrossRef]
  50. Noori, N.; Nikoorazm, M.; Ghorbani-Choghamarani, A. Synthesis and characterization of Ni and Zn Schiff base complexes supported on modified MCM-41 as reusable catalysts for various oxidation reactions. J. Porous Mater. 2015, 22, 1607–1615. [Google Scholar] [CrossRef]
  51. Muthusami, R.; Moorthy, M.; Kostova, I.; Anbarasu, G.; Manickam, C.; Rangappan, R. Designing a biomimetic catalyst for phenoxazinone synthase activity using mesoporous Schiff base Copper complex with a novel double-helix morphology. New J. Chem. 2018. [Google Scholar] [CrossRef]
  52. Bania, K.K.; Karunakar, G.V.; Satyanarayana, L. Oxidative coupling of 2-naphthol to (R)/(S)-BINOL by MCM-41 supported Mn-chiral Schiff base complexes. RSC Adv. 2015, 5, 33185–33198. [Google Scholar] [CrossRef]
  53. Yang, F.; Gao, S.; Xiong, C.; Wang, H.; Chen, J.; Kong, Y. Coordination of manganese porphyrins on amino-functionalized MCM-41 for heterogeneous catalysis of naphthalene hydroxylation. Cuihua Xuebao/Chinese J. Catal. 2015, 36, 1035–1041. [Google Scholar] [CrossRef]
  54. González, D.M.; Quijada, R.; Yazdani-Pedram, M.; Lourenço, J.P.; Ribeiro, M.R. Preparation of polypropylene-based nanocomposites using nanosized MCM-41 as support and in situ polymerization. Polym. Int. 2016, 65, 320–326. [Google Scholar] [CrossRef]
  55. Moritz, M.; Geszke-Moritz, M. Mesoporous materials as multifunctional tools in biosciences: Principles and applications. Mater. Sci. Eng. C 2015, 49, 114–151. [Google Scholar] [CrossRef] [PubMed]
  56. Carlsson, N.; Gustafsson, H.; Thörn, C.; Olsson, L.; Holmberg, K.; Åkerman, B. Enzymes immobilized in mesoporous silica: A physical–chemical perspective. Adv. Colloid Interface Sci. 2014, 205, 339–360. [Google Scholar] [CrossRef] [PubMed]
  57. Dang, P.T.; Le, H.G.; Hoang, V.-T.; Tran, H.T.H.; Dao, C.D.; Nguyen, K.T.; Le, G.H.; Nguyen, Q.K.; Nguyen, T.V.; Vu, T.A. Immobilization of D-Amino Acid Oxidase (DAAO) Enzyme on Hybrid Mesoporous MCF, SBA-15 and MCM-41 Nanomaterials. J. Nanosci. Nanotechnol. 2017, 17, 947–953. [Google Scholar] [CrossRef] [PubMed]
  58. Jia, F.; Narasimhan, B.; Mallapragada, S. Materials-based strategies for multi-enzyme immobilization and co-localization: A review. Biotechnol. Bioeng. 2014, 111, 209–222. [Google Scholar] [CrossRef] [PubMed]
  59. Díaz, J.F.; Balkus, K.J. Enzyme immobilization in MCM-41 molecular sieve. J. Mol. Catal. B Enzym. 1996, 2, 115–126. [Google Scholar] [CrossRef]
  60. Sadighi, A.; Motevalizadeh, S.F.; Hosseini, M.; Ramazani, A.; Gorgannezhad, L.; Nadri, H.; Deiham, B.; Ganjali, M.R.; Shafiee, A.; Faramarzi, M.A.; et al. Metal-Chelate Immobilization of Lipase onto Polyethylenimine Coated MCM-41 for Apple Flavor Synthesis. Appl. Biochem. Biotechnol. 2017, 182, 1371–1389. [Google Scholar] [CrossRef]
  61. Timofeeva, M. Acid catalysis by heteropoly acids. Appl. Catal. A Gen. 2003, 256, 19–35. [Google Scholar] [CrossRef]
  62. Patel, A.; Narkhede, N.; Singh, S.; Pathan, S. Keggin-type lacunary and transition metal substituted polyoxometalates as heterogeneous catalysts: A recent progress. Catal. Rev. 2016, 58, 337–370. [Google Scholar] [CrossRef]
  63. Kozhevinikov, I. Heteropoly Acids and Related Compounds as Catalysts for Fine Chemical Synthesis. Catal. Rev. 1995, 37, 311–352. [Google Scholar] [CrossRef]
  64. Coronel, N.C.; da Silva, M.J. Lacunar Keggin Heteropolyacid Salts: Soluble, Solid and Solid-Supported Catalysts. J. Clust. Sci. 2018, 29, 195–205. [Google Scholar] [CrossRef]
  65. Taguchi, A.; Schüth, F. Ordered mesoporous materials in catalysis. Microporous Mesoporous Mater. 2005, 77, 1–45. [Google Scholar] [CrossRef]
  66. Enferadi-Kerenkan, A.; Do, T.-O.; Kaliaguine, S. Heterogeneous catalysis by tungsten-based heteropoly compounds. Catal. Sci. Technol. 2018, 8, 2257–2284. [Google Scholar] [CrossRef]
  67. Ding, J.; Ma, T.; Yun, Z.; Shao, R. Heteropolyacid (H3PW12O40) supported MCM-41: An effective solid acid catalyst for the dehydration of glycerol to acrolein. J. Wuhan Univ. Technol.-Mater. Sci. Ed. 2017, 32, 1511–1516. [Google Scholar] [CrossRef]
  68. Kong, S.I.; Matei, D.; Cursaru, D.; Matei, V.; Ciuparu, D. Characterization and stability of heteropolyacid catalysts supported on MCM-41 materials synthesized by ultrasonic irradiation. Rev. Chim. 2017, 68, 101–107. [Google Scholar]
  69. Kocaman, E.; Akarçay, Ö.; Bağlar, N.; Çelebi, S.; Uzun, A. Isobutene oligomerization on MCM-41-supported tungstophosphoric acid. Mol. Catal. 2018, 457, 41–50. [Google Scholar] [CrossRef]
  70. Chen, Y.; Chen, X.; Dong, B.-B.; Wang, G.-H.; Zheng, X.-C. Facile synthesis and characterization of 12-tungstophosphoric acid anchoring MCM-41 mesoporous materials. Mater. Lett. 2014, 114, 72–75. [Google Scholar] [CrossRef]
  71. Liu, A.; Zhang, Z.; Fang, Z.; Liu, B.; Huang, K. Synthesis of 5-ethoxymethylfurfural from 5-hydroxymethylfurfural and fructose in ethanol catalyzed by MCM-41 supported phosphotungstic acid. J. Ind. Eng. Chem. 2014, 20, 1977–1984. [Google Scholar] [CrossRef]
  72. Chen, Y.; Zhang, X.; Dong, M.; Wu, Y.; Zheng, G.; Huang, J.; Guan, X.; Zheng, X. MCM-41 immobilized 12-silicotungstic acid mesoporous materials: Structural and catalytic properties for esterification of levulinic acid and oleic acid. J. Taiwan Inst. Chem. Eng. 2016, 61, 147–155. [Google Scholar] [CrossRef]
  73. Ma, T.; Ding, J.; Shao, R.; Xu, W.; Yun, Z. Dehydration of glycerol to acrolein over Wells–Dawson and Keggin type phosphotungstic acids supported on MCM-41 catalysts. Chem. Eng. J. 2017, 316, 797–806. [Google Scholar] [CrossRef]
  74. Ding, J.; Cui, M.; Ma, T.; Shao, R.; Xu, W.; Wang, P. Catalytic amination of glycerol with dimethylamine over different type of heteropolyacid/Zr-MCM-41 catalysts. Mol. Catal. 2018, 457, 51–58. [Google Scholar] [CrossRef]
  75. Dias, A.S.; Pillinger, M.; Valente, A.A. Mesoporous silica-supported 12-tungstophosphoric acid catalysts for the liquid phase dehydration of d-xylose. Microporous Mesoporous Mater. 2006, 94, 214–225. [Google Scholar] [CrossRef]
  76. Luo, G.; Kang, L.; Zhu, M.; Dai, B. Highly active phosphotungstic acid immobilized on amino functionalized MCM-41 for the oxidesulfurization of dibenzothiophene. Fuel Process. Technol. 2014, 118, 20–27. [Google Scholar] [CrossRef]
  77. Ding, J.; Ma, T.; Shao, R.; Xu, W.; Wang, P.; Song, X.; Guan, R.; Yeung, K.; Han, W. Gas phase dehydration of glycerol to acrolein on an amino siloxane-functionalized MCM-41 supported Wells–Dawson type H6P2W18O62 catalyst. New J. Chem. 2018, 42, 14271–14280. [Google Scholar] [CrossRef]
