Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Full Paper

Chemistry—A European Journal doi.org/10.1002/chem.202002996

& Polyhedral Silsesquioxanes

Zinc Imine Polyhedral Oligomeric Silsesquioxane as a Quattro-Site


Catalyst for the Synthesis of Cyclic Carbonates from Epoxides and
Low-Pressure CO2
Mateusz Janeta,* Tadeusz Lis, and Sławomir Szafert*[a]

Abstract: In the present research, the synthesis, spectro- ions like ZnII to form a unique coordination complex. The sil-
scopic characterization, and structural investigations of a sesquioxane core and the hindrance of the side arms (their
unique ZnII complex of imine-functionalized polyhedral oli- steric effect) influence the construction process of the ho-
gomeric silsesquioxane (POSS) is designed, and hereby de- moleptic Zn4@POSS-1 complex. The compound was charac-
scribed, as a catalyst for the synthesis of cyclic carbonates terized in solution by NMR (1H, 13C, 29Si), ESI-MS, UV/Vis spec-
from epoxides and CO2. The uncommon features of the de- troscopy and in the solid state by thermogravimetric/differ-
signed catalytic system is the elimination of the need for a ential thermal analysis (TG-DTA), elemental analysis, diffuse
high pressure of CO2 and the significant shortening of reac- reflectance infrared Fourier transform spectroscopy (DRIFTS),
tion times commonly associated with such difficult transfor- cross-polarization magic angle spinning (CP MAS) NMR (13C,
29
mations like that of styrene oxide to styrene carbonate. Our Si) spectroscopy, and X-ray crystallography.
studies have shown that imine-POSS is able to chelate metal

Introduction sist of a metalated T7 moiety, R7Si7O12M (where M = a d-block


element; R = an isobutyl or cyclopentyl group), often with ad-
Polyhedral silsesquioxanes, organosilicon compounds de- ditional ligands coordinated to the metal site (Figure 1 a).[10–14]
scribed by the chemical formula (RSiO3/2)n, are representatives For example, Duchateau et al. described the preparation of ZnII
of a class of cage-like compounds.[1–3] Polyhedral oligomeric sil- alkyl silsesquioxane complexes for the co-polymerization of cy-
sesquioxane (POSS) cages include usually T6, T8, T10, and T12, clohexene oxide and CO2.[15] Zirconium(IV) silsesquioxane com-
structures, however, the most widely studied are the cubic T8 plexes have been investigated as model catalysts for alkene
as well as, recently, the lantern,[4] butterfly,[5] Janus,[6] or double- polymerization and the same research group has also prepared
decker[7] types of cages. On the basis of their diameters, which similar SnII, AlIII, and GaIII species.[16–18] Silsesquioxanes contain-
range from 1 to 3 nm, they can be considered as the smallest ing Si-O-M (where M = zirconium,[19] titanium,[20] or iron[21]) con-
existing silica nanoparticles. Their unique star-shaped nano- nections have been studied as molecular models for heteroge-
structures make POSS an excellent building block for con- neous catalysis.
structing of multi-functional materials.[8, 9] The advantage of More recently, this interest has expanded to encompass oc-
POSS lies in its wide range of interesting physicochemical pa- tahedral systems. T8 compounds containing metals may be di-
rameters, that is, facile chemical modification, good pH toler- vided into two groups: (1) those with a metal ion directly
ance, high temperature and oxidation stability, mechanical re- bonded to the silsesquioxane core (mostly via an oxygen or sil-
sistance, optical transparency, reduced heat transfer coefficient, icon atom, Figure 1 b) and (2) those with seven inert organic
resistance to fire, and hardness of silsesquioxane-based materi- substituents and a single metalated site on one outside arm of
als. the cube (Figure 1 c).
Recently, in the field of silsesquioxanes, we have observed As said already, the metal sites in metal-silsesquioxane can
an increased interest in POSS-based systems, which could act as catalytic centers. A rare example of a crystallographically
serve as homogeneous catalysts. Such systems generally con- characterized metalated silsesquioxane-based catalytic system
is that of Maschmeyer et al., who have used silsesquioxanes as
soluble models for tethered OsIV and RhII complexes.[25, 26] Also,
[a] Dr. M. Janeta, Prof. T. Lis, Prof. S. Szafert
Faculty of Chemistry Ervithayasuporn et al.[24] reported a mononuclear organopalla-
University of Wrocław dium-functionalized T8-Pd POSS as a homogeneous catalyst for
F. Joliot-Curie 14, 50-383 Wrocław (Poland) the Suzuki–Miyaura reaction (Figure 1 c). Marciniec et al. de-
E-mail: mateusz.janeta@chem.uni.wroc.pl
scribed a ruthenium-silsesquioxyl complex active in silylative
slawomir.szafert@chem.uni.wroc.pl
coupling reactions (Figure 1 b).[22] Jones et al. synthesized and
Supporting information and the ORCID identification number(s) for the au-
thor(s) of this article can be found under: structurally characterized AlIII[23] and ZnII[27] POSS complexes
https://doi.org/10.1002/chem.202002996. and reported on their activity in the ring-opening polymeri-

Chem. Eur. J. 2020, 26, 13686 – 13697 13686 T 2020 Wiley-VCH GmbH
Full Paper
Chemistry—A European Journal doi.org/10.1002/chem.202002996

Cyclic carbonates are characterized by a range of applica-


tions, for instance: monomers for polymerization, substrates in
organic synthesis, green solvents, or electrolytes for lithium-ion
batteries.[48, 49] From the above reasons, a number of research
groups have focused their attention on the preparation of effi-
cient catalytic systems for cyclic carbonate synthesis including
metal complexes with imine,[50–54] aminophenolate,[55, 56] por-
phyrin,[57, 58] or other ligands.[59–62] Also, an organocatalytic ap-
proach has actively been investigated.[63–65]
It is also known that the cooperative effects of two or more
metal sites in salphen complexes play an important role in sev-
eral catalytic reactions. Therefore, efforts have been made to
synthesize molecular structures with multiple salphen/salen
binding pockets, creating a potential for placement of two or
more catalytically active metal centers in a predefined three-di-
Figure 1. Selected types of silsesquioxane complexes: a) T7 system, b) T8 mensional geometry.[66]
system with a metal ion directly bonded to the silsesquioxane core, c) sys- Over the last two decades, the North group has developed
tems with seven inert organic substituents and a single metalated site on aluminium-based salen complexes as highly effective catalysts
one outside arm of the cube.[21–24]
for the synthesis of cyclic carbonates from epoxides and
carbon dioxide.[67–69] Their studies resulted in the development
zation of rac-lactide. Grela et al.[28] reported a ruthenium POSS of a bimetallic complex (Figure 2 b), which was shown to be a
complex active in Hoveyda–Grubbs-type olefin metathesis, more efficient catalyst for the formation of cyclic carbonates
which can be recovered by using nanofiltration techniques. from epoxides than the monometallic aluminium(salphen)
Other types of catalytically active mononuclear silsesquioxane complexes.[70] The same group has also developed silica-sup-
metal complexes include a zirconium-functionalized POSS, re- ported catalysts with propylene oxide, which are reusable over
ported by Severn et al., which acts as a homogeneous model 30 times and could convert up to 98 % of CO2 into ethylene
for silica-supported olefin polymerization catalysts.[29] carbonate with use of a flow reactor.[71] Kleij et al.[72] has report-
To date, metal complexes of silsesquioxanes are created ed several Zn(salphen) complexes as effective, cheap, robust,
through either the introduction of metal ions by post-function- and relatively green catalysts for the cycloaddition of carbon
alization[30, 31] or the use of predefined metalated ligand build- dioxide to terminal epoxides to afford cyclic carbonates in
ing blocks as reactants.[32] To the best of our knowledge, there
is no example in the literature of a well-defined (structurally
characterized) T8-POSS that coordinates more than one metal
cation using side arms.[33]
On the other hand, carbon dioxide is a well-known green-
house gas that originates from the carbon footprint of human
activities and the natural carbon cycle.[34–36] Neutralization of
CO2 is of great significance for the deceleration of global
warming and the development of sustainable energy. One, and
a potentially long-term, solution to balance these emissions is
permanent CO2 trapping into minerals;[37–41] however, this
method requires considerable costs and energy. Recent prog-
ress in coordination chemistry and catalysis could provide an
effective means for the chemical transformation of CO2 and its
longer-term incorporation into organic molecules under mild
conditions.[42, 43] In this regard, much attention has been paid
to the synthesis of cyclic carbonates or polycarbonates from
epoxides and CO2 (Scheme 1) owing to the their useful proper-
ties.[44–47]

