Plan
Comptes Rendus

Charge transfer complexes and cation radical salts of azino-diselenadiazafulvalene
Comptes Rendus. Chimie, Volume 12 (2009) no. 9, pp. 1057-1065.

Résumés

The synthesis and the redox behaviour of electroactive donor molecules incorporating an azino spacer group between a benzoselenazole core and another heterocyclic moiety, either a benzoselenazole one or a thiazole one, are reported. Neutral complexes were obtained with TCNQ and, for the first time with dithiadiazafulvalene or diselenadiazafulvalene derivatives, cation radical salts by electrocrystallization. Crystal structures data of these complexes are presented and their geometries compared with those deduced from theoretical calculations.

La synthèse et les propriétés redox d'une famille de donneur électroactifs incorporant un espaceur azino entre une moitié benzosélénazole et un autre hétérocycle, soit benzosélénazole ou thiazole, sont décrites. Des complexes neutres ont été obtenus en présence de TCNQ et, pour la première fois avec des dérivés de type dithiadiazafulvalène ou disélénadiazafulvalène, des sels d'ions radicaux ont pu être isolés par électrocristallisation. Les données cristallographiques des complexes sont présentées et leurs géométries sont comparées à celles obtenues à partir de calculs théoriques.

Métadonnées
Reçu le :
Accepté le :
Publié le :
DOI : 10.1016/j.crci.2008.08.013
Keywords: Azino-diselenadiazafulvalene, Radical cation salt, Redox behaviour, Complex
Mots clés : Azino-disélénadiazafulvalène, Sel à radical cation, Comportement rédox, Complexe

Zdenko Časar 1, 2 ; Ivan Leban 2 ; Alenka Majcen-le Marechal 3 ; Lidia Piekara-Sady 1, 4 ; Dominique Lorcy 1

1 Sciences Chimiques de Rennes, UMR 6226 CNRS-Université de Rennes 1, Campus de Beaulieu, Bât 10A, 35042 Rennes cedex, France
2 Faculty of Chemistry and Chemical Technology, University of Ljubljana, Aškerčeva 5, 1000 Ljubljana, Slovenia
3 University of Maribor, Smetanova 17, 2000 Maribor, Slovenia
4 Institute of Molecular Physics, Polish Academy of Sciences, Smoluchowskiego 17, 60-179 Poznan, Poland
@article{CRCHIM_2009__12_9_1057_0,
     author = {Zdenko \v{C}asar and Ivan Leban and Alenka Majcen-le Marechal and Lidia Piekara-Sady and Dominique Lorcy},
     title = {Charge transfer complexes and cation radical salts of azino-diselenadiazafulvalene},
     journal = {Comptes Rendus. Chimie},
     pages = {1057--1065},
     publisher = {Elsevier},
     volume = {12},
     number = {9},
     year = {2009},
     doi = {10.1016/j.crci.2008.08.013},
     language = {en},
}
TY  - JOUR
AU  - Zdenko Časar
AU  - Ivan Leban
AU  - Alenka Majcen-le Marechal
AU  - Lidia Piekara-Sady
AU  - Dominique Lorcy
TI  - Charge transfer complexes and cation radical salts of azino-diselenadiazafulvalene
JO  - Comptes Rendus. Chimie
PY  - 2009
SP  - 1057
EP  - 1065
VL  - 12
IS  - 9
PB  - Elsevier
DO  - 10.1016/j.crci.2008.08.013
LA  - en
ID  - CRCHIM_2009__12_9_1057_0
ER  - 
%0 Journal Article
%A Zdenko Časar
%A Ivan Leban
%A Alenka Majcen-le Marechal
%A Lidia Piekara-Sady
%A Dominique Lorcy
%T Charge transfer complexes and cation radical salts of azino-diselenadiazafulvalene
%J Comptes Rendus. Chimie
%D 2009
%P 1057-1065
%V 12
%N 9
%I Elsevier
%R 10.1016/j.crci.2008.08.013
%G en
%F CRCHIM_2009__12_9_1057_0
Zdenko Časar; Ivan Leban; Alenka Majcen-le Marechal; Lidia Piekara-Sady; Dominique Lorcy. Charge transfer complexes and cation radical salts of azino-diselenadiazafulvalene. Comptes Rendus. Chimie, Volume 12 (2009) no. 9, pp. 1057-1065. doi : 10.1016/j.crci.2008.08.013. https://comptes-rendus.academie-sciences.fr/chimie/articles/10.1016/j.crci.2008.08.013/

