Mesoporous materials: versatile supports in heterogeneous catalysis for liquid phase catalytic transformations

Nabanita Pal *a and Asim Bhaumik *b
aSurface Physics Division, Saha Institute of Nuclear Physics, 1/AF Bidhannagar, Kolkata-700 064, India. E-mail: nabanita.pal@saha.ac.in; pal_nabamsc@yahoo.co.in
bDepartment of Materials Science, Indian Association for the Cultivation of Science, Jadavpur, Kolkata-700 032, India. E-mail: msab@iacs.res.in

Received 24th October 2014 , Accepted 9th February 2015

First published on 9th February 2015


Abstract

Due to their unprecedented intrinsic structural features, like tunable pore diameter of nanoscale dimensions, huge BET surface areas and good flexibility to recognize/accommodate various functional groups and metals onto the surface, an inevitable linkage of nanoporous materials with catalysis has been built-up over the past few decades. As a result of which, a huge number of communications and articles dealing with these materials with nanoscale porosity have come to light. In this review, our objective is to provide a comprehensive overview of mesoporous solids, the most remarkable member of the nanoporous family of materials, the general strategy for their syntheses and application of functionalized porous materials in liquid phase catalytic reactions. In the latter part, the role of catalytic centres in various organic transformations over these functionalized mesoporous materials and their economical, environmental and industrial aspects are described in detail.


image file: c4ra13077d-p1.tif

Nabanita Pal

Nabanita Pal is presently working as a Postdoctoral Research Fellow at Saha Institute of Nuclear Physics in Kolkata, India. She received her Bachelor (BSc) and Master (MSc) degrees in Chemistry from the University of Calcutta in 2004 and 2006, respectively. Then she joined the Indian Association for the Cultivation of Science in Kolkata in 2007 for her doctoral research. She obtained a PhD degree in 2012 with several publications on catalysis over nanoporous materials. During 2012–2014, she was a Dr D. S. Kothari postdoctoral fellow in Jadavpur University, India and later a postdoctoral fellow in Sungkyunkwan University, South Korea. Her research interest mainly focused on nanoporous materials and their applications in catalysis, biosensing and adsorption, etc.

image file: c4ra13077d-p2.tif

Asim Bhaumik

Asim Bhaumik is currently working as a senior professor at the Department of Materials Science, Indian Association for the Cultivation of Science (IACS), Kolkata, India. He received his PhD from NCL, Pune in 1997. After his postdoctoral research as a JSPS fellow at the University of Tokyo, Japan during 1997–1999 and as an associate researcher at Toyota Central R & D Labs. Inc., Japan during 1999–2001, in 2001 he joined IACS, India, as a faculty member. His research focuses on several aspects of energy, the environment and biomedical science, including designing porous nanomaterials for adsorption, gas storage, catalysis, sensing, photocatalysis, DSSCs and drug delivery applications. He is the coauthor of 270 research publications and the inventor of 13 patents. He is a board member of several journals and a Fellow of the Royal Society of Chemistry, UK.


1. Introduction

Today, worldwide over 60% of industrially important chemical products are produced by different chemical process and the majority of these processes are accelerated by using suitable catalysts.1 Wilhelm Ostwald rightly said that “there is probably no chemical reaction which cannot be influenced catalytically”.2 Catalysts play a crucial role in the manufacturing process for the synthesis of fine chemicals, pharmaceuticals, commodity chemicals, energy resources, fuels, biofuels and so on. The extraordinary advantage of heterogeneous catalysis over homogeneous catalysis is the separation, recoverability and recycling capability, which leads to compliance to green synthetic pathways involving minimum environmental pollution.

Porous nanostructured materials with a huge BET surface area are suitable for catalytic reactions and in most cases the nature of the catalytically active components (e.g. metals, organic functional groups etc.) is most crucial in heterogeneous catalysts. Hence, researchers are very keen to design hierarchical porous nanomaterials bearing catalytically active centers at their surface. A number of reviews by Pérez-Ramírez et al.,3 Pal et al.,4 Lopez-Orozco et al.5 and others on the synthesis and applications of porous solids have been reported in recent times. The introduction of microporous solids, particularly zeolites in industrial manufacturing processes, has brought an enormous economical and environmental revolution.6 These microporous materials play a key role in shape-selective catalysis on the industrial scale.7 However, their small pore opening has motivated researchers to focus their attention on designing materials with pores of still larger dimensions in the nanoscale region. This ultimately leads to the discovery of ordered mesoporous silicas through supramolecular templating pathways.8 Later, other mesoporous supports like mesoporous oxides, phosphates, carbons, etc. were invented and they play a significant role in industrial manufacturing processes for fine chemicals. Liquid phase catalytic transformations in heterogeneous media over mesoporous solids are more favorable, and they lead to minimum diffusion as compared to other materials belonging to the porous categories.

A number of comprehensive reviews in the literature related to the catalytic application of mesoporous solids by Corma,9 Taguchi et al.,10 Perego et al.11 and Martín-Aranda et al.12 have arisen, which mainly discussed catalysis over ordered mesoporous silica based materials. The other mesoporous materials like oxides, phosphates, carbons, polymers and organic–inorganic hybrid periodic mesoporous organosilicas,13 which are of parallel importance as far as industrial catalysis is concerned, have had less emphasis in these reviews. The present review illustrates an overall picture of heterogeneous catalytic reactions mediated by a wide range of mesoporous materials invented so far and their environmental aspects for clean chemical synthesis.

2. What are mesoporous materials?

Traditional porous materials, the more commonly used term being ‘nanoporous’ materials, are defined as a continuous and solid network material filled through voids of ‘nanoscale pores’ of the order of say 100 nm or smaller. The International Union of Pure and Applied Chemistry (IUPAC) have classified nanoporous solids into three categories according to the pore size they possess: microporous with a pore diameter of less than 2 nm; mesoporous having a pore diameter in the range of 2–50 nm; and macroporous with a greater than 50 nm pore size.14 There are some recent reports on catalytic activity over macroporous materials,15,16 but owing to the low surface area of nanoporous materials of this category they are not in such high demand for liquid phase heterogeneous catalysis. Also many microporous zeolites, aluminophosphates, organic–inorganic hybrid phosphonates, and metal organic frameworks (MOFs) are quite useful for environmentally benign green catalytic reactions,17,18 but those are out of the scope of this review, where our discussion will be focused on mesoporous solids and their potential in heterogeneous catalysis.

The history of mesoporous molecular sieves (where, ‘meso’ the Greek prefix, meaning “in between”, adopted by IUPAC for dimensions of pores typically between 2 and 50 nm) begins in the last decades of the twentieth century when materials like MCM (Mobil Composition of Matter)-type materials8 were successfully synthesized by Mobil scientists. The discovery of the M41S family of ordered mesoporous materials with a pore dimension of 2–10 nm, high surface area (∼1000 m2 g−1) and large pore volume (∼1.0 cm3 g−1), by using quaternary alkylammonium surfactants (e.g. cetyltrimethylammonium bromide, CH3(CH2)15N(CH3)3+Br) as a template, can be considered one of the most important milestones in the history of the porous materials world, resulting in huge expectations for their application as heterogeneous catalysts.19,20

Among the ordered mesoporous materials discovered initially, mesoporous silicas, e.g. MCM-41, MCM-48, MCM-50, FSM-16,21 SBA-15 (ref. 22) etc., are well-known. Though there is some difference in the synthesis conditions and structural properties of these silicas, the basic strategy of synthesis is similar in all cases, and is based on the supramolecular self-assembly of the surfactants (or templates).23 Rather than a small organic molecule as the single molecule template (SDA), long chain cationic surfactant molecules, like CTAB (cetyltrimethylammonium bromide), CPC (cetylpyridinium chloride) etc. or anionic surfactants like SDS (sodium dodecyl sulphate) are used during the synthesis of highly ordered silica based materials.8 In solution, co-operative self-assembly of surfactant molecules takes place at concentrations higher than the CMC when a spherical or rod shaped geometry is obtained around which hydrolyzed inorganic silica precursors (say, tetraethyl orthosilicate or TEOS) arrange themselves in order to form organic–inorganic metal-template composites in a highly basic medium. Through ageing for a particular time further condensation occurs and the composites take on an ordered arrangement such as cubic or hexagonal, etc. After removal of the surfactants from the composite materials through calcination or solvent-extraction we can obtain the final mesoporous solids free from organic templates. A detailed synthesis strategy for mesoporous silica and other mesoporous solids using a surfactant-templated route based on the selection of different types of surfactants has been extended in the review published recently.4 A typical synthesis scheme of highly ordered 2D hexagonal pure mesoporous silica MCM-41 mediated by a surfactant is shown in Fig. 1. Depending upon the type of surfactant used, the surfactant–silica ratio and the pH of the solution, the structure and mesopore size of the silica can be varied from hexagonal (e.g. MCM-41), to cubic (e.g. MCM-48), to lamellar (e.g. MCM-50). Mesoporous SBA-15 type materials are another family of silica with large mesopores, synthesized using non-ionic block copolymer surfactants under highly acidic conditions.22 In addition to the co-operative pathways, nanocasting using already formed ordered mesoporous materials as hard templates has been developed to synthesize mesoporous materials.24 This method of nanocasting is highly effective for the synthesis of other non-siliceous porous oxide and carbon materials with ordered pore arrangements which are difficult to prepare directly by the surfactant-assisted route.25


image file: c4ra13077d-f1.tif
Fig. 1 Surfactant assisted synthesis of mesoporous silica with a 2D ordered hexagonal pore arrangement.

Non-siliceous mesoporous materials like oxides, mixed oxides,26 metal sulphides,27 metal phosphates,28 polymers,29 carbons30 and carbon nitrides31 have also been synthesized successfully using surfactant templating routes. The long range applications of these materials extend from gas adsorption, ion-exchange, magnetism, sensing, catalysis and electrochemistry to many versatile fields of research. Some of these mesoporous materials are purely inorganic; some are organic–inorganic hybrid materials while the others are completely organic based. Types of mesoporous solids commonly known in these three categories are shown in Table 1.

Table 1 Possible types of mesoporous materials with examples
Framework composition Type of materials Examples
Purely inorganic Mesoporous silicas MCM-41/MCM-48,8 SBA-15 (ref. 22) etc.
Metal containing mesoporous silicas Ti-MCM-41, Zn–silica etc.
Mesoporous metal oxides and mixed metal oxides TiO2, AlO2, ZrO2, ZnTiO3 etc.
Mesoporous metallophosphates Silicotitanium phosphate, silicoalumino phosphate etc.
Organic–inorganic hybrid Periodic mesoporous organosilicas (PMOs) Various metal containing PMOs etc.
Metal oxophenylphosphates Iron phosphonate, chromium phosphonate etc.
Purely organic Mesoporous carbons CMK-3
Mesoporous polymers Triazine based polymer, porphyrine based polymer


3. Catalysis: homogeneous vs. heterogeneous

The phenomenon called ‘catalysis’ has been known since very ancient centuries, even though people knew nothing about the underlying mechanism and chemical process involved in it. The making of soap, the fermentation of wine to vinegar, and the leavening of bread are all processes involving catalysis. The name ‘catalysis’, first proposed in 1835 by Swedish chemist Jöns Jakob Berzelius (1779–1848), was taken from the Greek words kata, which means ‘down’ and lyein, meaning ‘loosen’. In a short paper summarising his concepts on ‘catalysis’ as a new force, Berzelius wrote: “It is, then, proved that several simple or compound bodies, soluble and insoluble, have the property of exercising on other bodies an action very different from chemical affinity. By means of this action they produce, in these bodies, decompositions of their elements and different recombination of these same elements to which they remain indifferent.” He called this new force a ‘catalytic force’.32 The actual basic role of a catalyst is to offer an alternative pathway of lower activation energy than the respective uncatalyzed reaction, resulting in a faster reaction rate without hampering the overall thermodynamics of the reaction. Enzymes are the most common and efficient catalysts found in nature. Not a single reaction in our body is possible without the aid of these biocatalysts. They have also been used as catalysts in food industries for many years. In the case of laboratory and industrial use, catalytic oxidation using oxygen, hydrogenation using metal catalysts, and many well known reactions are worthy of mention. Day by day the imperative role of catalysis in the manufacturing of the majority of chemicals used by our society is increasing rapidly. According to the survey report, in 2005, catalytic processes generated about $900 billion worth of products worldwide.

Catalytic reactions are usually categorized as either homogeneous or heterogeneous. Biocatalysis (enzymatic reactions) is often considered as a separate group. A homogeneous catalysis reaction is one where both the catalyst, the reactants on which it works and the products, are all in the same phase (solid, liquid, or gas), generally in the liquid phase. The first formal example of this type is the reaction involving the decomposition of starch into glucose in boiling water in the presence of a sulphuric acid catalyst studied by the Russian chemist Gottlieb Sigismund Constantin Kirchhoff (1764–1833) in 1812. Both the sulfuric acid and the starch were in aqueous solution during the reaction.33 Organometallic catalysts are another example of homogeneous catalysts.34 On the other hand, heterogeneous catalysis is that where the catalyst is generally solid and remains as a distinct phase from the liquid or gaseous reaction media on which it acts. Most of the reactions that occur in industry are driven by heterogeneous catalysts. The role of finely divided iron in the Haber process for synthesis of ammonia or the role of vanadium oxide for the production of sulphuric acid in the contact process, etc. are well-known examples of this type. In fact, with the production of sulphuric acid in the “contact” process solved by Knietsch in the year 1898 in Europe, the idea of industrial catalysis was developed.35 Today porous materials are recognized as one of the most important materials in heterogeneous catalysis and they play the major role in industrial fine chemical synthesis.9

Homogeneous catalysis has some important properties like high selectivity of the products and good accessibility to catalytically active sites, etc. but with increasing public awareness regarding environmental issues, heterogeneous catalysis has become much more important over homogeneous catalysis.36 This is because some major problems like corrosion, toxicity, the difficulty of catalyst separation, recovery, regeneration and reuse, high cost and the creation of huge amounts of solid waste, etc. in the case of homogeneous catalysis, make it unsuitable for application in industrial purposes in this environmentally conscious and economically pressured world. Although homogeneous catalysts are used in some industries like the food, fine chemical, pharmaceutical and agrochemical industry etc., in this century heterogeneous catalysis has been proven to be the ultimate goal of industrial catalysis and in engineering fields. The industrial processes which are highly dependent on homogeneous catalysts are recommended to be used only by ‘heterogenization’, which means immobilization of the catalysts on a solid insoluble support which possesses all the properties of the homogeneous counterpart, and also the advantage of recyclability like in heterogeneous catalysis.37

Catalytic reaction on a heterogeneous support involves fundamental steps like diffusion and adsorption of the reactants onto a solid surface, successful reaction on it, then desorption of the products in bulk with regeneration of the catalyst for the next cycle. A schematic representation of a heterogeneous catalytic cycle taking place through different steps is shown in Fig. 2. Since the surface of catalyst plays a pivotal role in the related process, materials having high surface area are more desirable than nonporous materials to accomplish a significant part in industrial catalysis.1,9 Although most of the chemical industries generally use zeolites or activated carbon based microporous materials as a stable support for catalysis, the number of fascinating organic transformations successfully carried out over different mesoporous supports is growing very rapidly. In light of this a systematic brief discussion on various mesopore mediated important heterogeneous catalytic transformations under liquid phase reaction conditions has been made and illustrated in the following part of this review.


image file: c4ra13077d-f2.tif
Fig. 2 Steps of a heterogeneous catalytic cycle over a mesoporous solid.