Figure 2. Selected complexes that are effective catalysts for the synthesis of
cyclic carbonates from epoxides and carbon dioxide: a) monometallic zinc-
(salphen) complexes,[72] b) bimetallic aluminium(salphen),[70] c) trimetallic
Scheme 1. Reaction between CO2 and epoxides. R can be H, alkyl, or aryl. metal-salen molecular cages.[48]

Chem. Eur. J. 2020, 26, 13686 – 13697 www.chemeurj.org 13687 T 2020 Wiley-VCH GmbH
Full Paper
Chemistry—A European Journal doi.org/10.1002/chem.202002996

good yields (Figure 2 a). Wu et al.[48] described AlIII and CoIII chains. In the 1H NMR spectrum, only one characteristic signal
salen based molecular cages, which have proven to be efficient of the imine group proton at d = 8.33 ppm was observed and
and recyclable heterogeneous catalysts for cycloaddition of the phenol OH signal was located at 13.95 ppm. Moreover, the
CO2 with epoxides under ambient conditions. presence of the iminopropyl group (-N = CH) was evidenced
Inspired by these reports, we designed, and hereby describe, by a single resonance at 158.4 ppm in the 13C NMR spectrum.
the synthesis, characterization, and structural investigations of The mass spectrum also unambiguously confirmed the forma-
a new ZnII complex with polyhedral oligomeric silsesquioxane tion of a closed-frame structure composed of eight silicon
as a catalyst for the synthesis of cyclic carbonates from epox- atoms possessing terminal alkyl-imine side chains showing a
ides and CO2. This search relates to the industry search for a signal at m/z = 2610.4979 (calcd 2610.4984) with a correct iso-
thermally stable systems that would allow the process to be tope envelope. In the diffuse reflectance infrared Fourier trans-
conducted at low CO2 pressures. According to this, we envi- form (DRIFT) spectrum of POSS-1, a sharp band appeared at
sioned POSS as perfect carriers for anchoring of a metal that is 1636 cm@1, which was attributed to the nC=N stretching vibra-
active in cyclic carbonates formation. tions of the imine group. Owing to an intramolecularly hydro-
Imine-POSS comprises a less-explored group of modified gen-bonded OH group, the nO-H vibrations displayed a band in
POSSs. Our previous studies demonstrated that imine-POSS are the 3100–2500 cm@1 region. Also, the Si-O-Si stretching gave
a versatile precursor for building supramolecular hybrid materi- rise to strong absorptions at around 1116 cm@1.
als.[73] Synthesis of well-defined metal complexes of POSS is Metalation of POSS-1 with zinc(II) acetate in a mixture of di-
challenging and to date, efficient methods for their prepara- chloromethane and methanol (v/v, 1:1) yielded a tetranuclear
tion have been lacking. At the same time, no papers have de- coordination compound Zn4@POSS-1 (Scheme 3). The complex
scribed the chemistry or the synthetic aspects of metal com- was also successfully synthesized in high yield by a one-step
plexes of imino-octafunctionalized POSS (so-called multifunc- protocol using ZnEt2 in toluene and POSS-1 with a precise 4:1
tional POSS). It was interesting to determine whether such molar ratio under dinitrogen at @78 8C. When POSS-1 was re-
POSS moieties can play the role of multifunctional ligands ca- acted with an excess (5–8 equivalents) of diethylzinc to obtain
pable of anchoring more than just one metal center. compounds with higher (than four) numbers of coordinated
zinc(II) ions, we invariably observed only the formation of
Zn4@POSS-1. Recrystallization of Zn4@POSS-1 from CHCl3 led
Results and Discussion to isolation of an analytically pure product, which was charac-
terized by NMR spectroscopy and mass spectrometry (see the
Synthesis
Experimental Section and Figures S9–S18 in the Supporting In-
A zinc(II) complex with Schiff base attached POSS was conven- formation). It should be noted that Zn4@POSS-1 is stable
iently prepared in two steps. First, octa-iminofunctionalized under ambient conditions and can be stored for several weeks
POSS derivative POSS-1 was obtained in high yield, by using without any signs of degradation.
octa(3-aminopropyl)silsesquioxane (OAS-POSS), 3,5-di-tert- The formation of Zn4@POSS-1 was confirmed by 1H NMR
butyl-2-hydroxybenzaldehyde, and triethylamine as a deproto- spectroscopy, which, in CDCl3, showed two sets of resonances
nation agent as shown in Scheme 2. The reaction was conduct- assigned to the phenyl ring and two assigned to the diastereo-
ed at room temperature in a mixture of CH2Cl2/methanol (v/v, topic -CH2- protons associated with iminopropyl functional
1:5) resulting in 91 % of POSS-1. groups arising from the ligand. Compared with the spectra of
The new imino-functionalized POSS-1 was characterized by the pure ligand, the lower chemical shifts observed for the
NMR (1H, 13C, 29Si), DRIFT spectroscopy, and ESI-MS spectrome- CH = N (at 8.17 and 8.10 ppm) moiety and the absence of the
try (Figures S3–S8 in the Supporting Information). In solution, broad OH signal suggested its coordination through the imine
it possesses Oh symmetry, manifested in its NMR (1H, 13C) spec- and phenoxo groups, indicating a lower symmetry than that of
tra by a set of symmetry-equivalent signals of organic side POSS-1. In the DRIFT spectrum of Zn4@POSS-1, new bands,

Scheme 2. Synthesis of POSS-1.

Chem. Eur. J. 2020, 26, 13686 – 13697 www.chemeurj.org 13688 T 2020 Wiley-VCH GmbH
Full Paper
Chemistry—A European Journal doi.org/10.1002/chem.202002996

Scheme 3. Preparation of Zn4@POSS-1.

which were absent in the spectra of the free ligand, were ob- mers with D4h or S4 symmetry (Figure 3). D4h symmetry of
served at 1531 and 1434 cm@1; these were attributed to the co- Zn4@POSS-1 was authenticated by NMR and single-crystal X-
ordinated nC-O mode. Bands at 539 and 517 cm@1 were as- ray diffraction studies as indicated below.
signed to the Zn@O and Zn@N bonds, respectively. The nC=N
band was shifted to 1613 cm@1, whereas the nO-H stretching vi-
brations disappeared, indicating that the ligand was coordinat-
ed in its phenol-imine form. In the DRIFT spectrum, no
CH3COO@ moieties were observed, proving that the ZnII ions
were connected by imine ligands only. This result was in agree-
ment with the NMR spectroscopy observations (see above).
Characteristic vibrations of the Si-O-Si moieties were observed
at 1097 cm@1. The absence of Si-OH group absorptions sup-
ported the preservation of the cage-like structure,[73] which
was further supported by our observations from solution and Figure 3. Schematic representation of the two possible connections of side
solid-state 29Si NMR analysis (see below). Tetranuclear stoichi- arms of imine-POSS.
ometry with the formula Zn4@POSS-1 was confirmed with the
ESI-MS spectrum of the isolated complex, which showed ad-
X-ray analysis
ducts with protons [Zn4@POSS-1 + 2 H]2 + (Figure S18 in the
Supporting Information). Elemental analyses were also consis- A single crystal of Zn4@POSS-1 suitable for X-ray diffraction
tent with the proposed formulations. Importantly, the POSS analysis was obtained by slow evaporation of a chloroform so-
cage did not react with either ZnEt2 or Zn(CH3COO)2 and did lution of the complex. Zn4@POSS-1 crystallizes in a monoclinic
not decompose during the complexation reaction. system in C2/c space group. The X-ray structure of Zn4@POSS-
To further confirm our observations, we acquired 13C and 29Si 1 is shown in Figure 4 and in Figure S1 (in the Supporting In-
NMR spectra in solution and in the solid state. The 29Si NMR formation). The solid-state structure revealed the formation of
spectrum of Zn4@POSS-1 in solution (Figure S11 in the Sup- a neutral four-coordinate tetranuclear complex. Each metal
porting Information) contained only one resonance at center is ligated by two imine nitrogens and two phenolate
@66.8 ppm, which means that cage rearrangement did not oxygens of two neighboring iminophenolate side arms of
occur. In addition, the 29Si cross-polarization magic angle spin- POSS-1. Each side arm plays the role of a ligand in a bidentate
ning (CP MAS) spectrum (Figure S16 in the Supporting Infor- (k2) fashion. Zinc(II) ions adopt a distorted tetrahedral geome-
mation) recorded for a bulk sample also showed only one sym- try. Two zinc atoms have a configuration D and the other two
metric resonance at @66.8 ppm. Thus, these results unambigu- L, resulting in an overall D4h molecule symmetry of meso-
ously confirmed that only a T8 cage is present in the resulting (D,D,L,L)-Zn4@POSS-1. Owing to steric hindrance by the tert-
structure. butyl groups, rotation around the Zn atom is blocked, which
Imine-POSS in which two side-chain iminopropyl group are prevents Zn4@POSS-1 from isomerization (Figure 4, Figure 5 a).
connected through a metal atom, can form one of the two iso- Zn@O bond lengths are in the 1.914(5)–1.924(4) a range and

Chem. Eur. J. 2020, 26, 13686 – 13697 www.chemeurj.org 13689 T 2020 Wiley-VCH GmbH
Full Paper
Chemistry—A European Journal doi.org/10.1002/chem.202002996

Figure 4. X-ray crystal structure of Zn4@POSS-1 (left). Hydrogen atoms and chloroform molecules have been omitted for clarity. View of the coordination envi-
ronments of the zinc center (right) in Zn4@POSS-1. Symmetry code: i3/2@x, 3/2@y, 1@z.

