Next Article in Journal
A Numerical Study Based on Haar Wavelet Collocation Methods of Fractional-Order Antidotal Computer Virus Model
Next Article in Special Issue
Electron Microscopy Study of Structural Defects Formed in Additively Manufactured AlSi10Mg Alloy Processed by Equal Channel Angular Pressing
Previous Article in Journal
A New Detection Function Model for Distance Sampling Based on the Burr XII Model
Previous Article in Special Issue
Metamaterial with Tunable Positive and Negative Hygrothermal Expansion Inspired by a Four-Fold Symmetrical Islamic Motif
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Molecular, and Supramolecular Structures of Two Azide-Bridged Cd(II) and Cu(II) Coordination Polymers

by
Mezna Saleh Altowyan
1,
Eman M. Fathalla
2,
Jörg H. Albering
3,
Assem Barakat
4,
Morsy A. M. Abu-Youssef
2,*,
Saied M. Soliman
2,* and
Ahmed M. A. Badr
2
1
Department of Chemistry, College of Science, Princess Nourah bint Abdulrahman University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
2
Department of Chemistry, Faculty of Science, Alexandria University, P.O. Box 426, Ibrahimia, Alexandria 21321, Egypt
3
Graz University of Technology, Mandellstr. 11/III, A-8010 Graz, Austria
4
Department of Chemistry, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Symmetry 2023, 15(3), 619; https://doi.org/10.3390/sym15030619
Submission received: 2 February 2023 / Revised: 21 February 2023 / Accepted: 22 February 2023 / Published: 1 March 2023
(This article belongs to the Special Issue Materials Science and Symmetry)

Abstract

:
Two 1D coordination polymers were synthesized by reaction of two ligands, 2-amino-4-picoline (2A4Pic) and quinoline-6-carboxylic acid (Qu-6-COOH) with two metal (II) nitrate (M = Cd and Cu) in the presence of azide as a linker. The synthesized metal complexes [Cd(2A4Pic)2(N3)2]n; (1) and [Cu(Qu-6-COO)(N3)(H2O)]n; (2) were isolated in single crystals and their X-ray structures revealed a 1D polymeric structure. Due to symmetry considerations, the asymmetric formula is half a [Cd(2A4Pic)2(N3)2] unit for 1 and one [Cu(Qu-6-COO)(N3)(H2O)] unit for 2. In complex 1, the Cd(II) is hexa-coordinated with two 2A4Pic molecules and four μ(1,1) azide units. Hence, the CdN6 coordination environment has a slightly distorted octahedral geometry. In 2, the Cu(II) is hexa-coordinated with three different ligands (Qu-6-COO¯, H2O and μ(1,1) N3¯) where all are connectors between the crystallographically related Cu(II) sites. Additionally, complex 2 distorted CuN2O4 octahedral geometry. In both complexes, the polymer arrays are connected by N…H hydrogen bonds and π–π stacking interactions. Based on Hirshfeld analysis, the percentages of N…H contacts are 43.1 and 27.4% for 1 and 2, respectively, while %C...C are 5.6 and 9.3%, respectively. Analysis of Cu-N, Cu-O, and Cd-N bonds using DFT calculations showed predominantly closed-shell coordination interactions with little covalent characters. Additionally, the negatively charged ligand groups were found to compensate the positive charge of the central metal ion to a larger extent than the electrically neutral ligands.

1. Introduction

Coordination polymers (CPs) attracted too much attention from researchers due to their wide range of applications [1,2,3,4,5,6]. These compounds were applied in a variety of fields, including gas adsorption [7,8], catalysis [9,10], and drug delivery [9,11]. Additionally, many CPs are interesting magnetic materials [12] and were used in device fabrication [13,14] and molecular sensing [7,15,16,17,18]. Other notable applications of CPs are their uses in solar cells, Schottky barrier diodes (SBD), and other electronic and opto-electronic devices [13,19,20,21,22].
On the other hand, structural flexibility and electronic characteristics of the organic ligand, and also nature of metal ion are important factors for construction of CPs [23,24]. In addition, the dimensionality of CPs could be enhanced by selection of suitable linkers. Azide and carboxylate anions are interesting bridging ligands because of their structural versatility in coordination chemistry [25,26,27,28,29,30,31]. Azide ion is a highly symmetric anion that has small size and linear shape. Hence, it has a high ability to propagate the magnetic interactions between paramagnetic centers, leading to CPs with interesting magnetic properties [6,32,33]. Additionally, the coordinated azide has two N-N bonds, which are less symmetric compared to the free one where the degree of asymmetry depends on the azide bonding mode. Among the pool of bonding modes of azide ion (Scheme 1), the end-on (μ(1,1) or EO) and end-to-end (μ(1,3) or EE) are the most prevalent [6,32,33]. These bridging modes are responsible for the construction of coordination compounds with varying nuclearity and dimensionality [32,33,34,35,36,37,38]. In this regard, many fascinating and structurally attractive Cu(II) [39,40,41] and Cd(II) [42,43,44,45,46,47,48,49,50] complexes were constructed. Additionally, carboxylate ligands are extremely versatile linkers because of their different coordination modes [31].
In the light of the exciting coordination chemistry of azide ion as a linker [51], the current work aims to synthesize new azido CPs comprising 2-amino-4-picoline (2A4Pic) and quinoline-6-carboxylic acid (Qu-6-COOH) (Scheme 2). Their molecular and supramolecular structure features were described. Theoretical studies were performed to describe the nature and strength of the metal–ligand interactions.

2. Materials and Methods

All chemicals, physicochemical characterizations, and details of solving the X-ray structure [52,53] are described in Supplementary data.

2.1. Synthesis of [Cd(2A4Pic)2(N3)2]n; (1)

2-Amino-4-picoline (21.6 mg, 0.2 mmol) in 10 mL EtOH was added to Cd(NO3)2.4H2O (30.8 mg, 0.1 mmol) solution in 10 mL distilled water followed by adding 0.5 mL saturated NaN3 aqueous solution drop wisely, then the clear mixture was left for slow evaporation at ambient conditions. After five days, complex 1 was formed as colorless needle crystals.
[Cd(2A4Pic)2(N3)2]n; (1) (69% yield). Anal. Calc. C12H16CdN10: C, 34.92; H, 3.91; N, 33.94; and Cd, 27.24%. Found: C, 34.74; H, 3.83; N, 33.61; and Cd, 27.07%.

