Next Article in Journal
Atomic Level Interface Control of SnO2-TiO2 Nanohybrids for the Photocatalytic Activity Enhancement
Next Article in Special Issue
Effective Separation of Prime Olefins from Gas Stream Using Anion Pillared Metal Organic Frameworks: Ideal Adsorbed Solution Theory Studies, Cyclic Application and Stability
Previous Article in Journal
Immobilization of Exfoliated g-C3N4 for Photocatalytical Removal of Organic Pollutants from Water
Previous Article in Special Issue
ZIF-67 Derived MnO2 Doped Electrocatalyst for Oxygen Reduction Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Metal(II) Coordination Polymers from Tetracarboxylate Linkers: Synthesis, Structures, and Catalytic Cyanosilylation of Benzaldehydes

1
Guangdong Research Center for Special Building Materials and Its Green Preparation Technology/Foshan Research Center for Special Functional Building Materials and Its Green Preparation Technology, Guangdong Industry Polytechnic, Guangzhou 510300, China
2
College of Chemistry and Chemical Engineering, Lanzhou University, Lanzhou 730000, China
3
Centro de Química Estrutural and Departamento de Engenharia Química, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001 Lisbon, Portugal
4
Research Institute of Chemistry, Peoples’ Friendship University of Russia (RUDN University), 6 Miklukho-Maklaya st., 117198 Moscow, Russia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 26 December 2020 / Revised: 19 January 2021 / Accepted: 29 January 2021 / Published: 3 February 2021
(This article belongs to the Special Issue MOFs for Advanced Applications)

Abstract

:
Three 2D coordination polymers, [Cu24-dpa)(bipy)2(H2O)]n∙6nH2O (1), [Mn2(µ6-dpa)(bipy)2]n (2), and [Zn24-dpa)(bipy)2(H2O)2]n·2nH2O (3), were prepared by a hydrothermal method using metal(II) chloride salts, 3-(2′,4′-dicarboxylphenoxy)phthalic acid (H4dpa) as a linker, as well as 2,2′-bipyridine (bipy) as a crystallization mediator. Compounds 13 were obtained as crystalline solids and fully characterized. The structures of 13 were established by single-crystal X-ray diffraction, revealing 2D metal-organic networks of sql, 3,6L66, and hcb topological types. Thermal stability and catalytic behavior of 13 were also studied. In particular, zinc(II) coordination polymer 3 functions as a highly active and recoverable heterogeneous catalyst in the mild cyanosilylation of benzaldehydes with trimethylsilyl cyanide to give cyanohydrin derivatives. The influence of various parameters was investigated, including a time of reaction, a loading of catalyst and its recycling, an effect of solvent type, and a substrate scope. As a result, up to 93% product yields were attained in a catalyst recoverable and reusable system when exploring 4-nitrobenzaldehyde as a model substrate. This study contributes to widening the types of multifunctional polycarboxylic acid linkers for the design of novel coordination polymers with notable applications in heterogeneous catalysis.

Graphical Abstract

1. Introduction

Functional coordination polymers (CPs) and derived materials have been of a special focus in recent years owing to important structural characteristics of these compounds [1,2,3], intrinsic properties [4,5], and a broad diversity of applications [6,7,8,9,10] including in the field of catalysis [11,12,13,14,15,16,17]. The development of new catalytic systems incorporating coordination polymers with target structures and functionalities continues to be a challenging area, since the assembly of CPs can be affected by a diversity of factors. These include the nature of metal centers, organic linkers and supporting ligands, stoichiometry, and various reaction conditions [18,19,20,21,22,23,24].
Aromatic carboxylic acids with several COOH groups are the most common building blocks for constructing functional CPs [14,15,17,19]. Within a diversity of such carboxylic acids, semi-flexible linkers are especially captivating due to unique geometrical arrangements and conformational flexibility, which may lead to crystallization of unexpected metal-organic architectures [21,22,23,25].
On the other hand, cyanosilylation of carbonyl substrates is an interesting reaction for C-C bond formation [26,27], which is used for the preparation of cyanohydrins—key precursors for some pharmaceutical and fine chemistry products [28,29]. The use of transition metal complexes or coordination polymers in cyanosilylation reactions is gaining relevance as these compounds can behave as low-cost, efficient, and recyclable catalysts [25,30,31].
Considering our research focus on the design of CPs and catalytic systems on their basis [14,15,25], the main goal of this study consisted in the preparation of new metal-organic architectures, followed by their characterization and catalytic application in cyanosilylation of benzaldehydes. Thus, we selected an unexplored ether-linked tetracarboxylic acid, 3-(2′,4′-dicarboxylphenoxy)phthalic acid (H4dpa, Scheme 1), and tested it as a principal building block for generating copper(II), manganese(II), and zinc(II) CPs. The use of H4dpa as a linker can be justified by a number of relevant features of this tetracarboxylic acid. These features include (i) up to 9 possible sites for coordination (eight O-carboxylate sites and an O-ether site); (ii) two aromatic rings that can provide certain spatial flexibility and conformational adaptation owing to their separation by O-ether group; (iii) the fact that this carboxylic acid possesses good stability under hydrothermal conditions and remains little-explored in the synthesis of CPs.
Thus, this work reports on the hydrothermal synthesis, characterization, thermal behavior, crystal structures, and catalytic application of 2D CPs derived from H4dpa as a linker and bipy as a crystallization mediator. The obtained products [Cu24-dpa)(bipy)2(H2O)]n∙6nH2O (1), [Mn26-dpa)(bipy)2]n (2), and [Zn24-dpa)(bipy)2(H2O)2]n·2nH2O (3) were also screened as potential catalysts for mild cyanosilylation of benzaldehydes into cyanohydrins. The influence of different reaction conditions and substrate scope was investigated, showing that the zinc(II) CP 3 is a particularly effective and recyclable heterogeneous catalyst.

