Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

DOI: 10.1002/ejic.

201700871 Full Paper

Metal–Organic Frameworks
A Ni(salen)-Based Metal–Organic Framework: Synthesis,
Structure, and Catalytic Performance for CO2 Cycloaddition with
Epoxides
Yamei Fan,[a] Jiawei Li,[a] Yanwei Ren*[a] and Huanfeng Jiang*[a]

Abstract: A three-dimensional (3D) chiral metal–organic ing a 2D lamellar structure, which are further linked by Ni(salen)
framework [Cd2{Ni(salen)}(DMF)3]·4DMF·7H2O (1) based on a into the 3D network with a 1D open channel (ca. 7.0 × 8.0 Å2)
new enantiopure tetracarboxyl-functionalized metallosalen along the a axis. On account of its porosity, Lewis acid sites,
Ni(H4salen) {where H6salen is (R,R)-N,N′-bis[3-tert-butyl-5-(3,5-di- and moderate uptake for CO2, 1 can be used as an efficient
carboxybenzyl)salicylidene]-1,2-diphenylethylenediame} was heterogeneous catalyst for the CO2 cycloaddition with epoxides
synthesized and characterized by infrared spectroscopy, ther- under relatively mild conditions. Moreover, the bulky epoxide
mogravimetric analysis, nitrogen and carbon dioxide adsorp- shows a decrease in activity with an increase in the alkyl chain
tion, and powder and single-crystal X-ray diffractions. In 1, the length of the substrate as a result of the confinement effect
dinuclear Cd2 cluster [Cd2(COO)4(DMF)3] as a node is cross- of 1, showing size-dependent selectivity.
linked by four isophthalate groups on the salen ligands, form-

Introduction an immense challenge to prepare more MOF catalysts for CO2


chemical fixation under mild conditions, including low tempera-
Metal–organic frameworks (MOFs) are exhibiting growing appli-
ture and atmospheric pressure.
cations in the field of heterogeneous catalysis due to their high
content of metal centers, tunable chemical functionality, flexi- Recently, functionalized metallosalen ligands have been
ble pore size, and high surface area.[1] Among these applica- commonly used in the construction of MOF catalysts because
tions, MOF-catalyzed conversion of carbon dioxide (CO2) into they have efficient catalytic performance in many organic trans-
cyclic carbonates has attracted extensive attention,[2] mainly formations. In addition, salen ligands also facilitate the intro-
based on the following virtues: (1) MOFs possess a large num- duction of various chiral groups on their structures. To date,
ber of catalytic sites from both metal nodes and organic li- many metallosalen-based MOFs have been reported such as
gands; the coordinatively unsaturated metal sites situated on Zn(salen),[3] Cu(salen),[4] Ni(salen),[5] Co(salen),[6] Mn(salen),[7]
inorganic clusters or organometallic linkers (e.g., metallosalens Fe(salen),[8] Cr(salen),[9] and VO(salen).[10] In this regard, porous
and metalloporphyrins) always serve as Lewis acids; (2) the pore Ni(salen)-based MOFs have received specific attention, and our
size (channels and cavities) and chemical environment can be recent studies[5b–5d] showed that these MOFs could be used as
regulated by the framework components, resulting in a size- or efficient and recyclable heterogeneous catalysts for the synthe-
shape-dependent selective catalyst; (3) MOFs easily separate sis of cyclic carbonates from the CO2 cycloaddition with epox-
from the reaction system, allowing for simple recycling; and (4) ides utilizing the Ni(salen) unit as a Lewis acid catalytic site.
most importantly, MOFs have displayed high CO2 adsorption Moreover, these MOF catalysts feature high local density of
uptake which can enhance the local concentration of CO2 Ni(salen) units and capacity for CO2 adsorption, exhibiting im-
around the catalytic active centers inside the pores of the proved catalytic performance than that of their corresponding
framework to improve catalytic efficiency. Although these vir- homogeneous catalyst.
tues have been utilized to design and synthesize efficient MOF Aiming at further exploring this field, herein we designed a
catalysts for using CO2 as a carbon building block, there is still new tetracarboxyl-functionalized chiral salen ligand, (R,R)-N,N′-
bis[3-tert-butyl-5-(3,5-dicarboxybenzyl)salicylidene]-1,2-diphen-
[a] Key Laboratory of Functional Molecular Engineering of Guangdong ylethylenediame (denoted as H6salen, Scheme 1), based on the
Province, School of Chemistry and Chemical Engineering, South China
University of Technology,
following considerations: (1) a high number of coordinating
Guangzhou 510640, P. R. China sites on the salen linker can improve structural stability and
E-mail: renyw@scut.edu.cn thus high reusability in the CO2–epoxide coupling reaction; (2)
jianghf@scut.edu.cn the introduction of a diphenyl group can enhance the hydro-
http://www.fmegd.com/col.jsp?id=164
http://www.fmegd.com/col.jsp?id=145
phobicity and thereby enhance the water-resistance stability of
Supporting information and ORCID(s) from the author(s) for this article are MOFs; (3) the diphenyl group may also facilitate the growth of
available on the WWW under https://doi.org/10.1002/ejic.201700871. high-quality MOF crystals by virtue of the space-filling effect

Eur. J. Inorg. Chem. 2017, 4982–4989 4982 © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper

Scheme 1. Synthesis of the metallosalen Ni(H4salen).

