Ole
Mallow
a,
Monther A.
Khanfar
bc,
Moritz
Malischewski
b,
Pamela
Finke
a,
Malte
Hesse
a,
Enno
Lork
a,
Timo
Augenstein
d,
Frank
Breher
d,
Jeffrey R.
Harmer
e,
Nadezhda V.
Vasilieva
f,
Andrey
Zibarev
fg,
Artem S.
Bogomyakov
h,
Konrad
Seppelt
b and
Jens
Beckmann
*a
aInstitut für Anorganische Chemie, Universität Bremen, Leobener Straße, 28359 Bremen, Germany. E-mail: j.beckmann@uni-bremen.de
bInstitut für Chemie und Biochemie, Freie Universität Berlin, Fabeckstraße 34/36, 14195 Berlin, Germany
cDepartment of Chemistry, The University of Jordan, Amman 11942, Jordan
dInstitut für Anorganische Chemie, Karlsruhe Institute of Technology, Engesserstr. 15, 76131 Karlsruhe, Germany
eCentre for Advanced Imaging, University of Queensland, St Lucia, Queensland 4072, Australia
fInstitute of Organic Chemistry, Russian Academy of Sciences, 630090 Novosibirsk, Russia
gDepartment of Physics, National Research University – Novosibirsk State University, 630090 Novosibirsk, Russia
hInternational Tomography Centre, Russian Academy of Sciences, 630090 Novosibirsk, Russia
First published on 21st October 2014
One-electron oxidation of two series of diaryldichalcogenides (C6F5E)2 (13a–c) and (2,6-Mes2C6H3E)2 (16a–c) was studied (E = S, Se, Te). The reaction of 13a and 13b with AsF5 and SbF5 gave rise to the formation of thermally unstable radical cations [(C6F5S)2]˙+ (14a) and [(C6F5Se)2]˙+ (14b) that were isolated as [Sb2F11]− and [As2F11]− salts, respectively. The reaction of 13c with AsF5 afforded only the product of a Te–C bond cleavage, namely the previously known dication [Te4]2+ that was isolated as [AsF6]− salt. The reaction of (2,6-Mes2C6H3E)2 (16a–c) with [NO][SbF6] provided the corresponding radical cations [(2,6-Mes2C6H3E)2]˙+ (17a–c; E = S, Se, Te) in the form of thermally stable [SbF6]− salts in nearly quantitative yields. The electronic and structural properties of these radical cations were probed by X-ray diffraction analysis, EPR spectroscopy, and density functional theory calculations and other methods.
The oxidation of the bulky bis(m-terphenyl)dichalcogenides were studied by cyclic voltammetry first. At a stationary Pt electrode, electrochemical oxidation of 16a–c in CH2Cl2/0.1 M [n-Bu4N][BF4] within the potential sweep range 0 < E < 1.5 V is characterized by a one-electron quasi-reversible peak (EOxp − ERedp = 0.07 − 0.08 V, EOxp − EOxp/2 = 0.06 V, IRedp/IOxp ≈ 0.8 − 0.9). The observed quasi-reversibility of the peak indicates relative stability of the radical cations. Significant differences in the peak currents for 16a–c can be tentatively attributed to the differences in their diffusion coefficients. The anodic peak potentials with respect to a saturated calomel electrode of (2,6-Mes2C6H3E)2 decrease from EOxp = 1.22 V (16a, E = S) over 1.09 V (16b, E = Se) to 0.79 V (16c, E = Te), respectively, which suggested nitrosonium salts to be suitable one-electron oxidizers.15 Indeed, the reaction of the bis(m-terphenyl)dichalcogenides (2,6-Mes2C6H3E)2 (16a, E = S; 16b, E = Se; 16c, E = Te) with [NO][SbF6] in propionitrile provided the corresponding radical cations [(2,6-Mes2C6H3E)2]˙+ (17a, E = S; 17b, E = Se; 17c, E = Te; counterion [SbF6]−) as dark blue crystals in very high yields, which showed no signs of decomposition for several months when isolated from the mother liquor and stored under argon (Scheme 3).
