Next Article in Journal
Cobalt(II) Terpyridin-4′-yl Nitroxide Complex as an Exchange-Coupled Spin-Crossover Material
Previous Article in Journal
Zinc Oxide Nanostructures: From Chestnut Husk-Like Structures to Hollow Nanocages, Synthesis and Structure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Structural and Luminescent Properties of Heterobimetallic Zinc(II)-Europium(III) Dimer Constructed from N2O2-Type Bisoxime and Terephthalic Acid

School of Chemical and Biological Engineering, Lanzhou Jiaotong University, Lanzhou 730070, China
*
Author to whom correspondence should be addressed.
Crystals 2018, 8(4), 154; https://doi.org/10.3390/cryst8040154
Submission received: 8 March 2018 / Revised: 29 March 2018 / Accepted: 30 March 2018 / Published: 2 April 2018
(This article belongs to the Section Crystal Engineering)

Abstract

:
A new heterohexanuclear ZnII–EuIII dimer [{(ZnL)2Eu}2(bdc)2]·2Cl constructed from a N2O2-type chelating ligand H2L (6,6′-Dimethoxy-2,2′-[ethylenedioxybis(nitrilomethylidyne)]diphenol), Zn(OAc)2·2H2O, EuCl3·6H2O and H2bdc (terephthalic acid) was synthesized, and characterized using elemental analyses, IR (Infrared), UV-Vis (Ultraviolet–visible) spectra and X-ray single crystal diffraction method. There are two crystallographically equivalent [(ZnL)2Eu] moieties in the ZnII–EuIII complex, the two [(ZnL)2Eu] moieties are linked by two bdc2– ligand leading to a heterohexanuclear dimer, in which the carboxylato groups bridge the ZnII and EuIII atoms. Furthermore, the luminescence properties of H2L and its ZnII–EuIII complex have been studied.

Graphical Abstract

1. Introduction

Salen-type N2O2 ligands are easily obtained via the reaction of salicylaldehyde or its derivatives with diamines, can coordinated to transition metal ions in a N2O2 tetradentate chelating mode to form stable mono- or multinuclear complexes [1,2,3,4,5,6]. Salen- and salamo-type complexes have been extensively investigated in organometallic chemistry and modern coordination chemistry for several decades [7,8,9,10,11], their complexes are well-known for their potential applications in many areas, such as catalysts [12,13], biological fields [14,15,16,17,18,19,20,21], supramolecular buildings [22,23,24,25,26], molecular recognitions [27,28,29,30,31], magnetic [32,33,34,35,36,37] and luminescence [38,39,40,41,42,43,44,45] materials and so forth. Self-assembling processes of a metallohost complex with auxiliary organic ligands are usually utilized in the building of metal-organic framework (MOF) materials [46,47]. When 3-alkoxy groups are introduced of salicylidene moieties, an O4 coordination site composed of the alkoxy and phenoxo oxygen atoms are produced in addition to the N2O2 site [48,49,50,51]. The O4 site of 3-alkoxy-induced Salen-type ligand is suitable for lanthanide(III) atoms to prepare 3d–4f heteronuclear complexes. These 3d–4f heterobimetallic Salen-type complexes have been diffusely studied [14,34,37,44] in which acetate (OAc-) combine plays a key role in assembling 3d and 4f metals [52], however, 3d–4f heterobimetallic Salamo-type complexes have been rarely reported, especially the auxiliary ligands were introduced [53].
Herein, a newly 3d–4f heterobimetallic dimer, [{(ZnL)2Eu}2(bdc)2]·2Cl has been obtained by a N2O2-type chelating ligand H2L and auxiliary ligand bdc2−. In addition, the luminescence properties of H2L and its ZnII–EuIII complex have been studied.

2. Experimental Section

2.1. Materials and Measurements

2-Hydroxy-3-methoxybenzaldehyde (99%) was purchased from Alfa Aesar (New York, NY, USA), while Tianjin Chemical Reagent Factory (Tianjin, China) supplied the remaining reagents. Elemental analyses for zinc and europium were detected by IRIS ER/S-WP-1 ICP atomic emission spectrometer (Elementar, Berlin, Germany), elemental analyses for Carbon, hydrogen and nitrogen were analyzed using GmbH VariuoEL V3.00 automatic elemental analysis instrument (Elementar, Berlin, Germany). IR spectra (4000–400 cm−1) were recorded on a Vertex 70 FT-IR (Fourier Transform infrared) spectrophotometer (Bruker, Billerica, MA, USA), with samples prepared as KBr pellets. UV-vis absorption spectra were measured on a Shimadzu UV-3900 spectrometer (Shimadzu, Tokyo, Japan). 1H NMR (Nuclear Magnetic Resonance) spectra were determined by German Bruker AVANCE DRX-400/600 spectroscopy (Bruker AVANCE, Billerica, MA, USA). Fluorescence spectra were recorded on a F-7000 FL spectrophotometer (Hitachi, Tokyo, Japan). X-ray single-crystal structure was determined on a SuperNova Dual (Cu at zero) four-circle diffractometer.