4. Heterogeneous catalysis over mesoporous materials

The most important feature of mesoporous materials for which they show great potential for catalytic applications, is the possibility of controlling the morphology and local environment at the catalytic site depending upon the requirements of a particular reaction. Basically the reasons for which mesoporous materials act as promising candidates for heterogeneous catalysis are, (i) they have a high surface area and narrow pore size distribution essential for acting as a catalyst bed, (ii) the dimension of the pores can be tuned by varying the functional groups and ligands to obtain good selectivity for the product of interest, (iii) they have good adsorption properties, which allow the facile diffusion of reactant and product molecules during reactions, (iv) the ability to vary the type and concentration of surface functionality to change the textural properties and (v) the ability to tailor the polarity/hydrophobicity/hydrophilicity of the surface in order to influence the yield of a catalytic reaction. Moreover, they are relatively non-toxic, non-corrosive, non-air sensitive, highly reusable, completely pollution-free, environmentally benign supports for catalytic transformation in the liquid phase.10 The role and function of mesoporous silica and other non-siliceous materials in mediating sustainable liquid phase reactions e.g. acid–base, redox, polymerization, condensation and other organic coupling reactions, etc. are discussed herein.

4.1 Redox reactions over mesoporous materials

Although a pure mesoporous silica surface has high BET surface area (ca. 800–1400 m2 g−1) and a tunable pore diameter (2–50 nm), the material is not so effective for doing any catalytic reactions. By introducing one or more heteroatom and metal complex into the silica framework (metal doped silica) or often by suitable immobilization of organic functionalities (organosilica or organic–inorganic hybrid silicas) the desired catalytic activity of the material can be achieved. The advantage of choosing mesoporous silica as a support for incorporating other foreign atoms or functionalities is the inertness of the inorganic silica walls, where a tetravalent Si atom can be replaced by other reactive metal centers whilst retaining considerably good surface area and porosity.10 In spite of these benefits, silica structures like SBA-15, KIT-6, etc. have the limitation of losing porosity at elevated temperatures, which restricts their application in large scale heterogeneous catalysis. Metal incorporated mesoporous silica can be prepared using both a direct (or one step) or a post synthesis method.4 A number of related reports of mesoporous metal doped silica as heterogeneous catalysts can be mentioned in this respect, e.g. Zn–silica,20 Cr–MCM-41,37 CeO2–silica,38 Ti–MCM-41,39 Ce–Fe-SBA-15,40 Mn doped CeO2–silica,41 Pt and Rh nanoparticles supported silica,42 and so on. Periodic mesoporous silica (PMO), prepared by immobilizing organic moieties onto the inorganic silica surface by a direct synthesis method, can also act as a good redox catalyst when grafted with transition metals like V, Ti, Au, Mo, etc.43

Further, additional advantages of mesoporous transition metal oxides and mixed oxide catalysts are their high crystallinity and also their capability of existing in various oxidation states due to having vacant d orbitals, which can take part in electron transfer from reactants during a given catalytic process.44 Thus, these metal oxides or metal doped silica based materials are attracting wide-spread attention in recent times as non-toxic and environmentally friendly mild catalysts replacing conventional homogeneous hazardous oxidants like chromate, permanganate or periodide. In the following section we summarize some redox reactions mediated by the mesoporous solids.

a. Oxidation of hydrocarbons. Oxidation, partial oxidation and epoxidation of hydrocarbons, especially olefins (Fig. 3), using peroxides or molecular oxygen under mild conditions are very important reactions in organic synthesis from both academic and industrial points of view, since a variety of chemicals in high demand like polyurethanes, nylons, unsaturated resins, glycols, etc. are manufactured by this pathway.9 Due to several over oxidation products, satisfactory conversion and isolation of epoxides from alkenes is quite difficult in many cases. Recently, a number of scientists have dealt with this difficulty and reported the epoxidation of various alkenes like styrene, cyclohexene, cyclooctene, norbornene etc. over Ti-PMO materials45 under mild conditions. Ti-PMO, Ti-containing periodic mesoporous organosilica, was obtained using a co-synthesis method using an organosilica, an inorganic silica precursor (TEOS) and a Ti-precursor in the presence of CTAB surfactant under hydrothermal conditions. This mesoporous Ti–silica was tested for epoxide preparation under liquid phase conditions in the presence of tert-butylhydroperoxide (TBHP) oxidant. On the contrary, Cu-complex embedded MCM-41 was prepared using a post-synthesis method by surface modification of MCM-41 and then applied for the epoxidation of various substrates in the presence of TBHP.46 Along with epoxide formation, partial oxidation of olefins sometimes produces other industrially used fine chemicals over various metal doped mesoporous silica supports. Another Cu(II)-complex grafted SBA-16 catalyst, synthesized in a similar way as reported by Jana et al.,46 has been used for the oxidation of cyclohexene in the presence of a mild oxidant, H2O2, where other oxidation products like cyclohexanol and cyclohexanone etc. are produced in a significant amount along with epoxide.47 Liquid phase allylic oxidation of highly strained α-pinene and β-pinene is also feasible over UO22+–MCM-41 producing a number of products along with the respective epoxides.48 In this respect, styrene oxidation to produce benzaldehyde as a major product over Fe and Ti-SBA-1 mesoporous silica using H2O2, reported by Tanglumlert et al., can also be mentioned.49 Instead of using any oxidant, oxidation of hydrocarbons using air or O2 under mild conditions is more environmentally-friendly and is in high demand. Mesoporous silica acts as a good support for Au nanoparticles which efficiently catalyze the selective oxidation of cyclohexane in the presence of molecular O2, as reported by Wu et al.50
image file: c4ra13077d-f3.tif
Fig. 3 Different oxidation products of cyclohexene, cyclohexanol and cyclohexanone with adipic acid as the final product.

Adipic acid is an industrially important dicarboxylic acid, which can be produced exclusively from the catalytic oxidation of cyclohexane, cyclohexene or cyclohexanone over mesoporous materials like WO3/SiO2,51 and the core–shell Fenton catalyst (Fe2O3/Al2O3)52 in the presence of H2O2 oxidant. In Fig. 3 the step by step formation of adipic acid from these hydrocarbons is shown. Over the core–shell Fenton catalyst, the process occurs through formation of a hydroxyl radical by the reaction of Fe3+ with H2O2, and that OH˙ helps to oxidize cyclohexanone to adipic acid through the intermediates caprolactone and 6-hydroxyhexanoic acid. Cr(VI) grafted mesoporous polyaniline, synthesized by radical polymerization of aniline with a Cr precursor, exhibits good conversion in the liquid phase oxidation of alkane and alkene in a non-toxic green solvent, water, with a reuse capability of more than five times.53 Highly selective oxidation of cyclohexane to produce cyclohexanone is a very challenging task since cyclohexanol is always obtained as a major side product. Very recently, another porous co-ordination polymer was reported by Zhang et al. who have employed atmospheric O2 as an oxidant.54 This catalytic pathway provides a useful route for the selective conversion of β-isophorone to ketoisophorone, a significant reaction for the synthesis of vitamin E.

Porous carbon based materials are also utilized as effective supports for the oxidation of hydrocarbons under liquid phase reaction conditions. N doped nanoporous carbon with supported Pd nanoparticles has shown high catalytic efficiency for the selective aerobic oxidation of indane, diphenyl methane and ethyl benzene etc. thus providing a potential route for large scale fine chemical synthesis.55 A new kind of boron- and fluorine-enriched polymeric mesoporous carbon nitride with high surface area has been prepared by Wang et al. who illustrate the potential of the material for selective cyclohexane to cyclohexanone conversion under mild conditions.56 This organic semiconductor with controlled mesopore structures and narrow pore size distribution greatly facilitates this selective oxidation reaction with reusability for several times without losing any activity. Heterogeneous organocatalysts are particularly high in demand as they are devoid of toxic metal and thus free from the possibility of metal-leaching during the catalytic reactions. In this context porous graphitic carbon nitride (g-C3N4) has shown high catalytic activity in the selective oxidation of many branched hydrocarbons under mild environmentally-friendly conditions.57,58

b. Oxidation of alcohols and sulfides. The aldehydes and ketones obtained from the oxidation of primary and secondary alcohols, respectively, are very essential intermediates in organic synthesis, and the pharmaceutical and agricultural industries, etc. Aerobic oxidation of long chain primary alcohols, like 1-octanol, has been achieved with excellent yields obtained using small Au nanoparticles deposited on ceria or iron oxide doped hexagonal mesoporous silica (HMS) materials.59 Synthesis of high grade benzaldehyde is particularly attractive to modern researchers for application in perfumery and food industries. A highly selective chlorine-free method of benzaldehyde production without the formation of over oxidized products like acid or ester at room temperature is observed over a highly ordered mesoporous ceria–silica composite which was able to convert above 50% of benzyl alcohol to benzaldehyde exclusively under solvent-free conditions using TBHP. The silica supported catalyst was truly heterogeneous and stable to exhibit high recyclability up to three times towards this alcohol oxidation reaction with minimum Ce leaching.38 A detailed study has also been undertaken by the authors to compare the product conversions at RT with those at an elevated temperature. A schematic representation is given in Fig. 4 to highlight the synthesis and catalysis conditions for this ceria–silica catalyst. Mesoporous metal oxides are also efficient in the oxidation of primary and secondary alcohols. This has been demonstrated by the catalytic potential of an interesting mesoporous nanocomposite of Cr2O3 and 12-phosphomolybdic acid (PMA), prepared via a “nanocasting” method using mesoporous SBA-15 as a hard template.60 The catalyst Cr2O3–PMA solid-solution shows an extraordinary green oxidative pathway for oxidation of 1-phenylethanol to produce acetophenone with up to 85–87% yield using non-toxic H2O2 at a very low temperature (323 K). The synergetic interaction between the PMA clusters and Cr2O3 matrix via Cr–O–Mo bonds as well as the dispersion of PMA over large effective surface area of Cr2O3 play the key role for this selective oxidation of secondary alcohol. Using molecular oxygen from air as a clean oxidant, oxidation of benzyl alcohol to benzaldehyde with 72% conversion and 100% selectivity is reported over highly dispersed mesoporous carbon supported iron oxide (FeOx/H-CMK-3) under mild reaction conditions.61 A Pd@core–shell nanosphere consisting of Pd metal clusters as the core and microporous silica as the outer shell also exhibits good catalytic efficiency and selectivity for the aerobic oxidation of long chain primary alcohols under solvent-free conditions.62
image file: c4ra13077d-f4.tif
Fig. 4 A schematic representation of the synthesis of a mesoporous ceria–silica composite and its catalytic application for oxidation of a primary alcohol (benzyl alcohol).

The oxidation of sulfides to obtain sulfoxide and sulfones; products playing an imperative role in organic and biological reactions as well as in anti-hypertension and as cardiotonic agents etc., using industrially important heterogeneous and mesoporous catalysts replacing the homogeneous ones like halogen compounds, nitrates and transition metal oxides etc., has won great scientific interest.63 A Cr-grafted organic–inorganic hybrid mesoporous polymer, synthesized using an anionic surfactant, efficiently catalyzes the partial oxidation of different aromatic and aliphatic sulfides to their corresponding oxides with impressive yields, having excellent recyclability and no significant loss in activity.63 The almost selective oxidation of sulfide to sulfoxide using H2O2 oxidant in methanol solvent was reported over WO3 nanoparticles supported MCM-48 heterogeneous catalyst in the year 2005.64 The activity and stability of Mn and Cu oxo complexes have been increased to a great extent by ‘heterogenization’ of these homogeneous catalysts, i.e. by post-grafting onto mesoporous MCM-41. The resulting supported catalyst was able to convert methyl phenyl sulphide to sulfoxide with a high yield and exhibited excellent reusability up to five times in the presence of TBHP.65 Here, MnV[double bond, length as m-dash]O species present in the Mn–oxo salen complex stabilized in the mesoporous silica framework play the active role for this sulphide oxidation. Mesoporous graphitic carbon nitride (g-C3N4) polymer has been established as a green oxidation catalyst for the selective oxidation of various sulphides using visible light at room temperature.66

