Next Article in Journal
DNA Methylation Episignatures in Neurodevelopmental Disorders Associated with Large Structural Copy Number Variants: Clinical Implications
Next Article in Special Issue
Behavior of Ni20Cr Alloy in Molten Nitrate Salts
Previous Article in Journal
PagGRF11 Overexpression Promotes Stem Development and Dwarfing in Populus
Previous Article in Special Issue
Insight and Recent Advances into the Role of Topography on the Cell Differentiation and Proliferation on Biopolymeric Surfaces
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural Diversity of Mercury(II) Halide Complexes Containing Bis-pyridyl-bis-amide with Bulky and Angular Backbones: Ligand Effect and Metal Sensing

1
Department of Chemistry, Chung-Yuan Christian University, Chung Li, Taoyuan 320, Taiwan
2
Department of Chemistry, National Taiwan Normal University, Taipei 106, Taiwan
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(14), 7861; https://doi.org/10.3390/ijms23147861
Submission received: 20 June 2022 / Revised: 12 July 2022 / Accepted: 15 July 2022 / Published: 16 July 2022
(This article belongs to the Collection Feature Papers in Materials Science)

Abstract

:
Hg(II) halide complexes [HgCl2] 2L1 [L1 = N,N’-bis(3-pyridyl)bicyclo(2,2,2,)oct-7-ene-2,3,5,6-tetracarboxylic diamide), 1, [HgBr2(L1)]n, 2, [HgI2(L1)], 3, [Hg2X4(L2)2] [X = Cl, 4, Br, 5, and I, 6; L2 = N,N’-bis(4-pyridylmethyl)bicyclo(2,2,2,)oct-7-ene-2,3,5,6-tetracarboxylic diamide] and {[HgX2(L3)]⋅H2O}n [X = Cl, 7, Br, 8 and I, 9; L3 = 4,4′-oxybis(N-(pyridine-3-yl)benzamide)] are reported and structurally characterized using single-crystal X-ray diffraction analyses. The linear HgCl2 units of complex 1 are interlinked by the L1 ligands through Hg---N and Hg---O interactions, resulting in 1D supramolecular chains. Complex 2 shows 1D zigzag chains interlinked through the Br---Br interactions to form 1D looped supramolecular chains, while the mononuclear [HgI2L2] molecules of 3 are interlinked through Hg---O and I---I interactions, forming 2D supramolecular layers. Complexes 46 are isomorphous dinuclear metallocycles, and 79 form isomorphous 1D zigzag chains. The roles of the ligand type and the halide anion in determining the structural diversity of 19 is discussed and the luminescent properties of 79 evaluated. Complexes 79 manifest stability in aqueous environments. Moreover, complexes 7 and 8 show good sensing towards Fe3+ ions with low detection limits and good reusability up to five cycles, revealing that the Hg-X---Fe3+ (X = Cl and Br) interaction may have an important role in determining the quenching effect of 7 and 8.

Graphical Abstract

1. Introduction

The investigation of the rational design and synthesis of novel coordination polymers (CPs) continues to be an intense area of research due to their interesting structural diversity and potential industrial applications [1,2,3]. Although many remarkable CPs have been reported, it remains elusive to predict the structural types of the various CPs prepared. While choosing the appropriate metal cations and organic spacers is essential, the structural diversity of CPs is also subject to the identity of the counterions and the reaction conditions involved, such as the metal-to-ligand ratio, the solvent system, and the reaction temperature. In the same way, the halide anions have shown significant influence on the structures of the Hg(II) complexes, but it is difficult to predict which anions give similar or different structures [4,5,6,7,8,9,10]. In some cases, the chloride and bromide anions have no contribution on structural diversity, but the iodide anion has; while in other cases, the halide anions yield the same contribution to the structures [11,12,13]. For example, the reactions of HgX2 (X = Cl, Br and I) with the ligand N,N,N′,N′-tetraisopropyl-3,4-pyridinedicarboxamide afforded 2D CPs for the three anions, whereas those with N,N,N′,N′-tetraisobutyl-3,4-pyridinedicarboxamide gave 2D CPs for the chloride and bromide anions and a dimeric complex for the iodide anion [6]. On the other hand, the complexes formed by the reactions of HgX2 with 2-pyridine piconyl hydrazone, 2-acetylpyridine piconyl hydrazone, or 2-phenylpyridine piconyl hydrazone are all mononuclear, showing interesting weak interactions that differentiate these complexes [7]. However, using the bis-(3-pyridyl)isophthalamide) ligand, bimetallic macrocycles for X = Cl and Br and 1D CP for X = I were produced, respectively [11]. The effect of the halide anion on the structural diversity of the formamidinate-based CPs has also been reported. While the 3D (X = Cl) and 2D (X = Br and I) heteronuclear CPs based on quadruple-bonded dimolybdenum units were obtained from the reactions of [Mo2(4-pyf)4] (4-Hpyf = 4-pyridylformamidine) with HgX2 [12], the reactions of 4-Hpyf with HgX2 afforded a 2D layer for X = Cl and 1D helical chains for X = Br and I, respectively [13].
Previously, we reported several bis-pyridyl-bis-amide (bpba)-based 1D Hg(II) halide CPs [14,15,16,17]. By using the rigid and isomeric 2,2-(1,2-phenylene)-bis(N-pyridin-3-yl)acetamide (1,2-pbpa), 2,2’-(1,3-phenylene)-bis(N-(pyridin-3-yl)acetamide (1,3-pbpa), and 2,2-(1,4-phenylene)-bis(N-(pyridin-3-yl)acetamide (1,4-pbpa), several CPs showing 1D zigzag, helical, mesohelical, and sinusoidal structures were prepared [14,15,16], while using the rigid N,N’-di(pyridin-3-yl)naphthalene-1,4-dicarboxamide (dpndc) afforded isostructural and helical CPs [17]. The structural diversity of these Hg(II) halide CPs containing rigid bpba are thus subject to the ligand types, whereas the role of the halide anions is only suggestive [14,15,16,17]. On the other hand, the pairs of supramolecular isomers with the flexible N,N’-di(3-pyridyl)adipoamide (L), [HgBr2(GAG-L)]n and [HgBr2(AAA-L)]n and [HgI2(GAG-L)]n and [HgI2(AAA-L)]n, exhibit mesohelical, helical, sinusoidal, and helical chains, respectively [18].
Although there are already several organic ligand-supported Hg(II) halide complexes reported, research towards understanding the general effect of the halide anion on their structural diversity is less. Moreover, it is well known that the bpba ligands may easily be tailored to form structures with different flexibility and different shapes [14,15,16,17,18]; thus, in this report, we intend to investigate the effect of the halide anion on the structural types of the Hg(II) halide complexes containing bpba ligands with a bulky and angular backbone. We prepared two bpba ligands with bulky backbones, namely N,N’-bis(3-pyridyl)bicyclo(2,2,2,)oct-7-ene-2,3,5,6-tetracarboxylic diamide (L1), Figure 1a, and N,N’-bis(4-pyridylmethyl)bicyclo(2,2,2,)oct-7-ene-2,3,5,6-tetracarboxylic diamide (L2), Figure 1b, as well as a bpba ligand with an angular backbone, namely 4,4′-oxybis(N-(pyridine-3-yl)benzamide) (L3), Figure 1c. Their corresponding reactions with Hg(II) halide salts were carried out. The synthesis and structures of [HgCl2]⋅2L1, 1, [HgBr2L1]n, 2, [HgI2L1], 3, [Hg2X4(L2)2] (X = Cl, 4; Br, 5; I, 6), and [HgX2(L3) H2O]n (X = Cl, 7, Br, 8 and I, 9) form the subject of this report, and the effect of the ligand type and halide anion on the structural diversity is discussed. The luminescence properties of 7 and 8 provide a unique opportunity to investigate the role of the halide anions in determining the sensing properties of the 1D bpba-based Hg(II) CPs.

