Next Article in Journal / Special Issue
Coordination-Induced Self-Assembly of a Heteroleptic Paddlewheel-Type Dirhodium Complex
Previous Article in Journal
A Highly Water-Soluble Fluorescent and Colorimetric pH Probe
Previous Article in Special Issue
Coordination Assemblies of Zn(II) Coordination Polymers: Positional Isomeric Effect and Optical Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Flexible Aromatic Tetracarboxylate as a New Linker for Coordination Polymers

1
School of Life Sciences, Fudan University, Shanghai 200438, China
2
Colleague of Chemistry and Chemical Engineering, Lanzhou University, Lanzhou 730000, China
3
Centro de Química Enstrutural, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001 Lisbon, Portugal
4
Research Institute of Chemistry, Peoples’s Friendship University of Russia (RUDN University), 6Miklukho-Maklaya st., 117198 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Crystals 2020, 10(2), 84; https://doi.org/10.3390/cryst10020084
Submission received: 17 December 2019 / Revised: 17 January 2020 / Accepted: 29 January 2020 / Published: 3 February 2020
(This article belongs to the Special Issue Crystal Structure and Thermal Studies of Coordination Compounds)

Abstract

:
Two new cadmium(II)-based materials, featuring two-dimensional (2D) [Cd26-deta)(bpy)(H2O)]n (1) and three-dimensional (3D) [Cd25-deta)(bpy)2(H2O)]n (2) structures, were constructed by the hydrothermal method from 2,3′,4′,5-diphenyl ether tetracarboxylic acid (H4deta) as an unexplored linker in research on coordination polymers (CPs) and 2,2′-bipyridine (bpy) as a mediator of crystallization. Microcrystalline samples of 1 and 2 were analyzed by IR/PXRD/EA/TGA and X-ray diffraction using single crystals. Structures and topologies of CPs 1 and 2 were established, revealing a 4,6L45 topological layer in 1 and a 3,5T1 topological framework in 2. Structural differences for 1 and 2 are attributed to distinct molar ratios between Cd2+ and bpy during the hydrothermal synthesis. Luminescence and thermal behavior of the obtained materials were also investigated. The present work opens up the use of an unexplored 2,3′,4′,5-diphenyl ether tetracarboxylic acid as a versatile and flexible linker toward the generation of functional coordination polymer materials.

Graphical Abstract

1. Introduction

Coordination polymers (CPs) and metal–organic frameworks (MOFs) have been extensively explored in different chemistry branches on account of their unique and diverse structures, remarkable functional properties, and important application in a wide variety of fields, which, among others, includes gas adsorption and storage [1,2,3,4], catalysis [5,6,7,8,9], molecular magnetism [10,11,12], and photochemistry [13,14,15,16]. In the synthesis of functional CPs and derived materials, various structural and experimental parameters, such as a metal–ion coordination environment and ligand type [17,18,19], pH value [20,21], solvent composition [22,23], and temperature [24,25,26], can play structure-guiding roles. In particular, the variation in the types of organic building blocks has been one of the most prolific routes in building up a large library of functional CPs/MOFs [13,14,15,25,26,27,28,29]. Different aromatic multicarboxylic acids embody a well-explored and adaptable class of ligands for constructing diverse CPs/MOFs, namely due to their ample coordination behavior, varying levels of deprotonation, good thermal stability, and involvement in H-bond interactions [1,12,13,20,27].
As an extension of our broad interest in bringing new types of multicarboxylic acids to construct metal–organic architectures [30,31,32], we have focused on 2,3′,4′,5-diphenyl ether tetracarboxylic acid (H4deta) as a main and still unstudied building block, and probed its reactivity toward cadmium(II) (Cd(II)) ions in the presence of different amounts of 2,2′-bipyridine (bpy) as a mediator of crystallization. H4deta has been selected due to several reasons: (A) this acid features some flexibility due to two phenyl rings interconnected via an O-ether group; (B) it contains two different functionalities (i.e., O-ether and COOH) with nine potential coordination sites that can lead to metal–organic networks of distinct dimensionality and topology; and (C) H4deta remains unexplored for the design of new CPs/MOFs [33].
We describe in the present study the hydrothermally assisted preparation, structural characterization, topologies, thermal stability, and luminescence of new cadmium CPs/MOFs built from H4deta, which feature highly complex metal–organic architectures. The obtained materials represent the unique Cd(II) derivatives assembled from 2,3′,4′,5-diphenyl ether tetracarboxylic acid (shown in Chart 1) and open up the use of such a multicarboxylate spacer toward assembling new metal–organic architectures.

2. Description of Experiments

2.1. Reagents and Equipment

The analytical (A.R.) grade reagents were used with no further purification. H4deta was supplied by Jinan Henghua Sci. & Tec. Co., Ltd. A Bruker EQUINOX 55 apparatus was utilized to measure infrared spectra. An Elementar Vario EL device was used to run elemental analyses. A Rigaku–Dmax/2400 diffractometer was employed for recording PXRD data (λ = 1.54060 Å, Cu-Kα). A LINSEIS STA/PT1600 apparatus was utilized for carrying out TGA measurements (an N2 gas atmosphere, a heating speed of 5 °C per minute, and a 20−900 °C temperature interval). An Edinburgh FLS920 device was used for measuring excitation/emission behavior. The quantum yields of the compounds were determined by an absolute method using an integrating sphere (150 mm diameter, BaSO4 coating) from the Edinburgh FLS920 instrument.