c. Oxidation of amines. Oxidation of primary amines, e.g., aniline and substituted anilines, to the corresponding azoxybenzene derivatives is a very common and fundamental reaction in organic chemistry, because azoxybenzene and its derivatives have great potential for application in the synthesis of industrially important fine chemicals.67 In 2009 a very interesting cobalt-bound polymer was prepared by Chang et al. and the polymer was used as a template to prepare mesoporous silica containing cobalt-oxide. The material exhibited good activity and reusability in the catalytic oxidation of aniline and its derivates to the respective azoxybenzenes with above 60% conversion achieved at 298 K as well as 100% conversion at an elevated temperature in the presence of the non-toxic H2O2 oxidant.67 Apart from Co, Ti- and Zr-containing mesoporous silicas have also been employed as catalysts for the oxidation of amines under variable reaction conditions.68 An advanced approach was also made at the end of the 20th century when hexagonal mesoporous silica HISiO2 was prepared using a non-ionic surfactant, Brij 76, at room temperature and then modified with 3-(aminopropyltriethoxysilane) (APTES), a well-known organosilica precursor to form a NH2-functionalized organic–inorganic hybrid silica, which was immobilized with various metal aqua complexes (e.g. MnII, CuII, CoII and ZnII). These metal complex loaded mesoporous silicas were tested for the oxidation of amines.69 Amine oxidation follows first order kinetics and the rate of the reaction is highly dependent on the oxidizing potential of the metal complex used in the catalyst.69 In Fig. 5, the synthesis of MnII-aqua-NH2 functionalized HISiO2 catalyst and its oxidizing role for the oxidation of o-aminophenol is shown by a simple schematic diagram.
image file: c4ra13077d-f5.tif
Fig. 5 The synthesis of MnII-aqua-NH2 functionalized mesoporous HISiO2 catalyst and its oxidizing role for amine (o-aminophenol) oxidation.
d. Ammoximation. Ammoximation of ketones with ammonia and dilute aqueous H2O2 oxidant to produce oximes is a very crucial reaction in modern synthetic chemistry because oximes play an intermediate role for the fine chemical synthesis of lactams and amides etc. via secondary chemical reactions. Many metallosilicates like TS-1 have been employed as efficient catalysts for the synthesis of oximes in the presence of NH3 and aqueous H2O2.70 The ammoximation of bulky ketones has been efficiently carried out over a Ti-grafted ethane bridged hybrid mesoporous silica support synthesized by a one pot surfactant assisted method.71 Cyclic ketones like cyclohexanone and cyclodecanone were catalytically converted with 100% selectivity for oxime using tert-butanol as a solvent at a reaction temperature of 353 K.
e. Hydroxylation of aromatics. Hydroxylation or introduction of a –OH group into various aromatic substrates like benzene, toluene, phenol etc. is a commercially important chemical reaction, because the resulting phenol and substituted phenol products are used in the fine chemical, agrochemical, photographic chemical, pharmaceutical and food industries.72 Fe-incorporated mesoporous silica was prepared by a simple co-precipitation method and employed for the catalytic hydroxylation of phenol to selectively produce dihydroxybenzene with proper optimization of the reaction conditions.72 Phenol hydroxylation was also successful over a redox active group ferrocene loaded nanocomposite made with mesoporous polymer and silica via a controlled co-polymerization route.73 Hydroxylation of benzene to phenol is sometimes inhibited by the formation of by-products like catechol and hydroquinone etc. The highly selective conversion of benzene to phenol is reported over V-complex incorporated PMO prepared by direct co-condensation of the metal grafted organosilica with the inorganic silica precursor in the presence of CTAB. In this fashion, VO(acac)2, a well-known homogeneous catalyst ‘heterogenized’ by immobilization onto a silica surface,74 catalyzes this benzene to phenol transformation in the presence of H2O2 oxidant and the reaction is proposed to be mediated by the peroxovanadyl radical (V5+–O–O˙) as the main active species.74 Another important example of heterogenization over a solid support is the loading of FeCl3 over mesoporous carbon nitride (g-C3N4). This hybrid catalyst shows up to 38% conversion of benzene to phenol with high selectivity (97%) under visible light irradiation and mild conditions.75 In Fig. 6, a schematic representation of the synthesis and catalytic role of mesoporous Zn doped aluminophosphate (ZnAlPO4) is shown. Highly selective formation of phenol from benzene is observed over this Zn-doped mesoporous aluminophosphate with a minor amount of hydroquinone produced as a by-product. The material has been synthesized by a method of physical mixing using tetrapropylammonium bromide as a template (Fig. 6).76
image file: c4ra13077d-f6.tif
Fig. 6 The synthesis scheme and catalytic performance of mesoporous Zn-doped aluminophosphate (ZnAlPO4) material.
f. Hydrogenation and other reductions. Hydrogenation of unsaturated compounds via treatment with hydrogen or reduction in the presence of Pt, Ir, Ru, Ni, Pd nanoparticles etc. over a heterogeneous support is an important chemical reaction industrially. An easy way to prepare a heterogeneous catalyst bearing the above mentioned reducing metals is via successful confinement of these metals into a mesoporous support. Hydrogenation of alkene, ketones, phenol and nitro substrates etc. are some potential examples of this category and worthy of mention in this . Instead of reduction in the gas phase under harsh conditions as reported in the case of Zr/Ni-doped mesoporous silica for the hydrogenation and ring opening of tetralin at a high temperature using H2 gas,77 liquid phase catalytic reactions under mild conditions are more desirable from an environmental view point. The asymmetric hydrogenation of various aryl substituted ketones and quinolones with high conversion was successful under very mild conditions over chiral Ir and Ru complexes supported mesoporous silica, synthesized through co-condensation of an organosilica, an organometallic complex, [Cp*IrCl2]2 or [RuCl2(-C6Me6)]2, and an inorganic silica precursor in the presence of non-ionic Pluronic (P123) and CTAB surfactants, respectively.78,79

Mesoporous oxides or mixed oxides can also act as a support similarly for thioether nanoparticles such as Au, Ni, Pt and Pd and these confined nanoparticles can catalyze the hydrogenation of aromatic nitro compounds. Mesoporous oxides like TiO2/Al2O3/SiO2/ZrO2 with confined PtPd/AuPd/AuPt thioether nanoparticles have been synthesized via a surfactant-assisted route and show highly selective synthesis of aniline from nitrobenzene under mild conditions using hydrazine hydrate as a hydrogenating agent.80 On the contrary, mesostructured nickel–aluminium mixed oxide hydrothermally synthesized using lauric acid as a structure directing agent, was itself efficient for the satisfactory reduction of nitrobenzene and its derivatives in 2-propanol solvent which acts as a hydride donor to generate substituted anilines.81 The synthetic pathway for the mesoporous self-assembled Ni–Al mixed oxide nanoparticles and their application for hydrogenation of nitrobenzene to aniline is shown in Fig. 7. Hydrogenation of D-glucose over Ru nanoparticles impregnated mesoporous hypercrosslinked polystyrene was reported by Sapunov et al.82 They also provided an elaborate mechanistic pathway and kinetics for the formation of D-sorbitol. In a non-toxic solvent like water, selective hydrogenation of phenol to cyclohexanone over ultra-small Pd nanoparticles grafted mesoporous carbon is a green synthesis in high demand. Pd supported N-functionalized ordered mesoporous carbon was prepared by nitration of a typical mesoporous carbon FDU-15 using NH3 followed by treatment with an aqueous solution of H2PdCl4.83 The catalyst was highly heterogeneous in nature and showed reusability up to 6 times with a good product yield each time.


image file: c4ra13077d-f7.tif
Fig. 7 A schematic representation showing the synthesis of mesoporous Ni–Al mixed oxide and its application for hydrogenation of nitrobenzene to aniline.

The above mentioned redox reactions are a few examples of a huge number of liquid phase catalytic reactions that have been carried out over the years using the mesoporous family of materials. In most of these cases the mechanisms of the reactions follow a free radical pathway mediated by one or two radicals generated in situ.53,74 In Table 2 we have summarized all the relevant reactions in tabular form for the convenience of the readers.

Table 2 Different redox reactions in liquid phase over mesoporous materials
Type of reactions Catalysts Reaction involved Product yield (%) Ref.
Oxidation of hydrocarbons Ti doped PMO Epoxidation of styrene, trans-stilbene norbonene, etc. 65–80 45
Cu-complex doped MCM-41 Epoxidation of cyclohexene, cyclooctene, styrene, etc. 92–99 46
Cu(II)-complex grafted SBA-16 Partial oxidation of cyclohexene 21–77 47
UO22+–MCM-41 Allylic oxidation of α-pinene, β-pinene etc. 80–88 48
Fe- & Ti-SBA-1 Partial oxidation of styrene 59–70 49
Au loaded mesoporous silica Oxidation of cyclohexane 10–34 50
Core–shell Fenton catalyst Oxidation of cyclohexanone 69 52
Cr–mesoporous polyaniline Oxidation of alkane, alkene 70–98 53
Porous co-ordination polymer Conversion of β-isophorone to ketoisophorone 85 54
Pd@N doped nanoporous carbon Selective aerobic oxidation of indane, ethyl benzene 14–30 55
Boron- and fluorine loaded mesoporous carbon nitride Selective oxidation of cyclohexane to cyclohexanone 2–8 56
g-C3N4 polymer Oxidation of strained hydrocarbons 26–100 57 and 58
Oxidation of alcohols Au–ceria or iron oxide–HMS Oxidation of 1-octanol 10–79 59
Mesoporous CeO2–silica Oxidation of benzyl alcohol 51 37
Mesoporous Cr2O3–PMA composite Oxidation of 1-phenylethanol 85–87 60
Iron oxide on mesoporous carbon (FeOx/H-CMK-3) Oxidation of benzyl alcohol to benzaldehyde 72 61
Pd@silica core–shell nanosphere Solvent-free aerobic oxidation of long chain primary alcohols 10–61 62
Oxidation of sulfides Cr–mesoporous polymer Partial oxidation of different aromatic and aliphatic sulfides 7–97 63
WO3 NP–MCM-48 Oxidation of different sulfides 100 64
Mn and Cu oxo complex–MCM-41 Oxidation of methyl phenyl sulphide 65–100 65
Mesoporous g-C3N4 polymer Selective oxidation of sulphides 39–98 66
Ammoximation Ti-grafted ethane bridged porous inorganic–organic hybrid silica Oxidation of cyclohexanone, cyclodecanone 42–90 71
Oxidation of amines Mesoporous cobalt oxide–silica composite Oxidation of aniline and its derivatives 26–100 67
Ti, Zr-doped mesoporous silica Oxidation of aniline 100 68
Mn, Cu, Co & Zn-NH2-grafted organic–inorganic hybrid silica Oxidation of amine 90–100 69
Hydroxylation of aromatics Fe incorporated mesoporous silica Hydroxylation of phenol 20–50 72
Ferrocene loaded nanocomposite of mesoporous polymer and silica Conversion of phenol to form dihydroxybenzene 25–30 73
VO(acac)2 incorporated PMO Selective conversion of benzene to phenol 12–38 74
FeCl3/g-C3N4 Conversion of benzene to phenol 38 75
Mesoporous Zn doped aluminophosphate Selective hydroxylation of benzene to phenol 99 76
Hydrogenation & other reductions Chiral Ir and Ru complexes-loaded mesoporous silica Asymmetric hydrogenation of aryl substituted ketones and quinolones 80–99 78 and 79
PtPd/AuPd/AuPt-embedded TiO2/Al2O3/SiO2/ZrO2 Selective conversion of nitrobenzene to aniline 65–87 80
Mesostructure NiO–Al2O3 Reduction of nitrobenzene and its derivatives 36–57 81
Ru–mesoporous polystyrene Hydrogenation of D-glucose 40–43 82
Pd nanoparticles-loaded mesoporous carbon Selective hydrogenation of phenol to cyclohexanone 23–80 83


4.2 Acid–base reactions mediated by mesoporous materials

Lewis and Brønsted acid or basic sites present in a solid catalyst are the main reason for the generation of the acid–base properties in a given solid. The Lewis acidity originates due to the presence of metal ions coordinated to the mesoporous support whereas –COOH, –SO3H, –NR3+ groups etc. are generally responsible for the Brønsted acidity. In the same way, –NH2, –NMe2 etc. and –COO groups account for Lewis and Brønsted basicity, respectively. The number and strength of these sites determine the catalytic activity of the concerned materials and also their selectivity for the organic transformations. Moreover, surface properties i.e. huge surface area and pore width of the mesoporous material give an additional advantage compared to the other nonporous materials. Though inherent weak acid sites restrict the application of mesoporous materials in many petrochemical reactions, they still demonstrate good potential for the reactions which require a lower level of acidity, e.g. Friedel–Crafts alkylation of organics.10 A lot of scientific reports such as those on mesoporous perovskite ZnTiO3,19 Zn–silica,20 –COOH grafted functionalized mesoporous silica,84 Mn-doped ceria–silica,42 sulphonated zinc phosphonate,85 zirconium oxophosphate,86 and so on have been published in the literature regarding the acid–base catalytic reactions over mesoporous supports. In this section we shall illustrate some of those heterogeneous reactions carried out in the liquid phase.
a. Friedel–Crafts reaction. Friedel–Crafts (FC) alkylation and acylation reactions are the set of organic reactions developed by Charles Friedel and James Crafts in 1877 to attach substituents to an aromatic ring and this is one of the most important chemical processes for laboratory synthesis and industrial production of value added organic fine chemicals.87 These liquid phase reactions proceed via electrophilic aromatic substitution and are usually catalyzed by Lewis acidic groups. Instead of using corrosive H2SO4 or solid acids like AlCl3, FeCl3 or BF3 etc. in the homogeneous phase, which suffers as a method from a lot of environmental hazards and inconvenience due to metal leaching, non-recyclability etc., mesoporous based catalysts can be a good replacement as non-toxic, heterogeneous, low-cost catalysts. Perovskite ZnTiO3, a mesoporous oxide, was reported to be highly efficient for FC benzylation of p-xylene, toluene and benzene using benzyl chloride as the alkylating agent at a temperature of 343–348 K to produce 1-(2,5-dimethylbenzyl)benzene, 2,4-methylbenzyl benzene and diphenyl methane, respectively with above 90% conversions.19 The presence of strong Lewis acidic metal sites (Zn and Ti) in the perovskite material along with a good surface area helps such excellent conversion to substituted aromatics. Zhang et al. have synthesized a new mesoporous iron phosphate Fe–P–O catalyst using four different methods and applied it for benzylation of benzene and other substrates.88 The alkylation carried out using benzyl chloride gives 100% selective formation of diphenyl methane after 3 h. On the contrary, an acylation reaction using acetic anhydride, benzoyl chloride or p-toluoyl chloride has been observed very frequently over functionalized mesoporous materials, e.g. p-toluoyl chloride is used for liquid phase FC benzoylation of anisole at 393 K over a silicotungstic acid loaded mesoporous alumina molecular sieve.89 Instead of these hazardous acylating and alkylating agents, green non-toxic reagents like aliphatic and aromatic alcohols are also very popular for use in FC reactions. Mesoporous Nb oxide when sulphated catalyzes the benzylation of toluene and anisole in the presence of benzyl alcohol to produce substituted aromatics.90 Alkylation of toluene using benzyl alcohol has been carried out efficiently over Mo-embedded mesoporous carbon microspheres which results in a 100% conversion of benzyl alcohol to produce p-, m-, and o-benzyl toluene after just 1 h at a reaction temperature of 383 K, and exhibits high reutilization for up to three cycles without any decay in activity. The high oxidation state of Mo(VI) along with the porous structure of carbon allowing free movement of reactants and products between embedded α-MoO3 and the liquid medium mainly accounts for this excellent catalytic performance.91 Different FC products obtained from different substrates using different alkylating and acylating reagents are shown in Fig. 8.
image file: c4ra13077d-f8.tif
Fig. 8 Friedel–Crafts products obtained from various substrates during reactions over different mesoporous catalysts.

