Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2021 Feb 5;60(4):2783–2796. doi: 10.1021/acs.inorgchem.0c03680

Electronic Communication in Binuclear Osmium- and Iridium-Polyhydrides

Lara Cancela , Miguel A Esteruelas †,*, Javier Galbán , Montserrat Oliván , Enrique Oñate , Andrea Vélez , Juan C Vidal §
PMCID: PMC9179948  PMID: 33543934

Abstract

graphic file with name ic0c03680_0013.jpg

Reactions of polyhydrides OsH6(PiPr3)2 (1) and IrH5(PiPr3)2 (2) with rollover cyclometalated hydride complexes have been investigated in order to explore the influence of a metal center on the MHn unit of the other in mixed valence binuclear polyhydrides. Hexahydride 1 activates an ortho-CH bond of the heterocyclic moiety of the trihydride metal–ligand compounds OsH32-C,N-[C5RH2N-py]}(PiPr3)2 (R = H (3), Me (4), Ph (5)). Reactions of 3 and 4 lead to the hexahydrides (PiPr3)2H3Os{μ-[κ2-C,N-[C5RH2N-C5H3N]-N,C2]}OsH3(PiPr3)2 (R = H (6), Me (7)), whereas 5 gives the pentahydride (PiPr3)2H3Os{μ-[κ2-C,N-[C5H3N-C5(C6H4)H2N]-C,N,C3]}OsH2(PiPr3)2 (8). Pentahydride 2 promotes C—H bond activation of 3 and the iridium-dihydride IrH22-C,N-[C5H3N-py]}(PiPr3)2 (9) to afford the heterobinuclear pentahydride (PiPr3)2H3Os{μ-[κ2-C,N-[C5H3N-C5H3N]-N,C2]}IrH2(PiPr3)2 (10) and the homobinuclear tetrahydride (PiPr3)2H2Ir{μ-[κ2-C,N-[C5H3N-C5H3N]-N,C2]}IrH2(PiPr3)2 (11), respectively. Complexes 68 and 11 display HOMO delocalization throughout the metal–heterocycle-metal skeleton. Their sequential oxidation generates mono- and diradicals, which exhibit intervalence charge transfer transitions. This notable ability allows the tuning of the strength of the hydrogen–hydrogen and metal–hydrogen interactions within the MHn units.

Short abstract

The influence of a metal center on the MHn unit of the other in binuclear osmium- and iridium-polyhydride complexes bearing rollover cyclometalated bipyridines as a bridging ligand has been studied.

Introduction

The hydrogen atoms of the MHn units of LmMHn transition metal polyhydride complexes interact with one another and with the metal center, forming Kubas type dihydrogens (dH–H = 0.8–1.0 Å), elongated dihydrogens (dH–H = 1.0–1.3 Å), compressed hydrides (dH–H = 1.3–1.6 Å), or classical hydrides (dH–H ≥ 1.6 Å). These interactions are governed by the electron density of the metal center, which is regulated by the coligands Lm.1 Thus, the ability of such compounds to reversibly release the H2 molecule requires Lm ligands, which are able to modify the electron density of the metal center, in order to allow reversible changes in the inner interactions of the MHn units. The design of such ligands is certainly a challenge of the first magnitude.

An attractive approach to the solution of this challenge is the use transition metal complexes, which should display frontier orbitals involving substantial mixing with a π-ligand backbone, whereas such a ligand should also bear atoms with free electrons. The coordination of this metal–ligand to an MHn unit would generate species with frontier orbitals delocalized between the two metal centers connected by the π-linker. Thus, the metals can be viewed as being electronically coupled and therefore the changes in electron density at one site should perturb the electron density at the other.2 The search for efficient π-linkers (bridging ligands), which promote the cooperative effect between the redox active centers through electronic coupling pathways, is central for success. Unsaturated carbon chains,3 aromatic polycycles,3c,4 aromatic N-heterocycles,5 bisdioxolenes,6 bisdithiolenes,7 dithiolate,8 cyanide,9 and cyanamides10 have been mainly employed so far, as bridging ligands, to provide electronical coupling between metals.

The interactions between the metal centers have been grouped into three categories, according to the Robin–Day classification: weak, moderate, and strong.11 Compounds displaying weak interaction form class I, and their redox centers mostly behave as separated sites. On the other hand, strong interaction affords a complete electron density delocalization, and complexes with this ability are grouped as class III. Species exhibiting moderate interaction between their redox centers constitute class II.3d,12 The degree of interaction is efficiently assessed by means of the analysis of the intervalence charge transfer (IVCT) band in the UV–vis–NIR spectra on the basis of the Marcus–Hush theory.13 At the electrochemistry level, the redox potential separation between successive redox processes is also a frequently used measure, although it often presents misinterpretation issues.14

We have recently shown that the platinum group polyhydrides OsH6(PiPr3)2 (1) and IrH5(PiPr3)2 (2) promote the activation of C—H bonds of the rings of 2,2′-bipyridines and related heterocycles to afford rollover cyclometalated trihydride- and dihydride-derivatives (Scheme 1),15 in agreement with the ability of polyhydrides of the platinum group metals to activate σ-bonds1f and in particular the d2-hexahydride OsH6(PiPr3)2.16 In the context of the rollover cyclometalation, we noted that in a few cases the resulting ligands underwent an additional cylometalation promoted by a second metal complex, to form binuclear species bearing a bridging rollover bis-cyclometalated heterocycle.17 Although evidence of the ability of these bridges to provide electronic coupling pathways has not been reported, these findings inspired us to prepare osmium- and iridium-polyhydrides with this class of bridging ligands and to use them as models to check our proposed approach toward the control of reversible changes in the existing interactions within the MHn units.

Scheme 1. C—H Bond Activation of 2,2′-Bipyridines.

Scheme 1

This paper proves that rollover cyclometalated 2,2′-bipyridine heterocycles provide electronic coupling pathways between the metals of (PiPr3)2HnOs(μ-L)OsHn(PiPr3)2 (n = 2 or 3) and (PiPr3)2H2Ir(μ-L)IrH2(PiPr3)2 complexes and that changes in the electron density of a metal center influence the inner interactions of the MHn unit of the other.

Results and Discussion

Metal–Ligand C–H Bond Activation

Osmium-hexahydride complex 1 is able to activate C—H bonds of the rollover cyclometalated trihydride derivatives OsH32-C,N-[C5RH2N-py]}(PiPr3)2 (R = H (3), Me (4), Ph (5)) in toluene under reflux (Scheme 2) and in agreement with its ability to promote σ-bond activation reactions. Complexes 3 and 4 afford the binuclear-hexahydride compounds (PiPr3)2H3Os{μ-[κ2-C,N-[C5RH2N–C5H3N]-N,C2]}OsH3(PiPr3)2 (R = H (6) Me (7)), as a result of the coordination of the free nitrogen atom of the rollover cyclometalated heterocycle and the ortho-CH bond activation of the other ring, whereas the reaction with the phenyl-derivative 5 leads to the pentahydride (PiPr3)2H3Os{μ-[κ2-C,N-[C5H3N-C5(C6H4)H2N]-C,N,C3]}OsH2(PiPr3)2 (8). In contrast to 6 and 7, complex 8 bears two different osmium(IV) OsHn(PiPr3)2 moieties, OsH3(PiPr3)2 and OsH2(PiPr3)2. In this case, the bridging ligand acts in a dual manner: monoanionic C,N-chelate with OsH3(PiPr3)2 and dianionic C,N,C-pincer with OsH2(PiPr3)2. The difference is a consequence of the hexahydride being also able to activate the phenyl substituent of the rollover cyclometalated heterocycle of 5. The three binuclear products can be also prepared by treatment of 1 with 0.5 equiv of the 2,2′-bipyridine. Both methods, via intermediates 35 and the one-pot synthesis procedures, afford the quantitative formation of the binuclear species, which were isolated as orange solids in about 80% yield. Complexes 6 and 8 were characterized by X-ray diffraction analysis.

Scheme 2. Formation of Binuclear Species 6, 7, and 8.

Scheme 2

Figure 1 shows the structure of 6, which can be described as two equivalent OsH3(PiPr3)2 units linked by a rollover bis-cyclometalated 2,2′-bipyridine. The coordination polyhedron around each osmium atom is the typical pentagonal bipyramid for osmium(IV) OsH3(Y-X)(PiPr3)2 species18 with axial phosphines (P(1)–Os-P(2) = P(2A)—Os(A)—P(1A) = 160.82(2)°), whereas the hydride ligands lie at the joint base of the bipyramid coplanar to the heterocycle. The Os—N and Os—C bond lengths of 2.1665(18) and 2.144(2) Å are similar to those of the precursor 3.15 In agreement with the high symmetry of the molecule, the 31P{1H} NMR spectrum of this compound in toluene-d8 displays a singlet at 23.1 ppm for the four equivalent phosphines. In the 1H NMR spectrum, the most noticeable feature is the hydride resonances, which appear between −5 and −13 ppm displaying the typical behavior observed for the inequivalent hydrides of OsH3(Y-X)(PiPr3)2 complexes, involved in a thermally activated site exchange process.18 The 31P{1H}, 1H, and 13C{1H} NMR spectra of 7 in toluene-d8 reflect the asymmetry imposed by the methyl substituent of the heterocycle. Thus, in contrast to 6 the 31P{1H} NMR spectrum shows two singlets at 22.7 and 21.4 ppm, whereas resonances corresponding to inequivalent OsH3(PiPr3)2 units are observed between −5 and −14 ppm in the 1H NMR spectrum. The 13C{1H} NMR spectrum displays two triplets (2JC–P ≈ 6 Hz) at 173.9 and 168.9 ppm for the inequivalent metalated carbon atoms.