2.2. Synthesis of H2L

The ligand 6,6′-dimethoxy-2,2′-[ethylenedioxybis(nitrilomethylidyne)]diphenol (H2L) was synthesized in accordance with the earlier reported method [54]. Yield: 83.6%. m.p: 131–132 °C. 1H NMR (400 MHz, CDCl3), δ 3.91 (s, 6H), 4.49 (s, 4H), 6.83 (dd, J = 7.9, 1.9 Hz, 2H), 6.86 (t, J = 7.9 Hz, 2H), 6.91 (dd, J = 7.9, 1.9 Hz, 2H), 8.26 (s, 2H), 9.74 (s, 2H). IR (KBr, cm–1): 3131 [ν(O–H)], 1618 [ν(C=N)], 1253 [ν(Ar-O)]. UV-Vis (CH3OH), λmax (nm) (εmax): 224, 270 and 318 nm (3.0 × 10−5 M). Anal. Calcd for C18H20N2O6 (%): C 59.99; H 5.59; N 7.77. Found: C 60.29; H 5.42; N 7.56.

2.3. Synthesis of the ZnII–EuIII Complex

To a solution of H2L (7.20 mg, 0.02 mmol) in CHCl3 (3 mL) was added Zn(OAc)2·2H2O (4.39 mg, 0.02 mmol) and EuCl3·6H2O (7.33 mg, 0.02 mmol) in CH3CH2OH (2 mL). After the mixture was stirred for about 15 min at r.t., a solution of H2bdc (3.32 mg, 0.02 mmol) in DMF (1 mL) was added dropwise and continued to stir for 15 min. The mixture was filtered, and the filtrate was obtained. Several yellow block single crystals were obtained via slow evaporation of the mixture solution in open atmosphere for almost two weeks. (Scheme 1) Yield: 72.6%. IR (KBr, cm−1): 1607 [ν(C=N)], 1217 [ν(Ar–O)]. UV-Vis (CH3OH), λmax (nm) (εmax): 233, 279 and 352 nm (3.0 × 10−5 M). Anal. Calcd for C88H80Cl2Eu2Zn4N8O32 (%): C, 44.08; H, 3.36; N, 4.67; Zn, 10.91; Eu, 12.67. Found: C, 44.29; H, 4.45; N, 4.51; Zn, 11.02; Eu, 12.43.

2.4. X-ray Crystal Structure Determinations for the ZnII–EuIII Complex

The single crystal diffractometer provides a monochromatic beam of Mo-Kα radiation (0.71073 Å) produced from a sealed Mo X-ray tube using Graphite monochromator and was used for obtaining crystal data for the ZnII–EuIII complex at 291(2) K, respectively. The LP factor and semi-empirical absorption were using SADABS. The structures of the ZnII–EuIII complex were solved via the direct methods (SHELXS-2016) [55], and all hydrogen atoms were included at the calculated positions and constrained to ride on their parent atoms. All non-hydrogen atoms were refined anisotropically using a full-matrix least-squares procedure on F2 with SHELXL-2016 [56]. In the X-ray structure refinement, however, the solvent molecules of the complex could not be located because of its high thermal disorder, and the final structure model was refined without the solvent molecules by using a SQUEEZE routine of PLATON. Table 1 shows the data collection and refinements of the ZnII–EuIII complex. Supplementary crystallographic data for this paper have been deposited at Cambridge Crystallographic Data Centre (1,828,198) and can be obtained free of charge via www.ccdc.cam.ac.uk/conts/retrieving.html.

3. Results and Discussion

3.1. IR Spectra

The IR spectra of H2L and its corresponding ZnII–EuIII complex exhibited different bands in the region of 4000–400 cm−1 region (Table 2). It is obvious that the ν(O–H) absorption band was near 3131 cm−1 in the spectrum of H2L. The free ligand exhibited a characteristic C=N stretching band at 1618 cm−1, while the ZnII–EuIII complex appeared at 1607 cm−1. The characteristic C=N stretching frequency is shifted to lower frequency, which indicating that the metal(II) atoms are bonded by oxime N atoms [57]. The Ar–O stretching frequencies appeared as a strong band within 1265–1213 cm–1 range as reported for similar Salen-type ligands [58]. Meanwhile, the free ligand H2L exhibited an Ar–O stretching frequency at 1253 cm−1, while the ZnII–EuIII complex appeared at 1217 cm−1, the Ar–O stretching frequency is shifted to lower frequency, indicating that the M–O bonds are formed [59]. In addition, characteristic absorption bands of the carboxylato ions are observed, the absorption bands νas(COO) and νs(COO) of the ZnII–EuIII complex are 1572 and 1468 cm–1, respectively, indicating the bidentate coordination mode of carboxylato ions [60], which is in consistent with the X-ray diffraction result obtained for the ZnII–EuIII complex.

