Next Article in Journal
Carboxylic Ester Hydrolases in Bacteria: Active Site, Structure, Function and Application
Next Article in Special Issue
Fluorescence Properties and Density Functional Theory Calculation of a Structurally Characterized Heterotetranuclear [ZnII2–SmIII2] 4,4′-Bipy-Salamo-Constructed Complex
Previous Article in Journal
Thermal Cyclic Properties of Ti-Pd-Pt-Zr High-Temperature Shape Memory Alloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Newly Synthesized Heterobimetallic NiII-GdIII Salamo-BDC-Based Coordination Polymer: Structural Characterization, DFT Calculation, Fluorescent and Antibacterial Properties

School of Chemical and Biological Engineering, Lanzhou Jiaotong University, Lanzhou 730070, Gansu, China
*
Author to whom correspondence should be addressed.
Submission received: 25 October 2019 / Revised: 12 November 2019 / Accepted: 12 November 2019 / Published: 14 November 2019
(This article belongs to the Special Issue Fluorescent Complexes)

Abstract

:
A unprecedented hetero-bimetallic 3d-4f BDC-salamo-based coordination polymer, [(L)Ni(BDC)Gd(NO3)(DMF)] was prepared and validated via elemental analyses, IR and UV–Visible absorption spectra, DFT calculation, and X-ray crystallography. The six-coordinated Ni1 ion lies at the N2O2 donor site of the L2− moiety, and one DMF O atom and carboxylate O atom occupy, collectively, the axial positions, and form a twisted octahedron. The nine-coordinated Gd1 ion consists of three oxygen atoms (O12, O13, and O14) of two carboxylate groups, two oxygen atoms (O8 and O9) derived from one bidentate nitrate group, and an O2O2 coordination site (O1, O2, O6, and O5) of the L2− unit, forming a twisted three-capped triangular prism coordination geometry. Compared to the ligand (H2L), the fluorescence intensity decreases due to the coordination of metal ions. Meanwhile, the antibacterial activities are researched.

1. Introduction

The salen-like compounds are an important class of Schiff bases having a C2 axis of symmetry [1,2,3,4,5], and the coordinating group of the salen-like ligands mostly contain two Schiff base N and phenol O atoms on the skeleton [6,7,8,9,10,11] which constitutes a tetradentate N2O2 coordination cavity, and is easy to bind with transition metal ions, alkaline earth metal ions, and rare earth metal ions to obtain various metal complexes such as mononuclear, binuclear, multinuclear, and heteropolynuclear complexes [12,13,14,15,16,17]. Salen and its metal complexes have been researched far and wide [18,19,20,21,22,23] for their excellent catalytic [24,25] and biological activities [26,27,28], magnetic materials [29,30,31,32,33], electrochemical researches [34,35,36,37], supramolecular buildings [38,39,40,41,42], and fluorescence properties [43,44,45,46]. The salamo-like ligands are a very important kind of versatile chelating bisoxime ligands including N2O2-donor cavity [47]. These ligands are synthesized by structural modification of the salen-like ligands, and are generally coordinated with metal ions to obtain multinuclear metal complexes. On account of the high coordination abilities of the phenoxy atoms, many central metal ions could be coordinated to prepare heteromultinuclear metal complexes [48,49,50]. However, transition metal complexes of salamo-based compounds have aroused wide concern on their photophysical properties and reported successively [51,52,53,54,55,56,57], and heterobimetallic 3d-4f salamo-like complexes, especially containing auxiliary ligands, have rarely been reported [58]. It is generally true that the self-assembly process of organometallic complexes is frequently used to construct metal organic framework (MOF) materials [59]. Here, an unprecedented heterobimetallic salamo-like coordination polymer, [(L)Ni(BDC)Gd(NO3)(DMF)], is synthesized, and the photophysical characteristics and antibacterial activities of H2L and its NiII-GdIII polymer are discussed. In addition, the paper also introduces DFT calculations to study the electronic structure and properties of the polymer.

2. Experimental Section

2.1. Materials and Physical Measurements

3-Ethoxybenzaldehyde (97%) was bought from Alfa Aesar and utilized without further purification. Terephthalic acid (BDC) and other analytically pure solvents and reagents were from Tianjin Chemical Reagent Factory. Elemental analyses were detected with an IRIS ER/S-WP-1 ICP atomic emission spectrometer and GmbH VarioEL V3.00 automatic elemental analysis instrument from Berlin, Germany, respectively. Melting points were measured via a micro melting point apparatus manufactured by Beijing Taike Instrument Limited company and were uncorrected. FT-IR spectra were recorded on a Vertex70 FT-IR spectrophotometer, with samples prepared as KBr (500–4000 cm−1) from Bruker, Germany. UV-vis spectra were obtained on a Shimadzu UV-3900 spectrometer from Hitachi, Tokyo, Japan. Fluorescence spectra were obtained on a F-7000 FL 220-240V spectrophotometer from Hitachi, Tokyo, Japan. 1H NMR spectra were determined by a German Bruker AVANCE DRX-400 spectrometer. X-ray structure determination was carried out on a SuperNova Dual (Cu at zero) Eos four-circle diffractometer (Bruker, Germany) and Mo-Kα (λ = 0.71073 Å) ray radiation was monochromated with graphite [60]. Molecular orbital calculations were performed by density functional theory (DFT). The DFT methods used were the gradient-corrected functional proposed by B3LYP; basis sets with SDD were used to expand the Kohn–Sham orbitals. These calculations were done using the Gaussian 09 (Shenzhen, P. R. China.) suite of programs. The conductivity was measured by DOS-110 conductivity meter produced by Shanghai Precision Scientific Instrument Co., Ltd. Powder X-ray diffraction (PXRD) data were recorded on a Rigaku D/Max-2400 X-ray diffractometer with Cu/Kα as the radiation source (λ = 0.15406 nm) in the angular range 2θ = 5–50° at room temperature.

2.2. Preparation and Characterization of H2L

H2L was prepared in the light of a similar method previously reported [61]. The synthetic route to H2L is shown in Scheme 1. An ethanol solution (10 mL) of 1,2-bis(aminooxy)ethane (276.0 mg, 3.0 mmol) was added to an ethanol solution (20 mL) of 2-hydroxy-3-ethoxybenzaldehyde (996.0 mg, 6.0 mmol), and the mixture was stirred at 60~55 °C for 5 h. Then, white crystalline H2L was separated and obtained. Yield: 784.0 mg, 67.3%. m.p. 462~463 K. Anal. calcd for C20H24N2O6 (%): C, 61.84; H, 6.23; N, 7.21. Found (%): C, 61.80; H, 6.30; N, 7.19. 1H NMR (500MHz, CDCl3): 1H NMR (500 MHz, CDCl3) δ 1.48 (t, J = 9.4 Hz, 6H, CH3), 4.12 (q, J = 7.0 Hz, 4H, OCH2), 4.46 (s, 4H, CH2), 6.83 (t, J = 3.1 Hz, 4H, ArH), 6.91 (dd, J = 4.9, 2.4 Hz, 2H, ArH), 8.26 (s, 2H, CH=N), 9.69 (s, 2H, OH).

