Next Article in Journal
Ultrafast Laser-Induced Crystallization of Lead Germanate Glass
Next Article in Special Issue
A Novel Fishbone-Like Lead(II) Supramolecular Polymer: Synthesis, Characterization, and Application for Producing Nano Metal Oxide
Previous Article in Journal
High-Durability Concrete with Supplementary Cementitious Admixtures Used in Corrosive Environments
Previous Article in Special Issue
IsoStar Program Suite for Studies of Noncovalent Interactions in Crystals of Chemical Compounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ultrasound-Assisted Synthesis and Crystal Structure of Novel 2D Cd (II) Metal–Organic Coordination Polymer with Nitrite End Stop Ligand as a Precursor for Preparation of CdO Nanoparticles

by
Younes Hanifehpour
1,*,
Jaber Dadashi
2 and
Babak Mirtamizdoust
2
1
Department of Chemistry, Sayyed Jamaleddin Asadabadi University, Asadabad 6541861841, Iran
2
Department of Chemistry, Faculty of Science, University of Qom, P.O. Box 37185-359, Qom, Iran
*
Author to whom correspondence should be addressed.
Crystals 2021, 11(2), 197; https://doi.org/10.3390/cryst11020197
Submission received: 23 December 2020 / Revised: 4 February 2021 / Accepted: 12 February 2021 / Published: 17 February 2021

Abstract

:
In the present research, a sonochemical approach was applied to prepare new cadmium(II) coordination 2D polymer, [Cd(L)(NO2)2]n (L = 1,2-bis(1-(pyridin-3-yl)ethylidene)hydrazine) and structurally characterized with various spectroscopic techniques including XRD, elemental analysis, SEM, and IR spectroscopy. The coordination number of cadmium (II) ions is seven (CdN2O5) by two nitrogen atoms from two organic Schiff base ligand and five oxygen of nitrite anions. The 2D sheet structures ended by nitrite anions and the nitrite anion displayed the end-stop role. The comprehensive system showed a three-dimensional structure with several weak interactions. The high-intensity ultrasound is regarded as an easy, environmentally-friendly, and flexible synthetic instrument for the compounds of coordination. CdO NPs was obtained by thermolysing 1 at 180 °C with oleic acid (as a surfactant). Further, the size and morphology of the produced CdO nanoparticles were investigated through SEM.

1. Introduction

Recently, the development of metal–organic coordination systems has become increasingly attractive for researchers [1,2,3]. Ascetic architecture, luminescence, topologic diversity, magnetic, storage molecular, adsorption, covering, various preparation manners, and other interesting possible usage are among the factors attracting attention and results in various examinations leading to many research works in the area of coordination polymers and novel group of material chemistry [4,5,6].
Schiff bases that are a significant series of organic ligands have the capability of coordination by metal ions for forming stable chelates through the N atom of the azo-group as well as other donors, like –NH2 and –OH. Schiff bases have advantages of satisfactory biological activity, simple synthetic pathways, high coordination potency, and photochemical features, and they include pharmaceuticals (e.g., antibiotics, anti-inflammatory drugs, and antivirals) and also photoelectric and dye materials [7,8,9,10,11,12]. Moreover, because of the variety of structure, potential features, and preparative accessibility, metal complexes with Schiff base ligands are considered a significant stereochemical model in metal coordination chemistry. The common coordination mode of these ligands is different and very interesting [13]. Because of the relatively simple yet robust synthetic procedure, Schiff base formation and complexation have formed the basis of a number of elegant reported undergraduate experiments that explore various aspects of Schiff base chemistry, from the formation of organometallic complexes to combinatorial synthesis to spectral analysis. In this experiment, Schiff base ligands are used to demonstrate the effect of metal ions on metal–ligand stoichiometry [14].
Metal–organic coordination systems can be used as templates in order to obtain the desired nano-materials. Via selecting proper metal–organic compounds precursors with special morphologies and under suitable experimental conditions, acquiring the desired morphologies will become possible. Nano metal–organic coordination compounds precursors may also be used to obtain nano-materials with better and more useful properties [15,16,17,18]. Depending on the conditions, metal–organic coordination compounds precursor may lead to the preparation of metals, metal oxides, metal sulfides, or other nano-sized materials. With this method, nano-materials can be obtained in different morphologies from different metal–organic coordination systems precursors.
A small number of the transition metals’ characteristic attributes are demonstrated by Cadmium and group 12 elements because they do not partially fill d or f electron shells. Cadmium, like zinc, has a preference for a + 2 oxidation state in the majority of its complexes. However, the more important point is that these elements with spherical d10 configuration are the facilitators of different coordination geometries, allowing forming of 1D, 2D, and 3D coordination compounds [19,20,21].
The focus of the present study is the preparation and design of metal–organic coordination systems, and the synthesis of some nano metal–organic coordination polymers have been recently reported by our research team [22,23,24,25,26]. The next phase of our team in this regard is an investigation of the behaviors of Cadmium (II) in terms of coordination by 2-bis(1-(pyridin-3-yl)ethylidene)hydrazine) Schiff-base ligand in the present work.
The near similar works on these compounds were reported in previous literature with various composite and applicant (see [27]). Especially, the ligand’s structural chemistry is attractive due to its multifunctional coordination state which four aromatic and aliphatic nitrogen can be coordinated to metals (Scheme 1). A simple sonochemical preparation process is described in this work for a nanostructure for this coordination system and its usage in the construction of Cadmium oxide (CdO) NPs.

