Next Article in Journal
Partial and Total Substitution of Zn by Mg in the Cu2ZnSnS4 Structure
Next Article in Special Issue
Structural Variations in Manganese Halide Chain Compounds Mediated by Methylimidazolium Isomers
Previous Article in Journal
Influence of Nonlocality on Transmittance and Reflectance of Hyperbolic Metamaterials
Previous Article in Special Issue
Zn/Ni and Zn/Pd Heterobimetallic Coordination Polymers with [SSC-N(CH2COO)2]3− Ligands
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Self-Assembled Hetero-Bimetallic [Ni(II)-Sm(III)] Coordination Polymer Constructed from a Salamo-Like Ligand and 4,4′-Bipyridine: Synthesis, Structural Characterization, and Properties

1
School of Automation and Electrical Engineering, Lanzhou Jiaotong University, Lanzhou 730070, China
2
PetroChina Lanzhou Huanqiu Petrochemical Engineering Company, Lanzhou 730060, China
3
School of Chemical and Biological Engineering, Lanzhou Jiaotong University, Lanzhou 730070, China
*
Authors to whom correspondence should be addressed.
Crystals 2020, 10(7), 579; https://doi.org/10.3390/cryst10070579
Submission received: 9 June 2020 / Revised: 2 July 2020 / Accepted: 2 July 2020 / Published: 4 July 2020
(This article belongs to the Special Issue Coordination Polymers: Structure, Bonding and Applications)

Abstract

:
An unusual self-assembled hetero-bimetallic [Ni(II)-Sm(III)] coordination polymer, [Ni(L)Sm(NO3)3(4,4′-bipy)]n, is prepared through a hexadentate chelating ligand 2,2′-[1,2-ethylenedioxybis(nitrilomethylidyne)]diphenol (H2L). The Ni(II)-Sm(III) coordination polymer is validated through elemental analyses, Fourier-transform infrared and UV-Visible spectroscopies, and X-ray single-crystal diffraction. The Ni(II) atom forms a twisted six-coordinated octahedron, and the Sm(III) atom is ten-coordinated, adopting a twisted bicapped square antiprism. An infinite three-dimensional-layer supramolecular structure is obtained through extensive π···π stacking and intermolecular hydrogen bonding interactions. The polymer has a good antibacterial effect against Staphylococcus aureus.

Graphical Abstract

1. Introduction

Salen-based compounds have been reported and studied previously [1,2,3,4,5,6]. Products gained by the reaction of salicylaldehyde with diamines are significant and common ligands for inorganic and coordination chemistry. The salen-based compounds and their transition metal complexes have attracted increased attention in the last few years. They often have unique crystal structures [7,8,9,10,11,12,13,14,15], interesting industrial catalyses, and ion recognitions [16,17,18,19], including their biological activities [20,21,22,23,24,25,26,27]. Based on the salen-based ligand, new salamo-based ligands were synthesized by introducing –C=N–O– groups [28,29,30,31,32,33,34,35,36,37,38,39,40,41,42]. Compared to salen-like compounds, the preparation of Salamo-like compounds is more laborious. Moreover, with the O-alkyloxime moiety [–CH=N–O–(CH2)n–O–N=CH–] replacing the [–CH=N–(CH2)n–N=CH–] part, the high electronegativity of the oxygen atom can strongly influence the electronic properties of the N2O2-donor environment, which can give rise to the novel structures and potential application values of the obtained complexes [1,13]. Their metal complexes studied include supramolecular modes of action, the growth of crystals, and optical and magnetic properties [43,44,45,46,47,48,49].
Symmetrical 3-alkoxy salamo-based ligands give relatively easy access to 3d–4f complexes, because the N2O2 cavity in the ligand can coordinate with transition metal ions, while the outer coordination site of the O4 cavity can accommodate rare-earth elements. Interestingly, anionic and solvent molecules are also coordinated with metal atoms. Transition metal atoms usually form distorted octahedral or triangular biconical configurations, while rare-earth atoms usually form a twisted tricapped trigonal prism with nine-coordination or form a twisted bicapped square antiprismatic coordination arrangement with ten-coordination [47].
According to the previous work of our research group [47], heterobimetallic [Ni(L)Ln] units can be adjusted to form a coordination polymer with a three-dimensional structure by adjusting the ratio of reactants and adding the auxiliary ligand 4,4′-bipyridine. In this paper, the symmetrical 3-MeO hexadentate chelating ligand H2L was prepared and used for the formation of 3d–4f coordination polymers, together with the exo-dentate auxiliary ligand 4,4′-bipy.

2. Experimental

2.1. Materials and Physical Measurements

3-Methoxysalicylaldehyde (99%) bought from Alfa Aesar (Tianjin, China) was used. The other reagents (Ni(OAc)2·4H2O, Sm(NO3)3·6H2O, 4,4′-bipy, and 1,2-dibromoethane) and solvents (ethanol, trichloromethane, etc.) were purchased from Shanghai Darui Chemical Fine Chemicals Company (Tianjin, China), and were of analytical purity.
Elemental analyses for Ni and Sm were carried out on an IRIS ER/SWP-1 ICP atomic emission spectrometer (Elementar, Berlin, Germany). C, H, and N analyses were determined on Elementar GmbH, VarioEL V3.00 automatic elemental analysis equipment (Elementar, Berlin, Germany). Melting points were tested through an X4 microscopic melting point apparatus (Beijing Taike Instrument Limited Company, Beijing, China).
FT-IR spectroscopy was carried out with a VERTEX70 FT-IR spectrophotometer (Bruker, Billerica, MA, USA, KBr, 400–4000 cm−1).
UV/Vis absorption spectra were recorded by a UV-3900 spectrophotometer (Hitachi, Tokyo, Japan). Fluorescence was measured on an F-7000 FL 220-240V spectrophotometer (Hitachi, Tokyo, Japan). 1H NMR spectra were measured by a Bruker AVANCE DRX-400 spectrometer (Bruker AVANCE, Billerica, MA, USA).
X-ray single-crystal structure determination was carried out on a Bruker APEX-II CCD diffractometer (Bruker AVANCE, Billerica, MA, USA).
The antimicrobial activities of H2L, Ni(II) acetate, and the title polymer were tested by a disk diffusion method, Staphylococcus aureus being selected as Gram-positive bacteria.