2.2. Synthesis of [Cd26-deta)(bpy)(H2O)]n (1)

CdCl2∙H2O (0.2 mmol, 40 mg), H4deta (0.1 mmol, 35 mg), bpy (0.1 mmol, 16 mg), NaOH (0.4 mmol, 16 mg), and H2O (10 mL) were placed into a stainless steel vessel (Teflon-lined, 25 mL) and stirred continuously for 15 minutes at 25 °C. After this stage, the vessel was closed and subjected to heating at 160 °C for three days, followed by cooling to room temperature (with a rate of 10 °C per hour). Colorless single crystals were manually recovered, rinsed by water (distilled), and left to dry to provide product 1 in a 48% yield relative to H4deta. Elemental analysis calculated for C52H32Cd4N4O20: C 42.13, H 2.18, N 3.78%. Experimental: C 41.95, H 2.16, N 3.76%. IR (main bands, cm−1, KBr): 3076 w, 1627 w, 1588 s, 1560 s, 1537 s, 1487 w, 1470 w, 1442 w, 1399 s, 1354 m, 1265 w, 1230 m, 1158 w, 1086 w, 1013 w, 1021 w, 968 w, 929 w, 890 w, 847 w, 829 w, 813 w, 767 m, 735 w, 651 w, 629 w, 590 w.

2.3. Synthesis of [Cd25-deta)(bpy)2(H2O)]n (2)

CP 2 was synthesized as colorless single crystals following the procedure for 1, but using a different amount of bpy (0.2 mmol, 31 mg). Yield was 38% relative to H4deta. Elemental analysis calculated for C36H24Cd2N4O10: C 48.18, H 2.70, N 6.24%. Experimental: C 48.37, H 2.71, N 6.21%. IR (main bands, cm−1, KBr): 3439 w, 3061 w, 1582 s, 1537 s, 1476 w, 1431 m, 1392 s, 1315 w, 1264 w, 1230 m, 1170 w, 1153 w, 1086 w, 1058 w, 1013 w, 974 w, 929 w, 901 w, 879 w, 851 w, 829 w, 776 m, 727 w, 710 w, 644 w, 616 w, 550 w.

2.4. Structural Determination (Single-crystals)

X-ray data for 1 and 2 were collected on a Bruker APEX-II/CCD diffractometer (λ = 0.71073 Å, Mo Kα). SADABS software was used for correcting absorption (the semi-empirical method). The crystal structure of each sample was determined using a direct method, followed by refinement (full-matrix/least-squares on F2) aided by the SHELXS-97/SHELXL-97 software [34,35]. Every non-hydrogen atom underwent an anisotropic refinement (full-matrix/least-squares on F2). Every hydrogen atom (excluding H2O) was added to a calculated position (fixed isotropic/thermal parameters) and taken into account at the ultimate stage when calculating structure factors. H atoms in water molecules were found using difference maps and fixed to the corresponding oxygen centers. Crystallographic parameters are given in Table 1, while the bonding data are provided in supplementary Tables S1 and S2. Analysis of topologies in 1 and 2 was carried out using an underlying network concept [36,37]. Simplified networks were created upon the elimination of bpy and H2O moieties, along with performing a reduction of the deta4− ligands to their centroids. CCDC-1,961,692 (1) and 1,961,693 (2) encompass all the structural parameters.

3. Discussion of Results

3.1. Preparation of 1 and 2

Owing to the unique conditions of pressure and temperature, the hydrothermal synthetic method is specifically attractive to form CPs/MOFs [38,39,40,41]. Hence, herein, we attempted the application of hydrothermal synthesis for the generation of coordination polymers using 2,3′,4′,5-diphenyl ether tetracarboxylic acid as an unexplored linker. Reactions involving the H2O solutions of CdCl2∙H2O, H4deta (the linker source), NaOH (a base to deprotonate the tetracarboxylic acid ligand), and 2,2′-bipyridine (the mediator of crystallization) resulted in two new products formulated as [Cd26-deta)(bpy)(H2O)]n (CP 1) and [Cd25-deta)(bpy)2(H2O)]n (CP 2). It should be mentioned that both compounds were synthesized using a similar procedure but different molar ratios between bpy and Cd2+, namely 1:2 in 1 and 1:1 in 2. No crystals were obtained in the absence of 2,2′-bipyridine (the crystallization mediator). Reactions using other molar ratios between bpy and Cd2+ (e.g., 2:1; 1:3) were also attempted but failed to produce suitable crystals or gave a mixture of products in low yields. Both products were obtained as analytically pure and stable microcrystals (including single crystals) and analyzed using conventional methods (IR/EA/PXRD/TGA) along with X-ray structural analysis (single crystals), which permitted us to establish their intricate two-dimensional (2D) and three-dimensional (3D) structures.

