Next Article in Journal
Topological Methods for Polymeric Materials: Characterizing the Relationship Between Polymer Entanglement and Viscoelasticity
Next Article in Special Issue
A Novel Polymethyl Methacrylate Derivative Grafted with Cationic Iridium(III) Complex Units: Synthesis and Application in White Light-Emitting Diodes
Previous Article in Journal
Influences of Hyperbranched Polyester Modification on the Crystallization Kinetics of Isotactic Polypropylene/Graphene Oxide Composites
Previous Article in Special Issue
Conducting Copper(I/II)-Metallopolymer for the Electrocatalytic Oxygen Reduction Reaction (ORR) with High Kinetic Current Density
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Dimensional Mercury Halide Coordination Polymers Based on A Semi-Rigid N-Donor Ligand: Reversible Structural Transformation

1
Department of Chemistry, Chung Yuan Christian University, Chung-Li 320, Taiwan
2
Department of Chemistry, Amrit Science Campus, Tribhuvan University, Kathmandu 44600, Nepal
*
Author to whom correspondence should be addressed.
Polymers 2019, 11(3), 436; https://doi.org/10.3390/polym11030436
Submission received: 15 February 2019 / Revised: 1 March 2019 / Accepted: 4 March 2019 / Published: 6 March 2019
(This article belongs to the Special Issue Metallopolymer, supramolecular chemistry and materials)

Abstract

:
Four one-dimensional (1D) mercury(II) halide coordination polymers have been synthesized by using a semi-rigid N-donor ligand, 2,2′-(1,4-phenylene)-bis(N-(pyridin-3-yl)acetamide) (1,4-pbpa). While [Hg(1,4-pbpa)Cl2·CH3OH]n, 1, forms a sinusoidal chain, the complexes [Hg(1,4-pbpa)X2]n (X = Cl, 2; Br, 3; I, 4) are helical. The sinusoidal 1 undergoes reversible structural transformation with helical 2 upon removal and uptake of CH3OH, which was accompanied with the conformation adjustment of the 1,4-pbpa ligand from trans anti-anti to trans syn-anti. Pyridyl ring rotation of the 1,4-pbpa ligand that results in the change of the ligand conformation is proposed for the initiation of the structural transformation.

Graphical Abstract

1. Introduction

The rational design and synthesis of novel coordination polymers (CPs) not only afford interesting structural topologies but also extend the range of potential applications in magnetism, luminescence, catalysis, and gas storage [1]. The preparation methods, metal-ligand ratio, and solvent combination play key roles in determining the structural diversity and properties, which can also be observed by the influence of the spacer ligands and metal identities. The structural transformations are intriguing in CPs due to their potential applications in switches and sensors, which can be triggered by exchange of guest molecules, removal and uptake of solvents and external stimuli like heat, light, and mechanical forces [2]. The synthesis and structures of a variety of one-dimensional (1D) mercury(II) CPs have been reported, however, it remains a challenge to elucidate their structure-ligand relationship and thereby the intrinsic properties [3,4,5,6,7,8,9,10,11,12,13,14].
We have made our effort to investigate the structural diversity and properties of the CPs constructed from the bis-pyridyl-bis-amide ligands [10] and several Hg(II) CPs that exhibited structural transformations have been reported [4,5,11]. The HgI2-containing CP [Hg(1,2-pbpa)I2]n [1,2-pbpa = 2,2′-(1,2-phenylene)-bis(N-pyridin-3-yl)acetamide] has been shown to display reversible structural transformation with [Hg(1,2-pbpa)I2·MeOH]n and [Hg(1,2-pbpa)I2·MeCN]n upon adsorption/desorption and exchange of methanol and acetonitrile molecules, which demonstrate the importance of N–H---X (X = halide anion) and Hg---X interactions in the evaluation of structural transformation [4]. Moreover, reversible structural transformations between [Hg(1,3-pbpa)X2]n [X = Br and I; 1,3-pbpa = 2,2′-(1,3-phenylene)-bis(N-(pyridin-3-yl)acetamide] and [Hg(1,3-pbpa)X2·MeCN]n were ascribed to the formation and breaking of the N–H---N hydrogen bonds to the acetonitrile molecules [5]. Although several HgCl2 CPs containing the bis-pyridyl-bis-amide ligands have been prepared [4,5,13,14,15], none of them has been reported to show structural transformation upon removal and uptake of solvent. The HgCl2-containing CPs constructed from 2-(2-hydroxyethyl)pyridine have been found to proceed reversible 1D-2D structural transformation through multiple covalent bond breaking and bond reforming [11].
To investigate the effect of ligand-isomerism on the formation mercury(II) halide-containing CPs as well as the impact on the properties, we have carried out the reactions of 2,2′-(1,4-phenylene)-bis(N-(pyridin-3-yl)acetamide (1,4-pbpa) with the Hg(II) halide salts. Herein, we report the syntheses, structures and emission properties of the 1D CPs [Hg(1,4-pbpa)Cl2·CH3OH]n, 1, and [Hg(1,4-pbpa)X2]n (X = Cl, 2; Br, 3; I, 4). Reversible structural transformation between sinusoidal 1 with the trans anti-anti ligand conformation and helical 2 with the trans syn-anti conformation upon removal and uptake of CH3OH is reported. The factors that govern the reversible structural transformation are also discussed.