  78. Dong, X.; Wang, D.; Li, K.; Zhen, Y.; Hu, H.; Xue, G. Vanadium-substituted heteropolyacids immobilized on amine- functionalized mesoporous MCM-41: A recyclable catalyst for selective oxidation of alcohols with H2O2. Mater. Res. Bull. 2014, 57, 210–220. [Google Scholar] [CrossRef]
  79. Popa, A.; Sasca, V.; Verdes, O.; Ianasi, C.; Banica, R. Heteropolyacids anchored on amino-functionalized MCM-41 and SBA-15 and its application to the ethanol conversion reaction. J. Therm. Anal. Calorim. 2017, 127, 319–334. [Google Scholar] [CrossRef]
  80. Thompson, D.J.; Zhang, Y.; Ren, T. Polyoxometalate [γ-SiW10O34(H2O)2]4− on MCM-41 as catalysts for sulfide oxygenation with hydrogen peroxide. J. Mol. Catal. A Chem. 2014, 392, 188–193. [Google Scholar] [CrossRef]
  81. Tayebee, R.; Amini, M.M.; Akbari, M.; Aliakbari, A. A novel inorganic–organic nanohybrid material H4SiW12O40/pyridino-MCM-41 as efficient catalyst for the preparation of 1-amidoalkyl-2-naphthols under solvent-free conditions. Dalton Trans. 2015, 44, 9596–9609. [Google Scholar] [CrossRef]
  82. Wang, S.; McGuirk, C.M.; D’Aquino, A.; Mason, J.A.; Mirkin, C.A. Metal-Organic Framework Nanoparticles. Adv. Mater. 2018, 30, 1800202. [Google Scholar] [CrossRef]
  83. Du, D.-Y.; Qin, J.-S.; Li, S.-L.; Su, Z.-M.; Lan, Y.-Q. Recent advances in porous polyoxometalate-based metal–organic framework materials. Chem. Soc. Rev. 2014, 43, 4615–4632. [Google Scholar] [CrossRef]
  84. Li, S.-W.; Gao, R.-M.; Zhang, R.-L.; Zhao, J. Template method for a hybrid catalyst material POM@MOF-199 anchored on MCM-41: Highly oxidative desulfurization of DBT under molecular oxygen. Fuel 2016, 184, 18–27. [Google Scholar] [CrossRef]
  85. Li, S.W.; Li, J.R.; Jin, Q.P.; Yang, Z.; Zhang, R.L.; Gao, R.M.; Zhao, J.S. Preparation of mesoporous Cs-POM@MOF-199@MCM-41 under two different synthetic methods for a highly oxidesulfurization of dibenzothiophene. J. Hazard. Mater. 2017, 337, 208–216. [Google Scholar] [CrossRef] [PubMed]
  86. Li, S.-W.; Yang, Z.; Gao, R.-M.; Zhang, G.; Zhao, J. Direct synthesis of mesoporous SRL-POM@MOF-199@MCM-41 and its highly catalytic performance for the oxidesulfurization of DBT. Appl. Catal. B Environ. 2018, 221, 574–583. [Google Scholar] [CrossRef]
  87. Oliveira, R.D.S.; Camilo, F.F.; Bizeto, M.A. Evaluation of the influence of sulfur-based functional groups on the embedding of silver nanoparticles into the pores of MCM-41. J. Solid State Chem. 2016, 235, 125–131. [Google Scholar] [CrossRef]
  88. Lin, X.-J.; Zhong, A.-Z.; Sun, Y.-B.; Zhang, X.; Song, W.-G.; Lu, R.-W.; Cao, A.-M.; Wan, L.-J. In situ encapsulation of Pd inside the MCM-41 channel. Chem. Commun. 2015, 51, 7482–7485. [Google Scholar] [CrossRef] [PubMed]
  89. Wu, N.; Zhang, W.; Li, B.; Han, C. Nickel nanoparticles highly dispersed with an ordered distribution in MCM-41 matrix as an efficient catalyst for hydrodechlorination of chlorobenzene. Microporous Mesoporous Mater. 2014, 185, 130–136. [Google Scholar] [CrossRef]
  90. Mori, K.; Yamaguchi, T.; Ikurumi, S.; Yamashita, H. Positive Effect of the Residual Templates within the MCM-41 Mesoporous Silica Channels in the Metal-Catalyzed Catalytic Reactions. Optoelectron. Adv. Mater. Rapid Commun. 2010, 4, 1166–1169. [Google Scholar]
  91. Feiz, A.; Bazgir, A. Gold nanoparticles supported on mercaptoethanol directly bonded to MCM-41: An efficient catalyst for the synthesis of propargylamines. Catal. Commun. 2016, 73, 88–92. [Google Scholar] [CrossRef]
  92. Dai, B.; Wen, B.; Zhu, M.; Kang, L.; Yu, F. Nickel catalysts supported on amino-functionalized MCM-41 for syngas methanation. RSC Adv. 2016, 6, 66957–66962. [Google Scholar] [CrossRef]
  93. Liao, H.; Ouyang, D.; Zhang, J.; Xiao, Y.; Liu, P.; Hao, F.; You, K.; Luo, H. Benzene hydrogenation over oxide-modified MCM-41 supported ruthenium-lanthanum catalyst: The influence of zirconia crystal form and surface hydrophilicity. Chem. Eng. J. 2014, 243, 207–216. [Google Scholar] [CrossRef]
  94. Azaroon, M.; Kiasat, A.R. An efficient and new protocol for the Heck reaction using palladium nanoparticle-engineered dibenzo-18-crown-6-ether/MCM-41 nanocomposite in water. Appl. Organomet. Chem. 2018, 32. n/a. [Google Scholar] [CrossRef]
  95. Demel, J.; Cejka, J.; Stepnicka, P. The use of palladium nanoparticles supported on MCM-41 mesoporous molecular sieves in the Heck reaction: A comparison of basic and neutral supports. J. Mol. Catal. A Chem. 2007, 274, 127–132. [Google Scholar] [CrossRef]
  96. Demel, J.; Park, S.-E.; Cejka, J.; Stepnicka, P. The use of palladium nanoparticles supported with MCM-41 and basic (Al)MCM-41 mesoporous sieves in microwave-assisted Heck reaction. Catal. Today 2008, 132, 63–67. [Google Scholar] [CrossRef]
  97. Chatterjee, M.; Ishizaka, T.; Suzuki, T.; Suzuki, A.; Kawanami, H. In situ synthesized Pd nanoparticles supported on B-MCM-41: An efficient catalyst for hydrogenation of nitroaromatics in supercritical carbon dioxide. Green Chem. 2012, 14, 3415–3422. [Google Scholar] [CrossRef]
  98. Chatterjee, M.; Ishizaka, T.; Kawanami, H. Preparation and characterization of PdO nanoparticles on trivalent metal (B, Al and Ga) substituted MCM-41: Excellent catalytic activity in supercritical carbon dioxide. J. Colloid Interface Sci. 2014, 420, 15–26. [Google Scholar] [CrossRef] [PubMed]
  99. Yan, K.; Lafleur, T.; Jarvis, C.; Wu, G. Clean and selective production of γ-valerolactone from biomass-derived levulinic acid catalyzed by recyclable Pd nanoparticle catalyst. J. Clean. Prod. 2014, 72, 230–232. [Google Scholar] [CrossRef]
  100. Wang, A.; Lu, Y.; Yi, Z.; Ejaz, A.; Hu, K.; Zhang, L.; Yan, K. Selective Production of g -Valerolactone and Valeric Acid in One-Pot Bifunctional Metal Catalysts. ChemistrySelect 2018, 1097–1101. [Google Scholar] [CrossRef]
  101. Ivashchenko, N.A.; Gac, W.; Tertykh, V.A.; Yanishpolskii, V.V.; Khainakov, S.A.; Dikhtiarenko, A.V.; Pasieczna-Patkowska, S.; Zawadzki, W. Preparation, characterization and catalytic activity of palladium nanoparticles embedded in the mesoporous silica matrices. World J. Nano Sci. Eng. 2012, 2, 117–125. [Google Scholar] [CrossRef]
  102. Chu, X.; Wang, C.; Guo, L.; Chi, Y.; Gao, X.; Yang, X. Mesoporous silica supported Au nanoparticles with controlled size as efficient heterogeneous catalyst for aerobic oxidation of alcohols. J. Chem. 2015, 2015. [Google Scholar] [CrossRef]