Figure 1.

Figure 1

Molecular diagram of complex 6 (ellipsoids shown at 50% probability). All hydrogen atoms (except the hydrides) are omitted for clarity. Selected bond distances (Å) and angles (deg): Os—P(1) = Os(A)—P(1A) = 2.3422(6), Os—P(2) = Os(A)—P(2A) = 2.3414(6), Os—C(2A) = Os(A)—C(2) = 2.144(2), Os—N(1) = Os(A)—N(1A) = 2.1665(18); P(1)—Os—P(2) = P(2A)—Os(A)—P(1A) = 160.82(2), N(1)—Os-C(2A) = N(1A)—Os(A)–C(2) = 76.14(7).

The structure of 8 (Figure 2) proves the dual coordination of the heterocycle in this complex, C,N-chelate to a metal center (Os(1)) and C,C,N-pincer to the other (Os(2)). The coordination polyhedron around both metal centers can be also idealized as a pentagonal bipyramid. However, there are significant differences between the bipyramids, which are associated with the acting fashion of the bridging ligand. The polyhedron around Os(1) resembles that of 6 with a P(1)—Os(1)—P(2) angle of 158.86(3)°. Phosphine ligands attached to Os(2) also occupy the axial positions of the bipyramid, forming a P(3)—Os(2)—P(4) angle of 164.05(3)°, whereas the pincer lies at the perpendicular joint base, coplanar to the hydride ligands, acting with a C(1)—Os(2)—C(12) angle of 150.41(12)°, which slightly deviates from the ideal value of 144°. The Os(1)—C(7) and Os—N(1) distances of 2.143(3) and 2.185(2) Å are similar to those found in 6, whereas the Os(2)—C(1), Os(2)—C(12), and Os(2)—N(2) bond lengths compare well with the observed ones for osmium compounds bearing C,N,C-pincer ligands.16c,15,1931P{1H}, 1H, and 13C{1H} NMR spectra of 8 in dichloromethane-d2 are consistent with the structure shown in Figure 2. Thus, the 31P{1H} NMR spectrum contains two singlets at 24.1 and 1.8 ppm, assigned to the OsH3(PiPr3)2 and OsH2(PiPr3)2 units, respectively. In the 1H NMR spectrum, the resonances of the OsH3(PiPr3)2 unit display the typical pattern for the cyclometalated OsH3(Y-X)(PiPr3)2 species, between −6 and −14 ppm, along with two temperature invariant doublets (2JH–H = 11.3 Hz) of triplets (2JH–P = 15.1 and 17.2 Hz) at −8.48 and −9.19 ppm corresponding to the hydride ligands of the OsH2(PiPr3)2 unit. The 13C{1H} NMR spectrum shows three triplets (2JC–P = 6.1–8.5 Hz) at 169.9, 168.0, and 165.5 ppm due to the metalated carbon atoms.

Figure 2.

Figure 2

Molecular diagram of complex 8 (ellipsoids shown at 50% probability). All hydrogen atoms (except the hydrides) are omitted for clarity. Selected bond distances (Å) and angles (deg): Os(1)—P(1) = 2.3415(8), Os(1)—P(2) = 2.3418(8), Os(2)—P(3) = 2.3752(8), Os(2)—P(4) = 2.3778(8), Os(1)—N(1) = 2.185(2), Os(2)—N(2) = 2.121(2), Os(1)—C(7) = 2.143(3), Os(2)—C(1) = 2.159(3), Os(2)—C(12) = 2.137(3); P(1)—Os(1)—P(2) = 158.86(3), P(3)—Os(2)—P(4) = 164.05(3), N(1)—Os(1)—C(7) = 76.41(10), C(1)—Os(2)—N(2) = 75.77(10), C(12)—Os(2)—N(2) = 74.89(11), C(1)—Os(2)—C(12) = 150.41(12).

The success of the reactions shown in Scheme 2 encouraged us to extend this synthetic methodology, involving polyhydride-mediated sequential rollover cyclometalation of 2,2′-bipyridines to other polyhydrides and to study its utility to generate heterobinuclear derivatives. Thus, we decided to also investigate the C—H bond activation of 3 and the iridium-dihydride IrH22-C,N-[C5H3N-py]}(PiPr3)2 (9), promoted by the iridium-pentahydride complex 2 (Scheme 3). Treatment of toluene solutions of 3 with 1.0 equiv of 2 under reflux for 16 h leads to the heterobinuclear pentahydride (PiPr3)2H3Os{μ-[κ2-C,N-[C5H3N-C5H3N]-N,C2]}IrH2(PiPr3)2 (10), which was isolated as an orange solid in 68% yield. Under the same conditions, the reaction of 9 and 2 affords the homobinuclear tetrahydride (PiPr3)2H2Ir{μ-[κ2-C,N-[C5H3N-C5H3N]-N,C2]}IrH2(PiPr3)2 (11), which can be also prepared by treatment of 2 with 0.5 equiv of 2,2′-bipyridine. Complex 11 was isolated as a yellow solid in almost quantitative yield.

Scheme 3. Formation of Complexes 10 and 11.

Scheme 3

The 31P{1H}, 1H, and 13C{1H} NMR spectra of 10 in toluene-d8 strongly support the structure proposed for this compound, in Scheme 3. The 31P{1H} NMR spectrum contains two singlets at 30.3 and 22.4 ppm, one for each group of equivalent phosphines, whereas the 1H NMR spectrum is consistent with the presence in the complex of two different classes of hydride ligands. In agreement with 8, hydrides attached to osmium give rise to resonances displaying the typical temperature-dependent pattern for a cyclometalated OsH3(XY)(PiPr3)2 species, between −5 and −13 ppm while hydrides of the IrH2(PiPr3)2 unit generate two temperature invariant doublets (2JH–H = 3.6 Hz) of triplets (2JH–P = 18.9 Hz) at −12.93 and −21.93 ppm. In the 13C{1H} NMR spectrum, the resonances due to the metalated carbon atoms appear at 173.4 and 163.2 ppm, as triplets with C–P coupling constants of about 6 Hz. Complex 11 was characterized by X-ray diffraction analysis. Figure 3 shows a view of the structure. The molecule is formed by two chemically equivalent IrH2(PiPr3)2 moieties connected to each other through a rollover bis-cyclometalated 2,2′-bipyridine linker. It is a d6d6 counterpart of the d4–d4 complex 6 and the d4–d6 derivative 10. The coordination polyhedron around each iridium center is the expected octahedron with trans phosphines (P—Ir—P = 156.94(3)°). In agreement with its structure, the 31P{1H} NMR spectrum of this highly symmetrical molecule shows a singlet at 29.9 ppm for the four equivalent phosphines, the 1H NMR spectrum contains two doublets (2JH–H = 4.1 Hz) of triplets (2JH–P = 21.2 and 19.3 Hz) at −12.93 and −22.00 ppm for the inequivalent hydrides of the equivalent IrH2(PiPr3)2 units, whereas the 13C{1H} NMR spectrum displays a triplet (2JC–P = 6.5 Hz) at 163.6 for the equivalent metalated carbon atoms.

Figure 3.

Figure 3

Molecular diagram of complex 11 (ellipsoids shown at 50% probability). All hydrogen atoms (except the hydrides) are omitted for clarity. Selected bond distances (Å) and angles (deg): Ir—P(1) = Ir—P(1A) = Ir(A)—P(1B) = Ir(A)—P(1C) = 2.2988(5), Ir—N(1) = Ir(A)—N(1A) = 2.1329(17), Ir—C(1) = Ir(A)—C(1B) = 2.1330(17); P(1)—Ir—P(1A) = P(1B)—Ir(A)—P(1C) = 156.94(3), N(1)—Ir—C(1) = N(1A)—Ir(A)—C(1B) = 78.47(9).

Frontier Orbitals and Photophysical Properties

The UV–vis spectra for 1 × 10–4 M solutions of the mononuclear precursors 35 and 9 and binuclear derivatives 68, 10, and 11 in 2-methyltetrahydrofuran (MeTHF) were recorded. Figure 4 shows the spectra of 10 and their mononuclear building blocks 3 and 9, whereas the rest are shown in Figures S19–S27. In addition, time-dependent DFT calculations (B3LYP-GD3//SDD(f)/6-31G**) were performed to their rationalization, considering tetrahydrofuran as solvent. Selected absorptions are collected in Table 1, whereas frontier orbitals are shown in Figures 5 and S28–S36.

Figure 4.

Figure 4

UV–vis spectra of complexes 3, 9, and 10 recorded in 2-methyltetrahydrofuran (1 × 10–4 M) at 298 K.

Table 1. Selected Experimental UV-vis Absorptions for 311 (in MeTHF) and Computed TD-DFT (in THF) Vertical Excitation Energies and Their Major Contributions.