2. Results and Discussion

2.1. Crystal Structure of 1

Crystals of 1 conform to triclinic space group Pī with a half Hg(II) ion, one chloride anion, and one L1 ligand in the asymmetric unit. Figure 2a shows the coordination environment of the Hg(II) metal center, which is two-coordinated by two symmetry-related chloride anions [Hg-Cl = Hg-Cl(A) = 2.3055(5) Å], resulting in a linear geometry for the Hg(II) ion [∠Cl-Hg-Cl(A) = 180°]. Moreover, the linear metal units interact with the L1 ligands in two directions, orthogonal to the linear metal unit through Hg---N [Hg---N(1) = Hg---N(1A) = 2.727(1) Å] and Hg---O [Hg---O(3B) = Hg---O(3C) = 3.122(1) Å] interactions, resulting in octahedral fashions for the Hg(II) ions, leading to the formation of a 1D linear supramolecular chain, Figure 2b. The sum of the van der Waals radius of Hg and N is 3.07 Å, and that of Hg and O is 3.10 Å (van der Waals radius: Hg = 1.55, N = 1.52, and O = 1.55 Å).

2.2. Crystal Structure of 2

Crystals of 2 conform to triclinic space group Pī with one Hg(II) ion, two bromide anions, and one L1 ligand in each asymmetric unit. Figure 3a shows the coordination environment of the Hg(II) metal center, which is four-coordinated by two nitrogen atoms from two L1 ligands [Hg-N(1) = 2.373(3); Hg-N(4A) = 2.435(3) Å] and two Br anions [Hg-Br(1) = 2.4472(5); Hg-Br(2) = 2.5180(4) Å], resulting in a distorted tetrahedral geometry (τ4 = 0.81) for the Hg(II) ion, with bond angles of N(1)-Hg-N(4A) = 89.77(11), N(1)-Hg-Br(1) = 110.48(8), N(4A)-Hg-Br(1) = 104.32(8), N(1)-Hg-Br(2) = 102.94(7), N(4A)-Hg-Br(2) = 105.28(8) and Br(1)-Hg-Br(2) = 134.87(2). The Hg(II) cations are connected by the L1 ligands to form 1D zigzag chains, which are further linked by the bromide anions through the Br---Br interactions of 3.646(1) Å, that is significantly shorter than the sum of two van der Waals radii of Br (3.70 Å), resulting in 1D looped supramolecular chains, Figure 3b.

2.3. Crystal Structure of 3

Crystals of 3 conform to triclinic space group Pī with one Hg(II) ion, two iodide anions, and one L1 ligand in each asymmetric unit. Figure 4a shows the coordination environment of the Hg(II) metal center of the mononuclear 3, which is three-coordinated by one nitrogen atom [Hg-N(1) = 2.362(3) Å] and two iodide anions [Hg-I(1) = 2.6272(3) Å; Hg-I(2) = 2.6290(3) Å], resulting in a trigonal planar geometry for the Hg(II) ion [∠I(1)-Hg-I(2) = 151.699(8)°; ∠N(1)-Hg-I(1) = 101.72(6)°; ∠N(1)-Hg-I(2) = 106.19(6)°]. Moreover, the molecules of 3 are interlinked through Hg---O [2.949(3) and 2.844(3) Å] and I---I [3.8791(4) Å] interactions, leading to the formation of a 2D supramolecular layer, Figure 4b. The I---I distance of 3.8791(4) Å is significantly shorter than the sum of two van der Waals radii of the iodo atom, which is 3.96 Å.

2.4. Crystal Structures of 46

Complexes 46 are isomorphous, and their crystals conform to monoclinic space group C2/c with one Hg(II) cation, two halide anions, and one L2 ligand in the asymmetric unit. Figure 5 shows a representative drawing for the dinuclear structures of 46 (X = Cl, 4; Br, 5; I, 6). The coordination environment of the Hg(II) metal center is four-coordinated by two nitrogen atoms from two L2 ligands and two halide anions, resulting in a distorted tetrahedral geometry (τ4 = 0.71 for 1; 0.73 for 2; 0.75 for 3). Table 1 lists the selected bond distances and angles for 46. The dihedral angles between the two phenyl rings of L2 are 59.24, 61.25, and 63.300, respectively. While the Hg-X distances and the dihedral angles increase from 4 to 6, the X-Hg-X angles decrease, showing the size effect of the halide anions. Moreover, the molecules of the dinuclear complexes 46 are linked by the intermolecular C-H---O (H---O = 2.509–2.539 Å, ∠C-H---O = 113.9–143.8° for 4; H---O = 2.522–2.587 Å, ∠C-H---O = 114.2–137.0° for 5; H---O = 2.532–2.638 Å, ∠C-H---O = 113.9–132.7° for 6), and 56 are also linked by the intermolecular C-H---X (H---Br = 3.069, 3.115 Å, ∠C-H---Br = 130.3, 118.5°; H---I = 3.186, 3.281 Å, ∠C-H---I = 132.5, 123.2°) interactions, resulting in 3D supramolecular structures, Figures S1–S3.

2.5. Crystal Structures of 79

Single-crystal X-ray diffraction analysis reveals that the isomorphous complexes 79 are crystallized in the monoclinic space group P21/c with one Hg(II) ion, one L3 ligand, two halide anions, and one co-crystallized water molecule in each asymmetric unit. Figure 6a depicts a representative drawing showing the coordination environment of the Hg(II) centers. The central Hg (II) ions are four-coordinated by two pyridyl nitrogen atoms from two L3 ligands and two halide anions, resulting in tetrahedral geometries, which are further linked by the L3 ligands to form 1D zigzag chains, Figure 6b. Table 2 lists the selected bond distances and angles for 79. While the Hg-N distances are similar, the Hg-X distances increase from Cl to I and the X-Hg-X angles decrease. Moreover, the zigzag chains of 79 are linked by the intermolecular C-H---O (H---O = 2.392, 2.479 Å, ∠C-H---O = 136.3, 141.6° for 7; H---O = 2.395, 2.522 Å, ∠C-H---O = 140.5, 142.8° for 8; H---O = 2.456, 2.599 Å, ∠C-H---O = 143.3, 147.9° for 9) and the intermolecular C-H---X (H---Cl = 2.851, 2.878 Å, ∠C-H---Cl = 150.7, 131.6°; H---Br = 2.916, 2.979 Å, ∠C-H---Br = 158.2, 132.9°; H---I = 3.065, 3.154 Å, ∠C-H---I = 143.3, 133.0°) interactions, as well as the N-H---O (H---O = 1.990 Å, ∠N-H---O = 161.1° for 7; H---O = 2.059 Å, ∠N-H---O = 159.4° for 8; H---O = 2.160 Å, ∠N-H---O = 156.7° for 9) interactions with the amide oxygen atoms, resulting in 3D supramolecular structures, Figures S4–S6.
Some additional parameters were manipulated to probe the structural differences in the isomorphous 79, where distances d1 and d2 are distances from the bridging oxygen atom to the two Hg(II) ions and d3 is the distance between the two Hg(II) ions bridged by the L ligand, and angles θ1, θ2, and θ3 are the dihedral angles differences (Figure 7 and Table 3). The dihedral angles θ2 and θ3 increased from 1 to 3, while the θ1 values are in a reverse order, indicating the effect of the size of the halide anion on the structures.