Figure 5. X-ray crystal structure of Zn4@POSS-1·10 CHCl3 : a) view of a chloroform molecule trapped in the cavity created by the side group of Zn4@POSS-1,
b) space-filling representation, c) surface plot of the pores formed by the arrangement of solvent molecules lying among Zn4@POSS-1 molecules. Representa-
tion of channels along the b-axis. The image shows a 1.5 V 1.5 V 1.5 array of unit cells.

Zn@N bonds in the 1.979(6)–2.016(5) a range, typical for ZnII < vdW sum), Si···HC (3.149 a, @4.58 % < vdW sum) to produce
Schiff base complexes, for which Zn@O and Zn@N bond an extended 3D network. Data on the geometry of the hydro-
lengths fall within the ranges of 1.890–2.139 a and 1.980– gen bonds and the close noncovalent contacts are presented
2.128 a, respectively.[74] The Si@O distances in the centrosym- in Table S3 and in Figure S2 (in the Supporting Information).
metric molecule fell within the range 1.611(5)–1.633(5) a. The Zn4@POSS-1 co-crystallized with ten chloroform molecules,
Si atoms tended to maintain tetrahedral geometry, with O-Si-O which were trapped between the side arms of Zn4@POSS-1
angles ranging from 107.8(2) to 109.9(2)8, however, the Si@O@ and connected by C@H···p and C@H···O interactions (Figure 5 a),
Si bond angles, strained by the cuboidal geometry imposed by creating an interconnected torus-like aggregate. The volume
the Si atoms, widened to 139.6(3)–167.3(3)8, representing a sig- of chloroform molecules (contact surface) in a Zn4@POSS-1
nificant deviation from the ideal oxygen bond angle of 1048 crystal was estimated at 32.3 % of the unit cell volume
(for more details, see the Supporting Information, Table S2 and (19 683 a3 per unit cell), with a solvent-accessible pore volume
Figure S1, and the CIF file). Nevertheless, all values concerning of 7.6 % (1489 a3 per unit cell; Figure 5 b, c). Zn4@POSS-1
the core fell within the standard range for T8-like silsesquiox- indeed appears to be a possible host for host/guest materials.
anes.[73] Compounds that possess noncovalent interactions may serve
Analysis of the crystal structure revealed that supramolecular as molecular containers for small molecules with potential uses
aggregation in Zn4@POSS-1 occurred through weak coopera- in so-called green reactions, in which selective trapping of a
tive noncovalent interactions CH2···H3C (2.146 a, @10.6 % below specific small molecules from a reaction system is a key per-
the sum of van der Waals radii), CH3···H3C (2.313 a, @3.63 % formance of the molecular reactor.

Chem. Eur. J. 2020, 26, 13686 – 13697 www.chemeurj.org 13690 T 2020 Wiley-VCH GmbH
Full Paper
Chemistry—A European Journal doi.org/10.1002/chem.202002996

Thermal properties UV/Vis absorption

The inorganic framework (Si-O-Si) ensures the chemical and The UV/Vis absorption spectra for POSS-1 and Zn4@POSS-1
thermal resistance of the POSS compounds. Compared with were investigated in CH2Cl2 solutions at a concentration of
pure organic compounds, hybrid organic–silsesquioxanes are 10@7 mol L@1 (Figure 7) and the corresponding data are com-
characterized by a much higher thermal stability. To character- piled in the Experimental Section. The UV/Vis spectrum of
ize the thermal stability of POSS-1 and Zn4@POSS-1, a thermal POSS-1 showed three bands; p!p* transitions of the C=N
gravimetric analysis (TGA) was carried out in an oxidative (O2/ groups are in the range 290–375 nm, whereas the bands in the
N2 = 40:60) and anaerobic (nitrogen flow) environment. The ranges 220–250 and 250–280 nm may be attributed to p!p*
TGA profiles are presented in Figure 6 (oxidative environment) and n!p* transitions of the aromatic rings. The UV/Vis spec-
and Figure S19 (nitrogen flow; in the Supporting Information). trum of Zn4@POSS-1 displays an absorption pattern similar to
The TGA curve of POSS-1 indicated its high thermal stability that of the ligand, but bathochromically shifted. The large
up to 346 (nitrogen flow) and 321 8C (oxidative environment), (compared with that observed for free Schiff base POSS-1) red-
which can be attributed to the presence of the rigid silses- shift can be assigned to the reduction of the transition energy
quioxane cage.[75] Oxidative decomposition of POSS-1 occurred levels owing to a high degree of electron density on the Schiff
in two main steps. The first weight loss was related to a de- base ligand.[76]
composition of organic side chains and the second to that of
the siloxane cage. Continuation of the analyses up to 1000 8C
under oxidative conditions enabled determination of the mate-
rials composition based on their ceramic yields.[75] The final
product of thermooxidative degradation was silicon oxide in
crystobalite form (ICSD number: 75483), as confirmed by X-ray
powder diffraction. The TGA curve of POSS-1 revealed a total
weight loss of 81.8 %, corresponding to 18.2 % of remaining
SiO2, which is close to the theoretical value (18.4 %).

Figure 7. The electronic absorption spectra (in dichloromethane) for POSS-1


(black) and Zn4@POSS-1 (red).

Catalytic test using Zn4@POSS-1 for the coupling of CO2


with epoxides
The effects of the nature of the co-catalyst, catalyst/co-catalyst
Figure 6. TGA trace graph (in an oxidative environment) for POSS-1 and ratio, solvent, and temperature were examined by using sty-
Zn4@POSS-1. Theoretical values of weight loss are given in parentheses. rene oxide as the model substrate. The initial conditions were
1 atm CO2, 100 8C, 1 mol % of catalyst, and 1 equivalent of TBAI
(relative to Zn), as a co-catalyst in neat substrate for 4 h. The
For Zn4@POSS-1, the initial weight loss of 3 % was related to results are summarized in Table 1. Under these conditions, no
the release of the remaining solvent molecules trapped be- catalytic activity was observed when using Zn4@POSS-1 alone
tween the side arms. The thermal behavior of Zn4@POSS-1 fol- (Table 1, entry 1). A control experiment without Zn4@POSS-1
lowing desolvation was similar to that for POSS-1, with a more catalyst resulted in negligible conversion of styrene oxide
important mass loss occurring in the second step as a result of (Table 1, entry 2).
decomposition of the organic arms (first step for POSS-1). The The optimal catalyst center/co-catalyst ratio was 1:1, as no
TGA results show that the thermal stability of desolvated improvement was achieved if the ratio was increased to 1:2
Zn4@POSS-1 at the 5 % weight reduction temperature (DT5 %) (Table 1, entry 7). An increase in temperature from 100 to
reached 346 8C under the nitrogen flow and 334 8C in an oxida- 130 8C resulted in a higher conversion (98 %, turnover frequen-
tive environment. The total weight loss of 75.7 % was close to cy (TOF) = 56 h@1, Table 1, entry 11), whereas at 80 8C a de-
the theoretical value of 74.1 % calculated for Zn4Si8O16. crease in activity was observed (45 %, TOF = 10 h@1, Table 1,
entry 3). When the catalyst loading of Zn4@POSS-1 was re-
duced to 0.1 mol %, the conversion of styrene oxide to styrene
carbonate decreased to 49 % (Table 1, entry 8). A longer reac-

