Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

Journal of Environmental Chemical Engineering 9 (2021) 105113

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


journal homepage: www.elsevier.com/locate/jece

Recent advances in the synthesis of cyclic carbonates via CO2 cycloaddition


to epoxides
Abdul Rehman a, b, *, Faisal Saleem a, b, Farhan Javed b, Amir Ikhlaq c, Syed Waqas Ahmad b,
Adam Harvey a
a
School of Engineering, Merz Court, Newcastle University, Newcastle Upon Tyne NE1 7RU, UK
b
Department of Chemical and Polymer Engineering, University of Engineering and Technology Lahore, Faisalabad Campus, Pakistan
c
Institute of Environmental Engineering & Research (IEER), University of Engineering & Technology, Lahore, Pakistan

A R T I C L E I N F O A B S T R A C T

Editor: Dr. Zhang Xiwang Global concerns about high CO2 levels and dwindling supplies of fossil resources are increasing. The utilization of
CO2 and waste biomass as renewable resources for valuable products is highly desirable to make various supply
Keywords: chains ‘greener’. Synthesis of cyclic carbonate from epoxide and CO2 is a highly promising reaction due to its
CO2 utilization 100% atom economy. Organic cyclic carbonates have a broad range of applications, such as high boiling point
Cyclic carbonates
green polar aprotic solvents, electrolytes in Li-ion batteries, the monomer for polycarbonates and polyurethanes
Catalysts
synthesis, intermediates for the manufacturing of pharmaceuticals and many other fine chemicals. Over the last
Reaction mechanism
Kinetic study 20 years significant progress has been made in this field of research. This review covers the recent developments
Flow chemistry in catalytic systems for the above-mentioned reaction that includes kinetics, mechanistic investigations and
highlights the parameters affecting their activity and selectivity. Moreover, the recent efforts on the utilization of
renewable resources as potential raw materials for the cyclic carbonate synthesis have also been discussed.
Nonetheless, to satisfy the demands of a multi-scale approach for industrial transformation, the use of flow
chemistry and novel reactor designs for the synthesis of cyclic carbonates has been also discussed.

1. Introduction remaining (65–70)% is going directly into the atmosphere as waste heat
[3]. Consequently, the resources of fossil fuels are depleting at a much
The diminishing national and global reserves of crude oil, coupled faster rate.
with fluctuating supplies of fossil resources, means that long-term sup­ The concentration of CO2 in the environment has been dramatically
ply and economic viability for petroleum-derived value-added products increased over the last two decades, causing severe climatic changes due
are at considerable risk. Additionally, many of the processes used to to global warming. The anthropogenic emission of CO2 into the atmo­
convert crude oil into valuable products utilize auxiliary chemical ele­ sphere now exceeds 36 Gt, which is 43% above the level since the
ments having sustainability issues (e.g. rare-earth and platinum group beginning of the industrial revolution [4]. Currently, many technologies
metals) [1], therefore it is likely that the ability of the chemical industry have been developed to reduce anthropogenic CO2 emissions. Generally,
to supply petroleum-derived products will come under ever-increasing these include ‘Carbon Capture and Storage’ (CCS) and ‘Carbon Capture
threat. Despite various efforts to develop alternative renewable energy and Utilization’ (CCU). CCS involves energy-intensive methods of CO2
resources, fossil fuels are still considered as the major source of power injection into exhausted oil and gas reservoirs under supercritical con­
generation, transportation fuel and raw materials for the production of ditions. The development of the less energy-intensive methodologies for
many other value-added products, and likely to remain so for the next the utilization of CO2 has gained much attention in both academia and
(20–40) years [2]. The demand for these products is continuously the chemical process industry. The utilization of a renewable source of
increasing due to the increased population. Only (30–35)% of the CO2 to produce other valuable products has attracted the attention of the
chemical energy produced by burning fossil fuel has been converted into scientific community. It can be used as a renewable, non-toxic, and
other forms of energy such as electrical, mechanical and others, and the economical source to produce many high-value products. In this aspect,

* Corresponding author at: School of Engineering, Merz Court, Newcastle University, Newcastle Upon Tyne NE1 7RU, UK.
E-mail address: a.rehman2@uet.edu.pk (A. Rehman).

https://doi.org/10.1016/j.jece.2021.105113
Received 27 September 2020; Received in revised form 18 December 2020; Accepted 20 January 2021
Available online 21 January 2021
2213-3437/© 2021 Elsevier Ltd. All rights reserved.
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

significant efforts have been devoted to academia and industry to


develop the methodologies for effective CO2 utilization. The utilization
of CO2 as a C1-building block is challenging due to its high thermody­
namic stability and kinetic inertness i.e. owing to its low standard heat
of formation i.e. ΔHf = − 394 kJ mol− 1 and standard Gibbs energy of
formation i.e. ΔGf = − 395 kJ/mol [5]. Due to the low reactivity of CO2,
this alternative approach of CO2 utilization as feedstock for organic Scheme 2. Synthesis of cyclic carbonates from epoxides and CO2.
synthesis requires high energy input and may result in the exhaust of
more CO2 than consumed by the reaction. Consequently, the reaction of
epoxides is a promising reaction in terms of green chemistry (Scheme 2)
CO2 should be carried out with those compounds which have relatively
[18].
high free energy to provide a thermodynamically feasible process. For
The formation of cyclic carbonate from epoxide and CO2 involves
instance, the synthesis of urea from ammonia and CO2 is a highly
fewer hazardous species, as it incorporates CO2 as a C1 feedstock source
exothermic reaction i.e. ΔHr = − 101 kJ mol− 1 [6]. Currently, the annual
rather than the conventional highly toxic and corrosive phosgene route
production of urea is around 157 Mt, corresponding to 115 Mt/year
[19]. Since CO2 is available in excess, with non-toxic and non-flammable
direct CO2 utilization [7]. Similarly, another example of effective CO2
nature, it can be used as a renewable source to produce cyclic carbon­
utilization is the production of salicylic acid by the reaction of pheno­
ates. Cyclic carbonates have a wide range of applications. These are
lates and CO2 i.e. ΔHr = − 31 kJ mol− 1, which is the first step in the
being increasingly used as ‘green’ polar aprotic solvents in the chemical
industrial production of aspirin [8]. This process was commercialized
process industry due to their high boiling point, high dipole moment,
since the 19th century [9]. Currently, the annual production of salicylic
high flash point, less toxicity and biodegradability [20]. These organic
acid is about 90 kt, which is directly consuming only 29 kt of CO2 [10].
solvents are considered as a potential replacement for traditionally used
Among these processes, synthesis of cyclic carbonate via CO2 cycload­
toxic polar aprotic solvents e.g. DMF which is likely to be banned ac­
dition to epoxide is another thermodynamically favorable reaction i.e.
cording to European REACH regulations due to its impact on human
ΔHr = − 144 kJ mol− 1 for ethylene carbonate (EC) synthesis from
health such as acute toxicity and reproductively toxicity [21]. Cyclic
ethylene oxide (EO) and CO2 [11]. The energy required for this reaction
carbonates can be used as electrolytes in the Li-ion batteries due to their
is provided by the discharge of ring-strain energy enclosed in the
high dielectric properties [22]. The demand for lithium-ion batteries has
three-membered epoxide used as a substrate.
increased rapidly due to their applications in many portable devices
Conventionally, the formation of cyclic carbonates was carried out
[23]. Cyclic carbonates can be used as monomers for polymers synthesis
by the reaction of diols with phosgene [12]. However, this reaction is
such as polycarbonate and non-isocyanate polyurethanes [24,25]. The
not environment friendly due to the toxic and corrosive nature of
use of organic cyclic carbonates as raw material to produce these
phosgene. Consequently, various alternative methods for cyclic car­
polymers have recently gained much attention in both academia and the
bonate synthesis have been proposed to replace phosgene with DMC
chemical process industry. These processes have the advantage of of­
[13], urea [14] or CO [15] as an alternative carbon source. However,
fering alternative routes for the production of polymers by avoiding the
these reactions are not feasible both ecologically and economically due
traditional use of highly volatile and toxic chemicals such as bisphenol A
to the additional steps involved in downstream processing such as sep­
(BPA) and phosgene [26]. Cyclic carbonates can also be used as in­
aration and handling of the by-products. Conversely, the synthesis of
termediates for the production of pharmaceuticals and many other
cyclic carbonate by CO2 cycloaddition to epoxides is a promising reac­
industrially important chemicals such as glycol or pyrimidines and
tion due to its 100% atom economy. Moreover, a life-cycle-assessment
carbamates [21,27].
study reveals that the amount of CO2 emitted by the traditional
Currently, the annual production of cyclic carbonates is around 80
methods of cyclic carbonate synthesis is significantly higher than the
Kt, corresponding to 40 Kt/year direct CO2 utilization [10]. However,
sustainable route of cyclic carbonate synthesis via CO2 cycloaddition to
the demand for cyclic carbonates has been increasing rapidly due to
epoxides (Scheme 1) [16,17].
their applications in Li-ion batteries that are an important part of our
modern life. Similarly, cyclic carbonates are being increasingly used as
2. Synthesis of cyclic carbonates from epoxides and CO2
intermediates for the preparation of many other useful products. For
instance, the synthesis of ethylene carbonate (EC) by CO2 cycloaddition
Synthesis of cyclic carbonates by the cycloaddition of CO2 to
to ethylene oxide (EO) and subsequent production of ethylene glycol

Scheme 1. Different routes for cyclic carbonate formation with CO2 emitted per unit of carbonate produced [16,17].

2
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

(EG) via hydrolysis has been already industrialized on large scale by be increased with an increase in CO2 pressure or a decrease in reaction
Shell Omega [28]. Similarly, the reaction between methanol and temperature. The decrease in reaction temperature can increase the CO2
ethylene carbonate to form ethylene glycol and dimethyl carbonate has solubility, but at the same time, it may decrease the solubility of the
been commercialized by Asahi Kasei [29]. Cyclic carbonates are highly catalyst and co-catalyst in the reaction mixture. Similarly, high CO2
stable compounds that can ensure long-term CO2 sequestration dissolution in the reaction mixture could also result in the decrease of
compared to other CO2 based products such as urea which readily re­ catalyst solubility and it may precipitate.
leases CO2 on its utilization as fertilizer [16].
2.2.1. Temperature
2.1. General mechanism It is very difficult to predict the influence of temperature on the
solubility of all components. However, the rate of catalytic CO2 cyclo­
The generally accepted reaction mechanism of cyclic carbonate addition to epoxides generally increases with temperature due to an
synthesis using an acid-base catalyst involves the following steps in increase in catalyst activity at a higher temperature. Moreover, the mass
Scheme 3 [21,30] Step 1) Epoxide activation by the interaction of Lewis transfer coefficient in a gas-liquid reaction can also be influenced by the
acid (A) with an oxygen atom. Step 2) Ring-opening of activated epoxide gas diffusivity in the liquid (DL) and the liquid viscosity (μL) [31].
by the nucleophilic attack of halide anion (X− ) on the least hindered C Increasing the temperature increases DL and decreases μL, both of which
atom of the epoxide to form an alkoxide intermediate. The stability of act to increase the gas-liquid mass transfer coefficient. Similarly, an
this ring-opened epoxide was provided by the counter cation. Step 3) increase in the reaction temperature can enhance the selectivity towards
Subsequently, CO2-insertion takes place by the nucleophilic attack of cyclic carbonates as these are thermodynamically benign products for a
negatively charged oxygen of alkoxide on an electrophilic carbon atom reaction between epoxide and CO2 [32]. The activation energy required
of CO2 to form a carbonate intermediate. CO2 is a linear apolar molecule for cyclic carbonate synthesis is significantly higher than for poly­
with two polar C˭O bonds resulting in a partial positive and partial carbonate formation [33]. However, the increase in temperature above a
negative charge on carbon and oxygen atoms, respectively. Thus, acti­ certain limit (i.e. >120 ◦ C for PC synthesis) may accelerate the side
vation of CO2 can be carried out by both nucleophilic and electrophilic reactions such as isomerization to acetone and hydrolysis to diols,
attack. Step 4) The resulting open-chain carbonate further undergoes causing a significant decrease in product selectivity [34–37].
intramolecular cyclic elimination (back-biting reaction). Step 5) Finally,
a five-membered cyclic carbonate is formed by the displacement of the 2.2.2. CO2 pressure
X− and the catalyst is regenerated. The nucleophile (halide anion) pro­ The effect of CO2 pressure on cyclic carbonate synthesis catalyzed by
vided by the catalyst should have good nucleophilicity and leaving various homogeneous and heterogeneous catalyst systems was exten­
group abilities so that it helps to open the epoxide ring and finally be sively studied resulting in an increase in reaction rate with the increase
displaced to allow cyclic carbonate formation. Similarly, to facilitate the in CO2 pressure up to a certain limit. However, a further increase in CO2
CO2 insertion, the cation should also have favorable interactions with an pressure after a certain limit (i.e. ρc = 0.47 g mL− 1) where the CO2
alkoxide to provide greater stability to a ring-opened epoxide [17]. insertion is not a rate-limiting step anymore, a sudden decrease in re­
action rate was commonly observed due to the dilution effect causing a
2.2. Reaction conditions decrease in epoxide and catalyst concentration in the reaction mixture
[38]. Moreover, an increase in the catalytic activity of longer alkyl chain
Synthesis of cyclic carbonates via CO2 cycloaddition to epoxides is ionic liquids for cyclic carbonate formation under supercritical CO2
strongly influenced by the reaction conditions. To achieve the high ef­ conditions was also reported [39]. The use of CO2 under supercritical
ficiency of any catalytic system, the maximum contact between all the conditions has the advantage of providing a homogeneous CO2-rich gas
components of a reaction mixture is highly desirable under the given phase which maximizes the contact between all the components of the
reaction conditions. For instance, solvent-free cyclic carbonate synthesis reaction mixture. This change in phase behavior for cyclic carbonate
catalyzed by a metal complex in combination with nucleophile additive synthesis was investigated using high-pressure reactors containing a
requires a uniform reaction mixture of all the components for high viewing window. Moreover, high CO2 pressure requires strict demands
catalytic performance. However, this can be a challenging task due to for the equipment and reduces substrate concentration causing a
the different range of polarities of all the species in the reaction mixture. decrease in reaction rate [36]. The use of too high CO2 pressure may also
Reaction conditions can affect CO2 dissolution in the reaction mixture slow down the interaction between the epoxide and catalyst due to a
due to a change in the density of the epoxide. The density of epoxide can decrease in the concentration of epoxide in the vicinity of the catalyst
[34,40]. A sharp decrease in the product yield due to the CO2-induced
expansion of the reaction mixture was reported for CO2 cycloaddition to
epoxides at higher pressure (>50 bar) [41,42]. The decrease in the
product yield due to an increase in CO2 pressure beyond the optimal
level was also reported by other authors [43,44].

2.2.3. Co-catalyst
The use of co-catalysts has been extensively studied in cyclic car­
bonate synthesis. Commonly used catalysts/co-catalyst for this reaction
are nucleophile additives such as quaternary ammonium and phospho­
nium salts and other Lewis bases as shown in (Scheme 4) [33].
The role of co-catalyst during cyclic carbonate formation is to pro­
vide a nucleophile (mostly halide anion) required to open the ring of the
epoxide. Moreover, co-catalyst is an essential part of the non-
nucleophilic catalyst systems e.g. Zn salts and Zn-salphen complexes
which show no catalytic activity in the presence of catalyst alone [45].
The use of TBAX as a co-catalyst exhibited variable catalytic activities
depending on the nature of the catalyst and reaction conditions. For
Scheme 3. General mechanism of cyclic carbonate synthesis using an acid-base example, the order of catalytic activity of TBAX in combination with
catalyst system [17,30]. bimetallic Al(III)-salen complexes was observed as Br− > I− > Cl− > F−

3
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

co-catalyst (nucleophile) helps to displace the metal-bond carbonate


intermediate favoring back-biting reaction to form cyclic carbonates
[51].

2.2.4. Solvent
The use of a reaction solvent is generally avoided for cyclic carbonate
synthesis if the catalyst is soluble in the epoxide under the reaction
conditions. However, in some cases, a reaction solvent is required to
avoid mass transfer limitations if the catalyst or co-catalyst is not
completely soluble in the given concentration of epoxide. However, the
dilution of the reaction mixture may also affect the rate of reaction due
to poor contact between the components of the reaction mixture. For
instance, the catalytic activity of the Zn-salphen complex in combination
with TBAI was higher under solvent-free conditions than using
dichloromethane (CH2Cl2) as reaction solvent under the same reaction
conditions [52].

Scheme 4. Structures of some commonly used co-catalysts for cyclic carbonate 2.3. Stereochemical information
and polycarbonate synthesis [33].
Synthesis of five-membered cyclic carbonates with high stereo­
selectivity has recently gained much attention [53]. The use of enan­
which was suggested as a good balance between nucleophilicity and
tiopure cyclic carbonates as intermediates are of high commercial
leaving group ability of halide anions [46]. Similarly, when TBAX
importance due to their versatile reactivity. These can be used as sub­
immobilized over chitosan supporting material was used as a hetero­
strates for the synthesis of optically pure cis-diols which are of particular
geneous catalyst for CO2 cycloaddition to terminal epoxides, the order of
interest in pharmaceutical applications [54]. For instance, the use of the
halide anion activity was observed to be I− > Br− > Cl− which is
drug in the enantiopure form can avoid unexpected side effects of
consistent with leaving group abilities of the halide anions [37]. Simi­
consuming undesired isomer [55]. Similarly, cyclic carbonates with
larly, the change in halide anion activity was also observed by changing
high trans-isomers have also shown high reactivity and better properties
the type of epoxide. The order of halide anion activity for terminal ep­
of limonene-based non-isocyanate polyurethane (NIPU) formation [56].
oxides was found to be consistent with the order of leaving group ability
Synthesis of enantiopure cyclic carbonates from low-cost racemic ep­
of the halide anions i.e. I− > Br− > Cl− > F− . However, the use of Br−
oxides with a high kinetic resolution is highly desirable from the com­
has shown higher catalytic activity than I− under the same reaction
mercial point of view. In this system, the use of a chiral complex (Lewis
conditions in the case of sterically hindered internal epoxides. This in­
acid) selectively reacts with one of the enantiomers of the racemic
crease in activity was attributed due to the smaller size of Br− than I−
oxirane. Subsequently, the activated epoxide undergoes
[47].
Synthesis of PC from PO and CO2 in the presence of TBAX as co-
catalysts has shown a change in selectivity by changing the halide
anion (X‾) [48]. This was observed due to the replacement of higher
leaving group ability anions such as I− and Br− with poor leaving group
ability anions such as Cl− , F− and CH3COO− which favors the poly­
carbonate formation by avoiding the back-biting reaction required to
close the ring of the carbonate intermediate. However, the reaction rate
can be decreased significantly using poor leaving group ability anions
due to reduced catalytic activity.
Similarly, the use of bis(triphenylphosphoranylide) ammonium ha­
lides (PPNX) as co-catalyst was preferred for polycarbonate synthesis.
For instance, PPNCl has poor leaving group ability, thus promotes pol­
ycarbonate synthesis by avoiding back-biting reaction to close the ring
of carbonate intermediate [49]. Moreover, in the case of the binary
catalyst system, the molar ratio between catalyst and co-catalyst, type of
co-catalyst and reaction conditions can also play an important role to
decide the selectivity of reaction between epoxide and CO2. For
instance, CO2 cycloaddition to cyclohexene oxide (CHO) catalyzed by Fe
(III) complex (0.1 mol%) without the use of co-catalyst favors the for­
mation of poly(cyclohexene) carbonate (PCHC) at high p (CO2) (10 bar).
However, the use of higher loading of Fe (III) complex (1 mol%) in
combination with PPNCl (2 mol%) as nucleophile additive resulted in
the switch of product selectivity from polycarbonate to cyclic carbonate
and cyclohexene carbonate (CHC) was achieved with high selectivity at
low p (CO2) (1 bar) [50]. Similarly, the reaction carried out in the
presence of Fe (III) amino triphenolate complex in combination with
PPNCl using the equimolar ratio (0.5/0.5 mol%) resulted in complete
selectivity towards PCHC at 85 ◦ C, 80 bar p (CO2) after 3 h. However,
the change in the molar ratio between catalyst and co-catalyst
(0.5/5 mol%) has switched the selectivity towards cyclic carbonates Scheme 5. Epoxide ring-opening positions in case of terminal epoxide and
under the same reaction conditions. The use of the higher molar ratio of possible back-biting reactions resulting in a cyclic carbonate formation [33].