3.2. Structural Features of [Cd26-deta)(bpy)(H2O)]n (CP 1)

This compound reveals a 2D coordination network (Figure 1). The asymmetric entity is composed of two cadmium(II) atoms (Cd1 and Cd2), one μ6-deta4– linker, one terminal bpy, and one H2O ligand. The seven-coordinate Cd1 center adopts a deformed pentagonal bipyramid (CdO7) environment that is occupied by carboxylate oxygen donors coming from four μ6-deta4– moieties. The six-coordinate Cd2 atom features a deformed octahedral (CdN2O4) geometry, occupied by three oxygen atoms of two μ6-deta4– blocks, one H2O ligand, and a pair of bpy nitrogen atoms (Figure 1a). The cadmium–oxygen (2.224(3)–2.483(10) Å) and cadmium–nitrogen (2.293(4)–2.321(4) Å) distances are usual for related types of compounds [12,13,20]. The deta4– ligand functions as a μ6-spacer (Scheme 1, mode-I), wherein the four carboxylate functionalities are bidentate, tridentate, or bidentate with a bridging mode. Also, μ6-deta4– is bent with an 84.67° angle (a dihedral one between two aromatic functionalities), while another relevant Carbon−Oether−Carbon angle is 117.21°. Four contiguous cadmium nodes are interlinked by six carboxylate functionalities (these come from four μ6-deta4– linkers), generating a tetracadmium(II) subunit (Figure 1b). These Cd4 subunits are additionally connected by the remaining carboxylate functionalities of μ6-deta4–, forming a two-dimensional net structure (Figure 1c). The structure of 1 has no porosity and solvent-accessible voids.
A topological examination of this 2D CP disclosed a two-nodal four, six-interconnected net (Figure 1d), which is assembled from four-linked Cd1 nodes, six-linked μ6-deta4– nodes, and two-interconnected Cd2 linkers. This layer belongs to a 4,6L45 topological type and is characterized by a point symbol of (3.44.6)(32.44.55.63.7); here, the (3.44.6) and (32.44.55.63.7) notations relate to Cd1 and μ6-deta4– nodes, respectively.

3.3. Structural Features of [Cd25-deta)(bpy)2(H2O)]n (2)

If compared to 1, compound 2 (Figure 2) discloses a more complex 3D coordination polymer structure. This asymmetric entity bears a couple of cadmium atoms (Cd1/Cd2), one µ5-deta4– block, two bpy moieties, and one terminal H2O ligand. As represented in Figure 2a, the six-coordinate Cd1 atom has a deformed octahedral (CdN2O4) geometry, taken by a pair of N donors from bpy and four carboxylate oxygen donors from three µ5-deta4– linkers. The five-coordinate Cd2 center features a deformed trigonal bipyramid (CdN2O3) geometry, composed of a couple of oxygen atoms from two µ5-deta4– linkers, a water ligand, and a couple of N atoms from bpy. The cadmium–oxygen distances span in the 2.228(10)−2.450(6) Å range, whereas the cadmium–nitrogen bonds vary between 2.288(9) and 2.388(7) Å; these lie within standard values [31,42]. In 2, the deta4– block takes a μ5-coordination mode (mode-II, Scheme 1) wherein the carboxylate functionalities are mono- or μ-bidentate. In µ5-deta4–, important angles are 88.92° (the dihedral angle between phenyl rings) and 117.67° (the C–Oether–C angle). The µ5-deta4– linkers interconnect the Cd(II) centers to a three-dimensional metal–organic net (Figure 2b). Upon removal of a coordinated water molecule, the PLATON analysis of 2 reveals a minor porosity with an effective free volume of 2.1% of the crystal volume [43].
Regarding the topological classification, this 3D CP is composed of three-linked Cd1 nodes, five-linked µ5-deta4– nodes, and two-interconnected Cd2 linkers (Figure 2c). The resulting underlying framework is a two-nodal three,five-interconnected net of 3,5T1 topological type. This is described by a (42.65.83)(42.6) point symbol where the (42.65.83) and (42.6) indices relate to µ5-deta4– and Cd1 nodes, respectively.

