Next Article in Journal
Biomedical Importance of Indoles
Previous Article in Journal
Synthesis and Characterization of Rice Straw/Fe3O4 Nanocomposites by a Quick Precipitation Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Self-Assembly of a 1D Hydrogen-Bonded Polymer from a Hexamethyltetraaza Macrocyclic Nickel(II) Complex and Isophthalic Acid

Department of Chemistry Education, Kongju National University, Kongju 314-701, Korea
*
Author to whom correspondence should be addressed.
Molecules 2013, 18(6), 6608-6619; https://doi.org/10.3390/molecules18066608
Submission received: 3 May 2013 / Revised: 31 May 2013 / Accepted: 3 June 2013 / Published: 6 June 2013
(This article belongs to the Section Molecular Diversity)

Abstract

:
The compound [Ni(L)(isoph)2][Ni(L)]·8H2O (1; L = C-meso-5,5,7,12,12,14-hexamethyl-1,4,8,11-tetraazacyclotetradecane; H2-isoph = isophthalic acid) has been synthesized and structurally characterized. Complex 1 exhibits a geometrically symmetric core with a {4/6} coordination number set. The coordination environment around the Ni(1) ion is a distorted octahedron, while the geometry around the four-coordinate Ni(2) is depicted as square planar in 1D hydrogen-bonded infinite chain. The compound crystallizes in the triclinic system P-1 with a = 8.602(2), b = 10.684(7), c = 16.550(3) Å, α = 91.04(4), β = 94.09(2), γ = 111.09(4)°, V = 1413.9(10) Å3, Z = 1. The cyclic voltammogram of 1 undergoes one-electron wave corresponding to NiII/NiI process. The electronic spectra, electrochemical and TGA behavior of the complex are significantly affected by the nature of the hexamethyltetraaza macrocycle and the axial isoph2− ligand.

1. Introduction

There has been considerable interest in macrocycles with C-alkyl groups on a polyaza macrocyclic ring and their metal complexes because of their structural and chemical properties, which are often quite different from those of the corresponding unalkylated macrocyclic ligands [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18]. The application of polyaza macrocycle precursors in the synthesis of transition metal(II) macrocyclic complexes, stems mainly from their use as models for protein-metal binding sites in biological systems [19], and as selective complexing agents for metal ions [20]. The structures and chemical properties of such complexes are affected by various factors, such as C-alkyl groups in the macrocycle and the type and synthetic differences between the axial ligands. For example, the compounds [Ni(cyclam)(NCO)(H2O)](ClO4) (cyclam = 1,4,8,11-tetraazacyclotetradecane) [8] and [Ni(dttd)(NCO)2] (dttd = 3,14-dimethyl-2,6,13,17-tetraazatricyclo[14,4,01.18,07.12]docosane) [9] exhibit a distorted octahedral geometry, in which the nickel(II) ion is coordinated by four secondary amines and two nitrogen atoms of the isocynate ligand. However, the complex [Ni(Me4cyclam)(NCO)](ClO4) (Me4cyclam = 1,4,8,11-tetramethyl-1,4,8,11-tetraazacyclotetradecane) [8] reveals a distorted square pyramidal geometry, which may be due to the steric constraints of the Me4cyclam ligand. On the other hand, the octahedral compound [Ni(L)(OCN)1.5(ClO4)0.5] (L = C-meso-5,5,7,12,12,14-hexamethyl-1,4,8,11-tetraazacyclotetradecane) [8] presents the peculiarity of the cyanato coordination mode instead of the habitual isocyanate mode. In contrast, the compound [{Ni(L1)}2(μ-OCN)2](ClO4)2 (L1 = dl-5,5,7,12,12,14-hexamethyl-1,4,8,11-tetraazacyclotetradecane) [10] shows that the cyanate ligand acts as an end-to-end bridge with nickel(II) ions in an NiN5O octahedral environment. The same synthetic strategy with the azido group leads end-to-end coordination, giving a one-dimensional compound [{Ni(L)(μ-N3)}n][ClO4]n [11]. The different molecular topologies in the complexes may be due to the C-alkyl groups into the macrocycle and different coordination modes of the ligands.
In the present work, we report the preparation and characterization of [Ni(L)(isoph)2][Ni(L)]·8H2O (1; L = C-meso -5,5,7,12,12,14-hexamethyl-1,4,8,11-tetraazacyclotetradecane; H2-isoph = isophthalic acid, Figure 1). To evaluate the effect of the hexamethyl tetraaza macrocycle and anionic group, we have investigated the X-ray crystal structure, spectroscopic, and electrochemical properties of the complex.
Figure 1. Structure of macrocycle and isophthalate ligand.
Figure 1. Structure of macrocycle and isophthalate ligand.
Molecules 18 06608 g001