λ exp (nm) ε (M–1 cm–1) excitation energy (nm) oscilator strength f transition
Complex 3
241 13310 232 0.1039 HOMO–6 → LUMO (47%)
427 3360 400 0.0582 HOMO–1 → LUMO (91%)
477 1730 444 (S1) 0.0272 HOMO → LUMO (91%)
512 290 498 (T1) 0 HOMO → LUMO (95%)
Complex 4
243 18290 235 0.1747 HOMO–6 → LUMO (56%)
426 4590 399 0.0599 HOMO–1 → LUMO (91%)
484 2350 446 (S1) 0.0327 HOMO → LUMO (91%)
506 1230 507 (T1) 0 HOMO → LUMO (95%)
Complex 5
243 18360 255 0.1933 HOMO–7 → LUMO (21%)
      HOMO–5 → LUMO (54%)
453 2430 405 0.0457 HOMO–1 → LUMO (88%)
494 1360 444 (S1) 0.0344 HOMO → LUMO (88%)
513 740 504 (T1) 0 HOMO → LUMO (95%)
Complex 6
282 28300 282 0.2481 HOMO–2 → LUMO+3 (51%)
408 5420 394 0.0653 HOMO–2 → LUMO (96%)
470 7470 440 (S1) 0.0966 HOMO → LUMO (96%)
506 3380 511 (T1) 0 HOMO → LUMO (96%)
Complex 7
274 33490 270 0.1805 HOMO–2 → LUMO+4 (65%)
416 2620 396 0.0531 HOMO–2 → LUMO (96%)
478 3810 447 (S1) 0.0949 HOMO → LUMO (97%)
524 320 525 (T1) 0 HOMO → LUMO (96%)
Complex 8
282 35260 278 0.0482 HOMO → LUMO+6 (59%)
402 9500 383 0.1737 HOMO–2 → LUMO (50%)
        HOMO → LUMO+1 (45%)
496 2380 481 (S1) 0.0347 HOMO → LUMO (95%)
532 2310 551 (T1) 0 HOMO → LUMO (96%)
Complex 9
276 20330 274 0.2307 HOMO–4 → LUMO (78%)
403 3560 381 (S1) 0.0657 HOMO → LUMO (94%)
463 500 445 (T1) 0 HOMO → LUMO (82%)
Complex 10
224 23460 247 0.0776 HOMO–4 → LUMO+3 (89%)
404 5330 390 0.0496 HOMO–1 → LUMO (95%)
450 6250 426 (S1) 0.0798 HOMO → LUMO (95%)
504 770 492 (T1) 0 HOMO → LUMO (94%)
Complex 11
273 24510 266 0.3100 HOMO–8 → LUMO (40%)
        HOMO → LUMO+3 (38%)
370 1900 338 0.0484 HOMO–4 → LUMO (97%)
431 3410 390 (S1) 0.1364 HOMO → LUMO (98%)
461 900 461 (T1) 0 HOMO → LUMO (88%)

Figure 5.

Figure 5

HOMO of the homobinuclear derivatives 68 and 11.

The spectra of the osmium mononuclear precursors 3-5 show bands in three different regions of the spectrum: <300, 300–500, and >500 nm. The absorptions at the highest energy region correspond mainly to 1π–π* intraheterocycle transitions, whereas the bands between 300 and 500 nm are due to transitions from the metal to the heterocycle mixed with from the heterocycle to the heterocycle. These bands mainly result from HOMO–1-to-LUMO, and HOMO-to-LUMO transitions. Both HOMO–1 and HOMO are essentially located at the metal center and the metalated heterocycle. For HOMO, the percentage of the former is between 52% and 59% and that of the second one lies in the range 28–38%. The LUMO is almost exclusively centered on the metalated heterocycle (95%). The very weak absorption tails after 500 nm are assigned to formally spin forbidden 3MLCT transitions caused by the large spin–orbit coupling introduced by osmium. The spectrum of the mononuclear iridium complex 9 is similar to those of 3-5. The absorptions of <300 nm should be assigned to 1π–π* ligand-to-ligand transitions, whereas those between 300 and 450 nm are due to spin-allowed iridium-to-heterocycle charge transfer (1MLCT) mixed with heterocycle-to-heterocycle transitions. The absorption tails after 450 nm correspond to formally spin forbidden 3MLCT transitions, which are produced by the large spin–orbit coupling introduced, in this case, by the iridium center.

Complexes 35 and 9 display a HOMO involving substantial mixing with a π-ligand backbone (Figures S28–S30 and S34). Thus, they fulfill the main requirement in order to serve as metal–ligand species, which allow building binuclear compounds bearing metals electronically coupled, where the new HOMO is delocalized between the metal centers connected by the π-linker. As a proof-of-concept validation, the HOMO of the homobinuclear derivatives 68 and 11 is clearly delocalized throughout the metal–heterocycle-metal system (Figure 5) with similar participation percentage of the three moieties. As in the mononuclear precursors, the LUMO is almost exclusively centered on the heterocycle. The UV–vis spectra of the binuclear osmium compounds 68 show bands between 274 and 496 nm corresponding to osmium-to-heterocycle charge transfer (1MLCT) mixed with heterocycle-to-heterocycle transitions and weak absorption tails after 500 nm due to formally spin forbidden 3MLCT transitions, whereas the spectrum of the binuclear iridium derivative 11 contains bands between 249 and 431 nm assigned to iridium-to-heterocycle charge transfer (1MLCT) mixed with heterocycle-to-heterocycle transitions and weak absorption tails after 440 nm due to formally spin forbidden 3MLCT transitions.

The HOMO delocalization throughout metal–heterocycle-metal of the binuclear complexes requires not only the HOMO delocalization along metal-heterocycle of the metal–ligand mononuclear precursor but also electronic compatibility between the metal fragments linked by the heterocycle. This is given in complexes 68, where the heterocycle links two d4-metal fragments, and in complex 11 formed by two d6-metal moieties. In contrast to complexes 68 and 11, the heterocycle of the heterobinuclear derivative 10 associates fragments of two different ions, d4 and d6, which appear to be inconsistent to produce electronic coupling. Thus, the HOMO of this compound (Figure S35) is essentially centered on the osmium atom (45%) and the heterocycle (33%), whereas the iridium center has only a residual contribution (5%). Despite this difference, the UV–vis spectrum of 10 can be rationalized in a similar manner to those of 7 and 8.

The mononuclear complexes 35 and binuclear derivatives 68 are osmium(IV) phosphorescent emitters in the orange-red spectral region (546–728 nm) upon photoexcitation. Emission spectra in doped poly(methyl methacrylate) (PMMA) film at 5 wt % at room temperature and MeTHF at room temperature and at 77 K are shown in Figures S37–S54, whereas Table 2 shows the experimental and calculated wavelengths, observed lifetimes, quantum yields, and radiative and nonradiative rate constants. The spectra of the six compounds are very similar, which is consistent with the scarce differences found for the DFT-calculated HOMO–LUMO gaps (3.26–3.54 eV, see Table 2). Because the emissions can be attributed to T1 excited states, there is good agreement between the experimental wavelengths and those calculated through the estimation of the difference in energy between the optimized triplet states T1 and the singlet states S0 in tetrahydrofuran. The observed lifetimes are in the range of 1.5–5.2 μs. Quantum yields are modest and higher for the binuclear compounds. This poor efficiency could be related to the low value of the radiative rate constants. Phosphorescent emitters based on osmium20 are comparatively much less frequent than those of iridium21 and platinum22 in particular the osmium(IV) ones.19,23 Complexes 68 are the first reported binuclear osmium(IV) emitters. In contrast to 38, the iridium derivatives 911 are not emissive.

Table 2. Emission Properties of Complexes 38a.

complex HOMO (eV) LUMO (eV) HLG (eV) calc λema (nm) media (T, K) λem (nm) τobs (μs) Φ krb (s–1) knrb (s–1) kr/knr
3         PMMA (298) 599 1.5 0.01 6.6 × 103 6.6 × 105 0.01
–4.81 –1.27 3.54 591 MeTHF (298) 614 2.3 0.01 4.3 × 103 4.3 × 105 0.01
        MeTHF (77) 578 4.1        
4         PMMA (298) 593 3.1 0.02 6.4 × 103 3.1 × 105 0.02
–4.75 –1.24 3.51 602 MeTHF (298) 610 3.1 0.02 6.4 × 103 3.1 × 105 0.02
        MeTHF (77) 550 5.2        
5         PMMA (298) 593 3.3 0.01 3.0 × 103 3.0 × 105 0.01
–4.78 –1.29 3.49 564 MeTHF (298) 611 2.7 0.01 3.7 × 103 3.6 × 105 0.01
        MeTHF (77) 574 3.4        
6         PMMA (298) 562, 599 4.4 0.03 6.8 × 103 2.2 × 105 0.03
–4.58 –1.11 3.47 577 MeTHF (298) 562 3.9 0.02 5.1 × 103 2.5 × 105 0.02
        MeTHF (77) 546, 592 3.9        
7         PMMA (298) 570, 611 3.7 0.06 1.6 × 104 2.5 × 105 0.06
–4.50 –1.08 3.42 596 MeTHF (298) 627 4.5 0.03 6.6 × 103 2.1 × 105 0.03
        MeTHF (77) 559, 620 3.5        
8         PMMA (298) 598, 658, 728 2.7 0.03 1.1 × 104 3.6 × 105 0.03
–4.35 –1.09 3.26 620 MeTHF (298) 602, 651 1.5 0.08 5.3 × 104 6.1 × 105 0.08
        MeTHF (77) 584, 641, 708 3.3        
a

Predicted from TD-DFT calculations in THF at 298 K by estimating the energy difference between the optimized T1 and singlet S0 state.

b

Calculated according to the equations kr = Φ/τobs and knr = (1 – Φ)/τobs, where kr is the radiative rate constant, knr is the nonradiative rate constant, Φ is the quantum yield, and τobs is the excited-state lifetime.