2. Experimental

2.1. Instruments and Materials

All chemical compounds were bought from the Aldrich and Merck (Germany and Merck Company, Darmstadt, Germany) chemical companies. A Bruker Vector 22 FT-IR (Richmond Scientific, Chorley, UK) spectrometer with KBr disks in the area of 400–4000 cm−1 was used for acquiring IR spectroscopy. An Electrothermal 9000 (Keison Products, Chelmsford, UK) apparatus determine the melting points. A multiwave ultrasonic generator (Sonicator_3000; Misonix, Inc., Farmingdale, NY, USA) was applied. X’pert diffractometer (produced by Panalytical, Panalytical, Malvern, UK) monochromatized Cu Kα radiation was employed for performing measurements of PXRD. Mercury 2.4 was used for preparing simulated PXRD powder patterns on the basis of single-crystal data [28]. Elemental analyses were carried out using a linked ISIS-300, Oxford EDS (Oxford Instruments, Abingdon, UK) (energy dispersion spectroscopy) detector. Chemical states and chemical compositions of the CdO nanoparticles were carried out through X-ray photoelectron spectroscopy (XPS) (Model: K-ALPHA, Thermo Fisher, Waltham, MA, USA).
The Scherrer formula was employed for estimating the crystallite size of chosen samples. Followed by gold coating Au, the samples’ morphology was assessed using an SEM (Hitachi, Japan). Using an XcaliburTM diffractometer with a Sapphire2 CCD detector and Mo Kα radiation (monochromator Enhance, Oxford Diffraction Ltd., Abingdon, UK) and ω-scan rotation methods at 150 K, diffraction data were collected for the single crystal. The CrysAlis program package (Oxford Diffraction Ltd., UK) was utilized for data reduction and gathering and refinement of cell parameters [29].
Multi-scan absorption modification with the CrysAlis (Rigaku, Tokyo, Japan) was used for the [Cd(L)(NO2)2]n data. With a full matrix weighted least-squares method (SHELX-2014) (Free Software Foundation, Boston, MA, US) with the w = 1/[σ2(Fo)2 + (0.035P)2] weight, where P = (Fo2 + 2Fc2)/3, we solved the structure with direct processes by SHELXS and adjusted on all F2 data anisotropically. Mercury 2.4 was applied for creating molecular exhibits. Crystal data and arrangement modification for 1 are provided in Table 1, and the designated bond lengths and angles can be observed in Table 2.

2.2. Preparation of 1,2-Bis(1-(pyridin-3-yl)ethylidene)hydrazine) Ligand

20 mmol (1.9 mL, 2.42 g) of 1-(pyridin-3-yl)ethan-1-one was dissolved in 50 mL of ethanol, then 10 mmol (0.9 mL, 0.91 g) of hydrazine hydrate (35 wt % solution in water), softened in 20 mL of ethanol, was mixed in a drop-wise manner; subsequently, two acetic acid drops were mixed with it too. The stirring of the reaction mix was done for 24 h at ambient temperature. Finally, the yellow sediments were filtered and air-dried (yield: 77%, 2.56 g product) (Comparable with previous synthesized [30,31]).
FT-IR (KBr, cm−1): 758(m), 823(w), 563(m), 1290(s), 1365(s), 1415(s), 1485(s), 1558(s), 1600(s).
Elemental analysis calculated for C14H14N4: C: 70.57, H: 5.92, N: 23.51%.
Experimental: C: 70.10, H: 6.00, N: 24.00%.

