You are on page 1of 8

Journal of Molecular Structure 1296 (2024) 136868

Contents lists available at ScienceDirect

Journal of Molecular Structure


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

Carboxylative cyclization of propargyl alcohols with carbon dioxide for the


synthesis of α-alkylidene cyclic carbonates in presence of Co(III) schiff base
complex catalyst
Juli Nanda Goswami a, Najirul Haque b, Asiful H. Seikh c, Biswajit Bhattacharya d,
Franziska Emmerling d, Nimai Bar e, Ahmad A. Ifseisi f, Surajit Biswas b, *, Malay Dolai a, *
a
Department of Chemistry, Prabhat Kumar College, Purba Medinipur 721404, West Bengal, India
b
Department of Chemistry, University of Kalyani, Kalyani, Nadia 741235, West Bengal, India
c
Mechanical Engineering Department, College of Engineering, King Saud University, Riyadh 11421, Saudi Arabia
d
BAM Federal Institute for Materials Research and Testing, Richard-Willstätter-Str, Berlin 11 12489, Germany
e
Raja Narendralal Khan Women’s College (Autonomous), Paschim Medinipur, West Bengal 721102, India
f
Department of Chemistry, College of Science, King Saud University, Riyadh 11451, Saudi Arabia

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

Keywords: A cobalt(III) complex, [Co(L)3](DMF) (1) of Schiff base ligand HL, 2-((E)-(benzylimino)methyl)-4-bromophenol
Cobalt (III) Schiff base complex is prepared and single crystal X-ray structural analysis have also been performed. The structures of complex 1
X-ray crystal analysis showed hexa-coordinated mononuclear systems that adopt octahedral geometry. The complex has also exhibited
Catalytic fixation of carbon dioxide
the supramolecular networks through non-covalent interactions like H-bonding, C–Hπ stacking. Moreover, the
Carboxylative cyclization of propargyl alcohols
А-alkylidene cyclic carbonates
complex 1 is very effective in the catalytic fixation of carbon dioxide in propergyl alcohols to produce α-alky­
DFT calculations lidene cyclic carbonates. The catalytic production of α-alkylidene cyclic carbonates have been carried out
through carboxylative cyclization of propargyl alcohols using CO2 balloon of 1 atm pressure at 80 ◦ C. Solvent
free condition (green synthesis) made this catalytic protocol eco-friendly towards the environment. Utilizing
various substrates of propargyl alcohols moderate to high percentage yield (62–95%) of respective α-alkylidene
cyclic carbonates product have been isolated over this catalytic reaction. Besides, the theoretical calculations
(DFT) was performed for the prediction of probable mechanism of the catalytic reaction.

1. Introduction pivotal role towards a clean environment [3–7]. Again, CO2 is


economical, ecological and most abundant C1-resourses as a result
Realistic resource management is the most important feasible exploitation of it in chemical fixation is gaining ample attention
development for the future generation. The present stage of sustain­ currently [8–11]. Chemical fixation of carbon dioxide in order to pro­
ability is silently required enormous advancement as it is outlying far duce valuable fine chemicals under mild reaction conditions is chal­
from the satisfactory point. Till date coal, crude oil and natural gas are lenging due to kinetic and thermodynamic stability of carbon dioxide. In
the prime resources of carbon. Thus carbon dioxide along with biomass this regards design, synthesis and utilization of suitable catalyst is very
can be an additional appropriate alternative for the renewable carbon much essential for the fixation of thermodynamically and kinetically
assets. Due to the fulfillment of global energy requirement a substantial stable CO2 [12–16].
quantity of carbon dioxide has been released from combustion of fossil Numerous catalytic protocols has been developed for the fixation of
fuels after the industrial revolution. CO2 is the principal constituent of carbon dioxide to produce cyclic carbonates [17–19], α-alkylidene cy­
greenhouse gasses, producing global warming which causes drastic clic carbonates [20–22] are organic carbamates [23–26], oxazolidinones
change of the climate [1,2]. As a result it is very urgent to reduce CO2 [27,28], benzimidazoles [29–31], quinazolidinone [32–34], N-methyl­
emission into the atmosphere by the means of storage or usages. Recy­ ated [35] and N-formylated products [36] etc. Among the before
cling and reduction of CO2 into valuable fine chemicals can play a mentioned CO2 fixation products, α-alkylidene cyclic carbonates is one

* Corresponding authors.
E-mail addresses: surajit.jrf@gmail.com (S. Biswas), dolaimalay@yahoo.in (M. Dolai).

https://doi.org/10.1016/j.molstruc.2023.136868
Received 26 July 2023; Received in revised form 23 September 2023; Accepted 16 October 2023
Available online 20 October 2023
0022-2860/© 2023 Elsevier B.V. All rights reserved.
J.N. Goswami et al. Journal of Molecular Structure 1296 (2024) 136868

Scheme 1. The schematic presentation of the formation of ligand HL.