2.6. Effect of Halide Anion and Ligand Type on Structural Diversity

The structural types of complexes 19 are listed in Table 4. While complexes 13 containing rigid-bulky L1 display different structures that are subject to the nature of the halide anion, the structural diversity of the flexible-bulky L2-based 46 and the angular L3-based 79 is independent of the nature of the halide anion. The halide anions play different structure-determining roles in 19 containing different bpba ligands. The halide anion effect on the structural types of the Hg(II) CPs is thus subject to the identity of the bpba ligand.

2.7. Luminescence Properties

Complexes 79 that contain d10 Hg(II) ions and organic ligand L3 with a large π-conjugated system may exhibit fluorescent properties [19]. Therefore, their solid-state emission spectra were examined at room temperature, as shown in Figure 8 and Table 5. The free L3 ligand shows an emission at 430 nm upon excitation at 330 nm, which could be due to the n → π* or π → π* transitions, while complexes 79 exhibit emission bands at 420, 416, and 400 nm, respectively. Due to the d10 electronic configuration of the Hg(II) metal ion that hardly undergoes either oxidation or reduction, the emissions of 79 may thus result from the organic linkers and are attributable to n → π* or π → π* transitions [20]. Noticeably, the emission wavelengths of 79 are similar with decreasing intensities from Cl, Br, to I, indicating the heavy atom effect of the halide anions [21]. It is noted that no detectable emission can be found for the L1, L2 and their complexes 16, which can be ascribed to the different natures of L1, L2, and L3.

2.8. Mechanochemical Synthesis and Stability of Complexes 79

To obtain complexes 79 efficiently, we studied their mechanochemical synthesis. Manual grinding of mercury(II) halide salts with L3 in methanol/H2O or ethanol/H2O afforded complexes 79, which were verified by characterization using PXRD. As shown in Figures S7–S9, the PXRD patterns of the samples prepared using the solvothermal reactions and mechanochemical reactions matched quite well, indicating the bulk purities of 79. The PXRD patterns of the mechanochemical products were comparatively broad, most probably due to the fact that mechanochemical products generally have lower crystallinity compared with the solvothermal product. Additionally, the mechanochemical method was only successful in the MeOH/H2O and EtOH/H2O solvent systems, while in various solvents such as pure H2O, CH2Cl2, and MeCN, different products were obtained, indicating the solvent selectivity of complexes 79 (Figures S10–S12). The stability of complexes 79 was studied by immersing them into H2O, EtOH, MeOH, CH2Cl2, and MeCN, respectively, for up to 7 days. Complexes 79 were then filtered and dried under vacuum and their PXRD patterns were measured. Figures S13–S15 show that the experimental PXRD patterns matched well with the simulated ones, indicating complexes 79 are stable in these solvents.

2.9. Halide Anion Effect on Metal Sensing

Complexes 7 and 8 provide a unique opportunity to investigate the effect of the halide anion on metal sensing. For the investigations, 25 mg samples of 7 and 8, respectively, was immersed into 10 mL aqueous solutions of nitrate salts M(NO3)x or acetate salts M(OAc)x (M = Ag+, Cd2+, Co2+, Mn2+, Ni2+, Mg2+, Cu2+, and Fe3+). After 10 h, the solids were filtered and their emission spectra were measured. As shown in Figures S16–S18, remarkable luminescence quenching of about 91% for 7 and 90% for 8 were found in the detection of Fe3+ ions. To further explore the quenching effect of Fe3+ ions, the sensing dependence of luminescence intensity on the concentration of Fe3+ was investigated by immersing finely ground samples (25 mg) of 7 or 8 into Fe3+ aqueous solutions with various concentrations (0.1 mM–1 mM) for 10 h. Figures S19 and S20 show that the emission intensities were getting lower and almost completely quenched upon increasing the concentration of Fe3+. Quantitatively, the quenching capacity of the Fe3+ ion can be rationalized by the Stern–Volmer equation: I0/I = 1 + Ksv × [Q], where [Q] is the concentration of Fe3+, Ksv is the quenching constant, and I0 and I are the emission intensities in the absence and presence of Fe3+, respectively [22]. As demonstrated in Figure 9, the titration curves for Fe3+ ions in 7 and 8 are virtually linear at low concentrations, which gave the linear correlation coefficient (R2) of 0.9714 for 7 and 0.9525 for 8, respectively, while the S-V curves at higher concentrations became nonlinear, affording Stern−Volmer constants (Ksv) of 2.48 × 104 M−1 for 7 and 1.2 × 104 M−1 for 8, respectively.
Furthermore, the detection limits were calculated according to the standard equation 3σ/k, where σ is the standard deviation from the blank measurements and k is the absolute value of the calibration curve at a lower concentration [23], giving 7.38 and 24 μM for 7 and 8, respectively. The recyclability test showed no significant changes in the PXRD patterns (Figure 10 and Figure 11), and the luminescence intensities (Figures S21 and S22) for five regeneration cycles were consistent, indicating the reusability of 7 and 8 as sensing materials toward Fe3+. This demonstrates that the luminescence quenching is not due to the framework collapse of 7 and 8, but upon the interactions with the Fe3+ ions. The use of Hg(II) CPs for sensing is rarely seen, and a 1D double-chain {[Hg(L)2]·(ClO4)2}n (L = 1,3,5-tris(benzimidazolylmethyl)benzene) has been reported to show multistimuli-responsive photoluminescence sensing properties toward anions, solvents, and nitroaromatic compounds [24].
Several mechanisms for luminescence quenching such as framework collapse, cation exchange, and interactions between the incoming metal ion and the organic linker that result in competitive absorption of the excitation energies have been suggested [25]. Since the quenching of the luminescence is not due to the framework collapse, the interactions between Fe3+ ions and complexes 7 and 8 are the main reasons leading to the luminescence quenching [26]. The UV-Vis absorption spectrum of Fe3+ in aqueous solution and the corresponding excitation and emission spectra of complexes 7 and 8 are shown in Figures S23 and S24, respectively. Partial overlaps between the absorption spectrum of the Fe3+ ion and the excitation spectra of complexes 7 and 8 are observed, indicating that the excitation energies of 7 and 8 can be partially absorbed by the Fe3+ ions, and the luminescence quenching can most probably be ascribed to competitive energy absorption [25].
Moreover, the Ksv values, 2.48 × 104 M−1 for 7 and 1.2 × 104 M−1 for 8, may indicate that the Fe3+ ion shows a better quenching effect to the chloride complex 7. In addition to the possible interactions between the metal ions and the amide carbonyl oxygen atoms of the L3 ligands [26], the halide anions of 7 and 8 may play an important role in determining the quenching effect. It is well known that the atomic radius of Cl is shorter than that of Br, and comparatively, Cl can be regarded as a harder Lewis base than Br. Since Fe3+ is a hard Lewis acid and interacts stronger with the Cl anion, the larger quenching effect to 7 is attributable to the formation of the stronger Hg-Cl---Fe3+ interaction upon the addition of the Fe3+ ion to complex 7. The different quenching effect exerted by the Fe3+ ion may thus be ascribed to the different Hg-X---Fe3+ (X = Cl and Br) interactions. Energy dispersive X-ray (EDX) analysis of complexes 78 was performed after Fe3+ sensing (Figures S25 and S26), confirming the Fe3+ uptake of 78.