Version originale du texte intégral

1 Introduction

Several electron rich olefins, mainly based on the tetrachalcogenofulvalene framework, have been used as precursors of organic materials [1]. Usually, the donor molecules used in this research area present three well defined oxidation state: neutral, cation radical and dication. Indeed, the physical properties of the organic materials (conductivity, magnetism…) are essentially ascribed to the presence of radical species. Therefore, it is of interest to work with donor molecules which exhibit an easily accessible cation radical state and a thermodynamic stability of this radical. Among the numerous variations realized on the fulvalene skeleton, the insertion of a conjugated spacer group such as an azino one between the two heterocyclic cores gave rise to an interesting family of compounds known as heterocyclic azines [2]. Hünig et al. developed this strategy some years ago and showed that these derivatives present two monoelectronic oxidation waves with excellent reversibility [3]. Additionally, the presence of an azino spacer group increases the window potential stability of the cation radical species. Within these azino derivatives, Hünig et al., showed that the three oxidation levels of this system could be isolated, but except one further work on conductivity measurements, no other studies were performed on these donors such as structural or EPR [4]. In an earlier study we examined the potential of azino-dithiadiazafulvalenes as precursors of charge transfer complexes in the presence of TCNQ [5]. Herein we decided to investigate the ability of the azinodibenzodiselenadiazafulvalene 1 as precursor of charge transfer salts. Furthermore, we also prepared unsymmetrical donors where we replaced one benzoselenazole moiety with thiazole cores in order to modulate the redox properties of these donor systems. In this article we will present also the X-ray structure of the complexes obtained with TCNQ oxidation or by using the electrocrystallisation technique as well as molecular geometry optimizations performed by DFT calculations in order to rationalize the geometries' evolutions and electrochemical properties.

2 Results and discussion

2.1 Synthesis and electrochemical studies

In order to prepare the symmetrical azino-DSeDAF 1, we followed Hünig's strategy which consists of the reaction of quaternary heterocyclic salts with a half equivalent of hydrazine in a basic medium (Scheme 1) [3]. First, we prepared 3-methyl-2-methylthio-benzoselenazolium salt 3 from 2-methylthio-benzoselenazole 2 [6] with a new one-step approach in 91% yield. The synthesis of N,N′-bis(benzo-3H-selenazol-2-ylidene)-hydrazine (azino-DSeDAF) 1 was performed by using salt 3 in the presence of NEt3 and half an equivalent of hydrazine hydrate, affording azino-DSeDAF 1 in 55% yield. For the synthesis of unsymmetrical azino donors, containing one selenazole and one thiazole cores, we first prepared the appropriate heterocyclic hydrazones 4 and 5 [5] which are consequently coupled with selenazolium salt 3 in basic medium as depicted in Scheme 1. Indeed, the reaction between selenazolium salt 3 and 3-methyl-3H-benzothiazol-2-one hydrazone 4 [7] in MeCN/EtOH in the presence of NEt3 afforded azinodibenzoselenathiadiazafulvalene (azino-dibenzoSeTDAF) 6 in 70% yield. Treatment of 3,4,5-trimethyl-3H-thiazol-2-one hydrazone 5 with salt 3 under the same conditions resulted in the formation of azino-benzotetramethylselenathiadiazafulvalene (azino-benzoTMSeTAF) 7 which was isolated in 83% yield.