2. Results and Discussion

2.1. Description of the Structure

An ORTEP drawing [21] of [Ni(L)(isoph)2][Ni(L)]·8H2O (1) with the atomic numbering scheme is shown in Figure 2. The selected bond lengths and angles are listed in Table 1. The macrocyclic ligand skeleton of the complex takes the most stable trans-III(R,R,S,S) configuration with two chair six-membered and two gauche five-membered chelate rings. The complex has a geometrically symmetric core with a {4/6} coordination number set. In both structures, the nickel atoms are located at inversion centers. The geometry around the four-coordinate Ni(2) is depicted as square planar with four secondary amines of the macrocycle, while the coordination environment around the Ni(1) ion is a distorted octahedron in which nickel ion is coordinated by four secondary amines of the macrocycle and two carboxylate oxygen atoms of the axial isoph2− ligands. Interestingly, the complex 1 has different topology compared with the compound {[Ni(dttd)(isoph)]·3H2O}n [18], which exhibits a 1D coordination polymer with a tetragonally distorted octahedral geometry. This fact may be due to the six methyl groups in the macrocycle, even though the complex 1 has the same axial ligand as the compound {[Ni(dttd)(isoph)]·3H2O}n. The average Ni(1)-N bond distance of 2.070 Å is typical for those of octahedral nickel(II) complexes with 14-membered tetraaza macrocyclic ligands [6,7,8,9,10]. The Ni(1)-O distance of 2.056(6) Å is shorter than that observed in the related system ({[Ni(dttd)(tp)]·2H2O}n [17]: Ni-O = 2.116(3) Å, where tp2− = terephthalate), which has a tetragonally distorted octahedral geometry. The average Ni(2)-N bond distance of 1.896 Å is in the normal range for square-planar nickel(II) complexes with 14-membered tetraaza macrocycles [11,12,13,14,15,16]. The N-Ni-N angles of the six-membered chelate rings are larger than those of the five-membered chelate rings. The intermolecular Ni(1)…Ni(2) distance is 11.641(8) Å, whereas the closest intermolecular Ni(1)…Ni(2)i distance between neighboring strands is 9.607(4) Å. The axial Ni(1)-O(1) bonds are not perpendicular to the NiN4 plane as the Oaxial-Ni-Nbasal angles ranging from 85.6(3) to 94.4(3)°. The dihedral angles (α) between the plane of the carboxylate group and benzene ring involving Ni(1) atom are 2.8(2) and 10.5(9)°, respectively. Moreover, the dihedral angle (β) between the plane of the carboxylate group and NiN2O4 plane involving Ni(1) atom is 14.6(9)°. The Ni(1)-O(1)-C(9) angle and C(9)-O(1) distance relative to the isophthalate ligand are 133.7(6)° and 1.254(11) Å, respectively. The deprotonated oxygen O(1) among the two isoph2− carboxylic groups is coordinated to the metal center. The secondary amine N(1) of the macrocycle is intramolecular hydrogen bonded to the uncoordinated carboxylic oxygen O(2) of the isoph2− ligand [N(1)-H(1)…O(2)iii 2.804(10) Å, 154.6°; symmetry code (iii) −x, −y+1, −z+1]. Interestingly, the