You are on page 1of 33

Applied Clay Science 53 (2011) 106–138

Contents lists available at ScienceDirect

Applied Clay Science


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / c l a y

Review Article

Organic synthesis using clay and clay-supported catalysts


Gopalpur Nagendrappa ⁎,1
Department of Chemistry, Bangalore University, Bangalore 560 001, India

a r t i c l e i n f o a b s t r a c t

Article history: Clays and modified clays are used to catalyze various types of organic reactions such as addition, Michael
Received 20 May 2010 addition, carbene addition and insertion, hydrogenation, allylation, alkylation, acylation, pericyclic reactions,
Received in revised form 17 October 2010 condensation reactions, aldol formation, imine synthesis, diazotization reactions, synthesis of heterocycles,
Accepted 19 October 2010
esterification reactions, rearrangement/isomerization reactions, cyclization reactions, oxidation of alcohols,
Available online 6 October 2010
dehydrogenation, epoxidation and several more. Clays function as Brønsted and/or Lewis acids, or as bases.
Keywords:
Clays with combined acidic and basic properties have been developed by simple procedures of modification.
Clay mineral Such clays are employed to catalyze a sequence of acid and base-catalyzed reactions in one pot. Good
Activated bentonite enantioselectivity and stereoselectivity are achieved using chiral organic compounds and chiral complexes
Montmorillonite intercalated between clay layers. Examples from recent literature are described here.
Saponite © 2010 Elsevier B.V. All rights reserved.
Organic synthesis
Heterogeneous catalyst

1. Introduction compatibility and cheapness, much effort is expended in discovering


newer methods of using clays in their native and modified forms as
Clays are widespread, easily available and low-cost chemical catalysts for diverse organic reactions.
substances. Both in their native state and in numerous modified Clays have a long history of use as catalysts and as supports in organic
forms, clays are versatile materials that catalyze a variety of chemical reactions (Vogels et al., 2005). Several excellent reviews on clay
reactions. Just as they can be molded into any shape, their micro catalyzed organic reactions have appeared in the recent past (Varma,
structure can be changed to suit chemists' needs to promote diverse 2002; Dasgupta and Török, 2008; Ranu and Chattopadhyay, 2009). Zhou
chemical reactions. It is convincingly argued that clays initiated, (2010) has briefly summarized the emerging trends in synthetic clay-
supported and sustained the process of formation of small molecules based materials. The present review attempts to report some of the
on the earth millions of years ago, which gradually developed into developments that have taken place in the area of organic synthesis
more complex molecules. In the course of time, there emerged from using clays and clay-supported catalysts during the past decade.
the latter the self replicating assemblies that evolved into simple life Much of the work on clays focus on the use of “normal” smectites,
forms and progressed to the present elaborate living world of plants mostly the commercially available K10 and KSF or native varieties with
and animals (Saladino et al., 2004; Stern and Jedrzejas, 2008; Ciciriello Brønsted or Lewis acid sites and enhancing their catalytic performance
et al., 2009). by pillaring techniques to manipulate the pore size, surface area and
Clays are nanoparticles with layered structures. The layers possess stability or replace interlayer cations to alter acid-base properties (Singh
net negative charge that is neutralized by cations such as Na+, K+, Ca2+, et al., 2007; Moronta et al., 2008). Clays have been intercalated with a
etc., which occupy the interlamellar space. The amazing amenability of variety of inorganic and organic ions, metal complexes, and organic
clays for modification lies in the fact that these interlamellar cations can compounds. These have brought about radical changes in the
be very easily replaced by other cations or other molecules. Molecules performance of clays in terms of increasing the rates of reactions, yields,
can be covalently anchored to layer atoms. All this can be done by very product selectivity, and stereoselectivity including enantioselectivity.
simple procedures. This provides tremendous scope for altering the Clays have been modified to act as acid-base combination catalysts
properties of clays like acidity, pore size, surface area, polarity and other which have been employed to carry out acid and base-catalyzed
characteristics that govern their performance as catalysts. Because of reactions in a sequence in one pot (Motokura et al., 2005, 2009). The
these wide ranging possibilities, in addition to their environmental possibilities seem to be limited only to the power of our imagination to
modify clays for any reaction.
The review describes seven types of organic reactions in as many
⁎ Permanent address: #13, Basappa Layout, Gavipuram Extension, Bangalore-560019,
India. Tel.: +91 80 26670899.
sections—Addition reactions, Condensation reactions, Diels–Alder and
E-mail address: gnagendrappa@gmail.com. related reactions, Esterification reactions, Friedel–Crafts and related
1
Retired from the organization. reactions, Isomerization reactions, and Oxidation reactions. It should be

0169-1317/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.clay.2010.09.016
G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 107

Scheme 1. Addition of allylsilanes to aromatic and aliphatic alkenes.

noted that the literature covered is essentially from articles in interlamellar space of the catalyst and is helped by its Lewis acid
mainstream journals published between 2000 and 2010, with a few character. This modified Hosomi–Sakurai reaction is environment
exceptions; the patent literature is completely omitted. As a result the friendly and delivers protected homoallylic ethers due to six membered
review is not exhaustive; some aspects and several types of reactions are pericyclic transition state of ene reaction (Scheme 3).
left out for various reasons. The reaction with aliphatic aldehydes and ketones was successful
in some cases, which are given below, with yields of the homoallylic
2. Addition reactions ether products mentioned in the parentheses.

In this section examples of addition reactions leading to carbon– O O


CHO
carbon and carbon–heteroatom bonds are considered. In the past ten CHO
years several groups of workers have reported a variety of addition CHO CHO
reactions efficiently facilitated by montmorillonite clays. They include (70%) (90%) (47%) (61%) (16%) (25%)

addition of allylsilanes to C=C and C=O bonds, carbene addition,


epoxidation, Michael addition, etc. Some of them are considered here.
Motokura et al. (2010) have found excellent catalytic performance
by proton-exchanged montmorillonite in the addition of allylsilanes Allylation of ketones and aldehydes has been carried out using
to aromatic and aliphatic alkenes (Scheme 1). potassium salts of allyl- and crotyltrifluoroborates using borontrifluoride
The mechanism has been studied in detail, a summary of which is etherate or montmorillonite K10 catalyst (Nowrouzi et al., 2009)
presented in Scheme 2. (Scheme 4). The authors find that K10 clay catalyzed reactions are robust,
Activated montmorillonite K10 clay was found to catalyze the straightforward and easy to work up, and scalable, which means the K10
reaction of allyl trimethylsilane with aromatic aldehydes to give catalyzed reaction is far superior to the Lewis acid-catalyzed one.
homoallylic silyl ethers (Dintzner et al., 2009). The authors suggest Other supports like alumina, silica gel and charcoal proved to be inferior.
that the reaction proceeds through a cyclic transition state in the The yields are generally excellent. In each case one stereoisomer is

Scheme 2. Mechanism of addition of allylsilanes to aromatic and aliphatic alkenes.

Scheme 3. Reaction of allyl trimethylsilane with aromatic aldehydes.


108 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 4. Allylation of ketones and aldehydes using potassium salts of allyl- and crotyltrifluoroborates.

Scheme 5. Addition of aniline derivatives to cinnamaldehyde.

predominantly more than the other. The K10 catalyzed reactions are dehydration and oxidation in the final step to deliver quinolines in good
better stereoregulated with the diastereomeric ratio being greater than in to excellent yields (De Paolis et al., 2009) (Scheme 5). The reaction is
the case of BF3.OEt2 catalyzed reactions. carried out under solvent-free condition and with the assistance of
Montmorillonite K10 clay catalyzes the addition of aniline deriva- microwave radiation.
tives to cinnamaldehyde in a Michael fashion as the first step of a A three-component reaction of enaminones, β-ketoesters/1,3-
domino process involving cyclization in the second step followed by diketones and ammonium acetate takes place under the catalytic

Scheme 6. Reaction of enaminones, β-ketoesters and ammonium acetate to form pyridines.


G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 109

Scheme 10. Addition of carbenes to C=N double bonds to produce aziridines.

Scheme 7. Synthesis of ethers by the addition of alcohols to olefins.

chiral bis(oxazaline)-copper (Box–Cu(II)) complexes supported on


influence of montmorillonite K10 clay in refluxing isopropyl alcohol to laponite clay-like solid and Nafion-like solid (Scheme 9). The
produce tri-substituted pyridines in good yields (Reddy et al., 2005c) suitability of the supported catalyst system is influenced by a variety
(Scheme 6). The reaction starts with the initial Michael addition of of factors. Better stereochemical control was possible on the laponite-
NH3 (from NH4OAc) to enaminone, followed by two condensation supported catalyst, compared to other supported catalysts studied.
reactions, and elimination of NMe2 to the final product. However, supported catalysts were inferior to the free chiral catalyst
Synthesis of ethers by the addition of alcohols to olefins is an in terms of yields and stereochemical control, though they could be
important reaction catalyzed by Brønsted acids. The problem, however, reused.
in such reactions is the possibility of extensive isomerization of the The enantiomeric excess depended on the alkyl group of the
double bond through the intermediate carbocation. Wang and Guin diazoester, and it was 83% in the case of menthyl ester, while in the
(2002) have found that sulphuric acid-treated montmorillonite clay is a case of n-butyl ester it was 74%. The phenyl substituted (R = Ph) Box
highly active catalyst for bringing about addition of methanol to 2,3- was a better catalyst than the t-butyl substituted one.
dimethyl-1-butene, and that it is far more selective to addition than Borkin et al. (2010) have prepared cis-aziridines in high diaster-
sulphated zirconia, Nafion, Amberlyst-15 or bentonite clay, which cause eoselectivity (N99%) and excellent yields (82–91%) by reacting Schiff
isomerization (Scheme 7). The activity of the catalyst depended on the bases with ethyl diazoacetate in the presence of montmorillonite K10
sulphuric acid content in the clay. as catalyst at room temperature for 2 h (Scheme 10). The K10 was the
Dintzner et al. (2006) have suggested an undergraduate green best catalyst among the several other acid catalysts, such as
experiment on the synthesis of a natural insecticide, methylenediox- H4W12SiO40, Nafion-H, Amberlist-15, and Nafion-H on silicon, in
yprecocene (MDP), with anti-juvenile hormone activity. It is a reaction achieving the highest diastereoselectivity. However, reactions con-
of sesamol with 3-methyl-2-butenol catalyzed by basic montmorillonite ducted using Nafion-H gave better yields of products (mixtures of cis-
K10–K+ (prepared by washing K10 with K2CO3 solution), and assisted and trans-aziridines). The efficiency of the catalyst remained the same
by microwave irradiation under solvent-free condition. The mechanism even after it was reused three times.
involves initial addition of the deprotonated sesamol to enal, the Box–Cu complexes (and Cu-complexes of three other ligands)
dehydration of the intermediate and finally an intramolecular hetero- immobilized in laponite clay are able to efficiently catalyze the
Diels–Alder reaction (Scheme 8). insertion of carbene formed from methyl phenyldiazoacetate into C–H
Addition of carbenes to carbon–carbon double bonds to produce bond of THF at the 2-position with high enantioselectivity (up to 88%
cyclopropanes is a well known reaction. Fraile et al. (2004) have ee). The immobilization not only allows recovery and reuse of the
studied the scope and limitations of the stereochemical course of the chiral catalyst, but also provides an improvement in selectivity over
addition of alkoxycarbonyl carbene generated from the corresponding the results obtained in solution, probably due to a confinement effect
diazoacetic esters. The reactions were carried out in the presence of of the bidimentional support (Fraile et al., 2007) (Scheme 11).

Scheme 8. Synthesis of methylenedioxyprecocene (MDP), a natural insecticide.

Scheme 9. Addition of carbenes to C=C double bonds to produce cyclopropanes.


110 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 11. Insertion of carbene into C–H bond.