The molecular structures of [(C6F5E)2]˙+ (14a, E = S; 14b, E = Se) and [(2,6-Mes2C6H3E)2]˙+ (17a, E = S; 17b, E = Se; 17c, E = Te) are shown in Fig. 1. Selected bond parameters are collected in Table 1 together with those of the neutral parent compounds. The radical cations 14a, 14b, 17a and 17b containing S and Se atoms adopt nearly Cs symmetric conformations. The phenyl rings comprising C10–C15 are almost coplanar with the C10–E1–E2 plane pointing to delocalization of unpaired electron spin density across the aromatic π-system, whereas the phenyl rings including C20–C25 are perpendicular to the C20–E2–E1 plane (E = S, Se). The radical cation 16c containing Te atoms is centrosymmetric and possesses C2 symmetry. Consequently, only one crystallographically independent Te atom is present. In the radical cations [(C6F5S)2]˙+ (14a) and [(2,6-Mes2C6H3S)2]˙+ (17a) the delocalization is also reflected in the S–C bond lengths; S1–C10 (14a, 1.727(8) Å; 17a, 1.762(5) Å) is significantly shorter than S1–C20 (14a, 1.764(8); 17a, 1.799(5) Å) pointing to a quinoid structure of the coplanar phenyl ring (C10–C15). Indeed, the quinoid character of the coplanar phenyl rings of 14a (Q = 0.036 (26%)), 14b (Q = 0.012 (9%)), 17a (Q = 0.019 (14%)) and 17b (Q = 0.019 (14%)) is substantially higher than that of the perpendicular phenyl rings of 14a (Q = 0.024 (17%)), 14b (Q = 0.001 (<1%)), 17a (Q = 0.001 (<1%)) and 17b (Q = 0.002 (1%)).16 The quinoid character of 17c (Q = 0.037 (27%)) adopts the highest value and nearly approaches that of the hexafluorobenzene radical cation [C6F6]˙+ (Q = 0.042 (30%)).17 Comparison of the parent compounds with the corresponding radical cations reveals a shortening of the E–E bonds by 0.008 Å for 13a/14a, 0.030 Å for 13b/14b, 0.075 Å for 16a/17a, 0.050 Å for 16b/17b and 0.049 Å for 16c/17c (E = Te), respectively. The E–E bond shortening unambiguously suggests that the bond order increased as electron density from π*-orbitals of the chalcogens has been depleted upon oxidation. The increase of the E–E bond orders should be also reflected by an increase of the E–E stretching vibrations, however, all attempts to obtain reasonable Raman spectra failed due to the intense colour of the compounds. The most striking structural difference upon going from the parent compounds to the radical cations is the dramatic increase of the C–E–E–C torsion angles from 84.6(2)° to 175.8(4)° for 13a/14a, 127.2(1) to 174.6(3)° for 16a/17a (E = S), 75.3(1) to 178.1(1)° for 13b/14b, 128.2(3) to 172.6(5)° for 16b/17b (E = Se) and 123.1(1) to 155.5(3)° for 16c/17c (E = Te), respectively. The positive charges of 17a–c seem to be compensated by intramolecular Menshutkin interactions between chalcogen atoms and mesityl groups of the m-terphenyl substituents (E–Zπca. 3 Å; Zπ = centroid of the phenyl ring).21 The neutral parent compounds exhibit interactions which are substantially longer (E–Zπca. 3.4 Å). Presumably for the same reason, 14a and 14b possesses a short intramolecular S⋯F (2.712(6) Å) and Se⋯F (2.770(2) Å) contacts. These structural changes upon oxidation were satisfactorily reproduced by DFT calculations on two series of parent compounds, namely (PhE)2 (1a–c) and (C6F5E)2 (13a–c), and radical cations, namely [(PhE)2]˙+ (2a–c) and [(C6F5E)2]˙+ (14a–c) for E = S, Se, Te. The adiabatic ionization energies of 1a–c (6.98–7.38 eV) are lower than those of 13a–c (7.64–8.16 eV) and follow the same trend as the cathodic peak potentials of series (2,6-Mes2C6H3E)2 (16a–c) for E = S, Se, Te. In contrast to the neutral (PhE)2 (1a–c) that all have a C2 (H2O2 type) structure with dihedral angles around 90°, all radical cations are calculated having much larger dihedral angles. The phenyl-substituted radical cations [(PhE)2]˙+ (2a–c), can be considered as essentially freely rotating around the E–E bonds, with a minimum energy at dihedral angles around 160° and a C2-symmetric structure. The experimentally found Cs symmetric structure is less than 1 kcal mol−1 higher in energy and corresponds to a calculated transition state (one imaginary frequency (>10 cm−1)). The size of the butterfly shaped m-terphenyl groups easily explains why the radical cations [(2,6-Mes2C6H3E)2]˙+ (17a–b; E = S, Se) have Cs symmetry, simply for reasons of sterical crowding. In the pentafluorophenyl substituted radical cations [(C6F5E)2]˙+ (14a–c) with E = S and Se, the experimentally observed Cs symmetry is the ground state, and the C2 symmetric structure is about 1.5 kcal mol−l higher in energy. Interestingly, [(C6F5Te)2]˙+ (14c) behaves differently. Its ground state is C2h symmetric, completely flat, with a dihedral angle of 180°. The Cs symmetric structure is again only <0.5 kcal mol−l higher in energy. It is unclear whether this peculiar structure is the reason why it has not been possible to isolate it. The HOMO of (C6F5Se)2 (13b) is an admixture of components situated at the Se atoms and the π-system of the pentafluorophenyl groups, whereas in the SOMO of the corresponding radical cation [(C6F5Se)2]˙+ (14b), the unpaired electron is strongly distributed over the coplanar phenyl group (Fig. 2).
Fig. 1 Molecular structures of the radical cations 14a,b and 17a–c; thermal ellipsoids are set at 30% probability. |
(RE)2[(RE)2]˙+ | 13a (ref. 18)/14a, E = S, R = C6F5 | 13b (ref. 18)/14b, E = Se, R = C6F5 | 16a/17a, E = S, R = m-Ter | 16b (ref. 19)/17b, E = Se, R = m-Ter | 16c (ref. 20)/17c, E = Te, R = m-Ter |
---|---|---|---|---|---|
a Quinoid character is defined as Q(C10–C15) = (dC10–C11 + dC12–C13 + dC13–C14 + dC15–C10)/4 − (dC11–C12 + dC14–C15)/2 and Q(C20–C25) = (dC20–C21 + dC22–C23 + dC23–C24 + dC25–C20)/4 − (dC21–C22 + dC24–C25)/2 and is 0 for a perfectly delocalized hexagonal benzene structure and 0.138 for a perfect quinoid structure where dC10–C11 = dC12–C13 = dC13–C14 = dC15–C10 = 1.455 Å and dC20–C21 = dC22–C23 = 1.317 Å. b Element distance to the centroid of the phenyl ring E–Zπ. | |||||
E–E | 2.022(2) | 2.319(4) | 2.073(1) | 2.339(2) | 2.711(1) |
2.014(3) | 2.289(1) | 1.998(2) | 2.289(7) | 2.662(1) | |
C–E | 1.785(6) | 1.90(2) | 1.790(2) | 1.926(6) | 2.144(3) |
1.796(6) | 1.92(1) | 1.787(2) | 1.926(7) | 2.151(3) | |
1.727(8) | 1.876(2) | 1.762(5) | 1.94(1) | 2.131(8) | |
1.764(8) | 1.904(2) | 1.799(5) | 1.947(8) | ||
C–E–E | 104.5(2) | 98.7(5) | 104.8(1) | 102.2(3) | 103.3(1) |
106.2(2) | 98.9(6) | 104.4(1) | 102.3(3) | 103.0(1) | |
96.2(3) | 93.71(7) | 95.4(2) | 95.9(3) | 98.5(2) | |