2.3. Preparation of the Coordination Polymer

A solution of gadolinium nitrate hexahydrate (27.5 mg, 0.05 mmol) and nickel acetate tetrahydrate (12.4 mg, 0.05 mmol) in ethanol (5.0 mL) were added dropwise to a solution (5.0 mL) of H2L in ethanol (5.0 mL), then a solution (5.0 mL) of terephthalic acid (16.6 mg, 0.1 mmol) in DMF was added to the mixed solution. The mixture color changed to green immediately, and the mixture was kept stirring for 1 h. Then the mixture was filtered and the filtrate was obtained. The crystals suitable for X-ray crystallographic analysis were obtained by vapor diffusion of diethyl ether into the resulting filtrate for a few weeks at room temperature. Yield: 54.8%. Anal. Calcd for [(L)Ni(BDC)Gd(NO3)(DMF)] (C31H33GdN4NiO14) (%): C, 41.31; H, 3.69; N, 6,21; Gd, 17.44, Ni, 6.51. Found: C, 41.48; H, 3.54; N, 6.12; Gd, 17.01, Ni, 6.10.

2.4. X-ray Crystallography

The single crystal of the polymer was collected by a SuperNova Dual Eos four-circle diffractometer, monochromatic Mo-Kα radiation (λ = 0.71073 Å) was carried out with a graphite monochromator and then processed with Olex2-2009. The crystal structure was solved with ShelXT-2015 and refined with ShelXL-2015 [62]. The anisotropic thermal parameters are assigned to all non–hydrogen atoms. CCDC 1957944. The crystallographic data for the NiII-GdIII coordination polymer was listed in Table 1.

3. Results and Discussion

3.1. Solubility and Molar Conductance

The NiII-GdIII polymer can be soluble in DMF and DMSO, slightly soluble in acetone, methanol, ethanol, trichloromethane, dichloromethane, and acetonitrile, and insoluble in ethyl ether, n-hexane, ethyl acetate, and water. The molar conductance of the NiII-GdIII polymer dissolved in DMF at 25 °C (1 × 10−3 mol L−1) is 174.2 Ω−1·cm2·mol−1. The molar conductivity result is close to the previously reported 1:2 electrolyte [63]. The NiII-GdIII polymer dissolved in DMF may be comprised of [Gd(DMF)2]3+ cation and [(L)Ni(BDC)]2− and NO3 anions. The observed conductance would then correspond to the following equilibria taking place in solution:
[(L)Ni(BDC)Gd(NO3)(DMF)] ⇌ [(L)Ni(BDC)]2− + [Gd(NO3)(DMF)]2+
[Gd(NO3)(DMF)]2+ + DMF ⇌ [Gd(DMF)2]3+ + NO3

3.2. PXRD Analysis

A PXRD experiment was performed for the NiII-GdIII coordination polymer to confirm whether the crystal structure is truly representative. The PXRD pattern of the NiII-GdIII coordination polymer is depicted in Figure 1. The experimental pattern is in good agreement with the simulated one through detailed comparison. Thus, the as-synthesized sample is pure enough for the further research of spectral characterization and fluorescence property.

3.3. FT-IR Spectra

The infrared spectra of H2L and its NiII-GdIII coordination polymer are given in Figure 2. The C=N stretching vibration band of H2L is observed at ca. 1611 cm−1, while that of the NiII-GdIII coordination polymer is found at ca. 1627 cm−1. The O–H stretching band of H2L is found at ca. 3452 cm−1 that belongs to the phenolic hydroxyl groups. Furthermore, the typical Ar–O stretching band of H2L emerges at approximately 1248 cm−1, and after coordination, the stretching band shifted to a low wave number emerges at approximately 1216 cm−1 [17,24]. In the coordination polymer, there is an unsymmetrical νasCOO stretching vibration band (1549 cm−1) and a symmetric νsCOO stretching band (1496 cm−1) of the terephthalic acid, indicating that the carboxylate is a bridge-shaped structure.

3.4. UV-Vis Spectra

The free ligand H2L and its NiII-GdIII coordination polymer were dissolved and made into 1 × 10−5 M methanol solution, and their UV-Vis absorption spectra are depicted in Figure 3. It can be seen that their UV-Vis absorption spectra are obviously different. The UV-Vis absorption spectrum of H2L possesses two strong absorption peaks at about 270 and 318 nm. The peak at 270 nm could be part of the π–π* transition of the benzene rings, and the peak at 318 nm can be part of the intra-ligand π–π* transition of the C=N bonds [11,37]. Compared with the peaks of H2L, a new absorption peak is found at about 320 nm upon coordination, which is part of the n–π* transition from the filled pπ orbital of the bridging phenolic O to the vacant d-orbital of the NiII ions [28,29].

3.5. Crystal Structure Description

The molecule structure of the NiII-GdIII coordination polymer and coordination geometries of metal(II/III) ions are depicted in Figure 4, and the essential bond lengths and angles are summarized in Table 2. The X-ray crystallographic result reveals that the NiII-GdIII coordination polymer crystallizes in the monoclinic system, space group P 21/n, an asymmetric unit of the coordination polymer contains one wholly deprotonated L2− unit, one NiII ion (Ni1), one GdIII ion (Gd1), one nitrate group, one DMF, and one terephthalic acid molecule.
The six-coordinated Ni1 ion lies at the N2O2 donor site of the ligand L2− unit, and one DMF O atom and carboxylate O atom occupy, collectively, the axial positions, and form a twisted octahedron. The nine-coordinated Gd1 ion consists of three O atoms (O12, O13, and O14) of two carboxylate molecules, two O atoms (O8 and O9) derived from one bidentate nitrate group, and an O2O2 coordination site (O1, O2, O6, and O5) of the ligand L2− unit, forming a twisted three-capped triangular prism coordination geometry. The salamo-like ligand H2L can be self-assembled with some bridging ligands and metal salts to obtain multiple complexes or polymers due to its structural specificity [64]; here, terephthalic acid, as a tetradentate connecting agent, can connect NiII and GdIII ions to form a chain coordination polymer.
Furthermore, the hydrogen bonding interactions of the NiII-GdIII coordination polymer are listed in Table 3. In the structure of the NiII-GdIII coordination polymer, there are seven intra-molecular hydrogen bonding interactions (C7–H7B···O13, C11–H11B···O11, C19–H19B···O13, C20–H20C···O8, C21–H21···O5, C23–H23A···O7, and C27–H27···O13) [65,66,67,68,69,70,71,72,73] (Figure 5a) and four inter-molecular hydrogen bonding interactions (C9–H9...O10, C10–H10A...O9, C19–H19B...O5, and C20–H20B...O3) (Figure 5b), and an infinite 1-D chain supramolecular structure is formed by inter-molecular hydrogen bonding interactions.