2. Results and Discussion

2.1. Hydrothermal Synthesis

Aqueous medium mixtures composed of Cu(II), Mn(II), or Zn(II) chlorides with H4dpa as a linker, NaOH as a base for deprotonation of carboxylic acid groups, and 2,2′-bipyridine as a mediator of crystallization were subjected to hydrothermal synthesis (3 days, 160 °C), resulting in the formation of three 2D coordination polymers as crystalline solids. These were formulated as [Cu24-dpa)(bipy)2(H2O)]n·6nH2O (1), [Mn26-dpa)(bipy)2]n (2), and [Zn24-dpa)(bipy)2(H2O)2]n·2nH2O (3) on the basis of standard solid-state characterization methods, namely infrared spectroscopy (IR), elemental analysis (EA), thermogravimetric analysis (TGA), powder (PXRD) and single-crystal X-ray diffraction. All compounds represent 2D metal-organic networks that are driven by ether-bridged tetracarboxylate nodes that show two distinct coordination modes, namely µ4-dpa4− in 1 and 3, or µ6-dpa4− in 2 (Scheme 2).

2.2. Structure of [Cu24-dpa)(bipy)2(H2O)]n·6H2O (1)

The structure of a 2D CP 1 (Figure 1) comprises two Cu(II) centers (Cu1, Cu2), one µ4-dpa4− linker, two bipy moieties, and one terminal H2O ligand per asymmetric unit. The Cu1 center is 4-coordinated and exhibits a distorted {CuN2O2} seesaw geometry. It is formed by two carboxylate O atoms from a pair of µ4-dpa4− ligands and two bipy N atoms (Figure 1a). The Cu2 atom is five-coordinated and features a distorted {CuN2O3} square-pyramidal geometry. It is constructed from two oxygen atoms from two µ4-dpa4− blocks, two bipy N atoms, and a terminal H2O ligand.
The Cu–O [1.942(2)–2.345(3) Å] and Cu–N [1.992(3)–2.058(3) Å] bonds are typical for such a type of compounds [14,18,30]. The dpa4− ligand behaves as a µ4-linker with carboxylate moieties being monodentate (Scheme 2, mode I). In μ4-dpa4−, a dihedral angle between aromatic cycles and a Car–Oether–Car angle are 86.34 and 115.93°, correspondingly. The μ4-dpa4− linkers connect four Cu atoms to assemble a 2D metal-organic layer (Figure 2b) which, after simplification, is described as a mononodal 4-linked net. It possesses a sql (Shubnikov tetragonal plane net) topology with a (44.62) point symbol (Figure 1c) [32,33].

2.3. Structure of [Mn26-dpa)(bipy)2]n (2)

The structure of 2 also shows a 2D coordination polymer network (Figure 2). Per asymmetric unit, there are two manganese(II) atoms, one µ6-dpa4− block, and two bipy ligands. The Mn1 center is five-coordinated and shows a distorted {MnN2O3} trigonal-bipyramidal environment. It is built from three oxygen atoms from three µ6-dpa4− linkers and a pair of bipy N donors (Figure 2a). The six-coordinate Mn2 center is bound by four oxygen atoms from three µ6-dpa4− ligands and a pair of bipy N atoms, thus creating a distorted {MnN2O4} octahedral geometry. The Mn–O [2.087(3)–2.217(4) Å] and Mn–N [2.226(4)–2.277(4) Å] bonds agree with those in related compounds [19,21,22]. The tetracarboxylate dpa4− moiety behaves as a μ6-linker (Scheme 2, mode II) with the carboxylate functionalities being monodentate or µ-bridging bidentate. In μ6-dpa4−, the relevant angles are 79.83° (dihedral angle among aromatic moieties) and 116.37° (Car–Oether–Car). The manganese(II) centers are held together by three carboxylate groups from three μ6-dpa4− blocks, thus generating a dimanganese(II) subunit [Mn1∙Mn2 3.4679(6) Å] (Figure 2b). Such Mn2 subunits are additionally connected by carboxylate groups of μ6-dpa4− to form a 2D metal-organic layer (Figure 2c). Regarding topology, this 2D layer is built from 3-linked Mn1/Mn2 nodes as well as the 6-linked µ6-dpa4− nodes (Figure 2d). The resulting net is thus classified as a binodal 3,6-linked layers of a 3,6L66 topological type [34]. It is described by a (43.612)(43)2 point symbol, wherein the (43.612) and (43) indices correspond to the Mn(II) and µ6-dpa4− nodes, respectively.

2.4. Structure of [Zn24-dpa)(bipy)2(H2O)2]n·2H2O (3)

In the structure of a 2D coordination polymer 3 (Figure 3), there are two zinc(II) atoms (Zn1, Zn2), one µ4-dpa4− block, two bipy moieties, two terminal water ligands and a couple of lattice H2O molecules. Both Zn atoms are five-coordinated, showing distorted {ZnN2O3} square pyramidal geometries. These are constructed from two O donors from a pair of µ4-dpa4− linkers, a water ligand, and a pair of bipy nitrogen atoms. The Zn–O [1.973(3)–2.093(4) Å] and Zn–N [2.119(3)–2.181(3) Å] bond lengths are typical for compounds having an O–Zn–N environment [2,30,35,36]. The dpa4− block adopts a µ4-coordination fashion (Scheme 2, mode I). In µ4-dpa4−, the relevant angles are 72.59° (dihedral angle among two aromatic moieties) and 117.72° (Car–Oether–Car). The µ4-dpa4− blocks interconnect four Zn(II) centers to furnish a 2D layer (Figure 3b). Its topological classification discloses an uninodal 3-linked layer with an hcb (Shubnikov hexagonal plane net/(6, 3)) topology and a (63) point symbol (Figure 3c) [37,38].