4
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

enantioselective, nucleophilic epoxide ring-opening by the attack of cycloaddition to sterically congested di- and tri-substituted internal
halide anion or activated CO2 to form the corresponding cyclic car­ epoxides with high stereoselectivity is still challenging. The cycloaddi­
bonate in enantiomerically pure form [57]. In the case of CO2 cyclo­ tion of CO2 to internal epoxides is of particular interest, as they can be
addition to terminal epoxides, the epoxide ring-opening was mostly derived from naturally occurring renewable sources, such as bio-based
carried out by the nucleophilic attack on the least sterically hindered β olefinic compounds extracted from terpenes and unsaturated fatty
(methine) carbon atom of the epoxide due to its high accessibility acids [61,62]. These bio-based olefinic compounds can be easily con­
(Scheme 5) [33]. verted to corresponding epoxides using standard epoxidation methods.
Similarly, the nature of the groups attached to epoxides may also In particular, cyclic carbonates formation from limonene epoxides i.e.
affect the selectivity of product formation. For instance, Styrene oxide limonene oxide (LO) and limonene dioxide (LDO) is of significant in­
(SO) has more affinity to make cyclic carbonates than CHO due to its terest as these can be used as a non-toxic monomer for the production of
electron-withdrawing nature which favors the back-biting reaction fully bio-based polymers [63,64]. Similarly, the by-products of
required for cyclic carbonate synthesis [58]. Although β-carbon is the bio-refineries can also be used as raw materials. One example is 1,
most favorable position for ring-opening in the case of terminal epox­ 4-cyclohexadiene, derived from unsaturated fatty acids, which can be
ides, the nucleophilic attack can also take place on the α-carbon converted into mono- or bis-epoxides to give sustainable CHO. Synthesis
(methylene) as suggested in the case of terminal epoxides having of CHC is generally considered to be more challenging than poly­
electron-withdrawing groups [49]. Moreover, the stereochemical in­ carbonate synthesis due to the strain offered by the CHO ring to
formation may partially change if the attack of the nucleophile is on the five-membered cyclic carbonate ring and high activation energy barrier
α-position [59]. i.e. >80 kJ mol− 1 compared to poly(cyclohexene) carbonate (PCHC)
The cycloaddition of CO2 to di-substituted epoxides with high ster­ [65].
eoselectivity can be achieved using an appropriate metal-based catalyst Some catalyst systems for CO2 coupling with internal epoxides have
system in combination with nucleophile additive. CO2 cycloaddition to a been reported in the literature. Taherimehr et al. [39] stated synthesis of
2,3-disubstituted epoxide (a mixture of cis- and trans-isomers) with CHC using Zn(II) salphen complex (0.05 mol%) in combination with
formal retention of configuration via a double inversion mechanism can TBAI (0.25 mol%) under supercritical conditions where all the reactants
be shown as Scheme 6 [60]. coexist in a single CO2-rich phase resulting in a 38% yield at 80 ◦ C,
Here, activation of epoxide takes place by the interaction of metal 80 bar p (CO2) after 5 h. Similarly, Buchard et al. [50] reported a
(Lewis acid) with the oxygen atom of an epoxide. The activated epoxide bimetallic Fe (III) complex (1 mol%) in combination with PPNCl (2 mol
undergoes a nucleophilic attack by the halide anion to form a metal- %), that can selectively transform CHO to CHC (90% conversion) at
bonded alkoxide intermediate causing an inversion of configuration at 80 ◦ C, 1 bar p (CO2) after 24 h. Interestingly, the selectivity can be
this C center. Subsequently, the CO2 insertion takes place by the switched to PCHC synthesis (70% conversion) using 0.1 mol% of the
nucleophilic attack of alkoxide on the electrophilic C atom of CO2 to complex without any co-catalyst at 80 ◦ C and higher p (CO2) i.e. 10 bar
form a carbonate intermediate. The resulting carbonate further un­ after 24 h (Scheme 7).
dergoes ring closure by intramolecular disassociation of halide anion by Moreover, in the case of CO2 cycloaddition to internal epoxides, the
SN1 mechanism resulting inversion of configuration again on the same C use of TBAB as a nucleophile additive has been preferred over TBAI due
center. Finally, a cyclic carbonate is formed with formal retention of to the reduced size of Br− compared to I− . Notably, MEK was used as a
configuration. reaction solvent in the case of both Fe (III) trisphenolate and Al (III)
triphenolate complexes as catalysts. Some other studies related to cyclic
2.4. Cyclic carbonate synthesis from internal epoxides carbonate synthesis from internal epoxides such as 1,2-dimethyloxirane
and 1,2-diphenyloxirane using various homogeneous catalyst systems
Although CO2 cycloaddition to terminal epoxides has been exten­ have been summarized in Table 1.
sively studied in the presence of various catalyst systems, the CO2
2.5. Cyclic carbonates synthesis from bio-based epoxides

Synthesis of cyclic carbonates is mostly carried out with


petrochemical-based terminal epoxides such as ethylene oxide (EO),
propylene oxide (PO) and styrene oxide (SO). Nevertheless, to satisfy the
demands of green chemistry, recent efforts have been focused on the
utilization of renewable resources as potential raw materials instead of
using crude oil as feedstock. The use of petroleum-based cyclic car­
bonate as a monomer for the production of polymers is also associated
with various environmental and health issues [26]. Moreover, the
growing demand for polymers in modern society is under threat due to
depleting global reserves of fossil fuel, environmental issues, and supply
security risks as many of the main oil producers are in politically un­
stable regions. There is a requirement to seek sustainable substitutes for
petroleum-derived polymers with naturally available feedstocks, espe­
cially from waste biomass.
Naturally occurring terpenes have been identified as a key starting
material for the production of bio-based epoxides [73]. Terpenes,
derived from ‘Turpentine’, are naturally occurring compounds. They are
mostly produced by coniferous plants. Terpenes are mainly cyclic
compounds containing highly unsaturated hydrocarbons with one or
more C–C double bonds. Terpenes are readily available at a large scale,
with worldwide production of about 350,000 tonnes/year [74]. Among
these unsaturated hydrocarbon monoterpenes (C10H16), D-limonene has
Scheme 6. General reaction mechanism for CO2 cycloaddition to 2,3-disubsti­ been extensively studied as a possible renewable platform chemical to
tuted epoxide with formal retention of configuration [60]. produce biopolymers. D-limonene is a monocyclic unsaturated terpene,

5
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Scheme 7. Switch in selectivity towards CHC and PCHC using Fe (III) complex under different reaction conditions [66].

Table 1
Summary of catalytic systems used for CO2 cycloaddition to internal epoxides.
Co-catalyst (mol
Epoxide Catalyst (mol%) Solvent Reaction conditions Conversion (%) Ref.
%)

85 ◦ C
Fe (III) amino (trisphenolate) complex Methyl Ethyl Ketone
1,2-Dimethyloxirane TBAB (5) 2 bar p (CO2) and 53 [47]
(0.5) (MEK)
18 h
85 ◦ C
Fe (III) amino (trisphenolate) complex
2,3-Dimethyloxirane TBAB (5) MEK 10 bar p (CO2) and 83 [67]
(0.5)
18 h
85 ◦ C,
1,2-Diphenyloxirane Al (III) salen complex (2.5) TBAB (2.5) No solvent 10 bar p (CO2) and > 99 [68]
72 h
85 ◦ C,
1,2-Dimethyloxirane Al (III) salen complex (2.5) TBAB (2.5) No solvent 10 bar p (CO2) and 49 [68]
24 h
90 ◦ C,
1,2-Dimethyloxirane Al (III) amino triphenolate complex (0.5) TBAB (5) MEK 10 bar p (CO2) and 99 [69]
42 h
90 ◦ C,
1,2-Diphenyloxirane Al (III) amino triphenolate complex (0.5) TBAB (5) MEK 10 bar p (CO2) and 82 [69]
42 h
25 ◦ C,
1,2-Dimethyloxirane Cu (II) complex (2.5) TBAB (5) No solvent 1 bar p (CO2) and 36 [70]
24 h
25 ◦ C,
1,2-Dimethyloxirane Zn (II) complex (2.5) TBAB (5) No solvent 1 bar p (CO2) and 42 [70]
24 h
150 ◦ C,
Co (III) based bisamino-bisamide complex
Cyclo(octane) oxide DMAP (0.002) No solvent 20 bar p (CO2) and 83 [71]
(0.001)
5h
85 ◦ C,
1,2-Dimethyloxirane (22:78 cis/ 66 (19:81 cis/
Polyphenolate complex (0.2) TBAB (0.8) No solvent 10 bar p (CO2) and [72]
trans) trans)
45 h

mainly extracted from the peel of citrus fruits (90 wt%) [75,76]. The 3 mol% TBAB at 140 ◦ C and 30 bar p (CO2) after 55 h. The reaction was
global production of D-limonene (also known as (+)-Limonene) was also performed using SiO2-supported pyrrolidinopyridinium iodide
approximately 70,000 tonnes, which is gradually increasing every year (SiO2-PPI) as a bifunctional heterogeneous catalyst. However, the het­
[77]. Due to its abundance as a waste by-product and suitability for erogeneous catalyst has exhibited a significantly lower reaction rate
organic synthesis due to the presence of two double bonds, it can be used than TBAB i.e., 78% epoxy group conversion after 120 h at the same
as a sustainable replacement for petrol-based epoxides without reaction conditions. The decrease in reaction rate was due to the steric
competing with food crops. The synthesis of cyclic carbonates from hindrance of LDO which shows less conversion in the presence of a bulky
D-limonene oxide, in particular, is of significant interest, as it can be used heterogeneous catalyst. LDC obtained from cycloaddition reaction was
as a bio-renewable monomer for the production of fully bio-based further used as a monomer for the synthesis of NIPUs by the reaction
polymers such as non-isocyanate polyurethanes (NIPUs), which have with polyfunctional amines (Scheme 8). The resulting NIPU has shown
potential applications as thermoset materials, elastomers, or thermo­ good mechanical and thermal properties. This novel approach to pro­
plastics [63,78]. The use of bio-based epoxides derived from renewable duce cyclic carbonates not only provides a substitute for the traditional
resources such as waste biomass and CO2 can provide a sustainable basis phosgene-based toxic isocyanate monomers with ‘green’ renewable
for the future polymer industry [79]. monomers but also reduces the strong reliability of petroleum-based raw
Synthesis of cyclic carbonates from bio-based epoxides is a promising materials [63].
approach for the sustainable production of cyclic carbonates. The utili­ Another study from the same group reported cyclic carbonate syn­
zation of CO2 and waste biomass for valuable products is highly desir­ thesis from bio-based epoxides derived from linseed oil (ELSO) and
able as part of a sustainable future of the chemical process industry [76, soybean oil (ESBO) [81]. The reaction was carried out using both ho­
80]. Bähr et al. reported the synthesis of limonene dicarbonate (LDC) mogeneous TBAB and heterogeneous SiO2-PPI catalysts. Again, the
from commercially available limonene dioxide (LDO) and CO2 catalyzed higher reaction rate was observed using homogeneous TBAB (3 mol%)
by TBAB alone resulting in almost complete conversion of LDO using resulting in complete conversion of ELSO at 140 ◦ C, 30 bar p (CO2) after

6
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Scheme 8. Reaction scheme for terpene based NIPUs formation from limonene [63].

20 h than 45 h using SiO2-PPI under the same reaction conditions.


Similar results have been also reported for CO2 coupling to ESBO. The
cyclic carbonates obtained from these reactions were further used to
produce NIPUs.
Fiorani et al. [82] reported bio-based cyclic carbonate synthesis from
terpene-based epoxides and CO2 in the presence of Al (III) triphenolate
complex in conjunction with bis(triphenylphosphine)iminium chloride
(PPNCl) as an active binary homogeneous catalyst system. For instance,
the synthesis of limonene carbonates (LC) from LO and CO2 was suc­
cessfully carried out using 1 mol% Al (III) triphenolate complex and
3 mol% PPNCl. Interestingly, the same catalyst system was used for poly
(limonene) carbonate (PLC) synthesis at low temperature (40–70) ◦ C
with no cyclic carbonate as a by-product. However, a further rise in
temperature up to 85 ◦ C resulted in a change of catalyst selectivity and
cyclic carbonate was found as the only product. These findings were
consistent with the previously reported DFT study describing the for­
mation of LC as a thermodynamically more favorable product due to a
higher kinetic energy barrier than PLC synthesis. This catalyst system
exhibited a 57% yield of LC using trans-limonene oxide as a substrate at Scheme 9. Proposed reaction mechanism of cyclic carbonate synthesis from
85 ◦ C and 10 bar p (CO2) after 66 h. The requirement of long reaction tri-substituted epoxide and CO2 [85].
time was due to the slow rate of reaction caused by the sterically hin­
dered nature of tri-substituted internal epoxides. Similarly, the catalyst following the reported protocols to achieve a high yield of the more
has shown a 27% isolated yield of limonene dicarbonate from limonene reactive trans-isomer (87 ± 2%). Moreover, a detailed kinetic study of
dioxide at 120 ◦ C and otherwise constant reaction conditions. This re­ CO2 cycloaddition to LO was carried out in the presence of propylene
action was performed at a higher temperature to enhance the carbonate as a greener polar aprotic solvent. The results indicate a
mass-transfer between CO2 and highly viscous limonene dioxide. first-order dependence of the reaction with respect to LO, TBAC and CO2
Moreover, a double inversion reaction mechanism was proposed concentrations. Furthermore, a general reaction mechanism of cyclic
explaining the complete retention of stereochemical information ob­ carbonate synthesis from tri-substituted epoxides and carbon dioxide
tained from experimental observation [82]. (CO2) in the presence of a TBAC catalyst was proposed (Scheme 9).
Martínez et al. [83] reported lanthanum heteroscorpionate based The halide anion (Cl− ) of tetrabutylammonium chloride (TBAC)
complex in conjunction with TBAX (1:4 molar ratio) as an effective undergoes a nucleophilic attack on the least hindered carbon atom to
catalyst system for LC synthesis via cycloaddition reaction of CO2 and ring-open the epoxide (Step 1). In the absence of the Lewis acid catalyst,
limonene oxide under solvent-free conditions. The reaction was carried there was no partial positive charge on any carbon atom and the 2 C
out using commercially available limonene oxide (40:60 cis/­ atoms in the epoxide ring have the same electrophilicity. The nucleo­
trans-isomer) at 100 ◦ C and 10 bar p (CO2) resulting in a 43% yield of LC philic attack takes place on the least sterically hindered carbon atom due
after 16 h. Similarly, the synthesis of LDC from LDO and CO2 was also to its higher accessibility. This results in the formation of a highly
carried resulting in 69% product yield at the same reaction conditions. reactive chloro-alkoxide intermediate which is stabilized by the counter
Hiroshi, Masato, Yuuta, Jun-ichi, Hisatoyo and Suguru [84] reported cations (TBA+) (Step 2). This ring-opened epoxide further coordinates
CO2 cycloaddition to LO catalyze by commercially available inexpensive with a carbon atom of CO2 to give carbonate anion (Step 3). Finally, the
tetrabutylammonium chloride (TBAC). The formation of cyclic carbon­ carbonate anion goes through intramolecular cyclic elimination to give a
ates was reported to be highly stereoselective, suggesting high reactivity five-membered cyclic carbonate (cis and trans-isomer) and the catalyst is
of trans-isomer than cis-isomer. Hence, commercial LO which is a regenerated (Step 4). The reported observations agree with the accepted
mixture of cis and trans-LO (40:60), is not economically feasible for this mechanism for cyclic carbonate formation and highlight the importance
reaction as it contains approximately 40% less reactive cis-isomer. of small nucleophiles like chloride when such reactions are performed
Recently, Rehman et al. [85] reported a 100% selective synthesis of with highly substituted epoxides. Moreover, the overall stereochemical
sustainable cyclic carbonates from bio-based LO and CO2 was carried information was retained due to two consecutive SN2 reactions at the
out using a commercially available, inexpensive TBAC as an effective same carbon center as shown in the ring-opening and ring closure steps
homogeneous catalyst. The initial studies of cycloaddition from a of the mechanism. This retention in the overall configuration was sug­
commercially available LO (a mixture of cis and trans-LO) revealed that gested for cyclic carbonate synthesis from internal epoxide using a
the reaction was stereoselective and the trans-isomer was found to be relatively high concentration of halide anion in the reaction mixture
more reactive than the cis-isomer in line with the previous report [86]. A [53].
stereoselective method of (R)-(+)-limonene epoxidation was carried out

7
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

2.6. Kinetic study ( ) ( )


k ΔH‡ ΔS‡ kB
ln = − + + ln (10)
To investigate the reaction mechanism involved in cyclic carbonate T RT R h
synthesis and to determine the general equation of the rate law, several
kinetic studies of cyclic carbonate synthesis by CO2 cycloaddition to ΔG‡ = ΔH‡ − TΔS‡ (11)
epoxide have been reported [87]. The general rate equation of cyclic
Where
carbonate synthesis in the presence of homogeneous catalysts can be
R = Universal gas constant (8.314 J mol− 1 K− 1).
written as Eq. (1), where [Epoxide], [CO2], [Catalyst] are the concen­
h = Planck’s constant (6.62608 × 1034 J s).
trations of epoxide, carbon dioxide and catalyst, respectively, and the
kB = Boltzmann constant (1.38065 × 10− 23 J K− 1).
superscripts a, b, c are the orders of reaction with respect to the epoxide,
Clegg et al. [46] reported a detailed kinetic study of SC synthesis
CO2 and catalyst. Literature precedent shows that cyclic carbonate for­
from SO and CO2 in the presence of bimetallic Al (III) salen complex and
mation via CO2 cycloaddition to epoxides has a high negative heat of
TBAB as the binary homogeneous catalyst system. The reaction has
formation, so it can be assumed that there is no back reaction [88].
shown first-order dependence on the epoxide, CO2 and salen complex
Moreover, the concentration of CO2 in the liquid phase remains constant
concentrations. However, an unexpected second-order reaction for
throughout the reaction due to a continuous supply of CO2. Similarly,
TBAB was observed suggesting a dual role of TBAB in the catalytic cycle
the concentration of the catalytic species also remains constant. Hence,
as shown in the reaction mechanism (Scheme 10).
Eq. (1) can be simplified to Eq. (2), where kobs is the observed
According to this mechanism, the activation of the epoxide was
pseudo-first-order rate constant and can be given as Eq. (3). Taking the
carried out by one of the metal centers due to its Lewis acidic properties
natural logarithm of both sides of Eq. (3) gives Eq. (4), which can be
(Step 1). The activated epoxide undergoes a nucleophilic attack by the
used to determine the orders of reaction by changing the concentration
Br− anion provided by the TBAB (first role) to form a metal-bound
of catalyst and CO2. By assuming the pseudo-first-order dependence of
alkoxide intermediate (Step 2). The in-situ tributylamine formation by
the reaction rate on epoxide concentration (a=1), the differential rate
the decomposition of TBAB (second role) was observed leading to CO2
law gives Eq. (5). Integrating both sides of Eq. (5) gives Eq. (6), which
activation by carbamate formation which further coordinates with one
can be used to determine the observed pseudo-first-order rate contestant
of the metal centers of the dimeric Al (III) complex at a much faster rate
for epoxide conversion.
than with CO2 only (Step 3). This results in the formation of two reaction
Rate = k[Epoxide]a [CO2 ]b [Catalyst]c (1) intermediates on the same side of the salen complex which easily
combines to form the corresponding carbonate (Step 4). Finally, the
Rate = kobs [epoxide]a (2) cyclic carbonate was formed, and the catalyst was regenerated (Step 5).
The inactivity of the mono-metallic salen complex under the same re­
Where action conditions also validates the participation of the two metallic
centers in the reaction mechanism. Al (III) complex has shown
kobs = k [CO2 ]b [catalyst]c (3)
remarkably high catalytic stability and can be recycled up to 60 times
without a decrease in catalytic activity. However, TBAB needs to be
ln kobs = ln k + b ln[CO2 ] + c ln [Catalyst] (4)
replaced during recycling experiments. The effect of halide anion on the
d[epoxide] catalytic activity was also examined. As a result, the order of halide
Rate = − = kobs [epoxide]1 (5) anion activity was found to be Br− > I− > Cl− > F− , which was
dt
described due to a balance between the leaving group ability and
− ln [epoxide] = kobs .t (6) nucleophilicity of halide anions.
Later on, Styring et al. Supasitmongkol and Styring [89] reported the
The temperature dependence of the reaction was studied to deter­
synthesis of cyclic carbonates using Al (III) salenac complex in
mine the activation and thermodynamic parameters of cyclic carbonate
conjunction with TBAB as an active binary homogenous catalyst. The
synthesis by CO2 cycloaddition to epoxide. The Ea for cyclic carbonate
effect of a change in temperature on the rate constants was studied by
formation was determined based on the Arrhenius equation Eq. (7). The
performing the experiments over the range of (80–150) ◦ C. The Ea for SC
Ea was calculated from the gradient of the graph between ln (kobs) and
formation catalyzed by Al (III)-salenac complex alone was determined to
reciprocal of absolute temperature (1/T) Eq. (8), where kobs is the
be 34 kJ mol− 1 which was further reduced to 23 kJ mol− 1 using TBAB as
observed pseudo-first-order rate constant (min− 1), A is the pre-
co-catalyst.
exponential factor (min− 1), R is the ideal gas law constant
Luo et al. [90] reported a kinetic study of SC synthesis from SO and
(8.314 J mol− 1 K− 1), T is the absolute temperature (K) [87]. Similarly,
CO2 using ionic liquid ([BMIm]Br) in combination with graphite oxide
the thermodynamic activation parameters such as enthalpy of activation
(GO) as a binary homogeneous catalyst. The experiments of the kinetic
(ΔH‡), the entropy of activation (ΔS‡) and Gibbs free energy of activa­
study were performed in the presence of the PC as the reaction solvent.
tion (ΔG‡) were determined from Eyring equation (Eq. 9), and the ΔH‡
( ) The progress of the reaction was monitored by in-situ IR spectroscopy.
and ΔS‡ were determined from gradient − ΔH and y-intercept (ΔSR + The effect of reaction temperature was studied over the range of
‡ ‡
RT
(75–95) ◦ C and Ea for SC synthesis was determined using the Arrhenius
ln khB ), respectively Eq. (10). Moreover, ΔG‡ was determined from the equation. The Ea for SC synthesis was determined to be 80.7 kJ mol− 1
fundamental thermodynamic equation Eq. (11) for all temperatures. using (BMIm)Br as catalyst alone. However, it was reduced by
(
Ea
) 25.4 kJ mol− 1 using GO as a co-catalyst. Moreover, the order of reaction
kobs = A. exp − (7) w.r.t epoxide (SO) and catalyst (BMIm)Br were determined to be first
R.T
order.
( ) Cuesta-Aluja et al. [91] reported CO2 cycloaddition to terminal ep­
Ea
lnkobs = lnA − (8) oxides at 80 ◦ C and 10 bar p (CO2) using Al (III)-salabza complexes in
R.T
conjunction with TBAB as a binary homogeneous catalyst. Moreover, a
kB T
(
ΔG‡
)
kB T
(
ΔH‡ ΔS‡
) detailed study of reaction kinetics was performed suggesting first-order
k= exp − = exp − + (9) reaction dependence on epoxide, CO2, TBAB and Al-salabza complex
h RT h RT R
concentrations. Moreover, the Ea for SC formation was determined to be
38 kJ mol− 1 by performing experiments over the range of (40–100) ◦ C

8
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Scheme 10. Reaction mechanism for CO2 cycloaddition to epoxides catalyzed by bimetallic Al (III) salen complex in combination with TBAB [46].