3.6. Fluorescence Properties

Photophysical properties of lanthanide complexes of salen-like and salamo-like ligands have been successfully reported previously [74]. The free ligand H2L and its NiII-GdIII coordination polymer were dissolved and made into a 2.5 × 10−5 M methanol solution, and their fluorescence spectra excited at 315 nm are revealed in Figure 6. H2L has the strongest absorption peak at about 404 nm, probably owing to the π–π* transition of the L2− unit. Compared to the free ligand H2L, a stronger emission peak at approximately 404 nm is observed upon coordination, which is attributable to the coordination of the L2− unit and metal ions; the conjugated system upon coordination decreases.

3.7. DFT Calculation

In order to understand the electronic structure and properties of the polymer, a DFT calculation was performed on the polymer using Gaussian 09 [75] (Figure 7). The graph of the highest occupied molecular orbitals (HOMOs) of the polymer indicates that the HOMOs are delocalized, mainly in the orbit of NiII-GdIII ions. LUMOs are relatively limited to the auxiliary ligands of terephthalic acid and NiII-GdIII orbitals. The frontier molecular orbital energy of the polymer is EHOMO = −4.4792 eV and ELUMO = −2.3518 eV. The polymer HOMO–LUMO gap is 2.12177 eV. The molecular orbital energy occupied by the polymer is all negative, indicating molecular chemical stability. The polymer has a lower HOMO–LUMO energy gap (ΔE = ELUMO − EHOMO), and the smaller the ΔE value, the more active the molecule.

3.8. Antibacterial Activities

This paper chooses DMF as the solvent and configured the ligand, nickel acetate, cerium nitrate, the NiII-GdIII polymer, and ampicillin as solutions to be tested. Four sets of different concentrations of DMF solutions were prepared–0.4, 0.8, 1.6, and 3.2 mg/mL. For ease of analysis, we distributed the bacterial inoculant evenly over a flat surface, placing the impregnated plate near the edge of the flat surface and maintaining a certain distance. Ampicillin was used as a positive experiment under the same condition and 200 μL sample was put into LB solid medium, and all samples were incubated at 32 °C for 12 h.
The ampicillin inhibition zone diameter was the best in various samples at a concentration of 3.2 mg/mL. As depicted in Figure 8, the inhibitory regions of DMF, nickel(II) acetate, gadolinium nitrate, H2L, ampicillin, and the coordination polymer dissolved in DMF gradually increased, and show that the coordination polymer dissolved in DMF has a good antibacterial activity. Firstly, due to the heavy metal ion effect force in the polymer, protein structure in the bacteria are destroyed. Secondly, the ligand also destroys part of the bacterial membrane, and the bacteria cannot further divide and multiply and thus die. These observations are similar to the biological activities of previously successfully reported complexes [76,77,78,79].

4. Conclusions

An unprecedented hetero-bimetallic 3d-4f coordination polymer, [(L)Ni(BDC)Gd(NO3)(DMF)] was designed and synthesized, and the single crystal structure was validated via X-ray crystallography. The coordination polymer forms a 1-D supra-molecular chain structure, and the nine-coordinated GdIII ion is located in an O9 coordination environment and forms a twisted three-capped triangular prism coordination geometry. The six-coordinated Ni1 ion lies at the N2O2 donor site of the L2− unit, and one DMF O atom and carboxylate O atom occupy, collectively, the axial positions, and form a twisted octahedron. In addition, the polymer dissolved in DMF has good antibacterial activity due to the heavy metal ion effect of the NiII-GdIII coordination polymer, and a stronger emission peak appears at about 404 nm. The structure of the polymer was rationalized by DFT calculation. In addition, the possible form of the polymer dissolved in DMF was investigated by measuring the molar conductivity.

Author Contributions

W.-K.D. conceived and designed the experiments; Y.-F.C. and Y.Z. performed the experiments; Y.Z. analyzed the data; K.-F.X., formal analysis; W.-K.D. contributed reagents/materials/analysis tools; W.-K.D. and Y.-F.C. wrote the paper.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 21761018), the Science and Technology Program of Gansu Province (Grant No. 18YF1GA057) and the Program for Excellent Team of Scientific Research in Lanzhou Jiaotong University (Grant No. 201706.

Acknowledgments

Computations were done using National Supercomputing Center in Shenzhen, P. R. China.

Conflicts of Interest

The authors declare no competing financial interests.