2.2. Preparation of H2L

The preparation of H2L is depicted in Scheme 1.
First, 1,2-di(aminooxy)ethane was prepared based on a similar method previously reported [47]. 3-Methoxysalicylaldehyde (304.2 mg, 0.02 mmol) was added to 1,2-di(aminooxy)ethane (92.1 mg, 0.01 mmol). The solvent of the two reactants was ethanol, each with a volume of 25 mL. At the temperature of 55– 60 °C, the product was gained by reaction for 7 h. Colorless crystals of the ligand H2L were obtained by recrystallization from ethanol, yield: 85%.m.p: 130–132 °C. 1H NMR (400 MHz, CDCl3, 25 °C, TMS): δ = 3.90 (s, 6H, –OMe), 4.48 (s, 4H, –CH2), 6.82 (dd, J = 7.9, 1.9 Hz, 2H, –ArH), 6.87 (t, J = 7.9 Hz, 2H, –ArH), 6.90 (dd, J = 7.9, 1.9 Hz, 2H, –ArH), 8.25 (s, 2H, –N=CH), 9.75 (s, 2H,–OH). Anal. Calcd. for C18H20N2O6 (%): C 59.99; H 5.59; N 7.77. Found C 60.22; H 5.67; N 7.53.

2.3. Preparation of the Ni(II)-Sm(III) Polymer

An ethanol solution (2 mL) of Ni(OAc)2·4H2O (2.49 mg, 0.01 mmol) and an ethanol solution (2 mL) of Sm(NO3)3·6H2O (4.45 mg, 0.01 mmol) were successively added to a trichloromethane solution (2 mL) of H2L (3.60 mg, 0.01 mmol), and the same amount of ethanol solution (2 mL) of 4,4′-bipy (1.56 mg, 0.01 mmol) was then added to the mixture. After stirring for ten minutes, the mixed solution was filtered. After standing for ten days, several colorless block-shaped single crystals were obtained by natural evaporation (Scheme 2). Yield: 37%. Anal. Calcd. for C28H26SmN7NiO15 (%): C 36.97; H 2.88; N 10.78; Ni 6.45; Sm 16.53. Found C 34.88; H 2.79; N 10.81; Ni 6.43; Sm 16.46.

2.4. Crystal Structure Determination of the Ni(II)-Sm(III) Polymer

The crystal structure was obtained by X-ray single-crystal diffraction on a Bruker APEX-II CCD diffractometer (Table 1). The X-ray single-crystal diffraction data of the Ni(II)-Sm(III) polymer were recorded and collected by using a Bruker APEX-II CCD diffractometer with Mo-Kα radiation (λ = 0.71073 Å), and corrected via the Lorentz and polarization factor with a multi-scan method. The program Shelxs-2018 and Fourier difference techniques were used to solve the structure. It was corrected via the full-matrix least-squares on F2. The structure included a large void, and the positive or negative ions and the solvent water molecules in the void could not be confirmed, owing to it being highly disordered. Thus, SQUEEZE in the PLATON program was used to remove the highly disordered ions and solvent. (Solvent Accessible Volume = 1275, Electrons Found in S.A.V. = 133). Anisotropic displacement parameters were applied for the non-hydrogen atoms and isotropic parameters for the hydrogen atoms. Hydrogen atoms were added geometrically and refined using a riding model.
CCDC 1939692 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.

3. Results and Discussion

H2L and the title polymer have been gained and validated by IR spectra, UV/Vis absorption spectra, and X-ray single-crystal diffraction and fluorescence spectra. Their antimicrobial activities were investigated. The compounds were stable in air.

3.1. PXRD Analysis

A PXRD test was conducted for the title polymer to demonstrate whether the structure is truly representative. The PXRD result of the title polymer is given in Figure 1. Through detailed comparison, the experimental result is coincident with the simulation result. Therefore, the synthesized samples have high purity and can be used for further study of spectral characterization and fluorescence properties.

3.2. FT-IR Spectra of H2L and the Title Polymer

The FT-IR spectra of H2L and the title polymer are depicted in Figure 2.
The band of H2L appearing at 3437 cm−1 is attributed to the O-H stretching band [47]. The characteristic C=N stretching band is found at approximately 1606 cm−1 in the spectrum of H2L, and moves to a lower frequency in the polymer, indicating the metal coordination at the nitrogen atoms of oxime groups [15,50]. The Ar–O stretching vibration band of H2L emerges as an intensive band at 1253 cm−1 as reported for analogous salen-based compounds [6,28,43], while the title polymer shows this band at 1210 cm−1, again due to complexation.

