Next Article in Journal
Incorporation of Carbon Nanofillers Tunes Mechanical and Electrical Percolation in PHBV:PLA Blends
Next Article in Special Issue
Paddlewheel SBU based Zn MOFs: Syntheses, Structural Diversity, and CO2 Adsorption Properties
Previous Article in Journal
Polymeric Nanocomposites Membranes with High Permittivity Based on PVA-ZnO Nanoparticles for Potential Applications in Flexible Electronics
Previous Article in Special Issue
Bimetallic Gold-Silver Nanoparticles Supported on Zeolitic Imidazolate Framework-8 as Highly Active Heterogenous Catalysts for Selective Oxidation of Benzyl Alcohol into Benzaldehyde
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sometimes the Same, Sometimes Different: Understanding Self-Assembly Algorithms in Coordination Networks

by
Y. Maximilian Klein
,
Alessandro Prescimone
,
Mariia Karpacheva
,
Edwin C. Constable
and
Catherine E. Housecroft
*
Department of Chemistry, University of Basel, BPR 1096, Mattenstrasse 24a, 4058 Basel, Switzerland
*
Author to whom correspondence should be addressed.
Polymers 2018, 10(12), 1369; https://doi.org/10.3390/polym10121369
Submission received: 23 November 2018 / Revised: 7 December 2018 / Accepted: 8 December 2018 / Published: 11 December 2018
(This article belongs to the Special Issue Advances in Coordination Polymers)

Abstract

:
The syntheses and characterizations of three new ligands containing two 4,2′:6′,4″-tpy or two 3,2′:6′,3″-tpy metal-binding domains are reported. The ligands possess different alkyloxy functionalities attached to the central phenylene spacer: n-pentyloxy in 3, 4-phenyl-n-butoxy in 4, benzyloxy in 5. Crystal growth under ambient conditions has led to the formation of {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n, with structures confirmed by single crystal X-ray diffraction. Both the cobalt(II) center and ligand 3 or 4 act as 4-connecting nodes and both {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n possess a 3D cds net despite the fact that 3 and 4 contain two 4,2′:6′,4″-tpy and two 3,2′:6′,3″-tpy units, respectively. Taken in conjunction with previously reported data, the results indicate that the role of the alkyloxy substituent is more significant than the choice of 4,2′:6′,4″- or 3,2′:6′,3″-tpy isomer in determining the assembly of a particular 3D net. The combination of Co(NCS)2 with 5 resulted in the formation of the discrete molecular complex [Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH in which 5 acts as a monodentate ligand. The pendant phenyls and both coordinated and non-coordinated 4,2′:6′,4″-tpy units are involved in efficient π-stacking interactions.

Graphical Abstract

1. Introduction

Polymeric materials are essential to mankind, both as the crucial biomolecules enabling life processes and as the materials at the core of our socioeconomic structures. Metallopolymers are an important subset of polymers, in which the incorporation of metal centers allows the blending of new properties which are not readily accessed through polymeric structures based only on C, H, N, O, S and P atoms. An important class of metallopolymers are the coordination polymers and networks, in which the metal centers are incorporated through coordination rather than covalent bonds. Specifically, coordination polymers and networks are infinite assemblies in which the metal centers are connected by coordinated organic molecules. The coordination number and coordination geometry of the metal centers provide an exquisite control over the dimensionality as well as the topographical and topological features of the coordination assemblies. Ligands with pyridine nitrogen donors comprise some of the most commonly utilized linkers, for example [1,2,3,4,5,6,7], with 4,4′-bipyridine (4,4′-bpy, Scheme 1) used extensively as a rigid organic linker in coordination assemblies. An [M(pytpy)2]n+ (pytpy = 4′-(4-pyridyl)-2,2′:6′,2″-terpyridine) is an expanded 4,4′-bpy since both can bind two metal ions and possess similar vectorial properties (Scheme 1) [8]. Newkome and coworkers have demonstrated the assembly of metallocages based upon [M(tpy)2]n+ building blocks with other vectorial properties [9]. By changing from the 2,2′:6′,2″-isomer of terpyridine to either 3,2′:6′,3″- or 4,2′:6′,4″-terpyridine (3,2′:6′,3″- or 4,2′:6′,4″-tpy, Scheme 2) it should be possible to modify the consequences of the coordination interactions on the network assembly in a controlled and predictable manner. The ligands 3,2′:6′,3″- or 4,2′:6′,4″-tpy both bind metal ions only through the four outer N-donors and there is no literature precedent for metal-coordination through the central pyridine ring. We and others have made widespread use of 4,2′:6′,4″-tpy ligands, typically functionalized in the 4′-position for the assembly of 1D-coordination polymers and 2D-networks [10,11,12]. Examples of 3D-architectures incorporating 4,2′:6′,4″-tpy ligands are rare [13] and only a very few examples of 2D- and 3D-assemblies directed by 3,2′:6′,3″-tpy ligands have been reported [14,15,16,17,18,19,20]. The most significant difference between a 4,2′:6′,4″-tpy and 3,2′:6′,3″-tpy is the effect that inter-ring C–C bond rotation has on the mutual directionality of the outer nitrogen donors. For 4,2′:6′,4″-tpy, there is no change in the vectorial orientation of the nitrogen donors, while for 3,2′:6′,3″-tpy there is significant variation with the interannular bond rotation [11]. This in turn increases the structural diversity to be expected with 3,2′:6′,3″-tpy, at the expense of the predictability of the self-assembly algorithm. In Scheme 2, the 3,2′:6′,3″-tpy unit is drawn in an orientation preorganized for the formation of metallomacrocycles, as observed with 1-(3,2′:6′,3′′-terpyridin-4′-yl)ferrocene [21] and 5-(3,2′:6′,3′′-terpyridin-4′-yl)-N,N-diphenylthiophen- 2-amine [22].
Covalent linkage of two 4,2′:6′,4″- or 3,2′:6′,3″-tpy domains combines two ditopic metal-binding domains into a tetratopic ligand as exemplified with the general ligands 1 and 2 in Scheme 2. We have developed a suite of such ligands, typically with 1,4-phenylene spacers functionalized with alkyloxy groups to increase solubility in organic solvents [23]. The nature of the OR group is critical in determining the outcome of the self-assembly. For example, the reaction of 1a or 1b (Scheme 2) with zinc(II) halides leads to (4,4) nets in which 1a or 1b is the 4-connecting node. However, in the case of 1a, simple 2D-sheets are observed, whereas the presence of the longer n-octyl groups in 1b leads to 2D→2D parallel interpenetration with the alkyloxy chains in extended conformations and directed within each corrugated sheet [23,24]. We have also demonstrated that introducing a pendant aryl group can facilitate face-to-face π-interactions within the solid-state assembly. In the case of [Zn2Br4(1c).H2O]n (see Scheme 2 for 1c), 2-fold interpenetrating nbo nets are observed with inter-net π-stacking resulting in tightly associated interpenetrated nets and, as a consequence, large void spaces in the lattice [25].
In this paper, we describe reactions of ligands 35 (Scheme 3) with Co(NCS)2. Upon interacting with pyridine donors, Co(NCS)2 typically generates octahedral cobalt(II) centers with trans-thiocyanato ligands and four pyridine nitrogen donors defining the equatorial plane. It is therefore expected to act as a 4-connecting node giving access to a range of different networks [26,27]. We have shown that the reaction of Co(NCS)2 with 1d (Scheme 2) leads to a 3D cds net in which both metal and ligand are 4-connecting nodes [28]. Increasing the length of the alkyloxy chain to R = n-octyl and replacing the 4,2′:6′,4″- by a 3,2′:6′,3″-tpy (2a, Scheme 2) switches the 3D assembly from a cds to an lvt net [20]. In the present work, 3 was selected in order to investigate the effect of lengthening the alkyloxy chain from three to five carbon atoms—would the cds net be retained? Ligand 4 is comparable with 2a in terms of the 3,2′:6′,3″-tpy donor set but introduces terminal phenyl domains in the alkyloxy substituents. Both 3 and 5 possess 4,2′:6′,4″-tpy units, but the inclusion of the pendant phenyls in 5 was expected to affect the assembly by enabling arene–arene π-interactions.

