DOI:
10.1039/D3DT00767G
(Paper)
Dalton Trans., 2023,
52, 6180-6186
A highly conducting mixed-valence nickel bis(dithiolene) salt [Et4N][Ni(Me-thiazSe-dt)2]2 with selone substitution†
Received
13th March 2023
, Accepted 14th April 2023
First published on 17th April 2023
Abstract
The prototypical [Ni(dmit)2] complex (dmit: 1,3-dithiole-2-thione-4,5-dithiolate) is modified here by combining the N–R substitution found in [Ni(R-thiazdt)2] complexes (R-thiazdt: N-alkyl-thiazoline-2-thione-4,5-dithiolate) and the selone substitution found in [Ni(dmiSe)2] complex (dmiSe: 1,3-dithiole-2-selone-4,5-dithiolate) to give a novel N-methyl substituted, radical anionic complex formulated as [Ni(Me-thiazSe-dt)2]1− (Me-thiazSe-dt: N-methyl-thiazoline-2-selone-4,5-dithiolate). Both this anionic complex and its mixed-valence Et4N+ salt crystallize with a rare cis arrangement of the two dithiolene ligands around the Ni atom. In the 1
:
2 [Et4N][Ni(Me-thiazSe-dt)2]2 salt, the complexes organize into dimerized chains well isolated from each other, giving the salt a strong one-dimensional character. It shows however a high RT conductivity of 4.6 S cm−1 and small activation energy of 33 meV, indicating a possible Mott insulator behavior, which is not suppressed under pressures up to 10 GPa.
Introduction
In the field of molecular conductors,1 the variability offered by organic chemistry gives endless possibilities to modify at will the electroactive molecules, be they organic donor or acceptor molecules or metal complexes with chemical variations on the nature and number of ligands. In that respect, square-planar metal bis(dithiolene) complexes2,3 provide a huge playfield as the two R1 and R2 substituents on the MS2C2R1R2 metallacycles can be engineered to control their shape and symmetry, particularly if R1 ≠ R2,4,5 or if R is chiral,6,7 but also their HOMO/LUMO energies and HOMO–LUMO gap8 (and associated optical properties in the NIR),9,10 the extend of intermolecular interactions such as the S⋯S overlap interactions but also hydrogen11 or halogen12 bonding interactions. The most successful complex in these series are the [M(dmit)2] ones (Chart 1, dmit: 1,3-dithiole-2-thione-4,5-dithiolate), which provided a large number of mixed-valence conducting (and even superconducting) salts,2,5 when associated with closed-shell (ammonium, phosphonium, sulfonium, …) or open-shell (tetrathiafulvalenium, ferricenium, …) cations. The dmit ligand itself can be chemically modified by atom substitution, as for example replacing the coordinating sulfur atoms by selenium ones in diselenolene dsit analogs,13 or replacing the outer thione moiety by a selone one in [M(dmiSe)2] complexes (Chart 1, dmiSe: 1,3-dithiole-2-selone-4,5-dithiolate).14
 |
| Chart 1 Chemical modifications of the dmit ligand in metal bis(dithiolene) complexes. | |
This introduction of the selone moiety in dmiSe complexes, in place of the outer thione moiety in dmit complexes, has been investigated with the aim to increase the dimensionality of the conducting systems by favoring direct Se⋯Se contacts, either between face-to-face complexes within a stack or slab, or in between conducting layers, as indeed observed first in 1991 in the mixed-valence salt [Me4N][Ni(dmiSe)2]2,15,16 followed by other salts such as [MexH4−xN][Ni(dmiSe)2]2 (x = 1–3) and Cs[Pd(dmiSe)2]2.17,18
Another modification of the dmit core is offered by the N-alkyl-thiazoline-2-thione-4,5-dithiolate ligand, abbreviated as R-thiazdt (Chart 1), where one of the sulfur atoms of the outer dithiole-2-thione ring is replaced by a N–R group. This substitution offers a rich palette of modifications of the dmit motif, by allowing a variety of R substituents on the nitrogen atom.19 Also these [M(R-thiazdt)2]1− complexes are known to systematically oxidize at lower potentials than their dmit counterparts and in most cases the oxidation of the nickel or gold R-thiazdt monoanionic complexes doesn't afford mixed-valence salts but the neutral complexes, either the closed-shell nickel [Ni(R-thiazdt)2]0 species20 or the open-shell [Au(R-thiazdt)2]˙ radicals,19d,e,21 known as single-component conductors. Rare examples of mixed-valence derivatives have been recently reported in R-thiazoline-diselenolene (R-thiazds) complexes such as [Ph4P][Au(Me-thiazds)2]2,22 or [Et4N][Ni(Me-thiazds)2]2.23 In that respect, introduction of the selone functionality in the R-thiazdt family to give the [M(R-thiazSe-dt)2] complexes (Chart 1), would combine the flexibility introduced by the R substituent on the nitrogen atom of R-thiazdt complexes with the possibility for extended intermolecular interactions observed earlier in dmiSe complexes.15–17
Only two such [M(R-thiazSe-dt)2] complexes have been reported earlier. The N-isopropyl gold complex24 [Au(iPr-thiazSe-dt)2]1− was prepared for comparison with the analogous thione derivative25 [Au(iPr-thiazdt)2]1−,˙, but its electrochemical oxidation failed to provide any crystalline material, neutral or mixed-valence. The N-ethyl nickel complex26 [Ni(Et-thiazSe-dt)2]1− was reported to oxidize to the neutral closed-shell complex [M(R-thiazSe-dt)2]0, a single component semiconductor with very low conductivity (σRT = 1.7 × 10−5 S cm−1), a consequence of its closed-shell character.