3.3. Description of the Molecular Structure of the Title Polymer

The structure of the heterobimetallic [Ni(II)-Sm(III)] salamo-based coordination polymer is depicted in Figure 3, complemented by data in Table 2.
The title polymer crystallizes in the monoclinic space group P21/c with Z = 4, consisting of one Sm(III) atom, one Ni(II) atom, one deprotonated (L)2− unit, one 4,4′-bipyridine ligand, three bidentate chelating nitrate anions, and three water molecules.
Each Ni(II) (Ni1) atom is sited in a N2O2-donor cavity of the salamo ligand, and is octahedrally hexacoordinated by two oxime N atoms (N2 and N1), two phenoxo O atoms (O5 and O2), and two N atoms (N4 or N3) from two 4,4′-bipyridine ligands [50]. The equatorial coordination sites have distances Ni1–N2 = 2.030(4) Å, Ni1–N1 = 2.072(4) Å, Ni1–O2 = 2.0370(19) Å, and Ni1–O5 = 2.0399(19) Å, and the axial positions are at the distances Ni1–N4 = 2.167(2) and Ni1–N3 = 2.172(2) Å (Table 2).
The Sm(III) is surrounded by an O10 environment containing four O atoms (O6, O5, O2, and O1) from the (L)2− units and six O atoms (O14, O13, O11, O10, O8, and O7) from the three bidentate nitrate ligands. The decacoordinated Sm(III) atom possesses a geometry of a twisted bicapped square antiprism. The Sm–O distance is in the normal range of 2.353(2)–2.653(4) Å.

3.4. Supramolecular Interactions in the Title Polymer

As is depicted in Figure 4, complemented by data in Table 3, in the title polymer structure, five of the couples are weak intramolecular hydrogen bonds (C1–H1B···O10, C9–H9A···N3, C23–H23···O2, C28–H28···O7, and C24–H24···N1) and two intermolecular hydrogen bonds (C4–H4···O9 and C9–H9B···O13), which indicate that the supramolecular interactions exist in the title polymer [51,52,53,54,55,56]. The donor (C1–H1B) from the 3-methoxy substituent group from the (L)2− moiety forms a hydrogen bond with an O atom (O10) of one bidentate chelating nitrate as a hydrogen bonding receptor. The donors (C23–H23, C24–H24, and C28–H28) of the 4,4′-bipy moieties form intermolecular hydrogen bond interactions with O and N atoms (O2, N1 and O7) of the fully deprotonated (L)2− moieties as hydrogen bond receptors. The donor (C9–H9A) of the (L)2− unit is contained in the hydrogen bond interaction with the N atom (N3) of the 4,4′-bipy ligand as a hydrogen bonding receptor. These ligands are connected by C−H···π (C19–H19···Cg2) interactions (Figure 5), forming a one-dimensional chain structure (Figure 6).
The π···π stacking interactions depicted in Figure 7 (Cg6···Cg6) (Cg6 = C17–C16–C15–C14–C13–C12), and the intermolecular hydrogen bond interactions (C4–H4···O9 and C9–H9B···O13) depicted in Figure 8 give rise to the formation of a 2D supramolecular network, further linking into a 3D supramolecular structure by C–H···π, π···π, and hydrogen-bond interactions [57,58,59,60,61,62,63,64] (Figure 9).

3.5. Molar Conductance and Solubility and Mass Spectrometry Analysis of the Title Polymer

The [Ni(II)-Sm(III)] coordination polymer could be soluble in DMSO and DMF, slightly soluble in ethanol, methanol, THF, dichloromethane, trichloromethane, acetonitrile, and acetone, and insoluble in n-hexane, ethyl ether, and water. Molar conductance of the title polymer in DMF at 25 °C (1 × 10−3 mol·L−1) is 179.6 Ω−1·cm2·mol−1. The result of molar conductivity is inconsistent with the 1:2 electrolyte reported previously [50].
According to the data analysis of mass spectrometry (Figure S1), when m/z = 418.9292, which is almost equal to the relative molecular mass of [NiL], when m/z = 158.9381, it is the relative molecular mass of 4,4′-bipy, indicating that 4,4′-bipy exists in the solution in the free form, while samarium nitrate might exist in the solution in the form of [Sm(NO3)(DMF)3]2+ due to the peak of m/z = 430.9834.
Through conductivity tests and mass spectrometry, we determined that the [Ni(II)-Sm(III)] coordination polymer dissolved in DMF may comprise [Ni(L)], 4,4′-bipy, [Sm(NO3)(DMF)3]2+ cation, and NO3 anions. The observed conductance would then result from the presence of [Sm(NO3)(DMF)3]2+ cation and NO3 anions, according to the following equilibria taking place in solution:
[Ni(L)Sm(NO3)3(4,4′-bipy)]n ⇌ n[Ni(L)Sm(NO3)3(4,4′-bipy)]
[Ni(L)Sm(NO3)3(4,4′-bipy)] + DMF ⇌ [Ni(L)] + 4,4′-bipy + [Sm(NO3)3(DMF)]
[Sm(NO3)3(DMF)] + 2DMF ⇌ [Sm(NO3)(DMF)3]2+ + 2NO3