3.2. UV–Vis Spectra

UV–Vis absorption spectra of H2L and its corresponding ZnII–EuIII complex were determined in 3.0 × 105 M methanol solution, as depicted in Figure 1. The absorption spectrum of the ZnII–EuIII complex is different from that of H2L. UV–Vis spectrum of H2L exhibited three absorptions at ca. 224, 270 and 318 nm. The absorptions at 224 and 270 nm can be assigned to ππ* transitions of the benzene rings, while the absorption at 318 nm can be attributed to ππ* transition of C=N groups [61]. Compared with the absorption peak of H2L, the corresponding absorption peaks at ca. 233 and 279 nm are observed in the ZnII–EuIII complex, which were bathochromically shifted, indicating the coordination of metal atoms with the ligand [62]. Meanwhile, a new absorption peak was observed at ca. 352 nm in the ZnII–EuIII complex, and assigned to L→M charge-transfer (LMCT) transitions, which is characteristic of the transition metal complexes with Salen-type N2O2 coordination spheres [63].

3.3. Description of the Crystal Structure

The crystal structure of the heterohexanuclear ZnII–EuIII complex is depicted in Figure 2 and Table 3, the crystallographic data revealed that the ZnII–EuIII complex crystallizes in the triclinic space group P-1 with Z = 1. By the reaction of H2L with Zn(OAc)2·2H2O, EuCl3·6H2O and H2bdc, a heterohexanuclear ZnII–EuIII complex is formed: [{(ZnL)2Eu}2(bdc)2]·2Cl, the molecular structure of the ZnII–EuIII complex composes of four ZnII atoms, two EuIII atoms, four deprotonated (L)2– ligands, two auxiliary ligand bdc2– moieties. Each ZnII atom (Zn1 or Zn2) is located at the N2O2 coordination spheres of Salamo moieties, they are penta-coordinated by two oxime nitrogen (N1, N2 or N3, N4) atoms, two deprotonated phenoxo-oxygen (O2, O5 or O8, O11) atoms of the (L)2− units and one oxygen (O16 or O13) atom comes from the coordinated terephthalic acid molecule. The coordination environment around all the ZnII atoms are best described as slightly distorted triangular bipyramid geometries [64], which were deduced by calculating the values of τ1 = 0.539 and τ2 = 0.6125 [65], respectively. The equatorial coordination sites with the distances of N2–Zn1 = 2.046(5) Å, O2–Zn1 = 2.000(4) Å and O16–Zn1 = 1.983(4) Å, and the axial position with the distances of O5–Zn1 = 2.052(4) and N1–Zn1 = 2.116(5) Å (Table 3). The dihedral angle between coordination planes of N1–Zn1–O2 and N2–Zn1–O5 is 51.84(5)°. The dihedral angle between coordination planes of N1–Zn2–O2 and N2–Zn2–O5 is 55.37(2)°.
In addition, the central EuIII atom is located in a deca-coordinated O10 environment, the coordination sphere of the central EuIII atom contains eight oxygen atoms (O1, O2, O5, O6 and O7, O8, O11, O12) from two deprotonated (L)2– units and two oxygen atoms (O14 and O15) comes from the coordinated bdc2– ions. Thus, the deca-coordinated EuIII atom adopts a distorted bicapped square antiprism coordination arrangement. The primary Eu–O distances are in the normal ranges of 2.378(4)–2.844(4) Å (Table 3). The dihedral angle between the O2–Zn1–O5 and O2–Eu1–O5 planes is 32.41(4)°.
As depicted in Figure 3 and Table 4, In the crystal structure of the ZnII–EuIII complex, five pairs of intramolecular C1–H1B···O8, C9–H9A···O16, C19–H19B···O2, C36–H36B···O15 and C43–H43···O13 interactions are formed [66,67,68], which plays a vital role in constructing and stabilizing the complex molecule. The donors (C1–H1B and C19–H19B) from 3-methoxy of the (L)2– units form hydrogen bonding with oxygen atoms (O8 and O2) of phenoxo-oxygen of the (L)2– units as hydrogen bonding receptors. The donors (C9–H9A and C36–H36B) from the (L)2– units form hydrogen bonding with oxygen atom (O16 and O15) of the bdc2− moiety as hydrogen bonding receptor. The donor (C43–H43) from bdc2− moiety formed hydrogen bonding with oxygen atom (O13) of bdc2− moiety as hydrogen bonding receptor.

3.4. Luminescence Spectra

The luminescence spectra of H2L and its ZnII–EuIII complex were measured in 3.0 × 10−5 M methanol solution (Figure 4). The ligand H2L exhibited an intense emission peak at ca. 442 nm upon excitation at 352 nm which should be assigned to intra-ligand π–π* transition [69,70]. The ZnII–EuIII complex showed slightly weak photoluminescence with maximum emission at ca. 445 nm upon excitation at 352 nm. In the luminescence spectra, only a band at ca. 375–650 nm instead of the f–f emission expected for EuIII ions [71]. In the ZnII–EuIII complex, the emission from EuIII is quenched which may due to thermal quenching of the 5D0 level of EuIII by LMCT process [72]. The concentration of H2L and its ZnII–EuIII complex used are all 3.0 × 10−5 M, indicating that the relative strengths of H2L and its ZnII–EuIII complex are independent of the concentration.

4. Conclusions

In summary, a new heterohexanuclear ZnII–EuIII dimer was synthesized and characterized. In the ZnII–EuIII complex, there are two crystallographically equivalent [(ZnL)2Eu] moieties which are linked by two bdc2– auxiliary ligand leading to a heterohexanuclear dimer. The ZnII atom possesses a penta-coordinated environment and adopts a slightly distorted triangular bipyramid geometry and the deca-coordinated EuIII atom adopts a distorted bicapped square antiprism. In addition, the luminescence spectrum of the ZnII–EuIII complex indicated that the coordination of EuIII atoms led to the fluorescence quenching of H2L.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (21761018) and the Program for Excellent Team of Scientific Research in Lanzhou Jiaotong University (201706), which is gratefully acknowledged.