107.5(3) | 103.9(1) | 110.4(2) | 105.6(3) | ||
C–E–E–C | 84.6(2) | 75.3(1) | 127.2(1) | 128.2(3) | 123.1(1) |
175.8(4) | 178.1(1) | 174.6(3) | 172.6(5) | 155.5(3) | |
Q (C10–C15) | 0.036 | 0.012 | 0.019 | 0.019 | 0.037 |
(26%) | (9%) | (14%) | (14%) | (27%) | |
Q (C20–C25) | 0.024 | 0.001 | 0.001 | 0.002 | — |
(17%) | (<1%) | (<1%) | (1%) | ||
E–Zπb | — | — | 3.439(1) | 3.452(2) | 3.377(1) |
— | — | 2.975(2) | 3.01(1) | 3.481(1) | |
3.161(2) | |||||
E⋯F | 2.712(6) | 2.770(2) | — | — | — |
Fig. 2 HOMO (left) of C2 symmetric (C6F5Se)2 (13b) and SOMO (right) of Cs symmetric [(C6F5Se)2]˙+ (14b). |
The Wiberg bond indices (WBIs) of the E–E bonds increase from about 1 for the parent compounds 1a–c and 13a–c to values between 1.216 and 1.315 for the [(PhE)2]˙+ (2a–c; E = S, Se, Te) series and between 1.176 and 1.282 for the [(C6F5E)2]˙+ (14a–c; E = S, Se, Te) series of radical cations. In the absence of electron delocalization across the coplanar phenyl rings the WBIs of the radical cations would have been expected to be close to 1.5.
The stability of 17a–c in solution dramatically depends on the solvents and in CH2Cl2 on the nature of the chalcogen. Ink-blue solutions of 17a and 17b in CH2Cl2 are stable for months, whereas purple-blue 17c decomposes within a few days. The blue colour arises from very broad absorptions in the near IR region, which stretch into the visible range. Only 17c shows an absorption maximum (CH2Cl2) at λmax = 583 nm in the visible range that is only slightly shifted compared to that of 16c (553 nm) and responsible for the purple tinge. The absorption is tentatively assigned to a n(Te) → σ* transition.22 Dark blue solutions of 17a–c in acetonitrile and propionitrile are stable only for a few days. Notably, after some time the parent compounds 16a–c slowly form back and eventually precipitate. Electrospray mass spectra (MeCN, positive mode) of 17a–c show prominent mass clusters at m/z = 690.4, 786.3 and 882.2, which were unambiguously assigned to the radical cations on the basis of the correct isotopic patterns. The molecular conductivities (MeCN, c = 5 × 10−7 mol L−1) of Λ = 1800, 600 and 540 Ω−1 cm2 mol−1 confirm a high concentration of electrolytes in solutions of 17a–c. The one-electron oxidation upon going from (2,6-Mes2C6H3E)2 (16b, E = Se; 16c, E = Te) to the radical cation [(2,6-Mes2C6H3E)2]˙+ (17b, E = Se; 17c, E = Te) is well reflected by multinuclear NMR spectroscopy. The 77Se NMR spectra show signals at δ = 426.2 (CDCl3) for 16b and at δ = 1362.3 (CDCl3, 223 K) and 1362.4 (MeCN) for 17b, respectively. The 125Te NMR spectra exhibit signals at δ = 322.2 (CDCl3) for 16c and at δ = 1703.8 (CDCl3) and 1698.7 (MeCN) for 17c, respectively. These values are in good agreement with those calculated for (PhSe)2 (1b, δ = 451.5), [(PhSe)2]˙+ (2b, δ = 1237.2 and 1493.0; average 1365.1), (PhTe)2 (1c, δ = 185.1) and [(PhTe)2]˙+ (1c, δ = 1346.8 and 1778.9; average 1562.9). 