Scheme 1

Electrochemical investigations were performed using tetrabutylamonium hexafluorophosphate as supporting electrolyte. In all cases, cyclic voltammetry revealed two monoelectronic reversible oxidation waves with a difference between the two oxidation potentials ΔE larger than 450 mV (ΔE = E2 − E1, K ≥ 4.23 × 107 Table 1). This indicates a very high thermodynamic stability of the cation radical species induced by the azino spacer group as for example the parent dibenzo-DSeDAF exhibits a smaller ΔE (K = 5.1 × 102, K = [D+]2/[D][D2+]). Comparison of the redox potentials between dibenzo-DSeDAF (Epa1 = −0.07 V and Epa2 = 0.09 V vs SCE in CH2Cl2) [6a] and azinodibenzo-DSeDAF 1 indicates that the presence of the azino spacer group induces an important anodic shift of the two oxidation potentials. Therefore, 1 is stable under atmospheric conditions while dibenzo-DSeDaF oxidizes readily upon air exposure. Furthermore, substitution of one selenium atom in the azino-dibenzo-DSeDAF 1 with sulfur one does not significantly modify the donor ability, since for 5 E1 is only 20 mV lower than for 1. Similarly, the effect on ΔE is negligible since a difference of 10 mV is observed (Table 1). The most pronounced effect on E1 was observed when one benzoselenazole core in 1 was replaced with the 3,4,5,-trimethyl-thiazole one. Indeed, azino-benzoTMSeTDAF 6 has the best electron donating ability in the synthesized azino series (E1 = 0.29 V vs SCE). This unambiguously demonstrates that benzo-fusion decreases the donor ability.

Table 1

Redox properties of azino-dichalcogenodiazafulvalenes n-Bu4NPF6, 100 mV/s.

DonorE1E2E2 − E1 mV (K)
1a0.531.07540 (1.42 × 109)
1b0.561.01450 (4.23 × 107)
5b0.541.00460 (6.26 × 107)
6b0.290.81520 (6.5 × 108)
Dibenzo-DSeDAF [6]−0.07c0.09c160 (5.1 × 102)
Dibenzo-TSeF [12]0.781.17390 (4.07 × 106)

a CH2Cl2.

b CH3CN.

c Oxidation peak potential.

2.2 TCNQ complexes

In order to form charge transfer salts with these azino derivatives, we tried both chemical and electrochemical pathways. For the chemical approach, we simply mixed a hot solution of the donor in CH2Cl2 with a hot solution of TCNQ in CH3CN. Treatment of azine 1, 6 and 7 solutions with TCNQ resulted in the formation of poorly soluble azine–TCNQ complexes which precipitated from the medium as microcrystalline materials in high yields for 1 and 6 (82% and 84% yields, respectively) while in low yield for 7 (19%). Elemental analysis shows that 1:1 complexes were formed in all three cases. FT-IR spectra revealed that all three isolated complexes exhibit a single sharp νCN band at 2218 cm−1 for (1)(TCNQ) and (6)(TCNQ) and at 2212 cm−1 for (7)(TCNQ). Based on the nitrile stretching absorption band, it is possible to infer the degree of charge transfer in TCNQ complexes [8]. The value obtained for the three complexes are close to the one observed for neutral TCNQ νCN at 2224 cm−1 which indicates that these materials are neutral charge transfer complexes [9]. This was confirmed by EPR experiments as no EPR signal was observed from room temperature to liquid helium temperature. This result is consistent with the presence of neutral complexes.

Crystals were obtained for complex (1)(TCNQ) and the crystal structure determination confirms a stoichiometry of one donor 1 for one TCNQ. It crystallizes in the triclinic system, space group P-1 with both the donor and TCNQ molecules located on an inversion center. Within this complex, donor molecules form alternated stacks with TCNQ. The plane of the donor is not fully parallel to the plane of the acceptor but forms an angle of 2.9(3)° with an interplanar separation of 3.47 Å (Fig. 1). This organization is reminiscent of what we previously observed with (azino-dithiadiazafulvalenes)(TCNQ) complexes [5]. Analysis of the bond lengths in the TCNQ skeleton shows that they are essentially similar to those reported for the neutral TCNQ [10] (see Table 2), confirming that the complex (1)(TCNQ) is neutral. Actually, the potential difference between the reduction potential of TCNQ (0.18 V) and the oxidation potential of 1 (0.56 V) is too large for allowing any redox reaction. Thus, in this case, we are dealing with co-crystallization of the two components rather than with an electron transfer process.

Fig. 1

(a) View of the molecular structure of 1 in (1)(TCNQ) showing the atom labelling. Hydrogen atoms are omitted for clarity (50% probability ellipsoids). (b) View of the unit cell along ab plane.

Table 2

Averaged distances (ad) in TCNQ.