We describe here a novel member of this R-thiazSe-dt family, namely the N-methyl nickel complex [Ni(Me-thiazSe-dt)2]1− and demonstrate that its electrochemical oxidation leads to a highly conducting, mixed-valence 2
:
1 salt [Et4N][Ni(Me-thiazSe-dt)2]2.
Results and discussion
Syntheses and properties in solution
The preparation of the N-methyl anionic complex (Scheme 1) is analog to that described for the N-ethyl nickel complex26 or the N-isopropyl gold complex.25 It starts with the selenation of the N-methyl-thiazole-2-thione 1 with successively MeI and NaHSe to provide the corresponding selone 2.27 Metalation of 2 with LDA followed by reaction with elemental sulfur and in situ alkylation of the air-sensitive dithiolate with 3-bromopropionitrile provides the proligand 3 which can be easily purified and stored. Deprotection of 3 with NaOMe, reaction with NiCl2, 6H2O and metathesis with Bu4NBr provides the dianionic complex which is directly oxidized in air during the recrystallization step to the monoanionic complex [Bu4N][Ni(Me-thiazSe-dt)2]. As other monoanionic nickel complex, [Ni(Me-thiazSe-dt)2]1− exhibits a strong absorption in the NIR range, observed here at 1306 nm (ε = 25
000 M−1 cm−1), identical to that reported for the N-ethyl analog (λmax = 1308 nm, ε = 26
900 M−1 cm−1).26
 |
| Scheme 1 Synthetic route to the anionic complex [NBu4][Ni(Me-thiazSe-dt)2]. | |
Redox properties were evaluated by cyclic voltammetry (Fig. S1 in ESI†) and results are collected in Table 1. The N-methyl complex exhibits similar redox potentials than its N-ethyl analog for the 2−/1− and 1−/0 redox processes, albeit the 1−/0 process is affected here by precipitation at the electrode, a phenomenon which was not observed with the N-ethyl analog. Also, the 1+/0 redox process identified in the N-ethyl complex at Epa/Epc = 0.76/0.63 V could not be observed here, a probable consequence of this decreased solubility. As already observed, the introduction of the outer selone moiety (rather than the thione one in R-thiazdt complexes) leads to a higher oxidation potential for 1−/0 redox process (cf.Table 1). Nevertheless, electrocrystallization of the monoanionic complex was attempted in the presence of Et4NPF6 salt as electrolyte. In strong contrast with the oxidation of the N-ethyl analog which afforded the 1e− oxidized neutral complex [Ni(Et-thiazSe-dt)2]0 (investigated as a single-component conductor),26 the electrocrystallization of [Ni(Me-thiazSe-dt)2]1− leads to a 1
:
2 mixed-valence tetraethylammonium salt formulated as [Et4N][Ni(Me-thiazSe-dt)2]2.
Table 1 Redox properties of [Ni(Me-thiazSe-dt)2]1− and reference complexes in CH2Cl2 with Bu4NPF6 0.1 M. E in V vs. SCE with scan rate of 100 mV s−1
Complex |
E
pa/Epc1(2−/1−) |
E
pa/Epc2(1−/0) |
E
pa/Epc3(0/1+) |
ΔE½ |
Ref. |
As Ph4P+ salt.