3. Materials and Methods

3.1. General Procedures

Elemental analyses of (C, H, N) were performed on a PE 2400 series II CHNS/O (PerkinElmer Instruments, Shelton, CT, USA) or an Elementar Vario EL-III analyzer (Elementar Analysensysteme GmbH, Hanau, Germany). Infrared spectra were obtained from a JASCO FT/IR-460 plus spectrometer with pressed KBr pellets (JASCO, Easton, MD, USA). Powder X-ray diffraction patterns were carried out with a Bruker D8-Focus Bragg–Brentano X-ray powder diffractometer equipped with a CuKα (λα = 1.54178 Å) sealed tube (Bruker Corporation, Karlsruhe, Germany). The UV-Vis spectrum was performed on a UV-2450 spectrophotometer (Dongguan Hongcheng Optical Products Co., Dongguan, China). Emission spectra were determined with a Hitachi F-4500 fluorescence spectrophotometer (Hitachi, Tokyo, Japan). Energy dispersive X-ray (EDX) analysis was performed by using a JEOL JSM-7600F Ultra-High Resolution Schottky Field Emission Scanning Electron Microscope with Oxford Xmax80 energy dispersive X-ray spectrometer (JEOL, Ltd., Tokyo, Japan).

3.2. Materials

The reagents HgCl2 and HgBr2 were purchased from Acros Organics (Themo Fisher Scientific, NJ, USA) and HgI2 from Aldrich Chemistry Co. (Milwaukee, WI, USA). The solvents CH3OH (99.5%) and CH3CH2OH (99.5%) were purchased from Echo Chemical Co., Ltd. (Toufen, Miaoli, Taiwan). The L1, L2, and L3 ligands were prepared according to published procedures with slight modification [27,28,29].

3.3. Preparations

3.3.1. [HgCl2]⋅2L1, 1

A mixture of HgCl2 (0.027 g, 0.10 mmol), L1 (0.040 g, 0.10 mmol), and 10 mL EtOH was sealed in a 23 mL Teflon-lined stainless steel autoclave, which was heated under autogenous pressure to 120 °C for two days, and then, the reaction system was cooled to room temperature at a rate of 2 °C per hour. The colorless crystals suitable for single-crystal X-ray diffraction were obtained. Yield: 0.043 g (40%). Anal. Calcd for C44H32Cl2HgN8O8 (MW = 1072.26): C, 49.29; H, 3.01; N, 10.45%. Found: C, 48.89; H, 2.88; N, 10.33%. FT-IR (cm−1): 3457 (s), 2363 (w), 2342 (w), 1715 (s), 1639 (m), 1566 (m), 1482 (m), 1429 (m), 1384 (s), 1178 (m), 1098 (w), 1048 (w), 1028 (w), 779 (m), 702 (m), 680 (m), 628 (m).

3.3.2. [HgBr2(L1)]n, 2

Complex 2 was prepared by using similar procedures for 1, except that HgBr2 (0.036 g, 0.10 mmol), L1 (0.040 g, 0.10 mmol), and 10 mL MeOH were used. Colorless crystals were obtained. Yield: 0.073 g (96%). Anal. Calcd for C22H16Br2HgN4O4 (MW = 760.78): C, 34.74; H, 2.11; N,7.37%. Found: C, 34.75; H, 1.98; N, 7.37%. IR (cm−1): 3588 (m), 2361 (m), 2340 (m), 1708 (s), 1484 (m), 1434 (m), 1374 (m), 1316 (w), 1232 (w), 1200 (m), 1080 (w), 1026 (w).

3.3.3. [HgI2(L1)], 3

Complex 3 was prepared by using similar procedures for 1, except that HgI2 (0.045 g, 0.10 mmol), L1 (0.040 g, 0.01 mmol), and 10 mL MeOH were used. Colorless crystals were obtained. Yield: 0.044 g (51%). Anal. Calcd for C22H16HgI2N4O4 (MW = 854.78): C, 30.91; H, 1.89; N, 6.55%. Found: C, 30.76; H, 1.84; N, 6.49%. FT-IR (cm−1): 3442 (s), 1722 (m), 1691 (m), 1638 (m), 1483 (w), 1430 (w), 1373 (m), 1201 (m), 1169 (w), 1050 (m), 776 (w), 729 (w), 695 (w), 680 (w), 575 (w).

3.3.4. [Hg2Cl4(L2)2], 4

Complex 4 was prepared by using similar procedures for 1, except that HgCl2 (0.027 g, 0.10 mmol), L2 (0.043 g, 0.10 mmol), and 10 mL MeOH were used. Colorless crystals were obtained. Yield: 0.053 g (76%). Anal. Calcd for C48H40Cl4Hg2N8O8 (MW = 1399.86): C, 41.14; H, 2.86; N, 8.00%. Found: C, 40.75; H, 2.89; N, 7.91%. IR (cm−1): 3074 (w), 2939 (w), 2361 (w), 1768 (m), 1703 (s), 1612 (m), 1565 (w), 1429 (m), 1400 (m), 1347 (m), 1318 (m), 1220 (w), 1171 (m), 1108 (w), 1010 (w), 914 (m), 877 (w), 800 (w), 771 (m), 731 (w), 667 (w), 633 (m), 585 (w), 492 (m).

3.3.5. [Hg2Br4(L2)2], 5

Complex 5 was prepared by following the similar procedures for 4, except that HgBr2 (0.036 g, 0.10 mmol) was used. Colorless crystals were obtained. Yield: 0.055 g (70%). Anal. Calcd for C48H40Br4Hg2N8O8 (MW = 1577.66): C, 36.55; H, 2.54; N, 7.11%. Found: C, 36.69; H, 2.39; N, 7.08%. IR (cm−1): 3072 (w), 2933 (w), 2361 (w), 1768 (w), 1702 (s), 1611 (m), 1565 (w), 1429 (m), 1399 (m),1346 (m), 1317 (m), 1219 (w), 1170 (m), 1107 (w), 1011 (w), 915 (m), 876 (w), 798 (w), 770 (m), 731 (w), 667 (w), 634 (m), 586 (w), 492 (m).