Scheme 2. Schematic diagram for synthesis of 1.

of the most important class of organic compounds. This compound are δ 13.56 (s, 1H), 8.71 (s, 1H), 7.72 (d, J = 2.5 Hz, 1H), 7.48 (dd, J = 8.8,
commonly found in pharmaceuticals, agrochemicals and natural prod­ 2.5 Hz, 1H), 7.43–7.27 (m, 5H), 6.87 (d, J = 8.8 Hz, 1H), 4.82 (s, 2H). IR
ucts such as Cycloolivil carbonate, Genkwanin I, Guaianolide (KBr, cm− 1): 1621 (s), 1498 (s), 1267 (s), 1131 (s) (Fig. S2).
holo-leucin, etc. [37–39]. In addition, they are very useful reagents for a
variety of organic transformations [40–43]. Synthesis of α-alkylidene 2.2. Synthesis of complex 1
cyclic carbonates by carboxylative cyclization of propargyl alcohols via
catalytic carbon dioxide fixation reaction is eco-friendly and promising Complex 1 is prepared by mixing CoCl2⋅6H2O (0.047 g, 0.2 mmol)
pathway developed recently. Several catalytic systems such as phos­ and HL (0.0656 g, 0.2 mmol) in a mixture solution of acetonitrile/DMF
phines [44], azole-anion-based aprotic ionic liquids [45], bicyclic gua­ (v/v = 15/5 mL). Then the triethylamine (0.0202 g, 0.2 mmol) was
nidines [46], N-heterocyclic carbene/carbon dioxide adducts [47], added and the mixture was continued to stir for 45 minutes more to give
N-heterocyclic olefin/carbon dioxide adducts [48], dark brown solution (Scheme 2). The solution has been left untouched
alkoxide-functionalized imidazolium betaines/carbon dioxide adducts for slow evaporation after filtration. The dark brown single crystals,
[49] etc. were used as metal free catalyst for this purpose. Besides, some suitable for single crystal X-ray diffraction obtained in few days. Then
metal-based catalysts like copper [50–53], cobalt [54], palladium [55], crystals were washed through cold MeCN and sent to SCXRD instrument.
zinc [56], tungsten [57], ruthenium [58], and silver [59–63] also been Anal.Calc. for C45H40Br3CoN4O4 Yield:0.519 g. (52%) (MW 999.47): C,
utilized for the same compound synthesis. However, most of the re­ 54.08; H, 4.03; N, 5.61. Found: C, 54.05; H, 4.02; N, 5.59%. 1HNMR
ported methods required either high pressure of carbon dioxide or (400 MHz, DMSO) ( Fig.S3). IR (KBr, cm− 1): 3475(b), 1619 (s), 1493 (s),
drastic reaction conditions. Recently, S. M. Islam and his group pub­ 1265 (s), 1130 (s) (Fig. S4).
lished a paper where silver nanoparticles embedded COF catalyst pro­
duced α-alkylidene cyclic carbonates through carboxylative cyclization 2.3. Catalytic procedure for the synthesis of α-alkylidene cyclic carbonate
of propargyl alcohols under mild reaction conditions [64]. from propargyl alcohols
Despite some advancement in this catalytic field, only one report has
been published earlier where cobalt metal-based catalyst is used for the Propargyl alcohol (8 mmol), DBU (Diazabicyclo(5.4.0)undec‑7-ene)
production of α-alkylidene cyclic carbonates [54]. Cobalt is relatively (8 mmol) and catalyst (0.005 mmol) were taken in a 25 ml rb flask. Then
low cost first row transition metal. In this article, a Schiff base ligand HL, CO2 (99.99%) balloon was connected with the flask and it was vac­
2-((E)-(benzylimino)methyl)− 4-bromophenol was used for the synthe­ uumed and purged with carbon dioxide three times to remove air from
sis of mononuclear robust cobalt(III) complex [Co(L)3](DMF) (1) which the reaction vessel. The mixture was stirred at 80 ◦ C under oil bath. The
was effective as catalyst for fixation of carbon dioxide to synthesize progress of reaction was monitored by TLC (thin layer chromatography)
α-alkylidene cyclic carbonates through carboxylative cyclization of technique. After completion of reaction, the mixture was filtered and
propargyl alcohols under mild condition. The complex 1 showed concentrated through rota-evaporator. The obtained crude product was
excellent catalytic activity with high selectivity. Moreover, theoretical purified through column chromatography on silica gel using ethyl ace­
calculations have performed to establish the probable mechanistic tate and petroleum ether mixture as eluent. The pure product was
pathway of the catalytic reaction. characterized by NMR (1H and 13C) spectroscopy.

2. Experimental section
3. Result and discussion

2.1. Synthesis of the ligand (HL)


3.1. Crystal structure description for [Co(L)3].DMF(1)

The ligand HL, was prepared by the Schiff base condensation reac­
The single crystal X-ray diffraction studies reveal that the complex 1
tion following the procedure of previous report (Scheme 1) [65]. Anal.
is mononuclear cobalt(III) complexes where the central atom is bounded
Calc. for C14H12BrNO (MW 290.15): Calc. C, 57.95; H, 4.17; N, 4.83;
by three bidentate NO donor Schiff base ligand (HL).
Found: C, 57.91; H, 4.15; N, 4.88%.1HNMR (400 MHz, DMSO) (Fig.-S1)
The complex 1 ([CoIII (L)3].DMF) crystallized in the monoclinic

2
J.N. Goswami et al. Journal of Molecular Structure 1296 (2024) 136868

combination of two C–H…π bonds C40-H40…Cg1(ring C29, C30, C31,


C32, C33, C34) and C11-H11…Cg1(ring C29, C30, C31, C32, C33, C34)
distance of 2.83 Å leads to the formation of supramolecular 1D chain
running along the crystallographic c axis. The third C–H…π bond C27-
H27…Cg2(ring C15, C16, C17, C18, C19, C20) distance of 2.79 Å sup­
ports the formation of 2D supramolecular layer parallel with the crys­
tallographic ac plane.

