C8CC08965E

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 4

ChemComm

View Article Online


COMMUNICATION View Journal
Published on 21 December 2018. Downloaded by Gothenburg University Library on 1/21/2019 1:40:24 PM.

A zeolite supramolecular framework with LTA


topology based on a tetrahedral metal–organic
Cite this: DOI: 10.1039/c8cc08965e
cage†
Received 11th November 2018,
Accepted 21st December 2018 Hai-Xia Zhang,‡a Xiaodong Yan, ‡a Yu-Xin Chen,a Shu-Heng Zhang,a Tao Li,a
Wang-Kang Han,a Ling-Yu Bao,a Rui Shena and Zhi-Guo Gu *ab
DOI: 10.1039/c8cc08965e

rsc.li/chemcomm

An unprecedented zeolite supramolecular framework featuring


truncated cuboctahedral and truncated octahedral cavities was
self-assembled from tetrahedral metal–organic cationic cages
and tetrahedral anions. This crystalline porous material could trap
iodine and organic dye molecules, and its solid state spin-crossover
behavior was affected by guest encapsulation.

Aluminosilicate Linde type A (LTA) zeolite is an industrially


important porous zeolite that is extensively used for adsorption
and separation processes, and as catalytic materials for petroleum
refining, and ion exchange materials for water softening and
effluent treatment.1 The LTA framework with cubic symmetry is
formed by two types of cages, truncated cuboctahedra (a-cage) and Fig. 1 The building units of the zeolite structure: (a) silica tetrahedron (Si:
truncated octahedra (b-cage), which are linked by double 4-rings.2 pale blue; O: red); (b) the ZIFs (C: grey; N: blue; Zn: turquoise); (c) tetrahedral
metal–organic M4 cage and tetrahedral anion X formed XM4 and MX4
This unique LTA topology has been established for inorganic
tetrahedrons (M: purple; X: green).
zeolites composed of TO4 tetrahedrons bridged by m2-O atoms
with a T–O–T angle of ca. 1451 (Fig. 1a), and zeolitic imidazolate
frameworks (ZIFs) composed of Zn(Im)4 tetrahedrons and m2-Im metal ions and organic ligands.5 The regular polyhedral geometry,
linkers with a Zn–Im–Zn angle also of 1451 (Fig. 1b).3 Recently, tunable cavity size, adjustable ligands, alterable counterions, and
supramolecular hydrogen-bond interactions between special optional metal centres of metal–organic cages make them ideal
organic molecules have been proved to be an effective strategy to building blocks for constructing supramolecular frameworks.
obtain an LTA structure, in which the configuration and geometry However, assembling metal–organic cages into a well-organized
of the building blocks play a decisive structure-directing role.4 architecture especially a zeolite supramolecular framework is still
However, there are great challenges in assembling a zeolite supra- challenging likely due to the uncontrollable and non-directional
molecular framework with LTA topology by using coordination intermolecular interactions.6 Based on inorganic zeolites and
compounds as building blocks. ZIFs, a tetrahedral building unit is an essential prerequisite for
Metal–organic cages are discrete, highly symmetric molecular constructing a zeolite framework.7 Therefore, tetrahedral cationic
cage compounds constructed by coordination bonds of inorganic metal–organic cages with tetrahedral anions inherently possess
fundamental building elements of a zeolite supramolecular
a
framework (Fig. 1c). Under the right circumstances, the positive
Key Laboratory of Synthetic and Biological Colloids, Ministry of Education,
School of Chemical and Material Engineering, Jiangnan University, Wuxi 214122,
charge centre of the metal ions (M) in tetrahedral metal–organic
P. R. China. E-mail: zhiguogu@jiangnan.edu.cn cages and the negative charge centre of the anions (X) may form
b
International Joint Research Center for Photoresponsive Molecules and Materials, MX4 and XM4 artificial tetrahedrons through electrostatic
School of Chemical and Material Engineering, Jiangnan University, Wuxi 214122, interaction, and it is expected to form a zeolite supramolecular
P. R. China
framework.
† Electronic supplementary information (ESI) available: Experimental details, general
characterization, additional plot and discussion. CCDC 1873319. For ESI and crystallo-
Herein, we report a tetrahedral iron(II) metal–organic cage,
graphic data in CIF or other electronic format see DOI: 10.1039/c8cc08965e which is further assembled with tetrahedral ClO4 anions to form
‡ These authors contributed equally. a rare LTA topological supramolecular framework. The high

This journal is © The Royal Society of Chemistry 2019 Chem. Commun.