2. Experimental Section

2.1. General Procedures

The elemental analyses (C, H, and N) were performed using a PE 2400 series II CHNS/O (PerkinElmer Instruments, Shelton, CT, USA) or an Elementar Vario EL-III analyzer (Elementar Analysensysteme GmbH, Hanau, German). The IR spectra (KBr disk) were recorded on a JASCO FT/IR-460 plus spectrometer ((JASCO, Easton, MD, USA). The powder X-ray diffraction was carried out by using a Bruker D2 PHASER diffractometer (Bruker Corporation, Karlsruhe, Germany) equipped with a CuKα (λ = 1.54 Å) radiation. The solid state emission spectroscopy was done using a Hitachi F-4500 spectrometer (Hitachi, Tokyo, Japan) with excitation slit = 5.0 nm and emission slit = 5.0 nm at room temperature.

2.2. Materials

The reagents 1,4-phenylenediacetic acid was purchased from ACROS Co. (Pittsburgh, PA, USA), and 3-aminopyridine, pyridine, tripenylphosphite, and mercury(II) halide salts from Alfa Aesar Co. (Heysham, UK). The ligand 2,2′-(1,4-phenylene)-bis(N-(pyridin-3-yl)acetamide) (1,4-pbpa) was prepared according to a published procedure [16].

2.3. Preparations

2.3.1. [Hg(1,4-pbpa)Cl2·CH3OH]n, 1

A 20 mL MeOH solution of HgCl2 (0.054 g, 0.20 mmol) was layered on top of a 20 mL THF solution of 1,4-pbpa (0.069 g, 0.20 mmol). After two weeks, colorless crystals of 1 were generated, which were then collected. Yield: 0.032 g (25%). Anal. calcd for C21H22Cl2HgN4O3 (MW = 649.91): C, 38.81; N, 8.62; H, 3.41%. Found: C, 38.48; N, 8.94; H, 2.91%. IR (cm−1): 560(w), 639(w), 701(m), 735(w), 781(w), 812(m), 968(w), 1026(w), 1047(w), 1105(w), 1148(m), 1196(m), 1244(w), 1300(m), 1345(s), 1427(s), 1478(s), 1540(s), 1593(m), 1675(s, C=O), 2918(w), 3030(w), 3063(w), 3131(w), 3190(w), 3303(w, N–H stretch).

2.3.2. [Hg(1,4-pbpa)Cl2]n, 2

Compound 2 was prepared by following the procedure described for 1, except a 20 mL EtOH solution of HgCl2 was used. Yield: 0.030 g (20%). Anal. calcd for C20H18Cl2HgN4O2 (MW = 617.87): C, 38.88; N, 9.07; H, 2.94%. Found: C, 38.95; N, 9.03; H, 2.85%. IR (cm−1): 416(w), 504(m), 557(m), 592(m), 636(m), 701(s), 739(m), 782(m), 811(s), 864(w), 923(w), 961(w), 1023(w), 1051(m), 1108(m), 1149(s), 1204(m), 1248(m), 1274(s), 1299(m), 1353(s), 1428(s), 1480(s), 1548(s), 1587(s), 1673(s, C=O), 2921(w), 3075(w), 3134(m), 3269(s, N–H stretch), 3358(s, N–H stretch).

2.3.3. [Hg(1,4-pbpa)Br2]n, 3

A 20 mL MeOH or EtOH solution of HgBr2 (0.072 g, 0.20 mmol) was layered on top of a 20 mL THF solution of 1,4-pbpa (0.069 g, 0.20 mmol). After two weeks, colorless crystals of 3 were generated, which were then collected. Yield: 0.047 g (33%). Anal. calcd for C20H18Br2HgN4O2 (MW = 706.79): C, 33.99; N, 7.93; H, 2.57%. Found: C, 34.36; N, 7.79; H, 2.22%. IR (cm−1): 416(w), 504(m), 557(m), 592(m), 636(m), 701(s), 739(m), 782(m), 811(s), 864(w), 923(w), 961(w), 1023(w), 1051(m), 1108(m), 1149(s), 1204(m), 1248(m), 1274(s), 1299(m), 1344(s), 1428(s), 1480(s), 1546(s), 1586(s), 1672(s, C=O), 2921(w), 3075(w), 3134(m) , 3262(s, N–H stretch), 3358(s, N–H stretch).

2.3.4. [Hg(1,4-pbpa)I2]n, 4

Compound 4 was prepared by following the procedure described for 3, except HgI2 (0.091 g, 0.20 mmol) was used. Yield: 0.061 g (38%). Anal. calcd for C20H18HgI2N4O2 (MW = 800.77): C, 29.99; N, 6.99; H, 2.27%. Found: C, 30.46; N, 6.95; H, 2.4%. IR (cm−1): 416(w), 504(m), 557(m), 592(m), 636(m), 701(s), 739(m), 782(m), 811(s), 864(w), 923(w), 961(w), 1023(w), 1051(m), 1108(m), 1149(s), 1204(m), 1248(m), 1274(s), 1299(m), 1342(s), 1427(s), 1478(s), 1546(s), 1586(s), 1670(s, C=O), 2921(w), 3075(w), 3134(m) , 3250(s, N–H stretch), 3355(s, N–H stretch).
The purities of complexes 14 have been examined by using the PXRD patterns, Figures S1–S4.