  103. Xu, J.; Qu, Z.; Wang, Y.; Huang, B. HCHO oxidation over highly dispersed Au nanoparticles supported on mesoporous silica with superior activity and stability. Catal. Today 2018. [Google Scholar] [CrossRef]
  104. Finashina, E.D.; Tkachenko, O.P.; Startseva, A.Y.; Redina, E.A.; Krasovsky, V.G.; Kustov, L.M.; Beletskaya, I.P. Intramolecular hydroamination of 2-(2-phenylethynyl)aniline catalyzed by gold nanoparticles. Russ. Chem. Bull. 2015, 64, 2821–2829. [Google Scholar] [CrossRef]
  105. Mukherjee, P.; Patra, C.R.; Kumar, R.; Sastry, M. Entrapment and catalytic activity of gold nanoparticles in amine-functionalized MCM-41 matrices synthesized by spontaneous reduction of aqueous chloroaurate ions. PhysChemComm 2001, 4, 24–25. [Google Scholar] [CrossRef]
  106. Patra, C.R.; Ghosh, A.; Mukherjee, P.; Sastry, M.; Kumar, R. Formation and stabilization of gold nanoparticles in organo-functionalized MCM-41 mesoporous materials and their catalytic applications. Stud. Surf. Sci. Catal. 2002, 141, 641–646. [Google Scholar]
  107. Joseph, T.; Kumar, K.V.; Ramaswamy, A.V.; Halligudi, S.B. Au-Pt nanoparticles in amine functionalized MCM-41: Catalytic evaluation in hydrogenation reactions. Catal. Commun. 2007, 8, 629–634. [Google Scholar] [CrossRef]
  108. Chilukuri, S.; Joseph, T.; Malwadkar, S.; Damle, C.; Halligudi, S.B.; Rao, B.S.; Sastry, M.; Ratnasamy, P. Au and Au-Pt bimetallic nanoparticles in MCM-41 materials: Applications in CO preferential oxidation. Stud. Surface Sci. Catal. 2002, 146, 573–576. [Google Scholar] [CrossRef]
  109. Yen, C.W.; Lin, M.L.; Wang, A.; Chen, S.A.; Chen, J.M.; Mou, C.Y. CO oxidation catalyzed by Au-Ag bimetallic nanoparticles supported in mesoporous silica. J. Phys. Chem. C 2009, 113, 17831–17839. [Google Scholar] [CrossRef]
  110. Wang, A.; Hsieh, Y.P.; Chen, Y.F.; Mou, C.Y. Au-Ag alloy nanoparticle as catalyst for CO oxidation: Effect of Si/Al ratio of mesoporous support. J. Catal. 2006, 237, 197–206. [Google Scholar] [CrossRef]
  111. Wang, A.Q.; Chang, C.M.; Mou, C.Y. Evolution of catalytic activity of Au-Ag bimetallic nanoparticles on mesoporous support for CO oxidation. J. Phys. Chem. B 2005, 109, 18860–18867. [Google Scholar] [CrossRef]
  112. Torres, C.C.; Alderete, J.B.; Pecchi, G.; Campos, C.H.; Reyes, P.; Pawelec, B.; Vaschetto, E.G.; Eimer, G.A. Heterogeneous hydrogenation of nitroaromatic compounds on gold catalysts: Influence of titanium substitution in MCM-41 mesoporous supports. Appl. Catal. A Gen. 2016, 517, 110–119. [Google Scholar] [CrossRef]
  113. Mori, K.; Araki, T.; Shironita, S.; Sonoda, J.; Yamashita, H. Supported Pd and PdAu nanoparticles on Ti-MCM-41 prepared by a photo-assisted deposition method as efficient catalysts for direct synthesis of H2O2 from H2 and O2. Catal. Lett. 2009, 131, 337–343. [Google Scholar] [CrossRef]
  114. Tang, Z.; Fiorilli, S.L.; Heeres, H.J.; Pescarmona, P.P. Multifunctional Heterogeneous Catalysts for the Selective Conversion of Glycerol into Methyl Lactate. ACS Sustain. Chem. Eng. 2018, 6, 10923–10933. [Google Scholar] [CrossRef] [PubMed]
  115. Ye, M.; Tao, Y.; Jin, F.; Ling, H.; Wu, C.; Williams, P.T.; Huang, J. Enhancing hydrogen production from the pyrolysis-gasification of biomass by size-confined Ni catalysts on acidic MCM-41 supports. Catal. Today 2018, 307, 154–161. [Google Scholar] [CrossRef] [Green Version]
  116. Moussa, S.; Arribas, M.A.; Concepcion, P.; Martinez, A. Heterogeneous oligomerization of ethylene to liquids on bifunctional Ni-based catalysts: The influence of support properties on nickel speciation and catalytic performance. Catal. Today 2016, 277, 78–88. [Google Scholar] [CrossRef]
  117. Al-Fatesh, A.; Ibrahim, A.; Abu-Dahrieh, J.; Al-Awadi, A.; El-Toni, A.; Fakeeha, A.; Abasaeed, A. Gallium-Promoted Ni Catalyst Supported on MCM-41 for Dry Reforming of Methane. Catalysts 2018, 8, 229. [Google Scholar] [CrossRef]
  118. Wang, X.; Zhu, L.; Liu, Y.; Wang, S. CO2 methanation on the catalyst of Ni/MCM-41 promoted with CeO2. Sci. Total Environ. 2018, 625, 686–695. [Google Scholar] [CrossRef] [PubMed]
  119. Pirouzmand, M.; Asadi, M.; Mohammadi, A. The remarkable activity of template-containing Mg/MCM-41 and Ni/MCM-41 in CO2 sequestration. Greenh. Gases Sci. Technol. 2018, 8, 462–468. [Google Scholar] [CrossRef]
  120. Ning, X.; Lu, Y.; Fu, H.; Wan, H.; Xu, Z.; Zheng, S. Template-Mediated Ni(II) Dispersion in Mesoporous SiO2 for Preparation of Highly Dispersed Ni Catalysts: Influence of Template Type. ACS Appl. Mater. Interfaces 2017, 9, 19335–19344. [Google Scholar] [CrossRef]
  121. Carrillo, A.I.; Schmidt, L.C.; Marin, M.L.; Scaiano, J.C. Mild synthesis of mesoporous silica supported ruthenium nanoparticles as heterogeneous catalysts in oxidative Wittig coupling reactions. Catal. Sci. Technol. 2014, 4, 435–440. [Google Scholar] [CrossRef] [Green Version]
  122. Mondal, D.K.; Mondal, C.; Roy, S. Catalytic wet air oxidation of aqueous solution of phenol in a fixed bed reactor over Ru catalysts supported on ceria promoted MCM-41. RSC Adv. 2016, 6, 114383–114395. [Google Scholar] [CrossRef]
  123. Zarrabi, M.; Entezari, M.H.; Goharshadi, E.K. Photocatalytic oxidative desulfurization of dibenzothiophene by C/TiO2@MCM-41 nanoparticles under visible light and mild conditions. RSC Adv. 2015, 5, 34652–34662. [Google Scholar] [CrossRef]
  124. Magatani, Y.; Kuwahara, Y.; Nishizawa, K.; Yamashita, H. Dramatically Enhanced Phenol Degradation on Alkali Cation-Anchored TiO2/SiO2 Hybrids: Effect of Cation-π Interaction as a Diffusion-Controlling Tool in Heterogeneous Catalysis. ChemistrySelect 2017, 2, 4332–4337. [Google Scholar] [CrossRef]
  125. Lin, K.; Pescarmona, P.P.; Houthoofd, K.; Liang, D.; Van Tendeloo, G.; Jacobs, P.A. Direct room-temperature synthesis of methyl-functionalized Ti-MCM-41 nanoparticles and their catalytic performance in epoxidation. J. Catal. 2009, 263, 75–82. [Google Scholar] [CrossRef]
  126. Astruc, D.; Lu, F.; Aranzaes, J.R. Nanoparticles as Recyclable Catalysts: The Frontier between Homogeneous and Heterogeneous Catalysis. Angew. Chem. Int. Ed. 2005, 44, 7852–7872. [Google Scholar] [CrossRef]
  127. Sharma, R.K.; Yadav, M.; Gawande, M.B. Silica-Coated Magnetic Nano-Particles: Application in Catalysis. In Ferrites and Ferrates: Chemistry and Applications in Sustainable Energy and Envirnomental Remediation; American Chemical Society: Washington, DC, USA, 2016; pp. 1–38. [Google Scholar]