Author Contributions

Wen-Kui Dong conceived and designed the experiments; Quan-Peng Kang and Xiao-Yan Li performed the experiments; Jian-Chun Ma analyzed the data; Wen-Ting Guo contributed reagents/materials/analysis tools; Xiao-Yan Li and Wen-Ting Guo wrote the paper.

Conflicts of Interest

The authors declare no competing financial interests.

References

  1. Dong, X.Y.; Akogun, S.F.; Zhou, W.M.; Dong, W.K. Tetranuclear Zn(II) complex based on an asymmetrical salamo-type chelating ligand: Synthesis, structural characterization, and fluorescence property. J. Chin. Chem. Soc. 2017, 64, 412–419. [Google Scholar] [CrossRef]
  2. 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]
  3. Chen, C.Y.; Zhang, J.W.; Zhang, Y.H.; Yang, Z.H.; Wu, H.L. Gadolinium(III) and dysprosium(III) complexes with a Schiff base bis(N-salicylidene)-3-oxapentane-1,5-diamine: Synthesis, characterization, antioxidation, and DNA-binding studies. J. Coord. Chem. 2015, 68, 1054–1071. [Google Scholar] [CrossRef]
  4. Sun, Y.X.; Zhang, Y.J.; Meng, W.S.; Li, X.R.; Lu, R.E. Synthesis and crystal structure of new nickel(II) complex with Salen-type bisoxime ligand. Asian J. Chem. 2014, 26, 416–418. [Google Scholar]
  5. Chai, L.Q.; Huang, J.J.; Zhang, H.S.; Zhang, Y.L.; Zhang, J.Y.; Li, Y.X. 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, 526–533. [Google Scholar] [CrossRef] [PubMed]
  6. 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]
  7. 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 Part A 2015, 135, 342–350. [Google Scholar] [CrossRef] [PubMed]
  8. Dong, X.Y.; Sun, Y.X.; Wang, L.; Li, L. Synthesis and structure of a penta- and hexa-coordinated tri-nuclear cobalt(II) complex. J. Chem. Res. 2012, 36, 387–390. [Google Scholar] [CrossRef]
  9. 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]
  10. 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]
  11. Zhang, H.; Dong, W.K.; Zhang, Y.; Akogun, S.F. Naphthalenediol-based bis(Salamo)-type homo- and heterotrinuclear cobalt(II) complexes: Syntheses, structures and magnetic properties. Polyhedron 2017, 133, 279–293. [Google Scholar] [CrossRef]
  12. 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]
  13. Li, L.H.; Dong, W.K.; Zhang, Y.; Akogun, S.F.; Xu, L. Syntheses, structures and catecholase activities of homo-and hetero-trinuclear cobalt(II) complexes constructed from an acyclic naphthalenediol-based bis(salamo)-type ligand. Appl. Organomet. Chem. 2017, 31, e3818. [Google Scholar] [CrossRef]
  14. 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]
  15. 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]
  16. Wu, H.L.; Pan, G.L.; Wang, H.; Wang, X.L.; Bai, Y.C.; Zhang, Y.H. 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 Biol. 2014, 135, 33–43. [Google Scholar] [CrossRef] [PubMed]
  17. 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]
  18. Wu, H.L.; Pan, G.L.; Bai, Y.C.; Wang, H.; Kong, J.; Shi, F.R.; Zhang, Y.H.; Wang, X.L. A Schiff base bis(N-salicylidene)-3-oxapentane-1,5-diamine and its yttrium(III) complex: Synthesis, crystal structure, DNA-binding properties, and antioxidant activities. J. Coord. Chem. 2013, 66, 2634–2646. [Google Scholar] [CrossRef]
  19. Li, X.Y.; Kang, Q.P.; Liu, L.Z.; Ma, J.C.; Dong, W.K. Trinuclear Co(II) and Mononuclear Ni(II) Salamo-Type Bisoxime Coordination Compounds. Crystals 2018, 8, 43. [Google Scholar] [CrossRef]
  20. 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]
  21. Wu, H.L.; Pan, G.L.; Bai, Y.C.; Wang, H.; Kong, J.; Shi, F.; 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]
  22. 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]
  23. 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]
  24. Dong, W.K.; Wang, Z.K.; Li, G.; Zhao, M.M.; Dong, X.Y.; Liu, S.H. Syntheses, crystal structures, and properties of a Salamo-type tetradentate chelating ligand and its pentacoordinated copper(II) complex. Z. Anorg. Allg. Chem. 2013, 639, 2263–2268. [Google Scholar] [CrossRef]