3.4. Discussion of TGA and PXRD, Porosity, and Gas Sorption

The stability of CPs 1 and 2 was investigated by TGA (Figure 3) upon heating from 25 °C to 800 °C using an inert gas flow (N2). As expected, compound 1 discloses the loss of an H2O ligand in the 157−199 °C temperature range (experimental, 2.3%; calculated, 2.4%); after dehydration, the remaining sample is thermally stable on heating till 334 °C. Similarly, compound 2 exhibits a mass decrease in the 162–195 °C range, which is associated with the release of one H2O ligand (experimental, 2.2%; calculated 2.0%), while the dehydrated sample maintains its integrity till 346 °C. CdO is expected as a final decomposition product of both coordination polymers. The decomposition is complete at 819 °C or 881 °C for 1 and 2, respectively. The experimentally observed remaining weights of 34.5% (1) and 28.8% (2) are in very good agreement with the values calculated for CdO, i.e., 34.6% (1) and 28.6% (2).
A bulk, crystalline sample of each coordination polymer (1 and 2) was studied using a PXRD method. Experimental (as-synthesized) and simulated PXRD patterns are given in (Supplementary Materials) Figures S1 and S2. Their comparison indicates a good match and confirms the phase purity of the synthesized products.
Given the presence of some porosity in 2 (2.1% of the effective void volume after activating and removing the water ligand), gas sorption measurements were also performed (for details, see Table S3). These were carried out on the activated samples by recording the N2 adsorption/desorption isotherms at 77 K, resulting in the adsorbed N2 amount of 1.0 cm3/g at 77 K and 1 atm. The BET surface area of 0.58 m2/g was also established.

4. Concluding Remarks

We have explored the simple, aqueous medium protocol (hydrothermal synthesis) to obtain two new two- or three-dimensional CPs/MOFs derived from 2,3′,4′,5-diphenyl ether tetracarboxylic acid (H4deta) as an unexplored linker in research on metal–organic frameworks and coordination polymers. In fact, CPs 1 and 2 represent unique examples of metal-containing derivatives constructed from H4deta. This work thus extends a variety of multifunctional carboxylic acids applied in the synthesis of coordination polymers [44,45,46,47,48,49,50].
The structures of metal–organic nets and the corresponding topologies in the synthesized compounds have also been underscored, revealing underlying networks of the 4,6L45 and 3,5T1 topological types, respectively. This study has also shown that the use of 2,2’-bipyridine as a crystallization mediator in a different stoichiometry to Cd2+ nodes permits the generation of products with distinct dimensionality, structural characteristics, and topologies. Both compounds also show promising luminescent properties toward potential applications as photoactive materials.
Future studies on extending the family of CPs and MOFs assembled from H4deta to other types of metals, as well as on exploring the functional properties of the obtained materials, are in progress.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4352/10/2/84/s1, Figures S1 and S2: PXRD patterns, Figure S3: emission spectra and discussion of luminescent properties, Tables S1 and S2: selected structural parameters for 1 and 2, Table S3: gas sorption data for 2.

Author Contributions

Conceptualization, J.G. and A.M.K.; data curation, W.G.; funding acquisition, J.G. and A.M.K.; investigation, W.G., J.G., and A.M.K.; methodology, J.G.; visualization, J.G. and A.M.K.; writing—original draft, W.G., J.G., and A.M.K.; writing—review and editing, A.M.K. All authors have read and agreed to the published version of the manuscript.