Chem. Eur. J. 2020, 26, 13686 – 13697 www.chemeurj.org 13691 T 2020 Wiley-VCH GmbH
Full Paper
Chemistry—A European Journal doi.org/10.1002/chem.202002996

and leaving group ability of iodide compared with bromide,


Table 1. Optimization of reaction conditions for the cycloaddition of sty-
rene oxide and CO2.[a] which enhanced the ring-opening and -closing catalytic
steps.[77]
To investigate the mechanism involved in the synthesis of
cyclic carbonate, a study of the reaction kinetics was per-
formed by using [D7]DMF as a solvent. Reaction progress was
observed with real-time NMR spectra. The conversion of sty-
rene oxide and the resulting yield increased along with in-
Entry Cat. Co-cat. Temp. [8C] Time [h] Conversion [%][b]
creasing reaction times, from 0 to 4 h (Figure 8). After 1 h (at
100 8C), more than 40 % of the substrates were converted and
1 Zn4@POSS-1 – 100 4 0
2 – TBAI 100 4 1
a styrene carbonate yield of 60 % was detected. Full conversion
3 Zn4@POSS-1 TBAI 80 4 45 of styrene oxide was achieved with the reaction time of 4 h
4 Zn4@POSS-1 TBAI (0.25)[c] 100 4 85 (Figure 8). For the conversion of styrene oxide with 1 atm CO2
5 Zn4@POSS-1 TBAI (0.5)[c] 100 4 90 pressure at three different temperatures (80 8C, 100 8C, and
6 Zn4@POSS-1 TBAI 100 4 96
7 Zn4@POSS-1 TBAI (2)[c] 100 4 96
130 8C), first-order kinetics were observed.
8 Zn4@POSS-1[d] TBAI 100 4 49 To investigate if the zinc(II) centers in Zn4@POSS-1 function
9 Zn4@POSS-1[d] TBAI 100 12 95 separately or cooperatively, we used the method described by
10 Zn4@POSS-1 TBAI 100 10 96 North et al.,[70] which is based on the investigation of the effect
11 Zn4@POSS-1 TBAI 130 4 98
of changing the metal active center/TBAX ratio (X can be Cl,
[a] Reaction conditions: styrene oxide (4.4 mmol), catalyst (44 mmol, Br, or I). If they function separately, then one equivalent of
1 mol %), co-catalyst (4 mol %, 1 equiv rel. to Zn), CO2 (1 atm). [b] Deter-
TBAX per metal center will be required, whereas if they func-
mined by using 1H NMR spectroscopy. [c] Number in brackets is the
equivalents per Zn. [d] Reaction conditions: styrene oxide (4.4 mmol), cat- tion cooperatively then less than one equivalent of TBAX per
alyst (4.4 mmol, 0.1 mol %), co-catalyst (0.4 mol %, 1 equiv rel. to Zn), CO2 metal will be needed. In the case of Zn4@POSS-1, using sty-
(1 atm). rene epoxide at 100 8C and 1 atm carbon dioxide pressure for
4 h, 96 % conversion was obtained when using 1 mol % of
Zn4@POSS-1 and 1 equiv (rel. to Zn) of TBAI whilst 90 % con-
tion time (12 instead of 4 h) at this reduced catalyst loading al- version was obtained when using 1 mol % of Zn4@POSS-1 and
lowed us to achieve 95 % conversion (Table 1, entry 9). 0.5 equiv (rel. to Zn) of TBAI. Further reduction of the amount
The influence of the co-catalyst was also investigated and of TBAI to 0.25 equiv (rel. to Zn) resulted in a reduction of the
the results are summarized in Table 2. It is well known that the conversion to 85 %, which may indicate that zinc atoms work
nucleophilicity and leaving group ability of the co-catalyst can cooperatively. These results are consistent with a mechanism
determine its activity.[69] N-Methylimidazole (NMI) and 4-dime- that was previously proposed for aluminium(salphen) com-
thylaminopyridine (DMAP) were weak co-catalysts as only 5 % plexes[70] in which one metal center activates the epoxide
and 8 % conversions were observed after 4 h (Table 2, entries 1 while the other activates the carbon dioxide to allow intramo-
and 2). When tetrabutylammonium fluoride (TBAF), chloride lecular carbonate formation. A possible mechanism for the
(TBACl), or tetrabutylammonium bromide (TBAB) were used as
a co-catalysts, low conversions and TOFs were also observed
(Table 2, entries 3–5). However, when tetrabutylammonium
iodide (TBAI) was used (Table 2, entry 6), an increase in conver-
sion up to 96 % (TOF = 25 h@1) was noticed. The higher activity
observed for iodide (96 %) compared with bromide (40 %) and
other halogens can be related to the higher nucleophilicity

Table 2. Influence of co-catalyst on the catalytic activity of Zn4@POSS-


1.[a]

Entry Co-cat. TOF [h@1][b] Conversion [%][c]


1 NMI 1.25 5
2 DMAP 2 8
3 TBAF 1.5 6
4 TBACl 2.5 10
5 TBAB 10 40
6 TBAI 25 96 Figure 8. Kinetic profiles for the formation of styrene carbonate from styrene
oxide and CO2 at different reaction temperatures when using the
[a] Reaction conditions: styrene oxide (4.4 mmol), Zn4@POSS-1 (44 mmol,
Zn4@POSS-1/TBAI system. Reaction conditions: substrate (0.22 mmol), cata-
1 mol %), co-catalyst (4 mol %, 1 equiv rel. to Zn), CO2 (1 atm), 100 8C, 4 h.
lyst Zn4@POSS-1 (1 mol %), CO2 (1 atm), co-catalyst TBAI (4 mol %, 1 equiv
[b] TOF = moles of product/(moles of active site of catalyst V time) at
rel. to Zn), solvent [D7]DMF (0.6 mL). The solid line represents the fit with
10 % conversion. [c] Determined by using 1H NMR spectroscopy.
the pseudo-first-order kinetic model.

Chem. Eur. J. 2020, 26, 13686 – 13697 www.chemeurj.org 13692 T 2020 Wiley-VCH GmbH
Full Paper
Chemistry—A European Journal doi.org/10.1002/chem.202002996

Scheme 4. Proposed cooperative mechanism for cyclic carbonate synthesis catalyzed by the Zn4@POSS-1/TBAI system.

conversion of styrene oxide by Zn4@POSS-1/TBAI is shown in perature was raised to 130 8C, the conversion only increased to
Scheme 4. First, the zinc-salen acts as a Lewis acid to coordi- 70 %. Thus, higher conversion (96 %) of styrene carbonate was
nate styrene oxide to activate the epoxide ring (polarizes it) for obtained with tetrametallic Zn4@POSS-1 than with the mono-
nucleophilic attack from the iodine anion from TBAI. Simulta- metallic analog 1 (60 %) under the same reaction conditions,
neously, the tetrabutylammonium iodide molecule forms tribu- supporting the occurrence of intramolecular cooperative catal-
tylamine,[78] which reacts reversibly with CO2 to create a carba- ysis with Zn4@POSS-1.
mate salt. The polymetallic nature of Zn4@POSS-1 allows it to The recyclability of Zn4@POSS-1 was finally evaluated.
coordinate to carbamate. Displacement of the tributylammoni- Zn4@POSS-1 retained most of its catalytic activity even after
um group generates the metal-coordinated carbonate and five runs, with the conversions decreasing from 99 % to 90 %
subsequent ring-closure forms the cyclic carbonate and regen- (Figure 10). The NMR spectra of Zn4@POSS-1 after the catalytic
erates both Zn4@POSS-1 and tetrabutylammonium iodide. reaction was the same as before, which may indicate that the
The proposed mechanism emphasizes that the role of the molecular structure is preserved. Mass spectrometry data con-
catalyst involves initial activation of the epoxide and stabiliza- firm that Zn4@POSS-1 retains its tetrametallic structure during
tion of its ring-opened form (iodo-alkoxide) and carbonate in- reactions and after its use in five consecutive reactions. More-
termediates formed during the reaction, which is important to over, the inductively coupled plasma optical emission spectros-
reduce the reaction pressure of CO2 and the reaction time. The copy (ICP-OES) analysis of the reaction mixture after filtering
catalytic performance depends on the nucleophilicity and leav- showed only a trace amount (below 8 ppm) of zinc atoms, in-
ing group properties of the halide, as well as the stabilizing dicating negligible leaching of Zn into the reaction mixture.
effect of the counterion. Moreover, the in situ formation of tri-
alkylamine followed by the formation of a carbamate salt with
CO2 was considered to play an additional catalytic role in this
process.[68]
To further study the cooperative nature of the catalyst, we
compared the obtained results with those for monometallic
zinc(salen)[79] (Figure 9). When 4 mol % of compound 1 and
1 equiv TBAI were used as the catalyst, at 1 atm of CO2 and
100 8C, after 4 h only 60 % conversion of styrene oxide into
cyclic carbonate was observed. Even when the reaction tem-

Figure 10. Reaction conditions: styrene oxide (0.22 mmol), catalyst


(2.2 mmol, 1 mol %), co-catalyst (4 mol %, 1 equiv rel. to Zn), CO2 (1 atm),
T = 100 8C, no solvent, 4 h.