Neutral TCNQ [10](1)(TCNQ)
a1.346(3)1.338(3)
b1.448(4)1.441(3)
c1.374(3)1.374(3)
d1.441(3)1.433(4)

2.3 Electrocrystallization

Then, we focused our attention on the electrochemical pathway which consists in using the electrocrystallization technique, an invaluable tool for the elaboration of ordered electroactive molecular charge transfer salts [11]. Deep blue needles grew on the anode within a couple of days from donor 1 by using n-Bu4NPF6 as the supporting electrolyte in CH2Cl2. The structure of the salt was determined by X-ray diffraction studies and revealed a stoichiometry of one donor for one PF6 anion, the first example of isolated cation radical salts in these series.

The salt (1)(PF6) crystallizes in the monoclinic system, space group C2/c with one donor molecule and one PF6 anion both in general position in the unit cell, hence a 1:1 stoichiometry. Within this salt, the oxidized donor molecules form segregated stacks (Fig. 2a). Two types of overlap are observed as depicted in Fig. 2b. Mean plane distances between donor molecules within one stack are, respectively, of 3.37 Å and 3.64 Å for overlap A and B. Furthermore, the donor core in its cation radical state is planar, as the structure of the donor 1 under neutral state in (1)(TCNQ). On the other hand, modification of the bond lengths of the donor skeleton is observed in the oxidized donor core essentially on the central spacer group. In the oxidized species the double bond CN is lengthened while the central N–N bond is shortened when compared with the neutral molecule (Table 3). The C–Se and C–N cyclic bonds are also shortened in the oxidized species compared with the neutral one.

Fig. 2

Crystal structure of complex (1)(PF6). (a) Projection onto the ac plane showing the segregated stack of donors. (b) Showing the overlap modes between two neighbouring donors 1.

Table 3

Averaged bond length of donor 1 in (1)(TCNQ) and in (1)(PF6) and B3LYP/LANL2DZ optimized bond lengths for 1 and 1+ in Å.

In complex (1)(TCNQ)In complex (1)(PF6)1 (DFT)1+(DFT)
a1.414(4)1.354(6)1.4141.361
b1.286(3)1.312(7)1.3011.339
1.322(7)
c1.366(3)1.357(6)1.3921.369
1.343(6)
d1.898(2)1.874(4)1.9671.937
1.864(4)

The EPR study was carried out on a polycrystalline sample of (1)PF6. A single Lorentzian shaped line was observed at g = 2.007 with a linewidth of 40 G at room temperature. Since PF6 is diamagnetic, this signal is due to the radical cation of 1. The temperature dependence of the signal was studied using the well-known approximation that spin susceptibility is proportional to the product: ABpp)2, where A is the line amplitude and ΔBpp – peak-to-peak linewidth. The result is displayed in Fig. 3 as χsT. From 70 K to ∼200 K spin susceptibility, within the experimental error, is constant, i.e. the Curie law is obeyed. Above 200 K there is a thermally activated slow increase in susceptibility which is a sign of a singlet–triplet behaviour associated with the dimerization of cations observed in the X-ray crystal structure (Fig. 2). Similarly, the linewidth decreases with lowering temperature from 40 G at room temperature to ∼20 G at 200 K. Further decrease in temperature does not change the linewidth significantly (Fig. 4). Both χsT and ΔBpp temperature evolution confirm a thermally activated susceptibility above 200 K. One can estimate singlet–triplet energy gap Δ/kB of ca. 900 K (630 cm−1) (Fig. 3, right).

Fig. 3

(Left) Temperature dependence of χsT derived from the EPR of (1)(PF6). (Right) Linear fit to ln(χsT) = −(Δ/kB)/T + ln C above 200 K.

Fig. 4

Temperature dependence of the EPR peak-to-peak linewidh of (1)(PF6).