1H and 13C NMR spectra gave expectedly broad signals. The paramagnetism of the radical cations was unambiguously confirmed by EPR spectroscopy and SQUID magnetometry. X- and Q-band field-sweep EPR spectra measured in frozen CH2Cl2/THF solution (1:1) show that all three species have g values that deviate significantly from the free-electron value. All radicals exhibit a rhombic g-matrix with the same g-value ordering, inferring the electronic structure of the three radicals is very similar. The g anisotropy for the radical cations [(2,6-Mes2C6H3E)2]˙+ increases in the order S (17a) < Se (17b) < Te (17c), as expected from the increasing spin–orbit coupling constant and for radicals where the spin density is predominately located on the central dichalcogen moiety. The EPR spectra of the [(2,6-Mes2C6H3E)2]˙+ radical cations feature a single proton hyperfine coupling resolved at g1 and g2 and this coupling was fully characterized by 1H Davies ENDOR spectra recorded at Q-band (Fig. 3, see ESI† for details). Simulation of the data yielded the hyperfine matrix (Table 2, see ESI† for details), which reveals a small but significant isotropic component of |aiso| = 7.1 MHz (aiso = (a1 + a2 + a3)/3), proving a small amount of spin density delocalises onto one of the m-terphenyl ligands.23 The findings were supported by spin density calculation (B3LYP/def2-TZVP) showing noticeable negative spin density on the para-H atom of one m-terphenyl ring (Fig. 4). The EPR spectra of the [(2,6-Mes2C6H3Se)2]˙+ (17b) at X- (Fig. 3 top) and Q-band (Fig. 3 bottom) exhibit two distinct Se hyperfine couplings along g1 (77Se, S = 1/2, 7.6% abundance), consistent with a radical containing a Se–Se moiety with a small asymmetry in the electronic structure. Along g2 and g3, 77Se hyperfine splittings were partially resolved and enabled an estimation of the remaining principal values of the two hyperfine interactions. Q-band Davies ENDOR was used to investigate the proton hyperfine interactions (see ESI† for details) which revealed (again) one largest resolved proton hyperfine interaction. The latter is, however, smaller than in the case of 17a, with |aiso| = 4.2 MHz. This result is consistent with the small asymmetry in the spin density inferred from the two inequivalent 77Se hyperfine couplings. The EPR spectra of [(2,6-Mes2C6H3Te)2]˙+ (17c) established principal values for the rhombic g-matrix (Fig. 3 and Table 2). Because of the low natural abundance of 125Te (S = 1/2, 7% abundance), the broad EPR spectrum and small background signals of the data only allowed estimates for the hyperfine couplings and were not of sufficient quality to resolve a potential inequivalence in the two Te nuclei. Along g1, the hyperfine coupling is resolved with |A(125Te)| = 1000 MHz; along g2 and g3, the 125Te hyperfine couplings were partially resolved and allowed the estimates shown in Table 2. Davies ENDOR was used to investigate the proton hyperfine interactions and showed that the largest proton hyperfine interaction has further decreased in comparison to 17a and 17b. Along g2, the coupling is |A| = 3.8 MHz, whereas for 17b: |A| = 6.4 MHz, and for 17a: |A| = 11.0 MHz. The experimental EPR data are very well modelled by DFT calculations (Fig. 4 and ESI† for details). The principal g-values are well reproduced as well as the two inequivalent hyperfine couplings for 77Se. The DFT data are able to unambiguously assign the largest proton coupling to the para-H atom of the central