2.5. TGA & PXRD

Thermal behavior of compounds 13 was investigated by TGA on gradual heating from 30 to 800 °C under nitrogen flow (Figure S2). CP 1 reveals a loss of six lattice and one coordinated H2O molecules in the 33–104 °C interval (calcd. 13.8%; exptl. 13.5%). After dehydration, the network of 1 maintains its stability until 220 °C. CP 2 does not encompass water as a lattice solvent or ligand and shows thermal stability up to 338 °C. For CP 3, a mass loss in the 86–171 °C window refers to an elimination of two crystallization and two coordinated H2O moieties (calcd. 8.4%; exptl. 8.6%); the obtained sample maintains stability until 218 °C.
To confirm a phase purity, the microcrystalline powders of 13 were analyzed by PXRD (Figure S2). The obtained experimental plots well agree with the patterns calculated from single-crystal X-ray diffraction data, what confirms a purity of the bulk solids 13 prepared via hydrothermal synthetic procedure.

2.6. Cyanosilylation of Benzaldehydes

Coordination polymers 13 were tested as catalysts in the mild, heterogeneous cyanosilylation of benzaldehyde substrates with TMSCN (trimethylsilyl cyanide) [33,35]. 4-Nitrobenzaldehyde was used as model substrate and converted into the corresponding cyanohydrin product, 2-(4-nitrophenyl)-2-[(trimethylsilyl)oxy]acetonitrile (Table 1, Scheme 3). An influence of catalyst loading, solvent type, reaction time, catalyst recycling and substrate scope was investigated.
The cyanosilylation reaction almost does not undergo when catalyst is absent or when using a carboxylic acid ligand or a metal salt precursor as potential catalysts (3−8% yields; Table 1, entries 1−3). The product yields are higher when using 1 (37%, entry 4) and 2 (27%, entry 5) as catalysts, and significantly higher in the presence of 3 (92%, entry 6). Although it is not possible to establish a clear relationship between structural features and catalytic activity of the obtained 2D compounds, we can speculate that a superior activity exhibited by the coordination polymer 3 may eventually be related to the presence of accessible and unsaturated zinc(II) centers on a surface of catalyst particles, together with a higher Lewis acidity of the zinc sites [30,39]. Given a superior activity of zinc(II) derivative 3, the reaction catalyzed by this CP was optimized further.
We found that there is an yield growth from 40 to 92% on increasing the time of reaction in the 1–12 h interval (Table 1, entries 6−12; Figure S7). The catalyst 3 loading of 2, 3, or 4 mol% has also a notable effect as attested by an yield increase from 73 to 92 and 93%, respectively (entries 6, 13, 14). Although dichloromethane (CH2Cl2) was used as a standard solvent to achieve the highest yield (92%), the cyanosilylation reaction also undergoes quite effectively in alternative solvents such as chloroform (CHCl3, 82%), tetrahydrofuran (THF, 68%), acetonitrile (CH3CN, 76%), and methanol (CH3OH, 80%) (Table 1, entries 15−18).
Furthermore, in the reactions catalyzed by 3, substrate scope was studied using different functionalized benzaldehydes (Table 2). The corresponding cyanohydrin products are obtained in yields ranging from 51 to 87%. The substrates with an electron-withdrawing group (R = NO2, Cl) generally show a higher reactivity (Table 2, entries 2−5), which can be explained by an increased substrate electrophilicity. As expected, benzaldehyde (R = H) and substrates with an electron-donating group (R = OH, CH3) reveal lower product yields (entries 6, 7).
Given a remarkable activity of CP 3 in these cyanosilylation reactions, the stability of the catalyst and its possible recycling were studied. Thus, in the end of the cyanosilylation of 4-nitrobenzaldehyde (conditions of entry 6, Table 1), the catalyst was isolated via centrifugation, washed by CH2Cl2, air-dried, and reused in the next reaction runs. These tests indicate that the coordination polymer 3 behaves as a stable catalyst which maintains its activity for a minimum of 4 cyanosilylation cycles, as attested by resembling product yields in the 92−87% range (Figure S8). In addition, PXRD data of the used catalyst (Figure S9) show that its metal-organic structure remains stable.

3. Experimental

3.1. Chemicals & Equipment

All chemicals were obtained from commercial sources. In particular, 3-(2′,4′-Dicarboxylphenoxy)phthalic acid (H4dpa) was acquired from Jinan Henghua Sci. & Tec. Co., Ltd. (Jinan, China). Infrared (IR) spectroscopy measurements (KBr discs) were performed on a Bruker EQUINOX 55 spectrometer (Bruker Corporation, Billerica, MA, USA). Elemental analyses (C, H, N) were carried out on an Elementar Vario EL device (Elementar, Langenselbold, Germany). LINSEIS STA PT1600 thermal analyzer (Linseis Messgeräte GmbH, Selb, Germany) was used for TGA (thermogravimetric analysis) measurements under N2 atmosphere at 10 °C/min heating rate. Excitation/emission spectroscopic data were obtained using an Edinburgh FLS920 fluorescence spectrometer (Edinburgh Instruments, Edinburgh, England). Rigaku-Dmax 2400 diffractometer (Cu-Kα radiation; λ = 1.54060 Å; Rigaku Corporation, Tokyo, Japan) was used to obtain powder X-ray diffraction (PXRD) patterns. Solution 1H NMR (nuclear magnetic resonance) spectra were measured on a JNM ECS 400M spectrometer (JEOL Ltd., Tokyo, Japan).