3.2. UV–Vis spectra for complex [Co(L)3].DMF(1)

Electronic spectra of complex 1 is shown in Fig. 3, were recorded in


MeCN (4 × 10− 5 (M)) solution. The first two peaks are at 226 nm and

Fig. 1. The ORTEP view of complex 1 (C atoms not lebelled for clarity). DMF
molecule is also omitted for clarity.

crystal system with space group P21/n. It is observed that the geometry
of 1 is octahedral where Co1 is hexa-coordinated by three deprotonated
HL ligands (namely N1, O1, N2, O2 and N3, O3). The DMF as solvent
molecule is situated in the unit cell of 1 (Fig. 1). The crystal data and
details of the structure determination for 1 are given in Table S1. The Co-
Oi bond lengths lie in the span of 1.888–1.893 Å as well as the Co-Ni
bond lengths lie in the range of 1.929–1.948 Å for 1. The certain bond
distances and angles for complex 1 are tabulated in Table S2 and S3.
The molecules of complex 1 connected by C40-H40…Cg1(ring C29,
C30, C31, C32, C33, C34) interaction with H40…Cg1(ring C29, C30,
C31, C32, C33, C34) distance of 2.856 Å generate supramolecular 1D
chain extending nearly parallel with [− 101] direction (Fig. 2). However, Fig. 3. Electronic spectra of [CoIII(L)].DMF in acetonitrile.

Fig. 2. The molecular view of (a) 1D network through CH… π interactions (green dotted line) extending nearly parallel with [− 101] direction and (b) 2D network
through CH… π interactions (skyish blue dotted line) in the crystallographic ac plane in complex 1.

3
J.N. Goswami et al. Journal of Molecular Structure 1296 (2024) 136868

After getting lesser yield from those reactions, we decided to use neat
reaction condition for further optimization of catalysis reaction. Utili­
zation of various base such as NEt3 (triethylamine), DABCO (1,4-dia­
zabicyclo[2.2. 2]octane), DIPEA (N, N-diisopropylethylamine), TBD
(1,5,7- triazabicyclo[4.4. 0]dec‑5-ene), TMEDA (tetramethylethylene­
diamine), t-BuOK (potassium tert‑butoxide), Cs2CO3 (caesium carbon­
ate), K2CO3 (potassium carbonate), PPh3 (triphenylphosphine) in place
Scheme 3. Catalytic carboxylative cyclization of propargyl alcohols in pres­ of DBU in the reaction, the observed results suggested us to use DBU for
ence of carbon dioxide for the synthesis of α-alkylidene cyclic carbonates.
the reaction (Table 1, entries 9–17). Then the reaction temperature was
enhanced stepwise from 60 to 90 ◦ C, increased of product yield was
253 nm is due to intra-ligand charge transfer transitions (ILCT). The noticed from 60 to 80 ◦ C (Table 1, entries 18, 19). Due to increase of
band at 401 nm in the electronic spectrum is due to ligand to metal temperature produced more number of collisions between the substrate
charge transfer transition (LMCT) which is indication of the metal and the active sites of the catalyst as a result yield percentage of product
complex formation. increased. Further increase of temperature from 80 to 90 ◦ C caused
slight reduction of desired product yield may be because of formation of
3.3. Catalytic synthesis of α-alkylidene cyclic carbonates through side product at that elevated temperature (Table 1, entry 20). Increment
carboxylative cyclization of propargyl alcohols of catalyst loading for the reaction did not alter the isolated yield per­
centage of the product (Table 1, entry 21). No conversion of the reaction
The catalytic activity of the synthesized robust cobalt (III) complex was found when the reaction was performed without base and catalyst
(1) have been checked in the production of α-alkylidene cyclic carbon­ (Table 1, entry 22). Various metal precursors such as cobalt chloride,
ates through carboxylative cyclization of propargyl alcohols with carbon cobalt nitrate, cobalt acetate were incorporated as catalyst in position of
dioxide (Scheme 3). First of all a variety of control experiments were complex 1 for the reaction, the obtained results did not compete with the
carried out to identify the optimized catalytic reaction conditions using result appeared by using our synthesized complex 1 catalyst (Table 1,
2-methyl-3-butyn-2-ol as model propargyl alcohol under 1 atm CO2 entries 23–25). In presence of ligand (HL) the catalytic reaction did not
pressure. Initial experiment was done by taking 2-methyl-3-butyn-2-ol proceeds at all as no product formation was observed after 6 h of reac­
(8 mmol), 1,8-diazabicyclo(5.4.0)undec‑7-ene (DBU) (8 mmol), in ab­ tion duration (Table 1, entry 26). Therefore ligand itself has no effect on
sent of any catalyst at 60 ◦ C under neat and homogeneous condition. the catalytic reaction. Again the same catalytic reaction was carried out
After 6 h of reaction 18% isolated yield of corresponding α-alkylidene in presence of ligand and metal precursor (CoCl2⋅6H2O) in 3:1 mole ratio
cyclic carbonate was observed (Table 1, entry 1). but the result obtained was almost similar to that of the result when only
Then the same reaction was monitored in presence of our synthesized metal precursor used as catalyst (Table 1, entry 27). This due to absence
cobalt complex 1 as catalyst (0.005 mmol), it was surprised that we got of suitable solvent (i.e. MeCN/DMF) in catalytic reaction medium, no
56% isolated yield of 4,4-dimethyl-5-methylene-1,3-dioxolan-2-one formation of complex 1 was took place as a result catalytic activity did
after 6 h of reaction (Table 1, entry 2). Use of different solvents in the not enhance.
reaction was checked for the expectation of better yield percentage of A kinetic graph of the reaction at 80 ◦ C in presence complex 1 as
desired product. Instead of increasing the yield percentage of the catalyst clearly depicted that 6 h reaction duration was enough for
product enormous decreased in the product yield was obtained (Table 1, maximum yield of desired product (Fig. 4).
entries 3–8).