and π–π stacking; and (4) the potential chiral source of the In order to optimize the conditions for the synthesis of the
salen linker can be introduced into MOFs for the asymmetric MOF, we carried out numerous parallel experiments by adjust-
cycloaddition of CO2 with epoxides. In this work, a 3D MOF, ing conditions such as the type of solvent, ratio of reaction
[Cd2{Ni(salen)}(DMF)3]·4DMF·7H2O (1), was successfully con- materials, and temperature. Maximum yields and good-quality
structed from metallosalen Ni(H4salen) and was structurally crystals of 1 could only be obtained using the synthetic condi-
characterized by single-crystal X-ray diffraction and other tions as listed in the Experimental Section. Compound 1 is
physicochemical methods. The catalytic performance of this stable in air and is insoluble in water and common organic
MOF as a heterogeneous catalyst for the coupling reaction of solvents. The solvent contents were established by a combina-
CO2 with epoxides in the presence of tetrabutylammonium tion of elemental analysis, thermogravimetric analysis (TGA),
bromide (Bu4NBr) was also investigated. and the electron count from SQUEEZE (see the CIF of 1). TGA
revealed that the guest water and DMF molecules were readily

Results and Discussion


Synthesis

The synthetic route to the H6salen ligand is illustrated in


Scheme 1. H6salen was obtained as yellow powder by Schiff
base condensation between (R,R)-1,2-diphenylethylenediamine
and 3-tert-butyl-5-(3,5-dicarboxybenzyl)salicylaldehyde that
was prepared by successive Pd-catalyzed Suzuki–Miyaura cou-
pling and base-promoted hydrolysis reactions. The metalation
of H6salen with Ni(OAc)2·4H2O afforded the metallosalen
Ni(H4salen). The infrared spectrum (Figure S2 in Supporting In-
formation) of Ni(H4salen) shows that the C=N stretching vibra-
tion shifted to about 1669.6 cm–1, as expected for the nickel
ion coordinated to H6salen. The UV/Vis spectrum of Ni(H4salen)
(Figure S3 in Supporting Information) presents one broad peak
in the visible region at 400–500 nm, attributed to the d–d tran-
sition of the square-planar four-coordinate Ni2+ ion, which is
the reason that Ni(H4salen) has a brown-red color.[11] This result
also indicates that the carboxybenzyl groups of H6salen are not Figure 1. PXRD patterns of 1. (a) Simulated from the CIF, (b) as-synthesized
involved in the coordination to the Ni2+ ion, and H6salen can (experimental) sample of 1, (c) sample of activated 1, (d) recovered 1 after
be utilized as a metalloligand to construct MOFs. the first catalysis, and (e) recovered 1 after the fifth catalysis.

Eur. J. Inorg. Chem. 2017, 4982–4989 www.eurjic.org 4983 © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper

removed in the temperature range 25–200 °C, and the frame- earlier reported for analogous systems,[5] each Ni2+ ion in the
work was stable up to 300 °C (Figure S12 in Supporting Infor- Ni(salen) ligands is coordinated in a nearly square-planar geom-
mation). The phase purity of the bulk sample was evidenced by etry with two nitrogen atoms and two oxygen atoms from the
experimental powder X-ray diffraction (PXRD), which matches salen ligand (Ni–Oavg = 1.842 Å, Ni–Navg = 1.828 Å) (Figure 2b).
the theoretical pattern simulated from the crystal data (Fig- The carboxylate groups in the Ni(salen) unit exhibit three types
ure 1), despite some peaks of the experimental PXRD being of coordination modes including one bis-bidentate chelating
weaker in intensity. In addition, the PXRD of activated 1, which mode, one bis-tridentate chelating-bridging mode, and two bi-
was obtained by heating 1 at 120 °C for four hours under vac- dentate bridging modes. In the linking node of a dinuclear Cd2
uum to remove the solvent molecules, showed a diffraction pat- cluster [Cd2(COO)4(DMF)3] (Figure 2a), Cd1 is coordinated by six
tern similar to that of the pristine sample, indicating no obvious oxygen atoms from four carboxylate groups with a distorted
structural change. octahedral geometry, while Cd2 adopts a quasi-octahedral ge-
ometry composed of six oxygen atoms from three carboxylate
groups and three coordinated DMF molecules. The Cd–O-
Crystal Structure (carboxylate) bond lengths range from 2.196 to 2.484 Å, and
the Cd–O(DMF) bond lengths are 2.229–2.285 Å, all of which
A single-crystal X-ray diffraction study shows that 1 crystallizes
are within the range of those observed for other Cd-based
in the chiral space group P212121 with absolute structure pa-
MOFs with oxygen-donor ligands.[5,12] The dinuclear Cd2 clus-
rameters of –0.001(5). The asymmetric unit contains one
ters are cross-linked by four isophthalate groups on salen li-
Ni(salen) unit, two Cd2+ ions, and three DMF molecules. As was
gands, forming a 2D lamellar structure (Figure 2c), which are
further linked by Ni(salen) into the 3D network (Figure 2d).
Compound 1 has open channels along the a axis with channel
cross-sections of about 7.0 × 8.0 Å2, which are filled with DMF
and water molecules. It should be noted that all Ni(salen) moie-
ties lie inside the 1D channels, and the empty coordination sites
of Ni2+ ions are oriented to the channels. Therefore, the trapped
substrate molecules are available for Lewis acid activation in
the 1D channels whose surfaces are uniformly arranged with
chiral Ni(salen) moieties with coordinatively unsaturated Ni2+
ions. The PLATON calculation reveals that 1 has about 42 % of
the total volume available for guest inclusion.