2.4. X-ray Crystallography

A Bruker AXS SMART APEX II CCD diffractometer (Bruker AXS, Madison, WI, USA) that was equipped with a graphite monochromated MoKα (λ = 0.71073 Å) radiation and operated at 50 kV and 30 mA was used to collect the diffraction data for complexes 14. Data reduction was performed by using well-established computational procedures with empirical absorption correction based on “multi-scan” [17]. Patterson or direct method was used to locate the positions of some of the heavier atoms and the remaining atoms were found in a series of alternating difference Fourier maps and least-square refinements, while hydrogen atoms were added by using the HADD command in SHELXTL [18]. The CH3OH molecule of 1 is disordered such that two sets of orientations of the oxygen atom can be shown. Basic crystal parameters and structure refinement are summarized in Table 1.

3. Results and Discussion

3.1. Synthesis

Complexes 14 were prepared by layering the MeOH or EtOH solution of mercury(II) halide on top of the THF solution of 1,4-pbpa. Using the combination of MeOH and THF solvents, the reaction of HgCl2 with 1,4-pbpa afforded the sinusoidal 1 containing the co-crystallized MeOH molecule, while that of EtOH and THF gave helical 2 without solvent co-crystallization. However, either in MeOH or EtOH, reactions of HgBr2 and HgI2 with 1,4-pbpa gave 3 and 4, respectively, without solvent co-crystallization. The powder patterns of helical 3 and 4 prepared from HgBr2 and HgI2 in MeOH/THF and EtOH/THF are shown in Figures S3 and S4, respectively, which indicate that formation of 3 and 4 is irrespective of the solvent system. Scheme 1 depicts the synthetic pathways for 14.

3.2. Structural Descriptions

3.2.1. Structure of 1

Single-crystal X-ray structural analysis shows that the crystals of 1 conform to the monoclinic space group P21/m with a half of Hg(II) ion, a half of 1,4-pbpa ligand, two halves of chloride anion and a half of methanol molecule in an asymmetric unit. Figure 1a depicts a drawing showing the coordination environment of metal ion, which is coordinated with two pyridyl nitrogen atoms [Hg-N = 2.432(3) Å] from two 1,4-pbpa ligands and two chloride anions [Hg–Cl = 2.3448(19) and 2.3700(18) Å], resulting in a distorted tetrahedral geometry (τ4 = 0.75) [19,20]. The 1,4-pbpa ligands bridge the adjacent metal ions to form a 1D sinusoidal chain, Figure 1b. The repeating unit reveals the period of 20.66 Å involving two 1,4-pbpa ligands and two Hg(II) ions, which is slightly less than the Hg---Hg distance of 20.82 Å separated by the 1,4-pbpa ligand, while the amplitude of the sinusoidal chain is 9.04 Å. The dihedral angle of pyridine-pyridine ring is 0° and that of benzene-pyridine ring is 69.44°. Two-dimensional supramolecular structure is stabilized by N–H---O [H---O = 2.16 Å; N---O = 2.96 Å; ∠N–H---O = 153.2°] hydrogen bonding originating from the amide hydrogen atoms to the amide oxygen atoms in the adjacent chains, as shown in Figure 1c. Figure 1d reveals that these interlinked chains stacked along the a axis and the methanol molecules are located in the concave-convex sites, which interact with the methylene hydrogen atoms of 1,4-pbpa ligands through weak C–H---O hydrogen bonds (H---O = 3.14 Å; C---O = 3.98 Å; ∠C–H---O = 146.6°) [21].

3.2.2. Structures of 24

Single crystal X-ray structural analyses revealed that isomorphous crystals of 24 conform to the monoclinic space group P21/c with one Hg(II) ion, one 1,4-pbpa ligand and two halide anions in an asymmetric unit. Figure 2a depicts a representative drawing showing the coordination environment of the metal ion, which is coordinated with two halide ions [Hg–X = 2.3552(1) and 2.3581(9) Å for 2; 2.4805(5) and 2.4846(5) Å for 3; 2.6488(5) and 2.6462(5) Å for 4] and two pyridyl nitrogen atoms [Hg–N = 2.415(2) and 2.452(2) Å for 2; 2.404(3) and 2.455(3) Å for 3; 2.416(5) and 2.477(5) Å for 4] from two 1,4-pbpa ligands, resulting in a distorted tetrahedral geometry (τ4 = 0.74, 2; 0.75, 3; 0.77, 4). Complexes 24 form 1D helical chains having the period of 16.29, 16.38 and 16.57 Å, respectively, involving one 1,4-pbpa ligand and two Hg(II) cations, Figure 2b. The 1D helical chains are supported by the N–H---O [H---O = 2.11 Å; N---O = 2.86 Å; ∠N–H---O = 144.3o, 2; H---O = 2.10 Å; N---O = 2.83 Å; ∠N–H---O = 142.2°, 3; [H---O = 2.07 Å; N---O = 2.80 Å; ∠N–H---O = 142.2°, 4] hydrogen bonds originating from the amide groups in 1,4-pbpa and forming 2D supramolecular structures in a crossover fashion, as shown in Figure 2c.