References

  1. Akine, S.; Varadi, Z.; Nabeshima, T. Synthesis of planar metal complexes and the stacking abilities of naphthaenediol-based acyclic and macrocyclic salen-type ligands. Eur. J. Inorg. Chem. 2013, 35, 5987–5998. [Google Scholar] [CrossRef]
  2. Kang, Q.P.; Li, X.Y.; Wei, Z.L.; Zhang, Y.; Dong, W.K. Symmetric containing-PMBP N2O2-donors nickel (II) complexes: Syntheses, structures, Hirshfeld analyses and fluorescent properties. Polyhedron 2019, 165, 38–50. [Google Scholar] [CrossRef]
  3. Wu, H.L.; Bai, Y.C.; Zhang, Y.H.; Pan, G.L.; Kong, J.; Shi, F.; Wang, X.L. Two lanthanide (III) complexes based on the schiff base N, N-Bis (salicylidene)-1,5-diamino-3-oxapentane: Synthesis, characterization, DNA-binding properties, and antioxidation. Z. Anorg. Allg. Chem. 2014, 640, 2062–2071. [Google Scholar] [CrossRef]
  4. Kang, Q.P.; Li, X.Y.; Wang, L.; Zhang, Y.; Dong, W.K. Containing-PMBP N2O2-donors transition metal (II) complexes: Synthesis, crystal structure, Hirshfeld surface analyses and fluorescence properties. Appl. Organomet. Chem. 2019, 33, e5013. [Google Scholar] [CrossRef]
  5. Wu, H.L.; Bai, Y.C.; Zhang, Y.H.; Li, Z.; Wu, M.C.; Chen, C.Y.; Zhang, J.W. Synthesis, crystal structure, antioxidation and DNA-binding properties of a dinuclear copper (II) complex with bis (N-salicylidene)-3-oxapentane-1, 5-diamine. J. Coord. Chem. 2014, 67, 3054–3066. [Google Scholar] [CrossRef]
  6. Gao, L.; Wang, F.; Zhao, Q.; Zhang, Y.; Dong, W.K. Mononuclear Zn(II) and trinuclear Ni(II) complexes derived from a coumarin-containing N2O2 ligand: Syntheses, crystal structures and fluorescence properties. Polyhedron 2018, 139, 7–16. [Google Scholar] [CrossRef]
  7. Liu, Y.A.; Wang, C.Y.; Zhang, M.; Song, X.Q. Structures and magnetic properties of cyclic heterometallic tetranuclear clusters. Polyhedron 2017, 127, 278–286. [Google Scholar] [CrossRef]
  8. Liu, P.P.; Sheng, L.; Song, X.Q.; Xu, W.Y.; Liu, Y.A. Synthesis, structure and magnetic properties of a new one dimensional manganese coordination polymer constructed by a new asymmetrical ligand. Inorg. Chim. Acta 2015, 434, 252–257. [Google Scholar] [CrossRef]
  9. Song, X.Q.; Cheng, G.Q.; Liu, Y.A. Enhanced Tb (III) luminescence by d 10 transition metal coordination. Inorg. Chim. Acta 2016, 450, 386–394. [Google Scholar] [CrossRef]
  10. Wang, P.; Zhao, L. An infinite 2D supramolecular cobalt (II) complex based on an asymmetric salamo-type ligand: Synthesis, crystal structure, and spectral properties. Synth. React. Inorg. Met. -Org. Nano-Met. Chem. 2016, 46, 1095–1101. [Google Scholar] [CrossRef]
  11. Zhang, Y.; Liu, L.Z.; Peng, Y.D.; Li, N.; Dong, W.K. Structurally characterized trinuclear nickel (II) and copper (II) salamo-type complexes: Syntheses, Hirshfeld analyses and fluorescent properties Transit. Met. Chem. 2019, 44, 627–639. [Google Scholar] [CrossRef]
  12. Sun, Y.X.; Zhang, S.T.; Ren, Z.L.; Dong, X.Y.; Wang, L. Synthesis, characterization, and crystal structure of a new supramolecular CdII complex with halogen-substituted salen-type bisoxime. Synth. React. Inorg. Met.-Org. Nano-Met. Chem. 2013, 43, 995–1000. [Google Scholar] [CrossRef]
  13. Chattopadhyay, S.; Drew, M.G.B.; Ghosh, A. Methylene spacer-regulated structural variation in cobalt(II/III) complexes with bridging acetate and salen- or salpn-type Schiff-base ligands. Eur. J. Inorg. Chem. 2008, 10, 1693–1701. [Google Scholar] [CrossRef]
  14. Liu, L.Z.; Wang, L.; Yu, M.; Zhao, Q.; Zhang, Y.; Sun, Y.X.; Dong, W.K. A highly sensitive and selective fluorescent “off-on-off” relay chemosensor based on a new bis (salamo)-type tetraoxime for detecting Zn2+ and CN. Spectrochim. Acta A 2019, 222, 7209. [Google Scholar] [CrossRef] [PubMed]
  15. Song, X.Q.; Peng, Y.J.; Chen, G.Q.; Wang, X.R.; Liu, P.P.; Xu, W.Y. Substituted group-directed assembly of Zn (II) coordination complexes based on two new structural related pyrazolone based salen ligands: Syntheses, structures and fluorescence properties. Inorg. Chim. Acta 2015, 427, 13–21. [Google Scholar] [CrossRef]
  16. Varga, Z.; Meana-Pañeda, R.; Song, G.L.; Paukku, Y.; Truhlar, D.G. Potential energy surface of triplet N2O2. J. Chem. Phys. 2016, 144, 8. [Google Scholar] [CrossRef]
  17. Zhao, Q.; An, X.X.; Liu, L.Z.; Dong, W.K. Syntheses, luminescences and Hirshfeld surfaces analyses of structurally characterized homo-trinuclear ZnII and hetero-pentanuclear ZnII-LnIII (Ln = Eu, Nd) bis(salamo)-like complexes. Inorg. Chim. Acta 2019, 490, 6–15. [Google Scholar] [CrossRef]
  18. Akine, S.; Utsuno, F.; Taniguchi, T.; Nabeshima, T. Dinuclear complexes of the N2O2 oxime chelate ligand with zinc(II)-lanthanide(III) as a selective sensitization system for Sm3+. Eur. J. Inorg. Chem. 2010, 20, 3143–3152. [Google Scholar] [CrossRef]
  19. Łępicka, K.; Pieta, P.; Francius, G.; Walcarius, A.; Kutner, W. Structure-reactivity requirements with respect to nickel-salen based polymers for enhanced electrochemical stability. Electrochim. Acta 2019, 315, 75–83. [Google Scholar] [CrossRef]
  20. Liu, X.; Manzurc, C.; Novoa, N.; Celedónc, S.; Carrilloc, D.; Hamon, J.R. Multidentate unsymmetrically substituted Schiff bases and their metal complexes: Synthesis, functional materials properties, and applications to catalysis. Coord. Chem. Rev. 2018, 357, 144–172. [Google Scholar] [CrossRef]
  21. Finelli, A.; Hérault, N.; Crochet, A.; Fromm, K.M. Threading salen-type Cu- and Ni-complexes into one-dimensional coordination polymers: Solution versus solid state, and the size effect of the alkali metal ion. Cryst. Growth Des. 2018, 18, 1215–1226. [Google Scholar] [CrossRef]