  128. Rossi, L.M.; Costa, N.J.S.; Silva, F.P.; Wojcieszak, R. Magnetic nanomaterials in catalysis: Advanced catalysts for magnetic separation and beyond. Green Chem. 2014, 16, 2906–2933. [Google Scholar] [CrossRef]
  129. Gawande, M.B.; Monga, Y.; Zboril, R.; Sharma, R.K. Silica-decorated magnetic nanocomposites for catalytic applications. Coord. Chem. Rev. 2015, 288, 118–143. [Google Scholar] [CrossRef]
  130. Zhu, K.; Ju, Y.; Xu, J.; Yang, Z.; Gao, S.; Hou, Y. Magnetic Nanomaterials: Chemical Design, Synthesis, and Potential Applications. Acc. Chem. Res. 2018, 51, 404–413. [Google Scholar] [CrossRef] [PubMed]
  131. Kudr, J.; Haddad, Y.; Richtera, L.; Heger, Z.; Cernak, M.; Adam, V.; Zitka, O. Magnetic Nanoparticles: From Design and Synthesis to Real World Applications. Nanomaterials 2017, 7, 243. [Google Scholar] [CrossRef]
  132. Ali, A.; Zafar, H.; Zia, M.; ul Haq, I.; Phull, A.R.; Ali, J.S.; Hussain, A. Synthesis, characterization, applications, and challenges of iron oxide nanoparticles. Nanotechnol. Sci. Appl. 2016, 9, 49–67. [Google Scholar] [CrossRef]
  133. Gotić, M.; Jurkin, T.; Musić, S. Factors that may influence the micro-emulsion synthesis of nanosize magnetite particles. Colloid Polym. Sci. 2007, 285, 793–800. [Google Scholar] [CrossRef]
  134. Zhou, Z.H.; Wang, J.; Liu, X.; Chan, H.S.O. Synthesis of Fe3O4 nanoparticles from emulsions. J. Mater. Chem. 2001, 11, 1704–1709. [Google Scholar] [CrossRef]
  135. Yu, W.W.; Falkner, J.C.; Yavuz, C.T.; Colvin, V.L. Synthesis of monodisperse iron oxide nanocrystals by thermal decomposition of iron carboxylate salts. Chem. Commun. 2004, 20, 2306–2307. [Google Scholar] [CrossRef] [PubMed]
  136. Nedkov, I.; Merodiiska, T.; Slavov, L.; Vandenberghe, R.E.; Kusano, Y.; Takada, J. Surface oxidation, size and shape of nano-sized magnetite obtained by co-precipitation. J. Magn. Magn. Mater. 2006, 300, 358–367. [Google Scholar] [CrossRef]
  137. Utkan, G.G.; Sayar, F.; Batat, P.; Ide, S.; Kriechbaum, M.; Pişkin, E. Synthesis and characterization of nanomagnetite particles and their polymer coated forms. J. Colloid Interface Sci. 2011, 353, 372–379. [Google Scholar] [CrossRef] [PubMed]
  138. Mizutani, N.; Iwasaki, T.; Watano, S.; Yanagida, T.; Tanaka, H.; Kawai, T. Effect of ferrous/ferric ions molar ratio on reaction mechanism for hydrothermal synthesis of magnetite nanoparticles. Bull. Mater. Sci. 2008, 31, 713–717. [Google Scholar] [CrossRef]
  139. Daou, T.J.; Pourroy, G.; Bégin-Colin, S.; Grenèche, J.M.; Ulhaq-Bouillet, C.; Legaré, P.; Bernhardt, P.; Leuvrey, C.; Rogez, G. Hydrothermal Synthesis of Monodisperse Magnetite Nanoparticles. Chem. Mater. 2006, 18, 4399–4404. [Google Scholar] [CrossRef]
  140. Zhang, H.; Zhong, X.; Xu, J.-J.; Chen, H.-Y. Fe3O4/Polypyrrole/Au Nanocomposites with Core/Shell/Shell Structure: Synthesis, Characterization, and Their Electrochemical Properties. Langmuir 2008, 24, 13748–13752. [Google Scholar] [CrossRef] [PubMed]
  141. Bourlinos, A.B.; Simopoulos, A.; Boukos, N.; Petridis, D. Magnetic modification of the external surfaces in the MCM-41 porous silica: Synthesis, characterization, and functionalization. J. Phys. Chem. B 2001, 105, 7432–7437. [Google Scholar] [CrossRef]
  142. Arruebo, M.; Ho, W.Y.; Lam, K.F.; Chen, X.; Arbiol, J.; Santamaría, J.; Yeung, K.L. Preparation of Magnetic Nanoparticles Encapsulated by an Ultrathin Silica Shell via Transformation of Magnetic Fe-MCM-41. Chem. Mater. 2008, 20, 486–493. [Google Scholar] [CrossRef]
  143. Zhang, L.; Papaefthymiou, G.C.; Ying, J.Y. Synthesis and Properties of γ-Fe2O3 Nanoclusters within Mesoporous Aluminosilicate Matrices. J. Phys. Chem. B 2001, 105, 7414–7423. [Google Scholar] [CrossRef]
  144. Shokouhimehr, M.; Piao, Y.; Kim, J.; Jang, Y.; Hyeon, T. A Magnetically Recyclable Nanocomposite Catalyst for Olefin Epoxidation. Angew. Chem. 2007, 119, 7169–7173. [Google Scholar] [CrossRef]
  145. MacLachlan, M.J.; Ginzburg, M.; Coombs, N.; Raju, N.P.; Greedan, J.E.; Ozin, G.A.; Manners, I. Superparamagnetic Ceramic Nanocomposites: Synthesis and Pyrolysis of Ring-Opened Poly(ferrocenylsilanes) inside Periodic Mesoporous Silica. J. Am. Chem. Soc. 2000, 122, 3878–3891. [Google Scholar] [CrossRef]
  146. Mori, K.; Kondo, Y.; Morimoto, S.; Yamashita, H. Synthesis and Multifunctional Properties of Superparamagnetic Iron Oxide Nanoparticles Coated with Mesoporous Silica Involving Single-Site Ti−Oxide Moiety. J. Phys. Chem. C 2008, 112, 397–404. [Google Scholar] [CrossRef]
  147. Alvaro, M.; Aprile, C.; Garcia, H.; Gómez-García, C.J. Synthesis of a Hydrothermally Stable, Periodic Mesoporous Material Containing Magnetite Nanoparticles, and the Preparation of Oriented Films. Adv. Funct. Mater. 2006, 16, 1543–1548. [Google Scholar] [CrossRef]
  148. Chen, X.; Lam, K.F.; Zhang, Q.; Pan, B.; Arruebo, M.; Yeung, K.L. Synthesis of Highly Selective Magnetic Mesoporous Adsorbent. J. Phys. Chem. C 2009, 113, 9804–9813. [Google Scholar] [CrossRef]
  149. Zanardi, F.B.; Barbosa, I.A.; de Sousa Filho, P.C.; Zanatta, L.D.; da Silva, D.L.; Serra, O.A.; Iamamoto, Y. Manganese porphyrin functionalized on Fe3O4@n SiO2@MCM-41 magnetic composite: Structural characterization and catalytic activity as cytochrome P450 model. Microporous Mesoporous Mater. 2016, 219, 161–171. [Google Scholar] [CrossRef]
  150. Shadjou, N.; Hasanzadeh, M. Amino functionalized mesoporous silica decorated with iron oxide nanoparticles as a magnetically recoverable nanoreactor for the synthesis of a new series of 2,4-diphenylpyrido[4,3-d]pyrimidines. RSC Adv. 2014, 4, 18117–18126. [Google Scholar] [CrossRef]
  151. Rostamizadeh, S.; Shadjou, N.; Isapoor, E.; Hasanzadeh, M. Catalytic activity of (Fe2O3)-MCM-41-nPrNH2 magnetically recoverable nanocatalyst for the synthesis of phenylpyrido[4,3-d]pyrimidins. J. Nanosci. Nanotechnol. 2013, 13, 4925–4933. [Google Scholar] [CrossRef]
  152. Khorshidi, A.; Shariati, S. Sulfuric acid functionalized MCM-41 coated on magnetite nanoparticles as a recyclable core-shell solid acid catalyst for three-component condensation of indoles, aldehydes and thiols. RSC Adv. 2014, 4, 41469–41475. [Google Scholar] [CrossRef]
  153. Rostamizadeh, S.; Azad, M.; Shadjou, N.; Hasanzadeh, M. (α-Fe2O3)-MCM-41-SO3H as a novel magnetic nanocatalyst for the synthesis of N-aryl-2-amino-1,6-naphthyridine derivatives. Catal. Commun. 2012, 25, 83–91. [Google Scholar] [CrossRef]