Gas Sorption

To verify the porosity of 1, the sorption isotherms of N2 and


CO2 were measured at different temperatures. The sample was
exchanged in methanol overnight and then activated under
vacuum at 120 °C for four hours, affording the desolvated 1. As
can be seen from Figure 3(top), the N2 adsorption isotherm of
1 reveals Type I behavior, indicative of a microporous material.
The N2 uptake at 77 K is 202 cm3 g–1. The pore-size distribution
obtained from the isotherm indicates that there is only one
type of pore with a diameter of about 7.9 Å in 1, which matches
well with the crystal structure analysis. The Brunauer–Emmett–
Teller (BET) and Langmuir surface areas of the sample were
calculated as 418.1 and 730.9 m2 g–1, respectively. The total
pore volume of 1 obtained from N2 isotherms is about
0.33 cm3 g–1. The low-pressure CO2 adsorption was also meas-
ured because unsaturated Ni2+ ions have high affinity for CO2
molecules.[2q] The results show that the maximal adsorbed
amounts at 1 atm are 32.0 cm3 g–1 at 273 K and 21.0 cm3 g–1
at 298 K [Figure 3(bottom)]. These values are comparable to
those at 273 K for previously reported Ni(salen)-based MOFs[5]
and lower than that of a discrete single-walled Ni-based metal–
organic nanotube Ni-TCPE1.[2i]
Figure 2. (a) The coordination environment of the dinuclear Cd2 cluster. (b)
View of the coordination mode of the Ni(salen) ligand. (c) Representation of
To further explore the affinity of 1 for CO2, the CO2 adsorp-
the 2D lamellar structure. (d) 3D network along the crystallographic a axis tion enthalpy (Qst) was calculated according to the Clausius–
(dinuclear Cd2 clusters are presented by polyhedrons). Clapeyron equation from the sorption isotherms at 273 and

Eur. J. Inorg. Chem. 2017, 4982–4989 www.eurjic.org 4984 © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper

CO2 cycloaddition with epoxides, tetrabutylammonium halides


(Bu4NX; X = Cl, Br, I) have been commonly employed as a Lewis
base to promote this reaction. Although tetrabutylammonium
iodide is the best promoter in accordance with its increased
nucleophilicity, in the presence of this microporous MOF, the
diffusion of the large iodide ion might be hampered,[2s,5c] and
thus we adopted Bu4NBr as co-catalyst in this work.
First, butylene oxide (BO) was selected as the benchmark
substrate, and the mildest reaction conditions involving 0.1 MPa
CO2 and room temperature were initially investigated for this
reaction using 1 (0.5 mol-% based on Ni) and Bu4NBr (0.5 mol-
%). However, to our disappointment, the yield of butylene carb-
onate (BC) was only 12 % (entry 4, Table 1) after 12 hours, and
no enantiomeric excess value was observed despite a chiral cat-
alyst being used. Similar experimental phenomenon was ob-
served for a previously reported chiral Ni(salen)-based MOF.[5d]
Although low yield was obtained, several control experiments
(entries 1–3, Table 1) indicate that both 1 and Bu4NBr play criti-
cal roles and are indispensable to this reaction, which indicate
that the Lewis base Br– from Bu4NBr and Lewis acid sites in 1
work together to enhance the catalytic activity. Next, the influ-
ence of reaction temperature on the yield of BC was studied,
and the results reveal that as the temperature increased, the
yield of BC improved, reaching 54 % at 80 °C. It has been re-
ported that the partition behavior of the substrates between
phases can affect the reaction rate of a biphasic reaction sys-
tem, and high CO2 pressure can improve the concentration of
CO2 in the liquid phase.[2a] Thus, we raised the reaction pressure
to 1 MPa, and a high yield of BC with no by-products was
achieved (entry 7, Table 1). Finally, in a typical experiment, the
optimized reaction conditions involve epoxide (10 mmol),
1 MPa of CO2 in the presence of 0.5 mol-% of 1, and 0.5 mol-
Figure 3. N2 adsorption/desorption isotherms of 1 at 77 K and its pore-size % of Bu4NBr at 80 °C under a solvent-free environment. As a
distribution (top), and CO2 sorption isotherms of 1 at 273 K and 298 K (bot- comparison, the catalytic activity of 1 was significantly higher
tom). than that of the corresponding homogeneous Ni(H4salen) [en-
try 8, Table 1, the molar amount of Ni(H4salen) was the same
298 K. At an initial coverage, Qst exhibits a maximum of as for 1], clearly indicating that a synergistic Lewis acidic activa-
39.6 kJ mol–1, which is superior to those of MOFs decorated tion of Ni(salen) units in 1 works during the catalytic process.
with Lewis basic sites, for example, PCN-88 (27 kJ mol–1), PCN- A careful examination of the crystal structure of 1 showed that
16 (22.5 kJ mol–1), and MOF-74 (30 kJ mol–1), but lower than
those of some high-polarity-functionalized MOFs, such as MIL-
Table 1. Optimization of reaction parameters for CO2 cycloaddition with BO
100 (60 kJ mol–1) and CuBTTri-en (80 kJ mol–1).[13] The relatively catalyzed by 1.[a]
high enthalpy of CO2 is probably mainly related to the Ni(salen)
Entry Catalyst Co-catalyst Pressure Temp. Yield[b]
units (Lewis acid) and the coordinated DMF molecules (Lewis
(MPa) (°C) (%)
base) located within the wall of the open channel, which facili-
1 – – 0.1 25 0
tate the framework with high polarity and then cause strong
2 – Bu4NBr 0.1 25 <5
interactions with CO2. 3 1 – 0.1 25 0
4 1 Bu4NBr 0.1 25 12
5 1 Bu4NBr 0.1 50 31
Cycloaddition Reaction of CO2 with Epoxides Catalyzed 6 1 Bu4NBr 0.1 80 54
7 (1st) 1 Bu4NBr 1 80 99
by 1
8 Ni(H4salen) Bu4NBr 1 80 65
9 (2nd) 1 Bu4NBr 1 80 94
As mentioned above, our catalytic experiments were focused
10 (3rd) 1 Bu4NBr 1 80 90
on the CO2 cycloaddition with epoxides. Prior to catalytic reac- 11 (4th) 1 Bu4NBr 1 80 87
tion, 1 was activated at 120 °C under vacuum for four hours to 12 (5th) 1 Bu4NBr 1 80 85
remove guest DMF and water molecules, generating the active [a] Other reaction conditions: BO (10 mmol), 1 (0.5 mol-%), Bu4NBr (0.5 mol-
catalyst. PXRD proved that the activated sample still retained %), and 12 h reaction time. [b] Yield and selectivity were determined by GC–
the original framework structure (Figure 1). Additionally, for the MS, and in all cases, the selectivity of the product is 99 %.