phenyl group of the one of the m-terphenyl substituents (Fig. 4). The trend to smaller 1H hyperfine values observed experimentally in the series 17a, 17b, and 17c follows the change in the orientation of two m-terphenyl substituents with respect to the central dichalcogen moiety: for 17a, the phenyl group π system of one m-terphenyl substituent is well orientated to allow overlap with the π-type orbitals carrying the unpaired electron on the S–S moiety and thus facilitates spin density delocalization. In contrast, the phenyl group of the second m-terphenyl substituent is approximately at 90° to this orientation and thus there is poor overlap with the S π-orbitals carrying the unpaired electron and less spin delocalisation. For 17c, the π-orbitals of both phenyl rings of the m-terphenyl substituent have essentially the same orientation with respect to the π-type orbitals of the Te–Te moiety carrying the unpaired electron but with relatively poor orbital overlap, resulting in an equivalent but relatively small delocalisation of the spin densities onto the two m-terphenyl substituents. As expected, 17b has a structure intermediate between 17a and 17c and this is reflected in the experimental 1H hyperfine coupling assigned to one para-H of the m-terphenyl group. The summed chalcogen-based spin densities from DFT calculations are 0.698, 0.734 and 0.778 for 17a, 17b, and 17c, respectively. These findings in conjunction with the EPR data leave no doubt that all three radical cations are characterized best as chalcogen-centred (i.e. ca. 70–80% spin density on the chalcogen atoms). The effective magnetic moments (μeff) of 17a–c at 300 K are 1.61, 1.48 and 1.41 μB, respectively, and reasonably close to the theoretically expected value of 1.744 μB, 1.7880 μB, 1.8625 μB, respectively, for a system of uncoupled paramagnetic centers with spins S = 1/2; the μeff decreases slightly with lowering temperature, that implies weak antiferromagnetic interactions in the solid state (see ESI† for details).
Fig. 3 Field-sweep EPR spectra for 17a (E = S) 17b (E = Se) and 17c (E = Te) measured in frozen CH2Cl2/THF solution (1:1), along with the corresponding simulations. |
Parameter | Principal values |
---|---|
17a | |
g | 2.0014, 2.0115, 2.0285 (2.0022, 2.0107, 2.0213) |
A(1H) | −3.0, −7.2, −11.0 (−1.8, −5.8, −8.5) |
17b | |
g | 1.9956, 2.0438, 2.1543 (2.0021, 2.0452, 2.1165) |
A(77Se) | −50, −100, 465 (−106, −116, 306) |
A(77Se) | −80, −115, 610 (−135, −140, 407) |
A(1H) | −1.8, −4.3, −6.4 (−1.2, −4.2, −6.1) |
17c | |
g | 1.9542, 2.0411, 2.4566 (2.0021, 2.0876, 2.3688) |
A(127Te) | 300, 350, −1000 (350, 370, −720) (352, 372, −724) |
A(1H) | −, −, −3.8 (−0.6, −2.6, −3.8) |
1H-NMR (CDCl3): δ = 7.27 (t, 1H, 3J = 8.1 Hz), 7.06 (d, 2H, 3J = 7.5 Hz), 7.02 (s, 4H), 3.07 (s, 1H, SH), 2.38 (s, 6H), 2.07 (s, 12H) ppm. 13C{1H}-NMR (CDCl3): δ = 138.7, 137.3, 136.2, 132.7, 128.4, 128.2, 128.0, 125.1, 21.2, 20.0 ppm.
1H-NMR (CDCl3): δ = 7.19 (t, 2H, 3J = 7.4 Hz), 6.88 (d, 4H, 3J = 7.5 Hz), 6.80 (8H), 2.35 (12H), 1.70 (24H) ppm. 13C{1H}-NMR (CDCl3): δ = 142.9, 137.4, 136.6, 136.5, 136.1, 129.2, 127.7, 127.1, 21.1, 20.3 ppm. Anal. calcd for C48H50S2 (691.06): C, 83.43; H, 7.29. Found C, 83.29; H, 6.91.