3.2.3. Ligand Conformation

The conformation of bis-pyridine-bis-amide has been determined by the orientations of C=O or N–H pairs and the orientations of the pyridyl nitrogen and amide oxygen. If the orientation of C=O or N–H pair are on the same direction, it is defined as “cis”, and if the pair is on the opposite direction, it is defined as “trans”. By considering the orientations of pyridyl nitrogen and amide oxygen, syn-syn, anti-anti and syn-anti are also specified [10,22,23]. Accordingly, the conformations of 1,4-pbpa are trans anti-anti in 1, Scheme 2a, and trans syn-anti in 24, Scheme 2b, indicating that the co-crystallized solvents play important role in determining the ligand conformation and, thus, the structural type. The detail of dihedral angles and conformations of 1,4-pbpa in 14 are listed in Table 2.

3.2.4. Structural Comparisons

The structural difference between sinusoidal 1 and helical 2 implies the influence of the cocrystallized MeOH solvent molecules on the structural diversity of the mercury(II) chloride CPs constructed from 1,4-pbpa ligands. A comparison of the structures of 24 indicates that the identity of the halide anions shows no effect on the structural type of the Hg(II) halide CPs based on 1,4-pbpa.
It is worthwhile to investigate the isomeric effect of 1,2-, 1,3-, and 1,4-pbpa ligands on the structural diversity of the Hg(II) halide CPs, Table 3. Solvothermal reactions of HgX2 with 1,2-pbpa in ethanol afforded the isostructural 1D zigzag chains [Hg(1,2-pbpa)X2]n (X = Cl, Br and I), while layering reactions of a ethanolic solution of 1,2-pbpa with a methanolic solution and an acetonitrile solution of HgI2 gave 1D helical chains [Hg(1,2-pbpa)I2·MeOH]n and [Hg(1,2-pbpa)I2·MeCN]n, respectively [4]. On the other hand, solvothermal reactions of HgX2 salts with 1,3-pbpa in acetonitrile afforded the 1D helical chains [Hg(1,3-pbpa)X2]n (X = Cl, Br and I), and the 1D mesohelical chains [Hg(1,3-pbpa)X2·MeCN]n (X = Br and I) were obtained by layering solutions of HgX2 and 1,3-pbpa at room temperature [5]. For those complexes without co-crystallized solvents, the 1,3-pbpa and 1,4-pbpa ligands direct the same structural type of helical chain, which are in marked contrast to the 1,2-pbpa ligands that result in the zigzag chain. Moreover, structural variations are observed upon solvent co-crystallization, leading to the formation of helical chains, mesohelical chains, and sinusoidal chain for the Hg(II) CPs constructed from, 1,2-pbpa, 1,3-pbpa and 1,4-pbpa, respectively. The structural diversity of these pbpa-based Hg(II) halide CPs are thus most probably governed by the co-crystallized solvent molecules and ligand-isomerism of the pbpa ligands, whereas the role of the halide anions is not influential.

3.2.5. Structural Transformation

The core structures of complexes 1 and 2 consist of the same formula and thus can be regarded as a pair of supramolecular isomers with and without the co-crystallized solvent molecules, which provide a unique opportunity for the study of the structural transformation between these two complexes upon removal and uptake of CH3OH. To investigate the structural transformation, we first checked feasibility of the structural change from 1 to 2 by heating 1 at variable temperatures, which was verified by using powder X-ray diffraction (PXRD) patterns. When complex 1 was heated from 50 to 100 °C, the PXRD patterns are significantly different from that of the simulation of 1, indicating structural transformation. Heating 1 at 120 and 150 °C, respectively, afforded PXRD patterns that are comparable to that of the simulation of 2 except some peaks around 2θ = 7.5 and 16°. Moreover, increasing the temperature to 180 °C gave a PXRD pattern well matched with the simulated pattern of 2, Figure 3. On the other hand, immersion of 2 into MeOH for one month showed no significant change on the PXRD pattern, Figure S5. However, when 2 was heated solvothermally at 120 °C, the pattern changed and well matched with that of the simulation of 1, Figure 4. To the best of our knowledge, the reversible structural transformation between 1 and 2 represents the first example of the bis-pyridyl-bis-amide-based HgCl2 CPs. Attempts to investigate the structural transformation by immersing 3 and 4 into MeOH, as well as heating solvothermally, led no significant change on their PXRD patterns, Figures S6–S9.
To propose the possible mechanism for the structural transformation between 1 and 2, we consider the change in ligand conformation. When the sinusoidal 1 was heated to remove the CH3OH molecules, cleavage of the N–H---O hydrogen bonds between the stacking chains, Figure 1c, and formation of the N–H---O hydrogen bonds between the crossover chains, Figure 2c, result in the rearrangement of the ligand conformation from trans anti-anti to trans syn-anti and the formation of helical 2, and vice versa, when 2 was solvothermally heated in CH3OH, the structure of 1 was recovered. Removal of the CH3OH from 1 thus triggers the reorientation of one of the pyridyl rings from “anti” to “syn”, leading to the conversion of the conformation trans anti-anti to trans syn-anti, Scheme 3. The rotation of the pyridyl rings drastically affects the Py-Py dihedral angles from 0 to 48.4° and C–C–C–O torsion angles (θ) from 1.3 to 51.8 and 47.8° at both sides of benzene rings and moderately modifies the Py-Ph, Py-NCO and NCO–NCO dihedral angles, Table 2. On the other hand, solvothermal reaction of 2 reversed the orientation of the pyridyl rings and recovered 1.