2.2. Synthesis of [Cu(Qu-6-COO)(N3)(H2O)]n; (2)

Quinoline-6-carboxylic acid (17.4 mg, 0.1 mmol) in 10 mL EtOH was added to Cu(NO3)2.3H2O (24.2 mg, 0.1 mmol) in 10 mL EtOH followed by dropwise addition of 0.5 mL of saturated NaN3 aqueous solution, then the slightly turbid solution was filtered and the clear filtrate was allowed to slowly evaporate at ambient conditions. Complex 2 was obtained as dark green plate crystals after one week.
[Cu(Qu-6-COO)(N3)(H2O)]n; (2) (66% yield) Anal. Calc. C10H8CuN4O3: C, 40.61; H, 2.73; N, 18.94; and Cu, 21.49%. Found: C, 40.49; H, 2.66; N, 18.72; and Cu, 21.35%.

2.3. Hirshfeld and DFT Calculations

Details regarding Hirshfeld [54] and DFT computations are described in Supplementary data [55,56,57,58,59,60,61].

3. Results and Discussion

3.1. X-ray Crystal Structure Description of 1

Structure of the asymmetric unit and the coordination environment of [Cd(2A4Pic)2(N3)2]n (1) is shown in Figure 1A and Figure 2A, respectively. The crystal system of 1 is monoclinic and the space group is P21/c. The unit cell parameters are a = 8.9157(3) Å, b = 3.69130(10) Å, c = 23.1784(7) Å, and β = 90.0980(10)°, while Z = 2 and V = 762.81(4) Å3 (Table 1). The structure of this highly symmetric complex comprised only a half [Cd(2A4Pic)2(N3)2] unit per asymmetric formula (Figure 1).
The Cd(II) ion is coordinated with two 2A4Pic ligand units and four N3¯ groups where each two trans Cd-N bonds are equidistant due to symmetry consideration. The Cd to N distance with 2A4Pic is 2.3305(15) Å. The azide group has a μ(1, 1) bonding mode and it acts as a linker between the Cd(II) centers via Cd1-N1 and Cd1-N1* bonds. The corresponding Cd-N distances are 2.3496(15) and 2.4198(15) Å, respectively. Similar to 1, the structurally related [Cd(3-acetylpyridine)2(N3)2]n 1D polymer has only μ(1,1) bridging azide, while the [Cd(4-ethylpyridine)2(N3)2]n, [Cd(4-hydroxymethylpyridine)2(N3)2]n, [Cd(3-hydroxymethylpyridine)2(N3)2]n [62], and [Cd(4-azidopyridine)2(N3)2]n complexes [63] are 1D polymer but with mixed μ(1,1) and μ(1,3) bridged azides. Hence, nature of the substituent at the pyridine ring has a significant impact on the azide bonding mode. In [Cd(4-azidopyridine)2(N3)2]n complex, the Cd-N distances with the μ(1,1) bridged azide are found to be 2.3371(17) and 2.3442(18) Å, which are very close to the corresponding Cd-N distances in 1 [63]. Hence, the structure of 1 is a 1D coordination polymer extended along the crystallographic b-direction (Figure 2). As can be seen from this figure, the coordinated azide groups form with Cd(II) the four membered (Cd1N1)2 ring, in which the N1-Cd1-N1 and Cd1-N1-Cd1 angles are 78.59(5)° and 101.41(5)°, respectively. The organic ligand units were found almost perpendicular to the extension of the coordination polymer and distributed above and below the 1D polymer array where the angle between the pyridine and the (Cd1N1)2 rings is 72.68°. All the trans N-Cd-N angles are perfectly the same as for ideal octahedron (180°) while the cis N-Cd-N angles are slightly deviated from the ideal value of 90° (Table 2).
As clearly seen in Figure 2, different levels of π–π stacking between the pyridine groups were detected. The shortest contact distances are C2…C1 (3.506 Å), C3…C2 (3.544 Å), C3…C6 (3.510 Å), and C4…C5 (3.509 Å), while the corresponding centroid–centroid distance is 3.691 Å. In addition, the coordination polymer arrays are interconnected with each other along the crystallographic a-direction via N5-H1A…N3 bridges between the amino group as the H-bond donor and the freely uncoordinated N-atom from the N3¯ group (Figure 3). The H1A…N3 and N5…N3 distances are 2.37(3) and 3.184(2) Å, respectively, while the N5-H1A…N3 angle is 167(2)°.

3.2. X-ray Crystal Structure Description of 2

The X-ray structure analysis showed a polymeric structure for the [Cu(Qu-6-COO)(N3)(H2O)]n complex (2). The asymmetric formula represents one [Cu(Qu-6-COO)(N3)(H2O)] unit as shown in Figure 1B. The crystal system of 2 is monoclinic and space group is P21/c. The unit cell parameters are a = 8.0773(5) Å, b = 6.4187(3) Å, c = 20.9267(12) Å, and β = 96.499(2)°, while Z = 4 and V = 1077.99(10) Å3 (Table 1). The Cu(II) is hexa-coordinated, but in this case, there are three types of ligands that are the quinoline-6-carboxylate anion, azide group, and the water molecule, where all are acting as bridging ligands connecting the crystallographically related Cu(II) sites along the b-direction. Each Cu(II) is coordinated with two sets of theses ligands, and due to symmetry consideration, each two similar ligand units are anti to one another (Figure 4A). The quinoline-6-carboxylate ligand is found coordinated by two Cu(1) ions via the carboxylate oxygen atoms O2 and O3 in the syn–syn geometry [64]. The corresponding Cu-O distances are 1.945(3) and 1.946(4) Å, respectively. In the structurally related [Cu(2-Chloro-6-methylnicotinate)(N3)(MeOH)]n complex, the bridged carboxylate forms two Cu-O bonds with very close distances as in complex 2. In this case, the Cu-O distances were found to be 1.953(1) to 2.010(1) Å [64]. Additionally, the water molecule is bridging the Cu(1) sites via the two significantly different Cu1-O1 (2.374(4) Å) and Cu12-O1 (2.771(3) Å) bonds (Table 3). Bridging water ligands in such complexes are not common. A bridging H2O ligand was found in the [Ni2(H2O)(4-amino-3,5-dimethyl-1,2,4-triazole)2(pivalate)4(H2O)2] complex [65]. It is worth noting that the bridged water metal complexes are not common in literature. For example, the bridged H2O was found in [Ni2(H2O)(4-amino-3,5-dimethyl-1,2,4-triazole)2(pivalate)4(H2O)2] complex [65]. Additionally, the penta-nuclear [Ag5(PTDM)4(H2O)6(ClO4)4]ClO4 complex, where PTDM is 4,4′-[6-(3,5-dimethyl-1H-pyrazol-1-yl)-1,3,5-triazine-2,4-diyl]dimorpholine, comprised a bridged water structure [66]. The coordination environment of Cu(1) in 2 is completed by interaction with two azide groups, which have a μ(1,1) mode. The two Cu-N bond distances that occurred between the bridged azide and Cu(II) are marginally different (Cu1-N1: 1.986(4) and Cu1-N11: 1.995(4) Å) [64]. Hence, the structure of 2 is a 1D coordination polymer extended along the b-direction (Figure 4B). The donor atoms N1 and O1 form with Cu1 the four membered ring N1Cu1O1Cu1, in which the O1-Cu1-N1 angles range from 78.1(1) to 88.5(1)° while the Cu-N-Cu and Cu-O-Cu angles are 108.7(2) and 77.6(1)°, respectively. The trans N1-Cu1-N1, O1-Cu1-O1, and O2-Cu1-O3 angles are 166.1(2), 175.7(1), and 178.2(1)°, respectively. Hence, the CuN2O4 coordination geometry is distorted octahedron.
The N4 atom of Qu-6-COO¯ and the free N3 atom of N3¯ ion do not participate in the coordination with Cu(II) ion. These N-sites are acting as H-bond acceptors, which form strong O1-H1...N3 and O1-H2...N4 H-bridges with the coordinated H2O molecule as the H-bond donor. The H1...N3 and H2...N4 distances are 2.26(3) and 1.98(4) Å, respectively, while the O1…N3 and O1…N4 donor–acceptor distances are 3.079(6) and 2.799(5) Å, respectively. The respective donor-H-acceptor angles are 163(5) and 169(4)°, respectively. The hydrogen bond network is presented in Figure 5.
In addition, there are a number of π–π stacking interactions which connect the 1D chains via short C5…C8 (3.396 Å), C3…C10 (3.327 Å), and C6…C8 (3.320 Å) interactions, while the centroid–centroid distance is 3.687 Å. These results confirm the presence of significant π–π stacking between the quinoline rings (Figure 6).