Table 1
Control experiments result for the catalytic reactiona.
Entry Catalyst (mmol) Base Solvent Temperature ( ◦ C) Time (h) Isolated Yield (%)

1 – DBU – 60 6 18
2 1 (0.005) DBU – 60 6 56
3 1 (0.005) DBU MeCN 60 6 28
4 1 (0.005) DBU DCM 60 6 21
5 1 (0.005) DBU EtOH 60 6 18
6 1 (0.005) DBU THF 60 6 32
7 1 (0.005) DBU DMF 60 6 30
8 1 (0.005) DBU 1,4-doixane 60 6 35
9 1 (0.005) NEt3 – 60 6 40
10 1 (0.005) DABCO – 60 6 30
11 1 (0.005) DIPEA – 60 6 32
12 1 (0.005) TBD – 60 6 28
13 1 (0.005) TMEDA – 60 6 37
14 1 (0.005) t-BuOK – 60 6 18
15 1 (0.005) Cs2CO3 – 60 6 19
16 1 (0.005) K2CO3 – 60 6 23
17 1 (0.005) PPh3 – 60 6 38
18 1 (0.005) DBU – 70 6 72
19 1 (0.005) DBU – 80 6 95
20 1 (0.005) DBU – 90 6 88
21 1 (0.007) DBU – 80 6 95
22 – – – 80 6 00
23 CoCl2⋅6H2O (0.005) DBU – 80 6 37
24 Co(NO₃)₂⋅6H₂O (0.005) DBU – 80 6 29
25 Co(OAc)2⋅4H2O (0.005) DBU – 80 6 34
26 ligand (HL) (0.005) DBU – 80 6 00
27 HL+ CoCl2⋅6H2Ob DBU – 80 6 40
a
Reaction conditions: 2-methyl-3-butyn-2-ol (8 mmol), CO2 balloon, base (8 mmol), homogeneous condition
b
3:1 mole ratio

4
J.N. Goswami et al. Journal of Molecular Structure 1296 (2024) 136868

Table 2
Substrate scope for the catalytic reactiona.
Entry Substrate Product Time Isolated
(h) Yield (%)

1 6 0

2 6 0

Fig. 4. Kinetic graph for the catalytic reaction [reaction conditions: 2-methyl- 3 6 95
3-butyn-2-ol (8 mmol), DBU (8 mmol), 1 (0.005 mmol), CO2 balloon, 80 ◦ C].

After reaching optimum reaction condition for the catalysis [i.e.


propargyl alcohols (8 mmol), DBU (8 mmol), catalyst [Complex 1]
(0.005 mmol), CO2 balloon, 80 ◦ C] we explored the substrate scope for
the reaction by changing R1 and R2 substituent of alcohol (Table 2). No
product was obtained from non substituted and mono substituted
4 7 91
propargyl alcohol (such as prop‑2-yn-1-ol and but‑3-yn-2-ol) (Table 2,
entries1–2). When both substituent were methyl (R1= R2= -Me) (i.e. 2-
methylbut-3-yn-2-ol), maximum product yield was obtained (95%)
through this catalytic reaction after 6 h (Table 2, entry 3). This phe­
nomenon occurred due to “gem effect” (or Thorpe− Ingold effect) where
two gem dimethyl group favoured cyclization due to their mutual
repulsion [66]. Replacement of one methyl substituent by ethyl
group/isopropyl group of propargyl alcohol influenced the reaction
5 8 87
duration as well as the product yield (Table 2, entries 4, 5). When both
R1= R2= -Et (ethyl) further decreased of desired product yield was
noticed (Table 2, entry 6). Thus increase of the size of substituents (i.e.
bulkiness of the group R1 and R2) greatly influenced the isolated product
yield percentage and reaction time [67]. 1-ethynylcyclohexan-1-ol
produced 78% isolated yield of respective product after 10 h of reac­
tion (Table 2, entry 7). When phenyl group was present in R2 position of
propargyl alcohol obtained moderate yield of corresponding carbonates
product (Table 2, entry 8). But the reactivity is not shown by the internal
propargyl alcohol (2-methyl-4-phenylbut-3-yn-2-ol) (Table 2, entry 9)
6 8 89
[64,67].
The literature survey for the comparison of catalytic production of
4,4-dimethyl-5-methylene-1,3-dioxolan-2-one from 2-methylbut-3-yn-
2-ol with previous reports is given in Table 3. Under both homoge­
neous and heterogeneous catalytic conditions those reported catalyst
produced high yield of 4,4-dimethyl-5-methylene-1,3-dioxolan-2-one
under high CO2 pressure [67–71]. Few reports were published where
1 bar CO2 pressure was used for the desired product synthesis [64,72,
73]. Only one cobalt metal containing catalyst was reported where 87%
7 10 78
GLC (Gas Liquid Chromatography) yield is produced for the respective
product under high carbon dioxide pressure in homogenous catalytic
medium [54]. Our synthesized Co(III) metal containing catalyst gave us
high isolated yield (95%) of 4,4-dimethyl-5-methylene-1,3-dioxola­
n-2-one in homogeneous medium under 1 bar CO2 pressure.