3.2. Hydrothermal Synthesis & Analytical Data

In a general procedure, metal(II) chloride (0.2 mmol: CuCl2·2H2O (34.1 mg) for 1, MnCl2·4H2O (39.6 mg) for 2, or ZnCl2 (27.3 mg) for 3), H4dpa (0.1 mmol, 34.6 mg), bipy (0.2 mmol, 31.2 mg), NaOH (0.4 mmol, 16.0 mg), and H2O (10 mL) were added into a Teflon-lined stainless steel vessel (volume: 25 mL) and stirred for 15 min at ambient temperature. Then, the vessel was closed and kept in an oven at 160 °C. After 3 days at this temperature, the vessel was gradually (10 °C/h) cooled down to ambient temperature. The reaction mixture was then transferred to a glass flask and the crystals of products were decanted or filtered off, followed by washing with H2O and drying in air to produce compounds 13.
[Cu24-dpa)(bipy)2(H2O)]n·6nH2O (1). Blue block-shaped crystals, yield: 55% based on H4dpa. Calcd for C36H36Cu2N4O16: C 47.63, H 4.00, N 6.17%. Found: C 47.86, H 3.98, N 6.21%. IR (KBr, cm−1): 3748 w, 3082 w, 1599 s, 1568 s, 1497 w, 1475 w, 1448 m, 1368 s, 1288 w, 1240 m, 1164 w, 1137 w, 1084 w, 1062 w, 1032 w, 974 w, 921w, 854 w, 823 w, 770 m, 735w, 663 w.
[Mn26-dpa)(bipy)2]n(2). Yellow block-shaped crystals, yield: 52% based on H4dpa. Calcd for C36H22Mn2N4O9: C 56.56, H 2.90, N 7.33%. Found: C 56.37, H 2.89, N 7.38%. IR (KBr, cm−1): 1639 s, 1598 s, 1569 m, 1470 w, 1438 w, 1376 s, 1240 m, 1182 w, 1157 w,1128 w, 1087 w, 1063 w, 1013 w, 968 w, 923 w, 848 w, 820 w, 770 m, 736 w, 696 w, 675 w, 646 w.
[Zn24-dpa)(bipy)2(H2O)2]n·2nH2O (3). Colorless block-shaped crystals, yield: 48% based on H4dpa. Calcd for C36H30Zn2N4O13: C 50.43, H 3.53, N 6.53%. Found: C 50.64, H 3.55, N 6.51%. IR (KBr, cm−1): 3383 w, 3106 w, 1614 s, 1569 s, 1470 m, 1438 s, 1392 s, 1322 w, 1244 w, 1157 w, 1083 w, 1063 w, 1021 w, 972 w, 918 w, 877 w, 766 m, 733 w, 687 w, 650 w.

3.3. Single-Crystal X-ray Diffraction

For CPs 13, data were collected on a Bruker APEX-II CCD diffractometer (Bruker, Karlsruhe, Germany) (graphite-monochromated CuKα radiation, λ = 1.54178 Å). SADABS was used for semi-empirical absorption corrections. Crystal structures were determined by direct methods, followed by the refinement (full-matrix least-squares on F2) with SHELXS-97 and SHELXL-97 [40]. With an exception of hydrogen atoms, all other atoms were subjected to an anisotropic refinement (full-matrix least-squares on F2). Except hydrogen atoms in water, all H atoms were placed in calculated positions (fixed isotropic thermal parameters); these were taken into account at the final stage of full-matrix least-squares refinement in structure factor calculations. Water hydrogen atoms were found by difference maps and constrained to the respective oxygen centers. In 1, some molecules of solvent are heavily disordered and thus were detached by applying SQUEEZE in PLATON [41]. Elemental and thermogravimetric analyses confirmed the number of crystallization H2O molecules. For 13, final crystal data are summarized in Table 3. Representative bond distances and angles (Table S1) as well as hydrogen bond parameters (Table S2) are listed in Supplementary Materials. CCDC-2043757–2043759 contain crystallographic data for 13.
For topological analysis of 2D layers in 13, an underlying net concept was followed [42]. Such simplified networks were constructed by reducing bridging ligands to their centroids and removing terminal ligands, preserving a ligand–metal center connectivity.

3.4. Catalytic Cyanosilylation

Typical reaction mixtures were prepared as follows: in a small vessel, solid catalyst (typically 3 mol%) was suspended in dichloromethane (2.5 mL) and then an aldehyde substrate (0.50 mmol) and a cyanosilylation agent trimethylsilyl cyanide (1.0 mmol) were added. The reaction was kept under stirring at room temperature (~25 °C) for the desired time. Then, the catalyst was separated by centrifugation and the filtrate was subjected to solvent evaporation under reduced pressure to form a crude solid. This solid was dissolved in CDCl3 and analyzed by 1H NMR spectroscopy for product quantification (for details, see Supplementary Materials, Figures S5 and S6). In the catalyst recycling experiments, the catalyst was centrifuged, washed with CH2Cl2, dried at room temperature, and reused in subsequent steps that were done as described above. Blank tests without any catalyst or using metal salt or H4dpa as catalyst were also carried out for comparative purposes.