View Article Online

Communication ChemComm

porosity of this supramolecular framework was confirmed by and coordination (Fe - N) bonds were formed during the one-
adsorption of iodine and dye molecules. The effect of guest on pot self-assembly process.8 The IR spectrum of 1 shows typical
the spin-crossover properties of the supramolecular framework absorptions for the stretching of the imidazole-imine (CQN)
was investigated. groups at 1607 and 1571 cm 1. The peaks at about 1097 cm 1
1 was synthesized by self-assembly of 3,5-di(2-(1-(imidazole- reveal the existence of ClO4 (Fig. S5, ESI†). The 1H NMR
2-carboxaldehyde))methyl)toluene, iron(II) perchlorate hydrate spectrum of 1 shows signals from two imidazole protons
and 1-phenylethylamine in acetonitrile solution (see the ESI†). at 10.65 and 10.87 ppm, and the CQN protons at 14.08 ppm
Published on 21 December 2018. Downloaded by Gothenburg University Library on 1/21/2019 1:40:24 PM.

Dark purple crystals were obtained in a few days by slow vapor (Fig. S6, ESI†). The high resolution mass spectra (HRMS) of 1
diffusion of diethyl ether into the reaction mixtures. reveal molecular ion peaks at m/z 1269.58 and 1954.09 corres-
The crystals of 1 remained stable in air and common solvents, ponding to [1(ClO4)5]3+ and [1(ClO4)6]2+ (Fig. S7, ESI†). 1 is
such as n-hexane and water (Fig. S4, ESI†). The covalent (CQN) stable at temperatures up to 195 1C as shown by TGA analysis

Fig. 2 Illustration of the X-ray crystal structures of 1. (a) Iron(II) tetrahedral cage (ORTEP, 20% ellipsoids). All H atoms and anions have been removed for
clarity; C: grey; N: blue; and Fe: purple. The green broken lines connect four vertices of the Fe4 tetrahedron. (b) One edge of the Fe4 tetrahedron and one
vertex of the Fe4 tetrahedron. (c) The two neighboring cages are linked by electrostatic interactions (indicated by the turquoise lines; O: red; and Cl:
green). (d) Artificial double 4-rings formed by a metal–organic cage with four anions situated on the windows of the Fe4 tetrahedron. (e) Artificial
truncated octahedron (b-cage) formed through assembling twelve iron(II) centers with twelve ClO4 anions. (f) The double 4-rings is connected with two
b-cages. (g) Artificial truncated cuboctahedra (a-cage) formed through assembling twenty four iron(II) centers with twenty four ClO4 anions. (h) Each
a-cage is surrounded by twelve double 4-rings. (i) Supramolecular LTA topology, only Fe atoms and Cl atoms are retained for clarity.

Chem. Commun. This journal is © The Royal Society of Chemistry 2019


View Article Online

ChemComm Communication

(Fig. S8, ESI†). The phase purity of 1 was checked by powder


X-ray diffraction (PXRD) (Fig. S9 and S10, ESI†).
Single-crystal X-ray diffraction analysis reveals that 1 crystal-
lized in the cubic F23 space groups with a giant unit cell volume
of 157931(18) Å3 (Table S1, ESI†). In each tetrahedral cage of 1
(Fig. 2a), four Fe(II) ions occupy the vertices of the tetrahedron
and six C2-symmetry ligands define the edges with an average
Published on 21 December 2018. Downloaded by Gothenburg University Library on 1/21/2019 1:40:24 PM.

Fe  Fe separation along the edges of 9.59 Å. Each iron(II)


center coordinates with three ligands forming a distorted
FeN6 octahedral geometry, while each ligand connects two
iron(II) centers (Fig. 2b). The average Fe–N bond length is
1.98 Å, indicating the low-spin state of the Fe(II) center.9 Rich
intramolecular p–p interactions exist between each imidazole
and phenyl ring of the adjacent ligand with a center–center
distance ranging from 3.552 to 3.690 Å (Fig. S11, ESI†). The
tetrahedral cages in 1 pack in window to vertex stacks and the Fig. 3 (a) Time-dependent UV-vis absorption spectra of the iodine/n-hexane
anions are decorated around the vertexes or the windows of solution (2 mmol L 1, 8 mL) upon addition of 1 (10 mg). (b) Gravimetric uptake
of iodine as a function of time at 75 1C. (c) Time-dependent UV-vis absorption
the Fe4 tetrahedron. Each ClO4 ion is surrounded by two
spectra of the Amaranth aqueous solution (AM, 5 mL, [AM] = 60 mmol L 1) upon
tetrahedral cages in the crystal lattice, and the positive Fe4 addition of 1 (15 mg). (d) Time-dependent UV-vis absorption spectra of the
cages and the negative ClO4 ions are connected through the Congo Red aqueous solution (CR, 5 mL, [CR] = 120 mmol L 1) upon addition of
electrostatic interactions (Fig. 2c). In addition, the anions also 1 (15 mg). The photographs of the bottles (insets) show the color of the iodine
act as hydrogen bond donors forming C–H  O hydrogen bonds and dye molecules before and after inclusion.

with the protons of methyl and phenyl (Fig. S12, ESI†).