and 10 bar p (CO2). easily scale-up and process optimization [94]. Moreover, the ability to
Castro-Osma et al.[92] reported a kinetic study of SC synthesis from straightforward scale-up using multiple flow devices is an attractive
SO and CO2 catalyzed by Cr (salophen) complex in combination with approach for organic synthesis. The high surface area to volume ratio in
TBAB which was performed at solvent-free conditions. The conversion flow reactors compared to batch reactors also helps to remove the heat
of SO to SC was determined by taking the sample from the reaction generated through an exothermic reaction to avoid thermal runaway.
mixture after a regular interval of 30 min followed by high-performance These reactions can be performed using solvent-free high concentrations
liquid chromatography (HPLC) analysis. The order of reaction for TBAB of reactants under continuous flow conditions [95]. These advantages of
was determined to be first order by experimenting with the range flow chemistry enable us to achieve the goals of green chemistry and
(175–437) mM TBAB concentrations. The order of the reaction with sustainability in the chemical process industry [96].
respect to Cr (salphen) complex in combination with TBAB was deter­ Flow reactors provide high interfacial surface areas which are
mined to be zero-order under the conditions when [TBAB] < [Complex] essential for high rates of mass transfer in the case of gas-liquid re­
and first order when [TBAB] > [Complex]. Moreover, the reaction actions. The use of microchannel gas-liquid reactors has shown a
mechanism was proposed suggesting the formation of six-coordinate significantly higher interfacial area among other types of gas-liquid
anionic [Cr(salophen)-Br2]− as an active catalytic species. contactors [97]. These reactors result in a significant enhancement in
Recently, Luo et al. [93] also reported CO2 cycloaddition to epoxides the rate of mass transfer for reactions that are generally considered as
using ionic liquid-functionalized Al (III) salen oligomers as an effective non-feasible in batch reactors. Moreover, the supply of gases under
bifunctional homogeneous catalyst. The activity of the catalyst system continuous flow conditions can be easily controlled by regulating the
can be further increased by using Al (III) salen oligomers in conjunction volume and by increasing the internal pressure to help the gas dissolu­
with ionic liquid ([BMIm]Cl) as the binary catalyst system. The kinetic tion. The use of flow technology by connecting multiple reactors in se­
study of SC synthesis has been investigating using in-situ IR spectros­ ries also has the advantage to reduce the chemical storage and
copy. As a result, the reaction was found to be first order to epoxide and transportation cost. Similarly, the scale-up of flow technology enables a
bifunctional catalyst. However, the order of the reaction for the binary significant reduction in transition time by replicating the pilot plant flow
catalyst system was found to non-first order, suggesting the formation of reactors by the numbering-up method or by simply extending the
a six-coordinated transition complex in the catalytic cycle as an active running time of the same reactor without many changes in the existing
species. Moreover, the Ea for SC formation catalyzed by the binary design. The ability to easily scale-up using multiple flow devices is an
catalyst was determined to be 64.9 kJ mol− 1 over the range of attractive approach for organic synthesis [98].
(80–140) ◦ C. Synthesis of cyclic carbonates from epoxides and CO2 is a typical gas-
liquid reaction involving a gas-liquid mass transfer in a reactor and the
catalytic cycloaddition reaction in the liquid phase [99]. Previous
2.7. Flow chemistry studies on cyclic carbonate synthesis performed in batch reactors sug­
gested that the rate of reaction was controlled by the rate of CO2 mass
Flow chemistry has many advantages over conventional batch re­ transfer from the gaseous phase to the reaction mixture [100]. Thus, it
actors, typically including enhanced rates of heat and mass transfer due was anticipated that an efficient reactor design in terms of the high rate
to higher surface area to volume ratios and efficient mixing, improved of heat/mass transfer and a continuous flow approach could mitigate
safety due to less use of inventories, easy and highly reproducible many of the shortcomings observed in a traditional batch reactor.
screening of reaction parameters, which may improve the reliability of

9
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Moreover, the reactor design can affect the rate of the gas-liquid cata­ further conveyed to the next column, in which EC was hydrolyzed to get
lytic reaction [101]. ethyl glycol (EG). For a heterogeneously catalyzed process, they
North et al. [102] reported a continuous method of EC synthesis from employed a fixed bed reactor involving liquid-gas phases to contact and
EO and waste CO2 in the presence of their highly active one component react at the surface of the immobilized ILs catalyst.
SiO2-supported bifunctional Al (III) (salen) complex as a heterogeneous Synthesis of cyclic carbonate was also demonstrated in a ‘tube-in-
catalyst. The activity of the catalyst was evaluated through experiments tube’ gas-liquid continuous flow reactor. The high permeation rate of
performed in a batch reactor for SC synthesis from SO and CO2 at CO2 through the inner Teflon® AF-2400 semipermeable tube combined
ambient temperature (25 ◦ C) and 1 bar p (CO2). The recyclability of the with the high surface area to volume ratio resulted in complete con­
catalyst was also tested for up to 32 cycles. The results have shown the version (100%) of SO to SC in a significantly reduced reaction time of
decrease in catalytic activity due to the dequaternisation of a quaternary 45 min. The continuous reactor generates homogeneous solutions of gas
ammonium salt, which can be reactivated by treating with benzyl bro­ in the liquid in a reliable and controlled manner, thereby facilitating
mide. However, the bimetallic Al (III) (salen) complex has shown high rapid optimization of the reaction, which can greatly enhance the effi­
stability and the catalytic activity of the complex was not affected even ciency (rate) of the reaction as compared to the conventional batch
operating at high temperatures (170 ◦ C), which shows its long life and reactor. A detailed kinetic study of styrene carbonate synthesis catalyzed
suitability for cyclic carbonate synthesis from high-temperature flue by a homogeneous ZnBr2/TBAB catalyst system was carried out to find
gases. These properties of the heterogeneous catalyst make it a highly the initial rate law equation. The activation energy of the reaction in the
desirable catalyst for cyclic carbonate synthesis under continuous flow presence of TBAB alone was calculated to be 55 kJ mol− 1, which was
conditions. reduced to 32 kJ mol− 1 when using ZnBr2 in combination with TBAB
Continuous flow reactor with a mixture of three gases (21% CO2, [106].
25% EO and 54% N2) passed through an immobilized catalyst packed
column held in a thermostatically controlled oven was applied to ach­ 3. Catalytic systems
ieve an almost quantitative conversion of EO to EC after 7 h at 60 ◦ C
[102]. In a separate study, the catalyst was also exposed to flue-gas from In the last 20 years, significant research has been focused on the
the combustion of coal for cyclic carbonate synthesis in the gas-phase development of highly efficient catalyst systems for CO2 cycloaddition
flow reactor and shown to be compatible with waste CO2 present in to epoxides to carry out the reaction under mild reaction conditions. As
the unpurified flue gas [103]. described earlier, the synthesis of cyclic carbonates from epoxides and
Zhao et al. [99] studied the use of a microreactor for the formation of CO2 was determined to be a highly exothermic reaction i.e. ΔHr
cyclic carbonates, resulting in an appreciable enhancement in produc­ = –144 kJ mol− 1 for EC synthesis [5]. However, unlike urea and sali­
tion rate than a conventional stirred tank reactor. Here, synthesis of PC cylic acid synthesis, this reaction does not occur spontaneously due to
from PO and CO2 was carried out in a microreactor using 2-hydroxye­ the high activation barrier of the uncatalyzed reaction i.e. (209–251)
thyl-tributyl ammonium bromide (HETBAB) as OH–functionalized kJ mol− 1 depending on the type of epoxide used [107]. Currently,
ionic liquid catalyst. The reaction was performed over the range of quaternary ammonium and phosphonium salts are being used as cata­
(140–190) ◦ C and 35 bar p (CO2). The high surface-to-volume ratio of lysts for commercial cyclic carbonate synthesis [92]. These catalysts
the microreactor resulted in a significant enhancement in heat and mass have relatively low catalytic activity, causing cyclic carbonate formation
transfer. The synthesis of cyclic carbonates in the microreactor resulted at elevated reaction temperature and CO2 pressure. As a result, the
in an appreciable decrease in the reaction to 14 s with >99% yield with a current commercial production of cyclic carbonate is emitting CO2 into
dramatic increase in TOF from (3000–14000) h− 1 compared to 60 h− 1 in the atmosphere rather than consuming i.e. 0.92 tonnes CO2 emitted per
a batch reactor. These results also suggested that the reaction performed each tonne of cyclic carbonate produced [16]. Consequently, a highly
in a microreactor can be intensified by increasing temperature. The in­ efficient catalyst system is required which can substantially decrease the
crease in reaction temperature has a significant effect on the activity of activation energy, allowing the reaction to take place at mild reaction
the catalyst by increasing the activation of OH and subsequently conditions. During the last two decades, significant research has been
nucleophilic attack of Br− to open the ring of the epoxide. Secondly, with conducted to develop various homogeneous and heterogeneous catalyst
the increase of temperature, the gas-liquid mass transfer coefficient systems. These mainly include organocatalysts such as ammonium salts,
increased significantly as a result of an increase in gas diffusivity in the phosphonium salts, imidazolium salts, ionic liquids and metal com­
liquid (DL) and the decrease of liquid viscosity (μL) [31]. plexes [108]. Among these metal complexes such as bimetallic Al (III)
Kozak et al. [104] reported CO2 cycloaddition to epoxides in a salen complexes in combination with nucleophilic co-catalysts have
gas-liquid continuous flow reactor catalyzed by N-bromosuccinimide attracted considerable attention due to their ability to catalyze the re­
(NBS) in conjunction with benzoyl peroxide (BPO) using DMF as the action under ambient conditions i.e. 25 ◦ C and 1 bar p (CO2) [109]. The
reaction solvent. Here, the simultaneous activation of epoxide by use of these catalysts on a commercial scale can help in the significant
bromine cation and CO2 activation by the amide provided by DMF has reduction of CO2 emissions related to cyclic carbonate synthesis.
exhibited a high cyclic carbonate yield at mild reaction conditions. A To develop a commercially viable catalytic system, the properties of
solution of epoxide and catalyst in DMF and CO2 gas stream was intro­ the catalyst system such as stability, recyclability, upscaling, ecologic
duced simultaneously under continuous flow conditions in a tubular and economic feasibility, and sustainability are important factors to
reactor, resulting in complete conversion of 1,2-epoxyoctane into cyclic evaluate the performance of a catalytic system. For homogeneous
carbonate at 120 ◦ C and 6.8 bar p (CO2) after 45 min reaction time. catalysis, the catalyst separation from the reaction mixture involved
Furthermore, a series of experiments were performed to study the ki­ different solvent extraction techniques and subsequent highly energy-
netics to investigate the mechanism involved and to determine the rate intensive distillation processes. Organic cyclic carbonates are high
of a reaction under continuous flow conditions. boiling point products e.g., PC having boiling point > 240 ◦ C, leading to
Xu et al. [105] reported EC synthesis from EO and CO2 using ILs as high energy requirements in separation processes. To make the process
catalysts using a bubbling column. To enhance the phase contact, CO2 sustainable, the use of heterogeneous catalysis with high catalytic ac­
gas was continuously injected to the bottom of the column and bubbling tivity is of high commercial importance. From the past few years, het­
up through the mixture of EO and ILs, resulting in an efficient conver­ erogeneous catalysts for cyclic carbonate synthesis have gained much
sion (99.5%) of EO into EC. The EC obtained from the first column was attention. Heterogeneous catalysts offer substantial advantages, such as

10
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

easy separation and reusability, thereby avoiding extensive use of sol­ nucleophilically on the least hindered carbon atom of the epoxide, (2)
vents involved in downstream processes [17,110]. To facilitate catalyst Epoxide ring-opening to form an alkoxide intermediate, (3) CO2-inser­
separation and to reuse the catalyst for consecutive runs, the immobi­ tion occurs in the ring-opened epoxide and (4) finally, a five-membered
lizing of active homogeneous catalysts on high surface area inorganic ring of corresponding cyclic carbonate was formed by intramolecular
supports such as silica and polymer-based materials was considered as a elimination of the halide anion due to its leaving group ability. Later on,
promising approach. To avoid leaching of the active catalytic species this mechanism was also supported by the density functional theory
and to improve the catalyst life, the grafting of the active homogeneous (DFT) [107]. The epoxide ring-opening was determined as the
catalyst over the supporting material was carried out through strong rate-determining step due to the high activation barrier. For EC syn­
covalent bonding. However, a decrease in CO2 cycloaddition reaction thesis from EO and CO2, the Ea required for epoxide ring-opening was
rate was generally observed using heterogeneous catalysts compared to determined to be 121 kJ mol− 1 using tetraethylammonium bromide
their homogeneous counterparts due to mass-transfer (diffusion) limi­ (Et4NBr) as a catalyst rather than 243 kJ mol− 1 required for the unca­
tations between reactants and active catalytic sites [111]. To minimize talyzed reaction. The catalytic activity of the TBAX varies depending on
this effect, a uniform distribution of the active catalytic species on high the nucleophilicity and leaving group ability of the halide anions. When
surface area materials is highly desirable. The supporting materials can TBAX were used as catalysts alone, the order of the reactivity of these
also play an important catalytic role by providing active catalytic sites in halide anions was observed as Cl− > Br− > I− > F− which agrees to the
combination with grafted catalytic species [112]. order of nucleophilicity of halide anions, except F− . Whereas, the less
catalytic activity of F− was explained by its poor leaving group ability
[115]. Similarly, the chain length of cations also plays an important role
3.1. Organocatalysts
in the catalytic activity of the alkylammonium halides for CO2 cyclo­
addition to terminal epoxides causing an increase in catalytic activity
The use of organocatalysts for CO2 cycloaddition to epoxides were
with the increase in the chain length. This increase in catalytic activity
extensively studied in both homogeneous and heterogeneous phases.
was explained by the weak electrostatic forces between the halide anion
Organocatalysts have advantages such as inexpensive commercial
and bulkier cation chain length leading to higher nucleophilicity of
availability, non-toxicity, and stable nature under the reaction condi­
halide anion which in turn favors the epoxide ring-opening [116].
tions [113].
Alkylammonium halides as heterogeneous catalyst prepared by
immobilizing on high surface area inorganic supports were extensively
3.1.1. Ammonium salt-based catalysts
studied. The catalysts have exhibited higher catalytic activities for CO2
Tetrabutylammonium halides (TBAX) are commonly used catalysts/
cycloaddition to terminal epoxides (Table 2). The activity of immobi­
co-catalysts for CO2 cycloaddition to epoxides due to their strong
lized alkylammonium halides was further enhanced by grafting OH-
nucleophilic and leaving group abilities. However, the use of TBAX
functionalized ammonium salts. These heterogeneous catalysts have
alone as catalysts for cyclic carbonate formation generally requires
the advantages of easy separation and good recyclability. However,
harsh reaction conditions due to the lack of an acidic group required for
cyclic carbonate formation in the presence of these catalysts still re­
the activation of an epoxide. The first use of TBAX as catalysts for CO2
quires high temperature, pressure, and catalyst loading, which can result
cycloaddition to terminal epoxides was reported by Calo et al. [114]. In
in the negative carbon footprint of the reaction. Moreover, the catalytic
this study, the synthesis of SC from SO and CO2 was carried out by two
activities were found to be significantly lower than their homogeneous
methods (a) using 10 wt% TBAI as a catalyst resulting in 80% SC yield at
analogs. Motokura et al. [117] reported SiO2-supported amino­
60 ◦ C and 1 bar p (CO2) after 22 h (b) using 1:1 wt% mixture of TBAB
pyridinium halides as a highly active catalyst for CO2 cycloaddition to
and TBAI resulting in 83% yield at 120 ◦ C and atmospheric pressure
epoxides under mild reaction conditions i.e. 100 ◦ C and 1 bar p (CO2)
after 4 h. Moreover, the reaction mechanism of cyclic carbonate syn­
(entry 5, Table 2). SiO2-supported aminopyridinium halides catalysts
thesis using TBAX catalyst was suggested for the first time (Scheme 11).
have unexpectedly exhibited higher catalytic activity than their homo­
According to this mechanism, the reaction takes place in four steps:
geneous counterparts. The enhancement in catalytic activity was
(1) The halide anion (X) provided by the catalyst attacks
observed due to the synergistic effect of the halide anions with Si–OH
(silanol) groups acting as weak acidic sites to activate epoxide. SiO2-­
supported aminopyridinium halides also have several advantages over
other alkylammonium halide-based catalysts, such as the presence of a
long chain of delocalized cations in resonance form, facile methods of
preparation and high stability in the air. Moreover, the catalyst exhibi­
ted good recyclability, with up to 3 cycles without a decrease in catalytic
activity. Two possible reaction mechanisms were suggested for styrene
carbonate (SC) from styrene oxide (SO) and CO2 in the presence of
silica-supported pyrrolidinopyridinium iodide as an acid-based hetero­
geneous catalyst [117,118]. First, the most accepted reaction mecha­
nism for epoxide/CO2 cycloaddition reaction in the presence of an
acid-base catalyst can be given as shown in Scheme 12. The reaction
initiates by the adsorption of SO over the silica surface whereas CO2 was
present in the bulk of the reaction mixture. The silanol groups (Si–OH)
present over the silica surface act as acidic sites to activate the SO by the
interaction with the oxygen atom of the epoxide. Subsequently, the
activated SO experiences a nucleophilic attack by the I− provided by the
immobilized PPI acting as the basic site of the catalyst. This results in the
formation of an iodo-alkoxide intermediate by the ring-opening of the
activated epoxide (Step 1). The synergistic role of Si–OH groups with
halide anions is already well established [117,119–121]. The significant
enhancement in reaction rate due to the synergistic effect of Si–OH
Scheme 11. General mechanism for cyclic carbonate synthesis by CO2 cyclo­ promotes epoxide ring-opening by the halide anion, suggesting
addition to epoxide catalyzed by tetrabutylammonium halides (TBAX). epoxide-opening is the rate-determining step. This intermediate further

11
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Table 2
Summary of ammonium salts as heterogeneous catalyst systems.
Sr. Yield Reusability
Catalyst Support Reaction conditions Reference
No. (%) (runs)

120 C, 45 bar p (CO2),



1 Hexaalkylguanidinium chloride (1 mol%) SiO2 99 5 [123]
4h
150 ◦ C, 80 bar p (CO2),
2 Quaternary ammonium salts (1 mol%) SiO2 97 8 [124]
6h
100 ◦ C, 80 bar p (CO2),
3 PEG-supported quaternary ammonium salts (5 mol%) PS 95 5 [125]
12 h
3-Chloro-2-(hydroxypropyl) trimethylammonium chloride 160 ◦ C, 40 bar p (CO2),
4 CS 88 5 [4]
(1.7 mol%) 6h
100 ◦ C, 1 bar p (CO2),
5 4-Pyrrolidinopyridinium iodide SiO2 89 3 [117]
20.5 h
110 ◦ C, 1 bar p (CO2),
6 Diethanolamine based ammonium halides (2 mol%) PS 96 6 [126]
6h
120 ◦ C, 12 bar p (CO2),
7 Hydroxypropyl trimethylammonium iodide (0.4 mol%) CL 88 6 [127]
6h
Nano-porous organic polymer 100 ◦ C, 1 bar p (CO2),
8 Quaternary ammonium salt 96 3 [128]
(COP-114) 24 h

Scheme 12. Proposed mechanism of SC formation based on the reaction of adsorbed SO with CO2 in the bulk stream[117,118].

coordinates with the CO2 present in the bulk of the reaction medium by [117]. The reaction progresses with the activation of adsorbed CO2 by
the nucleophilic attack of the iodo-alkoxide intermediate on the elec­ the iodide anion (I− ) to form a new ionic species [I–CO2]− , which is
trophilic C atom of CO2 to form a carbonate intermediate (Step 2). The more basic than the original halide anion (Scheme 13). The activated
increased stability of the ring-opened intermediate is required to in­ CO2 further attacks on the least hindered carbon atom of the epoxide
crease the reaction rate. The stability of this ring-opened epoxide was present in the bulk reaction medium to open the ring (Step 1). This re­
offered by the presence of large delocalized pyrrolidinopyridinium sults in the formation of a carbonate intermediate (Step 2). Finally, this
(PP+) cations to facilitate rapid CO2 insertion. Finally, this open-chain intermediate undergoes ring-closure by intramolecular cyclization and
carbonate intermediate undergoes ring-closure by the elimination of the elimination of iodide anion (I− ) (leaving group ability) leading to SC
iodide anion to form a five-membered ring of SC and the catalyst is re­ formation (Step 3). The literature precedent shows that this type of
generated in the catalytic cycle (Step 3). mechanism for CO2 cycloaddition to epoxide is most likely to occur in
The second pathway assumes the adsorption of CO2 over the catalyst the presence of a bulky anion based catalyst system where the interac­
surface whereas the epoxide (SO) remains in the bulk reaction medium tion between epoxide and anion is weak and the anion instead interacts

12
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Scheme 13. Proposed reaction mechanism of SC formation based on the reaction of adsorbed CO2 with SO in the bulk stream [117,118].

with CO2 [116]. This type of reaction pathway was also validated pre­
Table 3
viously via in-situ ATR infrared spectroscopy [122].
Summary of phosphonium salts as homogeneous catalysts.
The synergistic effect between hydroxyl and amino groups over the
Sr. Reaction Yield surface of the catalyst exhibited high catalytic activity. In addition to the
Catalyst Substrate Reference
No. conditions (%)
grafting of OH–functionalized groups, naturally occurring OH–group-
Butyl-
125 ◦ C, based biopolymers such as chitosan (CS) and cellulose (CL) were also
triphenylphosphonium
1 SO 20 bar p 83 [131] used as supporting materials for the immobilization of ammonium salts
Iodide (PPh3BuI)
(0.5 mol%)
(CO2), 1 h (entry 4 and 7, Table 2).
25 ◦ C,
2 Phosphorus ylides (4) SO 1 bar p 62 [132] 3.1.2. Phosphonium salt-based catalysts
(CO2), 6 h Phosphonium salts as catalysts in both homogeneous and heteroge­
120 ◦ C,
neous phases were also studied for cyclic carbonate formation due to
Tetraarylphosphonium 1 bar p
3 SO 91 [133] their high thermal stability [129]. Although the catalytic activity of
Salt (2) (15 mol%) (CO2),
12 h phosphonium salts was found to be identical with ammonium counter­
Bifunctional quaternary
60 ◦ C, parts, the use of phosphonium halides as catalysts was less commonly
1 bar p
4 phosphonium iodide (3) SO 92 [134] studied. The catalytic activity of phosphonium salts in conjunction with
(CO2),
(1 mol%)
24 h
Co (II) salts was even found to be higher than ammonium and
imidazolium-based ionic liquids [130]. The HBDs based bifunctional
phosphonium halides exhibited higher catalytic activity compared to

Table 4
Summary of phosphonium salt as heterogeneous catalysts.
Sr. Yield
Catalyst Substrate Reaction conditions Reusability (runs) Reference
No. (%)

100 ◦ C, 10 bar p (CO2), Fixed bed


1 Silica-supported-tributyl phosphonium iodide (1 mol%) PO >99 [135]
1h (1000 h)
Silica-supported-3-(triethoxysilyl) propyltriphenyl phosphonium bromide
2 SO 90 C, 10 bar p (CO2), 6 h

86 10 [121]
(1 mol%)
150 ◦ C, 50 bar p (CO2),
3 Polymer-supported-tributyl phosphonium chloride (0.08 g) SO 96 5 [43]
10 h
Cr-MIL-101-supported-4-(bromobutyl)triphenylphosphonium bromide 120 ◦ C, 20 bar p (CO2),
4 SO 69 5 [136]
(0.045 mol%) 2h

13
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Scheme 14. Proposed reaction mechanism for cyclic carbonate synthesis using SiO2-supported phosphonium salt as a heterogeneous catalyst [135].

monofunctional counterparts due to their ability of epoxide activation. heterogeneous catalysts under mild reaction conditions i.e. 100 ◦ C,
The use of phosphonium based homogeneous catalysts for CO2 cyclo­ 10 bar p (CO2) [135]. The sufficient increase in catalytic activity was
addition to epoxides is summarized in Table 3.. due to the synergistic effect between phosphonium halides and silanol

Similarly, cyclic carbonates synthesis catalyzed by phosphonium groups (Si–OH) acting as weak acids. SiO2-supported phosphonium salts
salts in the heterogeneous phase were summarized in Table 4. SiO2- were also investigated under continuous flow conditions showing good
supported phosphonium salts were reported as highly efficient catalytic activity even after 1000 h of reaction time. Moreover, a

Scheme 15. Proposed reaction mechanisms for cyclic carbonate synthesis using ILs as catalysts [116].