2.3. Synthesis of Nano-Structure of [Cd(L)(NO2)2]n

For synthesizing, the nanostructure of [Cd(L)(NO2)2]n a solution of cadmium(II) nitrate tetrahydrate (15 mL of a 0.1 M) in water was put in a high-density ultrasonic probe operating at 20 kHz by the highest power output of 60 W. 15 mL of a 0.1 M solution of the Schiff base ligand and 30 mL of a 0.1 M solution of sodium nitrite were added into this solution in a drop-wise manner. The resulting yellow precipitates were filtered, rinsed with H2O and MeOH, and dried at 50 °C (0.69 g product/yield: 65% decompose point: 210 °C).
Elemental analysis; calculated for C14H14CdN6O4: C: 37.98, H: 3.19, N: 18.98%.
Experimental: C: 38.00, H: 4.10, N: 20.00

2.4. Synthesis of Isolate Single Crystal of [Cd(L)(NO2)2]n

For isolating [Cd(L)(NO2)2]n single crystals, Schiff base ligand (1 mmol/0.24 g), sodium nitrite (2 mmol/0.14 g) and cadmium(II) nitrate tetrahydrate (1 mmol/0.31 g) were put in the primary arm of a branched tube for heating (Scheme 2) [32].
Both arms received methanol carefully. The tube was sealed, and the arm containing ligand was submerged in an oil bath at 60 °C; however, the other arm was stored at ambient temperature. Following two weeks, yellow crystals were precipitated in the colder arm for X-ray structure determination that was filtered, rinsed by acetone and ether, and being dried with exposure to air. (0.59 g product/Yield 89%—decompose point: 210 °C) Elemental analysis; calculated for C14H14CdN6O4: C: 37.98, H: 3.19, N: 18.98%. Experimental: C: 38.00, H: 4.00, N: 20.00

3. Synthesis of CdO Nanoparticles

The precursor of [Cd(L)(NO2)2]n (1) (0.02 mmol) was dissolved instantly in 1.0 mL of oleic acid, which formed a light yellow solution. This solution was degassed for 20 min and then heated to 180 °C for 2 h. At the end of the reaction, an orange precipitate was formed. A small amount of toluene and a large excess of EtOH were added to the reaction solution, and cadmium (II) oxide nanoparticles were separated by centrifugation. The light brown solids were washed with EtOH and dried under an air atmosphere (0.01 g product/yield: 58%).

4. Results and Discussion

The reaction of 1,2-bis(1-(pyridin-3-yl)ethylidene)hydrazine) Schiff base ligand with a mixture of sodium nitrite and cadmium nitrate, the novel Cd (II) metal–organic coordination polymer [Cd(L)(NO2)2]n was obtained. Using ultrasonic irradiation in a methanolic solution, the compound’s nanostructure was achieved; then, the single crystalline substance was attained through application of the heat gradients to the reagents solution or branched tube route. Scheme 3 presents the two various routes utilized for preparing [Cd(L)(NO2)2]n materials.
The surface morphology of the synthesized nano coordination compound was investigated through FE-SEM (field emission scanning microscopy) analysis. The morphology of the compound prepared by the sonochemical method is block-like aggregates (Figure 2). SEM assessments illustrate that achieved nanoparticles are of various sizes and have different shapes. Figure 3 showed the Simulated XPRD pattern of [Cd(L)(NO2)2]n in comparison with nanostructure XRD pattern, Acceptable matches with slight differences in were observed between the simulated and experimental XPRD patterns (Figure 3). This indicates that the synthesis obtained by the sonochemical method as nanoparticles is identical to that obtained by crystallography. The significant broadening of the peaks indicates that the particles are of nanometer dimensions. As it has been estimated by the Sherrer formula (D = 0.891λ/βcos·Ө, where D is the average grain size, λ the X-ray wavelength (0.15405 nm), and Ө and β the diffraction angle and full-width at half maximum of an observed peak, respectively) [33], the obtained value is D = 83 nm.
The determination of the structure of coordination polymer [Cd(L)(NO2)2]n displayed that the complex crystallizes in the Orthorhombic system with space group P212121, taking the form of a two-dimensional polymer in the solid-state (Figure 5). The Cd (II) atom is coordinated with two nitrogen atoms of two Schiff base ligands with the Cd–N distances of 2.356 (2), 2.358 (2) Å, five oxygen atoms of three nitrite anions. In addition distance of Cd–O bonds in this complex are around 2.368–2.247 Å. Therefore, the coordination number of the Cd (II) atom is seven, with an orthorhombic system with space group P212121 (Figure 4 and Figure 5) (See Appendix A for further details). The 2D sheet structures ended with nitrite anions. In this structure, the nitrite anion showed the end-stop role and by connecting directly to the metal, does not allow the system to expand to the third dimension.
The structure of other compounds reported from this ligand to Mn, Ag, Co, Zn, Bi, and Hg metals with varying coordination numbers for metals and the same coordination method for ligands is comparable [34,35,36,37,38].
There are many weak interactions exist in the structure; (like short contact involving C-H groups and aromatic interactions) (See Table 3. For example, there are pi–pi stacking interactions between the N1C5C4C3C3C1 plans about 3.770 Å and between the N4C12C13C14C10C11 plans about 3.474 Å. With several weak interactions, the 2D chains interact with neighbors, and the system extended to 3-dimensional structures (Figure 6 and Table 3).
The decomposition of precursor [Cd(L)(NO2)2]n in oleic acid functioning as a surfactant under an air atmosphere at 180 °C yields nano-powder of cadmium (II) oxide. The PXRD pattern of as-synthesized CdO nanopowder is seen in Figure 7. All the diffraction peaks are matched well with the cubic crystal structure of pure CdO with respect to their positions as found in JCPDS card No.05-0640.
Figure 8 displays the energy-dispersive X-ray spectroscopy (EDS) of the CdO nanoparticles and confirms the existence of the elements Cd/O and the formation of CdO as expected. The atomic percentage composition of the elements Cd and O were observed to be equal to 50.79 and 49.21, respectively.
Figure 9a displays the SEM image of as-prepared CdO nanoparticles. The obtained CdO has the regular nanoparticle shape with diameters of around 30–80 nm according to the size distribution graph of the product which was determined using Nahamin microstructure distance measurement software (Iran). (Figure 9b).
The XPS spectrum (Figure 10) confirms the presence of O and Cd in the sample. The binding energy corresponding to the peaks O1s, Cd 3d5/2, and 3d3/2 obtained by XPS analysis is 530.03, 405.6, and 412.9 eV, respectively [39].
The thermal gravimetric (TG) analysis was applied to evaluate the thermal stability of the [Cd(L)(NO2)2]n 1. TG was recorded in the temperature range of 20 and 800 °C. The diagram of 1 depicts that the compound remains stable up to 125 °C, and then decomposes up to 210 °C. The first weight loss is related to the removal of the NO2 unit. The step 2 in the range 300 °C to 380 °C with sharp weight loss is attributed to an organic moiety of 1 with a mass loss of about 45.3%. The remained solid around 800 °C is probably CdO (Figure 11).