2.4 Molecular geometry optimization

DFT calculations [B3LYP/LANL2DZ] were performed on the azinodibenzo-DSeDAF 1 and the oxidized species 1+, and, for comparison purpose, on the dibenzo-tetraselenafulvalene (TSeF) and the dibenzo-DSeDAF in order to determine the highest occupied molecular orbitals (HOMO) energies as well as their electron density contours. As shown in Fig. 5, the energy of the HOMO of 1 (E = −4.89 eV) is located in between those of dibenzo-TSeF (E = −5.00 eV) and dibenzo-DSeDAF (E = −4.42 eV), but closer in energy to the TSeF system. This evolution of the HOMOs was anticipated from their redox potentials determined by cyclic voltammetry, as for instance the redox potentials of the dibenzo-TSeF, E1 = 0.78 V and E2 = 1.17 V vs SCE in CH2Cl2, are closer to the redox potential of 1 compared with the dibenzo-DSeDAF [12]. These energies of the HOMO also explain the higher oxidation potentials observed for the 1 derivatives compared with DSeDAF. The topology of the HOMO for the selenazole rings in 1 and the dibenzo-DSeDAF is similar showing antibonding interactions between the selenium or the nitrogen and the carbon atoms. Concerning the conjugated spacer group in 1, bonding interactions are observed between the nitrogen and the carbon atoms while an antibonding interaction connect the two nitrogen atoms inducing a destabilization of the HOMO and generating a higher oxidation potential than for dibenzo-DSeDAF. It is interesting to note the small coefficient at the C4 carbon atom of the selenazole rings on both the DSeDAF and the azino derivative. This small coefficient on the carbon located alpha to the NMe group has already been noticed on the thiazole ring of sulfur analogues, DTDAF [13]. Optimized bond lengths calculated for azinodibenzo-DSeDAF 1 and the oxidized species 1+ are reported in Table 3. If we compare the optimized geometry and bond lengths calculated for 1 with the X-ray data in (1)(TCNQ), we can observe that there is a good agreement for the neutral state of the donor in the complex, (1)(TCNQ). Moreover the optimized bond length calculated for 1+ shows that oxidation affects essentially the bonds in which the electron density in the HOMO is concentrated, in agreement with the experimental evidence (1)(PF6).

Fig. 5

Molecular orbital surfaces and HOMO levels of dibenzo-DSeDAF (top left), dibenzo-TSeF (top right) and azinodibenzo-DSeDAF neutral (bottom).

3 Conclusion

In this study, a series of bis(benzoselenazolydene)hydrazine have been synthesized and their redox properties have been electrochemically determined. A first example of a cation radical salt of such azino derivatives has been electrochemically crystallized and structurally characterized. This is in sharp contrast with the recurrent decomposition observed with the “classical” DSeDAF or dithiadiazafulvalene (DTDAF) derivatives. Future work is in progress for exploring the potentialities of these azino donors upon electrocrystallization experiments.

4 Experimental

4.1 General

1H NMR spectra were recorded at 300 MHz and 13C NMR spectra at 75 MHz with tetramethylsilane as internal reference. Mass spectra were carried out at Centre de Mesures Physiques de l'Ouest, Rennes. Melting points were measured using a Kofler hot stage apparatus and are uncorrected. Elemental analysis results were obtained from the Laboratoire Central de Microanalyse du CNRS (Lyon). Cyclic voltammetry was carried out on a 10−3 M solution of metal complex derivative in CH2Cl2 containing 0.1 M n-Bu4NPF6 as supporting electrolyte. Voltammograms were recorded at 0.1 V s−1 on a platinum disk electrode (1 mm2). Potentials were measured vs Saturated Calomel Electrode (SCE). The EPR measurements were performed on Bruker ESP-300E X-band spectrometer equipped with a helium Oxford kriostat using liquid nitrogen as cooling agent.

4.2 Synthesis and characterization

4.2.1 3-Methyl-2-methylthio-benzoselenazolium tetrafluoroborate (3)

