Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2023 Dec 22;63(1):173–183. doi: 10.1021/acs.inorgchem.3c02928

Asymmetric by Design: Heteroleptic Coordination Compounds with Redox-Active Dithiolene and 1,2,4,5-Tetrakis(isopropylthio)benzene Ligands

Che Wu , Lakshmi Nishanth Kakarla , Chandru P Chandrasekaran , Xiaodong Zhang , Joel T Mague , Stephen Sproules §, James P Donahue †,*
PMCID: PMC10777400  PMID: 38134365

Abstract

graphic file with name ic3c02928_0008.jpg

The 1,2,4,5-tetrakis(alkylthio)benzenes are redox-active organosulfur molecules that support oxidation to a stable radical cation. Their utility as ligands for the assembly of multimetal complexes with tailored functionality/property is unexamined. Here, 1,2,4,5-tetrakis(isopropylthio)benzene (tptbz, 1) is shown to bind PdCl2 at either one end, leaving the other open, or at both ends to form centrosymmetric [Cl2Pd(tptbz)PdCl2], 4. Ligand metathesis between Na2[(N≡C)2C2S2] (Na2mnt) and [Cl2M(tptbz)] (M = Pd, 2; M = Pt, 3) yields [(mnt)M(tptbz)] (M = Pd, 5; M = Pt, 6), but an alternative route involving transmetalation with [(mnt)SnMe2] delivers substantially greater yield. The mixed dithiolene-dithioether compound [(Ph2C2S2)Pt(tptbz)] (8) is formed by a similar transmetalation protocol using [(Ph2C2S2)SnnBu2]. Compounds 5, 6, and 8 are the first such heteroleptic complexes prepared by deliberate synthesis. The cyclic voltammetry of 8 reveals anodic waves at +0.14 and +0.97 V vs Fc+/Fc, which are attributed to successive dithiolene oxidation processes. While oxidized at +0.73 V as a free ligand, the redox-active MO of tptbz is pushed to a higher potential upon coordination to Pt2+ and is inaccessible. Calculations of the structures of [8]+ and of [((Cl2-3,5-C6H3)2C2S2)Pt(tptbz)]+ show that, in the latter, the dithiolene MOs are drawn down in energy into proximity with the tptbz MOs.

Short abstract

Tetrakis-1,2,4,5-isopropylthiobenzene (tptbz), a redox-active molecule with oxidation at +0.73 V vs Fc+/Fc in CH2Cl2, binds MCl2 (M = Pd, Pt) to afford [Cl2M(tptbz)] or dinuclear [Cl2Pd(tptbz)PdCl2]. Treatment of [Cl2M(tptbz)] with [((N≡C)2C2S2)SnMe2] or [(Ph2C2S2)SnnBu2] produces heteroleptic dithiolene-dithioether compounds [((N≡C)2C2S2)M(tptbz)] and [(Ph2C2S2)Sn(tptbz)]. Two oxidations are observed for [(Ph2C2S2)Sn(tptbz)], both attributable to the [Ph2C2S2]2− ligand. Calculations suggest that electron-withdrawing substituents on [Ph2C2S2]2− draw the tptbz-based MOs close enough to the HOMO to be competitive as locus for oxidation.

Introduction

In a series of recent reports,14 we have described the utility of 1,2,4,5-tetrakis(diphenylphosphino)benzene (tpbz; Figure 1) as a rigid connector between redox-active metallodithiolene groups, which can be reversibly oxidized to access the radical monoanionic form of the dithiolene ligand ((b) in Scheme 1). The tpbz ligand is an effective electronic insulator that permits only weak dipolar coupling between the peripheral ligand radicals, thereby enabling the creation of a coherent quantum state that in principle could function as a qubit in quantum computing or memory applications. Improvement of coherence lifetimes in qubit candidates based upon electron-spins can be accomplished by minimizing the presence of spin-active nuclei in the vicinity,5 the effect of which is to induce decoherence in the entangled state. In this regard, 31P with I = 1/2 at 100% natural abundance is a less desirable nuclide for incorporation into a molecule-based qubit than is 32S, with I = 0 at 95% natural abundance, because its nuclear spin contributes to the decohering magnetic background “noise”.

Figure 1.

Figure 1

Structures of the tpbz and tptbz ligands.

Scheme 1. Redox Levels of the Dithiolene Ligand: (a) Fully Reduced Ene-1,2-dithiolate, (b) Half-Oxidized Radical Monoanion, and (c) Fully Oxidized α-Dithione.

Scheme 1

This figure is a modification of a scheme previously published.2

Quite apart from the removal of spin-active nuclei that undercut coherent quantum state lifetimes, substitution of the tpbz connector for a tetrathioarene bridge creates altogether new possibilities for qubit engineering in that it, just as the dithiolene ligand opposite it in a heteroleptic complex, can sustain reversible oxidation to a spin-delocalized radical. Indeed, the cations that are accessible from 1,2,4,5-tetrathioarenes have been the subject of spectroscopic study, first by Pedulli and co-workers,6 and later by Chen and associates,7 who also reported solid-state conductivity measurements and the first instances of structural characterization of such radical cationic salts by X-ray crystallography. Thus, if amenable to the formation of heteroleptic complexes with dithiolene ligands, tetrathioarenes might enable the formation of multiqubit systems whose component spins reside in chemically distinctive environments and thus might be separately addressable.

The foregoing considerations motivated us to examine the feasibility of substituting the tpbz connector with the 1,2,4,5-tetrakis(isopropylthioether)benzene (tptbz) platform. Like tpbz, this tetrathioether has the virtues of amenability to scaled synthesis and stability in air, but it enjoys the further advantage that the alkyl groups of the thioether groups can be readily varied for effect upon solubility and steric profile. However, thioethers are generally weak ligands with a decided binding preference for soft, later transition metals of the second and third rows. Nonetheless, the potential advantages offered by tptbz in contrast to tpbz in supporting ligand-based coherent quantum states are sufficient to justify efforts to synthesize mono- and dimetallic compounds of the forms [(R2C2S2)M(tptbz)] and [(R2C2S2)M(μ-tptbz)M(S2C2R2)]. In this report, we present the leading results of this exploratory synthetic foray.

Experimental Section

Physical Methods

All 1H spectra were recorded at 25 °C with a Bruker Avance spectrometer operating at 300.13 MHz and referenced to the protonated solvent residual. Mass spectra (ESI+) were obtained with a Bruker micrOTOF II mass spectrometer with an Agilent Technologies 1200 Series LC. The UV–vis spectra were acquired on a Hewlett-Packard 8752A diode array spectrometer. Elemental analyses were performed by Galbraith Laboratories, Inc. of Nashville, TN, or by Kolbe Microanalytical Laboratory in Oberhausen, Germany. Electrochemical measurements were performed using a CHI 620C electrochemical analyzer workstation with a Ag/AgCl reference electrode, glassy carbon working electrode, Pt wire as the auxiliary electrode, and [nBu4N][PF6] as the supporting electrolyte. Under these conditions, the Cp2Fe+/Cp2Fe couple consistently occurred at +0.50 V. Spectroelectrochemical measurements were made with an Ocean Optics HR2000 spectrophotometer along with a Pine Research Instruments platinum honeycomb working electrode and a Ag/AgCl reference electrode. Procedural details regarding crystal growth, X-ray diffraction data collection, data processing, and structure solution and refinement are deferred to the Supporting Information (SI). Unit cell data and selected refinement statistics for the compounds that have been structurally identified are presented in Table 1; more complete crystallographic data are summarized in Table S1.