Figure 9. Bis(salicylaldimine) zinc(II) complex Zn4@POSS-1 counterpart.[79]

Chem. Eur. J. 2020, 26, 13686 – 13697 www.chemeurj.org 13693 T 2020 Wiley-VCH GmbH
Full Paper
Chemistry—A European Journal doi.org/10.1002/chem.202002996

Substrate scope confirmed with X-ray analysis (see above). The distance be-
tween two zinc atoms is 11 a and is close enough for coopera-
We next investigated the scope and limitations of the tive interactions to take place between them. The metallic
Zn4@POSS-1/TBAI system for a variety of terminal and internal center is more easily accessible to terminal epoxides. The fact
epoxides, which included: 1-chloro-2,3-epoxypropane, 1,2-ep- that low conversion was observed for internal epoxides like cy-
oxyethylbenzene, 1,2-epoxybutane, 1,2-epoxycyclohexane, 1,2- clohexene oxide can suggest that the reaction occurs inside
epoxypropane, 1,2-epoxyhexane, and 1,2-epoxy-3-phenoxypro- the cavities.
pane (Figure 11). The influence of the substrate was explored The activity of the Zn4@POSS-1/TBAI system is significant
at a reaction temperature of 100 8C in the presence of compared with the most efficient amino- or imino-phenolate
Zn4@POSS-1 and TBAI under a pressure of 1 atm CO2. To deter- complexes reported heretofore (based on Zn,[82, 83] Al,[84]
mine isolated yields, the mixture after reaction was purified by Mn,[72, 85] Fe,[86] Co,[87, 88] and Cr[89, 90]), which usually operate in
means of flash chromatography (using a mixture of hexane/ the presence of CO2 under high pressure (up to 30 bar), with
EtOAc). For all studied epoxides, formation of cyclic carbonate catalyst loadings of 0.01–6.6 mol %, and which require a pro-
was unambiguously confirmed by NMR spectroscopy (Figur- longed reaction time (up to 48 h).[72] For example, for mononu-
es S20–S31 in the Supporting Information), which confirmed clear Zn(salphen), the efficiency after 18 h (under 10 mbar CO2)
the high selectivity of the catalytic system and an absence of is 66 %.[72] Moreover, we investigated new applications of
polycarbonate or hydrolysis products. Industrially, the most imine-POSS in catalysis with a green approach, preferring the
challenging substrate for cyclic carbonate synthesis is styrene use of relatively nontoxic metals such as zinc and avoiding the
oxide owing to its lower reactivity and selectivity.[80, 81] In the use of solvents that exert a negative impact on the environ-
case of the Zn4@POSS-1/TBAI system, high yields (85–99 %) of ment.
cyclic carbonates were achieved under mild conditions for ter-
minal epoxides, illustrating that this catalytic system was effec-
tive with both alkyl and aryl epoxides, tolerating functional Conclusion
groups such as halides and ethers (Figure 11). The modest A new type of tetranuclear ZnII coordination compound was
yields of propylene and butylene carbonate were due to the designed and synthesized utilizing a reaction of ZnII ions with
high volatility of propylene and butylene oxide even at room imine-functionalized polyhedral silsesquioxane. For the first
temperature. An internal epoxide, cyclohexene oxide, was con- time, a multifunctional imine-POSS-based metal complex was
verted into the corresponding cyclic carbonate with 5 % yield. structurally characterized. Zn4@POSS-1 was obtained in high
Even using TBAB as a co-catalyst did not increase the yield sig- yields through simultaneous complexation reactions of imine-
nificantly (7 % yield). The reduction of activity was in this case functionalized silsesquioxane with ZnII salts under mild condi-
caused by the steric hindrance induced by the silsesquioxane tions. Zn4@POSS-1 was successfully tested in the cycloaddition
substituents. X-ray crystallography showed that Zn4@POSS-1 is of CO2 with terminal epoxides, with tetrabutylammonium
characterized by large structural elements (silsesquioxane core) iodide as a co-catalyst, which ensured a high activity and selec-
erected over the active site region, allowing substrate access tivity of the tested catalytic system. Styrene oxide, which is
only through a deep, narrow channel. This architecture ac- usually a challenging substrate, was converted under mild con-
counts for the tendency of the described catalyst to specialize ditions (1 atm CO2) and with a short reaction time (4 h) into
in small substrates without bulky groups and to maintain the the corresponding cyclic carbonate, in high yield (96 %). For
reaction at the stage of formation of cyclic carbonate. The side terminal epoxides, high yields (85–99 %) of cyclic carbonates
arms of Zn4@POSS-1 form cavities that are large enough (12 a were achieved under mild conditions, illustrating that this cata-
in diameter) to accommodate small molecules, which has been lytic system was effective with both alkyl and aryl epoxides
while tolerating functionalities including halides and ethers.

Experimental Section
General procedures and chemicals
Dichloromethane (99 %, Chempur), chloroform (99 %, Chempur),
methanol (99 %, Chempur), 3,5-di-tert-butyl-2-hydroxybenzalde-
hyde (99.9 %, Aldrich), 1-chloro-2,3-epoxypropane (99 %, Acros),
1,2-epoxybutane (99 %, Sigma–Aldrich), 1,2-epoxycyclohexane
(98 %, EGA-chem), 1,2-epoxyethylbenzene (97 %, Sigma–Aldrich),
1,2-epoxyhexane (97 %, Acros), 1,2-epoxy-3-phenoxypropane (99 %,
Acros), 1,2-epoxypropane (+ 99.5, Sigma–Aldrich), tetrabutylammo-
nium bromide (99 %, Lancaster), tetrabutylammonium chloride
Figure 11. Isolated yields of cyclic carbonates formed by the cycloaddition (99 %, Acros), tetrabutylammonium fluoride (99 %, Sigma–Aldrich),
of CO2 with epoxides catalyzed by Zn4@POSS-1. Reaction conditions: sub- tetrabutylammonium iodide (99 %, Sigma–Aldrich), triethylamine
strate (4.4 mmol), Zn4@POSS-1 (1 %), TBAI (4 mol %, 1 equiv rel. to Zn), CO2 (99.5 %, Aldrich), N-methylimidazole (99 %, Sigma–Aldrich), 4-dime-
(1 atm), T = 100 8C, no solvent, 4 h. thylaminopyridine (98 %, Acros), 1,3,5-trimethylbenzene (99 %,

Chem. Eur. J. 2020, 26, 13686 – 13697 www.chemeurj.org 13694 T 2020 Wiley-VCH GmbH
Full Paper
Chemistry—A European Journal doi.org/10.1002/chem.202002996