It is interesting that these isolated tetrahedral iron(II) metal–
organic cages and tetrahedral ClO4 anions are further assembled gradually by weighing, and an iodine uptake of 281 wt% was
into a higher-order supramolecular framework. One Fe4 cage and determined, corresponding to 42.5 I2 per formula unit of 1
four adjacent anions form a cube or double 4-rings with four Fe (Fig. 3b). This is consistent with a theoretical void space of
atoms and four Cl atoms alternately occupying the vertexes 37.9% calculated from the crystal structure of 1 by PLATON.
(Fig. 2d). By linking the Fe atoms and Cl atoms, artificial truncated The iodine vapor uptake capacity is higher than those of the
octahedrons (b-cages) containing six quadrangles and eight previously reported solid porous adsorbents, including metal
hexagons can be obtained (Fig. 2e). It should be noted that organic frameworks (MOFs),10 amorphous porous organic poly-
the single crystal morphology of 1 is also truncated octahedron, mers (POPs),11 and zeolites.12 The dye adsorption capacity of 1
which is strikingly similar to the internal molecular structure was also investigated. The time-dependent UV-vis spectra
of the b-cages (Fig. S13, ESI†). Each b-cage is connected with clearly showed a continuous decrease of the absorbance in
six adjacent b-cages through double 4-ring units (Fig. 2f). In anionic dyes Amaranth (AM) and Congo Red (CR) (Fig. 3c and
addition, larger artificial truncated cuboctahedrons (a-cages) d), and the color of the aqueous solution gradually became
containing twelve quadrangles, eight hexagons and six octagons lighter and almost colorless after 60 h. The AM and CR removal
are formed (Fig. 2g). The shortest Fe  Fe and Cl  Cl distances capacity of 1 was 11.06 and 21.26 mg g 1, respectively (Fig. S17,
in 1 are about 12.97 and 11.52 Å, which generate large pores ESI†). The neutral dye neutral red (NR) and the cationic dye
with a diameter of 17.99 and 28.90 Å for the b-cage and a-cage, methylene blue (MB) could hardly be adsorbed in the cavities of
respectively. One a-cage is in contact with twelve double 4-rings 1 (Fig. S18, ESI†). These results show that the LTA topology
(Fig. 2h). In this way, a zeolite supramolecular framework with structure of 1 with various cavities, such as double 4-rings,
LTA topology is finally constructed in the form of a simple cubic b-cages and a-cages, was capable of capturing small iodine
arrangement of the b-cages connected by the double 4-rings, molecules and large cationic dye molecules.
and producing an a-cage in the center of the cell (Fig. 2i). This is The structure of 1 contains iron(II) centers with an N6
the first example of an LTA structure using metal–organic cages coordination environment, which have two electron configurations,
as building blocks. t42ge2g (high-spin HS) and t62ge0g (low-spin LS). Therefore, its magnetic
In order to confirm the porous structure of 1, the iodine susceptibility was measured in the temperature range of 2–400 K. As
capture capacity was examined in solution and in vapor. By shown in Fig. 4a, 1 exhibits a gradual and incomplete spin-crossover
soaking the crystalline samples of 1 in an iodine/n-hexane behavior. Upon increasing the temperature, the wMT value gradually
solution at room temperature, the purple solution gradually increased to ca. 3.92 cm3 k mol 1 at 200 K. As the temperature
faded until completely colorless (Fig. 3a). The iodine concentration increased to 400 K, the wMT value reached 11.20 cm3 k mol 1,
continuously decreased over time to nearly zero, corresponding to indicating that the four iron(II) centers are almost in the HS
a removal capacity of 20.12 wt% (Fig. S15, ESI†). An iodine capture state. And the thermal spin transition temperature T1/2 was
experiment in vapor was also carried out by exposing 1 to iodine estimated to be 321 K. In contrast, after loading the guest
vapor at 75 1C. Upon standing, the mass of 1 was found to increase molecules, 1@I2 (Fig. 4b), 1@AM (Fig. 4c) and 1@CR (Fig. 4d)

This journal is © The Royal Society of Chemistry 2019 Chem. Commun.