Table 1. Unit Cell Refinement Data for Crystallographically Characterized Compounds.

  tptbz C6(SiPr)6 [(pdt)SnnBu2] [Cl2Pd(tptbz)] [Cl2Pt(tptbz)]
formula C18H30S4 C24H42S6 C22H28S2Sn C18H30Cl2PdS4 C18H30Cl2PtS4
fw, g/mol 374.66 522.93 475.25 551.96 640.65
xtl system monoclinic triclinic tetragonal monoclinic monoclinic
space grp P21/c P I P21/c P21/c
a, Å 10.1189(12) 9.1370(2) 21.0089(3) 13.8803(12) 13.8706(8)
b, Å 8.3483(10) 9.6108(2) 21.0089(3) 9.7851(9) 9.6553(6)
c, Å 12.5378(15) 9.9046(2) 10.0948(2) 18.3062(16) 18.5310(11)
α, deg 90 72.062(1) 90 90 90
β, deg 98.834(2) 71.949(1) 90 98.921(3) 100.312(3)
γ, deg 90 63.875(1) 90 90 90
V, Å3 1046.6(2) 727.32(3) 4455.58(15) 2456.3(4) 2441.7(3)
Z 2 1 8 4 4
xtl size, mm 0.10 × 0.13 × 0.27 0.09 × 0.19 × 0.29 0.16 × 0.18 × 0.23 0.06 × 0.13 × 0.20 0.09 × 0.20 × 0.25
color, habit colorless column yellow plate colorless block yellow slat yellow plate
indices, h –13 ≤ h ≤ 13 –11 ≤ h ≤ 11 –32 ≤ h ≤ 32 –17 ≤ h ≤ 17 –18 ≤ h ≤ 18
indices, k –11 ≤ k ≤ 10 –12 ≤ k ≤ 12 –32 ≤ k ≤ 32 –12 ≤ k ≤ 12 –13 ≤ k ≤ 13
indices, l –16 ≤ l ≤ 16 –12 ≤ l ≤ 12 –15 ≤ l ≤ 15 –23 ≤ l ≤ 23 –25 ≤ l ≤ 25
2θ range, deg 2.074–57.586 4.818–53.106 4.476–66.450 4.374–54.526 4.774–58.666
refl. coll.a 19285 25064 128770 105892 116883
indp. refl. 19285 3018 7176 5476 6591
GooFb 1.038 1.038 1.045 1.124 1,018
R1,c,d wR2e,d 0.0446, 0.1185 0.0569, 0.1591 0.0236, 0.0649 0.0822, 0.2262 0.0453, 0.1137
R1,c,f wR2e,f 0.0598, 0.1270 0.0643, 0.1692 0.0250, 0.0653 0.1006, 0.2387 0.0618, 0.1269
  {[Cu(tptbz)][PF6]}n [Cl2Pd(tptbz)PdCl2] [(mnt)Pd(tptbz)] [(mnt)Pt(tptbz)] [(pdt)Pt(tptbz)]
solvent DMF 2DMI
formula C21H33CuF6NOPS4 C28H50Cl4N4O2Pd2S4 C22H30N2PdS6 C22H30N2PtS6 C32H40PtS6
fw, g/mol 652.23 957.56 621.24 709.93 812.09
xtl system monoclinic triclinic monoclinic monoclinic triclinic
space grp C2/m P P21/c P21/c P
a, Å 19.133(3) 7.2811(9) 5.7962(1) 5.7765(1) 15.9373(8)
b, Å 17.253(2) 10.6186(14) 19.2210(5) 19.2629(4) 16.3234(8)
c, Å 18.461(3) 12.9352(18) 24.1842(6) 24.2681(5) 22.6186(12)
α, deg 90 75.021(6) 90 90 85.242(2)
β, deg 107.016(4) 80.911(6) 94.386(1) 94.555(1) 78.027(2)
γ, deg 90 88.730(5) 90 90 62.321(2)
V, Å3 5827.3(14) 953.82 2686.44(11) 2691.83(9) 5096.9(5)
Z 8 1 4 4 6
2θ range, deg 2.306–66.588 5.784–52.984 4.562–63.908 2.276–41.251 1.84–61.276
xtl size, mm 0.08 × 0.21 × 0.23 0.10 × 0.18 × 0.41 0.06 × 0.26 × 0.37 0.06 × 0.07 × 0.31 0.07 × 0.18 × 0.30
color, habit colorless tablet yellow bar green plate yellow column yellow plate
indices, h –29 ≤ h ≤ 29 –8 ≤ h ≤ 9 –8 ≤ h ≤ 8 –10 ≤ h ≤ 10 –22 ≤ h ≤ 22
indices, k –26 ≤ k ≤ 26 –12 ≤ k ≤ 13 –28 ≤ k ≤ 28 –35 ≤ k ≤ 35 –23 ≤ k ≤ 23
indices, l –28 ≤ l ≤ 28 0 ≤ l ≤ 16 –35 ≤ l ≤ 35 –44 ≤ l ≤ 45 0 ≤ l ≤ 32
refl. coll.a 151087 3351 177124 213713 48672
indp. refl. 11481 3351 9273 18001 48672
GooFb 1.035 1.141 1.175 1.099 1.095
R1,c,d wR2e,d 0.0423, 0.1193 0.0876, 0.2270 0.0255, 0.0536 0.0433, 0.0874 0.0381, 0.0791
R1,c,f wR2e,f 0.0588, 0.1336 0.1110, 0.2386 0.0282, 0.0547 0.0578, 0.0925 0.0602, 0.0892
a

refl. coll. = total reflections collected.

b

GooF = {∑[w(Fo2Fc2)2]/(np)}1/2, where n is the number of reflections and p is the total number of parameters refined.

c

R1 = ∑||Fo| – |Fc||/∑|Fo|.

d

R indices for data cut off at I > 2σ(I).

e

wR2 = {∑[w(Fo2Fc2)2]/∑[w(Fo2)2]}1/2; w = 1/[σ2(Fo2) + (xP)2 + yP], where P = [2Fc2 + max(Fo2, 0)]/3.

f

R indices for all data.

General Considerations

Literature methods were implemented for the syntheses of the tptbz ligand, 1,8 disodium maleonitriledithiolate(2−),9 [(mnt)SnMe2],10 and 4,5-diphenyl-1,3-dithiol-2-one.11 All other reagents were purchased from commercial sources and used as received. Solvents were either dried with a system of drying columns from the Glass Contour Company (CH2Cl2, Et2O) or freshly distilled according to standard procedures (MeOH, CH3CN).12 All reactions described below were conducted under an atmosphere of N2, where mnt = maleonitriledithiolate(2−) = [(NC)2C2S2]2–; pdt = [Ph2C2S2]2– = 1,2-diphenylethylene-1,2-dithiolate(2−).

Syntheses

C6(SiPr)6

The following procedure is a modification of a published preparation of C6(SiPr)6.12 Sodium isopropyl thiolate (1.341 g, 13.66 mmol) was dissolved in 10 mL of dry DMF, and the solution was cooled in an ice bath to 0 °C. To this mixture, C6F6 (0.16 mL, 1.39 mmol) was added dropwise, and the mixture was allowed to warm to room temperature overnight, whereupon it was diluted with H2O (50 mL). The mixture was then extracted with Et2O (3 × 50 mL), and the combined organic extracts were washed with 2 × 50 mL of H2O and dried over anhydrous MgSO4. Hexakis(isopropylthio)benzene was obtained as a yellow crystalline solid after the removal of the solvent and trituration with 30 mL of cold MeOH. Yield: 0.4175 g, 58%. 1H NMR (δ, CDCl3): 3.75 (sept, 6 H, JHH = 6.7 Hz, −CH(CH3)2), 1.16 (d, 36 H, JHH = 6.7 Hz, −CH(CH3)2). 13C NMR (δ, CDCl3): 146.1, 41.0, 22.9.

[Cl2Pd(tptbz)], 2

To a 25 mL Schlenk flask charged with a stirring bar and PdCl2 (0.248 g, 1.41 mmol), 20 mL of MeCN were added under an active N2 flow. The mixture was heated to ∼70 °C until most of the solids dissolved. To this mixture was then added solid tptbz (0.528 g, 1.40 mmol), and the mixture was left to stir at ambient temperature overnight. All volatiles were removed from the reaction mixture under reduced pressure, and the orange solid residue was washed with Et2O (2 × 20 mL) and collected by filtration. Yield: 0.651 g, 1.18 mmol, 85%. 1H NMR (δ, CDCl3): 7.38 (s, 2 H, aromatic C–H), 3.98 (sept, 2 H, JHH = 6.8 Hz, PdSCH(CH3)2), 3.55 (sept, 2 H, JHH = 6.6 Hz, SCH(CH3)2)), 1.39 (d, 24 H, JHH = 6.5 Hz, −CH(CH3)2). Anal. Calcd for C18H30S4Cl2Pd: C, 39.16; H, 5.48; S, 23.24. Found: C, 39.17; H, 5.14; S, 20.84.