Electrochemical Properties

The redox properties of the osmium precursors 35, the mononuclear iridium complex 9, and the binuclear derivatives 68, 10, and 11 were evaluated by cyclic voltammetry performed under argon atmosphere in a 0.1 M [NBu4]PF6 dichloromethane solution, and the potentials were referenced versus Fc/Fc+. Table 3 summarizes the main findings.

Table 3. Electrochemical Data of Complexes 311.

complex Eox1 (V) E1/2ox1 (V) Eox2 (V) E1/2ox2 (V) Eox3 (V) E1/2ox3 (V) Kc(1–2) Kc(2–3)
3 0.31   0.77 0.72        
4 0.25   0.72 0.67        
5 0.29 0.25 0.76 0.69        
6 0.09 0.06 0.42 0.39 0.73 0.69 8.30 × 105 8.03 × 104
7 0.03 0.00 0.39 0.35 0.82   1.55 × 106 1.37 × 107
8 –0.41 –0.46 0.00 –0.05 0.42 0.38 2.03 × 107 6.84 × 106
9 0.02 –0.03 0.43 0.37 0.84 0.80    
10 –0.32 –0.35 0.07   0.31   2.86 × 105 3.24 × 103
11 –0.18 –0.20 0.15 0.08 0.38   2.90 × 106 9.86 × 103

Mononuclear complexes 35 exhibit Os(IV)/Os(V) and Os(V)/Os(VI) oxidations peaks between 0.20 and 0.80 V. The second oxidation is quasi-reversible for the three compounds, whereas the first one is irreversible for 3 and 4 and quasi-reversible for 5 (Figures S55–S57). The mononuclear iridium compound 9 displays three quasi-reversible Ir(III)/Ir(IV), Ir(IV)/Ir(V), and Ir(V)/Ir(VI) oxidation peaks at 0.02, 0.43, and 0.84 V, respectively (Figure S61).

The cyclic voltammograms of the homobinuclear osmium derivatives 68 (Figure 6) contain three quasi-reversible [Os2]/[Os2]+, [Os2]+/[Os2]2+, and [Os2]2+/[Os2]3+ oxidation peaks between −0.45 and 0.90 V. The first of them is observed in the range −0.45–0.10 V, the second one between 0.00 and 0.42 V, and the last one in the range 0.42–0.90 V. Both separations between the consecutive waves (ΔE) are long, yielding large values of Kc (Kc = e–nFΔE/RT),24 between 8.03 × 104 and 2.03 × 107. They in a first glance suggest class III radicals with the odd electron fully delocalized (eq 1; n = 4, 5).3d The homobinuclear iridium complex 11 (Figure 6d) also exhibits three oxidation peaks at −0.18, 0.15, and 0.38 V. However, the separations between them are in this case significantly different. The separation between the first oxidation and the second one is long, giving rise to a Kc value of 2.90 × 106, which lies within the range found for 68. On the other hand, the separation between the second oxidation peak and the third one is shorter. It only allows calculating a Kc value of 9.86 × 103. The heterobinuclear complex 10 (Figure S62) has three quasi-reversible oxidation peaks at −0.32, 0.07, and 0.31 V, corresponding to independent events on each metal. According to the contribution of the metal centers to the HOMO of the species generated in the process and their respective spin density maps (Figure S121), the first oxidation appears to take place on the osmium center, whereas the second and third ones should occur on the iridium center.

graphic file with name ic0c03680_m001.jpg 1

Figure 6.

Figure 6

Cyclic voltammograms of 10–3 M dichloromethane solutions of complexes 6 (a), 7 (b), 8 (c), and 11 (d). Supporting electrolyte: [Bu4N]PF6 (0.1 M). Scan rate: 100 mV s–1. The potentials were referenced to the ferrocene/ferrocenium (Fc/Fc+) couple.

UV–vis–NIR Spectra of the Oxidized Binuclear Species

UV–vis–NIR spectroelectrochemical investigations on 1 × 10–3 M solutions of the homobinuclear complexes 7, 8, and 11 and the heterobinuclear derivative 10 in dichloromethane and in the presence of 0.1 M of [NBu4]PF6 were carried out in order to corroborate the formation of mixed valence species, suggested by the electrochemical study, as a result of the performed oxidations. In contrast to 7, 8, 10, and 11, the solubility of the symmetrical complex 6 in the usual organic solvents is not enough to carry out the same spectroelectrochemical study with this compound.

The comparison of the spectra of the monocations [M2]+ with those of the neutral complexes reveals interesting findings. The spectrum of the monocation [7]+ (Figure S85) shows growing of the absorption bands in the visible region between 450 and 550 nm, with regard to that of 7 (Figure S84), together with the appearance of a broader absorption centered at 1746 nm in the NIR region. This behavior is ascribed to a HOMO(B)-to-LUMO(B) intervalence charge transfer transition (IVCT) signature by a mixed-valence species. An IVCT band is also observed in the spectrum of [8]+ (Figure 7, green line). It appears at 1705 nm, slightly red-shifted with regard to [7]+ by about 40 nm, being much more intense. In contrast, the spectrum of the diiridium cation [11]+ (Figure S93) has a much less intense IVCT band at 912 nm, blue-shifted. The spectrum of the heterobimetallic Os–Ir cation [10]+ (Figure S98) does not contain any perceptible IVCT band, in spite of that DFT calculations predict a weak transition at 1066 nm.

Figure 7.

Figure 7

UV–vis–NIR absorption spectra of complexes 8-[8]3+ in dichloromethane solutions.

The oxidation of the monocations to the [M2]2+ species gives rise to the disappearance of the IVCT band in some cases. Spectra of the dications [7]2+ and [11]2+ do not contain any IVCT band (Figures S86 and S94). However, an intense IVCT transition at 2000 nm is observed in the spectrum of [8]2+ (Figure 7, red line); which is about 300 nm red-shifted with regard to that of [8]+. DFT calculations on [8]2+ reveal that the triplet state is 3.9 kcal/mol more stable than the singlet state, so that [8]2+ should be described as a diradical. The oxidation from the dications [M2]2+ to the trications [M2]3+ regenerates mixed-valence species for [7]3+ and [8]3+. Thus, the spectra of [7]3+ (Figure S87) and [8]3+ (Figure 7, blue line) contain an IVCT transition centered at about 1824 and 1614 nm, red-shifted by 77 and 15 nm with regard to those of the respective monocations.

Mixed-valence transition metal complexes can be classified using the delocalization parameter Γ,12b which is calculated by means of eq 2(3d)

graphic file with name ic0c03680_m002.jpg 2

where Δν1/2 and Δνmax are the bandwidth at the half height and the maximum absorption, respectively, for a Gaussian-shaped ICTV band (cm–1).

Table 4 collects the values of the delocalization parameter, calculated according to eq 2, for the ITCV bands previously mentioned. Values of Γ < 0.5 indicate mixed-valence complexes of class II, while values of Γ > 0.5 are characteristic of compounds of class III. Complexes in the borderline class II/class III display values of Γ ≈ 0.5. According to this criteria, cation [7]3+ belongs to class II, whereas ITCV bands of the cations resulting from the three sequential oxidations of the asymmetrical homobinuclear osmium complex 8 and the diiridium cation [11]+ give Γ values, which fit to class III. Cation [7]+ appears to be a species of the borderline class II/class III with a Γ-value of 0.51.

Table 4. Mixed-Valence and IVCT Parameters.

complex νmax/cm–1a Δν1/2/cm–1a Δν1/20/cm–1a Γb
[7]+ 9676 2287 4728 0.51
[7]3+ 11088 3467 5061 0.31
[8]+ 6307 1427 3817 0.63
[8]2+ 5338 742 3511 0.78
[8]3+ 6481 1045 3869 0.72
[11]+ 11131 2001 5071 0.61
a

From Gaussian fit of ε/ν versus ν.

b

Parameters calculated using eq 2.