  22. Celedón, S.; Dorcet, V.; Roisnel, T.; Singh, A.; Ledoux-Rak, I.; Hamon, J.R.; Carrillo, D.; Manzur, C. Main-chain oligomers from NiII- and CuII-centered unsymmetrical N2O2 Schiff-base complexes: Synthesis and spectral, structural, and second-order nonlinear optical properties. Eur. J. Inorg. Chem. 2014, 29, 4984–4993. [Google Scholar]
  23. Akine, S. Novel ion recognition systems based on cyclic and acyclic oligo (salen)-type ligands. J. Incl. Phenom. Macrocycl. Chem. 2012, 72, 25–54. [Google Scholar] [CrossRef]
  24. Li, X.Y.; Chen, L.; Gao, L.; Zhang, Y.; Akogun, S.F.; Dong, W.K. Syntheses, crystal structures and catalytic activities of two solvent-induced homotrinuclear Co (II) complexes with a naphthalenediol-based bis(Salamo)-type tetraoxime ligand. RSC Adv. 2017, 7, 35905–35916. [Google Scholar] [CrossRef]
  25. Li, X.Y.; Kang, Q.P.; Liu, C.; Zhang, Y.; Dong, W.K. Structurally characterized homo-trinuclear ZnII and hetero- pentanuclear [ZnII4LnIII] complexes constructed from an octadentate bis(Salamo)-based ligand: Hirshfeld surfaces, fluorescence and catalytic properties. New J. Chem. 2019, 43, 4605–4619. [Google Scholar] [CrossRef]
  26. Wu, H.L.; Pan, G.L.; Bai, Y.C.; Zhang, Y.H.; Wang, H.; Shi, F.R.; Wang, X.L.; Kong, J. Study on synthesis, crystal structure, antioxidant and DNA-binding of mono-, di- and poly-nuclear lanthanides complexes with bis (N-salicylidene)-3-oxapentane-1,5-diamine. J. Photochem. Photobiol. B 2014, 135, 33–43. [Google Scholar] [CrossRef]
  27. Wu, H.L.; Pan, G.L.; Bai, Y.C.; Wang, H.; Kong, J.; Shi, F.R.; Zhang, Y.H.; Wang, X.L. Preparation, structure, DNA-binding properties, and antioxidant activities of a homodinuclear erbium (III) complex with a pentadentate Schiff base ligand. J. Chem. Res. 2014, 38, 211–217. [Google Scholar] [CrossRef]
  28. Wu, H.L.; Wang, C.P.; Wang, F.; Peng, H.P.; Zhang, H.; Bai, Y.C. A new manganese (III) complex from bis(5-methylsalicylaldehyde)-3-oxapentane-1,5-diamine: Synthesis, characterization, antioxidant activity and luminescence. J. Chin. Chem. Soc. 2015, 62, 1028–1034. [Google Scholar] [CrossRef]
  29. Song, X.Q.; Liu, P.P.; Liu, Y.A.; Zhou, J.J.; Wang, X.L. Two dodecanuclear heterometallic [Zn6Ln6] clusters constructed by a multidentate salicylamide salen-like ligand: Synthesis, structure, luminescence and magnetic properties. Dalton Trans. 2016, 45, 8154–8163. [Google Scholar] [CrossRef]
  30. Li, J.; Zhang, H.J.; Chang, J.; Jia, H.R.; Sun, Y.X.; Huang, Y.Q. Solvent-induced unsymmetric salamo-like trinuclear NiII complexes: Syntheses, crystal structures, fluorescent and magnetic properties. Crystals 2018, 8, 176. [Google Scholar] [CrossRef]
  31. Liu, P.P.; Wang, C.Y.; Zhang, M.; Song, X.Q. Pentanuclear sandwich-type ZnII-LnIII clusters based on a new Salen-like salicylamide ligand: Structure, near-infrared emission and magnetic properties. Polyhedron 2017, 129, 133–140. [Google Scholar] [CrossRef]
  32. Petrick, M.; Florian, B.; Katja, L.; Guntram, R.; Réne, P. Cooperative Al(Salen)-pyridinium catalysts for the asymmetric synthesis of trans-configured β-Lactones by [2+2]-cyclocondensation of acylbromides and aldehydes: Investigation of pyridinium substituent effects. Molecule 2012, 17, 7121–7150. [Google Scholar]
  33. Song, X.Q.; Liu, P.P.; Wang, C.Y.; Liu, Y.A.; Liu, W.S.; Zhang, M. Three sandwich-type zinc (II)-lanthanide (III) clusters: Structures, luminescence and magnetic properties. RSC Adv. 2017, 7, 22692–22698. [Google Scholar] [CrossRef] [Green Version]
  34. Chai, L.Q.; Huang, J.J.; Zhang, H.S. An unexpected cobalt (III) complex containing a Schiff base ligand: Synthesis, crystal structure, spectroscopic behavior, electrochemical property and SOD-like activity. Spectrochim. Acta A 2014, 131, 526–530. [Google Scholar] [CrossRef] [PubMed]
  35. Chai, L.Q.; Tang, L.J.; Chen, L.C.; Huang, J.J. Structural, spectral, electrochemical and DFT studies of two mononuclear manganese (II) and zinc (II) complexes. Polyhedron 2017, 122, 228–240. [Google Scholar] [CrossRef]
  36. Chai, L.Q.; Huang, J.J.; Zhang, J.Y.; Li, Y.X. Two 1-D and 2-D cobalt (II) complexes: Synthesis, crystal structures, spectroscopic and electrochemical properties. J. Coord. Chem. 2015, 68, 1224–1237. [Google Scholar] [CrossRef]
  37. Chai, L.Q.; Li, Y.X.; Chen, L.C.; Zhang, J.Y.; Huang, J.J. Synthesis, X-ray structure, spectroscopic, electrochemical properties and DFT calculation of a bridged dinuclear copper(II) complex. Inorg. Chim. Acta 2016, 444, 193–201. [Google Scholar] [CrossRef]
  38. Bhowmik, P.; Nayek, H.P.; Corbella, M.; Aliaga-Alcalde, N.; Chattopadhyay, S. Control of molecular architecture by steric factors: Mononuclear vs polynuclear manganese (III) compounds with tetradentate N2O2 donor Schiff bases. Dalton Trans. 2011, 40, 7916–7926. [Google Scholar] [CrossRef]
  39. Dong, W.K.; Ma, J.C.; Zhu, L.C.; Zhang, Y. Self-assembled zinc (II)-lanthanide (III) heteromultinuclear complexes constructed from 3-MeO salamo ligand: Syntheses, structures and luminescent properties. Cryst. Growth Des. 2016, 16, 6903–6914. [Google Scholar] [CrossRef]
  40. Dong, X.Y.; Kang, Q.P.; Li, X.Y.; Ma, J.C.; Dong, W.K. Structurally characterized solvent-induced homotrinuclear cobalt (II) N2O2-donor bisoxime-type complexes. Crystals 2018, 8, 139. [Google Scholar] [CrossRef] [Green Version]
  41. Novozhilova, M.V.; Smirnova, E.A.; Polozhentseva, J.A.; Danilova, J.A.; Chepurnaya, I.A.; Karushev, M.P.; Malev, V.V.; Timonov, A.M. Multielectron redox processes in polymeric cobalt complexes with N2O2 Schiff base ligands. Electrochim. Acta 2018, 282, 105–115. [Google Scholar] [CrossRef]