2. Materials and Methods

2.1. General

1H and 13C NMR spectra were recorded on a Bruker DRX-500 NMR spectrometer (Bruker BioSpin AG, Fällanden, Switzerland) with chemical shifts referenced to residual solvent peaks (TMS = δ 0 ppm). Electrospray ionization (ESI) mass spectra were measured on a Shimazu LCMS 2020 instrument (Shimadzu Schweiz GmbH, Roemerstr., Switzerland) and high resolution ESI (HR-ESI) mass spectra on a Bruker maXis 4G QTOF instrument (Bruker BioSpin AG, Fällanden, Switzerland); samples were introduced as MeCN solutions containing 0.1% formic acid or 0.1% trifluoroacetic acid (TFA) for electrospray mass spectrometry (ESI MS) and high-resolution ESI MS, respectively. Commercially available precursors were purchased from Fluorochem (Hadfield, UK), Sigma-Aldrich (Buchs SG, Switzerland), TCI (Eschborn, Germany) or Acros (Chemie Brunschwig AG, Basel, Switzerland) and used without further purification. Compound 4b was prepared following the literature and NMR spectroscopic data were in accord with those reported [29].

2.2. Compound 3a

Compound 3a was prepared by adapting a literature method [29]. 2,5-Dibromohydroquinone (2.0 g, 7.47 mmol), 1-bromopentane (2.36 mL, 2.88 g, 18.7 mmol) and anhydrous K2CO3 (3.1 g, 22.4 mmol) were dissolved in dry DMF (100 mL) and the mixture was heated at 100 °C for ca. 15 h. The reaction mixture was cooled to room temperature, added to a beaker containing 100 mL of ice-water and stirred for 30 min. The precipitate that formed was separated by filtration and washed with water (3 × 30 mL) and dried in vacuo. Compound 3a was isolated as a white powder (2.85 g, 6.98 mmol, 93.5%). 1H NMR (500 MHz, CDCl3) δ/ppm 7.09 (s, 2H, HC3), 3.95 (t, J = 6.5 Hz, 4H, Ha), 1.81 (m, 4H, Hb), 1.46 (m, 4H, Hc), 1.39 (m, 4H, Hd), 0.94 (t, J = 7.2 Hz, 6H). 13C{1H} NMR (126 MHz, CDCl3) δ/ppm 150.2 (CC2), 118.6 (CC3), 111.3 (CC1), 70.5 (Ca), 29.0 (Cb), 28.3 (Cc), 22.5 (Cd), 14.2 (Ce). In reference [29], 1H NMR signals were unassigned, and 13C NMR spectroscopic data were not reported.

2.3. Compound 4a

Compound 4a has been reported but we find the following method more convenient. The method was as for 3a starting with 2,5-dibromohydroquinone (1.5 g, 5.6 mmol), 1-bromo-4-phenylbutane (0.96 mL, 3.04 g, 14 mmol) and anhydrous K2CO3 (2.32 g, 16.8 mmol) in dry DMF (100 mL). Then, 4a was isolated as a white powder (2.67 g, 5.02 mmol, 89.6%). The NMR spectroscopic data agreed with those published [29].

2.4. Compound 5a

Compound 5a has previously been reported [30], but we find the following synthesis more convenient. The method was as for 3a starting with 2,5-dibromohydroquinone (1.5 g, 5.6 mmol), benzyl chloride (1.61 mL, 1.77 g, 14.0 mmol) and anhydrous K2CO3 (2.32 g, 16.8 mmol) in dry DMF (100 mL). Then, 5a was obtained as a white powder (2.19 g, 4.89 mmol, 87.3%). 1H NMR spectroscopic data were in accord with those reported [30]. 13C{1H} NMR (126 MHz, CDCl3) δ/ppm 150.2 (CC2c), 136.3 (CD1), 128.8 (CD3), 128.3 (CD4), 127.4 (CD2), 119.4 (CC3), 111.7 (CC1), 72.1 (Ca).

2.5. Compound 3b

Compound 3a (1.8 g, 4.41 mmol) and dry Et2O (150 mL) were added to a dried flask and cooled to 0 °C in an ice bath. nBuLi solution (1.6 M in hexanes, 8.27 mL, 13.2 mmol) was slowly added to the solution of 3a over 20 min and the temperature maintained at 0 °C for 6 h. Dry DMF (1.02 mL, 13.2 mmol) was then added and the solution stirred for 15 h, while warming up to room temperature. The reaction was neutralized with saturated aqueous NH4Cl solution and extracted with CH2Cl2 (200 mL). The organic phase was dried over MgSO4 and concentrated in vacuo. Compound 3b was obtained as a yellow solid (1.14 g, 3.72 mmol, 84.4%) and used without further purification. 1H NMR (500 MHz, CDCl3) δ/ppm 10.51 (s, 2H, HCHO), 7.42 (s, 2H, HC3), 4.08 (t, J = 6.5 Hz, 4H, Ha), 1.87–1.76 (m, 4H, Hb), 1.52–1.31 (m, 8H, Hc+d), 0.93 (t, J = 7.2 Hz, 6H, He). 13C{1H} NMR (126 MHz, CDCl3) δ/ppm 189.6 (CCHO), 155.3 (CC2), 129.4 (CC1), 111.7 (CC3), 69.3 (Ca), 28.8 (Cb), 28.3 (Cc), 22.5 (Cd), 14.1 (Ce).

2.6. Compound 5b

The method was the same as for 3b but starting with 5a (2.0 g, 4.46 mmol) and nBuLi (1.6 M in hexanes, 8.36 mL, 13.4 mmol). Compound 2b was obtained as a yellow solid (1.23 g, 3.55 mmol, 79.6%) and was used without further purification. 1H NMR spectroscopic data matched those reported [29]. 13C{1H} NMR (126 MHz, CDCl3) δ/ppm 189.2 (CCHO), 155.2 (CC2), 135.8 (CD1), 129.7 (CC1), 128.9 (CD2/D3), 128.5 (CD4), 127.7 (CD2/D3), 112.5 (CC3), 71.3 (Ca).

2.7. Compound 3

Compound 3b (0.3 g, 0.98 mmol) was dissolved in EtOH (100 mL), then 4-acetylpyridine (0.45 mL, 0.49 g, 3.92 mmol) and crushed solid KOH (0.22 g, 3.92 mmol) were added in one portion. Aqueous NH3 (32%, 7.8 mL) was added dropwise and the reaction mixture was stirred at room temperature for ca. 15 h. The precipitate that had formed was collected by filtration and washed with water, EtOH and Et2O (3 × 10 mL, each). Compound 3 was obtained as a white solid (0.13 g, 0.18 mmol, 18.8%). M.p. = 297 °C. 1H NMR (500 MHz, CDCl3) δ/ppm 8.81 (m, 8H, HA2), 8.15–8.07 (m, 12H, HB3+A3), 7.16 (s, 2H, HC3), 4.07 (t, J = 6.3 Hz, 4H, Ha), 1.87 (m, 4H, Hb), 1.37 (m, 4H, Hc), 1.32–1.19 (m, 4H, Hd), 0.77 (t, J = 7.3 Hz, 6H, He). 13C{1H} NMR (126 MHz, CDCl3) δ/ppm 154.8 (CB2), 150.7 (CC2), 150.6 (CA2), 148.4 (CB4), 146.4 (CA4), 129.2 (CC1), 121.7 (CB3), 121.3 (CA3), 115.3 (CC3), 69.9 (Ca), 29.27 (Cb), 28.61 (Cc), 22.53 (Cd), 14.02 (Ce). ESI-MS m/z 754.40 [M + H + MeCN]+ (base peak, calc. 754.39), 713.35 [M + H]+ (calc. 713.36). HR ESI-MS m/z 713.3600 [M + H]+ (calc. 713.3599).