Nature of the MHn Units upon Oxidation

The dissociation energy of a hydrogen molecule from a polyhydride complex depends upon the electron density of the metal. This energy increases as the hydrogen–hydrogen interaction decreases and therefore it is higher for hydride forms than for dihydrogen ones. This is a direct consequence of the metal-dihydrogen bonding situation. Similar for all σ-complexes, the interaction between the coordinated hydrogen molecule and the transition metal in the dihydrogen compounds involves σ-donation from the σ-orbital of the coordinated bond to empty orbitals of the metal and back bonding from the metal to the σ*-orbital of the bond. The balance between donation and back-donation determines the oxidative addition degree, which has been fit to the separation between the coordinated hydrogen atoms.1 To gain insight into the influence of the sequential oxidation of the binuclear complexes 68, 10, and 11 on the respective MHn units, we comparatively analyzed the hydrogen–hydrogen separations in the optimized structures of the generated cations (Figures S64–S83). Chart 1 gives a view of these structures, whereas Table 5 gathers the separation between the hydrogen atoms of the MHn units.

Chart 1. Nature of the MHn Units of the Neutral and Cationic Forms.

Chart 1

Table 5. Calculated (B3LYP-D3//SDD(f)/6-31G**) Separation between the Hydrogen Atoms Bonded to the Metal.

complex H1—H2 (Å) H2—H3 (Å) H4—H5 (Å) H5—H6 (Å) H3—H4 (Å)
6 1.6 1.8 1.6 1.8  
[6]+ 1.5 1.8 1.6 1.8  
[6]2+ 0.9 2.3 1.6 1.8  
[6]3+ 0.9 2.3 1.4 1.9  
7 1.6 1.8 1.6 1.7  
[7]+ 1.5 1.8 1.6 1.7  
[7]2+ 0.9 2.3 1.6 1.7  
[7]3+ 0.9 2.2 1.5 1.8  
8 1.6 1.8 1.6    
[8]+ 1.6 1.8 1.6    
[8]2+ 1.5 1.9 1.5    
[8]3+ 0.9 2.2 1.4    
10 1.6 1.8 2.3    
[10]+ 1.5 1.8 2.3    
[10]2+ 0.9 2.3 2.3    
[10]3+ 0.9 2.2 2.4    
11 2.3       2.3
[11]+ 2.3       2.4
[11]2+ 2.3       2.5
[11]3+ 2.3       2.5

The neutral complexes are in the four cases classical hydrides with separations between their hydride ligands longer than 1.6 Å. The monocations [M2]+ are also pure hydrides, although it should be mentioned that subtle but significant differences are observed between them. Two of the hydride ligands of a half of [6]+ approach about 0.1 Å to form a compressed dihydride (H(1) and H(2)). The same behavior is observed in the OsH3(PiPr3)2 moiety of [7]+ linked to the nitrogen atom of the unsubstituted pyridyl ring and in the OsH3(PiPr3)2 moiety of [10]+. In contrast, the hydrides of [8]+ and [11]+ are not affected. This difference in behavior appears to be connected with the distribution of the frontier orbitals of the cations (Figure 8). The SOMO of [6]+ (a), [7]+ (b), and [10]+ (d) is mainly centered on the heterocycle linker and the metal center keeping invariant the MHn unit, while the LUMO is distributed between the heterocycle linker and the metal center of the modified MHn unit. In contrast, both SOMO and LUMO of [8]+ (c) are delocalized on the heterocycle linker and the metals. The SOMO of [11]+ (e) is similarly distributed. However, the LUMO is mainly centered on the heterocycle linker and one of the metals.

Figure 8.

Figure 8

SOMO and LUMO of complexes [6]+ (a), [7]+ (b), [8]+ (c), [10]+ (d), and [11]+ (e).

The oxidation from the monocations to the [M2]2+ species enhances the approaching of the compressed dihydrides, which become a Kubas-type dihydrogen in [6]2+, [7]2+, and [10]2+. On the other hand, two hydrides of each OsHn unit of the cation [8]2+ approach 0.1 Å to generate a compressed dihydride attached to each metal. In contrast to [6]2+, [8]2+ and [10]2+, the hydrides of [11]2+ remain unaltered. The oxidation of the dications has also different implications depending upon the generated trication. Cations [6]3+ and [7]3+ undergo the transformation of two hydride ligands of the previously unaffected OsH3(PiPr3)2 moiety from classical to compressed, while the hydrogen atoms of the other are not affected. The OsH3(PiPr3)2 moiety of [8]2+ is more sensitive to the oxidation than the OsH2(PiPr3)2 one. Thus, while the compressed dihydrides of the OsH3(PiPr3)2 moiety of [8]2+ are transformed into a Kubas-type dihydrogen in [8]3+, those of the OsH2(PiPr3)2 moiety only experience a slight approach. The MHn units of cations [M2]2+ and [M2]3+ of 10 and 11 display similar parameters.

The previous observations suggest that the MHn units of d4-osmium fragments are more sensitive to the oxidation than those of d6-iridium fragments and that the metal center of the MHn unit that undergoes the transformation is that with the highest contribution to the LUMO of the binuclear species.

Concluding Remarks

The rollover cyclometalated hydride derivatives OsH32-C,N-[C5RH2N-py]}(PiPr3)2 (R = H, Me, Ph) and IrH22-C,N-[C5RH2N-py]}(PiPr3)2 display frontier orbitals involving substantial mixing of the metal center and the π-heterocycle backbone. The activation of an ortho-CH bond of the heterocyclic moiety of these metal–ligand units, promoted by the platinum group metals polyhydride complexes OsH6(PiPr3)2 and IrH5(PiPr3)2, gives rise to four different classes of binuclear derivatives: the hexahydride (PiPr3)2H3Os{μ-[κ2-C,N-[C5RH2N-C5H3N]-N,C2]}OsH3(PiPr3)2 compounds with two OsH3(PiPr3)2 halves, the homopentahydride (PiPr3)2H3Os{μ-[κ2-C,N-[C5H3N-C5(C6H4)H2N]-C,N,C3]}OsH2(PiPr3)2 derivative bearing OsH3(PiPr3)2 and OsH2(PiPr3)2 fragments, the heteropentahydride (PiPr3)2H3Os{μ-[κ2-C,N-[C5H3N-C5H3N]-N,C2]}IrH2(PiPr3)2 with OsH3(PiPr3)2 and IrH2(PiPr3)2 units, and the tetrahydride (PiPr3)2H2Ir{μ-[κ2-C,N-[C5H3N-C5H3N]-N,C2]}IrH2(PiPr3)2 complex formed by two IrH2(PiPr3)2 moieties. With the exception of the heterobinuclear pentahydride (PiPr3)2H3Os{μ-[κ2-C,N-[C5H3N-C5H3N]-N,C2]}IrH2(PiPr3)2, these compounds display HOMO delocalization throughout the metal–heterocycle-metal skeleton. This electronic situation lends them interesting electrochemical properties. Their sequential oxidation allows generating mixed valence species, including mono- and diradicals, which exhibit intervalence charge transfer transitions. This noticeable ability allows us to govern the strength of the hydrogen–hydrogen and metal–hydrogen interactions within the MHn units of these compounds. This finding should be of paramount importance for the attractive goal of reversibly controlling the coordination of the hydrogen molecule in transition metal polyhydride complexes.

Experimental Section

General Information

All reactions were carried out with exclusion of air using Schlenk-tube techniques or in a drybox. Instrumental methods and X-ray details are given in the Supporting Information. In the NMR spectra (Figures S1–S18) the chemical shifts (in ppm) are referenced to residual solvent peaks (1H, 13C{1H}) or external 85% H3PO4 (31P{1H}). Coupling constants J and N (N = JP–H + JP’-H for 1H and N = JP–C + JP’-C for 13C{1H}) are given in hertz.

Preparation of (PiPr3)2H3Os{μ-[κ2-C,N-[C5H3N-C5H3N]-N,C2]}OsH3(PiPr3)2 (6)

This compound can be prepared by two methods. Method a: A mixture of OsH6(PiPr3)2 (1, 77 mg, 0.149 mmol) and OsH32-C,N-[C5H3N-py]}(PiPr3)2 (3, 100 mg, 0.149 mmol) in toluene (6 mL) was refluxed for 16 h. The resulting suspension was cooled to room temperature, and the solvent was removed in vacuo. The addition of methanol (4 mL) caused the precipitation of a pale orange solid that was washed with further portions of methanol (3 × 3 mL) and finally it was dried in vacuo. Yield: 131 mg (74%). Method b: A mixture of 1 (150 mg, 0.29 mmol) and 2,2′-bipyridine (23 mg, 0.145 mmol) in toluene (6 mL) was refluxed for 16 h. The workup of the reaction is analogous as that described in Method a. Yield: 133 mg (78%). Anal. Calcd for C46H96N2Os2P4: C, 46.76; H, 8.19; N, 2.37. Found: C, 46.76; H 8.29; N, 2.31. IR (cm–1): ν(Os–H) 1985, 2102 (w). 1H NMR (300.13 MHz, toluene-d8, 298 K): δ 8.80 (d, 3JH–H = 5.4, 2H, py), 8.30 (d, 3JH–H = 7.3, 2H, py), 6.29 (m, 2H, py), 1.89 (m, 12H, PCH(CH3)2), 1.02 (dvt, 3JH–H = 6.6, N = 13.0, 36H, PCH(CH3)2), 0.99 (dvt, 3JH–H = 6.7, N = 13.4, 36H, PCH(CH3)2), −9.12 (br, 4H, Os–H), −12.30 (br, 2H, Os–H). 1H NMR (300.13 MHz, toluene-d8, high field region, 223 K): δ −5.92 (br, 2H, Os–H), −12.13 (br, 4H, Os–H). The low solubility of the solid precluded obtaining its 13C{1H} NMR spectrum. 31P{1H} NMR (121.50 MHz, toluene-d8, 298 K): δ 23.1 (s). T1(min) (ms, OsH. 300 MHz, toluene-d8, 223 K): 54 ± 5 (−5.92 ppm); 97 ± 10 (−12.13).