  154. Saadatjoo, N.; Golshekan, M.; Shariati, S.; Kefayati, H.; Azizi, P. Organic/inorganic MCM-41 magnetite nanocomposite as a solid acid catalyst for synthesis of benzo[α]xanthenone derivatives. J. Mol. Catal. A Chem. 2013, 377, 173–179. [Google Scholar] [CrossRef]
  155. Cai, Z.; Shu, C.; Peng, Y. Magnetically recoverable nano-sized mesoporous solid acid. Effective catalysts for the synthesis of 1-amidoalkyl-2-naphthols. Monatsh. Chem. 2014, 145, 1681–1687. [Google Scholar] [CrossRef]
  156. Campisciano, V.; Giacalone, F.; Gruttadauria, M. Supported Ionic Liquids: A Versatile and Useful Class of Materials. Chem. Rec. 2017, 17, 918–938. [Google Scholar] [CrossRef] [PubMed]
  157. Riisagera, A.; Fehrmanna, R.; Haumannb, M.; Wasserscheidb, P. Supported ionic liquids: Versatile reaction and separation media. Top. Catal. 2006, 40, 91–102. [Google Scholar] [CrossRef]
  158. Valkenberg, M.H.; DeCastro, C.; Hölderich, W.F. Immobilisation of ionic liquids on solid supports. Green Chem. 2002, 4, 88–93. [Google Scholar] [CrossRef]
  159. Kefayati, H.; Kohankar, A.M.; Ramzanzadeh, N.; Shariati, S.; Bazargard, S.J. Synthesis of spiro[benzochromeno[2,3-d]pyrimidin-indolines] using Fe3O4@MCM-41-SO3H@[HMIm][HSO4] as a magnetically separable nanocatalyst. J. Mol. Liq. 2015, 209, 617–624. [Google Scholar] [CrossRef]
  160. Rostamizadeh, S.; Zekri, N. An efficient, one-pot synthesis of 2-amino-4H-chromenes catalyzed by (α-Fe2O3)-MCM-41-supported dual acidic ionic liquid as a novel and recyclable magnetic nanocatalyst. Res. Chem. Intermed. 2016, 42, 2329–2341. [Google Scholar] [CrossRef]
  161. Nakagaki, S.; Ferreira, G.; Marcalb, A.; Ciuffi, K. Metalloporphyrins Immobilized on Silica and Modified Silica as Catalysts in Heterogeneous Processes. Curr. Org. Synth. 2014, 11, 67–88. [Google Scholar] [CrossRef]
  162. Hajian, R.; Ehsanikhah, A. Manganese porphyrin immobilized on magnetic MCM-41 nanoparticles as an efficient and reusable catalyst for alkene oxidations with sodium periodate. Chem. Phys. Lett. 2018, 691, 146–154. [Google Scholar] [CrossRef]
  163. Rayati, S.; Abdolalian, P. Heterogenization of a molybdenum Schiff base complex as a magnetic nanocatalyst: An eco-friendly, efficient, selective and recyclable nanocatalyst for the oxidation of alkenes. C. R. Chim. 2013, 16, 814–820. [Google Scholar] [CrossRef]
  164. Fernandes, C.I.; Stenning, G.B.; Taylor, J.D.; Nunes, C.D.; D. Vaz, P. Helical Channel Mesoporous Materials with Embedded Magnetic Iron Nanoparticles: Chiral Recognition and Implications in Asymmetric Olefin Epoxidation. Adv. Synth. Catal. 2015, 357, 3127–3140. [Google Scholar] [CrossRef]
  165. Nikoorazm, M.; Ghorbani, F.; Ghorbani-Choghamarani, A.; Erfani, Z. Synthesis and characterization of a Pd(0) Schiff base complex anchored on magnetic nanoporous MCM-41 as a novel and recyclable catalyst for Suzuki and Heck reactions under green conditions. Chin. J. Catal. 2017, 38, 1413–1422. [Google Scholar] [CrossRef]
  166. Nikoorazm, M.; Ghorbani, F.; Ghorbani-Choghamarani, A.; Erfani, Z. Pd(0)-S-propyl-2-aminobenzothioate immobilized onto functionalized magnetic nanoporous MCM-41 as efficient and recyclable nanocatalyst for the Suzuki, Stille and Heck cross coupling reactions. Appl. Organomet. Chem. 2018, 32. n/a. [Google Scholar] [CrossRef]
  167. Chen, H.; Zhang, P.; Tan, W.; Jiang, F.; Tang, R. Palladium supported on amino functionalized magnetic MCM-41 for catalytic hydrogenation of aqueous bromate. RSC Adv. 2014, 4, 38743–38749. [Google Scholar] [CrossRef]
  168. Rostamizadeh, S.; Estiri, H.; Azad, M. Au anchored to (α-Fe2O3)-MCM-41-HS as a novel magnetic nanocatalyst for water-medium and solvent-free alkyne hydration. Catal. Commun. 2014, 57, 29–35. [Google Scholar] [CrossRef]
  169. Rostamizadeh, S.; Estiri, H.; Azad, M. Ullmann homocoupling of aryl iodides catalyzed by gold nanoparticles stabilized on magnetic mesoporous silica. J. Iran. Chem. Soc. 2017, 14, 1005–1010. [Google Scholar] [CrossRef]
  170. Bian, J.; Zhang, Q.; Zhang, P.; Feng, L.; Li, C. Supported Fe2O3 nanoparticles for catalytic upgrading of microalgae hydrothermal liquefaction derived bio-oil. Catal. Today 2017, 293–294, 159–166. [Google Scholar] [CrossRef]
  171. Miguel-Sancho, N.; Martinez, G.; Sebastian, V.; Malumbres, A.; Florea, I.; Arenal, R.; Ortega-Liebana, M.C.; Hueso, J.L.; Santamaria, J. Pumping Metallic Nanoparticles with Spatial Precision within Magnetic Mesoporous Platforms: 3D Characterization and Catalytic Application. ACS Appl. Mater. Interfaces 2017, 9, 41529–41536. [Google Scholar] [CrossRef] [Green Version]
  172. Shao, Y.B.; Jing, T.; Tian, J.Z.; Zheng, Y.J.; Shang, M.H. Characterization and optimization of mesoporous magnetic nanoparticles for immobilization and enhanced performance of porcine pancreatic lipase. Chem. Pap. 2015, 69, 1298–1311. [Google Scholar] [CrossRef]
  173. Xie, W.; Zang, X. Immobilized lipase on core–shell structured Fe3O4–MCM-41 nanocomposites as a magnetically recyclable biocatalyst for interesterification of soybean oil and lard. Food Chem. 2016, 194, 1283–1292. [Google Scholar] [CrossRef]
  174. Ulu, A.; Noma, S.A.A.; Koytepe, S.; Ates, B. Chloro-Modified Magnetic Fe3O4@MCM-41 Core–Shell Nanoparticles for L-Asparaginase Immobilization with Improved Catalytic Activity, Reusability, and Storage Stability. Appl. Biochem. Biotechnol. 2018. Ahead of Print. [Google Scholar] [CrossRef]
  175. Ulu, A.; Ozcan, I.; Koytepe, S.; Ates, B. Design of epoxy-functionalized Fe3O4@MCM-41 core–shell nanoparticles for enzyme immobilization. Int. J. Biol. Macromol. 2018, 115, 1122–1130. [Google Scholar] [CrossRef] [PubMed]
  176. Ulu, A.; Noma, S.A.A.; Koytepe, S.; Ates, B. Magnetic Fe3O4@MCM-41 core–shell nanoparticles functionalized with thiol silane for efficient l-asparaginase immobilization. Artif. Cells Nanomed. Biotechnol. 2018, 1–11. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Preparation of MCM-41.
Figure 1. Preparation of MCM-41.
Catalysts 08 00617 g001
Figure 2. Regioselective di-functionalization of MCM-41.
Figure 2. Regioselective di-functionalization of MCM-41.
Catalysts 08 00617 g002
Figure 3. Synthesis of 4H-chromene and 4H-benzo[b]pyrans.
Figure 3. Synthesis of 4H-chromene and 4H-benzo[b]pyrans.
Catalysts 08 00617 g003
Figure 4. Synthesis of quinazoline-2,4(1H,3H)-dione derivatives.
Figure 4. Synthesis of quinazoline-2,4(1H,3H)-dione derivatives.