14
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

reaction mechanism was proposed (Scheme 14). Here, Si–OH groups Table 5
over the surface of SiO2 act as weak acids to activate the epoxide. The Summary of Imidazolium salts as homogeneous catalysts.
halide anion provided by the catalyst opens the ring of activated epoxide Sr. Reaction Yield
Catalyst Substrate Reference
by nucleophilic attack on least hindered carbon (Step I). The No. conditions (%)
ring-opened epoxide further interacts with CO2 to form carbonate (Step 70 C,

COOH-functionalized Bis
II). Finally, a five-membered cyclic carbonate is formed through a 4 bar p
1 Imidazolium Bromides SO 93 [144]
back-biting reaction (Step III). 5 mol%
(CO2),
The catalytic role of Si–OH (silanol groups) acting as active acidic 16 h
80 ◦ C,
sites is well established [117,119–121]. Previously heterogeneous cat­ Dicationic ILs (hexyl
10 bar p
alysts were prepared by grafting phosphonium [135], aminopyridinium 2 lateral alkyl chains) SO 55 [145]
(CO2),
C6(MIm)2 2Br 1 mol%
[117] and DABCO [120] based halides onto high surface area silica 20 h
support. These catalysts have unexpectedly exhibited a higher catalytic Imidazolium hydrogen
80 ◦ C,
5 bar p
activity than their homogeneous analogs. This increase in catalytic ac­ 3 carbonate ionic liquids SO 96 [146]
(CO2),
tivity was observed due to the synergistic effect of the halide anions with (C1C6 ImHCO3).
24 h
Si–OH (silanol) groups provided by silica acting as weak acids thereby 1-Benzyl-3- 90 ◦ C,
activating the epoxide [117,135]. Takahashi et al. investigated the CO2 4
butylimidazolium
SO
1 bar p
94 [147]
cycloaddition with PO over phosphonium salts-grafted silica catalysts. bromide (BnBImBr) and (CO2),
diethylamine (DEA) 24 h
From kinetic studies, the apparent activation energy for SiO2–C3H6–P 1-n-Butyl-3-methyl-
(n-Bu)3Br was 47 kJ/mol, which was significantly lower than that for P imidazolium chloride
40 ◦ C,
(n-Bu)4Br (86 kJ/mol). The results demonstrated the synergistic effect of and
2 bar p
the halide anion combined with the acidic site of silica. In another study, 5 μ-dichlorotetranitrosyl- SO
(CO2),
99 [148]
diiron (0.0832 mmol)
Hajipour et al. reported cyclic carbonate synthesis by CO2 cycloaddition 24 h
(BMIm) ([Fe(NO)2Cl]2)
to epoxides in the presence of silica-supported N-Benzyl DABCO bro­ 10 mol%
mide as a heterogeneous catalyst. From the results obtained, the Glycerol-tri (1-vinyl
silica-supported catalyst has shown a 91.4% yield of SC as compared to imidazolium) tri- 110 ◦ C,
1.4% using N-Benzyl DABCO bromide as a homogeneous analog under 6 mesylate ILs SO 20 bar p 88.7 [149]
(3.28 mmol)/DBU (CO2), 4 h
the same reaction conditions. A significant enhancement in catalytic
(6.56 mmol)
activity was observed due to the synergistic effect of Si–OH acidic sites 120 ◦ C,
1-Carboxypropyl
with halide anions [120]. 7 SO 15 bar p 82 [150]
imidazolium bromide
(CO2), 2 h
3-(2-Hydroxyethyl)-1-
3.1.3. Imidazolium salt-based catalysts 120 ◦ C,
vinyl-1H-imidazol-3-ium
The cycloaddition of CO2 to epoxides catalyzed by imidazolium- 8
chloride/DBU
SO 20 bar p 98 [151]
based ionic liquids (ILs) was also studied in both homogeneous and (CO2), 1 h
(1.66:1.9 mol%)
heterogeneous phases. ILs in the homogeneous phase can play a double
role as it can be used both as catalyst and solvent due to their ability to the ring to generate cyclic carbonate (Cycle 1). Another possible
dissolve a considerable amount of CO2. Cyclic carbonate formation pathway for cyclic carbonate formation was proposed using ILs con­
using ILs alone was mostly carried out at >80 ◦ C to achieve higher taining bulky anions such as BF−4 and PF−6 . These bulky anions interact
product yields. The reaction mechanism of CO2 cycloaddition to epox­ with CO2 directly to form a new ionic intermediate [X–CO2]− rather
ides in the presence of non-metallic ILs depends on the nature of cation than weaker interaction with the epoxides. Due to the more basicity of
and anion. Two types of reaction mechanisms for cyclic carbonate the ionic species compared to halide anion (X), this further interacts
synthesis catalyzed by ILs were proposed (Scheme 15) [33]. with an epoxide to form a carbonate intermediate. Finally, the cyclic
The first reaction mechanism is in agreement with a generally carbonate was obtained as the product (Cycle 2) [116]. This mechanism
accepted mechanism for cyclic carbonate synthesis as discussed previ­ was also supported by the mechanistic investigation carried out using
ously in the presence of an acid-base catalyst, explaining the ring- in-situ ATR-FTIR spectroscopy [122]. Some commonly known ILs used
opening of epoxides by nucleophilic attack of halide anion followed as catalysts for CO2 cycloaddition to terminal epoxides are shown in
by CO2 insertion into the ring-opened epoxide and finally the closure of Scheme 16.

Scheme 16. Structures of some commonly used imidazolium-based ILs as catalysts for cyclic carbonate synthesis [137].

15
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Peng and Deng [138] reported the first use of IL-based catalysts such
as 1-n-butyl-3-methylimidazolium halides (BMImX) for CO2 cycloaddi­
tion to epoxides resulting in the quantitative conversion of terminal
epoxides with 100% selectivity. Formation of PC from PO and CO2 was
carried out using [BMIm] PF6 (1) at 110 ◦ C and 25 bar p (CO2) for 6 h.
Moreover, the catalyst can be recycled after separation by distillation.
PC synthesis by CO2 cycloaddition to PO was also reported in the pres­
ence of 1-octyl-3-methylimidazolium tetrafluoroborate [C8-MIm]+ BF4
(2) as a homogeneous catalyst at supercritical conditions i.e., 100 ◦ C and
140 bar p (CO2). The high solubility of the epoxide and CO2 under su­
percritical conditions resulted in enhanced catalytic activity and a 100%
yield of PC in less than 2 h [139].
The use of hydroxyl-functionalized (OH) ionic liquids such as 1-(2-
hydroxyl-ethyl)-3-methylimidazolium bromide (HEMImBr) (3) has also
shown higher catalytic activities than their non-hydroxyl functionalized
analogs [140]. The quantitative yield of SC was achieved using HEM­
ImBr at 125 ◦ C and 20 bar p (CO2) after 1 h. Wang et al. [141] also re­
ported several hydroxymethyl-functionalized ILs as homogeneous
catalysts for cyclic carbonate synthesis under solvent-free conditions
resulting in a 90% yield of SC at 120 ◦ C and 20 bar p (CO2) for 2 h.
Similarly, the use of carboxyl-functionalized (COOH) Imidazolium ha­
Scheme 17. General mechanism for cyclic carbonate synthesis catalyzed by
lides (4) was also reported [142]. The activation of epoxide due to the
TBAX in conjunction with HBD’s [113].
HB donor ability of the COOH group resulted in higher catalytic activity.
Synthesis of PC in the presence of carboxylic-acid functionalized imi­
dazolium salt as an acid-base bifunctional catalyst was achieved with a Table 7
95% yield at 100 ◦ C and 20 bar p (CO2) for 1 h. Summary of HBDs based ammonium salts as homogeneous catalysts.
Anthofer et al. [143] also proposed hydroxyl-functionalized mono Sr. Reaction Yield
and bis-imidazolium halides (5) as the most active one-component cat­ Catalyst Substrate Reference
No. conditions (%)
alysts for CO2 cycloaddition to SO at 70 ◦ C and 4 bar p (CO2) resulting in 85 C,

96% yield after 22 h. The presence of the proton at the C2 position of the Choline iodide
1a SO 10 bar p 99 [163]
(6.08 mol%)
imidazolium ring causes epoxide activation and hence facilitates the (CO2), 6 h
epoxide ring-opening by the nucleophilic attack of halide anion. Tri-methylammonium 120 ◦ C,
2 histidine iodide SO 1 bar p 95 [164]
Furthermore, the presence of HB donors in the bifunctional structure of
(0.28 mol%) (CO2), 6 h
the catalyst provides the stability of the ring-opened epoxide in the Bifunctional
transition state. Table 5. 90 ◦ C, 5 bar
3 ammonium salts SO 95 [165]
p (CO2), 2 h
Due to the high product solubility of ILs in the homogeneous phase, (2.15 mol%)
highly energy-intensive downstream processes are required for the a
Ethanol was used as the reaction solvent.
separation of the catalyst. To make the process more economical and
environment-friendly, the immobilizing of ILs over the various sup­
porting materials were studied as a promising approach towards het­ divinylbenzene (DVB), and by immobilizing the active species of PS,
erogeneous catalysis. Particularly, the immobilization of –OH and SiO2 and biopolymeric (e.g., chitosan (CS), carboxymethylcellulose
–COOH group-containing ionic liquids are highly advantageous for CO2 (CMC)) supporting materials.
cycloaddition to epoxides under mild reaction conditions. SC formation
from SO and CO2 using various heterogeneous imidazolium-based ionic 3.1.4. Hydrogen bond donors-based catalysts
liquids are summarized in Table 6. The heterogenization of Hydrogen bond donors (HBDs) in combination with TBAX as binary
imidazolium-based ILs was mostly carried out by copolymerization with catalyst systems were also reported as efficient homogeneous catalytic

Table 6
Summary of imidazolium salts as heterogeneous catalysts.
Sr. Yield Reusability
Catalyst Reaction conditions Reference
No. (%) (runs)

1 SiO2-supported-1-butyl-3-methylimidazolium tetrafluoroborate (1.8 mol%) 160 ◦ C, 80 bar p (CO2), 4 h 96 3 [152]


2 DVB supported-3-butyl-1-vinylimidazolium chloride (0.7 mol%) 110 ◦ C, 60 bar p (CO2), 7 h 79 5 [40]
3 PS-supported-1-(2-hydroxyl-ethyl)-imidazolium bromide (1.6 mol%) 120 ◦ C, 25 bar p (CO2), 4 h 93 6 [121]
4 SiO2-supported-1-carboxyethyl ima imidazolium bromide (0.45 mol%) 110 ◦ C, 16.2 bar p (CO2), 5 h 96 4 [153]
5 PS-supported-2-3-(hydroxypropyl)-imidazolium bromide (13.8 wt%) 130 ◦ C, 20 bar p (CO2), 1 h 98 3 [154]
6 CS-supported-1-ethyl-3-methyl imidazolium bromide (1 mol%) 120 ◦ C, 20 bar p (CO2), 4 h 85 5 [34]
7 CMC-supported-1-butyl-3-triethoxysilylpropyl imidazolium iodide (1.2 mol%) 110 ◦ C, 18 bar p (CO2), 4 h 92 3 [155]
8 DVB-supported 3-(2-carboxyl-ethyl)-1- (3-amino-propyl) imidazole bromide 140 ◦ C, 20 bar p (CO2), 4 h 93.8 5 [156]
9 1-Vinyl-3-carboxyethyl imidazolium bromide (1 mol%) 130 ◦ C, 25 bar p (CO2), 6 h 90 5 [36]
10 DVB-supported-1-allyl-tetramethylguanidinium bromide (50 mg) 100 ◦ C, 1 bar p (CO2), 6 h 99 5 [157]
Porous hybrid polymers- supported-1,3-bis (3-bromobenzyl)-1H-imidazolium
11 80 ◦ C, 1 bar p (CO2), 72 h 84 5 [158]
bromide (50 mg)
12 SBA-15-supported-MIm-bromo ethane (50 mg) 110 ◦ C, 20 bar p (CO2), 5 h 99 – [159]
13 PS-supported-Imidazolium-based polymeric ionic liquids 130 ◦ C, 10 bar p (CO2), 8 h 92 6 [160]
120 ◦ C, 30 bar p (15% CO2 + 85%
14 PS-supported-1-benzyl-3-ethyl imidazolium bromide (10 mmol ionic sites) 87 5 [161]
N2), 32 h

16
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Table 8
Summary of two-component HBDs based catalyst systems.
Entry Catalyst Co-catalyst Reaction conditions Yield % References

1 p-Methoxyphenol (1) DMAP 120 C, 23 bar p (CO2), 48 h



96 [166]
2 Pyrogallol Ph (OH)3 (2) TBAI 45 ◦ C, 10 bar p (CO2), 18 h 52 [167]
3 Dinaphthylsilanediol (Si-diol) (3) TBAI 23 ◦ C, 1 bar p (CO2), 7 h 93 [168]
4 Tannic acid (4) TBAI 80 ◦ C, 10 bar p (CO2), 18 h 76 [169]
5 Cellulose DBU 120 ◦ C, 20 bar p (CO2), 2 h 85 [170]
6 Ethylenediaminetetraacetic acid (EDTA) (5) TBAI 50 ◦ C, 5 bar p (CO2), 18 h 94 [171]
7 2-Pyridine methanol (6) TBAI 45 ◦ C, 1 bar p (CO2), 20 h 85 [172]
8 (2,6-Dimethyl phenyl) boronic acid and water (7) TBAI 50 ◦ C, 10 bar p (CO2), 7 h 88 [173]
9 ascorbic acid TBAI 60 ◦ C, 1 bar p (CO2), 23 h 96 [174]

systems for cyclic carbonate formation [108,162]. Recently, consider­ Another class of nitrogen donor bases such as 4-(dimethylamino)
able attention has been focused to increase the activity of TBAX using pyridine (DMAP), 1,4-(diazabicyclco)octane (DABCO), 1,1,3,3-tetrame­
hydrogen bond donors (HBD’s) such as OH–functional groups as thylguanidine (TMG) and 1,5,7-(triazabicyclo)decene (TBD) were re­
co-catalyst. HBD’s cause epoxide activation by the interaction with its ported as active catalysts for the cyclic carbonate synthesis (Scheme 18).
oxygen atom through hydrogen bonding resulting in a polarization of Barbarini et al. [175] reported the SC synthesis from SO and CO2 in
the C–O–C bond and making it energetically less demanding. The acti­ the presence of methyl-1,5,7-(triazabicyclo) decene (MTBD) in both
vated epoxide facilitates ring-opening by the nucleophilic attack of homogeneous phase and heterogeneous phase prepared by covalent
halide anions followed by rapid CO2 insertion (Scheme 17). grafting over silica (SiO2). As a result, a 95% yield of SC was obtained
Wang et al. [124] reported OH-functionalised ammonium salt such using homogeneous catalysts at 140 ◦ C and 50 bar p (CO2) for 20 h.
as hydroxyethyl tributyl ammonium bromide (HETBAB) as an active Although heterogeneous analog of the catalyst has the advantage of
catalyst for cycloaddition reaction. The presence of an OH–group recycling, the catalytic activity decreased significantly, and a 90% yield
significantly enhances the rate of epoxide ring-opening by the nucleo­ of SC was achieved after 70 h under the same reaction conditions. Zhang
philic attack of the halide anion resulting in an improved yield of 96% et al. [176] reported SiO2-supported TBD and SiO2-supported DBU as
PC rather than 74% product yield when the reaction was catalyzed by heterogeneous catalysts for PC formation by CO2 cycloaddition to PO.
TBAB alone under the same reaction conditions. The presence of HBD in The quantitative yield of PC was achieved at 150 ◦ C and 15 bar p (CO2)
the catalyst system has two advantages: 1) epoxide activation through for 10 h. Both heterogeneous catalysts can be used for up to 5 cycles with
hydrogen bonding which favors epoxide ring-opening, 2) provides sta­ an appreciable decrease in activity and selectivity. Similarly, some other
bility to the ring-opened epoxide intermediate which is a key factor for TMG, DBU and DMAP based catalysts were also reported for CO2
rapid CO2 insertion. Some others OH–group functionalized ammonium cycloaddition to terminal epoxides [17]. However, the reduced catalytic
salts as HBD based homogenous catalyst for cyclic carbonate synthesis activity, harsh reaction conditions, and recycling problems were the key
have been summarized in Table 7. issues to limit the applications of these catalysts on a commercial scale.
The reaction carried out using HBD organocatalysts alone have Recently, Meng et al. [145] reported DBU-based protic ionic liquids
shown no or little catalytic activity. However, the presence of nucleo­ (DBPILs) as a bifunctional metal-free catalyst for cyclic carbonate syn­
phile additives in combination with HBD’s has resulted in a significant thesis via CO2 cycloaddition to various epoxides under mild reaction
increase in catalytic activity under mild reaction conditions due to the conditions. As a result, 87% yield of SC was obtained using 6 mol% of
synergistic effect [166]. Alkylammonium halides, KI, DMAP and DBU catalyst at 50 ◦ C and 1 bar p (CO2) for 6 h. Moreover, the catalyst can
were commonly used as nucleophile additives in combination with HBD also be used to produce a good yield of cyclic carbonates from simulated
organocatalysts for cyclic carbonates synthesis (Table 8). flue gas (15% CO2 and 85% N2 mixture). The reaction mechanism was
. also investigated using DFT and IR spectroscopy analysis showing the

17
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Scheme 18. Structure of nitrogen donor bases used as catalysts for cyclic carbonate synthesis [17].

Scheme 19. General reaction mechanism for CO2 cycloaddition to epoxides catalyzed by the monometallic complex [33].

role of DBPILs to activate both epoxide and CO2 by strong H-bonding this step energetically less demanding. This dual activation of epoxide
and alkoxy anions interactions. and CO2 has resulted in the higher catalytic activity of bimetallic com­
plexes than monometallic complexes [109].

3.2. Metal-based complexes

Metal-based salen and salphen complexes are widely recognized as


highly efficient catalysts for the cyclic carbonate synthesis under mild
reaction conditions. Salen ligands can be prepared by the condensation
of salicylaldehyde with a chiral or achiral diamine [60]. The use of
metal-based salen complexes for cyclic carbonate synthesis has several
advantages such as the facile method of preparation in enantiomerically Scheme 20. Structure of Al (III) salen complex as a catalyst for cyclic car­
pure form and easy availability. Due to their tunable structures, salen bonate synthesis [66].
complexes are highly desirable catalysts for kinetic resolution and
asymmetric cyclic carbonate formation [177]. The metal-based com­
plexes can be used as catalysts alone or in a combination with the
nucleophile additive co-catalyst. The reaction mechanism involved in
cyclic carbonate synthesis in the presence of metal complexes in com­
bination with nucleophile has been generally well established (Scheme
19) [33].
The reaction mechanism of cyclic carbonate synthesis in the presence
of monometallic complexes involves the activation of epoxide by the
interaction of a metal center with oxygen (M–O interaction), followed by
epoxide ring-opening by the nucleophilic attack of the halide anion on
the least hindered C atom of the epoxide to form a metal-bonded
alkoxide intermediate. This ring-opened intermediate further un­
dergoes CO2 insertion by the nucleophilic attack of metal-bonded
alkoxide on the carbon atom to form a metal-bonded carbonate.
Finally, the carbonate undergoes a back-biting reaction to form five-
membered cyclic carbonate. This mechanism was also supported by
the in-situ IR kinetic analysis and most likely to occur using a low
epoxide/catalyst molar ratio [178]. The bimetallic reaction mechanism
involving the activation of both epoxide and CO2 simultaneously was
also reported for the reactions carried out using metal complexes con­
taining two neighboring metal centers. The simultaneous activation of Scheme 21. Structure of bimetallic Al (III) salen complex as a catalyst for
CO2 helps to promote CO2 insertion to a ring-opened epoxide by making cyclic carbonate synthesis [109].

18
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Similarly, the steric and electronic properties of the complexes also


affect the catalytic activity and selectivity of the metal complexes. These
properties are mostly defined by the type of ligand attached to the metal
center. For instance, the use of ligands having high electron-donor
ability favors cyclic carbonate formation by promoting the back-biting
reaction [33]. Similarly, the steric properties of the complexes also
affect the catalytic activity of metal complexes and can be varied by the
replacement of substituents attached to the aromatic rings of the ligands
[180].
Various types of metal-based salen and salphen complexes with
tunable properties have been examined as highly efficient catalysts for
CO2 cycloaddition to epoxides. These were generally used in conjunc­
tion with nucleophile additives as an essential part of the catalyst or to
increase the catalytic activity of a nucleophile containing metal com­
plex. The role of the nucleophile is defined as opening the epoxide ring
Scheme 22. Structure of bimetallic Al (III) salen complex as a one-component by providing the halide anion. Moreover, it can also interact with the
bifunctional catalyst [182]. metal center of the complex to provide an electron-rich environment
leading to rapid epoxide or CO2 insertion. The properties of the metal
complexes can be further changed by the choice of the metal center.