Eur. J. Inorg. Chem. 2017, 4982–4989 www.eurjic.org 4985 © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper

the empty coordination site of the Ni2+ ions of the Ni(salen) phenomenon has commonly been observed in porous MOF ca-
units are all oriented to the 1D channel with a vertically face-to- talysis,[1,5b–5e] and it can be attributed to the confinement effect
face distance of about 8.0 Å, an ideal distance for the synergistic of the MOF channel (or cavity) that restricts the diffusion rate
activation of epoxide by two Ni2+ ions. Therefore, according to of larger substrates and products through the channel. These
the above experimental results, we propose that the plausible results further suggest that the catalytic reaction took place in
mechanism for the CO2 cycloaddition with BO catalyzed by 1 the channel, not on the catalyst surface. Indeed, upon extend-
is similar to that of previously reported catalytic systems with ing the reaction time to 24 hours, the yield of cyclic carbonates
Ni(salen)-based MOFs and Bu4NBr.[5b–5d,2o] In addition, we con- with longer alkyl chains reached 99 %. To our delight, the inter-
sider that the moderate CO2 uptake of this MOF is also responsi- nal epoxide, cyclohexene oxide, which is known to poorly un-
ble for the high reactivity as a result of the distinct difference dergo cycloaddition with CO2 because of the high steric hin-
of yields versus time for BO and CO2 catalyzed by 1 and homo- drance,[14] was converted into the corresponding cyclic carb-
geneous Ni(H4salen), respectively, under identical reaction con- onate with a moderate yield of 68 % (entry 7, Table 2), further
ditions (Figure S16 in Supporting Information). demonstrating the high catalytic performance of this MOF cata-
Reusability is an essential feature of any catalyst considered lyst.
for use in industrial applications. Thus, the reusability of 1 was Table 2. The synthesis of various cyclic carbonates catalyzed by 1.[a]
examined, and the results reveal a small decrease in the activity
of this MOF catalyst after being used five times (entries 7 and
9–12, Table 1). The main reason for the decrease is partly the
inevitable tiny loss of catalyst during the recovery process. Fur-
thermore, the catalyst recovered from the catalytic reaction ex-
hibited almost the same PXRD pattern as the as-synthesized 1
(Figure 1), unambiguously supporting the stability of the MOF
structure during the catalytic reactions. Another major issue for
heterogeneous catalysts is the possibility that some active sites
can migrate from the support into the liquid phase during the
reaction and that these leached species are actually responsible
for a significant part of the observed catalytic activity. With this
in mind, we conducted leaching measurements to determine
whether the Ni2+ or Cd2+ ions in 1 could leach into the liquid
phase during the course of the reaction. After a standard reac-
tion had taken place after six hours with a yield of about 60 %
of BC, the mother liquor was split into two fractions, in which
one fraction had the MOF catalyst while the other fraction was
without the MOF catalyst. After an additional six hours, the
yields of the two fractions were 92 % (with the MOF) and 60 %
(without the MOF). Additionally, inductively coupled plasma
atomic mass spectrometry (ICP-AMS) analysis of the filtrate of
the reaction mixture revealed negligible Ni2+ and Cd2+ leaching
(ca. 0.002 % for Ni2+ and ca. 0.003 % for Cd2+). These results
undoubtedly indicate the heterogeneity of 1 and no active spe-
cies leached into the reaction system during the above catalytic
condition. Furthermore, the recovered MOF catalyst still main-
tained a brown-red color, which also evidences that the Ni2+
coordination sphere in 1 was not destroyed.
Having established that 1 is a highly active catalyst for BC
formation, various epoxides were used to explore the applica-
bility of 1. As shown in Table 2, it is clear that electron-with- [a] Reaction conditions: epoxide (10 mmol), 1 (0.5 mol-%), Bu4NBr (0.5 mol-
%), CO2 (1 MPa), 80 °C, 12 h. [b] Determined by GC–MS (the selectivities of
drawing and electron-donating or aliphatic and aromatic termi-
products are 99 %). [c] TON: mol of cyclic carbonate per mol of catalyst used.
nal epoxides could be converted into the corresponding cyclic
carbonates in good yields. In all experiments, cyclic carbonates A literature survey[2,5] revealed that different types of epox-
were the sole products. As was reported in other literature, the ides and reaction conditions (such as catalyst loading, CO2 pres-
very good yield of epichlorohydrin (entry 3, Table 2) can be sure, reaction temperature, and time) were adopted, and there-
explained by the electron-withdrawing substituents,[14] which fore it is hard to critically compare the efficacy of reported MOF
facilitate nucleophilic attack of Br– during the ring opening of catalysts. Thus, we take styrene oxide as a substrate to compare
epoxide. It is worthwhile to note that the yield of cyclic carbon- the catalytic performance of reported MOFs with the Ni2+ ion
ates gradually decreased with the increase in alkyl chain length as a Lewis acid site (entries 1–6, Table 3). Furthermore, we also
of epoxides, exhibiting size-dependent selectivity. This summarize the catalytic performance of reported typical MOFs