Catalysts 08 00617 g004
Figure 5. Schematic illustration of reaction sequence and the structure of functionalized MCM-41. Reproduced with permission from [9]. Copyright 2014 Elsevier B.V.
Figure 5. Schematic illustration of reaction sequence and the structure of functionalized MCM-41. Reproduced with permission from [9]. Copyright 2014 Elsevier B.V.
Catalysts 08 00617 g005
Figure 6. Proposed plausible catalytic reaction pathway. Reproduced with permission from [9]. Copyright 2014 Elsevier B.V.
Figure 6. Proposed plausible catalytic reaction pathway. Reproduced with permission from [9]. Copyright 2014 Elsevier B.V.
Catalysts 08 00617 g006
Figure 7. Preparation of diamino-functionalised mesoporous material by anchoring 3-trimethoxysilylpropylethylenediamine moiety on calcine MCM-41. Reproduced with permission from [10]. Copyright 1999 Elsevier B.V.
Figure 7. Preparation of diamino-functionalised mesoporous material by anchoring 3-trimethoxysilylpropylethylenediamine moiety on calcine MCM-41. Reproduced with permission from [10]. Copyright 1999 Elsevier B.V.
Catalysts 08 00617 g007
Figure 8. Preparation of SO3H-MCM-41.
Figure 8. Preparation of SO3H-MCM-41.
Catalysts 08 00617 g008
Figure 9. SO3H-MCM-41-catalyzed synthesis of 2-amino-dihydropyrano[3,2-b]pyrane derivatives. Reproduced with permission from [20]. Copyright 2015 Springer.
Figure 9. SO3H-MCM-41-catalyzed synthesis of 2-amino-dihydropyrano[3,2-b]pyrane derivatives. Reproduced with permission from [20]. Copyright 2015 Springer.
Catalysts 08 00617 g009
Figure 10. Synthesis of dihydro-1H-pyrazolo[3,4-b]pyridines and 1H-pyrazolo[3,4-b]pyridines catalyzed by PTPSA@MCM-41. Reproduced with permission from [17]. Copyright 2017 Springer.
Figure 10. Synthesis of dihydro-1H-pyrazolo[3,4-b]pyridines and 1H-pyrazolo[3,4-b]pyridines catalyzed by PTPSA@MCM-41. Reproduced with permission from [17]. Copyright 2017 Springer.
Catalysts 08 00617 g010
Figure 11. Structure of PrSO3H-MCM-41 and MPrSO3H-MCM-41. Reproduced with permission from [18]. Copyright 2016 Elsevier B.V.
Figure 11. Structure of PrSO3H-MCM-41 and MPrSO3H-MCM-41. Reproduced with permission from [18]. Copyright 2016 Elsevier B.V.
Catalysts 08 00617 g011
Figure 12. Graphical representation. Reproduced with permission from [19]. Copyright 2016 Elsevier B.V.
Figure 12. Graphical representation. Reproduced with permission from [19]. Copyright 2016 Elsevier B.V.
Catalysts 08 00617 g012
Figure 13. DL-Alanine derivatized MCM-41.
Figure 13. DL-Alanine derivatized MCM-41.
Catalysts 08 00617 g013
Figure 14. MCM-41 functionalization.
Figure 14. MCM-41 functionalization.
Catalysts 08 00617 g014
Figure 15. Reaction mechanism to explain the enhanced efficiency of AP-E1 (a) in the Henry reaction relative to AP-T1 and AP-T2 (b). The presence of a significant number of spatially isolated silanol groups in AP-E1 leads to activation of the carbonyl group of benzaldehyde for nucleophilic attack. Reproduced with permission from [23]. Copyright 2007 Wiley-VCH Verlag GMBH & Co. KGAA, Weinheim.
Figure 15. Reaction mechanism to explain the enhanced efficiency of AP-E1 (a) in the Henry reaction relative to AP-T1 and AP-T2 (b). The presence of a significant number of spatially isolated silanol groups in AP-E1 leads to activation of the carbonyl group of benzaldehyde for nucleophilic attack. Reproduced with permission from [23]. Copyright 2007 Wiley-VCH Verlag GMBH & Co. KGAA, Weinheim.
Catalysts 08 00617 g015
Figure 16. Proposed mechanism for the aldol condensation of acetone and para-substituted benzaldehydes on amino-silica cooperative catalyst. Reproduced with permission from [24]. Copyright 2015 American Chemical Society.
Figure 16. Proposed mechanism for the aldol condensation of acetone and para-substituted benzaldehydes on amino-silica cooperative catalyst. Reproduced with permission from [24]. Copyright 2015 American Chemical Society.
Catalysts 08 00617 g016
Figure 17. MCM-PP (3) catalyzed synthesis of benzothiazoles. Reproduced with permission from [25]. Copyright 2015 American Chemical Society.
Figure 17. MCM-PP (3) catalyzed synthesis of benzothiazoles. Reproduced with permission from [25]. Copyright 2015 American Chemical Society.
Catalysts 08 00617 g017
Figure 18. Syntheses of bifunctional mesoporous silica nanoparticles having sulfonic acid groups on the internal surface and organic amine groups on the external surface. Reproduced with permission from [26]. Copyright 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 18. Syntheses of bifunctional mesoporous silica nanoparticles having sulfonic acid groups on the internal surface and organic amine groups on the external surface. Reproduced with permission from [26]. Copyright 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 08 00617 g018
Figure 19. Preparation of bifunctional silica particles 13. Reproduced with permission from [27]. Copyright 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 19. Preparation of bifunctional silica particles 13. Reproduced with permission from [27]. Copyright 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 08 00617 g019
Figure 20. MSNs 1719 bearing bipyridyl complexed Pd. Reproduced with permission from [28]. Copyright 2013 Royal society of Chemistry.
Figure 20. MSNs 1719 bearing bipyridyl complexed Pd. Reproduced with permission from [28]. Copyright 2013 Royal society of Chemistry.
Catalysts 08 00617 g020
Figure 21. Schematic representation of a bifunctional pentafluorophenyl propyl (PFP)/DAT-MSN. Reproduced with permission from [29]. Copyright 2011 American Chemical Society.
Figure 21. Schematic representation of a bifunctional pentafluorophenyl propyl (PFP)/DAT-MSN. Reproduced with permission from [29]. Copyright 2011 American Chemical Society.
Catalysts 08 00617 g021
Figure 22. Synthesis steps of Cu-salen-MCM-41. Reproduced with permission from [30]. Copyright 2014 Springer.
Figure 22. Synthesis steps of Cu-salen-MCM-41. Reproduced with permission from [30]. Copyright 2014 Springer.
Catalysts 08 00617 g022
Figure 23. Synthesis of MCM-41@Serine@Cu(II). Reproduced with permission from [31]. Copyright 2018 Springer.
Figure 23. Synthesis of MCM-41@Serine@Cu(II). Reproduced with permission from [31]. Copyright 2018 Springer.
Catalysts 08 00617 g023
Figure 24. Synthesis process of Cu(II)-Et-S@mcm-41 in the three separate steps. Reproduced with permission from [32]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 24. Synthesis process of Cu(II)-Et-S@mcm-41 in the three separate steps. Reproduced with permission from [32]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 08 00617 g024
Figure 25. Preparation of MCM-41-1, 10-phen–CuI. Reproduced with permission from [33]. Copyright 2016 Royal Society of Chemistry
Figure 25. Preparation of MCM-41-1, 10-phen–CuI. Reproduced with permission from [33]. Copyright 2016 Royal Society of Chemistry
Catalysts 08 00617 g025
Figure 26. Tricopper Cluster Complex CuIICuIICuII(7-N-Etppz)4+.
Figure 26. Tricopper Cluster Complex CuIICuIICuII(7-N-Etppz)4+.
Catalysts 08 00617 g026
Figure 27. Schematic pattern of immobilization of DFP and related complexes. Reproduced with permission from [35]. Copyright 2017 Springer.
Figure 27. Schematic pattern of immobilization of DFP and related complexes. Reproduced with permission from [35]. Copyright 2017 Springer.
Catalysts 08 00617 g027
Figure 28. Synthesis of Cu- Schiff base MCM-41 complex. Reproduced with permission from [36]. Copyright 2016 Springer.
Figure 28. Synthesis of Cu- Schiff base MCM-41 complex. Reproduced with permission from [36]. Copyright 2016 Springer.
Catalysts 08 00617 g028
Figure 29. MCM-41 anchored to palladium bipyridyl complex-catalyzed. Reproduced with permission from [41]. Copyright 2009 Elsevier B.V.