Preparation of (PiPr3)2H3Os{μ-[κ2-C,N-[C5H2MeN-C5H3N]-N,C2]}OsH3(PiPr3)2 (7)

This compound can be prepared by two methods. Method a: A mixture of 1 (76 mg, 0.146 mmol) and OsH32-C,N-[C5(Me)H2N-py]}(PiPr3)2 (4, 100 mg, 0.146 mmol) in toluene (4 mL) was refluxed for 16 h, giving a dark orange suspension. After cooling the mixture to room temperature, the solvent was removed in vacuo, affording an orange residue. Addition of cold methanol (3 mL) caused the precipitation of an orange solid that was washed with cold methanol (3 × 3 mL) and dried in vacuo. Yield: 136 mg (78%). Method b: A mixture of 1 (100 mg, 0.194 mmol) and 6-methyl-2,2′-bipyridine (14.7 μL, 0.095 mmol) in toluene (4 mL) was refluxed for 16 h, giving a dark orange solution. The workup of the reaction is analogous at that described in Method a. Yield: 96 mg (83%). Anal. Calcd for C47H98N2Os2P: C, 47.21; H, 8.26; N, 2.34. Found: C, 47.31; H, 7.96; N, 2.34. HRMS (electrospray, m/z): calculated for C47H98N2Os2P4 [M]+, 1198.5905, found, 1198.5911. IR (cm–1): ν(Os–H) 1978 (w). 1H NMR (300.13 MHz, toluene-d8, 298 K): δ 8.85 (d, 1H, 3JH–H = 5.0, py), 8.32 (d, 3JH–H = 7.4, 1H, Me-py), 8.21 (d, 3JH–H = 7.1, 1H, py), 6.62 (d, 3JH–H = 7.4, 1H, Me-py), 6.25 (m, 1H, py), 2.91 (s, 3H, CH3), 1.93 (m, 12H, PCH(CH3)2), 1.04 (dvt, 3JH–H = 6.5, N = 13.3, 36H, PCH(CH3)2), 1.09–0.92 (m, 72H, PCH(CH3)2), −9.21 (br, 4H, Os–H), −12.26 (br, 1H, Os–H), −13.17 (br, 1H, Os–H). 13C{1H}-apt NMR (75.48 MHz, toluene-d8, 298 K): δ 173.9 (t, 2JC–P = 6.8, Os–C py), 173.7 (s, C Me-py), 173.6 (s, C py), 168.9 (t, 2JC–P= 6.2, Os–C Me-py), 153.8 (s, CH Me-py), 150.9 (s, CH py), 150.0 (s, CH py), 149.9 (s, C Me-py), 122.0 (s, CH py), 121.4 (s, CH Me-py), 33.3 (s, CH3), 28.3 (vt, N = 23.0, PCH(CH3)2), 28.2 (vt, N = 23.2, PCH(CH3)2), 20.6, 20.5, and 20.4 (all s, PCH(CH3)2). 31P{1H} NMR (121.50 MHz, toluene-d8, 298 K): δ 22.7 (s), 21.4 (s). T1(min) (ms, OsH, 300 MHz, toluene-d8, 213 K): 57 ± 6 (−6.01 ppm); 57 ± 6 (−12.16 ppm); 48 ± 5 (−13.10 ppm).

Preparation of (PiPr3)2H3Os{μ-[κ2-C,N-[C5H3N-C5(C6H4)H2N]-C,N,C-κ3]}OsH2(PiPr3)2 (8)

This compound can be prepared by two methods. Method a: A mixture of 1 (66 mg, 0.128 mmol) and OsH32-C,N-[C5(Ph)H2N-py]}(PiPr3)2 (5, 96 mg, 0.128 mmol) in toluene (3 mL) was refluxed for 16 h, giving a dark orange suspension. After the mixture was cooled to room temperature, the solvent was removed in vacuo, affording an orange residue. Addition of cold methanol (3 mL) caused the precipitation of an orange solid that was washed with methanol (3 × 3 mL) and dried in vacuo. Yield: 122 mg (76%) Method b: A mixture of 1 (150 mg, 0.290 mmol) and 6-phenyl-2,2′-bipyridine (33.7 mg, 0.145 mmol) in toluene (5 mL) was refluxed for 16 h, giving a dark orange suspension. The workup of the reaction is analogous at that described in Method a. Yield: 147.5 mg (81%). Anal. Calcd for C52H98N2Os2P4: C, 49.74; H, 7.87; N, 2.23. Found: C, 49.48; H, 7.72; N, 2.14. HRMS (electrospray, m/z): calculated for C52H97N2Os2P4 [M – H]+, 1255.5826; found, 1255.5451. IR (cm–1): ν(Os–H) 2141, 2106 (w). 1H NMR (300.13 MHz, CD2Cl2, 298 K): δ 8.63 (d, 1H, 3JH–H = 5.5, py), 8.14 (d, 3JH–H = 7.8, 1H, central py), 7.83 (d, 3JH–H = 7.4, 1H, py), 7.78 (d, 3JH–H = 7.4, 1H, Ph), 7.51 (d, 3JH–H = 7.6, 1H, Ph), 7.10 (d, 3JH–H = 7.8, 1H, central py), 6.80 (t, 3JH–H= 7.4, 1H, Ph), 6.63 (t, 3JH–H = 7.2, 1H, Ph), 6.21 (t, 3JH–H = 5.6, 1H, py), 1.97 (m, 12H, PCH(CH3)2), 0.97 (dvt, 3JH–H = 6.4, N = 13, 36H, PCH(CH3)2), 0.81 (dvt, 3JH–H = 6.4, N = 12.2, 36H, PCH(CH3)2), −8,48 (dt, 2JH–H = 11.3, 2JH–P = 15.1, 1H, Os–H), −9.19 (dt, 2JH–H = 11.3, 2JH–P = 17.2, 1H, Os–H), −9.49 (br, 2H, Os–H), −12.47 (br, 1H, Os–H). 1H NMR (300.13 MHz, CD2Cl2, 203 K, high field region): δ – 6.23 (br, 1H, Os–H), −8.48 (dt, 2JH–H = 16.8, 2JH–P = 14.2, 1H, Os–H), −9.22 (dt, 2JH–H = 16.9, 2JH–P = 10.8, 1H, Os–H), −12.57 (br, 1H, Os–H), −12.89 (br, 1H, Os–H). 13C{1H}-apt NMR (75.48 MHz, CD2Cl2, 298 K): δ 175.2 (s, C py), 170.9 (s, C central py), 169.9 (t, 2JC–P = 8.3, Os–C Ph), 168.0 (t, 2JC–P= 6.1, Os–C central py), 165.5 (t, 2JC–P = 8.5, Os–C py), 156.7 (s, C central py), 151.9 (s, CH central py), 150.3 (s, C Ph), 149.9 (s, CH Ph), 149.5 (s, CH py), 146.9 (s, CH Ph), 126.9 (s, CH Ph), 122.4 (s, CH py), 122.2 (s, CH Ph), 119.8 (s, CH Ph), 113.8 (s, CH central py), 28.4 (vt, N = 23.4, PCH(CH3)2), 27.0 (vt, N = 23.6, PCH(CH3)2), 20.6, 20.3, 19.9, and 19.4 (all s, PCH(CH3)2). 31P{1H} NMR (121.50 MHz, CD2Cl2, 298 K): δ 24.1 (s), 1.8 (s).

Preparation of (PiPr3)2H3Os{μ-[κ2-C,N-[C5H3N-C5H3N]-N,C2]}IrH2(PiPr3)2 (10)