deprotonated oxygen atoms O(3) and O(4) of the isoph2− ligand form intermolecular hydrogen bonds to an adjacent secondary amines N(3) and N(4) of the macrocycle involving Ni(2) atom [N(3)-H(3)…O(4) 2.869(10) Å, 164.0°; N(4)-H(4)…O(3) 2.841(11) Å, 157.8°]. Furthermore, the water molecules also forms intermolecular hydrogen bonds with uncoordinated carboxylate oxygen atoms O(2) and O(4) of the isoph2− ligand [Ow(1)-HO(1)…O(4) 2.676(11)Å, 160(9)°; Ow(2)-HO(3)…O(2) 2.734(12) Å, 163(9)°] and the other water inclusions [Ow(2)-HO(4)…Ow(1) 2.706(14) Å, 147(11)°; Ow(4)-HO(8)…Ow(1)iv 2.829(17) Å, 145(13); symmetry code (iv) −x + 1, −y, −z]. This interaction gives rise to a 1D hydrogen-bonded polymer (Table 2 and Figure 3).
Figure 2. An ORTEP diagram of [Ni(L)(isoph)2][Ni(L)]∙8H2O (1) with the atom-numbering scheme (30% probability ellipsoids shown). The hydrogen atoms other than those participating in hydrogen bonding are omitted for clarity. Symmetry code: (iii) −x, −y+1, −z+1 (iv) −x+1, −y, −z.
Figure 2. An ORTEP diagram of [Ni(L)(isoph)2][Ni(L)]∙8H2O (1) with the atom-numbering scheme (30% probability ellipsoids shown). The hydrogen atoms other than those participating in hydrogen bonding are omitted for clarity. Symmetry code: (iii) −x, −y+1, −z+1 (iv) −x+1, −y, −z.
Molecules 18 06608 g002
Table 1. Selected bond lengths (Å) and angles (°).
Table 1. Selected bond lengths (Å) and angles (°).
Bond lengths and angles
Bond lengths
Ni(1)-N(1)2.130(7)Ni(1)-N(2)2.009(7)
Ni(1)-O(1)2.056(6)Ni(2)-N(3)1.913(7)
Ni(2)-N(4)1.878(8)C(9)-O(1)1.254(11)
C(9)-O(1)1.238(11)C(16)-O(3)1.196(12)
C(16)-O(4)1.239(12)Ni(1)…Ni(2)11.641(8)
Ni(1)…Ni(2)i9.607(4)
Bond angles
N(1)-Ni(1)-N(2)87.0(3)N(1)-Ni(1)-N(2)iii93.0(3)
N(1)-Ni(1)-O(1)89.7(3)N(1)iii-Ni(1)-O(1)90.3(3)
N(2)-Ni(1)-O(1)85.6(3)N(2)iii-Ni(1)-O(1)94.4(3)
N(3)-Ni(2)-N(4)91.3(4)N(3)-Ni(2)-N(4)iv88.7(4)
Ni(1)-O(1)-C(9)133.7(6)O(1)-C(9)-O(2)127.7(8)
O(3)-C(16)-O(4)122.6(9)
Symmetry codes: i x−1, y, z; iii −x, −y+1, −z+1; iv −x+1, −y, −z.
Table 2. Hydrogen bonding parameters (Å, °).
Table 2. Hydrogen bonding parameters (Å, °).
D-HAD-H (Å)HA (Å)DA (Å)D-HA (°)
N(1)-H(1)O(2)iii0.911.952.804(10)154.6
N(3)-H(3)O(4)0.911.982.869(10)164.0
N(4)-H(4)O(3)0.911.982.841(11)157.8
Ow(1)-HO(1)O(4)1.07(10)1.65(11)2.676(11)160(9)
Ow(2)-HO(3)O(2)1.00(11)1.76(11)2.734(12)163(9)
Ow(2)-HO(4)Ow(1)0.89(13)1.92(13)2.706(14)147(11)
Ow(4)-HO(8)Ow(1)iv0.82(13)2.12(14)2.829(17)145(13)
Symmetry codes: iii −x, −y+1, −z+1; vi −x+1, −y, −z.
Figure 3. Crystal packing diagram of [Ni(L)(isoph)2][Ni(L)]∙8H2O (1), showing the intermolecular hydrogen bonds as dashed lines. The hydrogen atoms other than those participating in hydrogen bonding are omitted for clarity.
Figure 3. Crystal packing diagram of [Ni(L)(isoph)2][Ni(L)]∙8H2O (1), showing the intermolecular hydrogen bonds as dashed lines. The hydrogen atoms other than those participating in hydrogen bonding are omitted for clarity.
Molecules 18 06608 g003