3.4. A probable reaction mechanism for carboxylative cyclization of


propargyl alcohols

The Co(III) center in complex 1 which was capable to activate (continued on next page)
propargyl alcohol via the C-atoms of triple bond that was the main
driving force for the effective catalytic cyclic carbonate synthesis. To
perceive the mechanism of the cyclic carbonate production reaction

5
J.N. Goswami et al. Journal of Molecular Structure 1296 (2024) 136868

Table 2 (continued ) Table 3


Entry Substrate Product Time Isolated
Comparison table for the catalytic production of 4,4-dimethyl-5-methylene-1,3-
(h) Yield (%) dioxolan-2-one from 2-methylbut-3-yn-2-ol with previous reports.

8 12 62 Catalyst Reaction conditions Catalytic Yield Reference


nature (%)a

ZnI2 substrate (1.5 mmol), Homogeneous 95 67


base (NEt3 1.5 mmol),
catalyst (0.3 mmol), 1
MPa CO2, 30 ◦ C, 14h
AgOAc substrate (5 mmol), Homogeneous 87 68
catalyst (1 mol%),
DavePhos (2 mol%),
CO2 (20 bar), CH3CN
9 18 0
(10 mL), RT, 16h
(nBu4N)2Ox substrate (0.03 mmol), Homogeneous 100 69
catalyst (2.5 mol%), (99)b
CO2 (80 bar), 80 ◦ C,
24h
Ag@RB-POP substrate (0.5 mmol), Heterogeneous 87 70
a
catalyst (20 mg, 0.01
Reaction condition: substrate (8 mmol), DBU (8 mmol), catalyst 1 (0.005 mol% based on Ag);
mmol), CO2 (1 atm/balloon), 80 ◦ C. base (DBU 0.5 mmol),
CO2 (1 MPa),
acetobitrile (2 mL),
between propargyl alcohol and CO2 catalyzed by complex 1 and the
30 ◦ C, 12 h
hypothetical results were performed using DFT/LanL2Dz for total cat­ Zn@BBAPA-1 substrate (1.5 mmol), Heterogeneous 98 71
alytic interpretation (the calculation methods are in ESI). From the base (DBU 1.5 mmol),
previous studies [64,67–73] and the experimental evidences, it was catalyst (50 mg), CO2
tallied that propargyl alcohol molecule was activated by the coordina­ (1 MPa), 30 ◦ C, 6h
AgN@COF substrate (3 mmol), Heterogeneous 95 64
tion of the cobalt(III) atom of complex 1 by the carbon atom of triple base (DBU 1 mmol),
bonded propargyl alcohol which was responsible for the start-up of the catalyst (25 mg), CO2
mentioned catalysis. The dissociation of one of the Co–O bond of the (1 atm), RT, 3h
phenolate of hexacoordinated octahedral Co(III) complexes is expected Zn@RIO-1 substrate (5 mmol), Heterogeneous 96 72
base (DBU 5 mmol),
to occur which provides a vacant site for the substrate to coordinate
CO2 (1 atm), catalyst
[74]. All the intermediates and transition states with their relative en­ (15 mg), CO2 (1 atm),
ergies for this catalysis are shown in Fig. 5. The first step of catalysis, the room temperature,
active site of catalyst as complex 1 was initially considered as zero for 10h
the full calculations. Basically, the complex 1 was coordinated via Co AgNPs@TzTa- substrate (5 mmol), Heterogeneous 93 73
POP base (DBU 5 mmol),
(III) ion with C-atoms of triple bond (dCo…Cpropargyl =3.28 Å) from catalyst (15 mg), CO2
propargyl alcohol and was forming the TS-I (− 29.65 kcal mol− 1). In (1 atm), RT, 4h
addition, the reaction was catalytically moved forward, in where the CoCp2 substrate (50 mmol), Homogeneous 87c 54
CO2 reacted with propargyl alcohol in the attendance of catalyst to form base (NEt3 3.6 mmol),
catalyst (1 mmol), CO2
simultaneously, Int-1 and TS –2 with respective energy of − 19.33 kcal
(50 kg/cm3), 80 ◦ C, 5h
mol− 1 and − 36.72 kcal mol− 1. Then, the Int-2 (20.29 kcal mol− 1) was 1 substrate (8 mmol), Homogeneous 95 This work
immediately released from Co(III) center to give regeneration of cata­ base (DBU 8 mmol),
lytic initiation by complex 1 and undergoing an intramolecular proton catalyst (0.005 mmol),
transfer in presence of DBU to form the desired product (− 10.89 kcal CO2 (1 atm/balloon),
80 ◦ C, 6h
mol− 1).
a
. Isolated yield;.
b
The plausible mechanism was constructed (Fig. 6) from the experi­ %Conversion (% selectivity);.
c
mental outcomes and theoretical calculations for the catalytic genera­ GLC yield.
tion of cyclic carbonate by the reaction of CO2 and propargyl alcohol.
The well agreed mechanism was compared with the former reports [64, based catalyst. To best of our knowledge, this is the second report where
67–73]. cobalt metal-based catalyst is used for the synthesis of α-alkylidene cy­
clic carbonates via carboxylative cyclization of propargyl alcohols in
4. Conclusion presence of CO2. Therefore, in near future exploration of this green
synthesis of α-alkylidene cyclic carbonates in presence of Co-Schiff base
The Cobalt (III) complex, [Co(L)3].(DMF) (1) of a Schiff base ligand catalyst encourages the researchers more.
was synthesized and structurally characterized through single crystal
XRD which have octahedral geometry. The complex 1 was very much Declaration of Competing Interest
active in the catalytic fixation of carbon dioxide in propargyl alcohols to
fabricate α-alkylidene cyclic carbonates. The catalytic carboxylative The authors declare that they have no known competing financial
cyclization of propargyl alcohols under 1 bar carbon dioxide pressure interests or personal relationships that could have appeared to influence
produce moderate to high yield (62–95%) of α-alkylidene cyclic car­ the work reported in this paper.
bonates from various substrate of propargyl alcohols at 80 ◦ C temper­
ature. No solvent was used during the catalytic syntheses. Again, DFT Data availability
calculations were performed in order to explain the probable mecha­
nism of the catalysis reaction theoretically. There is no very vast Data will be made available on request.
exploration of this type of catalytic reaction in presence of cobalt metal-