  25. Chai, L.Q.; Zhang, K.Y.; Tang, L.J.; Zhang, J.Y.; Zhang, H.S. Two mono- and dinuclear Ni(II) complexes constructed from quinazoline-type ligands: Synthesis, x-ray structures, spectroscopic, electrochemical, thermal, and antimicrobial studies. Polyhedron 2017, 130, 100–107. [Google Scholar] [CrossRef]
  26. 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]
  27. Wang, F.; Gao, L.; Zhao, Q.; Zhang, Y.; Dong, W.K.; Ding, Y.J. A highly selective fluorescent chemosensor for CN based on a novel bis(salamo)-type tetraoxime ligand. Spectrochim. Acta Part A 2018, 190, 111–115. [Google Scholar] [CrossRef] [PubMed]
  28. Dong, Y.J.; Li, X.L.; Zhang, Y.; Dong, W.K. A highly selective visual and fluorescent sensor for Pb2+ and Zn2+ and crystal structure of Cu2+ complex based-on a novel single-armed Salamo-type bisoxime. Supramol. Chem. 2017, 29, 518–527. [Google Scholar] [CrossRef]
  29. Dong, W.K.; Li, X.L.; Wang, L.; Zhang, Y.; Ding, Y.J. A new application of Salamo-type bisoximes: As a relay-sensor for Zn2+/Cu2+ and its novel complexes for successive sensing of H+/OH. Sens. Actuators B 2016, 229, 370–378. [Google Scholar] [CrossRef]
  30. Wang, B.J.; Dong, W.K.; Zhang, Y.; Akogun, S.F. A novel relay-sensor for highly sensitive and selective detection of Zn2+/Pic and fluorescence on/off switch response of H+/OH. Sens. Actuators B 2017, 247, 254–264. [Google Scholar] [CrossRef]
  31. Dong, W.K.; Akogun, S.F.; Zhang, Y.; Sun, Y.X.; Dong, X.Y. A reversible “turn-on” fluorescent sensor for selective detection of Zn2+. Sens. Actuators B 2017, 238, 723–734. [Google Scholar] [CrossRef]
  32. Song, X.Q.; Liu, P.P.; Xiao, Z.R.; Li, X.; Liu, Y.A. Four polynuclear complexes based on a versatile salicylamide salen-like ligand: Synthesis, structural variations and magnetic properties. Inorg. Chim. Acta 2015, 438, 232–244. [Google Scholar] [CrossRef]
  33. 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]
  34. 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] [PubMed]
  35. 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, 41, 4966–4973. [Google Scholar] [CrossRef]
  36. 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]
  37. 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]
  38. Dong, W.K.; Zheng, S.S.; Zhang, J.T.; Zhang, Y.; Sun, Y.X. Luminescent properties of heterotrinuclear 3d–4f complexes constructed from a naphthalenediol-based acyclic bis(salamo)-type ligand. Spectrochim. Acta Part A 2017, 184, 141–150. [Google Scholar] [CrossRef] [PubMed]
  39. Wang, L.; Li, X.Y.; Zhao, Q.; Li, L.H.; Dong, W.K. Fluorescence properties of heterotrinuclear Zn(II)–M(II) (M = Ca, Sr and Ba) bis(salamo)-type complexes. RSC Adv. 2017, 7, 48730–48737. [Google Scholar] [CrossRef]
  40. Yang, Y.H.; Hao, J.; Dong, Y.J.; Wang, G.; Dong, W.K. Two znic(II) complexes constructed from a bis(salamo)-type tetraoxime ligand: Syntheses, crystal structures and luminescence properties. Chin. J. Inorg. Chem. 2017, 33, 1280–1292. [Google Scholar]
  41. Dong, X.Y.; Li, X.Y.; Liu, L.Z.; Zhang, H.; Ding, Y.J.; Dong, W.K. Tri- and hexanuclear heterometallic Ni(II)–M(II) (M = Ca, Sr and Ba) bis(salamo)-type complexes: Synthesis, structure and fluorescence properties. RSC Adv. 2017, 7, 48394–48403. [Google Scholar] [CrossRef]
  42. 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]
  43. Dong, W.K.; Zhang, F.; Li, N.; Xu, L.; Zhang, Y.; Zhang, J.; Zhu, L.C. Trinuclear cobalt(II) and zinc(II) salamo–type complexes: Syntheses, crystal structures, and fluorescent properties. Z. Anorg. Allg. Chem. 2016, 642, 532–538. [Google Scholar] [CrossRef]
  44. Song, X.Q.; Cheng, G.Q.; Liu, Y.A. Enhanced Tb(III) luminescence by d10 transition metal coordination. Inorg. Chim. Acta 2016, 450, 386–394. [Google Scholar] [CrossRef]