4. Conclusions

In this study, we explored a still poorly studied tetracaboxylic acid, 3-(2′,4′-dicarboxylphenoxy)phthalic acid (H4dpa), as a multifunctional linker for the hydrothermal synthesis of new CPs. Hence, a simple synthetic procedure led to the formation of a series of copper(II), manganese(II), and zinc(II) coordination polymers. These compounds feature distinct types of 2D metal-organic layers driven by the µ4-dpa4− or µ6-dpa4− linkers, as well as different topologies ranging from sql (1) and 3,6L66 (2) to hcb (3).
Catalytic activity of 13 was also explored in the mild cyanosilylation of benzaldehyde substrates with TMSCN, showing that the zinc(II) CP 3 functions as an effective and reusable heterogeneous catalyst to give cyanohydrin products in up to 93% yields. The effects of various reactions parameters as well as substrate scope were studied.
As main novelty features of this work, we can highlight (i) an application of H4dpa as an underexplored multifunctional linker for design of new CPs, (ii) a synthesis of three structurally and topologically distinct metal-organic architectures, and (iii) notable catalytic activity of a zinc(II) derivative in the cyanosilylation reactions. Besides, the obtained results widen the applications of coordination polymers in heterogeneous catalysis [43,44]. Further research on assembling related types of CPs and exploring their potential in other catalytic transformations are currently underway.

Supplementary Materials

The following data are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2073-4344/11/2/204/s1. Figure S1: FT-IR spectra, Figure S2: TGA curves, Figures S3 and S9: PXRD patterns, Figure S4: emission spectra, Figures S5–S8: supplementary catalysis data, Tables S1 and S2: selected structural parameters for 13. CCDC-2043757–2043759.