As Bu4N+ salt.
|
[Ni(Et-thiazdt)2]1− a |
−0.30/−0.36 |
0.23/0.17 |
1.08/0.99 |
0.53 |
28
|
[Ni(Me-thiazdt)2]1− a |
−0.30/−0.36 |
0.18/−0.01 |
— |
— |
19c
|
[Ni(Et-thiazSe-dt)2]1− b |
−0.25/−0.31 |
0.27/0.21 |
0.76/0.63 |
0.52 |
25
|
[Ni(Me-thiazSe-dt)2]1− b |
−0.23/−0.29 |
0.27/0.03 |
— |
— |
This work |
Solid-state properties
The 1
:
1 salt [Bu4N][Ni(Me-thiazSe-dt)2] crystallizes in the triclinic system, space group P
, with both ions in general position in the unit cell (Fig. 2a). The ligands adopt the rare cis configuration around the nickel atom with a position disorder refined with a 85
:
15 occupation ratio (Fig. S2 in ESI†). The radical anions [Ni(Me-thiazSedt)2]−˙ organize into dimers, separated from each other in the structure by the bulky Bu4N+ cations (Fig. 1).
 |
| Fig. 1 (a) Projection view along the a axis of the unit cell of [Bu4N][Ni(Me-thiazSe-dt)2]; (b) organization of the dimers along the a axis. | |
 |
| Fig. 2 Temperature dependence of the magnetic susceptibility of [Bu4N][Ni(Me-thiazSe-dt)2]. In insert, projection view perpendicular to the molecular plane of the dimer showing the overlap pattern. | |
The plane-to-plane distance between anions within the dimers amounts to 3.56 Å, smaller than twice the sulfur van der Waals radius (3.60 Å). This lets us infer the possibility for direct antiferromagnetic interactions between the radical anions. This assumption is confirmed by the temperature dependence of the magnetic susceptibility.
As shown in Fig. 2, the susceptibility exhibits indeed an activated behavior above 100 K, together with a Curie tail at the lower temperatures. A fit considering the sum of both contributions (eqn (1)), i.e. the singlet–triplet contribution of the two anionic complexes and a Curie tail of magnetic defects gives a spin dimer contribution of 0.97 (close to expected S = 1 value, per dimer), with a J/kB value of −668 K. The Curie contribution accounts for x = 6.4% of S = 1/2 magnetic defects.
|  | (1) |
The Et4N+ 1
:
2 mixed-valence salt obtained by electrocrystallization crystallizes in the triclinic system, space group P
, with the partially oxidized complex in general position and the Et4N+ cation disordered on inversion center. As in the structure of the monoanionic salt (see above), the complex adopts the rare cis geometry, with both methyl groups pointing on the same side of the molecule. Evolutions of the intramolecular bond lengths (Table 2) from the anionic species to its partially oxidized analog are weak but doesn't follow the usual trends, i.e. shortening of the C–S(Ni) bonds (bonds b, b′) and lengthening of the C
C bond (bond c).
Table 2 Comparison of averaged bond distances (Å) in the monoanionic [Bu4N][Ni(Me-thiazSe-dt)2] and mixed-valence [Et4N][Ni(Me-thiazSe-dt)2]2 salts. Both complexes are in general position, only the major component of the disordered monoanion is reported

|
|
Monoanion |
Mixed-valence |
Dist. (Å) |
Aver. dist. (Å) |
Dist. (Å) |
Aver. dist. (Å) |
a |
2.169(4) |
2.164 |
2.161(2) |
2.162 |
2.159(3) |
2.162(2) |
a′ |
2.161(4) |
2.161 |
2.169(2) |
2.169 |
2.161(4) |
2.168(2) |
b |
1.719(10) |
1.670 |
1.682(8) |
1.695 |
1.622(9) |
1.708(7) |
b′ |
1.713(10) |
1.717 |
1.726(7) |
1.718 |
1.721(9) |
1.710(8) |
c |
1.373(16) |
1.386 |
1.362(11) |
1.365 |
1.400(11) |
1.368(10) |
d |
1.719(10) |
1.750 |
1.747(7) |
1.741 |
1.782(8) |
1.735(7) |
e |
1.719(9) |
1.732 |
1.745(9) |
1.748 |
1.746(9) |
1.751(9) |
f |
1.819(8) |
1.829 |
1.812(8) |
1.806 |
1.839(9) |
1.800(8) |
g |
1.326(10) |
1.325 |
1.349(9) |
1.350 |
1.325(10) |
1.351(10) |
h |
1.390(11) |
1.383 |
1.394(9) |
1.392 |
1.376(10) |
1.390(9) |
A projection view along the a axis (Fig. 3a) show the one-dimensional organization, with limited contacts between the stacks in the (b,c) place. Intra- and inter-stack Se⋯Se intermolecular distances exceeds 4.19 Å, well above the Se⋯Se van der Waals contact (1.90 × 2 = 3.80 Å). One single S⋯S inter-stack contact involving the sulfur atom of the thiazole rings is identified at 3.51 Å, close to the van der Waals contact distance. Within the stacks, the complexes form an alternated chain (Fig. 3b), with two different overlap patterns I and II (Fig. 3c), associated with different interplanar distances, respectively 3.54 Å (I) and 3.60 Å (II). Calculations of the βHOMO–HOMO interaction energies confirm this assumption, with |βI| = 0.23 eV and |βII| = 0.19 eV. The calculated band structure (Fig. 4) shows four bands formed out the HOMO and LUMO orbitals of both complex orientations in the solid. The third one is partially filled suggesting that this salt could be metallic, albeit its limited dispersion (0.27 eV) and strong one-dimensional character can also lead to a charge localization and associated Mott insulator behavior.
