Next Article in Journal
Phenolic Profile of Croton urucurana Baill. Leaves, Stems and Bark: Pairwise Influence of Drying Temperature and Extraction Solvent
Next Article in Special Issue
Interaction of Ampicillin and Amoxicillin with Mn2+: A Speciation Study in Aqueous Solution
Previous Article in Journal
Synthesis, Biological Evaluation and In Silico Studies of Certain Oxindole–Indole Conjugates as Anticancer CDK Inhibitors
Previous Article in Special Issue
Estimation of the Chelating Ability of an Extract from Aronia melanocarpa L. Berries and Its Main Polyphenolic Ingredients Towards Ions of Zinc and Copper
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A New Tetradentate Mixed Aza-Thioether Macrocycle and Its Complexation Behavior towards Fe(II), Ni(II) and Cu(II) Ions

1
Department of Chemistry, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong
2
State Key Laboratory of Terahertz and Millimeter Waves, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2020, 25(9), 2030; https://doi.org/10.3390/molecules25092030
Submission received: 5 April 2020 / Revised: 21 April 2020 / Accepted: 22 April 2020 / Published: 27 April 2020
(This article belongs to the Special Issue Chelating Agents towards Bioavailable Metal Ions)

Abstract

:
A new tetradentate mixed aza-thioether macrocyclic ligand 2,6-dithia[7](2,9)-1,10-phenanthrolinophane ([13]ane(phenN2)S2) was successfully synthesized. Reacting metal precursors [Fe(CH3CN)2(OTf)2], Ni(ClO4)2·6H2O, and Cu(ClO4)2·6H2O with one equivalent of [13]ane(phenN2)S2 afforded [Fe([13]ane(phenN2)S2)(OTf)2] (1), [Ni([13]ane(phenN2)S2)](ClO4)2 (2(ClO4)2), and [Cu([13]ane(phenN2)S2)(OH2)](ClO4)2 (3(ClO4)2), respectively. The structures of [13]ane(phenN2)S2 and all of its metal complexes were investigated by X-ray crystallography. The [13]ane(phenN2)S2 was found to behave as a tetradentate ligand via its donor atoms N and S.

Graphical Abstract

1. Introduction

Coordination chemistry of first-row transition metal ions has attracted considerable attention for decades due to their high biological relevance. Not only do the coordination complexes serve as models to study biologically related metallomolecules, but they also provide insights into biomimetic designs [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15]. Among a reservoir of ligand designs, macrocycles have been demonstrated to be a valuable type of auxiliary ligand for these studies because (1) many macrocycles are incorporated in naturally-occurring metalloproteins and metalloenzymes, such as porphyrin in haemoglobin and corrin in vitamin B12, and (2) the properties of the derived metal complexes can be modulated via modification in the ring size and donor atoms on the macrocyclic ligands.
Our group has recently initiated a paradigm in scrutinizing the reactivity between low-valent transition metal precursors and heteroatom-functionalized alkynes [16,17,18,19,20,21,22,23,24,25,26,27,28,29]. Gratifyingly, several Ru- and Os-heterocyclic complexes reported by our group exhibit interesting biological applications. The success of preparing these functional metalated heterocyclic complexes prompted us to expand the scope of our studies to other biologically relevant first-row transition metals. Moreover, we envision that employing multidentate macrocycles as auxiliary ligands in these studies can reduce the ambiguity in mechanistic elucidations because well-designed macrocycles can effectively minimize partial ligand dissociation. With these considerations in mind, we would like to prepare some new first-row transition metal complexes supported by macrocycles for analogous reactivity studies. Among a variety of macrocycles, we are interested in mixed phenanthroline-thioether ligands, not only because their preparation is straightforward but they are well-known to stabilize various low-valent first-row transition metal ions (Scheme 1) [30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47]. While most of the reported mixed phenanthroline-thioether macrocycles are a 15-membered ring with 5 donor atoms to behave as pentadentate ligands, analogs with smaller ring sizes that can increase their conformational rigidity are rare. In this context, we report the preparation and complexation ability of a new tetradentate macrocyclic ligand 2,6-dithia [7](2,9)-1,10-phenanthrolinophane (denoted as [13]ane(phenN2)S2). To our delight, this macrocycle allows the preparation of low-valent transition metal complexes [Fe([13]ane(phenN2)S2)(OTf)2], [Ni([13]ane(phenN2)S2)](ClO4)2 and [Cu([13]ane(phenN2)S2)(OH2)](ClO4)2 under mild reaction conditions with high yields. The solid-state structures of [13]ane(phenN2)S2 and its metal complexes have been determined by X-ray crystallography. The success of preparing these macrocyclic complexes undoubtedly paves the way for performing further reactivity studies.

2. Results and Discussion

2.1. Synthesis and Characterization of [13]ane(phenN2)S2

The new mixed aza-thioether macrocyclic ligand [13]ane(phenN2)S2 was prepared from cyclization between 2,9-bis(chloromethyl)-1,10-phenanthroline and 1,3-propanedithiol in the presence of Cs2CO3 as a base and source of template cation under high dilution conditions (Scheme 2). The 2,9-bis(chloromethyl)-1,10-phenanthroline can be synthesized in accordance with previously reported literature methods from 2,9-dimethyl-1,l0-phenanthroline as depicted in Scheme 2 [48,49,50,51,52]. The formation of this air-stable macrocycle was characterized by mass spectroscopy together with 1H and 13C-NMR spectroscopy.
The structure of [13]ane(phenN2)S2 was further confirmed by X-ray crystallography (Figure 1). While the phenanthroline moiety of this macrocycle is essentially planar, the 13-membered ring adopts a folded conformation in the solid state. More specifically, the dimethylphenanthroline moiety is essentially planar whereas the -S(CH2)3S- chain is tilted over the phenanthroline unit, with an angle of 51.6° between the mean planes constructed by C(1), C(10), C(13), C(17) and C(13), C(17), S(1), S(2). The C-C bonds along the -S(CH2)3S- unit take on anti conformations. All the C-S bond distances (1.822(2)-1.823(2) Å) are typical for thioethers [47]. Since the lone pairs of the S-donors adopt exodentate orientation pointing out of the ring cavity, large conformational change is expected for [13]ane(phenN2)S2 to act as a tetradentate ligand via donor atoms N and S.
To explore the possible low-energy conformers, the X-ray structure for [13]ane(phenN2)S2 was optimized at the density functional theory (DFT) level to yield the first conformer, Conformer I. The search for other low-energy conformers was then performed by exploring the torsional energy surface of the aliphatic chain, followed by structural optimization. Two other conformers, denoted as Conformers II and III, were obtained throughout the conformer search. These conformers IIII are labeled in ascending order of their relative energy (0.0, 1.2 and 4.4 kcal mol−1, respectively). Their structures are depicted in Figure 2, with selected bond lengths (Å) and angles (°) summarized in Table 1. Notably, the structure of Conformer I is in good agreement with the X-ray structure, in which the two S atoms are tilted over the dimethylphenanthroline moiety in the same direction and the two C-C bonds on the -S(CH2)3S- unit take on anti conformation. On the other hand, one of the S atoms in Conformer II is coplanar with the dimethylphenanthroline moiety, while another S atom keeps tilting over the phenanthroline unit. Although the two S atoms in Conformer III are tilted over the dimethylphenanthroline moiety in the same direction, the C-C bonds on the -S(CH2)3S- unit take on a combination of anti and gauche conformations, making it structurally distinct from Conformer I. As none of these conformers have their donor atoms pointing to a common center, large distortion is expected for the ligand to act as a tetradentate ligand.