Author Contributions

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

Funding

This work has been supported by Science and Technology Planning Project of Guangzhou (201904010381), the Foundation for Science and Technology (FCT) and Portugal 2020 (projects CEECIND/03708/2017, LISBOA-01-0145-FEDER-029697, PTDC/QUI-QIN/3898/2020, and UID/QUI/00100/2013). This paper has been supported by the RUDN University Strategic Academic Leadership Program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kaskel, S. (Ed.) The Chemistry of Metal-Organic Frameworks, 2 Volume Set: Synthesis, Characterization, and Applications; John Wiley & Sons: Weinheim, Germany, 2016. [Google Scholar]
  2. Batten, S.R.; Neville, S.M.; Turner, D.R. Coordination Polymers: Design, Analysis and Application; RSC: Cambridge, UK, 2009. [Google Scholar]
  3. MacGillivray, L.R.; Lukehart, C.M. (Eds.) Metal-Organic Framework Materials; John Wiley & Sons: Chichester, UK, 2014. [Google Scholar]
  4. Janiak, C. Engineering coordination polymers towards applications. Dalton Trans. 2003, 2003, 2781–2804. [Google Scholar] [CrossRef]
  5. Kitagawa, S.; Kitaura, R.; Noro, S.-I. Functional Porous Coordination Polymers. Angew. Chem. Int. Ed. 2004, 43, 2334–2375. [Google Scholar] [CrossRef] [PubMed]
  6. Tibbetts, I.; Kostakis, G.E. Recent Bio-Advances in Metal-Organic Frameworks. Molecules 2020, 25, 1291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Gómez-Avilés, A.; Muelas-Ramos, V.; Bedia, J.; Rodriguez, J.J.; Belver, C. Thermal post-treatments to enhance the water sta-bility of NH2-MIL-125(Ti). Catalysts 2020, 10, 603. [Google Scholar] [CrossRef]
  8. Yu, L.; Dong, X.L.; Gong, Q.H.; Acharya, S.R.; Lin, Y.H.; Wang, H.; Han, Y.; Thonhauser, T.; Li, J. Splitting mono- and di-branched alkane isomers by a robust aluminum-based metal-organic framework material with optimal pore dimensions. J. Am. Chem. Soc. 2020, 142, 6925–6929. [Google Scholar] [CrossRef] [Green Version]
  9. Wang, H.; Li, J. Microporous Metal–Organic Frameworks for Adsorptive Separation of C5–C6 Alkane Isomers. Acc. Chem. Res. 2019, 52, 1968–1978. [Google Scholar] [CrossRef] [PubMed]
  10. Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent Functional Metal–Organic Frameworks. Chem. Rev. 2011, 112, 1126–1162. [Google Scholar] [CrossRef]
  11. Bedia, J.; Muelas-Ramos, V.; Peñas-Garzón, M.; Gómez-Avilés, A.; Rodriguez, J.J.; Belver, C. A Review on the Synthesis and Characterization of Metal Organic Frameworks for Photocatalytic Water Purification. Catalysts 2019, 9, 52. [Google Scholar] [CrossRef] [Green Version]
  12. Karabach, Y.Y.; Guedes da Silva, M.F.C.; Kopylovich, M.N.; Gil-Hernández, B.; Sanchiz, J.; Kirillov, A.M.; Pombeiro, A.J.L. Self-assembled 3D heterometallic CuII/FeII coordination polymers with octahedral net skeletons: Structural features, molecular magnetism, thermal and oxidation catalytic properties. Inorg. Chem. 2010, 49, 11096–11105. [Google Scholar] [CrossRef]
  13. Xiao, J.-D.; Jiang, H. Metal–Organic Frameworks for Photocatalysis and Photothermal Catalysis. Acc. Chem. Res. 2019, 52, 356–366. [Google Scholar] [CrossRef]
  14. Gu, J.Z.; Wen, M.; Cai, Y.; Shi, Z.F.; Arol, A.S.; Kirillova, M.V.; Kirillov, A.M. Metal−organic architectures assembled from multifunctionalpolycarboxylates: Hydrothermal self-assembly, structures, andcatalytic activity in alkane oxidation. Inorg. Chem. 2019, 58, 2403–2412. [Google Scholar] [CrossRef] [PubMed]
  15. Gu, J.Z.; Wen, M.; Cai, Y.; Shi, Z.F.; Nesterov, D.S.; Kirillova, M.V.; Kirillov, A.M. Cobalt(II) Coordination Polymers Assem-bled from UnexploredPyridine-Carboxylic Acids: Structural Diversity and CatalyticOxidation of Alcohols. Inorg. Chem. 2019, 58, 5875–5885. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, Y.Y.; Huang, C.; Mi, L.W. Metal-organic frameworks as acid- and/or base-functionalized catalysts for tandem reac-tions. Dalton Trans. 2020, 49, 14723–14730. [Google Scholar] [CrossRef] [PubMed]
  17. Agarwal, R.A.; Gupta, A.K.; De, D. Flexible Zn-MOF Exhibiting Selective CO2 Adsorption and Efficient Lewis Acidic Catalytic Activity. Cryst. Growth Des. 2019, 19, 2010–2018. [Google Scholar] [CrossRef]
  18. Liu, J.-J.; Lu, Y.; Lu, W. Metal-dependent photosensitivity of three isostructural 1D CPs based on the 1,1′-bis(3-carboxylatobenzyl)-4,4′-bipyridinium moiety. Dalton Trans. 2020, 49, 4044–4049. [Google Scholar] [CrossRef]
  19. Gu, J.Z.; Cui, Y.H.; Liang, X.X.; Wu, J.; Lv, D.Y.; Kirillov, A.M. Srtructurally distinct metal−organic and H-bonded networks derivedfrom 5-(6-carboxypyridin-3-yl)isophthalic acid: Coordination and template effect of 4,4′-bipyridine. Cryst. Grwoth Des. 2016, 16, 4658–4670. [Google Scholar] [CrossRef]
  20. Han, S.-D.; Chen, Y.-Q.; Zhao, J.-P.; Liu, S.-J.; Miao, X.-H.; Hu, T.-L.; Bu, X.-H. Solvent-induced structural diversities from discrete cup-shaped Co8 clusters to Co8 cluster-based chains accompanied by in situ ligand conversion. Cryst. Eng. Comm. 2014, 16, 753–756. [Google Scholar] [CrossRef]
  21. Zou, X.Z.; Wu, J.; Gu, J.Z.; Zhao, N.; Feng, A.S.; Li, Y. Syntheses of two nickel(II) coordination compounds based on a rigid linear tricarboxylic acid. Chin. J. Inorg. Chem. 2019, 35, 1705–1711. [Google Scholar]
  22. Gu, J.; Gao, Z.; 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]