  45. Wang, L.; Hao, J.; Zhai, L.X.; Zhang, Y.; Dong, W.K. Synthesis, crystal structure, luminescence, electrochemical and antimicrobial properties of bis(salamo)-based Co(II) complex. Crystals 2017, 7, 277. [Google Scholar] [CrossRef]
  46. Lu, W.G.; Su, C.Y.; Lu, T.B.; Jiang, L.; Chen, J.M. Two stable 3D metal-organic frameworks constructed by nanoscale cages via sharing the single-layer walls. J. Am. Chem. Soc. 2006, 128, 34–35. [Google Scholar] [CrossRef] [PubMed]
  47. Zhou, H.C.; Kitagawa, S. Metal-organic frameworks (MOFs). Chem. Soc. Rev. 2014, 43, 5415–5418. [Google Scholar] [CrossRef] [PubMed]
  48. Hao, J.; Li, L.L.; Zhang, J.T.; Akogun, S.F.; Wang, L.; Dong, W.K. Four homo- and hetero-bismetallic 3d/3d-2s complexes constructed from a naphthalenediol-based acyclic bis(salamo)-type tetraoxime ligand. Polyhedron 2017, 134, 1–10. [Google Scholar] [CrossRef]
  49. Xu, L.; Zhu, L.C.; Ma, J.C.; Zhang, Y.; Zhang, J.; Dong, W.K. Syntheses, structures and spectral properties of mononuclear CuII and dimeric ZnII complexes based on an asymmetric Salamo-type N2O2 ligand. Z. Anorg. Allg. Chem. 2015, 641, 2520–2524. [Google Scholar] [CrossRef]
  50. Dong, X.Y.; Gao, L.; Wang, F.; Zhang, Y.; Dong, W.K. Tri- and mono-nuclear zinc(II) complexes based on half- and mono-salamo chelating ligands. Crystals 2017, 7, 267. [Google Scholar] [CrossRef]
  51. Dong, Y.J.; Ma, J.C.; Zhu, L.C.; Dong, W.K.; Zhang, Y. Four 3d–4f heteromultinuclear zinc(II)–lanthanide(III) complexes constructed from a distinct hexadentate N2O2-type ligand: Syntheses, structures and photophysical properties. J. Coord. Chem. 2017, 70, 103–115. [Google Scholar] [CrossRef]
  52. 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 O-donor auxiliary ligand: Syntheses, structures and magnetic properties. New J. Chem. 2016, 40, 6998–7010. [Google Scholar] [CrossRef]
  53. Dong, W.K.; Ma, J.C.; Zhu, L.C.; Zhang, Y. Self-assembled zinc(II)-lanthanide(III) heteromultinuclear complexes constructed from 3-MeOsalamo ligand: Syntheses, structures and luminescent properties. Cryst. Growth Des. 2016, 16, 6903–6914. [Google Scholar] [CrossRef]
  54. 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]
  55. Sheldrick, G.M. SHELXS-2016, Program for Crystal Structure Solution; University of Göttingen: Göttingen, Germany, 2016. [Google Scholar]
  56. Sheldrick, G.M. SHELXL-2016, Program for Crystal Structure Refinement; University of Göttingen: Göttingen, Germany, 2016. [Google Scholar]
  57. Wang, P.; Zhao, L. Synthesis and crystal structure of supramolecular copper(II) complex based on N2O2 coordination sphere. Asian J. Chem. 2015, 4, 1424–1426. [Google Scholar] [CrossRef]
  58. Chai, L.Q.; Wang, G.; Sun, Y.X.; Dong, W.K.; Zhao, L.; Gao, X.H. Synthesis, crystal structure, and fluorescence of an unexpected dialkoxo-bridged dinuclear copper(II) complex with bis(salen)-type tetraoxime. J. Coord. Chem. 2012, 65, 1621–1631. [Google Scholar] [CrossRef]
  59. Wang, L.; Kang, Q.P.; Hao, J.; Zhang, Y.; Bai, Y.; Dong, W.K. Two trinuclear cobalt(II) salamo–type complexes: Syntheses, crystal structures, solvent effect and fluorescent properties. Chin. J. Inorg. Chem. 2018, 34, 525–533. [Google Scholar]
  60. Wang, L.; Ma, J.C.; Dong, W.K.; Zhu, L.C.; Zhang, Y. A novel Self–assembled nickel(II)–cerium(III) heterotetranuclear dimer constructed from N2O2-type bisoxime and terephthalic acid: Synthesis, structure and photophysical properties. Z. Anorg. Allg. Chem. 2016, 642, 834–839. [Google Scholar] [CrossRef]
  61. 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]
  62. 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]
  63. Dong, X.Y.; Kang, Q.P.; Jin, B.X.; Dong, W.K. A dinuclear nickel(II) complex derived from an asymmetric Salamo-type N2O2 chelate ligand: Synthesis, structure and optical properties. Z. Naturforsch. 2017, 72, 415–420. [Google Scholar] [CrossRef]