An Iranian natural bentonite modified by N-alkylated 1,4- proposed mechanism is a bit elaborate; the interested reader may
diazabicyclo[2,2,2]octane quaternary salt was found to act as a good refer to the original paper).
catalyst in triphase reactions consisting of the solid catalyst and the Dithiocarbamic acid A adds to arylideneoxazalones B to form the
aqueous and organic phases (Ghiaci et al., 2005). When a mixture of Michael adducts C which undergo cyclization to 1,3-thiazinan
cyclooctene, the catalyst, chloroform and an aqueous solution of derivatives D. The reactions are carried out by microwave irradiation
sodium hydroxide was refluxed, dichlorocarbene was generated, of the reactants adsorbed on montmorillonite K10, basic and neutral
which added to the cyclooctene present in the reaction mixture, alumina, and silica gel. The best yields (76–91%) were obtained in the
forming dichlorobicyclo[6.1.0]nonane as the product in almost shortest reaction times from the reactions performed on K10 clay.
quantitative yield (Scheme 12). Reactions carried out on other solid catalysts were inferior. Micro-
The triphase systems (water-petroleum ether-catalyst) with water wave irradiation without using the adsorbent K10 clay was found to
soluble nucleophiles work well for nucleophilic substitution reactions be ineffective, demonstrating the significant role of clay for the
also (Scheme 13). success of the reactions (Siddiqui et al., 2010) (Scheme 15).
Triazenes have been synthesized (Scheme 14) by adding p-
aminobenzene-1-sulfonyl azide or amide to a cold mixture of sodium 3. Condensation reactions
nitrite and acid-treated clay (K10, bentonite, or kaolin), followed by a
cyclic secondary amine, which adds to the diazo intermediate formed Carbon–carbon bond forming reactions are of primary importance in
in the previous step (Dabbagh et al., 2007). The yields are moderate in organic synthesis. Among the numerous procedures developed for this
all the cases and are similar to the yields obtained using HZSM-5 and purpose aldolization/aldol condensation occupies an important posi-
sulfated zirconia. The mechanism does not involve formation of the tion. The aldol reaction is catalyzed by acids as well as bases. A base is the
conventional diazonium salt intermediate. Instead it consists of initial preferred catalyst to obtain aldol. Because of the importance of these
interaction of nitrite with protonated silicate of clay followed by a reactions attention is being paid to develop environment friendly
series of nucleophilic addition and elimination processes. (The procedures using heterogeneous clay catalysts.
Hydrotalcites (HT) as solid base catalysts have been successfully
used in bringing about the aldol reactions. For example, Roelofs et al.
(2000) have reported the self condensation of acetone (1) to give
aldol 2 and cross condensation of acetone with citral (3) on modified
hydrotalcite catalysts at 0 °C (Scheme 16). The catalyst shows high
activity and a small amount (5%) of it is enough to effect the aldol
formation. The cross aldol (4) is formed with very high selectivity.
These reactions are 100% atom economic and are examples for good
“green” procedures.
Scheme 12. Addition of dichlorocarbene in triphase system. Acetaldehyde (5) condenses with heptanal (6) in the presence of
hydrotalcite-type catalysts to give cross condensation product none-
nal (7) in ethanol at 100 °C (Tichit et al., 2003) (Scheme 17). If the
Brønsted base strength is increased due to a higher proportion of MgO
or hydrated sites, the reaction takes a different course through the
intermediate α-anion of heptanal (10) to give 8 and 9.
In an interesting study an acid-layered clay was combined with a
basic layered clay to bring about sequential acid and base-catalyzed
reactions in one pot. Motokura et al. (2005) mixed Ti4+ intercalated
Scheme 13. Nucleophilic substitution in triphase system. montmorillonite with surface tunable basic hydrotalcites. The acid-

Scheme 14. Triazenes from p-aminobenzene-1-sulfonyl azide.


G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 111

Scheme 15. 1,3-Thiazinans from dithiocarbamic acid and arylideneoxazalones.

the desired enantiomers by employing suitable chiral catalysts. Such


catalysts immobilized in clays through intercalation have shown
excellent results. Srivastava et al. (2009) report using montmorillonite
clay in which the sodium ions have been exchanged with prolium
(protonated proline) by treating the clay with prolium chloride in
methanol. The immobilized prolium acts as a chiral molecular catalyst to
induce chirality in the newly formed asymmetric carbon of the aldol
product 15 (Scheme 19). They found that intercalated hydroxyproline
behaves in a similar manner. They further report that pillaring the
prolium embedded (Pro-Mont) clay with trimethylbutylammonium
bistriflimide leads to significantly higher yields and enantiomeric
excesses, and that the solvents too have influence on the yields and
Scheme 16. Aldol reaction of acetone with itself and citral.
enantiomeric excesses.
Montmorillonite K10 has proved to be an effective catalyst in bringing
base catalyst combination acts sequentially in multistep reactions about Mukaiyama crossed aldol condensation of silyl enol ethers with
that require both acid and base catalysis. The authors report the various aldehydes (Loh and Li, 1999). If the clay is intercalated with a
following multistep reactions, (i) deacetalization–aldol condensation, chiral catalyst an excess of one enantiomer of the aldol is obtained. For
Michael addition–acetalization, (ii) esterification–aldol condensation– example, Fabra et al. (2008) describe the use of the chiral diphenylbis
epoxidation, and (iii) esterification–aldol condensation–Michael addi- (oxazoline)-Cu2+ complex 16 immobilized on laponite clay to bring
tion. Each one of these combined three-reaction sequences has been about the Mukaiyama aldol reaction in a stereo-controlled manner
performed in one pot resulting in excellent yields. The aldol conden- (Scheme 20). They found that 2-(trimethylsilyloxy)furan (17) adds to
sation reactions are initiated by deacetalization effected by Ti4+-mont, the α-ketoesters 18 to give the aldols 19a,b with the anti-isomers as
followed by aldol condensation catalyzed by basic hydrotalcite. The major products (up to 90% ee; dr, 86:14).
reaction sequences are depicted in Scheme 18. However, after surveying several such chiral Cu2+-complex-clay
The next four reactions start with Michael addition catalyzed by catalyzed reactions, Fraile et al. (2009) offer a note of caution that the
HT, followed by acetal formation catalyzed by Ti4+-montmorillonite. application of these catalysts is limited despite excellent results in
In the last reaction sequence, the first reaction is acid-catalyzed some cases.
esterification, the second is acid-catalyzed deacetalization, followed Knoevenagel condensation of malononitrile with carbonyl com-
immediately by base-catalyzed aldol condensation. The olefinic pounds has been found to be activated by ultrasound and catalyzed by
product 11 is then epoxidized to 12 by hydrogen peroxide under alkaline-doped saponites. The dopant ions were Li+ and Cs+. The Cs+
basic catalysis. Or the olefin 11 can be hydrogenated to 13, which doped clay was far superior to the Li+ doped one in providing higher
undergoes Michael addition to acrylonitrile in the same pot to give 14. yields of products (Martin-Aranda et al., 2005) (Scheme 21). Thermal
None of the intermediate products in any reaction depicted in reactions with the same basic saponites as catalysts but without
Scheme 18 was isolated; each step was performed one after the other ultrasound were inferior.
in one pot. Montmorillonite K10 catalyzes the condensation of primary mono-
Aldolization creates a chiral centre in aldol product. This provides an (20) and diamines (21) with β-hydroxy-β-bis(trifluoromethyl) ketones
opportunity to design enantio-selective synthetic procedures to deliver (22), obtained by aldol reaction of methyl ketones with hexafluoro

Scheme 17. Aldol condensation of heptanal.


112 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 18. Sequential multiple reactions using combined acid and basic layered clay.

acetone hexahydrate, to give mono- (23) and di-imines (24). The


reaction is conducted in chloroform at 70 °C for 4–5 days. A number of
di-imines have been synthesized which find use in the preparation of
fluorinated chelate complexes for vapour deposition in microelectronics
(Marquet et al., 2008) (Scheme 22).
Motokura et al. (2009) have succeeded in preparing catalysts with
coexisting acidic and basic sites in the montmorillonite clay interlayer
by treating the clay first with hydrochloric acid and then with
triethoxysilylalkyl amines. Silicon of –Si(OEt)3 covalently binds with –
SiO– in clay by displacing an EtO group, and thus the alkyl amine is
immobilized. The acid sites and basic sites act independently to bring
about both the acid-catalyzed deacetalyzation and the base-catalyzed
Knoevenagal condensation reactions in tandem in one pot (Scheme 23).
A three-step reaction of addition of phenylhydrazines to dihydro-
4-H-pyranone derivatives, followed by ring opening and intramolec-
Scheme 19. Enantio-selective synthesis of aldols. ular condensation to deliver pyrazines in good yields is promoted by
G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 113

Scheme 20. Mukaiyama reaction of silyl enol ethers with aldehydes.

catalyst to give tripyrrane (27) in N90% yield (Okujima et al., 2004)


(Scheme 26).
The Biginelli reaction, which involves a one-pot condensation of β-
ketoesters (30)/β-dicarbonyl compounds (31) with aldehydes (32)
and ureas (33) to give dihydropyrimidones (34), is conventionally
conducted in ethanol with strong protic acid or Lewis acid as catalyst.
Salmon et al. (2001) have shown that the reaction takes place on
commercially available bentonite clay TAFF as catalyst under infrared
irradiation and in the absence of solvent (Scheme 27).
Earlier Li and Bao (2003) had demonstrated that the three-
Scheme 21. Knoevenagel condensation of malononitrile with carbonyl compounds.
component Biginelli reaction can be performed efficiently using
samarium trichloride supported on montmorillonite clay (clay:
SmCl3 = 10:1) as catalyst with microwave irradiation under solvent-
montmorillonite KSF clay in ethanol solvent. The stereochemistry of free condition (Scheme 28). The catalyst was found to be reusable.
the substituents is retained in the products (Yadav et al., 2004b) Hydrotalcite-like materials have been used by Zhu et al. (2009a) as
(Scheme 24). catalysts for the condensation of aromatic as well as aliphatic primary
A microwave-assisted montmorillonite KSF catalyzed, solventless amines with aromatic and aliphatic aldehydes, and cyclohexanone to
aldol condensation of aromatic aldehydes with aryl methyl ketones obtain Schiff bases (35). The reaction is carried out under solvent-free
has been developed by Chtourou et al. (2010) (Scheme 25). The condition by stirring together the reactants and the catalyst at room
reaction takes just about 1 h for completion. It is highly selective (87– temperature (Scheme 29). The catalyst is recyclable. The yields in
98% trans chalcones) and the yields are good to excellent (80–95%). most cases are excellent. Aldehydes reacted faster than ketones.
The authors maintain that their method is far more efficient than Malonic acid undergoes Knoevenagel condensation with salicylic
other known methods which use a number of acid-treated supports, aldehydes in the presence of montmorillonite KSF clay on refluxing in
alkalies and other catalysts in various solvent media. water for 24 h. The condensation products initially formed cyclize to
In a series of reactions leading to the synthesis of conjugation- give coumarin-3-carboxylic acids, which are isolated by filtering off
expanded carba- and azuliporphyrins (28 and 29) an intermediate the catalyst, evaporation of the solvent and finally recrystallization.
step of condensation of bicyclo[2.2.2]octadiene-fused pyrrole (25) The selectivity was 95% and the yields were N90% in most cases.
with bicyclo[2.2.2]octadiene-fused 2-acetoxymethyl-4-t-butoxycar- However, if diethyl malonate was used in place of malonic acid the
bonylpyrrole (26) was carried out using montmorillonite K10 as yields of the coumarin-3-carboxylate esters were only 30–48% (Bigi

Scheme 22. Condensation of primary mono- and diamines with ketones.


114 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 23. Reaction using coexisting acidic and basic sites in tandem.

commonly prepared by the condensation of 1,2-diketones with o-


phenylene diamines. A variety of catalysts are used, particularly acidic
or oxidizing reagents. In most cases an organic solvent is also used. Huang
et al. (2008) have overcome the eco-disadvantages associated with using
such catalysts and organic solvents by making use of montmorillonite K10
Scheme 24. Pyrazines from dihydropyranones and phenylhydrazines. as catalyst in water at room temperature. The reaction takes about 3 h and
gives excellent yields (90–100%; in one case, it was 70%, because the
diamine has an electron withdrawing NO2 substituent) (Scheme 31). The
catalyst can be reused without much loss in its activity.
et al., 1999) (Scheme 30). Montmorillonite K10 clay was much less The authors propose a mechanism suggesting an initial proton-
effective than the KSF as catalyst. Since the Knoevenagel reaction ation of the diketone by the Brønsted acidic K10 clay. The nucleophilic
requires basic conditions and the subsequent cyclization needs diamine then adds to the protonated diketone intermediate followed
protons for its initiation, the authors presume that the clay catalyst by dehydration to give the quinoxaline products.
must be ditopic and hence should contain both acid and basic sites. Isobezofuran-1(3H)-one derivatives have been prepared by
The other important feature of the catalyst was that it retained its microwave irradiation of a mixture of phthalaldehydic acid and
activity even after five reaction cycles. ketones well dispersed in montmorillonite K10 catalyst under
Quinoxalines, which find applications in the area of medicines, dyes, solventless conditions (Landge et al., 2008) (Scheme 32). The best
electron luminescent materials, organic semiconductors and others, are results were obtained when the irradiation is carried out for 10–

Scheme 25. Microwave-assisted solventless aldol condensation.

Scheme 26. Synthesis of conjugation-expanded carba- and azuliporphyrins.


G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 115

reported by Kulkarni and Török (2010). The three-component system


consists of an aromatic aldehyde, aniline or its derivative and phenyl
acetylene or its derivative. Microwave irradiation of the reaction
mixture combined with K10 at 100 °C for about 10 min was found to
be the optimum condition for obtaining the best yields (Scheme 33).
The reaction is a multistep domino process that starts with the
formation of a Schiff base in the first step. In the next step, phenyl
acetylene adds to the base, then the adduct cyclizes followed by
deprotonation to give the observed quinoline product.
Microwave irradiation of a mixture of tryptamines and aromatic
aldehydes dispersed in combined 10% Pd/C/K10 catalyst delivered
Scheme 27. Solventless method for Biginelli reaction under IR irradiation.
β-carbolines in good to excellent yields. The optimum reaction
temperature was 130 °C. The selectivity for β-carbolines was
100% when 10% Pd/C with K10 was used (Kulkarni et al., 2009a)
(Scheme 34). If Pd/C was less than 10%, the intermediate tetrahydro-
β-carbolines were isolated along with β-carbolines, and if only
K10 was used as catalyst, only the former compounds were formed
as products. These results indicate that the cyclization process
is catalyzed by K10, while the dehydrogenation is brought about by
Pd/C. Use of other acid catalysts, such as Al2O3 and acetic acid, gave
very low yields. Tryptaphan as substrate in place of triptamine or Pt/C
instead of Pd/C gave the same β-carboline products.