3.6. UV/Vis Absorption Spectra of H2L and the Title Polymer

The UV/Vis absorption spectra of H2L and the title polymer (in 10−5 m ethanol solution) are depicted in the Figure 10.
The characteristic peaks of H2L appear at approximately 317, 271, and 223 nm, respectively, the peaks at 271 and 223 nm could be ascribed to the π–π* transition of the benzene ring, and the peak at 317 nm could be attributed to the intraligand π–π* transition of the oxime-like group [20,47]. Compared to the peaks of H2L, the absorptions of the polymer bathochromically move to approximately 228 and 276 nm, exhibiting coordination of the (L)2− unit to the Ni(II) atom [15,50]. Upon coordination, the intra-ligand π–π* transition of the oxime-like group disappeared again, indicating the complexation [47,50]. A new characteristic peak at approximately 350 nm is found for the coordination polymer, belonging to an n-π* charge transfer of the imino group [43,50].
In order to further prove that the polymer in the form of a solution is not affected by metal ions, 4,4′-bipy, we conducted UV spectrum experiments on its solution (Ni(OAc)2·4H2O, Sm(NO3)3·6H2O, and 4,4′-bipy in 10−5 m EtOH). There was almost no change in the absorbance value of metal ions and no characteristic absorption peak. Further, 4,4′-bipy showed a characteristic absorption peak at 239 nm.

3.7. Fluorescence Properties of the Title Polymer

The fluorescence spectra of H2L and the [Ni(II)–Sm(III)] polymer in EtOH (1.0 × 10−5 m) are shown in Figure 11. The results show that H2L presents an intensive photoluminescence with maximum emission at approximately 395 nm at 311 nm excitation, which can be explained as the π–π* transition in H2L. In the title polymer, the emission peak appears at approximately 382 nm, and there is a slight blue shift compared to H2L [50], the fluorescence intensity of the title polymer appears to be quenched, and at the same time, it can be seen from the data of the mass spectrum that the polymer solution is almost in the form of [Ni(L)] + 4,4′-bipy + [Sm(NO3)(DMF)3], indicating that the coordinated Ni(II) ion has a heavy atom effect [62]. The high nuclear charge of the Ni(II) causes the electronic energy levels of the phosphorescent molecules to be staggered, which enhances the spin–orbit coupling of the phosphorescent molecules, thereby increasing the probability of inter-system hopping (ISC) of S1→Tl, thereby quenching the fluorescence. In the fluorescence spectra, only the band at approximately 375–650 nm instead of the f–f emission is expected for Sm(III) ions.

3.8. Antimicrobial Activity

Firstly, the Staphylococcus aureus on the plate was inoculated into Agar (2%) (LB) medium for overnight culture, and 0.1 mL of night-cultured fresh bacterial suspension was added to the LB medium after autoclaving and cooling to 50 °C. Secondly, after the mixture was mixed well, LB solid AGAR was poured into a sterile Petri dish. After being completely solidified, we punched holes in the LB medium with a perforator. Finally, DMF sample solutions were configurated with solution gradients of four various concentrations (0.35, 0.7, 1.4, and 2.8 mg mL−1).
Then, 70 µL of samples with different concentrations were added to the samples with a pipette gun. After an eight-hour incubation period at 37 °C, diameters of the inhibition zones were measured. The experimental results are compared to Ampicillin as a reference standard with various concentrations. The diameters of the inhibition zones of H2L, nickel acetate, and the title polymer are shown in Figure 12. The title polymer shows a more enhanced antibacterial effect than H2L and the metal salt solution under the same conditions. At the same time, the complex [NiL] also has a certain antibacterial activity, but it is not as high as that of the coordination polymer, which may be due to the antibacterial activity of the dissolved polymer being the sum of the comprehensive activities of various components. It can be conjectured from the experimental results and previous literature [54], first of all, that this is the result of the destruction of the protein structure in bacteria due to the action of heavy metal ions in coordination polymers [32]. Secondly, the ligand H2L destroys part of the cell membrane, making the bacteria unable to further divide and reproduce, leading to death. These results are similar to biological activities of related Schiff base complexes previously reported [65].

4. Conclusions

A new heterobimetallic Ni(II)-Sm(III) polymer has been prepared and structurally validated, in which H2L represents a symmetric salamo-based bisoxime ligand. In the Ni(II)-Sm(III) polymer, the hexacoordinated nickel(II) atom bears a slightly twisted octahedral geometry, and the decacoordinated Sm(III) atom possesses a twisted bicapped square antiprismatic arrangement. In the crystal, the polymer forms a self-assembling infinite 2D network further linking into a 3D supramolecular structure by C–H···π, π···π, and hydrogen-bond interactions. Compared to H2L, the Ni(II)-Sm(III) polymer exhibits only a very slightly hypsochromically shifted fluorescence of lower intensity, indicating that the coordinated Ni(II) cation has a minor heavy atom effect. Antimicrobial experiments have shown that the title polymer demonstrates stronger antimicrobial activity than H2L under the same conditions.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4352/10/7/579/s1, Figure S1: Mass spectrogram of Ni(II)-Sm(III) coordination polymer solution.

Author Contributions

W.-K.D. and Y.-P.Z. conceived and designed the experiments; X.-X.A. and R.-Y.L. performed the experiments; J.-F.W. and H.-R.M. analyzed the data; Y.Z. contributed reagents/materials/analysis tools; R.-Y.L., X.-X.A. and H.-R.M. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 21761018) and the Program for Excellent Team of Scientific Research in Lanzhou Jiaotong University (Grant No. 201706).

Acknowledgments

This work was supported by the National Natural Science Foundation of China (21761018), Science and Technology Program of Gansu Province (18YF1GA057), and the Program for Excellent Team of Scientific Research in Lanzhou Jiaotong University (201706), three of which are gratefully acknowledged.