3.2.6. Emission Properties

The photoluminescence (PL) property of d10 metal complexes have potential applications as chemical sensors and fluorescent materials [24,25,26,27]. The solid state PL of 14, and 1,4-pbpa were obtained at room temperature, Figure 5. The maximum excitation and emission wavelengths of 14, and 1,4-pbpa are listed in Table 4, showing that the emissions of 14 and 1,4-pbpa appear at 448, 438, 455, 461, and 330 nm upon the excitations at 372, 367, 376, 396 and 273 nm, respectively. Due to the d10 electronic configuration of the Hg(II) metal ion that undergoes hardly either oxidation or reduction, the fluorescence emissions of 14 may result from the organic linkers and are probably attributable to π* → n or π* → π transitions [28]. The red shifts of the emission wavelengths for complexes 24 (Cl < Br < I) may be ascribed to the different identity (electronegativity and size) of the halide anions, while coordination environment of the metal centers and arrangement of linkers also influence the luminescence [29].

4. Conclusions

Four Hg(II) halide CPs based on 1,4-pbpa have been synthesized and structurally characterized. While complex 1 forms a 1D sinusoidal chain, 24 are isostructural 1D helical chains. A comparison of the structures of the Hg(II) halide complexes involving 1,2-, 1,3-, and 1,4-pbpa ligands demonstrates that the co-crystallized solvent molecules and the ligand-isomerism of the spacer ligands play important roles in determining the structural diversity of the pbpa-based Hg(II) halide CPs, whereas the structure-directing role of the halide anion is not influential. Complex 1 undergoes reversible structural transformation with 2 upon removal and uptake of CH3OH, which represents the first example of HgCl2-containing CPs constructed from the bis-pyridyl-bis-amide ligands. The structural transformation can be ascribed to the pyridyl ring rotation of the 1,4-pbpa ligand that rearranges reversibly the ligand conformation between trans anti-anti and trans syn-anti. This study provides an insight into understanding the reversibility of the structural transformation invoked by the conformational change.

Supplementary Materials

Supplementary materials can be found at https://www.mdpi.com/2073-4360/11/3/436/s1. Crystallographic data for 14 have been deposited with the Cambridge Crystallographic Data Centre, CCDC no. 1891685-1891688. Figure S1. Powder X-ray patterns of 1. (a) Simulation and (b) experimental, Figure S2. Powder X-ray patterns of 2. (a) Simulation and (b) experimental; Figure S3. Powder X-ray patterns of 3. (a) Simulation (b) experimental of 3 obtained from THF/MeOH and (c) experimental of 3 obtained from THF/EtOH; Figure S4. Powder X-ray patterns of 4. (a) Simulation (b) experimental of 4 obtained from THF/MeOH and (c) experimental of 4 obtained from THF/EtOH; Figure S5. Powder X-ray patterns of 2. (a) Simulation (b) experimental and (c) in MeOH after one month; Figure S6. Powder X-ray patterns of 3. (a) Simulation (b) experimental and (c) in MeOH after one month; Figure S7. Powder X-ray patterns of 3. (a) Simulation (b) experimental and (c) in solvothermal reaction with MeOH; Figure S8. Powder X-ray patterns of 4. (a) Simulation (b) experimental and (c) in MeOH after one month; Figure S9. Powder X-ray patterns of 4. (a) Simulation (b) experimental and (c) in solvothermal reaction with MeOH.

Author Contributions

Investigation, P.M.C.; data curation, X.-K.Y.; validation, C.-T.Y.; supervision, J.-D.C.

Funding

This research was funded by Ministry of Science and Technology of the Republic of China, grant number MOST 107-2113-M-033-002.