  64. Gao, L.; Liu, C.; Wang, F.; Dong, W.K. Tetra-, penta- and hexa-coordinated transition metal complexes constructed from coumarin-containing N2O2 ligand. Crystals 2018, 8, 77. [Google Scholar] [CrossRef]
  65. Addison, A.W.; Rao, T.N.; Reedijk, J.; Rijn, J.V.; Verschoor, G.C. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen–sulphur donor ligands; the crystal and molecular structure of aqua[1,7-bis(N-methylbenzimidazol-2′-yl)-2,6-dithiaheptane]copper(II) perchlorate. J. Chem. Soc. Dalton Trans. 1984, 7, 1349–1356. [Google Scholar] [CrossRef]
  66. 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]
  67. Li, G.; Hao, J.; Liu, L.Z.; Zhou, W.M.; Dong, W.K. Syntheses, crystal structures and thermal behaviors of two supramolecular salamo-type cobalt(II) and zinc(II) complexes. Crystals 2017, 7, 217. [Google Scholar] [CrossRef]
  68. Chen, L.; Dong, W.K.; Zhang, H.; Zhang, Y.; Sun, Y.X. Structural variation and luminescence properties of tri- and dinuclear CuII and ZnII complexes constructed from a naphthalenediol-based bis(Salamo)-type ligand. Cryst. Growth Des. 2017, 17, 3636–3648. [Google Scholar] [CrossRef]
  69. Ma, J.C.; Dong, X.Y.; Dong, W.K.; Zhang, Y.; Zhu, L.C.; Zhang, J.T. An unexpected dinuclear Cu(II) complex with a bis(Salamo) chelating ligand: Synthesis, crystal structure, and photophysical properties. J. Coord. Chem. 2016, 69, 149–159. [Google Scholar] [CrossRef]
  70. Tao, C.H.; Ma, J.C.; Zhu, L.C.; Zhang, Y.; Dong, W.K. Heterobimetallic 3d–4f Zn(II)–Ln(III) (Ln=Sm, Eu, Tb and Dy) complexes with a N2O4 bisoxime chelate ligand and a simple auxiliary ligand Py: Syntheses, structures and luminescence properties. Polyhedron 2017, 128, 38–45. [Google Scholar] [CrossRef]
  71. Yang, X.P.; Ben, J.M.; Hahn, P.; Jones, R.A.; Wong, W.K.; Stevenson, K.J. Synthesis of an octanuclear Eu(III) cage from Eu42+: Chloride anion encapsulation, luminescence, and reversible MeOH adsorption via a porous supramolecular architecture. Inorg. Chem. 2007, 46, 7050–7054. [Google Scholar] [CrossRef] [PubMed]
  72. Dong, Y.J.; Dong, X.Y.; Dong, W.K.; Zhang, Y.; Zhang, L.S. Three asymmetric Salamo-type copper(II) and cobalt(II) complexes: Syntheses, structures, fluorescent properties. Polyhedron 2017, 123, 305–315. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of the ZnII–EuIII complex.
Scheme 1. Synthesis of the ZnII–EuIII complex.
Crystals 08 00154 sch001
Figure 1. The UV–Vis spectra of H2L and its ZnII–EuIII complex (cm−1).
Figure 1. The UV–Vis spectra of H2L and its ZnII–EuIII complex (cm−1).
Crystals 08 00154 g001
Figure 2. (a) Molecular structure and atom numberings of the ZnII–EuIII complex with 30% probability displacement ellipsoids (hydrogen atoms are omitted for clarity); (b) coordination polyhedrons for ZnII and EuIII atoms.
Figure 2. (a) Molecular structure and atom numberings of the ZnII–EuIII complex with 30% probability displacement ellipsoids (hydrogen atoms are omitted for clarity); (b) coordination polyhedrons for ZnII and EuIII atoms.
Crystals 08 00154 g002
Figure 3. View of the intramolecular hydrogen bondings of the ZnII–EuIII complex.
Figure 3. View of the intramolecular hydrogen bondings of the ZnII–EuIII complex.
Crystals 08 00154 g003
Figure 4. Luminescence spectra of H2L and its ZnII–EuIII complex in methanol (3.0 × 105 M) upon excitation at 352 nm.
Figure 4. Luminescence spectra of H2L and its ZnII–EuIII complex in methanol (3.0 × 105 M) upon excitation at 352 nm.
Crystals 08 00154 g004
Table 1. Crystal data and refinement parameters for the ZnII–EuIII complex.
Table 1. Crystal data and refinement parameters for the ZnII–EuIII complex.
ZnII–EuIII Complex
FormulaC88H80Cl2Eu2Zn4N8O32
Formula weight2397.90
Temperature (K)291(2)
Wavelength (Å)0.71073
Crystal systemtriclinic
Space groupP-1
a (Å)12.5053(15)