2.8. Compound 4

The method was as for compound 3 but starting with 4b (0.69 g, 1.62 mmol) and 3-acetylpyridine (0.82 mL, 0.89 g, 7.25 mmol), crushed KOH (0.41 g, 7.25 mmol) and aqueous NH3 (32%, 12.4 mL). Compound 4 was isolated as a white solid (0.47 g, 0.56 mmol, 34.7%). M.p. = 253 °C. 1H NMR (500 MHz, CDCl3) δ/ppm 9.38 (d, J = 2.2 Hz, 4H, HA2), 8.73 (m, 4H, HA6), 8.50 (dd, J = 7.9, 2.0 Hz, 4H, HA4), 8.01 (s, 4H, HB3), 7.45 (dd, J = 8.0, 7.9 Hz, 4H, HA5), 7.17–7.12 (m, 6H, HC3+D3), 7.09 (m, 2H, HD4), 6.98 (m, 4H, HD2), 4.07 (t, J = 6.1 Hz, 4H, Ha), 2.55 (t, J = 7.5 Hz, 4H, Hd), 1.87–1.78 (m, 4H, Hb), 1.75–1.67 (m, 4H, Hc). 13C{1H} NMR (126 MHz, CDCl3) δ/ppm 154.9 (CB2), 150.7 (CC2), 150.4 (CA6), 148.6 (CA2), 148.1 (CB4), 141.9 (CD1), 134.8 (CA3), 134.6 (CA4), 129.5 (CC1), 128.45 (CD3), 128.4 (CD2), 125.9 (CD4), 123.8 (CA5), 120.3 (CB3), 115.5 (CC3), 69.7 (Ca), 35.5 (Cd), 29.1 (Cb), 28.0 (Cc). ESI-MS m/z 837.40 [M + H]+ (base peak, calc. 837.39). HR ESI-MS m/z 837.3901 [M + H]+ (calc. 837.3912).

2.9. Compound 5

The method was as for 3, starting with 5b (1.06 g, 3.06 mmol), 4-acetylpyridine (1.56 mL, 1.7 g, 13.8 mmol), crushed KOH (0.77 g, 13.8 mmol), and aqueous NH3 (32%, 24 mL). Compound 5 was isolated as a white solid (0.87 g, 1.16 mmol, 37.8%). Decomp. > 280 °C. 1H NMR (500 MHz, CDCl3) δ/ppm 8.77 (m, 8H, HA2), 8.07 (s, 4H, HB3), 7.96 (m, 8H, HA3), 7.40–7.33 (m, 10H, HD2+D3+D4), 7.31 (s, 2H, HC3), 5.19 (s, 4H, Ha). 13C{1H} NMR (126 MHz, CDCl3) δ/ppm 154.9 (CB2), 150.7 (CC2), 150.5 (CA2), 147.7 (CB4), 146.3 (CA4), 136.2 (CD1), 129.5 (CC1), 129.0 (CD3), 128.7 (CD4), 127.9 (CD2), 121.5 (CB3), 121.4 (CA3), 116.1 (CC3), 72.1 (Ca). ESI-MS m/z 794.30 [M + H + MeCN]+ (calc. 794.32), 753.35 [M + H]+ (calc. 753.30). HR ESI-MS m/z 753.2967 [M + H]+ (calc. 753.2973).

2.10. {[Co(NCS)2(3)]·0.8C6H4Cl2}n

A solution of Co(NCS)2 (1.75 mg, 0.01 mmol) in MeOH (8 mL) was layered over a solution of 3 (7.13 mg, 0.01 mmol) in 1,2-dichlorobenzene (5 mL). A few pink crystals of [Co(NCS)2(3)·0.8C6H4Cl2]n were obtained after 2–4 weeks. Insufficient material was obtained for powder X-ray diffraction.

2.11. {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n

A solution of Co(NCS)2 (3.5 mg, 0.02 mmol) in MeOH (8 mL) was layered over a solution of 4 (8.37 mg, 0.01 mmol) in 1,2-dichlorobenzene (5 mL). Pink crystals of [Co(NCS)2(4)·1.6H2O·1.2C6H4Cl2]n (2.2 mg, 0.0018 mmol, 18% based on 5) were obtained after 2–4 weeks. The bulk sample was characterized by powder X-ray diffraction (Figure S1).

2.12. [Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH

A solution of Co(NCS)2 (3.5 mg, 0.02 mmol) in MeOH (8 mL) was layered over a solution of 5 (15.1 mg, 0.02 mmol) in CHCl3 (5 mL). A small crop of pink crystals of [Co(NCS)2(MeOH)2(5)2·2CHCl3·2MeOH] was obtained after 2–4 weeks. Insufficient material was obtained for powder X-ray diffraction

2.13. Crystallography

Single crystal data were collected on a Bruker APEX-II diffractometer (Bruker Biospin AG, Fällanden, Switzerland); data reduction, solution and refinement used APEX2, SuperFlip and CRYSTALS respectively [31,32,33]. Structure analysis used Mercury v. 3.10 [34,35]. In {[Co(NCS)2(3)]·0.8C6H4Cl2}n, the dichlorobenzene molecule was refined as rigid body (occupancy = 0.8 per formula unit). In addition, the aliphatic chain showed high thermal motion, and was refined isotropically using chemically reasonable restraints to bond distances and angles. In {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n, the side chain of the organic ligand is disordered with high thermal motion; as a result, the phenyl ring was refined with rigid body restraints and the whole chain was refined isotropically. Powder diffraction data were collected on a Stoe Stadi P powder diffractometer (Stoe & Cie GmbH, Darmstadt, Germany).
{[Co(NCS)2(3)]·0.8C6H4Cl2}n: C52.80H47.20Cl1.60CoN8O2S2, M = 1005.60, pink block, monoclinic, space group P21/c, a = 10.3756(6), b = 19.1855(11), c = 16.2699(9) Å, β = 106.881(3)o, U = 3099.2(3) Å3, Z = 2, Dc = 1.078 Mg m–3, μ(Cu-Kα) = 3.749 mm−1, T = 123 K. Total 31,013 reflections, 5731 unique, Rint = 0.033. Refinement of 4672 reflections (282 parameters) with I > 2σ (I) converged at final R1 = 0.1176 (R1 all data = 0.1247), wR2 = 0.1340 (wR2 all data = 0.1341), gof = 0.9888. CCDC 1877183.
{[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n: C65.20H56.00Cl2.40CoN8O3.60S2, M = 1217.36, pink block, monoclinic, space group P21/c, a = 13.7465(6), b = 15.7832(7), c = 16.2872(8) Å, β = 112.147(2)o, U = 3273.0(3) Å3, Z = 2, Dc = 1.235 Mg m–3, μ (Cu-Kα) = 3.953 mm−1, T = 123 K. Total 23,855 reflections, 5937 unique, Rint = 0.035. Refinement of 5459 reflections (313 parameters) with I >2σ (I) converged at final R1 = 0.1460 (R1 all data = 0.1460), wR2 = 0.1415, gof = 0.9888. CCDC 1877182.
[Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH: C108H90Cl6CoN14O8S2, M = 2047.77, pink block, triclinic, space group P–1, a = 14.1006(10), b = 14.5250(10), c = 14.5512(11) Å, α =62.394(3), β = 89.108(3), γ = 64.293(3)o, U = 2314.5(3) Å3, Z = 1, Dc = 1.469 Mg m–3, μ(Cu-Kα) = 4.035 mm−1, T = 123 K. Total 26,118 reflections, 8388 unique, Rint = 0.026. Refinement of 7965 reflections (618 parameters) with I > 2σ (I) converged at final R1 = 0.1090 (R1 all data = 0.2656), wR2 = 0.1117 (wR2 all data = 0.2664), gof = 0.9524. CCDC 1877184.