6
J.N. Goswami et al. Journal of Molecular Structure 1296 (2024) 136868

Fig. 5. Possible intermediates and transition states with the relative energies catalyzed carboxylative cyclization of propargyl alcohols with carbon dioxide by
complex 1.

Fig. 6. Plausible mechanism for the carboxylative cyclization of propargyl alcohols catalyzed by complex 1 (mentioned bond lengths are in angstrom).

Acknowledgements Data centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: (+44)
1223–336–033; or e-mail: deposit@ccdc.cam.ac.uk.
M.D gratefully acknowledges and thanks the I-STEM/catalytic grant/ Supplementary material associated with this article can be found, in
acad-08/2021–2022 for the research support. AHS is thankful to Dep­ the online version, at doi:10.1016/j.molstruc.2023.136868.
utyship for Research and Innovation, Ministry of Education in Saudi
Arabia for funding the project (IFKSUOR3-318-3). References

Supplementary materials [1] A. Otto, T. Grube, S. Schiebahna, D. Stolten, Closing the loop: captured CO2 as a
feedstock in the chemical industry, Energy Environ. Sci. 8 (2015) 3283–3297.
[2] M. Aresta, Carbon Dioxide As Chemical Feedstock, John Wiley & Sons, Weinheim,
CCDC 2,237,135 contains the supplementary crystallographic data 2010. ISBN: 9783527324750.
for 1. These data can be obtained free of charge via www.ccdc.cam.ac. [3] S. Dabral, T. Schau, The use of carbon dioxide (CO2) as a building block in organic
synthesis from an industrial perspective, Adv. Synth. Catal. 361 (2019) 223–246.
uk/conts/retrieving.html, or from the Cambridge Crystallographic