Sigma–Aldrich), zinc(II) acetate (99.9 %, Alfa Aesar), ZnEt2 (1.0 m so- UV/Vis (dichloromethane, 293 K) l (log e): 232 (6.21), 248 (6.20), 279
lution in hexanes, Aldrich) were used as received. Toluene was puri- (5.96), 383 nm (5.73 m@1 cm@1); decomposition onset temperature
fied and dried by using a solvent-purification system that con- (DTG, 10 8C min@1): 334 8C (air), 346 8C (N2); temperature of decom-
tained activated alumina. Octa(3-aminopropyl)silsesquioxane hy- position to Zn2SiO4 (determined by TGA measurement, air,
drochloride[91] and 1,2-ethanediamino-N,N’-bis(3,5-di-tert-butylsali- 10 8C min@1), residue yield: 520 8C, 24.52 % (calcd 25.85 %).
cylidene) zinc(II)[79] were prepared with a method based on a previ- Zn4@POSS-1: Method B. Under nitrogen, ZnEt2 (0.77 mL,
ously reported procedure. CO2 used for the catalytic tests was 0.7657 mmol, 4.0 equiv) was added dropwise to a solution of
technical grade and was obtained from Linde Gaz Polska. POSS-1 (0.500 g, 0.1914 mmol) in toluene (50 mL) at @78 8C. The
reaction mixture was stirred overnight at ambient temperature.
Synthesis After that, the solvent was evaporated. The resulting yellow solid
was recrystallized from chloroform at @15 8C to yield 84 % (0.461 g,
POSS-1: 3,5-Di-tert-butyl-2-hydroxybenzaldehyde (0.799 g, 0.161 mmol) Zn4@POSS-1.
3.41 mmol, 8.0 equiv) was added dropwise to a suspension of
octa(3-aminopropyl)silsesquioxane hydrochloride (0.500 g,
0.426 mmol), triethylamine (0.475 mL, 0.345 g, 3.41 mmol, Coupling reaction of epoxide and CO2
8.0 equiv), and MgSO4 (0.050 g) in a mixture of methanol (50 mL)
General procedure: A 20 mL Pyrex vial was charged with epoxide
and chloroform (10 mL). The resulting suspension was stirred for
(4.4 mmol), Zn4@POSS-1 (1 mol %), and co-catalyst (4 mol %,
24 h at room temperature. After that time, the solvent was evapo-
1 equiv rel. to Zn) and filled with CO2. The vial was then closed.
rated. The crude product was purified by washing with water (3 V
The mixture was stirred at a constant temperature during the time
20 mL), then with methanol (3 V 10 mL), and after that it was dried
of the reaction. The vial was then cooled in an ice bath. The con-
in vacuo (25 8C, 0.5 mbar) to give the expected product in 91 %
version of epoxide to cyclic carbonate was confirmed through
yield (1.010 g, 0.387 mmol) as a yellow solid. 1H NMR (500 MHz,
analysis of the post-reaction mixture by means of 1H NMR spectros-
CDCl3, 300 K): d = 13.95 (s, 8 H, OH), 8.33 (s, 8 H, CHN), 7.35 (d,
4
copy, using mesitylene as an internal standard. To determine isolat-
JHH = 2.3 Hz, 8 H, 6-Ph), 7.06 (d, 4JHH = 2.4 Hz, 8 H, 4-Ph), 3.55 (t,
3
ed yields, following the reaction the mixture was purified by
JHH = 6.5 Hz, 16 H, CH2N), 1.84–1.78 (m, 16 H, -CH2-), 1.43 (s, 72 H,
means of flash chromatography (n-hexane/EtOAc (3:1) for en-
tBu), 1.28 (s, 72 H, tBu) 0.71 ppm (t, 3JHH = 8.4 Hz, 16 H, SiCH2);
13 1
tries 1–3, 5, and 6, hexane/EtOAc (1:1) for entries 4 and 7). Cyclic
C{ H} NMR (126 MHz, CDCl3, 300 K): d = 166.0 (Ph-OH), 158.4 (C=
carbonates were characterized in accordance with literature
N), 139.9 (3-Ph), 136.7 (5-Ph), 126.8 (4-Ph), 125.9 (6-Ph), 118.0 (1-
data.[92, 93]
Ph), 62.0 (CH2N), 35.2 (CMe3), 34.2 (CMe3), 31.7 (-(CH3)3), 29.6
(-(CH3)3), 24.6 (CH2), 9.7 ppm (SiCH2); 29Si{1H} NMR (59.6 MHz, CDCl3, Yield–time plot of the reaction: A J. Young NMR tube was
300 K): d = @66.6 (s); DRIFT: ṽ = 3100–2500 (m, nO-H), 2954 (s, nC-H), charged with epoxide (0.22 mmol), Zn4@POSS-1 (1 mol %), co-cata-
2908 (m, nC-H), 2866 (m, nC-H), 1636 (s, nC=N), 1471 (m, dC-O), 1361 (m, lyst (4 mol %, 1 equiv rel. to Zn) and [D7]DMF (0.6 mL) and then it
nC-N), 1249 (m, nC-O), 1116 cm@1 (s, nSi-O-Si); HRMS (ESI + , TOF), m/z: was purged with CO2. The reaction was heated in an oil bath at a
2610.4979 [M+ +H] + (calcd 2610.4984), 1305.7700 [M+ +2 H]2 + (calcd constant temperature. The 1H NMR spectrum of the reaction was
1305.7528), 870.8397 [M+ 3+
+3 H] (calcd 870.8376); elemental analy- collected at regular time intervals using a 10 s relaxation delay. Re-
sis calcd (%) for C144H224N8O20Si8 (2612.05): C 66.21, H 8.64, N 4.29; action plots were generated by integrating the resonances of the
found: C 66.36, H 8.83, N 4.21; UV/Vis (dichloromethane, 293 K) l reactants and products for each time measurement, with the inte-
(log e): 231 (6.49), 262 (6.27), 328 nm (5.83 m@1 cm@1); decomposi- grations normalized to mesitylene as an internal standard present
tion onset temperature (DTG, 10 8C min@1): 321 8C (air), 346 8C (N2); in the reaction. The average of two resonances for each species
temperature of decomposition to SiO2 (determined by TGA mea- was used to determine the molar concentration with an estimated
surement, air, 10 8C min@1): 633 8C, residue yield: 18.23 % (calcd error of less than 5 %.
18.40 %). The TOF of epoxide conversion was estimated by using the data
Zn4@POSS-1: Method A. Zn(CH3COO)2 (0.281 g, 1.531 mmol, obtained at epoxide conversions of less than 10 %.
4.0 equiv), POSS-1 (1.0 g, 0.383 mmol), and a mixture of dichloro- Recycling experiments: For catalyst recycling experiments,
methane (20 mL) and methanol (20 mL) were stirred for 24 h at Zn4@POSS-1 was recycled by filtration, washing with hexane, and
room temperature. Subsequently, the solvent was evaporated. The then it was dried under vacuum. The recycled catalyst was then
resulting yellow solid was recrystallized from chloroform at @15 8C reused for the next run without further purification. In each subse-
to yield 60 % (0.658 g, 0.230 mmol) Zn4@POSS-1. Single crystals quent run, a fresh portion of TBAI was added.
were grown by means of a slow-evaporation solution growth tech-
nique using chloroform as a solvent. After a week, yellow, transpar-
ent, needle-like crystals were obtained. 1H NMR (500 MHz, CDCl3, Acknowledgments
300 K): d = 8.17–8.10 (m, 8 H, CHN), 7.48–7.36 (m, 8 H, 6-Ph), 6.99–
6.89 (m, 8 H, 4-Ph), 4.23–4.03 and 3.14–3.00 (m, 2 V 8 H, CH2N), 1.81– The authors acknowledge the National Science Centre, Poland
1.74 and 1.52–1.50 (m, 2 V 8 H, -CH2-), 1.40–1.27 (m, 16 V 9 H, tBu), UMO-2016/21/N/ST5/03293 (M.J.) and National Centre for Re-
0.73–0.63 and 0.52–0.39 ppm (m, 2 V 8 H, SiCH2); 13C NMR (126 MHz, search and Development (Grant TANGO1/266660/NCBR/2015)
CDCl3): d = 172.8 (C=N), 169.2 (C-O), 141.6 (3-Ph), 135.2 (5-Ph),
(S.S.) for support of this research. M.J. gratefully acknowledges
129.7 (4-Ph), 129.1 (6-Ph), 116.7 (1-Ph), 63.1 (CH2N), 35.7 (CMe3),
33.9 (CMe3), 31.6 (-(CH3)3), 29.6 (-(CH3)3), 24.6 (CH2), 8.4 ppm (SiCH2);
financial support by the Foundation for Polish Science (FNP)
29
Si{1H} CP MAS NMR (59.6 MHz): d = @66.8 (s); DRIFT: ṽ = 2952 (s, via the START stipend program.
nC-H), 2904 (m, nC-H), 2866 (m, nC-H), 1613 (s, nC=N), 1531 (m), 1434 (m,
dC-O), 1254 (s, nC-O), 1097 (s, nSi-O-Si), 539 (w, nZn-O), 517 cm@1 (w, nZn-N);
HRMS (ESI + , FT), m/z: 1433.5799 [M+ +2 H]2 + (calcd 1433.5784); ele- Conflict of interest
mental analysis calcd (%) for C144H216N8O20Si8Zn4 (2865.50): C 60.36,
H 7.60, N 3.91, Zn 9.13; found: C 60.48, H 7.54, N 3.86, Zn 9.258; The authors declare no conflict of interest.

Chem. Eur. J. 2020, 26, 13686 – 13697 www.chemeurj.org 13695 T 2020 Wiley-VCH GmbH
Full Paper
Chemistry—A European Journal doi.org/10.1002/chem.202002996

[34] S. Chu, Science 2009, 325, 1599.