FC reactions are also very common over mesoporous silica based materials as revealed by many recent reports. A 3D network of mesoporous gallosilicate (GaSBA-1) was prepared in an acidic medium using the cationic surfactant CTAB and an appropriate amount of Ga salt. This material has been tested for the benzylation of various aromatics using benzyl chloride. The catalytic role of the Ga species along with a schematic representation of a suitable mechanism has been mentioned and explained by the authors.92 A Cu(II) complex containing a chiral bis(oxazoline) ligand has been immobilized on mesoporous silica and the catalyst was employed for asymmetric benzoylation of 1,2-diphenyl-1,2-ethanediol which produces a highly enantioselective product using benzoyl chloride under mild conditions.93 There are many other reports on the alkylation of aromatics using alcohols over metal doped mesoporous silica. Sulphated Zr–MCM–41 and Zr–MCM–48 solid acid catalysts have successfully catalyzed attached substituent of phenol aromatic ring in presence of tert-butyl alcohol to produce 4-tert-butyl phenol as major product with above 90% yield at 413 K temperature.94

b. Esterification and transesterification reactions. Esterification and transesterification reactions are widely employed today for biofuel production, which is an alternative and renewable source of fuel in most of the developed countries. The condensation of carboxylic acid with alcohol in the presence of any acid catalyst was named as esterification by the German chemist Emil Fischer whereas the conversion of one ester to another carboxylic acid ester using alcohol and acid or base as a catalyst is called transesterification (Fig. 9). Generally, the reactions of triglycerides i.e. triesters of glycerol or any other fatty acids with short chain alcohols like methanol or ethanol are categorized as biodiesel reactions.95
image file: c4ra13077d-f9.tif
Fig. 9 (A) Esterification and (B) transesterification reactions for biodiesel production by acid or base catalysts.

Both Lewis and Brønsted acids are used as catalysts for the esterification of long chain fatty acids and the equilibrium of the reversible reaction can be right shifted towards the ester conversion by removing water from the reaction medium through azeotropic distillation or using molecular sieves or using an excess of alcohol.19 Replacing previously used strong Brønsted acids like H2SO4, p-TsOH (p-toluenesulphonic acid) various environmentally non-hazardous, recyclable functionalized mesoporous materials have been designed, which are employed successfully as heterogeneous catalyst for the biodiesel production. The mechanism of an esterification reaction over solid acid catalysts can proceed via activation of the carboxylic acid by the catalyst, then nucleophilic attack by the alcohol on the activated carbonyl carbon to form a tetrahedral intermediate and finally formation of an ester with elimination of one molecule of water.19 Esterification of a long chain fatty acid with alcohol and vice versa is observed over mesoporous zirconium oxophosphate material which not only produces the corresponding ester in high yield but also shows high reusability up to the 5th cycle, the catalytic activity being attributed to the large surface area and large number of acidic sites located at the surface of the phosphate material.86 In Fig. 10, a simple scheme for the synthesis of this Zr-oxophosphate material is shown along with its application in the proposed mechanistic pathway for an acid catalyzed reaction. The presence of the acidic sites necessary for catalytic activity in the Zr-catalyst has been investigated and quantitatively estimated by using the temperature-programmed desorption of ammonia (NH3-TPD) measurement.86 The TPD result suggested that with the increase in surface Brønsted acidity a higher yield of the biodiesel products was obtained. Esterification over periodic mesoporous silica materials has been described in the review published by Yang et al. in 2009.96 MCM-48 supported tungstophosphoric acid acts as a heterogeneous catalyst for esterification of a number of fatty acids like palmitic, lauric, stearic, isostearic, myristic etc. with alcohols, e.g. cetyl, butanol, hexanol, octanol etc. in 11 MPa super critical CO2 medium at a temperature of 373 K to produce esters in an appreciable amount which increases with the increasing chain length of the fatty acids.97


image file: c4ra13077d-f10.tif
Fig. 10 A simple schematic pathway of the synthesis of mesoporous zirconium oxophosphate and a proposed catalytic cycle for the esterification reaction over this catalyst.

Transesterification reactions involve an acid or base mediated reaction where increasing the electrophilicity of a carbonyl carbon of fatty acids or the nucleophilicity of an attacking alcohol can improve the reaction rate considerably.95 Sulphonated porous zinc phosphonate is a example of a green catalyst, which provides an eco-friendly route for biodiesel production via transesterification of long chain fatty acids in the presence of methanol at room temperature.85 Thus, ordered Zn-doped mesoporous silica20 or amine-functionalized mesoporous organosilica98 or mesoporous polyoxometalate–tantalum pentoxide composite, H3PW12O40/Ta2O5 (ref. 99), are very active acid catalysts to give a satisfactory yield of the respective esters under suitable conditions. In the mesoporous H3PW12O40/Ta2O5 composite, ([triple bond, length as m-dash]TaOH2)n+[H3−nPW12O40]n species formed at the catalyst surface of the composite via a Ta–O–W bond facilitates the electron transfer from the terminal oxygen atoms of W[double bond, length as m-dash]O groups to Ta2O5; encouraging a good amount of proton release to ensure high Brønsted acidity of the catalyst responsible for the biofuel conversion.99 On the contrary, transesterification using base catalysts has been efficiently carried out over Mn-doped ceria–silica and MgO-functionalized mesoporous silica. etc. Mn doped ceria–silica catalyzes the transesterification of methyl benzoate, ethyl cyanoacetate and ethyl chloroacetate with methanol, butanol and octanol under solvent-less mild reaction conditions. The presence of basic sites in the MgO–silica was determined by CO2 sorption measurements and the sites help to produce biofuel from vegetable oil in the presence of ethanol, achieving a conversion level of above 90% within 5 h reaction time.100

c. Acetalization reactions. Acetalization is widely used in the synthesis of natural products or industrially important chemicals which are employed for the protection of active functional groups like aldehydes or ketones in a molecule and thus useful in designing a multi-functional molecule. Acetal formation occurs through the nucleophilic addition of an alcohol to carbonyl compounds in the presence of an acid.101,102 Mesoporous solids are used extensively as acid catalysts for acetalization, e.g. methanol was used to acetalize cyclohexanone at room temperature over a mesoporous silica catalyst, which shows remarkable shape selectivity and a good correlation of catalytic activity with pore diameter of the corresponding mesopores.103 Acetalization of cyclohexanone using methanol to produce dimethyl ketal is also successfully carried out over mesoporous zirconia functionalized with –SO4 gr. and a conversion of over 90% with 100% selectivity is observed within a very short reaction period (45 min).101 Glycerol is also used for acetalization of acetone over a stable Ni–Zr supported mesoporous carbon catalyst, when selectively two products, 5-membered solketal and six-membered acetal, were obtained under solvent-less conditions (Fig. 11).104 Another interesting acetalization process is the reaction of heptanal with 1-butanol over various sulphonic acid functionalized mesoporous organosilicas.105 Three types of organosilica, viz. arene sulfonic acid ethane-silica (AS-MES), mercaptopropyl ethane-silica (Pr-SH-MES) and arene sulfonic acid functionalized SBA-15 were prepared using Pluronic non-ionic surfactant, P123 and then post-functionalized with an –SO3H group via treatment with a H2O2–H2SO4 solution. These acid sites in the organosilica help to acetalize heptanal with butanol to form the corresponding acetal.105
image file: c4ra13077d-f11.tif
Fig. 11 Acetalization reaction between glycerol and acetone over a Ni–Zr supported mesoporous activated carbon catalyst.
d. Aldol condensation. Carbon–carbon bond formation reactions in various large biomolecule and industrial fine chemical syntheses are largely dependent on aldol condensation reactions, a well-known organic reaction of an enol or enolate ion with a carbonyl compound in order to form a β-hydroxyaldehyde or β-hydroxyketone or ‘aldol’ (aldehyde + alcohol), followed by dehydration of the intermediate to produce an α,β-unsaturated carbonyl compound. The reaction is catalyzed by both acid and base to form enol and enolate ion, respectively. Conventionally used Brønsted bases like NaOH, Ca(OH)2, Ba(OH)2, etc. and acids such as acetic acid, and Lewis acids like M(Otf)n (M = Zn, Cu, Pb, Sc etc.) produce hazardous, toxic by-products causing a lot of environmental pollution. Hence, aldol reactions over mesoporous catalysts in heterogeneous media are in high demand.106

Mesoporous carbon can act as a good support for metal nanoparticles or metal oxides to improve catalytic performances, e.g. Mg–Zr mixed oxide supported on mesoporous carbon catalyzes the cross-aldol reaction of furfural and acetone followed by hydrogenation–dehydration to form C8 and C13 alkanes at a temperature of 323 K, the conversion and selectivity being higher than when bulk Mg–Zr mixed oxide is used.107 The reaction mechanism and the kinetic model of aldol condensation over the distributed basic sites of this supported catalyst are also explained by the authors. Bai et al. have synthesized mesoporous γ-Al2O3 based on a cation–anion double hydrolysis method and the material was proven to be truly efficient for the aldol reaction of cyclohexanone carried out in a water-separating device in the form of water–cyclohexanone azeotrope at a temperature of 438 K for 2 h when the conversion was as high as ∼70%. The remarkably high catalytic activity of crystalline γ-Al2O3 could be attributed to the huge surface area, precise mesopore diameter, absence of anionic species as well as the surface basicity generated due to aluminium and oxygen vacancies.108 Based on the experimental results they also proposed a suitable mechanism for the formation of α,β-unsaturated ketone involving the absorption of α-H of cyclohexanone by the aluminium vacancies. Aldol condensation of benzaldehyde and heptanal to form jasminaldehyde under solvent free conditions is reported over novel Mg–Al mixed oxide supported on hexagonal mesoporous silica (Fig. 12).109 The bi-functional behaviour of this catalyst, where the weak acid sites activate the aldehyde by protonation and the basic sites favour the formation of the enolate heptanal intermediate, helps in carrying out the reaction successfully.109 In this context a completely metal-free organocatalyst has been developed via heterogenizing (S)-proline on MCM-41, which catalyzes the asymmetric aldol condensation of p-nitrobenzaldehyde and 2,2-dimethyl-1,3-dioxan-5-one under the influence of different solvents ranging from hydrophobic to hydrophilic, proving hydrophilic polar media as more suitable for catalysis.110


image file: c4ra13077d-f12.tif
Fig. 12 The synthesis of hexagonal mesoporous silica supported Mg–Al mixed oxide catalyst and its application for the aldol condensation of benzaldehyde with heptanal.
e. Knoevenagel condensation. Unlike aldol reactions involving the condensation of two carbonyl compounds, Knoevenagel condensation involves the reaction of a carbonyl compound with a compound containing an active methylene group followed by dehydration to form a C–C bond resulting in an α,β-unsaturated compound. The reaction is generally mediated by a basic catalyst and is claimed to be one of the most important environmentally benign C–C bond formation reactions, because water is the only by-product produced in this condensation.1 Instead of using soluble amine bases in homogeneous media, any amine based heterogeneous support or metal support with basic sites will be more suitable as a non-toxic pollution free catalyst for such reactions.

Mesoporous CexZr1−xO2 solid consisting of nanometer sized particles was synthesized via a sol–gel method and it showed excellent chemoselectivity for the classical Knoevenagel condensation of benzaldehyde and malononitrile owing to the presence of pores of a large size (∼10 nm) and both surface acidic and basic sites.111 The Knoevenagel reaction under mild conditions is also reported over ordered mesoporous Ni–Al mixed oxide obtained from Ni–Al-layered double hydroxides (LDHs) by ultrasonic irradiation using Pluronic-F127 surfactant. The catalyst exhibits high selectivity for the product with good reusability without significant loss of activity.112 Mesoporous oxides without basic sites can also be used for Knoevenagel reaction after proper functionalization of the oxide surface with amine groups. By co-condensation of 3-aminopropyltrimethoxy silane and zirconium precursors in CTAB surfactant solution, mesoporous amine grafted zirconia can be prepared and used for condensation of benzaldehyde and a malonic ester, diethyl malonate, in methanol solvent at room temperature to produce cinammic acid exclusively.113 Besides this, as metal-free organocatalysts, guanidine base immobilized on mesoporous silica SBA-15 and urea embedded mesoporous polymer are highly recommended for Knoevenagel reaction.114,115 Two types of Knoevenagel condensation, one between benzaldehyde and ethyl cyanoacetate, and another between cyclohexanone and benzylcyanide have been reported over basic guanidine functionalized silica at RT as well as at an elevated temperature.114 On the contrary, mesoporous urea-functionalized polymers synthesized through the non-ionic F127 surfactant-directed urea–phenol–formaldehyde oligomer self-assembly approach have been utilized for Knoevenagel reaction of aromatic aldehyde and ethyl cyanoacetate at mild temperatures in a non-hazardous aqueous medium. In Fig. 13, a simple pictorial representation for the synthesis of this mesoporous urea-polymer along with a proposed mechanism for the catalytic reaction over this polymer is shown. The synergic effect between the urea active sites embedded on the mesopore wall and their neighbouring surface phenolic –OH groups in the mesoporous support accounts for the success of this reaction along with the good stability of the catalyst inhibiting the leaching of active species as documented through successful reuse up to six reaction cycles.115