To a stirred solution of 2-methylthio-benzoselenazole 2 (2.25 g, 9.86 mmol) in CH2Cl2 (10 mL) was added trimethyloxonium tetrafluoroborate (1.46 g, 9.86 mmol) in CH2Cl2/MeNO2 (5 mL/9 mL). The reaction mixture was stirred at room temperature for 52 h. The solvent was evaporated and CH2Cl2 (10 mL) was added to the residue. The mixture was treated with Et2O (35 mL). The precipitate was filtered washed with Et2O and dried to give 3-methyl-2-methylthio-benzoselenazolium tetrafluoroborate 3 as white powder in 91% yield (2.97 g) as a white powder, mp 187 °C. 1H NMR (CD3CN): δ 8.13–8.00 (1H, dd, J1 = 8.1 Hz, J2 = 1.0 Hz), 7.85–7.72 (1H, dd, J1 = 8.3 Hz, J2 = 0.8 Hz), 7.70–7.58 (1H, td, J1 = 7.3 Hz, J2 = 1.3 Hz), 7.56–7.42 (1H, td, J1 = 7.3 Hz, J2 = 1.3 Hz), 3.87 (3H, s), 2.91 (3H, s). 13C NMR (CD3CN): δ 189.4, 143.8, 129.6, 128.8, 126.9, 125.9, 116.6, 37.5, 19.5. Anal. Calcd for C9H10NSSeBF4: C, 32.76; H, 3.05; N, 4.24. Found: C, 32.50; H, 2.94; N, 4.05.

4.2.2 Azino-dibenzoDSeTDAF (1)

To a stirred suspension of 3-methyl-2-methylthio-benzoselenazolium tetrafluoroborate 2 (1.20 g, 3.64 mmol) in absolute EtOH (8 mL) were added NEt3 (0.74 g, 7.27 mmol) in absolute EtOH (2 mL) and H2N–NH2·H2O (88.2 μL, 1.82 mmol). The mixture was stirred at room temperature for 40 min. The precipitated product was filtered off and dried to give pale yellow powder. The crude product was suspended in 100 mL of water and suspension was stirred at room temperature for 30 min. Undissolved white solid was filtered and dried on air at room temperature to give white powder, which was recrystallized from petroleum ether/CH2Cl2 to furnish azino-DSeDAF 1 in 55% yield (0.42 g) as white powder, Rf (CH2Cl2) = 0.72, mp 273 °C (lit. [2] 270–272 °C). 1H NMR (acetone-d6): δ 7.66–7.57 (2H, dd, J1 = 7.6 Hz, J2 = 1.0 Hz), 7.39–7.27 (2H, td, J1 = 7.8 Hz, J2 = 1.3 Hz), 7.17–7.09 (2H, dd, J1 = 8.1 Hz, J2 = 1.0 Hz), 7.06–7.96 (2H, td, J1 = 7.6 Hz, J2 = 1.0 Hz), 3.58 (6H, s). 13C NMR (CDCl3): δ 159.6, 143.8, 126.7, 126.0, 122.8, 121.3, 110.0, 31.5. MS m/z (relative intensity) 422 (M+, 100), 374 (47), 225 (10), 197 (24), 184 (37). HRMS: calcd for C16H14N4Se2: 421.9549. Found: 421.9565. Anal. Calcd for C16H14N4Se2: C, 45.73; H, 3.36; N, 13.33. Found: C, 45.82; H, 3.53; N, 13.37.

4.2.3 Azino-dibenzoSeTDAF (6)

To a stirred suspension of 3-methyl-3H-benzothiazol-2-one hydrazone 4 (0.18 g, 1.00 mmol) in absolute EtOH (4 mL) were added NEt3 (0.10 g, 1.00 mmol) in absolute EtOH (2 mL) and 3-methyl-2-methylthio-benzoselenazolium tetrafluoroborate 3 (0.33 g, 1.00 mmol) in MeCN (6 mL). The mixture was stirred under nitrogen at room temperature for 45 min. The precipitated product was filtered off, washed with absolute EtOH (10 mL) and dried to give white crystalline powder, which was recrystallized from petrol ether/CH2Cl2 to furnish azine 6 in 70% yield (0.26 g), white powder, Rf (CH2Cl2) = 0.74, mp 268 °C. 1H NMR (DMSO-d6): δ 7.69–7.62 (1H, dd, J1 = 7.8 Hz, J2 = 1.3 Hz), 7.59–7.52 (1H, dd, J1 = 7.6 Hz, J2 = 0.8 Hz), 7.37–7.25 (2H, m), 7.22–7.11 (2H, td, J1 = 8.1 Hz, J2 = 1.0 Hz), 7.09–6.94 (2H, m), 3.52 (3H, s), 3.49 (3H, s). 13C NMR (CDCl3): δ 159.6, 159.3, 143.9, 142.1, 126.7, 126.4, 126.0, 124.8, 122.8, 122.4, 121.2, 121.1, 109.9, 108.7, 31.9, 30.9. MS m/z (relative intensity) 374 (M+, 100), 326 (25), 212 (5), 197 (10), 164 (18), 149 (22). HRMS: calcd for C16H14N4SSe: 374.0104. Found: 374.0116. Anal. Calcd for C16H14N4SSe: C, 51.47; H, 3.78; N, 15.00. Found: C, 51.76; H, 3.83; N, 15.23.