4. Diels–Alder and related reactions

Pericyclic reactions are one of the most important ways of constructing


more complex organic molecules from the simpler ones. Among these the
most celebrated are those that involve the 6-electron cyclic transition
Scheme 28. Microwave-assisted Biginelli reaction without solvent. state, which include electrocyclic and cycloaddition reactions. Since they
are thermally allowed processes, their energy demand can be easily met.
Many of them, in fact, occur at room temperature and even below that in a
number of cases. The Diels–Alder reaction is a cycloaddition process
between a 4-electron and a 2-electron component which could be starting
compounds or intermediates formed in a multistep reaction sequence.
Many such reactions are assisted or initiated by acid catalysts. This has
provided an opportunity to exploit clays, particularly montmorillonites, as
Brønsted or Lewis acid catalysts for Diels–Alder reactions. In this section
some recent examples of such catalyzed reactions are presented.
A major problem in developing a suitable chiral organocatalyst
entrapped in the montmorillonite interlayer is its instability due to
leaching. This problem has been overcome by Mitsudome et al. (2008)
by entrapping the cationic chiral organocatalyst (5S)-2,2,3-trimethyl-
5-phenylethyl-4-imidazolinone hydrochloride by replacing the inter-
layer cations in montmorillonite clay. The catalyst is very effective in
carrying out asymmetric Diels–Alder reactions (Scheme 35). Several
Scheme 29. Condensation of primary amines with aldehydes and ketones.
other inorganic solids, such as silica, titania, zeolite, MCM-41,
hydroxyapatite and γ-ZrP failed to function as support. The
30 min at 170 °C. The catalyst is reusable and its performance in the montmorillonite supported catalyst is stable and could be reused a
fifth trial was as good as in the first one. few times.
A microwave-assisted, montmorillonite K10 catalyzed three- Lopez et al. (2007) have made an interesting observation of
component reaction for the preparation of quinolines has been montmorillonite K10 exerting considerable influence in terms of

Scheme 30. Coumarin-3-carboxylic acids from malonic acid and salicylic aldehydes.
116 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 31. Quinoxalines by condensing 1,2-diketones with o-phenylene diamines.

stereochemical outcome and product yields on Diels–Alder reaction produced oligomeric and polymeric products rather than [4+2] addition
carried out in ionic liquid, 1-hexyl-3-methylimidazolium (HMI) products. p-Chloro and p-nitrobenzaldehyde gave very poor yields of the
tetrafluoroborate. The results with other supports like silica or corresponding dihydropyrans as compared to o-chloro- and o-
alumina or without support in the same medium were not good. nitrobenzaldehyde.
Microwave irradiation was also not as good as K10 clay. With K10 clay Chiba et al. (1999) have generated in situ the highly reactive o-
as catalyst the yields were as high as 99% and endo:exo ratio was 93:7. quinomethanes from o-hydroxybenzyl alcohols at room temperature
The reaction took just 30 min at room temperature (Scheme 36). The using wet monmorillonite K10 and lithium perchlorate in nitrometh-
same medium after work up was reused four times without much loss ane. The presence of air is necessary. The quinomethanes formed add
in yields. instantly to olefins to give benzodihydropyrans (Scheme 38).
Several clay catalyzed hetero-Diels Alder reactions have been reported Methylene cyclopropanes function as dienophiles in their aza-
in the last few years. Dintzner et al. (2007) found that the carbonyl group Diels–Alder addition reaction with ethyl (arylimino)acetates under
of benzaldehyde and its derivatives adds to 2,3-dimethyl-1,3-butadiene the catalytic influence of montmorillonite K10 in dichloroethane at
under the influence of montmorillonite K10 clay in carbon tetrachloride at room temperature to produce tetrahydroquinolines. The K10 clay
25 °C to form dihydropyrans. The clay was preheated to 250 °C which performs as good as triflic acid in catalyzing these reactions (Zhu et al.,
enabled the collapse of the interior structure of clay by extrusion of water 2009b) (Scheme 39).
leading to a decrease in Brønsted acidity but an increase in Lewis acidity Two of the authors of this group, Shao and Shi (2003) had reported
that is responsible for the catalytic activity of the clay, suggest the authors. similar aza-Diels–Alder reactions of methylene cyclopropanes with
Based on the fact that benzaldehydes with ortho-substituents having lone Schiff bases. In this case they had employed monmorillonite KSF clay
pair electrons give far better yields of dihydropyran products, the authors or scandium triflate as catalyst (Scheme 40). The Schiff bases were
propose a clay metal ion coordinated transition state that favours the produced in situ by the reaction of aryl aldehydes and aryl amines
reaction (Scheme 37). The less substituted dienes (e.g., isoprene) employed directly in the 3-phase reaction.

Scheme 32. Isobezofuranones from phthaldehydic acid and ketones without solvent.
G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 117

Scheme 33. Three-component reaction for the preparation of quinolines.

Aryl amines add to endocyclic enecarbonates resulting in the latter moderate to good yields varying from 42–81% depending on the
undergoing ring opening to give aralkyl imines under the influence of substituents (Scheme 43). The hetero-Diels–Alder reaction proceeds
montmorillonite KSF clay in THF at room temperature. The imines so by a multistep process in which the acetylene adds to the Schiff base
formed are highly reactive and undergo rapid aza-Diels–Alder formed by the condensation of the amine and the aldehyde. Slightly
reaction with a second molecule of the endocyclic enecarbonate better yields were obtained in oxygen atmosphere instead of open air.
present to finally give hexahydro-1H-pyrrolo(3,2-c) quinoline deri- The research group also found that, among some twenty different
vatives (Yadav et al., 2004a) (Scheme 41). solid supports, the HClO4 treated montmorillonite worked the best in
The same group of workers had previously reported (Yadav et al., terms of yields, reaction time, work up procedure, etc.
2002) similar reaction of aryl amines with dihydrofuran and Cyclopentadiene and furan undergo Diels–Alder addition at room
dihydropyran to produce furano- and pyrano-quinolines (Scheme 42). temperature with trans-2-methylene-1,3-dithiolane-1,3-dioxide in
A similar mechanistic pathway is proposed for this reaction also. the presence of Fe3+-doped montmorillonite K10 combined with
Guchhait et al. (2009) have reported a multicomponent Povarov 2,6-di-tert-butyl-4-methylphenol (butylated hydroxytoluene, BHT) to
reaction involving aromatic amine, aromatic aldehyde and terminal give the product in a overall yield of 64% (Gültekin, 2004)
acetylene. The three compounds were allowed to react in the (Scheme 44).
presence of perchloric acid-treated montmorillonite clay at 70 °C in Aldimines generated in situ from aliphatic aldehydes and p-
open air. Substituted quinolines were obtained as products in anisidine add to Danishefsky diene in the presence of montmorillonite

Scheme 34. Condensation of tryptamines and aldehydes to give β-carbolines.


118 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 35. Entrapped cationic chiral organocatalyst in asymmetric Diels–Alder reactions.

K10 in aqueous or aqueous acetonitrile medium to produce 2-alkyl- particularly as oils, fats and waxes. They are a major component of all
2,3-dihydro-4-pyridones in excellent yields (Akiyama et al., 2002) living organisms. They have numerous applications in industries and as
(Scheme 45). foods and fuels, for which esters obtained from both natural as well as
Montmorillonite K10, filtrol-24, bentonite and pyrophillite clays synthetic sources are employed. There are several ester forming
were used as catalysts to bring about Diels–Alder addition of 1,4- reactions available, such as (i) reaction of alcohols or phenols with
naphthoquinone and N-phenylmaleimide to 4,6-bis(4-methoxyphenyl)- carboxylic acids, carboxylic acid anhydrides or acyl halides, (ii) addition
and 4-(4-methoxyphenyl)-6-methylpyran-2(H)-ones under a dry state of carboxylic acids to olefins, (iii) addition of alcohols to ketenes,
adsorbed condition. Filtrol-24 performed the best under this condition. (iv) substitution of alkyl halides/tosylates with carboxylates, (v) Baeyer–
Further, when montmorillonite K10 and bentonite were modified Villiger oxidation of ketones, etc. The most common and expedient
by impregnating with AlCl3, ZnCl2 and FeCl3 using their aqueous or method is the acid-catalyzed reaction of a carboxylic acid with an
nonaqueous solution (PhNO2 for AlCl3, MeCN for ZnCl2 and FeCl3), alcohol, called esterification, which is usually carried out in homoge-
followed by washing with water and drying, the resulting modified K10 neous media. Though several Brønsted acids are used as catalysts for this,
and bentonite catalysts performed as well as filtrol-24. Among the doped the most convenient and commonly used one is concentrated sulfuric
catalysts FeCl3-containing ones worked the best (Kamath et al., 2000) acid. The disadvantages of using this acid, like corrosion problems,
(Scheme 46). handling difficulties, waste disposal hassles, environmental hazards etc.
are well known. This opens up a good opportunity for acid clays to
5. Esterification reactions replace the hazardous mineral acid in ester forming processes, and a
substantial amount of research activity is going on in this area.
Esters, which are normally understood to be alkyl or aryl carbox- Clays available commercially or occurring naturally, including
ylates, are an important class of naturally occurring compounds, those found in the geographical region of the researchers, have been

Scheme 36. Diels–Alder reaction in ionic liquid. K10 is the best catalyst. Scheme 37. Hetero-Diels–Alder reaction of benzaldehyde with dimethylbutadiene.
G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 119

Scheme 38. o-Quinomethanes from o-hydroxybenzyl alcohols and their D–A addition.

found to work well with or without suitable modifications. The most 2 h reaction time). They also discovered that the supporting K10 clay is
successful esterification catalysts have been the montmorillonites, essential for the reaction, since InCl3 alone was not effective as catalyst.
normally those that possess high Brønsted acidity. Some selected The yield of benzyl benzoate was only 11% after 1 h of reaction at 50 °C
examples from recent literature are described here. between benzyl alcohol and benzoyl chloride with InCl3 as catalyst,
Kaolin, montmorillonite K10 and KSF supported with transition while with equivalent amount of InCl3/K10 as catalyst the ester yield
metal chlorides, InCl3, GaCl3, FeCl3, and ZnCl2 were employed to was 96%. The authors further found that the InCl3/K10 catalyst was
esterify tert-butanol with acetic anhydride to tert-butyl acetate with almost as effective in its fifth time reuse as in the beginning.
more than 98% selectivity. The K10 clay was found to be the best Srinivas and Das (2003) have demonstrated that ferric chloride
support and K10 supported InCl3was the best catalyst followed by supported on montmorillonite K10 clay (Fe3+/K10) is an efficient
GaCl3/K10, FeCl3/K10 and ZnCl2/K10, in that order. A noteworthy esterification catalyst. The catalyst is highly selective in esterifying
feature was the low activity of the catalysts for the dehydration of tert- both saturated and unsaturated aliphatic carboxylic acids, while the
butanol below 50 °C (Choudhary et al., 2001a) (Scheme 47). aromatic acids are unreactive, as demonstrated by using mixtures of
In a later study employing the InCl3/K10 catalyst Choudhary et al. aromatic and aliphatic acids. The yields of esters are high and the
(2004) reported the preparation of thirteen esters by the reaction of catalyst is reusable. The catalyst is shown to be useful also in the
benzyl alcohol, phenol, 4-nitrophenol, 1-naphthol and 2-naphthol with preparation of amides of aliphatic carboxylic acids with aliphatic as
benzoyl chloride, acetyl chloride and n-butyryl chloride. The reaction well as aromatic amines, but ineffective in the preparation of the
gave high yields of esters (up to 98%) under mild conditions (50 °C, 0.2– amides of aromatic acids with aromatic amines (Scheme 48).

Scheme 39. Aza-Diels–Alder reaction of arylimino esters with methylene cyclopropanes.

Scheme 40. Aza-Diels–Alder reaction of Schiff bases with methylene cyclopropanes.


120 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 41. Hexahydropyrroloquinolines from aryl amines and enecarbonates.

Scheme 42. Furano-/pyrano-quinolines from aryl amines and dihydrofuran/dihydropyran.

Some Brazilian natural clays (smectite, atapulgite and vermicu- Vijayakumar et al. (2004, 2005a,b) have shown that acid activated
lite), without pretreatment or activation, have been demonstrated by Indian bentonite is a good catalyst, in some cases better than zeolites,
Silva et al. (2004) to act as good catalysts for transesterification of for the preparation of aryl and alkyl esters of fourteen different
ethyl acetoacetate and ethyl bezoylacetate by six carbohydrate- aromatic and aliphatic carboxylic acids. Particularly, the yields of aryl
acetonides. Refluxing a mixture of the reactants and the catalyst in esters of long chain fatty acids are very impressive, which were more
toluene for ~48 h produced the acetonide esters in good yields by than 90%, are very impressive. Results of a part of their work are
replacing the ethyl group of the β-keto esters. The reaction mixture on depicted in Scheme 50.
cooling to room temperature was stirred (~24 h) with bezylamine to Earlier Kantam et al. (2002) observed that Fe3+-impregnated
produce β-benzyl enamino esters (Scheme 49). montmorillonite clay catalyzed the esterification of various aliphatic

Scheme 43. Multicomponent Povarov reaction—synthesis of quinolines.