3.3.6. [Hg2I4(L2)2], 6

Complex 6 was prepared by following the similar procedures for 4, except that HgI2 (0.045 g, 0.10 mmol) was used. Colorless crystals were obtained. Yield: 0.039 g (44%). Anal. Calcd for C48H40Hg2I4N8O8 (MW = 1765.66): C, 32.58; H, 2.26; N, 6.33%. Found: C, 32.72; H, 2.16; N, 6.31%. IR (cm−1): 3736 (w), 3567 (w), 3067 (w), 2967 (w), 2927 (w), 2362 (m), 2340 (w), 1942 (w), 1768 (m), 1701 (s), 1611 (m), 1565 (w), 1427 (m), 1396 (s), 1344 (m), 1316 (m), 1217 (m), 1169 (m), 1070 (m) 1009 (m), 914 (m), 875 (m), 796 (m), 769 (m), 729 (m), 696 (w), 667 (m).

3.3.7. {[HgCl2(L3)]⋅H2O}n, 7

Complex 7 was prepared by using similar procedures for 1, except HgCl2 (0.028 g, 0.10 mmol), L3 (0.043 g. 0.10 mmol) in 8 mL EtOH, and 2 mL H2O were used. Colorless crystals were obtained. Yield: 0.057 g (81%). Anal. Calcd for C24H20Cl2HgN4O4 (MW = 699.93): C, 41.18; H, 2.88; N, 8.00%. Found: C, 41.02; H, 2.28; N, 7.56%. IR (cm−1): 3292 (s), 3046 (w), 1920 (m), 1660 (s), 1535 (s), 1505 (s), 1496 (s), 1414 (m), 1231 (s), 1170 (m), 1116 (s), 1097 (m), 1051 (w), 944 (m), 863 (m), 843 (m), 803 (m), 757 (m), 698 (m), 643 (w), 590 (m), 518 (w), 499 (w).

3.3.8. {[HgBr2(L3)]⋅H2O}n, 8

Complex 8 was prepared by following the similar procedures for 7, except HgBr2 (0.036 g, 0.10 mmol) was used. Colorless crystals were obtained. Yield: 0.058 g (74%). Anal. Calcd for C24H20Br2HgN4O4 (MW = 788.85): C, 36.54; H, 2.55; N, 7.10%. Found: C, 36.49; H, 2.88; N, 7.04%. IR (cm−1): 3280 (s), 3046 (w), 1914 (m), 1656 (s), 1539 (s), 1498 (s), 1479 (s), 1416 (m), 1233 (s), 1174 (m), 1109 (s), 1051 (m), 937 (m), 869 (m), 844 (m), 804 (m), 758 (m), 694 (m), 638 (w), 592 (m), 532 (w), 493 (w).

3.3.9. {[HgI2(L3)]⋅H2O}n, 9

Complex 9 was prepared by following the similar procedures for 7, except HgI2 (0.045 g, 0.10 mmol) was used. Colorless crystals were obtained. Yield: 0.046 g (52%). Anal. Calcd for C24H20HgI2N4O4 (MW = 882.83): C, 32.65; H, 2.28; N, 6.34%. Found: C, 33.18; H, 1.99; N, 6.06%. IR (cm−1): 3297 (s), 3057 (w), 1920 (m), 1653 (s), 1544 (s), 1496 (s), 1480 (s), 1414 (m), 1228 (s), 1171 (m), 1115 (s), 1052 (m), 1048 (w), 942 (m), 866 (m), 842 (m), 803 (m), 759 (m), 697 (m), 636 (w), 585 (m), 535 (w), 499 (w).

3.4. Powder X-ray Analysis

In order to check the phase purity of the product, powder X-ray diffraction (PXRD) experiments were carried out for complexes 19. As shown in Figures S7–S9 and S27–S32, the peak positions of the experimental and simulated PXRD patterns were in good agreement with each other, indicating the bulk purities.

3.5. X-ray Crystallography

Single-crystal X-ray diffraction data for complexes 19 were collected on a Bruker AXS SMART APEX II CCD diffractometer with graphite-monochromated MoKα (λα = 0.71073 Å) radiation at 296 K [30]. Data reduction and absorption correction were performed by using standard methods with well-established computational procedures [31]. Some of the heavier atoms were located by the direct or Patterson method, and the remaining atoms were found in a series of Fourier maps and least-squares refinements, while the hydrogen atoms were added by using the HADD command in SHELXTL. Basic information pertaining to crystal parameters and structure refinement is listed in Table 6.

4. Conclusions

Nine Hg(II) halide complexes containing bpba ligands with bulky and angular spacers were successfully synthesized. Complexes 13 containing the rigid L1 ligands with bulky backbones showed bizarre supramolecular structures that were dependent on the identity of the halide anions, whereas the structural types of 46 containing the flexible L2 ligands with bulky spacer and 79 constructed from the angular L3 ligands showed minimal dependence on the halide anions. The structural diversity of the bpba-based Hg(II) halide complexes and the effect of the halide anion are thus subject to the identities of the bpba ligands. Moreover, the bulkiness and the flexibility of the bpba ligands may also determine the effect of the halide anion. To further investigate the effect of halide anion on the structural diversity of flexible bpba-based Hg(II) complexes, future works can be geared towards the preparation of bpba ligands with more methylene groups [-(CH2)n-] in the backbone that can link the amide groups. Furthermore, reactions of mercury halide salts with flexible bpba ligands with C3 and C4 symmetries may also be investigated to afford Hg(II) complexes with interesting structural topologies. The sensing properties of 78 provide a unique insight into understanding the role of the halide anion in determining the quenching effect of the Hg(II) CPs by the metal ions, and the Hg-X---Fe3+ (X = Cl and Br) interactions may govern the quenching efficiency. Although toxic, Hg(II) halide complexes provide opportunities for the investigation of metal sensing.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms23147861/s1.

Author Contributions

Investigation, M.G. and W.-C.H.; data curation, C.-Y.L., V.L., Y.-H.L. and P.B.S.; review and supervision, C.-H.L. and J.-D.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology, Taiwan: MOST 109-2113-M-033-009.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article or Supplementary Materials.