4.2.4 Azino-benzoTMSeTDAF (7)

To a stirred solution of 3,4,5-trimethyl-3H-thiazol-2-one hydrazone 5 (0.26 g, 1.67 mmol) in absolute EtOH/MeCN = 1:1 (16 mL) was added 3-methyl-2-methylthio-benzoselenazolium tetrafluoroborate 3 (0.55 g, 1.67 mmol) and NEt3 (0.17 g, 1.67 mmol). The mixture was stirred under nitrogen at room temperature for 60 min. The precipitated product was filtered off, washed with absolute EtOH (25 mL) and dried to give green powder, which was recrystallized from petrol ether/CH2Cl2 to furnish azine 7 in 83% yield (0.49 g), light green cubes, Rf (CH2Cl2) = 0.45, mp 249 °C. 1H NMR (CDCl3): δ 7.50–7.42 (1H, dd, J1 = 7.6 Hz, J2 = 1.3 Hz), 7.28–7.21 (1H, dd, J1 = 7.8 Hz, J2 = 1.3 Hz), 6.99–6.86 (2H, m), 3.56 (3H, s), 3.35 (3H, s), 2.09 (3H, s), 2.04 (3H, s). 13C NMR (pyridine-d5): δ 164.2, 158.6, 146.4, 132.1, 129.1, 128.4, 125.0, 123.2, 112.2, 106.5, 33.8, 33.5, 14.2, 13.1. MS m/z (relative intensity) 352 (M+, 100), 304 (15), 197 (5), 184 (6), 176 (8), 156 (7), 141 (12), 127 (12). HRMS: calcd for C14H16N4SSe: 352.0261. Found: 352.0275. Anal. Calcd for C14H16N4SSe: C, 47.86; H, 4.59; N, 15.95. Found: C, 48.06; H, 4.56; N, 16.19.

4.2.5 TCNQ complexes

To a solution of azino donor (0.2 mmol) in CH2Cl2 (9 mL) was added a hot solution of TCNQ (0.2 mmol) in MeCN (8 mL). The mixture was left to stand at room temperature for 2–3 days. The precipitate was filtered and dried to give the azine·TCNQ complexes.

(1)(TCNQ): 82% yield, brown-red needles, mp 283 °C (dec). IR (KBr): νCN/cm−1 2218. Anal. Calcd for C28H18N8Se2: C, 53.86; H, 2.91; N, 17.95. Found: C, 54.15; H, 2.83; N, 17.68.

(6)(TCNQ): 84% yield, brown-red needles, mp 279 °C (dec). IR (KBr): νCN/cm−1 2218. Anal. Calcd for C28H18N8SSe: C, 58.23; H, 3.14; N, 19.40. Found: C, 58.48; H, 2.93; N, 19.17.

(7)(TCNQ): 19% yield, violet-red microcrystalline solid, mp 250 °C (dec). IR (KBr): νCN/cm−1 2212. Anal. Calcd for C26H20N8SSe: C, 56.21; H, 3.63; N, 20.17. Found: C, 56.57; H, 3.39; N, 20.21.

4.3 Electrocrystallization

Into a 30 mL H-shaped glass cell with a fine frit dividing the anodic and cathodic compartments equipped with platinum wire electrodes was placed a solution of tetrabutylammonium hexafluorophosphate (∼3 g) in CH2Cl2 (30 mL). The azine donor 1 (10 mg) was added into the anode compartment. The solution was electrolyzed under a constant current of 2 μA at room temperature. Blue crystals of radical cation salts gradually grew on the anode electrode within several days. The crystals were collected by filtration, washed with cold CH2Cl2 and dried. The obtained crystals were utilized for X-ray crystallographic analyses.