2.2. Chemical Properties

The IR spectrum of 1 shows a band at 3148 cm−1 corresponding to the ν(NH) of the coordinated secondary amines of the macrocycle. Two strong bands exhibit νas(COO) stretching frequency at 1606 cm−1 and νsym(COO) at 1357 cm−1, respectively. The value of Δν (249 cm−1) indicates that the carboxylate groups are coordinated to the nickel(II) ion only as monodentate ligands [22,23]. In addition, a sharp band at 3420 cm−1 is associated to the ν(OH) stretching vibration of the water molecule. The UV-Vis spectral data of 1 are listed in Table 3. The solid reflectance spectrum of 1 is also shown in Figure 4. The UV spectrum of 1 in the water solution shows an absorption maximum in the region 260 nm attributed to a ligand-metal charge transfer associated with the nitrogen and oxygen donors [23]. As shown in Figure 4, the solid state electronic spectrum of 1 in the visible region shows three absorption bands at 325(sh), 474, and 696 nm assignable to the 3B1g3Eg(T1g(P)), 1B2g1B1g, 3B1g3B2g(T2g(F)) transitions [3,24]. The transitions at 325 and 696 nm arise from the tetragonal complex while the transition at 474 nm results from the square planar species. However, the visible spectrum of 1 in water solution displays a broad band 458 nm, which has a low-spin d8 nickel(II) ion in a square-planar environment [Ni(L)](ClO4)2 (463nm) [5]. This fact can be understood in terms of the decomposition of the hydrogen bonding or the building block in water solution. The electronic spectrum for 1 clearly support the structure determined by the X-ray diffraction study.
Table 3. Electronic spectral data a.
Table 3. Electronic spectral data a.
ComplexStateλmax/nm (ε/M−1 cm−1)
[Ni(L)](ClO4)2 bH2O463(80)
CH3CN469(63)
Solid262, 468
[Ni(L)(isoph)2][Ni(L)]∙8H2O (1)H2O260(5.2 × 103), 458(71)
Solid325(sh), 474, 696
a Solution = H2O at 20 ± 0.1 °C; Solid = diffuse reflectance; b Ref. [5].
Figure 4. Solid state electronic absorption spectrum of [Ni(L)(isoph)2][Ni(L)]∙8H2O (1) by the diffuse reflectance method at 20.0 ° 0.1 °C.
Figure 4. Solid state electronic absorption spectrum of [Ni(L)(isoph)2][Ni(L)]∙8H2O (1) by the diffuse reflectance method at 20.0 ° 0.1 °C.
Molecules 18 06608 g004
Cyclic voltammetric data for 1 in 0.10 M TEAP-DMSO solution are given in Table 4. Cyclic voltammogram of 1 in 0.1 M TEAP-DMSO solution is shown in Figure 5. The H2-isoph ligand is electrochemically active in the range of potential studied. Cyclic voltammetry of H2-isoph shows two irreversible one-electron reductions at −1.33 and −1.61 V vs. the Ag/AgCl reference electrode. The reduction potential for 1 gives the irreversible one-electron process at −0.76 V vs. the Ag/AgCl refernce electrode, assigned to the NiII/NiI process. The reduction potentials for 1 are considerably more positive than that for the square-planar [Ni(L)](ClO4)2 (−1.29V) [5]. This fact may be due to the coordination of the axial isoph2− ligand, which is in agreement with the crystal structure of 1.
Table 4. Cyclic voltametric dataa.
Table 4. Cyclic voltametric dataa.
ComplexPotentials(V) Ag/AgCl
Ni(II)/Ni(I)
[Ni(L)](ClO4)2 b,a−1.29
[Ni(L)(isoph)2][Ni(L)]∙8H2O (1)−0.76(i) c
a Measured in 0.10 M TEAP-DMSO solution at 20.0 ± 0.1°C; b Ref. [5]; c i = irreversible.
Figure 5. Cyclic voltammograms of [Ni(L)](ClO4)2 and [Ni(L)(isoph)2][Ni(L)]∙8H2O (1) in 0.1 M TEAP-DMSO solution at 20.0 ± 0.1 °C. The scan rate is 100 mV/s.
Figure 5. Cyclic voltammograms of [Ni(L)](ClO4)2 and [Ni(L)(isoph)2][Ni(L)]∙8H2O (1) in 0.1 M TEAP-DMSO solution at 20.0 ± 0.1 °C. The scan rate is 100 mV/s.
Molecules 18 06608 g005
The TGA diagram of 1 is shown in Figure 6. The compound was heated in the temperature range 50–1,000 °C in nitrogen gas. The first weight loss is observed from 30 to 139 °C, which is due to the loss of six water molecules (observed 8.1%, calculated 9.3%).
Figure 6. Thermogravimetric curve of [Ni(L)(isoph)2][Ni(L)]∙8H2O (1).
Figure 6. Thermogravimetric curve of [Ni(L)(isoph)2][Ni(L)]∙8H2O (1).
Molecules 18 06608 g006
A second weight loss corresponding to the macrocycle (observed 51.8%, calculated 49.1%) is found in the temperature range 382–405 °C. On further heating, 1 lost weight between 405 and 756 °C corresponding to two isophthalate ligands (observed 26.7%, calculated 28.3%). Further weigh loss is not observed in the temperature range 756 to 1000 °C corresponding to the greenish black residue of NiO (observed 13.3%, calculated 12.9%).