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. Lustig, W.P.; Mukherjee, S.; Rudd, N.D.; Desai, A.V.; Li, J.; Ghosh, S.K. Metal-organic frameworks: Functional luminescent and photonic materials for sensing applications. Chem. Soc. Rev. 2017, 46, 3242–3285. [Google Scholar] [CrossRef] [PubMed]
  2. Li, B.; Wen, H.-M.; Cui, Y.; Zhou, W.; Qian, G.; Chen, B. Emerging Multifunctional Metal-Organic Framework Materials. Adv. Mater. 2016, 28, 8819–8860. [Google Scholar] [CrossRef] [PubMed]
  3. Wales, D.J.; Grand, J.; Ting, V.P.; Burke, R.D.; Edler, K.J.; Bowen, C.R.; Mintova, S.; Burrows, A.D. Gas sensing using porous materials for automotive applications. Chem. Soc. Rev. 2015, 44, 4290–4321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Morsali, A.; Masoomi, M.Y. Structures and properties of mercury(II) coordination polymers. Coord. Chem. Rev. 2009, 253, 1882–1905. [Google Scholar] [CrossRef]
  5. 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] [Green Version]
  6. Rana, L.K.; Sharma, S.; Hundal, G. First report on crystal engineering of Hg(II) halides with fully substituted 3,4-pyridinedi carboxamides: 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]
  7. 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]
  8. 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]
  9. Mahmoudi, G.; Khandar, A.A.; White, J.; Mitoraj, M.P.; Jena, H.S.; Der Voort, P.V.; Qureshi, N.; Kirillov, A.M.; Robeyns, K.; Safin, D.A. Polar protic solvent-trapping polymorphism of the HgII-hydrazone coordination polymer: Experimental and theoretical findings. CrystEngComm 2017, 19, 3017–3025. [Google Scholar] [CrossRef]
  10. Yuan, Z.Z.; Luo, F.; Song, Y.-M.; Sun, G.-M.; Tian, X.-Z.; Huang, H.-X.; Zhu, Y.; Feng, X.-F.; Luo, M.-B.; Liu, S.-J.; et al. Solvent-induced supramolecular isomers, structural diversity, and unprecedented in situ formation of both inorganic and organic ions in inorganic–organic mercury(ii) complexes. Dalton Trans. 2012, 41, 12670–12673. [Google Scholar] [CrossRef]
  11. Wang, H.; Wang, P.; Huang, C.; Chang, L.; Wu, J.; Hou, H.; Fan, Y. Construction of a series of mercury(II) complexes based on a bis-pyridyl-bis-amide ligand: Effect of counter anions, interactions on the supermolecular structures. Inorg. Chim. Acta 2011, 378, 326–332. [Google Scholar] [CrossRef]
  12. Hsu, W.; Li, Y.-S.; He, H.-Y.; Chen, K.-T.; Wu, H.-S.; Proserpio, D.M.; Chen, J.-D.; Wang, J.-C. Stepwise formation of heteronuclear coordination networks based on quadruple-bonded dimolybdenum units containing formamidinate ligands. CrystEngComm 2014, 16, 7385–7388. [Google Scholar] [CrossRef]
  13. Hsu, W.; Yang, X.-K.; Chhetri, P.M.; Chen, J.-D. Hg(II) coordination polymers based on N,N’-bis(pyridine-4-yl)formamidine. Polymers 2016, 8, 137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. 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]
  15. 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]
  16. Mahat Chhetri, P.; Yang, X.-K.; Yang, C.-Y.; Chen, J.-D. One-dimensional Mercury Halide Coordination Polymers Based on a Semi-rigid N-donor Ligand: Reversible Structural Transformation. Polymers 2019, 11, 436. [Google Scholar] [CrossRef] [Green Version]
  17. Mahat Chhetri, P.; Yang, X.-K.; Chen, J.-D. Isostructural Hg(II) halide coordination polymers: A comparison of powder XRD, IR, emission and Hirshfeld Surface Analysis. J. Mol. Struct. 2021, 1239, 130543. [Google Scholar] [CrossRef]
  18. 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]
  19. 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]
  20. 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]
  21. Mac, M.; Danel, A.; Kizior, K.; Nowak, P.; Karocki, A.; Tokarczyk, B. Investigations of the heavy atom effect occurring in bianthryl and 10,10′-dibromobianthryl. Fluorescence, cyclovoltamperometric and actinometric studies. Phys. Chem. Chem. Phys. 2003, 5, 988–997. [Google Scholar] [CrossRef]
  22. Ge, F.-Y.; Sun, G.-H.; Meng, L.; Ren, S.-S.; Zheng, H.-G. Four New Luminescent Metal–Organic Frameworks as Multifunctional Sensors for Detecting Fe3+, Cr2O72– and Nitromethane. Cryst. Growth Des. 2020, 20, 1898–1904. [Google Scholar] [CrossRef]
  23. Zhang, S.-H.; Zhang, S.-Y.; Li, J.-R.; Huang, Z.-Q.; Yang, J.; Yue, K.-F.; Wang, Y.-Y. Rational synthesis of an ultra-stable Zn(ii) coordination polymer based on a new tripodal pyrazole ligand for the highly sensitive and selective detection of Fe3+ and Cr2O72− in aqueous media. Dalton Trans. 2020, 49, 11201–11208. [Google Scholar] [CrossRef] [PubMed]
  24. Tripathi, S.; Bardhan, D.; Chand, D.K. Multistimuli-Responsive Hydrolytically Stable “Smart” Mercury(II) Coordination Polymer. Inorg. Chem. 2018, 57, 11369–11381. [Google Scholar] [CrossRef]
  25. Pamei, M.; Puzari, A. Luminescent transition metal–organic frameworks: An emerging sensor for detecting biologically essential metal ions. Nano-Struct. Nano-Objects 2019, 19, 100364. [Google Scholar] [CrossRef]
  26. Yang, X.-K.; Lee, W.-T.; Hu, J.-H.; Chen, J.-D. Zn(II) and Cd(II) coordination polymers with a new angular bis-pyridyl-bis-amide: Synthesis, structures and sensing properties. CrystEngComm 2021, 23, 4486–4493. [Google Scholar] [CrossRef]
  27. Yu, Z.-Q.; Pan, M.; Jiang, J.-J.; Liu, Z.-M.; Su, C.-Y. Anion Modulated Structural Diversification in the Assembly of Cd(II) Complexes Based on a Balance-like Dipodal Ligand. Cryst. Growth Des. 2012, 12, 2389–2396. [Google Scholar] [CrossRef]
  28. Ganguly, S.; Parveen, R.; Dastidar, P. Rheoreversible Metallogels Derived from Coordination Polymers. Chem. Asian J. 2018, 13, 1474–1484. [Google Scholar] [CrossRef]