4.4 Crystallography

Single-crystal diffraction data were collected on a Nonius KappaCCD diffractometer from the University of Ljubljana for complex (1)(TCNQ) and on a Stoe IPDS diffraction system from the University of Angers for compound (1)(PF6). Details of the crystallographic are given in Table 4.

Table 4

Crystal data and structure refinement parameters for (1)(TCNQ) and (1)(PF6).

Structure parameter(1)(TCNQ)(1)(PF6)
Empirical formulaC16H14N4Se2, C12H4N4C16H14N4Se2, PF6
Molecular weight624.42565.20
Crystal systemTriclinicMonoclinic
Space groupP-1C2/c
a (Å)8.5649(2)7.1183(5)
b (Å)9.1230(2)20.4029(11)
c (Å)9.8916(2)28.750(2)
α (°)68.4690(10)90
β (°)72.1990(10)91.610(8)
γ (°)67.8490(10)90
V (Å3)653.23(3)4173.8(5)
T (K)293(2)293(2)
Z18
F0003102200
Dcalc (g/cm3)1.5871.799
μ (mm−1)2.8643.682
Total number of measured intensities14,98313,528
Abs corrMultiscanMultiscan
Number of unique data2961(0.0285)3912(0.0771)
Observed reflections (I > 2σ(I))26772267
Number of refined variables (n)180298
R1, wR20.0356, 0.09560.0419, 0.0912
R1, wR2 (all data)0.0309, 0.09050.0874, 0.1029
GoF1.0540.876

5 Supplementary material

Crystallographic data for structural analysis have been deposited with the Cambridge Crystallographic Data Centre, CCDC nos. 695874–695875 for compound (1)(TCNQ) and (1)(PF6), respectively. Copies of this information may be obtained free of charge from The Director, CCDC, 12 Union road, Cambridge CB2 1EZ, UK (fax: +44 1223 336033; e-mail: deposit@ccdc.cam.ac.uk or http://www.ccdc.cam.ac.uk).

Acknowledgements

L.P.-S. thanks the CNRS for financial support for her stay in Rennes. Z.C. thanks the Ministry of Education, Science and Sport of the Republic of Slovenia for young researcher fellowship (T19-103-002/19281/99).


Bibliographie

[1] P. Batail Chem. Rev., 104 (2004), p. 4887 (Special issue on “Molecular conductors”)

[2] S. Hünig; G. Kießlich; F. Linhart; H. Schlaf Justus Liebigs Ann. Chem., 752 (1971), p. 182

[3] S. Hünig; G. Kießlich; F. Linhart; H. Schlaf Justus Liebigs Ann. Chem., 752 (1971), p. 196

[4] W.A. Barlow; G.R. Davies; E.P. Goodings; R.L. Hand; G. Owen; M. Rhodes Mol. Cryst. Liq. Cryst., 33 (1976), p. 123

[5] D. Guérin; D. Lorcy; R. Carlier; S. Los; L. Piekara-Sady J. Solid State Chem., 168 (2002), p. 590

[6] Z. Časar; P. Bénard-Rocherullé; A. Majcen-Le Maréchal; D. Lorcy; Z. Časar; I. Leban; A. Majcen-Le Maréchal; D. Lorcy J. Chem. Soc., Perkin Trans. 1 (2001), p. 1336

[7] S. Hünig; H. Balli Justus Liebigs Ann. Chem., 609 (1957), p. 160

[8] J.S. Chappel; A.N. Bloch; W.A. Bryden; M. Maxfield; T.O. Poehler; D.O. Cowan J. Am. Chem. Soc., 103 (1981), p. 2442

[9] F.H. Herbstein (J.D. Dunitz; J.A. Ibers, eds.), Perspectives in Structural Chemistry, vol. 4, Wiley, New York, 1971, p. 166

[10] R.E. Long; R.A. Sparks; K.N. Trueblood Acta Crystallogr., 18 (1965), p. 932

[11] P. Batail; M. Fourmigué; K. Boubekeur; J.-C. Gabriel Chem. Mater., 10 (1998), p. 3005

[12] K. Lerstrup; D. Talham; A. Bloch; T. Poehler; D. Cowan J. Chem. Soc., Chem. Commun. (1982), p. 332

[13] S. Eid; M. Guerro; T. Roisnel; D. Lorcy Org. Lett., 8 (2006), p. 2377


Commentaires - Politique