3. Experimental

3.1. Materials and Physical Measurements

The chemicals and solvent NiCl2·6H2O (99.99%), isophthalic acid (99%), and acetonitrile (99.93%) were purchased from Aldrich (Milwaukee, WI, USA) and used without futher purification. The C-meso-5,5,7,12,12,14-hexamethyl-1,4,8,11-tetraazacyclotetradecane ligand (L) was prepared according to literature method [4,5]. IR spectra were recorded with a Jasco FT/IR-4100 spectrophotometer using KBr pellets. Electrochemical measurements were accomplished with a three electrode potentiostat BAS-100BW system. A 3-mm Pt disk was used as the working electrode. The counter electrode was a coiled Pt wire and an Ag/AgCl electrode was used as a reference electrode. Cyclic voltammetric data were obtained in DMSO solution using 0.10 M tetraethylammonium perchlorate (TEAP) as supporting electrolyte at 20.0 ± 0.1 °C. The solution was degassed with high purity N2 prior to carrying out the electrochemical measurements. The DSC and TGA were performed under flowing nitrogen at a heating rate of 10 °C mim−1 using a SDT 2960 Thermogravimetric Analyzer. Elemental analyses (C, H, N) were performed on a Perkin-Elmer CHN-2400 analyzer.

3.2. Synthesis of [Ni(L)(isoph)2][Ni(L)]·8H2O (1)

To a methanol solution of (20 mL) of NiCl2·6H2O (119 mg, 0.5 mmol) was added L (142 mg, 0.5 mmol) and sodium isophthalate (Na2isoph) (105 mg, 0.5 mmol) and the mixture was stirred for 1 hr at room temperature. The solution was filtered and left at room temperature until the violet crystals formed. The product was recrystallized from a hot water/acetonitrile (1:1 v/v, 10 mL) mixture. Yield: 56%. Calc. (found) for C48H96N8Ni2O16: C, 49.76 (49.62); H, 8.35 (8.46); N, 9.67 (9.75)%. IR (KBr, cm−1): 3,420 (s), 3269 (m), 3,148 (m), 3,068 (m), 2,962 (m), 1,606 (s), 1,561 (s), 1,460 (w), 1,422 (w), 1,373 (s), 1,357 (s), 1,304 (w), 1,274 (w), 1,186 (m), 1,156 (m), 1,095 (m), 1,073 (m), 1,053 (m), 997 (w), 958 (w), 928 (w), 810 (w), 743 (m), 710 (m), 692 (w), 641 (w).