  29. Kumar, G.; Turnbull, M.M.; Thakur, V.; Gupta, S.; Trivedi, M.; Kumar, R.; Husain, A. Solvothermal synthesis, structural characterization, DFT and magnetic studies of a dinuclear paddlewheel Cu(II)-metallamacrocycle. J. Mol. Struct. 2020, 1201, 127193. [Google Scholar] [CrossRef]
  30. Bruker AXS. APEX2, V2008.6, SADABS V2008/1, SAINT V7.60A, SHELXTL V6.14; Bruker AXS Inc.: Madison, WI, USA, 2008. [Google Scholar]
  31. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. 2008, A64, 112–122. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Structures of (a) L1, (b) L2, and (c) L3.
Figure 1. Structures of (a) L1, (b) L2, and (c) L3.
Ijms 23 07861 g001
Figure 2. (a) Coordination environment of Hg(II) ion in 1. Symmetry transformations used to generate equivalent atoms: (A) −x, −y + 2, −z + 1; (B) −x + 1, −y + 1, −z + 1; (C) x − 1, y + 1, z. (b) A depiction showing the 1D supramolecular structure of 1.
Figure 2. (a) Coordination environment of Hg(II) ion in 1. Symmetry transformations used to generate equivalent atoms: (A) −x, −y + 2, −z + 1; (B) −x + 1, −y + 1, −z + 1; (C) x − 1, y + 1, z. (b) A depiction showing the 1D supramolecular structure of 1.
Ijms 23 07861 g002
Figure 3. (a) Coordination environment about the Hg(II) ion in 2. Symmetry transformations used to generate equivalent atoms: (A) x − 1, y, z + 1. (b) A drawing showing the 1D looped supramolecular chain.
Figure 3. (a) Coordination environment about the Hg(II) ion in 2. Symmetry transformations used to generate equivalent atoms: (A) x − 1, y, z + 1. (b) A drawing showing the 1D looped supramolecular chain.
Ijms 23 07861 g003
Figure 4. (a) Coordination environment of Hg(II) ion in 3. Symmetry transformations used to generate equivalent atoms: (A) −x + 1, −y + 1, −z + 2; (B) −x + 2, −y + 1, −z + 2. (b) A drawing showing the 2D supramolecular layer.
Figure 4. (a) Coordination environment of Hg(II) ion in 3. Symmetry transformations used to generate equivalent atoms: (A) −x + 1, −y + 1, −z + 2; (B) −x + 2, −y + 1, −z + 2. (b) A drawing showing the 2D supramolecular layer.
Ijms 23 07861 g004
Figure 5. A representative drawing showing the dinuclear structures of 4 (X = Cl), 5 (X = Br), and 6 (X = I). Symmetry transformations used to generate equivalent atoms: (A) −x + 1/2, −y + 1/2, −z + 1 for 4, −x + 1/2, −y + 3/2, −z + 1 for 5 and 6.
Figure 5. A representative drawing showing the dinuclear structures of 4 (X = Cl), 5 (X = Br), and 6 (X = I). Symmetry transformations used to generate equivalent atoms: (A) −x + 1/2, −y + 1/2, −z + 1 for 4, −x + 1/2, −y + 3/2, −z + 1 for 5 and 6.
Ijms 23 07861 g005
Figure 6. (a) A representative drawing showing the coordination environment about the Hg(II) ion for 79. Symmetry transformations used to generate equivalent atoms: (A) x + 1, −y + 3/2, z + 1/2. (b) A drawing showing the 1D linear chain.
Figure 6. (a) A representative drawing showing the coordination environment about the Hg(II) ion for 79. Symmetry transformations used to generate equivalent atoms: (A) x + 1, −y + 3/2, z + 1/2. (b) A drawing showing the 1D linear chain.
Ijms 23 07861 g006
Figure 7. A schematic diagram defining the distances and the dihedral angle.
Figure 7. A schematic diagram defining the distances and the dihedral angle.
Ijms 23 07861 g007
Figure 8. Excitation and emission spectra of complexes 7–9 and L3.
Figure 8. Excitation and emission spectra of complexes 7–9 and L3.
Ijms 23 07861 g008
Figure 9. The Stern–Volmer plot of I0/I versus Fe3+ ions’ concentration for (a) 7ex = 325 nm) and (b) 8ex = 326 nm).
Figure 9. The Stern–Volmer plot of I0/I versus Fe3+ ions’ concentration for (a) 7ex = 325 nm) and (b) 8ex = 326 nm).
Ijms 23 07861 g009
Figure 10. PXRD patterns before and after treatments with Fe3+ ions for 7. (a) Simulated, (b) experimental, (c) 1st cycle, (d) 2nd cycle, (e) 3rd cycle, (f) 4th cycle, and (g) 5th cycle.
Figure 10. PXRD patterns before and after treatments with Fe3+ ions for 7. (a) Simulated, (b) experimental, (c) 1st cycle, (d) 2nd cycle, (e) 3rd cycle, (f) 4th cycle, and (g) 5th cycle.
Ijms 23 07861 g010
Figure 11. PXRD patterns before and after treatments with Fe3+ ions for 8. (a) Simulated, (b) experimental, (c) 1st cycle, (d) 2nd cycle, (e) 3rd cycle, (f) 4th cycle, and (g) 5th cycle.
Figure 11. PXRD patterns before and after treatments with Fe3+ ions for 8. (a) Simulated, (b) experimental, (c) 1st cycle, (d) 2nd cycle, (e) 3rd cycle, (f) 4th cycle, and (g) 5th cycle.
Ijms 23 07861 g011
Table 1. Selected bond lengths (Å) and angles (°) for complexes 4 (X = Cl), 5 (X = Br), and 6 (X = I).
Table 1. Selected bond lengths (Å) and angles (°) for complexes 4 (X = Cl), 5 (X = Br), and 6 (X = I).
456
Hg-N(1)2.452(4)2.503(4)2.507(3)
Hg-N(4A)2.510(4)2.453(4)2.459(3)
Hg-X(1)2.3554(14)2.4681(8)2.6386(4)
Hg-X(2)2.3461(15)2.4670(8)2.6322(4)
∠N(1)-Hg-N(4A)84.27(13)84.31(15)83.92(12)
∠N(1)-Hg-X(1)100.98(11)98.54(10)99.37(8)
∠N(1)-Hg-X(2)97.22(10)97.14(11)100.10(9)
∠N(4A)-Hg-X(1)97.47(10)101.43(12)102.67(9)
∠N(4A)-Hg-X(2)94.91(10)99.21(12)100.18(9)
∠X(1)-Hg-X(2)158.88(5)155.15(2)151.300(12)
Table 2. Selected bond distance (Å) and bond angles (°) for 7 (X = Cl), 8 (X = Br), and 9 (X = I).
Table 2. Selected bond distance (Å) and bond angles (°) for 7 (X = Cl), 8 (X = Br), and 9 (X = I).
789
Hg-X(1)2.3374(11)2.4771(13)2.6270(6)
Hg-X(2)2.3858(10)2.5213(12)2.6732(6)
Hg-N(4A)2.413(3)2.419(9)2.435(5)
Hg-N(1)2.418(3)2.388(8)2.405(5)
∠X(1)-Hg-X(2)147.79(4)143.29(5)139.976(18)
∠X(1)-Hg-N(4A)105.38(8)106.6(2)107.22(12)