[Cl2Pt(tptbz)], 3

A 50 mL Schlenk flask with PtCl2 (0.241 g, 0.906 mmol) was charged with MeCN (15 mL) via a syringe, and the mixture was heated at ∼70 °C until most of the solids dissolved and a yellow solution was formed. To this mixture was added solid tptbz (0.345 g, 0.921 mmol) under an outward flow of N2, and the mixture was left to stir at ambient temperature overnight. All volatiles were then removed in vacuo. The solid residue was redissolved in 2 mL of CH2Cl2 and then precipitated by the addition of 50 mL of Et2O. The resulting pale yellow solid was isolated by cannula filtration, washed with an additional portion of Et2O (25 mL), and then dried under vacuum. Yield: 0.348 g, 60% yield. 1H NMR (δ, CDCl3): 7.47 (s, 2 H, aromatic C–H), 3.98–3.81 (m, 2 H, PtSCH(CH3)2), 3.63–3.51 (m, 2 H, SCH(CH3)2), 1.48–1.29 (m, 24 H, SCH(CH3)2). Anal. Calcd for C18H30S4Cl2Pt: C, 33.74; H, 4.72. Found: C, 33.71; H, 4.77.

[Cl2Pd(tptbz)PdCl2], 4

Method A

To a 50 mL Schlenk flask charged with PdCl2 (0.040 g, 0.22 mmol) was added MeCN (15 mL) via a syringe, and the mixture was heated at ∼70 °C until most solids dissolved. To this mixture was added a solution of [(tptbz)PdCl2] (0.124 g, 0.225 mmol) in CH2Cl2 (15 mL) via a cannula, which induced the immediate formation of a yellow precipitate. Stirring was continued at ambient temperature for 12 h. All volatile materials were removed in vacuo, and the remaining solids were collected as the product (0.165 g, 100%). 1H NMR (δ, DMSO-d6): 7.79 (s, 2 H, aromatic C–H), 3.83 (septet, 4 H, JH–H = 6.5 Hz, SCH(CH3)2), 1.46 (s, br, 12 H, SCH(CH3)2), 1.30 (d, 12 H, JH–H = 6.5 Hz, SCH(CH3)2). Calcd for [C18H28S4Cl4Pd2 + Na+]+: m/z 752.7897; Observed: m/z 752.7932; Error (δ) 4.72 ppm.

Method B

In a 50 mL Schlenk flask charged with PdCl2 (0.112 g, 0.632 mmol), MeCN (10 mL) was added under an active N2 flow. The mixture was heated to ∼70 °C until most of the solids dissolved, and the solvent was then removed under reduced pressure. Dichloromethane (20 mL) and solid tptbz (0.120 g, 0.320 mmol) were added, and the mixture was stirred at ambient temperature overnight. The solvent was again removed under vacuum, and the yellow solid residue was collected as the product.

[(mnt)Pd(tptbz)], 5

Method A

In a 25 mL Schlenk tube, Na2[mnt] (0.042 g, 0.23 mmol) was combined with dry CH2Cl2 (10 mL) under an atmosphere of N2. To this mixture, solid [Cl2Pd(tptbz)] (0.104 g, 0.188 mmol) was added along with an additional portion of dry CH2Cl2 (5 mL). The mixture was then stirred at ambient temperature overnight, at which point it was then filtered. The filtrate was reduced to a volume of ∼0.5 mL before being combined with dry MeOH (∼15 mL). The resulting solid precipitate was collected and recrystallized as green plates from dry CH2Cl2 (2 mL) by slow introduction of dry MeOH (50 mL). Yield: 0.0329 g, 0.0530 mmol, 28.1%. 1H NMR (δ, CDCl3): 7.49 (s, 2 H, aromatic C–H), 3.56 (sept, 2 H, JH–H = 6.6 Hz, PdSCH(CH3)2), 3.47, (sept, 2 H, JH–H = 6.6 Hz, SCH(CH3)2), 1.45 (d, 12 H, JHH = 6.7 Hz, PdSCH(CH3)2), 1.41 (d, 12 H, JHH = 6.7 Hz, SCH(CH3)2). UV–vis [CH2Cl2, λmax, nm (ε, M–1·cm–1)]: ∼338 (sh, ∼6600), 378 (2300). MS (ESI+) Calcd for [C22H30N2PdS6]+: m/z 620.9851; Observed: m/z 620.9839; Error (δ) 2.07 ppm. Anal. Calcd for C22H30N2S6Pd: C, 42.53; H, 4.87; N, 4.51. Found: C, 42.49; H, 4.86; N, 4.50.

Method B

In a 50 mL Schlenk flask, 2 (0.1080 g, 0.1957 mmol) was dissolved in 10 mL of dry CH2Cl2 under an atmosphere of N2. To this mixture, solid [(mnt)SnMe2] (0.0577 g, 0.200 mmol) was added under an outward flow of N2. While the mixture was stirred at room temperature overnight, it turned dark green. All volatiles were removed in vacuo, and the solid residue was purified on silica chromatography column packed with 1:1 hexanes/CH2Cl2 and eluted with 1:2 hexanes/CH2Cl2. The product was collected as the leading deep blue band. After the removal of the solvents, the solid residue was crystallized by a layered diffusion of MeOH into a CH2Cl2 solution. Yield: 0.0782 g, 0.126 mmol, 64.3%.

[(mnt)Pt(tptbz)], 6

In a 25 mL Schlenk flask, PtCl2 (0.1046 g, 0.393 mmol) was combined with MeCN (10 mL) under a N2 atmosphere and heated until a clear yellow solution was attained. This solution was then cooled to room temperature, whereupon solid tptbz (0.1493 g, 0.398 mmol) was added, and stirring was continued at ambient temperature overnight. The solvent was then removed in vacuo, and the solid residue was redissolved in CH2Cl2 (10 mL). Solid [(mnt)SnMe2] (0.1212 g, 0.419 mmol) was added, and stirring was maintained for an additional 12 h. The reaction mixture was then gravity filtered through paper, and the solid material thus separated was washed with further CH2Cl2 (40 mL). The filtrate was taken to dryness under reduced pressure, and the resulting solid residue was purified on a silica gel column that was packed as a slurry in, and eluted with, 1:1 CH2Cl2/hexanes. Compound 6 was collected as the leading orange-red band and crystallized by diffusion of Et2O vapor into a CH2Cl2 solution. Yield: 0.1516 g, 54.3%. 1H NMR (δ, CDCl3): 7.59 (s, 2 H aromatic C–H), 3.66–3.55 (m, 2 H, SCH(CH3)2, JH–H = 6.7 Hz), 3.53–3.46 (m, 2 H, SCH(CH3)2, JH–H = 6.7 Hz), 1.41 (d, 20 H, SCH(CH3)2, JH–H = 6.7 Hz), 1.21–1.17 (m, 4 H, SCH(CH3)2). UV–vis [CH2Cl2, λmax, nm (ε, M–1·cm–1)]: ∼294 (sh, ∼32,000), ∼333 (sh, 8200), 368 (6800). Calcd m/z for [M + H]+: 711.045431; Found: 711.04456; Error: 1.22 ppm. Anal. Calcd for C22H30N2S6Pt: C, 37.22; H, 4.26; N, 3.95; S, 27.10. Found: C, 37.21; H, 4.27; N, 3.94; S, 27.06.

[(pdt)SnnBu2], 7

To a solution of NaOMe (0.151 g, 2.80 mmol) in 20 mL of dry MeOH in a 50 mL Schlenk flask was added solid 4,5-diphenyl-1,3-dithiol-2-one (0.300 g, 1.11 mmol) under an active flow of N2. The mixture was left to stir at ambient temperature for 2 h, at which point solid nBu2SnCl2 (0.337 g, 1.11 mmol) was then added. Stirring was continued overnight, and all volatiles were then removed in vacuo. The solid residue was dissolved in CH2Cl2 (20 mL). This solution was filtered through packed Celite to remove the finely suspended NaCl byproduct, and the filtrate was then reduced to a volume of ∼4 mL. Hexanes (40 mL) were slowly introduced via a syringe, which induced the formation of the product as a yellow crystalline solid. Yield: 0.115 g, 21.8%. 1H NMR (δ, CDCl3): 7.17–6.96 (m, 10 H, aromatic C–H), 1.92–1.75 (m, 4 H, CH2), 1.74–1.64 (m, 4 H, CH2), 1.54–1.34 (m, 4 H, CH2), 0.97 (t, 6 H, JH–H = 7.3 Hz, CH3). MS (ESI+) Calcd for [C22H28S2Sn]+: m/z 477.0733; Observed: m/z 477.0735; Error (δ) 0.32 ppm.