3. Results

3.1. Synthesis and Characterization of Ligands

Compounds 35 were prepared following the one-pot strategy of Wang and Hanan [36] (right-side of Scheme 4), an approach that requires the appropriate dicarbaldehyde. Precursors 3a5a (Scheme 4) were prepared by treatment of 2,5-dibromohydroquinone with 1-bromopentane, 1-bromo-4-phenylbutane or benzyl chloride under basic conditions. Subsequent lithiation and treatment with DMF followed by addition of aqueous NH4Cl yielded the dicarbaldehydes 3b5b. After work-up, compounds 35 were isolated as white solids in yields ranging from 18.8 to 37.8%. In the HR-ESI mass spectrum of each compound, the highest-mass peak corresponded to the [M + H]+ ion (m/z = 713.3600, 837.3901 and 753.2967, respectively). For 3 and 5, this was the base peak, and lower intensity peaks at 357.1841 and 377.1522, respectively, arose from the [M + 2H]2+ ion. In the HR-ESI mass spectrum of 4, the [M + 2H]2+ ion gave the base peak (m/z = 419.1994), and a peak assigned to the [M + 3H]3+ ion (m/z = 279.8021) was also observed. The mass spectra are shown in Figures S2–S4. The 1H and 13C NMR spectra of the ligands were assigned by COSY, NOESY, HMQC and HMBC methods, and the atom labelling scheme is defined in Scheme 4. NOESY cross peaks between the signals for protons HA3 and HB3 in 3 and 5 distinguished HA3 from HA2. Figure 1 displays the aromatic regions of the 1H NMR spectra and Figures S5–S10 show full 1H and 13C NMR spectra.

3.2. Crystal Growth

All crystal growth experiments were carried out under ambient conditions by layering a methanol solution of Co(NCS)2 over a solution of the ligand in either 1,2-dichlorobenzene or chloroform (see Materials and methods). Parallel crystal-setups using 1,2-dichlorobenzene or chloroform for each ligand/Co(NCS)2 combination were made, and the ratio of Co(NCS)2: ligand was either 1:1 or 2:1. For ligands 3 and 4, pink crystals were harvested after 2–4 weeks, although in the case of 3, few crystals formed. X-ray quality crystals were only obtained when 1,2-dichlorobenzene was used during crystal growth, and single crystal X-ray diffraction established the assembly of coordination networks (see below). For ligand 5, crystals were only obtained when using a solution of 5 in chloroform, and X-ray diffraction revealed the unexpected discrete coordination compound [Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH.

3.3. {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n

Single crystal X-ray diffraction confirmed that reactions of Co(NCS)2 with 3 and 4 yielded the coordination assemblies {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n. For the latter, verification that the single crystal was representative of the bulk sample was obtained from powder X-ray diffraction (Figure S1). Both {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n crystallize in the monoclinic space group P21/c. The cobalt(II) center in each compound is in a similar six-coordinate environment with trans-thiocyanato ligands and nitrogen donors from four different ligands defining the equatorial sites. The asymmetric unit in each structure contains one cobalt atom and half of a ligand 3 or 4. In {[Co(NCS)2(3)]·0.8C6H4Cl2}n, the asymmetric unit contains 0.4 C6H4Cl2, while in {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n, the asymmetric unit has 0.6 C6H4Cl2 disordered over two orientations modelled with equal occupancies. For the compound with ligand 4, the pyridine ring containing N1 is disordered and was modelled over two sites of occupancies 0.6 and 0.4. The O atom of the 4-phenyl-n-butoxy chain is also disordered, and was again modelled over two sites with occupancies 0.6 and 0.4. In the discussion below, we focus only on the major occupancy sites in {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n. Figure 2 and Figure 3 illustrate that each of ligands 3 and 4 binds four cobalt (II) centers, thereby acting as a 4-connecting node. Table 1 gives bond parameters within each cobalt (II) coordination sphere. These parameters and all other bond distances and angles in the structures are unremarkable, and do not warrant further comment.
The dihedral angles between the planes through adjacent aromatic rings in coordinated 3 and 4 are given in the first two entries in Table 2. The twist of the ring with N2 with respect to the phenylene ring in each compound is typical of covalently bonded arene rings and minimizes H…H repulsions. There are small differences in the conformations of the 4,2′:6′,4″- and 3,2′:6′,3″-tpy units in coordinated 3 and 4, as displayed in Figure 4a in the overlay of the {Co4(3)} and {Co4(4)} units. However, both 3 and 4 function as planar 4-connecting nodes and, despite the isomeric donor sets, each propagates into a {65.8} cds net (Figure 5 and Figure S12) in which half of the adjacent nodes are mutually perpendicular and half are coplanar [26]. This resembles the 3D net observed in {[Co(NCS)2(1d)]·2C6H4Cl2}n [28] and, for comparison, the corresponding dihedral angles for this structure are given in Table 1. In fact, the unit cell dimensions of {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(1d)]·2C6H4Cl2}n (both in the space group P21/c) are very similar (a = 10.3756(6), b = 19.1855(11), c = 16.2699(9) Å, β = 106.881(3)o, U = 3099.2(3) Å3 versus a = 10.2136(9), b = 19.3452(17), c = 16.2214(15) Å, β = 107.027(3)o, U = 3064.6(5) Å3), revealing that the network can accommodate either n-pentyloxy or n-propoxy chains without significant perturbation (Figure S11).
The phenyl groups in the 4-phenylbutoxy chains in {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n do not play a critical role in the assembly process and are not involved in π-stacking interactions. This is in contrast to the 3-phenylpropoxy groups in [Zn2Br4(1c).H2O]n (see introduction) [25]. Instead, the 1,2-dichlorobenzene solvate in {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n engages in a face-to-face π-interaction with the pyridine ring containing atom N3 (Figure S13). Pertinent parameters are centroid…centroid and centroid…pyridine-plane separations of 3.74 and 3.49 Å, respectively, and an angle between the ring planes of 9.7°. However, since the solvent site is only partially occupied, one must be cautious in over-discussing both these interactions and the Cl…H–C contacts shown in Figure S13.
It is instructive to compare the structures of {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n (this work) and {[Co(NCS)2(2a)]·4CHCl3}n [20]. These compounds both have 3,2′:6′,3″-tpy donor sets, but differ in the nature of the alkyloxy chain (4-phenylbutoxy versus n-octyloxy). A second difference is the solvent system for crystal growth (1,2-dichlorobenzene versus chloroform). Figure 4b illustrates an overlay of the {Co4(4)} and {Co4(2a)} units in {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n and {[Co(NCS)2(2a)]·4CHCl3}n, and reveals a significant difference in ligand conformation. This arises from the larger dihedral angles between the 3,2′:6′,3″-tpy ring with N2 and the phenylene ring (Table 1). As previously described, {[Co(NCS)2(2a)]·4CHCl3}n possesses a {42.84} lvt net. It is now clear that the assembly of the lvt net is not simply a consequence of introducing the 3,2′:6′,3″-tpy in place of the 4,2′:6′,4″-tpy domain. Unfortunately, we have not been able to obtain good quality X-ray diffraction data from crystals grown from a combination of Co(NCS)2 with 1b (Scheme 2) and are unable, therefore, to verify whether the switch from the cds to lvt net is a consequence of the steric demands of the long n-octyl chain.

3.4. [Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH

The discussion above demonstrates that consistent networks can be obtained with a combination of Co(NCS)2 and tetratopic ligands incorporating either 3,2′:6′,3″- or 4,2′:6′,4″-tpy metal-binding domains. The observations suggest that the nature of the alkyloxy substituents, rather than the choice of tpy isomer, may be the critical factor in determining the outcome of the coordination assembly. On going from ligand 3 to 5, we gain the potential for π-stacking without significantly lengthening the alkyloxy chain. Layering an MeOH solution of Co(NCS)2 over a CHCl3 solution of 5 produced only a few crystals, and there was insufficient material to obtain powder XRD data for the bulk sample. Single crystal X-ray diffraction revealed the unexpected formation of the discrete molecular coordination compound [Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH. The compound crystallizes in the triclinic space group P–1, and the centrosymmetric structure is shown in Figure 6 with selected bond parameters given in the figure caption. Each ligand 5 acts as a monodentate N-donor ligand and MeOH molecules complete the octahedral coordination sphere of Co1. While {Co(NCS)2(MeOH)2(py)2} (py = pyridine) motifs are well established [37,38,39,40,41,42], there is only one example in the CSD (v. 5.39 with updates [43]) featuring a monotopic 4,2′:6′,4″-tpy ligand 6 (Scheme 5) [44]. In this work, crystal growth by layering a MeOH/Co(NCS)2 solution over a solution of 6 in MeOH/CH2Cl2 with an intermediate MeOH/CH2Cl2 layer resulted in the formation of [Co(NCS)2(MeOH)2(6)2]. The preference for this discrete molecular assembly was attributed to the introduction of the sterically demanding pyrenyl group which is involved in efficient intermolecular pyrene…4,2′:6′,4″-tpy π-stacking interactions [44]. In the case of [Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH, packing interactions (Figure 7) involve centrosymmeric pairs of phenyl rings containing C20 and C20ii (symmetry code ii = 2 − x, 2 − y, 1 − z), pairs of 4,2′:6′,4″-tpy units containing N2/N4 and N2iii/N4iii (symmetry code iii = 1 − x, 2 − y, −z), and pairs of 4,2′:6′,4″-tpy units containing N6/N7 and N6iv/N7iv (symmetry code iv = 2 − x, −y, 2 − z). The first two sets of interactions lock together four molecules (Figure 7a). For the phenyl…phenyl π-stacking interaction, the separations of the ring planes and centroids are 3.35 and 3.65 Å, respectively. The coordinated 4,2′:6′,4″-tpy unit with N2/N3/N4 is virtually planar (dihedral angles between the ring planes = 7.4 and 6.0°), and 4,2′:6′,4″-tpys with N2/N4 and N2iii/N4iii engage in a π-stacking interaction (Figure 7a) with an angle between the ring planes of 6.0° and centroid…centroid distance of 3.60 Å. For the non-coordinated 4,2′:6′,4″-tpy unit, the dihedral angles between pairs of rings with N5/N6 and N6/N7 are 24.1 and 3.2°, respectively. Figure 7b depicts the face-to-face π-stacking interaction pairs of 4,2′:6′,4″-tpys containing N6/N7 and N6iv/N7iv (centroid…centroid = 3.83 Å). The overall packing in the lattice is, therefore, dominated by efficient π-stacking [45], but, because of the low yield of crystals, we are unable to say whether this leads to a preference for a discrete molecular structure in which three of the four potential metal binding sites of ligand 5 remain non-coordinated, or whether the isolation of crystals of [Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH is serendipitous.

4. Conclusions

We have prepared three new ligands containing two 4,2′:6′,4″-tpy or two 3,2′:6′,3″-tpy metal-binding domains with the expectation that they would function as 4-connecting nodes in coordination networks. The ligands possess different alkyloxy functionalities attached to the central phenylene spacer: n-pentyloxy in 3, 4-phenylbutoxy in 4, benzyloxy in 5. Reactions with Co(NCS)2 under ambient conditions with MeOH and 1,2-C6H4Cl2 as solvents resulted in the formation of {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n, both of which exhibit 3D cds nets. These data taken along with previously reported related structures suggests that the assembly of a particular net (cds or lvt) with 4-connecting Co and bis(tpy) ligands is independent of the choice of 4,2′:6′,4″- or 3,2′:6′,3″-tpy. The role of the alkyloxy substituent appears to be more significant. The combination of Co(NCS)2 with ligand 5 resulted in the formation of the discrete molecular complex [Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH in which 5 behaves as a monodentate ligand. The pendant phenyls and both coordinated and non-coordinated 4,2′:6′,4″-tpy units engage in efficient π-stacking interactions. However, we are unable to identify whether the isolation of the mononuclear complex is preferred over network assembly, or is a serendipitous result. Further systematic investigations are needed in order to delineate the assembly algorithms in these systems.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4360/10/12/1369/s1, Figure S1: Powder diffraction data for the bulk sample of {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n. Figures S2–S4: High resolution electrospray (HR-ESI) mass spectra of 3, 4 and 5. Figures S5–S10: 1H and 13C NMR spectra of 3, 4 and 5. Figure S11: Overlay of {Co4(3)} and {Co4(1d)} units in {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(1d)]·2C6H4Cl2}n. Figure S12: Topological representation of the 3D net in {[Co(NCS)2(3)]·0.8C6H4Cl2}n. Figure S13: Close contacts involving the 1,2-dichlorobenzene molecule in {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n.

Author Contributions

Y.M.K. and M.K.: synthetic work; A.P.: crystallography; E.C.C.: group leader, project concepts, contributions to manuscript preparation; C.E.H.: group leader, project concepts, manuscript preparation.

Funding

This research was funded by the Swiss National Science Foundation (Grant number 200020_162631) and the University of Basel.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Biradha, K.; Sarkar, M.; Rajput, L. Crystal engineering of coordination polymers using 4,4′-bipyridine as a bond between transition metal ions. Chem. Commun. 2006, 4169–4179. [Google Scholar] [CrossRef]
  2. Ye, B.-H.; Tong, M.-L.; Chen, X.-M. Metal-organic molecular architectures with 2,2′-bipyridyl-like and carboxylate ligands. Coord. Chem Rev. 2005, 249, 545–565. [Google Scholar] [CrossRef]
  3. Constable, E.C.; Housecroft, C.E. Ligand and Metalloligand Design for Macrocycles, Multimetallic Arrays, Coordination Polymers and Assemblies. In Elsevier Reference Module in Chemistry, Molecular Sciences and Chemical Engineering; Reedijk, J., Ed.; Elsevier: Waltham, MA, USA, 2016. [Google Scholar]
  4. Byrne, J.P.; Kitchen, J.A.; Gunnlaugsson, T. The btp [2,6-bis(1,2,3-triazol-4-yl)pyridine] binding motif: A new versatile terdentate ligand for suprmolecualr and coordination chemistry. Chem. Soc. Rev. 2014, 43, 5302–5325. [Google Scholar] [CrossRef] [PubMed]
  5. Adarsh, N.N.; Dastidar, P. Coordination polymers: What has been achieved in going from innocent 4,4′-bipyridine to bis-pyridyl ligands having a non-innocent backbone? Chem. Soc. Rev. 2012, 41, 3039–3060. [Google Scholar] [CrossRef]
  6. Constable, E.C. Coordination Polymers. In Supramolecular Chemistry: From Molecules to Nanomaterials; Reedijk, J., Ed.; Elsevier: Waltham, MA, USA, 2016; Volume 6, p. 3073. [Google Scholar]
  7. Janiak, C.; Veith, J.K. MOFs, MILs and more: Concepts, properties and applications for porous coordination networks (PCNs). New J. Chem. 2010, 34, 2366–2388. [Google Scholar] [CrossRef]
  8. Constable, E.C. Expanded ligands–An assembly principle for supramolecular chemistry. Coord. Chem. Rev. 2008, 252, 842–855. [Google Scholar] [CrossRef]
  9. Chakroborty, S.; Newkome, G.R. Terpyridine-based metallosupramolecular constructs: Tailored monomers to precise 2D-motifs and 3D-metallocages. Chem. Soc. Rev. 2018, 47, 3991–4016. [Google Scholar] [CrossRef]
  10. Housecroft, C.E. 4,2′:6′,4′′-Terpyridines: Diverging and diverse building blocks in coordination polymers and metallomacrocycles. Dalton Trans. 2014, 43, 6594–6604. [Google Scholar] [CrossRef]
  11. Housecroft, C.E. Divergent 4,2′:6′,4′′- and 3,2′:6′,3′′-terpyridines as linkers in 2- and 3-dimensional architectures. CrystEngComm 2015, 17, 7461–7468. [Google Scholar] [CrossRef] [Green Version]
  12. Constable, E.C.; Housecroft, C.E. Tetratopic bis(4,2′:6′,4′′-terpyridine) and bis(3,2′:6′,3′′-terpyridine) ligands as 4-connecting nodes in 2D-coordination networks and 3D-frameworks. J. Inorg. Organomet. Polym. Mater. 2018, 28, 414–427. [Google Scholar] [CrossRef]
  13. Klein, Y.M.; Prescimone, A.; Pitak, M.B.; Coles, S.J.; Constable, E.C.; Housecroft, C.E. Constructing chiral MOFs by functionalizing 4,2′:6′,4′′-terpyridine with long-chain alkoxy domains: Rare examples of neb nets. CrystEngComm 2016, 18, 4704–4707. [Google Scholar] [CrossRef]
  14. Liu, C.; Ding, Y.-B.; Shi, X.-H.; Zhang, D.; Hu, M.-H.; Yin, Y.-G.; Li, D. Interpenetrating Metal−Organic Frameworks Assembled from Polypyridine Ligands and Cyanocuprate Catenations. Cryst. Growth Des. 2009, 9, 1275–1277. [Google Scholar] [CrossRef]
  15. Yang, P.; Wang, M.-S.; Shen, J.-J.; Li, M.-X.; Wang, Z.-X.; Shao, M.; He, X. Seven novel coordination polymers constructed by rigid 4-(4-carboxyphenyl)-terpyridine ligands: Synthesis, structural diversity, luminescence and magnetic properties. Dalton Trans. 2014, 43, 1460–1470. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, L.; Li, C.-J.; He, J.-E.; Chen, Y.-Y.; Zheng, S.-R.; Fan, J.; Zhang, W.-G. Construction of New Coordination Polymers from 4′-(2,4-disulfophenyl)- 3,2′:6′,3′′-terpyridine: Polymorphism, pH-dependent syntheses, structures, and properties. J. Solid State Chem. 2016, 233, 444–454. [Google Scholar] [CrossRef]
  17. Cheng, Y.; Yang, M.-L.; Hu, H.-M.; Xu, B.; Wang, X.; Xue, G. Syntheses, structures and luminescence for zinc coordination polymers based on a multifunctional 4′-(3-carboxyphenyl)-3,2′:6′,3′′-terpyridine ligand. J. Solid State Chem. 2016, 239, 121–130. [Google Scholar] [CrossRef]
  18. Li, N.; Zhu, Q.-E.; Hu, H.-M.; Guo, H.-L.; Xie, J.; Wang, F.; Dong, F.-X.; Yang, M.-L.; Xue, G.-L. Hydrothermal syntheses, crystal structures and luminescence properties of zinc(II) coordination polymers constructed by bifunctional 4′-(4-carboxyphenyl)-3,2′:6′,3′′-terpyridine. Polyhedron 2013, 49, 207–215. [Google Scholar] [CrossRef]
  19. Chen, N.; Li, M.-X.; Yang, P.; He, X.; Shao, M.; Zhu, S.-R. Chiral coordination polymers with SHG-active and luminescence: An unusual homochiral 3D MOF constructed from achiral components. Cryst. Growth Des. 2013, 13, 2650–2660. [Google Scholar] [CrossRef]
  20. Klein, Y.M.; Constable, E.C.; Housecroft, C.E.; Prescimone, A. A 3-dimensional {42.84} lvt net built from a ditopic bis(3,2′:6′,3′′-terpyridine) tecton bearing long alkyl tails. CrystEngComm 2015, 17, 2070–2073. [Google Scholar] [CrossRef]
  21. Klein, Y.M.; Lanzilotto, A.; Prescimone, A.; Krämer, K.W.; Decurtins, S.; Liu, S.-X.; Constable, E.C.; Housecroft, C.E. Coordination behaviour of 1-(3,2′:6′,3′′-terpyridin-4′-yl)ferrocene: Structure and magnetic and electrochemical properties of a tetracopper dimetallomacrocycle. Polyhedron 2017, 129, 71–76. [Google Scholar] [CrossRef]
  22. Zhao, M.; Tan, J.; Su, J.; Zhang, J.; Zhang, S.; Wu, J.; Tian, Y. Syntheses, crystal structures and third-order nonlinear optical properties of two series of Zn(II) complexes using the thiophene-based terpyridine ligands. Dyes Pigments 2016, 130, 216–225. [Google Scholar] [CrossRef]
  23. Constable, E.C.; Housecroft, C.E.; Vujovic, S.; Zampese, J.A. 2D→2D Parallel interpenetration of (4,4) sheets constructed from a ditopic bis(4,2′:6′,4′′-terpyridine). CrystEngComm 2014, 16, 3494–3497. [Google Scholar] [CrossRef]
  24. Vujovic, S.; Constable, E.C.; Housecroft, C.E.; Morris, C.D.; Neuburger, M.; Prescimone, A. Engineering 2D→2D parallel interpenetration using long alkoxy-chain substituents. Polyhedron 2015, 92, 77–83. [Google Scholar] [CrossRef] [Green Version]
  25. Klein, Y.M.; Prescimone, A.; Neuburger, M.; Constable, E.C.; Housecroft, C.E. What a difference a tail makes: 2D→2D parallel interpenetration of sheets to interpenetrated nbo networks using ditopic-4,2′:6′,4′′-terpyridine ligands. CrystEngComm 2017, 19, 2894–2902. [Google Scholar] [CrossRef]
  26. Batten, S.R.; Neville, S.M.; Turner, D.R. Coordination Polymers: Design, Analysis and Application; RSC Publishing: Cambridge, UK, 2009; Chapter 2; ISBN 978-0-85404-837-3. [Google Scholar]
  27. Li, D.-S.; Wua, Y.-P.; Zhao, J.; Zhang, J.; Lu, J.Y. Metal-organic frameworks based upon non-zeotype 4-connected topology. Coord. Chem. Rev. 2014, 261, 1–27. [Google Scholar] [CrossRef]
  28. Klein, Y.M.; Prescimone, A.; Constable, E.C.; Housecroft, C.E. 4,2′:6′,4′′- and 3,2′:6′,3′′-terpyridines: The conflict between well-defined vectorial properties and serendipity in the assembly of 1D-, 2D- and 3D-architectures. Materials 2017, 10, 728. [Google Scholar] [CrossRef]
  29. Prasad, T.K.; Suh, M.P. Metal-organic frameworks incorporating various alkoxy pendant groups: Hollow tubular morphologies, X-ray single-crystal structures, and selective carbon dioxide adsorption properties. Chem. Asian J. 2015, 10, 2257–2263. [Google Scholar] [CrossRef]
  30. Kuhnert, N.; Lopez-Periago, A.; Rossignolo, G.M. The synthesis and conformation of oxygenated trianglimine macrocycles. Org. Biomol. Chem. 2005, 3, 524–537. [Google Scholar] [CrossRef]
  31. Bruker Analytical X-ray Systems, Inc. APEX2, version 2 User Manual, M86-E01078; Bruker Analytical X-ray Systems, Inc.: Madison, WI, USA, 2006. [Google Scholar]
  32. Palatinus, L.; Chapuis, G. SUPERFLIP—A computer program for the solution of crystal structures by charge flipping in arbitrary dimensions. J. Appl. Cryst. 2007, 40, 786–790. [Google Scholar] [CrossRef]
  33. Betteridge, P.W.; Carruthers, J.R.; Cooper, R.I.; Prout, K.; Watkin, D.J. CRYSTALS version 12: Software for guided crystal structure analysis. J. Appl. Cryst. 2003, 36, 1487. [Google Scholar] [CrossRef]
  34. Bruno, I.J.; Cole, J.C.; Edgington, P.R.; Kessler, M.K.; Macrae, C.F.; McCabe, P.; Pearson, J.; Taylor, R. New software for searching the Cambridge Structural Database and visualizing crystal structures. Acta Cryst. B 2002, 58, 389–397. [Google Scholar] [CrossRef] [Green Version]
  35. Macrae, C.F.; Bruno, I.J.; Chisholm, J.A.; Edgington, P.R.; McCabe, P.; Pidcock, E.; Rodriguez-Monge, L.; Taylor, R.; van de Streek, J.; Wood, P.A. Mercury CSD 2.0—New features for the visualization and investigation of crystal structures. J. Appl. Cryst. 2008, 41, 466–470. [Google Scholar] [CrossRef]
  36. Wang, J.; Hanan, G.S. A facile route to sterically hindered and non-hindered 4′-aryl-2,2′:6′,2′′-terpyridines. Synlett 2005, 1251–1254. [Google Scholar] [CrossRef]
  37. Lee, I.S.; Shin, D.M.; Chung, Y.K. Novel supramolecular isomerism in coordination polymer synthesis from unsymmetrical bridging ligands: Solvent influence on the ligand placement orientation and final network structure. Chem. Eur. J. 2004, 10, 3158–3165. [Google Scholar] [CrossRef]
  38. Suckert, S.; Jess, I.; Näther, C. Crystal structure of bis (3,5-di methyl pyridine-κN)bis (methanol-κO)bis (thiocyanato-κN)cobalt(II). Acta Crystallogr. Sect. E 2016, 72, 1824–1826. [Google Scholar] [CrossRef]
  39. Suckert, S.; Werner, J.; Jess, I.; Näther, C. Crystal structure of bis(4-benzoylpyridine-κN)bis(methanol-κO)bis(thiocyanato-κN)cobalt(II). Acta Crystallogr. Sect. E 2017, 73, 616–619. [Google Scholar] [CrossRef]
  40. Yang, G.-R.; Ren, J.; Li, G.-T. Bis{N(2),N(6)-bis-[(pyridin-3-yl)methyl]pyridine-2,6-dicarboxamide- κN}bis(methanol-κO)bis(thiocyanato-κN)cobalt(II). Acta Crystallogr. Sect. E 2012, 68, m765. [Google Scholar] [CrossRef]
  41. Robert, F.; Naik, A.D.; Tinant, B.; Robiette, R.; Garcia, Y. Insights into the origin of solid-state photochromism and thermochromism of N-salicylideneanils: The intruging case of aminopyridines. Chem. Eur. J. 2009, 15, 4327–4342. [Google Scholar] [CrossRef]
  42. Zhang, P.; Wu, B.-L.; Niu, Y.-Y.; Zhang, H.-Y.; Niu, C.-Y.; Hou, H.-W. Synthesis, crystal structure and properties of the enantiotopic complex constructed from chiral ligand H2bpb, [Co(H2bpb)2(NCS)2(CH3OH)2] (H2bpb = 1,2-bis(3-pyridylcarboxamide)benzene). Synth. React. Inorg. Met. Org. Nano-Met. Chem. 2007, 37, 577–582. [Google Scholar] [CrossRef]
  43. Groom, C.R.; Bruno, I.J.; Lightfoot, M.P.; Ward, S.C. The Cambridge Structural Database. Acta Crystallogr. Sect. B 2016, 72, 171–179. [Google Scholar] [CrossRef] [Green Version]
  44. Yin, Z.; Zhang, S.; Zheng, S.; Golen, J.A.; Rheingold, A.L.; Zhang, G. Cobalt(II) coordination polymers versus discrete complex with 4,2′:6′,4′′-terpyridine ligands: The role of a pyrenyl substituent. Polyhedron 2015, 101, 139–145. [Google Scholar] [CrossRef]
  45. Janiak, C. A critical account on π-π stacking in metal complexes with aromatic nitrogen-containing ligands. J. Chem. Soc. Dalton Trans. 2000, 3885–3896. [Google Scholar] [CrossRef]
Scheme 1. An [M(tpy)2]n+ complex is considered as an expanded 4,4′-bpy.
Scheme 1. An [M(tpy)2]n+ complex is considered as an expanded 4,4′-bpy.
Polymers 10 01369 sch001
Scheme 2. Structures of 4,2′:6′,4″- and 3,2′:6′,3″-tpy of general ligand types 1 (with 4,2′:6′,4″-tpy units) and 2 (with 3,2′:6′,3″-tpy units), and structures of ligands 1a1d and 2a. The green arrows show the vectoral orientation of the lone pairs involved in formation of the coordination network.
Scheme 2. Structures of 4,2′:6′,4″- and 3,2′:6′,3″-tpy of general ligand types 1 (with 4,2′:6′,4″-tpy units) and 2 (with 3,2′:6′,3″-tpy units), and structures of ligands 1a1d and 2a. The green arrows show the vectoral orientation of the lone pairs involved in formation of the coordination network.
Polymers 10 01369 sch002
Scheme 3. Structures of ligands 35.
Scheme 3. Structures of ligands 35.
Polymers 10 01369 sch003
Scheme 4. Synthetic routes to 3–5 with numbering for NMR spectroscopic assignments in the experimental section; CH2 units in alkyloxy chains are labelled a, b, c… etc. starting at the OCH2 unit and the terminal phenyl rings are labelled D. Conditions: (i) RBr or benzyl chloride (see Materials and Methods), anhydrous K2CO3, dry DMF, 100 °C, 15 h; (ii) nBuLi, Et2O, 0 °C; DMF, warmed to room temperature, 15 h; (iii) 4-acetylpyridine, KOH, EtOH, aqueous NH3, room temperature, 15 h; (iv) 3-acetylpyridine, KOH, EtOH, aqueous NH3, room temperature, 15 h.
Scheme 4. Synthetic routes to 3–5 with numbering for NMR spectroscopic assignments in the experimental section; CH2 units in alkyloxy chains are labelled a, b, c… etc. starting at the OCH2 unit and the terminal phenyl rings are labelled D. Conditions: (i) RBr or benzyl chloride (see Materials and Methods), anhydrous K2CO3, dry DMF, 100 °C, 15 h; (ii) nBuLi, Et2O, 0 °C; DMF, warmed to room temperature, 15 h; (iii) 4-acetylpyridine, KOH, EtOH, aqueous NH3, room temperature, 15 h; (iv) 3-acetylpyridine, KOH, EtOH, aqueous NH3, room temperature, 15 h.
Polymers 10 01369 sch004
Figure 1. Aromatic regions of the 1H NMR spectra of ligands 35 (500 MHz, CDCl3, 298 K). See Scheme 4 for atom labels. * = residual CHCl3.
Figure 1. Aromatic regions of the 1H NMR spectra of ligands 35 (500 MHz, CDCl3, 298 K). See Scheme 4 for atom labels. * = residual CHCl3.
Polymers 10 01369 g001
Figure 2. The asymmetric unit in {[Co(NCS)2(3)]·0.8C6H4Cl2}n and symmetry-generated atoms to show ligand 3 as a 4-connecting node. Symmetry codes: i = −1 − x, 1 – y, 1 − z; ii = −1 + x, y, 1 + z; iii = −1 + x, 3/2y, ½ + z; iv = −x, 1 − y, −z; v = x, ½ − y, −1/2 + z; vi = −x, ½ + y, ½ – z; vii = x, ½ − y, ½ + z.
Figure 2. The asymmetric unit in {[Co(NCS)2(3)]·0.8C6H4Cl2}n and symmetry-generated atoms to show ligand 3 as a 4-connecting node. Symmetry codes: i = −1 − x, 1 – y, 1 − z; ii = −1 + x, y, 1 + z; iii = −1 + x, 3/2y, ½ + z; iv = −x, 1 − y, −z; v = x, ½ − y, −1/2 + z; vi = −x, ½ + y, ½ – z; vii = x, ½ − y, ½ + z.
Polymers 10 01369 g002
Figure 3. The asymmetric unit in {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n and symmetry-generated atoms to show ligand 4 as a 4-connecting node. Symmetry codes: i = −1 − x, 2 − y, 1 − z; ii = −1 + x, 1 + y, z; iii = −1 − x, ½ + y, ½ − z; iv = −x, 1 − y, 1 − z; v = −x, −1/2 + y, 3/2z; vi = x, 3/2y, −1/2 + z; vii = −x, ½ + y, 3/2z.
Figure 3. The asymmetric unit in {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n and symmetry-generated atoms to show ligand 4 as a 4-connecting node. Symmetry codes: i = −1 − x, 2 − y, 1 − z; ii = −1 + x, 1 + y, z; iii = −1 − x, ½ + y, ½ − z; iv = −x, 1 − y, 1 − z; v = −x, −1/2 + y, 3/2z; vi = x, 3/2y, −1/2 + z; vii = −x, ½ + y, 3/2z.
Polymers 10 01369 g003
Figure 4. (a) Overlay of {Co4(3)} and {Co4(4)} units in {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n; (b) Overlay of {Co4(4)} and {Co4(2a)} units in {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n (this work) and {[Co(NCS)2(2a)]·4CHCl3}n [20]. In each overlay, the atoms of the central phenylene ring superimposed. For clarity, only the O atoms of the alkyloxy chains are shown.
Figure 4. (a) Overlay of {Co4(3)} and {Co4(4)} units in {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n; (b) Overlay of {Co4(4)} and {Co4(2a)} units in {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n (this work) and {[Co(NCS)2(2a)]·4CHCl3}n [20]. In each overlay, the atoms of the central phenylene ring superimposed. For clarity, only the O atoms of the alkyloxy chains are shown.
Polymers 10 01369 g004
Figure 5. Topological representation of the 3D net in {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n generated using Mercury [34,35] with 4-connecting cobalt and ligand nodes. The net in {[Co(NCS)2(3)]·0.8C6H4Cl2}n is the same (see Figure S12).
Figure 5. Topological representation of the 3D net in {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n generated using Mercury [34,35] with 4-connecting cobalt and ligand nodes. The net in {[Co(NCS)2(3)]·0.8C6H4Cl2}n is the same (see Figure S12).
Polymers 10 01369 g005
Figure 6. Molecular structure of [Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH with solvent molecules and H atoms omitted. Selected bond distances and angles: Co1–O1 = 2.138(3), Co1–N1 = 2.090(4), Co1–N2 = 2.167(3) Å, O1–Co1–N2 = 86.80(12), O1–Co1–N1 = 88.39(15), N2–Co1–N1 = 88.06(13), N1–Co1–N2i = 91.94(13), N1–Co1–O1i = 91.61(15), O1–Co1–N2i = 93.20(12)°. Symmetry code i = 1 − x, 3 − y, −z.
Figure 6. Molecular structure of [Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH with solvent molecules and H atoms omitted. Selected bond distances and angles: Co1–O1 = 2.138(3), Co1–N1 = 2.090(4), Co1–N2 = 2.167(3) Å, O1–Co1–N2 = 86.80(12), O1–Co1–N1 = 88.39(15), N2–Co1–N1 = 88.06(13), N1–Co1–N2i = 91.94(13), N1–Co1–O1i = 91.61(15), O1–Co1–N2i = 93.20(12)°. Symmetry code i = 1 − x, 3 − y, −z.
Polymers 10 01369 g006
Scheme 5. The structure of ligand 6 from reference [44].
Scheme 5. The structure of ligand 6 from reference [44].
Polymers 10 01369 sch005
Figure 7. Packing interactions in [Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH (solvent molecules omitted) (a) between pairs of phenyl rings containing C20 and C20ii (symmetry code ii = 2 − x, 2 − y, 1 − z) and pairs of 4,2′:6′,4″-tpy units containing N2/N4 and N2iii/N4iii (symmetry code iii = 1 − x, 2 − y, −z) and (b) pairs of 4,2′:6′,4″-tpy units containing N6/N7 and N6iv/N7iv (symmetry code iv = 2 − x, −y, 2 − z).
Figure 7. Packing interactions in [Co(NCS)2(MeOH)2(5)2]·2CHCl3·2MeOH (solvent molecules omitted) (a) between pairs of phenyl rings containing C20 and C20ii (symmetry code ii = 2 − x, 2 − y, 1 − z) and pairs of 4,2′:6′,4″-tpy units containing N2/N4 and N2iii/N4iii (symmetry code iii = 1 − x, 2 − y, −z) and (b) pairs of 4,2′:6′,4″-tpy units containing N6/N7 and N6iv/N7iv (symmetry code iv = 2 − x, −y, 2 − z).
Polymers 10 01369 g007
Table 1. Important bond parameters in {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n.
Table 1. Important bond parameters in {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n.
Bond Parameter{[Co(NCS)2(3)]·0.8C6H4Cl2}n{[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n
Distance/Å
Co1–N12.208 (2)2.238 (4)
Co1–N4 2.062 (3)2.065 (4)
Co1–N3v2.179 (2)2.169 (4)
Angle/deg
N1–Co1–N490.18 (10)88.67 (15)
N3v–Co1–N184.73 (9)93.81 (17)
N3v–Co1–N490.20 (11)90.32 (16)
N3vi–Co1–N195.27 (9)86.19 (17)
N3vi–Co1–N489.80 (11)89.68 (16)
N1iv–Co1–N489.82 (10)91.33 (15)
Table 2. Dihedral angles between pairs of bonded aromatic rings in {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n, and in {[Co(NCS)2(1d)]·2C6H4Cl2}n [28] and {[Co(NCS)2(2a)]·4CHCl3}n [20].
Table 2. Dihedral angles between pairs of bonded aromatic rings in {[Co(NCS)2(3)]·0.8C6H4Cl2}n and {[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n, and in {[Co(NCS)2(1d)]·2C6H4Cl2}n [28] and {[Co(NCS)2(2a)]·4CHCl3}n [20].
CompoundDihedral Angle between Planes/Deg
Ring with N1/Ring with N2Ring with N2/Ring with N3Ring with N2/Phenylene Ring
{[Co(NCS)2(3)]·0.8C6H4Cl2}n25.330.740.2
{[Co(NCS)2(4)]·1.6H2O·1.2C6H4Cl2}n 34.66.444.3
{[Co(NCS)2(1d)]·2C6H4Cl2}n19.531.340.5
{[Co(NCS)2(2a)]·4CHCl3}n29.9; 22.712.7; 12.2 58.5; 55.9

Share and Cite

MDPI and ACS Style

Klein, Y.M.; Prescimone, A.; Karpacheva, M.; Constable, E.C.; Housecroft, C.E. Sometimes the Same, Sometimes Different: Understanding Self-Assembly Algorithms in Coordination Networks. Polymers 2018, 10, 1369. https://doi.org/10.3390/polym10121369

AMA Style

Klein YM, Prescimone A, Karpacheva M, Constable EC, Housecroft CE. Sometimes the Same, Sometimes Different: Understanding Self-Assembly Algorithms in Coordination Networks. Polymers. 2018; 10(12):1369. https://doi.org/10.3390/polym10121369

Chicago/Turabian Style

Klein, Y. Maximilian, Alessandro Prescimone, Mariia Karpacheva, Edwin C. Constable, and Catherine E. Housecroft. 2018. "Sometimes the Same, Sometimes Different: Understanding Self-Assembly Algorithms in Coordination Networks" Polymers 10, no. 12: 1369. https://doi.org/10.3390/polym10121369

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