G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 121

Scheme 44. Diels–Alder reaction of 2-methylene-1,3-dithiolane-1,3-dioxide.


Scheme 47. Esterification of tert-butanol with acetic anhydride.

acids, including the long chain fatty acids, aromatic acids, and α,β-
unsaturated mono- and dicarboxylic acids with alcohols, under mild
reaction conditions. The yields were obtained in the range of good to
excellent.
In three reports Reddy et al. (2004, 2005a,b) describe their study
on the esterification of monocarboxylic and dicarboxylic acids with
phenols and alcohols including ethylene glycol. They found that Al3+
exchanged montmorillonite GK-129 was the best catalyst compared
with the same clay exchanged with divalent and other trivalent metal
cations. It is better than the similarly treated montmorillonite K10 or
Indian bentonite. Its performance is comparable to the results Scheme 48. Ferric chloride supported on K10 for esterification.
obtained using zeolite H-β or p-toluene sulfonic acid catalyst. They
discuss various experimental conditions such as temperature,
reaction time, solvents, catalyst preparation, etc., that give the best Methyl mandelate, used in flavouring and perfumery, has been
results, as well as the catalyst's performance in its reuse. Their results prepared by esterification of mandelic acid with methyl alcohool in
are consolidated in Scheme 51. Reddy et al. (2007) have carried out the presence of montmorillonite K10 supported dodecatungstopho-
esterification of succinic acid with iso-butyl alcohol to di-iso-butyl sphoric acid (DTP/K10) and its cesium salt (Cs-DTP/K10) (Yadav and
succinate in connection with the evaluation of surface activity of Mn+- Bhagat, 2005) (Scheme 52). The Cs-DTP/K10 catalyst was found to be
montmorillonite clay catalysts, where Mn+ = Al3+, Fe3+, Cr3+, Zn2+, better than K10 and other solid catalysts like S-ZrO2. The catalyst was
Ni2+, Cu2+ and H+. The reactions performed on Al3+-mont and H+- shown to be recyclable. Using the same catalyst (Cs-DTP/K10 clay)
mont gave the best yields (96% and 97%), Fe3+-mont and Cr3+-mont Yadav and George (2008) have carried out the esterification of
gave good yields (74% and 51%), while the other Mn+-mont catalysts benzoic acid with phenol to phenyl benzoate which then undergoes
gave poor yields (23–27%). The results were rationalized based on Fries rearrangement to give 2- and 4-hydroxybenzophenones. The
various properties of catalysts due to exchanged ions. selectivity of the products depended on a number of experimental
As noted in the section on condensation reactions, Motokura et al. parameters.
(2005) have esterified cyanoacetic acid with methanol to methyl The esterification strategy was used by Mittal (2007) to increase
cyanoacetate over the combined Ti4+-mont, HT catalyst (Section 3, the basal plane spacing in clays in order to achieve shear induced
Scheme 18). exfoliation. This was accomplished by modifying the clay platelets with

Scheme 45. Alkyldihydropyridones from Danishefsky diene and aldimines formed in situ.

Scheme 46. Diels–Alder addition of 1,4-naphthoquinone to N-phenylmaleimide.


122 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 49. Transesterification of esters by carbohydrate-acetonides.

quaternary ammonium ions carrying hydroxyalkyl groups, which were catalytic activity of different batches of commercially available K10
then reacted with long chain fatty acids (Scheme 53). The paper reports montmorillonite clays. They have found that acid treatment improves
various aspects of catalyst modification and resulting properties. the activity of the clay and suggest that esterification is suitable for
Wallis et al. (2007) have employed the esterification of maleic determining the degree of clay delamination. They suggest that loss of
anhydride with p-cresol, along with a transacetalization reaction, layer stacking and increase in available exchange sites for protonation
diacetylation of bezaldehyde with acetic anhydride and tetrahydro- are responsible for clay activity enhancement. They compare this with
pyranylation of ethanol (Scheme 54), to assess and improve the an earlier observation by Reddy et al. (2005a) who had observed

Scheme 50. Esterification using Indian bentonite.


G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 123

Scheme 51. Preparation of diesters using Al3+ exchanged montmorillonite GK-129.

Scheme 52. K10 supported dodecatungstophosphoric acid-Cs salt for esterification.

dramatic improvement in the catalytic activity of K10 montmorillon- p-Hydroxybenzoic acid esters, with the sobriquet parabens, find
ite clay with much reduced specific surface area on acid treatment. application in cosmetic, pharmaceutical and food industries. Hazarika
Thermally activated Nigerian Ukpor kaolinite clay and Udi clay et al. (2007) have prepared the methyl, ethyl and n-propyl p-hydroxy
were shown to be good catalysts for the preparation of n-propyl benzoates by refluxing the mixtures of the aromatic acid and
acetate (Igbokwe et al., 2008). Various experimental parameters were appropriate alcohol on montmorillonite K10 clay for 10–15 h, and
measured to obtain the best yield of the ester. obtained the esters in 81–90% yields (Scheme 56). The acid-treated
Esterification of long chain fatty acids, stearic, oleic and palmitic clay gave slightly better yields (~2% more).
acid, with short chain alcohols, methanol, ethanol, 1-propanol, 1- Acetic acid reacts efficiently with 2-methoxyethanol in the
butanol and 2-butanol has been carried out using a series of presence of clay catalyst treated with sulfuric acid and aluminium
montmorillonite based clay catalysts, KSF/0, KP10, K10 (Neji et al., salts and calcined at 313–633 K, to give methoxyethyl acetate. Both
2009) (Scheme 55). The product esters are meant to be used as Lewis acid and Brønsted acid sites are active in catalyzing the
biodiesel. A variety of reaction conditions were investigated. KSF/0, esterification process (Wang and Li, 2000) (Scheme 57).
which had the lowest pH value, was found to be the best catalyst. The The waxy stearyl stearate ester was synthesized by reacting stearic
yields in the case of primary alcohols were almost quantitative, but in acid with stearyl alcohol on montmorillonite clay under solvent-free
the case of 2-butanol the ester was obtained in only 40% yield. The condition. The temperature was strictly maintained at 170 °C
catalyst was recycled twice without significant loss in its activity. throughout the bulk of the reaction mixture in the pilot scale reactor
by using microwave irradiation. The pure ester is obtained in 95%
yield on filtering off the solid catalyst. The use of microwave radiation
reduces the reaction time by a factor of 20–30 times compared to the
time required by the reaction done in conventional reactor (Esveld
et al., 2000) (Scheme 58).

6. Friedel–Crafts and related reactions

The Friedel–Crafts reaction is an important carbon–carbon bond


forming process. It enjoys numerous applications in the synthesis of
Scheme 53. Strategy to increase the basal plane spacing in clays by esterification. bulk chemicals, fine chemicals, pharmaceutical and perfumery
124 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 54. Esterification for determining the degree of clay delamination.

chemicals, and several others. It is an electrophilic substitution reaction indeed catalysts, they are not really used in catalytic quantities, but in
of usually aromatic and heteroaromatic compounds by alkyl or acyl much larger amounts. This obviously poses serious problems of
groups catalyzed by a number of Lewis and Brønsted acids under handling, recovery and disposal of the waste products. In this respect,
homogeneous or heterogeneous conditions. Alkyl halides, alcohols, Friedel–Crafts reactions are an antithesis of Green chemistry principles.
sulfonates and olefins are used as alkylating agents, and acyl halides or This has provided challenging opportunity to explore alternative
acid anhydrides are generally used for acylation. The most commonly catalytic procedures. Thus reusable, environmentally benign solid acid
used catalysts are AlCl3, ZnCl2, SnCl4, BF3, FeCl3, SbCl5, H2SO4, H3PO4, etc., catalysts including clays are being explored for these reactions. Clays
mostly under homogeneous conditions. Though these Lewis acids are and modified clays have been used successfully as catalysts for bringing
about Friedel–Crafts alkylation and acylation.
Cyclopentyl and cyclohexyl derivatives of benzene, toluene, o-
and p-xylene, mesitylene and anisole have been prepared in 85–95%
yield by refluxing a solution of the respective aromatic substrate
with cyclopentanol or cyclohexanol in 1,2-dichloroethane on Fe3+-
montmorillonite with 10 mol% of TsOH or MsOH as cocatalyst
(Chaudary et al., 2002) (Scheme 59). The suggested mechanism
involves the formation of sulfonate ester of the alcohol as intermediate
which alkylates the arene selectively. In the absence of the cocatalyst
TsOH or MsOH, the reaction with cyclopentanol produces only a minor
amount of the expected alkylated product (~30%), while the major
product is dicyclopentyl ether (~60%). The reaction does not take place
Scheme 55. Esters for using as biodiesel. in the absence of Fe3+–montmorillonite.
An interesting case of Fe–Mg–hydrotalcite anionic clay being used
for the Friedel–Crafts alkylation has been described by Choudhary
et al. (2005a) (Scheme 60). They have benzylated anisole, mesitylene,
p-xylene, toluene and naphthalene with high degree of conversions,
using benzyl choride. They observed that calcining increases the
activity of the catalyst, and that the higher the calcining temperature
the more active the catalyst is. They attribute this to dehydration at
200 °C, formation of metal oxides on calcining at 500 °C and higher
temperatures up to 800 °C. They also found that the used catalyst is
Scheme 56. Preparation of p-hydroxybenzoic acid esters. more active than the one used for the initial reaction. This observation
is explained as due to a possible formation of Lewis acid sites resulting
from the reaction of HCl liberated in the benzylation process.
In–Mg hydrotalcite anionic clay is found to function in a similar
manner (Choudhary et al., 2005b). Ga–Mg–hydrotalcite anionic clay
was used earlier for benzylation and benzoylation of benzene
Scheme 57. Preparation of methoxyethyl acetate. (Choudhary et al., 2001b).

Scheme 58. Microwave-assisted esterification without solvent.


G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 125

Scheme 59. Mesilates and tosylates in F–C alkylation.

Scheme 60. Anionic clay for the Friedel–Crafts alkylation.

Benzene has been alkylated with propylene to cumene on Al3+- the presence of montmorillonite K10 clay-POCl3 under microwave
exchanged synthetic Zn-saponite with 99% conversion (Scheme 61), irradiation. The effect of this catalyst was comparable to the silica
compared to just 0.3% conversion when the same reaction was gel-POCl3 catalyzed reaction (Devi, 2006).
performed using commercial solid phosphoric acid (SPA), (Vogels A number of other clay-based catalytic systems have been
et al., 2005). developed by several groups of researchers for bezylation reactions.
To introduce an isopropyl group on xylenes, isopropyl alcohol is The reactions are conducted in liquid phase or solvent-free conditions
used by Yadav and Kamble (2009). They have obtained dimethylcu- with or without microwave irradiation. Studies were directed to find
menes by reacting xylenes with isopropyl alcohol in an autoclave in the right experimental conditions to obtain the best results by
the presence of cesium substituted dodecatungstophosphoric acid preparing the most active catalyst.
supported on K10 as catalyst (Scheme 62). The Cs-DTP/K10 was the Ga/AlClx-grafted montmorillonite-K10 was found to be an efficient
most active among the five catalysts studied, including K10, DTP/K10, catalyst for benzylation as well as benzoylation of benzene, substituted
sulfated zirconia and filtrol-24. Various aspects of catalytic activity benzenes and naphthalene, using benzyl chloride and benzoyl chloride
and kinetics of reactions are considered. Selectivity for mono- respectively. The catalyst is highly active, and as such benzoylation
isopropyl products is very high, with the byproducts, diisopropylated occurs even if a strong electron withdrawing group like NO2 is present
xylenes, diisopropyl ether and propylene being formed in small on benzene ring (Choudhary and Jha, 2008) (Scheme 65).
amounts. Clay catalyists obtained from Pakistani clay minerals were found
Cs-DTP/K10 has been used fruitfully in several other Friedel–Crafts to be useful in benzylation of toluene, naphthalene, anthracene,
type reactions. For example, 1,3-dibenzyloxybenzene has been quinoline, 8-hydroxyquinoline and pyridine (Ehsan et al., 2006).
acetylated to 3,5-dibenzyloxyacetophenone with acetic anhydride as
shown in Scheme 63 (Yadav and Badure, 2008).
Phenol has been benzoylated by benzoic acid to 4-hydroxybenzo-
phenone via the formation of phenyl benzoate as intermediate
which underwent Fries rearrangement, a variant of Friedel–Crafts
acylation (Yadav and George, 2008) (Scheme 64). Cs-DTP/K10 has
been found to be a good catalyst for benzoylation of p-xylene to
2,5-dimethylbezophenone (Yadav et al., 2003). Similar results were
obtained when resorcinol was treated with phenylacetic acid in

Scheme 61. Al3+-exchanged synthetic Zn-saponite for alkylation of benzene. Scheme 62. Alkylation of xylenes using Cs-DTP/K10.
126 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

benzylation of toluene where toluene is first brominated by bromine


either in situ or separately to generate benzyl bromide, which then
benzylates toluene to benzyl toluenes. Both the reactions are
catalyzed by ion-exchanged bentonite and montmorillonite K10
(Scheme 66).
Friedel–Crafts acylation of arenes has been carried out using
chloroacetyl chloride on Fe3+-exchanged montmorillonite K10 in
Scheme 63. Friedel–Crafts acetylation using Cs-DTP/K10.
liquid phase (Paranjape et al., 2008) (Scheme 67). The yields are good
only with polymethylated benzene derivatives, namely, durene,
Durap et al. (2006) have used Fe3+, Cr3+ and Al3+ pillared bentonite mesitylene, p-xylene and m-xylene.
for benzylation of benzene and toluene. The Fe3+-pillared clay was Liu et al. (2009) have prepared an efficient montmorillonite K10
found to be the most active catalyst. Kurian and Sugunan (2005) have supported antimony trichloride catalyst for alkylation of nitrogen
carried out benzylation of benzene on alumina pillared transition heterocycles, pyrrole, indole and indole derivatives, using epoxides as
metal cation (Cr3+, Fe3+, etc.) exchanged clays. Mechanistic investi- alkylating agents. The reaction occurs at room temperature under
gation indicated the formation of benzyl cation as intermediate. solvent-free conditions and takes less than an hour for completion in
Ahmed and Dutta (2005) have prepared acid-treated Zn2+ and Cd2+ most cases (Scheme 68). The yields are good to excellent.
ion-exchanged clay composites which act as very good catalysts for Acid-washed montmorillonite K10 was found to be a good catalyst
benzylation of benzene. Barton et al. (2003) developed a process for for alkylation as well as acylation of aromatic and heterocyclic

Scheme 64. Benzoylation of aromatic compounds.