3.3. Hirshfeld Analysis of Molecular Packing

X-ray structures of 1 and 2 revealed 1D coordination polymeric structures for both complexes. Hence, Hirshfeld calculations were performed on the [Cd(2A4Pic)2(N3)4] and [Cu(Qu-6-COO)(N3)(H2O)] complex units. Decomposition analysis of the different intermolecular contacts is shown in Figure 7, while their percentages are listed in Table S1 (Supplementary Materials). It is shown that the largest contributions are related to the N…H (43.1, 27.4%) and H...H (34.5, 27.1%) interactions for complexes 1 and 2, respectively.
For complex 1, many N…H and H...H contacts are contributed in the molecular packing and crystal stability. Their dnorm surfaces and fingerprint (FP) plots showed large intense red spots and sharp spikes, respectively (Figure S1, Supplementary Materials). The H...H contacts appeared as red spots, where all are related to H7…H8 interaction with interaction distance of 2.101 Ǻ. On the other hand, the N1…H2 interaction is the shortest N…H contact (2.164 Ǻ). Other longer contacts, such as N3…H1 (2.203Ǻ) and N2…H5 (2.565Ǻ) were also observed as red spots in the dnorm surface. It is worth mentioning that the shape index and curvedness maps can give good indications on the presence of aromatic π–π stacking interactions. The red/blue triangles in the former and the flat green area in the latter are clearly evident from Figure S2, (Supplementary Materials). These criteria are achieved in complex 1, which revealed the π–π stacking between the organic ligand units. Additionally, the Cd-N coordination interactions represent 4.3% from the total fingerprint area of complex 1. It is clear from the dnorm surface that the Cd-N interactions appeared as intense red spots, which confirm the polymeric structure of this complex via the azide ion as a linker (Figure 8).
For complex 2, the decomposition of the different contacts also revealed the polymeric nature of this complex via Cu-N and Cu-O coordination interactions. The intense red regions in the dnorm map close to the central Cu(II) are related to the coordination interactions with the bridged azido, carboxylate, and water ligand groups, which form the 1D polymeric chain of this complex. The %Cu-O and %Cu-N interactions are 5.3 and 3.5%, respectively. Additionally, the Cu-O and Cu-N interactions appeared as sharp spikes in the FP plots (Figure 9). Other contacts, such as N…H (27.4%), C...C (9.3%), and N...O (8.1%) are important for the crystal stability of complex 2 (Figure S3, Supplementary Materials). Additionally, the presence of the significant amount of C...C contacts, along with the red/blue triangles in the SI map related to the coordinated quinoline moiety confirm the presence of π–π stacking interactions (Figure S4, Supplementary Materials). The C10…C3 (3.329 Å) and C6…C8 (3.319 Å) are the shortest. Other important short interactions are listed in Table 4.

3.4. The Atoms in Molecules (AIM) Analysis

The AIM analysis is widely used to describe the nature and strength of bonding in different types of molecules [67,68]. The AIM theory is extensively applied to differentiate between covalent and noncovalent interactions [69,70]. The properties of the coordinate bonds were described based on electron density ρ(r), its laplacian ∇2ρ(r), kinetic energy density G(r), potential energy density V(r), and kinetic energy of the Hamiltonian H(r) = G(r) + V(r) at the bond critical points (BCPs) [71,72]. These topological parameters are listed in Table S2 (Supplementary Materials). Our results generally demonstrate the mainly closed-shell character of all coordinate bonds in both complexes because ρ(r) values are less than 0.1 a.u. along with the small positive value of ∇2ρ(r). In accordance with the observed shorter Cu-O(carboxylate) (1.945 and 1.946 Å) bond distances than the Cu-O(water) (2.771 and 2.374 Å) bonds in complex 2, the ρ(r) and ∇2ρ(r) of the former are larger than those for the latter [73]. Additionally, the H(r) parameter at the BCPs has small negative values, which are close to zero, indicating the presence of little covalent characters in the studied Cd-N, Cu-N, and Cu-O coordinate bonds. Additionally, the results reveal the higher covalent characters for the Cu-O(carboxylate) bonds than the Cu-O(water) bonds in complex 2, where the H(r) values for the former are significantly more negative (−0.0153 and −0.0157 a.u) than those for the latter.