Figure 29. MCM-41 anchored to palladium bipyridyl complex-catalyzed. Reproduced with permission from [41]. Copyright 2009 Elsevier B.V.
Catalysts 08 00617 g029
Figure 30. Operation of catalyst in presence of nucleophile salt. Reproduced with permission from. [42] Copyright 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 30. Operation of catalyst in presence of nucleophile salt. Reproduced with permission from. [42] Copyright 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 08 00617 g030
Figure 31. Synthesis of palladium anchored on functionalized MCM-41. steps. Reproduced with permission from [43]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 31. Synthesis of palladium anchored on functionalized MCM-41. steps. Reproduced with permission from [43]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 08 00617 g031
Figure 32. Possible surface Ti species. Reproduced with permission from [44]. Copyright 2016 Elsevier B.V.
Figure 32. Possible surface Ti species. Reproduced with permission from [44]. Copyright 2016 Elsevier B.V.
Catalysts 08 00617 g032
Figure 33. Preparation of M–PrPPh2Rh catalyst. Reproduced with permission from [45]. Copyright 2008 Springer.
Figure 33. Preparation of M–PrPPh2Rh catalyst. Reproduced with permission from [45]. Copyright 2008 Springer.
Catalysts 08 00617 g033
Figure 34. Zr(IV)-salen-MCM-41 synthesis to catalyze fructose. Reproduced with permission from [47]. Copyright 2018 Elsevier B.V.
Figure 34. Zr(IV)-salen-MCM-41 synthesis to catalyze fructose. Reproduced with permission from [47]. Copyright 2018 Elsevier B.V.
Catalysts 08 00617 g034
Figure 35. Synthesis of Materials MCM-Pr, MCM-L, and MCM-C1 to MCM-C5. Reproduced with permission from [48]. Copyright 2015 American Chemical Society.
Figure 35. Synthesis of Materials MCM-Pr, MCM-L, and MCM-C1 to MCM-C5. Reproduced with permission from [48]. Copyright 2015 American Chemical Society.
Catalysts 08 00617 g035
Figure 36. Preparation of ordered MCM-41 mesoporous silica containing Cobalt, Niquel, Iron, and Zinc Schiff base complex.
Figure 36. Preparation of ordered MCM-41 mesoporous silica containing Cobalt, Niquel, Iron, and Zinc Schiff base complex.
Catalysts 08 00617 g036
Figure 37. Synthesis of MCM-41 supported Mn-Schiff base complexes. Reproduced with permission from [52]. Copyright 2015 Royal Society of Chemistry.
Figure 37. Synthesis of MCM-41 supported Mn-Schiff base complexes. Reproduced with permission from [52]. Copyright 2015 Royal Society of Chemistry.
Catalysts 08 00617 g037
Figure 38. Synthesis of the Mn(TF5 PP)-MCM-41 catalyst. Reproduced with permission from [53]. Copyright 2015 Elsevier B.V.
Figure 38. Synthesis of the Mn(TF5 PP)-MCM-41 catalyst. Reproduced with permission from [53]. Copyright 2015 Elsevier B.V.
Catalysts 08 00617 g038
Figure 39. Methodologies of impregnation for Me2Si(Ind)2ZrCl2/NMCM-41. Reproduced with permission from [54]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 39. Methodologies of impregnation for Me2Si(Ind)2ZrCl2/NMCM-41. Reproduced with permission from [54]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 08 00617 g039
Figure 40. Schematic procedure for the surface modification of MCM-41 silica and the subsequent immobilization of 12-tungstophosphoric acid (HPW). Reproduced with permission from [76]. Copyright 2010 Elsevier B.V.
Figure 40. Schematic procedure for the surface modification of MCM-41 silica and the subsequent immobilization of 12-tungstophosphoric acid (HPW). Reproduced with permission from [76]. Copyright 2010 Elsevier B.V.
Catalysts 08 00617 g040
Figure 41. Schematic representation of vanadium-substituted phosphotungstic acids immobilized on amine-functionalized mesoporous MCM-41. Reproduced with permission from [78]. Copyright 2014 Elsevier B.V.
Figure 41. Schematic representation of vanadium-substituted phosphotungstic acids immobilized on amine-functionalized mesoporous MCM-41. Reproduced with permission from [78]. Copyright 2014 Elsevier B.V.
Catalysts 08 00617 g041
Figure 42. Functionalization and impregnation of MCM-41 silica. Conditions: (a) 6 M HCl reflux; (b) dried in oven at 100 °C and refluxed in toluene with attached Dean–Stark trap; (c) reflux in toluene with APTES; (d) suspend in DCM with 2 equiv. HNO3; (e) suspend in acetonitrile solution of [γ-SiW10O34(H2O)2]4−. Reproduced with permission from [80]. Copyright 2014 Elsevier B.V.
Figure 42. Functionalization and impregnation of MCM-41 silica. Conditions: (a) 6 M HCl reflux; (b) dried in oven at 100 °C and refluxed in toluene with attached Dean–Stark trap; (c) reflux in toluene with APTES; (d) suspend in DCM with 2 equiv. HNO3; (e) suspend in acetonitrile solution of [γ-SiW10O34(H2O)2]4−. Reproduced with permission from [80]. Copyright 2014 Elsevier B.V.
Catalysts 08 00617 g042
Figure 43. Schematic diagram for the pyridine functionalization of MCM-41 mesoporous silica and immobilization of heteropolyacid (HPA). Reproduced with permission from [81]. Copyright 2015 Royal Society of Chemistry.
Figure 43. Schematic diagram for the pyridine functionalization of MCM-41 mesoporous silica and immobilization of heteropolyacid (HPA). Reproduced with permission from [81]. Copyright 2015 Royal Society of Chemistry.
Catalysts 08 00617 g043
Figure 44. Schematic representation of POM@MOF-199@MCM-41 hybrid system and dibenzothiophene (DBT) oxidative desulfurization. Reproduced with permission from [84]. Copyright 2016 Elsevier B.V. BTC: 1,3,5-benzenetricarboxylic acid.
Figure 44. Schematic representation of POM@MOF-199@MCM-41 hybrid system and dibenzothiophene (DBT) oxidative desulfurization. Reproduced with permission from [84]. Copyright 2016 Elsevier B.V. BTC: 1,3,5-benzenetricarboxylic acid.
Catalysts 08 00617 g044
Figure 45. One-pot synthesis route for Pd/MCM-41. Reproduced with permission from [88]. Copyright 2015 Royal Society of Chemistry.
Figure 45. One-pot synthesis route for Pd/MCM-41. Reproduced with permission from [88]. Copyright 2015 Royal Society of Chemistry.
Catalysts 08 00617 g045
Figure 46. Synthetic route for preparation of MCM-41-Crown.Pd. Reproduced with permission from [94]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 46. Synthetic route for preparation of MCM-41-Crown.Pd. Reproduced with permission from [94]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 08 00617 g046
Figure 47. Schematic description for preparation of Au@HS-MCM nanocomposite. Reproduced with permission from [91]. Copyright 2015 Elsevier B.V.
Figure 47. Schematic description for preparation of Au@HS-MCM nanocomposite. Reproduced with permission from [91]. Copyright 2015 Elsevier B.V.
Catalysts 08 00617 g047
Figure 48. Schematic representation of A3 coupling supported in Au-SH/MCM-41. Reproduced with permission from [91]. Copyright 2015 Elsevier B.V.
Figure 48. Schematic representation of A3 coupling supported in Au-SH/MCM-41. Reproduced with permission from [91]. Copyright 2015 Elsevier B.V.
Catalysts 08 00617 g048
Figure 49. The probable structure of the NH2-MCM-41 material before and after immersion in HAuCl4 solution. Reproduced with permission from [105]. Copyright 2001 Royal society of chemistry.
Figure 49. The probable structure of the NH2-MCM-41 material before and after immersion in HAuCl4 solution. Reproduced with permission from [105]. Copyright 2001 Royal society of chemistry.
Catalysts 08 00617 g049
Figure 50. The possible mechanism of amine functionalization and subsequent introduction of the metals into Si-MCM-41. Reproduced with permission from [107]. Copyright 2006 Elsevier B.V.
Figure 50. The possible mechanism of amine functionalization and subsequent introduction of the metals into Si-MCM-41. Reproduced with permission from [107]. Copyright 2006 Elsevier B.V.
Catalysts 08 00617 g050
Figure 51. The formation of charge transfer excited state with tetrahe- drally coordinated Ti oxide moiety within Ti-MCM-41 by UV light irradiation and its application for the synthesis of nano-sized metal particles. Reproduced with permission from [113]. Copyright 2009 Springer.