image file: c4ra13077d-f13.tif
Fig. 13 The synthesis of urea functionalized mesoporous polymer and Knoevenagel condensation of substituted benzaldehyde with ethyl cyanoacetate over this polymer.
f. Baeyer–Villiger oxidation. Baeyer–Villiger (BV) oxidation involves the oxidative cleavage of a carbon–carbon bond adjacent to a carbonyl group, which forms esters and lactones from acyclic ketones and cyclic ketones, respectively. This reaction is important for the industrial production of bioactive chemicals and generally proceeds in the presence of peracids like C6H5CO3H and CF3CO3H etc. or non-toxic, less hazardous hydrogen peroxide along with Lewis acidic or sometimes basic catalysts.116 The reaction has been successfully employed for the conversion of cyclohexanone, cyclooctanone, ethyl methyl ketone and methyl isopropyl ketone to their respective lactones and esters in the presence of a non-siliceous mesoporous Mg–Al mixed oxide catalyst using H2O2.117 The high surface area and nanoscale porosity alone with a considerably good amount of basic sites present in the porous MgAl2O4 catalyst accelerate the reaction process. However, the above reaction was carried out in benzonitrile as a co-solvent at an elevated temperature (343 K) for which formation of benzamide is observed as a by-product and this minimises the selectivity of the desired lactone and ester. The conversion and selectivity of cyclic and acyclic ketones with or without using benzonitrile is shown in Fig. 14. More convenient and mild conditions have been employed by Jeong et al. who carried out the BV reaction over metal doped PMO at quite a lower temperature (313 K) in the presence of molecular oxygen (O2)/benzaldehyde system. For highly selective conversion of cyclohexanone to ε-caprolactone they used different metalloporphyrin (Fe, Cu, Sn) bridged periodic mesoporous organosilica materials among which an Fe-catalyst has been proven to be the most promising and efficient catalyst for this reaction.118 Using O2/benzaldehyde as an oxidizing agent, an environmentally benign green catalytic route has been reported over a mesoporous zirconium phosphate material which exhibits a strong ability to catalyze the conversion of cyclic ketones, even bulkier ones such as adamantanone, to the corresponding lactones with a high yield and 100% selectivity at RT under solvent-free conditions.119 A free-radical reaction pathway facilitated by the Lewis acidic sites present in this mesoporous Zr-phosphate catalyst has been suggested. The catalyst is chemically stable and showed good recyclability, suggesting that the procedure is economically advantageous for industrial purposes.
image file: c4ra13077d-f14.tif
Fig. 14 Baeyer–Villiger oxidation of cyclic and acyclic ketones over a mesoporous Mg–Al mixed oxide along with the conversion–selectivity representation in a bar diagram.
g. Beckmann rearrangement. Beckmann rearrangement, named after the German chemist Ernst Otto Beckmann, is another important acid-induced reaction in organic chemistry to form lactams and amides from cyclic and acyclic oximes, respectively. The reaction is generally catalyzed by strong acids like polyphosphoric acid and sulphuric acid etc. Instead of using these corrosive acids there are some reports of Beckmann rearrangements over mesoporous acid catalysts, but most of those reactions proceed under harsh conditions like high temperatures in the vapor or gas phase.120 Liquid phase rearrangement of cyclohexanone oxime to form ε-caprolactam (with about 50% yield) was observed in a series of solvents over acidic sites of Al doped mesoporous MCM-41 molecular sieves by Ngamcharussrivichai et al. although the conversion and selectivity level are low.121 A few years later, sulphonic acid functionalized mesoporous organosilica with an SBA-15 type nanostructure was employed for this reaction, but the yield and selectivity of lactam did not improve by much. But the material is quite interesting as reports on heterogeneous metal free organocatalyst mediated catalytic transformations are comparatively scarce.122 The above acid–base reactions are summarized in Table 3. These acid–base reactions mentioned here are a few examples of a wide range of mesoporous catalyst-mediated liquid phase organic transformations reported in the literature.
Table 3 Different liquid phase acid–base reactions mediated by mesoporous materials
Type of reactions Catalysts Reaction involved Product yield (%) Ref.
Friedel–Crafts reaction Mesoporous perovskite ZnTiO3 Reaction of p-xylene, toluene, benzene with benzyl chloride 90–93 19
Mesoporous iron phosphate Benzylation of benzene, toluene, p-xylene 100 88
Silicotungstic acid-loaded mesoporous alumina Benzoylation of anisole with p-toluoyl chloride 80–100 89
Mesoporous sulphated Nb oxide Reaction of toluene and anisole with benzyl alcohol 100 90
Mesoporous Mo-doped carbon sphere Alkylation of toluene using benzyl alcohol 100 91
Mesoporous GaSBA-1 Benzylation of various aromatics using benzyl chloride 95–100 92
Cu(II) complex–bis(oxazoline) loaded mesoporous silica Asymmetric benzoylation of 1,2-diphenyl-1,2-ethanediol 29–40 93
Sulphated Zr–MCM-41 and Zr–MCM-48 Reaction of phenol with tert-butyl alcohol 18–91 94
Esterification Mesoporous ZnTiO3 Reaction of palmitic, lauric, oleic acid with methanol 73–92 19
Mesoporous zirconium oxophosphate Esterification of long chain fatty acid with alcohol and vice versa 46–91 85
Tungstophosphoric acid-loaded MCM-48 Esterification of palmitic, lauric, stearic, isostearic, myristic acids with different alcohols 48–97 97
Trans-Esterification Sulphonated porous zinc phosphonate Transesterification of long chain fatty acids with methanol 35–81 84
Zn-doped mesoporous silica Transesterification of ECA, EClA, EAA, EA etc. 25–94 20
Amine-grafted OMS Transesterification of glyceryl tributyrate with methanol 75–95 98
H3PW12O40/Ta2O5 porous composite Reaction of palmitic, oleic, linoleic acid with methanol 20–68 99
MgO-grafted mesoporous silica Vegetable oil with ethanol 72–96 100
Acetalization reactions Mesoporous silica Acetalization of cyclohexanone with methanol 87–89 103
Mesoporous sulphated zirconia Reaction of cyclohexanone with methanol, ethylene glycol 72–98 101
Ni–Zr supported mesoporous carbon Reaction of acetone and glycerol 68–100 104
Mesoporous SO3H-loaded organosilica Reaction of heptanal with 1-butanol 39–97 105
Aldol condensation Mg–Zr supported mesoporous carbon Cross-aldol reaction of furfural and acetone 98 107
Mesoporous γ-Al2O3 Homo aldol reaction of cyclohexanone 60–80 108
Mg–Al mixed oxide loaded HMS Condensation of benzaldehyde and heptanal 100 109
(S)-Proline loaded MCM-41 Reaction of p-nitrobenzaldehyde & 2,2-dimethyl-1,3-dioxan-5-one 20–96 110
Knoevenagel condensation Mesoporous CexZr1−xO2 Condensation of benzaldehyde and malononitrile 72–90 111
Mesoporous Ni–Al mixed oxide Reaction of benzaldehyde and malononitrile 99 112
Mesoporous amine grafted zirconia Condensation of benzaldehyde with diethyl malonate 80–98 113
Guanidine base loaded SBA-15 Condensation of cyclohexanone with benzyl cyanide 30–65 114
Urea embedded porous polymer Condensation of benzaldehyde with ECA, EAA, etc. 21–99 115
Baeyer–Villiger oxidation Mesoporous Mg–Al mixed oxide BV oxidation of cyclohexanone, cyclooctanone, methyl isopropyl ketone, ethyl methyl ketone 62–89 117
Metalloporphyrin-bridged PMO Selective conversion of cyclohexanone to ε-caprolactone 10–100 118
Mesoporous Zr-phosphate Conversion of cyclic ketones like adamantanone to lactones 96 119
Beckmann rearrangement Al doped MCM-41 Cyclohexanone oxime 46–50 121
SO3H-grafted SBA-15 Cyclohexanone oxime to ε-caprolactam 14–52 122


4.3 Different coupling reactions over mesoporous catalysts

The involvement of mesoporous solids for hydrosilylation, cyanosilylation, Diels–Alder, Suzuki reactions etc. in mediating C–C, C–N, C–H, C–S and C–O coupling and cross coupling reactions in heterogeneous media to synthesize various value-added organic fine chemicals is a widely recognized area of research nowadays. Transition metals like Cu, Ni, Fe etc. or noble metals like Au, Pt, Pd etc. supported mesoporous siliceous and non-siliceous materials, polymers or carbons, play an imperative role in these types of organic transformations.123 A concise report on all these reactions is mentioned in the following sections and these are also summarized in Table 4.
Table 4 Various organic coupling reactions carried out under liquid phase conditions over mesoporous catalysts
Type of reactions Reaction name Reaction involved Catalysts used Product yield (%) Ref.
C–C coupling reaction Sonogashira coupling Reaction of iodobenzene & phenylacetylene PdNP loaded mesoporous starch 90–99 126
Substituted aryl halides with phenylacetylene Pd-grafted triazine functionalized mesoporous polymer 60–90 125
Reaction of substituted aryl halides & phenylacetylene without Cu co-catalyst Pd-loaded periodic mesoporous silica 72–90 124
Reaction of substituted benzoyl chloride and phenyl acetylene Mesoporous tin silicates 76–97 127
Suzuki or coupling Reaction of substituted aryl halide with sodium salt of phenyltrihydroxyborate Pd-incorporated triazine functionalized mesoporous polymer 60–95 125
Reaction of substituted aryl halides with substituted phenylboronic acids Pd(II) complex anchored 2D-HMS 63–99 128
Reaction of different aryl halides with phenyl boronic acids Pd@imidazolium-functionalized SBA-15 87–99 129
Substituted aryl halides with substituted phenyl boronic acids PdNP@mesoporous carbon >99 130
Coupling of substituted aryl halides with different substituted phenyl boronic acids Pd grafted mesoporous ionic-liquid framework organosilica 55–99 131
Coupling of substituted aryl halides with different substituted phenyl boronic acids Au Schiff-base complex anchored mesoporous support 84–95 132
Heck reaction Coupling of substituted aryl halides with acrylic acid or styrene Pd loaded mesoporous polymer 50–95 125
Coupling of substituted aryl iodides with styrene Pd@mesoporous triallylamine polymer 88–95 133
Reaction of substituted aryl iodides with substituted olefins Pd supported TMG ionic liquid modified SBA-15 60–94 134
Reaction of substituted aryl halides (Cl, Br, I) with Styrene Pd supported NiFe2O4 72–97 135
Hiyama coupling Cross couplings of various substituted aryl halides with alkyl and aryl trimethoxysilane Pd loaded phloroglucinoldiimine modified PMO 65–95 124
Substituted aryl halides with different trimethoxysilane Pd NP loaded SBA-15 70–95 136
Ullmann reaction Reaction of iodobenzene to form biaryl Pd doped Ph-functionalized MCM-41 70–88 137
Homo coupling of chlorobenzene in water Pd@mesoporous silica–carbon nanocomposites 100 138
Reaction of various aryl iodides AuNP@phenylene containing PMO 80–95 139
Stille reaction Organotin compound with a sp2-hybridized organic halide Pd loaded triazole & benzene grafted PMO 60–80 140
Negishi coupling Cross coupling of an organozinc compound with an organic halide 141
C–S coupling reaction Reaction of 4-chlorothiophenol & substituted aryl iodides Mesoporous NiO–ZrO2 nanocrystals 39–89 145
Reaction of thiols with different aryl halides Mesoporous Cu–Fe-hydrotalcite 50–90 143
S-arylation of aryl iodides with thiols CuO well-dispersed on mesoporous silica 66–97 147
Reaction of substituted aryl iodide with thiophenol Cu(OAc)2 loaded mesoporous furfural functionalized silica 75–88 144
Coupling of aryl halides, benzyl bromides with thiourea Cu-grafted imine functionalized SBA-15 80–88 146
Three component coupling Biginelli condensation One pot reaction of aryl aldehyde, EAA and urea Al–MCM-41 68–89 152
Reaction of substituted aryl aldehyde, EAA and urea Fe3O4 nanoparticles@cysteine loaded SBA-15 78–85 153
One pot reaction of aryl aldehyde, EAA and urea Phosphonic acid functionalized 2D HMS 83–92 154
A3 coupling Coupling of aldehyde, alkyne and amine in glycol AgNP doped mesoporous SBA-15 40–95 150
Three component reaction of aldehyde, alkyne and amine Oxidized Cu NP supported on titania 50–88 149
Coupling of substituted aryl aldehyde, substituted phenylacetylene with different amine bases Au@ionic liquid framework PMO heterogeneous support 75–88 151


a. Reactions involving C–C coupling. Replacing metal based homogeneous catalysts which suffer from the drawbacks of low catalyst recovery, lack of regeneration of the original activity and corrosiveness, metal based heterogeneous mesoporous supports with exceptionally high surface area and tunable pore opening in nanoscale dimensions are highly desirable for different C–C homo or cross coupling reactions124 as mentioned below.

The Sonogashira coupling reaction is a highly demanding reaction in organic synthesis, which is usually catalyzed by Pd metal to form a C–C bond between a terminal alkyne and an aryl or vinyl halide in the presence of an amine base and with or without a Cu co-catalyst.124 Instead of using Pd metal directly, Pd-grafted mesoporous silica, polymer or carbons are highly appreciated as non-corrosive, stable, environmentally-friendly, highly reusable catalysts for the production of aryl or vinyl substituted alkynes using the Sonogashira reaction. Using Cu catalyst the mechanism is initiated with the simultaneous formation of copper acetylide complex when the copper ion catalyst attacks on the terminal alkyne in presence of base and oxidative addition of organic halides to Pd(0) occurs to form a Pd(II) intermediate complex. The Pd(II) intermediate on reaction with copper acetylide followed by reductive elimination gives the final C–C coupled product with regeneration of the Pd(0) catalyst.125 The reaction is reported to be highly successful over a Pd-grafted triazine functionalized ordered mesoporous polymer material125 and a Pd nanoparticles loaded mesoporous polysaccharide-derived material,126 etc. The catalysis in the former case (Pd–polysaccharide) was carried out without a Cu co-catalyst under microwave conditions at a temperature of 403 K, while the latter (Pd–triazine polymer) employs a much lower temperature (363 K) and a green solvent, water, for the coupling of substituted aryl halides with phenylacetylene in the presence of triethanol amine and a Cu catalyst. A similar experiment was carried out very efficiently by Modak et al. with a Pd-loaded periodic mesoporous silica catalyst in aqueous media at a temperature of 393 K using hexamine base and without any Cu catalyst.124 The yield was very satisfactory in this Cu-free Sonogashira reaction and a suitable mechanism has also been proposed showing the coupling between a Pd(II) intermediate and an alkyne to form an alkyne co-ordinated Pd complex which was followed by the attack of a base producing the final product with regeneration of the catalyst.124 A comparative mechanism of Sonogashira coupling without and with a Cu co-catalyst over a Pd@PMO material and a mesoporous Pd@triazine polymer, respectively, has been proposed in Fig. 15. An extraordinary approach was made by Reddy et al. who tested an acyl Sonogashira coupling reaction using acyl halide (instead of aryl halide) over a mesoporous tin silicate material (an alternative to a Pd catalyst) synthesized by a simple surfactant assisted soft-templating route.127 The reaction was performed at room temperature under solvent-free conditions using triethyl amine as a base as well as a solvent and the catalyst was highly stable for reuse for several cycles to produce a high yield of ynones.


image file: c4ra13077d-f15.tif
Fig. 15 A comparative mechanism proposed for Sonogashira coupling without and with a Cu co-catalyst over Pd@PMO and mesoporous Pd@triazine polymer, respectively.