  42. Zhao, L.; Dang, X.T.; Chen, Q.; Zhao, J.X.; Wang, L. Synthesis, crystal structure and spectral properties of a 2D supramolecular copper (II) complex with 1-(4-{[(E)-3-ethoxyl-2-hydroxybenzylidene]amino}phenyl)ethanone oxime. Synth. React. Inorg. Met.-Org. Nano-Met. Chem. 2013, 43, 1241–1246. [Google Scholar] [CrossRef]
  43. Yamamura, M.; Takizawa, H.; Sakamoto, N.; Nabeshima, T. Monomeric and dimeric red/NIR-fluorescent dipyrrin-germanium complexes: Facile monomer-dimer interconversion driven by acid/base additions. Tetrahedron Lett. 2013, 54, 7049–7052. [Google Scholar] [CrossRef]
  44. Manna, A.K.; Mondal, J.; Chandra, R.; Rout, K.; Patra, G.K. A fluorescent colorimetric azo dye based chemosensor for detection of S2- in perfect aqueous solution and its applications in real sample analysis and molecular logic gate. Sens. Actuators B 2018, 10, 2317–2326. [Google Scholar] [CrossRef]
  45. Zhou, L.; Cai, P.; Fen, Y.; Cheng, J.; Xiang, H.; Liu, J.; Wu, D.; Zhou, X. Synthesis and photophysical properties of water-soluble sulfonato-Salen-type Schiff bases and their applications of fluorescence sensors for Cu2+ in water and living cells. Anal. Chim. Acta 2012, 735, 96–106. [Google Scholar] [CrossRef] [PubMed]
  46. Pushkarev, A.P.; Balashova, T.V.; Kukinov, A.A.; Arsenyev, M.V.; Yablonskiy, A.N.; Kryzhkov, D.I.; Andreev, B.A.; Rumyantcev, R.V.; Fukin, G.K.; Bochkarev, M.N. Sensitization of NIR luminescence of Yb3+ by Zn2+ chromophores in heterometallic complexes with a bridging Schiff-base ligand. Dalton Trans. 2017, 46, 10408–10417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Peng, Y.D.; Li, X.Y.; Kang, Q.P.; An, G.X.; Zhang, Y.; Dong, W.K. Synthesis and fluorescence properties of asymmetrical salamo-type tetranuclear zinc (II) complex. Crystals 2018, 8, 107. [Google Scholar] [CrossRef] [Green Version]
  48. Peng, Y.D.; Wang, F.; Gao, L.; Dong, W.K. Structurally characterized dinuclear zinc (lI) bis(salamo)-type tetraoxime complex possessing square pyramidal and trigonal bipyramidal geometries. J. Chin. Chem. Soc. 2018, 65, 893–899. [Google Scholar] [CrossRef]
  49. Wang, F.; Liu, L.Z.; Gao, L.; Dong, W.K. Unusual constructions of two salamo-based copper (Il) complexes. Spectrochim. Acta A 2018, 203, 56–64. [Google Scholar] [CrossRef]
  50. Li, X.Y.; Kang, Q.P.; Liu, L.Z. Trinuclear Co (II) and mononuclear Ni (II) salamo-type bisoxime coordination compounds. Crystals 2018, 8, 43. [Google Scholar] [CrossRef] [Green Version]
  51. Dong, X.Y.; Zhao, Q.; Wei, Z.L.; Mu, H.R.; Zhang, H.; Dong, W.K. Synthesis and fluorescence propertiesof structurally characterized heterobimetalic Cu (II)-Na (I) bis (salamo)-based complex bearing square planar, square pyramid and triangular prism geometries of metal centers. Molecules 2018, 23, 1006. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Gao, L.; Liu, C.; Wang, F.; Dong, W.K. Tetra-, penta- and hexa-coordinated transition metal complexes constructedfrom coumarin-containing N2O2 ligand. Crystals 2018, 8, 77. [Google Scholar] [CrossRef] [Green Version]
  53. Wang, L.; Kang, Q.P.; Hao, J.; Dong, W.K. Two trinuclear cobalt(Il) salamo-type complexes: Syntheses, crystal structures, solvent effect and fluorescent properties. Chin. J. Inorg. Chem. 2018, 34, 525–533. [Google Scholar]
  54. Kang, Q.P.; Li, X.Y.; Zhao, Q.; Ma, J.C.; Dong, W.K. Structurally characterized homotrinuclear Salamo-type nickel (lI) complexes: Synthesis, solvent effect and fluorescence properties. Appl. Organomet. Chem. 2018, 32, e4379. [Google Scholar] [CrossRef]
  55. Hao, J.; Li, X.Y.; Zhang, Y.; Dong, W.K. A reversible bis(salamo)-based fluorescence sensor for selective detection of Cd2+ in water-containing systems and food samples. Materials 2018, 11, 523. [Google Scholar] [CrossRef] [Green Version]
  56. Zhang, L.W.; Li, X.Y.; Kang, Q.P. Liu, L.Z. Ma, J.C. Dong, W.K. Structures and fluorescent and magnetic behaviors of newly synthesized NiII and CuII coordination compounds. Crystals 2018, 8, 173. [Google Scholar] [CrossRef] [Green Version]
  57. Ren, Z.L.; Hao, J.; Hao, P.; Dong, X.Y.; Bai, Y.; Dong, W.K. Synthesis, crystal structure, luminescence and electrochemical properties of a salamo-type trinuclear cobalt (II) complex. Z. Nat. B 2018, 73, 203–210. [Google Scholar] [CrossRef]
  58. Zhang, L.W.; Liu, L.Z.; Wang, F.; Dong, W.K. Unprecedented fluorescent dinuclear CoII and ZnII coordination compounds with a symmetric bis(salamo)-like tetraoxime. Molecules 2018, 23, 1141. [Google Scholar] [CrossRef] [Green Version]
  59. Gulzhian, I.D.; Igor E, U. Design and synthesis of coordination polymers with chelated units and their application in nanomaterials science. RSC Adv. 2017, 7, 42242–42288. [Google Scholar]
  60. Yang, Y.H.; Zhang, Y.; Yu, M.; Zheng, S.S.; Dong, W.K. Heterbimetallic Cu (II)-Nd (III) and Zn (II)-Ce (III) salamo-type complexes: Syntheses, crystal structure and fluorescence properties. Chin. J. Inorg. Chem. 2018, 34, 997–1006. [Google Scholar]
  61. Akine, S.; Taniguchi, T.; Dong, W.K.; Masubuchi, S.; Nabeshima, T. Oxime-based salen-type tetradentate ligands with high stability against imine metathesis reaction. J. Org. Chem. 2005, 70, 1704–1711. [Google Scholar] [CrossRef] [PubMed]
  62. Sheldrick, G.M. Acta Crystallogr. Sect. C: Cryst. Struct. Commun. 2015, 71, 3–8. [Google Scholar]
  63. Geary, W.J. The use of conductivity measurements in organic solvents for the characterization of coordination compounds. Coord. Chem. Rev. 1971, 7, 81–122. [Google Scholar] [CrossRef]
  64. Dong, W.K.; Ma, J.C.; Zhu, L.C.; Zhang, Y. Nine self–assembled nickel (II)-lanthanide (III) heterometallic complexes constructed from a salamo–type bisoxime and bearing N– or –donor auxiliary ligand: Syntheses, structures and magnetic properties. New J. Chem. 2016, 40, 6998–7010. [Google Scholar] [CrossRef]