 |
| Fig. 3 Structure of [Et4N][Ni(Me-thiazSe-dt)2]2 with (a) a projection view of the structure along the a axis, (b) a detail of the alternated stack of radical anions along the a axis with intra (I) and inter (II) dimer interactions, and (c) a detail of the I and II overlap patterns. | |
 |
| Fig. 4 Calculated band structure of [Et4N][Ni(Me-thiazSe-dt)2]2, with the Fermi level represented by a dashed red line, assuming a metallic filling of the levels. Γ, X, Y, Z, M, refer respectively to Γ = (0, 0, 0), X = (1/2, 0, 0), Y = (0, 1/2, 0), Z = (0, 0, 1/2), M = (1/2, 1/2, 0), of the Brillouin zone of the triclinic lattice. | |
The temperature dependence of the resistivity is shown in Fig. 5. It exhibits an activated behavior characteristic of a semi-conductor. At ambient pressure, the room temperature conductivity amounts to σRT = 4.6 S cm−1, with a very small activation energy (Eact = 0.033 eV), in line with a Mott insulator behavior. Under such circumstances, it is often possible to close the Mott gap under pressure, as observed for example in single-component conductors derived from neutral radical gold bis(dithiolene) complexes,19e but also from closed-shell nickel complexes.20,29 The evolution of the resistivity has been evaluated under pressure using a Diamond Anvil Cell (DAC) set up and is reported in Fig. 5. The RT conductivity increases by one order of magnitude (up to 53 S cm−1) at 5.6 GPa, with an activation energy divided by 2 (14 meV). At higher pressures between 5.6 and 9.8 GPa, the conductivity slightly decreases (Fig. S3 in ESI†).
 |
| Fig. 5 Temperature and pressure dependence of the resistivity of the mixed-valence salt [NEt4][Ni(Me-thiazSe-dt)2]2. | |
The inability of this system to exhibit a metallic state under high pressures despite its mixed-valence character can be attributed here to its strong one-dimensional nature, but could be also a consequence of the disorder brought by the Et4N+ cation. Indeed, there are two possible routes for the occurrence of an electron localization in such systems. In the Mott-type mechanism, the electronic repulsions drive the localization of electrons, as often observed in such narrow-band systems.30 On the other hand, the Anderson-type mechanism31,32 is known to induce electronic localization in the presence of an extrinsic disorder, as already observed in molecular conductors.33 Both mechanisms can be simultaneously operative in this salt.
In conclusion, we have shown here that, based on the prototypical [M(dmit)2] complexes, the combination of both the N–Me substitution found in [M(Me-thiazdt)2] complexes and the selone substitution found in [M(dmiSe)2] complexes could afford novel nickel dithiolene complexes [Ni(Me-thiazSe-dt)2] able to crystalize, upon oxidation, into mixed-valence highly conducting salts. A metallic state could not be reached in this Et4N+ salt, even under very high pressures (up to 10 GPa), a probable consequence of its strong 1D electronic character and possibly of the disorder introduced by the Et4N+ cation. We believe that the use of smaller onium salts such as Me4N+, Me4P+ or Me3S+ could favor stronger lateral interactions between anionic stacks and hence more two-dimensional electronic structures. Also, they might suppress the disorder problem observed here, as often, with the Et4N+ cation. Such systems will be investigated in a close future.
Experimental section
General
Chemicals and materials from commercial sources were used without further purification. All the reactions were performed under an argon atmosphere. Melting points were measured on a Kofler hot-stage apparatus and are uncorrected. Mass spectra were recorded by the CRMPO, Rennes. Methanol, acetonitrile and dichloromethane were dried using an Inert pure solvent column device.
Syntheses
N-Me-1,3-thiazoline-2-selone (2).