Eur. J. Inorg. Chem. 2017, 4982–4989 www.eurjic.org 4986 © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper

Table 3. Cycloaddition reaction of CO2 with styrene oxide (or propylene oxide) catalyzed by different MOFs.

with different valence metal ions (divalent, trivalent, and tetra- ter. Infrared (IR) spectra were measured by using KBr pellets
valent ions) as Lewis acid sites for the CO2 cycloaddition with with a Nicolet Nexus 470 FTIR spectrometer in the range of
propylene oxide (entries 7–12, Table 3), since it is generally be- 4000–400 cm–1. Gas chromatography–mass spectroscopy (GC–MS)
lieved that higher valence of metal ions corresponds with was conducted by using a Shimadzu GC–MS-QP5050A system that
was equipped with a 0.25 mm × 30 m DB-WAX capillary column.
stronger Lewis acidity. As can be seen from Table 3, the catalytic
Gas chromatography (GC) was conducted by using an Agilent
performance observed in this work is comparable to that of Technologies7820A GC system that was equipped with a
some MOFs reported before. 0.25 mm × 30 m Beta DEX chiral column. The content of metal ions
was determined with an Agilent 7700 equipment by inductively
coupled plasma atomic mass spectroscopy (ICP-AMS). The gas ad-
sorption measurements were performed by using a MicroActive
Conclusions ASAP2460 system under N2 (273 K) or CO2 (273 and 298 K).
In summary, we report the synthesis and structural characteriza- Synthesis of the Metallosalen
tion of a chiral 3D porous MOF using a tetracarboxyl-functional-
3-tert-Butyl-5-(3,5-dimethoxycarbonyl)phenylsalicylaldehyde:
ized metallosalen ligand. Based on the Ni(salen) Lewis acid site, (3,5-Dimethoxycarbonyl)phenylboronic acid (1.309 g, 5.5 mmol), 2 M
open channel, and moderate uptake for CO2, this MOF can act NaHCO3 aqueous solution (5.5 mL, 11 mmol), and Pd(PPh3)4 (0.35 g,
as an efficient catalyst for the heterogeneous cycloaddition of 0.3 mmol) were successively added to a solution of 5-bromo-3-tert-
CO2 with epoxides, resulting in cyclic carbonates with high butyl-salicylaldehyde (1.285 g, 5 mmol) in dimethoxyethane (DME,
yields and size-dependent selectivity. The catalytic activity of 100 mL). The resulting solution was heated to reflux for 24 h under
this MOF is higher than that of the corresponding homogene- nitrogen. The reaction mixture was diluted with H2O (60 mL) and
ous species, demonstrating the merits of MOF catalysts for CO2 extracted with EtOAc (60 mL × 3). The organic phase was washed
conversion reactions. Moreover, this MOF can be facilely sepa- with brine and dried with anhydrous MgSO4. The organic solvent
was evaporated in vacuo, and the residue was then purified by
rated and reused for five successive cycles without significant
column chromatography (eluent: Hex/EtOAc = 25:1) to give the de-
loss of activity and structural integrity. Although this chiral MOF
sired product. Yield: 1.2 g, (64.86 %). 1H NMR (CDCl3): δ = 11.87 (s,
was not suitable for the enantioselectivity cycloaddition of CO2 1 H), 9.99 (s, 1 H), 8.65 (s, 1 H), 8.41 (d, 2 H), 7.78 (d, 1 H), 7.68 (d, 1
with epoxides, we envision that it might have potential value H), 3.99 (s, 6 H), 1.49 (s, 9 H) ppm. 13C NMR (CDCl3): δ = 197.11,
in other asymmetric catalytic reactions. Relative exploration of 166.18, 161.28, 140.96, 139.39, 132.83, 131.80, 131.34, 130.35,
the enantioselective activity of this MOF is ongoing in our labo- 129.14, 120.82, 52.53, 35.13, 29.23 ppm.
ratory.
3-tert-Butyl-5-(3,5-dicarboxybenzyl)salicylaldehyde: 3-tert-Butyl-
5-(3,5-dimethoxycarbonyl)phenylsalicylaldehyde (1.11 g, 3 mmol)
was dissolved in a solution of LiOH (0.63 g, 15 mmol), THF (15 mL),
Experimental Section and H2O (15 mL). The mixture was stirred at room temp. for 24 h
in the dark. After removing THF in vacuo, the residue was diluted
Materials and Methods: All of the reagents were commercially with H2O (60 mL) and washed with CH3Cl (60 mL × 3). The aqueous
available and were used without further purification. PXRD patterns phase was acidified (pH = 1) with 3 M HCl (5 mL). The precipitate
were collected with a Bruker D8 powder diffractometer at 40 kV, was extracted with EtOAc (60 mL × 3), and then the organic phase
40 mA with Cu-Kα radiation (λ = 1.5406 Å), with a scan speed of was washed with water and dried with anhydrous MgSO4. The or-
45.65 s per step and a step size of 0.0131303(2θ). Thermogravimet- ganic solvent was evaporated in vacuo to afford a pale-yellow solid.
ric analyses (TGA) were performed with a Q600SDT instrument un- Yield: 1.005 g (98 %). 1H NMR ([D6]dmso): δ = 13.40 (s, 2 H), 11.94
der a flow of nitrogen at a heating rate of 10 °C min–1. NMR spectro- (s, 1 H), 10.13 (s, 1 H), 8.45 (s, 1 H), 8.40 (s, 2 H), 8.13 (s, 1 H), 7.83
scopy was performed with a Bruker AM-400 (400 MHz) spectrome- (s, 1 H), 1.46 (s, 9 H) ppm. 13C NMR ([D6]dmso): δ = 199.63, 166.99,