3.5. Natural Population Analysis

One of the best models for calculating the charge population is the natural charge analysis [74]. In transition metal complexes, the interaction between the ligand groups and metal ion affect the charges of these fragments. The charges of the central metal ion in the two complexes are changed significantly after complexation. The charge of the Cd(II) and Cu(II) ions are changed to 1.0374 e and 0.9199 e in complexes 1 and 2, respectively. As a result, the amounts of electron density transferred from the ligands to metal ion are 0.9626 and 1.0801 e, respectively. In complex 1, the calculated charges transferred from each N3¯ and 2A4Pic units were estimated to be 0.1948 and 0.0918 e (average values), respectively. In complex 2, the net charge of the Qu-6-COO¯, N3, and H2O ligand groups are estimated to be −0.8032, −0.7176, and 0.0608e, respectively. As a result, the water molecule transferred the lowest amount (0.0608 e) of electron density to Cu(II) cation, while the N3¯ ligand transferred the largest amount (0.2824 e). In addition, each Qu-6-COO¯ transferred 0.1968 e to the Cu(II) ion. It is also noticed that each coordinated azide transferred a higher amount of electron density to the metal ion in complex 1 than 2. The reason might be explained in terms of the higher compensation of the positive charge of Cu(II) ion due to the presence of another negatively charged ligand group (Qu-6-COO¯) in 2, while the other ligand groups in 1 are electrically neutral (2A4Pic).

4. Conclusions

The supramolecular structures of two highly symmetric 1D coordination polymers [Cd(2A4Pic)2(N3)2]n; (1) and [Cu(Qu-6-COO)(N3)(H2O)]n (2) were presented. Both complexes were assembled from the reaction of the organic ligand with the corresponding metal(II) nitrate in the presence of azide as a linker. The CdN6 and CuN2O4 coordination spheres in complexes 1 and 2 have distorted octahedral configuration. In both complexes, the azide ligand has a μ(1,1) bridging mode. In addition, the Qu-6-COO ¯ and water in complex 2 are connectors between the Cu(II) sites. Hirshfeld calculations revealed the 1D polymeric backbone of both complexes where the 1D backbone structures are interconnected by N…H H-bonds and π–π interactions. The N…H contacts are contributed by 43.1 and 27.4% in the molecular packing of 1 and 2, respectively. In addition, the %C...C contacts are 5.6 and 9.3%, respectively. AIM analysis showed predominant closed-shell interactions for the Cd-N, Cu-N, and Cu-O coordinate bonds. Additionally, natural charge analysis sheds the light on the higher compensation of the central metal ion positive charge in 2 than in 1.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/sym15030619/s1, Chemicals and Physicochemical characterizations; Crystal structure analysis; DFT calculations; CheckCIF reports; Figure S1. The FP plots and dnorm surface maps for the N…H and H…H contacts in complex 1; Figure S2. Shape index (SI; left) and curvedness (right) maps reveal the π–π stacking interactions in complex 1; Figure S3. The FP plots and dnorm maps of the significant contacts in 2; Figure S4. Red/blue triangles in SI map reveal the presence of π–π stacking interactions in complex 2; Table S1. The %contacts in compounds 1 and 2; Table S2 AIM parameters (a.u.) at BCPs for complexes 1 and 2.

Author Contributions

Conceptualization, S.M.S., M.A.M.A.-Y. and A.M.A.B.; methodology, E.M.F. and J.H.A.; software, S.M.S., M.A.M.A.-Y., E.M.F. and A.M.A.B.; formal analysis, E.M.F. and J.H.A.; investigation, S.M.S., M.A.M.A.-Y., E.M.F. and A.M.A.B.; resources, M.S.A., A.B., S.M.S., M.A.M.A.-Y. and A.M.A.B.; data curation, M.S.A., A.B., S.M.S., M.A.M.A.-Y., E.M.F. and A.M.A.B.; writing—original draft preparation, S.M.S., M.A.M.A.-Y., J.H.A., E.M.F. and A.M.A.B.; writing—review and editing, S.M.S., M.A.M.A.-Y., J.H.A., E.M.F. and A.M.A.B.; supervision, S.M.S., M.A.M.A.-Y. and A.M.A.B. All authors have read and agreed to the published version of the manuscript.

Funding

Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2023R86), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Data Availability Statement

Not applicable.