The Suzuki or Suzuki–Miyuara cross coupling reaction first published by Akira Suzuki in 1979 is the organic reaction of an organoboronic acid with an aryl or alkyl halide in the presence of Pd(0) to form a C–C coupled product with or without using a base.124 Compared to the homogeneous counterparts, Pd-grafted mesoporous catalysts provide ease of product separation, catalyst recovery by simple filtration and excellent recycling efficiency. Recently, a number of reports have been revealed regarding this Suzuki coupling reaction studied in the presence of Pd immobilized heterogeneous mesoporous silica or polymer using mild reaction conditions. The reaction proceeds through oxidative addition of a halide to Pd(0) to form an organopalladium(II) halide. This species on reaction with an organic base followed by a nucleophilic attack by a boronate species via transmetalation forms an organopalladium(II) species. Finally, following reductive elimination of the desired biaryl or divinyl the original palladium(0) catalyst is restored.128,129 Ordered mesoporous carbon with loaded Pd nanoparticles has also been utilized for a Suzuki coupling reaction in a DMF–water solvent at a temperature of 423 K (Fig. 16).130 A base free route is suggested using a sodium salt of a boronic species instead of an organoboron compound (Fig. 16). Karimi et al. have carried out the reaction with Pd-grafted mesoporous ionic-liquid framework organosilica in aqueous media in the presence of K2CO3 under very mild conditions.131 The reaction is very successful at room temperature giving a satisfactory yield. In addition to Pd, Au metal anchored mesoporous support also has high efficiency and true heterogeneity for Suzuki homo coupling reactions under mild conditions, reports of which are not quite rare in the literature.132


image file: c4ra13077d-f16.tif
Fig. 16 Suzuki–Miyaura cross coupling reactions (A) with K2CO3 base over Pd doped ordered mesoporous carbon and (B) without base over Pd grafted mesoporous triazine polymer.

Another example of C–C cross coupling is the Heck or Mizoroki–Heck reaction involving an interaction between an unsaturated halide (or triflate) with an alkene in the presence of a base and a palladium catalyst or a palladium thioetherific-tethered catalyst to form a substituted alkene via C–C bond formation. Allowing substitution on a planar system, the reaction is very much important in organic synthesis with applications ranging from natural product and bioactive compound synthesis to the development of organic electronics.133 Bases used in this reaction are generally triethylamine, potassium carbonate and sodium acetate, etc. The reaction occurs through the oxidative addition of a halide with a Pd(0) species to form a Pd(II) halide intermediate, then formation of a π-complex with an alkene, next some rearrangement and β-elimination to form a new palladium-alkene π-complex and finally reductive elimination to give the final product with regeneration of the catalyst.133 The reaction in a non-hazardous water–ethanol mixture using a Pd loaded mesoporous polymer shows a high percentage of conversion in a very short time,125 however there is always a positive effort to use more green solvents like water for such a reaction and a bit lower reaction temperature. Mondal et al. have successfully employed a Pd-loaded triallylamine based mesoporous polymer catalyst for the Heck coupling reaction.133 The scheme in Fig. 17 illustrates a simple route for the synthesis of this Pd-based mesoporous polymer and a Heck coupling reaction over this catalyst. Another convenient and environmentally friendly method is the reaction in solvent-free conditions as reported by Ma et al. They employed a highly reusable catalyst, which consisted of Pd supported on 1,1,3,3-tetramethylguanidinium (TMG) ionic liquid modified mesoporous SBA-15, for Heck coupling of different aryl halides with substituted olefins at a temperature of 413 K to produce C–C coupled arylated olefins. The yield was high enough with respect to the reaction time and the amount of catalyst showing excellent activity for several reaction cycles without significant metal leaching.134 On the contrary, Pd supported on NiFe2O4 catalyst, highly efficient for Mizoroki–Heck reactions of chloro, bromo and iodo substituted aryl halides with styrene in DMF solvent, has the advantage of magnetic separation from the reaction mixture.135


image file: c4ra13077d-f17.tif
Fig. 17 Mizoroki–Heck coupling reaction over mesoporous Pd doped triallylamine based polymer.

Hiyama coupling, another palladium-catalyzed cross-coupling reaction used in organic natural product synthesis involves reaction of organosilanes with organic halides to form a carbon–carbon bond (C–C bond) in the presence of an activating agent such as a fluoride ion or a base. The use of an expensive fluoride source and hazardous organic solvents like DMF or DMSO etc. has recently been modified by Modak et al. who suggested a suitable mechanistic pathway for the Hiyama coupling reaction without the need for any fluoride source at a moderate temperature over a newly developed Pd-grafted PMO.124 They used NaOH to maintain the alkaline conditions and water as a solvent to make the process more environmentally-friendly. The heterogeneity of the Pd-catalyst was proven by different tests like a hot filtration test, three phase test and a solid phase poisoning test which showed minimum metal leaching with retention of catalytic activity. Recently, another interesting catalyst is a Pd nanoparticles loaded SBA-15 material fabricated through deposition of Pd nanoparticles in the micropores of SBA-15, consisting of hydrophobic trimethylsilyl (TMS) or triphenylsilyl (TPS) group functionalized mesopores.136 In Fig. 18, we have shown a scheme for the synthesis of this novel catalyst. This nanocatalyst shows excellent activity for the Hiyama cross-coupling between various aryl triethoxysilanes and aryl halides under relatively mild reaction conditions (373 K) to form different biphenyl derivatives.


image file: c4ra13077d-f18.tif
Fig. 18 A simple synthesis scheme for Pd nanoparticles doped trimethylsilyl group functionalized SBA-15 and the Hiyama coupling reaction over this material.

The Ullmann reaction, named after scientist Frietz Ullmann, is a homo coupling reaction of aryl halides to form a symmetric biaryl derivative in the presence of Cu, Pd or Au catalysts (Fig. 19). A number of reports have been published about the Ullmann reaction in aqueous media to provide a green route for the synthesis of biphenyls over heterogeneous catalysts. Ullmann reaction of iodobenzene to form a biaryl with a maximum ∼80% yield and high selectivity in water as a solvent was carried out over a novel Pd doped Ph and Al-functionalized MCM-41 catalyst.137 Coupling of substituted chlorobenzene over heterogeneous palladium catalysts supported on ordered mesoporous silica–carbon nanocomposites was carried out both at normal and elevated temperatures showing product formation increasing with temperature.138 The reaction is also catalyzed by a novel recyclable catalyst, comprising Au nanoparticles embedded bifunctional phenyl containing periodic mesoporous organosilica (Au@PMO),139 with an unexpected high yield (above 90%) and negligible leaching of metal (Fig. 19).


image file: c4ra13077d-f19.tif
Fig. 19 Ullmann homo coupling reaction over Au incorporated PMO and Stille reaction over Pd@PMO.

The Stille reaction, or Migita–Kosugi–Stille coupling, used extensively in organic synthesis is the reaction of an organotin or organo stannanes compound with an sp2-hybridized organic halide catalyzed by palladium usually performed under an inert atmosphere (Fig. 19). The mechanism involves three basic steps like in the coupling reactions: oxidative addition, transmetalation and reductive elimination. A highly efficient, reusable Pd mesoporous catalyst reported by Singh et al. was synthesized by grafting a Pd anchored triazole ligand onto the surface of a functionalized benzene containing periodic mesoporous organosilica with a uniform hexagonal arrangement prepared under basic conditions using CTAB surfactant.140

Beside these reactions, there are also Negishi coupling involving C–C cross coupling of an organozinc compound with an organic halide over a nickel or palladium based homogeneous or heterogeneous catalyst,141 and pinacol coupling, which is an acid catalyzed homo coupling of a carbonyl compound to form a 1,2-diol.142 Though the reactions over metal doped mesoporous supports have not yet been reported, extensive research is ongoing in an effort to design suitable mesoporous catalysts for carrying out these coupling reactions under mild conditions.

b. Reaction involving C–S coupling. In order to design new target molecules in synthetic organic, biological and pharmaceutical research via carbon–sulfur bond formation, the C–S coupling reaction of aryl or substituted aryl halides with different sulfur containing reagents, e.g. thiol,143 thiophenol,144 thiourea,145 etc., has drawn great attention from modern scientists. Replacing hazardous homogeneous catalysts containing transition metals like Fe, Ni, Pd, Cu and Co; metal based heterogeneous supports are much more effective for C–S coupling reactions to form diaryl sulfides in various polar organic solvents and an inert atmosphere (Fig. 20). Various metals, mixed metal oxide nanoparticles and metal complex supported mesoporous catalysts have been reported up to now suggesting that they provide an easy, reusable and simple mechanistic pathway for this aryl-sulfur coupling reaction.144
image file: c4ra13077d-f20.tif
Fig. 20 A general scheme for a C–S coupling reaction over a mesoporous solid catalyst.

The mesoporous Cu–Fe-hydrotalcite solid catalyst is a ligand free, novel, environmentally green catalyst for C–S coupling reactions using thiols with different aryl halides in DMF solvent and with K2CO3 base. The homogeneous mixed phase of two oxides (CuO and Fe2O3), good BET surface area (123 m2 g−1) and large pore diameter (∼15 nm) mainly contribute to the observed high conversion of products achieved within a reaction time of 4–10 h.143 CuO well-dispersed on mesoporous silica prepared by a sol–gel route is a highly reusable heterogeneous support for S-arylation of aryl iodides with thiols in the presence of Cs2CO3 base and DMSO solvent. Cu(Oac)2 loaded on organic inorganic hybrid mesoporous silica functionalized with furfural groups is also applied for aryl-sulfur coupling of thiophenol with aryl iodide in the presence of K2CO3 base and DMF solvent.144 The yield observed for different substituted diaryl sulfides was very satisfactory with proper reusable efficiency shown up to the 5th cycle. A synthesis scheme of mesoporous Cu grafted furfural functionalized organic–inorganic hybrid silica is shown in Fig. 21. A probable mechanism has also been suggested showing generation of Cu(0) active species from Cu(II) grafted organosilica in the presence of a base, followed by the consecutive oxidation addition of aryl iodide and thiophenol to form another intermediate. As observed in the catalytic cycle the intermediate formed releases an aryl-sulfur coupled product with regeneration of the active species via a reductive elimination step (Fig. 21).144


image file: c4ra13077d-f21.tif
Fig. 21 Synthesis scheme of mesoporous Cu grafted furfural functionalized organosilica and its role in the mechanistic catalytic C–S cross coupling reaction cycle.

To avoid the toxicity and hazards of the polar solvents used in different C–S coupling reactions, we explored the reaction in a completely green solvent, water, over NiO–ZrO2 self-assembled mesoporous nanocrystals at a relatively lower temperature (353 K).145 4-Chlorothiophenol is used as a sulfur source and different substituted aryl iodides are used to produce value added aryl-sulfur compounds with an appreciable yield. NiO species present in the material mainly take the active part in the catalysis which is illustrated with a proposed mechanism in Fig. 20. As an alternative to volatile and toxic thiol or foul-smelling thiophenol sulfur sources, which lead to enormous environmental and safety problems, thiourea is widely used nowadays for large scale operations. Three component one-pot coupling of aryl halides, benzyl bromides and thiourea in aqueous media is another significant approach towards green organic synthesis.146 A Cu-grafted mesoporous organo functionalized SBA-15 catalyst synthesized via a simple post-grafting technique was employed for this one-pot thioetherification carried out at a temperature of 373 K. A significant yield of aryl sulfides is obtained using this heterogeneous, non-toxic Cu-mesoporous catalyst.146

There are also other coupling reactions like C–N and C–O, etc. which to the best of our knowledge have not yet been observed to proceed over heterogeneous mesoporous materials.

c. Three component coupling reaction. There are some three component one-pot coupling reactions which are observed to be catalyzed by mesoporous catalysts in liquid phase heterogeneous media. These reactions of a variety of organic reagents to synthesize new heterocyclic compounds are highly important in different fields for synthesis of natural products, fine chemicals and pharmaceutical components, etc. One such reaction is the one-pot thioetherification involving C–S coupling which has already been mentioned in the previous section.147

Another highly significant three component coupling reaction is A3 coupling, i.e. the coupling of aldehyde, alkyne and amine, used extensively in the manufacturing of various nitrogen-containing heterocycles, natural products and biologically active compounds, etc. to form propargylamine in a convenient approach. Though many metals viz. Au, Ag, Cu, In, etc. based homogeneous catalysts showed efficiency for A3 coupling reactions,148 but due to some disadvantages for e.g. difficulty of separation of those catalysts from reaction mixture, lack of recycling efficiency and the possibility of metal leaching etc. metal immobilized heterogeneous supports are much more desirable for such reactions.149 Ag nanoparticle supported mesoporous SBA-15 has been proven to be highly efficient for A3 coupling reactions in glycol as a ‘green’ solvent at a temperature of 373 K. The product conversion observed was believed to be highly dependent on Ag particle size and a high yield of propargylamines was obtained with ca. 8 nm optimized size Ag-loaded SBA-15.150 A 2D-hexagonal periodic mesoporous organosilica containing imidazolium ionic liquid framework synthesized in acidic conditions with P123 surfactant has been functionalized with 0.2 mol% of Au and a thorough investigation was performed to study the catalytic ability of that Au loaded PMO heterogeneous support for the A3 coupling reaction under mild conditions.151 In Fig. 22 we have illustrated the synthesis of the Au@ionic liquid PMO catalyst and its application in an A3 coupling reaction. Chloroform has been proven to be the most appropriate solvent for the reaction at a relatively lower temperature (333 K) with a high average yield of around 70–80% for the corresponding propargylamine derivatives.


image file: c4ra13077d-f22.tif
Fig. 22 A synthesis scheme for mesoporous Au nanoparticles doped imidazole functionalized PMO and an A3 coupling reaction over this catalyst.