  65. Sun, Y.X.; Xu, L.; Zhao, T.H.; Liu, S.H.; Liu, G.H.; Dong, X.T. Synthesis and crystal structure of a 3D supramolecular copper (II) complex with 1-(3-{[(E)-3-bromo-5-chloro-2-hydroxybenzylidene]amino}phenyl) ethanone oxime. Synth. React. Inorg. Met. -Org. Nano-Met. Chem. 2013, 43, 509–513. [Google Scholar] [CrossRef]
  66. An, X.X.; Zhao, Q.; Mu, H.R.; Dong, W.K. A new half-salamo-based homo-trinuclear nickel (II) complex: Crystal structure, Hirshfeld surface analysis, and fluorescence properties. Crystals 2019, 9, 101. [Google Scholar] [CrossRef] [Green Version]
  67. Sun, Y.X.; Gao, X.H. Synthesis, characterization, and crystal structure of a new CuII complex with salen-type ligand. Synth. React. Inorg. Met. -Org. Nano-Met. Chem. 2011, 41, 973–978. [Google Scholar] [CrossRef]
  68. Wang, P.; Zhao, L. Synthesis, structure and spectroscopic properties of the trinuclear cobalt (II) and nickel (II) complexes based on 2-hydroxynaphthaldehyde and bis (aminooxy)alkane. Spectrochim. Acta A 2015, 135, 342–350. [Google Scholar] [CrossRef]
  69. Sun, Y.X.; Wang, L.; Dong, X.Y.; Ren, Z.L.; Meng, W.S. Synthesis, characterization, and crystal structure of a supramolecular CoII complex containing salen-type bisoxime. Synth. React. Inorg. Met.-Org. Nano-Met. Chem. 2013, 43, 599–603. [Google Scholar] [CrossRef]
  70. Chai, L.Q.; Liu, G.; Zhang, Y.L.; Huang, J.J.; Tong, J.F. Synthesis, crystal structure, fluorescence, electrochemical property, and SOD-like activity of an unexpected nickel (II) complex with a quinazoline-type ligand. J. Coord. Chem. 2013, 66, 3926–3938. [Google Scholar] [CrossRef]
  71. Jia, H.R.; Chang, J.; Zhang, H.J.; Li, J.; Sun, Y.X. Three polyhydroxyl-bridged defective dicubane tetranuclear MnIII complexes: Synthesis, crystal structures, and spectroscopic properties. Crystals 2018, 8, 272. [Google Scholar] [CrossRef] [Green Version]
  72. Wu, H.L.; Bai, Y.; Yuan, J.K.; Wang, H.; Pan, G.L.; Fan, X.Y.; Kong, J. A zinc (II) complex with tris (2-(N-methyl) benzimidazlylmethyl) amine and salicylate: Synthesis, crystal structure, and DNA-binding. J. Coord. Chem. 2012, 65, 2839–2851. [Google Scholar] [CrossRef]
  73. Wu, H.L.; Pan, G.L.; Bai, Y.C.; Wang, H.; Kong, J. Synthesis, structure, antioxidation, and DNA-bindingstudies of a binuclear ytterbium (III) complex with bis(N-salicylidene)-3-oxapentane-1,5-diamine. Res. Chem. Intermed. 2015, 41, 3375–3388. [Google Scholar] [CrossRef]
  74. Zheng, S.S.; Dong, W.K.; Zhang, Y.; Chen, L.; Ding, Y.J. Four salamo-type 3d-4f hetero-bimetallic [ZnIILnIII] complexes: Syntheses, crystal structures, and luminescent and magnetic properties. New J. Chem. 2017, 44, 4966–4973. [Google Scholar] [CrossRef]
  75. Chai, L.Q.; Hu, Q.; Zhang, K.Y.; Zhou, L.; Huang, J.J. Synthesis, structural characterization, spectroscopic, and DFT studies of two pentacoordinated zinc (II) complexes containing quinazoline and 1, 10-phenanthroline as mixed ligands. J. Lumin. 2018, 203, 234–246. [Google Scholar] [CrossRef]
  76. Purkayastha, A.; Debnath, D.; Majumder, M.; Ortega-Castro, J.; Kirillov, A.M.; Ganguly, R.; Misra, T.K. Nickel (II) based homo- vs heterometallic 1D coordination polymers derived from a novel 6-aminouracil building block: Structures, topologies, non-covalent interactions, magnetism, and antibacterial activity. Inorg. Chim. Acta 2018, 482, 384–394. [Google Scholar] [CrossRef]
  77. Tabong, C.D.; Yufanyi, D.M.; Paboudam, A.G.; Nono, K.N.; Eni, D.B.; Agwara, M.O. Synthesis, crystal structure, and antimicrobial properties of [diaquabis(hexamethylenetetramine)diisothiocyanato-κN]nickel (II) complex. Adv. Chem. 2016, 1–8. [Google Scholar] [CrossRef] [Green Version]
  78. Colinas, I.R.; Rojas-Andrade, M.D.; Chakraborty, I.; Oliver, S.R.J. Two structurally diverse Zn-based coordination polymers with excellent antibacterial activity. CrystEngComm 2018, 20, 3353–3362. [Google Scholar] [CrossRef]
  79. Rauf, A.; Ye, J.; Zhang, S.; Shi, L.; Akram, M.A.; Ning, G. Synthesis, structure and antibacterial activity of a copper (II) coordination polymer based on thiophene-2,5-dicarboxylate ligand. Polyhedron 2019, 166, 130–136. [Google Scholar] [CrossRef]
Scheme 1. Synthetic route to the salamo-like ligand H2L.
Scheme 1. Synthetic route to the salamo-like ligand H2L.
Crystals 09 00596 sch001
Figure 1. Comparing the simulated PXRD (black) and experimental patterns of the NiII-GdIII coordination polymer.
Figure 1. Comparing the simulated PXRD (black) and experimental patterns of the NiII-GdIII coordination polymer.
Crystals 09 00596 g001
Figure 2. IR spectra of H2L and its NiII-GdIII coordination polymer.
Figure 2. IR spectra of H2L and its NiII-GdIII coordination polymer.
Crystals 09 00596 g002
Figure 3. The UV-visible absorption spectra of H2L and its NiII-GdIII coordination polymer dissolved in methanol.
Figure 3. The UV-visible absorption spectra of H2L and its NiII-GdIII coordination polymer dissolved in methanol.
Crystals 09 00596 g003
Figure 4. (a) Molecule structure and atom numberings of the NiII-GdIII coordination polymer with 30% probability displacement ellipsoids; (b) coordination polyhedra for NiII and GdIII ions of the coordination polymer; (c) view of 1-D chain structure of the NiII-GdIII coordination polymer.
Figure 4. (a) Molecule structure and atom numberings of the NiII-GdIII coordination polymer with 30% probability displacement ellipsoids; (b) coordination polyhedra for NiII and GdIII ions of the coordination polymer; (c) view of 1-D chain structure of the NiII-GdIII coordination polymer.