Acknowledgments

Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2023R86), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mir, M.H.; Koh, L.L.; Tan, G.K.; Vittal, J.J. Single-crystal to single-crystal photochemical structural transformations of interpenetrated 3D coordination polymers by [2 + 2] cycloaddition reactions. Angew. Chem. Int. Ed. 2010, 49, 390–393. [Google Scholar] [CrossRef] [PubMed]
  2. Das, L.K.; Gómez-García, C.J.; Ghosh, A. Influence of the central metal ion in controlling the self-assembly and magnetic properties of 2D coordination polymers derived from [(NiL)2M]2+ nodes (M = Ni, Zn and Cd) (H2L = salen-type di-Schiff base) and dicyanamide spacers. Dalton Trans. 2015, 44, 1292–1302. [Google Scholar] [CrossRef] [PubMed]
  3. Das, L.K.; Diaz, C.; Ghosh, A. Antiferromagnetic mixed-valence Cu(I)–Cu(II) two-dimensional coordination polymers constructed by double oximato bridged Cu(II) dimers and CuISCN based one-dimensional anionic chains. Cryst. Growth Des. 2015, 15, 3939–3949. [Google Scholar] [CrossRef]
  4. Yamada, T.; Otsubo, K.; Makiura, R.; Kitagawa, H. Designer coordination polymers: Dimensional crossover architectures and proton conduction. Chem. Soc. Rev. 2013, 42, 6655–6669. [Google Scholar] [CrossRef] [PubMed]
  5. Li, H.; Wang, Y.; He, Y.; Xu, Z.; Zhao, X.; Han, Y. Synthesis of several novel coordination complexes: Ion exchange, magnetic and photocatalytic studies. New J. Chem. 2017, 41, 1046–1056. [Google Scholar] [CrossRef]
  6. Mondal, M.; Jana, S.; Drew, M.G.; Ghosh, A. Application of two Cu(II)-azido based 1D coordination polymers in optoelectronic device: Structural characterization and experimental studies. Polymer 2020, 204, 122815. [Google Scholar] [CrossRef]
  7. Zhang, Z.; Zhao, Y.; Gong, Q.; Li, Z.; Li, J. MOFs for CO2 capture and separation from flue gas mixtures: The effect of multifunctional sites on their adsorption capacity and selectivity. Chem. Commun. 2013, 49, 653–661. [Google Scholar] [CrossRef]
  8. Zeng, L.W.; Hu, K.Q.; Mei, L.; Li, F.Z.; Huang, Z.W.; An, S.W.; Chai, Z.F.; Shi, W.Q. Structural diversity of bipyridinium-based uranyl coordination polymers: Synthesis, characterization, and ion-exchange application. Inorg. Chem. 2019, 58, 14075–14084. [Google Scholar] [CrossRef]
  9. Farrusseng, D. (Ed.) Metal-Organic Frameworks: Applications from Catalysis to Gas Storage; John Wiley & Sons: Hoboken, NJ, USA, 2011. [Google Scholar]
  10. Ganguly, S.; Kar, P.; Chakraborty, M.; Ghosh, A. The first alternating MnII–MnIII 1D chain: Structure, magnetic properties and catalytic oxidase activities. New J. Chem. 2018, 42, 9517–9529. [Google Scholar] [CrossRef]
  11. Kundu, T.; Mitra, S.; Patra, P.; Goswami, A.; Díaz Díaz, D.; Banerjee, R. Mechanical downsizing of a gadolinium(III)-based metal–organic framework for anticancer drug delivery. Chem. Eur. J. 2014, 20, 10514–10518. [Google Scholar] [CrossRef]
  12. Kar, P.; Guha, P.M.; Drew, M.G.; Ishida, T.; Ghosh, A. Spin-canted antiferromagnetic phase transitions in alternating phenoxo-and carboxylato-bridged MnIII-salen complexes. Eur. J. Inorg. Chem. 2011, 2011, 2075–2085. [Google Scholar] [CrossRef]
  13. Dutta, B.; Jana, R.; Bhanja, A.K.; Ray, P.P.; Sinha, C.; Mir, M.H. Supramolecular aggregate of cadmium(II)-based one-dimensional coordination polymer for device fabrication and sensor application. Inorg. Chem. 2019, 58, 2686–2694. [Google Scholar] [CrossRef]
  14. Ghorai, P.; Dey, A.; Brandão, P.; Benmansour, S.; Gómez García, C.J.; Ray, P.P.; Saha, A. Multifunctional Ni(II)-based metamagnetic coordination polymers for electronic device fabrication. Inorg. Chem. 2020, 59, 8749–8761. [Google Scholar] [CrossRef]
  15. Liu, J.Q.; Luo, Z.D.; Pan, Y.; Singh, A.K.; Trivedi, M.; Kumar, A. Recent developments in luminescent coordination polymers: Designing strategies, sensing application and theoretical evidences. Coord. Chem. Rev. 2020, 406, 213145. [Google Scholar] [CrossRef]
  16. Zhang, X.; Wang, W.; Hu, Z.; Wang, G.; Uvdal, K. Coordination polymers for energy transfer: Preparations, properties, sensing applications, and perspectives. Coord. Chem. Rev. 2015, 284, 206–235. [Google Scholar] [CrossRef]
  17. Xie, Z.; Ma, L.; deKrafft, K.E.; Jin, A.; Lin, W. Porous phosphorescent coordination polymers for oxygen sensing. J. Am. Chem. Soc. 2010, 132, 922–923. [Google Scholar] [CrossRef]
  18. Alsharabasy, A.M.; Pandit, A.; Farràs, P. Recent advances in the design and sensing applications of hemin/coordination polymer-based nanocomposites. Adv. Mater. 2021, 33, 2003883. [Google Scholar] [CrossRef]
  19. Stavila, V.; Talin, A.A.; Allendorf, M.D. MOF-based electronic and opto-electronic devices. Chem. Soc. Rev. 2014, 43, 5994–6010. [Google Scholar] [CrossRef] [Green Version]
  20. Dhakshinamoorthy, A.; Garcia, H. Catalysis by metal nanoparticles embedded on metal–organic frameworks. Chem. Soc. Rev. 2012, 41, 5262–5284. [Google Scholar] [CrossRef]
  21. Allendorf, M.D.; Bauer, C.A.; Bhakta, R.K.; Houk, R.J.T. Luminescent metal–organic frameworks. Chem. Soc. Rev. 2009, 38, 1330–1352. [Google Scholar] [CrossRef]
  22. Roy, S.; Halder, S.; Drew, M.G.; Ray, P.P.; Chattopadhyay, S. Fabrication of an active electronic device using a hetero-bimetallic coordination polymer. ACS Omega 2018, 3, 12788–12796. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. He, Y.; Li, B.; O’Keeffe, M.; Chen, B. Multifunctional metal–organic frameworks constructed from meta-benzenedicarboxylate units. Chem. Soc. Rev. 2014, 43, 5618–5656. [Google Scholar] [CrossRef] [PubMed]