3.2.1. Al (III) salen complexes


The use of Al (III) salen complex in combination with TBAX as a
binary catalyst system for CO2 cycloaddition to epoxides was initially
reported by Lu et al. [181]. In this study, an almost quantitative yield of
SC was obtained using Al (III) salen complex (0.125 mol%) in combi­
nation with TBAI (0.125 mol%) at 25 ◦ C and 6 bar (CO2) after 15 h
(Scheme 20).
Scheme 23. Structure of Al (III) amino triphenolate complex as a catalyst for Later on, Melendez et al. [109] reported bimetallic Al (III) salen
cyclic carbonate synthesis [66].
complex (Scheme 21) in combination with TBAB as a highly effective
The nature of the metal center and the ligand connected with the binary homogeneous catalyst for the CO2 cycloaddition to terminal ep­
metal center plays an important role to decide the catalytic activity of oxides at ambient conditions i.e. at 25 ◦ C and 1 bar p (CO2). SC forma­
the metal complexes. Moreover, the strength of the M–O bond of tion from SO and CO2 using this binary homogeneous catalyst has
alkoxide and carbonate intermediates also plays a key role to determine resulted in almost quantitative conversion after 24 h reaction time. The
the activity of the catalyst [179]. For instance, a strong M–O bond makes presence of two metallic centers in the catalyst structure facilitates the
the alkoxide challenging to interact with CO2 to form carbonate. simultaneous interaction with one or both reactants and intermediates
Conversely, a weak M–O bond can be easily replaced by the nucleophile leading to a higher reaction rate and product selectivity.
or a solvent molecule resulting in low catalytic activity, but also pro­ Subsequently, to avoid the use of TBAB as a co-catalyst, the use of a
moting the back-biting reaction at the same time to form cyclic car­ one-component bifunctional catalyst was also reported by the covalent
bonates. Hence, intermediate bond strength is more favored for the bonding of the TBAB with Al (III) salen complex (Scheme 22) [182]. The
better performance of the catalyst. use of a one-component bifunctional catalyst has shown comparable

Table 9
Summary of Al (III) salen complex based catalyst systems.
Yield
Entry Catalyst Co-catalyst Substrate Reaction conditions References
%

25 ◦ C, 6 bar p (CO2),
1 Schiff-base aluminum complexes ((Salen)AlX) (0.125 mol%) TBAI (0.125 mol%) PO 99 [181]
14 h
25 ◦ C, 1 bar p (CO2),
2 Dimetallic aluminum(salen) complex 1a (2.5 mol%) TBAB (2.5 mol%) SO 98 [188]
24 h
25 ◦ C, 1 bar p (CO2),
3 One-component bimetallic Al (III) salen complex (2.5 mol%) – SO 90 [189]
24 h
25 ◦ C, 1 bar p (CO2),
4 Bimetallic µ-oxoaluminium(acen) complex (2.5 mol%) TBAB (2.5 mol%) SO 93 [190]
24 h
Polystyrene supported one-component bimetallic Al (III) 25 ◦ C, 1 bar p (CO2),
5 – SO 90 [103]
salen complex (2.5 mol%) 24 h
Silica-supported one-component bimetallic Al (III) salen 25 ◦ C, 1 bar p (CO2),
6 – SO 86 [103]
complex (2.5 mol%) 20 h
Phosphonium salt based one-component bimetallic Al (III) 25 ◦ C, 1 bar p (CO2),
7 – SO 91 [185]
salen complex (2.5 mol%) 24 h
50 ◦ C, 50 bar p
8 Bimetallic Al (salen) complex (0.5 mol%) – SO 78 [41]
(CO2), 24 h
25 ◦ C, 1 bar p (CO2),
9 Al(salphen) complex (1.5 mol%) TBAB (1.5 mol%) SO 83 [191]
24 h
25 ◦ C, 1 bar p (CO2),
10 Bimetallic Al(salphen) complex (0.25 mol%) TBAB (0.5 mol%) SO 85 [186]
24 h
1,2- 90 ◦ C, 10 bar p
11 Amino triphenolate based Al (III) complex (0.05 mol%) TBAI (0.25 mol%) 96 [69]
Epoxyhexane (CO2), 2 h
Imidazolium hydrogen carbonate 25 ◦ C, 1 bar p (CO2),
12 Al-salen complex (0.25 mol%) SO 98 [187]
(0.25 mol%) 24 h

19
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Table 10 in catalytic activity was observed due to the dequaternization of quater­


Summary of Cr (III)-based salen and salphen complexes as homogeneous nary ammonium groups connected with the complex scaffold. This het­
catalysts. erogeneous catalyst has shown high-temperature stability and can be used
Co- Reaction Yield up to 170 ◦ C. Due to these properties, the heterogeneous catalyst was also
Entry Catalyst Substrate References
catalyst conditions (%) tested in a flow reactor for cyclic carbonate synthesis under continuous
Cr (III) flow conditions [184]. The catalyst has also shown a reasonable conver­
salen 85 ◦ C, sion of SO to SC using simulated flue gases under continuous flow con­
DMAP
1 complex
1 mol%
SO 3.4 bar p [192] ditions. Moreover, the catalyst deactivation was observed after extended
(1) 1 mol (CO2), 7 h
use which can be reversed by the quaternization of the deactivated cata­
%
Cr (III) lyst by the addition of benzyl bromide.
100 ◦ C, Another one-component Al (III) salen complex based homogeneous
salphen
DMAP 2 bar p
2 complex
2 mol%
PO
(CO2),
100 [193] catalyst with the covalent bonding of quaternary phosphonium halides
(4) 1 mol instead of quaternary ammonium halides was also reported by the same
1.4 h
%
Cr (III)
group [185]. The use of this catalyst for CO2 cycloaddition to terminal
25 ◦ C, epoxide has also shown comparable conversion at mild reaction condi­
salphen TBAB
1 bar p
3 complex 2.5 mol SO 92% [92] tions. Recently, Wu and North [186] proposed Al (III) salphen complex
(CO2),
(2) %
24 h made by the connection of two salen scaffolds with the phenyl ring. This
2.5 mol%
catalyst was prepared with the aim to increase activity and to further
Cr (III)
salphen TBAB
25 ◦ C, reduce the amount of catalyst used during the reaction. The experi­
1 bar p mental results have shown that only 0.25 mol% of the complex in
4 complex 1.5 mol SO 100 [191]
(CO2),
(3) %
24 h
combination with 0.5 mol% TBAB has resulted in 90% conversion of SO
1.5 mol% to SC at 50 ◦ C and 10 bar p (CO2).
Whiteoak et al. [69] reported an amino triphenolate based Al (III)
results to binary catalyst system i.e., quantitative conversion of SO to SC complex (0.5 mol%) (Scheme 23) in combination with TBAI (0.25 mol
at 25 ◦ C and 1 bar p (CO2) after 24 h. However, the use of a %) as being highly effective for CO2 cycloaddition to several terminal
one-component catalyst has a drawback of high catalyst loading due to epoxides. Remarkably, this catalyst has shown high activity for cyclic
its high molecular weight i.e., about 1600 g/mol. For example, 2.5 mol carbonate synthesis from terminal epoxides (initial TOF = 36,000 h− 1).
% catalyst loading means 4 g of the one-component catalyst is required The formation of cyclic carbonate from CO2 and 1,2-epoxyhexane
for each 12 g (100 mmol) of SO. To reduce the molecular weight of the catalyzed by this binary catalyst has shown 96% conversion at 90 ◦ C
one-component catalyst prepared by the replacement of tert-butyl and 10 bar p (CO2) after 2 h. Due to its trigonal geometry, this catalyst
groups (tBu) attached to each phenyl ring with H was also investi­ has also shown higher catalytic activity for cyclic carbonates formation
gated. However, the catalyst has shown less catalytic activity due to from more sterically congested internal epoxides.
solubility issues [183]. Similarly, the presence of Br− anion on the salen Recently, Liu et al. [187] reported Al-salen complex in combination
scaffold has shown higher catalytic activity than Cl− due to a good with imidazolium hydrogen carbonate ILs as an efficient binary catalyst
balance of its nucleophilicity and leaving group ability. system for cycloaddition of CO2 and various terminal epoxides at room
To make the process sustainable, the use of one-component catalysts temperature. For instance, synthesis of SC from SO and CO2 using 1:1
in the heterogeneous phase was considered as a promising approach molar ratio Al-salen complex and imidazolium hydrogen carbonate ILs
towards its commercial applications. It can be prepared by grafting of resulted in 98% yield of SC at 25 ◦ C, 10 bar CO2 pressure and 24 h re­
ammonium moiety of the one-component bifunctional catalyst over high action time. The use of novel co-catalyst used in this study has shown
surface areas inorganic supports such as a polymer (Merrifield resin) and higher catalytic activity as compared to traditionally used TBAX and
amorphous SiO2 [182]. The supported catalysts have shown comparable organic bases. The use of this non-corrosive co-catalyst has the advan­
results to their homogenous analogs. For instance, the use of 2.5 mol% tage over halide anions containing a corrosive catalyst which is a
polymer-supported catalysts have shown 100% SO conversion to SC at growing concern in their commercial applications. Moreover, this
25 ◦ C, 1 bar p (CO2) and 24 h. The heterogeneous catalysts can be easily catalyst system can also be recycled up to 4 cycles without a substantial
separated and recycled. However, a significant decrease in the catalytic decrease in catalytic activity and selectivity towards the final product. A
activity of the immobilized catalyst was observed and the product yield summary of Al (III) salen complex based catalyst systems used for the
was decreased to 70% after two successive runs. To improve the recy­ synthesis of cyclic carbonates from CO2 cycloaddition to various ter­
clability of the heterogeneous catalyst, another one-component catalyst minal epoxides are given in Table 9.
having four quaternary ammonium groups attached with the scaffold of
Al (III) salen complex was prepared by immobilizing over the polymer 3.2.2. Cr (III)-based complexes
resin. However, no significant improvement in the reusability of the Cr (III) salen and salphen complexes in conjunction with nucleophile
heterogeneous catalyst was observed. additives were identified as highly active catalysts for cyclic carbonate
The heterogeneous catalysts were also prepared by the immobilization formation at low p (CO2) (<10 bar). Moreover, these complexes can
of bimetallic catalysts over amorphous SiO2 and MCM-41 supports. SiO2- easily tuneable structures which can be efficiently used for asymmetric
supported catalyst has shown a 78% conversion of SO to SC under ambient cyclic carbonate synthesis with high stereoselectivity. The use of Cr (III)
conditions [103]. Moreover, the catalyst has shown good recyclability up salen and salphen complexes for cyclic carbonate synthesis is summa­
to 11 runs with a progressive decrease in catalytic activity. The reduction rized in Table 10..

Table 11
Summary of Zn-based complexes as homogeneous catalysts.
Entry Catalyst Co-catalyst Substrate Reaction conditions Yield (%) Reference

1 Zn (II) salphen complex (1) (2.5 mol%) TBAI (2.5 mol%) SO 80 ◦ C, 80 bar p (CO2), 3 h 80 [45]
2 Zn(salpyr) catalyst (2) (0.5 mol%) _ SO 80 ◦ C, 10 bar p (CO2), 18 h 88 [194]
3 Salen-based zinc-coordinated complex (3) (0.1 mol%) TBAB (3.6 mol%) SO 120 ◦ C, 30 bar p (CO2), 1 h 96 [195]
4 Zn-based complex (4) (0.2 mol%) TBAI (0.2 mol%) PO 40 ◦ C, 20 bar p (CO2), 20 h 88 [196]
5 Heterometallic RE-Zn complexes (1 mol%) TBAB (3 mol%) SO 25 ◦ C, 1 bar p (CO2), 24 h 84 [197]

20
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

3.2.3. Zinc (II)-based complexes substituents to stabilize the ring-opened epoxide. The reaction was
Decortes et al. [45] reported zinc-based monometallic complexes performed using dichloromethane CH2 Cl 2 as a solvent. The syn­
with high catalytic activity at mild reaction conditions i.e. thesis of SC from SO and CO 2 in the presence of the above-mentioned
(25–45 ◦ C and 2–10 bar p (CO 2 )). These complexes have the binary catalyst system has shown 80% yield at 45 ◦ C and 10 bar p
advantage of bulky groups connected with salphen scaffolds on (CO 2 ) after 18 h. However, the use of solvent was omitted under
ortho-positions that prevent catalytic deactivation due to unwanted supercritical conditions i.e., 80 ◦ C and 80 bar p (CO 2 ) and reaction
dimerization of the complex. Nevertheless, the use of these com­ time was decreased from 18 to 3 h. The use of a solvent in this re­
plexes in combination with TBAX was limited for cyclic carbonate action has shown a decrease in the reaction rate as compared to a
synthesis from terminal epoxides. The use of some zinc-based sal­ CO 2 rich solvent-free environment. Later, it was observed that the
phen complexes as catalysts for cycloaddition of CO 2 to epoxides is reaction can be further optimized using methyl ethyl ketone (MEK)
summarized in Table 11. The catalytic activity of the Zn (II) salphen as the reaction solvent. The synthesis of SC carried out in the pres­
complex was attributed to its unique structure offering increased Zn ence of MEK has resulted in 89% isolated yield at 25 ◦ C and 2 bar p
ion Lewis acid properties in combination with bulky tert-butyl (tBu) (CO 2 ) after 18 h reaction [45]..

21
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Table 12
Summary of MOFs and ZIFs-based catalysts.
Sr. Reaction Yield
Catalyst Substrate Reference
No. conditions (%)

MOF-5/n- 50 C, 60 bar p

1 PO 98 [200]
Bu4NBr (CO2), 4 h
100 ◦ C, 20 bar p
2 Co-MOF-74 SO 96 [213]
(CO2), 4 h
100 ◦ C, 20 bar p
3 Mg-MOF-74 SO 94 [214]
(CO2), 4 h
25 ◦ C, 8 bar p
4 Cr-MIL-101 SO 95 [215]
(CO2), 48 h
25 ◦ C, 8 bar p
5 Fe-MIL-101 SO 93 [202]
(CO2), 48 h
25 ◦ C, 1 bar p
6 MOF-505 PO 48 [216]
(CO2), 48 h
MOF-5/n- 50 ◦ C, 4 bar p
7 PO 45 [217]
Bu4NBr (CO2), 4 h
80 ◦ C, 8 bar p
8 ZnMOF-1-NH2 SO 88 [218]
(CO2), 8 h
70 ◦ C, 1 bar p
9 Zn(II)-MOF SO 99 [219]
(CO2), 12 h
100 ◦ C, 20 bar p
10 Zn(II)-MOF SO 99 [220]
(CO2), 6 h
Scheme 24. General reaction mechanism for CO2 cycloaddition to epoxides
25 ◦ C, 1 bar p
11 JUC-1000 PO 96 [221] catalyzed by Zn-salts in combination with ILs [105].
(CO2), 48 h
100 ◦ C, 20 bar p
12 HKUST-1 SO 48 [222]
(CO2), 4 h combination with MOF-5 resulted in a 92% yield of SC under mild re­
150 ◦ C, 8 bar p
13 MIL-68(In) 39 [223] action conditions i.e. 1 bar CO2 pressure and 50 C. Simultaneously,
(CO2), 8 h
MIL-101(Cr)- 110 ◦ C, 20 bar p Baiker and coworkers [201] reported the use of mixed-linker MOFs in
14 SO 95 [224]
TSIL (CO2), 6 h combination with TBAX as a binary catalyst system for CO2 cycloaddi­
15
Co-POM@MIL-
SO
110 ◦ C, 20 bar p
88 [225] tion to PO. Later on, Mg-MOF and CO-MOF were also implied as het­
101(Cr) (CO2), 6 h
erogeneous catalysts for cycloaddition reaction. Zalomaeva et al.
35 ◦ C, 1.5 bar p
16 MIL-101(Cr) PO
(CO2), 24 h
99 [226] reported Cr-MIL-101 in combination with TBAB as promising catalysts
MIL-101(Cr)/n- 60 ◦ C, 80 bar p [202] for CO2 cycloaddition to epoxides as well as direct cyclic car­
17 SO 99 [227]
Bu4NBr (CO2), 3 h bonate synthesis via oxidative carboxylation olefins. Cr-MIl-101 in
100 ◦ C, 7 bar p combination with TBAB resulting 95% SC held at 25 ◦ C and 8 bar CO2
18 ZIF-8 SO 54 [228]
(CO2), 5 h
pressure in 48 h. However, Cr-MIl-101has shown poor recyclability as
120 ◦ C, 10 bar p
19 ZIF-68 SO
(CO2), 12 h
93 [206] the surface area has been decreasing remarkably from 3270 to 860 m2/g
120 ◦ C, 20 bar p resulting in a 63% yield after the first cycle. Similarly, the use of
20 gea-MOF-1 SO 85 [204]
(CO2), 6 h Fe-MIL-101 has been also reported in combination with TBAB resulting
100 ◦ C, 20 bar p
21 IRMOF-3 SO 33 [222] in a 93% yield of SC under the same reaction conditions. Some other
(CO2), 4 h
100 ◦ C, 10 bar p
MOFs-based catalysts of MIL series such as MIL-125, amine-functional­
22 Ni-TCPE1 SO 99 [229] ized (NH2-MIL-125) [203], MIL-68 and NH2-MIL-68 [204] were also
(CO2), 12 h
reported as heterogeneous catalysts for cyclic carbonate synthesis. These
MOFs are derived from benzene dicarboxylate (BDC) linkers having
3.2.4. MOFs and ZIFs-based catalysts different metal centers.
The use of metal-organic frameworks (MOFs) as heterogeneous cat­ Meanwhile, Zeolitic imidazolate frameworks (ZIFs) were also intro­
alysts for cyclic carbonate synthesis has recently gained much attention duced as catalysts for CO2 cycloaddition to epoxides [205]. ZIFs are
[198]. MOFs is a novel class of highly porous organic-inorganic hybrid prepared by functionalizing the MOFs with imidazolate-based linkers
crystalline materials consists of metal clusters in combination with and divalent metal nodes such as Zn+2. Carreon and coworkers reported
organic linkers. The high surface area, well-ordered pores, high the first use of ZIF-8 as catalyst resulted in a 54% yield of SC at 100 ◦ C
adsorption capacity and several means of changing the surface chem­ 7 bar CO2 pressure after 4 h. The catalyst has shown good recyclability
istry are the major characteristics of these materials. MOFs are also used with a slight decrease in the catalyst surface area. In 2014, Yang et al.
as CO2 gas storage and separation materials. The ability of these mate­ [206] proposed cyclic carbonate synthesis in the presence of ZIF-68 as
rials to capture CO2 and simultaneous transformation into cyclic car­ an acid-base heterogeneous catalyst. This was prepared by the func­
bonates makes them an excellent choice as a catalyst for CO2 tionalization of 2-nitroimidazole with benzimidazole linkers. Due to the
cycloaddition to epoxides. The catalytic sites can be homogeneously high CO2 adsorption ability of ZIF-68, 93% yield of SC was achieved at
distributed over the MOFs as compared to other supporting materials, 120 ◦ C, 10 bar CO2 pressure after 12 h. Moreover, the catalyst can be
resulting in high product selectivity comparable to their homogeneous recycled up to 3 cycles without a significant decrease in catalytic ac­
counterparts. The presence of lewis acidic active metal site in the tivity. Similarly, ZIF-90 and F-ZIF-90 were also implied as catalysts
combination of high positive charge metal ions significantly enhances [207,208]. The role of OH and COOH-functionalised MOFs to enhance
the reaction under optimized reaction conditions. Moreover, the effi­ the catalytic activity was also investigated [209]. Some other types of
cient design of the MOF catalyst alone can effectively catalyze the re­ carboxylate-based MOFs such as Ni(salphen)and Ni(saldpen) and
action without the use of expensive co-catalysts [199]. zirconium-based MOF were also reported as highly efficient heteroge­
Han and coworkers [200] reported the first use of MOF-5 in com­ neous catalysts for CO2 cycloaddition to various epoxides [210–212].
bination with TBAB as a binary heterogeneous catalyst for cyclic car­ These catalysts favor the CO2-insertion to alkoxide intermediate in the
bonate synthesis. MOF-5 was prepared from benzene-1,4-dicarboxylate reaction mechanism. The applications of MOFs as catalysts for this re­
and Zn4O. The synergic effect of using quaternary ammonium salts in action and recent advances in this area are summarized in Table 12.

22
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

Table 13
Summary of Zn-salts based binary homogeneous catalysts.
Yield
Entry Catalyst system Substrate Reaction conditions Reference
(%)

1 ZnCl2 (0.05 mmol) and [BMIm]Br (0.30 mmol) SO 100 C, 15 bar p (CO2), 1 h

86 [231]
2 ZnBr2 (0.4 mmol) and [BMIm]Cl (0.8 mmol) SO 80 ◦ C, 40 bar p (CO2), 1 h 93 [232]
90 ◦ C, 80 bar p (CO2),
3 ZnBr2 (0.1 mmol) and TBAI (0.2 mmol) SO 98 [230]
30 min
4 ZnBr2 (0.015 mmol) and tetraphenylphosphonium iodide [Ph4P] I (0.09 mmol) SO 120 ◦ C, 25 bar p (CO2), 1 h 86 [233]
5 ZnI2 (0.1 mmol) and dicationic ionic liquid (0.05 mmol) SO 110 ◦ C, 15 bar p (CO2), 3 h 94 [234]
6 ZnBr2 (1 mol%) poly(4-vinylimidazolium) (1 mol%) and DBU (2 mol%) SO 80 ◦ C, 1 bar p (CO2), 10 h 92 [235]
ZnBr2 (0.09 mmol) methylimidazole (MIm) based N‑heterocyclic compounds
7 PO 130 ◦ C, 30 bar p (CO2), 1 h 94 [236]
(0.36 mmol)
8 Imidazolium zinc tetrahalide (0.02 mol%) EO 100 ◦ C, 35 bar p (CO2), 1 h 72 [237]
Epichlorohy-drin
9 ZnBr2 (1.5 mol%) and N-dodecyl- N-methylmorpholinium bromide (3 mol%) 25 ◦ C, 1 bar p (CO2), 24 h >99 [238]
(ECH)
10 ZnI2 (2.0 mmol%) and polyethylenimine (0.15 g) PO 100 C, 10 bar p (CO2), 3 h

95 [239]
11 ZnI2 and amidinothiourea (ATUI) (0.1 g) SO 110 ◦ C, 10 bar p (CO2), 4 h 71 [240]