7
J.N. Goswami et al. Journal of Molecular Structure 1296 (2024) 136868

[4] B. Grignard, S. Gennen, C. Je´roˆme, A.W. Kleij, C. Detrembleur, Advances in the use [38] L. Maisonneuve, O. Lamarzelle, E. Rix, E. Grau, H. Cramail, Chem. Rev. 115 (2015)
of CO2 as a renewable feedstock for the synthesis of polymers, Chem. Soc. Rev. 48 12407–12439.
(2019) 4466–4514. [39] Q.W. Song, L.N. He, Adv. Synth. Catal. 358 (2016) 1251–1258.
[5] G. Fiorani, W. Guo, A.W. Kleij, Sustainable conversion of carbon dioxide: the [40] K. Ohe, H. Matsuda, T. Ishihara, S. Ogoshi, N. Chatani, S. Murai, J. Org. Chem. 58
advent of organocatalysis, Green Chem. 17 (2015) 1375–1389. (1993) 1173–1177.
[6] Q.W. Song, Z.H. Zhou, L.N. He, Efficient, selective and sustainable catalysis of [41] K. Ohe, H. Matsuda, T. Morimoto, S. Ogoshi, N. Chatani, S. Murai, J. Am. Chem.
carbon dioxide, Green Chem. 19 (2017) 3707–3728. Soc. 116 (1994) 4125–4126.
[7] L. Guo, K.J. Lamb, M. North, Recent developments in organocatalysed [42] V. Besse, F. Camara, C. Voirin, R. Auvergne, S. Caillol, B. Boutevin, Polym. Chem. 4
transformations of epoxides and carbon dioxide into cyclic carbonates, Green (2013) 4545–4561.
Chem. 23 (2021) 77–118. [43] Y. Kayaki, M. Yamamoto, T. Ikariya, J. Org. Chem. 72 (2007) 647–649.
[8] B. Hu, C. Guild, S.L. Suib, J. CO2 Util. 1 (2013) 18–27. [44] H. Zhang, H.B. Liu, J.M. Yue, Chem. Rev. 114 (2014) 883–898.
[9] J. Jia, C. Qian, Y. Dong, Y.F. Li, H. Wang, M. Ghoussoub, K.T. Butler, A. Walsh, G. [45] Y. Zhao, Y. Wu, G. Yuan, L. Hao, X. Gao, Z. Yang, B. Yu, H. Zhang, Z. Liu, Chem.
A. Ozin, Chem. Soc. Rev. 46 (2017) 4631–4644. Asian. J. 11 (2016) 2735. -2340.
[10] Q.W. Song, Z.H. Zhoua, L.N. He, Green Chem. 19 (2017) 3707–3728. [46] N.D. Ca, B. Gabriele, G. Ruffolo, L. Veltri, T. Zanetta, M. Costa, Adv. Synth. Catal.
[11] P. Bhanja, A. Modak, A. Bhaumik, ChemCatChem 11 (2019) 244–257. 353 (2011) 133–146.
[12] P. Bhanja, A. Modak, A. Bhaumik, Chem. Eur. J. 24 (2018) 7278–7297. [47] Y. Kayaki, M. Yamamoto, T. Ikariya, Angew. Chem. 121 (2009) 4258–4261.
[13] N. Haque, S. Biswas, P. Basu, I.H. Biswas, R. Khatun, A. Khan, S.M. Islam, New J. [48] Y.B. Wang, Y.M. Wang, W.Z. Zhang, X.B. Lu, J. Am. Chem. Soc. 135 (2013)
Chem. 44 (2020) 15446–15458. 11996–12003.
[14] M. Dolai, U. Saha, S. Biswas, S. Chatterjee, G.S. Kumar, CrystEngComm. 22 (2020) [49] Y.B. Wang, D.S. Sun, H. Zhou, W.Z. Zhang, X.B. Lu, Green Chem 16 (2014)
8374–8386. 2266–2272.
[15] P. Sarkar, I.H. Chowdhury, S. Das, S.M. Islam, Mater. Adv. (2022), https://doi.org/ [50] H.S. Kim, J.W. Kim, S.C. Kwon, S.C. Shim, T.J. Kim, J. Organomet. Chem. 545
10.1039/D2MA00600F. (1997) 337–344.
[16] R. Sani, T.K. Dey, M. Sarkar, P. Basu, S.M. Islam, Mater. Adv. 3 (2022) 5575–5597. [51] Y. Gu, F. Shi, Y. Deng, J. Org. Chem. 69 (2004) 391–394.
[17] S. Biswas, D. Roy, S. Ghosh, S.M. Islam, J. Organomet. Chem. 898 (2019), 120877. [52] H.F. Jiang, A. Wang, H.L. Liu, C.R. Qi, Eur. J. Org. Chem. 2008 (2008) 2309–2312.
[18] A.K. Ghosh, U. Saha, S. Biswas, Z.A. ALOthman, M.A. Islam, M. Dolai, Ind. Eng. [53] L. Ouyang, X. Tang, H. He, C. Qi, W. Xiong, Y. Ren, H. Jiang, Adv. Synth. Catal. 357
Chem. Res. 61 (2022) 175–186. (2015) 2556–2565.
[19] V.B. Saptal, B.M. Bhanage, Bifunctional ionic liquids for the multitask fixation of [54] Y. Inoue, J. Ishikawa, M. Taniguchi, H. Hashimoto, Bull. Chem. Soc. Jpn. 60 (1987)
carbon dioxide into valuable chemicals, ChemCatChem 8 (2016) 244–250. 1204–1206.
[20] Y. Yuan, Y. Xie, C. Zeng, D. Song, S. Chaemchuen, C. Chen, F. Verpoort, Green [55] K. Iritani, N. Yanagihara, K. Utimoto, J. Org. Chem. 51 (1986) 5499–5501.
Chem. 19 (2017) 2936–2940. [56] J. Hu, J. Ma, Q. Zhu, Q. Qian, H. Han, Q. Mei, B. Han, Green Chem. 18 (2016)
[21] C. Xie, J. Song, H. Wu, Y. Hu, H. Liu, Y. Yang, Z. Zhang, B. Chen, B. Han, Green 382–388.
Chem. 20 (2018) 4655–4661. [57] T. Kimura, K. Kamata, N. Mizuno, Angew. Chem. Int. Ed. 51 (2012) 6700–6703.
[22] S. Dabral, B. Bayarmagnai, M. Hermsen, J. Schießl, Mormul V, A.S.K. Hashmi, [58] (a) Y. Sasaki, Tetrahedron Lett. 27 (1986) 1573–1574;
T. Schaub, Org. Lett. 21 (2019) 1422–1425. (b) C. Bruneau, P.H. Dixneuf, J. Mol. Catal. 74 (1992) 97–107.
[23] S. Biswas, R. Khatun, M. Sengupta, S.M. Islam, Polystyrene supported Zinc complex [59] J.M. Liu, X.G. Peng, J.H. Liu, S.Z. Zheng, W. Sun, C.G. Xia, Tetrahedron Lett. 48
as an efficient catalyst for cyclic carbonate formation via CO2 fixation under (2007) 929–932.
atmospheric pressure and organic carbamates production, Mol.Catal. 452 (2018) [60] X. Peng, F. Li, C. Xia, Synlett (2006) 1161–1164.
129–137. [61] Y. Yuan, Y. Xie, C. Zeng, D. Song, S. Chaemchuen, C. Chen, F. Verpoort, Catal. Sci.