Scheme 65. Alkylation and benzoylation of aromatic compounds.


G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 127

Scheme 66. Benzylation of aromatic compounds.

compounds under microwave irradiation and solvent-free condition Singh et al. (2002) synthesized three sesquiterpenes, elvirol,
(Devi and Ganguly, 2008). The yields are good and comparable to or curcuphenol and sesquichemaenol by Friedel–Crafts alkylation of
better than the yields obtained from reactions using similarly treated appropriate cresols using suitable alkylating agents. The reaction was
silica gel catalyst (Scheme 69). brought about by heating the reaction mixtures in presence of
montmorillonite K10 as catalyst. The targeted sesquiterpenes were
formed in major amounts accompanied by minor quantities of
isomeric compounds (Scheme 70).
Zhang et al. (2008) have prepared a number of dihydrocoumarin
derivatives by a microwave-assisted reaction of cinnamoyl chloride
with phenol and its several derivatives using montmorillonite K10 as
catalyst in chlorobenzene as solvent (Scheme 71). The reaction
proceeds by initial formation of phenyl ester of cinnamic acid
Scheme 67. Chloroacetylation of alkylbenzenes. followed by alkylation-cyclization of the ester to give the observed

Scheme 68. Alkylation of nitrogen heterocycles by epoxides.


128 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 71. Dihydrocoumarins by acylation–cyclization.

shape selectivity and di- and tri-substituted products are also


obtained as well (Binitha and Sugunan, 2008) (Scheme 72). The
catalyst was regenerated by heating and reused four times with little
loss of activity. The authors propose that the nature of porosity in the
pillared clay is responsible for selectivity and that methylation occurs
by an Eley–Rideal type mechanism.
A microwave-assisted montmorillonite K10 catalyzed alkylation of
indoles by tert-butyl alcohol and hexane-2,5-diol has been reported
by Kulkarni et al. (2009b). The alkylation occurs selectively to form 3-
Scheme 69. Alkylation and acylation of aromatic and heterocyclic compounds. tert-butylindole. In the case of the diol, the alkylation is followed by
cyclization to give finally carbazole derivatives (Scheme 73). Good
results are obtained when the irradiation is carried out at 130 °C for 8–
15 min. Though the GC yields are reported to be good, the isolated
products. The microwave irradiation reduces the reaction time from yields are moderate.
several hours to just a few minutes. The yields vary from poor to
excellent. 7. Isomerization reactions
Methylation of toluene with methanol in the presence of chromia-
pillared montmorillonite K10 gives xylenes selectively. Mixed pillar- Isomerization or rearrangement reactions are important in
ing with chromia and titania or zirconia or alumina was found to producing a large number of bulk as well as fine chemicals. A majority
produce more efficient catalysts than pillaring with a single metal of such transformations are brought about by acid-catalyzed process-
oxide. In contrast the natural (unmodified) clay does not exhibit any es using conventional Brønsted and Lewis acids. A variety of acid clays

Scheme 70. Synthesis of sesquiterpenes by Friedel–Crafts alkylation.


G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 129

Epoxides are good substrates for isomerization under a variety of


acid catalytic conditions. The epoxide ring opens under the influence of
Brønsted as well as Lewis acid catalysis. The incipient cation may
undergo skeletal rearrangement through bond migrations and/or
experience nucleophilic attack. The products have diverse applications.
The diastereomeric (R)-(+)-limonene diepoxides (36a,b) under-
went isomerisation at room temperature on synthetic K10 clay
calcined at 100 °C or natural ascanite-bentonite clay to give 37, 38, 39
and 40a,b in a ratio of ~4:3:7:2 (Salomatina et al., 2005) (Scheme 74).
Il'ina et al. (2007) have reported an extensive study of the isomeri-
zation of allyl alcohols of the pinene series and their epoxides using
askanite–bentonite clay (calcined at 110 °C) (Scheme 75). Treatment
of (+)-trans-pinocarveol ((+)-41) with askanite–bentonite at
room temperature led to the formation of isomeric allyl alcohol
(−)-myrtenol ((−)-42), and the dimeric ether (+)-43, selectively. In
contrast, the allylic alcohol (+)-41 gives Wagner–Meerwein rear-
rangement product 44, under homogeneous acidic conditions, due to
protonation of the exocyclic double bond. On the other hand, when
(−)-42 was kept on clay for one hour at room temperature (+)-41
and the rearranged products (+)-45 and 46 were obtained. The
authors suggest the formation of a common intermediate A from (+)-
Scheme 72. Methylation of toluene with methanol and its mechanism. 41 and (−)-42.
To explain the formation of (+)-45 and 46, the authors propose
that (−)-42 gets protonated at –OH to give the carbocation A as well
as at the double bond to give B, while proposing protonation of only –
OH group in the case of (+)-41.
A similar treatment of the epoxides (+)-47 and (−)-48 of these
two allyl alcohols on askanite–bentonite clay yielded slightly differing
results, though the two epoxides were visualized to give the same
intermediate C Scheme 76. The pinocarpeol epoxide (+)-47 yielded,
among other unidentified products, the α,β-unsaturated ketone (+)-
49 and the monocyclic keto alcohol (+)-50. Myrtenol epoxide (−)-48
produced the aromatic alcohol 4-isopropylbenzyl alcohol 46 and the
hydroxy aldehyde (+)-50. Interestingly, the epoxide (+)-47 did not
produce 46, and myrtenol epoxide (−)-48 did not give the
unsaturated ketone (+)-49 (Scheme 76).
(+)-trans-Verbinol ((+)-51) and the epoxide (−)-53 of (−)-cis-
verbinol ((−)-52) were similarly treated with askanite–bentonite
clay. The results obtained are shown in Scheme 77.
Scheme 73. Microwave-assisted alkylation of indoles by alcohols. Storing (−)-cis-verbenol ((−)-52) and its epoxide on askanite–
bentonite clay in the presence of aldehydes produced heterocyclic
compounds 57–65 as depicted in Scheme 78.
also have served as good catalysts for functional group transformation Montmorillonite K10 treated with the heteropolyacid dodeca-
and skeletal rearrangements. The clay catalyzed conversion of low tungstophosphoric acid has been found to be effective in opening 1,2-
octane hydrocarbon fuels to high octane grade is an old well known epoxyoctane (66) to octanal (67), octenol (68) and other products
reaction. Due to enormous progress achieved in the use of clay- (Yadav, 2005) (Scheme 79).
supported catalysts in various areas of organic synthesis, much Isolongifolene (70) is a commercially important tricyclic sesqui-
attention is being focused on the study of clay catalyzed isomeriza- terpene used extensively in the perfumery industry. It is obtained by
tion/rearrangement reactions, in order to replace the hazardous acid-catalyzed isomerization of longifolene (69). A number of
conventional Brønsted and Lewis acids. Many times the products homogeneous acid conditions are used. However, as isolongifolene
formed from clay catalyzed reactions are quite different from those has the tendency to undergo further isomerization the conditions
that are formed in homogenous mineral acid or Lewis acid-catalyzed have to be carefully maintained. Singh et al. (2007) have used alumina
reactions. This aspect gives an opportunity to choose catalysts and and zirconia pillared clays and Ce3+ and La3+ modified montmoril-
reaction conditions to prepare the desired compounds. lonite clays for the conversion of longifolene to isolongifolene. They

Scheme 74. Isomerization of diastereomeric (R)-(+)-limonene diepoxides.


130 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 75. Isomerization of allyl alcohols of pinenes.

found that Al3+ pillared montmorillonite clay shows the highest pillared clay is the best catalyst for the conversion of 1-butene to trans-
conversion rate and reasonable selectivity (Scheme 80). 2-butene and cis-2-butene accompanied by small amounts of cracking
Hydroisomerization of n-heptane to isoheptane was carried out by products (Scheme 82).
Vogels et al. (2005) on synthetic Co-, Mg- and Co/Mg-saponites. They Nucleophilic substitution at allylic position (SN1' or SN2' type)
relate the isomerization to the Lewis acidity of the catalyst (Scheme 81). with isomerization of double bond has been observed by Shanmugam
Moronta et al. (2008) have studied the isomerization of 1-butene and Vaithiyanathan (2008) on K10 clay under neat conditions in the
over natural smectite clay (STx-1, USA) ion-exchanged with Al3+, Fe3+ case of the derivatives of oxindole 71 (Scheme 83). The products
or pillared with Al and Fe polyoxocations. The results show that Al- (72a–c) were further used for construction of spirolactone ring (not

Scheme 76. Isomerization of epoxy alcohols of pinenes.


G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 131

Scheme 80. Longifolene to isolongifolene.

Scheme 77. Isomerization of trans-verbinol and cis-verbinol epoxide.

Scheme 81. Hydroisomerization of n-heptane to isoheptane.

phenylacetylene (73) carried out on Zn2+-exchanged montmorillon-


ite K10 clay by the addition of aniline (74) to the triple bond, followed
by isomerization of the product enamine (75) to imine (76)
(Scheme 84). The authors suggest the activation of the triple bond
by π-complexing with Zn2+ ions in K10, which enables the
nucleophilic addition. The isomerization is due to protonation of the
double bond followed by removal of proton on nitrogen.
Ortega et al. (2006) have achieved success in enriching the cis-
isomer 77a in an isomeric mixture of cis- and trans-lauthisan (77a and
77b) from an isomer ratio of cis:trans = 1:1.7 to 17:1, on montmo-
rillonite K10. It should be noted with much appreciation that many
other acid catalysts including BF3.OEt2, CF3COOH and TsOH had either
failed to bring about any reaction or had led to decomposition of 77a,
b. This K10 catalyzed enrichment procedure enabled the authors to
develop a short route to the desired isomer (+)-cis-lauthisan in good
yield (Scheme 85).
Tomooka et al. (2000) have observed an interesting case of
stereoregulated substitutive migration of a phenyl group from TBDPS
ether 78 to give 79, thereby achieving an asymmetric synthesis of α-
aryl and β-hydroxycyclic amines and silanols (Scheme 86).
Two highly branched 25-carbon isoprenoid olefins, a diene (80)
and a triene (81), isolated from Haslea ostrearia were treated with
montmorillonite K10 clay at room temperature. The diene was found
to unergo isomerization of a disubstituted double bond to a tri-
substituted double bond, whereas the triene underwent cyclization.
The results were rationalized by assuming that K10 clay acts as
Brønsted acid and hence protonates one of the double bonds to give
tertiary cations 82 and 83, which finally gave the observed products
84, 85 and 86 (Belt et al., 2000) (Scheme 87). The authors claim that
the work has implications in understanding the presence of the
observed organic compounds in sedimentary materials.
Dintzner et al. (2004) have investigated the migration of isoprenyl
Scheme 78. Reactions of cis-verbinol and its epoxide with aldehydes. group in phenyl isoprenyl ether (87) using montmorillonite K10 and
montmorillonite KSF clays. With K10 clay the isoprenyl group moves
shown in the scheme) through Morita–Baylis–Hillman reaction mainly to ortho-position by a [1,3] migration to give ortho-prenylated
followed by hydrolysis. phenol 88 (Scheme 88). o-Prenylated phenols show broad range of
Isomerization of enamine 75 to imine 76 has been observed by pharmacological activity, and therefore the reaction has the potential as
Shanbhag and Halligudi (2004). They report hydroamination of a ‘green chemical’ method of synthesis of o-prenyl phenolic derivatives.

Scheme 79. 1,2-Epoxyoctane ring opening.


132 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 82. Isomerization and cracking of 1-butene.

Scheme 85. cis-Lauthisan to trans-lauthisan isomerization.