Acknowledgments

This work was supported by the Foundation for Science and Technology (FCT) and Portugal 2020 (LISBOA-01-0145-FEDER-029697, UIDB/00100/2020). The publication was also prepared with the support of the RUDN University Program 5-100.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Glover, T.G.; Mu, B. Gas Adsorption in Metal-Organic Frameworks: Fundamentals and Applications; CRC Press: Boca Raton, FL, USA, 2018. [Google Scholar]
  2. Hashemi, L.; Morsali, A. Pillared Metal-Organic Frameworks: Properties and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2019. [Google Scholar]
  3. Fu, H.R.; Zhao, Y.; Zhou, Z.; Yang, X.G.; Ma, L.F. Neutral ligand TIPA-based two 2D metal-organic frameworks: Ultrahigh selectivity of C2H2/CH4 and efficient sensing and sorption of Cr(VI). Dalton Trans. 2018, 47, 3725–3732. [Google Scholar] [CrossRef] [PubMed]
  4. Zhu, J.; Usov, P.M.; Xu, W.Q.; Celis-Salazar, P.J.; Lin, S.Y.; Kessinger, M.C.; Landaverde-Alvarado, C.; Cai, M.; May, A.M.; Slebodnick, C.; et al. A new class of metal-cyclam-based zirconium metal-organic frameworks for CO2 adsorption and chemical fixation. J. Am. Chem. Soc. 2018, 140, 993–1003. [Google Scholar] [CrossRef] [PubMed]
  5. Gu, J.Z.; Wen, M.; Cai, Y.; Shi, Z.F.; Arol, A.S.; Kirillova, M.V.; Kirillov, A.M. Metal-organic architectures assembled from multifunctional polycarboxylates: hydrothermal self-assembly, structures, and catalytic activity in alkane oxidation. Inorg. Chem. 2019, 58, 2403–2412. [Google Scholar] [CrossRef] [PubMed]
  6. Sotnik, S.A.; Polunin, R.A.; Kiskin, M.A.; Kirillov, A.M.; Dorofeeva, V.N.; Gavrilenko, K.S.; Eremenko, I.L.; Novotortsev, V.M.; Kolotilov, S.V. Heterometallic Coordination Polymers Assembled from Trigonal Trinuclear Fe2Ni-Pivalate Blocks and Polypyridine Spacers: Topological Diversity, Sorption, and Catalytic Properties. Inorg. Chem. 2015, 54, 5169–5181. [Google Scholar] [CrossRef]
  7. Chen, Y.Z.; Zhang, R.; Jiao, L.; Jiang, H.L. Metal-organic framework-derived porous materials for catalysis. Coord. Chem. Rev. 2018, 362, 1–23. [Google Scholar] [CrossRef]
  8. Hu, M.L.; Safarifard, V.; Doustkhah, E.; Rostamnia, S.; Morsali, A.; Nouruzi, N.; Beheshti, S.; Akhbari, K. Taking organic reactions over metal-organic frameworks as heterogeneous catalysis. Micropor. Mesopor. Mat. 2018, 256, 111–127. [Google Scholar] [CrossRef]
  9. Kirillov, A.M.; Coelho, J.A.S.; Kirillova, M.V.; Guedes da Silva, M.F.C.; Nesterov, D.S.; Gruenwald, K.R.; Haukka, M.; Pombeiro, A.J.L. Bringing an “old” biological buffer to coordination chemistry: New 1D and 3D coordination polymers with [Cu4(Hbes)4] cores for mild hydrocarboxylation of alkanes. Inorg. Chem. 2010, 49, 6390–6392. [Google Scholar] [CrossRef]
  10. Espallargas, G.M.; Coronado, E. Magnetic functionalities in MOFs: From the framework to the pore. Chem. Soc. Rev. 2018, 47, 533–557. [Google Scholar] [CrossRef] [Green Version]
  11. Karabach, Y.Y.; Kirillov, A.M.; Haukka, M.; Sanchiz, J.; Kopylovich, M.N.; Pombeiro, A.J.L. Multicopper(II) pyromellitate compounds: Self-assembly synthesis, structural topologies and magnetic features. Cryst. Growth Des. 2008, 8, 4100–4108. [Google Scholar] [CrossRef]
  12. Xu, G.C.; Zhang, W.; Ma, X.M.; Chen, Y.H.; Zhang, L.; Cai, H.L.; Wang, Z.M.; Xiong, R.G.; Gao, S. Coexistence of magnetic and electric orderings in the metal-formate frameworks of [NH4][M(HCOO)3]. J. Mater. Chem. A 2011, 133, 14948–14951. [Google Scholar] [CrossRef]
  13. Cui, Y.L.; Yue, Y.F.; Qian, G.D.; Chen, B.L. Luminescent functional metal-organic frameworks. Chem. Rev. 2012, 112, 1126–1162. [Google Scholar] [CrossRef] [PubMed]
  14. Gao, L.L.; Zhang, J.; Zhai, L.J.; Niu, X.Y.; Hu, T.P. Gas adsorption and fluorescent sensing properties of three metal-organic frameworks based on 3,5-di(2′,5′-dicarboxylphenyl)benozoic acid and cadmium ions. J. Solid State Chem. 2019, 271, 346–353. [Google Scholar] [CrossRef]
  15. Huang, R.W.; Wei, Y.S.; Dong, X.Y.; Wu, X.H.; Du, C.X.; Zang, S.Q.; Mak, T.C. Hypersensitive dual-function luminescence switching of a silver-chalcogenolate cluster-based metal-organic framework. Nat. Chem. 2017, 9, 689–697. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, W.M.; Meng, X.L.; Zhuang, G.L.; Wang, Z.; Kurmoo, M.; Zhao, Q.Q.; Wang, X.P.; Shan, B.R.; Tung, C.H.; Sun, D. A superior fluorescent sensor for Al3+ and UO22+ based on a Co(II) metal-organic framework with exposed pyrimidyl Lewis base sites. J. Mater. Chem. A 2017, 5, 13079–13085. [Google Scholar] [CrossRef]