[(pdt)Pt(tptbz)], 8

In a 50 mL Schlenk flask, PtCl2 (0.1262 g, 0.4745 mmol) and 20 mL of MeCN were combined under a N2 atmosphere and heated at ∼80 °C until a clear yellow solution formed. To this mixture was added solid tptbz (0.1782 g, 0.476 mmol), and the mixture was left to stir overnight at this same temperature. The solvent was then removed in vacuo, and the remaining solid residue was combined with dry CH2Cl2 (20 mL) followed by solid [(pdt)SnnBu2] (0.2295 g, 0.483 mmol). The resulting red mixture was stirred at ambient temperature overnight under N2, after which time the solvent was removed under reduced pressure and the remaining solid residue was washed with MeOH (20 mL) and Et2O (2 × 20 mL). Purification was accomplished on a silica gel column eluted with 1:1 CH2Cl2/hexanes. Following the collection of a leading purple band, 8 emerged as the subsequent orange-red band. After the removal of the solvent mixture, 8 was crystallized as deep orange-red crystals after diffusion of Et2O into a CH2Cl2 solution. Yield: 0.1099 g, 0.1353 mmol, 28.6%. Rf = 0.71 (CH2Cl2). 1H NMR (δ, CDCl3): 7.58 (s, 2 H, tptbz aromatic C–H), 7.37–7.29 (m, 4 H, phenyl C–H), 7.18–7.05 (m, 6 H, phenyl C–H), 3.57 (sept, 4 H, JH–H = 6.7 Hz, SCH(CH3)2), 1.44–1.32 (m, 24 H, SCH(CH3)2). UV–vis [CH2Cl2, λmax, nm (ε, M–1·cm–1)]: ∼335 (sh, 8050). MS (ESI+) Calcd for [C32H40S6Pt]+: m/z = 812.1101; Observed: m/z 812.108; Error (δ) 2.61 ppm.

[[Cu(tptbz)][PF6]]n, 9

To a Schlenk flask with a stirring solution of [Cu(MeCN)4][PF6] (0.398 g, 1.07 mmol) in CH2Cl2 (10 mL), solid tptbz (0.4009 g, 1.070 mmol) was added under an outward flow of N2 followed by additional CH2Cl2 (10 mL). The reaction mixture became turbid, and a white precipitate quickly formed. The mixture was left to stir at ambient temperature overnight, after which time the solid residual was collected by filtration, washed with fresh CH2Cl2 (2 × 10 mL), and dried under vacuum. Yield: 0.2763 g, 0.474 mmol, 44.3%. Crystallization of polymeric 9 as colorless tablets was accomplished by the diffusion of Et2O into a DMF solution. 1H NMR (δ, DMSO-d6): 7.78 (s, 2 H, aromatic C–H), 3.95–3.64 (m, 4 H, SCH(CH3)2), 1.24 (d, 24 H, JH–H = 6.5 Hz, SCH(CH3)2).

Syntheses and Structures

The tptbz ligand (1; Scheme 2) is attained in good yield from the corresponding 1,2,4,5-tetrachlorobenzene by straightforward reaction with iPrSNa+ in N,N-dimethylformamide8 or hexamethylphosphoramide.13 Although this approach is general in leading to arene tetrathioethers, other alkyl thiolates are less conveniently obtained (e.g., MeS) or less conducive to the formation of the crystalline material (e.g., nBuS). Slow evaporation of an acetonitrile solution of 1 reliably deposited columnar crystals in monoclinic P21/c (no. 14) with the central arene ring coincident with a crystallographic inversion center (Figure 2). The S–Carene bond lengths in 1 are ∼0.06 Å shorter than the S–CiPr bond lengths, likely reflecting the effect of S p-π arene-π interaction. An earlier structural characterization reported for 1 was conducted at a somewhat higher temperature (200 vs 150 K) than that used here.14

Scheme 2. Synthesis of 1,2,4,5-Tetrakis(isopropylthio)benzene Compounds.

Scheme 2

Figure 2.

Figure 2

Thermal ellipsoid plots (50% ellipsoids) of selected compounds characterized by X-ray diffraction. For clarity, all H atoms are omitted, and disorder in the alkyl groups of 5 and 7 is not shown.

Surprisingly little well-defined coordination chemistry has been described with 1,2,4,5-arene tetrathioethers in general, and with tptbz in particular and its simpler homologue 1,2-bis(isopropylthio)benzene, no metal complexes of which we are aware have been reported. As expected for a chelating dithioether, the tptbz ligand shows affinity primarily for soft or borderline soft late transition metals. Thus, when introduced to 1 equiv of tptbz, both PdCl2 and PtCl2 are readily coordinated at one end, while NiCl2 resists binding under the same conditions. Dithioether ligation for Ni2+ is known, but the majority of well-defined complexes that are known have this donor atom set incorporated within tetradendate1517 or macrocyclic hexadentate ligands18,19 whose kinetic and thermodynamic advantages overcome the otherwise tepid Ni–Sthioether interaction. The use of 2 equiv of PdCl2 in the reaction with 1 equiv of tptbz led to the formation of the dimetallic [Cl2Pd(tptbz)PdCl2] (4), as marked by its immediate precipitation due to lowered solubility in consequence of its centrosymmetry.

Compounds 2 and 3 are isostructural (Table 1 and Figure 2) and reveal quite similar bond lengths and degrees of planarity (Table 2). The M–Sthioether bond lengths in these complexes are, within experimental resolution, shorter than the M–Cl bond lengths by 0.035 and 0.059 Å for Pd and Pt, respectively. This difference appears to reflect a modest contribution from the chelate effect, as the (Pt–Cl)ave – (Pt–S)ave difference in cis-[PtCl2(SMe2)2] is 0.047 Å.20 A modest difference with no clearly discernible explanation is the S–Cisopropyl distance at the metalated end (∼1.86 Å) vs the open end (∼1.79 Å). Compared to the S–Cisopropyl distance in the free tptbz ligand (∼1.83 Å), this difference appears to have equal contribution from S–Cisopropyl elongation at the metalated end and S–Cisopropyl contraction at the open end. The structure of 4 occurs on an inversion center in triclinic P-1, its unique half being effectively indistinguishable from 2 in all metrical details (Table 2). The intramolecular Pd···Pd separation of 8.651(1) Å is slightly shorter than the 8.857 Å observed for [Cl2Pd(tpbz)PdCl2].2

Table 2. Selected Interatomic Distances (Å) and Angles (deg.) for Crystallographically Characterized Compoundsa.

  tptbz C6(SiPr)6 [(pdt)SnnBu2]
S–Carene 1.7695[14] 1.779[1] Sn–S 2.4401[5]
S–CiPr 1.8303[14] 1.838[1] S–C 1.775[2]
δb 0.074 0.119 Sn–C 2.142[3]
σc 0.038–0.109 0.057–0.164 θ (deg)d 28.3
  [Cl2M(tptbz)]
   
  M = Pd M = Pt [Cl2Pd(tpbzPdCl2]
M–S 2.2718[13] 2.2528[12] M–S 2.276[2]
M–Cl 2.3070[14] 2.3122[13] M–Cl 2.307[2]
MS–CiPr 1.859[6] 1.858[5] MS–CiPr 1.857[9]
MS–Carene 1.775[5] 1.788[4] MS–Carene 1.789[9]
openS–CiPr 1.779[8] 1.778[7] Pd···Pd 8.651(1)
openS–Carene 1.765[5] 1.766[5] Cl–M–Cl 90.77(11)
Cl–M–Cl 91.98(9) 90.35(8) S–M–S 89.99(10)
S–M–S 90.13(6) 90.45(6) θe 4.9(2)
θe 6.2(1) 4.9(1)    
  [(mnt)M(tptbz)]
     