5. Conclusions

In the present study, a novel cadmium (II) coordination polymer [Cd(L)(NO2)2]n nanostructure was prepared in an ultrasound-assisted manner and also compared by its crystal structure. The structural characterization of new coordination compounds was carried out by elemental analysis, IR spectroscopy, SEM, and single-crystal crystallography. The complex’s structure creates a 2D coordination polymer in solid-state. It was determined that the coordination number of Cd (II) ions is 7 with five oxygen of nitrite anions and two nitrogen atoms from two Schiff base ligand. With several weak interactions, the comprehensive system displayed a three-dimensional structure. Thermolysis of coordination polymer 1 results in CdO nanoparticles.

Author Contributions

Y.H.: supervision, writing—original draft; B.M.: investigation, writing—review and editing; J.D.: writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by grant number 32/1374 of the Sayyed Jamaleddin Asadabadi University.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to express their gratitude to University of Qom for supporting this research.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Crystallographic data for the structure reported in this paper have been deposited with the Cambridge Crystallographic Data Centre as supplementary publication CCDC-2045446 for [Cd(L)(NO2)2]n (1). Copies of the data can be obtained by request from the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK (Fax: +44­1223­336033, e­Mail: de­[email protected]).

References

  1. Emhoff, K.A.; Balaraman, L.; Salem, A.M.H.; Mudarmah, K.I.; Boyd, W.C. Coordination chemistry of organic nitric oxide derivatives. Coord. Chem. Rev. 2019, 396, 124–140. [Google Scholar] [CrossRef]
  2. Liu, X.; Hamon, J.R. Recent developments in penta-, hexa- and heptadentate Schiff base ligands and their metal complexes. Coord. Chem. Rev. 2019, 389, 94–118. [Google Scholar] [CrossRef]
  3. Mirtamizdoust, B. Sonochemical synthesis of nano lead(II) metal-organic coordination polymer; New precursor for the preparation of nano-materials. Ultrason Sonochem. 2017, 35, 263–269. [Google Scholar] [CrossRef]
  4. Rasheed, T.; Hassan, A.A.; Bilal, M.; Hussain, T.; Rizwan, K. Metal-organic frameworks based adsorbents: A review from removal perspective of various environmental contaminants from wastewater. Chemosphere. 2020, 259, 127369. [Google Scholar] [CrossRef]
  5. Engel, E.R.; Scott, J.L. Advances in the green chemistry of coordination polymer materials. Green Chem. 2020, 22, 3693–3715. [Google Scholar] [CrossRef]
  6. Mishra, A.; Jo, J.H.; Kim, H.; Woo, S.; Chi, K.W. A Discrete Cobalt Complex Obtained from a 1D Coordination Polymer for Highly Selective Sensing of the Mercury (II) Ion. ChemPlusChem 2014, 79, 925–928. [Google Scholar] [CrossRef]
  7. Guo, Y.; Hu, X.; Zhang, X.; Pu, X.; Wang, Y. The synthesis of a Cu(II) Schiff base complex using a bidentate N2O2 donor ligand: Crystal structure, photophysical properties, and antibacterial activities. RSC Adv. 2019, 9, 41737. [Google Scholar] [CrossRef] [Green Version]
  8. Chen, S.; Liu, X.; Ge, X.; Wang, Q.; Xie, Y.; Hao, Y.; Zhang, Y.; Zhang, L.; Shang, W.; Liu, Z. Lysosome-targeted iridium(III) compounds with pyridine-triphenylamine Schiff base ligands: Syntheses, antitumor applications and mechanisms. Inorg. Chem. Front. 2020, 7, 91–100. [Google Scholar] [CrossRef]
  9. Che, C.M.; Huang, J.S. Metal complexes of chiral binaphthyl Schiff-base ligands and their application in stereoselective organic transformations. Coord. Chem. Rev. 2003, 242, 97–113. [Google Scholar] [CrossRef]