Keywords: carbon dioxide · cyclic carbonates · epoxides · [35] K. Sumida, D. L. Rogow, J. A. Mason, T. M. McDonald, E. D. Bloch, Z. R.
polyhedral silsesquioxanes · zinc complexes Herm, T.-H. Bae, J. R. Long, Chem. Rev. 2012, 112, 724 – 781.
[36] C. Hepburn, E. Adlen, J. Beddington, E. A. Carter, S. Fuss, N. Mac Dowell,
J. C. Minx, P. Smith, C. K. Williams, Nature 2019, 575, 87 – 97.
[1] D. B. Cordes, P. D. Lickiss, F. Rataboul, Chem. Rev. 2010, 110, 2081 – 2173. [37] J. Lyu, X. Zhang, K. Otake, X. Wang, P. Li, Z. Li, Z. Chen, Y. Zhang, M. C.
[2] R. H. Baney, M. Itoh, A. Sakakibara, T. Suzuki, Chem. Rev. 1995, 95, 1409 – Wasson, Y. Yang, P. Bai, X. Guo, T. Islamoglu, O. K. Farha, Chem. Sci.
1430. 2019, 10, 1186 – 1192.
[3] P. D. Lickiss, F. Rataboul, Advances in Organometallic Chemistry (Eds.: [38] A. Rafiee, K. Rajab Khalilpour, D. Milani, M. Panahi, J. Environm. Chem.
A. F. Hill, M. J. Fink), Academic Press, Cambridge, 2008, pp. 1 – 116. Eng. 2018, 6, 5771 – 5794.
[4] T. Uchida, Y. Egawa, T. Adachi, N. Oguri, M. Kobayashi, T. Kudo, N. [39] D. Yu, S. P. Teong, Y. Zhang, Coord. Chem. Rev. 2015, 293 – 294, 279 – 291.
Takeda, M. Unno, R. Tanaka, Chem. Eur. J. 2019, 25, 1683 – 1686. [40] J. Klankermayer, S. Wesselbaum, K. Beydoun, W. Leitner, Angew. Chem.
[5] N. Oguri, Y. Egawa, N. Takeda, M. Unno, Angew. Chem. Int. Ed. 2016, 55, Int. Ed. 2016, 55, 7296 – 7343; Angew. Chem. 2016, 128, 7416 – 7467.
9336 – 9339; Angew. Chem. 2016, 128, 9482 – 9485. [41] Q.-W. Song, Z.-H. Zhou, L.-N. He, Green Chem. 2017, 19, 3707 – 3728.
[6] A. Bl#zquez-Moraleja, M. E. P8rez-Ojeda, J. R. Su#rez, M. L. Jimeno, J. L. [42] M. D. Burkart, N. Hazari, C. L. Tway, E. L. Zeitler, ACS Catal. 2019, 9,
Chiara, Chem. Commun. 2016, 52, 5792 – 5795. 7937 – 7956.
[7] B. Dudziec, B. Marciniec, Curr. Org. Chem. 2017, 21, 2794 – 2813. [43] J. Artz, T. E. Meller, K. Thenert, J. Kleinekorte, R. Meys, A. Sternberg, A.
[8] H. Yang, H. Liu, Microporous Mesoporous Mater. 2020, 300, 110135. Bardow, W. Leitner, Chem. Rev. 2018, 118, 434 – 504.
[9] M. Janeta, W. Bury, S. Szafert, ACS Appl. Mater. Interfaces 2018, 10, [44] X. Zhang, M. Fevre, G. O. Jones, R. M. Waymouth, Chem. Rev. 2018, 118,
19964 – 19973. 839 – 885.
[10] R. Duchateau, Chem. Rev. 2002, 102, 3525 – 3542. [45] H. Bettner, L. Longwitz, J. Steinbauer, C. Wulf, T. Werner, Top. Curr.
[11] F. J. Feher, T. L. Tajima, J. Am. Chem. Soc. 1994, 116, 2145 – 2146. Chem. 2017, 375, 50.
[12] F. J. Feher, J. F. Walzer, Inorg. Chem. 1991, 30, 1689 – 1694. [46] M. Aresta, A. Dibenedetto, A. Angelini, Chem. Rev. 2014, 114, 1709 –
[13] F. J. Feher, D. A. Newman, J. F. Walzer, J. Am. Chem. Soc. 1989, 111, 1742.
1741 – 1748. [47] V. B. Saptal, B. M. Bhanage, Curr. Opin. Green Sus. Chem. 2017, 3, 1 – 10.
[14] E. A. Quadrelli, J.-M. Basset, Coord. Chem. Rev. 2010, 254, 707 – 728. [48] C. K. Ng, R. W. Toh, T. T. Lin, H.-K. Luo, T. S. A. Hor, J. Wu, Chem. Sci. 2019,
[15] R. Duchateau, W. J. van Meerendonk, S. Huijser, B. B. P. Staal, M. A. van 10, 1549 – 1554.
Schilt, G. Gerritsen, A. Meetsma, C. E. Koning, M. F. Kemmere, J. T. F. [49] D. J. Darensbourg, W.-C. Chung, A. D. Yeung, M. Luna, Macromolecules
Keurentjes, Organometallics 2007, 26, 4204 – 4211. 2015, 48, 1679 – 1687.
[16] R. Duchateau, T. W. Dijkstra, R. A. van Santen, G. P. A. Yap, Chem. Eur. J. [50] G.-P. Wu, D. J. Darensbourg, X.-B. Lu, J. Am. Chem. Soc. 2012, 134,
2004, 10, 3979 – 3990. 17739 – 17745.
[17] R. Duchateau, R. J. Harmsen, H. C. L. Abbenhuis, R. A. van Santen, A. [51] S. A. Kuznetsova, Y. A. Rulev, V. A. Larionov, A. F. Smol’yakov, Y. V. Zubavi-
Meetsma, S. K.-H. Thiele, M. Kranenburg, Chem. Eur. J. 1999, 5, 3130 – chus, V. I. Maleev, H. Li, M. North, A. S. Saghyan, Y. N. Belokon, Chem-
3135. CatChem 2019, 11, 511 – 519.
[18] G. Gerritsen, R. Duchateau, G. P. A. Yap, Organometallics 2003, 22, 100 – [52] A. Thevenon, J. A. Garden, A. J. P. White, C. K. Williams, Inorg. Chem.
110. 2015, 54, 11906 – 11915.
[19] V. Lorenz, S. Gießmann, Y. K. Gun’ko, A. K. Fischer, J. W. Gilje, F. T. Edel- [53] S. He, F. Wang, W.-L. Tong, S.-M. Yiu, M. C. W. Chan, Chem. Commun.
mann, Angew. Chem. Int. Ed. 2004, 43, 4603 – 4606; Angew. Chem. 2004, 2016, 52, 1017 – 1020.
116, 4703 – 4706. [54] Y. Ren, J. Chen, C. Qi, H. Jiang, ChemCatChem 2015, 7, 1535 – 1538.
[20] F. T. Edelmann, S. Gießmann, A. Fischer, Chem. Commun. 2000, 2153 – [55] V. Laserna, G. Fiorani, C. J. Whiteoak, E. Martin, E. Escudero-Ad#n, A. W.
2154. Kleij, Angew. Chem. Int. Ed. 2014, 53, 10416 – 10419; Angew. Chem. 2014,
[21] F. Liu, K. D. John, B. L. Scott, R. T. Baker, K. C. Ott, W. Tumas, Angew. 126, 10584 – 10587.
Chem. Int. Ed. 2000, 39, 3127 – 3130; Angew. Chem. 2000, 112, 3257 – [56] T. S. Anderson, C. M. Kozak, Eur. Polym. J. 2019, 120, 109237.
3260. [57] X.-M. Hu, M. H. Rønne, S. U. Pedersen, T. Skrydstrup, K. Daasbjerg,
[22] P. Żak, M. Kubicki, B. Marciniec, S. Rogalski, C. Pietraszuk, D. Fra˛ckowiak, Angew. Chem. Int. Ed. 2017, 56, 6468 – 6472; Angew. Chem. 2017, 129,
Dalton Trans. 2014, 43, 7911 – 7916. 6568 – 6572.
[23] C. Di Iulio, M. D. Jones, M. F. Mahon, J. Organomet. Chem. 2012, 718, [58] S. Jayakumar, H. Li, L. Tao, C. Li, L. Liu, J. Chen, Q. Yang, ACS Sustainable
96 – 100. Chem. Eng. 2018, 6, 9237 – 9245.
[24] V. Ervithayasuporn, K. Kwanplod, J. Boonmak, S. Youngme, P. Sangtrirut- [59] J. Mart&nez, F. de la Cruz-Mart&nez, M. A. Gaona, E. Pinilla-PeÇalver, J.
nugul, J. Catal. 2015, 332, 62 – 69. Fern#ndez-Baeza, A. M. Rodr&guez, J. A. Castro-Osma, A. Otero, A. Lara-
[25] P. P. Pescarmona, A. F. Masters, J. C. van der Waal, T. Maschmeyer, J. Mol. S#nchez, Inorg. Chem. 2019, 58, 3396 – 3408.
Catal. Chem. 2004, 220, 37 – 42. [60] J. Mart&nez, J. Fern#ndez-Baeza, L. F. S#nchez-Barba, J. A. Castro-Osma,
[26] M. Nowotny, T. Maschmeyer, B. F. G. Johnson, P. Lahuerta, J. M. Thomas, A. Lara-S#nchez, A. Otero, ChemSusChem 2017, 10, 2886 – 2890.
J. E. Davies, Angew. Chem. Int. Ed. 2001, 40, 955 – 958; Angew. Chem. [61] Z. Zhao, J. Qin, C. Zhang, Y. Wang, D. Yuan, Y. Yao, Inorg. Chem. 2017,
2001, 113, 981 – 984. 56, 4568 – 4575.
[27] C. Di Iulio, M. D. Jones, M. F. Mahon, D. C. Apperley, Inorg. Chem. 2010, [62] R. Ma, L.-N. He, Y.-B. Zhou, Green Chem. 2015, 17, 226 – 231.
49, 10232 – 10234. [63] S. SopeÇa, E. Martin, E. C. Escudero-Ad#n, A. W. Kleij, ACS Catal. 2017, 7,
[28] A. Kajetanowicz, J. Czaban, G. R. Krishnan, M. Malińska, K. Woźniak, H. 3532 – 3539.
Siddique, L. G. Peeva, A. G. Livingston, K. Grela, ChemSusChem 2013, 6, [64] J. A. Castro-Osma, J. Mart&nez, F. de la Cruz-Mart&nez, M. P. Caballero, J.
182 – 192. Fern#ndez-Baeza, J. Rodr&guez-Ljpez, A. Otero, A. Lara-S#nchez, J.
[29] J. R. Severn, R. Duchateau, R. A. van Santen, D. D. Ellis, A. L. Spek, Orga- Tejeda, Catal. Sci. Technol. 2018, 8, 1981 – 1987.
nometallics 2002, 21, 4 – 6. [65] H. Bettner, J. Steinbauer, C. Wulf, M. Dindaroglu, H.-G. Schmalz, T.
[30] H.-L. Au-Yeung, S. Y.-L. Leung, A. Y.-Y. Tam, V. W.-W. Yam, J. Am. Chem. Werner, ChemSusChem 2017, 10, 1076 – 1079.
Soc. 2014, 136, 17910 – 17913. [66] M. Mastalerz, I. M. Oppel, Eur. J. Org. Chem. 2011, 5971 – 5980.
[31] S. Li, Z.-Y. Wang, G.-G. Gao, B. Li, P. Luo, Y.-J. Kong, H. Liu, S.-Q. Zang, [67] J. Mel8ndez, M. North, R. Pasquale, Eur. J. Inorg. Chem. 2007, 3323 –
Angew. Chem. Int. Ed. 2018, 57, 12775 – 12779; Angew. Chem. 2018, 130, 3326.
12957 – 12961. [68] M. North, R. Pasquale, Angew. Chem. Int. Ed. 2009, 48, 2946 – 2948;
[32] J. Sun, Y. Chen, L. Zhao, Y. Chen, D. Qi, K.-M. Choi, D.-S. Shin, J. Jiang, Angew. Chem. 2009, 121, 2990 – 2992.
Chem. Eur. J. 2013, 19, 12613 – 12618. [69] J. A. Castro-Osma, M. North, X. Wu, Chem. Eur. J. 2016, 22, 2100 – 2107.
[33] C. R. Groom, I. J. Bruno, M. P. Lightfoot, S. C. Ward, Acta Crystallogr. Sect. [70] X. Wu, M. North, ChemSusChem 2017, 10, 74 – 78.
B 2016, 72, 171 – 179. [71] M. North, P. Villuendas, C. Young, Chem. Eur. J. 2009, 15, 11454 – 11457.