  23. Li, Y.; Wu, J.; Gu, J.Z.; Qiu, W.D.; Feng, A.S. Temperature-dependent syntheses of two manganese(II) coordination com-pounds based on an ether-bridged tetracarboxylic acid. Chin. J. Struct. Chem. 2020, 39, 727–736. [Google Scholar]
  24. Hou, Y.L.; Peng, Y.L.; Diao, Y.X.; Liu, J.; Chen, L.Z.; Li, D. Side chain induced self-assembly and selective catalytic oxidation activity of copper(I)-coper(II)-N4 complexes. Cryst. Growth Des. 2020, 20, 1237–1241. [Google Scholar] [CrossRef]
  25. Gu, J.Z.; Wan, S.M.; Kirillova, M.V.; Kirillov, A.M. H-bonded and metal(II)-organic architectures assembled from an unex-plored aromatic tricarboxylic acid: Structural variety and functional properties. Dalton Trans. 2020, 49, 7197–7209. [Google Scholar] [CrossRef] [PubMed]
  26. Gregory, R.J.H. Cyanohydrins in Nature and the Laboratory: Biology, Preparations, and Synthetic Applications. Chem. Rev. 1999, 99, 3649–3682. [Google Scholar] [CrossRef] [PubMed]
  27. Brunel, J.M.; Holms, I.P. Chemically catalyzed asymmetric cyanohydrin syntheses. Angew. Chem. Int. Ed. 2004, 43, 2752–2778. [Google Scholar] [CrossRef]
  28. Lacour, M.-A.; Rahier, N.J.; Taillefer, M. Mild and Efficient Trimethylsilylcyanation of Ketones Catalysed by PNP Chloride. Chem. Eur. J. 2011, 17, 12276–12279. [Google Scholar] [CrossRef]
  29. Mowry, D.T. The preparation of nitriles. Chem. Rev. 1948, 42, 189–283. [Google Scholar] [CrossRef]
  30. Karmakar, A.; Paul, A.; Rúbio, G.M.D.M.; Da Silva, M.F.C.G.; Pombeiro, A.J.L. Zinc(II) and Copper(II) Metal-Organic Frameworks Constructed from a Terphenyl-4,4′′-dicarboxylic Acid Derivative: Synthesis, Structure, and Catalytic Application in the Cyanosilylation of Aldehydes. Eur. J. Inorg. Chem. 2016, 2016, 5557–5567. [Google Scholar] [CrossRef]
  31. Xi, Y.M.; Ma, Z.Z.; Wang, L.N.; Li, M.; Li, Z.J. Three-dimensional Ni(II)-MOF containing anasymmetric pyridyl-carboxylate ligand: Catalytic cyanosilylation of aldehydes and inhibits human promyelocytic leukemia cancer cells. J. Clust. Sci. 2019, 30, 1455–1464. [Google Scholar] [CrossRef]
  32. 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]
  33. 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]
  34. Chen, J.; Li, Y.; Gu, J.; Kirillova, M.V.; Kirillov, A.M. Introducing a flexible tetracarboxylic acid linker into functional coordination polymers: Synthesis, structural traits, and photocatalytic dye degradation. New J. Chem. 2020, 44, 16082–16091. [Google Scholar] [CrossRef]
  35. Yu, Y. Auxiliary ligands induced two new Zn(II) compounds with the structural variations from 2D layer to 3D framework: Syntheses, structures and photoluminescent properties. J. Mol. Struct. 2017, 1149, 761–765. [Google Scholar] [CrossRef]
  36. Wang, L.-N.; Zhang, Y.-H.; Jiang, S.; Liu, Z.-Z. Three coordination polymers based on 3-(3′,5′-dicarboxylphenoxy)phthalic acid and auxiliary N-donor ligands: Syntheses, structures, and highly selective sensing for nitro explosives and Fe3+ ions. CrystEngComm 2019, 21, 4557–4567. [Google Scholar] [CrossRef]
  37. 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]
  38. Jaros, S.W.; Da Silva, M.F.C.G.; Florek, M.; Smoleński, P.; Pombeiro, A.J.L.; Kirillov, A.M. Silver(I) 1,3,5-Triaza-7-phosphaadamantane Coordination Polymers Driven by Substituted Glutarate and Malonate Building Blocks: Self-Assembly Synthesis, Structural Features, and Antimicrobial Properties. Inorg. Chem. 2016, 55, 5886–5894. [Google Scholar] [CrossRef] [PubMed]
  39. Cui, X.; Xu, M.-C.; Zhang, L.-J.; Yao, R.-X.; Zhang, X.-M. Solvent-free heterogeneous catalysis for cyanosilylation in a dynamic cobalt-MOF. Dalton Trans. 2015, 44, 12711–12716. [Google Scholar] [CrossRef]
  40. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  41. Van de Sluis, P.; Spek, A.L. Platon Program. Acta Cryst. 1990, 46, 194–201. [Google Scholar]
  42. 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]
  43. Llabrés i Xamena, F.X.; Gascon, J. (Eds.) Metal-Organic Frameworks as Heterogeneous Catalysts; Royal Society of Chemistry: Cambridge, UK, 2013. [Google Scholar]
  44. García, H.; Navalón, S. (Eds.) Metal-Organic Frameworks: Applications in Separations and Catalysis; John Wiley & Sons: Weinheim, Germany, 2018. [Google Scholar]
Scheme 1. Structural formulae of H4dpa and bipy.
Scheme 1. Structural formulae of H4dpa and bipy.
Catalysts 11 00204 sch001
Scheme 2. Coordination modes of µ4- or µ6-dpa4− linkers in 13.
Scheme 2. Coordination modes of µ4- or µ6-dpa4− linkers in 13.
Catalysts 11 00204 sch002
Figure 1. Structural fragments of 1. (a) Coordination environment around Cu(II) atoms; H atoms are omitted for clarity. Symmetry codes: i = x + 1, y, z + 1; ii = x + 1, y, z. (b) 2D metal-organic layer; view along the b axis. (c) Topological view of 2D layer with a sql topology; view along the b axis; Cu centers (green balls), centroids of µ4-dpa4− nodes (gray).
Figure 1. Structural fragments of 1. (a) Coordination environment around Cu(II) atoms; H atoms are omitted for clarity. Symmetry codes: i = x + 1, y, z + 1; ii = x + 1, y, z. (b) 2D metal-organic layer; view along the b axis. (c) Topological view of 2D layer with a sql topology; view along the b axis; Cu centers (green balls), centroids of µ4-dpa4− nodes (gray).
Catalysts 11 00204 g001