Acknowledgments

We are grateful to the Ministry of Science and Technology of the Republic of China for support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Batten, S.R.; Neville, S.M.; Turner, D.R. Coordination Polymers: Design, Analysis and Application; Royal Society of Chemistry: Cambridge, UK, 2009. [Google Scholar]
  2. Kole, G.K.; Vittal, J.J. Solid-state reactivity and structural transformations involving coordination polymers. Chem. Soc. Rev. 2013, 42, 1755–1775. [Google Scholar] [CrossRef] [PubMed]
  3. Morsali, A.; Masoomi, M.Y. Structures and properties of mercury(II) coordination polymers. Coord. Chem. Rev. 2009, 253, 1882–1905. [Google Scholar] [CrossRef]
  4. Mahat Chhetri, P.; Yang, X.-K.; Chen, J.-D. Solvent-mediated reversible structural transformation of mercury iodide coordination polymers: Role of halide anions. Cryst. Growth Des. 2017, 17, 4801–4809. [Google Scholar] [CrossRef]
  5. Mahat Chhetri, P.; Yang, X.-K.; Chen, J.-D. Mercury halide coordination polymers exhibiting reversible structural transformation. CrystEngComm 2018, 20, 2126–2134. [Google Scholar] [CrossRef]
  6. Mahmoudi, G.; Zaręba, J.K.; Bauzá, A.; Kubicki, M.; Bartyzel, A.; Keramidas, A.D.; Butusov, L.; Mirosławh, B.; Frontera, A. Recurrent supramolecular motifs in discrete complexes and coordination polymers based on mercury halides: Prevalence of chelate ring stacking and substituent effects. CrystEngComm 2018, 20, 1065–1076. [Google Scholar] [CrossRef]
  7. Cheng, P.-C.; Yeh, C.-W.; Hsu, W.; Chen, T.-R.; Wang, H.-W.; Chen, J.-D.; Wang, J.-C. Ag(I) complexes containing flexible N,N′-di(3-pyridyl)adipoamide ligands: Syntheses, structures, ligand conformations, and crystal-to-crystal transformations. Cryst. Growth Des. 2012, 12, 943–953. [Google Scholar] [CrossRef]
  8. Rana, L.K.; Sharma, S.; Hundal, G. First report on crystal engineering of Hg(II) halides with fully substituted 3,4-pyridinedicarboxamides: Generation of two-dimensional coordination polymers and linear zig-zag chains of mercury metal ions. Cryst. Growth Des. 2016, 16, 92–107. [Google Scholar] [CrossRef]
  9. Mahmoudi, G.; Bauzá, A.; Gurbanov, A.V.; Zubkov, F.I.; Maniukiewicz, W.; Rodriguez-Diéguez, A.; López-Torres, E.; Frontera, A. The role of unconventional stacking interactions in the supramolecular assemblies of Hg(II) coordination compounds. CrystEngComm 2016, 18, 9056–9066. [Google Scholar] [CrossRef]
  10. Thapa, K.B.; Chen, J.-D. Crystal engineering of coordination polymers containing flexible bis-pyridyl-bis-amide ligands. CrystEngComm 2015, 17, 4611–4626. [Google Scholar] [CrossRef]
  11. Thapa, K.B.; Hsu, Y.-F.; Lin, H.-C.; Chen, J.-D. Hg(II) supramolecular isomers: Structural transformation and photoluminescence change. CrystEngComm 2015, 17, 7574–7582. [Google Scholar] [CrossRef]
  12. Mobin, S.M.; Srivastava, A.K.; Mathur, P.; Lahiri, G.K. Reversible single-crystal to single-crystal transformations in a Hg(II) derivative. 1D-polymeric chain ⇌ 2D-networking as a function of temperature. Dalton Trans. 2010, 39, 8698–8705. [Google Scholar] [CrossRef] [PubMed]
  13. Burchell, T.J.; Eisler, D.J.; Puddephatt, R.J. Self-assembly using dynamic coordination chemistry and hydrogen bonding: Mercury(II) macrocycles, polymers and sheets. Inorg. Chem. 2004, 43, 5550–5557. [Google Scholar] [CrossRef] [PubMed]
  14. Burchell, T.J.; Puddephatt, R.J. Self-assembly of chiral coordination polymers and macrocycles: A metal template effect on the polymer−macrocycle equilibrium. Inorg. Chem. 2005, 44, 3718–3730. [Google Scholar] [CrossRef] [PubMed]
  15. Hu, H.-L.; Hsu, Y.-F.; Wu, C.-J.; Yeh, C.-W.; Chen, J.-D.; Wang, J.-C. Structural diversity in the d10 metal complexes containing N,N′-di(3-pyridyl)oxamide. Polyhedron 2012, 33, 280–288. [Google Scholar] [CrossRef]
  16. Maity, K.; Kundu, T.; Banerjee, R.; Biradha, K. One-dimensional water cages with repeat units of (H2O)24 resembling pagodane trapped in a 3D coordination polymer: Proton conduction and tunable luminescence emission by adsorption of anionic dyes. CrystEngComm 2015, 17, 4439–4443. [Google Scholar] [CrossRef]
  17. Bruker AXS, APEX2, V2008.6; SAD ABS V2008/1; SAINT+ V7.60A; SHELXTL V6.14; Bruker AXS Inc.: Madison, WI, USA, 2008.
  18. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. 2008, A64, 112–122. [Google Scholar] [CrossRef] [PubMed]
  19. Yang, L.; Powell, D.R.; Houser, R.P. Structural variation in copper(I) complexes with pyridylmethylamide ligands: Structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans. 2007, 9, 955–964. [Google Scholar] [CrossRef] [PubMed]