2.2. Synthesis and Characterization of [13]ane(phenN2)S2-Containing Transition Metal Complexes

The coordination behavior of [13]ane(phenN2)S2 ligand was explored with various first-row transition metal ions including Fe(II), Ni(II) and Cu(II) (Scheme 3). Reacting metal precursors [Fe(CH3CN)2(OTf)2], Ni(ClO4)2·6H2O, and Cu(ClO4)2·6H2O with one equivalent of [13]ane(phenN2)S2 afforded a six-coordinated Fe complex [Fe([13]ane(phenN2)S2)(OTf)2] (1), four-coordinated Ni complex [Ni([13]ane(phenN2)S2)](ClO4)2 (2(ClO4)2), and five-coordinated Cu complex [Cu([13]ane(phenN2)S2)(OH2)](ClO4)2 (3(ClO4)2), respectively. All of them are isolable in good yields (70–80%). It is noteworthy that both 2(ClO4)2 and 3(ClO4)2 are very stable in CH3CN as their X-ray structures can be obtained upon recrystallization under ambient conditions. On the other hand, 1 exhibits a certain degree of instability towards air- and moisture-sensitive conditions. For example, the color of 1 changed from yellow to deep brown during recrystallization in CH3CN under ambient conditions, leading to an ill-defined product. Despite this observation, 1 cannot be regarded as very air-sensitive since its UV-visible absorption profile in non-degassed CH3CN remained unchanged for 16 h (Figure S1). The solubility of all metal complexes was also investigated in common organic solvents (CH3CN, acetone, CH2Cl2, THF and MeOH) and H2O. All of the three complexes were found to be soluble in CH3CN. In addition, 1 is soluble in acetone and CH2Cl2, whereas 2(ClO4)2 and 3(ClO4)2 are soluble in H2O. The solution magnetic moments for these complexes were determined via 1H-NMR spectroscopy using the Evans method [53,54]; the Fe(II) and Ni(II) complexes were determined to possess high-spin electronic states (S = 2 for Fe and S = 1 for Ni), whereas the electronic spin determined for the Cu(II) complex was the expected S = 1/2. The molecular structures for all these metal complexes were determined by X-ray crystallography. Their perspective views and selected structural parameters are depicted in Figure 3, Figures 5 and 6. It is evident that [13]ane(phenN2)S2 possesses structural flexibility to behave as a tetradentate ligand, and the energy required for the conformational change is expected to be small as complexes 1, 2(ClO4)2 and 3(ClO4)2 can be prepared at room temperature.
In high-spin (S = 2) Fe complex 1, the [13]ane(phenN2)S2 acts as a tetradentate ligand, with two N and S atoms coordinating to the Fe center (Figure 3). Meanwhile, the Fe center is coordinated with two O-bound OTf. The coordination of OTf to the Fe center in solution was revealed by 19F-NMR study, with the observation of a single peak at −72.05 ppm indicating the coordinating nature of OTf [55,56,57]. The coordination geometry of 1 is best described as trigonal prismatic rather than octahedral. As depicted in Figure 4, the angles formed by the lateral faces f1 and f2 (61.5°), f2 and f3 (59.1°), f3 and f1 (60.5°) are all close to 60°, and the angle between the two triangular basal planes of the trigonal prism is small (7.9o), indicating that these basal planes are nearly parallel to each other. The distortion from a perfect trigonal prismatic geometry is characterized by the average twist angle of 14.9°. Meanwhile, the Fe-N (2.156(2)-2.209(2) Å) and Fe-S (2.486(1)-2.531(1) Å) bond distances are consistent with those of high-spin Fe(II) species bearing diimine moiety [58] and thioether linker [59]. The Fe-O bond distances (2.126(2)-2.134(2) Å) of 1 are typical of known high-spin Fe(II) complexes bearing coordinated OTf (2.025-2.211 Å) [60,61,62].
In high-spin (S = 1) 2(ClO4)2, the Ni atom adopts a square planar geometry with the N and S atoms of [13]ane(phenN2)S2 occupying the equatorial position (Figure 5). The Ni atom only lies above the mean plane constructed by the four coordinating atoms of [13]ane(phenN2)S2 by 0.073 Å, and the geometry index τ [63,64] calculated for this complex is 0.11 (cf. τ ≈ 0 and τ ≈ 1 for square planar and tetrahedral geometry, respectively, in four-coordinated compounds). In comparison with [Ni([15]ane(phenN2)S3(CH3CN)]2+ where the Ni atom adopts a distorted octahedral geometry [46], the Ni-N and Ni-S bond distances in 2(ClO4)2 (1.849(2)-1.855(2) Å for Ni-N, 2.155(1)-2.156(1) Å for Ni-S) are shorter than those in [Ni([15]ane(phenN2)S3(CH3CN)]2+ (2.013(2)–2.025(2) Å for Ni-N, 2.434(1)-2.444(1) Å for Ni-S). These observations are probably originated from the restricted flexibility of the shorter aliphatic linker in [13]ane(phenN2)S2.
In 3(ClO4)2, the Cu atom adopts a square pyramidal geometry (τ = 0.02; cf. τ ≈ 0 and τ ≈ 1 for square pyramidal and trigonal bipyramidal geometry, respectively, in five-coordinated compounds), with the N and S atoms of [13]ane(phenN2)S2 occupying the equatorial position, and the O atom of H2O in the axial position (Figure 6). The Cu atom is found to be located above the mean plane constructed by the four coordinating atoms of [13]ane(phenN2)S2 by 0.444 Å. The Cu-N bond distances of 3(ClO4)2 (1.950(3)-1.963(3) Å) are comparable to that in [Cu([15]ane(phenN2)S2)(ClO4)]+ (1.956(5) Å) [47]. On the other hand, the Cu-S bond distances in 3(ClO4)2 (2.292(1)-2.296(1) Å) are shorter than those in [Cu([15]ane(phenN2)S2)(ClO4)]+ (2.347(2) Å). This again reflects the restricted flexibility of the shorter aliphatic linker in [13]ane(phenN2)S2.

3. Materials and Methods

3.1. General Procedures

All reactions were performed under an argon atmosphere using standard Schlenk techniques unless otherwise stated. All reagents were used as received, and solvents for reactions were purified by a PureSolv MD5 solvent purification (Amesbury, MA, USA) system. The 2,9-dicarbaldehyde-1,10-phenanthroline [50,52], 2,9-bis(hydroxymethyl)-1,10-phenanthroline [51], 2,9-bis(chloromethyl)-1,10-phenanthroline [48,49] and Fe(CH3CN)2(OTf)2 [65], were prepared in accordance with the literature methods. 1H, 13C{1H}, 1H-1H COSY, 1H-1H-ROESY, 1H–13C-HSQC and 1H-13C-HMBC-NMR spectra were recorded on a Bruker 600 AVANCE III FT-NMR spectrometer (Billerica, MA, USA). Peak positions were calibrated with solvent residue peaks as internal standard. Labeling scheme for H and C atoms in the NMR assignments is shown in Scheme 4. Electrospray mass spectrometry was performed on a PE-SCIEX API 3200 triple quadrupole mass spectrometer (Tokyo, Japan). Elemental analyses were done on an Elementar Vario Micro Cube carbon-hydrogen-nitrogen elemental microanalyzer (Okehampton, UK). UV-visible spectra were recorded on a Shimadzu UV-1800 spectrophotometer (San Francisco, CA, USA). Neocuproine (98%), nickel(II) perchlorate hexahydrate (Ni(ClO4)2·6H2O) and copper(II) perchlorate hexahydrate (Cu(ClO4)2·6H2O) were purchased from J&K Scientific Ltd (Hong Kong). Iron(II) chloride (FeCl2, anhydrous, 98%) and trimethylsilyl trifluoromethanesulphonate ((CH3)3SiOTf, 98%) were purchased from Fluorochem Ltd. (Derbyshire, UK)

3.2. Solution Magnetic Susceptibility Measurements

The solution state molar magnetic susceptibility ( χ m ) of 1, 2(ClO4)2 and 3(ClO4)2 was measured according to the Evans method [53,54]. This involves the placement of a coaxial insert containing sample solution (0.010 M) into a standard 5 mm NMR tube containing only solvent. CD3CN containing one drop of benzene as internal reference was used as the solvent throughout. Based on the chemical shift difference of the benzene internal reference between inner and outer tubes, values of molar magnetic susceptibility ( χ m ) and magnetic moment ( μ e f f ) were calculated using the equations shown below [66,67]:
χ m   =   3 · Δ f / 1000 · f · c
μ e f f   =   798 T · χ m  
where χ m = molar magnetic susceptibility of the sample (m3 mol−1); △f = difference of the chemical shift (Hz) between the internal references of inner and outer tubes; f = operating frequency of the NMR spectrometer (Hz); c = concentration of the metal complex (mol dm−3); T = probe temperature of the measurement (K).

3.3. Geometry Index

The geometry index (τ) is a parameter ranging from 0 to 1 to describe the geometry of the coordination center in four- and five-coordinated compounds.
For four-coordinated compounds, the τ is used to distinguish whether the coordination geometry of the metal center is square planar (τ ≈ 0) or tetrahedral (τ ≈ 1) [64]. The formula is shown below:
τ = β α / 360 ° θ + 180 ° β / 180 ° θ 0.00399 α 0.01019 β + 2.55
where α and β are the two largest valence angles of the coordination center (β > α) and θ = cos−1 (−1/3) ≈ 109.5° is a tetrahedral angle.
For 2(ClO4)2, α(N(2)-Ni(1)-S(2)) = 171.9° and β(N(1)-Ni(1)-S(1)) = 172.0°.
For five-coordinated compounds, the τ is used to distinguish whether the coordination geometry of the metal center is square pyramidal (τ ≈ 0) or trigonal bipyramidal (τ ≈ 1) [63]. The formula is shown below:
τ = β α / 60 ° 0.01667 α + 0.01667 β
where α and β are the two largest valence angles of the coordination center (β > α).
For 3(ClO4)2, α(N(1)-Cu(1)-S(1)) = 154.0° and β(N(2)-Cu(1)-S(2)) = 155.0°.