A mixture of 2 (100 mg, 0.193 mmol) and OsH32-C,N-(C5H3N-py)}(PiPr3)2 (3, 129 mg, 0.193 mmol) in toluene (4 mL) was refluxed for 16 h. After this time, the resulting dark orange solution was cooled to room temperature, filtered through Celite and the solvent was removed in vacuo. The addition of pentane (5 mL) caused the precipitation of an orange solid; this was washed with further portions of pentane (3 × 2 mL) and finally dried in vacuo. Yield: 155 mg (68%). Anal. Calcd for C46H95IrN2OsP4: C, 46.72; H, 8.10; N, 2.37. Found: C, 46.72; H, 8.12; N, 2.36. HRMS (electrospray, m/z) calcd for C46H94IrN2OsP4 [M – H]+, 1183.5585; found, 1183.5529. IR (cm–1): ν(Ir–H) 2141 (m), ν(Os–H) 2102 (m), 1988 (m). 1H NMR (300.13 MHz, C6D6, 298 K): δ 8.93 (d, 3JH–H = 5.5, 1H, CH py), 8.47 (m, 2H, CH py), 8.11 (d, 3JH–H = 7.3, 1H, CH py), 6.38 (m, 2H, CH py), 2.04 (m, 6H, PCH(CH3)2), 1.93 (m, 6H, PCH(CH3)2), 1.05 (m, 72H, PCH(CH3)2), −8.98 (br, 2H, Os–H), −12.24 (br, 1H, Os–H), −13.05 (dt, 2JH–H = 4.1, 2JH–P = 21.4, 1H, Ir–H), −22.16 (dt, 2JH–H = 4.1, 2JH–P = 19.3, 1H, Ir -H). 1H NMR (300.13 MHz, toluene-d8, 223 K, high field region): δ – 5.80 (br, 1H, Os–H), −12.10 (br, 2H, Os–H), −12.93 (br t,2JH–P = 18.9, 1H, Ir–H), −21.93 (dt,2JH–H = 3.6, 2JH–P = 18.9, 1H, Ir–H). 13C{1H}-apt NMR (75.48 MHz, C6D6, 298 K): δ 176.1, 174.0 (both s, C py), 173.4 (t, 2JC–P = 6.4, Os–C), 163.2 (t, 2JC–P = 6.2, Ir–C), 152.1, 150.1, 150.0, 147.7, 122.2, 122.1 (all s, CH py), 28.0 (vt, N = 23.3, PCH(CH3)2), 27.4 (vt, N = 26.8, PCH(CH3)2), 20.4, 20.3, 20.3, 20.1 (all s, PCH(CH3)2). 31P{1H} NMR (161.99 MHz, C6D6, 298 K): δ 30.3 (s, Ir–P), 22.4 (s, Os–P). T1(min) (ms, OsH, 300 MHz, toluene-d8, 243 K): 66 ± 7 (−12.15 ppm), value of the resonance at −5.98 ppm could not be calculated due to the broadness of it.

Preparation of (PiPr3)2H2Ir{μ-[κ2-C,N-[C5H3N-C5H3N]-N,C2]}IrH2(PiPr3)2 (11)

This compound can be prepared by two methods. Method a: A mixture of 2 (115 mg, 0.224 mmol) and IrH22-C,N-[C5H3N-py]}(PiPr3)2 (9, 150 mg, 0.224 mmol) in toluene (8 mL) was refluxed for 16 h. After this time, the resulting yellow dark solution was cooled to room temperature, filtered through Celite, and the solvent was removed in vacuo. The addition of pentane (5 mL) caused the precipitation of a yellow solid, which was washed with further portions of pentane (3 × 5 mL), and finally, it was dried in vacuo. Yield: 225 mg (85%). Method b: A mixture of 2 (200 mg, 0.386 mmol) and 2,2′-bipyridine (30 mg, 0.193 mmol) in toluene (8 mL) was refluxed for 16 h. The workup of the reaction is analogous at that described in Method a. Yield: 204 mg (89%). Anal. Calcd for C46H94Ir2N2P4: C, 46.68; H, 8.01; N, 2.37. Found: C, 46.83; H, 8.18; N, 2.36. HRMS (electrospray, m/z) calcd for C46H93Ir2N2P4 [M – H]+, 1183.5542; found, 1183.5351. IR (cm–1): ν(Ir–H) 2151 (m), 1931 (m). 1H NMR (300 MHz, C6D6, 298 K): δ 8.57 (d, 3JH–H = 5.1, 2H, CH py), 8.22 (d, 3JH–H = 7.1, 2H, CH py), 6.46 (dd, 3JH–H = 7.1, 3JH–H = 5.1, 2H, CH py), 2.04 (m, 12H, PCH(CH3)2), 1.08 (dvt, 3JH–H = 6.6, N = 13.5, 36H, PCH(CH3)2), 1.04 (dvt, 3JH–H = 6.7, N = 13.1, 36H, PCH(CH3)2), −12.93 (dt, 2JH–H = 4.1, 2JH–P = 21.2, 1H, Ir–H), −22.00 (dt, 2JH–H = 4.1, 2JH–P = 19.3, 1H, Ir–H). 13C{1H}-apt NMR (75.45 MHz, C6D6, 298 K): δ 177.3 (s, C py), 163.6 (t, 2JC–P = 6.5, Ir–C py), 150.8, 148.5, 121.9 (all s, CH py), 27.4 (vt, N = 26.8, PCH(CH3)2), 20.4, 20.1 (both s, PCH(CH3)2). 31P{1H} NMR (121.5 MHz, C6D6, 298 K): δ 29.9 (s).

UV–vis–NIR Spectroelectrochemical Investigations

Spectroelectrochemical experiments combine UV–vis–NIR spectroscopic measurements and redox processes at the same time. Thus, they allow obtaining the spectra of specific controlled oxidation states. The electrochemical measurements were performed with a micro-Autolab FRA2 Type III (Methrom, Utrecht, Netherlands) potentiostat controlled by NOVA (v.2.1.4) software. For the optical measurements, a JASCO V670 spectrophotometer using quartz (1 mm optical path length) was used. The spectroelectrochemical cell (1 mL volume) was a DRP-PTGRID-TRANSCELL (DropSens). It contains an optically transparent Pt grid working electrode (0.6 × 0.4 cm) which allows the bulk electrolysis of the solution contained in the cell, a Ag/AgCl reference electrode, and a platinum counter electrode. The experiments were performed under argon and protected from the light in dichloromethane solution (10–3 M) with [Bu4N]PF6 as a supporting electrolyte (0.1 M). To obtain the UV–vis–NIR spectra, anodic potentials according to the previously measured cyclic voltammograms were applied for the corresponding oxidations to [M2]+, [M2]2+, and [M2]3+ during the wavelength scan.

Acknowledgments

Financial support from the MINECO of Spain (Projects CTQ2017-82935-P and RED2018-102387-T (AEI/FEDER, UE)), Gobierno de Aragón (Group E06_20R and project LMP148_18), FEDER, and the European Social Fund is acknowledged.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.0c03680.

  • Experimental details; crystallographic data; NMR spectra; experimental and computed UV/vis spectra; photophysical, UV–vis–NIR spectroelectrochemistry, electrochemical and theoretical studies (PDF)

  • Cartesian coordinates of the optimized structures (XYZ)

The authors declare no competing financial interest.

Supplementary Material

ic0c03680_si_001.pdf (12.2MB, pdf)
ic0c03680_si_002.xyz (317.9KB, xyz)