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. 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]
  3. 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]
  4. 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]
  5. 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 A 2018, 190, 111–115. [Google Scholar] [CrossRef] [PubMed]
  6. 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]
  7. 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]
  8. 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]
  9. 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] [Green Version]
  10. Łḛpicka, 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]
  11. 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]
  12. 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] [Green Version]
  13. 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]
  14. 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] [CrossRef]
  15. 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]
  16. 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]
  17. Wei, Z.L.; Wang, L.; Wang, J.F.; Guo, W.T.; Zhang, Y.; Dong, W.K. Two highly sensitive and efficient salamo-like copper(II) complex probes for recognition of CN. Spectrochim. Acta A 2020, 228, 117775. [Google Scholar] [CrossRef]
  18. Wang, L.; Wei, Z.L.; Chen, Z.Z.; Liu, C.; Dong, W.K.; Ding, Y.J. A chemical probe capable for fluorescent and colorimetric detection to Cu2+ and CN based on coordination and nucleophilic addition mechanism. Microchem. J. 2020, 155, 104801. [Google Scholar] [CrossRef]
  19. Liu, C.; Wei, Z.L.; Mu, H.R.; Dong, W.K.; Ding, Y.J. A novel unsymmetric bis(salamo)-based chemosensor for detecting Cu2+ and continuous recognition of amino acids. J. Photochem. Photobio. A 2020, 397, 112569. [Google Scholar] [CrossRef]
  20. 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] [Green Version]
  21. Jaros, S.W.; Guedes da Silva, M.F.C.; Florek, M.; Oliveira, M.C.; Smoleński, P.; Pombeiro, A.J.L.; Kirillov, A.M. Aliphatic dicarboxylate directed assembly of Silver(I) 1,3,5-triaza-7-phosphaadamantane coordination networks: Topological versatility and antimicrobial activity. Cryst. Growth Des. 2014, 14, 5408–5417. [Google Scholar] [CrossRef]
  22. Horcajada, P.; Serre, C.; Vallet-Regí, M.; Sebban, M.; Taulelle, F.; Férey, G. Metal–organic frameworks as efficient materials for drug delivery. Angew. Chem. Int. Ed. 2006, 45, 5974–5978. [Google Scholar] [CrossRef] [PubMed]
  23. Jaros, W.S.; Guedes da Silva, M.F.C.; Król, J.M.; Oliveira, M.C.; Smoleński, P.; Pombeiro, A.L.; Kirillov, A.M. Bioactive silver–organic networks assembled from 1,3,5-triaza-7-phosphaadamantane and flexible cyclohexanecarboxylate blocks. Inorg. Chem. 2016, 55, 1486–1496. [Google Scholar] [CrossRef] [PubMed]
  24. Jaros, W.S.; Smoleński, P.; Guedes da Silva, M.F.C.; Florek, M.; Król, J.; Staroniewicz, Z.; Pombeiro, A.J.L.; Kirillov, A.M. New silver bioMOFs driven by 1,3,5-triaza-7-phosphaadamantane-7-sulfide (PTAS): Synthesis, topological analysis and antimicrobial activity. CrystEngComm 2013, 15, 8060–8064. [Google Scholar] [CrossRef]
  25. Lu, X.Y.; Ye, J.W.; Zhang, D.K.; Xie, R.X.; Bogale, R.F.; Sun, Y.; Zhao, L.M.; Zhao, Q.; Ning, G.L. Silver carboxylate metal–organic frameworks with highly antibacterial activity and biocompatibility. J. Inorg. Biochem. 2013, 138, 114–121. [Google Scholar] [CrossRef] [PubMed]
  26. Wyszogrodzka, G.; Marszałek, B.; Gil, B.; Dorożyński, P. Metal-organic frameworks: Mechanisms of antibacterial action and potential applications. Drug Discov. Today 2016, 21, 1009–1018. [Google Scholar] [CrossRef] [PubMed]
  27. Rodríguez, H.S.; Hinestroza, J.P.; Puentes, C.O.; Sierra, C.; Soto, C.Y. Antibacterial activity against Escherichia coli of Cu-BTC (MOF-199) metal-organic framework immobilized onto cellulosic fibers. J. Appl. Polym. Sci. 2014, 131, 40815. [Google Scholar]
  28. 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]
  29. 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. Naturforsch B 2018, 73, 203–210. [Google Scholar] [CrossRef]
  30. 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]
  31. 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] [PubMed] [Green Version]
  32. 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]
  33. 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, 117209. [Google Scholar] [CrossRef]
  34. 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]
  35. 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]
  36. Yu, M.; Zhang, Y.; Pan, Y.Q.; Wang, L. Two novel copper(II) salamo-based complexes: Syntheses, X-ray crystal structures, spectroscopic properties and Hirshfeld surfaces analyses. Inorg. Chim. Acta 2020, 509, 119701. [Google Scholar] [CrossRef]
  37. Zhang, S.Z.; Chang, J.; Zhang, H.J.; Sun, Y.X.; Wu, Y.; Wang, Y.B. Synthesis, crystal structure and spectral properties of binuclear Ni(II) and cubane-like Cu4(μ3-O)4 cored tetranuclear Cu(II) complexes based on coumarin Schiff base. Chin. J. Inorg. Chem. 2020, 36, 503–514. [Google Scholar]
  38. Chang, J.; Zhang, S.Z.; Wu, Y.; Zhang, H.J.; Sun, Y.X. Three supramolecular trinuclear nickel(II) complexes based on Salamo-type chelating ligand: Syntheses, crystal structures, solvent effect, Hirshfeld surface analysis and DFT calculation. Transit. Met. Chem. 2020, 45, 279–293. [Google Scholar] [CrossRef]
  39. 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]