View Article Online

Communication ChemComm

Conflicts of interest
There are no conflicts to declare.

Notes and references


1 (a) L. V. Nadal and L. Cronin, Nat. Rev. Mater., 2017, 2, 17054;
(b) Introduction to Zeolite Science and Practice, ed. H. Van Bekkum,
E. M. Flanigen, P. A. Jacobs and J. C. Jansen, Elsevier, Amsterdam,
Published on 21 December 2018. Downloaded by Gothenburg University Library on 1/21/2019 1:40:24 PM.

2001; (c) A. S. Huang, F. Y. Liang, F. Steinbach and J. Caro, J. Membr. Sci.,


2010, 350, 5–9; (d) J. Čejka, A. Corma and S. Zones, Zeolites and Catalysis:
Synthesis Reactions and Applications, Wiley, Weinheim, 2010.
2 (a) M. Kumar, M. K. Choudhary and J. D. Rimer, Nat. Commun., 2018,
9, 2129; (b) A. Corma, F. Rey, J. Rius, M. J. Sabater and S. Valencia,
Nature, 2004, 431, 287–290; (c) T. Pham, K. A. Forrest, H. Furukawa,
J. Eckert and B. Space, J. Phys. Chem. C, 2018, 122, 15435–15445.
3 (a) H. Hayashi, A. P. Côté, H. Furukawa, M. O’Keeffe and O. M. Yaghi,
Nat. Mater., 2007, 6, 501–506; (b) S. R. Venna, J. B. Jasinski and
Fig. 4 Plots of wMT versus T for (a) 1, (b) 1@I2, (c) 1@AM, and (d) 1@CR. M. A. Carreon, J. Am. Chem. Soc., 2010, 132, 18030–18033.
4 (a) M. Handke, T. Adachi, C. Hu and M. D. Ward, Angew. Chem., Int.
Ed., 2017, 56, 14003–14006; (b) S. Calero and P. G. Álvarez, J. Phys.
Chem. C, 2014, 118, 9056–9065.
displayed paramagnetic behaviors with no SCO properties. This 5 (a) D. A. Roberts, B. S. Pilgrim and J. R. Nitschke, Chem. Soc. Rev., 2018,
phenomenon can also be confirmed by IR spectroscopy, Raman 47, 626–644; (b) N. Ahmad, H. A. Younus, A. H. Chughtai and F. Verpoort,
Chem. Soc. Rev., 2015, 44, 9–25; (c) H. Vardhan, M. Yusubov and
spectroscopy and solid UV-vis spectroscopy13 (Fig. S19–S21, F. Verpoort, Coord. Chem. Rev., 2016, 306, 171–194; (d) M. D. Ward,
ESI†). The disappearance of the SCO behaviors of 1@I2, C. A. Hunter and N. H. Williams, Acc. Chem. Res., 2018, 51, 2073–2082.
1@AM and 1@CR could be attributed to the following reasons: 6 (a) Y. Inokuma, T. Arai and M. Fujita, Nat. Chem., 2010, 2, 780–783;
(b) D. Luo, X. P. Zhou and D. Li, Angew. Chem., Int. Ed., 2015, 54, 1–7;
(I) the donor–acceptor interactions between the electron-rich (c) Z. F. Ju, G. L. Liu, Y. S. Chen, D. Q. Yuan and B. L. Chen, Chem. –
Fe(II) framework and the electron-deficient iodine molecules; Eur. J., 2017, 23, 1–5; (d) M. Käseborn, J. J. Holstein, G. H. Clever and
(II) the electrostatic interactions between the cationic frame- A. Lützen, Angew. Chem., Int. Ed., 2018, 130, 12349–12353.
7 (a) Y. X. Tan, F. Wang and J. Zhang, Chem. Soc. Rev., 2018, 47,
work and the anionic dye molecules; and (III) the intermolecular 2130–2144; (b) H. X. Zhang, F. Wang, H. Yang, Y. X. Tan, J. Zhang
interactions (such as hydrogen-bond interactions, p–p interactions, and X. H. Bu, J. Am. Chem. Soc., 2011, 133, 11884–11887.
and CH  p interactions) between the abundant phenyl groups 8 (a) D. A. Roberts, B. S. Pilgrim, G. Sirvinskaite, T. K. Ronson and
J. R. Nitschke, J. Am. Chem. Soc., 2018, 140, 9616–9623;
on the cages and the dye molecules. This is in accordance (b) T. K. Ronson, W. J. Meng and J. R. Nitschke, J. Am. Chem. Soc.,