To a solution of N-Me-1,3-thiazoline-2-thione271 (2 g, 15.2 mmol) in dry acetonitrile (50 mL), iodomethane (1.9 mL, 30.4 mmol) was slowly added. The reaction was stirred at 50 °C for 12 hours under inert atmosphere. The solution was then evaporated affording a crystalline solid. The solid was added to a freshly prepared solution of NaHSe prepared from sodium borohydride (1.3 g, 34.3 mmol) and selenium powder (2.4 g, 30.4 mmol) in ethanol under inert atmosphere. After stirring for 1 hour, the solution was added to a 10% acetic acid solution (100 mL). Ethanol was evaporated and the medium was extracted with dichloromethane. The organic phase was washed with water (2 × 200 mL) and dried over MgSO4. The solvent was then evaporated affording 2 as white powder. Yield: 96%; mp = 70 °C (lit.27 69–70 °C); Rf = 0.48 (SiO2, CH2Cl2); 1H NMR (300 MHz) δ: 7.19 (d, J = 4.4 Hz, 1H, CH), 6.83 (d, J = 4.4 Hz, 1H, CH), 3.80 (s, 3H, CH3); 13C NMR (75 MHz) δ: 40.6 (CH3), 116.1 (C
C), 134.9 (C
C), 180.5 (C
Se).
4,5-Bis(2-cyanoethylthio)-N-methyl-1,3-thiazoline-2-selone (3).
To a −10 °C cooled solution of 2 (1 g, 5.4 mmol) in dry THF (60 mL) was added a freshly prepared LDA solution in THF (8.3 mmol). After stirring for 30 min at −10 °C, sulfur (0.26 g, 8.2 mmol) was added and the solution was stirred for an additional 30 min. To the medium, LDA (11.1 mmol) was added, the reaction was then stirred at −10 °C for 3 h, followed by addition of sulfur (0.35 g, 11.1 mmol). After 30 min, 3-bromopropionitrile (4.5 mL, 54 mmol) was added dropwise and the reaction mixture was stirred overnight at RT. The solvent was evaporated in vacuo, and the residue was extracted with CH2Cl2. The organic phase was then washed with water and dried over MgSO4. The solvent was evaporated and the concentrated solution was purified by flash chromatography on silica gel using CH2Cl2 as eluent to afford 3 as a dark brown powder which was recrystallized from ethanol to obtain yellowish crystals. Yield: 41%; mp = 124 °C; Rf = 0.23 (SiO2, CH2Cl2); 1H NMR (300 MHz) δ 3.92 (s, 3H, CH3), 3.12 (m, 4H, CH2), 2.73 (m, 4H, CH2); 13C NMR (75 MHz) δ 182.1 (C
Se), 137.1 (C
C), 130.0 (C
C), 117.0 (CN), 39.2 (NCH3), 31.8 (SCH2), 31.6 (SCH2), 18.7 (![[C with combining low line]](https://www.rsc.org/images/entities/char_0043_0332.gif)
2CN), 18.6 (![[C with combining low line]](https://www.rsc.org/images/entities/char_0043_0332.gif)
2CN); HRMS (ESI) calcd for C10H11N3NaS380Se+: 371.91725, found: 371.9172; Elem. anal. calcd for C10H11N3S3Se: C, 34.48; H, 3.18; N, 12.06. Found: C, 34.56; H, 2.99; N, 12.08.
[Bu4N][Ni(Me-thiazSe-dt)2].
Under inert atmosphere, a solution of sodium methanolate in MeOH (40 mg, 1.74 mmol in MeOH: 20 mL) was added to the proligand 3 (190 mg, 0.55 mmol). After complete dissolution, the solution was stirred at room temperature for 30 mn. Then a solution of NiCl2,6H2O (66 mg, 0.27 mmol) in MeOH (5 mL) was added, followed 6 hours later by the addition of Bu4NBr (180 mg, 0.56 mmol). After stirring for 15 h, the formed precipitate was filtered and recrystallized under air from CH2Cl2/MeOH 20/80 to afford the monoanionic complex as dark red crystals. Yield: 54% (120 mg); mp = 188 °C; MS (TOF MS ES) calcd for [C8H6N2NiS6Se2]−: 539.65 found: 539.65; Elem. anal. calcd for C24H42N3NiS6Se2: C, 36.88; H, 5.42; N, 5.38. Found: C, 36.32; H, 5.18; N, 5.36.