3.3. X-ray Crystallography

Single crystal X-ray diffraction measurement for 1 was carried out on an Enraf-Nonius CAD4 diffractometer using graphite-monochromated Mo-Kα radiation (λ = 0.71073 Å). Intensity data were measured at 293(2) K by ω-2θ technique. Accurate cell parameters and an orientation matrix were determined by the least-squares fit of 25 reflections. The intensity data were corrected for Lorentz and polarization effects. Empirical absorption correction was carried out using ψ-scan [25]. The structure was solved by direct methods and the least-squares refinement of the structure was performed by the SHELXL-97 program [26]. All the non-hydrogen atoms were refinded anisotropically. The hydrogen atoms were placed in calculated positions, allowing them to ride on their parent C and N atoms with Uiso(H) = 1.2Ueq(C or N). A summary of the data collections and details of the structure refinement is given in Table 5.
Crystallographic data for the structural analysis have been deposited with the Cambridge Crystallographic Data Center, CCDC No. 846954 for 1. Copies of this information may be obtained free of charge from the Director, CCDC, 12 Union Road, Cambridge, CB2, 1EZ, UK (fax: +44-1223-336033; e-mail: [email protected] or http://www.ccdc.cam.ac.uk).
Table 5. Crystallographic data.
Table 5. Crystallographic data.
Data and refinement
Empirical formulaC48H96N8Ni2O16
Formula weight1157.94
Temperature (K)293(2)
Crystal color/habitBrown/prism
Crystal systemTriclinic
Space groupP-1
Unit cell dimensions
a (Å)8.602(2)
b (Å)10.684(7)
c (Å)16.550(3)
α (°)91.04(4)
β (°)94.09(2)
γ (°)111.09(4)
V3)1413.9(10)
Z1
Dcalc (Mg m−3)1.360
Absorption coefficient (mm−1)0.738
F(000)624
Crystal size (mm3)0.40 × 0.30 × 0.10
θ range (°)1.23 to 26.37
Limiting indices−10 ≤ h ≤ 10, −13 ≤ k≤ 13, −1 ≤ l≤ 19
Reflection collected/unique5729/5137 (Rint = 0.0817)
Absorption correctionψ-scan
Max./min. transmission0.9145 and 0.7122
Data/restraints/parameters5137/0/369
Goodness of fit on F21.184
Final R indices (I>2σ(I))R1a = 0.0811, wR2b = 0.2224
R indices (all data)R1 = 0.1742, wR2 = 0.3391
Weighting schemew = 1/[σ2(Fo2) + (0.20000P)2 +0.0000P]
rgest difference peak and hole (eÅ−3)with P = (Fo2 + 2Fc2)/3
1.731 and −1.199
a R1 = Σ|| Fo|−| Fc||/Σ| Fo|; b wR2 = [Σ[w(Fo2-Fc2)2]/Σ[w(Fo2)2]]1/2.