3.3. Zinc salt-based catalysts 4. Summary and future trends

The use of commercially available zinc salts in combination with The use of CO2 as a C1-building block has attracted the attention of
onium salts was also reported as a simple acid-base bifunctional ho­ both academia and the chemical process industry. Among the other
mogeneous catalyst for CO2 cycloaddition to epoxides. Zn-salts such as methods of CO2 utilization, the synthesis of cyclic carbonates from ep­
ZnX2 (where X = I− , Br− or Cl− ) provides a Lewis acidic site to activate oxides and CO2 has been extensively studied using numerous catalyst
the epoxide which readily undergoes ring-opening by nucleophilic systems in the last 20 years. Although the synthesis of cyclic carbonates
attack of halide anion provided by the ammonium salts [230]. The has been extensively studied using various catalyst systems with higher
catalytic role of imidazolium-based ionic liquids in combination with conversion and yield of the product, a detailed study of reaction kinetics
Lewis acid compounds such as zinc halides have been also investigated is required to investigate the reaction mechanism. The use of organo­
as binary homogeneous catalysts for cyclic carbonate synthesis. PC catalysts such as ammonium, phosphonium, and imidazolium salts have
synthesis carried out using BMImBr alone has shown TOF 37 h− 1 at the advantage of commercial availability. However, cyclic carbonate
100 ◦ C and 35 bar p (CO2) after 1 h. However, a significant enhance­ synthesis in the presence of most of the organocatalysts still requires
ment in catalytic activity was reported (TOF 1679 h− 1) using ZnBr2 as harsh reaction conditions due to the lack of epoxide activation groups.
co-catalyst at the same reaction conditions. The reaction mechanism for Consequently, efforts have been made to increase the catalytic activity
CO2 cycloaddition to epoxides using Zn-salts in combination with ionic by combining the organocatalysts with –OH, –COOH groups and zinc
liquids was also proposed (Scheme 24). salts (Lewis acids) which can activate the epoxide. Overall catalytic
The Lewis acid site provided by the Zn salts activates the epoxide by performance evaluation of these catalyst systems under variable reac­
the interaction with the oxygen atom of the epoxide. Subsequently, the tion conditions is a difficult task.
ring-opening of the activated epoxide takes place by the Br− provided by Heterogeneous catalysis for cyclic carbonate synthesis was less
the ionic liquid catalyst. Li et al. [231] also described the increase in extensively studied than their homogeneous counterparts. The use of
catalytic activity using imidazolium-based ionic liquids in combination heterogeneous catalysts still had some potential short-comings due to
with zinc halides. PC synthesis carried out in the presence of BMImBr in the requirement of harsh reaction conditions, the use of co-catalysts
a combination of ZnCl2 has shown 86% yield (3015 h− 1) at 100 ◦ C and (nucleophiles) in a homogeneous phase and loss in catalytic activity in
15 bar p (CO2) after 1 h. Moreover, a reaction mechanism was proposed consecutive runs. To solve these issues, the immobilizing of active ho­
explaining the synergistic effect between ionic liquids and zinc salts. mogenous catalysts on mesoporous materials e.g., MCM-41, SBA 15 and
Similarly, the use of TBAX in combination with zinc halides as a highly novel MOFs as high surface area supports are required. Furthermore,
effective acid-base binary homogeneous catalyst was also reported. SC immobilizing with more strong covalent grafting methods is highly
formation from SO and CO2 catalyzed by TBAI/ZnBr2 has shown a 98% desirable to avoid leaching of the catalyst under reaction conditions.
yield of SC with a turnover frequency (TOF) of 6861 h− 1 using 1:2 molar Moreover, the development of the robust active heterogeneous catalyst
ratio under supercritical conditions i.e. 90 ◦ C and 80 bar p (CO2) after with the ability to tolerate the impurities present in flue gases is desir­
30 min [230]. This enhanced catalytic activity was attributed to the able to start the industrial application of cyclic carbonate synthesis from
supercritical reaction conditions where a homogeneous mixture of waste CO2 as a feedstock.
epoxide, CO2 and catalyst was achieved resulting in good mixing. The Generally, the catalyst systems described in the literature have
effect of a change in CO2 pressure on cyclic carbonate synthesis was also shown higher catalytic activity for mono-substituted terminal epoxides.
investigated. The results have shown that SC yield increases gradually Conversely, the use of internal epoxides (di- and tri-substituted) for
by increasing pressure from 20 to 80 bar. However, the change in p cyclic carbonate synthesis under mild reaction conditions is still chal­
(CO2) from 80 bar to 100 bar resulted in a sharp decrease in product lenging. The advancement in catalytic systems is highly desirable for
yield which was explained by the dilution effect leading to a decrease in cycloaddition of CO2 to internal epoxides at mild reaction conditions.
epoxide concentration around the catalyst in the expanded reaction The use of bio-based epoxides as a potential replacement of petroleum-
medium at higher pressure i.e., >80 bar. The decrease in the rate of based epoxides is highly desirable for the sustainable future of cyclic
reaction due to the similar dilution effect at very high pressure was also carbonate synthesis. The attention of future research needs to be focused
observed by some other authors. Moreover, a reaction mechanism was on the development of active catalyst systems for CO2 cycloaddition to
proposed explaining the increase in catalytic activity due to a synergistic renewable bio-based epoxides to make the process 100% sustainable.
effect between ZnBr2 and TBAI. The use of zinc salts in conjunction with The optimization of reaction conditions also needs to be a focus to
onium salts for cyclic carbonate synthesis has been summarized in overcome heat and mass transfer limitations to improve the overall
Table 13. performance of this reaction. The use of bio-based resources will not
only help to make processes more sustainable but also should generate

23
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

new products such as bio-based polymers and green solvents. Bio- [16] M. Aresta, A. Dibenedetto, A. Angelini, Catalysis for the valorization of exhaust
carbon: from CO2 to chemicals, materials, and fuels. Technological use of CO2,
derived polymers frequently have desirable properties such as biode­
Chem. Rev. 114 (3) (2013) 1709–1742.
gradability, hydrophobicity, bioactivity, and liquid crystallinity. [17] M. Cokoja, M.E. Wilhelm, M.H. Anthofer, W.A. Herrmann, F.E. Kuehn, Synthesis
Synthesis of cyclic carbonates from epoxides and CO2 is a typical gas- of cyclic carbonates from epoxides and carbon dioxide by using organocatalysts,
liquid multiphase catalytic process involving a gas-liquid mass transfer ChemSusChem 8 (15) (2015) 2436–2454.
[18] M. Yoshida, M. Ihara, Novel methodologies for the synthesis of cyclic carbonates,
in a reactor and the catalytic cycloaddition reaction in the liquid phase. Chem. Eur. J. 10 (12) (2004) 2886–2893.
Therefore, reactions carried out in an “efficient” reactor design (in terms [19] R. Martin, A.W. Kleij, Myth or reality? Fixation of carbon dioxide into complex
of the high rate of heat/mass transfer and continuous flow approach) organic matter under mild conditions, ChemSusChem 4 (9) (2011) 1259–1263.
[20] M. North, F. Pizzato, P. Villuendas, Organocatalytic, asymmetric aldol reactions
could mitigate many of the short-comings observed in conventional with a sustainable catalyst in a green solvent, ChemSusChem 2 (9) (2009)
batch reactors. This approach will not only help to reduce the emission 862–865.
of greenhouse gases into the atmosphere but also has the benefit of [21] J.W. Comerford, I.D.V. Ingram, M. North, X. Wu, Sustainable metal-based
catalysts for the synthesis of cyclic carbonates containing five-membered rings,
making useful products from the industrial flue gases. Green Chem. 17 (4) (2015) 1966–1987.
Future research also needs to focus on the development of direct [22] V. Etacheri, R. Marom, R. Elazari, G. Salitra, D. Aurbach, Challenges in the
cyclic carbonate synthesis by oxidative carboxylation of bio-derived development of advanced Li-ion batteries: a review, Energy Environ. Sci. 4 (9)
(2011) 3243–3262.
alkenes. This could be achieved by the development of a multifunc­ [23] V. Aravindan, J. Gnanaraj, S. Madhavi, H.K. Liu, Lithium-ion conducting
tional catalyst having the ability to catalyze both oxidation and cyclo­ electrolyte salts for lithium batteries, Chem. Eur. J. 17 (51) (2011) 14326–14346.
addition reactions with high selectivity. Synthesis of cyclic carbonates [24] H.R. Gillis, D. Stanssens, R. De Vos, A.R. Postema, D. Randall, Polymeric Foams,
Google Patents, 1997.
via this route will help to make the process more economical by
[25] J.H. Clements, Reactive applications of cyclic alkylene carbonates, Ind. Eng.
decreasing the cost involved in epoxidation, purification, handling of Chem. Res. 42 (4) (2003) 663–674.
epoxide and thus increase the sustainability of the process. [26] B. Geueke, C.C. Wagner, J. Muncke, Food contact substances and chemicals of
concern: a comparison of inventories, Food Addit. Contam. Part A 31 (8) (2014)
1438–1450.
Declaration of Competing Interest [27] G. Laugel, C.C. Rocha, P. Massiani, T. Onfroy, F. Launay, Homogeneous and
heterogeneous catalysis for the synthesis of cyclic and polymeric carbonates from
CO2 and epoxides: a mechanistic overview, Adv. Chem. Lett. 1 (3) (2013)
The authors declare that they have no known competing financial
195–214.
interests or personal relationships that could have appeared to influence [28] Z. Han, L. Rong, J. Wu, L. Zhang, Z. Wang, K. Ding, Catalytic hydrogenation of
the work reported in this paper. cyclic carbonates: a practical approach from CO2 and epoxides to methanol and
diols, Angew. Chem. 124 (52) (2012) 13218–13222.
[29] S. Fukuoka, M. Kawamura, K. Komiya, M. Tojo, H. Hachiya, K. Hasegawa,
Acknowledgments M. Aminaka, H. Okamoto, I. Fukawa, S. Konno, A novel non-phosgene
polycarbonate production process using by-product CO2 as starting material,
This work is supported by the Engineering and Physical Science Green Chem. 5 (5) (2003) 497–507.
[30] A. Rehman, V.C. Eze, M.G. Resul, A. Harvey, Kinetics and mechanistic
Research Council (EPSRC) funding for Sustainable Polymers (Project investigation of epoxide/CO2 cycloaddition by a synergistic catalytic effect of
reference EP/L017393/1). We are also grateful for the support of the pyrrolidinopyridinium iodide and zinc halides, J. Energy Chem. 37 (2019) 35–42.
University of Engineering and Technology, Lahore in providing the [31] A. Ferreira, C. Ferreira, J.A. Teixeira, F. Rocha, Temperature and solid properties
effects on gas–liquid mass transfer, Chem. Eng. J. 162 (2) (2010) 743–752.
scholarship to one of the authors. [32] T. Sakakura, K. Kohno, The synthesis of organic carbonates from carbon dioxide,
Chem. Commun. 11 (2009) 1312–1330.
References [33] P.P. Pescarmona, M. Taherimehr, Challenges in the catalytic synthesis of cyclic
and polymeric carbonates from epoxides and CO2, Catal. Sci. Technol. 2 (11)
(2012) 2169–2187.
[1] J.W. Comerford, I.D. Ingram, M. North, X. Wu, Sustainable metal-based catalysts
[34] J. Sun, J. Wang, W. Cheng, J. Zhang, X. Li, S. Zhang, Y. She, Chitosan
for the synthesis of cyclic carbonates containing five-membered rings, Green
functionalized ionic liquid as a recyclable biopolymer-supported catalyst for
Chem. 17 (4) (2015) 1966–1987.
cycloaddition of CO2, Green Chem. 14 (3) (2012) 654–660.
[2] M. North, Synthesis of cyclic carbonates from epoxides and carbon dioxide using
[35] C. Jing-Xian, J. Bi, D. Wei-Li, D. Sen-Lin, C. Liu-Ren, C. Zong-Jie, L. Sheng-Lian,
bimetallic aluminum(salen) complexes, ARKIVOC Online J. Org. Chem. 2012
L. Xu-Biao, T. Xin-Man, A. Chak-Tong, Catalytic fixation of CO2 to cyclic
(2012) 610–628.
carbonates over biopolymer chitosan-grafted quarternary phosphonium ionic
[3] M. Aresta, A. Dibenedetto, Utilisation of CO2 as a chemical feedstock:
liquid as a recylable catalyst, Appl. Catal. A Gen. 484 (2014) 26–32.
opportunities and challenges, Dalton Trans. 28 (2007) 2975–2992.
[36] T.-Y. Shi, J.-Q. Wang, J. Sun, M.-H. Wang, W.-G. Cheng, S.-J. Zhang, Efficient
[4] IPCC, https://www.co2.earth/global-co2-emissions, 2018.
fixation of CO 2 into cyclic carbonates catalyzed by hydroxyl-functionalized poly
[5] NIST database, NIST database: http://webbook.nist.gov/chemistry/name-ser.
(ionic liquids), RSC Adv. 3 (11) (2013) 3726–3732.
html. 2018.
[37] Y. Zhao, J.-S. Tian, X.-H. Qi, Z.-N. Han, Y.-Y. Zhuang, L.-N. He, Quaternary
[6] F. Barzagli, F. Mani, M. Peruzzini, From greenhouse gas to feedstock: formation of
ammonium salt-functionalized chitosan: an easily recyclable catalyst for efficient
ammonium carbamate from CO2 and NH3 in organic solvents and its catalytic
synthesis of cyclic carbonates from epoxides and carbon dioxide, J. Mol. Catal. A:
conversion into urea under mild conditions, Green Chem. 13 (5) (2011)
Chem. 271 (2007) 284–289.
1267–1274.
[38] T. Guadagno, S.G. Kazarian, High-pressure CO2 -expanded solvents: simultaneous
[7] I. Omae, Recent developments in carbon dioxide utilization for the production of
measurement of CO2 sorption and swelling of liquid polymers with in-situ near-IR
organic chemicals, Coord. Chem. Rev. 256 (13–14) (2012) 1384–1405.
spectroscopy, J. Phys. Chem. B 108 (37) (2004) 13995–13999.
[8] M. Aresta, A. Dibenedetto, The Contribution of the Utilization Option to Reducing
[39] M. Taherimehr, A. Decortes, S.M. Al-Amsyar, W. Lueangchaichaweng, C.
the CO2 Atmospheric Loading: Research Needed to Overcome Existing Barriers
J. Whiteoak, E.C. Escudero-Adán, A.W. Kleij, P.P. Pescarmona, A highly active Zn
for A Full Exploitation of the Potential of the CO2 Use, Elsevier, 2004.
(salphen) catalyst for production of organic carbonates in a green CO2 medium,
[9] C. Federsel, R. Jackstell, M. Beller, Moderne Katalysatoren zur Hydrierung von
Catal. Sci. Technol. 2 (11) (2012) 2231–2237.
Kohlendioxid, Angew. Chem. 122 (36) (2010) 6392–6395.
[40] Y. Xie, Z. Zhang, T. Jiang, J. He, B. Han, T. Wu, K. Ding, CO2 cycloaddition
[10] A.W. Kleij, M. North, A. Urakawa, CO2 catalysis, ChemSusChem 10 (6) (2017)
reactions catalyzed by an ionic liquid grafted onto a highly cross-linked polymer
1036–1038.
matrix, Angew. Chem. Int. Ed. 46 (38) (2007) 7255–7258.
[11] M. North, Synthesis of cyclic carbonates from epoxides and carbon dioxide using
[41] J.A. Castro-Osma, M. North, X. Wu, Development of a halide-free aluminium-
bimetallic aluminium (salen) complexes, ARKIVOC 1 (2012) 610–628.
based catalyst for the synthesis of cyclic carbonates from epoxides and carbon
[12] M. Aresta, E. Quaranta, Carbon dioxide: a substitute for phosgene, CHEMTECH 27
dioxide, Chem. Eur. J. 20 (46) (2014) 15005–15008.
(3) (1997).
[42] F. Jutz, A. Buchard, M.R. Kember, S.B. Fredriksen, C.K. Williams, Mechanistic
[13] M. Selva, A. Caretto, M. Noè, A. Perosa, Carbonate phosphonium salts as catalysts
investigation and reaction kinetics of the low-pressure copolymerization of
for the transesterification of dialkyl carbonates with diols. The competition
cyclohexene oxide and carbon dioxide catalyzed by a dizinc complex, J. Am.
between cyclic carbonates and linear dicarbonate products, Org. Biomol. Chem.
Chem. Soc. 133 (43) (2011) 17395–17405.
12 (24) (2014) 4143–4155.
[43] Y. Xiong, F. Bai, Z. Cui, N. Guo, R. Wang, Cycloaddition reaction of carbon
[14] V.A. Kuznetsov, M.G. Pervova, A.V. Pestov, Synthesis of alkylene carbonates in
dioxide to epoxides catalyzed by polymer-supported quaternary phosphonium
ionic liquid, Russ. J. Org. Chem. 12 (49) (2013) 1859–1860.
salts, J. Chem. 2013 (2013) (2013) 1–9.
[15] B. Gabriele, R. Mancuso, G. Salerno, L. Veltri, M. Costa, A. Dibenedetto, A general
[44] S. Liang, H. Liu, T. Jiang, J. Song, G. Yang, B. Han, Highly efficient synthesis of
and expedient synthesis of 5–and 6–membered cyclic carbonates by palladium-
cyclic carbonates from CO2 and epoxides over cellulose/KI, Chem. Commun. 47
catalyzed oxidative carbonylation of 1, 2–and 1, 3–diols, ChemSusChem 4 (12)
(7) (2011) 2131–2133.
(2011) 1778–1786.

24
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

[45] A. Decortes, M.M. Belmonte, J. Benet-Buchholz, A.W. Kleij, Efficient carbonate the ligand on catalytic cyclic carbonate synthesis from epoxide and carbon
synthesis under mild conditions through cycloaddition of carbon dioxide to dioxide, Dalton Trans. 42 (36) (2013) 13151–13160.
oxiranes using a Zn (salphen) catalyst, Chem. Commun. 46 (25) (2010) [72] J. Qin, P. Wang, Q. Li, Y. Zhang, D. Yuan, Y. Yao, Catalytic production of cyclic
4580–4582. carbonates mediated by lanthanide phenolates under mild conditions, Chem.
[46] W. Clegg, R.W. Harrington, M. North, R. Pasquale, Cyclic carbonate synthesis Commun. 50 (75) (2014) 10952–10955.
catalysed by bimetallic aluminium–salen complexes, Chem. Eur. J. 16 (23) (2010) [73] S.L. Kristufek, K.T. Wacker, Y.-Y.T. Tsao, L. Su, K.L. Wooley, Monomer design
6828–6843. strategies to create natural product-based polymer materials, Nat. Prod. Rep. 34
[47] C.J. Whiteoak, E. Martin, M.M. Belmonte, J. Benet-Buchholz, A.W. Kleij, An (4) (2017) 433–459.
efficient iron catalyst for the synthesis of five-and six-membered organic [74] E. de Jong, A. Higson P. Walsh, M. Wellisch, Bio-based chemicals value added
carbonates under mild conditions, Adv. Synth. Catal. 354 (2–3) (2012) 469–476. products from biorefineries, IEA Bioenergy, Task42 Biorefinery, 2012.
[48] X.B. Lu, Y. Wang, Highly active, binary catalyst systems for the alternating [75] M. Firdaus, L. Montero de Espinosa, M.A.R. Meier, Terpene-based renewable
copolymerization of CO2 and epoxides under mild conditions, Angew. Chem. Int. monomers and polymers via thiol–ene additions, Macromolecules 44 (18) (2011)
Ed. 43 (27) (2004) 3574–3577. 7253–7262.
[49] M.H. Chisholm, Z. Zhou, Concerning the mechanism of the ring opening of [76] A. Rehman, E. Russell, F. Saleem, F. Javed, S. Ahmad, V.C. Eze, A. Harvey,
propylene oxide in the copolymerization of propylene oxide and carbon dioxide Synthesis of trans-limonene bis-epoxide by stereoselective epoxidation of (R)-(+)-
to give poly (propylene carbonate), J. Am. Chem. Soc. 126 (35) (2004) limonene, J. Environ. Chem. Eng. (2020), 104680.
11030–11039. [77] R. Ciriminna, M. Lomeli-Rodriguez, P.D. Carà, J.A. Lopez-Sanchez, M. Pagliaro,
[50] A. Buchard, M.R. Kember, K.G. Sandeman, C.K. Williams, A bimetallic iron (III) Limonene: a versatile chemical of the bioeconomy, Chem. Commun. 50 (97)
catalyst for CO2/epoxide coupling, Chem. Commun. 47 (1) (2011) 212–214. (2014) 15288–15296.
[51] M. Taherimehr, S.M. Al-Amsyar, C.J. Whiteoak, A.W. Kleij, P.P. Pescarmona, [78] M.F.M.G. Resul, A.M.L. Fernández, A. Rehman, A.P. Harvey, Development of a
High activity and switchable selectivity in the synthesis of cyclic and polymeric selective, solvent-free epoxidation of limonene using hydrogen peroxide and a
cyclohexene carbonates with iron amino triphenolate catalysts, Green Chem. 15 tungsten-based catalyst, React. Chem. Eng. 3 (5) (2018) 747–756.
(11) (2013) 3083–3090. [79] Y. Zhu, C. Romain, C.K. Williams, Sustainable polymers from renewable
[52] A. Decortes, A.W. Kleij, Ambient fixation of carbon dioxide using a ZnIIsalphen resources, Nature 540 (7633) (2016) 354–362.
catalyst, ChemCatChem 3 (5) (2011) 831–834. [80] C. Ampelli, S. Perathoner, G. Centi, CO2 utilization: an enabling element to move
[53] C.J. Whiteoak, E. Martin, E. Escudero-Adán, A.W. Kleij, Stereochemical to a resource-and energy-efficient chemical and fuel production, Philos. Trans. R.
divergence in the formation of organic carbonates derived from internal epoxides, Soc. A 373 (2037) (2015), 20140177.
Adv. Synth. Catal. 355 (11–12) (2013) 2233–2239. [81] M. Bähr, R. Mülhaupt, Linseed and soybean oil-based polyurethanes prepared via
[54] X. Sun, H. Lee, S. Lee, K.L. Tan, Catalyst recognition of cis-1, 2-diols enables site- the non-isocyanate route and catalytic carbon dioxide conversion, Green Chem.
selective functionalization of complex molecules, Nat. Chem. 5 (9) (2013) 14 (2) (2012) 483–489.
790–795. [82] G. Fiorani, M. Stuck, C. Martín, M.M. Belmonte, E. Martin, E.C. Escudero-Adán, A.
[55] A. Archelas, R. Furstoss, Synthesis of enantiopure epoxides through biocatalytic W. Kleij, Catalytic coupling of carbon dioxide with terpene scaffolds: access to
approaches, Annu. Rev. Microbiol. 51 (1) (1997) 491–525. challenging bio-based organic carbonates, ChemSusChem 9 (11) (2016)
[56] V. Schimpf, B.S. Ritter, P. Weis, K. Parison, R. Mu¨lhaupt, High purity limonene 1304–1311.
dicarbonate as versatile building block for sustainable non-isocyanate [83] J. Martínez, J. Fernández-Baeza, L.F. Sánchez-Barba, J.A. Castro-Osma, A. Lara-
polyhydroxyurethane thermosets and thermoplastics, Macromolecules 50 (3) Sánchez, A. Otero, An efficient and versatile lanthanum heteroscorpionate
(2017) 944–955. catalyst for carbon dioxide fixation into cyclic carbonates, ChemSusChem 10 (14)
[57] X. Wu, J. Castro-Osma, M. North, Synthesis of chiral cyclic carbonates via kinetic (2017) 2886–2890.
resolution of racemic epoxides and carbon dioxide, Symmetry 8 (1) (2016) 4. [84] M. Hiroshi, M. Masato, G. Yuuta, Y. Jun-ichi, M. Hisatoyo, M. Suguru, Two
[58] G.-P. Wu, S.-H. Wei, X.-B. Lu, W.-M. Ren, D.J. Darensbourg, Highly selective diastereomers of d-limonene-derived cyclic carbonates from d-limonene oxide
synthesis of CO2 copolymer from styrene oxide, Macromolecules 43 (21) (2010) and carbon dioxide with a tetrabutylammonium chloride catalyst, Bull. Chem.
9202–9204. Soc. Jpn. 91 (1) (2018) 92–94.
[59] F. Castro-Gómez, G. Salassa, A.W. Kleij, C. Bo, A DFT study on the mechanism of [85] A. Rehman, A.M.L. Fernández, M.G. Resul, A. Harvey, Highly selective,
the cycloaddition reaction of CO2 to epoxides catalyzed by Zn (Salphen) sustainable synthesis of limonene cyclic carbonate from bio-based limonene oxide
complexes, Chem. - Eur. J. 19 (20) (2013) 6289–6298. and CO2: a kinetic study, J. CO2 Util. 29 (2019) 126–133.
[60] C. Martin, G. Fiorani, A.W. Kleij, Recent advances in the catalytic preparation of [86] H. Morikawa, M. Minamoto, Y. Gorou, J.-i Yamaguchi, H. Morinaga,
cyclic organic carbonates, ACS Catal. 5 (2) (2015) 1353–1370. S. Motokucho, Two diastereomers of d-limonene-derived cyclic carbonates from
[61] C.M. Byrne, S.D. Allen, E.B. Lobkovsky, G.W. Coates, Alternating d-limonene oxide and carbon dioxide with a tetrabutylammonium chloride
copolymerization of limonene oxide and carbon dioxide, J. Am. Chem. Soc. 126 catalyst, Bull. Chem. Soc. Jpn. 91 (1) (2018) 92–94.
(37) (2004) 11404–11405. [87] A. Rehman, M.G. Resul, V.C. Eze, A. Harvey, A kinetic study of Zn halide/TBAB-
[62] J. Langanke, L. Greiner, W. Leitner, Substrate dependent synergetic and catalysed fixation of CO2 with styrene oxide in propylene carbonate, Green
antagonistic interaction of ammonium halide and polyoxometalate catalysts in Process. Synth. 8 (1) (2019) 719–729.
the synthesis of cyclic carbonates from oleochemical epoxides and CO2, Green [88] M. North, R. Pasquale, Mechanism of cyclic carbonate synthesis from epoxides
Chem. 15 (5) (2013) 1173–1182. and CO2, Angew. Chem. 121 (16) (2009) 2990–2992.
[63] M. Bähr, A. Bitto, R. Mülhaupt, Cyclic limonene dicarbonate as a new monomer [89] S. Supasitmongkol, P. Styring, A single centre aluminium (III) catalyst and TBAB
for non-isocyanate oligo-and polyurethanes (NIPU) based upon terpenes, Green as an ionic organo-catalyst for the homogeneous catalytic synthesis of styrene
Chem. 14 (5) (2012) 1447–1454. carbonate, Catal. Sci. Technol. 4 (6) (2014) 1622–1630.
[64] A. Durkin, I. Taptygin, Q. Kong, M.F.G. Resul, A. Rehman, A.M. Fernández, A. [90] R. Luo, X. Zhou, Y. Fang, H. Ji, Metal-and solvent-free synthesis of cyclic
P. Harvey, N. Shah, M. Guo, Scale-up and sustainability evaluation of biopolymer carbonates from epoxides and CO2 in the presence of graphite oxide and ionic
production from citrus waste offering carbon capture and utilisation pathway, liquid under mild conditions: a kinetic study, Carbon 82 (2015) 1–11.
ChemistryOpen 8 (6) (2019) 668–688. [91] L. Cuesta-Aluja, J. Castilla, A.M. Masdeu-Bultó, Aluminium salabza complexes for
[65] D.J. Darensbourg, J.C. Yarbrough, C. Ortiz, C.C. Fang, Comparative kinetic fixation of CO2 to organic carbonates, Dalton Trans. 45 (37) (2016)
studies of the copolymerization of cyclohexene oxide and propylene oxide with 14658–14667.
carbon dioxide in the presence of chromium salen derivatives. In situ FTIR [92] J.A. Castro-Osma, K.J. Lamb, M. North, Cr (salophen) complex catalyzed cyclic
measurements of copolymer vs cyclic carbonate production, J. Am. Chem. Soc. carbonate synthesis at ambient temperature and pressure, ACS Catal. 6 (8) (2016)
125 (25) (2003) 7586–7591. 5012–5025.
[66] C. Martín, G. Fiorani, A.W. Kleij, Recent advances in the catalytic preparation of [93] R. Luo, W. Zhang, Z. Yang, X. Zhou, H. Ji, Synthesis of cyclic carbonates from
cyclic organic carbonates, ACS Catal. 5 (2) (2015) 1353–1370. epoxides over bifunctional salen aluminum oligomers as a CO2 -philic catalyst:
[67] C.J. Whiteoak, B. Gjoka, E. Martin, M.M. Belmonte, E.C. Escudero-Adán, C. Zonta, catalytic and kinetic investigation, J. CO2 Util. 19 (2017) 257–265.
G. Licini, A.W. Kleij, Reactivity control in iron (III) amino triphenolate complexes: [94] J.C. Pastre, D.L. Browne, S.V. Ley, Flow chemistry syntheses of natural products,
comparison of monomeric and dimeric complexes, Inorg. Chem. 51 (20) (2012) Chem. Soc. Rev. 42 (23) (2013) 8849–8869.
10639–10649. [95] H. Löwe, V. Hessel, P. Löb, S. Hubbard, Addition of secondary amines to α,
[68] C. Beattie, M. North, P. Villuendas, C. Young, Influence of temperature and β-unsaturated carbonyl compounds and nitriles by using microstructured
pressure on cyclic carbonate synthesis catalyzed by bimetallic aluminum reactors, Org. Process Res. Dev. 10 (6) (2006) 1144–1152.
complexes and application to overall syn-bis-hydroxylation of alkenes, J. Org. [96] S.G. Newman, K.F. Jensen, The role of flow in green chemistry and engineering,
Chem. 78 (2) (2013) 419–426. Green Chem. 15 (6) (2013) 1456–1472.
[69] C.J. Whiteoak, N. Kielland, V. Laserna, E.C. Escudero-Adán, E. Martin, A.W. Kleij, [97] J. Yue, G. Chen, Q. Yuan, L. Luo, Y. Gonthier, Hydrodynamics and mass transfer
A powerful aluminum catalyst for the synthesis of highly functional organic characteristics in gas–liquid flow through a rectangular microchannel, Chem.
carbonates, J. Am. Chem. Soc. 135 (4) (2013) 1228–1231. Eng. Sci. 62 (7) (2007) 2096–2108.
[70] H.B. Vignesh, K. Muralidharan, Zn (II), Cd (II) and Cu (II) complexes of 2, 5-bis [98] C.J. Mallia, I.R. Baxendale, The use of gases in flow synthesis, Org. Process Res.
{N-(2, 6-diisopropylphenyl) iminomethyl} pyrrole: synthesis, structures and their Dev. 20 (2) (2015) 327–360.
high catalytic activity for efficient cyclic carbonate synthesis, Dalton Trans. 42 (4) [99] Y. Zhao, C. Yao, G. Chen, Q. Yuan, Highly efficient synthesis of cyclic carbonate
(2013) 1238–1248 (Cambridge, England: 2003). with CO2 catalyzed by ionic liquid in a microreactor, Green Chem. 15 (2) (2013)
[71] P. Ramidi, N. Gerasimchuk, Y. Gartia, C.M. Felton, A. Ghosh, Synthesis and 446–452.
characterization of Co (III) amidoamine complexes: influence of substituents of