∠X(2)-Hg-N(4A)97.52(7)99.1(2)100.84(12)
∠X(1)-Hg-N(1)104.26(7)106.4(2)106.63(11)
∠X(2)-Hg-N(1)99.16(7)100.2(2)102.16(11)
∠N(4A)-Hg-N(1)87.10(10)88.0(3)88.18(18)
Table 3. Comparison of angles and distances for complexes 79.
Table 3. Comparison of angles and distances for complexes 79.
d1 (Å) d2 (Å) d3 (Å)θ1 (°)θ2 (°)θ3 (°)C-O-C
711.0511.5019.8021.7577.792.11118.3(2)
811.0611.4719.6019.5281.093.21117.3(7)
911.0611.4019.4116.8984.515.94117.3(4)
Table 4. Structural types of 19.
Table 4. Structural types of 19.
ComplexStructure
[HgCl2]⋅2L1, 11D supramolecular chain
[HgBr2(L1)]n, 21D zigzag chain
[HgI2(L1)], 32D supramolecular layer
[Hg2Cl4(L2)2], 4Dinuclear metallocycle
[Hg2Br4(L2)2], 5Dinuclear metallocycle
[Hg2I4(L2)2], 6Dinuclear metallocycle
{[HgCl2(L3)] H2O}n, 71D zigzag chain
{[HgBr2(L3)] H2O}n, 81D zigzag chain
{[HgI2(L3)] H2O}n, 91D zigzag chain
Table 5. Luminescent properties of 79.
Table 5. Luminescent properties of 79.
CompoundExcitation
λex (nm)
Emission
λem (nm)
L3330430
7325420
8326416
9316400
Table 6. Crystallographic data for 19.
Table 6. Crystallographic data for 19.
Compound123
FormulaC44H32Cl2HgN8O8C22H16Br2HgN4O4C22H16HgI2N4O4
Formula weight1072.26760.80854.78
Crystal systemTriclinicTriclinicTriclinic
Space groupPīPīPī
a, Å8.7814(14)6.80779(8)9.7275(5)
b, Å10.5498(16)11.83139(14)11.5101(8)
c, Å10.9631(16)14.71857(16)12.4521(7)
α, °85.121(4)80.4699(6)106.3362(18)
β, °74.166(4)85.6414(6)98.0437(13)
γ, °81.889(4)81.8897(6)114.9454(11)
V, Å3966.2(3)1155.78(2)1158.30(12)
Z122
Dcalc, Mg/m31.8432.1862.451
F(000)530716788
µ (Mo Kα), mm−14.19210.1539.347
Range (2θ) for data collection, deg3.86 ≤ 2θ ≤ 56.733.52 ≤ 2θ ≤ 56.603.57 ≤ 2θ ≤ 52.00
Independent reflections4816
[R(int) = 0.0350]
5727
[R(int) = 0.0298]
4540
[R(int) = 0.0317]
Data/restraints/parameters4816/0/2865727/0/2984540/0/298
quality-of-fit indicator c1.0881.0701.029
Final R indices
[I > 2σ(I)] a,b
R1 = 0.0142,
wR2 = 0.0368
R1 = 0.0289,
wR2 = 0. 696
R1 = 0.0158,
wR2 = 0.0396
R indices (all data)R1 = 0.0142,
wR2 = 0.0368
R1 = 0.0327,
wR2 = 0.0714
R1 = 0.0161,
wR2 = 0.0397
Compound456
FormulaC48H40Cl4Hg2N8O8C48H40Br4Hg2N8O8C48H40Hg2I4N8O8
Formula weight1399.861577.701765.66
Crystal systemMonoclinicMonoclinicMonoclinic
Space groupC2/cC2/cC2/c
a, Å25.8153(6)26.208(2)26.7552(5)
b, Å7.0635(2)7.1272(7)7.2351(1)
c, Å27.1652(7)27.046(3)27.0340(5)
α, °909090
β, °97.5487(12)97.866(5)98.0783(9)
γ, °909090
V, Å34910.5(2)5004.3(8)5181.22(15)
Z444
Dcalc, Mg/m31.8932.0942.264
F(000)270429923280
µ (Mo Kα), mm−16.5259.3838.362
Range (2θ) for data collection, deg3.024 ≤ 2θ ≤ 56.6563.04 ≤ 2θ ≤ 56.893.044 ≤ 2θ ≤ 56.678
Independent reflections6126
[R(int) = 0.0372]
6265
[R(int) = 0.0588]
6439
[R(int) = 0.0371]
Data/restraints/parameters6126/0/3166265/0/2986439/0/317
quality-of-fit indicator c1.0591.0251.033
Final R indices
[I > 2σ(I)] a,b
R1 = 0.0380,
wR2 = 0.0951
R1 = 0.0397,
wR2 = 0.0815
R1 = 0.0294,
wR2 = 0.0674
R indices (all data)R1 = 0.0506,
wR2 = 0.1001
R1 = 0.0717,
wR2 = 0.0904
R1 = 0.0363,
wR2 = 0.0706
Compound789
FormulaC24H20Cl2HgN4O4C24H20 Br2HgN4O4C24H20HgI2N4O4
Formula weight699.93788.85882.83
Crystal systemMonoclinicMonoclinicMonoclinic
Space groupP21/cP21/cP21/c
a, Å18.4292(3)18.2680(16)18.052(3)
b, Å13.6689(2)13.9475(11)14.386(3)
c, Å9.80579.9861(8)10.1560(18)
α, °909090
β, °90.3635(9)91.326(4)91.810(7)
γ, °909090
V, Å32470.09(7)2543.7(4)2636.3(8)
Z444
Dcalc, Mg/m31.8822.0602.224
F(000)135214961640
µ(Mo Kα), mm−16.4869.2308.217
Range (2θ) for data collection, deg3.71 ≤ 2θ ≤ 56.703.67 ≤ 2θ ≤ 52.113.62 ≤ 2θ ≤ 56.70
Independent reflections6145
[R(Int) = 0.0357]
4984
[R(Int) = 0.0512]
6571
[R(Int) = 0.0448]
Data/restraints/parameters6145/0/3244984/0/3216571/0/316
quality-of-fit indicator c1.0481.0441.034
Final R indices
[I > 2σ(I)] a,b
R1 = 0.0285,
wR2 = 0.0689
R1 = 0.0535,
wR2 = 0.1408
R1 = 0.0396,
wR2 = 0.0929
R indices (all data)R1 = 0.0366,
wR2 = 0.0722
R1 = 0.0760,
wR2 = 0.1506
R1 = 0.0592,
wR2 = 0.1008
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 = 0.0141, b = 0.4522 for 1; a = 0.0348, b = 1.304 for 2; a = 0.0127, b = 3.7802 for 3; a = 0.0453; b = 2.4854 for 4; a = 0.0341 b = 5.0712 for 5; a = 0.0296 b = 13.8466 for 6; a = 0.0331, b = 2.1145 for 7; a = 0.0712, b = 13.7591 for 8; a = 0.0468, b = 4.1899 for 9; c quality-of-fit = [Σw(|Fo2| − |Fc2|)2]/(Nobserved − Nparameters)1/2.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Govindaraj, M.; Huang, W.-C.; Lee, C.-Y.; Lakshmanan, V.; Liu, Y.-H.; So, P.B.; Lin, C.-H.; Chen, J.-D. Structural Diversity of Mercury(II) Halide Complexes Containing Bis-pyridyl-bis-amide with Bulky and Angular Backbones: Ligand Effect and Metal Sensing. Int. J. Mol. Sci. 2022, 23, 7861. https://doi.org/10.3390/ijms23147861

AMA Style

Govindaraj M, Huang W-C, Lee C-Y, Lakshmanan V, Liu Y-H, So PB, Lin C-H, Chen J-D. Structural Diversity of Mercury(II) Halide Complexes Containing Bis-pyridyl-bis-amide with Bulky and Angular Backbones: Ligand Effect and Metal Sensing. International Journal of Molecular Sciences. 2022; 23(14):7861. https://doi.org/10.3390/ijms23147861

Chicago/Turabian Style

Govindaraj, Manivannan, Wei-Chun Huang, Chia-Yi Lee, Venkatesan Lakshmanan, Yu-Hsiang Liu, Pamela Berilyn So, Chia-Her Lin, and Jhy-Der Chen. 2022. "Structural Diversity of Mercury(II) Halide Complexes Containing Bis-pyridyl-bis-amide with Bulky and Angular Backbones: Ligand Effect and Metal Sensing" International Journal of Molecular Sciences 23, no. 14: 7861. https://doi.org/10.3390/ijms23147861

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