  40. Li, J.; Zhang, H.J.; Chang, J.; 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] [Green Version]
  41. Chang, J.; Zhang, H.J.; Jia, H.R.; Sun, Y.X. Binuclear nickel(II) and zinc(II) complexes based on 2-amino-3-hydroxy-pyridine Schiff base: Syntheses, supramolecular structures and spectral properties. Chin. J. Inorg. Chem. 2018, 34, 2097–2107. [Google Scholar]
  42. 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]
  43. 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]
  44. Liu, L.Z.; Yu, M.; Li, X.Y.; Kang, Q.P.; Dong, W.K. Syntheses, structures, Hirshfeld analyses and fluorescent properties of two Ni(II) and Zn(II) complexes constructed from a bis(salamo)-like ligand. Chin. J. Inorg. Chem. 2019, 35, 1283–1294. [Google Scholar]
  45. 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]
  46. Peng, Y.D.; Wang, F.; Gao, L.; Dong, W.K. Structurally characterized dinuclear zinc(II) bis(salamo)-type tetraoxime complex possessing square pyramidal and trigonal bipyramidal geometries. J. Chin. Chem. Soc. 2018, 65, 893–899. [Google Scholar] [CrossRef]
  47. 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]
  48. 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]
  49. Yu, M.; Mu, H.R.; Liu, L.Z.; Li, N.; Bai, Y.; Dong, X.Y. Syntheses, structures and Hirshfeld analyses of trinuclear Ni(II) salamo-type complexes. Chin. J. Inorg. Chem. 2019, 35, 1109–1120. [Google Scholar]
  50. 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]
  51. Sun, Y.X.; Pan, Y.Q.; Xu, X.; Zhang, Y. Unprecedented dinuclear CuII N,O-donor complex: Synthesis, structural characterization, fluorescence property, and Hirshfeld analysis. Crystals 2019, 9, 607. [Google Scholar] [CrossRef] [Green Version]
  52. Zhang, Y.; Yu, M.; Pan, Y.Q.; Zhang, Y.; Xu, L.; Dong, W.K. Three rare heteromultinuclear 3d-4f salamo-like complexes constructed from auxiliary ligand 4,4′-bipy: Syntheses, structural characterizations, fluorescence and antimicrobial properties. Appl. Organomet. Chem. 2020, 34, e5442. [Google Scholar] [CrossRef]
  53. Pan, Y.Q.; Zhang, Y.; Yu, M.; Zhang, Y.; Wang, L. Newly synthesized homomultinuclear Co(II) and Cu(II) bissalamo-like complexes: Structural characterizations, Hirshfeld analyses, fluorescence and antibacterial properties. Appl. Organomet. Chem. 2020, 34, e5441. [Google Scholar] [CrossRef]
  54. Liu, C.; An, X.X.; Cui, Y.F.; Xie, K.F.; Dong, W.K. Novel structurally characterized hetero-bimetallic [Zn(II)2M(II)] (M = Ca and Sr) bis(salamo)-type complexes: DFT calculation, Hirshfeld analyses, antimicrobial and fluorescent properties. Appl. Organomet. Chem. 2020, 34, e5272. [Google Scholar] [CrossRef]
  55. Pan, Y.Q.; Xu, X.; Zhang, Y.; Zhang, Y.; Dong, W.K. A highly sensitive and selective bis(salamo)-type fluorescent chemosensor for identification of Cu2+ and the continuous recognition of S2−, Arginine and Lysine. Spectrochim. Acta A 2020, 229, 117927. [Google Scholar] [CrossRef]
  56. 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]
  57. 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]
  58. 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]
  59. Zhou, L.; Hu, Q.; Chai, L.Q.; Mao, K.H.; Zhang, H.S. X-ray characterization, spectroscopic, DFT calculations and Hirshfeld surface analysis of two 3-D supramolecular mononuclear zinc(II) and trinuclear copper(II) complexes. Polyhedron 2019, 158, 102–116. [Google Scholar] [CrossRef]
  60. Wang, F.; Liu, L.Z.; Gao, L.; Dong, W.K. Unusual constructions of two Salamo-based copper(II) complexes. Spectrochim. Acta A 2018, 203, 56–64. [Google Scholar] [CrossRef]
  61. 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]
  62. Zhang, L.W.; Zhang, Y.; Cui, Y.F.; Yu, M.; Dong, W.K. Heterobimetallic [NiIILnIII] (Ln = Sm and Tb) N2O4-donor coordination polymers: Syntheses, crystal structures and fluorescence properties. Inorg. Chim. Acta 2020, 506, 119534. [Google Scholar] [CrossRef]
  63. Zhang, H.J.; Chang, J.; Jia, H.R.; Sun, Y.X. Syntheses, supramolecular structures and spectroscopic properties of Cu(II) and Ni(II) complexes with Schiff base containing oxime group. Chin. J. Inorg. Chem. 2018, 34, 2261–2270. [Google Scholar]
  64. An, X.X.; Chen, Z.Z.; Mu, H.R.; Zhao, L. Investigating into crystal structures and supramolecular architectures of four newly synthesized hetero-octanuclear [CuII4-LnIII4] (Ln = Sm, Eu, Tb and Dy) complexes produced by a hexadentate bisoxime chelate ligand. Inorg. Chim. Acta 2020, 511, 119823. [Google Scholar] [CrossRef]
  65. 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]
Scheme 1. Synthetic route to H2L.
Scheme 1. Synthetic route to H2L.
Crystals 10 00579 sch001
Scheme 2. Synthetic route to the title polymer.
Scheme 2. Synthetic route to the title polymer.
Crystals 10 00579 sch002
Figure 1. Comparing the simulated PXRD and experimental patterns of the [Ni(II)-Sm(III)] coordination polymer.
Figure 1. Comparing the simulated PXRD and experimental patterns of the [Ni(II)-Sm(III)] coordination polymer.
Crystals 10 00579 g001
Figure 2. Infrared spectra of H2L and its Ni(II)-Sm(III) coordination polymer.
Figure 2. Infrared spectra of H2L and its Ni(II)-Sm(III) coordination polymer.
Crystals 10 00579 g002
Figure 3. (a) Molecular structure and atom numberings of the title polymer. (b) Coordination polyhedra for Sm(III) and Ni(II) atoms in the title polymer.
Figure 3. (a) Molecular structure and atom numberings of the title polymer. (b) Coordination polyhedra for Sm(III) and Ni(II) atoms in the title polymer.