Electrocrystallization
Under inert conditions in a U shaped electrocrystallization cell equipped with Pt electrodes, Et4NPF6 (200 mg) and the electro active monoanionic complex [Bu4N][Ni(Me-thiazSe-dt)2] (10 mg) were dissolved in CH3CN/CH2Cl2 (90
:
10, 15 mL). The current intensity was adjusted to 0.5 μA between the electrodes, and the reaction was left during five days at room temperature. Crystals of the mixed-valence salt were collected on the anode as black crystalline needles.
X-ray crystallography
Data collections were performed on XtaLAB AFC11 Rigaku diffractometer for [Bu4N][Ni(Me-thiazSe-dt)2] and on an APEXII Bruker-AXS diffractometer equipped with a CCD camera for [Et4N][Ni(Me-thiazSe-dt)2]2. Structures were solved by direct methods using the SIR97 program,34 and then refined with full-matrix least-square methods based on F2 (SHELXL-97)35 with the aid of the WINGX program.36 All non-hydrogen atoms were refined with anisotropic atomic displacement parameters. H atoms were finally included in their calculated positions. Details of the final refinements are summarized in Table 3.
Table 3 Crystallographic dataa,b
|
[Bu4N] [Ni(Me-thiazSe-dt)2] |
[Et4N] [Ni(Me-thiazSe-dt)2]2 |
R
1 = ||Fo| − |Fc||/|Fo|.
wR2 = [w(Fo2 − Fc2)2]/[w(Fo2)2]1/2.
|
CCDC |
2246654† |
2246655† |
Formulae |
C24H42N3NiS6Se2 |
C12H16N2.50NiS6Se2 |
FW (g mol−1) |
781.59 |
604.26 |
System |
Triclinic |
Triclinic |
Space group |
P![[1 with combining macron]](https://www.rsc.org/images/entities/char_0031_0304.gif) |
P![[1 with combining macron]](https://www.rsc.org/images/entities/char_0031_0304.gif) |
a (Å) |
8.5113(2) |
8.2371(11) |
b (Å) |
13.2483(4) |
11.5377(17) |
c (Å) |
15.9417(4) |
12.3255(16) |
α (deg) |
78.770(3) |
65.007(5) |
β (deg) |
79.406(2) |
87.675(5) |
γ (deg) |
73.437(3) |
70.068(5) |
V (Å3) |
1674.19(8) |
990.7(2) |
T (K) |
293(2) |
293(2) |
Z
|
2 |
2 |
D
calc (g cm−3) |
1.550 |
2.026 |
μ (mm−1) |
3.147 |
5.285 |
Total refls |
30 948 |
14 609 |
Abs corr. |
Multi-scan |
Multi-scan |
Uniq refls (Rint) |
8134 (0.0343) |
4521 (0.0858) |
Uniq refls (I > 2σ(I)) |
4282 |
2054 |
R
1, wR2 |
0.0703, 0.1832 |
0.0638, 0.1186 |
R
1, wR2 (all data) |
0.1431, 0.2156 |
0.1803, 0.1486 |
GOF |
1.02 |
0.951 |
Band structure calculations
The tight-binding band structure calculations were of the extended Hückel type37 and used a modified Wolfsberg–Helmholtz formula38 to calculate the non-diagonal Hμν values. All valence electrons were considered in the calculations and the basis set consisted of Slater-type orbitals of double-ζ quality for Ni, S and Se atoms, and single-ζ quality for C and H atoms. The ionization potentials, contraction coefficients and exponents were taken from previous work.39
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This work was supported in part by Universite de Rennes (PhD grant to H. H.). The stay of H. H. to RIKEN was supported in part by RIKEN International Program Associate and by Rennes Métropole.
References
- P. Batail, Chem. Rev., 2004, 104, 4887–4890 CrossRef CAS.
- R. Kato, Chem. Rev., 2004, 104, 5319–5346 CrossRef CAS PubMed.
- N. Robertson and L. Cronin, Coord. Chem. Rev., 2002, 227, 93–127 CrossRef CAS.
-
(a) D. Schallenberg, N. Pardemann, A. Villinger and W. Seidel, Dalton Trans., 2022, 51, 13681–13691 RSC;
(b) O. Jeannin, J. Delaunay, F. Barrière and M. Fourmigué, Inorg. Chem., 2005, 44, 9763–9770 CrossRef CAS PubMed;
(c) H. Benjamin, M. L. Müller, S. Afanasjevs, K. V. Kamenev and N. Robertson, Dalton Trans., 2020, 49, 13786–13796 RSC.
-
(a) D. Belo and M. Almeida, Coord. Chem. Rev., 2010, 254, 1479–1492 CrossRef CAS;
(b) D. Belo, H. Alves, E. B. Lopes, M. T. Duarte, V. Gama, R. T. Henriques, M. Almeida, A. Pérez-Benítez, C. Rovira and J. Veciana, Chem. – Eur. J., 2001, 7, 511–519 CrossRef CAS.