Scheme 83. Allylic nucleophilic substitution–isomerization.

The isoprenyl group also migrates to the para-position by a [1,5]


migration to give a small amount of p-prenylated phenol 89. o-Prenyl Scheme 86. Stereoregulated substitutive migration of phenyl group.
phenol 88 undergoes further protonation and cyclization to dihydro-
pyran derivative 90. The authors observed that the mont.KSF clay also
catalyzes the isomerisation but at a very slow rate. For example, with
mont.K10, the reaction is complete in just 30 min, while with mont.KSF clay, microwave irradiation of a mixture of 91a and 92 delivers
it takes 15 h to complete. The K10 catalyzed reaction was faster in intermediate prenyl ether (not shown in the scheme) and not 93. The
CH2Cl2, but in CCl4 the selectivity was higher. authors have not suggested any mechanism for the reaction, but it is
Prenylated phenolic compounds, such as prenylated xanthones conceivable that initially prenyl ethers are formed which then
have been prepared by microwave irradiation of hydroxyxanthones undergo [1,3] migration, followed by proton catalyzed cyclization of
(91a–c) with prenyl bromide (92) in chloroform solution in the the intermediates to the observed products 93–98 (Scheme 89).
presence of montmorillonite K10 clay. The reaction with 91a takes A [3,3]-sigmatropic shift following the addition of arylhydrazine
just 20 minutes to give about 86% yield of the final dihydropyran hydrochlorides (99) to cyclic enol ethers and enol lactones (100) takes
product 93 (Castanheiro et al., 2009) (Scheme 89). The conventional place, in a manner exactly similar to the one in Fischer indole reaction, to
heating at 100 °C takes 1 h and at room temperature the reaction give a variety of substituted indoles (101) (Scheme 90). The reaction is
takes 5 days to give 63% yield of the product. When the reaction was catalyzed by several Brønsted acids. However, montmorillonite K10
carried out at 200 °C with ZnCl2 as catalyst, the yield of the product promoted reaction in 50% aqueous N,N-dimethylacetamide at 80 °C
was only 22%. The other two hydroxyxanthones 91b and 91c give gave the best yields. Amberlyst-15 and amberlyst-120 gave reasonably
mixtures of products because two ortho-positions are available for good results, but zeolite-HY and silica performed poorly (Singh et al.,
cyclization in the prenyl ether intermediates. In the absence of K10 2008).

Scheme 84. Addition–isomerization reaction.


G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 133

Scheme 87. Isomerization of isoprenoid olefins.

8. Oxidation reactions clay mainly catalyzes the trimerisation of aldehydes to trialkyl 1,3,5-
trioxanes, which is a temperature dependent equilibrium reaction
In this section oxidation reactions that are brought about by a few with trimer-to-aldehyde ratio of 3.9:1 attained at −20 °C. However,
different types of oxidizing reagents in the presence of clay catalysts are when KSF clay is used, oxidation to carboxylic acid takes place under
reviewed. The reactions described include oxidation of aldehydes to aerobic conditions. Only aliphatic aldehydes undergo oxidation, but
carboxylic acids, alcohols to carbonyl compounds, epoxidation, Baeyer– not the aromatic or the unsaturated aliphatic aldehydes.
Villiger oxidation, oxidation and dehydrogenation of hydrocarbons. The common procedures for the oxidation of alcohols to ketones or
An interesting difference between montmorillonite K10 and aldehydes make use of compounds of transition metals, such as
montmorillonite KSF in their interaction with aldehydes has been chromium, manganese, vanadium, etc., which are toxic. It is desirable
observed by Dintzner et al. (2010) (Scheme 91). For example, the K10 to reduce their use or replace them by oxidants that are less hazardous

Scheme 88. [1,3] Migration of prenyl group-cyclization of product.


134 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Scheme 89. Preparation of prenylated xanthones and their cyclization.

to health and environment. Many clay-based oxidizing agents that fulfill nitrate and acetoacetate added catalysts give higher yields of products.
this Green chemistry condition have been developed. Eftekhari-Sis et al. The oxidation takes place by a free radical mechanism.
(2007) reported the oxidation of alcohols to aldehydes and ketones Cyclohexanone undergoes Baeyer–Villiger oxidation to ε-caprolactam
using hydrogen peroxide as oxidizing agent in the presence of lithium by hydrogen peroxide in the presence of Brazilian kaolinite interca-
chloride supported on montmorillonite K10. The proposed mechanism lated with a porphyrin derivative, [meso-tetrakis(pentafluorophenyl)
involves the formation of lithium hypochlorite, which is the active porphyrin]-iron(II), Fe(TPFPP). The porphyrin moiety is introduced
oxidizing agent (Scheme 92). between the clay layers after expanding and functionalizing the
Hydrogen peroxide or iodosylbenzene epoxidizes cyclooctene to interlayer space by pretreatment of the clay. The authors (Bizaia et al.,
cyclooctene epoxide and cyclohexane to cyclohexanone in the presence 2009) claim that this is the first example of a porphyrin-in-clay
of modified natural saponite clay (Scheme 93). The modification catalyzed Baeyer–Villiger reaction. In the presence of the same catalyst
procedure involves first, the intercalation of the clay with aluminium iodosylbenzene brings about the epoxidation of cyclooctene and
polycation, followed by calcination at 500 °C to get alumina pillared clay oxidation of cylcohexane to cyclohexanone (Scheme 94). The catalyst
and then impregnation with nickel nitrate or chloride or acetoacetate was reused up to five times without significant diminution in the
(Mata et al., 2009). The amount of nickel loaded may be varied. The yields of products.

Scheme 90. Arylhydrazine addition to cyclic enol ethers and enol lactones to form indoles.
G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 135

Scheme 91. Difference between K10 and KSF action on aldehydes.

Scheme 94. Porphyrin-in-clay catalyzed oxidation reactions.

Scheme 95. V2O5-on-clay/H2O2-AcOH for hydroxylation of benzene to phenol.

Scheme 92. Oxidation of alcohols to aldehydes and ketones.

Gao and Xu (2006) have used vanadium oxide catalyst supported


on clay (chlorite, illite, and atapulgite from Inner Mongolia) for the
oxidation of benzene to phenol using hydrogen peroxide as co-
oxidant. The presence of acetic acid enhances the hydrogen peroxide Scheme 96. Substrate-based oxidation or nitration using nitric acid or nitrate salts.
reaction as it avoids phase separation problem. A 14% conversion with
94% selectivity was achieved (Scheme 95). Several other oxides of
metals, such as copper, iron, manganese, chromium, molybdenum, dilute nitric acid in the presence of natural bentonite oxidizes benzylic
tungsten, etc., were tried in place of vanadium, but were found to be alcohols to the corresponding carbonyl derivatives. In contrast, the same
not as effective. However, the TS-1 was found to be equally good. reagent-catalyst system nitrates activated aromatic compounds regio-
Nitric acid and metal nitrate salts are used as both oxidizing and selectively (Bahulayan et al., 2002) (Scheme 96).
nitrating agents depending on the nature of the substrate and reaction Ethylbenzene has been dehydrogenated to styrene using a
conditions. In most cases the reaction is slow, but if vigorous conditions processed Venezuelan smectite clay intercalated with trinuclear iron
are used to hasten the reaction the substrate may decompose. However, complex, [Fe3O(OAc)6.3H2O]+. Ethylbenzene was heated over the
reports show that such reactions can be accomplished under very mild catalyst for 1 h at 410 °C. There was 50% increase in styrene formation
conditions, if the reagents are supported on smectite clays. For example, with the use of this catalyst (Huerta et al., 2003) (Scheme 97).

Scheme 93. Cyclooctene epoxidation and cyclohexane oxidation by H2O2 or C6H5IO.


136 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

References

Ahmed, O.S., Dutta, D.K., 2005. Friedel–Crafts benzylation of benzene using Zn and Cd
ions exchanged clay composites. J. Mol. Catal. A Chem. 229, 227–231.
Akiyama, T., Matsuda, K., Fuchibe, K., 2002. Montmorillonite K10-catalyzed aza-Diels–Alder
reaction of Danishefsky's diene with aldimines generated in situ from aliphatic aldehydes
and an amine, in aqueous media. Synlett 1898–1900.
Bahulayan, D., Narayan, G., Sreekumar, V., Lalithambika, M., 2002. Natural bentonite
Scheme 97. Ethylbenzene dehydrogenation to styrene.
clay/dilute HNO3 (40%)—a mild, efficient, and reusable catalyst/reagent system for
selective mononitration and benzylic oxidations. Synth. Commun. 32, 3565–3574.
Barton, B., Hlohloza, N.S., McInnes, S.M., Zeelie, B., 2003. Catalysts and process for the
production of benzyl toluenes. Org. Proc. Dev. 7, 571–576.
Belt, S.T., Guy Allard, W., Rintatalo, J., Johns, L.A., van Dun, A.C.T., Rowland, S.J., 2000.
Clay and acid catalyzed isomerization and cyclization of highly branched
isoprenoid (HBI) alkenes: implications for sedimentary reactions and distributions.
Geochim. Cosmochim. Acta 64, 3337–3345.
Bigi, F., Chesini, L., Maggi, R., Sartori, G., 1999. Montmorillonite KSF as an inorganic,
water stable, and reusable catalyst for Knoevenagel synthesis of coumarin-3-
carboxylic acids. J. Org. Chem. 64, 1033–1035.
Scheme 98. Pt-in-synthetic TS-1 in dehydrogenation of ethylbenzene to styrene. Binitha, N.N., Sugunan, S., 2008. Shape selective toluene methylation over chromia
pillared montmorillonites. Catal. Commun. 9, 2376–2380.
Bizaia, N., de Faria, E.H., Ricci, G.P., Calefi, P.S., Nassar, E.J., Castro, K.A.D.F., Nakagaki, S.,
Ciuffi, K.J., Trujillano, R., Vicente, M.A., Gil, A., Korili, S.A., 2009. Porphyrin-kaolinite
as efficient catalyst for oxidation reactions. Appl. Mater. Interfaces 1, 2667–2678.
Borkin, D., Carlson, A., Török, B., 2010. K10-catalyzed highly diastereoselective
Dehydrogenation of ethylbenzene has been carried out at 400 °C synthesis of aziridines. Synlett 0745–0748.
using natural or aluminium pillared Venezuelan clay impregnated with Castanheiro, R.A.P., Pinto, M.M.M., Cravo, S.M.M., Pinto, D.C.G.A., Silva, A.M.S., Kijjoa, A.,
cobalt nitrate or cobalt acetate. The selectivity for styrene varies from 75 2009. Improved methodologies for synthesis of prenylated xanthones by
microwave irradiation and combination of heterogeneous catalysis (K10 clay)
to 90% and the conversion is 20% (Gonzalez and Moronta, 2004). with microwave irradiation. Tetrahedron 65, 3848–3857.
Moran et al. (2007) have dehydrogenated ethylbenzene to styrene Chaudary, B.M., Rao, B.P.C., Chowdari, N.S., Kantam, M.L., 2002. Fe3+–montmorillonite:
using synthetic clay TS-1 impregnated with 0.5 or 1.0 wt.% of platinum. a bifunctional catalyst for one pot Friedel–Crafts alkylation of arenes with alcohols.
Catal. Commun. 3, 363–367.
A 50% conversion was achieved (Scheme 98). The activity of Pt-TS-1 Chiba, K., Hirano, T., Kitano, Y., Tada, M., 1999. Montmorillonite-mediated hetero-Diels–
catalyst is attributed to its high thermal stability and improved surface Alder reaction of alkenes and o-quinomethanes generated in situ by dehydration of
area of the support. o-hydroxybenzyl alcohols. Chem. Commun. 691–692.
Choudhary, V.R., Jha, R., 2008. GaAlClx-grafted Mont.K10 clay: highly active and stable
solid catalyst for the Friedel–Crafts type benzylation and acylation reactions. Catal.
Commun. 9, 1101–1105.
9. Conclusions Choudhary, V.R., Mantri, K., Jana, S.K., 2001a. Selective esterification of tert-butanol by
acetic acid anhydride over clay supported InCl3, GaCl3, FeCl3 and InCl2 catalysts.
Catal. Commun. 2, 57–61.
Natural as well as commercially available clays and their diverse Choudhary, V.R., Jana, S.K., Mandale, A.B., 2001b. Highly active, reusable and moisture
modified forms are good catalysts for accomplishing a large variety of insensitive catalyst obtained from basic Ga–Mg–hydrotalcite anionic clay for
organic reactions. Since the procedures employed for their modification Friedel–Crafts type benzylation and acylation reactions. Catal. Lett. 74, 95–98.
Choudhary, V.R., Patil, K., Jana, S.K., 2004. Acylation of aromatic alcohols and phenols
are usually simple chemical operations, there is great scope to prepare over InCl3/montmorillonite K10 catalysts. J. Chem. Sci. 116, 175–177.
newer clay-supported catalysts that are capable of steering organic Choudhary, V.R., Jha, R., Choudhari, P.A., 2005a. Highly active and reusable catalyst from
reactions in any desired direction to achieve higher yields and greater Fe–Mg–hydrotalcite anionic clay for Friedel–Crafts type benzylation reactions.
J. Chem. Sci. 117, 635–639.
selectivity including diastereo- and enantioselectivity. The present Choudhary, V.R., Jha, R., Narkhede, V.S., 2005b. In–Mg–hydrotalcite anionic clay as
knowledge about the application of a modified catalyst for a particular catalyst or catalyst precursor for Friedel–Crafts type benzylation reactions. J. Mol.
reaction is still empirical; it is difficult to predict accurately whether Catal. A Chem. 239, 76–81.
Chtourou, M., Abdelhedi, R., Frikha, M.H., Trabelsi, M., 2010. Solvent free synthesis of 1,
newly modified clay can catalyze a specific reaction in a desired manner. 3-diaryl-2-propenones catalyzed by commercial acid-clays under ultrasound
Some kind of structure activity correlation studies could throw some irradiation. Ultrason. Sonochem. 17, 246–249.
light on this aspect. Ciciriello, F., Costanzo, G., Pino, S., Di Mauro, E., 2009. Spontaneous generation revisited
at molecular level. In: Pontarotti, P. (Ed.), Evolutionary Biology: Concept, Modeling,
Clays are ubiquitous. Many scientific groups from different parts of the
and Application. Springer, Berlin, pp. 3–22.
Globe have shown that the clay varieties available in their geographical Dabbagh, H.A., Teimouri, A., Najafi Chermahini, A., 2007. Environmentally friendly
locality are as good as the commercially available ones. This provides a efficient synthesis and mechanism of triazenes derived from cyclic amines on clays
HZSM-5 and sulfated zirconia. Appl. Catal. 76, 24–33.
great opportunity to synthetic chemists to get involved in low budget
Dasgupta, S., Török, B., 2008. Application of clay catalysts in organic synthesis. A review.
research activity, even in undergraduate colleges, and yet achieve Org. Prep. Proced. Int. 40, 1–65.
significant results. This is particularly relevant as part of Green chemistry De Paolis, O., Teixeira, L., Török, B., 2009. Synthesis of quinolines by a solid acid-
curriculum. catalyzed microwave-assisted dominocyclization–aromatization approach. Tetra-
hedron Lett. 50, 2939–2942.
There is much scope and need to translate laboratory scale Devi, N., 2006. Green synthesis of deoxybenzoins. Res. J. Chem. Environ. 10, 59–61.
procedures to industrial scale procedures, as there is very little activity Devi, N., Ganguly, M., 2008. Friedel–Crafts reaction in dry media under microwave
at present in this aspect. This is particularly important because clay irradiation. Indian J. Chem. 47B, 153–154.
Dintzner, M.R., Morse, K.M., McClelland, K.M., Coligado, D.M., 2004. Investigation of the
catalysts are environmentally benign, recyclable and economical, and montmorillonite clay-catalyzed [1, 3] shift reaction of 3-methyl-2-butenyl phenyl
there is urgent need to replace the not-so-desirable conventional ether. Tetrahedron Lett. 45, 79–81.
catalysts. There can probably be no other catalytic system that would Dintzner, M.R., Wucka, P.R., Lyons, T.W., 2006. Microwave assisted synthesis of a
natural insecticide on basic montmorillonite K10 clay. J. Chem. Educ. 83, 270–272.
qualify to be called ‘Green’ than the clay-based one. Dintzner, M.R., Little, A.J., Pacilli, M., Pileggi, D.J., Osner, Z.R., Lyons, T.W., 2007.
Montmorillonite clay catalyzed hetero-Diels–Alder reaction of 2,3-dimethyl-1,3-
butadiene with benzaldehydes. Tetrahedron Lett. 48, 1577–1579.
Dintzner, M.R., Mondjnou, Y.A., Unger, B., 2009. Montmorillonite K10 clay-catalyzed
Acknowledgements
synthesis of homoallylic silyl ethers: an efficient and environmentally friendly
Hosomi–Sakurai reaction. Tetrahedron Lett. 50, 6639–6641.
The author is much indebted to the following persons for their Dintzner, M.R., Mondjnou, Y.A., Pileggi, D.J., 2010. Montmorillonite clay-catalyzed
support: Mr. R.K. Rao, Dr. S. Hari Prasad, Dr. M.N. Mallya, Dr. Ravi Kalyan cyclotrimerization and oxidation of aliphatic aldehydes. Tetrahedron Lett. 51,
826–827.
and Ms. Divya Jyothi in literature search and my wife Dr. Renukarani and Durap, F., Akcay, M., Yurdakoc, K., 2006. Benzylation of benzene and toluene with
my son Mr. Suchit Gopalpur in preparing the manuscript. benzyl chloride over clay based acid catalysts. Asian J. Chem. 18, 1803–1807.
G. Nagendrappa / Applied Clay Science 53 (2011) 106–138 137

Eftekhari-Sis, B., Khalili, B., Abdollahifar, A., Hashemi, M.M., 2007. Transition metal free cyclohexane oxidation on Ni/alumina-pillared clay catalysts. Microporous Mesoporous
oxidation of alcohols to carbonyl compounds using hydrogen peroxide catalyzed Mater. 124, 218–226.
with LiCl on montmorillonite K10. Acta Chim. Slov. 54, 635–637. Mitsudome, T., Nose, K., Mizugaki, T., Jitsukawa, K., Kaneda, K., 2008. Reusable
Ehsan, A.M., Ehsan, S., Khan, S., Khan, M.S., 2006. Friedel–Crafts benzylation using clay mineral montmorillonite-entrapped organocatalyst for asymmetric Diels–Alder reaction.
catalysts and new synthesis of metal complex dyes. J. Chem. Soc. Pak. 28, 489–493. Tetrahedron Lett. 49, 5464–5466.
Esveld, E., Chemat, F., Van Haveran, J., 2000. Pilot scale continuous microwave dry- Mittal, V., 2007. Esterification reactions on the surface of layered silicate clay platelets.
media reactor—Part-II: application to waxy ester production. Chem. Eng. Technol. J. Colloid Interface Sci. 315, 135–141.
23, 429–435. Moran, C., Gonzalez, E., Sanchez, J., Solano, R., Carruyo, G., Moronta, A., 2007.
Fabra, M.J., Fraile, J.M., Herrerias, C.I., Lahoz, F.J., Mayoral, J.A., Perez, I., 2008. Surface- Dehydrogenation of ethylbenzene to styrene using Pt, Mo, and Pt–Mo catalysts
enhanced stereoselectivity in Mukaiyama aldol reactions catalyzed by clay- supported on clay nanocomposites. J. Colloid Interface Sci. 315, 164–169.
supported bis(oxazoline)-copper complexes. Chem. Commun. 5402–5404. Moronta, A., Oberto, T., Carruyo, G., Solano, R., Sanchez, J., Gonzalez, E., Huerta, L., 2008.
Fraile, J.M., Garcia, J.I., Herrerias, C.I., Mayoral, J.A., Harmer, M.A., 2004. Bis(oxazoline)– Isomerization of 1-butene catalyzed by ion-exchanged, pillared and ion-exchanged/
copper complexes supported by electrostatic interactions: scope and limitations. pillared clays. Appl. Catal. A Gen. 334, 173–178.
J. Catal. 22, 532–540. Motokura, K., Fujita, N., Mori, K., Mizugaki, T., Ebitani, K., Kaneda, K., 2005. An acidic
Fraile, J.M., Garcia, J.I., Mayoral, J.A., Roldan, M., 2007. Simple and efficient layered clay is combined with a basic layered clay for one-pot sequential reactions.
heterogeneous copper catalysts for enantioselective C–H carbene insertion. Org. J. Am. Chem. Soc. 127, 9674–9675.
Lett. 9, 731–733. Motokura, K., Tada, M., Iwasawa, Y., 2009. Layered materials with coexisting acidic and
Fraile, J.M., Garcia, J.I., Herrerias, C.I., Mayoral, J.A., Pires, E., Salvatella, L., 2009. Beyond reuse basic sites for catalytic one-pot reaction sequences. J. Am. Chem. Soc. 131, 7944–7945.
in chiral immobilized catalysis: the bis(oxazoline) case. Catal. Today 140, 44–50. Motokura, K., Matsunaga, S., Miyaji, A., Sakamoto, Y., Baba, T., 2010. Heterogeneous
Gao, X., Xu, J., 2006. A new application of clay-supported vanadium oxide catalyst to allylsilation of aromatic and aliphatic alkenes catalyzed by proton-exchanged
selective hydroxylation of benzene to phenol. Appl. Clay Sci. 33, 1–6. montmorillonite. Org. Lett. 12, 1508–1511.
Ghiaci, M., Sedaghat, M.E., Kalbasi, R.J., Abbaspur, A., 2005. Application of surfactant- Neji, S.B., Trabelsi, M., Frikha, M.H., 2009. Esterification of fatty acids with short-chain alcohols
modified clays to synthetic organic chemistry. Tetrahedron 61, 5529–5534. over commercial acid clays in a semi-continuous reactor. Energies 2, 1107–1117.
Gonzalez, E., Moronta, A., 2004. The dehydrogenation of ethylbenzene to styrene Nowrouzi, F., Thadani, A.N., Batey, R.A., 2009. Allylation and crotylation of ketones and
catalyzed by a natural and Al-pillared clays impregnated with cobalt compounds: a aldehydes using potassium organotrifluoroborate salts under Lewis acid and
comparative study. Appl. Catal. A Gen. 258, 99–105. montmorillonite K10 catalyzed conditions. Org. Lett. 11, 2631–2634.
Guchhait, S.K., Jadeja, K., Madaan, C., 2009. A new process of multicomponent Povarov Okujima, T., Komobuchi, N., Shimiju, Y., Uno, H., Ono, N., 2004. An efficient synthesis of
reaction-aerobic dehydrogenation: synthesis of polysubstituted quinolines. Tetra- conjugation-expanded carba- and azuliporphyrins using a bicyclo[2.2.2]octadiene-
hedron Lett. 50, 6861–6865. fused tripyrrane. Tetrahedron Lett. 45, 5461–5464.
Gültekin, Z., 2004. Iron(III)-doped montmorillonite catalysis of alkenes bearing Ortega, N., Martin, T., Martin, V.S., 2006. Stereoselective synthesis of eight-membered
sulfoxide groups in Diels–Alder reactions. Clay Miner. 39, 345–348. cyclic ethers by tandem Nicholas reaction/ring-closing metathesis: a short
Hazarika, M.K., Parajuli, R., Phukan, P., 2007. Synthesis of parabens using montmoril- synthesis of (+)-cis-lauthisan. Org. Lett. 8, 871–873.
lonite K10 clay as catalyst: a green protocol. Indian J. Chem. Technol. 14, 104–106. Paranjape, T.B., Gokhale, G.D., Samant, S.D., 2008. Chloroacetylation of arenes using
Huang, T.-K., Wang, R., Shi, L., Lu, X.-X., 2008. Montmorillonite K10: an efficient and choroacetyl chloride in the presence of FeCl3 modified montmorillonite K10. Indian
reusable catalyst for the synthesis of quinoxaline derivatives in water. Catal. J. Chem. 47B, 310–314.
Commun. 9, 1143–1147. Ranu, B.C., Chattopadhyay, K., 2009. Green procedures for the synthesis of useful
Huerta, L., Meyer, A., Choren, E., 2003. Synthesis, characterization and catalytic molecules avoiding hazardous solvents and toxic catalysts. Eco-friendly synthesis
application for ethylbenzene dehydrogenation of an iron pillared clay. Microporous of fine chemicals. Royal Society of Chemistry, Cambridge, U.K., pp. 186–219.
Mesoporous Mater. 57, 219–227. Reddy, C.R., Vijayakumar, B., Iyengar, P., Nagendrappa, G., Jai Prakash, B.S., 2004.
Igbokwe, P.K., Ugonabo, V.I., Iwegbu, N.A., Akachukwu, P.C., Olisa, C.J., 2008. Kinetics of Synthesis of phenyl acetates using aluminium-exchanged montmorillonite clay
the catalytic esterification of propanol with ethanoic acid using catalysts obtained catalyst. J. Mol. Catal. A Chem. 223, 117–122.
from Nigerian clays. J. Univ. Chem. Technol. Metall. 43, 345–348. Reddy, C.R., Iyengar, P., Nagendrappa, G., Jai Prakash, B.S., 2005a. Esterification of
Il'ina, I.V., Volcho, K.P., Korchagina, D.V., Barkhash, V.A., Salakhutdinov, N.F., 2007. succinic anhydride to di-(p-cresyl) succinate over Mn+-montmorillonite clay
Reactions of allyl alcohols of the pinene series and of their epoxides in the presence catalysts. J. Mol. Catal. A Chem. 229, 31–37.
of montmorillonite clay. Helv. Chim. Acta 90, 353–368. Reddy, C.R., Iyengar, P., Nagendrappa, G., Jai Prakash, B.S., 2005b. Esterification of
Kamath, C.R., Shinde, A.B., Samant, S.D., 2000. Diels–Alder reaction of pyran-2(H)-ones. Part carboxylic acids to diesters over Mn+-montmorillonite clay catalysts. Catal. Lett.
5. Diels–Alder reaction of 4, 6-disubstituted pyran-2(H)-ones with 1, 4-naphthoqui- 101, 87–91.
none and N-phenylmaleimide under dry state adsorbed condition (DSAC) on Reddy, G.J., Latha, D., Thirupathaiah, C., Rao, K.S., 2005c. A facile synthesis of 2, 3-
montmorillonite K10, filtrol-24, bentonite, pyrophillite; and Al3+, Zn2+, Fe3+ disubstituted-6-arylpyridines from enaminones using montmorillonite K10 as
exchanged montmorillonite K10 and bentonite. Indian J. Chem. 39B, 270–276. solid support. Tetrahedron Lett. 46, 301–302.
Kantam, M.L., Bhaskar, V., Choudary, B.M., 2002. Direct condensation of carboxylic acids Reddy, C.R., Nagendrappa, G., Jai Prakash, B.S., 2007. Surface activity study of Mn+-
with alcohols: the atom economic protocol catalyzed by Fe3+-montmorillonite. montmorillonite clay catalysts by FT-IR spectroscopy: correlation with esterifica-
Catal. Lett. 78, 185–188. tion activity. Catal. Commun. 8, 241–246.
Kulkarni, A., Török, B., 2010. Microwave-assisted multicomponent domino cyclization– Roelofs, J.C.A.A., van Dillen, A.J., de Jong, K.P., 2000. Base catalyzed condensation of citral
aromatization: an efficient approach for the synthesis of substituted quinolines. and acetone at low temperature using modified hydrotalcite catalysts. Catal. Today
Green Chem. 12, 875–878. 60, 297–303.
Kulkarni, A., Abid, M., Török, B., Huang, X., 2009a. A direct synthesis of β-carbolines via a Saladino, R., Crestini, C., Ciambecchini, U., Ciciriello, F., Costanzo, G., Di Mauro, E., 2004.
three-step one-pot domino approach with a bifunctional Pd/C/K10 catalyst. Synthesis and degradation of nucleobases and nucleic acids by formamide in the
Tetrahedron Lett. 50, 1791–1794. presence of montmorillonites. Chembiochem 5, 1558–1566.
Kulkarni, A., Quang, P., Török, B., 2009b. Microwave-assisted solid acid-catalyzed Salmon, M., Osnaya, R., Gomez, L., Arroyo, G., Delgado, F., Miranda, R., 2001.
Friedel–Crafts alkylation and electrophilic annulation of indoles using alcohols as Contribution to the Biginelli reaction, using bentonitic clay as catalyst and a
alkylating agents. Synthesis 4010–4014. solventless procedure. J. Mexican Chem. Soc. 45, 206–207.
Kurian, M., Sugunan, S., 2005. Selective benzylation of benzene over alumina pillared Salomatina, O.V., Yarovaya, O.I., Korchagina, D.V., Polovinka, M.P., Barkhash, V.A., 2005.
clays. Indian J. Chem. 44A, 1772–1781. Solid acid- catalyzed isomerization of R(+)-limonone diepoxides. Mendeleev
Landge, S.M., Berryman, M., Török, B., 2008. Microwave-assisted solid acid-catalyzed Commun. 15, 59–61.
one-pot synthesis of isobenzofuran-1(3H)-ones. Tetrahedron Lett. 49, 4505–4508. Shanbhag, G.V., Halligudi, S.B., 2004. Intermolecular hydroamination of alkynes catalyzed
Li, Y.X., Bao, W.L., 2003. Microwave-assisted solventless Biginelli reaction catalyzed by by zinc-exchanged montmorillonite clay. J. Mol. Catal. A Chem. 222, 223–228.
montmorillonite clay-SmCl3.6 H2O system. Chin. Chem. Lett. 14, 993–995. Shanmugam, P., Vaithiyanathan, V., 2008. Stereoselective synthesis of 3-spiro-α-
Liu, Y.-H., Liu, Q.-S., Zhang, Z.-H., 2009. An efficient Friedel–Crafts alkylation of nitrogen methylene-γ-butyrolactone oxindoles from Morita–Baylis–Hillman adducts of
heterocycles catalyzed by antimony trichloride/montmorillonite K10. Tetrahedron isatin. Tetrahedron 64, 3322–3330.
Lett. 50, 916–921. Shao, L.-X., Shi, M., 2003. Montmorillonite KSF-catalyzed one-pot, three component,
Loh, T.P., Li, X.-R., 1999. Clay montmorillonite K10 catalyzed aldol-type reaction of aza-Diels–Alder reactions of methylene cyclopropanes with arene carbaldehydes
aldehydes with silyl enol ethers in water. Tetrahedron 55, 10789–10802. and arylamines. Adv. Synth. Catal. 345, 963–966.
Lopez, I., Silvero, G., Arevalo, M.J., Babiano, R., Palacios, J.C., Bravo, J.L., 2007. Enhanced Siddiqui, I.R., Singh, P.K., Srivastava, V., Singh, J., 2010. Facile synthesis of acyclic analogues
Diels–Alder reactions: on the role of mineral catalysts and microwave irradiation in of carbocyclic nucleoside as potential anti-HIV prodrug. Indian J. Chem. 49B, 512–520.
ionic liquids as recyclable media. Tetrahedron 63, 2901–2906. Silva, F.C., de Souza, M.C.B.V., Ferreira, V.F., Sabino, S.J., Antunes, O.A.C., 2004. Natural clays
Marquet, N., Grunova, E., Kirillov, E., Bouyahyi, M., Thomas, C.M., Carpentier, J.-F., 2008. as efficient catalysts for obtaining chiral β-enaminoesters. Catal. Commun. 5, 151–155.
Convenient synthesis of mono- and di-β-hydroxy-β-bis(trifluoromethyl)-(di)imines Singh, V., Khurana, A., Kaur, I., Sapehiyia, V., Kad, G.L., Singh, J., 2002. Microwave
from β-hydroxy-β-bis(trifluoromethyl)-ketones and (di)amines. Tetrahedron 64, assisted facile synthesis of elvirol, curcuphenol and sesquichamaenol using
75–83. montmorillonite K10 clay in dry media. J. Chem. Soc. Perkin Trans. 1, 1766–1768.
Martin-Aranda, R.M., Ortega-Cantero, E., Rojas-Cervantes, M.L., Vicente-Rodriguez, Singh, B., Patial, J., Sharma, P., Agarwal, S.G., Qazi, G.N., Maity, S., 2007. Influence of
M.A., Banares-Munoz, M.A., 2005. Ultrasound activated Knoevenagel condensation acidity of montmorillonite and modified montmorillonite clay minerals for the
of malononitrile with carbonyl compounds catalyzed by alkaline-doped saponites. conversion of longifolene to isolongifolene. J. Mol. Catal. A Chem. 266, 215–220.
J. Chem. Technol. Biotechnol. 80, 234–238. Singh, P.R., Surpur, M.P., Patil, S.B., 2008. An expeditious and environmentally benign
Mata, G., Trujillano, R., Vicente, M.A., Korili, S.A., Gil, A., Belver, C., Ciuffi, K.J., Nassar, methodology for the synthesis of substituted indoles from cyclic enol ethers and
E.J., Ricci, G.P., Cestari, A., Nakagaki, S., 2009. (Z)-Cyclooctene epoxidation and enol lactones. Tetrahedron Lett. 49, 3335–3340.
138 G. Nagendrappa / Applied Clay Science 53 (2011) 106–138

Srinivas, K.V.N.S., Das, B., 2003. A highly convenient, efficient, and selective process for Yadav, G.D., Badure, O.V., 2008. Selective acylation of 1, 3-dibenzyloxybenzene to 3,5-
preparation of esters and amides from carboxylic acids using Fe3+-K10 montmo- dibenzyloxyacetophenone over cesium modified dodecatungstophosphoric acid
rillonite clay. J. Org. Chem. 68, 1165–1167. (DTP) on clay. Appl. Catal. A Gen. 348, 16–25.
Srivastava, V., Gaubert, K., Pucheault, M., Vaultier, M., 2009. Organic–inorganic hybrid Yadav, G.D., Bhagat, R.D., 2005. Clean esterification of mandelic acid over Cs2.5H0.5
materials for enantioselective organocatalysis. ChemCatChem 1, 94–98. PW12O40 supported on acid treated clay. Clean Technol. Environ. Policy 7, 245–251.
Stern, R., Jedrzejas, M.J., 2008. Carbohydrate polymers at the center of life's origins: the Yadav, G.D., George, G., 2008. Single step synthesis of 4-hydroxybenzophenone via
importance of molecular processivity. Chem. Rev. 108, 5061–5085. esterification and Fries rearrangement: novelty of cesium substituted heteropoly
Tichit, D., Lutic, D., Coq, B., Durand, R., Teissier, R., 2003. The aldol condensation of acid supported on clay. J. Mol. Catal. A Chem. 292, 54–61.
acetaldehyde and heptanal on hydrotalcite-type catalysts. J. Catal. 219, 167–175. Yadav, G.D., Kamble, S.B., 2009. Alkylation of xylenes with isopropyl alcohol over acidic
Tomooka, K., Nakazaki, A., Nakai, T., 2000. A novel migration from silicon to carbon: an clay supported catalysts: efficacy of 20% w/w Cs2.5H0.5PW12O40/K10 clay. Ind. Eng.
efficient approach to asymmetric synthesis of α-aryl β-hydroxy cyclic amines and Chem. Res. 48, 9383–9393.
silanols. J. Am. Chem. Soc. 122, 408–409. Yadav, J.S., Reddy, B.V.S., Sadasiv, K., Reddy, P.S.R., 2002. Montmorillonite clay-catalyzed
Varma, R.S., 2002. Clay and clay supported reagents in organic synthesis. Tetrahedron [4 + 2] cycloaddition reactions: a facile synthesis of pyrano- and furanoquinolines.
58, 1235–1255. Tetrahedron Lett. 43, 3853–3856.
Vijayakumar, B., Iyengar, P., Nagendrappa, G., Jai Prakash, B.S., 2004. A facile synthesis of Yadav, G.D., Asthana, N.S., Kamble, V.S., 2003. Friedel–Crafts benzoylation of p-xylene
fatty acid esters of p-cresol catalyzed by acid activated Indian bentonite. Indian J. over clay-supported catalysts: novelty of cesium substituted dodecatungstopho-
Chem. Technol. 11, 565–568. sphoric acid on K10 clay. Appl. Catal. A Gen. 240, 53–69.
Vijayakumar, B., Reddy, C.R., Iyengar, P., Nagendrappa, G., Jai Prakash, B.S., 2005a. Yadav, J.S., Reddy, B.V.S., Sunitha, V., Reddy, K.S., Ramakrishna, K.V.S., 2004a.
Synthesis of p-tolyl stearate catalyzed by acid activated Indian bentonite. Indian J. Montmorillonite KSF-catalyzed one-pot synthesis of hexahydro-1H-pyrrolo(3, 2-
Chem. Technol. 12, 316–321. c)quinoline derivatives. Tetrahedron Lett. 45, 7947–7950.
Vijayakumar, B., Iyengar, P., Nagendrappa, G., Jai Prakash, B.S., 2005b. An eco-friendly Yadav, J.S., Reddy, B.V.S., Srinivas, M., Prabhakar, A., Jagadeesh, B., 2004b. Montmoril-
method for the synthesis of aryl and alkyl esters of carboxylic acids using acid lonite KSF clay-promoted synthesis of enantiomerically pure 5-substituted
activated Indian bentonite. J. Indian Chem. Soc. 82, 922–925. pyrazoles from 2,3-dihydro-4H-pyran-4-ones. Tetrahedron Lett. 45, 6033–6036.
Vogels, R.J.M.J., Kloprogge, J.T., Geus, J.W., 2005. Catalytic activity of synthetic saponite Zhang, Z., Ma, Y., Zhao, Y., 2008. Microwave-assisted one-pot synthesis of dihydrocou-
clays: effects of tetrahedral and octahedral composition. J. Catal. 231, 443–452. marins from phenols and cinnamoyl chloride. Synlett 1091–1095.
Wallis, P.J., Gates, W.P., Patti, A.F., Scott, J.L., Teoh, E., 2007. Assessing and improving the Zhou, C.-H., 2010. Emerging trends and challenges in synthetic clay-based materials
catalytic activity of K10 montmorillonite. Green Chem. 9, 980–986. and layered double hydroxides. Appl. Clay Sci. 48, 1–4.
Wang, S., Guin, J.A., 2002. Etherification of dimethylbutene with methanol over clay- Zhu, J., Chen, L., Wu, H., Yang, J., 2009a. Highly efficient procedure for the synthesis of Schiff
based acid catalysts. React. Kinet. Catal. Lett. 75, 169–175. bases using hydrotalcite-like materials as catalyst. Chin. J. Chem. 27, 1868–1870.
Wang, Y., Li, W., 2000. Kinetics of acetic acid esterification with 2-methoxyethanol over Zhu, Z.-B., Shao, L.-X., Shi, M., 2009b. Brønsted acid or solid acid catalyzed azo-Diels–Alder
a pillared clay catalyst. React. Kinet. Catal. Lett. 69, 169–176. reactions of methylenecyclopropanes with ethyl (arylimino)acetates. Eur. J. Org.
Yadav, G.D., 2005. Synergism of clay and heteropoly acids as nano-catalysts for the Chem. 2576–2580.
development of green processes with potential industrial applications. Catal. Surv. 9,
117–137.

You might also like