  20. Okuniewski, A.; Rosiak, D.; Chojnacki, J.; Becker, B. Coordination polymers and molecular structures among complexes of mercury(II) halides with selected 1-benzoylthioureas. Polyhedron 2015, 90, 47–57. [Google Scholar] [CrossRef]
  21. Desiraju, G.R.; Steiner, T. The Weak Hydrogen Bond in Structural Chemistry and Biology; Oxford Science Publications: Oxford, UK, 2001. [Google Scholar]
  22. Chang, M.-N.; Yang, X.-K.; Mahat Chhetri, P.; Chen, J.-D. Metal and ligand effects on the construction of divalent coordination polymers based on bis-pyridyl-bis-amide and polycarboxylate ligands. Polymers 2017, 9, 691. [Google Scholar] [CrossRef]
  23. Thapa, K.B.; Yang, X.-K.; Chen, J.-D. Mg(II) coordination polymers based on flexible isomeric tetracarboxylate ligands: Syntheses, structures, structural transformation and luminescent properties. Polymers 2018, 10, 371. [Google Scholar] [CrossRef]
  24. Kreno, L.E.; Leong, K.; Farha, O.K.; Allendorf, M.; Van Duyne, R.P.; Hupp, J.T. Metal–organic framework materials as chemical sensors. Chem. Rev. 2012, 112, 1105–1125. [Google Scholar] [CrossRef] [PubMed]
  25. Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent functional metal–organic frameworks. Chem. Rev. 2012, 112, 1126–1162. [Google Scholar] [CrossRef] [PubMed]
  26. Sun, S.-S.; Lees, A.J. Transition metal based supramolecular systems: Synthesis, photophysics, photochemistry and their potential applications as luminescent anion chemosensors. Coord. Chem. Rev. 2002, 230, 171–192. [Google Scholar] [CrossRef]
  27. Zhang, K.Y.; Yu, Q.; Wei, H.; Liu, S.; Zhao, Q.; Huang, W. Long-lived emissive probes for time-resolved photoluminescence bioimaging and biosensing. Chem. Rev. 2018, 118, 1770–1839. [Google Scholar] [CrossRef] [PubMed]
  28. Allendorf, M.D.; Bauer, C.A.; Bhakta, R.K.; Houk, R.J.T. Luminescent metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1330–1352. [Google Scholar] [CrossRef] [PubMed]
  29. Zeng, F.; Ni, J.; Wang, Q.; Ding, Y.; Ng, S.W.; Zhu, W.; Xie, Y. Synthesis, structures, and photoluminescence of zinc(II), cadmium(II), and mercury(II) coordination polymers constructed from two novel tetrapyridyl ligands. Cryst. Growth Des. 2010, 10, 1611–1622. [Google Scholar] [CrossRef]
Scheme 1. Synthetic pathways for complexes 14.
Scheme 1. Synthetic pathways for complexes 14.
Polymers 11 00436 sch001
Figure 1. (a) A representative drawing showing coordination environment about the Hg(II) ion for 1. Symmetry transformations used to generate equivalent atoms: (A) x, −y + 3/2, z. (b) A drawing showing the 1D sinusoidal chain. (c) A drawing showing the N–H---O (red dash) hydrogen bonds. (d) A view looking down the a axis.
Figure 1. (a) A representative drawing showing coordination environment about the Hg(II) ion for 1. Symmetry transformations used to generate equivalent atoms: (A) x, −y + 3/2, z. (b) A drawing showing the 1D sinusoidal chain. (c) A drawing showing the N–H---O (red dash) hydrogen bonds. (d) A view looking down the a axis.
Polymers 11 00436 g001aPolymers 11 00436 g001b
Figure 2. (a) A representative drawing showing coordination environment about the Hg(II) ion for 24. Symmetry transformations used to generate equivalent atoms: (A) x-1, y -1, z. (b) A representative drawing showing the 1D helical chain for 24. (c) A drawing showing the N–H---O (red dash) hydrogen bonds that link the chains in a crossover fashion.
Figure 2. (a) A representative drawing showing coordination environment about the Hg(II) ion for 24. Symmetry transformations used to generate equivalent atoms: (A) x-1, y -1, z. (b) A representative drawing showing the 1D helical chain for 24. (c) A drawing showing the N–H---O (red dash) hydrogen bonds that link the chains in a crossover fashion.
Polymers 11 00436 g002
Scheme 2. Ligand conformations in complexes 14. (a) trans anti-anti conformation of 1,4-pbpa in 1 and (b) trans syn-anti conformation of 1,4-pbpa in 24.
Scheme 2. Ligand conformations in complexes 14. (a) trans anti-anti conformation of 1,4-pbpa in 1 and (b) trans syn-anti conformation of 1,4-pbpa in 24.
Polymers 11 00436 sch002
Figure 3. PXRD pattern for structural transformation from 1 to 2. (a) Simulation of 1; (b) experiment of 1; (c) 1 at 50 °C; (d) 1 at 70 °C; (e) 1 at 100 °C; (f) 1 at 120 °C; (g) 1 at 150 °C; (h) 1 at 180 °C; and (i) simulation of 2.
Figure 3. PXRD pattern for structural transformation from 1 to 2. (a) Simulation of 1; (b) experiment of 1; (c) 1 at 50 °C; (d) 1 at 70 °C; (e) 1 at 100 °C; (f) 1 at 120 °C; (g) 1 at 150 °C; (h) 1 at 180 °C; and (i) simulation of 2.