25
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

[100] I.S. Metcalfe, M. North, P. Villuendas, Influence of reactor design on cyclic [128] D. Kim, S. Subramanian, D. Thirion, Y. Song, A. Jamal, M.S. Otaibi, C.T. Yavuz,
carbonate synthesis catalysed by a bimetallic aluminium (salen) complex, J. CO2 Quaternary ammonium salt grafted nanoporous covalent organic polymer for
Util. 2 (2013) 24–28. atmospheric CO2 fixation and cyclic carbonate formation, Catal. Today, 2020.
[101] D. Lokhat, A.K. Domah, K. Padayachee, A. Baboolal, D. Ramjugernath, Gas–liquid [129] R.E. Del Sesto, C. Corley, A. Robertson, J.S. Wilkes, Tetraalkylphosphonium-based
mass transfer in a falling film microreactor: effect of reactor orientation on liquid- ionic liquids, J. Organomet. Chem. 690 (10) (2005) 2536–2542.
side mass transfer coefficient, Chem. Eng. Sci. 155 (2016) 38–44. [130] A. Sibaouih, P. Ryan, M. Leskelä, B. Rieger, T. Repo, Facile synthesis of cyclic
[102] M. North, P. Villuendas, C. Young, A gas-phase flow reactor for ethylene carbonates from CO2 and epoxides with cobalt (II)/onium salt based catalysts,
carbonate synthesis from waste carbon dioxide, Chem. Eur. J. 15 (43) (2009) Appl. Catal. A Gen. 365 (2) (2009) 194–198.
11454–11457. [131] J. Sun, J. Ren, S. Zhang, W. Cheng, Water as an efficient medium for the synthesis
[103] J. Meléndez, M. North, P. Villuendas, C. Young, One-component bimetallic of cyclic carbonate, Tetrahedron Lett. 50 (4) (2009) 423–426.
aluminium (salen)-based catalysts for cyclic carbonate synthesis and their [132] H. Zhou, G.-X. Wang, W.-Z. Zhang, X.-B. Lu, CO2 adducts of phosphorus ylides:
immobilization, Dalton Trans. 40 (15) (2011) 3885–3902. highly active organocatalysts for carbon dioxide transformation, ACS Catal. 5 (11)
[104] J.A. Kozak, J. Wu, X. Su, F. Simeon, T.A. Hatton, T.F. Jamison, Bromine-catalyzed (2015) 6773–6779.
conversion of CO2 and epoxides to cyclic carbonates under continuous flow [133] Y. Toda, Y. Komiyama, A. Kikuchi, H. Suga, Tetraarylphosphonium salt-catalyzed
conditions, J. Am. Chem. Soc. 135 (49) (2013) 18497–18501. carbon dioxide fixation at atmospheric pressure for the synthesis of cyclic
[105] B.-H. Xu, J.-Q. Wang, J. Sun, Y. Huang, J.-P. Zhang, X.-P. Zhang, S.-J. Zhang, carbonates, ACS Catal. 6 (10) (2016) 6906–6910.
Fixation of CO2 into cyclic carbonates catalyzed by ionic liquids: a multi-scale [134] S. Liu, N. Suematsu, K. Maruoka, S. Shirakawa, Design of bifunctional quaternary
approach, Green Chem. 17 (1) (2015) 108–122. phosphonium salt catalysts for CO2 fixation reaction with epoxides under mild
[106] A. Rehman, A.M.L. Fernandez, M.G. Resul, A. Harvey, Kinetic investigations of conditions, Green Chem. 18 (17) (2016) 4611–4615.
styrene carbonate synthesis from styrene oxide and CO2 using a continuous flow [135] T. Takahashi, T. Watahiki, S. Kitazume, H. Yasuda, T. Sakakura, Synergistic
tube-in-tube gas-liquid reactor, J. CO2 Util. 24 (2018) 341–349. hybrid catalyst for cyclic carbonate synthesis: remarkable acceleration caused by
[107] J.-Q. Wang, K. Dong, W.-G. Cheng, J. Sun, S.-J. Zhang, Insights into quaternary immobilization of homogeneous catalyst on silica, Chem. Commun. 15 (2006)
ammonium salts-catalyzed fixation carbon dioxide with epoxides, Catal. Sci. 1664–1666.
Technol. 2 (7) (2012) 1480–1484. [136] W. Dai, P. Mao, Y. Liu, S. Zhang, B. Li, L. Yang, X. Luo, J. Zou, Quaternary
[108] H. Buettner, L. Longwitz, J. Steinbauer, C. Wulf, T. Werner, Recent developments phosphonium salt-functionalized Cr-MIL-101: a bifunctional and efficient catalyst
in the synthesis of cyclic carbonates from epoxides and CO2, Top. Curr. Chem. 375 for CO2 cycloaddition with epoxides, J. CO2 Util. 36 (2020) 295–305.
(3) (2017) 50 (Z.). [137] F.D. Bobbink, P.J. Dyson, Synthesis of carbonates and related compounds
[109] J. Melendez, M. North, R. Pasquale, Synthesis of cyclic carbonates from incorporating CO2 using ionic liquid-type catalysts: state-of-the-art and beyond,
atmospheric pressure carbon dioxide using exceptionally active aluminium J. Catal. 343 (2016) 52–61.
(salen) complexes as catalysts, Eur. J. Inorg. Chem. 2007 (21) (2007) 3323–3326. [138] J. Peng, Y. Deng, Cycloaddition of carbon dioxide to propylene oxide catalyzed by
[110] M. Liu, J. Lan, L. Liang, J. Sun, M. Arai, Heterogeneous catalytic conversion of ionic liquids, New J. Chem. 25 (4) (2001) 639–641.
CO2 and epoxides to cyclic carbonates over multifunctional tri-s-triazine terminal- [139] H. Kawanami, A. Sasaki, K. Matsui, Y. Ikushima, A rapid and effective synthesis of
linked ionic liquids, J. Catal. 347 (2017) 138–147. propylene carbonate using a supercritical CO2 –ionic liquid system, Chem.
[111] K. Yu, C.W. Jones, Silica-immobilized zinc β-diiminate catalysts for the Commun. 7 (2003) 896–897.
copolymerization of epoxides and carbon dioxide, Organometallics 22 (13) [140] J. Sun, S. Zhang, W. Cheng, J. Ren, Hydroxyl-functionalized ionic liquid: a novel
(2003) 2571–2580. efficient catalyst for chemical fixation of CO2 to cyclic carbonate, Tetrahedron
[112] M. Liu, X. Lu, L. Shi, F. Wang, J. Sun, Periodic mesoporous organosilica with a Lett. 49 (22) (2008) 3588–3591.
basic urea-derived framework for enhanced carbon dioxide capture and [141] J.-Q. Wang, W.-G. Cheng, J. Sun, T.-Y. Shi, X.-P. Zhang, S.-J. Zhang, Efficient
conversion under mild conditions, ChemSusChem 10 (6) (2017) 1110–1119. fixation of CO2 into organic carbonates catalyzed by 2-hydroxymethyl-function­
[113] M. Cokoja, M.E. Wilhelm, M.H. Anthofer, W.A. Herrmann, F.E. Kühn, Synthesis of alized ionic liquids, RSC Adv. 4 (5) (2014) 2360–2367.
cyclic carbonates from epoxides and carbon dioxide by using organocatalysts, [142] J. Sun, L. Han, W. Cheng, J. Wang, X. Zhang, S. Zhang, Efficient acid–base
ChemSusChem 8 (15) (2015) 2436–2454. bifunctional catalysts for the fixation of CO2 with epoxides under metal-and
[114] V. Calo, A. Nacci, A. Monopoli, A. Fanizzi, Cyclic carbonate formation from solvent-free conditions, ChemSusChem 4 (4) (2011) 502–507.
carbon dioxide and oxiranes in tetrabutylammonium halides as solvents and [143] M.H. Anthofer, M.E. Wilhelm, M. Cokoja, I.I.E. Markovits, A. Pöthig, J. Mink, W.
catalysts, Org. Lett. 4 (15) (2002) 2561–2563. A. Herrmann, F.E. Kühn, Cycloaddition of CO2 and epoxides catalyzed by
[115] J. Steinbauer, C. Kubis, R. Ludwig, T. Werner, Mechanistic study on the addition imidazolium bromides under mild conditions: influence of the cation on catalyst
of CO2 to epoxides catalyzed by ammonium and phosphonium salts: a combined activity, Catal. Sci. Technol. 4 (6) (2014) 1749–1758.
spectroscopic and kinetic approach, ACS Sustain. Chem. Eng. 2018. [144] Y. Li, B. Dominelli, R.M. Reich, B. Liu, F.E. Kühn, Bridge-functionalized
[116] F. Jutz, J.-M. Andanson, A. Baiker, Ionic liquids and dense carbon dioxide: a bisimidazolium bromides as catalysts for the conversion of epoxides to cyclic
beneficial biphasic system for catalysis, Chem. Rev. 111 (2) (2010) 322–353. carbonates with CO2, Catal. Commun. 124 (2019) 118–122.
[117] K. Motokura, S. Itagaki, Y. Iwasawa, A. Miyaji, T. Baba, Silica-supported [145] X. Meng, Z. Ju, S. Zhang, X. Liang, N. von Solms, X. Zhang, X. Zhang, Efficient
aminopyridinium halides for catalytic transformations of epoxides to cyclic transformation of CO2 to cyclic carbonates using bifunctional protic ionic liquids
carbonates under atmospheric pressure of carbon dioxide, Green Chem. 11 (11) under mild conditions, Green Chem. 21 (12) (2019) 3456–3463.
(2009) 1876–1880. [146] G. Long, D. Wu, H. Pan, T. Zhao, X. Hu, Imidazolium hydrogen carbonate ionic
[118] A. Rehman, F. Saleem, F. Javed, H. Qutab, V.C. Eze, A. Harvey, Kinetic study for liquids: versatile organocatalysts for chemical conversion of CO2 into valuable
styrene carbonate synthesis via CO2 cycloaddition to styrene oxide using silica- chemicals, J. CO2 Util. 39 (2020), 101155.
supported pyrrolidinopyridinium iodide catalyst, J. CO2 Util. 43 (2021), 101379. [147] L. Ji, Z. Luo, Y. Zhang, R. Wang, Y. Ji, F. Xia, G. Gao, Imidazolium ionic liquids/
[119] T. Takahashi, T. Watahiki, S. Kitazume, H. Yasuda, T. Sakakura, Synergistic organic bases: efficient intermolecular synergistic catalysts for the cycloaddition
hybrid catalyst for cyclic carbonate Synthesis, Synfacts 2006 (07) (2006), 0736- of CO2 and epoxides under atmospheric pressure, Mol. Catal. 446 (2018)
0736. 124–130.
[120] A.R. Hajipour, Y. Heidari, G. Kozehgary, Silica grafted ammonium salts based on [148] M.K. Leu, I. Vicente, J.A. Fernandes, I. De Pedro, J. Dupont, V. Sans, P. Licence,
DABCO as heterogeneous catalysts for cyclic carbonate synthesis from carbon A. Gual, I. Cano, On the real catalytically active species for CO2 fixation into
dioxide and epoxides, RSC Adv. 5 (29) (2015) 22373–22379. cyclic carbonates under near ambient conditions: dissociation equilibrium of
[121] T. Sakai, Y. Tsutsumi, T. Ema, Highly active and robust organic–inorganic hybrid [BMIm][Fe (NO) 2Cl2] dependant on reaction temperature, Appl. Catal. B
catalyst for the synthesis of cyclic carbonates from carbon dioxide and epoxides, Environ. 245 (2019) 240–250.
Green Chem. 10 (3) (2008) 337–341. [149] A.A. Pawar, H. Kim, Reaction parameters dependence of the CO2/epoxide
[122] T. Seki, J.-D. Grunwaldt, A. Baiker, In situ attenuated total reflection infrared coupling reaction catalyzed by tunable ionic liquids, optimization of comonomer-
spectroscopy of imidazolium-based room-temperature ionic liquids under alternating enhancement pathway, J. CO2 Util. 33 (2019) 500–512.
“supercritical” CO2, J. Phys. Chem. B 113 (1) (2008) 114–122. [150] T. Wang, X. Zhu, L. Mao, Y. Liu, T. Ren, L. Wang, J. Zhang, Synergistic
[123] H. Xie, H. Duan, S. Li, S. Zhang, The effective synthesis of propylene carbonate cooperation of bi-active hydrogen atoms in protic carboxyl imidazolium ionic
catalyzed by silica-supported hexaalkylguanidinium chloride, New J. Chem. 29 liquids to push cycloaddition of CO2 under benign conditions, J. Mol. Liq. 296
(9) (2005) 1199–1203. (2019), 111936.
[124] J.-Q. Wang, D.-L. Kong, J.-Y. Chen, F. Cai, L.-N. He, Synthesis of cyclic carbonates [151] R.B. Mujmule, M.R. Rao, P.V. Rathod, V.G. Deonikar, A. Chaugule, H. Kim,
from epoxides and carbon dioxide over silica-supported quaternary ammonium Synergistic effect of a binary ionic liquid/base catalytic system for efficient
salts under supercritical conditions, J. Mol. Catal. A Chem. 249 (2006) 143–148. conversion of epoxide and carbon dioxide into cyclic carbonates, J. CO2 Util. 33
[125] Y. Du, J.-Q. Wang, J.-Y. Chen, F. Cai, J.-S. Tian, D.-L. Kong, L.-N. He, A poly (2019) 284–291.
(ethylene glycol)-supported quaternary ammonium salt for highly efficient and [152] J.-Q. Wang, X.-D. Yue, F. Cai, L.-N. He, Solventless synthesis of cyclic carbonates
environmentally friendly chemical fixation of CO2 with epoxides under from carbon dioxide and epoxides catalyzed by silica-supported ionic liquids
supercritical conditions, Tetrahedron Lett. 47 (8) (2006) 1271–1275. under supercritical conditions, Catal. Commun. 8 (2) (2007) 167–172.
[126] X. Chen, J. Sun, J. Wang, W. Cheng, Polystyrene-bound diethanolamine based [153] L. Han, H.-J. Choi, S.-J. Choi, B. Liu, D.-W. Park, Ionic liquids containing carboxyl
ionic liquids for chemical fixation of CO2, Tetrahedron Lett. 53 (22) (2012) acid moieties grafted onto silica: synthesis and application as heterogeneous
2684–2688. catalysts for cycloaddition reactions of epoxide and carbon dioxide, Green Chem.
[127] J. Tharun, D.W. Kim, R. Roshan, Y. Hwang, D.-W. Park, Microwave assisted 13 (4) (2011) 1023–1028.
preparation of quaternized chitosan catalyst for the cycloaddition of CO2 and [154] R.A. Watile, K.M. Deshmukh, K.P. Dhake, B.M. Bhanage, Efficient synthesis of
epoxides, Catal. Commun. 31 (2013) 62–65. cyclic carbonate from carbon dioxide using polymer anchored diol functionalized