  M = Pd M = Pt [(pdt)Pt(tptbz)]f [Cu(tptbz)]+n
M–Sdithiolene 2.2693[3] 2.2712[4] 2.2614[5] Cu–S 2.2980[3]
M–Sthioether 2.3246[3] 2.3025[4] 2.3095[5] S–Cthioether 1.7764[8]
Sdithiolene–C 1.7370[11] 1.738[2] 1.764[2] S–CiPr 1.851[1]
MSthioether–Carene 1.7797[10] 1.780[1] 1.781[2] Cu···Cu 8.619, 8.635
MSthioether–CiPr 1.8721[11] 1.860(2)g 1.866[3] S–Cu–Schelate 91.90[1]
openSthioether–Carene 1.7662[9] 1.757[1] 1.762[2] θ (deg)h 88.96
openSthioether–CiPr 1.8178[10] 1.825[2] 1.824[2]    
θi 6.69(2) 5.46(4) 2.83[5]    
a

Uncertainties are propagated according to Taylor, J. R. An Introduction to Error Analysis; 2nd ed.; University Science Books: Sausalito, CA, 1997, pp 73–77. The square bracket represents the uncertainties propagated in the averaging of chemically identical values.

b

δ = mean displacement of S atoms from the C6 arene plane.

c

σ = range of S atom displacements from the C6 arene plane.

d

θ = fold angle between S2Sn and S2C2 planes.

e

θ = angle between Cl2M and S2M planes.

f

All values are averages for three independent molecules in the asymmetric unit of the cell.

g

Because of static disorder afflicting one of the SiPr groups, this table entry is a single, unaveraged bond length.

h

θ = angle between S2Cu chelate planes.

i

θ = angle between S2,dithioleneM and S2,dithioetherM planes.

When introduced to [Cu(MeCN)4][PF6] in CH2Cl2, tptbz forms a 1D cationic coordination polymer (9; Scheme 2 and Figure 2) similar to that formed by hexakis(methylthio)benzene and Cu(I) and Ag(I).21 The intraligand tptbz bond lengths in 9 differ little from those of the free ligand. With the 1,2,4,5-tetramethylmercaptobenzene (tmmb) ligand, Cu(I) halide precursors yield halide-linked 1D or 2D polymers,22 as well as discrete halide-bridged dimetallic species,23 instead of homoleptic thioether polymers.

Compounds 2 and 3 are readily subject to halide substitution for dithiolene reactions via either salt metathesis or through the agency of tin-transmetallating compounds, whose usefulness in chloride-for-dithiolene exchange is well-established.13,2429 Thus, [(mnt)Pd(tptbz)] (5) is produced in modest yield (28%) from the heterogeneous reaction between 2 and Na2mnt in CH2Cl2 but, as has been our consistent observation, in appreciably better yield (64%) via chloride-for-mnt2– exchange with [(mnt)SnMe2]. The better margins produced by such tin reagents are likely due to the effect of greater homogeneity in the reaction solution. Similar reactions of 3 with [(mnt)SnMe2] and [(pdt)SnnBu2] (7) afforded [(mnt)Pt(tptbz)] (6) and [(pdt)Pt(tptbz)] (8). Compounds 5 and 6, as well as their precursors 2 and 3, are indefinitely stable to moisture and air, both in the solid state and in solution. In contrast, while 8 is amenable to purification by column chromatography and to crystallization in the open air, its solutions slowly deteriorate to a dark orange oily material over a period of days if not maintained in the dark under a protecting N2 atmosphere. Considering the quite extensive body of heteroleptic dithiolene compounds of the group 10 metals, which includes many dithiolene-dithione compounds,3033 it is noteworthy that 5, 6, and 8 appear to be first heteroleptic dithiolene-dithioether complexes made by deliberate synthesis. A platinum bis(trifluoromethyl)dithiolene (tfd) dithioether complex, formed serendipitously by the addition of 2 equiv of 2,3-dimethyl-1,3-butadiene to [Pt(tfd)2], stands as the only prior example.34

Compounds 5, 6, and 8 retain the near ideal planarity of 2 and 3, as gauged by θ values of ∼3–7° (Table 2). The principal metrical change in the immediate metal ion environment is a substantial lengthening of the M–Sthioether bond lengths by ∼0.05 Å compared to the precursor dichlorides (Table 2), which reflects the appreciably greater ligand field strength of the chelating dithiolene ligand. The Pt–Sdithiolene bond length in 8 is shorter than that observed in a series of [(pdt)Pt(C≡NR)2] complexes by a modest, but significant, amount within experimental resolution, ∼0.01 Å.35 The comparative binding weakness of tptbz that is implicated by this difference is borne out in the finding that 8 is not directly accessible from [(pdt)2Pt] by dithiolene displacement with tptbz, though such an approach is efficacious for the synthesis of a range of other heteroleptic dithiolene complexes of the group 10 metals.33,35,36

In the crystalline state, the packing arrangement for isostructural 5 and 6 is such that the square planar molecules form orderly, columnar stacks that arrange approximately along the a-axis of the monoclinic cell (Figure S23). Compound 8, however, occurs in triclinic P-1 with an atypical three full molecules in the asymmetric unit such that Z = 6. The disposition of these separate molecules relative to one another is not simply described, as they are not coplanar or mutually orthogonal or related by any pseudosymmetry operation. Rather, the planes defined by their PtS4 coordination cores form angles of ∼56°, ∼64°, and ∼67° with one another that appear to be guided by intermolecular Sthioether···Sthioether close contacts (Figure S24). Distances of 3.414 Å separating S(4) of molecule 1 from S(10) from molecule 2 and 3.508 Å between S(9) of molecule 2 and S(15) of molecule 3 are less than twice the crystallographic van der Waals radius for sulfur (1.8 Å)37 and suggest that soft–soft dispersion-type forces may be operative in the crystal packing. Furthermore, C–H···arenecentroid hydrogen bonding contacts likely play an additional role in dictating the packing, as seen in the pair of H-bonds that relate molecule 3 of compound 8 to its inversion counterpart (Figure S25).

Electrochemistry and Electronic Structure

The tptbz ligand sustains a reversible oxidation at +0.73 V vs Fc+/Fc (Figure 3, top) because the four thioether sulfur atoms collectively impart appreciable π-electron density to the arene ring. The oxidation potential for this ring system, when it is implemented as a chelating ligand, is anticipated to shift to higher potential owing to the diversion of two sulfur lone pairs in σ donation to a metal cation.

Figure 3.

Figure 3

Cyclic voltammogram of the free tptbz ligand in CH2Cl2 with the [nBu4N][PF6] supporting electrolyte (top). Cyclic voltammogram of [(mnt)Pt(tptbz)] revealing a 1-e oxidation that is largely dithiolene ligand-based (middle). Cyclic voltammogram of [(pdt)Pt(tptbz)] revealing successive oxidations that are assigned to the [Ph2C2S2]2– ligand (bottom). Potentials are referenced to [Cp2Fe]+/Cp2Fe, but decamethyl ferrocene was used as an internal additive to avoid interference with the analyte.

For [(mnt)Pd(tptbz)] (5) and [(mnt)Pt(tptbz)] (6), the nearly ∼3.5 V potential window that is accessible in CH2Cl2 reveals a single oxidation wave that is reversible in appearances for 6 at +0.94 V vs Fc+/Fc (Figure 3, middle) but only quasireversible for 5 at +1.00 V (Figure S47). Upon reversal of the scanning potential, 5 reveals a minor current maximum at +0.71 V, which, owing to the absence of a comparable feature on the initial anodic pass, is attributable to decomposition product forming on the timescale of the oxidative scanning. These oxidation processes for 5 and 6 are ene-1,2-dithiolate to radical monoanion oxidations (Scheme 1, (a) → (b)), as confirmed by the DFT calculation (Figure S49). The ∼0.06 V milder potential for the oxidation in 6 vs 5 likely is due to a greater degree of the modest metal d character that is mixed into this largely ligand-based HOMO. In the cathodic direction, 5 and 6 reveal irreversible reduction processes that are of a qualitatively similar irreversible appearance but positioned at quite different potentials. The cathodic current maximum for this wave occurs at −2.17 V for 6 vs −1.40 V for 5(Figure S48). The LUMOs for 5 and 6, which are the expected M-dx2–y2 S-p σ* interaction, reflect this difference in reduction potentials by their energies relative to the HOMO. The HOMO–LUMO gap calculated for 6 in the gas phase exceeds that for 5 by more than 0.5 eV. The higher energy for this σ* MO in 6 is fully in accordance with the greater ligand field strength for the third row metal vs its second row counterpart. For both 5 and 6, irreversibility upon reduction likely is due to dissociation of the tptbz ligand.