  10. Himed, Y.; Komatsuzaki, N.O.; Sugihara, H.; Arakawa, H.; Kasuga, K. Transfer hydrogenation of a variety of ketones catalyzed by rhodium complexes in aqueous solution and their application to asymmetric reduction using chiral Schiff base ligands. J. Mol. Catal A Chem. 2003, 195, 95–100. [Google Scholar] [CrossRef]
  11. Miroslaw, B. Homo-and Hetero-Oligonuclear Complexes of Platinum Group Metals (PGM) Coordinated by Imine Schiff Base Ligands. Int. J. Mol. Sci. 2020, 21, 3493. [Google Scholar] [CrossRef]
  12. Vigato, P.A.; Tamburini, S. The challenge of cyclic and acyclic Schiff bases and related derivatives. Coord. Chem. Rev. 2004, 248, 1717–2128. [Google Scholar] [CrossRef]
  13. Leovac, V.M.; Jevtovic, V.S.; Jovanovic, L.S.; Bogdanovic, G.A. Metal complexes with Schiff-base ligands—Pyridoxal and semicarbazide-based derivatives. J. Serb. Chem. Soc. 2005, 70, 393–422. [Google Scholar] [CrossRef]
  14. Wilkinson, S.M.; Sheedy, T.M.; New, E.J. Synthesis and Characterization of Metal Complexes with Schiff Base Ligands. J. Chem. Educ. 2016, 93, 351−354. [Google Scholar] [CrossRef]
  15. Masoomi, M.Y.; Morsali, A. Applications of metal–organic coordination polymers as precursors for preparation of nano-materials. Coord. Chem. Rev. 2012, 256, 2921–2943. [Google Scholar] [CrossRef]
  16. Malik, M.A.; O’Brien, P. Organometallic and Metallo-Organic Precursors for Nanoparticles. In Precursor Chemistry of Advanced Materials; Springer: Berlin/Heidelberg, Germany, 2005; p. 173. [Google Scholar]
  17. Leong, W.L.; Vittal, J.J. One-Dimensional Coordination Polymers: Complexity and Diversity in Structures, Properties, and Applications. Chem. Rev. 2010, 111, 688–764. [Google Scholar] [CrossRef]
  18. Vittal, J.J.; Ng, M.T. Chemistry of Metal Thio- and Selenocarboxylates:  Precursors for Metal Sulfide/Selenide Materials, Thin Films, and Nanocrystals. Acc. Chem. Res. 2006, 39, 869–877. [Google Scholar] [CrossRef] [PubMed]
  19. Das, K.; Datta, A.; Massera, C.; Sinha, C. Structural diversity, topology and luminescent properties of a two-dimensional Cd(II) coordination polymer incorporating 4,4′-dipyridyl and 4,4′-sulfonyldibenzoic. J. Mol. Struct. 2019, 1179, 618–622. [Google Scholar] [CrossRef]
  20. Yadav, C.L.; Anamika; Rajput, G.; Kumar, K.; Drew, M.G.B.; Singh, N. Effect of Substituents on the Crystal Structures, Optical Properties, and Catalytic Activity of Homoleptic Zn(II) and Cd(II) β-oxodithioester Complexes. Inorg. Chem. 2020, 59, 11417–11431. [Google Scholar] [CrossRef] [PubMed]
  21. Dutta, B.; Jana, R.; Sinha, C.; Ray, P.P.; Mir, M.H. Synthesis of Cd(II) based 1D coordination polymer by in situ ligand generation and fabrication of photosensitive electronic device. Inorg. Chem. Front. 2018, 5, 1998–2005. [Google Scholar] [CrossRef]
  22. Mirtamizdoust, B.; Hanifehpour, Y.; Behzadfar, E.; Roodsari, M.S.; Jung, J.H.; Joo, S.W. A novel nano-structured three-dimensional supramolecular metal-organic framework for cadmium (II): A new precursor for producing nano cadmium oxide. J. Mol. Struct. 2020, 1201, 127191. [Google Scholar] [CrossRef]
  23. Hanifehpour, Y.; Morsali, A.; Mirtamizdoust, B.; Joo, S.W. Sonochemical synthesis of tri-nuclear lead(II)-azido nano rods coordination polymer with 3,4,7,8-tetramethyl-1,10-phenanthroline (tmph): Crystal structure determination and preparation of nano lead(II) oxide. J. Mol. Struct. 2015, 1079, 67–73. [Google Scholar] [CrossRef]