3.4. Synthesis of ([13]ane(phenN2)S2) and [13]ane(phenN2)S2-containing Fe(II), Ni(II), Cu(II) complexes

Synthesis of 2,9-dicarbaldehyde-1,10-phenanthroline. The ligand was synthesized from a modified literature procedure [52]: To a solution of selenium oxide (2.26 g, 20.41 mmol) in deionized water (5 mL) and 1,4-dioxane (70 mL), 2,9-dimethyl-1,10-phenanthroline (2.00 g, 9.60 mmol) in 1,4-dioxane (20 mL) was added in a dropwise manner. The reaction mixture was refluxed for 0.5 h, during which the color of the reaction mixture changed from yellow to reddish brown with black deposits. Upon cooling to room temperature, the brown solids were collected by suction filtration and washed with Et2O (10 mL × 3). These solids were further purified by recrystallization with warm CHCl3 and allowed to cool and stand overnight. Yellowish orange solids were collected by suction filtration and washed with Et2O (10 mL × 3) to give 2,9-dicarbaldehyde-1,10-phenanthroline. 1H-NMR signals are consistent with that of the literature reported [50]. Yield: 1.95 g, 86%.
Synthesis of 2,9-bis(hydroxymethyl)-1,10-phenanthroline. The ligand was synthesized from a modified literature procedure [51]: To a solution of 2,9-dicarbaldehyde-1,10-phenanthroline (1.60 g, 6.77 mmol) in absolute EtOH (120 mL), anhydrous NaBH4 (1.02 g, 27.09 mmol) was added. The reaction mixture was refluxed under argon for 1.5 h, during which the color of the reaction mixture changed from yellow to brown. Upon cooling to room temperature, the reaction mixture was removed by reduced pressure. The aqueous layer was extracted with EtOAc (50 mL × 2). The organic phases were combined, washed with brine (50 mL × 2), dried over anhydrous MgSO4, and concentrated to give 2,9-bis(hydroxymethyl)-1,10-phenanthroline as pale pink solids. 1H-NMR signals are consistent with that of literature reported. Yield: 1.20 g, 74%.
Synthesis of 2,9-bis(chloromethyl)-1,10-phenanthroline. The ligand was synthesized from a modified literature procedure [49]: To a SOCl2 solution (16 mL) at 0 °C, 2,9-bis(hydroxymethyl)-1,10-phenanthroline (1.60 g, 6.66 mmol) was added. The reaction mixture was stirred for 2 h at 0 °C, during which the color of the reaction mixture changed from colorless to yellow. Upon adding n-hexane (30 mL) to the reaction mixture, the deep yellow precipitates were collected, dissolved in CHCl3 and extracted with an aqueous layer containing 5% (w/v) Na2CO3 and 5% (w/v) NaHCO3 for neutralization. The organic phases were combined, washed with brine (50 mL × 2), dried over anhydrous MgSO4 and concentrated to give 2,9-bis(chloromethyl)-1,10-phenanthroline as yellow solids. 1H-NMR signals are consistent with that of literature reported [48]. Yield: 1.52 g, 82%.
Synthesis of 2,6-dithia[7](2,9)-1,10-phenanthrolinophane ([13]ane(phenN2)S2). To a well-stirred suspension of dry Cs2CO3 (3.34 g, 10.24 mmol) in anhydrous DMF (224 mL) at 55 °C, a mixture of 1,3-propanedithiol (0.51 mL, 5.12 mmol) and 2,9-bis(chloromethyl)-1,10-phenanthroline (1.43 g, 5.12 mmol) in anhydrous DMF (112 mL) was added under argon for 42 h at a rate of 2.67 mL/min, during which the color of the reaction mixture changed from colorless to pale yellow suspension. Upon cooling to room temperature over a period of 24 h, the reaction mixture was removed by vacuum distillation. After dissolving the residue in CH2Cl2, the pale orange suspension was filtered to remove Cs2CO3. The filtrate was concentrated to give a red oil. The [13]ane(phenN2)S2 was obtained as white solids after silica gel column chromatography (silica gel, gradual elution with CH2Cl2/EtOAc (8:2, v/v); Rf value = 0.70 (silica gel TLC; CH2Cl2/EtOAc (8:2, v/v))). Recrystallization by slow diffusion of Et2O solution into a CH2Cl2 solution of the white solids gave analytically pure [13]ane(phenN2)S2 as white crystals. Yield: 0.45 g, 30%. Anal. Calcd. for C17H16N2S2: C, 65.35; H, 5.16; N, 8.97. Found: C, 65.38; H, 5.18; N, 8.99. ESI-MS found (calcd.): m/z 313.30 (313.46) [C17H16N2S2 + H]+. 1H-NMR (600 MHz, CDCl3): δ 2.83–2.98 (m, 2H, Hf), 3.11–3.22 (m, 4H, He), 4.12–4.29 (m, 4H, Hd), 7.43 (d, J = 6.0 Hz, 2H, Hb), 7.73 (s, 2H, Ha), 8.14 (d, J = 6.0 Hz, 2H, Hc). 13C-NMR (150 MHz, CDCl3): δ 34.57 (Cf), 36.72 (Ce), 41.81 (Cd), 121.71 (Cb), 125.73 (Ca), 127.37 (CII), 136.35 (Cc), 146.07 (CI), 160.79 (CIII).
Synthesis of [Fe(CH3CN)2(OTf)2]. The metal precursor was synthesized from a modified literature procedure [65]: To a solution of FeCl2 (0.50 g, 3.94 mmol) in anhydrous CH3CN (20 mL), (CH3)3SiOTf (4 mL in 10 mL anhydrous CH3CN, 22.10 mmol) was added in a dropwise manner. The reaction mixture was stirred under argon for 16 h, during which the color of the reaction mixture changed from pale yellow to colorless. The colorless solution was concentrated to about 5 mL by reduced pressure, and then added to Et2O (150 mL) to give white precipitates. The solids were collected by suction filtration, washed with Et2O (10 mL × 3) and dried by reduced pressure to give [Fe(CH3CN)2(OTf)2]. Yield: 1.20 g, 70%.
Synthesis of 1. A mixture of [Fe(CH3CN)2(OTf)2] (0.40 g, 0.91 mmol) and [13]ane(phenN2)S2 (0.28 g, 0.91 mmol) was stirred in CH3CN (30 mL) under argon for 16 h, during which the color of the reaction mixture changed from red to deep red with pale yellow precipitates. Upon filtration for removal of these precipitates, the supernatant was collected and concentrated to give deep red precipitates. The solids were collected by suction filtration and washed with Et2O (10 mL × 3). Recrystallization by slow diffusion of Et2O solution into a CH3CN solution of the deep red solids under argon gave analytically pure 1 as orange crystals. Yield: 0.43 g, 71%. Anal. Calcd. for C19H16N2S4FeF6O6: C, 34.24; H, 2.42; N, 4.20. Found: C, 34.28; H, 2.44; N, 4.21. 19F-NMR (400 MHz, CD3CN): δ −72.05. ESI-MS found (calcd.): m/z 517.30 (517.38) [C18H16N2S3FeF3O3]+. Evans NMR at 298 K: μeff = 4.98 μB.
Synthesis of 2(ClO4)2. To a solution of Ni(ClO4)2·6H2O (0.04 g, 0.10 mmol) in nitromethane (15 mL) and acetic anhydride (0.62 mL), [13]ane(phenN2)S2 (0.03 g, 0.10 mmol) was added. The reaction mixture was stirred under argon for 2 h, during which the color of the reaction mixture changed from pale green to yellow. The reaction mixture was concentrated to about 5 mL by reduced pressure and then added to Et2O (150 mL) to give orange precipitates. The solids were collected by suction filtration, washed with deionized water (5 mL × 2), cold EtOH (5 mL × 2), and finally Et2O (10 mL × 3). Recrystallization by slow diffusion of Et2O into a CH3CN solution of the orange precipitates gave analytically pure 2(ClO4)2 as brown crystals. Yield: 0.05 g, 80%. Anal. Calcd. for C17H16N2S2NiCl2O8: C, 35.82; H, 2.83; N, 4.91. Found: C, 35.85; H, 2.86; N, 4.93. ESI-MS found (calcd.): m/z 470.30 (470.60) [C17H16N2S2NiClO4]+. Evans NMR at 298 K: μeff = 2.77 μB.
Synthesis of 3(ClO4)2. To a solution of a mixture of [13]ane(phenN2)S2 (0.03 g, 0.10 mmol) in EtOH/CH2Cl2 (10 mL, 2:1 (v/v)), Cu(ClO4)2·6H2O (0.89 g, 2.40 mmol) in EtOH (5 mL) was added. The reaction mixture was stirred under argon for 4 h, during which the color of the reaction mixture changed from pale blue to pale purple with bluish green precipitates. The solids were collected by suction filtration, washed with cold EtOH (5 mL × 2), and finally Et2O (10 mL × 3). Recrystallization by slow diffusion of Et2O into a CH3CN solution of the bluish green precipitates gave analytically pure 3(ClO4)2 as purple crystals. Yield: 1.00 g, 70%. Anal. Calcd. for C17H18N2S2CuCl2O9: C, 34.44; H, 3.06; N, 4.72. Found: C, 34.48; H, 3.08; N, 4.76. ESI-MS found (calcd.): m/z 475.30 (475.45) [C17H16N2S2CuClO4]+. Evans NMR at 298 K: μeff = 1.69 μB.