[24] S. Arshadi, E. Vessally, A. Hosseinian, S. Soleimani-amiri, L. Edjlali, Three- Technol. 7 (2017) 2935–2939.
component coupling of CO2, propargyl alcohols, and amines: an environmentally [62] Z. Zhou, C. He, L. Yang, Y. Wang, T. Liu, C. Duan, ACS Catal 7 (2017) 2248–2256.
benign access to cyclic and acyclic carbamates, Journal of CO₂ Util. 21 (2017) [63] S. Dabral, B. Bayarmagnai, M. Hermsen, J. Schießl, V. Mormul, A.S.K. Hashmi,
108–118. T. Schaub, Org. Lett. 21 (2019) 1422–1425.
[25] S. Ghosh, A. Ghosh, S. Biswas, M. Sengupta, D. Roy, S.M. Islam, ChemistrySelect 4 [64] S.S. Islam, S. Biswas, R.A. Molla, N. Yasmin, S.M. Islam, ChemNanoMat 6 (2020)
(2019) 3961–3972. 1386–1397.
[26] P. Sarkar, A.H. Chowdhury, S. Riyajuddin, S. Biswas, K. Ghosh, S.M. Islam, New J. [65] A. Dutta, S. Biswas, M. Dolai, B.K. Shaw, A. Mondal, S.K. Saha, M. Ali, RSC Adv. 5
Chem. 44 (2020) 744–752. (2015) 23855–23864.
[27] X.D. Li, Q.W. Song, X.D. Lang, Y. Chang, L.N. He, AgI/TMG-promoted cascade [66] M.E. Jung, G. Piizi, Disubstituent effect: theoretical basis and synthetic
reaction of propargyl alcohols, carbon dioxide, and 2-aminoethanols to 2- applications, Chem. Rev. 105 (2005) 1735–1766.
oxazolidinones, ChemPhysChem 18 (2017) 3182–3188. [67] J. Hu, J. Ma, Q. Zhu, Q. Qian, H. Han, Q. Mei, B. Han, Zinc(II)-catalyzed reactions
[28] R. Khatun, S. Biswas, I.H. Biswas, S. Riyajuddin, N. Haque, K. Ghosh, S.M. Islam, J. of carbon dioxide and propargylic alcohols to carbonates at room temperature,
CO2 Util. 40 (2020), 101180. Green Chem. 18 (2016) 382–385.
[29] Z. Zhang, Q. Sun, C. Xia, W. Sun, CO2 as a C1 Source: b(C6F5)3-Catalyzed [68] S. Dabral, B. Bayarmagnai, M. Hermsen, J. Schießl, V. Mormul, A.S.K. Hashmi,
Cyclization of o-Phenylene-diamines To Construct Benzimidazoles in the Presence T. Schaub, Silver-catalyzed carboxylative cyclization of primary propargyl alcohols
of Hydrosilane, Org. Lett. 18 (2016) 6316–6319. with CO2, Org. Lett. 21 (2019) 1422–1425.
[30] I.H. Biswas, S. Biswas, M.S. Islam, S. Riyajuddin, P. Sarkar, K. Ghosh, S.M. Islam, [69] B. Grignard, C.G. Ngassamtounzoua, S. Gennen, B. Gilbert, R. Mereau, C. Jerome,
Catalytic synthesis of benzimidazoles and organic carbamates using a polymer T. Tassaing, C. Detrembleur, Boosting the catalytic performance of organic salts for
supported zinc catalyst through CO2 fixation, New J. Chem. 43 (2019) the fast and selective synthesis of α-alkylidene cyclic carbonates from CO2 and
14643–14652. propargylic alcohols, ChemCatChem 10 (2018) 2584–2592.
[31] P. Basu, T.K. Dey, S. Riyajuddin, S. Biswas, K. Ghosh, S.M. Islam, New J. Chem. 44 [70] X. Yu, Z. Yang, F. Zhang, Z. Liu, P. Yang, H. Zhang, B. Yu, Y. Zhao, Z. Liu, A rose
(2020) 12680–12691. bengal-functionalized porous organic polymer for carboxylative cyclization of
[32] W.Z. Zhang, H. Li, Y. Zeng, X. Tao, X. Lua, Palladium-catalyzed cyclization reaction propargyl alcohols with CO2, Chem. Commun. 55 (2019) 12475. -1247.
of o -Haloanilines, CO2 and isocyanides: access to quinazoline-2,4(1H,3H)-diones, [71] S.S. Islam, N. Salam, R.Ali Molla, S. Riyajuddin, N. Yasmin, D. Das, K. Ghosh, S.
Chin.J. Chem. 36 (2018) 112–118. M. Islam, Zinc (II) incorporated porous organic polymeric material (POPs): a mild
[33] S. Biswas, R. Khatun, M. Dolai, I.H. Biswas, N. Haque, M. Sengupta, M.S. Islam, S. and efficient catalyst for synthesis of dicoumarols and carboxylative cyclization of
M. Islam, Catalytic formation of N3-substituted quinazoline-2,4(1H,3H)-diones by propargyl alcohols and CO2 in ambient conditions, Mol. Cat. 477 (2019), 110541.
Pd(II)EN@GO composite and its mechanistic investigations through DFT [72] N. Haque, S. Biswas, S. Ghosh, A.H. Chowdhury, A. Khan, S.M. Islam, ACS Appl.
calculations, New J. Chem. 44 (2020) 141–151. Nano Mater. 4 (2021) 7663–7674.
[34] T.K. Dey, P. Basu, S. Riyajuddin, S. Biswas, A. Khan, K. Ghosh, S.M. Islam, [73] A. Roy, N. Haque, R. Chatterjee, S. Biswas, A. Bhaumik, M. Sarkar, S.M. Islam,
ChemistrySelect 5 (2020) 10355–10366. AgNPs supported over porous organic polymers for the fixation of CO2 on
[35] P. Sarkar, A.H. Chowdhury, S. Riyajuddin, S. Biswas, K. Ghosh, S.M. Islam, New J. propargyl alcohols and amines under solvent-free conditions, New J. Chem. 47
Chem. 44 (2020) 744–752. (2023) 6673–6684.
[36] R. Khatun, S. Biswas, S. Islam, I.H. Biswas, S. Riyajuddin, K. Ghosh, S.M. Islam, [74] H. Ullah, B. Mousavi, H.A. Younus, Z.A.K. Khattak, S. Chaemchuen, S. Suleman,
ChemCatChem 11 (2019) 1303–1312. F. Verpoort, Chemical fixation of carbon dioxide catalyzed via cobalt (III) ONO
[37] A.A.G. Shaikh, S. Sivaram, Chem. Rev. 96 (1996) 951–976. pincer ligated complexes, Commun. Chem. 2 (2019) 42.

You might also like