  24. Hanifehpour, Y.; Mirtamizdoust, B.; Hatami, M.; Khomami, B.; Joo, S.W. Synthesis and structural characterization of new bismuth (III) nano coordination polymer: A precursor to produce pure phase nano-sized bismuth (III) oxide. J. Mol. Struct. 2015, 1091, 43–48. [Google Scholar] [CrossRef]
  25. Hanifehpour, Y.; Morsali, A.; Soltani, B.; Mirtamizdoust, B.; Joo, S.W. Ultrasound-assisted fabrication of a novel nickel(II)-bis-pyrazolyl borate twonuclear discrete nano-structured coordination compound. Ultrason Sonochem. 2017, 34, 519–524. [Google Scholar] [CrossRef] [PubMed]
  26. Hammud, H.H.; Kortz, U.; Bhattacharya, S.; Demirdjian, S.; Hariri, E.; Isber, S.; Choi, E.S.; Mirtamizdoust, B.; Mroueh, M.; Daher, C.F. Structure, DFT studies, Magnetism and Biological activity of Bis[(μ 2-azido)-chloro-(1,10-phenanthroline)-copper(II)] complex. Inorganica Chim. Acta. 2020, 506, 119533. [Google Scholar] [CrossRef]
  27. Kravtsov, V.C.; Lozovan, V.; Siminel, N.; Coropceanu, E.B.; Kulikova, O.V.; Costriucova, N.V.; Fonari, M.S. Fonari, From 1D to 2D Cd(II) and Zn(II) coordination networks by replacing monocarboxylate with dicarboxylates in partnership with azine ligands: Synthesis, crystal structures, inclusion, and emission properties. Polyhedron 2020, 180, 114411. [Google Scholar] [CrossRef]
  28. Mercury 2.4, Copyright Cambridge Crystallographic Data Centre: 12 Union Road. Cambridge, UK; pp. 2001–2010.
  29. Oxford Diffraction. CrysAlis RED and CrysAlis CCD software (Ver. 1.171.32.5); Oxford Diffraction Ltd.: Abingdon, Oxfordshire, UK, 2010. [Google Scholar]
  30. Pal, A.; Chand, S.; Elahi, S.M.; Das, M.C. A microporous MOF with a polar pore surface exhibiting excellent selective adsorption of CO2 from CO2–N2 and CO2–CH4 gas mixtures with high CO2 loading. Dalton Trans. 2017, 46, 15280–15286. [Google Scholar] [CrossRef]
  31. Dong, Y.B.; Smith, M.D.; Loye, H.C.Z. New Inorganic/Organic Coordination Polymers Generated from Bidentate Schiff-Base Ligands. Inorg. Chem. 2000, 39, 4927–4935. [Google Scholar] [CrossRef]
  32. Harrowfield, J.M.; Miyamae, H.; Skelton, B.W.; Soudi, A.A.; White, A.H. Lewis-Base Adducts of Lead (II) Compounds. XX. Synthesis and Structure of the 1:1 Adduct of Pyridine With Lead (II) Thiocyanate. Aust. J. Chem. 1996, 49, 1165–1169. [Google Scholar] [CrossRef]
  33. Patterson, A.L. The Scherrer formula for X-ray particle size determination. Phys. Rev. 1939, 56, 978. [Google Scholar] [CrossRef]
  34. Gao, E.Q.; Cheng, A.L.; Xu, Y.X.; Yan, C.H.; He, M.Y. New Inorganic–Organic Hybrid Supramolecular Architectures Generated from 2, 5-Bis (3-pyridyl)-3, 4-diaza-2, 4-hexadiene. Cryst. Growth Des. 2005, 5, 1005–1011. [Google Scholar] [CrossRef]
  35. Lee, G.H.; Wang, H.T. Synthesis and Crystal Structure of a Three-Dimensional Nickel (II) Coordination Polymer with 1, 4-Bis (3-pyridyl)-2, 3-diazo-1, 3-butadiene as a Ligand. Anal. Sci. X-ray Struct. Anal. Online 2007, 23, x1–x2. [Google Scholar] [CrossRef] [Green Version]
  36. Soltanzadeh, N.; Morsali, A. Metal–organic supramolecular assemblies generated from bismuth (III) bromide and polyimine ligands. Polyhedron 2009, 28, 703–710. [Google Scholar] [CrossRef]
  37. Khanpour, M.; Morsali, A. Solid state crystal-to-crystal transformation from a monomeric structure to 1-D coordination polymers on anion exchange. CrystEngComm 2009, 11, 2585–2587. [Google Scholar] [CrossRef]
  38. Mahmoudi, G.; Morsali, A. Three new HgII metal–organic polymers generated from 1, 4-bis (n-pyridyl)-3, 4-diaza-2, 4-hexadiene ligands. Inorg. Chim. Acta 2009, 362, 3238–3246. [Google Scholar] [CrossRef]