  24. Lee, M.M.; Kim, H.Y.; Hwang, I.H.; Bae, J.M.; Kim, C.; Yo, C.H.; Kim, Y.; Kim, S.J. CdII MOFs constructed using succinate and bipyridyl ligands: Photoluminescence and heterogeneous catalytic activity. Bull. Korean Chem. Soc. 2014, 35, 1777–1783. [Google Scholar] [CrossRef] [Green Version]
  25. Escuer, A.; Aromí, G. Azide as a bridging ligand and magnetic coupler in transition metal clusters. Eur. J. Inorg. Chem. 2006, 2006, 4721–4736. [Google Scholar] [CrossRef]
  26. Zeng, Y.F.; Hu, X.; Liu, F.C.; Bu, X.H. Azido-mediated systems showing different magnetic behaviors. Chem. Soc. Rev. 2009, 38, 469–480. [Google Scholar] [CrossRef]
  27. Adhikary, C.; Koner, S. Structural and magnetic studies on copper(II) azido complexes. Coord. Chem. Rev. 2010, 254, 2933–2958. [Google Scholar] [CrossRef]
  28. Zheng, Y.Z.; Zheng, Z.; Chen, X.M. A symbol approach for classification of molecule-based magnetic materials exemplified by coordination polymers of metal carboxylates. Coord. Chem. Rev. 2014, 258, 1–15. [Google Scholar] [CrossRef]
  29. Kurmoo, M. Magnetic metal—Organic frameworks. Chem. Soc. Rev. 2009, 38, 1353–1379. [Google Scholar] [CrossRef]
  30. Escuer, A.; Esteban, J.; Perlepes, S.P.; Stamatatos, T.C. The bridging azido ligand as a central “player” in high-nuclearity 3D-metal cluster chemistry. Coord. Chem. Rev. 2014, 275, 87–129. [Google Scholar] [CrossRef]
  31. Yue, Q.; Gao, E.Q. Azide and carboxylate as simultaneous coupler for magnetic coordination polymers. Coord. Chem. Rev. 2019, 382, 1–31. [Google Scholar] [CrossRef]
  32. Bhowmik, P.; Biswas, S.; Chattopadhyay, S.; Diaz, C.; Gómez-García, C.J.; Ghosh, A. Synthesis, crystal structure and magnetic properties of two alternating double μ1,1 and μ1,3 azido bridged Cu(II) and Ni(II) chains. Dalton Trans. 2014, 43, 12414–12421. [Google Scholar] [CrossRef]
  33. Mukherjee, S.; Mukherjee, P.S. CuII–azide polynuclear complexes of Cu4 building clusters with Schiff-base co-ligands: Synthesis, structures, magnetic behavior and DFT studies. Dalton Trans. 2013, 42, 4019–4030. [Google Scholar] [CrossRef]
  34. Mautner, F.A.; Fischer, R.C.; Williams, B.R.; Massoud, S.S.; Salem, N.M. Hexanuclear cadmium(II) cluster constructed from tris (2-methylpyridyl) amine (TPA) and azides. Crystals 2020, 10, 317. [Google Scholar] [CrossRef] [Green Version]
  35. Mautner, F.A.; Fischer, R.C.; Reichmann, K.; Gullett, E.; Ashkar, K.; Massoud, S.S. Synthesis and characterization of 1D and 2D cadmium(II)-2,2′-bipyridine-N,N′-dioxide coordination polymers bridged by pseudohalides. J. Mol. Struct. 2019, 1175, 797–803. [Google Scholar] [CrossRef]
  36. Boonmak, J.; Nakano, M.; Chaichit, N.; Pakawatchai, C.; Youngme, S. Spin canting and metamagnetism in 2D and 3D cobalt(II) coordination networks with alternating double end-on and double end-to-end azido bridges. Inorg. Chem. 2011, 50, 7324–7333. [Google Scholar] [CrossRef]
  37. Mautner, F.A.; Louka, F.R.; Hofer, J.; Spell, M.; Lefèvre, A.; Guilbeau, A.E.; Massoud, S.S. One-dimensional cadmium polymers with alternative di (EO/EE) and di(EO/EO/EO/EE) bridged azide bonding modes. Cryst. Growth Des. 2013, 13, 4518–4525. [Google Scholar] [CrossRef]
  38. Massoud, S.S.; Louka, F.R.; Obaid, Y.K.; Vicente, R.; Ribas, J.; Fischer, R.C.; Mautner, F.A. Metal ions directing the geometry and nuclearity of azido-metal (II) complexes derived from bis (2-(3, 5-dimethyl-1 H-pyrazol-1-yl) ethyl) amine. Dalton Trans. 2013, 42, 3968–3978. [Google Scholar] [CrossRef]
  39. Lazari, G.; Stamatatos, T.C.; Raptopoulou, C.P.; Psycharis, V.; Pissas, M.; Perlepes, S.P.; Boudalis, A.K. A metamagnetic 2D copper (II)-azide complex with 1D ferromagnetism and a hysteretic spin-flop transition. Dalton Trans. 2009, 17, 3215–3221. [Google Scholar] [CrossRef]
  40. You, Y.S.; Yoon, J.H.; Kim, H.C.; Hong, C.S. Chiral azide-bridged two-dimensional Cu(II) compounds showing a field-induced spin–flop transition. Chem. Commun. 2005, 32, 4116–4118. [Google Scholar] [CrossRef]
  41. Gao, E.Q.; Bai, S.Q.; Wang, C.F.; Yue, Y.F.; Yan, C.H. Structural and magnetic properties of three one-dimensional azido-bridged copper(II) and manganese(II) coordination polymers. Inorg. Chem. 2003, 42, 8456–8464. [Google Scholar] [CrossRef]
  42. Sun, D.; Han, L.L.; Yuan, S.; Deng, Y.K.; Xu, M.Z.; Sun, D.F. Four new Cd(II) coordination polymers with mixed multidentate N-donors and biphenyl-based polycarboxylate ligands: Syntheses, structures, and photoluminescent properties. Cryst. Growth Des. 2013, 13, 377–385. [Google Scholar] [CrossRef]
  43. Wei, G.; Shen, Y.F.; Li, Y.R.; Huang, X.C. Synthesis, crystal structure, and photoluminescent properties of ternary Cd(II)/triazolate/chloride system. Inorg. Chem. 2010, 49, 9191–9199. [Google Scholar] [CrossRef] [PubMed]
  44. Du, M.; Jiang, X.J.; Zhao, X.J. Direction of unusual mixed-ligand metal–organic frameworks: A new type of 3-D polythreading involving 1-D and 2-D structural motifs and a 2-fold interpenetrating porous network. Chem. Commun. 2005, 44, 5521–5523. [Google Scholar] [CrossRef]
  45. Chen, F.; Wu, M.F.; Liu, G.N.; Wang, M.S.; Zheng, F.K.; Yang, C.; Xu, Z.N.; Liu, Z.F.; Guo, G.C.; Huang, J.S. Zinc(II) and cadmium(II) coordination polymers based on 3-(5H-Tetrazolyl)benzoate ligand with different coordination modes: Hydrothermal syntheses, crystal structures and ligand-centered luminescence. Eur. J. Inorg. Chem. 2010, 2010, 4982–4991. [Google Scholar] [CrossRef]
  46. Huang, F.P.; Zhang, Q.; Yu, Q.; Bian, H.D.; Liang, H.; Yan, S.P.; Liao, D.Z.; Cheng, P. Coordination assemblies of CoII/NiII/ZnII/CdII with succinic acid and bent connectors: Structural diversity and spin-canted antiferromagnetism. Cryst. Growth Des. 2012, 12, 1890–1898. [Google Scholar] [CrossRef]
  47. Liu, X.; Zhang, N.; Zhou, J.; Chang, T.; Fang, C.; Shangguan, D. A turn-on fluorescent sensor for zinc and cadmium ions based on perylene tetracarboxylic diimide. Analyst. 2013, 138, 901–906. [Google Scholar] [CrossRef]
  48. Majumdar, D.; Dey, S.; Sreejith, S.S.; Biswas, J.K.; Mondal, M.; Shukla, P.; Das, S.; Pal, T.; Das, D.; Bankura, K.; et al. Syntheses, crystal structures and photophysical aspects of azido-bridged tetranuclear cadmium(II) complexes: DFT/TD-DFT, thermal, antibacterial and anti-biofilm properties. J. Mol. Struct. 2019, 1179, 694–708. [Google Scholar] [CrossRef]
  49. Majumder, S.; Mandal, L.; Mohanta, S. Syntheses, structures, and steady state and time resolved photophysical properties of a tetraiminodiphenol macrocyclic ligand and its dinuclear zinc(II)/cadmium(II) complexes with coordinating and noncoordinating anions. Inorg. Chem. 2012, 51, 8739–8749. [Google Scholar] [CrossRef]
  50. 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]
  51. Ma, Y.; Cheng, A.L.; Tang, B.; Gao, E.Q. Copper(II) coordination polymers with azide and bipyridine-based zwitterionic carboxylate ligands: Structures and magnetism. Dalton Trans. 2014, 43, 13957–13964. [Google Scholar] [CrossRef]
  52. Sheldrick, G. SADABS, Program for Empirical Absorption Correction of Area Detector Data; University of Gottingen: Gottingen, Germany, 1996. [Google Scholar]
  53. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  54. Spackman, M.A.; Jayatilaka, D. Hirshfeld surface analysis. Cryst. Eng. Comm. 2009, 11, 19–32. [Google Scholar] [CrossRef]
  55. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. GAUSSIAN 09; Revision A02; Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  56. Glendening, E.D.; Reed, A.E.; Carpenter, J.E.; Weinhold, F. NBO Version 3.1, CI; University of Wisconsin: Madison, MI, USA, 1998. [Google Scholar]
  57. Adamo, C.; Barone, V. Exchange functionals with improved long-range behavior and adiabatic connection methods without adjustable parameters: The mPW and mPW1PW models. J. Chem. Phys. 1998, 108, 664–675. [Google Scholar] [CrossRef]
  58. Feller, D. The role of databases in support of computational chemistry calculations. J. Comp. Chem. 1996, 17, 1571–1586. [Google Scholar] [CrossRef]
  59. Schuchardt, K.L.; Didier, B.T.; Elsethagen, T.; Sun, L.; Gurumoorthi, V.; Chase, J.; Li, J.; Windus, T.L. Basis set exchange: A community database for computational sciences. J. Chem. Inf. Model. 2007, 47, 1045–1052. [Google Scholar] [CrossRef] [Green Version]
  60. Bader, R.F.W. Atoms in Molecules: A Quantum Theory; Oxford University Press: Oxford, UK, 1990. [Google Scholar]
  61. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comp. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  62. Goher, M.A.S.; Mautner, F.A.; Gatterer, K.; Abu-Youssef, M.A.M.; Badr, A.M.A.; Sodin, B.; Gspan, C. Four [Cd(L)2(N3)2]n 1D systems with different azide bridging sequences: Synthesis, spectral and structural characterization. J. Mol. Struct. 2008, 876, 199–205. [Google Scholar] [CrossRef]
  63. Mautner, F.A.; Scherzer, M.; Berger, C.; Fischer, R.C.; Vicente, R.; Massoud, S.S. Synthesis and characterization of three new 1-D polymeric [M2(4-azidopyridine)4(μ1,1-N3)2(μ1,3-N3)2]n (M = Ni, Co, Cd) complexes. Polyhedron 2015, 85, 329–336. [Google Scholar] [CrossRef]
  64. Louka, F.R.; Massoud, S.S.; Haq, T.K.; Koikawa, M.; Mikuriya, M.; Omote, M.; Fischer, R.C.; Mautner, F.A. Synthesis, structural characterization and magnetic properties of one-dimensional Cu(II)-azido coordination polymers. Polyhedron 2017, 138, 177–184. [Google Scholar] [CrossRef]
  65. Zhou, J.-H.; Cheng, R.-M.; Song, Y.; Li, Y.-Z.; Yu, Z.; Chen, X.-T.; You, X.-Z. Syntheses, structures and magnetic properties of two new water bridged dinuclear nickel(II) complexes containing derivatives of 1,2,4-triazole and pivalate ligands. Polyhedron 2006, 25, 2426–2432. [Google Scholar] [CrossRef]
  66. Soliman, S.M.; Albering, J.H.; Sholkamy, E.N.; El-Faham, A. Mono- and penta-nuclear self-assembled silver(I) complexes of pyrazolyl s-triazine ligand; synthesis, structure and antimicrobial studies. Appl. Organomet. Chem. 2020, 24, e5603. [Google Scholar] [CrossRef]
  67. Bohórquez, H.J.; Boyd, R.J.; Matta, C.F. Molecular model with quantum mechanical bonding information. J. Phys. Chem. A. 2011, 115, 12991–12997. [Google Scholar] [CrossRef] [PubMed]
  68. Bader, R.F. A quantum theory of molecular structure and its applications. Chem. Rev. 1991, 91, 893–928. [Google Scholar] [CrossRef]
  69. Bhadane, S.A.; Lande, D.N.; Gejji, S.P. Understanding binding of cyano-adamantyl derivatives to pillar [6] arene macrocycle from density functional theory. J. Phys. Chem. A 2016, 120, 8738–8749. [Google Scholar] [CrossRef]
  70. Marana, N.L.; Casassa, S.M.; Sambrano, J.R. Adsorption of NH3 with different coverages on single-walled ZnO nanotube: DFT and QTAIM study. J. Phys. Chem. C 2017, 121, 8109–8119. [Google Scholar] [CrossRef]
  71. Akman, F.; Issaoui, N.; Kazachenko, A.S. Intermolecular hydrogen bond interactions in the thiourea/water complexes (Thio-(H2O)n)(n = 1, …, 5): X-ray, DFT, NBO, AIM, and RDG analyses. J. Mol. Model. 2020, 26, 1–16. [Google Scholar] [CrossRef]
  72. Venkataramanan, N.S.; Suvitha, A.; Kawazoe, Y. Unravelling the nature of binding of cubane and substituted cubanes within cucurbiturils: A DFT and NCI study. J. Mol. Liq. 2018, 260, 18–29. [Google Scholar] [CrossRef]
  73. Nimmermark, A.; Öhrström, L.; Reedijk, J. Metal-ligand bond lengths and strengths: Are they correlated? A detailed CSD analysis. Z. Krist. 2013, 228, 311–317. [Google Scholar] [CrossRef] [Green Version]
  74. Reed, A.E.; Weinstock, R.B.; Weinhold, F. Natural population analysis. J. Chem. Phys. 1985, 83, 735–746. [Google Scholar] [CrossRef]
Scheme 1. Different bonding modes of the azide ion.
Scheme 1. Different bonding modes of the azide ion.
Symmetry 15 00619 sch001
Scheme 2. Structure of the ligands used.
Scheme 2. Structure of the ligands used.
Symmetry 15 00619 sch002
Figure 1. Structure of asymmetric unit of complexes 1 (A) and 2 (B).
Figure 1. Structure of asymmetric unit of complexes 1 (A) and 2 (B).
Symmetry 15 00619 g001
Figure 2. The coordination environment (A) and structure of the 1D polymer (B) of 1. Symmetry code for N1* is 2-x,2-y,-z. All H-atoms were omitted from the right part of this illustration for better clarity.
Figure 2. The coordination environment (A) and structure of the 1D polymer (B) of 1. Symmetry code for N1* is 2-x,2-y,-z. All H-atoms were omitted from the right part of this illustration for better clarity.
Symmetry 15 00619 g002
Figure 3. The hydrogen bond polymer in [Cd(2A4Pic)2(N3)2]n complex.
Figure 3. The hydrogen bond polymer in [Cd(2A4Pic)2(N3)2]n complex.
Symmetry 15 00619 g003
Figure 4. The coordination environment (A) and structure of the 1D polymer (B) of 2. All H-atoms were omitted from the right part of this illustration for better clarity.
Figure 4. The coordination environment (A) and structure of the 1D polymer (B) of 2. All H-atoms were omitted from the right part of this illustration for better clarity.
Symmetry 15 00619 g004
Figure 5. The hydrogen bond polymer in [Cu(Qu-6-COO)(N3)(H2O)]n complex.
Figure 5. The hydrogen bond polymer in [Cu(Qu-6-COO)(N3)(H2O)]n complex.
Symmetry 15 00619 g005
Figure 6. The π–π stacking interactions in [Cu(Qu-6-COO)(N3)(H2O)]n complex.
Figure 6. The π–π stacking interactions in [Cu(Qu-6-COO)(N3)(H2O)]n complex.
Symmetry 15 00619 g006
Figure 7. All intermolecular contacts in compounds 1 and 2.
Figure 7. All intermolecular contacts in compounds 1 and 2.
Symmetry 15 00619 g007
Figure 8. The overall dnorm map (right) and decomposed fingerprint plot (left) for the Cd-N coordination interactions in complex 1.
Figure 8. The overall dnorm map (right) and decomposed fingerprint plot (left) for the Cd-N coordination interactions in complex 1.
Symmetry 15 00619 g008
Figure 9. Hirshfeld dnorm surface (lower) and FP plot (upper) reveal the polymeric structure of complex 2 via Cu-N and Cu-O coordination interactions.
Figure 9. Hirshfeld dnorm surface (lower) and FP plot (upper) reveal the polymeric structure of complex 2 via Cu-N and Cu-O coordination interactions.
Symmetry 15 00619 g009
Table 1. Crystallographic details for 1 and 2.
Table 1. Crystallographic details for 1 and 2.
Complex12
Empirical formulaC12 H16 Cd N10C10 H8 Cu N4 O3
F.Wt412.75 g/mol295.74 g/mol
T100(2) K100(2) K
λ0.71073 Å0.71073 Å
Crystal systemMonoclinicMonoclinic
Space groupP21/cP21/c
Unit cell dimensions (Å, °) a = 8.9157(3)a = 8.0773(5)
b = 3.69130(10)b = 6.4187(3)
c = 23.1784(7)c = 20.9267(12)
β = 90.0980(10)β = 96.499(2)
V (Å3)762.81(4)1077.99(10)
Z24
ρcalc (g/cm3)1.7971.822
μ(Mo Kα) (mm−1)1.4482.032
Θ-range (°)2.88 to 26.432.54 to 23.35
No. Reflns703010351
Indep. reflns1543 [R(int) = 0.0248]1542 [R(int) = 0.0775]
%Completeness to Θ97.8098.80
GOOF (F2)1.0981.050
Final R [I > 2sigma(I)]R1 = 0.0166, wR2 = 0.0387R1 = 0.0367, wR2 = 0.0770
R (all data)R1 = 0.0190, wR2 = 0.0398R1 = 0.0721, wR2 = 0.0925
CCDC22070792207080
Table 2. The important distances (Å) and angles (°) for complex 1.
Table 2. The important distances (Å) and angles (°) for complex 1.
BondDistanceBondDistance
Cd1-N42.3305(15)Cd1-N1 12.4198(15)
Cd1-N12.3496(15)
BondsAngleBondsAngle
N1-Cd1-N1 2180.00(8)N4-Cd1-N1 192.79(5)
N41-Cd1-N4180N4-Cd1-N1 387.21(5)
N4-Cd1-N191.80(5)N1-Cd1-N1 178.59(5)
N4-Cd1-N1 288.20(5)N1-Cd1-N1 4101.41(5)
1 2−X,3−Y, −Z; 2 2−X,2−Y, −Z; 3 +X, −1+Y,+Z; and 4 +X,1+Y,+Z.
Table 3. The important distances (Å) and angles (°) for complex 2.
Table 3. The important distances (Å) and angles (°) for complex 2.
BondDistanceBondDistance
Cu1-O21.945(3)Cu1-N1 11.995(4)
Cu1-O31.946(3)Cu1-O12.374(4)
Cu1-N11.988(4)Cu1 2-O12.771(3)
BondsAngleBondsAngle
O2-Cu1-O3178.22(14)N1-Cu1-N1 1166.14(12)
O2-Cu1-N186.97(15)O2-Cu1-O183.93(13)
O3-Cu1-N191.46(15)O3-Cu1-O195.69(13)
O2-Cu1-N1 191.34(15)N1-Cu1-O1105.02(15)
O3-Cu1-N1 190.39(15)N11-Cu1-O188.46(15)
1 2−X,1/2+Y,1/2-Z; 2 2−X, −1/2+Y,1/2−Z.
Table 4. The short interactions and their distances in complex 2.
Table 4. The short interactions and their distances in complex 2.
ContactDistanceContactDistance
C10…C33.327(7)N4…H21.827
C6…C83.320(7)N2…H102.586
C5…C83.396(7)N3…H12.121
Cu1-N11.995(4)N2…O22.828(5)
Cu1-O21.945(3)N2…O32.857(5)
Cu1-O12.771(3)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Altowyan, M.S.; Fathalla, E.M.; Albering, J.H.; Barakat, A.; Abu-Youssef, M.A.M.; Soliman, S.M.; Badr, A.M.A. Synthesis, Molecular, and Supramolecular Structures of Two Azide-Bridged Cd(II) and Cu(II) Coordination Polymers. Symmetry 2023, 15, 619. https://doi.org/10.3390/sym15030619

AMA Style

Altowyan MS, Fathalla EM, Albering JH, Barakat A, Abu-Youssef MAM, Soliman SM, Badr AMA. Synthesis, Molecular, and Supramolecular Structures of Two Azide-Bridged Cd(II) and Cu(II) Coordination Polymers. Symmetry. 2023; 15(3):619. https://doi.org/10.3390/sym15030619

Chicago/Turabian Style

Altowyan, Mezna Saleh, Eman M. Fathalla, Jörg H. Albering, Assem Barakat, Morsy A. M. Abu-Youssef, Saied M. Soliman, and Ahmed M. A. Badr. 2023. "Synthesis, Molecular, and Supramolecular Structures of Two Azide-Bridged Cd(II) and Cu(II) Coordination Polymers" Symmetry 15, no. 3: 619. https://doi.org/10.3390/sym15030619

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