Figure 2. Structural fragments of 2. (a) Coordination environment around Mn(II) atoms; H atoms are omitted for clarity. Symmetry codes: i = x + 1, y, z; ii = x + 1, −y + 3/2, z + 1/2. (b) Mn2 subunit. (c) 2D metal-organic layer; view along the b axis. (d) Topological representation of 2D layer with a 3,6L66 topology; view along the b axis; Mn(II) nodes (turquoise balls), centroids of µ6-dpa4− nodes (gray).
Figure 2. Structural fragments of 2. (a) Coordination environment around Mn(II) atoms; H atoms are omitted for clarity. Symmetry codes: i = x + 1, y, z; ii = x + 1, −y + 3/2, z + 1/2. (b) Mn2 subunit. (c) 2D metal-organic layer; view along the b axis. (d) Topological representation of 2D layer with a 3,6L66 topology; view along the b axis; Mn(II) nodes (turquoise balls), centroids of µ6-dpa4− nodes (gray).
Catalysts 11 00204 g002
Figure 3. Structural fragments of 3. (a) Coordination environment around Zn(II) atoms; H atoms are omitted for clarity. Symmetry codes: i = −x, −y + 1, −z; ii = −x + 1/2, y − 1/2, −z + 1/2. (b) 2D metal-organic layer; view along the a axis. (c) Topological representation of 2D layer with an hcb topology; view along the a axis; Zn(II) centers (cyan balls), centroids of µ4-dpa4− nodes (gray).
Figure 3. Structural fragments of 3. (a) Coordination environment around Zn(II) atoms; H atoms are omitted for clarity. Symmetry codes: i = −x, −y + 1, −z; ii = −x + 1/2, y − 1/2, −z + 1/2. (b) 2D metal-organic layer; view along the a axis. (c) Topological representation of 2D layer with an hcb topology; view along the a axis; Zn(II) centers (cyan balls), centroids of µ4-dpa4− nodes (gray).
Catalysts 11 00204 g003
Scheme 3. Cyanosilylation of model substrate(4-nitrobenzaldehyde) into the corresponding cyanohydrin.
Scheme 3. Cyanosilylation of model substrate(4-nitrobenzaldehyde) into the corresponding cyanohydrin.
Catalysts 11 00204 sch003
Table 1. Cyanosilylation of 4-nitrobenzaldehyde with TMSCN catalyzed by coordination polymers. a
Table 1. Cyanosilylation of 4-nitrobenzaldehyde with TMSCN catalyzed by coordination polymers. a
EntryCatalystTime, hCatalyst Loading, mol%SolventYield b, %
1blank12-CH2Cl23
2H3dpna123.0CH2Cl26
3ZnCl2123.0CH2Cl28
41123.0CH2Cl237
52123.0CH2Cl227
63123.0CH2Cl292
7313.0CH2Cl240
8323.0CH2Cl262
9343.0CH2Cl273
10363.0CH2Cl280
11383.0CH2Cl284
123103.0CH2Cl288
133122.0CH2Cl273
143124.0CH2Cl293
153123.0CHCl382
163123.0CH3CN76
173123.0THF68
183123.0CH3OH80
a Conditions: 4-nitrobenzaldehyde (0.5 mmol), TMSCN (1.0 mmol), solvent (2.5 mL), room temperature (~25 °C). b Yields based on 1H NMR (nuclear magnetic resonance) analysis: (moles of product per mol of benzaldehyde substrate) × 100%.
Table 2. Substrate scope in cyanosilylation of various benzaldehydes with TMSCN catalyzed by 3. a
Table 2. Substrate scope in cyanosilylation of various benzaldehydes with TMSCN catalyzed by 3. a
Catalysts 11 00204 i001
EntrySubstrate (R-C6H4CHO)Yield b, %
1R = H61
2R = 2-NO282
3R = 3-NO287
4R = 4-NO292
5R = 4-Cl62
6R = 4-OH56
7R = 4-CH351
a Conditions: functionalized benzaldehyde (0.5 mmol), TMSCN (1.0 mmol), catalyst 3 (3.0 mol.%), CH2Cl2 (2.5 mL), room temperature (~25 °C). b Yields based on 1H NMR analysis: (moles of product per mol of functionalized benzaldehyde substrate) × 100%.
Table 3. Crystal data for CPs 13.
Table 3. Crystal data for CPs 13.
Compound123
Chemical formulaC36H36Cu2N4O16C36H22Mn2N4O9C36H30Zn2N4O13
Molecular weight907.78764.45857.38
Crystal systemMonoclinicMonoclinicMonoclinic
Space groupP21/nP21/cP21/n
a10.3607 (5)10.5312 (3)7.9352 (2)
b/Å35.8638 (11)21.2195 (5)13.1264 (4)
c/Å11.3823 (5)15.1835 (4)34.5048 (10)
α/(°)909090
β/(°)116.385 (6)107.751 (3)94.158 (3)
γ/(°)909090
V3788.8 (3)3231.47 (17)3584.58 (18)
Z444
F (000)162415521752
Crystal size/mm0.25 × 0.23 × 0.210.18 × 0.16 × 0.150.26 × 0.24 × 0.20
θ range for data collection4.508–69.9933.699–66.9923.604–70.060
Limiting indices−12 ≤ h ≤ 12, −43 ≤ k ≤ 30, −13 ≤ l ≤ 13−12 ≤ h ≤ 11, −25 ≤ k≤ 25, −18 ≤ l ≤ 17−8 ≤ h ≤ 9, −1 ≤ 15, −41 ≤ l ≤ 40
Reflections collected/unique (Rint)19,512/7079 (0.0494)20,432/5753 (0.0993)13,030/6658 (0.0337)
Dc/(Mg·cm−3)1.4021.5711.589
μ/mm−11.8926.9172.285
Data/restraints/parameters7079/0/4695753/0/4606658/0/496
Goodness-of-fit on F21.0471.0741.017
Final R indices [(I ≥ 2σ(I))] R1, wR20.0537, 0.14440.0707, 0.16800.0515, 0.1169
R indices (all data) R1, wR20.0710, 0.15460.0901, 0.18000.0619, 0.1238
Largest diff. peak and hole/(e·Å−3)1.195 and −0.6480.859 and −1.0890.857 and −1.096
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Y.; Liang, C.; Zou, X.; Gu, J.; Kirillova, M.V.; Kirillov, A.M. Metal(II) Coordination Polymers from Tetracarboxylate Linkers: Synthesis, Structures, and Catalytic Cyanosilylation of Benzaldehydes. Catalysts 2021, 11, 204. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020204

AMA Style

Li Y, Liang C, Zou X, Gu J, Kirillova MV, Kirillov AM. Metal(II) Coordination Polymers from Tetracarboxylate Linkers: Synthesis, Structures, and Catalytic Cyanosilylation of Benzaldehydes. Catalysts. 2021; 11(2):204. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020204

Chicago/Turabian Style

Li, Yu, Chumin Liang, Xunzhong Zou, Jinzhong Gu, Marina V. Kirillova, and Alexander M. Kirillov. 2021. "Metal(II) Coordination Polymers from Tetracarboxylate Linkers: Synthesis, Structures, and Catalytic Cyanosilylation of Benzaldehydes" Catalysts 11, no. 2: 204. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020204

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