4. Conclusions

Complex 1 reveals a geometrically symmetric core with a {4/6} coordination number set. The coordination environment around the Ni(1) ion is a distorted octahedron, while the geometry around the four-coordinate Ni(2) is depicted as square planar in 1D hydrogen-bonded infinite chain. The solid state electronic spectrum of 1 in the visible region shows three absorption bands assignable to the tetragonal and the square planar species. The cyclic voltammogram of 1 displays a one-electron wave corresponding to the NiII/NiI redox process. The reduction potentials for 1 are considerably more positive than that for the square-planar [Ni(L)](ClO4)2. The TGA behavior of the complex 1 is significantly affected by the nature of the hexamethyltetraaza macrocycle and the axial isoph2− ligand.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/18/6/6608/s1.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hay, R.W.; Pujari, M.P. The preparation of cis-folded macrocyclic nickel(II) complexes and the kinetics of their base-catalysed conversion to planar species. J. Chem. Soc. Dalton Trans. 1986, 1485–1489. [Google Scholar]
  2. Black, D.G.; Jordan, R.F.; Rogers, R.D. Structural Trends in Group 4 Metal Tetraaza Macrocycle Complexes. Molecular Structures of (Me4taen)Zr(OtBu)2 and (Me4taen)Hf(NMe2)2. Inorg. Chem. 1997, 36, 103–108. [Google Scholar]
  3. Misra, T.K.; Chung, C.S.; Cheng, J.; Lu, T.H. Isolation and structural characterization of trans-β-{(C-meso-5,5,7,12,12,14-hexamethyl-1,4,8,11-tetraazacyclotetradecane)(diaquo)}nickel(II) perchlorate and cis-α-{(C-racemic-5,5,7,12,12,14-hexamethyl-1,4,8,11-tetraazacyclotetradecane) (diisothiocyanato)}nickel(II). Polyhedron 2001, 20, 3149–3154. [Google Scholar]
  4. Curtis, N.F. Some cyclic tetra-amines and their metal-ion complexes. Part I. Two isomeric hexamethyltetra-azacyclotetradecanes and their copper(II) and nickel(II) complexes. J. Chem. Soc. 1964, 2644–2650. [Google Scholar]
  5. Warner, L.G.; Busch, D.H. Stereoisomers of a macrocyclic nickel(II) complex containing six asymmetric centers. Factors determining the stabilities of configurations and conformations. J. Am. Chem. Soc. 1969, 91, 4092–4101. [Google Scholar]
  6. Escuer, A.; Vicente, R.; Salah El Fallah, M.; Solans, X.; Font-Bardia, M. Synthesis, structure and magnetic properties of several nickel-azido complexes with tetraamines as blocking ligands. Inorg. Chim. Acta 1996, 247, 85–91. [Google Scholar]
  7. Escuer, A.; Vicente, R.; Ribas, J.; Salah J Fallah, M.; Solans, X. The first 1D nickel(II) complex with a single azido bridge: structure and magnetic behavior of catena-(μ-N3)[Ni(1,4,8,11-tetraazacyclotetradecane)](ClO4)•H2O. Inorg. Chem. 1993, 32, 1033–1035. [Google Scholar]
  8. Vicente, R.; Escuer, A.; Salah El Fallah, M.; Solans, X.; Font-Bardia, M. Three new mononuclear nickel(II) cyanate and isocyanate compounds derived from macrocyclic ligands: [Ni(TMCY) (NCO)](ClO4), [Ni(m-CTH)(OCN)1.5(ClO4)0.5]. Inorg. Chim. Acta 1997, 261, 227–237. [Google Scholar]
  9. Choi, K.Y.; Kim, Y.J.; Ryu, H.; Suh, I.H. Synthesis and characterization of nickel(II) complexes of a tetraaza macrocycle containing axial ligands. Inorg. Chem. Commun. 1999, 2, 176–180. [Google Scholar]
  10. Escuer, A.; Vicente, R.; Salah El Fallah, M.; Solans, X.; Font-Bardia, M. Structure and magnetic behaviour of the first singly bridged nickel cyanate chain and a new dinuclear complex: an approximation to the superexchange mechanism for the nickel pseudohalide system. J. Chem. Soc. Dalton Trans. 1996, 1013–1019. [Google Scholar] [CrossRef]
  11. Escuer, A.; Vicente, R.; Salah El Fallah, M.; Ribas, J.; Solans, X.; Font-Bardia, M. Unusual magnetic and structural one-dimensional nickel(II) azido systems. J. Chem. Soc. Dalton Trans. 1993, 2975–2976. [Google Scholar] [CrossRef]
  12. Hay, R.W.; Jeragh, B.; Ferguson, G.; Kaither, B.; Ruhl, B.L. Axial additions in nickel(II) complexes of (5SR,7RS,12RS,14SR)-tetra-methyl-1,4,8,11-tetra-azacyclotetradecane (Lα) and crystal-structure analysis of the planar orange and octahedral violet perchlorate salts NiLα,(ClO4)2. J. Chem. Soc. Dalton Trans. 1982, 1531–1539. [Google Scholar] [CrossRef]