Eur. J. Inorg. Chem. 2017, 4982–4989 www.eurjic.org 4987 © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper
160.44, 140.64, 138.55, 132.59, 132.53, 131.46, 131.36, 130.05, Table 4. Crystallographic data and structure refinement for 1.
128.98, 121.54, 35.16, 29.46 ppm.
1
(R,R)-N,N′-Bis[3-tert-butyl-5-(3,5-dicarboxybenzyl)salicylidene]- Empirical formula C61H63Cd2N5NiO13
1,2-diphenylethylenediame (H6salen): A solution of (R,R)-1,2-di- Formula weight 1357.67
phenylethylenediamine (0.129 g, 0.608 mmol) in THF (20 mL) was T (K) 100(2)
added dropwise to a solution of 3-tert-butyl-5-(3,5-dicarboxy- Wavelength (Å) 1.54178
benzyl)salicylaldehyde(0.45 g, 1.216 mmol) in THF (50 mL). The reac- Crystal system orthorhombic
tion mixture was stirred at room temp. for 24 h. THF was removed Space group P212121
by evaporation in vacuo, and the precipitate was sonicated with a (Å) 12.0167(5)
Et2O, collected by filtration, and dried in air. A light-brown product b (Å) 15.5049(6)
c (Å) 43.5389(15)
was obtained. Yield: 0.5 g (95.6 %). 1H NMR ([D6]dmso): δ = 14.45
α (°) 90
(s, 1 H), 8.76 (s, 1 H), 8.45 (s, 1 H), 8.29 (s, 1 H), 7.58 (s, 1 H), 7.53 β (°) 90
(s, 1 H), 7.43 (d, 1 H), 7.33 (t, 1 H), 7.24 (t, 1 H), 5.19 (s, 1 H), γ (°) 90
1.41 (s, 1 H) ppm. 13C NMR ([D6]dmso): δ = 167.73, 166.73, 160.18, V (Å3) 8112.1(5)
140.84, 139.45, 137.41, 132.23, 130.59, 128.70, 128.51, 128.25, Z 4
128.13, 127.84, 127.65, 118.71, 77.38, 34.65, 30.49, 29.05 ppm. IR: ρ (g cm–3) 1.112
ν̃ = 2953.9, 1708.5, 1620.3, 1429.8, 1269.8, 1095.7, 889.5, 769.1, μ (mm–1) 4.837
699.9, 516.7 cm–1. F(000) 2768
θ range data collection 3.022–66.253
Ni(H4salen): A solution of Ni(OAc)2·4H2O (0.083 g, 0.332 mmol) in Reflections collected 12860
MeOH (30 mL) was added dropwise to a solution of the above Unique reflections 11876
H6salen ligand (0.264 g, 0.302 mmol) in MeOH (30 mL). The reaction Data/restraints/parameters 12860/21/739
mixture was stirred at room temp. for 2 h and then heated to reflux GOF on F2 1.030
for 3 h. The red-powder precipitate was collected by centrifugation, R1, wR2 [I > 2σ(I)] 0.0493, 0.1301
R1, wR2 (all data) 0.0530,0.1335
washed with MeOH, and dried under reduced pressure. Yield: 0.26 g
(93.9 %). IR: ν̃ = 3521.5, 2934.3, 1669.6, 1499.1, 1393.3, 1255.3,
1095.5, 661.0 cm–1. ESI-MS: m/z (%) = 940.2428 [Ni(H4salen) + Na]+.
CCDC 1540442 (for 1) contains the supplementary crystallographic
Synthesis of [Cd2{Ni(salen)}(DMF)3]·4DMF·7H2O (1): Ni(H4salen) data for this paper. These data can be obtained free of charge from
(10 mg, 0.01 mmol) and CdCl2 (20 mg, 0.1 mmol) were dissolved in The Cambridge Crystallographic Data Centre.
DMF (3 mL) and H2O (1 mL) in a 5-dramscrew-capped vial. The vial
was heated at 80 °C for 4 d. Brown, blocklike crystals of 1 were
filtered, washed with EtOH and Et2O, and dried at room temp. Yield: Acknowledgments
12 mg (68 %). IR: ν̃ = 2940.4, 1652.6, 1543.8, 1384.9, 1256.9, 1170.4, We are grateful to the National Key Research and Development
1092.6, 1010.4, 928.5, 856.5, 778.6, 715.8, 572.8 cm–1. Program of China (2016YFA0602900) and the National Natural
C73H105Cd2N9NiO24 (1776.19): calcd. C 49.36, H 5.96, N 7.10; found
Science Foundation of China (No. 21372087) for financial sup-