3.5. X-ray Crstayllographic Data

CCDC 1994366–1994369 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html (or from the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK; Fax: +44 1223 336033; E-mail: [email protected]).
Crystal Data for [13]ane(phenN2)S2 (C17H16N2S2; M = 312.44 g/mol): triclinic, space group P-1 (no. 2), a = 8.7121(4) Å, b = 10.2500(6) Å, c = 10.4864(5) Å, α = 62.718(2)°, β = 68.073(2)°, γ = 70.672(2)°, V = 757.41(7) Å3, Z = 2, T = 212.99 K, μ(MoKα) = 0.346 mm−1, Dcalc = 1.370 g/cm3, 9505 reflections measured (4.558° ≤ 2Θ ≤ 52.754°), 3087 unique (Rint = 0.0502, Rsigma = 0.0514) which were used in all calculations. The final R1 was 0.0393 (I > 2σ(I)) and wR2 was 0.0973 (all data). (CCDC 1994366)
Crystal Data for 1 (C19H16F6O6S4N2Fe; M = 666.43 g/mol): monoclinic, space group C2/c (no. 15), a = 28.3742(8) Å, b = 9.9821(3) Å, c = 18.6161(5) Å, β = 112.3340(10)°, V = 4877.2(2) Å3, Z = 8, T = 172.98 K, μ(MoKα) = 1.048 mm−1, Dcalc = 1.815 g/cm3, 29,372 reflections measured (4.366° ≤ 2Θ ≤ 52.764°), 4989 unique (Rint = 0.0523, Rsigma = 0.0336) which were used in all calculations. The final R1 was 0.0367 (I > 2σ(I)) and wR2 was 0.0898 (all data). (CCDC 1994367)
Crystal Data for 2(ClO4)2 (C17H16N2O8S2Cl2Ni; M = 570.05 g/mol): triclinic, space group P-1 (no. 2), a = 8.8218(7) Å, b = 10.0103(8) Å, c = 12.7982(9) Å, α = 108.789(7)°, β = 106.996(7)°, γ = 92.621(6)°, V = 1010.95(13) Å3, Z = 2, T = 293(2) K, μ(CuKα) = 6.231 mm−1, Dcalc = 1.873 g/cm3, 6468 reflections measured (7.72° ≤ 2Θ ≤ 134.98°), 3576 unique (Rint = 0.0202, Rsigma = 0.0256) which were used in all calculations. The final R1 was 0.0295 (I > 2σ(I)) and wR2 was 0.0810 (all data). (CCDC 1994368)
Crystal Data for 3(ClO4)2 (C17H18N2O9S2Cl2Cu; M = 592.89 g/mol): monoclinic, space group P21/c (no. 14), a = 26.9718(6) Å, b = 11.0279(3) Å, c = 14.7341(3) Å, β = 99.779(2)°, V = 4318.87(18) Å3, Z = 8, T = 293(2) K, μ(MoKα) = 1.508 mm−1, Dcalc = 1.824 g/cm3, 20,460 reflections measured (6.62° ≤ 2Θ ≤ 53°), 8925 unique (Rint = 0.0213, Rsigma = 0.0291) which were used in all calculations. The final R1 was 0.0487 (I > 2σ(I)) and wR2 was 0.1291 (all data). (CCDC 1994369)

3.6. Computational Methodology

The structures and relative energies of Conformers IIII were calculated by density functional theory (DFT) calculations using the ORCA software package (version 4.1.1, Max-Planck-Institut für Chemische Energiekonversion, Mülheim an der Ruhr, North Rhine-Westphalia, Germany) [68]. All the calculated structures were optimized in the gas phase at the BP86/def2-SVP level with the Grimme’s DFTD3 method to account for the dispersion effects, and the Resolution of Identity (RI) approximation to speed up the calculations. Tight self-consistent field (SCF) convergence criteria were used throughout. The optimized conformers were confirmed to be local minima via frequency calculations (no negative frequency).

4. Conclusions

A new tetradentate mixed aza-thioether macrocyclic ligand 2,6-dithia[7](2,9)-1,10-phenanthrolinophane ([13]ane(phenN2)S2) was successfully prepared. Three low-valent first-row transition metal complexes supported by [13]ane(phenN2)S2, namely [Fe([13]ane(phenN2)S2)(OTf)2], [Ni([13]ane(phenN2)S2)](ClO4)2 and [Cu([13]ane(phenN2)S2)(OH2)](ClO4)2, were synthesized from reactions between corresponding transition metal precursors and [13]ane(phenN2)S2 under mild reaction conditions. The X-ray structures of the metal complexes revealed that [13]ane(phenN2)S2 acts as a tetradentate ligand by coordinating to the metal centers via donor atoms N and S. The success of preparing these macrocyclic complexes undeniably opens new opportunities for further reactivity studies and possible biological applications.

Supplementary Materials

The following are available online, Table S1–S3: Cartesian coordinates of calculated structures of Conformers IIII. Figure S1: UV-visible absorption spectra of 1, 2(ClO4)2 and 3(ClO4)2 in CH3CN at 298 K, Figure S2: 1H-NMR spectrum of [13]ane(phenN2)S2, Figure S3: 13C{1H}-NMR spectrum of [13]ane(phenN2)S2, Figure S4: 1H-1H COSY of [13]ane(phenN2)S2, Figure S5: 1H-1H ROESY of [13]ane(phenN2)S2, Figure S6: 1H-13C HSQC of [13]ane(phenN2)S2, Figure S7: 1H-13C HMBC of [13]ane(phenN2)S2.

Author Contributions

Conceptualization, S.-W.N., S.-C.C., C.-Y.W.; investigation, S.-W.N., S.-C.C., C.-F.Y., S.-M.Y.; writing—original draft preparation, S.-W.N., S.-C.C.; writing—review and editing, S.-W.N., C.-Y.W.; supervision, C.-Y.W.; funding acquisition, C.-Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

The work described in this paper was supported by the Research Grants Council of Hong Kong SAR (CityU 11207117 and T42-103/16-N) and City University of Hong Kong (7005096, 7005257, 9680206).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hubin, T.J. Synthesis and coordination chemistry of topologically constrained azamacrocycles. Coord. Chem. Rev. 2003, 241, 27–46. [Google Scholar] [CrossRef]
  2. MacKay, B.A.; Fryzuk, M.D. Dinitrogen Coordination Chemistry: On the Biomimetic Borderlands. Chem. Rev. 2004, 104, 385–401. [Google Scholar] [CrossRef]
  3. Friedle, S.; Reisner, E.; Lippard, S.J. Current challenges of modeling diiron enzyme active sites for dioxygen activation by biomimetic synthetic complexes. Chem. Soc. Rev. 2010, 39, 2768–2779. [Google Scholar] [CrossRef] [Green Version]
  4. Marchetti, L.; Levine, M. Biomimetic Catalysis. ACS Catal. 2011, 1, 1090–1118. [Google Scholar] [CrossRef]
  5. Zhao, M.; Wang, H.-B.; Ji, L.-N.; Mao, Z.-W. Insights into metalloenzyme microenvironments: Biomimetic metal complexes with a functional second coordination sphere. Chem. Soc. Rev. 2013, 42, 8360–8375. [Google Scholar] [CrossRef]
  6. Ray, K.; Pfaff, F.F.; Wang, B.; Nam, W. Status of Reactive Non-Heme Metal-Oxygen Intermediates in Chemical and Enzymatic Reactions. J. Am. Chem. Soc. 2014, 136, 13942–13958. [Google Scholar] [CrossRef]
  7. Rebilly, J.-N.; Colasson, B.; Bistri, O.; Over, D.; Reinaud, O. Biomimetic cavity-based metal complexes. Chem. Soc. Rev. 2015, 44, 467–489. [Google Scholar] [CrossRef] [Green Version]
  8. Joshi, T.; Graham, B.; Spiccia, L. Macrocyclic Metal Complexes for Metalloenzyme Mimicry and Sensor Development. Acc. Chem. Res. 2015, 48, 2366–2379. [Google Scholar] [CrossRef]
  9. Sahu, S.; Goldberg, D.P. Activation of Dioxygen by Iron and Manganese Complexes: A Heme and Nonheme Perspective. J. Am. Chem. Soc. 2016, 138, 11410–11428. [Google Scholar] [CrossRef] [Green Version]
  10. Engelmann, X.; Monte-Pérez, I.; Ray, K. Oxidation Reactions with Bioinspired Mononuclear Non-Heme Metal-Oxo Complexes. Angew. Chem. Int. Ed. 2016, 55, 7632–7649. [Google Scholar] [CrossRef]
  11. Baglia, R.A.; Zaragoza, J.P.T.; Goldberg, D.P. Biomimetic Reactivity of Oxygen-Derived Manganese and Iron Porphyrinoid Complexes. Chem. Rev. 2017, 117, 13320–13352. [Google Scholar] [CrossRef]
  12. Joshi, T.; Kubeil, M.; Nsubuga, A.; Singh, G.; Gasser, G.; Stephan, H. Harnessing the Coordination Chemistry of 1,4,7-Triazacyclononane for Biomimicry and Radiopharmaceutical Applications. ChemPlusChem 2018, 83, 554–564. [Google Scholar] [CrossRef]
  13. Sun, W.; Sun, Q. Bioinspired Manganese and Iron Complexes for Enantioselective Oxidation Reactions: Ligand Design, Catalytic Activity, and Beyond. Acc. Chem. Res. 2019, 52, 2370–2381. [Google Scholar] [CrossRef]
  14. Guo, M.; Corona, T.; Ray, K.; Nam, W. Heme and Nonheme High-Valent Iron and Manganese Oxo Cores in Biological and Abiological Oxidation Reactions. ACS Cent. Sci. 2019, 5, 13–28. [Google Scholar] [CrossRef]
  15. Dearle, A.E.; Cutler, D.J.; Fraser, H.W.L.; Sanz, S.; Lee, E.; Dey, S.; Diaz-Ortega, I.F.; Nichol, G.S.; Nojiri, H.; Evangelisti, M.; et al. An [FeIII34] Molecular Metal Oxide. Angew. Chem. Int. Ed. 2019, 58, 16903–16906. [Google Scholar] [CrossRef]
  16. Wong, C.-Y.; Lai, L.-M.; Lam, C.-Y.; Zhu, N. Ruthenium Carbene and Allenylidene Complexes Supported by the Tertiary Amine-Aromatic Diimine Ligand Set: Structural, Spectroscopic, and Theoretical Studies. Organometallics 2008, 27, 5806–5814. [Google Scholar] [CrossRef]
  17. Wong, C.-Y.; Lai, L.-M.; Pat, P.-K. Ruthenium Acetylide Complexes Supported by Trithiacyclononane and Aromatic Diimine: Structural, Spectroscopic, and Theoretical Studies. Organometallics 2009, 28, 5656–5660. [Google Scholar] [CrossRef]
  18. Wong, C.-Y.; Lai, L.-M.; Chan, S.-C.; Tai, L.-H. Photophysical and Theoretical Studies of Ruthenium(II)-Acetylide and -Cyanide Complexes with Aromatic Diimine and Trithiacyclononane. Organometallics 2010, 29, 6259–6266. [Google Scholar] [CrossRef]
  19. Chung, L.-H.; Wong, C.-Y. Isolation of Ruthenium-Indolizine Complexes: Insight into the Metal-Induced Cycloisomerization of Propargylic Pyridines. Organometallics 2013, 32, 3583–3586. [Google Scholar] [CrossRef]
  20. Chung, L.-H.; Yeung, C.-F.; Ma, D.-L.; Leung, C.-H.; Wong, C.-Y. Metal-Indolizine Zwitterion Complexes as a New Class of Organometallic Material: A Spectroscopic and Theoretical Investigation. Organometallics 2014, 33, 3443–3452. [Google Scholar] [CrossRef]
  21. Tsui, W.-K.; Chung, L.-H.; Tsang, W.-H.; Yeung, C.-F.; Chiu, C.-H.; Lo, H.-S.; Wong, C.-Y. Synthesis, Spectroscopic and Theoretical Studies of Ruthenafuran and Osmafuran Prepared by Activation of Ynone in Alcohol. Organometallics 2015, 34, 1005–1012. [Google Scholar] [CrossRef]
  22. Yeung, C.-F.; Chung, L.-H.; Lo, H.-S.; Chiu, C.-H.; Cai, J.; Wong, C.-Y. Isolation of Ruthenium-Indoline and -Indole Zwitterion Complexes: Insight into the Metal-Induced Cyclization of Aniline-Tethered Alkynes and Strategy to Lower the Activation Barrier of Metal-Vinylidene Formation. Organometallics 2015, 34, 1963–1968. [Google Scholar] [CrossRef]
  23. Ng, S.-W.; Chung, L.-H.; Yeung, C.-F.; Lo, H.-S.; Shek, H.-L.; Kang, T.-S.; Leung, C.-H.; Ma, D.-L.; Wong, C.-Y. Metalated Chromene and Chromone Complexes: pH Switchable Metal-Carbon Bonding Interaction, Photo-triggerable Chromone Delivery Application, and Antioxidative Activity. Chem. Eur. J. 2018, 24, 1779–1783. [Google Scholar] [CrossRef]
  24. Chung, L.-H.; Ng, S.-W.; Yeung, C.-F.; Shek, H.-L.; Tse, S.-Y.; Lo, H.-S.; Chan, S.-C.; Tse, M.-K.; Yiu, S.-M.; Wong, C.-Y. Ruthenium-indolizinone complexes as a new class of metalated heterocyclic compounds: Insight into unconventional alkyne activation pathways, revelation of unexpected electronic properties and exploration of medicinal application. Dalton Trans. 2018, 47, 12838–12842. [Google Scholar] [CrossRef]
  25. Chung, L.-H.; Wong, C.-Y. Ruthenium-Induced Alkyne Cycloisomerization: Construction of Metalated Heterocycles, Revelation of Unconventional Reaction Pathways, and Exploration of Functional Applications. Chem. Eur. J. 2019, 25, 2889–2897. [Google Scholar] [CrossRef]
  26. Yeung, C.-F.; Chung, L.-H.; Ng, S.-W.; Shek, H.-L.; Tse, S.-Y.; Chan, S.-C.; Tse, M.-K.; Yiu, S.-M.; Wong, C.-Y. Phosphonium-Ring-Fused Bicyclic Metallafuran Complexes of Ruthenium and Osmium. Chem. Eur. J. 2019, 25, 9159–9163. [Google Scholar] [CrossRef]
  27. Chung, L.-H.; Yeung, C.-F.; Shek, H.-L.; Wong, C.-Y. Isolation of a C3-metalated indolizine complex and a phosphonium ring-fused bicyclic metallafuran from the osmium-induced transformation of pyridine-tethered alkynes. Faraday Discuss. 2019, 220, 196–207. [Google Scholar] [CrossRef]
  28. Aoki, Y.; Bauer, M.; Braun, T.; Cadge, J.A.; Clarke, G.E.; Durand, D.J.; Eisenstein, O.; Gallarati, S.; Greaves, M.; Harvey, J.; et al. Mechanistic insight into organic and industrial transformations: General discussion. Faraday Discuss. 2019, 220, 282–316. [Google Scholar] [CrossRef]
  29. Ng, S.-W.; Tse, S.-Y.; Yeung, C.-F.; Chung, L.-H.; Tse, M.-K.; Yiu, S.-M.; Wong, C.-Y. Ru(II)- and Os(II)-Induced Cycloisomerization of Phenol-Tethered Alkyne for Functional Chromene and Chromone Complexes. Organometallics 2020, 39, 1299–1309. [Google Scholar] [CrossRef]
  30. Contu, F.; Demartin, F.; Devillanova, F.A.; Garau, A.; Isaia, F.; Lippolis, V.; Salis, A.; Verani, G. Conformationally locked mixed aza-thioether macrocycles: Synthesis and structures of complexes of PdII, PtII and RhIII of 2,5,8-trithia-[9](2,9)-1,10-phenanthrolinophane. J. Chem. Soc. Dalton Trans. 1997, 4401–4405. [Google Scholar] [CrossRef]
  31. Blake, A.J.; Casabò, J.; Devillanova, F.A.; Escriche, L.; Garau, A.; Isaia, F.; Lippolis, V.; Kivekas, R.; Muns, V.; Schröder, M.; et al. Mixed aza–thioether crowns containing the 1,10-phenanthroline sub-unit. Substitution reactions in [NiL(MeCN)][BF4]2 {L = 2,5,8-trithia[9](2,9)-1,10-phenanthrolinophane}. J. Chem. Soc. Dalton Trans. 1999, 1085–1092. [Google Scholar] [CrossRef]
  32. Shamsipur, M.; Javanbakht, M.; Mousavi, M.F.; Ganjali, M.R.; Lippolis, V.; Garau, A.; Tei, L. Copper(II)-selective membrane electrodes based on some recently synthesized mixed aza-thioether crowns containing a 1,10-phenanthroline sub-unit. Talanta 2001, 55, 1047–1054. [Google Scholar] [CrossRef]
  33. Arca, M.; Blake, A.J.; Casabò, J.; Demartin, F.; Devillanova, F.A.; Garau, A.; Isaia, F.; Lippolis, V.; Kivekas, R.; Muns, V.; et al. Conformationally locked pentadentate macrocycles containing the 1,10-phenanthroline unit. Synthesis and crystal structure of 5-oxa-2,8-dithia[9](2,9)-1,10-phenanthrolinophane (L) and its coordination properties to NiII, PdII, PtII, RhIII and RuII. J. Chem. Soc. Dalton Trans. 2001, 1180–1188. [Google Scholar] [CrossRef]
  34. Shamsipur, M.; Javanbakht, M.; Lippolis, V.; Garau, A.; De Filippo, G.; Ganjali, M.R.; Yari, A. Novel Ag+ ion-selective electrodes based on two new mixed azathioether crowns containing a 1,10-phenanthroline sub-unit. Anal. Chim. Acta 2002, 462, 225–234. [Google Scholar] [CrossRef]
  35. Aragoni, M.C.; Arca, M.; Demartin, F.; Devillanova, F.A.; Isaia, F.; Garau, A.; Lippolis, V.; Jalali, F.; Papke, U.; Shamsipur, M.; et al. Fluorometric Chemosensors. Interaction of Toxic Heavy Metal Ions PbII, CdII, and HgII with Novel Mixed-Donor Phenanthroline-Containing Macrocycles: Spectrofluorometric, Conductometric, and Crystallographic Studies. Inorg. Chem. 2002, 41, 6623–6632. [Google Scholar] [CrossRef] [PubMed]
  36. Shamsipur, M.; Javanbakht, M.; Ganjali, M.R.; Mousavi, M.F.; Lippolis, V.; Garau, A. Mixed Aza-Thioether Crowns Containing a 1,10-Phenanthroline Sub-Unit as Neutral Ionophores for Silver Ion. Electroanalysis 2002, 14, 1691–1698. [Google Scholar] [CrossRef]
  37. Blake, A.J.; Caçote, M.H.M.; Devillanova, F.A.; Garau, A.; Isaia, F.; Lippolis, V.; Pereira, C.M.; Silva, F.; Tei, L. Coordination Chemistry of 2,5,8-Trithia[9],(2,9)-1,10-phenanthrolinophane (L) toward Rhodium(III) at the Polarised Water/1,2-Dichloroethane Interface—A Possible New Approach to the Problem of Separating RhIII from Chloride Media. Eur. J. Inorg. Chem. 2002, 1816–1822. [Google Scholar] [CrossRef]
  38. Shamsipur, M.; Kazemi, S.Y.; Azimi, G.; Madaeni, S.S.; Lippolis, V.; Garau, A.; Isaia, F. Selective transport of silver ion through a supported liquid membrane using some mixed aza-thioether crowns containing a 1,10-phenanthroline sub-unit as specific ion carriers. J. Membr. Sc. 2003, 215, 87–93. [Google Scholar] [CrossRef]
  39. De Filippo, G.; Demartin, F.; Garau, A.; Lippolis, V.; Yari, A.; Shokrollahi, A.; Shamsipur, M. Complexes of Ag+ with mixed donor phenanthroline-containing macrocycles: Spectrofluorimetric, spectrophotometric, conductometric and potentiometric studies. Inorg. Chim. Acta 2005, 358, 801–807. [Google Scholar] [CrossRef]
  40. Shamsipur, M.; Hashemi, O.R.; Lippolis, V. A supported liquid membrane system for simultaneous separation of silver(I) and mercury(II) from dilute feed solutions. J. Membr. Sc. 2006, 282, 322–327. [Google Scholar] [CrossRef]
  41. Ferreira, E.S.; Garau, A.; Lippolis, V.; Pereira, C.M.; Silva, F. Electrochemistry of 2,8-dithia[9],(2,9)-1,10-phenanthrolinophane (L) at the polarized water/1,2-dichloroethane interface: Evaluation of the complexation properties towards transition and post-transition metal ions. J. Electroanal. Chem. 2006, 587, 155–160. [Google Scholar] [CrossRef]
  42. Aragoni, M.C.; Arca, M.; Bencini, A.; Biagini, S.; Blake, A.J.; Caltagirone, C.; Demartin, F.; De Filippo, G.; Devillanova, F.A.; Garau, A.; et al. Interaction of Mixed-Donor Macrocycles Containing the 1,10-Phenanthroline Subunit with Selected Transition and Post-Transition Metal Ions: Metal Ion Recognition in Competitive Liquid-Liquid Solvent Extraction of CuII, ZnII, PbII, CdII, AgI, and HgII. Inorg. Chem. 2008, 47, 8391–8404. [Google Scholar] [CrossRef] [PubMed]
  43. Gulaboski, R.; Ferreira, E.S.; Pereira, C.M.; Cordeiro, M.N.D.S.; Garau, A.; Lippolis, V.; Silva, A.F. Coupling of Cyclic Voltammetry and Electrochemical Impedance Spectroscopy for Probing the Thermodynamics of Facilitated Ion Transfer Reactions Exhibiting Chemical Kinetic Hindrances. J. Phys. Chem. C 2008, 112, 153–161. [Google Scholar] [CrossRef]
  44. Shamsipur, M.; Hashemi, B.; Dehdashtian, S.; Mohammadi, M.; Gholivand, M.B.; Garau, A.; Lippolis, V. Silver ion imprinted polymer nanobeads based on a aza-thioether crown containing a 1,10-phenanthroline subunit for solid phase extraction and for voltammetric and potentiometric silver sensors. Anal. Chim. Acta 2014, 852, 223–235. [Google Scholar] [CrossRef]
  45. Casula, A.; Nairi, V.; Fernández-Moreira, V.; Laguna, A.; Lippolis, V.; Garau, A.; Gimeno, M.C. Re(I) derivatives functionalised with thioether crowns containing the 1,10-phenanthroline subunit as a new class of chemosensors. Dalton Trans. 2015, 44, 18506–18517. [Google Scholar] [CrossRef]
  46. Blake, A.J.; Demartin, F.; Devillanova, F.A.; Garau, A.; Isaia, F.; Lippolis, V.; Schröder, M.; Verani, G. A new class of mixed aza-thioether crown containing a 1,10-phenanthroline sub-unit. J. Chem. Soc. Dalton Trans. 1996, 3705–3712. [Google Scholar] [CrossRef]
  47. Arca, M.; Azimi, G.; Demartin, F.; Devillanova, F.A.; Escriche, L.; Garau, A.; Isaia, F.; Kivekas, R.; Lippolis, V.; Muns, V.; et al. Complexes of CuII with mixed-donor phenanthroline-containing macrocycles: Analysis of their structural, redox and spectral properties in the context of Type-1 blue copper proteins biomimetic models. Inorg. Chim. Acta. 2005, 358, 2403–2412. [Google Scholar] [CrossRef]
  48. Newkome, G.R.; Kiefer, G.E.; Puckett, W.E.; Vreeland, T. α-Methyl Functionalization of Electron-Poor Heterocycles: 2,9-Bis(chloromethyl)-1,10-phenanthroline. Synthesis of a [3.3]Cyclophane Containing the 1,10-Phenanthrolino Moiety. J. Org. Chem. 1983, 48, 5112–5114. [Google Scholar] [CrossRef]
  49. Weijnen, J.G.J.; Engbersen, J.F.J. Catalytic hydrolysis of phosphate esters by metallocomplexes of 1,10-phenanthroline derivatives in micellar solution. Recl. Trav. Chim. Pays-Bas 1993, 112, 351–357. [Google Scholar] [CrossRef]
  50. De Cian, A.; DeLemos, E.; Mergny, J.-L.; Teulade-Fichou, M.-P.; Monchaud, D. Highly Efficient G-Quadruplex Recognition by Bisquinolinium Compounds. J. Am. Chem. Soc. 2007, 129, 1856–1857. [Google Scholar] [CrossRef]
  51. Higashi, T.; Inami, K.; Mochizuki, M. Synthesis and DNA-binding Properties of 1,10-Phenanthroline Analogues as Intercalating-Crosslinkers. J. Heterocyclic Chem. 2008, 45, 1889–1892. [Google Scholar] [CrossRef]
  52. Yang, Y.; Liu, J.; Yang, L.; Li, K.; Zhang, H.; Luo, S.; Rao, L. Probing the difference in covalence by enthalpy measurements: A new heterocyclic N-donor ligand for actinide/lanthanide separation. Dalton Trans. 2015, 44, 8959–8970. [Google Scholar] [CrossRef] [PubMed]
  53. Evans, D. 400. The Determination of the Paramagnetic Susceptibility of Substances in Solution by Nuclear Magnetic Resonance. J. Chem. Soc. 1959, 2003–2005. [Google Scholar] [CrossRef]
  54. Grant, D.H. Paramagnetic Susceptibility by NMR: The “Solvent Correction” Reexamined. J. Chem. Educ. 1995, 72, 39–40. [Google Scholar] [CrossRef]
  55. Britovsek, G.J.P.; England, J.; White, A.J.P. Non-heme Iron(II) Complexes Containing Tripodal Tetradentate Nitrogen Ligands and Their Application in Alkane Oxidation Catalysis. Inorg. Chem. 2005, 44, 8125–8134. [Google Scholar] [CrossRef]
  56. England, J.; Gondhia, R.; Bigorra-Lopez, L.; Petersen, A.R.; White, A.J.P. Towards robust alkane oxidation catalysts: Electronic variations in non-heme iron(II) complexes and their effect in catalytic alkane oxidation. Dalton Trans. 2009, 5319–5334. [Google Scholar] [CrossRef] [Green Version]
  57. Ayad, M.; Klein Gebbink, R.J.M.; Le Mest, Y.; Schollhammer, P.; Le Poul, N.; Pétillon, F.Y.; Mandon, D. Mononuclear iron(II) complexes containing a tripodal and macrocyclic nitrogen ligand: Synthesis, reactivity and application in cyclohexane oxidation catalysis. Dalton Trans. 2018, 47, 15596–15612. [Google Scholar] [CrossRef]
  58. Ossinger, S.; Naggert, H.; Bill, E.; Näther, C.; Tuczek, F. Electronic Structure, Vibrational Spectra, and Spin-Crossover Properties of Vacuum-Evaporable Iron(II) Bis(dihydrobis(pyrazolyl)borate) Complexes with Diimine Coligands. Origin of Giant Raman Features. Inorg. Chem. 2019, 58, 12873–12887. [Google Scholar] [CrossRef]
  59. Hughes, D.L.; Jimenez-Tenorio, M.; Leigh, G.J.; Houlton, A.; Silver, J. Iron Complexes with Polythioether Ligands: The Relation of Unusually Large Mössbauer Quadrupole Splittings to Structure. J. Chem. Soc. Dalton Trans. 1992, 2033–2037. [Google Scholar] [CrossRef]
  60. Benhamou, L.; Thibon, A.; Brelot, L.; Lachkar, M.; Mandon, D. Structural bases of oxygen-sensitivity in Fe(II) complexes with tripodal ligands. Steric effects, Lewis acidity and the role of ancillary ligands. Dalton Trans. 2012, 41, 14369–14380. [Google Scholar] [CrossRef]
  61. Panchbhai, G.; Singh, W.M.; Das, B.; Jane, R.T.; Thapper, A. Mononuclear Iron Complexes with Tetraazadentate Ligands as Water Oxidation Catalysts. Eur. J. Inorg. Chem 2016, 3262–3268. [Google Scholar] [CrossRef]
  62. Li, J.; Molenda, M.A.; Biros, S.M.; Staples, R.J.; Chavez, F.A. Assembly of a mononuclear ferrous site using a bulky aldehyde-imidazole ligand. Inorg. Chim. Acta 2017, 464, 152–156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Addison, A.W.; Nageswara Rao, T.; Reedijk, J.; Van Rijn, J.; Verschoor, G.C. Synthesis, Structure, and Spectroscopic Properties of Copper(II) Compounds containing Nitrogen-Sulphur Donor Ligands; the Crystal and Molecular Structure of Aqua[l,7-bis(N-methylbenzimidazol-2’-yl)-2,6-dithiaheptane]copper(II) Perchlorate. J. Chem. Soc. Dalton Trans. 1984, 1349–1356. [Google Scholar] [CrossRef]
  64. Okuniewski, A.; Rosiak, D.; Chojnacki, J.; Becker, B. Coordination polymers and molecular structures among complexes of mercury(II) halides with selected 1-benzoylthioureas. Polyhedron 2015, 90, 47–57. [Google Scholar] [CrossRef]
  65. Chan, S.-C.; Gupta, P.; Engelmann, X.; Ang, Z.Z.; Ganguly, R.; Bill, E.; Ray, K.; Ye, S.; England, J. Observation of Carbodicarbene Ligand Redox Noninnocence in Highly Oxidized Iron Complexes. Angew. Chem. Int. Ed. 2018, 57, 15717–15722. [Google Scholar] [CrossRef] [PubMed]
  66. Britovsek, G.J.P.; Gibson, V.C.; Spitzmesser, S.K.; Tellmann, K.P.; White, A.J.P.; Williams, D.J. Cationic 2,6-bis(imino)pyridine iron and cobalt complexes: Synthesis, structures, ethylene polymerisation and ethylene/polar monomer co-polymerisation studies. J. Chem. Soc. Dalton Trans. 2002, 1159–1171. [Google Scholar] [CrossRef]
  67. Chan, S.-C.; Ang, Z.Z.; Gupta, P.; Ganguly, R.; Li, Y.; Ye, S.; England, J. Carbodicarbene Ligand Redox Noninnocence in Highly Oxidized Chromium and Cobalt Complexes. Inorg. Chem. 2020, 59, 4118–4128. [Google Scholar] [CrossRef]
  68. Neese, F. Software update: The ORCA program system, version 4.0. WIREs Comput. Mol. Sci. 2018, 8, e1327. [Google Scholar] [CrossRef]
Sample Availability: Samples of the newly prepared compounds are available from the authors.
Scheme 1. Mixed phenanthroline-thioether macrocycles reported in previous works and this work.
Scheme 1. Mixed phenanthroline-thioether macrocycles reported in previous works and this work.
Molecules 25 02030 sch001
Scheme 2. Synthesis of [13]ane(phenN2)S2.
Scheme 2. Synthesis of [13]ane(phenN2)S2.
Molecules 25 02030 sch002
Figure 1. Perspective view of [13]ane(phenN2)S2 as represented by 30% probability ellipsoids. Selected bond lengths (Å) and angles (°): N(1)-C(1), 1.326(2); N(1)-C(5), 1.352(2); C(5)-C(6), 1.456(2); N(2)-C(6), 1.349(3); N(2)-C(10), 1.323(2); C(1)-C(17), 1.498(3); C(10)-C(13), 1.501(3); S(1)-C(13), 1.822(2); S(1)-C(14), 1.823(2); S(2)-C(16), 1.819(2); S(2)-C(17), 1.822(2); C(1)-C(17)-S(2), 114.6(1), C(10)-C(13)-S(1), 115.3(1); C(13)-S(1)-C(14), 103.5(1); C(16)-S(2)-C(17), 103.8(1).
Figure 1. Perspective view of [13]ane(phenN2)S2 as represented by 30% probability ellipsoids. Selected bond lengths (Å) and angles (°): N(1)-C(1), 1.326(2); N(1)-C(5), 1.352(2); C(5)-C(6), 1.456(2); N(2)-C(6), 1.349(3); N(2)-C(10), 1.323(2); C(1)-C(17), 1.498(3); C(10)-C(13), 1.501(3); S(1)-C(13), 1.822(2); S(1)-C(14), 1.823(2); S(2)-C(16), 1.819(2); S(2)-C(17), 1.822(2); C(1)-C(17)-S(2), 114.6(1), C(10)-C(13)-S(1), 115.3(1); C(13)-S(1)-C(14), 103.5(1); C(16)-S(2)-C(17), 103.8(1).
Molecules 25 02030 g001
Figure 2. (a) Front views and (b) side views of three possible low-energy conformers of [13]ane(phenN2)S2.
Figure 2. (a) Front views and (b) side views of three possible low-energy conformers of [13]ane(phenN2)S2.
Molecules 25 02030 g002
Scheme 3. Synthesis of [13]ane(phenN2)S2-containing complexes 1, 2(ClO4)2 and 3(ClO4)2.
Scheme 3. Synthesis of [13]ane(phenN2)S2-containing complexes 1, 2(ClO4)2 and 3(ClO4)2.
Molecules 25 02030 sch003
Figure 3. Perspective view of 1 as represented by 30% probability ellipsoids (hydrogen atoms on [13]ane(phenN2)S2 are omitted for clarity). Selected bond lengths (Å) and angles (°): Fe(1)-N(1), 2.156(2); Fe(1)-N(2), 2.209(2); Fe(1)-S(1), 2.486(1); Fe(1)-S(2), 2.531(1); Fe(1)-O(1), 2.134(2); Fe(1)-O(4), 2.126(2); N(1)-Fe(1)-N(2), 72.3(1); N(2)-Fe(1)-S(1), 74.5(1); S(1)-Fe(1)-S(2), 92.10(3); N(1)-Fe(1)-S(2), 75.4(1); O(1)-Fe(1)-N(1), 98.8(1); O(1)-Fe(1)-N(2), 81.9(1); O(4)-Fe(1)-S(1), 97.7(1); O(4)-Fe(1)-S(2), 82.6(1).
Figure 3. Perspective view of 1 as represented by 30% probability ellipsoids (hydrogen atoms on [13]ane(phenN2)S2 are omitted for clarity). Selected bond lengths (Å) and angles (°): Fe(1)-N(1), 2.156(2); Fe(1)-N(2), 2.209(2); Fe(1)-S(1), 2.486(1); Fe(1)-S(2), 2.531(1); Fe(1)-O(1), 2.134(2); Fe(1)-O(4), 2.126(2); N(1)-Fe(1)-N(2), 72.3(1); N(2)-Fe(1)-S(1), 74.5(1); S(1)-Fe(1)-S(2), 92.10(3); N(1)-Fe(1)-S(2), 75.4(1); O(1)-Fe(1)-N(1), 98.8(1); O(1)-Fe(1)-N(2), 81.9(1); O(4)-Fe(1)-S(1), 97.7(1); O(4)-Fe(1)-S(2), 82.6(1).
Molecules 25 02030 g003
Figure 4. (a) Two different views of the coordination environment surrounding the Fe center of 1 and (b) illustrations of the three lateral faces (f1f3), centroids (Ct1 and Ct2) and twist angles (θ1θ3) of the two triangular basal planes on the twisted trigonal prism.
Figure 4. (a) Two different views of the coordination environment surrounding the Fe center of 1 and (b) illustrations of the three lateral faces (f1f3), centroids (Ct1 and Ct2) and twist angles (θ1θ3) of the two triangular basal planes on the twisted trigonal prism.
Molecules 25 02030 g004
Figure 5. Perspective view of 2 as represented by 30% probability ellipsoids (hydrogen atoms on [13]ane(phenN2)S2 are omitted for clarity). Selected bond lengths (Å) and angles (°): Ni(1)-N(1), 1.849(2); Ni(1)-N(2), 1.855(2); Ni(1)-S(1), 2.155(1); Ni(1)-S(2), 2.156(1); N(1)-Ni(1)-N(2), 85.3(1); N(2)-Ni(1)-S(1), 87.5(1); S(1)-Ni(1)-S(2), 99.10(2); S(2)-Ni(1)-N(1), 87.8(1).
Figure 5. Perspective view of 2 as represented by 30% probability ellipsoids (hydrogen atoms on [13]ane(phenN2)S2 are omitted for clarity). Selected bond lengths (Å) and angles (°): Ni(1)-N(1), 1.849(2); Ni(1)-N(2), 1.855(2); Ni(1)-S(1), 2.155(1); Ni(1)-S(2), 2.156(1); N(1)-Ni(1)-N(2), 85.3(1); N(2)-Ni(1)-S(1), 87.5(1); S(1)-Ni(1)-S(2), 99.10(2); S(2)-Ni(1)-N(1), 87.8(1).
Molecules 25 02030 g005
Figure 6. Perspective view of 3 as represented by 30% probability ellipsoids (hydrogen atoms on [13]ane(phenN2)S2 are omitted for clarity). Only one of the independent cations in [3(ClO4)2]2 is depicted. Selected bond lengths (Å) and angles (°) are listed in the order of Cu(1) moiety and then Cu(2) moiety: Cu(1)-N(1), 1.963(3), 1.950(3); Cu(1)−N(2), 1.959(3), 1.955(3); Cu(1)-S(1), 2.293(1), 2.293(1); Cu(1)-S(2), 2.292(1), 2.296(1); Cu(1)-O(1), 2.160(3), 2.165(3); N(1)-Cu(1)-N(2), 81.6(1), 82.1(2); N(2)-Cu(1)-S(1), 84.5(1), 84.4(1); S(1)-Cu(1)-S(2), 99.72(4), 100.51(4); S(2)-Cu(1)-N(1), 84.1(1), 84.2(1).
Figure 6. Perspective view of 3 as represented by 30% probability ellipsoids (hydrogen atoms on [13]ane(phenN2)S2 are omitted for clarity). Only one of the independent cations in [3(ClO4)2]2 is depicted. Selected bond lengths (Å) and angles (°) are listed in the order of Cu(1) moiety and then Cu(2) moiety: Cu(1)-N(1), 1.963(3), 1.950(3); Cu(1)−N(2), 1.959(3), 1.955(3); Cu(1)-S(1), 2.293(1), 2.293(1); Cu(1)-S(2), 2.292(1), 2.296(1); Cu(1)-O(1), 2.160(3), 2.165(3); N(1)-Cu(1)-N(2), 81.6(1), 82.1(2); N(2)-Cu(1)-S(1), 84.5(1), 84.4(1); S(1)-Cu(1)-S(2), 99.72(4), 100.51(4); S(2)-Cu(1)-N(1), 84.1(1), 84.2(1).
Molecules 25 02030 g006
Scheme 4. Labeling scheme for H and C atoms in [13]ane(phenN2)S2.
Scheme 4. Labeling scheme for H and C atoms in [13]ane(phenN2)S2.
Molecules 25 02030 sch004
Table 1. Selected bond lengths (Å) and angles (°) for Conformers IIII.
Table 1. Selected bond lengths (Å) and angles (°) for Conformers IIII.
ConformersIIIIII
N(1)-C(1)1.3321.3321.331
N(1)-C(5)1.3491.3481.348
N(2)-C(6)1.3491.3531.343
N(2)-C(10)1.3321.3271.330
C(1)-C(17)1.5111.5061.512
C(10)-C(13)1.5111.5221.511
S(1)-C(13)1.8441.8251.850
S(1)-C(14)1.8451.8501.846
S(2)-C(16)1.8461.8381.851
S(2)-C(17)1.8461.8511.856
C(1)-C(17)-S(2)114.5112.9111.6
C(10)-C(13)-S(1)114.4116.9111.9
C(13)-S(1)-C(14)102.7101.7101.8
C(16)-S(2)-C(17)102.6103.6105.1

Share and Cite

MDPI and ACS Style

Ng, S.-W.; Chan, S.-C.; Yeung, C.-F.; Yiu, S.-M.; Wong, C.-Y. A New Tetradentate Mixed Aza-Thioether Macrocycle and Its Complexation Behavior towards Fe(II), Ni(II) and Cu(II) Ions. Molecules 2020, 25, 2030. https://doi.org/10.3390/molecules25092030

AMA Style

Ng S-W, Chan S-C, Yeung C-F, Yiu S-M, Wong C-Y. A New Tetradentate Mixed Aza-Thioether Macrocycle and Its Complexation Behavior towards Fe(II), Ni(II) and Cu(II) Ions. Molecules. 2020; 25(9):2030. https://doi.org/10.3390/molecules25092030

Chicago/Turabian Style

Ng, Sze-Wing, Siu-Chung Chan, Chi-Fung Yeung, Shek-Man Yiu, and Chun-Yuen Wong. 2020. "A New Tetradentate Mixed Aza-Thioether Macrocycle and Its Complexation Behavior towards Fe(II), Ni(II) and Cu(II) Ions" Molecules 25, no. 9: 2030. https://doi.org/10.3390/molecules25092030

Article Metrics

Back to TopTop