Biginelli condensation is another three component one-pot reaction which creates 3,4-dihydropyrimidin-2(1H)-ones from aryl aldehyde, ethyl acetoacetate (EAA) and urea in the presence of a Brønsted or Lewis acid as a catalyst. In 2010, an Al–MCM-41 mesoporous catalyst was employed for this reaction in octane solvent to obtain various essential substituted dihydropyrimidinones with high yields and a high catalytic yield was obtained with a small loss of activity after repeated use.152 Later in 2012, superparamagnetic Fe3O4 nanoparticles impregnated on cysteine functionalized mesoporous SBA-15 was designed and it has been proven to be a highly reusable and magnetically recoverable catalyst for the one-pot synthesis of 3,4-dihydropyrimidin-2(1H)-ones via the Biginelli reaction in pollution free ethanol solvent.153 A method of synthesis of the Fe3O4@cysteine functionalized SBA-15 material is shown in Fig. 23. An idea of a one-pot reaction and separation of the catalyst with a large bar magnet is also illustrated in this figure. The catalyst allowed for a high yield of products using different substituted aldehydes within a very short reaction time and a suitable catalytic cycle was suggested to be taking place over the Lewis acidic Fe3O4 active sites of the catalyst. A more environment friendly approach was made by Pramanik et al. who synthesized a novel phosphonic acid functionalized 2D hexagonal mesoporous organosilica and applied this metal free catalyst for a Biginelli condensation reaction under solvent-free conditions at 333 K.154 The reaction pathway is very clean with a very high yield of the products obtained and good recycling efficiency. The high activity of this heterogeneous organocatalyst can be attributed to the high Brønsted acidity of the phosphonic group along with the huge surface area and large pore diameter of the silica support.


image file: c4ra13077d-f23.tif
Fig. 23 Magnetically recoverable Fe3O4@cysteine grafted mesoporous SBA-15 catalyst and its role in a one-pot Biginelli condensation reaction.

Besides the above mentioned organic reactions taking place over mesoporous supports, other reactions like Diels–Alder reactions, hydroformylation of olefins, aldehydes, polymerization reactions etc. are also found to be efficiently catalyzed by some metal supported or metal-free mesoporous materials under mild reaction conditions.155–157 Further, hierarchical porosity can be introduced in the ordered porous silica composites by using anionic polystyrene spheres and triblock copolymers as templates158 and the resulting materials can potentially be utilized as supports in heterogeneous catalysis for bulkier molecules where diffusion of the reactant molecules could be facilitated.

5. Future perspectives and conclusions

To give an overview of the potential of mesoporous siliceous and non-siliceous materials in liquid phase heterogeneous catalysis, the present review offers a concise idea showing numerous examples of organic transformations with proper illustrations. Additionally, an outline of the synthesis strategies of the functionalized mesoporous materials presented here helps with suggestive information about how to design and prepare the mesoporous catalysts, tune their pore structure and specific surface area as well as how to impregnate them with various active transition metal nanoparticles, complexes or organic functional groups at the pore surface. Based on this one can prepare a number of new metal doped and organically functionalized mesoporous materials and explore their potential in versatile catalytic transformations under liquid phase reaction conditions. Replacing conventional hazardous, toxic homogeneous catalysts, which create serious environmental problems and ecological imbalance, many reactions mediated by those metallocatalysts or metal-free heterogeneous organocatalysts are carried out in aqueous solvents or without solvent, at lower temperatures or without the use of any toxic reagents. These green catalytic pathways are in high demand for the development of industrial manufacturing processes. Thus, it is expected that in future, more and more applications of these green catalysts in each and every industrial fine chemical synthesis will create a pollution free world for the benefit of all human beings.

List of abbreviations

nmNanometer
3D3 dimensional
CMCCritical micelle concentration
$Dollar sign
VO(acac)2Vanadyl(IV) acetylacetonate
Cp*Cyclopentadienyl
MPaMegapascal
RTRoom temperature
HMSHexagonal mesoporous silica
EAEthyl acrylate
ECAEthyl cyanoacetate
EclAEthyl chloroacetate
EAAEthyl acetoacetate
TMG1,1,3,3-Tetramethylguanidinium
NPNanoparticle
BETBrunauer–Emmett–Teller
DMFDimethyl formamide
DMSODimethyl sulfoxide
AIBN2,2′-Azobis(isobutyronitrile)

Acknowledgements

AB wishes to thank DST, New Delhi for financial support through DST-SERB and DST-UKIERI research grants. NP is grateful to Department of Atomic Energy, Government of India for financial support.