  17. Gu, J.Z.; Cui, Y.H.; Liang, X.X.; Wu, J.; Lv, D.Y.; Kirillov, A.M. Structurally distinct metal-organicand H-bonded networks derived from 5-(6-carboxypyridin-3-yl)isophthalic acid: coordination and template effect of 4,4’-bipyridine. Cryst. Growth Des. 2016, 16, 4658–4670. [Google Scholar] [CrossRef]
  18. Zhai, Z.W.; Yang, S.H.; Luo, P.; Li, L.K.; Du, C.X.; Zang, S.Q. Zn(II)/Cd(II) metal-organic frameworks based on the 2,5-bis(4-pyridyl)thiazolo[5,4-d]thiazole ligand. Eur. J. Inorg. Chem. 2019, 2725–2734. [Google Scholar] [CrossRef]
  19. Baladi, E.; Nobakht, V.; Tarassoli, A.; Proserpio, D.M.; Carlucci, L. Three cationic, nonporous Cu-I-coordination polymers: structural investigation and vapor iodine capture. Cryst. Growth Des. 2018, 18, 7207–7218. [Google Scholar] [CrossRef] [Green Version]
  20. Gu, J.Z.; Gao, Z.Q.; Tang, Y. pH and auxiliary ligand influence on the structural variations of 5(2’-carboxylphenyl) nicotate coordination polymers. Cryst. Growth Des. 2012, 12, 3312–3323. [Google Scholar] [CrossRef]
  21. An, R.; Zhao, H.; Hu, H.M.; Wang, X.F.; Yang, M.L.; Xue, G.L. Synthesis, structure, white-light emission, and temperature recognition properties of Eu/Tb mixed coordination polymers. Inorg. Chem. 2016, 55, 871–876. [Google Scholar] [CrossRef]
  22. Sánchez-Férez, F.; Bayés, L.; Font-Bardia, M.; Pons, J. Solvent dependent formation of Cu(II) complexes based on isonicotinamide ligand. Inorg. Chim. Acta 2019, 494, 112–122. [Google Scholar] [CrossRef]
  23. Jaremko, Ł.; Kirillov, A.M.; Smoleński, P.; Pombeiro, A.J.L. Engineering coordination and supramolecular copper-organic networks by aqueous medium self-assembly with 1,3,5-triaza-7-phosphaadamantane (PTA). Cryst. Growth Des. 2009, 9, 3006–3010. [Google Scholar] [CrossRef]
  24. Dong, Y.B.; Jiang, Y.Y.; Li, J.; Ma, J.P.; Liu, F.L.; Tang, B.; Huang, R.Q.; Batten, S.R. Temperature-dependent synthesis of metal-organic frameworks based on a flexible tetradentate ligand with bidirectional coordination donors. J. Am. Chem. Soc. 2007, 129, 4520–4521. [Google Scholar] [CrossRef] [PubMed]
  25. Paz, F.A.A.; Klinowski, J.; Vilela, S.M.F.; Tomé, J.P.C.; Cavaleiro, J.A.S.; Rocha, J. Ligand design for functional metal–organic frameworks. Chem. Soc. Rev. 2012, 41, 1088–1110. [Google Scholar]
  26. Wan, J.; Cai, S.L.; Zhang, K.; Li, C.J.; Feng, Y.; Fan, J.; Zheng, S.R.; Zhang, W.G. Anion- and temperature-dependent assembly, crystal structures and luminescence properties of six new Cd(II) coordination polymers based on 2,3,5,6-tetrakis(2-pyridyl)pyrazine. CrystEngComm 2016, 18, 5164–5176. [Google Scholar] [CrossRef]
  27. Fernandes, T.A.; Santos, C.I.M.; André, V.; Kłak, J.; Kirillova, M.V.; Kirillov, A.M. Copper(II) Coordination Polymers Self-assembled from Aminoalcohols and Pyromellitic Acid: Highly Active Pre-catalysts for the Mild Water-promoted Oxidation of Alkanes. Inorg. Chem. 2016, 55, 125–135. [Google Scholar] [CrossRef] [PubMed]
  28. Li, L.; Li, C.X.; Ren, Y.L.; Song, M.; Ma, Y.; Huang, R.D. Novel luminescent metal-organic frameworks based on rigid carboxylate ligands for highly selective sensing of Fe3+ ions. CrystEngComm 2016, 18, 7787–7795. [Google Scholar] [CrossRef]
  29. Zhao, D.; Timmons, D.J.; Yuan, D.; Zhou, H.-C. Tuning the Topology and Functionality of Metal−Organic Frameworks by Ligand Design. Acc. Chem. Res. 2011, 44, 123–133. [Google Scholar] [CrossRef] [PubMed]
  30. Zhao, S.Q.; Gu, J.Z. Synthesis, crystal structures and magnetic properties of Mn(II) and Ni(II) dinuclear coordination compounds constructed from biphenyl-2,4,4’-tricarboxylate and phenanthroline. Chin. J. Inorg. Chem. 2016, 32, 1611–1618. [Google Scholar]
  31. Zhao, N.; Li, Y.; Gu, J.Z.; Kirillova, M.V.; Kirillov, A.M. Hydrothermal generation, structural versatility and properties of metal(II)-organic architectures driven by a pyridine-tricarboxylic acid. Dalton. Trans. 2019, 48, 8361–8374. [Google Scholar] [CrossRef]
  32. Zhao, N.; Li, Y.; Gu, J.Z.; Fernandes, T.A.; Kirillova, M.V.; Kirillov, A.M. New copper(II) coordination compounds assembled from multifunctional pyridine-carboxylate blocks: Synthesis, structures, and catalytic activity in cycloalkane oxidation. Molecules 2019, 24, 6. [Google Scholar] [CrossRef] [Green Version]