C 49.80, H 6.12, N 7.08.
port.
Cycloaddition Reactions of Epoxides and CO2: A 10-mL stainless-
steel reactor was charged with butylene oxide (10.0 mmol), 1
(65 mg, 0.5 mol-%), Bu4NBr (16 mg, 0.5 mol-%), and then CO2 Keywords: Metal–organic frameworks · Metallosalen
(1 MPa). The mixture was stirred at 80 °C for 12 h. After the reaction ligands · Heterogeneous catalysis · Carbon dioxide
was complete, the remaining CO2 was carefully released, the cata- fixation · Cyclic carbonates
lyst was recovered by centrifugation, and the filtrate was analyzed
by GC–MS.
[1] a) Y. Liu, W. Xuan, Y. Cui, Adv. Mater. 2010, 22, 4112; b) J. W. Liu, L. F.
Recyclability Experiments: The recyclability of 1 was tested with Chen, H. Cui, J. Y. Zhang, L. Zhang, C. Y. Su, Chem. Soc. Rev. 2014, 43,
five consecutive runs. The solid phase was collected by centrifuga- 6011; c) M. Zhao, C. D. Wu, ChemCatChem 2017, 9, 1192; d) Y. B. Huang,
tion, rinsed with methanol, dried in air, and reused for subsequent J. Liang, X. S. Wang, R. Cao, Chem. Soc. Rev. 2017, 46, 126; e) S. M. J.
reactions without further treatment. Rogge, A. Bavykina, J. Hajek, H. Garcia, A. I. Olivos-Suarez, A. Sepúlveda-
Escribano, A. Vimont, G. Clet, P. Bazin, F. Kapteijn, M. Daturi, E. V. Ramos-
Crystal Structure Determination and Refinement Fernandez, F. X. Llabrés i Xamena, V. Van Speybroeck, J. Gascon, Chem.
Soc. Rev. 2017, 46, 3134.
Single-crystal XRD analysis of 1 was performed with an Xcalibur [2] a) J. L. Song, Z. F. Zhang, S. Q. Hu, T. B. Wu, T. Jiang, B. X. Han, Green
Onyx Nova four-circle diffractometer using Cu-Kα radiation (λ = Chem. 2009, 11, 1031; b) W. Kleist, F. Jutz, M. Maciejewski, A. Baiker, Eur.
1.54178 Å) at 100 K. The empirical absorption correction was per- J. Inorg. Chem. 2009, 3552; c) D. A. Yang, H. Y. Cho, J. Kim, S. T. Yang,
formed using the CrystalClear program. The structure was solved W. S. Ahn, Energy Environ. Sci. 2012, 5, 6465; d) C. M. Miralda, E. E. Macias,
by direct methods and refined on F2 by the full-matrix least-squares M. Zhu, P. Ratnasamy, M. A. Carreon, ACS Catal. 2012, 2, 180; e) X. Zhou,
technique using the SHELXL-2016/6 program package.[15] The Y. Zhang, X. G. Yang, L. Z. Zhao, G. Y. Wang, J. Mol. Catal. A 2012, 361,
SQUEEZE subroutine of the PLATON software suite[16] was applied 12; f) J. Kim, S. N. Kim, H. G. Jang, G. Seo, W. S. Ahn, Appl. Catal. A 2013,
453, 175; g) W. J. Gao, Y. Chen, Y. H. Niu, K. Williams, L. Cash, P. J. Perez,
to remove the scattering from the highly disordered guest mol-
L. Wojtas, J. F. Cai, Y. S. Chen, S. Q. Ma, Angew. Chem. Int. Ed. 2014,
ecules. The resulting new HKL file was used to further refine the 53, 2615; Angew. Chem. 2014, 126, 2653; h) M. H. Beyzavi, R. C. Klet, S.
structure. Hydrogen atoms attached to carbon atoms were placed Tussupbayev, J. Borycz, N. A. Vermeulen, C. J. Cramer, J. F. Stoddart, J. T.
in geometrically idealized positions and refined using a riding Hupp, O. K. Farha, J. Am. Chem. Soc. 2014, 136, 15861; i) Z. Zhou, C. He,
model. The crystal data and refinement parameters are listed in J. H. Xiu, L. Yang, C. Y. Duan, J. Am. Chem. Soc. 2015, 137, 15066; j) Q. X.
Table 4. Han, B. Qi, W. Ren, C. He, J. Y. Niu, C. Y. Duan, Nat. Commun. 2015, 6,