26
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

ionic liquids as a highly active heterogeneous catalyst, Catal. Sci. Technol. 2 (5) [181] X.-B. Lu, Y.-J. Zhang, B. Liang, X. Li, H. Wang, Chemical fixation of carbon
(2012) 1051–1055. dioxide to cyclic carbonates under extremely mild conditions with highly active
[155] K.R. Roshan, G. Mathai, J. Kim, J. Tharun, G.-A. Park, D.-W. Park, A biopolymer bifunctional catalysts, J. Mol. Catal. A Chem. 210 (1–2) (2004) 31–34.
mediated efficient synthesis of cyclic carbonates from epoxides and carbon [182] J. Melendez, M. North, P. Villuendas, One-component catalysts for cyclic
dioxide, Green Chem. 14 (10) (2012) 2933–2940. carbonate synthesis, Chem. Commun. 18 (2009) 2577–2579.
[156] Y. Zhang, S. Yin, S. Luo, C.T. Au, Cycloaddition of CO2 to epoxides catalyzed by [183] M. North, C. Young, Reducing the cost of production of bimetallic aluminium
carboxyl-functionalized imidazolium-based ionic liquid grafted onto cross-linked catalysts for the synthesis of cyclic carbonates, ChemSusChem 4 (11) (2011)
polymer, Ind. Eng. Chem. Res. 51 (10) (2012) 3951–3957. 1685–1693.
[157] W. Hui, X.-M. He, X.-Y. Xu, Y.-M. Chen, Y. Zhou, Z.-M. Li, L. Zhang, D.-J. Tao, [184] M. North, B. Wang, C. Young, Influence of flue gas on the catalytic activity of an
Highly efficient cycloaddition of diluted and waste CO2 into cyclic carbonates immobilized aluminium (salen) complex for cyclic carbonate synthesis, Energy
catalyzed by porous ionic copolymers, J. CO2 Util. 36 (2020) 169–176. Environ. Sci. 4 (10) (2011) 4163–4170.
[158] G. Chen, Y. Zhang, J. Xu, X. Liu, K. Liu, M. Tong, Z. Long, Imidazolium-based [185] M. North, P. Villuendas, C. Young, Inter-and intramolecular phosphonium salt
ionic porous hybrid polymers with POSS-derived silanols for efficient cocatalysis in cyclic carbonate synthesis catalysed by a bimetallic aluminium
heterogeneous catalytic CO2 conversion under mild conditions, Chem. Eng. J. 381 (salen) complex, Tetrahedron Lett. 53 (22) (2012) 2736–2740.
(2020), 122765. [186] X. Wu, M. North, A bimetallic aluminium (salphen) complex for the synthesis of
[159] Z. Dokhaee, M. Ghiaci, H. Farrokhpour, G. Buntkowsky, H. Breitzke, SBA-15 cyclic carbonates from epoxides and carbon dioxide, ChemSusChem 10 (1) (2017)
supported imidazolium ionic liquid through different linkers as a sustainable 74–78.
catalyst for the synthesis of cyclic carbonates: a kinetic study and theoretical DFT [187] J. Liu, G. Yang, Y. Liu, D. Zhang, X. Hu, Z. Zhang, Efficient conversion of CO2 into
calculations, Ind. Eng. Chem. Res. 2020. cyclic carbonates at room temperature catalyzed by Al-salen and imidazolium
[160] Y. Wang, J. Nie, C. Lu, F. Wang, C. Ma, Z. Chen, G. Yang, Imidazolium-based hydrogen carbonate ionic liquid, Green Chem. 2020.
polymeric ionic liquids for heterogeneous catalytic conversion of CO2 into cyclic [188] J. Meléndez, M. North, R. Pasquale, Synthesis of cyclic carbonates from
carbonates, Microporous Mesoporous Mater. 292 (2020), 109751. atmospheric pressure carbon dioxide using exceptionally active aluminium
[161] W. Zhang, F. Ma, L. Ma, Y. Zhou, J. Wang, Imidazolium-functionalized ionic (salen) complexes as catalysts, Eur. J. Inorg. Chem. 2007 (21) (2007) 3323–3326.
hypercrosslinked porous polymers for efficient synthesis of cyclic carbonates from [189] J. Meléndez, M. North, P. Villuendas, One-component catalysts for cyclic
simulated flue gas, ChemSusChem 13 (2) (2020) 341–350. carbonate synthesis, Chem. Commun. 18 (2009) 2577–2579.
[162] M. Liu, X. Wang, Y. Jiang, J. Sun, M. Arai, Hydrogen bond activation strategy for [190] M. North, C. Young, Bimetallic aluminium (acen) complexes as catalysts for the
cyclic carbonates synthesis from epoxides and CO2: current state-of-the art of synthesis of cyclic carbonates from carbon dioxide and epoxides, Catal. Sci.
catalyst development and reaction analysis, Catal. Rev. 2018. Technol. 1 (1) (2011) 93–99.
[163] A.J.R. Amaral, J.F.J. Coelho, A.C. Serra, Synthesis of bifunctional cyclic [191] J.A. Castro-Osma, M. North, X. Wu, Synthesis of cyclic carbonates catalysed by
carbonates from CO2 catalysed by choline-based systems, Tetrahedron Lett. 54 chromium and aluminium salphen complexes, Chem. Eur. J. 22 (6) (2016)
(40) (2013) 5518–5522. 2100–2107.
[164] K.R. Roshan, T. Jose, D. Kim, K.A. Cherian, D.W. Park, Microwave-assisted one [192] R.L. Paddock, S.T. Nguyen, Chemical CO2 fixation: Cr (III) salen complexes as
pot-synthesis of amino acid ionic liquids in water: simple catalysts for styrene highly efficient catalysts for the coupling of CO2 and epoxides, J. Am. Chem. Soc.
carbonate synthesis under atmospheric pressure of CO2, Catal. Sci. Technol. 4 (4) 123 (46) (2001) 11498–11499.
(2014) 963–970. [193] R.L. Paddock, S.T. Nguyen, Chiral (salen) Co III catalyst for the synthesis of cyclic
[165] H. Büttner, K. Lau, A. Spannenberg, T. Werner, Bifunctional one-component carbonates, Chem. Commun. 14 (2004) 1622–1623.
catalysts for the addition of carbon dioxide to epoxides, ChemCatChem 7 (3) [194] C. Martin, C.J. Whiteoak, E. Martin, M.M. Belmonte, E.C. Escudero-Adán, A.
(2015) 459–467. W. Kleij, Easily accessible bifunctional Zn (salpyr) catalysts for the formation of
[166] Y.M. Shen, W.L. Duan, M. Shi, Phenol and organic bases co-catalyzed chemical organic carbonates, Catal. Sci. Technol. 4 (6) (2014) 1615–1621.
fixation of carbon dioxide with terminal epoxides to form cyclic carbonates, Adv. [195] Y. Xie, T.T. Wang, R.X. Yang, N.Y. Huang, K. Zou, W.Q. Deng, Efficient fixation of
Synth. Catal. 345 (3) (2003) 337–340. CO2 by a zinc-coordinated conjugated microporous polymer, ChemSusChem 7 (8)
[167] C.J. Whiteoak, A. Nova, F. Maseras, A.W. Kleij, Merging sustainability with (2014) 2110–2114.
organocatalysis in the formation of organic carbonates by using CO2 as a [196] M.A. Fuchs, S. Staudt, C. Altesleben, O. Walter, T.A. Zevaco, E. Dinjus, A new air-
feedstock, ChemSusChem 5 (10) (2012) 2032–2038. stable zinc complex based on a 1, 2-phenylene-diimino-2-cyanoacrylate ligand as
[168] A.M. Hardman-Baldwin, A.E. Mattson, Silanediol-catalyzed carbon dioxide an efficient catalyst of the epoxide–CO2 coupling, Dalton Trans. 43 (6) (2014)
fixation, ChemSusChem 7 (12) (2014) 3275–3278. 2344–2347.
[169] M.E. Wilhelm, M.H. Anthofer, M. Cokoja, I.I.E. Markovits, W.A. Herrmann, F. [197] L. Qu, I. del Rosal, Q. Li, Y. Wang, D. Yuan, Y. Yao, L. Maron, Efficient CO2
E. Kühn, Cycloaddition of carbon dioxide and epoxides using pentaerythritol and transformation under ambient condition by heterobimetallic rare earth
halides as dual catalyst system, ChemSusChem 7 (5) (2014) 1357–1360. complexes: experimental and computational evidences of a synergistic effect,
[170] J. Sun, W. Cheng, Z. Yang, J. Wang, T. Xu, J. Xin, S. Zhang, Superbase/cellulose: J. CO2 Util. 33 (2019) 413–418.
an environmentally benign catalyst for chemical fixation of carbon dioxide into [198] F. Zhang, Y. Wang, X. Zhang, X. Zhang, H. Liu, B. Han, Recent advances in the
cyclic carbonates, Green Chem. 16 (6) (2014) 3071–3078. coupling of CO2 and epoxides into cyclic carbonates under halogen-free
[171] X.-F. Liu, Q.-W. Song, S. Zhang, L.-N. He, Hydrogen bonding-inspired condition, Green Chem. Eng. 2020.
organocatalysts for CO2 fixation with epoxides to cyclic carbonates, Catal. Today [199] T.K. Pal, D. De, P.K. Bharadwaj, Metal–organic frameworks for the chemical
263 (2016) 69–74. fixation of CO2 into cyclic carbonates, Coord. Chem. Rev. 408 (2020), 213173.
[172] L. Wang, G. Zhang, K. Kodama, T. Hirose, An efficient metal-and solvent-free [200] J. Song, Z. Zhang, S. Hu, T. Wu, T. Jiang, B. Han, MOF-5/n-Bu 4 NBr: an efficient
organocatalytic system for chemical fixation of CO2 into cyclic carbonates under catalyst system for the synthesis of cyclic carbonates from epoxides and CO2
mild conditions, Green Chem. 18 (5) (2016) 1229–1233. under mild conditions, Green Chem. 11 (7) (2009) 1031–1036.
[173] J. Wang, Y. Zhang, Boronic acids as hydrogen bond donor catalysts for efficient [201] W. Kleist, F. Jutz, M. Maciejewski, A. Baiker, Mixed-linker metal-organic
conversion of CO2 into organic carbonate in water, ACS Catal. 6 (8) (2016) frameworks as catalysts for the synthesis of propylene carbonate from propylene
4871–4876. oxide and CO2, Eur. J. Inorg. Chem. 2009 (24) (2009) 3552–3561.
[174] S. Arayachukiat, C. Kongtes, A. Barthel, S.V.C. Vummaleti, A. Poater, [202] O.V. Zalomaeva, N.V. Maksimchuk, A.M. Chibiryaev, K.A. Kovalenko, V.P. Fedin,
S. Wannakao, L. Cavallo, V. D’Elia, Ascorbic acid as a bifunctional hydrogen bond B.S. Balzhinimaev, Synthesis of cyclic carbonates from epoxides or olefins and
donor for the synthesis of cyclic carbonates from CO2 under ambient conditions, CO2 catalyzed by metal-organic frameworks and quaternary ammonium salts,
ACS Sustain. Chem. Eng. 5 (8) (2017) 6392–6397. J. Energy Chem. 22 (1) (2013) 130–135.
[175] A. Barbarini, R. Maggi, A. Mazzacani, G. Mori, G. Sartori, R. Sartorio, [203] S.-N. Kim, J. Kim, H.-Y. Kim, H.-Y. Cho, W.-S. Ahn, Adsorption/catalytic
Cycloaddition of CO2 to epoxides over both homogeneous and silica-supported properties of MIL-125 and NH2-MIL-125, Catal. Today 204 (2013) 85–93.
guanidine catalysts, Tetrahedron Lett. 44 (14) (2003) 2931–2934. [204] V. Guillerm, J. Weseliński, Y. Belmabkhout, A.J. Cairns, V. D’elia, Ł. Wojtas,
[176] X. Zhang, N. Zhao, W. Wei, Y. Sun, Chemical fixation of carbon dioxide to K. Adil, M. Eddaoudi, Discovery and introduction of a (3, 18)-connected net as an
propylene carbonate over amine-functionalized silica catalysts, Catal. Today 115 ideal blueprint for the design of metal–organic frameworks, Nat. Chem. 6 (8)
(2006) 102–106. (2014) 673–680.
[177] L. Canali, D.C. Sherrington, Utilisation of homogeneous and supported chiral [205] C.M. Miralda, E.E. Macias, M. Zhu, P. Ratnasamy, M.A. Carreon, Zeolitic
metal (salen) complexes in asymmetric catalysis, Chem. Soc. Rev. 28 (2) (1999) imidazole framework-8 catalysts in the conversion of CO2 to chloropropene
85–93. carbonate, ACS Catal. 2 (1) (2012) 180–183.
[178] D.J. Darensbourg, J.C. Yarbrough, Mechanistic aspects of the copolymerization [206] L. Yang, L. Yu, G. Diao, M. Sun, G. Cheng, S. Chen, Zeolitic imidazolate
reaction of carbon dioxide and epoxides, using a chiral salen chromium chloride framework-68 as an efficient heterogeneous catalyst for chemical fixation of
catalyst, J. Am. Chem. Soc. 124 (22) (2002) 6335–6342. carbon dioxide, J. Mol. Catal. A Chem. 392 (2014) 278–283.
[179] R. Srivastava, T.H. Bennur, D. Srinivas, Factors affecting activation and utilization [207] T. Jose, Y. Hwang, D.-W. Kim, M.-I. Kim, D.-W. Park, Functionalized zeolitic
of carbon dioxide in cyclic carbonates synthesis over Cu and Mn peraza imidazolate framework F-ZIF-90 as efficient catalyst for the cycloaddition of
macrocyclic complexes, J. Mol. Catal. A Chem. 226 (2) (2005) 199–205. carbon dioxide to allyl glycidyl ether, Catal. Today 245 (2015) 61–67.
[180] G.-P. Wu, S.-H. Wei, W.-M. Ren, X.-B. Lu, B. Li, Y.-P. Zu, D.J. Darensbourg, [208] J. Tharun, G. Mathai, A.C. Kathalikkattil, R. Roshan, Y.-S. Won, S.J. Cho, J.-
Alternating copolymerization of CO2 and styrene oxide with Co (iii)-based S. Chang, D.-W. Park, Exploring the catalytic potential of ZIF-90: solventless and
catalyst systems: differences between styrene oxide and propylene oxide, Energy co-catalyst-free synthesis of propylene carbonate from propylene oxide and CO2,
Environ. Sci. 4 (12) (2011) 5084–5092. ChemPlusChem 80 (4) (2015) 715–721.
[209] A.C. Kathalikkattil, R. Roshan, J. Tharun, H.G. Soek, H.S. Ryu, D.W. Park, Pillared
cobalt–amino acid framework catalysis for styrene carbonate synthesis from CO2

27
A. Rehman et al. Journal of Environmental Chemical Engineering 9 (2021) 105113

and epoxide by metal–sulfonate–halide synergism, ChemCatChem 6 (1) (2014) [225] A. Marandi, M. Bahadori, S. Tangestaninejad, M. Moghadam, V. Mirkhani,
284–292. I. Mohammadpoor-Baltork, R. Frohnhoven, S. Mathur, A. Sandleben, A. Klein,
[210] Y. Ren, Y. Shi, J. Chen, S. Yang, C. Qi, H. Jiang, Ni (salphen)-based metal–organic Cycloaddition of CO2 with epoxides and esterification reactions using the porous
framework for the synthesis of cyclic carbonates by cycloaddition of CO2 to redox catalyst Co-POM@ MIL-101 (Cr), New J. Chem. 43 (39) (2019)
epoxides, RSC Adv. 3 (7) (2013) 2167–2170. 15585–15595.
[211] Y. Ren, X. Cheng, S. Yang, C. Qi, H. Jiang, Q. Mao, A chiral mixed metal–organic [226] E. Akimana, J. Wang, N.V. Likhanova, S. Chaemchuen, F. Verpoort, MIL-101 (Cr)
framework based on a Ni (saldpen) metalloligand: synthesis, characterization and for CO2 conversion into cyclic carbonates, under solvent and Co-catalyst free mild
catalytic performances, Dalton Trans. 42 (27) (2013) 9930–9937. reaction conditions, Catalysts 10 (4) (2020) 453.
[212] D. Feng, W.-C. Chung, Z. Wei, Z.-Y. Gu, H.-L. Jiang, Y.-P. Chen, D.J. Darensbourg, [227] M. Taherimehr, B. Van de Voorde, L.H. Wee, J.A. Martens, D.E. De Vos, P.
H.-C. Zhou, Construction of ultrastable porphyrin Zr metal–organic frameworks P. Pescarmona, Strategies for enhancing the catalytic performance of
through linker elimination, J. Am. Chem. Soc. 135 (45) (2013) 17105–17110. metal–organic frameworks in the fixation of CO2 into cyclic carbonates,
[213] H.-Y. Cho, D.-A. Yang, J. Kim, S.-Y. Jeong, W.-S. Ahn, CO2 adsorption and ChemSusChem 10 (6) (2017) 1283–1291.
catalytic application of Co-MOF-74 synthesized by microwave heating, Catal. [228] M. Zhu, D. Srinivas, S. Bhogeswararao, P. Ratnasamy, M.A. Carreon, Catalytic
Today 185 (1) (2012) 35–40. activity of ZIF-8 in the synthesis of styrene carbonate from CO2 and styrene oxide,
[214] Y.-J. Kim, D.-W. Park, Functionalized IRMOF-3: an efficient heterogeneous Catal. Commun. 32 (2013) 36–40.
catalyst for the cycloaddition of allyl glycidyl ether and CO2, J. Nanosci. [229] Z. Zhou, C. He, J. Xiu, L. Yang, C. Duan, Metal–organic polymers containing
Nanotechnol. 13 (3) (2013) 2307–2312. discrete single-walled nanotube as a heterogeneous catalyst for the cycloaddition
[215] O.V. Zalomaeva, A.M. Chibiryaev, K.A. Kovalenko, O.A. Kholdeeva, B. of carbon dioxide to epoxides, J. Am. Chem. Soc. 137 (48) (2015) 15066–15069.
S. Balzhinimaev, V.P. Fedin, Cyclic carbonates synthesis from epoxides and CO2 [230] J. Sun, S.-I. Fujita, F. Zhao, M. Arai, A highly efficient catalyst system of ZnBr2/n-
over metal–organic framework Cr-MIL-101, J. Catal. 298 (2013) 179–185. Bu4NI for the synthesis of styrene carbonate from styrene oxide and supercritical
[216] W.Y. Gao, Y. Chen, Y. Niu, K. Williams, L. Cash, P.J. Perez, L. Wojtas, J. Cai, Y. carbon dioxide, Appl. Catal. A Gen. 287 (2) (2005) 221–226.
S. Chen, S. Ma, Crystal engineering of an nbo topology metal–organic framework [231] F. Li, L. Xiao, C. Xia, B. Hu, Chemical fixation of CO2 with highly efficient ZnCl2/
for chemical fixation of CO2 under ambient conditions, Angew. Chem. Int. Ed. 53 [BMIm] Br catalyst system, Tetrahedron Lett. 45 (45) (2004) 8307–8310.
(10) (2014) 2615–2619. [232] J. Sun, S.-i Fujita, F. Zhao, M. Arai, Synthesis of styrene carbonate from styrene
[217] H. Kim, H.-S. Moon, M. Sohail, Y.-N. Yoon, S.F.A. Shah, K. Yim, J.-H. Moon, Y. oxide and carbon dioxide in the presence of zinc bromide and ionic liquid under
C. Park, Synthesis of cyclic carbonate by CO2 fixation to epoxides using mild conditions, Green Chem. 6 (12) (2004) 613–616.
interpenetrated MOF-5/n-Bu 4 NBr, J. Mater. Sci. 54 (18) (2019) 11796–11803. [233] S.-S. Wu, X.-W. Zhang, W.-L. Dai, S.-F. Yin, W.-S. Li, Y.-Q. Ren, C.-T. Au,
[218] P. Patel, B. Parmar, R.I. Kureshy, H.K. Noor-ul, E. Suresh, Amine-functionalized ZnBr2–Ph4PI as highly efficient catalyst for cyclic carbonates synthesis from
Zn (ii) MOF as an efficient multifunctional catalyst for CO2 utilization and terminal epoxides and carbon dioxide, Appl. Catal. A Gen. 341 (2008) 106–111.
sulfoxidation reaction, Dalton Trans. 47 (24) (2018) 8041–8051. [234] M. Liu, L. Liang, T. Liang, X. Lin, L. Shi, F. Wang, J. Sun, Cycloaddition of CO2 and
[219] M. Gupta, D. De, K. Tomar, P.K. Bharadwaj, From Zn (II)-carboxylate to double- epoxides catalyzed by dicationic ionic liquids mediated metal halide: influence of
walled Zn (II)-carboxylato phosphate MOF: change in the framework topology, the dication on catalytic activity, J. Mol. Catal. A Chem. 408 (2015) 242–249.
capture and conversion of CO2, and catalysis of Strecker reaction, Inorg. Chem. 56 [235] U.R. Seo, Y.K. Chung, Poly (4–vinylimidazolium) s/diazabicyclo [5.4. 0] undec-
(23) (2017) 14605–14611. 7–ene/Zinc (II) bromide-catalyzed cycloaddition of carbon dioxide to epoxides,
[220] A. Verma, D. De, K. Tomar, P.K. Bharadwaj, An amine functionalized Adv. Synth. Catal. 356 (9) (2014) 1955–1961.
metal–organic framework as an effective catalyst for conversion of CO2 and [236] M. Liu, B. Liu, S. Zhong, L. Shi, L. Liang, J. Sun, Kinetics and mechanistic insight
Biginelli reactions, Inorg. Chem. 56 (16) (2017) 9765–9771. into efficient fixation of CO2 to epoxides over N-heterocyclic compound/ZnBr2
[221] H. He, Q. Sun, W. Gao, J.A. Perman, F. Sun, G. Zhu, B. Aguila, K. Forrest, B. Space, catalysts, Ind. Eng. Chem. Res. 54 (2) (2015) 633–640.
S. Ma, A stable metal–organic framework featuring a local buffer environment for [237] H.S. Kim, J.J. Kim, H. Kim, H.G. Jang, Imidazolium zinc tetrahalide-catalyzed
carbon dioxide fixation, Angew. Chem. 130 (17) (2018) 4747–4752. coupling reaction of CO2 and ethylene oxide or propylene oxide, J. Catal. 220 (1)
[222] J. Kim, S.-N. Kim, H.-G. Jang, G. Seo, W.-S. Ahn, CO2 cycloaddition of styrene (2003) 44–46.
oxide over MOF catalysts, Appl. Catal. A Gen. 453 (2013) 175–180. [238] Y. Shang, Q. Gong, M. Zheng, H. Zhang, X. Zhou, An efficient morpholinium ionic
[223] T. Lescouet, C. Chizallet, D. Farrusseng, The origin of the activity of amine- liquid based catalyst system for cycloaddition of CO2 and epoxides under mild
functionalized metal–organic frameworks in the catalytic synthesis of cyclic conditions, J. Mol. Liq. 283 (2019) 235–241.
carbonates from epoxide and CO2, ChemCatChem 4 (11) (2012) 1725–1728. [239] C. Liu, Y. Ye, Z. Jiang, P. Xu, J. Zhang, J. Sun, Carbon dioxide activation and
[224] M. Bahadori, S. Tangestaninejad, M. Bertmer, M. Moghadam, V. Mirkhani, conversion by hyperbranched polyethylenimine/ZnI2 catalysts, Ind. Eng. Chem.
I. Mohammadpoor-Baltork, R. Kardanpour, F. Zadehahmadi, Task-specific ionic Res. 58 (2) (2018) 872–878.
liquid functionalized–MIL–101 (Cr) as a heterogeneous and efficient catalyst for [240] C. Yang, Y. Chen, P. Xu, L. Yang, J. Zhang, J. Sun, Facile synthesis of zinc halide-
the cycloaddition of CO2 with epoxides under solvent free conditions, ACS based ionic liquid for efficient conversion of carbon dioxide to cyclic carbonates,
Sustain. Chem. Eng. 7 (4) (2019) 3962–3973. Mol. Catal. 480 (2020), 110637.

28

You might also like