Polymers 11 00436 g003
Figure 4. PXRD pattern for structural transformation from 2 to 1. (a) Simulation of 2; (b) experiment of 2; (c) 2 heated with MeOH at 120 °C in hydrothermal process; and (d) simulation of 1.
Figure 4. PXRD pattern for structural transformation from 2 to 1. (a) Simulation of 2; (b) experiment of 2; (c) 2 heated with MeOH at 120 °C in hydrothermal process; and (d) simulation of 1.
Polymers 11 00436 g004
Scheme 3. Pyridyl ring rotation and torsional angle change in 1 and 2 during structural transformation.
Scheme 3. Pyridyl ring rotation and torsional angle change in 1 and 2 during structural transformation.
Polymers 11 00436 sch003
Figure 5. The solid state photoluminescence (PL) of 14 and 1,4-pbpa.
Figure 5. The solid state photoluminescence (PL) of 14 and 1,4-pbpa.
Polymers 11 00436 g005
Table 1. Crystal data for complexes 14.
Table 1. Crystal data for complexes 14.
Compound1234
FormulaC21H22Cl2HgN4O3C20H18Cl2HgN4O2C20H18Br2HgN4O2C20H18HgI2N4O2
Fw649.91617.87706.79800.77
Crystal SystemMonoclinicMonoclinicMonoclinicMonoclinic
Space GroupP21/mP21/cP21/cP21/c
a, Å5.0705(1)9.3764(5)9.3217(1)9.3679(2)
b, Å20.6636(5)13.3232(7)13.4660(1)13.6718(3)
c, Å11.4320(3)17.1229(10)17.2458(2)17.3825(5)
α,°90909090
β,°92.3088(18)97.490(2)96.741(1)95.555(1)
ϒ,°90909090
V, Å31196.81(5)2120.8(2)2149.83(4)2215.83(9)
Temperature, K296 (2)299 (2)296 (2)296 (2)
Dcal, g cm−31.8031.9352.1842.400
F(000)628118413281472
Z2444
µ(Mo Kα), mm−16.6827.53310.99.756
Range(2θ) for data collection, °4.07 to 56.743.88 to 56.633.84 to 56.603.79 to 56.68
Reflections collected11011194952115420206
Independent reflections3057
[R(int) = 0.0491]
5275
[R(int) = 0.0260]
5335
[R(int) = 0.0353]
5516
[R(int) = 0.0323]
Data/restraints/parameters3057/1/1435275/0/2625335/0/2625516/0/262
Quality-of-fit
Indicator c
1.0091.0321.0221.022
Final R indices
[I>2σ(I)] a,b
R1 = 0.0364,
wR2 = 0.0711
R1 = 0.0255,
wR2 = 0.0492
R1 = 0.0312,
wR2 = 0.0556
R1 = 0.0337,
wR2 = 0.0840
R indices
(all data)
R1 = 0.0558,
wR2 = 0.0783
R1 = 0.0379,
wR2 = 0.0524
R1 = 0.0537,
wR2 = 0.0608
R1 = 0.0421,
wR2 = 0.0882
a R1 = Σ||Fo| − |Fc||/Σ|Fo|. b wR2 = [Σw(Fo2 − Fc2)2/Σw(Fo2)2]1/2. w = 1/[σ2(Fo2) + (ap)2 + (bp)], p = [max(Fo2 or 0) + 2(Fc2)]/3. a = 0.0359, b = 0 for 1; a = 0.0215, b = 0.6032 for 2; a = 0.0220, b = 1.5402 for 3; a = 0.0367. b = 10.3217 for 4. c quality-of-fit = [Σw(|Fo2| − |Fc2|)2/Nobserved − Nparameters )]1/2
Table 2. Selected parameters for complexes 1 and 2.
Table 2. Selected parameters for complexes 1 and 2.
ComplexDihedral Angle/°ConformationStructure
Py-PyPh-PyPy-NCONCO–NCO
1069.428.40trans anti-antisinusoidal
248.457.0, 81.632.6, 22.78.6trans syn-antihelical
Table 3. Structural diversity of these pbpa-based Hg(II) halide CPs.
Table 3. Structural diversity of these pbpa-based Hg(II) halide CPs.
ComplexStructureReference
[Hg(1,2-pbpa)Cl2]nZigzag chain[4]
[Hg(1,2-pbpa)Br2]nZigzag chain[4]
[Hg(1,2-pbpa)I2]nZigzag chain[4]
{[Hg(1,2-pbpa)I2]·MeCN}nHelical chain[4]
{[Hg(1,2-pbpa)I2]·MeOH}nHelical chain[4]
[Hg(1,3-pbpa)Cl2]nHelical chain[5]
[Hg(1,3-pbpa)Br2]nHelical chain[5]
[Hg(1,3-pbpa)I2]nHelical chain[5]
{[Hg(1,3-pbpa)Br2]·MeCN}nMesohelical chain[5]
{[Hg(1,3-pbpa)I2]·MeCN}nMesohelical chain[5]
{[Hg(1,4-pbpa)Cl2]·MeOH}n, 1Sinusoidal chainThis work
[Hg(1,4-pbpa)Cl2]n, 2Helical chainThis work
[Hg(1,4-pbpa)Br2]n, 3Helical chainThis work
[Hg(1,4-pbpa)I2]n, 4Helical chainThis work
Table 4. Excitation and emission of complexes 14 and 1,4-pbpa.
Table 4. Excitation and emission of complexes 14 and 1,4-pbpa.
CompoundExcitation
λmax
Emission
λmax
1372448
2367438
3376455
4396461
1,4-pbpa273330

Share and Cite

MDPI and ACS Style

Mahat Chhetri, P.; Yang, X.-K.; Yang, C.-T.; Chen, J.-D. One-Dimensional Mercury Halide Coordination Polymers Based on A Semi-Rigid N-Donor Ligand: Reversible Structural Transformation. Polymers 2019, 11, 436. https://doi.org/10.3390/polym11030436

AMA Style

Mahat Chhetri P, Yang X-K, Yang C-T, Chen J-D. One-Dimensional Mercury Halide Coordination Polymers Based on A Semi-Rigid N-Donor Ligand: Reversible Structural Transformation. Polymers. 2019; 11(3):436. https://doi.org/10.3390/polym11030436

Chicago/Turabian Style

Mahat Chhetri, Pradhumna, Xiang-Kai Yang, Chih-Tung Yang, and Jhy-Der Chen. 2019. "One-Dimensional Mercury Halide Coordination Polymers Based on A Semi-Rigid N-Donor Ligand: Reversible Structural Transformation" Polymers 11, no. 3: 436. https://doi.org/10.3390/polym11030436

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