Eur. J. Inorg. Chem. 2017, 4982–4989 www.eurjic.org 4988 © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper
10007; k) V. Guillerm, L. J. Weseliński, Y. Belmabkhout, A. J. Cairns, V. D′ 2010, 132, 15390; c) A. Bhunia, M. A. Gotthardt, M. Yadav, M. T. Gamer,
Elia, L. Wojtas, K. Adil, M. Eddaoudi, Nat. Chem. 2014, 6, 673; l) Z. R. Jiang, A. Eichhőfer, W. Kleist, P. W. Roesky, Chem. Eur. J. 2013, 19, 1986.
H. Wang, Y. Hu, J. Lu, H. L. Jiang, ChemSusChem 2015, 8, 878; m) A. C. [8] a) Z. W. Yang, C. F. Zhu, Z. J. Li, Y. Liu, G. H. Liu, Y. Cui, Chem. Commun.
Kathalikkattil, R. Roshan, J. Tharun, R. Babu, G. S. Jeong, D. W. Kim, S. J. 2014, 50, 8775; b) J. Li, J. Yang, Y. Y. Liu, J. F. Ma, Chem. Eur. J. 2015, 21,
Cho, D. W. Park, Chem. Commun. 2016, 52, 280; n) R. Babu, A. C. Kathalik- 4413; c) L. Ma, P. Du, J. Yang, Y. Y. Liu, X. L. Liu, J. F. Ma, RSC Adv. 2016,
kattil, R. Roshan, J. Tharun, D. W. Kim, D. W. Park, Green Chem. 2016, 18, 6, 98611.
232; o) H. M. He, J. A. Perman, G. S. Zhu, S. Q. Ma, Small 2016, 12, 6309; [9] Q. C. Xia, Y. Liu, Z. J. Li, W. Gong, Y. Cui, Chem. Commun. 2016, 52, 13167.
p) L. Xu, M. K. Zhai, X. C. Lu, H. B. Du, Dalton Trans. 2016, 45, 18730; q) [10] a) W. Q. Xi, Y. Liu, Q. C. Xia, Z. J. Li, Y. Cui, Chem. Eur. J. 2015, 21, 12581;
B. Ugale, S. S. Dhankhar, C. M. Nagaraja, Inorg. Chem. 2016, 55, 9757; r) b) C. F. Zhu, Q. C. Xia, X. Chen, Y. Liu, X. Du, Y. Cui, ACS Catal. 2016, 6,
J. W. Li, Y. W. Ren, C. R. Qi, H. F. Jiang, Eur. J. Inorg. Chem. 2017, 1478; s) 7590; c) A. Bhunia, S. Dey, J. M. Moreno, U. Diaz, P. Concepcion, K. V.
M. Taherimehr, B. V. de Voorde, L. H. Wee, J. A. Martens, D. E. D. Vos, P. P. Hecke, C. Janiak, P. V. D. Voort, Chem. Commun. 2016, 52, 1401.
Pescarmona, ChemSusChem 2017, 10, 1283; t) J. Ai, X. Min, C. Y. Gao, [11] Y. W. Ren, J. X. Lu, B. W. Cai, D. D. B. Shi, H. F. Jiang, J. Chen, D. Zheng, B.
H. R. Tian, S. Dang, Z. M. Sun, Dalton Trans. 2017, 46, 6756. Liu, Dalton Trans. 2011, 40, 1372.
[3] a) X. B. Xi, T. W. Dong, G. Li, Y. Cui, Chem. Commun. 2011, 47, 3831; b) [12] a) X. Y. Zhao, Y. Li, Z. Chang, L. Chen, X. H. Bu, Dalton Trans. 2016, 45,
W. S. Kim, K. Y. Lee, E. H. Ryu, J. M. Gu, Y. Kim, S. J. Lee, S. Huh, Eur. J. 14888; b) W. Jiang, J. Yang, Y. Y. Liu, S. Y. Song, J. F. Ma, Inorg. Chem.
Inorg. Chem. 2013, 4228; c) S. Bhunia, R. A. Molla, V. Kumari, S. M. Islamb, 2017, 56, 3036.
A. Bhaumik, Chem. Commun. 2015, 51, 15732. [13] a) P. L. Llewellyn, S. Bourrelly, C. Serre, A. Vimont, M. Daturi, L. Hamon,
[4] a) S. C. Xiang, Z. J. Zhang, C. G. Zhao, K. L. Hong, X. B. Zhao, D. R. Ding, G. D. Weireld, J. S. Chang, D. Y. Hong, Y. K. Hwang, S. H. Jhung, G. Ferey,
M. H. Xie, C. D. Wu, M. C. Das, R. Gill, K. M. Thomas, B. L. Chen, Nat. Langmuir 2008, 24, 7245; b) K. Sumida, D. L. Rogow, J. A. Mason, T. M.
Commun. 2011, 2, 204; b) Y. Liu, Z. J. Li, G. Z. Yuan, Q. H. Xia, C. Yuan, Y. McDonald, E. D. Bloch, Z. R. Herm, T. H. Bae, J. R. Long, Chem. Rev. 2012,
Cui, Inorg. Chem. 2016, 55, 12500; c) P. Müller, R. Grünker, V. Bon, M. 112, 724; c) J. R. Li, J. M. Yu, W. G. Lu, L. B. Sun, J. Sculley, P. B. Balbuena,
Pfeffermann, I. Senkovska, M. S. Weiss, X. L. Feng, S. Kaskel, CrystEng- H. C. Zhou, Nat. Commun. 2013, 4, 1538.
Comm 2016, 18, 8164. [14] a) Y. W. Ren, J. G. Chen, C. R. Qi, H. F. Jiang, ChemCatChem 2015, 7, 1535;
[5] a) Y. B. Huang, T. F. Liu, J. X. Lin, J. Lü, Z. J. Lin, R. Cao, Inorg. Chem. 2011, b) M. H. Anthofer, M. E. Wilhelm, M. Cokoja, M. Drees, W. A. Herrmann,
50, 2191; b) Y. W. Ren, X. F. Cheng, S. R. Yang, C. R. Qi, H. F. Jiang, Q. P. F. E. Kühn, ChemCatChem 2015, 7, 94; c) Y. W. Ren, O. Jiang, H. Zeng,
Mao, Dalton Trans. 2013, 42, 9930; c) Y. W. Ren, Y. C. Shi, J. X. Chen, S. R. Q. P. Mao, H. F. Jiang, RSC Adv. 2016, 6, 3243; d) C. Maeda, J. Shimonishi,
Yang, C. R. Qi, H. F. Jiang, RSC Adv. 2013, 3, 2167; d) J. W. Li, Y. W. Ren, R. Miyazaki, J. Hasegawa, T. Ema, Chem. Eur. J. 2016, 22, 6556.
C. R. Qi, H. F. Jiang, Dalton Trans. 2017, 46, 7821; e) J. W. Li, Y. W. Ren, [15] G. M. Sheldrick, Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2015, 71,
C. R. Qi, H. F. Jiang, Chem. Commun. 2017, 53, 8223. 3.
[6] C. F. Zhu, G. Z. Yuan, X. Chen, Z. W. Yang, Y. Cui, J. Am. Chem. Soc. 2012, [16] A. L. Spek, Acta Crystallogr., Sect. D: Biol. Crystallogr. 2009, 65, 148.
134, 8058.
[7] a) F. J. Song, C. Wang, W. B. Lin, Chem. Commun. 2011, 47, 8256; b) F. J.
Song, C. Wang, J. M. Falkowski, L. Q. Ma, W.-B. Lin, J. Am. Chem. Soc. Received: July 18, 2017

Eur. J. Inorg. Chem. 2017, 4982–4989 www.eurjic.org 4989 © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like