Scanning anodically, the first redox process observed for [(pdt)Pt(tptbz)] occurs at +0.14 V versus Fc+/Fc and is attributed to the [Ph2C2S2]2– – e→ [Ph2C2SS]1– ligand oxidation (Figure 3, bottom). In heteroleptic [(pdt)ML2] complexes (M = Ni2+, Pd2+, Pt2+; L = phosphine, isonitrile), this oxidation occurs at ∼0 V vs Fc+/Fc with only modest variation as the identity of L is changed.33,35,36 The second, quasireversible anodic process for [(pdt)Pt(tptbz)] at +0.98 V has a plausible assignment either as a second oxidation of the phenyl dithiolene ligand ([Ph2C2SS]1– – e→ [Ph2(C=S)2]0), which is precedented in related compounds,1,2 or as a first oxidation of the tptbz ligand.

Spectroscopically, the generation of [8]+ by a controlled-potential one-electron oxidation is attended by the onset of broad, low-energy absorptions at ∼692 and ∼865 nm that likely conceal the presence of multiple unresolved excitations (Figure 4). Subsequent one-electron oxidation to [8]2+ at a higher potential induces the disappearance of these features. A TD-DFT simulation of this absorption spectrum points to the contribution of multiple transitions to these low-energy features, most of them involving the [Ph2C2SS]1–-based SOMO as the acceptor (Figure S50). The lowest-energy absorption is described most simply as an intraligand excitation from the π-systems of the Ph substituents, with a minor admixture of metal 5d character, to the C2S2 fragment of the ligand.

Figure 4.

Figure 4

Spectroelectrochemically generated UV–vis absorption spectrum of [8]+ in CH2Cl2.

To assist the interpretation of the second oxidation process observed for 8, calculations were undertaken upon [(pdt)Pt(tptbz)]2+ as a closed shell singlet, where the second oxidation is dithiolene based and produces the α-dithione form of the ligand, as a triplet with one spin on each ligand, and as a singlet diradical (broken symmetry) with an unpaired electron on each organic ligand but with opposite spin. Of these several scenarios, the closed shell singlet arising from successive oxidation of [Ph2C2S2]2– to the dithione is assessed as modestly lower in energy by ∼0.5 kcal/mol than the open shell singlet diradical, and the desired triplet state is the highest energy configuration by ∼7 kcal/mol. Regardless of the net spin state, a dicationic form of [(pdt)Pt(tptbz)] that could be formulated as [(Ph2C2SS)Pt2+(tptbz•,+)]2+ is fundamentally interesting from the perspective of materials engineering and the potential for insight into how physical properties might be tailored with better control.

One suggestion arising from the electrochemistry depicted in Figure 3 is that oxidations from each of the two different organosulfur ligands might be observable if either the aromatic dithioether ligand is made more electron rich or the phenyl dithiolene ligand is rendered more electron deficient such that its second oxidation to the dithione ((b) → (c), Scheme 1) is disfavored relative to radical cation formation in the tetrathioarene ring. The known 1,2,3,4,5,6-hexakis(isopropylthio)benzene, C6(SiPr)6, is a plausible candidate as more electron rich relative to tptbz and is prepared by a similar route. However, the oxidation potential found for C6(SiPr)6 is nearly identical to that found for tptbz (Figure S46). That C6(SiPr)6 is not more easily oxidized in proportion to its greater number of pendant sulfur atoms is possibly due to steric crowding of the SiPr substituents around the ring periphery such that optimal S p-π arene-π overlap is impeded. The greater value of δ for C6(SiPr)6 vs tptbz (Table 2) aligns with this supposition.

Complexes featuring polychlorophenyl-substituted dithiolene ligands have not been reported, but they are likely accessible via the same benzoin/P4S10 route that has been implemented to prepare [Ni(S2C2(C6H4-p-Cl)2)2]38,39 and other complexes with arene-substituted dithiolene ligands. We have computationally investigated the 3,5-dichlorophenyl analogue of 8 for its effect in drawing downward in energy the dithiolene-based C2S2–HOMO and making the tptbz-based ligand competitive as the site of a second oxidation. Figure 5 illustrates the frontier MOs in the structures of gas-phase optimizations of both [(pdt)Pt(tptbz)]+ and [(Cl2-3,5-pdt)Pt(tptbz)]+. In [(pdt)Pt(tptbz)]+, the MO housing the dithiolene radical is ∼1.72 eV higher in energy than the tptbz-based HOMO–1 (Figure 5, left), but in [(Cl2-3,5-pdt)Pt(tptbz)]+, while the qualitative ordering is the same, the energy difference is substantially narrowed to ∼0.13 eV (Figure 5, right). This outcome suggests that, in [(Cl3-2,4,6-pdt)Pt(tptbz)]+, where the number and placement of chlorine substituents are such as to further lower the dithiolene-based MOs, successive one-electron oxidations would occur on the opposing dithiolene and dithioether ligands.

Figure 5.

Figure 5

Relative energies of the highest occupied MOs for [(pdt)Pt(tptbz)]+ (left side) and [(3,5-Cl2-pdt)Pt(tptbz)]+ (right side). Orbital images are drawn at the 0.04 contour level.

Summary and Conclusions

The majority of coordination complexes featuring arene-1,2-dithioether ligation have this fragment incorporated within a larger, multidentate ligand architecture as typified by 1,2-bis(2-mercaptophenylthio)phenylene(2−),40 1,2-bis(2-pyridylthio)phenylene,40 and various cyclic multithioether ligands.18,19 Simple bidentate arene dithioether ligands, such as 1,2-bis(2-methoxyethylthio)benzene, have been investigated for their potential in the selective extraction and recovery of Pd(II) from waste streams.41 However, perhaps because they are seen as ligands that favor the formation of inorganic polymers, as seen with [[M(C6(SMe)6)][PF6]]n (M = Cu+, Ag+),21 1,2,4,5-tetrathioether benzenes have received scant attention, if indeed any at all, for their possibilities in the systematic, “bottom-up” assembly of multimetal arrays or for their potential redox activity while chelated to a metal ion.

In this report, we have detailed the syntheses, structures, and properties of the monometallic, open-ended complexes [Cl2M(tptbz)] (M = Pd, Pt), dimetallic [Cl2Pd(tptbz)PdCl2], and the heteroleptic, open-ended dithiolene-dithioether complexes [(mnt)M(tptbz)] (M = Pd, Pt) and [(pdt)Pt(tptbz)]. Surprisingly, these complexes appear to be the first with tptbz itself and join only [(tmmb)Cu]2(μ-I)223 as structurally authenticated coordination compounds with such 1,2,4,5-substituted arene tetrathioethers. While oxidation of coordinated tptbz to the radical cation is not observed in [(pdt)Pt(tptbz)], the observable redox chemistry being attributable to [Ph2C2S2]2– alone, computational work suggests that phenyldithiolene variants whose oxidation potentials are rendered more positive may make tptbz oxidation experimentally accessible. In continuing work, we both explore the use of phenyldithiolene derivatives bearing electron-withdrawing groups for this end and target the synthesis of symmetric dimetallic [(R2C2S2)M(tptbz)M(S2C2R2)] complexes for comparison to their [(R2C2S2)M(tpbz)M(S2C2R2)] tetraphosphine analogues.

Acknowledgments

The Louisiana Board of Regents (LEQSF-(2002-03)-ENH-TR-67) and the National Science Foundation (MRI-1228232 and 0619770) are thanked for funding of Tulane University’s X-ray crystallography and mass spectrometry instrumentation, and Tulane University is acknowledged for its ongoing assistance with operational costs for the X-ray diffraction facility. The authors gratefully acknowledge support for this project from the National Science Foundation (CHE-1836589 for C.W. and J.P.D.). Professor R. H. Schmehl and Qingxin Chen are thanked for assistance with the spectroelectrochemistry measurements.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.3c02928.

  • Procedures for crystal growth, X-ray diffraction data collection, and structure solution and refinement; description of computational procedures; tables summarizing unit cell and refinement data (Tables S1–S4); thermal ellipsoid plots with complete atom labeling (Figures S1–S22); crystal packing images (Figures S23–S25); analytical, spectroscopic, and electrochemical data for compounds reported (Figures S26–S48); MO energy level diagrams for 5 and 6 (Figure S49); calculated UV–vis spectrum for [8]+ (Figure S50); coordinates for optimized geometries (Tables S5–S18) (PDF)

The authors declare no competing financial interest.