- R. Perochon, C. Poriel, O. Jeannin, L. Piekara-Sady and M. Fourmigué, Eur. J. Inorg. Chem., 2009, 5413–5421 CrossRef CAS.
- F. Pop and N. Avarvari, Coord. Chem. Rev., 2017, 346, 20–31 CrossRef CAS.
- E. Canadell, Coord. Chem. Rev., 1999, 185–186, 629–651 CrossRef CAS.
- P. Deplano, L. Pilia, D. Espa, M.-L. Mercuri and A. Serpe, Coord. Chem. Rev., 2010, 254, 1434–1447 CrossRef CAS.
- F. Camerel and M. Fourmigué, Eur. J. Inorg. Chem., 2020, 508–522 CrossRef CAS.
-
(a) S. Yokomori, S. Dekura, A. Ueda, R. Kumai, Y. Murakami and H. Mori, J. Mater. Chem. C, 2021, 9, 10718–10726 RSC;
(b) H. Hachem, N. Bellec, M. Fourmigué and D. Lorcy, Dalton Trans., 2020, 49, 6056–6064 RSC;
(c) S. Yokomori, A. Ueda, T. Higashino, R. Kumai, Y. Murakami and H. Hatsumi, CrystEngComm, 2019, 21, 2940–2948 RSC.
-
(a) H. Hachem, O. Jeannin, M. Fourmigué, F. Barriere and D. Lorcy, CrystEngComm, 2020, 22, 3579–3587 RSC;
(b) R. Kato, Bull. Chem. Soc. Jpn., 2014, 87, 355–374 CrossRef CAS.
-
(a) R. M. Olk, A. Roehr, J. Sieler, K. Koehler, R. Kirmse, W. Dietzsch, E. Hoyer and B. Olk, Z. Anorg. Allg. Chem., 1989, 577, 206–216 CrossRef CAS;
(b) S. Curreli, P. Deplano, M.-L. Mercuri, L. Pilia, A. Serpe, F. Bigoli, M. A. Pellinghelli, E. Coronado, C. J. Gomez-Garcia and E. Canadell, J. Solid State Commun., 2002, 168, 653–660 CrossRef CAS.
-
(a) H. Kobayashi, R. Kato and A. Kobayashi, Synth. Met., 1991, 42, 2495–2498 CrossRef CAS;
(b) B. Olk, R. M. Olk, J. Sieler and E. Hoyer, Synth. Met., 1991, 42, 2585–2588 CrossRef CAS.
- J. P. Cornelissen, D. Reefman, J. G. Haasnoot, A. L. Spek and J. Reedijk, Rec. Trav. Chim. Pays-Bas, 1991, 110, 345–346 CrossRef CAS.
- J. P. Cornelissen, B. Pomarède, A. L. Spek, D. Reefman, J. G. Haasnoot and J. Reedijk, Inorg. Chem., 1993, 32, 3720–3726 CrossRef CAS.
- A. Sato, H. Kobayashi, T. Naito, F. Sakai and A. Kobayashi, Inorg. Chem., 1997, 36, 5262–5269 CrossRef CAS.
- R. Kato, Y.-L. Liu, H. Sawa, S. Aonuma, A. Ichikawa, H. Takahashi and N. Mori, Solid State Commun., 1995, 94, 973–977 CrossRef CAS.
-
(a) S. Eid, M. Guerro and D. Lorcy, Tetrahedron Lett., 2006, 47, 8333–8336 CrossRef CAS;
(b) M. C. Aragoni, M. Arca, F. A. Devillanova, F. Isaia, V. Lippolis, A. Mancini, L. Pala, A. M. Z. Slawin and J. D. Woollins, Inorg. Chem., 2005, 44, 9610–9612 CrossRef CAS PubMed;
(c) S. Eid, M. Fourmigué, T. Roisnel and D. Lorcy, Inorg. Chem., 2007, 46, 10647–10654 CrossRef CAS PubMed;
(d) N. Tenn, N. Bellec, O. Jeannin, L. Piekara-Sady, P. Auban-Senzier, J. íñiguez, E. Canadell and D. Lorcy, J. Am. Chem. Soc., 2009, 131, 16961–16967 CrossRef CAS PubMed;
(e) G. Yzambart, N. Bellec, N. Ghassan, O. Jeannin, T. Roisnel, M. Fourmigué, P. Auban-Senzier, J. íñiguez, E. Canadell and D. Lorcy, J. Am. Chem. Soc., 2012, 134, 17138–17148 CrossRef CAS PubMed.