  13. Hay, R.W.; Danby, A.; Lighfoot, P.; Lampeka, Y.D. Padlock macrocyclic complexes. The synthesis of a range of nickel(II) complexes of N-alkyl azacyclams and the crystal structure of (3-ethyl-1,3,5,8,12-penta-azacyclotetradecane) nickel(II) perchlorate. Polyhedron 1997, 16, 2777–2783. [Google Scholar] [CrossRef]
  14. Hay, R.W.; Crayston, J.A.; Cromie, T.J.; Lighfoot, P.; de Alwis, D.C.L. The preparation, chemistry and crystal structure of the nickel(II) complex of N-hydroxyethylazacyclam [3-(2′-hydoxyethyl)-1,3,5,8,12-pentaazacyclotetradecane nickel(II) perchlorate]. A new electrocatalyst for CO2 reduction. Polyhedron 1997, 16, 3557–3563. [Google Scholar] [CrossRef]
  15. Choi, K.Y.; Kim, J.C.; Jensen, W.P.; Suh, I.H.; Choi, S.-S. Copper(II) and Nickel(II) Complexes with 3,14-Dimethyl-2,6,13,17-tetraazatricyclo[16.4.0.07,12]docosane, [M(C20H40N4)]Cl2 • 2H2O (M = CuII and NiII). Acta Crystallogr. 1996, C52, 2166–2168. [Google Scholar]
  16. Curtis, N.F.; Wikaira, J. Nickel(II) compounds of a tri-amine mono-imine macrocycle: Preparations and structures of (5,7,7,12,14,14-hexamethyl-1,4,8,11-tetraazacyclotetradec-4-ene)nickel(II) compounds. Polyhedron 2011, 30, 895–902. [Google Scholar] [CrossRef]
  17. Choi, K.Y.; Chun, K.M.; Suh, I.H. Synthesis and characterization of one-dimensional nickel(II) macrocyclic complexes with bridging organic ligands. Polyhedron 2001, 20, 57–65. [Google Scholar] [CrossRef]
  18. Choi, K.Y.; Kim, K.J. Self-assembly of 1D coordination polymers from nickel(II) tetraaza macrocycle and carboxylate ligands. Polyhedron 2008, 27, 1310–1316. [Google Scholar] [CrossRef]
  19. Fenton, D.D.; Okawa, H. Dalton perspective. The emergence of trinuclear constellations at metallobiosites. J. Chem. Soc. Dalton Trans. 1993, 1349–1357. [Google Scholar] [CrossRef]
  20. Tsukube, H.; Yoden, T.; Iwachido, T.; Zenki, M. Lipid-bound macrocycles as new immobilized ligand systems for effective separation of metal cations. J. Chem. Soc. Chem. Commun. 1991, 1069–1070. [Google Scholar] [CrossRef]
  21. Farrugia, L.J. ORTEP-3 for windows - a version of ORTEP-III with a graphical user interface (GUI). J. Appl. Crystallogr. 1997, 30, 565. [Google Scholar] [CrossRef]
  22. Bakalbassis, E.G.; Tsipis, C.A.; Bozopoulos, A.P.; Dreissing, D.W.; Hartl, H.; Mrozinski, J. Strong ferromagnetism between copper(II) ions separated by 6.7.ANG. in a new phthalato-bridged copper(II) binuclear complex. Inorg. Chem. 1991, 30, 2801–2806. [Google Scholar]
  23. Cao, R.; Shi, Q.; Sun, D.; Hong, M.; Bi, W.; Zhao, Y. Syntheses and characterizations of copper(II) polymeric complexes constructed from 1,2,4,5-benzenetetracarboxylic acid. Inorg. Chem. 2002, 41, 6161–6168. [Google Scholar] [CrossRef] [PubMed]
  24. Vigee, G.S.; Watkins, C.L.; Bowen, H.F. The chemical dynamics of 1,4,8,11-tetraazacyclotetradecane nickel(II) perchlorate with several solvents. I. Thermodynamics of base adduct formation. Inorg. Chim. Acta 1979, 35, 255–259. [Google Scholar] [CrossRef]
  25. North, A.C.T.; Phillips, D.C.; Mathews, F.S. A semi-empirical method of absorption correction. Acta Crystallogr. 1968, A24, 351–359. [Google Scholar] [CrossRef]
  26. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. 2008, A64, 112–122. [Google Scholar] [CrossRef] [PubMed]
  • Sample Availability: Sample of the [Ni(L)(isoph)2][Ni(L)]·8H2O is available from the authors.

Share and Cite

MDPI and ACS Style

Lim, I.-T.; Choi, K.-Y. Self-Assembly of a 1D Hydrogen-Bonded Polymer from a Hexamethyltetraaza Macrocyclic Nickel(II) Complex and Isophthalic Acid. Molecules 2013, 18, 6608-6619. https://doi.org/10.3390/molecules18066608

AMA Style

Lim I-T, Choi K-Y. Self-Assembly of a 1D Hydrogen-Bonded Polymer from a Hexamethyltetraaza Macrocyclic Nickel(II) Complex and Isophthalic Acid. Molecules. 2013; 18(6):6608-6619. https://doi.org/10.3390/molecules18066608

Chicago/Turabian Style

Lim, In-Taek, and Ki-Young Choi. 2013. "Self-Assembly of a 1D Hydrogen-Bonded Polymer from a Hexamethyltetraaza Macrocyclic Nickel(II) Complex and Isophthalic Acid" Molecules 18, no. 6: 6608-6619. https://doi.org/10.3390/molecules18066608

Article Metrics

Back to TopTop