1H-NMR (CDCl3): δ = 7.19 (t, 1H, 3J = 7.5 Hz), 7.09 (d, 2H, 3J = 7.4 Hz), 7.04 (s, 4H), 2.41 (s, 6H), 2.10 (s, 12H), 1.11 (s, 1H, SeH), 1J (77Se–1H = 63.4 Hz) ppm. 13C{1H}-NMR (CDCl3): δ = 141.1, 138.9, 137.3, 136.0, 129.1, 128.4, 128.0, 126.0, 21.2, 20.0 ppm. 77Se-NMR (CDCl3): δ = 71.8 (d, 1J (1H–77Se = 63.4 Hz) ppm.
1H-NMR (CDCl3): δ = 7.19 (t, 2H, 3J = 7.5 Hz), 6.82 (d, 4H, 3J = 7.5 Hz), 6.74 (8H), 2.32 (12H), 1.72 (24H) ppm. 13C{1H}-NMR (CDCl3): δ = 144.1, 138.2, 136.4, 136.2, 132.8, 129.0, 127.9, 127.3, 20.9, 20.4 ppm. 77Se-NMR (CDCl3): δ = 426.2 ppm.
ESI MS (CH3CN, positive mode): m/z = 690.4 [C48H50S2]˙+ for 17a. Molar conductivity (CH3CN, c = 5 × 10−7 mol L−1): Λ = 1800 Ω−1 cm2 mol−1. SQUID: μeff (300 K) = 1.61 μB. Anal. calcd for C48H50F6S2Sb (926.80): C, 62.21; H, 5.44. Found C, 62.57; H, 5.62.
77Se-NMR (CDCl3, r.t.): δ = no signal. 77Se-NMR (CDCl3, 223 K): δ = 1362.3 ppm. 77Se-NMR (CD3CN): δ = 1362.4 ppm. ESI MS (CH3CN, positive mode): m/z = 786.3 [C48H50Se2]˙+ for 17b. UV-vis (CH2Cl2, c = 1 × 10−3 mol L−1): λmax = 710 nm. Molar conductivity (CH3CN, c = 5 × 10−7 mol L−1): Λ = 600 Ω−1 cm2 mol−1. SQUID: μeff (300 K) = 1.48 μB. Anal. calcd for C48H50F6Se2Sb (1105.59): C, 56.49; H, 4.94. Found C, 56.57; H, 5.12.
125Te-NMR (CD3CN): δ = 1698.7 ppm. 125Te-NMR (CDCl3): δ = 1703.8 ppm. ESI MS (CH3CN, positive mode): m/z = 882.2 [C48H50Te2]˙+ for 17c. UV-vis (CH2Cl2, c = 1 × 10−3 mol L−1): λmax = 583 nm. Molar conductivity (CH3CN, c = 5 × 10−7 mol L−1): Λ = 540 Ω−1 cm2 mol−1. SQUID: μeff (300 K) = 1.41 μB. Anal. calcd for C51H55F6NTe2Sb (1172.94): C, 52.22; H, 4.73; N, 1.19. Found C, 52.38; H, 4.62; N, 1.24.
Footnotes |
† Electronic supplementary information (ESI) available: General experimental considerations, cyclic voltammograms of 16a–c; UV-vis spectra of 16a–c and 17a–c; ESI-MS spectra of 17a–c; experimental and DFT EPR parameters of 17a–c; SQUID magnetic moments of 17a–c; crystal and refinement data of [14a][Sb2F11], [14b][As2F11], 16a, [17a][SbF6], [17b][SbF6]·CH2Cl2, [17c][SbF6]·NCCH2CH3; molecular structure of 16a; results from the conformational analysis of 1a–c, 2a–c, 13a–c and 14a–c; ionisation energies of 1a–c and 13a–c; Wiberg bond indices (WBIs), NBO and Mulliken charges of 1a–c, 2a–c, 13a–c and 14a–c. CCDC 999869–999873 and 1028017. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c4sc02964j |
‡ Dedicated to Professor Hubert Schmidbaur on the occasion of his 80th birthday. |
This journal is © The Royal Society of Chemistry 2015 |