Crystals 09 00596 g004aCrystals 09 00596 g004b
Figure 5. View of the intramolecular (a) and intermolecular (b) hydrogen bonding interactions of the NiII-GdIII coordination polymer.
Figure 5. View of the intramolecular (a) and intermolecular (b) hydrogen bonding interactions of the NiII-GdIII coordination polymer.
Crystals 09 00596 g005
Figure 6. Emission spectra of H2L and its NiII-GdIII coordination polymer dissolved in CH3OH.
Figure 6. Emission spectra of H2L and its NiII-GdIII coordination polymer dissolved in CH3OH.
Crystals 09 00596 g006
Figure 7. Molecular orbital map of DFT calculation of the NiII-GdIII polymer.
Figure 7. Molecular orbital map of DFT calculation of the NiII-GdIII polymer.
Crystals 09 00596 g007
Figure 8. The inhibition zones diameter of ampicillin in various concentrations of different samples.
Figure 8. The inhibition zones diameter of ampicillin in various concentrations of different samples.
Crystals 09 00596 g008
Table 1. Crystallographic data for the NiII-GdIII coordination polymer.
Table 1. Crystallographic data for the NiII-GdIII coordination polymer.
CompondThe NiII-GdIII Coordination Complex
FormulaC31H33GdN4NiO14
Formula weight901.57
Temperature (K)100.00(10)
Radiation (Å)0.71073
Crystal systemmonoclinic
Space groupP 21/n
a (Å)10.8324 (7)
b (Å)18.4853 (11)
c (Å)17.7452 (12)
α (°)90
β (°)104.790 (7)
γ (°)90
V3)3435.6 (4)
Z4
Dc (g·cm−3)1.743
μ (mm−1)2.536
F (000)1804
Crystal size (mm)0.11 × 0.09 × 0.08
θ Range (°)2.203–29.566
Index ranges−14 ≤ h ≤ 11
−23 ≤ k ≤ 23
−22 ≤ l ≤ 19
Completeness to θ84.3% (θ = 25.242)
Tot. Date18,620
Uniq. Date8118
R (int)0.0397
Observed Date6605
Nref/Npar8118/464
GOF1.020
R[I/>2σ(I)]R1 = 0.0396, R2 = 0.0846
Largest diff. peak
and hole (e·Å−3)
−0.88, 1.45
R1 = Σ||Fo| − |Fc||/Σ|Fo|; ωR2 = [Σω(Fo2Fc2)2/Σω(Fo2)2]1/2; GOF = [Σω(Fo2Fc2)2/nobs − nparam)]1/2.
Table 2. Essential bond lengths (Å) and angles (°) for the NiII-GdIII coordination polymer.
Table 2. Essential bond lengths (Å) and angles (°) for the NiII-GdIII coordination polymer.
BondDist.BondDist.BondDist.
Gd1–O12.378(2)Gd1–O13 #12.416(3)Ni1–N12.063(3)
Gd1–O22.521(3)Gd1–O14 #12.482(3)Ni1–N22.060(3)
Gd1–O52.336(3)Ni1–O12.019(3)Ni1–O72.091(3)
Gd1–O62.599(3)Ni1–O52.025(2)Ni1–O112.063(3)
Gd1–O82.473(3)Gd1–O92.520(3)
BondAnglesBondAnglesBondAngles
O1–Gd1–O263.76(9)O2–Gd1–O6149.31(10)O5–Gd1–O1282.51(9)
O1–Gd1–O568.34(8)O2–Gd1–O8122.89(10)O5–Gd1–O13 #1131.02(9)
O1–Gd1–O6123.24(8)O2–Gd1–O971.92(9)O5–Gd1–O14 #190.05(9)
O1–Gd1–O8145.56(10)O2–Gd1–O1287.78(10)O5–Gd1–C31 #1110.41(10)
O1–Gd1–O9122.76(9)O1–Gd1–O13 #178.28(10)O7–Ni1–N286.07(12)
O1–Gd1–O1273.76(9)O1–Gd1–O14 #180.07(10)O6–Gd1–O870.28(10)
O1–Gd1–O13 #1126.03(9)O1–Gd1–C31 #179.59(11)O6–Gd1–O9113.51(9)
O1–Gd1–O14 #181.99(9)O5–Gd1–O662.21(8)O6–Gd1–O12122.81(10)
O1–Gd1–C31 #1105.41(10)O5–Gd1–O898.84(9)O6–Gd1–O13 #174.78(10)
O2–Gd1–O5131.95(9)O5–Gd1–O9143.56(9)O6–Gd1–O14 #172.24(10)
O6–Gd1–C31 #169.72(11)O9–Gd1–O13 #173.97(10)O13#1–Gd1–31 #126.98(10)
O8–Gd1–O951.00(10)O9–Gd1–O14 #1124.44(9)O14#1–Gd1–31 #126.72(10)
O8–Gd1–O1272.90(10)O9–Gd1–C31 #1100.01(10)O1–Ni1–O581.82(10)
O8–Gd1–O13 #186.97(10)O12–Gd1–O13 #1143.91(10)O1–Ni1–O791.33(10)
O8–Gd1–O14 #1131.28(10)O12–Gd1–O14 #1155.71(10)O1–Ni1–O1190.02(11)
O8–Gd1–C31 #1109.03(11)O12–Gd1–C31 #1166.02(11)O1–Ni1–N189.95(12)
O9–Gd1–O1270.06(10)O13#1–Gd1–14 #153.60(9)O1–Ni1–N2169.75(12)
O5–Ni1–O792.62(10)O7–Ni1–O11175.67(11)O11–Ni1–N293.31(12)
O5–Ni1–O1191.64(11)O7–Ni1–N187.35(12)N1–Ni1–N299.83(13)
O5–Ni1–N1171.76(12)O7–Ni1–N286.07(12)Gd1 #2–C31–O1462.0(2)
O5–Ni1–N288.39(12)O11–Ni1–N188.54(13)Gd1 #2–C31–C28171.4(3)
Gd1 #2–C31–O1359.03(19)Gd1 #2–O13–C3194.0(2)Gd1 #2–O13–C3191.3(2)
Symmetry transformations used to generate equivalent atoms: #1 x + 1/2, −y + 1/2, z − 1/2, #2 x − 1/2, −y + 1/2, z + 1/2.
Table 3. Hydrogen bond parameters (Å, °) for the NiII-GdIII coordination polymer.
Table 3. Hydrogen bond parameters (Å, °) for the NiII-GdIII coordination polymer.
D–H···Ad(D–H)d(H···A)d(D···A)∠D–H···ASymmetry Code
C7–H7B···O130.972.553.163(5)121–1/2 + x, 1/2–y, 2 + z
C9–H9···O100.932.563.456(5)1633/2 − x, 1/2+ y, 3/2 − z
C10–H10A···O90.972.543.259(6)1313/2 − x, 1/2+y, 3/2 − z
C11–H11B···O110.972.243.169(5)159
C19–H19B···O40.972.503.105(5)1201/2 − x, −1/2 + y,3/2 − z
C19–H19B··O130.972.473.146(6)128–1/2 + x, 1/2 − y, 2 + z
C20–H20B···O30.962.493.396(6)1571/2 − x, 1/2 + y, 3/2 − z
C20–H20C···O80.962.413.185(6)138
C21–H21···O50.932.563.090(4)116
C21–H21···O140.932.462.258(5)148–1/2 + x, 1/2 − y, 2 + z
C23–H23A···O70.962.442.810(5)103
C23–H27···O130.962.492.796(5)100

Share and Cite

MDPI and ACS Style

Cui, Y.-F.; Zhang, Y.; Xie, K.-F.; Dong, W.-K. A Newly Synthesized Heterobimetallic NiII-GdIII Salamo-BDC-Based Coordination Polymer: Structural Characterization, DFT Calculation, Fluorescent and Antibacterial Properties. Crystals 2019, 9, 596. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst9110596

AMA Style

Cui Y-F, Zhang Y, Xie K-F, Dong W-K. A Newly Synthesized Heterobimetallic NiII-GdIII Salamo-BDC-Based Coordination Polymer: Structural Characterization, DFT Calculation, Fluorescent and Antibacterial Properties. Crystals. 2019; 9(11):596. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst9110596

Chicago/Turabian Style

Cui, Yong-Fan, Yu Zhang, Ke-Feng Xie, and Wen-Kui Dong. 2019. "A Newly Synthesized Heterobimetallic NiII-GdIII Salamo-BDC-Based Coordination Polymer: Structural Characterization, DFT Calculation, Fluorescent and Antibacterial Properties" Crystals 9, no. 11: 596. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst9110596

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