Crystals 10 00579 g003
Figure 4. View of the intramolecular hydrogen bond interactions of the Ni(II)-Sm(III) polymer (hydrogen atoms are omitted for clarity, except those forming hydrogen bonding).
Figure 4. View of the intramolecular hydrogen bond interactions of the Ni(II)-Sm(III) polymer (hydrogen atoms are omitted for clarity, except those forming hydrogen bonding).
Crystals 10 00579 g004
Figure 5. View of the intramolecular C–H···π interactions.
Figure 5. View of the intramolecular C–H···π interactions.
Crystals 10 00579 g005
Figure 6. View of one-dimensional chain-like structure of the title polymer.
Figure 6. View of one-dimensional chain-like structure of the title polymer.
Crystals 10 00579 g006
Figure 7. View of the π···π stacking interactions.
Figure 7. View of the π···π stacking interactions.
Crystals 10 00579 g007
Figure 8. View of two-dimensional supramolecular network of the title polymer, combined by intermolecular hydrogen bondings.
Figure 8. View of two-dimensional supramolecular network of the title polymer, combined by intermolecular hydrogen bondings.
Crystals 10 00579 g008
Figure 9. View of the three-dimensional supramolecular structure of the title polymer.
Figure 9. View of the three-dimensional supramolecular structure of the title polymer.
Crystals 10 00579 g009
Figure 10. The UV/Vis absorption spectra of H2L, its Ni(II)-Sm(III) coordination polymer, Ni(OAc)2·4H2O, Sm(NO3)3·6H2O, and 4,4′-bipy.
Figure 10. The UV/Vis absorption spectra of H2L, its Ni(II)-Sm(III) coordination polymer, Ni(OAc)2·4H2O, Sm(NO3)3·6H2O, and 4,4′-bipy.
Crystals 10 00579 g010
Figure 11. Photophysical properties of H2L and its corresponding Ni(II)-Sm(III) polymer in ethanol (1 × 10−5 mol L−1) at 311 nm excitation.
Figure 11. Photophysical properties of H2L and its corresponding Ni(II)-Sm(III) polymer in ethanol (1 × 10−5 mol L−1) at 311 nm excitation.
Crystals 10 00579 g011
Figure 12. Inhibition of the Staphylococcus aureus at different concentrations.
Figure 12. Inhibition of the Staphylococcus aureus at different concentrations.
Crystals 10 00579 g012
Table 1. Crystal data and structure refinement for the title polymer.
Table 1. Crystal data and structure refinement for the title polymer.
PolymerThe [Ni(II)-Sm(III)] Coordination Polymer
Empirical formulaC28H26N7NiO15Sm
Formula weight909.62
T, K173(2)
Radiation; wavelength (Å)Mo Kα; 0.71073
Crystal systemmonoclinic
Space groupP21/c
a, Å11.3928(4)
b, Å19.4539(7)
c, Å19.4229(8)
β, deg98.689(2)
Volume, Å34255.4(3)
Z4
Dcalcd, g cm−31.42
Crystal size, mm30.18 × 0.20 × 0.22
Absorption coefficient, mm−11.9
range data collection, deg2.400 to 26.000
F(000), e1812
hklrange±14, ±24, ±24
Refl. collected/unique/Rint60159/9380/0.0196
Data/restraints/ref. parameters9380/11/471
Final R1/wR2 [I > 2 σ(I)] a,b0.0332/0.1039
Final R1 wR2(all data) a,b0.0405/0.1063
GoF c0.989
Δρmax/min,e Å−31.02/–1.01
a R1 = Σ||Fo| − |Fc||/Σ|Fo|; b wR2 = [ÿ w(Fo2 − Fc2)2/ÿ w(Fo2)2]1/2, w = [σ2(Fo2) + (AP)2 + BP]−1, where P = (Max(Fo2, 0) + 2Fc2)/3; c GoF = S = [ÿ w(Fo2 − Fc2)2/(nobs − nparam)]1/2.
Table 2. Selected bond lengths (Å) and angles (deg) for the title polymer.
Table 2. Selected bond lengths (Å) and angles (deg) for the title polymer.
BondLengthsBondLengthsBondLengths
Sm1-O12.653(4)Sm1-O102.617(4)Ni1-N12.072(4)
Sm1-O22.353(2)Sm1-O112.518(4)Ni1-N22.030(4)
Sm1-O52.416(2)Sm1-O132.510(3)Ni1-N32.172(2)
Sm1-O62.6013(19)Sm1-O142.481(2)Ni1-N4 #12.167(2)
Sm1-O72.529(3)Ni1-O22.0370(19)
Sm1-O82.463(4)Ni1-O52.0399(19)
BondAnglesBondAnglesBondAngles
O1-Sm1-O261.11(7)O5-Sm1-O10173.74(9)O10-Sm1-O1363.93(10)
O1-Sm1-O5117.08(7)O5-Sm1-O11124.32(10)O10-Sm1-O14102.11(9)
O1-Sm1-O6143.61(9)O5-Sm1-O13114.44(8)O11-Sm1-O1373.92(12)
O1-Sm1-O771.32(11)O5-Sm1-O1473.25(7)O11-Sm1-O1473.63(10)
O1-Sm1-O878.12(11)O6-Sm1-O773.41(10)O13-Sm1-O1451.02(7)
O1-Sm1-O1064.15(10)O6-Sm1-O873.12(10)O2-Ni1-O578.33(8)
O1-Sm1-O1172.52(11)O6-Sm1-O10121.02(9)O2-Ni1-N190.31(12)
O1-Sm1-O13128.07(9)O6-Sm1-O11140.94(10)O2-Ni1-N2167.66(12)
O1-Sm1-O14143.78(9)O6-Sm1-O1369.80(10)O2-Ni1-N390.61(12)
O2-Sm1-O565.35(7)O6-Sm1-O1472.59(7)O2-Ni1-N4 #189.60(12)
O2-Sm1-O6126.96(7)O7-Sm1-O851.43(10)O5-Ni1-N1168.08(12)
O2-Sm1-O788.84(9)O7-Sm1-O10108.73(9)O5-Ni1-N290.42(12)
O2-Sm1-O8131.26(10)O7-Sm1-O11143.58(12)O5-Ni1-N387.88(12)
O2-Sm1-O10111.99(10)O7-Sm1-O13127.65(10)O5-Ni1-N4 #195.18(12)
O2-Sm1-O1177.41(10)O7-Sm1-O14142.47(9)N1-Ni1-N2101.20(15)
O2-Sm1-O13143.33(9)O8-Sm1-O1066.71(12)N1-Ni1-N388.64(14)
O2-Sm1-O1499.12(7)O8-Sm1-O11116.21(12)N1-Ni1-N4 #188.27(14)
O5-Sm1-O662.10(6)O8-Sm1-O1382.62(10)O10-Sm1-O1394.07(14)
O5-Sm1-O777.17(7)O8-Sm1-O14129.42(10)O10-Sm1-O1486.35(14)
O5-Sm1-O8119.44(10)O10-Sm1-O1149.60(12)O11-Sm1-O13176.91(14)
Symmetry transformations used to generate equivalent atoms: #1 1 + x, y, z.
Table 3. Hydrogen bonding interactions (Å, °) and C-H···π stacking interactions for the Ni(II)-Sm(III) polymer.
Table 3. Hydrogen bonding interactions (Å, °) and C-H···π stacking interactions for the Ni(II)-Sm(III) polymer.
D–H···A.d(D-H)d(H···A)d(D···A)∠D–H···ASymmetry Code
C1-H1B···O100.982.392.897(7)111
C4-H4···O90.952.553.365(5)143x, 3/2 − y, 1/2 + z
C9-H9A···N30.992.443.357(5)154
C9-H9B···O130.992.403.230(5)1411 − x, −1/2 + y, 3/2 − z
C23-H23···O20.952.583.093(4)114
C24-H24···N10.952.613.104(5)113−1 + x, y, z
C28-H28···O70.952.253.120(5)151−1 + x, y, z
C19-H19···Cg20.982.562.897(7)114x, y, z
Cg6···Cg6 4.065(2)
Cg6 = C12–C13–C14–C15–C16–C.