with other reported porous SCO compounds, in which the 2017, 139, 9698–9707; (c) Z. X. Wu, K. Zhou, A. V. Ivanov, M. Yusobov
guest such as benzene or iodine molecules caused T1/2 to and F. Verpoort, Coord. Chem. Rev., 2017, 353, 180–200.
9 (a) C. Bressler, C. Milne, V. T. Pham, A. Elnahhas, R. M. van der Veen,
move to low temperature and even made the SCO properties W. Gawelda, S. Johnson, P. Beaud, D. Grolimund, M. Kaiser, C. N. Borca,
disappear.14 G. Ingold, R. Abela and M. Chergui, Science, 2009, 323, 489–492; (b) L. S.
In conclusion, we have presented the construction of a Xie, L. Sun, R. M. Wan, S. S. Park, J. A. Degayner, C. H. Hendon and
M. Dincǎ, J. Am. Chem. Soc., 2018, 140, 7411–7414; (c) Z. Yan, M. Li,
tetrahedral metal–organic cage that is able to self-assemble H. L. Gao, X. C. Huang and D. Li, Chem. Commun., 2012, 48, 3960–3962.
into a rare supermolecular zeolite framework with LTA topology. 10 (a) D. F. Sava, K. W. Chapman, M. A. Rodriguez, J. A. Greathouse,
Such supermolecular zeolite frameworks possess much larger P. S. Crozier, H.-Y. Zhao, P. J. Chupas and T. M. Nenoff, Chem.
Mater., 2013, 25, 2591–2596; (b) Q. K. Liu, J. P. Ma and Y. B. Dong,
a-cage and b-cage cavities compared to traditional inorganic Chem. Commun., 2011, 47, 7185–7187.
zeolites and ZIFs, thereby showing good ability to adsorb large 11 (a) Z. Yan, Y. Yuan, Y. Tian, D. Zhang and G. Zhu, Angew. Chem., Int.
size guest molecules. We believe that this new approach will Ed., 2015, 54, 12733–12737; (b) C. Y. Pei, T. Ben, S. X. Xu and
S. L. Qiu, J. Mater. Chem. A, 2014, 2, 7179–7187; (c) Y. H. Abdelmoaty,
initiate the search for other supermolecular zeolite frameworks T. Tessema, F. A. Choudhury, O. M. El-Kadri and H. M. El-Kaderi,
with metal–organic polyhedral compounds as building blocks. ACS Appl. Mater. Interfaces, 2018, 10, 16049–16058.
Simultaneously, the fascinating optical, electrical, magnetic, 12 K. W. Chapman, P. J. Chupas and T. M. Nenoff, J. Am. Chem. Soc.,
2010, 132, 8897–8899.
and catalytic functions of metal–organic cages can be introduced 13 (a) W. K. Han, L. F. Qin, C. Y. Pang, C. K. Cheng, W. Zhu, Z. H. Li, Z. J. Li,
into the supermolecular porous framework to realize multi- X. H. Ren and Z. G. Gu, Dalton Trans., 2017, 46, 8004–8008; (b) N. Suemura,
functional metal–organic materials. M. Ohama and S. Kaizaki, Chem. Commun., 2001, 1538–1539;
(c) A. Ferguson, M. A. Squire, D. Siretanu, D. Mitcov, C. Mathonière,
This work was supported by the National Natural Science R. Clérac and P. E. Kruger, Chem. Commun., 2013, 49, 1597–1599.
Foundation of China (21771089), the Fundamental Research 14 (a) M. Ohba, K. Yoneda, G. Agustı́, M. C. Muñoz, A. B. Gaspar, J. A. Real,
Funds for the Central Universities (JUSRP51725B, JUSRP51513), M. Yamasaki, H. Ando, Y. Nakao, S. Sakaki and S. Kitagawa, Angew.
Chem., 2009, 121, 4861–4865; (b) P. D. Southon, L. Liu, E. A. Fellows,
the project for Jiangsu Scientific and Technological Innovation D. J. Price, G. J. Halder, K. W. Chapman, B. Moubaraki, K. S. Murray,
Team, and the MOE & SAFEA for the 111 Project (B13025). J. F. Létard and C. J. Kepert, J. Am. Chem. Soc., 2009, 131, 10998–11009.

Chem. Commun. This journal is © The Royal Society of Chemistry 2019

You might also like