  39. Velusamy, P.; Ramesh Babu, R.; Ramamurthi, K.; Elangovan, E.; Viegas, J.; Dahlem, M.S.; Arivanandhan, M. Characterization of spray pyrolytically deposited high mobility praseodymium doped CdO thin films. Ceram. Int. 2016, 42, 12675–12685. [Google Scholar] [CrossRef]
Scheme 1. Schematic of 2-bis (1-(pyridin-3-yl)ethylidene)hydrazine) Schiff-base ligand.
Scheme 1. Schematic of 2-bis (1-(pyridin-3-yl)ethylidene)hydrazine) Schiff-base ligand.
Crystals 11 00197 sch001
Scheme 2. Brached tube method [32].
Scheme 2. Brached tube method [32].
Crystals 11 00197 sch002
Scheme 3. The synthetic techniques and the produced materials. The FT-IR spectra of synthesized crystal and nano-structure of coordination polymer ([Cd(L)(NO2)2]n) indicate the characteristic absorption bands for organic Schiff base ligand. The weak band near 3000–3100 cm−1 is assigned to aromatic C–H groups’ stretching vibrations. The C=N tensile band is a sharp and strong peak in the 1622 area. The absorption band in the 1400–1600 cm−1 range is attributed to the Schiff base ligand’s aromatic ring vibrations. There is a good match between FT-IR spectra of the complex and nano-structure in all areas (Figure 1).
Scheme 3. The synthetic techniques and the produced materials. The FT-IR spectra of synthesized crystal and nano-structure of coordination polymer ([Cd(L)(NO2)2]n) indicate the characteristic absorption bands for organic Schiff base ligand. The weak band near 3000–3100 cm−1 is assigned to aromatic C–H groups’ stretching vibrations. The C=N tensile band is a sharp and strong peak in the 1622 area. The absorption band in the 1400–1600 cm−1 range is attributed to the Schiff base ligand’s aromatic ring vibrations. There is a good match between FT-IR spectra of the complex and nano-structure in all areas (Figure 1).
Crystals 11 00197 sch003
Figure 1. FT–IR spectra of the complex and nano-structures of coordination polymer.
Figure 1. FT–IR spectra of the complex and nano-structures of coordination polymer.
Crystals 11 00197 g001
Figure 2. SEM images of the [Cd(L)(NO2)2]n.
Figure 2. SEM images of the [Cd(L)(NO2)2]n.
Crystals 11 00197 g002
Figure 3. The PXRD patterns of compound [Cd(L)(NO2)2]n (1): (a) Simulated from single-crystal X-ray data and (b) nanostructure of (1) prepared by sonochemical method. Based on the crystal structure of 1, the composition and stereochemistry of a building block [Cd(L)(NO2)2]n) were disclosed. A view of the molecular structure is shown in Figure 4.
Figure 3. The PXRD patterns of compound [Cd(L)(NO2)2]n (1): (a) Simulated from single-crystal X-ray data and (b) nanostructure of (1) prepared by sonochemical method. Based on the crystal structure of 1, the composition and stereochemistry of a building block [Cd(L)(NO2)2]n) were disclosed. A view of the molecular structure is shown in Figure 4.
Crystals 11 00197 g003
Figure 4. Asymmetric unit of the [Cd(L)(NO2)2]n.
Figure 4. Asymmetric unit of the [Cd(L)(NO2)2]n.
Crystals 11 00197 g004
Figure 5. The coordination polymer fragment showing the 2D arrangement of complex (Left), The coordination sphere of Cd(II) (Right).
Figure 5. The coordination polymer fragment showing the 2D arrangement of complex (Left), The coordination sphere of Cd(II) (Right).
Crystals 11 00197 g005
Figure 6. From di nuclear complexes to 3D networks.
Figure 6. From di nuclear complexes to 3D networks.
Crystals 11 00197 g006
Figure 7. XRD pattern of cadmium oxide obtained by decomposition of (1).
Figure 7. XRD pattern of cadmium oxide obtained by decomposition of (1).
Crystals 11 00197 g007
Figure 8. EDX of the as-prepared CdO nanoparticles.