Notes and references

  1. M. H. Valkenberg and W. F. Hölderich, Catal. Rev., 2002, 44, 321 CAS.
  2. B. Lindström and L. J. Pettersson, A brief history of catalysis, CATTECH, 2003, 7, 4130 Search PubMed.
  3. J. Pérez-Ramírez, C. H. Christensen, K. Egeblad, C. H. Christensen and J. C. Groen, Chem. Soc. Rev., 2008, 37, 2530 RSC.
  4. N. Pal and A. Bhaumik, Adv. Colloid Interface Sci., 2013, 189–190, 21 CrossRef CAS PubMed.
  5. S. Lopez-Orozco, A. Inayat, A. Schwab, T. Selvam and W. Schwieger, Adv. Mater., 2011, 23, 2602 CrossRef CAS PubMed.
  6. Z. M. Li, Y. Zhou, D. J. Tao, W. Huang, X. S. Chen and Z. Yang, RSC Adv., 2014, 4, 12160 RSC.
  7. P. B. Venuto, Microporous Mater., 1994, 2, 297 CrossRef CAS.
  8. C. T. Kresge, M. E. Leonowicz, W. J. Roth, C. E. Vartuli and J. S. Beck, Nature, 1992, 359, 710 CrossRef CAS.
  9. A. Corma, Chem. Rev., 1997, 97, 2373 CrossRef CAS PubMed.
  10. A. Taguchi and F. Schüth, Microporous Mesoporous Mater., 2005, 77, 1 CrossRef CAS PubMed.
  11. C. Perego and R. Millini, Chem. Soc. Rev., 2013, 42, 3956 RSC.
  12. R. M. Martín-Aranda and J. Cčjka, Top. Catal., 2010, 53, 141 CrossRef.
  13. S. Fujita and S. Inagaki, Chem. Mater., 2008, 20, 891–908 CrossRef CAS.
  14. J. Rouquerol, D. Avnir, C. W. Fairbridge, D. H. Everett, J. M. Haynes, N. Pernicone, J. D. F. Ramsay, K. S. W. Sing and K. K. Unger, Pure Appl. Chem., 1994, 66, 1739 CrossRef CAS.
  15. D. Sengupta, J. Saha, G. De and B. Basu, J. Mater. Chem. A, 2014, 2, 3986 CAS.
  16. C. M. A. Parlett, K. Wilson and A. F. Lee, Chem. Soc. Rev., 2013, 42, 3876 RSC.
  17. T. Yokoi, M. Yoshioka, H. Ima and T. Tatsumi, Angew. Chem., Int. Ed., 2009, 48, 9884–9887 CrossRef CAS PubMed.
  18. M. Pramanik and A. Bhaumik, Chem.–Eur. J., 2013, 19, 8507 CrossRef CAS PubMed.
  19. N. Pal, M. Paul and A. Bhaumik, Appl. Catal., A, 2011, 393, 153 CrossRef CAS PubMed.
  20. N. Pal, M. Paul and A. Bhaumik, J. Solid State Chem., 2011, 184, 1805 CrossRef CAS PubMed.
  21. T. Yanagisawa, T. Shimizu, K. Kuroda and C. Kato, Bull. Chem. Soc. Jpn., 1990, 63, 988 CrossRef CAS.
  22. D. Zhao, Q. Huo, J. Feng, B. F. Chmelka and G. D. Stucky, J. Am. Chem. Soc., 1998, 120, 6024 CrossRef CAS.
  23. U. Ciesla and F. Schüth, Microporous Mesoporous Mater., 1999, 27, 131 CrossRef CAS.
  24. R. Ryoo, S. H. Joo and S. Jun, J. Phys. Chem. B, 1999, 103, 7743 CrossRef CAS.
  25. X. Lai, X. Li, W. Geng, J. Tu, J. Li and S. Qiu, Angew. Chem., Int. Ed., 2007, 46, 738 CrossRef CAS PubMed.
  26. C. Liang, Z. Li and S. Dai, Angew. Chem., Int. Ed., 2008, 47, 3696 CrossRef CAS PubMed.
  27. X. P. Fang, X. Q. Yu, S. F. Liao, Y. F. Shi, Y. S. Hu, Z. X. Wang, G. D. Stucky and L. Q. Chen, Microporous Mesoporous Mater., 2012, 151, 418–423 CrossRef CAS PubMed.
  28. A. Bhaumik and S. Inagaki, J. Am. Chem. Soc., 2001, 123, 691–696 CrossRef CAS PubMed.
  29. J. Liu, T. Y. Yang, D. W. Wang, G. Q. M. Lu, D. Y. Zhao and S. Z. Qiao, Nat. Commun., 2013, 4, 2798 Search PubMed.
  30. K. Ariga, A. Vinu, Y. Yamauchi, Q. M. Ji and J. P. Hill, Bull. Chem. Soc. Jpn., 2012, 85, 1–32 CrossRef CAS.
  31. Y. Zheng, Y. Jiao, J. Chen, J. Liu, J. Liang, A. Du, W. M. Zhang, Z. H. Zhu, S. C. Smith, M. Jaroniec, G. Q. Lu and S. Z. Qiao, J. Am. Chem. Soc., 2011, 133, 20116–20119 CrossRef CAS PubMed.
  32. A. J. B. Robertson, Platinum Met. Rev., 1975, 19, 64 CAS.
  33. C. Elschenbroich, Organometallics, Wiley-VCH, Weinheim, 2006, ISBN 978-3-527-29390-2 Search PubMed.
  34. W. K. Lewis and E. D. Ries, Ind. Eng. Chem., 1925, 17, 593 CrossRef CAS.
  35. X. S. Zhao, X. Y. Bao, W. Guo and F. Y. Lee, Mater. Today, 2006, 9, 32 CrossRef CAS.
  36. K. Wilson and J. H. Clark, Pure Appl. Chem., 2000, 72, 1313 CrossRef CAS.
  37. S. Samanta, N. K. Mal and A. Bhaumik, J. Mol. Catal. A: Chem., 2005, 236, 7 CrossRef CAS PubMed.
  38. N. Pal, E.-B. Cho and D. Kim, RSC Adv., 2014, 4, 9213 RSC.
  39. A. Bhaumik and T. Tatsumi, J. Catal., 2000, 189, 31 CrossRef CAS.
  40. Y. Zhang, F. Gao, H. Wan, C. Wu, Y. Kong, X. Wu, B. Zhao, L. Dong and Y. Chen, Microporous Mesoporous Mater., 2008, 113, 393 CrossRef CAS PubMed.
  41. N. Pal, E.-B. Cho, D. Kim and M. Jaroniec, J. Phys. Chem. C, 2014, 118, 15892 CAS.
  42. W. Huang, J. N. Kuhn, C.-K. Tsung, Y. Zhang, S. E. Habas, P. Yang and G. A. Somorjai, Nano Lett., 2008, 8, 2027 CrossRef CAS PubMed.
  43. S. Verma, M. Nandi, A. Modak, S. L. Jain and A. Bhaumik, Adv. Synth. Catal., 2011, 353, 1897–1902 CrossRef CAS.
  44. Y. Rao and D. M. Antonelli, J. Mater. Chem., 2009, 19, 1937 RSC.
  45. A. Modak, M. Nandi and A. Bhaumik, Catal. Today, 2012, 198, 45 CrossRef CAS PubMed.
  46. S. Jana, B. Dutta, R. Bera and S. Koner, Langmuir, 2007, 23, 2492 CrossRef CAS PubMed.
  47. E. A. Prasetyanto and S.-E. Park, Bull. Korean Chem. Soc., 2008, 29, 1033 CrossRef CAS.
  48. P. Selvam, V. M. Ravat and V. Krishna, J. Phys. Chem. C, 2011, 115, 1922 CAS.
  49. W. Tanglumlert, T. Imae, T. J. White and S. Wongkasemjit, Catal. Commun., 2009, 10, 1070 CrossRef CAS PubMed.
  50. P. Wu, Z. Xiong, K. P. Loh and X. S. Zhao, Catal. Sci. Technol., 2011, 1, 285 CAS.
  51. Z. Bohström, I. Rico-Lattes and K. Holmberg, Green Chem., 2010, 12, 1861 RSC.
  52. A. K. Patra, A. Dutta and A. Bhaumik, Chem.–Eur. J., 2013, 19, 12388 CrossRef CAS.
  53. U. Mandi, M. Pramanik, A. S. Roy, N. Salam, A. Bhaumik and S. M. Islam, RSC Adv., 2014, 4, 15431 RSC.
  54. P. Zhang, H. Li, G. M. Veith and S. Dai, Adv. Mater., 2015, 27, 234 CrossRef CAS PubMed.
  55. P. Zhang, Y. Gong, H. Li, Z. Chen and Y. Wang, Nat. Commun., 2013, 4, 1593 CrossRef PubMed.
  56. Y. Wang, J. Zhang, X. Wang, M. Antonietti and H. Li, Angew. Chem., Int. Ed., 2010, 49, 3356 CrossRef CAS PubMed.
  57. P. Zhang, H. Li and Y. Wang, Chem. Commun., 2014, 50, 6312 RSC.
  58. P. Zhang, Y. Wang, J. Yao, C. Wang, C. Yan, M. Antonietti and H. Li, Adv. Synth. Catal., 2011, 353, 1447 CrossRef CAS.
  59. S. Martínez-González, A. Gómez-Avilés, O. Martynyuk, A. Pestryakov, N. Bogdanchikova and V. Cortés Corberán, Catal. Today, 2014, 227, 65 CrossRef PubMed.
  60. I. Tamiolakis, I. N. Lykakis, A. P. Katsoulidis, M. Stratakis and G. S. Armatas, Chem. Mater., 2011, 23, 4204 CrossRef CAS.
  61. L. Geng, X. Zhang, W. Zhang, M. Jia and G. Liu, Chem. Commun., 2014, 50, 2965 RSC.
  62. Z.-A. Qiao, P. Zhang, S.-H. Chai, M. Chi, G. M. Veith, N. C. Gallego, M. Kidder and S. Dai, J. Am. Chem. Soc., 2014, 136, 11260 CrossRef CAS PubMed.
  63. N. Salam, P. Mondal, J. Mondal, A. S. Roy, A. Bhaumik and S. M. Islam, RSC Adv., 2012, 2, 6464 RSC.
  64. D. H. Koo, M. Kim and S. Chang, Org. Lett., 2005, 7, 5015 CrossRef CAS PubMed.
  65. V. Ayala, A. Corma, M. Iglesias and F. Sánchez, J. Mol. Catal. A: Chem., 2004, 221, 201 CAS.
  66. P. Zhang, Y. Wang, H. Li and M. Antonietti, Green Chem., 2012, 14, 1904 RSC.
  67. C.-F. Chang and S.-T. Liu, J. Mol. Catal. A: Chem., 2009, 299, 121 CrossRef CAS PubMed.
  68. S. Gontier and A. Tuel, J. Catal., 1995, 157, 124 CrossRef CAS.
  69. J. Evans, A. B. Zaki, M. Y. El-Sheikh and S. A. El-Safty, J. Phys. Chem. B, 2000, 104, 10271 CrossRef CAS.
  70. C. Perego, A. Carati, P. Ingallina, M. A. Mantegazza and G. Bellussi, Appl. Catal., A, 2001, 221, 63 CrossRef CAS.
  71. A. Bhaumik, M. P. Kapoor and S. Inagaki, Chem. Commun., 2003, 470 RSC.
  72. X. Liang, R. Yang, G. Li and C. Hu, Microporous Mesoporous Mater., 2013, 182, 62 CrossRef CAS PubMed.
  73. A. Dubey, M. Choi and R. Ryoo, Green Chem., 2006, 8, 144 RSC.
  74. P. Borah, X. Ma, K. T. Nguyen and Y. Zhao, Angew. Chem., Int. Ed., 2012, 51, 7756 CrossRef CAS PubMed.
  75. P. Zhang, Y. Gong, H. Li, Z. Chen and Y. Wang, RSC Adv., 2013, 3, 5121 RSC.
  76. P. Sreenivasulu, D. Nandan, M. Kumar and N. Viswanadham, J. Mater. Chem. A, 2013, 1, 3268 CAS.
  77. E. Rodríguez-Castellón, L. Díaz, P. Braos-García, J. Mérida-Robles, P. Maireles-Torres, A. Jiménez-López and A. Vaccari, Appl. Catal., A, 2003, 240, 83 CrossRef.
  78. G. Liu, J. Wang, T. Huang, X. Liang, Y. Zhang and H. Li, J. Mater. Chem., 2010, 20, 1970 RSC.
  79. R. Liu, T. Cheng, L. Kong, C. Chen, G. Liu and H. Li, J. Catal., 2013, 307, 55 CrossRef CAS PubMed.
  80. J. Liu, S. Zou, L. Xiao and J. Fan, Catal. Sci. Technol., 2014, 4, 441 CAS.
  81. M. Paul, N. Pal and A. Bhaumik, Eur. J. Inorg. Chem., 2010, 5129 CrossRef CAS.
  82. V. N. Sapunov, M. Y. Grigoryev, E. M. Sulman, M. B. Konyaeva and V. G. Matveeva, J. Phys. Chem. A, 2013, 117, 4073 CrossRef CAS PubMed.
  83. Z. Li, J. Liu, C. Xia and F. Li, ACS Catal., 2013, 3, 2440 CrossRef CAS.
  84. M. Nandi, J. Mondal, K. Sarkar, Y. Yamauchi and A. Bhaumik, Chem. Commun., 2011, 47, 6677 RSC.
  85. M. Pramanik, M. Nandi, H. Uyama and A. Bhaumik, Green Chem., 2012, 14, 2273 RSC.
  86. S. K. Das, M. K. Bhunia, A. K. Sinha and A. Bhaumik, ACS Catal., 2011, 1, 493 CrossRef CAS.
  87. T. Tsuchimoto, K. Tobita, T. Hiyama and S.-I. Fukuzawa, J. Org. Chem., 1997, 62, 6997 CrossRef CAS.
  88. X. Zhang, T. Lin, R. Li, T. Bai and G. Zhang, Ind. Eng. Chem. Res., 2012, 51, 3541 CrossRef CAS.
  89. S. Selvakumar and A. P. Singh, Catal. Lett., 2009, 128, 363 CrossRef CAS PubMed.
  90. Y. Rao, M. Trudeau and D. Antonelli, J. Am. Chem. Soc., 2006, 128, 13996 CrossRef CAS PubMed.
  91. J. Dou and H. C. Zeng, J. Phys. Chem. C, 2012, 116, 7767 CAS.
  92. C. Anand, B. Sathyaseelan, L. Samie, A. Beitollahi, R. Pradeep Kumar, M. Palanichamy, V. Murugesan, E.-R. Kenawy, S. S. Al-Deyab and A. Vinu, Microporous Mesoporous Mater., 2010, 134, 87 CrossRef CAS PubMed.
  93. A. R. Silva, L. Carneiro, A. P. Carvalho and J. Pires, Catal. Sci. Technol., 2013, 3, 2415 CAS.
  94. Q. Zhao, Q. Wang, D. Wu, X. Fu, T. Jiang and H. Yin, J. Ind. Eng. Chem., 2012, 18, 1326 CrossRef CAS PubMed.
  95. A. K. Singh, S. D. Fernando and R. Hernandez, Energy Fuels, 2007, 21, 1161 CrossRef CAS.
  96. Q. Yang, J. Liu, L. Zhang and C. Li, J. Mater. Chem., 2009, 19, 1945 RSC.
  97. A. Sakthivel, K. Komura and Y. Sugi, Ind. Eng. Chem. Res., 2008, 47, 2538 CrossRef CAS.
  98. V. V. Guerrero and D. F. Shantz, Ind. Eng. Chem. Res., 2009, 48, 10375 CrossRef CAS.
  99. L. Xu, Y. Wang, X. Yang, X. Yu, Y. Guo and J. H. Clark, Green Chem., 2008, 10, 746 RSC.
  100. E. Li and V. Rudolph, Energy Fuels, 2008, 22, 145 CrossRef CAS.
  101. A. Sinhamahapatra, N. Sutradhar, M. Ghosh, H. C. Bajaj and A. B. Panda, Appl. Catal., A, 2011, 402, 87 CrossRef CAS PubMed.
  102. S. B. Umbarkar, T. V. Kotbagi, A. V. Biradar, R. Pasricha, J. Chanale, M. K. Dongare, A.-S. Mamede, C. Lancelot and E. Payen, J. Mol. Catal. A: Chem., 2009, 310, 150 CrossRef CAS PubMed.
  103. M. Iwamoto, Y. Tanaka, N. Sawamur and S. Namba, J. Am. Chem. Soc., 2003, 125, 13032 CrossRef CAS PubMed.
  104. M. S. Khayoon and B. H. Hameed, Appl. Catal., A, 2013, 464–465, 191 CrossRef CAS PubMed.
  105. M. Rat, M. H. Zahedi-Niaki, S. Kaliaguine and T. O. Do, Microporous Mesoporous Mater., 2008, 112, 26 CrossRef CAS PubMed.
  106. C. Noda Pérez, C. A. Pérez, C. A. Henriques and J. L. F. Monteiro, Appl. Catal., A, 2004, 272, 229 CrossRef PubMed.
  107. L. Faba, E. Díaz and S. Ordóñez, ChemSusChem, 2013, 6, 463 CrossRef CAS PubMed.
  108. P. Bai, P. Wu, Z. Yan and X. S. Zhao, J. Mater. Chem., 2009, 19, 1554 RSC.
  109. G. D. Yadav and P. Aduri, J. Mol. Catal. A: Chem., 2012, 355, 142 CrossRef CAS PubMed.
  110. E. G. Doyagüez, F. Calderón, F. Sánchez and A. Fernández-Mayoralas, J. Org. Chem., 2007, 72, 9353 CrossRef PubMed.
  111. G. Postole, B. Chowdhury, B. Karmakar, K. Pinki, J. Banerji and A. Auroux, J. Catal., 2010, 269, 110 CrossRef CAS PubMed.
  112. M. N. Pahalagedara, L. R. Pahalagedara, C.-H. Kuo, S. Dharmarathna and S. L. Suib, Langmuir, 2014, 30, 8228 CrossRef CAS PubMed.
  113. K. M. Parida, S. Mallick, P. C. Sahoo and S. K. Rana, Appl. Catal., A, 2010, 381, 226 CrossRef CAS PubMed.
  114. K.-S. Kim, J. H. Song, J.-H. Kim and G. Seo, Stud. Surf. Sci. Catal., 2003, 146, 505 CrossRef CAS.
  115. R. Zhu, J. Shen, Y. Wei and F. Zhang, New J. Chem., 2011, 35, 1861 RSC.
  116. A. Baeyer and V. Villiger, Chem. Ber., 1899, 32, 3625 CrossRef.
  117. M. Paul, N. Pal, J. Mondal, M. Sasidharan and A. Bhaumik, Chem. Eng. Sci., 2012, 71, 564 CrossRef CAS PubMed.
  118. E.-Y. Jeong, M. B. Ansari and S.-E. Park, ACS Catal., 2011, 1, 855 CrossRef CAS.
  119. A. Sinhamahapatra, A. Sinha, S. K. Pahari, N. Sutradhar, H. C. Bajaja and A. B. Panda, Catal. Sci. Technol., 2012, 2, 2375 CAS.
  120. R. Palkovits, C.-M. Yang, S. Olejnik and F. Schüth, J. Catal., 2006, 243, 93 CrossRef CAS PubMed.
  121. C. Ngamcharussrivichai, P. Wu and T. Tatsumi, J. Catal., 2004, 227, 448 CrossRef CAS PubMed.
  122. X. Wang, C.-C. Chen, S.-Y. Chen, Y. Mou and S. Cheng, Appl. Catal., A, 2005, 281, 47 CrossRef CAS PubMed.
  123. A. Suzuki, Angew. Chem., Int. Ed., 2011, 50, 6722 CrossRef CAS PubMed.
  124. A. Modak, J. Mondal and A. Bhaumik, Green Chem., 2012, 14, 2840 RSC.
  125. A. Modak, J. Mondal, M. Sasidharan and A. Bhaumik, Green Chem., 2011, 13, 1317 RSC.
  126. V. L. Budarin, J. H. Clark, R. Luque, D. J. Macquarrie and R. J. White, Green Chem., 2008, 10, 382 RSC.
  127. K. R. Reddy, M. Suresh, M. L. Kantam, S. K. Bhargava and P. Srinivasu, Ind. Eng. Chem. Res., 2014, 53, 18630 CrossRef CAS.
  128. K. Sarkar, M. Nandi, M. Islam, M. Mubarak and A. Bhaumik, Appl. Catal., A, 2009, 352, 81 CrossRef CAS PubMed.
  129. P. Han, H. Zhang, X. Qiu, X. Ji and L. Gao, J. Mol. Catal. A: Chem., 2008, 295, 57 CrossRef CAS PubMed.
  130. H. R. Choi, H. Woo, S. Jang, J. Y. Cheon, C. Kim, J. Park, K. H. Park and S. H. Joo, ChemCatChem, 2012, 4, 1587 CrossRef CAS.
  131. B. Karimi, D. Elhamifar, J. H. Clark and A. J. Hunt, Chem.–Eur. J., 2010, 16, 8047 CrossRef CAS PubMed.
  132. K. M. Parida, S. Singha, P. C. Sahoo and S. Sahu, J. Mol. Catal. A: Chem., 2011, 342–343, 11 CrossRef CAS PubMed.
  133. J. Mondal, A. Modak and A. Bhaumik, J. Mol. Catal. A: Chem., 2011, 350, 40 CrossRef CAS PubMed.
  134. X. Ma, Y. Zhou, J. Zhang, A. Zhu, T. Jiang and B. Han, Green Chem., 2008, 10, 59 RSC.
  135. B. Baruwati, D. Guin and S. V. Manorama, Org. Lett., 2007, 9, 5377 CrossRef CAS PubMed.
  136. S.-H. Huang, C.-H. Liu and C.-M. Yang, Green Chem., 2014, 16, 2706 RSC.
  137. H. Li, J. Chen, Y. Wan, W. Chai, F. Zhang and Y. Lu, Green Chem., 2007, 9, 273 RSC.
  138. Y. Wan, H. Wang, Q. Zhao, M. Klingstedt, O. Terasaki and D. Zhao, J. Am. Chem. Soc., 2009, 131, 4541 CrossRef CAS PubMed.
  139. B. Karimi and F. K. Esfahani, Chem. Commun., 2011, 47, 10452 RSC.
  140. A. P. Singh and P. Sharma, RSC Adv., 2014, 4, 43070 RSC.
  141. S. W. Smith and G. C. Fu, Angew. Chem., 2008, 120, 9474 CrossRef.
  142. Y.-S. Yang, Z.-L. Shen and T.-P. Loh, Org. Lett., 2009, 11, 2213 CrossRef CAS PubMed.
  143. D. K. Dumbrea, P. R. Selvakannana, S. K. Patil, V. R. Choudhary and S. K. Bhargava, Appl. Catal., A, 2014, 476, 54 CrossRef PubMed.
  144. J. Mondal, A. Modak, A. Dutta and A. Bhaumik, Dalton Trans., 2011, 5228 RSC.
  145. N. Pal and A. Bhaumik, Dalton Trans., 2012, 9161 RSC.
  146. J. Mondal, A. Modak, A. Dutta, S. Basu, S. N. Jha, D. Bhattacharyya and A. Bhaumik, Chem. Commun., 2012, 48, 8000 RSC.
  147. C.-K. Chen, Y.-W. Chen, C.-H. Lin, H.-P. Lin and C.-F. Lee, Chem. Commun., 2010, 46, 282 RSC.
  148. V. A. Peshkov, O. P. Pereshivko and E. V. Van der Eycken, Chem. Soc. Rev., 2012, 41, 3790 RSC.
  149. M. J. Albaladejo, F. Alonso, Y. Moglie and M. Yus, Eur. J. Org. Chem., 2012, 3093 CrossRef CAS.
  150. G.-P. Yonga, D. Tiana, H.-W. Tonga and S.-M. Liu, J. Mol. Catal. A: Chem., 2010, 323, 40 CrossRef PubMed.
  151. B. Karimi, M. Gholinejad and M. Khorasani, Chem. Commun., 2012, 48, 8961 RSC.
  152. H. Murata, H. Ishitani and M. Iwamoto, Org. Biomol. Chem., 2010, 8, 1202 CAS.
  153. J. Mondal, T. Sen and A. Bhaumik, Dalton Trans., 2012, 6173 RSC.
  154. M. Pramanik and A. Bhaumik, ACS Appl. Mater. Interfaces, 2014, 6, 933 CAS.
  155. S. Minakata, T. Nagamachi, K. Nakayama, T. Suzuki and T. Tanaka, Chem. Commun., 2011, 47, 6338 RSC.
  156. W. Zhou and D. He, Green Chem., 2009, 11, 1146 RSC.
  157. M. Pramanik, M. Nandi, H. Uyama and A. Bhaumik, Catal. Sci. Technol., 2012, 2, 613 CAS.
  158. T. Sen, G. J. T. Tiddy, J. L. Casci and M. W. Anderson, Chem. Mater., 2004, 16, 2044 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.