  33. Groom, C.R.; Bruno, I.J.; Lightfoot, M.P.; Ward, S.C. The Cambridge Structural Database. Acta Cryst. 2016, B72, 171–179. [Google Scholar] [CrossRef] [PubMed]
  34. Sheldrick, G.M. SHELXS-97; Program for X-ray Crystal Structure Determination; University of Gottingen: Göttingen, Germany, 1997. [Google Scholar]
  35. Sheldrick, G.M. SHELXL-97; Program for X-ray Crystal Structure Refinement; University of Gottingen: Göttingen, Germany, 1997. [Google Scholar]
  36. Blatov, V.A. Multipurpose crystallochemical analysis with the program package TOPOS. IUCr CompComm Newsletter 2006, 7, 4–38. [Google Scholar]
  37. Blatov, V.A.; Shevchenko, A.P.; Proserpio, D.M. Applied topological analysis of crystal structures with the program package topospro. Cryst. Growth Des. 2014, 14, 3576–3586. [Google Scholar] [CrossRef]
  38. Gu, J.Z.; Wen, M.; Liang, X.X.; Shi, Z.F.; Kirillova, M.V.; Kirillov, A.M. Multifunctional aromatic carboxylic acids as versatile building blocks for hydrothermal design of coordination polymers. Crystals 2018, 8, 83. [Google Scholar] [CrossRef] [Green Version]
  39. Gu, J.Z.; Liang, X.X.; Cui, Y.H.; Wu, J.; Shi, Z.F.; Kirillov, A.M. Introducing 2-(2-carboxyphenoxy)terephthalic acid as a new versatile building block for design of diverse coordination polymers: synthesis, structural features, luminescence sensing, and magnetism. CrystEngComm 2017, 19, 2570–2588. [Google Scholar] [CrossRef]
  40. Gu, J.Z.; Liang, X.X.; Cai, Y.; Wu, J.; Shi, Z.F.; Kirillov, A.M. Hydrothermal assembly, structures, topologies, luminescence, and magnetism of a novel series of coordination polymers driven by a trifunctional nicotinic acid building block. Dalton Trans. 2017, 46, 10908–10925. [Google Scholar] [CrossRef]
  41. Gu, J.Z.; Cui, Y.H.; Wu, J.; Kirillov, A.M. A series of mixed-ligand 2D and 3D coordination polymers assembled from a novel multifunctional pyridine-tricarboxylate building block: Hydrothermal syntheses, structural and topological diversity, and magnetic and luminescent properties. RSC Adv. 2015, 5, 78889–78901. [Google Scholar] [CrossRef]
  42. Gu, J.Z.; Cai, Y.; Wen, M.; Shi, Z.F.; Kirillov, A.M. A new series of Cd(II) metal-organic architectures driven by soft ether-bridged tricarboxylate spacers: synthesis, structural and topological versatility, and photocatalytic properties. Dalton Trans. 2018, 47, 14327–14339. [Google Scholar] [CrossRef]
  43. Van de Sluis, P.; Spek, A.L. Bypass-an effective merhve method for the refinement of crystal-structures containing disordered solvent regions. Acta Crystallogr. Sect. A 1990, 46, 194–201. [Google Scholar] [CrossRef]
  44. Gu, J.Z.; Cai, Y.; Qian, Z.Y.; Wen, M.; Shi, Z.F.; Lv, D.Y.; Kirillov, A.M. A new series of Co, Ni, Zn, and Cd metal-organic architectures driven by an unsymmetrical biphenyl-tricarboxylic acid: Hydrothermal assembly, structural features and properties. Dalton Trans. 2018, 47, 7431–7444. [Google Scholar] [CrossRef]
  45. Kirillov, A.M.; Karabach, Y.Y.; Kirillova, M.V.; Haukka, M.; Pombeiro, A.J.L. Topologically unique 2D heterometallic CuII/Mg coordination polymer: synthesis, structural features, and catalytic use in alkane hydrocarboxylation. Cryst. Growth Des. 2012, 12, 1069–1074. [Google Scholar] [CrossRef]
  46. Jaros, S.W.; Guedes da Silva, M.F.C.; Florek, M.; Oliveira, M.C.; Smoleński, P.; Pombeiro, A.J.L.; Kirillov, A.M. Aliphatic dicarboxylate directed assembly of silver(I) 1,3,5-triaza-7-phosphaadamantane coordination networks: topological versatility and antimicrobial activity. Cryst. Growth Des. 2014, 14, 5408–5417. [Google Scholar] [CrossRef]
  47. Zhang, S.W.; Shi, W.; Cheng, P. The coordination chemistry of N-heterocyclic carboxylic acid: A comparison of the coordination polymers constructed by 4,5-imidazoledicarboxylic acid and 1H-1,2,3-triazole-4,5-dicarboxylic acid. Coord. Chem. Rev. 2017, 352, 108–150. [Google Scholar] [CrossRef]
  48. Gangu, K.K.; Maddila, S.; Jonnalagadda, S.B. A review on synthesis, crystal structure and functionality of naphthalenedicarboxylate ligated metal-organic frameworks. Inorg. Chim. Acta. 2017, 466, 308–323. [Google Scholar] [CrossRef]