Supplementary Material

ic3c02928_si_001.pdf (7.7MB, pdf)

References

  1. Arumugam K.; Shaw M. C.; Mague J. T.; Bill E.; Sproules S.; Donahue J. P. Long-Range Spin Coupling: A Tetraphosphine-Bridged Palladium Dimer. Inorg. Chem. 2011, 50, 2995–3002. 10.1021/ic102493g. [DOI] [PubMed] [Google Scholar]
  2. Arumugam K.; Selvachandran M.; Obanda A.; Shaw M. C.; Chandrasekaran P.; Caston Good S. L.; Mague J. T.; Sproules S.; Donahue J. P. Redox-Active Metallodithiolene Groups Separated by Insulating Tetraphosphinobenzene Spacers. Inorg. Chem. 2018, 57, 4023–4038. 10.1021/acs.inorgchem.8b00201. [DOI] [PubMed] [Google Scholar]
  3. Kumar S.; Selvachandran M.; Arumugam K.; Shaw M. C.; Wu C.; Maurer M.; Zhang X.; Sproules S.; Mague J. T.; Donahue J. P. Open-Ended Metallodithiolene Complexes with the 1,2,4,5-Tetrakis(diphenylphosphino)benzene Ligand: Modular Building Elements for the Synthesis of Multimetal Complexes. Inorg. Chem. 2021, 60, 13177–13192. 10.1021/acs.inorgchem.1c01573. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Kumar S.; Selvachandran M.; Wu C.; Pascal R. A. Jr.; Zhang X.; Grusenmeyer T.; Schmehl R. H.; Sproules S.; Mague J. T.; Donahue J. P. Heterotrimetallic Assemblies with 1,2,4,5-Tetrakis(diphenylphosphino)benzene Bridges: Constructs for Controlling the Separation and Spatial Orientation of Redox-Active Metallodithiolene Groups. Inorg. Chem. 2022, 61, 17804–17818. 10.1021/acs.inorgchem.2c03112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Sproules S. Molecules as Electron Spin Qubits. Electron Paramag. Reson. 2016, 25, 61–97. 10.1039/9781782629436-00061. [DOI] [Google Scholar]
  6. Alberti A.; Pedulli G. F.; Tiecco M.; Testaferri L.; Tingoli M. An Electron Spin Resonance Investigation of the Radical Cations from Poly(alkylthio)benzenes. J. Chem. Soc., Perkin Trans. 1984, 2, 975–979. 10.1039/p29840000975. [DOI] [Google Scholar]
  7. Chen X.; Gao F.; Yang W. Single-Crystal X-ray Structures of Conductive π-Stacking Dimers of Tetrakis(alkylthio)benzene Radical Cations. Sci. Rep. 2016, 6, 29314. 10.1038/srep29314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Larsen J.; Bechgaard K. 1,3,5,7-Tetrathia-s-indacene-2,6-dione Chemistry. Synthesis of New Multisulfur Donor Molecules and Nickel-Dithiolene Electron-Transfer Complexes. J. Org. Chem. 1987, 52, 3285–3288. 10.1021/jo00391a019. [DOI] [Google Scholar]
  9. Davison A.; Holm R. H.; Benson R. E.; Mahler W.; Muetterties E. L. Metal Complexes Derived from cis-1,2-Dicyano-1,2-ethylenedithiolate and Bis(trifluoromethyl)-1,2-dithiete. Inorg. Synth. 1967, 10, 8–25. 10.1002/9780470132418.ch3. [DOI] [Google Scholar]
  10. Abel E. W.; Jenkins C. R. 1,3-Dithia-2-stannacyclopentenes. J. Chem. Soc. (A). 1967, 1344–1346. 10.1039/j19670001344. [DOI] [Google Scholar]
  11. Bhattacharya A. K.; Hortmann A. G. A General Synthesis of 1,3-Dithiol-2-ones. J. Org. Chem. 1974, 39, 95–97. 10.1021/jo00915a021. [DOI] [Google Scholar]
  12. Armarego W. L. F.; Perrin D. D.. Purification of Laboratory Chemicals, 4th ed.; Oxford, U.K, Butterworth-Heinemannn., 2000. [Google Scholar]
  13. Testaferri L.; Tingoli M.; Tiecco M. Reactions of Polychlorobenzenes with Alkanethiol Anions in HMPA. A Simple, High-Yield Synthesis of Poly(alkylthio)benzenes. J. Org. Chem. 1980, 45, 4376–4380. 10.1021/jo01310a022. [DOI] [Google Scholar]
  14. Bock H.; Rauschenbach A.; Näther C.; Kleine M.; Havlas Z. Crystallization and Structure Determination of Donor/Acceptor Complexes between 1,2,4,5-Tetrakis(alkylthio)benzene Derivatives and Bromine or Iodine. Liebigs. Ann. 1996, 1996, 2185–2194. 10.1002/jlac.199619961232. [DOI] [Google Scholar]
  15. Patra A.; Mondal B.; Sen B.; Zangrando E.; Chattopadhyay P. Nickel(II) Complexes of N2S2 Donor Set Ligand and Halide/Pseudohalides: Synthesis, Crystal Structure, DNA and Bovine/Human Serum Albumin Interaction. J. Chem. Sci. 2015, 127 (11), 1883–1893. 10.1007/s12039-015-0962-3. [DOI] [Google Scholar]
  16. Song L.-C.; Sun X.-J.; Zhao P.-H.; Li J.-P.; Song H.-B. Synthesis, Characterization and Some Properties of Mononuclear Ni and Trinuclear NiFe2 Complexes Related to the Active Sirte of [NiFe]-Hydrogenases. Dalton Trans. 2012, 41, 8941–8950. 10.1039/c2dt30609c. [DOI] [PubMed] [Google Scholar]
  17. Berghof C.; Grosu I. G.; Lönnecke P.; Gómez-Ruiz S.; Silaghi-Dumitrescu L.; Hey-Hawkins E. Coordination Chemistry of the Heterotopic 1,2-Phenylenebis(thio)diacetic Acid Ligand: Rhodium(I), Palladium(II), and Nickel(II) Complexes. Inorg. Chim. Acta 2011, 374, 127–133. 10.1016/j.ica.2010.12.004. [DOI] [Google Scholar]
  18. de Vries N.; Reedijk J. Nickel Template Syntheses of Pentadentate and Macrocyclic Nitrogen-Sulfur Ligands: Single-Crystal X-Ray Structure of (1,10-Diaza-4,7,13,16-tetrathia-5,6:14,15-dibenzocyclooctadecane)nickel(II) Bis(tetraphenyl)borate–Acetonitrile(1/3). Inorg. Chem. 1991, 30, 3700–3703. 10.1021/ic00019a026. [DOI] [Google Scholar]
  19. Sellmann D.; Neuner H. P.; Knoch F. Transition Metal Complexes with Sulfur Ligands Part LXXIX*. Crown Thioether Complexes of Iron, Cobalt, Nickel and Copper with bzo2-18S6. X-ray Structure Analysis of [Ni(bzo2-18S6)][B(C6H5)4]2 (bzo2-18S6 = 2,3,11,12-Dibenzo-1,4,7,10,13,16-hexathiacyclooctadeca-2,11-diene). Inorg. Chim. Acta 1991, 190, 61–69. 10.1016/S0020-1693(00)80232-7. [DOI] [Google Scholar]
  20. Horn G. W.; Kumar R.; Maverick A. W.; Fronczek F. R.; Watkins S. F. Structure of cis-Dichlorobis(dimethyl sulfide)platinum(II). Acta Crystallogr., Sect. C 1990, 46, 135–136. 10.1107/S0108270189009893. [DOI] [Google Scholar]
  21. Suenaga Y.; Kuroda-Sowa T.; Munakata M.; Maekawa M.; Morimoto H. Syntheses and Structures of Copper(I) and Silver(I) Coordination Polymers with Hexakis(methylthio)benzene. Polyhedron 1998, 18, 429–436. 10.1016/S0277-5387(98)00314-3. [DOI] [Google Scholar]