Share and Cite

MDPI and ACS Style

Li, R.-Y.; An, X.-X.; Wang, J.-F.; Mu, H.-R.; Zhang, Y.-P.; Zhang, Y.; Dong, W.-K. A Self-Assembled Hetero-Bimetallic [Ni(II)-Sm(III)] Coordination Polymer Constructed from a Salamo-Like Ligand and 4,4′-Bipyridine: Synthesis, Structural Characterization, and Properties. Crystals 2020, 10, 579. https://doi.org/10.3390/cryst10070579

AMA Style

Li R-Y, An X-X, Wang J-F, Mu H-R, Zhang Y-P, Zhang Y, Dong W-K. A Self-Assembled Hetero-Bimetallic [Ni(II)-Sm(III)] Coordination Polymer Constructed from a Salamo-Like Ligand and 4,4′-Bipyridine: Synthesis, Structural Characterization, and Properties. Crystals. 2020; 10(7):579. https://doi.org/10.3390/cryst10070579

Chicago/Turabian Style

Li, Ruo-Yan, Xiao-Xin An, Ji-Fa Wang, Hao-Ran Mu, You-Peng Zhang, Yang Zhang, and Wen-Kui Dong. 2020. "A Self-Assembled Hetero-Bimetallic [Ni(II)-Sm(III)] Coordination Polymer Constructed from a Salamo-Like Ligand and 4,4′-Bipyridine: Synthesis, Structural Characterization, and Properties" Crystals 10, no. 7: 579. https://doi.org/10.3390/cryst10070579

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