- H. Hachem, H.-B. Cui, T. Tsumuraya, R. Kato, O. Jeannin, M. Fourmigué and D. Lorcy, J. Mater. Chem. C, 2020, 8, 11581–11592 RSC.
- Y. Le Gal, T. Roisnel, P. Auban-Senzier, N. Bellec, J. íñiguez, E. Canadell and D. Lorcy, J. Am. Chem. Soc., 2018, 140, 6998–7004 CrossRef CAS PubMed.
- Y. Le Gal, H.-B. Cui, P. Alemany, E. Canadell, R. Kato, T. Roisnel, V. Dorcet, M. Fourmigué and D. Lorcy, J. Mater. Chem. C, 2021, 9, 12291–12302 RSC.
- H. Hachem, H.-B. Cui, R. Kato, P. Alemany, E. Canadell, O. Jeannin, M. Fourmigué and D. Lorcy, Inorg. Chem., 2023, 62, 4197–4209 CrossRef CAS PubMed.
- A. Filatre-Furcate, F. Barrière, T. Roisnel, O. Jeannin and D. Lorcy, Inorg. Chim. Acta, 2018, 469, 255–263 CrossRef CAS.
- A. Filatre-Furcate, N. Bellec, O. Jeannin, P. Auban-Senzier, M. Fourmigué, J. íñiguez, E. Canadell, B. Brière, V. Ta Phuoc and D. Lorcy, Inorg. Chem., 2016, 55, 6036–6046 CrossRef CAS PubMed.
- H. Hachem, H.-B. Cui, R. Kato, O. Jeannin, F. Barrière, M. Fourmigué and D. Lorcy, Inorg. Chem., 2021, 60, 7876–7886 CrossRef CAS PubMed.
- N. Bellec, D. Lorcy, K. Boubekeur, R. Carlier, A. Tallec, S. Los, W. Pukacki, M. Trybula, L. Piekara-Sady and A. Robert, Chem. Mater., 1999, 11, 3147–3153 CrossRef CAS.
- A. Filatre-Furcate, N. Bellec, O. Jeannin, P. Auban-Senzier, M. Fourmigué, A. Vacher and D. Lorcy, Inorg. Chem., 2014, 53, 8681–8690 CrossRef CAS PubMed.
- H. Cui, T. Tsumuraya, T. Miyazaki, Y. Okano and R. Kato, Eur. J. Inorg. Chem., 2014, 3837–3840 CrossRef CAS.
-
N. F. Mott, Metal-Insulator Transitions, Barnes and Noble, New York, 1977 Search PubMed.
- P. W. Anderson, Phys. Rev., 1958, 109, 1492–1505 CrossRef CAS.
-
W. Hayes and A. M. Stoneham, Defects and Defect Processes in Non -metallic Solids, Wiley, New York, 1985, ch. 8 Search PubMed.
-
(a) K. Riedl, E. Gati and R. Valentí, Crystals, 2022, 12, 1689 CrossRef CAS;
(b) M. Čulo, E. Tafra, B. Mihaljević, M. Basletić, M. Kuveždić, T. Ivek, A. Hamzić, S. Tomić, T. Hiramatsu, Y. Yoshida, G. Saito, J. A. Schlueter, M. Dressel and B. Korin-Hamzić, Phys. Rev. B, 2019, 99, 045114 CrossRef.
- A. Altomare, M. C. Burla, M. Camalli, G. Cascarano, C. Giacovazzo, A. Guagliardi, A. G. G. Moliterni, G. Polidori and R. Spagna, J. Appl. Crystallogr., 1999, 32, 115–119 CrossRef CAS.
- G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. Crystallogr., 2008, 64, 112–122 CrossRef CAS PubMed.
- L. J. Farrugia, J. Appl. Crystallogr., 2012, 45, 849–854 CrossRef CAS.
- M.-H. Whangbo and R. Hoffmann, J. Am. Chem. Soc., 1978, 100, 6093–6098 CrossRef CAS.
- J. H. Ammeter, H.-B. Bürgi, J. Thibeault and R. Hoffmann, J. Am. Chem. Soc., 1978, 100, 3686–3692 CrossRef CAS.
- S. Curreli, P. Deplano, M.-L. Mercuri, L. Pilia, A. Serpe, F. Bigoli, M. A. Pellinghelli, E. Coronado, C. J. Gómez-García and E. Canadell, J. Solid. State Chem., 2002, 168, 653–660 CrossRef CAS.
|
This journal is © The Royal Society of Chemistry 2023 |
Click here to see how this site uses Cookies. View our privacy policy here.