  22. Suenaga Y.; Maekawa M.; Kuroda-Sowa T.; Munakata M.; Morimoto H.; Hiyama N.; Kitagawa S. Comparative X-Ray Studies of a Copper(I) Coordination Polymer with 1,2,4,5-Tetramethylmercaptobenzene (tmmb), [(CuX)2(tmmb)]n (X = Br, I). Anal. Sci. 1997, 1, 1047–1049. 10.2116/analsci.13.1047. [DOI] [Google Scholar]
  23. Suenaga Y.; Maekawa M.; Kuroda-Sowa T.; Munakata M.; Morimoto H.; Hiyama N.; Kitagawa S. Crystal Structure of Dinuclear Copper(I) Complex with 1,2,4,5-Tetramethylmercaptobenzene, [Cu(tmmb)I]2. Anal. Sci. 1997, 13 (4), 651–652. 10.2116/analsci.13.651. [DOI] [Google Scholar]
  24. Usón R.; Vicente J.; Oro J. Maleonitriledithiolate Complexes of Au(III), Au(I), Pd(II) and Pt(II) Containing Neutral or Anionic Ligands. Inorg. Chim. Acta 1981, 52, 29–34. 10.1016/S0020-1693(00)88567-9. [DOI] [Google Scholar]
  25. Cerrada E.; Fernández E. J.; Gimeno M. C.; Laguna A.; Laguna M.; Terroba R.; Villacampa M. D. Synthesis of Dithiolate Gold(III) Complexes by Dithiolate Transfer Reactions. X-ray Structure of [Au(C6F5)(S2C6H4)(PPh3)]. J. Organometall. Chem. 1995, 492, 105–110. 10.1016/0022-328X(95)05291-V. [DOI] [Google Scholar]
  26. Cerrada E.; Fernández E. J.; Jones P. G.; Laguna A.; Laguna M.; Terroba R. Synthesis and Reactivity of Trinuclear Gold(III) Dithiolate Complexes. X-ray Structure of [Au(C6F5)(S2C6H4)]3 and [Au(C6F5)(S2C6H4)(SC6H4SPPh3)]. Organometallics 1995, 14, 5537–5543. 10.1021/om00012a021. [DOI] [Google Scholar]
  27. Nomura M.; Fourmigué M. Dinuclear Cp* Cobalt Complexes of the 1,2,4,5-Benzenetetrathiolate Bischelating Ligand. Inorg. Chem. 2008, 47, 1301–1312. 10.1021/ic700778b. [DOI] [PubMed] [Google Scholar]
  28. Arumugam K.; Shaw M. C.; Chandrasekaran P.; Villagrán D.; Gray T. G.; Mague J. T.; Donahue J. P. Synthesis, Structures, and Properties of 1,2,4,5-Benzenetetrathiolate Linked Group 10 Metal Complexes. Inorg. Chem. 2009, 48, 10591–10607. 10.1021/ic901257s. [DOI] [PubMed] [Google Scholar]
  29. Chandrasekaran P.; Arumugam K.; Jayarathne U.; Pérez L. M.; Mague J. T.; Donahue J. P. Synthesis, Structures, and Properties of Mixed Dithiolene-Carbonyl and Dithiolene-Phosphine Complexes of Tungsten. Inorg. Chem. 2009, 48, 2103–2113. 10.1021/ic802016b. [DOI] [PubMed] [Google Scholar]
  30. Pilia L.; Espa D.; Barsella A.; Fort A.; Makedonas C.; Marchiò L.; Mercuri M. L.; Serpe A.; Mitsopoulou C. A.; Deplano P. Combined Theoretical and Experimental Study on Redox-Active d8 Metal Dithione–Dithiolato Complexes Showing Molecular Second-Order Nonlinear Optical Activity. Inorg. Chem. 2011, 50, 10015–10027. 10.1021/ic2007765. [DOI] [PubMed] [Google Scholar]
  31. Espa D.; Pilia L.; Marchiò L.; Mercuri M. L.; Serpe A.; Barsella A.; Fort A.; Dalgleish S. J.; Robertson N.; Deplano P. Redox-Switchable Chromophores Based on Metal (Ni, Pd, Pt) Mixed-Ligand Dithiolene Complexes Showing Molecular Second-Order Nonlinear-Optical Activity. Inorg. Chem. 2011, 50, 2058–2060. 10.1021/ic102343g. [DOI] [PubMed] [Google Scholar]
  32. Espa D.; Pilia L.; Marchiò L.; Attar S. S.; Barsella A.; Fort A.; Mercuri M. L.; Serpe A.; Deplano P. Structural Changes in MII Dithione/Dithiolato Complexes (M = Ni, Pd, Pt) on Varying the Dithione Functionalization. CrystEngComm 2015, 17, 4161–4171. 10.1039/C5CE00469A. [DOI] [Google Scholar]
  33. Obanda A.; Martinez K.; Schmehl R. H.; Mague J. T.; Rubtsov I. V.; MacMillan S. N.; Lancaster K. M.; Sproules S.; Donahue J. P. Expanding the Scope of Ligand Substitution from [(Ph2C2S2)2M] (M = Ni2+, Pd2+, Pt2+) to Afford New Heteroleptic Dithiolene Complexes. Inorg. Chem. 2017, 56, 10257–10267. 10.1021/acs.inorgchem.7b00971. [DOI] [PubMed] [Google Scholar]
  34. Kerr M. J.; Harrison D. J.; Lough A. J.; Fekl U. Ligand-Based Reactivity of a Platinum Bisdithiolene: Double Diene Addition Yields a New C2-Chiral Chelate Ligand. Inorg. Chem. 2009, 48, 9043–9045. 10.1021/ic901116c. [DOI] [PubMed] [Google Scholar]
  35. Obanda A.; Valerius K.; Mague J. T.; Sproules S.; Donahue J. P. Group 10 Metal Dithiolene Bis(isonitrile) Complexes: Synthesis, Structures. Properties and Reactivity. Organometallics 2020, 39, 2854–2870. 10.1021/acs.organomet.0c00375. [DOI] [Google Scholar]
  36. Chandrasekaran P.; Greene A. F.; Lillich K.; Capone S.; Mague J. T.; DeBeer S.; Donahue J. P. A Structural and Spectroscopic Investigation of Octahedral Platinum Bis(dithiolene)phosphine Complexes: Platinum Dithiolene Internal Redox Chemistry Induced by Phosphine Association. Inorg. Chem. 2014, 53, 9192–9205. 10.1021/ic501273b. [DOI] [PubMed] [Google Scholar]
  37. Batsanov S. S. Waals radii of Elements. Inorg. Mater. 2001, 37 (9), 871–885. 10.1023/A:1011625728803. [DOI] [Google Scholar]
  38. Schrauzer G. N.; Mayweg V. P. Preparation, Reactions, and Structure of Bisdithio-α-diketone Complexes of Nickel, Palladium, and Platinum. J. Am. Chem. Soc. 1965, 87, 1483–1489. 10.1021/ja01085a014. [DOI] [Google Scholar]
  39. Koehne S.; Mirmelli B.; Mague J. T.; Donahue J. P. Bis[1,2-bis(4-chlorophenyl)ethylene-1,2-dithiolato(1−)]nickel(II). IUCrData 2022, 7, 220148. 10.1107/S2414314622001481. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Sellmann D.; Engl K.; Gottschalk-Gaudig T.; Heinemann F. W. Transition Metal Complexes with Sulfur Ligands, 134[≠] Synthesis of the Tetradentate Thioether Amine and Thioether Thiolate Ligands tpS2(NH2)2 and tpS4-H2 and Some Representative Six-Coordinate Ruthenium and Osmium Complexes [tpS2(NH2)2 = 1,2-Bis(2-aminophenylthio)phenylene; tpS42– = 1,2-Bis(2-mercaptophenylthio)phenylene(2−)]. Eur. J. Inorg. Chem. 1999, 333–339. . [DOI] [Google Scholar]
  41. Traeger J.; Klamroth T.; Kelling A.; Lubahn S.; Cleve E.; Mickler W.; Heydenreich M.; Müller H.; Holdt H.-J. Complexation of Palladium(II) with Unsaturated Dithioethers – A Systematic Development of Highly Selective Ligands for Solvent Extraction. Eur. J. Inorg. Chem. 2012, 2012, 2341–2352. 10.1002/ejic.201101406. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ic3c02928_si_001.pdf (7.7MB, pdf)

Articles from Inorganic Chemistry are provided here courtesy of American Chemical Society

RESOURCES