Figure 8. EDX of the as-prepared CdO nanoparticles.
Crystals 11 00197 g008
Figure 9. (a) SEM photograph and (b) size distribution graph of nano cadmium oxide (obtained by thermolysis of precursor 1).
Figure 9. (a) SEM photograph and (b) size distribution graph of nano cadmium oxide (obtained by thermolysis of precursor 1).
Crystals 11 00197 g009
Figure 10. CdO XPS spectra (a) Cd scan and (b) O scan obtained by thermolysis of 1.
Figure 10. CdO XPS spectra (a) Cd scan and (b) O scan obtained by thermolysis of 1.
Crystals 11 00197 g010
Figure 11. Thermal gravimetric (TG) thermogram of [Cd(L)(NO2)2]n.
Figure 11. Thermal gravimetric (TG) thermogram of [Cd(L)(NO2)2]n.
Crystals 11 00197 g011
Table 1. Crystal data and structure refinements of complex.
Table 1. Crystal data and structure refinements of complex.
Chemical FormulaC14H14CdN6O4
Mr442.7
Crystal system, space groupOrthorhombic, P212121
Temperature (K)120
a, b, c (Å)4.5208 (1), 16.2510 (2), 22.0539 (4)
V3)1620.25 (5)
Z4
Radiation typeCu Kα
µ (mm−1)11.12
Crystal size (mm)0.38 × 0.10 × 0.03
DiffractometerGoniometer Xcalibur, detector: Atlas (Gemini ultra Cu)
Absorption correctionAnalytical CrysAlis PRO, Agilent Technologies (2010), Analytical numeric absorption correction using a multifaceted crystal model
Tmin, Tmax0.181, 0.723
No. of measured, idenpendent and observes [1 > 2σ(1) reflections20706, 2895, 2742
Rint0.039
(sinθ/λ)max−1)0.598
Refinement R[F2 > 2σ(F2)], wR(F2), S0.020, 0.057, 1.13
No. of reflections2895
No. of parameters227
H-atom treatmentH-atom parameters constrained
Δ›max, Δ›min (e Å−3)0.30, −0.37
Table 2. Chosen bond lengths [Å] & angles [°] for 1.
Table 2. Chosen bond lengths [Å] & angles [°] for 1.
1.234 (4)O2—N52.390 (2)Cd1—O1
1.252 (5)O3—N62.368(2)Cd1—O3
1.255 (5)O4—N62.470(2)Cd1—O2
1.335 (5)N1—C12.389(2)Cd1—O4
1.340 (5)N1—C52.328 (2)Cd1—O1i
1.397 (5)N2—N32.358 (2)Cd1—N1i
1.291 (5)N2—C62.356 (2)Cd1—N4
1.346 (5)N4—C121.288 (5)N3—C8
120.2 (2)Cd1ii—N1—C51.339 (5)N4—C11
154.44 (9)N1i—Cd1—N487.27 (11)O1—Cd1—N1i
121.9 (2)Cd1ii—N1—C189.09 (10)O1—Cd1—N4
Symmetry code(s): (i) −x, y + 1/2, −z + 1/2; (ii) −x, y−1/2, −z + 1/2.
Table 3. Selected non-covalent contacts in the crystal structure of 1 [Å and °].
Table 3. Selected non-covalent contacts in the crystal structure of 1 [Å and °].
D-H...Ad(D-H)d(H...A)d(D...A)<(DHA)
C(13)-H(13)...O(4)0.962.5253.237(4)130.9
C(13)-H(7)...C(12)0.962.7082.789(4)155.5
C(4)-H(4)...O(3)0.952.5723.442(5)151.2
C(1)-H(1)...N(5)0.962.5143.328(4)142.6
C(5)-H(5)...O(3)0.952.6413.674(5)145.7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hanifehpour, Y.; Dadashi, J.; Mirtamizdoust, B. Ultrasound-Assisted Synthesis and Crystal Structure of Novel 2D Cd (II) Metal–Organic Coordination Polymer with Nitrite End Stop Ligand as a Precursor for Preparation of CdO Nanoparticles. Crystals 2021, 11, 197. https://doi.org/10.3390/cryst11020197

AMA Style

Hanifehpour Y, Dadashi J, Mirtamizdoust B. Ultrasound-Assisted Synthesis and Crystal Structure of Novel 2D Cd (II) Metal–Organic Coordination Polymer with Nitrite End Stop Ligand as a Precursor for Preparation of CdO Nanoparticles. Crystals. 2021; 11(2):197. https://doi.org/10.3390/cryst11020197

Chicago/Turabian Style

Hanifehpour, Younes, Jaber Dadashi, and Babak Mirtamizdoust. 2021. "Ultrasound-Assisted Synthesis and Crystal Structure of Novel 2D Cd (II) Metal–Organic Coordination Polymer with Nitrite End Stop Ligand as a Precursor for Preparation of CdO Nanoparticles" Crystals 11, no. 2: 197. https://doi.org/10.3390/cryst11020197

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