b (Å)16.125(3)
c (Å)16.5068(13)
α (°)68.518(13)
β (°)82.216(12)
γ (°)86.716(14)
V3)3068.6(8)
Z1
Dcalc (g∙cm−3)1.298
μ (mm−1)1.885
F (000)1200
Crystal size (mm)0.18 × 0.20 × 0.22
θ Range (°)2.123–25.682
Index ranges−14 ≤ h ≤ 15
–18 ≤ k ≤ 19
0 ≤ l ≤ 20
Reflections collected11,612
Independent reflections11,611
Rint0.0726
Completeness to θ99.6% (θ = 25.242)
Data/restraints/parameters11611/0/627
GOF1.029
Final R1, wR2 indices0.0482, 0.1182
R1, wR2 indices (all data)0.0644, 0.1215
Largest differences peak and hole (e Å−3)1.696/−0.554
GOF: Goodness-of-fit on F.
Table 2. The major IR spectra of H2L and its ZnII–EuIII complex (cm−1).
Table 2. The major IR spectra of H2L and its ZnII–EuIII complex (cm−1).
Compoundν(O–H)ν(C=N)ν(Ar–O)ν(C=C)νs(COO)νas(COO)
H2L3131161812531374
Complex-16071217131414681572
Table 3. Selected bond lengths (Å) and angles (°) of the ZnII–EuIII complex.
Table 3. Selected bond lengths (Å) and angles (°) of the ZnII–EuIII complex.
BondLengthsBondLengths
N1–Zn12.116(5)N2–Zn12.046(5)
O2–Zn12.000(4)O5–Zn12.052(4)
O16–Zn11.983(4)N3–Zn22.135(5)
N4–Zn22.005(5)O8–Zn21.965(4)
O11–Zn22.056(3)O13–Zn21.986(3)
Eu1–O52.383(4)Eu1–O22.385(3)
Eu1–O152.418(3)Eu1–O82.445(4)
Eu1–O112.452(4)Eu1–O12.637(3)
Eu1–O72.670(4)Eu1–O122.832(4)
Eu1–O62.844(4)Eu1–O142.378(4)
BondAnglesBondAngles
O16–Zn1–O2111.62(15)O16–Zn1–O597.39(15)
O2–Zn1–O576.97(14)O16–Zn1–N2118.07(18)
O2–Zn1–N2129.54(18)O5–Zn1–N288.3(2)
O16–Zn1–N197.88(17)O2–Zn1–N188.36(17)
O5–Zn1–N1161.88(16)N2–Zn1–N193.1(2)
O8–Zn2–O13114.67(16)O8–Zn2–N4125.97(18)
O13–Zn2–N4118.96(19)O8–Zn2–O1179.89(15)
O13–Zn2–O1197.49(15)N4–Zn2–O1187.20(18)
O8–Zn2–N386.80(19)O13–Zn2–N398.07(17)
N4–Zn2–N391.8(2)O11–Zn2–N3162.72(17)
O14–Eu1–O5107.44(12)O14–Eu1–O2151.04(12)
O5–Eu1–O263.88(13)O14–Eu1–O1572.43(11)
O5–Eu1–O1571.56(12)O2–Eu1–O1578.67(12)
O14–Eu1–O877.38(12)O5–Eu1–O8117.81(13)
O2–Eu1–O8131.54(12)O15–Eu1–O8149.80(12)
O14–Eu1–O1170.71(11)O5–Eu1–O11177.52(11)
O2–Eu1–O11116.88(12)O15–Eu1–O11106.14(12)
O8–Eu1–O1163.68(13)O14–Eu1–O1138.67(11)
O5–Eu1–O1113.38(12)O2–Eu1–O160.79(11)
O15–Eu1–O1126.52(11)O8–Eu1–O178.26(11)
O11–Eu1–O168.66(11)O14–Eu1–O7125.52(12)
O5–Eu1–O769.27(12)O2–Eu1–O779.07(12)
O15–Eu1–O7140.37(12)O8–Eu1–O760.58(12)
O11–Eu1–O7113.11(12)O1–Eu1–O766.58(12)
O14–Eu1–O1297.83(12)O5–Eu1–O12121.21(12)
O2–Eu1–O1268.48(12)O15–Eu1–O1267.04(12)
O8–Eu1–O12119.14(12)O11–Eu1–O1257.94(11)
O1–Eu1–O1266.45(11)O7–Eu1–O12131.59(12)
O14–Eu1–O665.87(13)O5–Eu1–O659.08(12)
O2–Eu1–O6120.77(13)O15–Eu1–O697.08(12)
O8–Eu1–O669.06(12)O11–Eu1–O6120.81(12)
O1–Eu1–O6132.51(11)O7–Eu1–O667.58(12)
O12–Eu1–O6160.75(11)
Table 4. Hydrogen bonding interactions (Å, °) of the ZnII–EuIII complex.
Table 4. Hydrogen bonding interactions (Å, °) of the ZnII–EuIII complex.
D–H···AD–HH···AD···AD–H···A
C1–H1B···O80.962.563.281(6)132
C9–H9A···O160.972.413.311(10)154
C19–H19B···O20.962.553.280(8)133
C36–H36B···O150.962.543.148(7)122
C43–H43···O130.932.462.773(7)100

Share and Cite

MDPI and ACS Style

Guo, W.-T.; Li, X.-Y.; Kang, Q.-P.; Ma, J.-C.; Dong, W.-K. Structural and Luminescent Properties of Heterobimetallic Zinc(II)-Europium(III) Dimer Constructed from N2O2-Type Bisoxime and Terephthalic Acid. Crystals 2018, 8, 154. https://doi.org/10.3390/cryst8040154

AMA Style

Guo W-T, Li X-Y, Kang Q-P, Ma J-C, Dong W-K. Structural and Luminescent Properties of Heterobimetallic Zinc(II)-Europium(III) Dimer Constructed from N2O2-Type Bisoxime and Terephthalic Acid. Crystals. 2018; 8(4):154. https://doi.org/10.3390/cryst8040154

Chicago/Turabian Style

Guo, Wen-Ting, Xiao-Yan Li, Quan-Peng Kang, Jian-Chun Ma, and Wen-Kui Dong. 2018. "Structural and Luminescent Properties of Heterobimetallic Zinc(II)-Europium(III) Dimer Constructed from N2O2-Type Bisoxime and Terephthalic Acid" Crystals 8, no. 4: 154. https://doi.org/10.3390/cryst8040154

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