  49. Razavi, S.A.A.; Morsali, A. Linker functionalized metal-organic frameworks. Coord. Chem. Rev. 2019, 399, 213023. [Google Scholar] [CrossRef]
  50. Xie, X.X.; Yang, Y.C.; Dou, B.H.; Li, Z.F.; Li, G. Proton conductive carboxylate-based metal-organic frameworks. Coord. Chem. Rev. 2020, 403, 213100. [Google Scholar] [CrossRef]
Chart 1. 2,3′,4′,5-diphenyl ether tetracarboxylic acid (H4deta).
Chart 1. 2,3′,4′,5-diphenyl ether tetracarboxylic acid (H4deta).
Crystals 10 00084 ch001
Scheme 1. Distinct modes of coordination for µ6-deta4− in 1 (mode-I) and µ5-deta4− in 2 (mode-II).
Scheme 1. Distinct modes of coordination for µ6-deta4− in 1 (mode-I) and µ5-deta4− in 2 (mode-II).
Crystals 10 00084 sch001
Figure 1. Representations of the structure of 1. (a) Coordination environment and connectivity of cadmium (Cd) atoms. (b) A tetracadmium(II) subunit. (c) A two-dimensional (2D) coordination polymer structure (view along the a axis). (d) Topological view of a simplified 2D net representing a two-nodal 4,6-c layer of the 4,6L45 topological type. View along the a axis; 2-c Cd2 linkers and 4-c Cd1 nodes (balls, turquoise), centroids of 6-c µ6-deta4– nodes (gray sticks).
Figure 1. Representations of the structure of 1. (a) Coordination environment and connectivity of cadmium (Cd) atoms. (b) A tetracadmium(II) subunit. (c) A two-dimensional (2D) coordination polymer structure (view along the a axis). (d) Topological view of a simplified 2D net representing a two-nodal 4,6-c layer of the 4,6L45 topological type. View along the a axis; 2-c Cd2 linkers and 4-c Cd1 nodes (balls, turquoise), centroids of 6-c µ6-deta4– nodes (gray sticks).
Crystals 10 00084 g001aCrystals 10 00084 g001b
Figure 2. Representations of the structure of 2. (a) Coordination environment and connectivity of Cd atoms. (b) Three-dimensional (3D) coordination polymer structure (view along the a axis). (c) Topological view of a simplified 3D network representing a two-nodal 3,5-c net of the 3,5T1 topological type. View along the a axis; 2-c Cd2 linkers and 3-c Cd1 nodes (balls, turquoise), centroids of 5-c µ5-deta4– nodes (gray sticks).
Figure 2. Representations of the structure of 2. (a) Coordination environment and connectivity of Cd atoms. (b) Three-dimensional (3D) coordination polymer structure (view along the a axis). (c) Topological view of a simplified 3D network representing a two-nodal 3,5-c net of the 3,5T1 topological type. View along the a axis; 2-c Cd2 linkers and 3-c Cd1 nodes (balls, turquoise), centroids of 5-c µ5-deta4– nodes (gray sticks).
Crystals 10 00084 g002aCrystals 10 00084 g002b
Figure 3. TGA curves for compounds 1 and 2.
Figure 3. TGA curves for compounds 1 and 2.
Crystals 10 00084 g003
Table 1. Crystal data for coordination polymers (CPs) 1 and 2.
Table 1. Crystal data for coordination polymers (CPs) 1 and 2.
Compound12
Chemical formulaC52H32Cd4N4O20C36H24Cd2N4O10
Molecular weight1482.41897.39
Crystal systemMonoclinicMonoclinic
Space groupP21/cP21/c
a11.5673(3)7.1639(3)
b18.4564(4)17.8025(8)
c11.9067(3)26.3248(10)
α/ (°)9090
β/(°)93.468(3)93.142(4)
γ/(°)9090
V32537.31(11)3352.3(2)
Z24
F(000)14481776
Crystal size/mm0.23×0.22×0.200.24×0.23×0.21
θ range for data collection3.360–25.0503.508–25.050
Limiting indices–13 ≤ h ≤ 13, −20 ≤ k ≤ 21, −14 ≤ l ≤ 6–8 ≤ h ≤ 8, −21 ≤ k ≤ 19, −31 ≤ l ≤ 13
Reflections collected/unique (Rint)9068/4479 (0.0382)12407/5932 (0.0514)
Dc/ (Mg·cm−3)1.9401.778
μ/mm−11.7391.335
Data/restraints/parameters4479/0/3615932/16/470
Goodness-of-fit on F21.0681.020
Final R indices [(I≥2σ(I))] R1, wR20.0370, 0.06250.0662, 0.1341
R indices (all data) R1, wR20.0525, 0.07150.1018, 0.1595
Largest diff. peak and hole/(e·Å−3)0.473 and −0.6081.661 and −1.078

Share and Cite

MDPI and ACS Style

Gu, W.; Gu, J.; Kirillov, A.M. A Flexible Aromatic Tetracarboxylate as a New Linker for Coordination Polymers. Crystals 2020, 10, 84. https://doi.org/10.3390/cryst10020084

AMA Style

Gu W, Gu J, Kirillov AM. A Flexible Aromatic Tetracarboxylate as a New Linker for Coordination Polymers. Crystals. 2020; 10(2):84. https://doi.org/10.3390/cryst10020084

Chicago/Turabian Style

Gu, Wenjun, Jinzhong Gu, and Alexander M. Kirillov. 2020. "A Flexible Aromatic Tetracarboxylate as a New Linker for Coordination Polymers" Crystals 10, no. 2: 84. https://doi.org/10.3390/cryst10020084

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop