Next Article in Journal
Crystal Structure of the 5-Chloro Salicylamides: Three Different Types of the H-bonding Influenced Linear Chain Formation in the Solid State
Next Article in Special Issue
Polyoxotungstate-Surfactant Layered Crystal toward Conductive Inorganic-Organic Hybrid
Previous Article in Journal / Special Issue
Halogen Bonding Interactions in DDQ Charge Transfer Salts with Iodinated TTFs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electroactive Bisiminopyridine Ligands: Synthesis and Complexation Studies

CIMI group, Laboratoire MOLTECH-Anjou, CNRS UMR 6200, Université d’Angers, LUNAM Université, 2 bd Lavoisier, 49045 ANGERS cedex, France
*
Authors to whom correspondence should be addressed.
Crystals 2012, 2(2), 338-348; https://doi.org/10.3390/cryst2020338
Submission received: 22 March 2012 / Revised: 6 April 2012 / Accepted: 9 April 2012 / Published: 24 April 2012
(This article belongs to the Special Issue Molecular Conductors)

Abstract

:
The condensation reaction of (4-(6,7-dimethyldithio-tetrathiafulvalene)-aniline) with 2,6-diformylpyridine afforded an electroactive Schiff base (N,N,N) pincer (3). This pincer was reacted with Zn(II) cation to yield the corresponding Zinc chloride complex (4). The crystal structure of the newly prepared electroactive zinc complex reveals that the tetrathiafulvalene (TTF) is neutral and the zinc cation is pentacoordinated. The two chlorines are involved in a set of hydrogen bonds giving rise to a 2D supramolecular grid arrangement. The electronic absorption properties and the electrochemical behavior have been elucidated. These two compounds are promising for the construction of crystalline radical cation salts.

1. Introduction

Tetrathiafulvalene (TTF) and its derivatives have attracted much interest because of their electron-donating ability and their attractive reversible redox properties. They have therefore been widely used as donor components in the preparation of molecular conductors and superconductors [1,2,3]. However, one challenge is the construction of multifunctional molecular materials that can exhibit interplay between one or more physical properties such as magnetic properties and electrical conductivity [4,5]. In order to address this challenge, considerable effort has been devoted in associating the TTF moiety with an inorganic component and various organic mono- or polydentate ligands, and their corresponding electroactive metal complexes have therefore been reported [6]. For example, metal complexes of phosphines [7,8,9,10,11,12], dithiolates [13], acetylacetonates [14], pyridines [15,16,17], bipyridines [18,19] and more recently of Schiff bases [20,21,22] have been synthesized. Pyridine based Schiff bases ligands such as 2,6-bis(imino)pyridyl, with chelating abilities, form stable complexes, with various transition metals, which have been extensively used as catalysts for olefin polymerization [23,24].
In order to obtain electroactive transition metal complexes with original structural and electronic properties, we report herein the synthesis of the TTF2-BisIm-Py (3) that can act as an (N,N,N) pincer. The TTF based donor with monoaniline moiety, (4-(6,7-dimethyldithio-tetrathiafulvalene)-aniline) (2), has been chosen for the Schiff base condensation reaction with 2,6-diformylpyridine. The reaction of this TTF based tridentate pincer with zinc metal cation afforded a new electroactive metal complex.

2. Results and Discussion

2.1. Synthesis and Characterization

The protocol followed for the synthesis of the new (2,6-bis(4-(4',5'-bis(methylthio)-[2,2'-bi(1,3-dithiolylidene)]-4-yl)aniline)-iminopyridine) pincer (3, TTF2-BisIm-Py) is presented in Figure 1.
Figure 1. Synthesis scheme of the TTF2-BisIm-Py Pincer (3).
Figure 1. Synthesis scheme of the TTF2-BisIm-Py Pincer (3).
Crystals 02 00338 g001
The starting stannic derivative (6,7-dimethyldithio-2-trimethyltin-tetrathiafulvalene) was synthesized as previously described [25] and was reacted with 1-iodo-4-nitrobenzene under Stille coupling conditions to afford (1-(6,7-dimethyldithio-tétrathiafulvalene)-4-nitro-benzene) (1) in 67% yield [26]. Reduction of the nitro group to the amine was achieved using a mixture of HCl ethanolic solution in the presence of metallic tin [26]. The obtained amine (2) was further used in a condensation reaction with 2,6-diformylpyridine to afford the TTF2-BisIm-Py pincer (3).
The TTF2-BisIm-Py pincer was reacted with one equivalent of zinc chloride (ZnCl2) in a dicloromethane/acetonitrile solvents mixture to afford good quality single crystals of [Zn(TTF2-BisIm-Py) Cl2] (4) (Figure 2) that were suitable for the X-ray crystal structure determination.
Figure 2. Schematic drawing of the zinc metal complex (4).
Figure 2. Schematic drawing of the zinc metal complex (4).
Crystals 02 00338 g002
Single-crystal diffraction analysis revealed that the asymmetric unit (Figure 3) contains half of the molecule of complex (4). Within this complex the zinc ion is pentacoordinated, the coordination sphere being formed by the three nitrogen atoms of the pincer and two chlorines ions. For a more accurate description of the stereochemistry around the zinc ion, the τ parameter (τ = [(αα')/60] [27], where α and α' are the transoid angles formed by the metal ion and the donor atoms within the basal plane), also known as Adisson’s parameter, was calculated to be τ = 0.47. This indicates an intermediate between the two geometries corresponding number of five, square pyramidal and trigonal bipyramidal.
Figure 3. The crystal structure of complex 4 with the labeling scheme of the atoms. Hydrogen atoms have been omitted for clarity.
Figure 3. The crystal structure of complex 4 with the labeling scheme of the atoms. Hydrogen atoms have been omitted for clarity.
Crystals 02 00338 g003
In the coordination polyhedron, Zn–(N/Cl)ligand distances are in the 2.052(5)–2.351(3) Å range. A selection of relevant bond lengths and angles are given in Table 1. In the complex, the planarity of the ligand is disrupted by the TTF-(SMe)2 units, as proven by the value of the dihedral angle least-squares planes of the pyridyl group and the dithiole ring (23.55°). In the solid state, each of the two chlorine atoms coordinated to the zinc ion are involved in non-conventional hydrogen bonds with hydrogen atoms from other two neighboring molecules to generate a grid-like sheet represented in Figure 4.
Table 1. Selected bond lengths (Å) and angles (°) for compound 4.
Table 1. Selected bond lengths (Å) and angles (°) for compound 4.
bondlength (Å)bondlength (Å)bondlength (Å)
C11–C121.330(6)C14–S41.761(5)N1–Zn12.052(5)
C11–S11.774(4)C15–C161.345(6)N2–Zn12.351(3)
C12–S21.720(4)C15–S61.743(4)Cl1–Zn12.2345(13)
C13–C141.340(5)C15–S31.769(4)Zn1–Cl1 i2.2345(13)
C13–S21.756(4)C16–S51.742(4)Zn1–N2 i2.351(3)
C13–S11.759(4)C18–S51.794(5)H10–Cl1 i3.1536(15)
C14–S31.758(4)N1–C1i1.342(5)Cl1–Cl1 i3.8584(19)
anglevalue (°)anglevalue (°)anglevalue (°)
N1–Zn1–Cl1120.30(4)N1–Zn1–N2 i74.31(9)Cl1–Zn1–N297.74(9)
N1–Zn1–Cl1 i120.30(4)Cl1–Zn1–N2 i97.94(9)Cl1 i–Zn1–N297.94(9)
Cl1–Zn1–Cl1 i119.40(9)N1–Zn1–N274.31(9)N2 i–Zn1–N2148.63(18)
i 2 − x, y, 0.5 − z.
Figure 4. Front view of the 2D supramolecular layer in 4 reinforced by non-conventional hydrogen bonds (2.69 Å) (red dashed lines) and π–π stacking (highlighted by the blue circle). The TTF-(SMe)2 units were omitted for clarity.
Figure 4. Front view of the 2D supramolecular layer in 4 reinforced by non-conventional hydrogen bonds (2.69 Å) (red dashed lines) and π–π stacking (highlighted by the blue circle). The TTF-(SMe)2 units were omitted for clarity.
Crystals 02 00338 g004
Within this supramolecular layer developed in the ab crystallographic plane, each of the three six-membered aromatic rings of the ligand are stacked in a face to face manner (centroid···centroid 3.56 Å) with two other rings forming the triads highlighted in Figure 5. The crystal packing (Figure 5) reveals a segregation between the TTF units and the rest of the molecule that can be associated with the presence of numerous sulfur-sulfur interactions (intralayer S···S: S2···S3 3.59 Å, S2···S6 3.66 Å, S4···S6 3.58 Å) reinforcing the above-mentioned layers but also connecting them into a 3D network (interlayer S···S: S3···S5 3.73 Å, S5···S5 3.30 Å).
Figure 5. The crystal packing in 4 showing the non-conventional hydrogen bonds (red dashed lines) and S···S (yellow dashed lines) interactions.
Figure 5. The crystal packing in 4 showing the non-conventional hydrogen bonds (red dashed lines) and S···S (yellow dashed lines) interactions.
Crystals 02 00338 g005

2.2. UV-Vis Spectroscopy

The electronic absorption spectra of the TTF2-BisIm-Py pincer (3) and of the [Zn(TTF2-BisIm-Py) Cl2] (4) complex were recorded in dichloromethane solution (~2·105 M at room temperature; Figure 6). The TTF2-BisIm-Py pincer exhibits two strong electronic absorption bands at λ = 269 nm and 336 nm which are assigned to the π→π* absorption bands of the TTF and the phenyl ring. The broad band observed in the visible region at λmax = 433 nm is characteristic of the intramolecular charge transfer transition (ICT) from the highest occupied molecular orbital in the two TTFs to the lowest unoccupied molecular orbital in the electron-accepting pyridyl unit [20,28] and is responsible for the dark red color of this compound. Upon complexation of the pincer with zinc chloride (ZnCl2), the electronic absorption spectrum presents the same features as the free ligand in the high energy region (<400 nm). However, the ICT transition is red shifted by about 80 nm indicating some effect on the electron acceptor behavior of this ligand upon complexation of zinc chloride. In addition, a shoulder (around 385 nm) is appearing, which is probably due to conformational change (all trans to all cis) of the pincer after complexation with zinc chloride.
Figure 6. Electronic absorption spectra of the pincer 3 and the zinc complex 4.
Figure 6. Electronic absorption spectra of the pincer 3 and the zinc complex 4.
Crystals 02 00338 g006

2.3. Electrochemistry

The electrochemical behaviors of the new pincer (3) and of the new complex (4) were examined by cyclic voltammetry (CV). The CV was performed in a three-electrode cell equipped with a platinum millielectrode, a platinum wire counter-electrode and a silver wire used as quasi-reference electrode. The electrochemical experiments were carried out under dry and oxygen-free atmosphere (H2O < 1 ppm, O2 < 1 ppm) CH3CN (5.10−4 M) with Bu4NPF6 (TBAP) (0.1 M) as supporting electrolyte. The voltammograms were recorded on an EGG PAR 273A potentiostat with positive feedback compensation. Based on repetitive measurements, absolute errors on potentials were estimated around ±5 mV.
Cyclic voltammetry measurements (Figure 7) in the case of the pincer TTF2-BisIm-Py (3) show two reversible oxidations waves (at E1/21 = 0.400 V and E1/22 = 0.758 V vs. SCE). These potentials are anodically shifted from those of the free TTF (0.37 V vs. SCE) [1c)] because of the presence of the electrodeficient pyridine ring. The two oxidation bands, being broad, suggest very small differences between the oxidation potentials of the two TTF units present in the molecule. This behavior is similar to the one observed for bis(TTF) donors, described earlier by Avarvari et al. [29], separated by a triazine ring.
Figure 7. Cyclic voltammograms of both the pincer (3) (left) and of the zinc complex (4) (right) (2 × 10−5 M) in CH3CN with nBu4NPF6 (0.1 M).
Figure 7. Cyclic voltammograms of both the pincer (3) (left) and of the zinc complex (4) (right) (2 × 10−5 M) in CH3CN with nBu4NPF6 (0.1 M).
Crystals 02 00338 g007
After complexation, the oxidation potentials of the zinc metal complex (4) are slightly positively shifted (about 20 mV and 25 mV), suggesting that the zinc metal cation is affecting the oxidation potential of the TTF moieties by decreasing the electron density on the TTF units, a fact also suggested by the electronic absorption spectra. As a consequence of these electrochemical behaviors, we expect radical cation salts of both the TTF pincer and the corresponding zinc metal complex to form readily upon electrochemical oxidation into air-stable crystals [15].

3. Experimental Section

3.1. X-Ray Structure Determinations

Details about data collection and solution refinement are given in Table 2. X-ray diffraction measurements were performed on a Bruker Kappa CCD diffractometer, operating with a Mo-Kα (λ = 0.71073 Å) X-ray tube with a graphite monochromator. The structures were solved (SHELXS-97) by direct methods and refined (SHELXL-97) by full matrix least-square procedures on F2. All non-hydrogen atoms were refined anisotropically. Hydrogen atoms were introduced at calculated positions (riding model), included in structure factor calculations but not refined. CCDC reference number: 871185 for 4. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.
Table 2. Crystal Data and Structure Refinement for compound 4.
Table 2. Crystal Data and Structure Refinement for compound 4.
complex4
empirical formula C35H27Cl2N3S12Zn
fw 1010.59
T (K) 293(2)
wavelength (Å)0.71073
crystal systemmonoclinic
space group C2/c
unit cell dimensions:
a (Å) 11.8152(16)
b (Å) 10.100(2)
c (Å) 35.579(6)
α(deg) 90
β(deg) 98.111(13)
γ(deg) 90
V3) 4202.2(12)
Z4
Dc (g cm−3) 1.597
abs coeff (mm−1) 1.340
θ range for data collection (deg) 1.16–27.54
reflections collected 4760
indep reflns3371
completeness (%) 98.5
data/restraints/param4760/0/243
GOF on F 21.197
final R indices [I > 2σ(I)] R1 = 0.0597, wR2 = 0.1051
R indices (all data) R1 = 0.0968, wR2 = 0.1227
largest diff. peak and hole (e Å−3) 0.394 and −0.416

3.2. General

All the reactions were carried out under Argon and using HPLC solvents. Nuclear magnetic resonance spectra were recorded on a Bruker Avance DRX 500 spectrometer (operating at 500.04 MHz for 1H and 125 MHz for 13C) and Bruker Avance DRX 300 automatic spectrometer (operating at 300 MHz for 1H, and 75 MHz for 13C). Chemical shifts are expressed in parts per million (ppm) downfield from external TMS. The following abbreviations are used: s, singlet; d, doublet; t, triplet. MALDI- TOF MS spectra were recorded on Bruker Biflex-IIITM apparatus, equipped with a 337 nm N2 laser. Elemental analyses were recorded using a Flash 2000 Fisher Scientific Thermo Electron analyzer.

3.3. Synthesis of (1-(6,7-Dimethyldithio-tétrathiafulvalene)-4-nitro-benzene) (1)

1-Iodo-4-nitro-benzene (0.44 g, 1.78 mmol), 4-(trimethylstannyl)-4',5'-bis(methylthio)-tetrathiafulvalene (0.98 g, 2.14 mmol), and Pd(PPh3)4 as catalyst (10% mmol) were mixed in 40 mL of dry toluene and refluxed overnight under argon. The mixture was filtered over Celite and silica, washed with toluene and dichloromethane and dried under vacuum. The residue was purified by column chromatography (pentane/dichloromethane 1/1) to give 0.50 g of dark blue solid, yield: 67%. 1H-NMR (300 MHz, CDCl3) δ/ppm: 8.22 (d, 2H, J3 = 9 Hz), 7.53 (d, 2H, J3 = 9 Hz), 6.78 (s, 1H), 2.45 (s, 3H), 2.44 (s, 3H); 13C-NMR (75 MHz, CDCl3) δ/ppm: 147.0, 138.0, 133.7, 127.7, 126.6, 124.3, 118.3, 112.2, 109.6, 19.2; MS (MALDI-TOF): m/z: 416.9 (Mth = 416.91)

3.4. Synthesis of (4-(6,7-Dimethyldithio-tétrathiafulvalene)-aniline) (2)

To a solution of compound 1 (0.25 g, 0.59 mmol in 20 mL ethanol), Sn (0.14 g, 1.2 mmol) and HCl 35% (0.31 mL) were added and the mixture was refluxed for 4 h under argon. After the removal of the solvent, the crude was solubilised in ethyl acetate and washed several times with a NaOH solution (1 M) and water. The residue was purified by column chromatography with dichloromethane as eluent to give 0.23 g of an orange solid, yield: 99%. 1H-NMR (300 MHz, CDCl3) δ/ppm: 7.18 (d, 2H, J3 = 8.7 Hz), 6.62 (d, 2H, J3 = 8.4 Hz), 6.26 (s, 1H), 3.79 (broad s, 2H) 2.44 (s, 3H), 2.43 (s, 3H); 13C-NMR (75 MHz, CDCl3) δ/ppm:146.9, 136.2, 127.6, 127.5, 122.8, 115.4, 114.9, 109.4, 106.3, 29.7, 19.2; MS (MALDI-TOF): m/z: 386.9 (Mth = 386.94).

3.5. (2,6-Bis(4-(4',5'-bis(Methylthio)-[2,2'-bi(1,3-dithiolylidene)]-4-yl)aniline)-iminopyridine) (3)

Compound 2 (0.115 g, 0.3 mmol) was solubilised in 50 mL ethanol, 2,6-Pyridinedicarboxaldehyde (0.019 g, 0.14 mmol) was added, followed by 3 drops of acetic acid and the mixture was refluxed overnight. After cooling the reaction, a red precipitate was formed which was filtered, washed with additional ethanol and dried to afford 116 mg of a red solid, yield: 94%. 1H-NMR (300 MHz, CDCl3) δ/ppm: 8.67 (s, 2H), 8.29 (d, 2H, J3 = 7.7 Hz), 7.95 (t, 1H, J3 = 7.7 Hz), 7.46 (d, 4H, J3 = 8.4 Hz), 7.31 (d, 4H, J3 = 8.4 Hz), 6.55 (s, 2H), 2.45 (s, 6H), 2.44 (s, 6H); 13C-NMR (75 MHz, CDCl3) δ/ppm:160.2, 154.4, 150.6, 135.4, 130.9, 127.2, 123.5, 121.7, 113.2, 19.2; λmax(CH2Cl2)/nm 269 (ε/dm3 mol1 cm1 49260), 336 (60870), 433 (12670). Selected IR bands (cm−1): 2916, 1621, 1567, 1454, 1411, 1212, 958, 835, 769. MS (MALDI-TOF): m/z: 873.6 (Mth = 872.89).

3.6. (2,6-Bis(4-(4',5'-bis(methylthio)-[2,2'-bi(1,3-dithiolylidene)]-4-yl)aniline)-iminopyridine) Zinc Chloride (4)

In a test tube, a solution of the pincer 3 (20 mg, 0.023 mmol) in CH2Cl2 (5 mL) was mixed with a solution of ZnCl2 (3.1 mg, 0.023 mmol) in CH3CN (3 mL) and ultrasonicated for 2 min. On top of the resulting solution a layer of 1:1 diethyl-ether mixture was added which led to the formation of good quality single crystals of complex 3 after 1 week. Yield: 18 mg (77%). λmax(CH2Cl2)/nm 267 (ε/dm3 mol1 cm1 40750), 333 (42040), 385 sh, 550 (7290). Selected IR bands (cm−1): 2916, 1622, 1568, 1457, 1418, 1215, 966, 888, 773.

4. Conclusions

As a first step towards the realization of multifunctional materials, a new TTF based pincer has been synthesized and its chelating ability has been investigated in respect to transition metal cations such as Zn(II). The crystal structure of the newly prepared electroactive zinc complex reveals that the TTF is neutral and the zinc cation is pentacoordinated. The two chlorines are involved in a set of hydrogen bonds giving rise to a 2D supramolecular grid arrangement. The electrochemical behavior of both the ligand and the neutral metal complex suggests that these compounds are useful candidates for the preparation of chemically and/or electrochemically oxidized radical cation crystalline salts. Further studies will be performed in order to investigate the complexation ability of this novel electroactive pincer on other metal cations such as Cu(II), Fe(II), Co(II), etc.

Acknowledgments

Financial support from the CNRS, the University of Angers, the Region Pays de la Loire (grant for D.B.) and from the National Agency for Research (ANR, Project 09-BLAN-0045-01) is gratefully acknowledged. ERASMUS program exchange (between University of Angers and Technical Military Academy of Bucharest) is also acknowledged (grant for G.N.).

Conflict of Interest

The authors declare no conflict of interest.

References

  1. Williams, J.M.; Ferraro, J.R.; Thorn, R.J.; Carlson, K.D.; Geiser, U.; Wang, H.H.; Kini, A.M.; Whangbo, M.-H. Organic Superconductors (Including Fullerenes), Synthesis, Structure, Properties and Theory; Prentice Hall: Upper Saddle River, NJ, USA, 1992. [Google Scholar]
  2. Bryce, M.R.; Murphy, L.C. Organic metals. Nature 1984, 309, 119–126. [Google Scholar]
  3. Segura, J.L.; Martin, N. New concepts in tetrathiafulvalene chemistry. Angew. Chem. Int. Ed. 2001, 40, 1372–1409. [Google Scholar] [CrossRef]
  4. Coronado, E.; Galan-Marcos, J.R.; Gomez-Garcia, C.J.; Laukhin, V. Coexistence of ferromagnetism and metallic conductivity in a molecule-based layered compound. Nature 2000, 408, 447–449. [Google Scholar]
  5. Coronado, E.; Day, P. Magnetic molecular conductors. Chem. Rev. 2004, 104, 5419–5448. [Google Scholar] [CrossRef]
  6. Lorcy, D.; Bellec, N.; Fourmigué, M.; Avarvari, N. Tetrathiafulvalene-based group XV ligands: Synthesis, coordination chemistry and radical cation salts. Coord. Chem. Rev. 2009, 253, 1398–1438. [Google Scholar] [Green Version]
  7. Avarvari, N.; Martin, D.; Fourmigué, M. Structural and electrochemical study of metal carbonyl complexes with chelating bis- and tetrakis(diphenylphosphino)tetrathiafulvalenes. J. Organomet. Chem. 2002, 643/644, 292–300. [Google Scholar]
  8. Pellon, P.; Gachot, G.; Le Bris, J.; Marchin, S.; Carlier, R.; Lorcy, D. Complexing ability of the versatile, redox-active, 3-[3-(Diphenylphosphino)propylthio]-3',4,4'-trimethyl-tetrathiafulvalene Ligand. Inorg. Chem. 2003, 42, 2056–2060. [Google Scholar] [CrossRef]
  9. Avarvari, N.; Fourmigué, M. First radical cation salt of a tetrathiafulvalene-based phosphine metal complex. Chem. Commun. 2004, 2, 1300–1301. [Google Scholar] [CrossRef]
  10. Réthoré, C.; Fourmigué, M.; Avarvari, N. Tetrathiafulvalene based phosphino-oxazolines: A new family of redox active chiral ligands. Chem. Commun. 2004, 34, 1384–1385. [Google Scholar]
  11. Devic, T.; Batail, P.; Fourmigué, M.; Avarvari, N. Unexpected reactivity of PdCl2 and PtCl2 complexes of the unsaturated diphosphine o-Me2TTF(PPh2)2 towards chloride abstraction with thallium triflate. Inorg. Chem. 2004, 43, 3136–3141. [Google Scholar]
  12. Gouverd, C.; Biaso, F.; Cataldo, L.; Berclaz, T.; Geoffroy, M.; Levillain, E.; Avarvari, N.; Fourmigué, M.; Sauvage, F.X.; Wartelle, C. Tetrathiafulvalene–phosphine-based iron and ruthenium carbonyl complexes: Electrochemical and EPR studies. Phys. Chem. Chem. Phys. 2005, 7, 85–93. [Google Scholar]
  13. Kobayashi, A.; Fujiwara, E.; Kobayashi, H. Single-component molecular metals with extended-TTF dithiolateligands. Chem. Rev. 2004, 104, 5243–5264. [Google Scholar] [CrossRef]
  14. Massue, J.; Bellec, N.; Chopin, S.; Levillain, E.; Roisnel, T.; Clérac, R.; Lorcy, D. Electroactiveligands:  The first metal complexes of tetrathiafulvenyl−acetylacetonate. Inorg. Chem. 2005, 44, 8740–8748. [Google Scholar]
  15. Setifi, F.; Ouahab, L.; Golhen, S.; Yoshida, Y.; Saito, G. First radical cation salt of paramagnetic transition metal complex containing TTF as ligand, [CuII(hfac)2(TTF-py)2](PF6)·2CH2Cl2 (hfac = Hexafluoroacetylacetonate and TTF-py = 4-(2-Tetrathiafulvalenyl-ethenyl)pyridine). Inorg. Chem. 2003, 42, 1791–1793. [Google Scholar]
  16. Liu, S.-X.; Dolder, S.; Franz, P.; Neels, A.; Stoeckli-Evans, H.; Decurtins, S. Structural studies of transition metal complexes with 4,5-Bis(2-pyridylmethylsulfanyl)-4',5'-ethylenedithiotetrathiafulvalene:  Probing their potential for the construction of multifunctional molecular assemblies. Inorg. Chem. 2003, 42, 4801–4803. [Google Scholar]
  17. Devic, T.; Avarvari, N.; Batail, P. A series of redox active, tetrathiafulvalene-based amidopyridines and bipyridinesligands: Syntheses, crystal structures, a radical cation salt and group 10 transition-metal complexes. Chem. Eur. J. 2004, 10, 3697–3707. [Google Scholar] [CrossRef]
  18. Devic, T.; Rondeau, D.; Şahin, Y.; Levillain, E.; Clérac, R.; Batail, P.; Avarvari, N. Copper(I/II) complexes of a bis(tetrathiafulvalene)-2,2'-bipyridine: Synthesis, characterization, magnetic and electrochemical properties. Dalton Trans. 2006, 10, 1331–1337. [Google Scholar]
  19. Hervé, K.; Liu, S.-X.; Cador, O.; Golhen, S.; le Gal, Y.; Bousseksou, A.; Stoeckli-Evans, H.; Decurtins, S.; Ouahab, L. Synthesis of a BEDT-TTF bipyridine organic donor and the first FeII coordination complex with a redox-active ligand. Eur. J. Inorg. Chem. 2006, 2006, 3498–3502. [Google Scholar]
  20. Chahma, M.; Hassan, N.; Alberola, A.; Stoeckli-Evans, H.; Pilkington, M. Preparation and coordination complex of the first imine-bridged tetrathiafulvalene–pyridine donor ligand. Inorg. Chem. 2007, 46, 3807–3809. [Google Scholar] [CrossRef]
  21. Qin, J.; Qian, C.-X.; Zhou, N.; Zhu, R.-M.; Li, Y.-Z.; Zuo, J.-L.; You, X.-Z. Metal complexes based on tetrathiafulvalene-fused π-extended schiff base ligands—Syntheses, characterization, and properties. Eur. J. Inorg. Chem. 2012, 234–245. [Google Scholar]
  22. Wu, J.-C.; Liu, S.-X.; Keene, T.D.; Neels, A.; Mereacre, V.; Powell, A.K.; Decurtins, S. Coordination chemistry of a π-extended, rigid and redox-active tetrathiafulvalene-fused schiff-base ligand. Inorg. Chem. 2008, 47, 3452–3459. [Google Scholar] [CrossRef]
  23. Britovsek, G.J.P.; Gibson, V.C.; Kimberley, B.S.; Maddox, P.J.; McTavish, S.J.; Solan, G.A.; White, A.J.P.; Williams, D.J. Novel olefin polymerization catalysts based on iron and cobalt. Chem. Commun. 1998, 849–850. [Google Scholar]
  24. Small, B.L.; Brookhart, M.; Bennett, A.M.A. Highly active iron and cobalt catalysts for the polymerization of ethylene. J. Am. Chem. Soc. 1998, 120, 4049–4050. [Google Scholar]
  25. Iyoda, M.; Hara, K.; Kuwatani, Y.; Nagase, S. Helical tetrathiafulvalene oligomers. Synthesis and properties of Bi-, Ter-, and quatertetrathiafulvalenes. Org. Lett. 2000, 2, 2217–2220. [Google Scholar]
  26. Lamère, J.F.; Malfant, I.; Sournia-Saquet, A.; Lacroix, P.G. Quadric non linear optical response in partially charged donor-substituted tetrathiafulvalene: From a computational investigation to a rational synthetic feasibility. Chem. Mater. 2007, 19, 805–815. [Google Scholar] [CrossRef]
  27. Addison, A.W.; Rao, T.N.; 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[1,7-bis(N-methylbenzimidazol-2'-yl)-2,6-dithiaheptane]copper(II) perchlorate. J. Chem. Soc. Dalton Trans. 1984, 1349–1356. [Google Scholar]
  28. Andreu, R.; Malfant, I.; Lacroix, P.G.; Cassoux, P. New tetrathiafulvalene-π-spacer-acceptor derivatives: Synthesis, crystal structure, optical and electrochemical properties. Eur. J. Org. Chem. 2000, 2000, 737–741. [Google Scholar]
  29. Riobé, F.; Grosshans, P.; Sidorenkova, H.; Geoffroy, M.; Avarvari, N. Mono- and Bis(tetrathiafulvalene)-1,3,5-triazines as covalently linked donor–acceptor systems: Structural, spectroscopic, and theoretical investigations. Chem. Eur. J. 2009, 15, 380–387. [Google Scholar] [CrossRef] [Green Version]

Share and Cite

MDPI and ACS Style

Nita, G.; Branzea, D.; Pop, F.; El-Ghayoury, A.; Avarvari, N. Electroactive Bisiminopyridine Ligands: Synthesis and Complexation Studies. Crystals 2012, 2, 338-348. https://doi.org/10.3390/cryst2020338

AMA Style

Nita G, Branzea D, Pop F, El-Ghayoury A, Avarvari N. Electroactive Bisiminopyridine Ligands: Synthesis and Complexation Studies. Crystals. 2012; 2(2):338-348. https://doi.org/10.3390/cryst2020338

Chicago/Turabian Style

Nita, Gabriela, Diana Branzea, Flavia Pop, Abdelkrim El-Ghayoury, and Narcis Avarvari. 2012. "Electroactive Bisiminopyridine Ligands: Synthesis and Complexation Studies" Crystals 2, no. 2: 338-348. https://doi.org/10.3390/cryst2020338

Article Metrics

Back to TopTop