Figure 51. The formation of charge transfer excited state with tetrahe- drally coordinated Ti oxide moiety within Ti-MCM-41 by UV light irradiation and its application for the synthesis of nano-sized metal particles. Reproduced with permission from [113]. Copyright 2009 Springer.
Catalysts 08 00617 g051
Figure 52. Schematic representation of NiO/NH2-MCM-41. Reproduced with permission from [92]. Copyright 2016 Royal Society of Chemistry.
Figure 52. Schematic representation of NiO/NH2-MCM-41. Reproduced with permission from [92]. Copyright 2016 Royal Society of Chemistry.
Catalysts 08 00617 g052
Figure 53. Proposed mechanism for the alcohol oxidation-Wittig olefination showing alcohol oxidation and olefination as coupled processes. Reproduced with permission from [121]. Copyright 2017 Royal Society of Chemistry.
Figure 53. Proposed mechanism for the alcohol oxidation-Wittig olefination showing alcohol oxidation and olefination as coupled processes. Reproduced with permission from [121]. Copyright 2017 Royal Society of Chemistry.
Catalysts 08 00617 g053
Figure 54. Zirconia-modified MCM-41 supported Ru-La catalyst. Reproduced with permission from [93] Copyright 2014 Elsevier B.V.
Figure 54. Zirconia-modified MCM-41 supported Ru-La catalyst. Reproduced with permission from [93] Copyright 2014 Elsevier B.V.
Catalysts 08 00617 g054
Figure 55. Schematic illustration of alkali metal cations-anchored mesoporous silica with catalytic centres. Reproduced with permission from [124]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 55. Schematic illustration of alkali metal cations-anchored mesoporous silica with catalytic centres. Reproduced with permission from [124]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 08 00617 g055
Figure 56. Synthetic route for the preparation of Fe3O4@nSiO2@MCM-41-MnP and Fe3O4@nSiO2@MCM-41(E)-MnP catalysts. Reproduced with permission from [149]. Copyright 2016 Elsevier B.V. TEOS: tetraethyl orthosilicate; MnP: manganese porphyrin.
Figure 56. Synthetic route for the preparation of Fe3O4@nSiO2@MCM-41-MnP and Fe3O4@nSiO2@MCM-41(E)-MnP catalysts. Reproduced with permission from [149]. Copyright 2016 Elsevier B.V. TEOS: tetraethyl orthosilicate; MnP: manganese porphyrin.
Catalysts 08 00617 g056
Figure 57. Preparation of (α-Fe2O3)-MCM-41-SO3H. Reproduced with permission from [153]. Copyright 2012 Elsevier B.V.
Figure 57. Preparation of (α-Fe2O3)-MCM-41-SO3H. Reproduced with permission from [153]. Copyright 2012 Elsevier B.V.
Catalysts 08 00617 g057
Figure 58. Synthesis of Fe3O4@MCM-41-SO3H by thiol oxidation (R:CH2CH2SH). Reproduced with permission from [154]. Copyright 2013 Elsevier B.V.
Figure 58. Synthesis of Fe3O4@MCM-41-SO3H by thiol oxidation (R:CH2CH2SH). Reproduced with permission from [154]. Copyright 2013 Elsevier B.V.
Catalysts 08 00617 g058
Figure 59. Synthesis of SO3H-MCM-41@NiFe2O4 by grafting and oxidation of MCM-41@NiFe2O4. Reproduced with permission from [155]. Copyright 2014 Springer.
Figure 59. Synthesis of SO3H-MCM-41@NiFe2O4 by grafting and oxidation of MCM-41@NiFe2O4. Reproduced with permission from [155]. Copyright 2014 Springer.
Catalysts 08 00617 g059
Figure 60. Synthesis of 2-amino-4H chromenes in the presence of (α-Fe2O3)-MCM-41-DAIL catalyst. Reproduced with permission from [160]. Copyright 2016 Springer.
Figure 60. Synthesis of 2-amino-4H chromenes in the presence of (α-Fe2O3)-MCM-41-DAIL catalyst. Reproduced with permission from [160]. Copyright 2016 Springer.
Catalysts 08 00617 g060
Figure 61. Schematic diagram for preparation of Fe3O4@MCM-41-Im@MnPor. Reproduced with permission from [162]. Copyright 2018 Elsevier B.V.
Figure 61. Schematic diagram for preparation of Fe3O4@MCM-41-Im@MnPor. Reproduced with permission from [162]. Copyright 2018 Elsevier B.V.
Catalysts 08 00617 g061
Figure 62. Synthesis and derivatization of magMSh material with bipyridine ligand and Mo(II) center. Reproduced with permission from [164]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 62. Synthesis and derivatization of magMSh material with bipyridine ligand and Mo(II) center. Reproduced with permission from [164]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 08 00617 g062
Figure 63. Synthesis of Fe3O4@MCM-41@Pd-SPATB. Reproduced with permission from [165]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 63. Synthesis of Fe3O4@MCM-41@Pd-SPATB. Reproduced with permission from [165]. Copyright 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Catalysts 08 00617 g063
Figure 64. Preparation of (α-Fe2O3)-MCM-41-SH-Au. Reproduced with permission from [169]. Copyright 2014 Elsevier B.V.
Figure 64. Preparation of (α-Fe2O3)-MCM-41-SH-Au. Reproduced with permission from [169]. Copyright 2014 Elsevier B.V.
Catalysts 08 00617 g064
Figure 65. Pt nanoparticles pumped within the channels of the magnetic MCM-41 nanospheres. Reproduced with permission from [171]. Copyright 2017 American Chemical Society.
Figure 65. Pt nanoparticles pumped within the channels of the magnetic MCM-41 nanospheres. Reproduced with permission from [171]. Copyright 2017 American Chemical Society.
Catalysts 08 00617 g065
Figure 66. Schematic representation of Fe3O4-MCM-41 nanoparticles formation and enzyme immobilization. Reproduced with permission from [172]. Copyright 2015 Springer.
Figure 66. Schematic representation of Fe3O4-MCM-41 nanoparticles formation and enzyme immobilization. Reproduced with permission from [172]. Copyright 2015 Springer.
Catalysts 08 00617 g066
Figure 67. Synthesis of the magnetic MCM-41 immobilized Candida rugosa lipase. Reproduced with permission from [173]. Copyright 2016 Springer.
Figure 67. Synthesis of the magnetic MCM-41 immobilized Candida rugosa lipase. Reproduced with permission from [173]. Copyright 2016 Springer.
Catalysts 08 00617 g067
Figure 68. Schematic illustrates stapes of magnetic Fe3O4@MCM-41-Cl nanoparticle preparation, modification, and the immobilization of L-ASNase. Reproduced with permission from [174]. Copyright 2018 Springer.
Figure 68. Schematic illustrates stapes of magnetic Fe3O4@MCM-41-Cl nanoparticle preparation, modification, and the immobilization of L-ASNase. Reproduced with permission from [174]. Copyright 2018 Springer.
Catalysts 08 00617 g068
Figure 69. Schematic depiction of magnetic Fe3O4@MCM-41 sieves preparation, epoxy modification and L-ASNase immobilization. Reproduced with permission from [175]. Copyright 2018 Elsevier B.V.
Figure 69. Schematic depiction of magnetic Fe3O4@MCM-41 sieves preparation, epoxy modification and L-ASNase immobilization. Reproduced with permission from [175]. Copyright 2018 Elsevier B.V.
Catalysts 08 00617 g069

Share and Cite

MDPI and ACS Style

Martínez-Edo, G.; Balmori, A.; Pontón, I.; Martí del Rio, A.; Sánchez-García, D. Functionalized Ordered Mesoporous Silicas (MCM-41): Synthesis and Applications in Catalysis. Catalysts 2018, 8, 617. https://doi.org/10.3390/catal8120617

AMA Style

Martínez-Edo G, Balmori A, Pontón I, Martí del Rio A, Sánchez-García D. Functionalized Ordered Mesoporous Silicas (MCM-41): Synthesis and Applications in Catalysis. Catalysts. 2018; 8(12):617. https://doi.org/10.3390/catal8120617

Chicago/Turabian Style

Martínez-Edo, Gabriel, Alba Balmori, Iris Pontón, Andrea Martí del Rio, and David Sánchez-García. 2018. "Functionalized Ordered Mesoporous Silicas (MCM-41): Synthesis and Applications in Catalysis" Catalysts 8, no. 12: 617. https://doi.org/10.3390/catal8120617

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop