You are on page 1of 47

Catalysis Reviews

Science and Engineering

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/lctr20

A comprehensive review on catalytic O-alkylation


of phenol and hydroquinone

Priyanka Bhongale, Sunil Joshi & Nilesh Mali

To cite this article: Priyanka Bhongale, Sunil Joshi & Nilesh Mali (2021): A comprehensive
review on catalytic O-alkylation of phenol and hydroquinone, Catalysis Reviews, DOI:
10.1080/01614940.2021.1930490

To link to this article: https://doi.org/10.1080/01614940.2021.1930490

View supplementary material

Published online: 22 Jun 2021.

Submit your article to this journal

Article views: 114

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=lctr20
CATALYSIS REVIEWS
https://doi.org/10.1080/01614940.2021.1930490

A comprehensive review on catalytic O-alkylation of phenol


and hydroquinone
Priyanka Bhongalea,b, Sunil Joshia,b, and Nilesh Malia,b
a
Chemical Engineering and Process Development Division, CSIR-National ChemicalLaboratory, , Pune-
411008, India; bAcademy of Scientific and Innovative Research (AcSIR), Ghaziabad-201002, India

ABSTRACT ARTICLE HISTORY


The alkylation process involves two competitive paths of O- and Received 14 September 2020
C-alkylation and achieving better selectivity for desired products Revised 16 March 2021
is a very challenging problem. The development of new process Accepted 06 May 2021
for synthesis of O-methylated products of phenol and dihydric KEYWORDS
phenols is a subject of high industrial and academic interest. O-alkylation; phenol;
Alkyl phenyl ethers, especially anisole and 4-methoxyphenol, hydroquinone; dimethyl
have captivated significant interest due to their increasing appli­ carbonate; methanol
cations in pharmaceutical industries. The main emphasis of the
present review is to explore the recent development in two
catalytic O-alkylation processes. The first process is
O-methylation of phenol into anisole and another is selective
mono O-methylation of hydroquinone into 4-methoxyphenol.
The present article covers O-alkylation methods with methanol
and dimethyl carbonate as alkylating agent over various acidic
and basic catalytic systems. The catalyst systems analyzed
involves Bronsted and Lewis acidic and basic ionic liquids, con­
ventional acids, metal oxides, solid acid and basic catalysts,
hydrotalcites, various zeolites and heteropolyacids. The
mechanistic behavior of alkylation reactions in presence of
different catalytic system is reviewed critically which is impor­
tant to design new and/or modified catalyst in order to max­
imize the yield of desired product. Additionally, an influence of
reaction parameters, role of catalyst and their active sites on
product distribution is described. The review paper gives useful
insight for researchers in the field of catalysis and reaction
engineering of alkylation reactions. Understandings of the reac­
tion pathways will help in developing reliable kinetic models
necessary for process scale-up to industrial scale reactor system.

Contents
(1) Introduction
(2) O-alkylation of phenol into anisole
(1) Methanol as methylating agent
1. Zeolites
2. Metal oxides

CONTACT Sunil Joshi ss.joshi@ncl.res.in Academy of Scientific and Innovative Research (Acsir), CSIR-NCL
Campus, Pune, India
Supplemental data for this article can be accessed on the publisher’s website.
© 2021 Taylor & Francis
2 P. BHONGALE ET AL.

3. Phosphates and sulfates


4. Other solid catalysts
(2) Dimethyl carbonate as methylating agent
1. Ionic liquids
2. Phase transfer catalysts and base
3. Other solid basic catalysts
1 Basic zeolites
2 Hydrotalcites
3 Metal oxides
(3) Selective mono O-alkylation of hydroquinone into 4-methoxyphenol
(1) Methanol as methylating agent
(2) Dimethyl carbonate as methylating agent
(4) Challenges in process development
(5) Conclusions

Acknowledgment
References

1. Introduction
Alkylation of phenol and substituted phenols is an important organic reaction,
which leads to either O-alkylation (ether products) or C-alkylation (ring
alkylated products). These two reactions are competitive and thus post
major challenges in controlling the rate of formation toward one of the
alkylated product. Selective synthesis of monoalkylated products is a subject
of high industrial interest due to their vital role in production of fine and
specialty chemicals.[1–4] The interest in improving selectivity toward the
desired product, studying an effect of substrate variation and catalyst activity
by taking into consideration environmental and economic aspects has moti­
vated researchers to study catalysis of aromatic alkylation and related process
optimization.[5]
Alkylation of phenol and its derivative can be accomplished using different
alkylation agents such as alcohols, dimethyl ether, dimethyl carbonate,
dimethyl sulfate and alkyl halide.[6] Commercially, monoethers are prepared
by Williamson ether synthesis reaction. Industrial process for production of
O-methylated product of phenol and dihydroxybenzenes involves the use of
toxic and corrosive reagents like dimethyl sulfate or alkyl halide in presence of
stoichiometric amount of sodium hydroxide as catalyst.[7] The strong acids
such as sulfuric acid and hydrohalic acids are used as homogeneous catalysts
in industrial processes.[8–11] In 1981, Eastman Kodak company patented
mono-O-methylation of hydroquinone by methanol in presence of sulfuric
acid and benzoquinone. This process can be used to develop industrial pro­
duction of 4-methoxyphenol.[12] These methods are environmentally
CATALYSIS REVIEWS 3

hazardous and also produce waste which needs further treatment and thus
there is a high demand to develop a green catalytic based processes.[13–15] In
order to overcome these problems and considering an enormous interest in
green chemistry, different research approaches for catalyst development have
been employed and various new homogeneous and heterogeneous catalysts
have been developed for alkylation of phenols.[16,17] It includes homogenous
catalysts like ionic liquids with good separation and reusability property. The
heterogenization of homogeneous catalyst research have attracted much atten­
tion which offers various advantages such as ease of separation, selective
formation of targeted products, reusability and waste minimization. The
global heterogeneous catalyst demand in 2017 was approximately US
$24.29 billion and it is expected to grow up to US$35.63 billion by 2025.[18]
Petroleum refining industries are prominent application segment in chemical
industry which has approximate 23% global catalyst market value share and
almost 16% of petroleum refining industries involves alkylation reactions.[19]
These numbers are significant and indirectly showed high demand for devel­
opment of improved or and new heterogeneous catalyst for alkylation
reactions.
O-alkylated product of phenol named anisole has captivated immense
interest due to its increasing significance as building blocks for production
of dyestuff, cosmetics and fragrances.[20,21] Additionally it is an important
intermediate in agrochemical industries used in herbicides, pesticides and
insecticides productions.[22] Another vital role of anisole is as an additive in
gasoline for octane booster. It is used in huge volume as antioxidants in oils
and greases manufacturing and stabilizers for plastics and polymers.[23] Owing
to the numerous applications of anisole, researcher have been attempted to
synthesize anisole using dimethyl sulfate under microwave conditions, with
methanol and dimethyl carbonate as methylating agent and under solvent free
conditions.[24–27]
It is simple to alkylate hydroxyl group in phenol in which an initial forma­
tion of alkali phenolate react with alkylating agent to produce etherified
phenol. While in case of dihydric phenol viz., hydroquinone, catechol and
resorcinol, many difficulties can be occurred to obtain an alkylation of single
hydroxyl group because monoalkylated product formed can readily react with
alkylating agent to produce dialkylated product. The reaction mixture com­
prised of dihydric phenol, mono and di alkylated product is difficult and costly
to separate. Among three dihydric phenols, hydroquinone is the most tedious
to alkylate into monoalkyl ether due to presence of hydroxyl groups in para
position.[28]
Mono and di O-methylated products of 1,4-hydroquinone, named 4-meth­
oxyphenol and 1,4-dimethoxybenzene have an enormous attention from both
an industrial and a biological point of view.[29] 4-methoxyphenol is an active
ingredient in topical drugs used for skin depigmentation.[30] It also has wide
4 P. BHONGALE ET AL.

applications such as an antioxidant, a polymerization inhibitor in vinyl and


acrylic monomers, an intermediate in manufacturing of dyes and pharmaceu­
ticals, a stabilizer of photosensitive materials and acrylic acid.[31,32] It is
a valuable intermediate for perfume and flavor industries, especially for food
additives like vanillin.[33] 1,4-dimethoxybenzene is used as an intermediate in
dyestuff and pharmaceuticals as well as in production of perfumes and
soaps.[23,30] 4-methoxyphenol has demand in the international market as it
is considered as a high volume chemical in USA and has a product capacity of
more than 500000 kg/annum.[34]
The alkylation of phenol with dimethyl carbonate or methanol in presence
of catalyst gave several products such as anisole, cresols, methylanisole and
xylenols. The alkylation of hydroquinone resulted in 2-methyl hydroquinone
via C-methylation. The choice of catalyst, type of acidic and basic sites on the
catalyst surface, operating conditions of reaction are influential factors affect­
ing on the product distribution. Several papers and patents have been pub­
lished on O-alkylation of phenol and hydroquinone with dimethyl carbonate
and methanol as methylating agent in presence of different catalysts such as
mineral acids, ionic liquids, ion exchange resins, zeolites, solid acid and basic
catalysts, mixed metal oxides and hydrotalcites.[29,35–40]
The present work reviews the progress on alkylation of phenol and hydro­
quinone using several catalysts employing methanol and dimethyl carbonate
as methylating agent. Influence of reaction parameters and active sites on
catalyst surface on conversion of reactant and product selectivity is elaborated.
The detail mechanistic perspectives of alkylation with methanol and dimethyl
carbonate in presence of acidic and basic catalytic system is emphasized in the
article. The understanding of catalytic behavior through reaction kinetic
analysis is described. The selection of suitable catalyst with appropriate pro­
cess parameters is very important in commercialization of any process. These
insights are useful to develop synthesis strategy toward new and or improved
catalysts, predict more realistic reaction conditions for selective product for­
mation and design industrial catalytic reactor systems.

2. O-alkylation of phenol
The alkylation of phenol is an electrophilic substitution reaction, in which the
alkylating agent is preferentially attacked by electrons in the ring or the oxygen
atom in the hydroxyl group.[16] It is easy to break O-H bond in phenol due to
partial positive polarization of oxygen and stabilization by resonance of the
conjugated base as shown in Figure 1. The hydroxyl group in phenol is
directing the electrophile to the ortho and para positions. Therefore possible
products formed during phenol alkylation are anisole, cresols, substituted
anisole and di/tri methylphenols. The electron density is higher at oxygen
atom in hydroxyl group as compared to the carbon atoms in ortho and para
CATALYSIS REVIEWS 5

Figure 1. Resonance of phenol

positions and thus it favors O-alkylation rather than C-alkylation, at least


initially.
The acid-base catalyst is essential to perform an alkylation reaction of
phenol and product selectivity can be correlated on the basis of acid-base
properties of the catalyst. The orientation of phenol on the surface of catalyst,
type of methylating agent, reaction conditions, and acid-base strength, deac­
tivation rate of catalyst are major factors which controls the behavior of
a catalyst.[41]

2.1. Methanol as methylating agent


The alkylation of phenol with methanol can be accomplished in liquid or
vapor phase depending on the reaction conditions. Very few report presented
the liquid phase methylation of phenol due to various limitations associated to
it such as long reaction time, high temperature, high pressure and secondary
isomerization reactions.[36] Most of researchers have preferably studied vapor
phase mode methylation since it is promising, economical and simpler in
engineering design.[42] Mostly, acid as well as basic catalysts were reported for
methylation of phenol with methanol which gave various products such as
anisole, cresols, methylanisoles and xylenols. The activity of methanol is much
lower as compared to dimethyl carbonate and thus high temperature is desir­
able for conducting methylation reaction using methanol as methylating agent
where difficulties such as coke formation and catalyst deactivation were
occurred frequently.[35,43,44]
In section 2, we stated that acid-base catalyst is essential which can be
referred as nature of acid or base (Lewis or Bronsted) and strength (strong
or moderate or weak). The presence of basic site on the catalyst surface is
important since it facilitates the generation of phenolate species which is
supposed to be adsorbed on the catalyst surface. However, this adsorption
geometry moves the para-position of the aromatic ring far from the surface
which results in more difficulty to attack by methanol species adsorbed on the
surface. Lewis or Bronsted acidic sites are responsible for activation of metha­
nol which is well explained through Figure 2. The plausible reaction mechan­
ism for formation of anisole in presence of basic site and Lewis or Bronsted
6 P. BHONGALE ET AL.

Figure 2. Reaction mechanism for O-methylation of phenol with methanol over Bronsted acid and
basic site; Lewis acid and basic site

acidic site are depicted in Figure 2. Methanol activated on Lewis acid sites
undergoes a direct attack by phenolate via SN2 mechanism with water as
leaving group.[45]

Figure 3. An alkylation of phenol with methanol and decomposition of methanol


CATALYSIS REVIEWS 7

Figure 3 represents a reaction pathway for formation of primary products


(anisole, o-cresol and p-cresol) of phenol methylation and decomposition of
methanol. Methanol undergoes decomposition at high temperature; dimethyl
ether, carbon dioxide, carbon monoxide and hydrogen were formed over
acidic sites whereas formaldehyde, dimethyl formate was observed over basic
sites. The presence of Lewis and Bronsted acidic sites and its nature viz.,
whether it is strong or weak plays a significant role in product selectivities.[46]
A schematic representation of the various reaction pathways for formation
mechanism of secondary products is presented in Figure 4 (from anisole) and
Figure 5 (from cresol). Anisole can undergo various reactions such as alkyla­
tion, dealkylation, rearrangement and disproportionation. Phenol and olefins
may be obtained from anisole by dealkylation in presence of acid. Anisole
disproportionation gives 2-/4-methylanisole and phenol. The direct isomer­
ization of anisole leads to o- or p-cresol. Anisole itself acts as effective methy­
lating agent by alkylating phenol into cresols over solid acid catalyst.[35] The
2-/4-methylanisole can be achieved via alkylation of anisole using methanol.
The C-alkylation of o-cresol results in 2,4-xylenol and 2,6-xylenol while

Figure 4. Possible reaction pathways from anisole


8 P. BHONGALE ET AL.

Figure 5. Possible reaction pathways from o- and p-cresol

O-alkylation gives 2-methylanisole. The O- and C-alkylation of p-cresol pro­


duces 4-methylanisole and 2,4-xylenol, respectively. The m-cresol is obtained
by isomerization of o- and p-cresol as it is thermodynamically favored cresol
isomer, but it is not kinetically favored by electrophilic substitution. Further,
alkylation of xylenol forms 2,4,6-trimethylphenol. A wide variety of catalysts
have been explored for the alkylation of phenol with methanol such as zeolites,
sulfates and phosphates, metal oxides.[35,36,44,47,48]

2.1.1. Zeolites
Zeolites are typically defined as crystalline molecular sieves or aluminosili­
cates, with a three dimensional framework structure of intracrystalline chan­
nels and cages. The cages are built from SiO44- and AlO43- tetrahedral sharing
of oxygen atoms. Zeolites are extensively used as catalysts in various organic
reactions as they are nontoxic and non-corrosive and easy to recover for
reuse.[49] The properties of zeolites depend on the framework structure,
CATALYSIS REVIEWS 9

shape, size, connectivity between channels and cages, composition of


framework.[50] The acidity and acid strength of a zeolite can be modified by
changing preparation method, exchanging the cations, decreasing Si/Al ratio
or by isomorphous substitution of Al and Si.[51] However, the basicity can be
increased with increasing more Si-O-Al oxygens and by ionic exchange with
larger cation, for instance exchanging Na+ with Cs+ results in increase in
basicity.[52,53] Another major importance of zeolites is to provide a stereo-
and regio- control in chemical reactions. The presence of acidic and basic sites
in zeolites plays a significant role in synthesizing various alkylated products.
Zeolites are also finds wide industrial applications due to its high thermal
stability and easy method to regenerate the spent catalyst.[54] Anisole, o-cresol
and p-cresol are primary products of methylation of phenol with methanol
while xylenols, methylanisoles, m-cresol are secondary products which are
formed by reaction of primary products with phenol and methanol.
In 1980, Namba et al.[55] studied an effect of acid strength of Y zeolite by
introduction of Na and K in HY zeolite on methylation of phenol where the
moderate acid sites (−3.0> H0>-8.2) on HY zeolite resulted in p-cresol
selectively and weaker acidic sites (1.5> H0>-3.0) selectively yielded anisole.
The performance of K-HY was more effective than Na-HY for selective
formation of p-cresol due to more electronegativity of potassium than
sodium.
Balsama et al.[56] explored several zeolites such as X, Y, ZSM for phenol
methylation and also aimed to evaluate the reactivity of methanol. They tested
the stability of methanol at 523 K over all zeolites in absence of phenol and it
showed 15–65% loss of methanol in presence of hydrogen as carrier gas. The
HZSM-5 catalyst exhibited higher rate toward conversion of methanol into
hydrocarbons. The loss of methanol into side reactions over zeolites was
increasing in the order of X < Y< ZSM. Balsama et al.[56] showed that phenol
could be alkylated not only by methanol, but also by anisole. Hence, in the
alkylation of phenol by methanol, the possibility of secondary reactions of
phenol with anisole should be considered.
Renaud et al.[4] performed an alkylation of phenol in microcatalytic reactor
in presence of helium as carrier gas using H+ and Na+ forms of ZSM-5 zeolite.
Maximum selectivity for anisole obtained was 80% with only 2% phenol
conversion at 473 K temperature and 1.2 h−1 liquid hour space velocity
(LHSV) over HZSM-5 catalyst. A catalytic activity of NaZSM-5 toward methy­
lation and coke formation was much lower than HZSM-5 due to presence of
lower acidic sites on the surface of NaZSM-5. The basic form of ZSM-5 was
found to be selective toward O-methylation while acidic form yielded
C-methylated products. With increase in temperature, both forms of ZSM-5
showed increase in conversion of phenol with decrement in anisole selectivity
and higher rate of coke formation due to presence of Bronsted acidic sites on
the catalyst surface.[4]
10 P. BHONGALE ET AL.

Marczewski et al.[57] confirmed the role of anisole as alkylating agent while


developing the reaction network over ultrastable HY zeolite which showed
that rate of O-methylation is approximately three times the rate C-alkylation.
Cresol was not the primary product formed over ultrastable HY zeolite, but it
was observed that cresol formation was mainly due to the alkylation of phenol
by anisole. An addition of methanol in anisole did not significantly affect the
selectivity of methylanisoles because rate of formation of methylanisoles via
disproportionation and dealkylation of anisole was higher as compared to
alkylation of anisole by methanol.[57]
Xu et al.[58] investigated alkylation of phenol with methanol on HBEA
catalyst with Si/Al ratio 14.1 which showed maximum 55.7% selectivity of
anisole at 473 K. They found significant effect of reaction temperature on
rate of reaction with enhanced conversion of phenol from 46.4 to 76.3%,
while selectivity of anisole dropped to 8.8% with increase in formation of
xylenols and trimethylphenols as temperature was varied from 473 to 573 K.
A commercial NaX zeolite showed higher selectivity of O-methylated pro­
duct when higher ratio of methanol to phenol in range of 5–8:1 was used in the
reaction. A higher concentration of alkylating agent increased the O-alkylation
rate due to the greater possibility of nucleophilic attack of phenoxide ion over
the alkylating agents.[59] Almost 99.8% phenol conversion and 90.2% anisole
selectivity was observed at 593 K temperature, 8:1 methanol to phenol mole
ratio with 0.5 h−1 space velocity over NaX zeolite.
Reddy et al.[60] improved the selectivity toward O-methylated product,
anisole over molybdenum oxide supported on NaY zeolite by suppressing
rate of ring alkylation, but an increase in Mo loading showed decreasing trend
in phenol conversion due to drop in surface areas and total acidity of catalyst.
Highest anisole selectivity (92.0%) was observed over Mo-NaY zeolite among
all reported zeolites at 673 K temperature.
Barman et al.[61] studied the influence of cerium exchanged with NaX
zeolite on methylation of phenol where high selectivity of o-cresol (70.0%)
and xylenol (20.0%) was examined due to increase in acidity of NaX zeolite.
The turnover frequency of Ce-exchanged NaX was decreasing with Ce content
which indicate that cerium in the catalyst does not enhance the catalyst activity
toward O-methylation. A kinetic analysis of the process was also reported
assuming pseudo first order of reaction, plug flow through the catalytic bed
and differential equations were used to estimate the rate constant (9.04 × 10−3
kgmol/kg.h) and the activation energy (57.2 kJ/mol).[61]
Bhattacharyya et al.[62] found the acidity of MCM-41 low as compared to
zeolites due to disordered wall structure. They incorporated Al atom into
MCM-41 in order to increase acidity of the catalyst and examined its impact
on methylation of phenol which showed higher rate of formation of
C-methylated product mainly o-cresol with 62.0% selectivity.
CATALYSIS REVIEWS 11

Further Bregolato et al.[63] synthesized HBEA having high Si/Al (130–154)


ratio with different crystal size and tested its impact on catalyst behavior. The
surface acidity of the catalyst was exhibited due to medium to low strength
silanols. Higher Si/Al ratio showed lower phenol conversion but rate of anisole
formation was found to be higher. Larger crystal size of HBEA revealed more
initial catalytic activity due to higher intraparticle residence time of reactants.
This leads to lower ratio of anisole to cresol because of higher rate for
transformation of anisole to cresols by consecutive intramolecular rearrange­
ment and intermolecular alkylation of anisole. Catalyst deactivation rate was
found increasing with temperature because methanol undergoes self oligo­
merization, cyclization to olefins and aromatics. The time on stream studies
indicated blocking of the strong acid sites by coke deposition but weak acid
sites were not blocked and thus higher formation of anisole was reported with
time on stream.[63]
The role of methanol in acid and base catalyzed gas-phase methylation of
phenol was studied by Ballarini et al.[64] in a flow reactor. The formation of
electrophilic species in acid media and generation of active methylating species
by dehydrogenation of methanol over basic catalyst was responsible for
methylation of phenol.[64]
Sad et al.[35] reported gas phase alkylation of phenol with methanol on
various zeolites such as HBEA, HZSM-5 and HMCM-22 in fixed bed reactor
and also established the reaction network by performing reaction of primary
products with phenol and methanol. In case of zeolites, Table 1 indicates
dependency of product distribution and catalyst deactivation rate on physico­
chemical properties. The crystal structure of zeolites were determined through
X-ray diffraction, surface area were estimated by BET method, ammonia TPD
was used to determine the density and strength of acid sites. Pyridine FTIR
was used to conclude the nature of surface acid sites and coke formed was
measured by TPO analysis. The strength and nature of acidic sites showed
drastic change in O- and C-alkylated product distribution as listed in the Table
1. O/C-alkylation ratio is defined as ratio of selectivity of O-alkylated product
to C-alkylated products. An equal distribution of Bronsted and Lewis acidic
sites on catalyst surface showed anisole and o-cresol as major products in case
of HBEA and HZSM-5. The HBEA having lower acidity as compared to
HZSM-5 and HMCM-22 directed the reaction toward dialkylation and thus
high selectivity of xylenol was observed over HBEA. As mentioned in Table 1,
total acidity of HZSM-5 is higher which showed significant decrement in
selectivity for xylenol and suppress the rate of isomerization reaction. The
kinetics and mechanism of coke formation and catalyst deactivation was
studied using temperature programmed oxidation (TPO) and Fourier trans­
form infrared (FTIR) analysis.[65] The reaction phase played a significant role
in methylation of phenol which was clearly seen by Table S1 where gas phase
phenol methylation with HZSM-5 resulted in 23.0% conversion of phenol with
12
P. BHONGALE ET AL.

Table 1. Dependency of product distribution and catalyst deactivation rate on physicochemical properties of zeolites and other solid catalysts [35,65]
Zeolite type TPD B/L ratio Pore size O/C-alkylation ratio Deactivation rate×103(min−1)
(Si/Al ratio) (µmol/m2) (Å)
HBEA (12.5) 0.90 1 6.6×6.7;5.6×5.6 0.49 10.5
HZSM-5 (20.0) 2.20 1 5.1×5.5;5.3×5.6 0.70 14.6
HMCM-22 (15.0) 1.18 3.2 4.0×5.5;4.1×5.1 0.09 18.8
HY (2.4) 2.10 0.67 7.4×7.4 0.45 27.0
SiO2-Al2O3(11.3) 1.80 0.33 45 0.51 2.1
TPA/SiO2(a28wt.%) 0.80 6 225 1.86 17.5
a
tungstophosphoric acid (TPA) loading on SiO2
CATALYSIS REVIEWS 13

higher rate toward C-methylation. However, liquid phase methylation showed


lower conversion with 80.2% anisole selectivity.[4,65]
The H-MCM-22 showed almost 60.0% anisole selectivity with 2.0% phenol
conversion at 573 K but p-cresol to o-cresol ratio was much higher as com­
pared to other zeolites.[66] Higher selectivity of p-cresol was obtained over
HMCM-22 due to small pore size and high density of Bronsted acid sites.[35]
Moon et al.[67] confirmed that HMCM-22 with a narrow channels were also
responsible toward higher selective formation of p-cresol.
Kirichenko et al.[42] proposed the reaction scheme of chemical transforma­
tion of phenol during methylation and developed a kinetic model which
suggests methanol was predominantly adsorbing on the catalyst surface.
They assumed reactions R(1) and R(2) in which adsorption of both reactants
and adsorption of only methanol was considered, respectively, whereas
Churkin et al.[68] considered adsorption of individual reactants for estimating
kinetic parameters. Table 2 lists rate constants and activation energies for
various catalysts. The activation energy for backward reaction was found to be
higher for both catalytic systems and thus lower temperature was favorable for
accelerating rate of forward reactions. Zeolite provides almost 10 times faster
rate toward anisole formation and gives higher conversion at a lower tem­
perature. Also, considerable high difference in activation energy for forward
and backward reaction is observed over zeolite catalyst which is beneficial for
controlling the reaction rate for formation of anisole.
Phenolads þ Methanolads $ Anisole Rð1Þ

Phenol þ Methanolads $ Anisole Rð2Þ

Phenolads þ Methanol $ Anisole :Rð3Þ


The ratio para/ortho selectivity for cresol was much lower on amorphous
acid catalyst or wide pore zeolites such as H-mordenite, HY and HBEA.[63]
Namba et al.[55] correlated the product distribution with acidity strength of
catalyst; weaker acid sites are effective for O-alkylation, medium acid sites are
helpful for ring alkylation and secondary products are formed over strong
acidic sites. In contrast, numerous researcher reported that O-alkylation of
phenol is favored with strong acid sites while a weak acid and or strong basic
site promotes formation of methylated products.[44,56,69–71] Malshe et al.[72]
concluded that the basicity or very low acidity favors O-alkylation at lower

Table 2. Kinetic parameter for reaction R(1), R(2) and R(3) [42,68]
Catalyst Reaction Temperature kfor kback Efor Eback
K mol/(gcat.Pa.s) kJ/mol
NaX R(1) 533-673 1.4×10−8 2.4×10−8 42.5 50.5
NaX R(2) 533-673 2.2×10−8 7.7×10−8 22.4 60.5
Zeolite R(3) 573-633 2.1×10−7 3.5×10−7 42.5 127.1
Zeolite R(2) 573-633 1.6×10−7 3.6×10−7 22.4 96.3
14 P. BHONGALE ET AL.

temperatures and moderate or high acidity favors C-alkylation. The literature


data are much scattered and various reported conclusions are conflicting due
to drastic variation in behavior of catalyst activity. Many researchers consid­
ered particular catalyst properties such as nature of acidic or basic sites,
strength of acidity or basicity, pore size, surface area of catalyst. There is
need to consider all these affecting factors in multiple combinations of para­
meters and design experiments accordingly which will be useful to understand
an effect of physicochemical properties on activity and selectivities in the
reaction.
The time on stream data for all zeolites reported in literature showed loss in
its activity due to higher rate of coke formation which results in blockage of
catalyst pores. The kinetics and mechanism of coke formation and catalyst
deactivation during methylation of phenol were investigated on zeolites, silica-
alumina and tungustophosphoric acid (TPA) supported on silica. When
methanol was passed through catalytic bed at 473 K temperature, only 0.6 to
6.2% carbon content was obtained and almost 3.7 to 14.9% coke formation was
detected during methylation reaction.[65] Hence, it is concluded that lower
temperature, short residence time, lower acidity, higher Bronsted to Lewis acid
ratio favored formation of O-methylated product of phenol, anisole with
higher selectivity. The activity of zeolite for methylation of phenol is strong
function of various properties and there is no satisfactory correlation between
each property and product selectivity which is a major reason for researchers
to conduct detail parametric study for phenol methylation with methanol over
zeolite catalyst. Comparative catalytic activity data of different zeolites for
methylation of phenol with methanol is presented in Table S1. (Supporting
information document) The main drawback of zeolite is the low selectivity
obtained to a one specific methylated product, since several reaction pathways
resulted in O-methylation, C-methylation and combination of O- and
C-methylation. The comparative pros and cons for each zeolite for methyla­
tion of phenol are depicted in Table 3.

2.1.2. Metal oxides


Metal oxides are ubiquitous materials in heterogeneous catalysts which serve
as catalysts, supports, modifiers and promoters. The active center on the
surface of metal oxides arises due to electronegative oxygen atom and electro­
positive metal ion. The oxygen anion and metal cation are acting as Lewis base
and acid, respectively while the presence of hydroxyl group on the surface
exhibits Bronsted acid sites.[73] These properties of metal oxides are helpful to
design a most desirable catalytic system. The catalytic activity of metal oxide
can be enhanced by reducing their particle size to nano scale which resulted in
improved catalytic activity per unit mass. The use of supports such as silica,
alumina, carbon, zeolites while synthesizing metal oxides are beneficial as it
provides high dispersion, synergistic activation of substrates and thermal
Table 3. Pros and cons over zeolite catalysts for phenol methylation using methanol
Zeolite type Pros Cons
HBEA ● Highest initial phenol conversion ● Higher rate for disprortionation and dialkylation reactions which leads poor stability of anisole
● Rapid declination in activity with time on stream
HZSM-5 ● Narrow channels hampered the formation of bulky intermediates ● Least active as compared to HBEA and HY in spite of their strong acidity
● Quite unstable under reaction conditions
HMCM-22 ● Methylation rate is controlled by intracrystalline diffusion due to small ● Significant coke formation due to presence of strong Bronsted acid sites and thus progressive
pore size loss in activity with time
HY ● Selective formation of p-cresol ● Selective site blocking (higher deactivation rate) due to presence of large cages
HX, NaX, Ce- ● Lower loss of methanol over HX zeolite into side reactions among all ● Ce-NaX (Ce loading led to increase in acidity) is not suitable as it resulted in decrease in turnover
NaX zeolites frequency
● Sodium exchange zeolite (increase in basicity) shows higher rate
towards O-methylation
Mo-NaY ● Increase in anisole selectivity with Mo loading ● Decrease in phenol conversion with increase in Mo loading
Al-MCM41 ● High selectivity to o-cresol ● Buildup of a layer of pseudographic carbon which resulted in rapid deactivation
TPA/SiO2 ● Selective formation of anisole ● Much higher deactivation rate and coke formation
CATALYSIS REVIEWS
15
16 P. BHONGALE ET AL.

conductivity.[74] The basicity of metal oxides can be increased by impregna­


tion of various alkali metals such as Li, Na, K and Cs.[23] Catalytic properties of
metal oxides were determined using XRD, BET surface area, ammonia and
carbon dioxide TPD, FTIR and SEM analytical methods.
Searle patented a process for preparing anisole in vapor phase with a solid
dehydrating catalyst, named alumina at temperature range of 473 to 673 K.[75]
Sharp and Dean[76] reported vapor phase methylation of phenol over titanium
dioxide and also claimed that titanium dioxide catalyst is showing stable
catalytic activity for prolonged period, 300 h. Another solid dehydrating
catalyst, thoria was found to be highly selective for formation of methylphenyl
ether, anisole (86% selectivity) but its radioactive nature makes it
unattractive.[77]
Alumina is one of widely used support for synthesizing various catalysts and
it was found that the alumina is more active in O-methylation of phenol as
compared to HZSM-5.[75,78] The presence of strong Lewis acid sites on γ-
alumina orients phenol molecules vertically which prevents direct interaction
of the aromatic ring. This leads to lowering of deactivation rate and enhanced
formation of alkylated products by decreasing selectivity for ether products
with increase in temperature and contact time.[43] Tleimat-Manzalji et al.[78]
investigated the influence of nature of porosity of catalyst on reaction rate.
They synthesized two types of alumina by sol-gel method; one with micro­
porous, amorphous nature named xerogel and second having macroporous,
poorly crystallized nature named aerogel. Xerogel resulted in only 1.7% phenol
conversion with 51.0% anisole selectivity while aerogel showed 11.0% conver­
sion with 40.0% selectivity of anisole at 523 K. Authors found no significant
influence on the reaction rate. Langmuir-Hinshelwood reaction mechanism
was considered to estimate kinetic parameters by Tleimat-Manzalji et al.[78]
which showed higher adsorption constant for methanol as compared to
phenol. The deactivation rate was found to be much higher over silica-
alumina than alumina due to presence of Bronsted acidic sites and thus
mixed product stream was observed initially which diverted toward ortho
product formation with time due to deactivation of catalyst.[43]
Gandhe et al.[79] studied alkylation of phenol using a highly active anatase
titanium oxide, but it resulted in mainly ring alkylated products. Various metal
oxides such as V2O5, TiO2 (anatase and rutile), SiO2, MgO, CeO2 and its
combinations were explored for methylation of phenol into ortho-methylated
products by researchers.[71,80] Cerium oxide showed 13.0% conversion of
phenol with selective formation of o-cresol (91.9%) at 773 K under flow
condition whereas batch liquid phase phenol methylation over titanium
oxide exhibited higher selectivity toward anisole (28.0%) with 51.0% selectivity
of o-cresol at 573 K temperature.[79,80] The SiO2 catalyst showed only 3–4%
conversion of phenol with 6.9% anisole selectivity at 673 K temperature but
activity of SiO2 can be increased toward O-alkylation producing arylalkyl
Table 4. Pros and cons of metal oxides for phenol methylation using methanol
Metal oxide Pros Cons
Thoria ● Selective formation of anisole ● Radioactive in nature
Cs-SiO2 ● Increase in phenol conversion and decrease in deactivation rate ● Catalyst stability needs to be tested over time on stream
Borate ● Catalyst stable up to 150 h ● Only 1-2% formation of anisole
zirconia ● Higher rate for C-methylation
TiO2 ● Much better conversion of phenol as compared to other metal ● All competitive reactions are occurred and thus low selectivity of specific product can be
oxides achieved
CATALYSIS REVIEWS
17
18 P. BHONGALE ET AL.

ethers by impregnation of Li, Na, K and Cs. The conversion of phenol was
improved by increasing basicity of catalyst which also decreases the deactiva­
tion rate. Phenol molecules may get adsorb over the acidic catalyst by inter­
action of its π–electron cloud, while over basic catalyst it may get adsorb by –
OH group.[23] Two different zinc aluminate spinels prepared from aluminum
isopropoxide (Cat-A) and basic aluminum nitrate (Cat-B) were explored for
methylation of phenol which showed 17.2% conversion at 613 K and 58.0% at
543 K with approximate 30.0% selectivity of anisole.[81] It was found that
a drastic change in reactivity of both zinc aluminates (Phenol starts getting
consumed at 603 K and 488 K over Cat-A and Cat-B, respectively) was
observed due to variation in physicochemical properties of these catalysts.
Cat-A has lower surface area (67 m2/g), absence of acidic sites, presence of zinc
aluminate spinel only and Cat-B has higher surface area (123 m2/g), presence
of Lewis acidic sites and presence of zinc aluminate spinel with alumina
phase.[81]
Ballarini et al.[64] reported the transformation of methanol into formalde­
hyde, methylformate and decomposition products like CO, CO2, CH4, H2 over
MgO, Fe2O3 and Mg/Fe/O catalysts. The catalytic activity of different metal
oxides for methylation of phenol with methanol is presented in Table S2. The
pros and cons of metal oxides for phenol methylation are summarized in
Table 4.

2.1.3. Phosphates and sulfates


The sulfates and phosphates are major class of low acidic compounds which
can be used in unsupported form or supported on alumina, silica.[82] Nozaki
and Kimura[70] investigated vapor phase methylation of phenol over metal
phosphates such as Ca3(PO4)2, BPO4 and CaHPO4 at temperature ranging
from 623 to 773 K in a flow reactor and found excellent activity for ortho
methylation. Among these metal phosphate BPO4 showed higher anisole
selectivity (25.4%) but conversion of methanol is 95.8% at 733 K.
Eskinazi[83] claimed the use of numerous sulfates and phospates, La2

Table 5. Pros and cons for phenol methylation using methanol over phosphates and sulfates
Phosphates and Pros Cons
sulfates
La2(HPO4)3, ● Selective formation of anisole ● Lower activity for phenol methylation
BaSO4 and ● Liquid phase is favorable for
SrSO4 higher yield of anisole
Cs-Sm- ● Much higher phenol conversion ● Chances of precipitation formations when both
phosphate towards anisole formation metal and phosphorus precursors are added
● Long life cycle with no loss in ● Difficult to control specific morphologies
catalytic activity
Ca3(PO4)2 ● Selective formation of ● Progressive loss in catalytic activity due to deposition
C-methylated product of carboneous substance
K2SO4 on γ- ● High selectivity of anisole ● Low phenol conversion
alumina ● Coke formation
● Catalyst stability test was not performed
CATALYSIS REVIEWS 19

(HPO4)3, Sr(HSO4)2 and Ba(HSO4)2 in either gas phase or liquid phase


methylation of phenol to selectively alkyl aryl ethers. The activity of phenol
methylation was found to be lower with higher selectivity of anisole over
pure sulfates and phosphates as compared to alumina and zeolites.
Mossman[84–86] claimed that catalysts with promoters, comprising sulfated
oxides of metals from the Groups VIIB (Mn, Re), IIB (Zn, Cd, Hg) and alkali
metals (K, Li, Na), on different supports such as basic oxides, alumina, silica,
carbon, improves O-methylation rate in methylation of phenol. Pierantozzi
and Nordquist[87] studied the alkylation of phenol with methanol over La2
(HPO4)3, BaSO4 and SrSO4 catalysts in liquid phase batch reactor and vapor
phase flow reactor. Batch liquid phase phenol methylation over BaSO4
possessed higher reactivity toward anisole formation with 98.9% selectivity
and 30.3% phenol conversion whereas gas phase showed only 7.1% conver­
sion and 90.9% selectivity at 573 K. Authors reported that due to lower
acidity BaSO4 gave significantly lower reaction rate than that of La2(HPO4)3,
but lead to higher selectivity to anisole. Devi et al.[36] improved an activity of
La, Ce, Sm and Sb phosphate catalysts by impregnation of Cs atoms which
acts as promoter. The Sm phosphate catalyst showed 80% conversion of
phenol into 85% selective anisole at 623 K and addition of Cs promoter
enhanced anisole selectivity (93%) by suppressing ring alkylation. The
increase in the rate of O-methylation was due to much higher decrement
in acidity of Sm phosphate after incorporation with Cs promoter. Another
significant effect of impregnation was observed through a time on stream
data which showed stability of Cs-Sm phosphate without loss in its activity
up to 8 h time.[36]
Two new classes of molecular sieves named Aluminophosphate (AIPO) and
Silicoaluminophosphate (SAPO) were explored for vapor phase methylation
of phenol in the temperature region 523–573 K by Durgakumari et al.[88] The
AIPO catalyst has a large pore size with few Bronsted sites and the majority of
acid site is associated with Lewis centers and thus higher ring alkylation rate
was observed over AIPO. The SAPO catalyst gave 98% selectivity of anisole
with maximum 1–2% conversion of phenol.
Samolada et al.[44] reported a substantial increment in selectivity of anisole
from 68.5 to 90.8% by incorporating a less acidic potassium sulfate on γ-
alumina. The aluminum phosphate loaded on γ-alumina (AlPO4-Al2O3)
exhibited 65% selectivity of anisole and further F – and SO4 – anion treated

Table 6. Pros and cons for phenol methylation using methanol over other solid catalysts
Other solid Pros Cons
catalysts
Nafion-H ● Higher selectivity of anisole at lower temperature
● Higher deactivation rate
● Thermal stability up to 473 K
Hydrotalcites ● Mg-Al and Zn-Al favored O-methylation at lower ● Poor kinetics and stability of
temperature anisole
20 P. BHONGALE ET AL.

with AlPO4-Al2O3 showed declination in catalytic activity due to higher acidic


and less basic nature of catalyst.[46] The higher selectivity of anisole (96.3%)
over F – anion treated with AlPO4-Al2O3 (higher acidic than SO4 – treated)
and selective formation of xylenols and trimethylphenols over SO4 – anion
treated with AlPO4-Al2O3 was noted. The kinetic runs were conducted by
Bautista et al.[46] in absence of internal and external mass transfer resistance
and in order to derive kinetic analysis, boundary layer disturbance was
neglected. The apparent rate constants, reaction rate constant, k and adsorp­
tion constant K were estimated by plotting ln(1/(1-Xp)) versus F−1 where Xp is
conversion of phenol and F is flowrate of the carrier gas. The activation energy
and pre-exponential factor were obtained through a linear plot of lnkK versus
T−1, while the activation enthalpy and activation entropy were estimated
through Eyring equation from plot of lnkK/T versus T−1. The activation
energy showed an increasing trend with anion loading due to decrease in
conversion of phenol. While developing a kinetic model, author considered
only conversion of phenol and no competitive reactions were taken into
account.[46] An effect of various sulfates and phosphates on conversion of
phenol and their product distribution for methylation of phenol with metha­
nol are presented in Table S3. Table 5 indicates pros and cons for various
phosphates and sulfates reported for phenol methylation with methanol.

2.1.4. Other solid catalysts


Other solid catalysts include ion exchange resins and acid loaded on different
supports. The strong Bronsted acidic ion exchange resin, Nafion-H was
explored for alkylation of phenol which resulted in higher selectivity of anisole
(88.8%) with slight formation of methylanisole and cresol at 473 K
temperature.[43] The main drawback of using Nafion-H as catalyst was higher
deactivation rate due to coke formation over strong Bronsted acidic sites and
also there was a limitation for temperature for Nafion-H due to its thermal
stability up to 473 K temperature. The phosphoric acid supported on
Kieselguhr showed 90.5% anisole selectivity with 7.5% conversion only at

Figure 6. Reaction mechanism for methoxycarbonylation and methylation using dimethyl


carbonate
CATALYSIS REVIEWS 21

Figure 7. Reaction network for reaction between phenol and dimethyl carbonate

498 K temperature but reaction was directed toward ring alkylation at higher
reaction temperature (573 K) due to presence of weak and strong Bronsted
acid.[43] A series of hydrotalcite (HT) M(II)Al; where M(II) is Mg, Mn, Co, Ni,
Cu, Zn were tested in vapor phase methylation of phenol by Velu and
Swamy.[89] Among these Mg-Al HT (100.0% anisole selectivity) and Zn-Al
HT favored predominantly O-methylated products at low temperature
(523 K), whereas Cu, Ni, Co, Mn-Al HT exclusively formed C-alkylated
products. In order to investigate kinetic parameters, pseudo-first order differ­
ential equation in a plug flow reactor was assumed, where only conversion of
phenol was taken into consideration and no product distribution was
22 P. BHONGALE ET AL.

considered. The kinetic analysis showed rate of phenol conversion was first
order with respect to phenol concentration and zero order with respect to
methanol concentration. A linear free energy relationship represented as
Ea = a + b lnA (where a and b are constants called compensation parameters)
was observed between activation energy and frequency factor and it revealed
a best fitted correlation with the experimental data. The Mg-Al HT showed
lowest activation energy (36.4 kJ/mol) and higher rate constant (8.9 × 10−4
mol/s.kg).[89]
Sad et al.[35] found 20.0% phenol conversion with 65.0% anisole selectivity
over tungustophosphoric acid (TPA) supported on silica catalyst due to the
presence of lower total acidity with higher Bronsted to Lewis acid type
distribution as compared to zeolites and silica-alumina catalyst. But higher
deactivation rate was also reported by authors as listed in the Table 1.[65]
Comparative catalytic activity data of other solid catalysts for methylation of
phenol with methanol are presented in Table S4. The advantages and draw­
backs of other solid catalysts for methylation of phenol with methanol process
are illustrated in Table 6.

2.2. Dimethyl carbonate as methylating agent


Dimethyl carbonate is nontoxic, environmentally innocuous, biodegradable,
relatively inert under ambient conditions that can be used as green substitute
for conventional methylating agents such as dimethyl sulfate, methyl halide,
and triflate in alkylation reactions.[90] Dimethyl carbonate is a valuable green
reagent having versatile and tunable chemical reactivity.[91] Dimethyl carbo­
nate have functional groups, such as carbonyl, methyl and methoxy which
makes it an important intermediate for various organic syntheses.[92] The
unique selective behavior that follows the Hard-Soft Acid-Base theory is one
of major advantage of dimethyl carbonate as depicted in the Figure 6.[93] The
high selectivity in methylation reactions is due to the ambident electrophilic
character of dimethyl carbonate. It reacts on its hard center (carbonyl group:
due to polarized positive charge and presence of sp2 hybridization) with
harder nucleophiles and on its soft center (methyl group: saturated carbon
atom and sp3 hybridization) with softer nucleophiles.[94] Dimethyl carbonate
can also be used as methoxycarbonylating agent by nucleophilic attack at its
carbonyl carbon at lower temperature.[95] It is also used as methylating agent
at higher temperature (above 393 K) where nucleophilic attack at the methyl
carbon of dimethyl carbonate occurs.[96] The Figure 7 showed transesterifica­
tion between phenol and dimethyl carbonate which is thermodynamically
unfavorable reaction (equilibrium constant = 3 × 10−4 at 453 K).[97]
The thermodynamic behavior of transesterification process of dimethyl
carbonate with phenol and O-methylation process of phenol with dimethyl
carbonate at atmospheric pressure helps to estimate experimental process
CATALYSIS REVIEWS 23

parameters. The transesterification reaction is endothermic at temperatures


below 453 K, but it direct toward exothermic due to vaporization of phenol
above 453 K. An O-methylation of phenol with dimethyl carbonate is
endothermic reaction and thus higher temperature is desirable. The Gibbs
free energy for transesterification reaction is positive in temperature range of
298 to 523 K which showed reaction is thermodynamically unfavorable even at
high temperature.
Fuming et al.[39] studied transesterification of dimethyl carbonate with
phenol over uncalcined Mg-Al (2:1) hydrotalcite and found to yield 14.7 and
11.6% for diphenyl carbonate and monomethyl carbonate, respectively at
453 K, 1:2 mole ratio of dimethyl carbonate to phenol and 10 h of reaction
time. The strong Bronsted basic sites in double layers of uncalcined Mg-Al
hydrotalcite are responsible for higher rate toward transesterification reaction
than calcined Mg-Al hydrotalcite in which unsaturated O2- ion acts as Lewis
basic sites.[39] The reactivity of dimethyl carbonate is higher as compared to
methanol under relatively lower temperature with improved selectivity of the
desired O-alkylated products.[98] The methylation of phenol in presence of
dimethyl carbonate is depicted in Figure 7 which shows strongly dependency
on acid-base properties, pore structure of catalyst and operating parameters of
the reaction for product distribution.[44,88] The presence of basic sites over
catalyst led to formation of anisole whereas acidic sites favored C-alkylated
product, mainly o-cresol. Other major byproducts produced are 4-methylani­
sole and xylenols. Various catalyst such as ionic liquids, phase transfer catalyst
(PTC), basic zeolites and hydrotalcites have been explored for methylation of
phenol with dimethyl carbonate.

2.2.1. Ionic liquids


Ionic liquid (IL) as catalyst is a versatile chemical in diverse fields of the
synthetic chemistry.[99] The unique properties of ionic liquids such as having
high thermal stability, negligible vapor pressure, excellent solubility for a wide
range of substances and the potential for recycling are responsible for growing
extensive use of ILs as catalysts in various organic reactions.[100,101] The most
interesting and powerful property is their combinatorial flexibility which
allows to design specific IL by varying the anion and cation
combinations.[102] Heteroatomic cations such as imidazolium, pyridinium,
ammonium and phosphonium ions are the most widely used cations in
synthetic methods.[103–105] The chemo physical properties of ILs are depen­
dent on both the nature of the cation and the anion.[106] In addition, the length
of the alkyl chain and functionality have a striking influence on their
properties.[107] The functionalized ionic liquids are promising new materials
as they combine the advantage of a liquid (large contact area), with the
advantages of solid-state acids (negligible vapor pressure and reusability).
A huge number of ILs approximately 1018 can be designed by considering
24 P. BHONGALE ET AL.

the combination of available anions and cations.[102] The Bronsted acids are
formed in ionic liquids by introduction of a protic molecule into a liquid salt
or by reaction with an appropriate base; for instance pyridine and 1-methyli­
midazole were responsible for accelerating Bronsted acidity in ionic
liquids.[108] Recent developments of ionic liquids for O-alkylation of phenol
have been highlighted in this section. Ionic liquids were characterized by using
NMR, FTIR and thermal stability was determined by thermal gravimetric
analysis.
An excellent yield of anisole (more than 99%) was obtained over 1-N-butyl-
3-methylimidazolium hydrogen chloride (IL-1) at 393 K temperature.[109]
Shen et al.[110] synthesized several imidazolium salts having same cation
(BMIm) but different anions (Cl, Br, BF4, PF6) and examined its activity for
O-methylation of phenol with dimethyl carbonate which showed significance
of imidazolium moiety of ionic liquid. However, the amount of onium salt
used was 50 mol% of phenol. The activity of IL-1 catalyst showed no signifi­
cant loss even after fifth catalytic cycle.[110] Das and Das[111] reported pyridi­
nium based ionic liquid named 1-Butyl-4-methylpyridinuium bromide (IL-2)
with 87% yield of anisole when 1 mmol of IL was used for 1 mmol of phenol at
443 K temperature. In addition, IL-2 catalyst has shown same performance
even after third recycle run.[111] Recently Kabra et al.[40] attempted to review
the efficacy of phosphonium based ionic liquids as base catalysts. The trihexyl
(tetradecyl) phosphonium hexafluorophosphate and tetrabutyl phosphonium
bromide showed lowest activity due to presence of asymmetric cations as
compared to trihexyl (tetradecyl) phosphonium bromide (IL-3), trihexyl (tet­
radecyl) phosphonium chloride and trihexyl (tetradecyl) phosphonium decan­
oate. A complete conversion of phenol with 97% selectivity of anisole was
found at 473 K, a 1:6 mole ratio of phenol to dimethyl carbonate over IL-3
catalyst. Fresh and third reused catalyst found to result in same conversion of
phenol.[40] The efficacies of various ILs as catalysts for methylation of phenol
with dimethyl carbonate are depicted in Table S5.

2.2.2. Phase transfer catalyst and base


The phase transfer catalyst (PTC) is special form of heterogeneous catalysis
which helps to transfer active ions from one phase to another. The reaction of
phenol and dimethyl carbonate under homogenous conditions and lower
temperature preferably follow the reaction path 1 of Figure 6 with attack at
carbonilic center. In presence of PTC, the reaction mechanism has been
changed from path 1 to path 2 of Figure 6 due to anion activation by PTC in
absence of solvation. As shown in path 2, CO2 is generated which means
acidity is produced and thus use of potassium carbonate (K2CO3) as a base,
does not change the composition of solid and catalytic bed.[112] The activity of
PTC for methylation of phenol in presence of dimethyl carbonate was
reported by Lissel et al.[113] for the first time. They used K2CO3 as a base
CATALYSIS REVIEWS 25

and 18-crown-6 (crown ether) as a PTC at 373 K for 8 h under atmospheric


pressure in a batch reactor and it showed 40% yield for anisole.[113] Merger
et al.[114] patented a novel process for anisole formation using various base
such as amines, tertiary phosphine and p-dimethylaminopyridine as a catalyst.
Several pentaalkylguanidines found to be superior catalysts for preparation of
ethers through methylation of phenol and dimethyl carbonate.[115]
A continuous flow synthesis of anisole, operated under gas-liquid-PTC
conditions in presence of poly(ethylene glycol) (PEG) and K2CO3 was
reported by Tundo.[116] The K2CO3 helps to produce reactive nucleophile
and PEG acts as an anionic activator which represents aprotic environment
where reaction takes place.[117] The thermal and chemical stability of catalytic
bed was tested up to 7 days which showed no change in its efficiency. Reaction
at 453 K temperature, phenol to dimethyl carbonate mole ratio of 0.2 and
9 × 10−2 h−1 WHSV in continuous stirred tank reactor (CSTR) resulted in
complete conversion of phenol into anisole with 96–100% purity.[118,119]
Further, in order to develop a continuous process in a plug flow reactor
(PFR), PEG-6000 and potassium carbonate were physically adsorbed on α-
alumina spheres and these coated spheres were packed in a reactor to perform
the reactions. At 453 K, PEG-6000 gets melted and forms a liquid film on
alumina which adsorbs reactants and desorbs gaseous products.[112,120]
Ouk et al.[20] developed an effective process yielding 100 and 68.5% of
anisole in presence of tetrabutylammonium bromide (TBAB) and tetrabutyl­
phosphonium bromide (1 mol/mol of phenol), respectively at higher tempera­
ture (403 K) under semi-continuous mode of operation. At the reaction
temperature, ammonium quaternary salt reacts with phenol to form
a complex containing nucleophilic anion which further reacts with dimethyl
carbonate to produce anisole. In order to recover catalyst from reaction
stream, aqueous hydrochloric acid and tert-butyl methyl ether were added to
the mixture and catalyst was extracted in aqueous phase.[20] Further, Ouk
et al.[21] performed the reactions in presence of K2CO3 and tetrabutylammo­
nium bromide at reflux temperature. A reaction mixture comprising 16 mol of
dimethyl carbonate/mol of phenol, 0.2 mol of PTC/mol of phenol and
0.75 mol of K2CO3/mol of phenol gave 31 and 80% yield of anisole within 5
and 13 h reaction time, respectively. Further variation in concentration of
dimethyl carbonate (30 mol/mol of phenol) and PTC (0.6 mol/mol of phenol)

Table 7. Pros and cons for phenol methylation using dimethyl carbonate over ionic liquids and
phase transfer catalysts
Catalysts Pros Cons
IL-1, IL-2, ● Excellent yield of anisole ● Much higher catalyst loading is required
IL-3 ● Easy recycle, regeneration and reuse with no loss which is not economical feasible
in chemical activity
PTC ● Effective process operated at semi-continuous and ● Poor catalyst recovery
continuous mode ● Difficulties in product separation
● Short catalyst life-span
26 P. BHONGALE ET AL.

showed 99% yield of product within 5 h reaction time.[21] The rate of anisole
formation per mole of catalyst was significantly low for these methods and
thus Ouk et al.[21] tried to improve the known process by performing reactions
in presence of potassium carbonate alone at relatively higher temperature
(423 K) under atmospheric pressure in a continuous flow of dimethyl carbo­
nate. In order to maintain reaction temperature at 423 K under atmospheric
pressure, methanol formed during reaction and unreacted dimethyl carbonate
were progressively distilled from reaction mixture. The maximum 70% yield
was obtained at 3.5 h−1 molar hourly space velocity of anisole per mole of
catalyst.[121] The use of PTC has been widely reported for methylation of
phenol, but there is no data reported for activity of recovered PTC from the
reaction mixture.
The polarity of solvent was significantly affecting on selectivity of methyla­
tion where dimethylformamide (DMF) and triglyme showed higher rate of
methylation.[122] The 1,8-Diazabicyclo[5.4.0]undec-7-ene DBU and micro­
wave accelerated green chemistry in methylation of phenols with dimethyl
carbonate as methylating agent and DMF as solvent was studied by Shieh
et al.[123] and reported 80-fold enhancement in the rate of reaction.
Furthermore, incorporation of PTC named tetrabutylammonium iodide to
the reaction mixture speed up rate by 1900-fold. The DBU (nucleophilic
catalyst) reacts with dimethyl carbonate to produce a more activated methy­
lating agent and thus the activation energy required for the reaction was
reduced.[123] Tilstam[22] reported a continuous process in steel braided PTFE
tube reactor using GC oven as source of heating instead of microwave heating.
The DBU was used as a base in catalytic amount along with phenol and
dimethyl carbonate with molar ratio of 1:3. A complete conversion of phenol
with 95% yield of anisole was obtained at 493 K in 0.2 h residence time.[22]
Although high catalytic activity was observed over PTC, but there are
obvious disadvantages such as poor catalyst recovery and difficulties in pro­
duct separation. A summary of catalytic activity data of different PTC for
methylation of phenol with dimethyl carbonate are presented in Table S6. The
advantages and disadvantages for ionic liquid and phase transfer catalysts are
mentioned in Table 7.

2.2.3. Other solid basic catalysts


The use of ionic liquids and PTC as catalyst for methylation of phenol with
dimethyl carbonate resulted in higher rate toward formation of anisole but
these catalysts also found to have numerous drawbacks such as difficulties in
separation and short catalyst life-span. Thus there is high demand to design
other solid basic catalytic system which helps to overcome these problems
associated with homogeneous catalysts. Other solid catalysts reported in the
literature such as zeolites, metal oxides and hydrotalcites for phenol alkylation
with DMC are summarized here.
CATALYSIS REVIEWS 27

Table 8. Comparative pros and cons for phenol methylation using dimethyl carbonate over solid
basic catalysts
Other solid Pros Cons
basic
catalysts
Basic zeolites ● NaX and KNaX zeolite showed 93% and ● Use of additional solvent which is toxic in
85% selectivity nature
● Higher phenol conversion
● Better controlled reactivity
Hydrotalcites ● High selectivity (100% for Fluorine mod­ ● Use of large amount of salt, alkaline solution
ified MgAl) and water during washing process
● Catalyst stability with marginal loss in its ● High production cost and low raw material
activity utilization rate
Metal oxides ● Higher rate of phenol consumption with ● Use of additional solvent
selective formation of anisole ● Poor catalyst stability
● Low turnover number

2.2.3.1. Basic zeolites. Fu and Ono[124] reported methylation over various


basic zeolites such as Li, K, Na, Cs, Ca, Mg- X and Y zeolites. The highest
basic strength of CsX zeolite showed lower activity because steric hindrance
due to the high density of the large cations in the supercages in CsX.[125] The
higher stability of NaX resulted in 93% selectivity and 76% yield of anisole at
553 K.[124] Further, spectroscopic analysis was studied by Beutel[126] in order
to examine the influence of acid-case sites in NaX zeolite through FTIR, UV
and mass spectroscopy. The FTIR analysis showed phenol adsorption on
external surface of NaX by hydrogen bonding to basic oxygen atom at room
temperature. Further, heating to 473 K temperature showed partial deproto­
nation of phenol over basic sites to form zeolitic hydroxyl groups and pheno­
late ion, NaOPh and generate surface ether. The reaction mechanism
proposed suggest dimethyl carbonate activation on Lewis acid site by its
carbonyl oxygen and phenol on an adjacent Lewis base site by H- bonding.
Process kinetic measurements have been studied by conducting experiments
in a flow reactor which showed strong dependency of reaction rate on partial
pressure of each reactant. Dimethyl carbonate and phenol showed +1 and −1
reaction orders, respectively which signifies strong adsorption of phenol on
catalyst surface which blocks sites for the adsorption of dimethyl
carbonate.[126] Additionally, Romero et al.[47] investigated O-methylation of
phenol with dimethyl carbonate as methylating agent and dimethyl sulfoxide
as a solvent in presence of KNaX zeolite impregnated with alkaline metal
hydroxides. They found basic zeolites improve anisole formation because of
the carbanion formed by abstracting a hydrogen atom from basic site. It
showed better controlled reactivity than carbocation formed in an acidic site.
A reaction at 438 K temperature with dimethyl carbonate to phenol mole ratio
2 resulted in complete consumption of phenol into 85% anisole, 15% cresol
and 4-methylanisole over KNaX zeolite.[47]
28 P. BHONGALE ET AL.

2.2.3.2. Hydrotalcites. Reactions in the temperature range of 513 to 548 K


over calcined Mg-Al (3:1) hydrotalcite reported by Jyothi et al.[48] to give
anisole as major product along with o-cresol and 2,6-xylenol as by-products.
The basicity of Mg-Al hydrotalcite was increased by incorporating Fluorine
which reported to have linear correlation with the conversion of phenol. The
Mg-Al (3:1) catalyst gave 59.6% conversion of phenol and 87.0% selectivity of
anisole with remaining o-cresol and methylphenyl carbonate, while the fluor­
ine modified mesoporous Mg-Al mixed oxide showed 99.3% conversion with
single product anisole at 473 K temperature within 8 h reaction time. Fluorine
modified Mg-Al hydrotalcite having 155.8 μmol/g CO2 uptake with 76.6%
strong sites exhibited much higher catalytic activity than Mg-Al hydrotalcite
having 91.1 μmol/g CO2 uptake with 75.5% strong sites.[39] Further
Subramanian et al.[127] synthesized sixteen different amino acids which are
intercalated into Mg-Al layered double hydroxide and was investigated for
chemoselective synthesis of anisole. Higher activity toward O-methylation was
attained due to physical adsorption of amino acid over layered double hydro­
xide and presence of large hydrophobic pockets in amino acid. A plausible
reaction mechanism was proposed in detail for Leucine Mg-Al layered double
hydroxides (Leu-LDH) catalyst which showed marginal drop in its activity
after fourth cycle.[127]

2.2.3.3. Metal oxides. Porous silica exhibited no methylation activity toward


phenol but when it is treated with primary, secondary and tertiary amine, it
showed substantial increment in its activity. Xue-Hong et al.[128] reported
a drop in phenol conversion from 99.8 to 13.8% after fifth recycle run with
100% selectivity of anisole over porous silica modified with primary amine.
Chen et al.[129] studied spectroscopic analysis of methylation of phenol with
dimethyl carbonate over KBr/SiO2 using in-situ FTIR at 473 K. Silica revealed
no activity toward methylation, whereas KBr and KBr/SiO2 resulted in 94%
conversion of phenol and 94% selectivity of anisole at 473 K and within 8 h
reaction time.[129,130] An activity of MgO catalyst was examined using
dimethylformamide as a solvent and microwave heat treatment at 443 K for
0.5 h time. Anisole yield reported was 94% when catalyst is 4 equivalent of
substrate.[131] Manganese, tungsten and cobalt carbonyls have also been
explored for methylation of substituted phenols with dimethyl carbonate
and 98% yield of anisole was reported over manganese carbonyl at 453 K
temperature.[132] Comparative catalytic activity data of different solid basic
catalysts for methylation of phenol with dimethyl carbonate are presented in
Table S7. The concluding remark for the use of solid basic catalysts for
methylation of phenol with dimethyl carbonate is summarized in terms of
pros and cons and depicted in Table 8.
CATALYSIS REVIEWS 29

Figure 8. Reaction pathways for methylation of hydroquinone with methanol at reflux and above
473 K temperature

3. O-alkylation of hydroquinone
O-alkylation of hydroquinone can be carried out using various alkylating
agents such as methanol, dimethyl carbonate, dimethyl ether, and dialkyl
sulfate and alkyl chloride. Joseph et al.[28] studied an effect of various
solvents such as trichloroethylene, benzene, and nitrobenzene, anisole for
alkylation of dihydric phenols (hydroquinone, catechol and resorcinol) in
30 P. BHONGALE ET AL.

Table 9. An effect of different acid on product distribution [8]


Reaction mixture and conditions Acid Yield (%)
4-methoxyphenol 1,4-dimethoxy benzene Hydroquinone
Hydroquinone+Methanol + No 52 21 27
trimethylamine Acetic acid 67 23 10
T: 483 K; Time: 6 h; Reactor: Hydrogen 45 52 3
Autoclave chloride
Sulfuric acid 71 29 0

presence of dialkyl sulfate. It was observed that hydroquinone was the


most difficult to alkylate to form monoalkyl ether and thus ratio of mono
to di-alkylated product of hydroquinone was lower as compared to other
dihydric phenols. Gradeff and Bertrand[3] reported monoalkylation of
hydroquinone with alkyl chloride in presence of sodium or potassium
carbonate or bicarbonate as a condensing agent. In this process, rate of
dialkylation reaction was found to be higher than monoalkylation and thus
the maximum 32% purity of monoalkyl ether of hydroquinone was
achieved. Amongst various alkylating agents, methanol and dimethyl car­
bonate reported to have higher rate of mono-O-alkylation and hence, the
present review is limited to only these two alkylating agents.[3] A various
routes for preparation of alkoxyphenols have been reported in number of
patents.[133–135] However, limited work is reported on selective mono
O-methylation of hydroquinone using acidic and basic catalysts.

3.1. Methanol as methylating agent

The liquid phase and vapor phase methylation of hydroquinone with methanol
was studied in both, batch and continuous mode of operation. There are two
different paths of conversion of hydroquinone into 4-methoxyphenol; one is
direct conversion of hydroquinone at higher reaction temperature (above
473 K) and second is an oxidation of hydroquinone to p-benzoquinone fol­
lowed by conversion to p-alkoxyphenol at reflux temperature. The possible
reaction pathways at reflux and higher (above 473 K) temperature are men­
tioned in Figure 8. The 4-methoxyphenol, 2,4-dimethoxyphenol and 1,2,4-tri­
methoxybenzene are the products formed when reactions were carried out at
reflux temperature and in presence of catalytic amount of p-benzoquinone.
However, 4-methoxyphenol, 1,4-dimethoxybenzene and 2-methylhydroqui­
none were obtained at high temperature and high pressure reaction conditions.
The methylation of hydroquinone with methanol in presence of various
heterogeneous catalysts such as alkali loaded silica and zeolites, alumina, ion
exchange resins, acid loaded clays and mineral acids such as sulfuric acid,
perchloric acid and hydrohalic acid has been explored
[9,29,33,136]
experimentally.
CATALYSIS REVIEWS 31

Grote[137] claimed a continuous process for etherification of water soluble


hydroquinone using methanol or dimethyl ether in a fixed bed reactor over
silica and alumina as catalyst in temperature range of 505–645 K. Author
reported maximum 70% yield of 4-methoxyphenol along with 1,4-dimethox­
ybenzene as byproduct. Etherification of phenolic compounds with methanol
in presence of tertiary amines or corresponding salts at higher temperature
was studied by McCloud et al.[8] and also the influence of different acid such as
sulfuric acid, acetic acid and hydrochloric acid was examined on yield of
hydroquinone mono and di-methyl ether and results for the same are sum­
marized in Table 9. The tertiary amine was also found to act as a catalyst as in
the absence of acid giving 52% yield of 4-methoxyphenol. The relation
between the conversion of hydroquinone and acidity was found to be directly
proportional. Cupric chloride, anhydrous iron trichloride, ferric sulfate, sul­
fonated styrene resin and its combination were explored as catalysts for
developing a process for monoalkylethers of hydroquinone.[136] Correale et ­
al.[9] developed a novel process in which in situ benzoquinone was produced
via oxidation of hydroquinone in presence of hydrogen peroxide which led to
higher yield for 4-methoxyphenol.
Bal et al.[15] investigated vapor phase O-methylation of hydroquinone with
methanol over alkali loaded fumed silica in a fixed bed reactor in the tem­
perature range of 573–673 K. The Cs loaded fumed silica resulted in 100%
selectivity toward O-methylated product, but with higher selectivity of
1,4-dimethoxybenzene. Due to differences in mesomeric effects of – OCH3
and – OH group, activity of mono-methoxyphenol was higher than corre­
sponding hydroxybenzene.[15] The less basic catalyst would expect adsorption
of hydroquinone by ring electron leading to C-alkylation products and thus
higher basicity suppresses attack of carbenium ion at carbon atom and direct
reaction toward O-methylation. It showed O-methylation of both OH group
and thus higher selectivity of 1,4-dimethoxybenzene was observed. Bal et al.[13]
also developed a kinetic model by assuming pseudo-first order kinetics.
A liquid phase O-alkylation of hydroquinone with methanol was performed
by using alkali metal (Na, K, Rb, Cs) ion-exchanged zeolite catalysts in a high
pressure batch reactor. The strong basicity and shape selectivity of Cs-NaX
zeolite resulted in 84% selectivity of 4-methoxyphenol and 95% conversion of
hydroquinone at 513 K temperature and 16 h reaction time. The mole ratio of
methanol to hydroquinone was 1000 which was much higher and catalyst was
used in three times as that of substrate. Higher temperature (above 513 K)
resulted in increase in rate of formation of 2-methylhydroquinone and
1,4-dimethoxybenzene.[33]
Further an extended research was carried out to study an effect of phase
transition of methanol on the reaction rate by Park et al.[138] An increase in
mole ratio of methanol to hydroquinone from 300 to 1000 showed a drastic
reduction in the conversion of hydroquinone from 93 to 10%. The conversion
32 P. BHONGALE ET AL.

Figure 9. Methylation of hydroquinone with dimethyl carbonate

was found to be inversely proportional to reaction pressure and thus faster


reaction rate in gas phase was observed than that in liquid-vapor equilibrium
phase which was due to the faster diffusion rate of reactants in the gas
phase.[138]
O-methylation of hydroquinone with methanol in presence of p-benzoqui­
none and acid catalyst showed higher selectivity toward mono-O-methylated
product, 4-methoxyphenol at mild reaction conditions. Yadav et al.[32]
explored various heteropolyacids supported on clay for synthesizing 4-meth­
oxyphenol in isolated yield. The 40% tungustophosphoric acid/montmorillo­
nite clay (TPA/K10) was found to be most effective catalyst which gave 100%
selectivity for 4-methoxyphenol when benzoquinone was used as co-catalyst at
483 K temperature. The reaction does not proceed in the absence of benzo­
quinone as confirmed experimentally and through proposed mechanism by
Yadav et al. .[32] Reusability of catalyst was tested up to third recycle run and it
showed reduction in conversion from 39 to 33%. An ionic reaction mechanism
was suggested where protonated benzoquinone couples with protonated
methanol on a catalyst site to yield an intermediate, which further reacts
with protonated hydroquinone and results in 4-methoxyphenol. Yadav
CATALYSIS REVIEWS 33

et al.[32] also confirmed absence of dimethylated product, 1,4-dimethoxyben­


zene when reaction was performed in presence of benzoquinone which might
be due to presence of excess of methanol. Further, Gambarotti et al.[29] studied
selective monoalkylation of hydroquinone promoted by sodium nitrite
(NaNO2) which acts as oxidizing agent instead of benzoquinone. A plausible
reaction mechanism was proposed which suggest formation of semiquinone
from hydroquinone by oxidation in presence of NaNO2. The 4-methoxyphe­
nol is less acidic than semiquinone radical which is responsible for inhibition
of alkylation of second – OH group. Sulfuric acid and acidic ion exchange
resins were explored as a catalyst at room and reflux temperatures. If benzo­
quinone was used instead of catalytic amount NaNO2, there is formation of
2,4-dimethoxyphenol by Michael substitution.[29] No data was reported for
consumption of benzoquinone even if it was used in catalytic amount at reflux
temperature or at high temperature.[29,32] However, Mondal et al.[139] reported
conversion of p-benzoquinone to 4-methoxyphenol with methanol using
amberlyst-15 catalyst at reflux temperature with very good yield (72%).
Authors also suggested a plausible reaction mechanistic aspect of conversion
of benzoquinone in presence of excess of methanol which showed a reducing
property of alcohol. Various – SO3H functionalized ionic liquids were inves­
tigated for O-methylation of hydroquinone in presence of benzoquinone at
reflux temperature.[140] The highly acidic ionic liquid named 1,3-Disulfonic
acid imidazolium hydrogen sulfate (IL-4) resulted in single product formation
with 93.79% yield of 4-methoxyphenol. The higher activity of protonated
benzoquinone (pKa-4) toward alkylation than 4-methoxyphenol (pKa-10.21)
and higher dissociation energy of O-H bond in 4-methoxyphenol (349 kJ/mol)
than protonated benzoquinone (226 kJ/mol) resulted in no methylation
of second – OH group in hydroquinone at reflux temperature.[29,141,142]
A detail reaction mechanism for O-methylation of hydroquinone in pre­
sence of benzoquinone at reflux temperature was reported by Bhongale et al.­
[140]
and it showed multiple parallel and series reactions and also involved
regeneration of some intermediates. Therefore, it is difficult and challenging
task to study reaction kinetics for this system by considering detailed
mechanism.

3.2. Dimethyl carbonate as methylating agent

Relatively limited work has been published on mono-O-methylation of hydro­


quinone with dimethyl carbonate due to higher selective formation of desired
product using methanol as methylating agent at reflux temperature in pre­
sence of p-benzoquinone and at higher temperature in range of 473–513 K.
Possible products formation in presence of dimethyl carbonate as methylating
agent are shown in the Figure 9.
34 P. BHONGALE ET AL.

Vijayaraj and Gopinath[143] studied an influence of alkali metal ion loaded


MgO on methylation of hydroquinone with dimethyl carbonate as methylat­
ing agent and methanol as solvent in a flow reactor at 583 K temperature. It
was confirmed that methanol has no methylation activity over K-MgO cata­
lyst. Lower dimethyl carbonate to hydroquinone mole ratio (2:1) showed
higher selectivity of 4-methoxyphenol (95%) but at 4:1 mole ratio, higher
dimethyl carbonate content accelerated the rate of di methylation. Also sig­
nificant loss of dimethyl carbonate in side reactions such as dissociation and
gasification was observed.[143] Khusnutdinov et al.[132] synthesized aryl methyl
ether by reaction of hydroquinone with dimethyl carbonate in presence of
manganese, tungsten and cobalt carbonyls. Manganese carbonyl showed 100%
conversion of hydroquinone with formation of 1,4-dimethoxybenzene and
4-methoxyphenol in molar ratio of 1:0.8, whereas tungsten carbonyl resulted
in 72% conversion with 1:3.5 mole ratio of di to mono methylated product at
453 K temperature with 1 h reaction time.[132]
Lui et al.[144] The 1,8-Diazabicyclo[5.4.0]undec-7-ene DBU as a base cata­
lyst for conducting reaction of hydroquinone with dimethyl carbonate in
a microwave reactor at 443 K temperature. A complete conversion of hydro­
quinone into bis-methylated product, 1,4-dimethoxybenzene with 96.1%
selectivity was observed when the substrate to DBU ratio used was 1:2.[144]
Merger et al.[114] patented preparation of mono methyl ether of hydroquinone
by reacting hydroquinone with dimethyl carbonate in presence of pyrrolodi­
nopyridine and triphenylphosphine as catalyst. Tundo et al.[97] tested methy­
lation of hydroquinone with dimethyl carbonate over solid bed composed of
PEG 6000 and potassium carbonate which showed 98.2% conversion and
77.9% yield of 4-methoxyphenol within 13 s of contact time.

4. Challenges in process development


At the end, this section involves the challenges occurred while studying an
O-alkylation of phenol and hydroquinone. Various catalysis systems have been
reported for production of anisole using methanol and dimethyl carbonate as
methylating agents. However, in case of methanol, catalyst deactivation rate
was much higher and in case of dimethyl carbonate carbon dioxide was
evolved which is a greenhouse gas. Another disadvantage in case of using
dimethyl carbonate is formation of minimum boiling azeotrope between
methanol formed during reaction and unreacted dimethyl carbonate which
creates separation problems. Therefore, there is high demand to develop a new
catalytic system which has higher life cycle and higher rate toward
O-methylation phenol. The review will be helpful for understanding properties
of catalyst which can be manipulated through various methods. However, in
case of selective mono O-methylation of hydroquinone, it is difficult to control
alkylation of one of – OH group and very few studies have been conducted for
CATALYSIS REVIEWS 35

the same. Thus, there is scope for designing new catalytic system for
O-methylation of hydroquinone and explore its effect along with reaction
parameters on product distribution. It is interesting to study an influence of
the catalytic properties such as nature of catalyst sites, strength of acidity or
basicity, physical properties like surface area, pore size on product distribution
which helps to develop a detail reaction mechanism at higher reaction
temperatures.

5. Conclusions
O-methylation of phenol with two different methylating agents, methanol and
dimethyl carbonate produced various industrial useful chemicals such as
anisole, cresols, methylanisole and xylenols. The activity and selectivity for
O-methylation of dimethyl carbonate was much higher as compared to metha­
nol in case of phenol O-methylation to anisole. The phase transfer catalysts
and ionic liquids showed formation of anisole as isolated product whereas
other basic catalyst yielded byproducts such as o-cresol, 4-methylanisole and
xylenol in minor selectivity. The mechanistic aspects suggested by various
researcher helps to design a catalytic system by taking into consideration the
reaction network. In case of methanol as methylating agent, major limitations
are faced such as numerous side reactions, such as anisole dealkylation and
disproportionation, isomerization of cresols, decomposition of methanol;
higher rate of catalyst deactivation; lower atom economy. Thus, dimethyl
carbonate was found to be highly efficient in case of O-methylation of phenol
to anisole. The spectroscopic analysis and reaction kinetics are found useful
for designing a new catalytic and reactor system, respectively.
The main conclusions that can be drawn based on catalyst synthesis, its
characterization, reaction parameter optimization, reaction kinetics and
mechanism and catalyst deactivation rate for methylation of phenol process
are summarized as follows:

● Among all catalytic systems used for methylation of phenol with metha­
nol, Ce-Sm phosphate synthesized through wet impregnation method was
found to be best suited for achieving better catalytic performance toward
formation of anisole.
● However, NaX and KNaX basic zeolites showed higher activity for
O-methylation of phenol using dimethyl carbonate.
● Deactivation rate and methanol loss through side reactions was found to
affect significantly on catalytic activity which leads to poisoning of active
sites by carbonaceous residues. Even if poisoning resulted in decrease in
phenol conversion but possible increase in anisole selectivity was found to
increase.
36 P. BHONGALE ET AL.

● Lower reaction temperature, short residence time, lower acidity, higher


Bronsted to Lewis acid ratio and high methanol to phenol mole ratio
favored formation of anisole with higher selectivity.
● Since methylation of phenol with methanol involves various competitive
reactions such as O-alkylation, C-alkylation, isomerization, disproporta­
tion, dialkylation and methanol decomposition, it is difficult to control
reactivity and getting a specific product with maximum yield.

However, the performance of methanol was distinguished over dimethyl


carbonate in case of selective mono O-methylation of hydroquinone. From the
research available in the literature, two different routes for synthesizing
4-methoxyphenol were analyzed in the article. Reaction at mild conditions
in presence of benzoquinone resulted in selectively desired product, while
dimethylated and C-methylated product of hydroquinone were formed at
higher reaction temperatures. The plausible reaction mechanism at different
reaction conditions, nature of catalyst along with reported kinetic models is
discussed in detail which is necessary in designing of an industrial reactor
system. A complete understanding of catalysis, reaction kinetics and mechan­
istic aspects for O-methylation of phenol and hydroquinone using methanol
and dimethyl carbonate as methylating agent is also reviewed critically. For
designing a catalyst of desired properties for selective product formation, it is
necessary to correlate catalytic function, their chemical and physical proper­
ties, flow and mechanical properties. These correlations will be helpful to
design a shape selective catalyst for a desired product. A more active catalyst
can be operated at milder conditions which indirectly minimize the required
volume of reactor and cost. The arrangement of rapid and convenient regen­
eration of catalyst is necessary while designing a reactor system in
O-methylation of phenol with methanol where the catalyst deactivation rate
was much higher.

Nomenclature
AIPO: Aluminophosphate
AlPO4: Aluminium phosphate
B: Bronsted
Ba(HSO4)2: Barium hydrogen sulfate
BaSO4: Barium sulfate
BPO4: Boron phosphate
CaHPO4: Calcium hydrogen phosphate
Ca3(PO4)2: Calcium phosphate
Cat-A: Zinc aluminate spinels prepared from aluminium isopropoxide
Cat-B: Zinc aluminate spinels prepared from basic aluminium nitrate
Ce: Cerium
Cs: Cesium
CATALYSIS REVIEWS 37

CSTR: Continuous stirred tank reactor


DBU: 1,8-diazabicyclo[5.4.0]undec-7-ene
DMF: Dimethylformamide
F: Flow-rate of carrier gas
FTIR: Fourier-transform infrared
IL: Ionic liquid
IL-1: 1-N-butyl-3-methylimidazolium hydrogen chloride
IL-2: 1-Butyl-4-methylpyridinuium bromide
IL-3: Trihexyl (tetradecyl) phosphonium bromide
IL-4: 1,3-Disulfonic acid imidazolium hydrogen sulfate
K: Kelvin
K: Adsorption constant
k: Rate constant
K2CO3: Potassium carbonate
KBr: Potassium bromide
K10: Montmorillonite clay
L: Lewis
La: Lanthanum
La2(HPO4)3: Lanthanum hydrogen phosphate
LDH: Layered double hydroxides
Leu: Leucine
LHSV: Liquid hour space velocity
MgO: Magnesium oxide
NaNO2: Sodium nitrite
PEG: Poly(ethylene glycol)
PFR: Plug flow reactor
PTC: Phase transfer catalyst
SAPO: Silicoaluminophosphate
Sb: Antimony
SiO2: Silica
Sm: Samarium
Sr(HSO4)2: Strontium hydrogen sulfate
SrSO4: Strontium sulfate
TPA: Tungustophosphoric acid
TPD: Temperature programmed desorption
TPO: Temperature programmed oxidation
TBAB: Tetrabutylammonium bromide
UV: Ultraviolet
WHSV: Weight hour space velocity
Xp: Conversion of phenol

Acknowledgments
Priyanka Bhongale acknowledges Council of Scientific and Industrial Research (CSIR), New
Delhi, India for Senior Research Fellowship. The authors thank Academy of Scientific &
Innovative Research (AcSIR) and CSIR-National Chemical Laboratory for additional support.
38 P. BHONGALE ET AL.

References
[1] Newman, M. S.; Cella, J. A. Studies on the Monoalkylation of Hydroquinone. J. Org.
Chem. 1974, 39(2), 214–215. DOI: 10.1021/jo00916a020.
[2] Ashtekar, S.; Kumbhar, P.; Mahalingam, R. J.; Thampi, J. Process for the Monoalkylation
of Dihydroxy Aromatic Compounds. US7087705B2, 2006 (accessed Aug 8, 2006).
[3] Gradeff, P. S.; Bertrand, C. Monoalkylation of Unsubstituted Dihydric Phenols with
Lower Alkyl Chlorides Using Alkali Metal Carbonated or Bicarbonated as a Condensing
Agent. US3689570, 1972 (accessed Sep 5, 1972).
[4] Renaud, M.; Chantal, P. D.; Kaliaguine, S. Anisole Production by Alkylation of Phenol
over ZSM5. Can. J. Chem. Eng. 1986, 64(5), 787–791. DOI: 10.1002/cjce.5450640511.
[5] Titze-Frech, K.; Ignatiev, N.; Uerdingen, M.; Schulz, P. S.; Wasserscheid, P. Highly
Selective Aromatic Alkylation of Phenol and Anisole by Using Recyclable Bronsted
Acidic Ionic Liquid Systems. European. J. Org. Chem. 2013, 2013(30), 6961–6966.
DOI: 10.1002/ejoc.201300579.
[6] Lamoureux, G.; Agüero, C. A. Comparison of Several Modern Alkylating Agents. Spec.
Issue Rev. And Account, ARKIVOC. 2009, 251–264. DOI: 10.3998/
ark.5550190.0010.108.
[7] Lewis, H. F.; Shaffer, S.; Trieschmann, W.; Cogan, H. Methylation of Phenol by Dimethyl
Sulfate. Ind. Eng. Chem. 1930, 22(1), 34–36. DOI: 10.1021/ie50241a009.
[8] McCloud, G. T.; Brimer, M. R.; Gibson, C. L. Etherification of Phenolic Compounds.
US3911022, 1975 (accessed Oct 7, 1975).
[9] Correale, M.; Panseri, P.; Romano, U.; Minisci, F. Process for the Preparation of
Mono-ethers of Hydroquinones. US4933504, 1990 (accessed Jun 12, 1990).
[10] Clark, J. H. Green Chemistry: Challenges and Opportunities. Green. Chem. 1999, 1(1),
1–8. DOI: 10.1039/A807961G.
[11] Kocal, J. A.; Vora, B. V.; Imai, T. Production of Linear Alkylbenzenes. Appl. Catal.
A. Gen. 2001, 221(1), 295–301. DOI: 10.1016/S0926-860X(01)00808-0.
[12] Bellas, M.; Cahill, R. Process for Monoalkylation of Dihydric Phenols. US4294991, 1981
(accessed Oct 13, 1981).
[13] Bal, R.; Mayadevi, S.; Sivasanker, S. O-methylation of Dihydroxybenzenes with Methanol
in the Vapour Phase over Alkali-loaded SiO2 Catalysts : A Kinetic Analysis. Org. Process.
Res. Dev. 2003, 7(1), 17–21. DOI: 10.1021/op020054o.
[14] Jafari, A. A.; Khodadadi, A.; Mortazavi, Y. Vapor-phase Selective O-alkylation of
Catechol with Methanol over Lanthanum Phosphate and Its Modified Catalysts with
Ti and Cs. J. Mol. Catal. A Chem. 2013, 372, 79–83. DOI: 10.1016/j.molcata.2013.02.011.
[15] Bal, R.; Tope, B. B.; Sivasanker, S. Vapour Phase O-methylation of Dihydroxy Benzenes
with Methanol over Cesium-loaded Silica, a Solid Base. J. Mol. Catal. A Chem. 2002, 181,
161–171. DOI: 10.1016/S1381-1169(01)00361-2.
[16] González-Borja, M. A.; Resasco, D. E. Reaction Pathways in the Liquid Phase Alkylation
of Biomass-derived Penolic Compounds. AIChE. J. 2015, 61(2), 598–609. DOI: 10.1002/
aic.14658.
[17] Elavarasan, P.; Kondamudi, K.; Upadhyayula, S. Comparative Kinetic Study on
Alkylation of P-cresol with Tert-butyl Alcohol Using Different SO3-H Functionalized
Ionic Liquid Catalysts. Int. J. Chem. Mol. Eng. 2010, 4(6), 384–389.
[18] Markets, F. Catalyst Market by Material (Metal, Chemical, Zeolites, Organometallic
Materials), Type, Application, Regions, Global Industry Analysis, Market Size, Share,
Growth, Trends, and Forecast 2018 to 2025. Fior Markets, Report Database, 2019.
[19] Bravo-Suarez, J. J.; Chaudhari, R. V.; Subramaniam, B. Design of Heterogeneous
Catalysts for Fuels and Chemicals Processing: An Overview. Novel Materials for
CATALYSIS REVIEWS 39

Catalysis and Fuel Processing, ACS symposium Series. 2013, Chapter 1(1132), 3–68. DOI:
10.1021/bk-2013-1132.ch001
[20] Ouk, S.; Thiébaud, S.; Borredon, E.; Gars, P. L. High Performance Method for
O-methylation of Phenol with Dimethyl Carbonate. Appl. Catal. A. 2003, 241,
227–233. DOI: 10.1016/S0926-860X(02)00467-2.
[21] Ouk, S.; Thiébaud, S.; Borredon, E.; Gars, P. L.; Lecomte, L. O-methylation of Phenolic
Compounds with Dimethyl Carbonate under Solid/Liquid Phase Transfer System.
Tetrahedron. Lett. 2002, 43(14), 2661–2663. DOI: 10.1016/S0040-4039(02)00201-0.
[22] Tilstam, U. Continuous Base-catalyzed, A. Methylation of Phenols with Dimethyl
Carbonate. Org. Process. Res. Dev. 2012, 16(5), 1150–1153. DOI: 10.1021/op200379j.
[23] Bal, R.; Sivasanker, S. Vapour Phase Selective O-alkylation of Phenol over Alkali Loaded
Silica. Appl. Catal. A. 2003, 246(2), 373–382. DOI: 10.1016/S0926-860X(03)00082-6.
[24] Heravi, M. M.; Ahari, N. Z.; Oskooie, H. A.; Ghassemzadeh, M. Solid State S-methylation
of Thiols and O-methylation of Phenols and Naphthols with Dimethyl Sulfate under
Microwave Irradiation. Phosphorus, Sulfur, Silicon. Relat. Elem. 2005, 180(7), 1701–1712.
DOI: 10.1080/10426500590885165.
[25] Parrish, J. P.; Sudaresan, B.; Jung, K. W. Improved Cs2CO3 Promoted O-alkylation of
Phenols. Synth. Commun. 1999, 29(24), 4423–4443. DOI: 10.1080/00397919908086606.
[26] Bu, X.; Jing, H.; Wang, L.; Chang, T.; Jin, L.; Liang, Y. Organic Base Catalyzed
O-alkylation of Phenols under Solvent-free Condition. J. Mol. Catal. A Chem. 2006,
259(1), 121–124. DOI: 10.1016/j.molcata.2006.06.009.
[27] Brieger, G.; Hachey, D.; Nestrick, T. Convenient O-alkylation of Phenols. J. Chem. Eng.
Data. 1968, 13(4), 581–582. DOI: 10.1021/je60039a048.
[28] Joseph, L.; Paramus, N. J.; Alvin, F. Method for Producing Monoalkyl Ethers of Dihydric
Phenols. US3274260, 1966 (accessed Sept 20, 1966).
[29] Gambarotti, C.; Melone, L.; Punta, C.; Shisodia, S. U. Selective Monoetherification of
1,4-hydroquinone Promoted by NaNO2. Curr. Org. Chem. 2013, 17, 1108–1113. DOI:
10.2174/1385272811317100011.
[30] Radoslaw, M.; Sebastian, S. W, S. J. Simple Synthesis of Non-symmetric
1,4-dialkoxybenzenes via 4-alkoxyphenols. Org. Commun. 2016, 9(1), 23–31.
[31] Becker, H.; Vogel, H. The Role of Hydroquinone Monomethyl Ether in the Stabilization
of Acrylic Acid. Chem. Eng. Technol. 2006, 29, 1227–1231. DOI: 10.1002/ceat.200500401.
[32] Yadav, G. D.; Deshmukh, S. A. R. K.; Asthana, N. S. Synthesis of Hydroquinone
Monomethyl Ether from Hydroquinone and Methanol over Heteropolyacids
Supported on Clay : Kinetics and Mechanism. Ind. Eng. Chem. Res. 2005, 44,
7969–7977. DOI: 10.1021/ie050519.
[33] Lee, S. S.; Lee, S. K.; Kim, J. C. Selective O-alkylation Reaction of Hydroquinone with
Methanol over Cs Ion-exchanged Zeolites. Korean J. Chem. Eng. 2002, 19(3), 406–410.
DOI: 10.11405/nisshoshi1964.62.850.
[34] Indurkar, J. R. Selective Para-functionalization of Phenol. Nelson Mandela Metropolitan
University, 2008, January 1–195.
[35] Sad, M. E.; Padro, C. L.; Apesteguia, C. R. Study of the Phenol Methylation Mechanism
on Zeolites HBEA, HZSM5 and HMCM22. J. Mol. Catal. A Chem. 2010, 327, 63–72.
DOI: 10.1016/j.molcata.2010.05.014.
[36] Devi, G. S.; Giridhar, D.; Reddy, B. M. Vapour Phase O-alkylation of Phenol over Alkali
Promoted Rare Earth Metal Phosphates. J. Mol. Catal. A. Chem. 2002, 181, 173–178.
DOI: 10.1016/S1381-1169(01)00362-4.
[37] Chakrabarti, A.; Sharma, M. M. State-of-the-Art Report Cationic Ion Exchange Resins as
Catalyst. React. Polym. 1993, 20, 1–45. DOI:10.1016/0923-1137(93)90064-M.
40 P. BHONGALE ET AL.

[38] Wu, G.; Wang, X.; Chen, B.; Li, J.; Zhao, N.; Wei, W.; Sun, Y. Fluorine-modified
Mesoporous Mg–Al Mixed Oxides: Mild and Stable Base Catalysts for O-methylation
of Phenol with Dimethyl Carbonate. Appl. Catal. A. 2007, 329, 106–111. DOI: 10.1016/j.
apcata.2007.06.031.
[39] Fuming, M.; Zhi, P.; Guangxing, L. The Transesterification of Dimethyl Carbonate with
Phenol over Mg-Al Hydrotalcite Catalyst. Org. Process. Res. Dev. 2004, 8(3), 372–375.
DOI: 10.1021/op0302098.
[40] Kabra, S. K.; Huuhtanen, M.; Keiski, R. L.; Yadav, G. D. Selectivity Engineering of
O-methylation of Hydroxybenzenes with Dimethyl Carbonate Using Ionic Liquid as
Catalyst. React. Chem. Eng. 2016, 1, 330–339. DOI: 10.1039/C6RE00016A.
[41] Gandhe, A. R.; Fernandes, J. B. Highly ortho-selective TiO2 Catalyst for the Methylation
of Phenol. Catal. Commun, 2004, 5(2), 89–94. DOI: 10.1016/j.catcom.2003.11.017.
[42] Kirichenko, G. N.; Glazunova, V. I.; Balaev, A. V.; Dzhemilev, U. M. Catalytic
Vapor-phase Alkylation of Phenol with Methanol. Pet. Chem. 2008, 48(5), 389–392.
DOI: 10.1134/s0965544108050095.
[43] Santacesaria, E.; Grasso, D.; Gelosa, D.; Carrà, S. Catalytic Alkylation of Phenol with
Methanol: Factors Influencing Activities and Selectivities. Appl. Catal. 1990, 64,
101–117. DOI: 10.1016/s0166-9834(00)81556-0.
[44] Samolada, M. C.; Grigoriadou, E.; Kiparissides, Z.; Vasalos, I. A. Selective O-alkylation of
Phenol with Methanol over Sulfates Supported on γ-Al2O3. J. Catal. 1995, 152(1), 52–62.
DOI: 10.1006/jcat.1995.1059.
[45] Crocella, V.; Cerrato, G.; Magnacaa, G.; Morterra, C.; Cavani, F.; Maselli, L.; Passeri, S.
Gas-phase Phenol Methylation over Mg/Me/O (Me = Al, Cr, Fe) Catalysts: Mechanistic
Implications Due to Different Acid-base and Dehydrogenating Properties. Dalton Trans.
2010, 39, 8527–8537. DOI: 10.1039/C002490B.
[46] Bautista, F. M.; Campelo, J. M.; Garcia, A.; Luna, D.; Marinas, J. M.; Romero, A.;
Navio, J. A.; Macias, M. Anion Treatment of AlPO4-Al2O3(25 Wt.% Al2O3) Catalyst.
IV. Catalytic Perfomance in the Alkylation of Phenol with Methanol. Appl. Catal. A.
1993, 99(2), 161–173. DOI: 10.1016/0926-860X(93)80097-A.
[47] Romero, M. D.; Ovejero, G.; Rodríguez, A.; Gómez, J. M.; Águeda, I. O-methylation of
Phenol in Liquid Phase over Basic Zeolites. Ind. Eng. Chem. Res. 2004, 43(26),
8194–8199. DOI: 10.1021/ie049461.
[48] Jyothi, T. M.; Raja, T.; Talawar, M. B.; Sreekumar, K.; Sugunan, S.; Rao, B. S. Selective
Methylation of Phenol, Aniline and Catechol with Dimethyl Carbonate over Calcined
Mg-Al Hydrotalcites. Synth. Commun. 2000, 30(21), 3929–3934. DOI: 10.1080/
00397910008086951.
[49] Guo, F.; Fang, Z.; Xu, C. C.; Smith Jr, R.; Solid Acid, L. Mediated Hydrolysis of Biomass
for Producing Biofuels. Prog. Energy Combust. Sci. 2012, 38(5), 672–690. DOI: 10.1016/j.
pecs.2012.04.001.
[50] Moshoeshoe, M.Nadiye-Tabbiruka, M. S.Obuseng, V. A Review of the Chemistry,
Structure, Properties and Applications of Zeolites. American J. Mater. Sci. White Rose
eTheses Online. 2017, 7(5), 196–221. DOI: 10.5923/j.materials.20170705.12.
[51] Deka, R. C. Acidity in Zeolites and Their Characterization by Different Spectroscopic
Methods. Indian. J. Chem. Technol. 1998, 5(3), 109–123. DOI: 10.1002/chin.19995124.
[52] Schoonheydt, R. A.; Geerlings, P.; Pidko, E. A.; Van Santen, R. A. The Framework
Basicity of Zeolites. J. Mater. Chem. 2012, 22(36), 18705–18717. DOI: 10.1039/
C2JM31366A.
[53] Busca, G. Acidity and Basicity of Zeolites: A Fundamental Approach. Microporous
Mesoporous Mater. 2017, 254, 3–16. DOI: 10.1016/j.micromeso.2017.04.007.
CATALYSIS REVIEWS 41

[54] Liu, X. Li, LYang, T.Yan, Z. Zeolite Y Synthesized with FCC Spent Catalyst Fines:
Particle Size Effect on Catalytic Reactions. J. Porous Mater. Shodhganga: A reservoir
of Indian theses. 2012,19, 133–139. DOI: 10.1007/s10937-011-9534-1.
[55] Namba, S.; Yashima, T.; Itaba, Y.; Hara, N. Selective Formation of p-cresol by Alkylation
of Phenol with Methanol over Y Type Zeolite. Studie.Surf. Sci. Catal. 1980, 5, 105–111.
DOI: 10.1016/S167-2991(08)64870-3.
[56] Balsama’, S.; Beltrame, P.; Beltrame, P. L.; Carniti, L.; Forni, L.; Zuretti, G. Alkylation of
Phenol with Methanol over Zeolites. Appl. Catal. 1984, 13(1), 161–170. DOI: 10.1016/
S0166-9834(00)83334-5.
[57] Marczewski, M.; Bodibo, J.-P.; Perot, G.; Guisnet, M. Alkylation of Aromatics. Part
I. Reaction Network of the Alkylation of Phenol by Methanol on USHY Zeolite. J. Mol.
Catal. 1989, 50(2), 211–218. DOI: 10.1016/0304-5102(89)85064-3.
[58] Xu, J.; Yan, A.-Z.; Xu, Q.-H. Alkylation of Phenol with Methanol on H-beta Zeolite.
React. Kinet. Catal. Lett. 1997, 62(1), 71–74. DOI: 10.1007/BF02475715.
[59] Cheralathan, K. K.; Kumar, I. S.; Palanichamy, M.; Murugesan, V. Liquid Phase
Alkylation of Phenol with 4-hydroxybutan-2-one in the Presence of Modified Zeolite
HBEA. Appl. Catal. A. Gen. 2003, 241,247–260. DOI: 10.1016/S0926-860X(02)00472-6.
[60] Reddy, K. R.; Ramesh, K.; Seela, K. K.,Vattikonda, V. R.; Komandur, V. R. Chary
Alkylation of Phenol with Methanol over Molybdenum Oxide Supported on NaY
Zeolite. Catal. Commun. 2003, 4(3), 112–117. DOI: 10.1016/S1566-7367(03)00006-2.
[61] Barman, S.; Pradhan, N. C.; Basu, J. K. Kinetics of Alkylation of Phenol with Methanol
over Ce-exchanged NaX Zeolite. Catal. Letters. 2006, 111, 67–73. DOI: 10.1007/s10562-
006-0135-6.
[62] Bhattacharyya, K. G.; Talukdar, A. K.; Das, P.; Sivasanker, S. Al-MCM-41 Catalysed
Alkylation of Phenol with Methanol. J. Mol. Catal. A Chem. 2003, 197, 255–262. DOI:
10.1016/S1381-1169(02)00624-6.
[63] Bregolato, M.; Bolis, V.; Busco, C.; Ugliengo, P.; Bordinga, S.; Cavani, F.; Ballarini, N.;
Maselli, L.; Passeri, S.; Rosseri, I.; et al. Methylation of Phenol over High-silica Beta
Zeolite : Effect of Zeolite Acidity and Crystal Size on Catalyst Behaviour. J. Catal. 2007,
245, 285–300. DOI: 10.1016/j.jcat.2006.10.024.
[64] Ballarini, N.; Cavani, F.; Maselli, L.; Montaletti, A.; Passeri, S.; Scagliarini, D.; Flego, C.;
Perego, C. The Transformations Involving Methanol in the Acid- and Base-catalyzed
Gas-phase Methylation of Phenol. J. Catal. 2007, 251(2), 423–436. DOI: 10.1016/j.
jcat.2007.07.033.
[65] Sad, M. E.; Padró, C. L.; Apesteguia, C. R. Phenol Methylation on Acid Catalysts: Study
of the Catalyst Deactivation Kinetics and Mechanism. Appl. Catal. A. Gen. 2014, 475,
305–313. DOI: 10.1016/j.apcata.2014.01.039.
[66] Moon, G.; Bohringer, W.; O’Connor, C. T. Pressure Induced Enhancement of Shape
Selective Phenol Methylation. Studie.Surf. Sci. Catal. 2004, 154(PartC), 2255–2260. DOI:
10.1016/S167-2991(04)80483-X.
[67] Moon, G.; Möller, K. P.; Böhringer, W.; O’Connor, C. T. Alkylation of Phenol with
Methanol over Zeolite H-MCM-22 for the Formation of p-cresol. Studie.Surf. Sci. Catal.
2002, 142, 635–642. DOI: 10.1016/S0167-2991(02)80083-0.
[68] Churkin, U. V.; Chvalyuk, L. A.; Akhunova, L. V.; Kirichenko, G. N.; Spivak, S. I. Kinetic
Model for Anisol Synthesis by Methanol Alkylation of Phenol on a Zeolite Catalyst.
React. Kinet. Catal. Lett. 1985, 29,145–151. DOI: 10.1007/BF02067961.
[69] Bezouhanova, C.; Al-Zihari, M. A. Alkylation of Phenol with Methanol over Mn3O4.
Appl. Catal. A. 1992, 83(1), 45–49. DOI: 10.1016/0926-860X(92)80024-7.
42 P. BHONGALE ET AL.

[70] Nozaki, F.; Kimura, I. A. Study of Catalysis by Metal Phosphates. IV. The Alkylation of
Phenol with Methanol over Metal Phosphate Catalysts. Bull. Chem. Soc. Jpn. 1977, 50(3),
614–619. DOI: 10.1246/bcsj.50.614.
[71] Rao, V. V.; Chary, K. V. R.; Durgakumari, V.; Narayanan, S. Alkylation of Phenol over
Simple Oxides and Supported Vanadium Oxides. Appl. Catal. 1990, 61, 89–97. DOI:
10.1016/S0166-9834(00)82137-5.
[72] Malshe, K. M.; Patil, P. T.; Umbarkar, S. B.; Dongare, M. K. Selective C-methylation of
Phenol with Methanol over Borate Zirconia Solid Catalyst. J. Mol. Catal. A Chem. 2004,
212, 337–344. DOI: 10.1016/j.molcata.2003.11.016.
[73] Védrine, J. C. Importance, Features and Uses of Metal Oxide Catalysts in Heterogeneous
Catalysis. Chin. J. Catal. 2019, 40(19), 1627–1636. DOI: 10.1016/S1872-2067(18)63162-
6.
[74] Védrine, J. C. Metal Oxides in Heterogeneous Oxidation Catalysis: State of the Art and
Challenges for a More Sustainable World. Chem. Sus. Chem. 2019, 12(3), 577–588. DOI:
10.1002/cssc.201802248.
[75] Searle, N. E. Process for Preparing Anisole. US2487832, 1949 (accessed Nov 15, 1949).
[76] Sharp, J. A. Dean, R. E. Alkylation of Phenols. US3642912, 1972 (accessed Feb 15, 1972).
[77] Karuppannasamy, S.; Narayanan, K.; Pillai, C. N. Reactions of Phenols and Alcohols over
Thoria: Mechanism of Ether Formation. J. Catal. 1980, 66(2), 281–289. DOI: 10.1016/
0021-9517(80)90032-9.
[78] Tleimat-Manzalji, R.; Bianchi, D.; Pajonk, G. M. Catalytic Alkylation of Phenol with
Methanol on a Microporous Xerogel and a Macroporous Aerogel of High Surface Area
Aluminas. Appl. Catal. A. Gen. 1993, 101(2), 339–350. DOI: 10.1016/0926-860X(93)
80278-X.
[79] Gandhe, A. R.; Naik, S. P.; Kakodkar, S. B.; Fernandes, J. B.; Highly Active, A. Anatase
TiO2 Catalyst for Alkylation of Phenol with Methanol. Catal. Commun. 2006, 7(5),
285–288. DOI: 10.1016/j.catcom.2005.09.013.
[80] Sato, S.; Koizumi, K.; Nozaki, F. Ortho-selective Methylation of Phenol over CeO2
Catalyst. Appl. Catal. A. 1995, 133(1), 1–4. DOI: 10.1016/0926-860X(95)00225-1.
[81] Grabowska, H.; Miśta, W.; Trawczyński, J.; Wrzyszcz, J.; Zawadzki, M. Catalytic
Alkylation of Phenol with Methanol over Zinc Aluminate. Res. Chem. Intermed. 2001,
27(3), 305–313. DOI: 10.1163/156856701300356527.
[82] Wells, J. E.; Eskinazi, V. Alkylation of Phenols to Alkyl Aryl Ethers Using Phosphate
Catalysts. US4533758, 1985 (accessed by Aug 6, 1958).
[83] Eskinazi, V. Selective Catalytic Alkylation of Phenols to Alkyl Aryl Ethers. US4450306,
1984 (accessed May 22, 1984).
[84] Mossman, A. B. Catalytic Etherification of Phenols to Alkyl Aryl Ethers. US4638098,
1987 (accessed Jan 20, 1987).
[85] Mossman, A. B. Catalytic Etherification of Phenols to Alkyl Aryl Ethers. US4611084,
1986 (accessed Sep 9, 1986).
[86] Mossman, A. B. Catalytic Etherification of Phenols to Alkyl Aryl Ethers. US4675456,
1987 (accessed Jun 23, 1987).
[87] Pierantozzi, R.; Nordquist, A. Selective O-alkylation of Phenol with Methanol. Appl.
Catal. 1986, 21(2), 263–271. DOI: 10.1016/S0166-9834(00)81359-7.
[88] Durgakumari, V.; Narayanan, S.; Guczi, L. Alkylation of Phenol with Methanol over
AlPO and SAPO Molecular Sieves. Catal. Lette. 1990, 5(4), 377–384. DOI: 10.1007/
BF00765180.
[89] Velu, S.; Swamy, C. S. Kinetics of the Alkylation of Phenol with Methanol over Catalysts
Derived from Hydrotalcite-like Anionic Clays. React. Kinet. Catal. Lett. 1997, 62(2),
339–346. DOI: 10.1007/BF02475473.
CATALYSIS REVIEWS 43

[90] Sharma, S.; Ameta, S. C.; Sharma, V. K. Use of Dimethyl Carbonate (DMC) as
Methylating Agent under Microwave Irradiation-A Green Chemical Approach.
Proceedings of the World Congr. Eng. Comput. Sci San Francisco, USA. 2010, 2, 721–723.
[91] Ono, Y. Catalysis in the Production and Reactions of Dimethyl Carbonate, an
Environmentally Benign Building Block. Appl. Catal. A. Gen. 1997, 155(2), 133–166.
DOI: 10.1016/S0926-860X(96)00402-4.
[92] Wang, Q.; Sun, M. M.; Guo, M.; Luo, S. J.; Li, C. H.; Hu, C. H. Heterogeneous
Transesterification of Dimethyl Carbonate and Phenol to Diphenyl Carbonate over
Vanadium Pentoxide Microstructures. Integr. Ferroelectr. 2015, 164(1), 154–164. DOI:
10.1080/10584587.2015.1045399.
[93] Tundo, P.; Musolino, M.; Aricò, F. The Reactions of Dimethyl Carbonate and Its
Derivatives. Green Chem. 2018, 20(1), 28–85. DOI: 10.1039/C7GC01764B.
[94] Tundo, P.; Aricò, F. Dimethyl Carbonate: Green Solvent and Ambient Reagent. Green.
Chem. React. 2010, 79, 213–232. DOI: 10.1007/978-1-4020-8457-7_10.
[95] Aricò, F.; Tundo, P. Dimethyl Carbonate as a Modern Green Reagent and Solvent. Russ.
Chem. Rev. 2010, 79, 479–489. DOI: 10.1070/rc2010v079n06abeh004113.
[96] Glasnov, T. N.; Holbrey, J. D.; Kappe, C. O.; Seddon, K. R.; Yan, T. Methylation Using
Dimethylcarbonate Catalysed by Ionic Liquids under Continuous Flow Conditions.
Green. Chem. 2012, 14(11), 3071–3076. DOI: 10.1039/c2gc36226k.
[97] Tundo, P.; Trotta, F.; Moraglio, G.; Ligorati, F. Continuous-flow Processes under
Gas-liquid Phase-transfer Catalysis (GL-PTC) Conditions: The Reaction of Dialkyl
Carbonates with Phenols, Alcohols, and Mercaptans. Ind. Eng. Chem. Res. 1988, 27(9),
1565–1571. DOI: 10.1021/ie00081a002.
[98] Talawar, M. B.; Jyothi, T. M.; Sawant, P. D.; Raja, T.; Rao, B. S. Calcined Mg-Al
Hydrotalcite as an Efficient Catalyst for the Synthesis of Guaiacol. Green. Chem. 2000,
2(6), 266–268. DOI: 10.1039/b006077l.
[99] Vafaeezadeh, M.; Alinezhad, H. Brønsted Acidic Ionic Liquids: Green Catalysts for
Essential Organic Reactions. J. Mol. Liq. 2016, 218, 95–105. DOI: 10.1016/j.
molliq.2016.02.017.
[100] Khaligh, N. G. Preparation, Characterization and Use of 3-methyl-1-sulfonic Acid
Imidazolium Hydrogen Sulfate as an Eco-benign, Efficient and Reusable Ionic Liquid
Catalyst for the Chemoselective Trimethylsilyl Protection of Hydroxyl Groups. J. Mol.
Catal. A Chem. 2011, 349, 63–70. DOI: 10.1016/j.molcata.2011.08.021.
[101] Hawker, R. R.; Harper, J. B. Chapter Two - Organic Reaction Outcomes in Ionic Liquids.
Advances in Physical Organic Chemistry, 2018; 2015, 52, 49–85. DOI: 10.1016/bs.
apoc.2018.09.001.
[102] Gazitúa, M.; Fuentealba, P.; Contreras, R.; Ormazabal-Toledo, R. Lewis Acidity/Basicity
Changes in Imidazolium Based Ionic Liquids Brought about by Impurities. J. Phys.
Chem. B. 2015, 119(41), 13160–13166. DOI: 10.1021/acs.jpcb.5b05305.
[103] Dharaskar, S. A.; Wasewar, K. L.; Varma, M. N.; Shende, D. Z.; Yoo, C. Synthesis,
Characterization and Application of 1-butyl-3-methylimidazolium Tetrafluoroborate
for Extractive Desulfurization of Liquid Fuel. Arab. J. Chem. 2016, 9(4), 578–587.
DOI: 10.1016/j.arabjc.2013.09.034.
[104] Joni, J. Development of a Sustainable Technology Platform for the Homogeneous Friedel-
Crafts Alkylation Using Acidic Ionic Liquid Catalyst. Friedrich-Alexander-Universität
Erlangen-Nürnberg. 2009.
[105] Tao, D.; Wu, J.; Wang, Z.; Lu, Z.; Yang, Z.; Chen, X. SO3H-functionalized Bronsted
Acidic Ionic Liquids as Efficient Catalysts for the Synthesis of Isoamyl Salicylate. RSC
Adv. 2014, 4(1), 1–7. DOI: 10.1039/c3ra45921g.
44 P. BHONGALE ET AL.

[106] Welton, T. Ionic Liquids in Catalysis. Coord. Chem. Rev. 2004, 248, 2459–2477. DOI:
10.1016/j.ccr.2004.04.015.
[107] Li, M.; Zang, H.; Feng, J.; Yan, Q.; Yu, N.; Shi, X.; Cheng, B. Efficient Conversion of
Chitosan into 5-hydroxymethylfurfural via Hydrothermal Synthesis in Ionic Liquids
Aqueous Solution. Polym. Degrad. Stab. 2015, 121, 331–339. DOI: 10.1016/j.
polymdegradstab.2015.09.009.
[108] Mihichuk, L.; Driver, G.; Johnson, K. Acidity in Ionic Liquids. Proceedings in A Clean
Environment, Fray Int. Symp., Cancun, Mexico. 2011, 3, 307–315.
[109] Nie, J.; Chen, H.; Song, Q.; Liao, B.; Guo, Q. O-Methylation of the Phenolic Bio-oil with
Dimethyl Carbonate in an Ionic Liquid [Bmim]cl. Energy Fuels. 2010, 24(10),
5722–5726. DOI: 10.1021/ef1008659.
[110] Shen, Z. L.; Jiang, X. Z.; Mo, W. M.; Hu, B. X.; Sun, N. Catalytic O-methylation of
Phenols with Dimethyl Carbonate to Aryl Methyl Ethers Using [Bmim]cl. Green Chem.
2005, 7(2), 97–99. DOI: 10.1039/b411201f.
[111] Das, P. J.; Das, J. Selective O-methylation of Phenols and Benzyl Alcohols in Simple
Pyridinium Based Ionic Liquids. J. Mol. Liq. 2015, 209, 94–98. DOI: 10.1016/j.
molliq.2015.05.028.
[112] Tundo, P.; Trotta, F.; Moraglio, G. Reactions of Dimethyl Carbonate with Nucleophiles
under Gas-liquid Phase-transfer Catalysis (GL-PTC) Conditions. React. Polym. 1989, 10,
185–188. DOI: 10.1016/0923-1137(89)90025-0.
[113] Lissel, M.; Schmidt, S.; Neumann, B. Use of Dimethyl Carbonate as a Methylating Agent
under Phase Transfer-catalysed Conditions. Synthesis (Stuttg). 1986, 5, 382–383. DOI:
10.1055/s-1986-31642.
[114] Merger, F.; Towae, F.; Schroff, L. Preparation of Aralkyl Phenyl Ethers and Alkyl Phenyl
Ethers. US4192949, 1980 (accessed Mar 11, 1980).
[115] Barcelo, G.; Grenouillat, D.; Senet, J.; Sennyey, G. Pentaalkylguanidines as Etherification
and Esterification Catalysts. Tetrahedron. 1990, 46(6), 1839–1848. DOI: 10.1016/S0040-
4020(01)89753-2.
[116] Tundo, P. New Developments in Dimethyl Carbonate Chemistry. Pure Appl. Chem.
2001, 73, 1117–1124. DOI: 10.1351/pac200173071117.
[117] Tundo, P.; Selva, M. The Chemistry of Dimethyl Carbonate. Acc. Chem. Res. 2002, 35(9),
706–716. DOI: 10.1021/ar010076f.
[118] Bomben, A.; Selva, M.; Tundo, P.; Valli, L. A Continuous-flow O-methylation of Phenols
with Dimethyl Carbonate in A Continuously Fed Stirred Tank Reactor. Ind. Eng. Chem.
Res. 1999, 38(5), 2075–2079. DOI: 10.1021/ie9806444.
[119] Bomben, A.; Selva, M.; Tundo, P. Dimethyl Carbonate as a Methylating Agent. The
Selective Mono-C-methylation of Alkyl Aryl Sulfones. J. Chem. Res. Synopses. 1997,
448–449. DOI: 10.1039/A703510A.
[120] Memoli, S.; Selva, M.; Tundo, P. Dimethylcarbonate for Eco-friendly Methylation
Reactions. Chemosphere. 2001, 43(1), 115–121. DOI: 10.1016/S0045-6535(00)00331-3.
[121] Ouk, S.; Thiébaud, S.; Borredon, E.; Gars, P. L. Dimethyl Carbonate and Phenols to Alkyl
Aryl Ethers via Clean Synthesis. Green Chem. 2002, 4(5), 431–435. DOI: 10.1039/
b203353b.
[122] Perosa, A.; Selva, M.; Tundo, P.; Zordan, F. Alkyl Methyl Carbonates as Methylating
Agents. The O-Methylation of Phenols. Synlett. 2000, 2000, 272–274. DOI: 10.1055/
s-2000-6488.
[123] Shieh, W.; Dell, S.; Repič, O. 1,8-Diazabicyclo[5.4.0]undec-7-ene (DBU) and
Microwave-accelerated Green Chemistry in Methylation of Phenols, Indoles and
Benzimidazoles with Dimethyl Carbonate. Org. Lett. 2001, 3(26), 4279–4281. DOI:
10.1021/ol016949n.
CATALYSIS REVIEWS 45

[124] Fu, Z.; Ono, Y. Selective O-methylation of Phenol with Dimethyl Carbonate over
X-zeolites. Catal. Lett. 1993, 21(1), 43–47. DOI: 10.1007/BF00767369.
[125] Barthomeuf, D.;. Spectroscopic Investigations of Zeolite Properties NATO ASI Series
(Series C: Mathematical and Physical Sciences). Zeolite Microporous Solids: Synth. Struct.
React. 1992, 352, 193–223. DOI: 10.1007/978-94-011-2604-5_10.
[126] Beutel, T. Spectroscopic and Kinetic Study of the Alkylation of Phenol with Dimethyl
Carbonate over NaX Zeolite. J. Chem. Soc. Faraday Trans. 1998, 94(7), 985–993. DOI:
10.1039/A706356C.
[127] Subramanian, T.; Dhakshinamoorthy, A.; Pitchumani, K. Amino Acid Intercalated
Layered Double Hydroxide Catalyzed Chemoselective Methylation of Phenols and
Thiophenols with Dimethyl Carbonate. Tetrahedron Lett. 2013, 54(52), 7167–7170.
DOI: 10.1016/j.tetlet.2013.10.098.
[128] Zhang, X.; Bing, C.; Wei, W.; Sun, Y. Preparation of Functionalized Porous Silica by
Ultrasonic Technique for the Methylation Reaction of Phenol with Dimethyl Carbonate.
Chinese J. Chem. 2005, 23, 1376–1380. DOI: 10.1002/cjoc.200591376.
[129] Chen, B.; Dong, Q.; Zhao, N.; Wei, W.; Sun, Y. FTIR Study of the Methylation of Phenol
with Dimethyl Carbonate over KBr/SiO2. J. Chem. Soc. Pakistan. 2009, 31, 552–558.
[130] Chen, B.; Zhao, N.; Wei, W.; Sun, Y. Mechanism Study of Anisole Synthesis from
O-methylation of Phenol with DMC over KBr Catalyst. J. Fuel Chem. Technol. 2011,
39, 144–148.
[131] Gadg, S. T.; Mishra, A.; Gajengi, A. L.; Shahi, N. V.; Bhanage, B. M. Magnesium Oxide as
a Heterogeneous and Recyclable Base for the N-methylation of Indole and
O-methylation of Phenol Using Dimethyl Carbonate as a Green Methylating Agent.
RSC Adv. 2014, 4(91), 50271–50276. DOI: 10.1039/C4RA07240E.
[132] Khusnutdinov, R. I.; Shchadneva, N. A.; Mayakova, Y. Y. Methylation of Phenol and Its
Derivatives with Dimethyl Carbonate in the Presence of Mn2(CO)10, W(CO)6, and Co2
(CO)8. Russ. J. Org. Chem. 2015, 51(3), 330–334. DOI: 10.1134/S1070428015030070.
[133] Pansegrau, P. D.; Munson, B. P. Method for the Production of Alkoxy and Aryloxy
Phenols. US5840997, 1998 (accessed Nov 24, 1998).
[134] Satoh, T.; Matsumoto, H.; Niiro, Y. Hydroquinone Derivatives. US5416242, 1995
(accessed May 16, 1995).
[135] Torii, S.; Tanaka, H.; Akada, M. Process for Preparing Alkoxyphenols. US4588835, 1986
(accessed May 13, 1986).
[136] Rivetti, F.; Romano, U.; Muzio, N. D. Process for Preparing Monoalkylethers of
Hydroquinone and Its Derivatives. US4469897, 1984 (accessed Sep 4, 1984).
[137] Grote, H. W. Etherification of Hydroquinone. US2615051, 1952 (accessed Oct 15, 1952).
[138] Park, J. J.; Lee, S. C.; Lee, S. S.; Jung, S. Y.; Lee, S. J.; Kim, J. C. The Effect of Phase
Transition of Methanol on the Reaction Rate in the Alkylation of Hydroquinone. Korean
J. Chem. Eng. 2009, 26(3), 649–653. DOI: 10.1007/s11814-009-0108-8.
[139] Mondal, R.; Guha, C.; Mallik, A. K. Facile Conversion of Para-benzoquinones to
Para-alkoxyphenols with Primary/Secondary Alcohols and Amberlyst-15: A Process
Showing Novel Reducing Property of Such Alcohols. Tetrahedron Lett. 2014, 55(1),
86–89. DOI: 10.1016/j.tetlet.2013.10.119.
[140] Bhongale, P. V.; Joshi, S. S.; Mali, N. A. Selective Monoalkylation of Hydroquinone in the
Presence of SO3H-functionalized Ionic Liquids as Catalysts. Chem. Pap. 2020, 74,
4461–4471. DOI: 10.1007/s11696-020-01230-1.
[141] Hasegawa, R.; Saito, K.; Takaoka, T.; Ishikita, H. pKa of Ubiquinone, Menaquinone,
Phylloquinone, Plastoquinone, and Rhodoquinone in Aqueous Solution. Photosynth.
Res. 2017, 133(1), 297–304. DOI: 10.1007/s11120-017-0382-y.
46 P. BHONGALE ET AL.

[142] Evgeny, D.; Taisa, D. Application of Thermodynamics to Biological and Materials Science:
Dissociation Energies of O−H Bonds of Phenols and Hydroperoxides; 2011, 1858–1866.
IntechOpen. doi:10.5772/13290
[143] Vijayaraj, M.; Gopinath, C. S. Selective Production of Methoxyphenols from
Dihydroxybenzenes on Alkali Metal Ion-loaded MgO. J. Catal. 2006, 243(2), 376–388.
DOI: 10.1016/j.jcat.2006.08.009.
[144] Lui, M. Y.; Lokare, K. S.; Hemming, E.; Stanley, J. N. G.; Perosa, A.; Selva, M.;
Masters, A. F.; Maschmeyer, T. Microwave-assisted Methylation of Dihydroxybenzene
Derivatives with Dimethyl Carbonate. RSC Adv. 2016, 6(63), 58443–58451. DOI:
10.1039/C6RA09841J.

You might also like