Chem. Eur. J. 2020, 26, 13686 – 13697 www.chemeurj.org 13696 T 2020 Wiley-VCH GmbH
Full Paper
Chemistry—A European Journal doi.org/10.1002/chem.202002996

[72] A. Decortes, M. M. Belmonte, J. Benet-Buchholz, A. W. Kleij, Chem. [83] S. Bhunia, R. A. Molla, V. Kumari, Sk. M. Islam, A. Bhaumik, Chem.
Commun. 2010, 46, 4580 – 4582. Commun. 2015, 51, 15732 – 15735.
[73] M. Janeta, Ł. John, J. Ejfler, T. Lis, S. Szafert, Dalton Trans. 2016, 45, [84] W. Clegg, R. W. Harrington, M. North, R. Pasquale, Chem. Eur. J. 2010, 16,
12312 – 12321. 6828 – 6843.
[74] The values were derived from the crystallographic data of molecules [85] F. Jutz, J.-D. Grunwaldt, A. Baiker, J. Mol. Catal. A: Chem. 2008, 279, 94 –
deposited in the Cambridge Structural Database. 103.
[75] M. Janeta, S. Szafert, J. Organomet. Chem. 2017, 847, 173 – 183. [86] C. K. Karan, M. Bhattacharjee, Inorg. Chem. 2018, 57, 4649 – 4656.
[76] J. Zhao, F. Dang, B. Liu, Y. Wu, X. Yang, G. Zhou, Z. Wu, W.-Y. Wong, [87] R. L. Paddock, S. T. Nguyen, Chem. Commun. 2004, 1622 – 1623.
Dalton Trans. 2017, 46, 6098 – 6110. [88] G.-P. Wu, S.-H. Wei, W.-M. Ren, X.-B. Lu, T.-Q. Xu, D. J. Darensbourg, J.
[77] F. Castro-Gjmez, G. Salassa, A. W. Kleij, C. Bo, Chem. Eur. J. 2013, 19, Am. Chem. Soc. 2011, 133, 15191 – 15199.
6289 – 6298. [89] J. A. Castro-Osma, K. J. Lamb, M. North, ACS Catal. 2016, 6, 5012 – 5025.
[78] 1H NMR analysis of post-reaction mixtures showed the presence of trib- [90] R. L. Paddock, S. T. Nguyen, J. Am. Chem. Soc. 2001, 123, 11498 – 11499.
utylamine. The ability to form tributylamine in the presence of [91] M. Janeta, Ł. John, J. Ejfler, S. Szafert, Chem. Eur. J. 2014, 20, 15966 –
Zn4@POSS-1 was confirmed by an experiment in which TBAI and a stoi- 15974.
chiometric amount of Zn4@POSS-1 in styrene oxide were heated at [92] C. Maeda, S. Sasaki, K. Takaishi, T. Ema, Catal. Sci. Technol. 2018, 8,
100 8C for 4 h. When TBAI was heated in the absence of Zn4@POSS-1, 4193 – 4198.
no tributylamine formation was observed. [93] L. Qin, B. Wang, Y. Zhang, L. Chen, G. Gao, Chem. Commun. 2017, 53,
[79] G. A. Morris, H. Zhou, C. L. Stern, S. T. Nguyen, Inorg. Chem. 2001, 40, 3785 – 3788.
3222 – 3227.
[80] S. Subramanian, J. Park, J. Byun, Y. Jung, C. T. Yavuz, ACS Appl. Mater. In-
terfaces 2018, 10, 9478 – 9484. Manuscript received: June 22, 2020
[81] X.-B. Lu, D. J. Darensbourg, Chem. Soc. Rev. 2012, 41, 1462 – 1484. Revised manuscript received: July 22, 2020
[82] E. Mercad8, E. Zangrando, C. Claver, C. Godard, ChemCatChem 2016, 8, Accepted manuscript online: July 24, 2020
234 – 243. Version of record online: September 24, 2020

Chem. Eur. J. 2020, 26, 13686 – 13697 www.chemeurj.org 13697 T 2020 Wiley-VCH GmbH

You might also like