References

  1. a Kubas G. J. Metal-dihydrogen and σ-bond coordination: the consummate extension of the Dewar-Chatt-Duncanson model for metal-olefin π bonding. J. Organomet. Chem. 2001, 635, 37–68. 10.1016/S0022-328X(01)01066-X. [DOI] [Google Scholar]; b Kubas G. J. Fundamentals of H2 Binding and Reactivity on Transition Metals Underlying Hydrogenase Function and H2 Production and Storage. Chem. Rev. 2007, 107, 4152–4205. 10.1021/cr050197j. [DOI] [PubMed] [Google Scholar]; c Morris R. H. Dihydrogen, dihydride and in between: NMR and structural properties of iron group complexes. Coord. Chem. Rev. 2008, 252, 2381–2394. 10.1016/j.ccr.2008.01.010. [DOI] [Google Scholar]; d Kubas G. J. Activation of dihydrogen and coordination of molecular H2 on transition metals. J. Organomet. Chem. 2014, 751, 33–49. 10.1016/j.jorganchem.2013.07.041. [DOI] [Google Scholar]; e Crabtree R. H. Dihydrogen Complexation. Chem. Rev. 2016, 116, 8750–8769. 10.1021/acs.chemrev.6b00037. [DOI] [PubMed] [Google Scholar]; f Esteruelas M. A.; López A. M.; Oliván M. Polyhydrides of Platinum Group Metals: Nonclassical Interactions and σ-Bond Activation Reactions. Chem. Rev. 2016, 116, 8770–8847. 10.1021/acs.chemrev.6b00080. [DOI] [PubMed] [Google Scholar]; g Special issue Metal Hydrides, Chem. Rev. 2016, 116, 8315−9000. [DOI] [PubMed] [Google Scholar]
  2. Kaim W. Manifestations of Noninnocent Ligand Behavior. Inorg. Chem. 2011, 50, 9752–9765. 10.1021/ic2003832. [DOI] [PubMed] [Google Scholar]
  3. a Paul F.; Meyer W. E.; Toupet L.; Jiao H.; Gladysz J. A.; Lapinte C. A “Conjugal” Consanguineous Family of Butadiynediyl-Derived Complexes: Synthesis and Electronic Ground States of Neutral, Radical Cationic, and Dicationic Iron/Rhenium C4 Species. J. Am. Chem. Soc. 2000, 122, 9405–9414. 10.1021/ja0011055. [DOI] [Google Scholar]; b Jiao H.; Costuas K.; Gladysz J. A.; Halet J.-F.; Guillemot M.; Toupet L.; Paul F.; Lapinte C. Bonding and Electronic Structure in Consanguineous and Conjugal Iron and Rhenium sp Carbon Chain Complexes[MC4M’]n+: Computational Analyses of the Effect of the Metal. J. Am. Chem. Soc. 2003, 125, 9511–9522. 10.1021/ja034619n. [DOI] [PubMed] [Google Scholar]; c Ceccon A.; Santi S.; Orian L.; Bisello A. Electronic communication in heterobinuclear organometallic complexes through unsaturated hydrocarbon bridges. Coord. Chem. Rev. 2004, 248, 683–724. 10.1016/j.ccr.2004.02.007. [DOI] [Google Scholar]; d Aguirre-Etcheverry P.; O’Hare D. Electronic Communication through Unsaturated Hydrocarbon Bridges in Homobimetallic Organometallic Complexes. Chem. Rev. 2010, 110, 4839–4864. 10.1021/cr9003852. [DOI] [PubMed] [Google Scholar]; e Halet J.-F.; Lapinte C. Charge delocalization vs localization in carbon-rich iron mixed-valence complexes: A subtle interplay between the carbon spacer and the (dppe)Cp*Fe organometallic electrophore. Coord. Chem. Rev. 2013, 257, 1584–1613. 10.1016/j.ccr.2012.09.007. [DOI] [Google Scholar]; f Zhang D.-B.; Wang J.-Y.; Wen H.-M.; Chen Z.-N. Electrochemical, Spectroscopic, and Theoretical Studies on Diethynyl Ligand Bridged Ruthenium Complexes with 1,3-Bis(2-pyridylimino)isoindolate. Organometallics 2014, 33, 4738–4746. 10.1021/om401235p. [DOI] [Google Scholar]; g Yu C.-H.; Yang X.; Ji X.; Wang C.-H.; Lai Q.; Bhuvanesh N.; Ozerov O. V. Redox Communication between Two Diarylamido/Bis(phosphine) (PNP)M Moieties Bridged by Ynediyl Linkers (M = Ni, Pd, Pt). Inorg. Chem. 2020, 59, 10153–10162. 10.1021/acs.inorgchem.0c01281. [DOI] [PubMed] [Google Scholar]
  4. a Barlow S.; O’Hare D. Metal-Metal Interactions in Linked Metallocenes. Chem. Rev. 1997, 97, 637–669. 10.1021/cr960083v. [DOI] [PubMed] [Google Scholar]; b Hildebrandt A.; Miesel D.; Lang H. Electrostatic interactions within mixed-valent compounds. Coord. Chem. Rev. 2018, 371, 56–66. 10.1016/j.ccr.2018.05.017. [DOI] [Google Scholar]
  5. a Forster R. J.; Keyes T. E. Tetrazine Bridged Osmium Dimers: Electrochemical vs Photoinduced Electron Transfer. J. Phys. Chem. B 2001, 105, 8829–8837. 10.1021/jp010948v. [DOI] [Google Scholar]; b Kaim W.; Klein A.; Glöckle M. Exploration of Mixed-Valence Chemistry: Inventing New Analogues of the Creutz-Taube Ion. Acc. Chem. Res. 2000, 33, 755–763. 10.1021/ar960067k. [DOI] [PubMed] [Google Scholar]; c Ghumaan S.; Lahiri G. K. Tuning intermetallic electronic coupling in polyruthenium systems via molecular architecture. Proc. - Indian Acad. Sci., Chem. Sci. 2006, 118, 537–545. 10.1007/BF02703951. [DOI] [Google Scholar]; d Browne W. R.; Hage R.; Vos J. G. Tuning interaction in dinuclear ruthenium complexes: HOMO versus LUMO mediated superexchange through azole and azine bridges. Coord. Chem. Rev. 2006, 250, 1653–1668. 10.1016/j.ccr.2005.12.008. [DOI] [Google Scholar]; e Sarkar B.; Patra S.; Fiedler J.; Sunoj R. B.; Janardanan D.; Lahiri G. K.; Kaim W. Mixed-Valent Metals Bridged by a Radical Ligand: Fact orFiction Based on Structure-Oxidation State Correlations. J. Am. Chem. Soc. 2008, 130, 3532–3542. 10.1021/ja077676f. [DOI] [PubMed] [Google Scholar]; f Kaim W.; Sarkar B. Mixed valency of a 5d element: The osmium example. Coord. Chem. Rev. 2013, 257, 1650–1659. 10.1016/j.ccr.2012.08.026. [DOI] [Google Scholar]; g Kubiak C. P. Inorganic Electron Transfer: Sharpening a Fuzzy Border in Mixed Valency and Extending Mixed Valency across Supramolecular Systems. Inorg. Chem. 2013, 52, 5663–5676. 10.1021/ic302331s. [DOI] [PubMed] [Google Scholar]; h Hu Y. X.; Zhang J.; Zhang F.; Wang X.; Yin J.; Hartl F.; Liu S. H. Electronic Properties of Oxidized Cyclometalated Diiridium Complexes: Spin Delocalization Controlled by the Mutual Position of the Iridium Centers. Chem. - Eur. J. 2020, 26, 4567–4575. 10.1002/chem.201904894. [DOI] [PubMed] [Google Scholar]
  6. Gransbury G. K.; Livesay B. N.; Janetzki J. T.; Hay M. A.; Gable R. W.; Shores M. P.; Starikova A.; Boskovic C. Understanding the Origin of One- or Two-Step Valence TautomericTransitions in Bis(dioxolene)-Bridged Dinuclear Cobalt Complexes. J. Am. Chem. Soc. 2020, 142, 10692–10704. 10.1021/jacs.0c01073. [DOI] [PubMed] [Google Scholar]
  7. a Tsukada S.; Shibata Y.; Sakamoto R.; Kambe T.; Ozeki T.; Nishihara H. Ir3Co6 and Co3Fe3 Dithiolene Cluster Complexes: Multiple Metal-Metal Bond Formation and Correlation between Structure and Internuclear Electronic Communication. Inorg. Chem. 2012, 51, 1228–1230. 10.1021/ic202548g. [DOI] [PubMed] [Google Scholar]; b Vacher A.; Le Gal Y.; Roisnel T.; Dorcet V.; Devic T.; Barrière F.; Lorcy D. Electronic Communication within Flexible Bisdithiolene Ligands Bridging Molybdenum Centers. Organometallics 2019, 38, 4399–4408. 10.1021/acs.organomet.9b00485. [DOI] [Google Scholar]
  8. Begum N.; Hyder M. I.; Kabir S. E.; Hossain G. M. G.; Nordlander E.; Rokhsana D.; Rosenberg E. Dithiolate Complexes of Manganese and Rhenium: X-ray Structure and Properties of an Unusual Mixed Valence Cluster Mn3(CO)6(μ-η2-SCH2CH2CH2S)3. Inorg. Chem. 2005, 44, 9887–9894. 10.1021/ic050987b. [DOI] [PubMed] [Google Scholar]
  9. a Sheng T.; Vahrenkamp H. Long Range Metal-Metal Interactions Along Fe-NC-Ru-CN-Fe Chains. Eur. J. Inorg. Chem. 2004, 2004, 1198–1203. 10.1002/ejic.200300656. [DOI] [Google Scholar]; b Endicott J. F.; Chen Y.-J. Electronic coupling between metal ions in cyanide-bridged ground state and excited state mixed valence complexes. Coord. Chem. Rev. 2013, 257, 1676–1698. 10.1016/j.ccr.2012.10.018. [DOI] [Google Scholar]; c Xiao Y.; Cheung A. W.-Y.; Lai S.-W.; Cheng S.-C.; Yiu S.-M.; Leung C.-F.; Ko C.-C. Electronic Communication in Luminescent Dicyanorhenate-Bridged Homotrinuclear Rhenium(I) Complexes. Inorg. Chem. 2019, 58, 6696–6705. 10.1021/acs.inorgchem.9b00072. [DOI] [PubMed] [Google Scholar]
  10. Zhang L.-Y.; Shi L.-X.; Chen Z.-N. Syntheses, Structures, and Electronic Interactions of Dicyanamide/Tricyanometanide Bridged Binuclear Organometallic Complexes. Inorg. Chem. 2003, 42, 633–640. 10.1021/ic025879t. [DOI] [PubMed] [Google Scholar]
  11. a Robin M. B.; Day P. Mixed Valence Chemistry: A Survey and Classification. Adv. Inorg. Chem. Radiochem. 1968, 10, 247–422. 10.1016/S0065-2792(08)60179-X. [DOI] [Google Scholar]; b Parthey M.; Kaupp M.. Quantum-chemical insights into mixed-valence systems: within and beyond the Robin-Day scheme. Chem. Soc. Rev. 2014, 43, 5067–5088. 10.1039/C3CS60481K [DOI] [PubMed] [Google Scholar]
  12. a Demadis K. D.; Hartshorn C. M.; Meyer T. J. The Localized-to-Delocalized Transition in Mixed-Valence Chemistry. Chem. Rev. 2001, 101, 2655–2685. 10.1021/cr990413m. [DOI] [PubMed] [Google Scholar]; b Brunschwig B. S.; Creutz C.; Sutin N. Optical transitions of symmetrical mixed-valence systems in the Class II–III transition regime. Chem. Soc. Rev. 2002, 31, 168–184. 10.1039/b008034i. [DOI] [PubMed] [Google Scholar]
  13. a Allen G. C.; Hush N. S. Intervalence-Transfer Absorption. Part 1. Qualitative Evidence for Intervalence-Transfer Absorption in Inorganic Systems in Solution and in the Solid State. Prog. Inorg. Chem. 1967, 8, 357–389. 10.1002/9780470166093.ch6. [DOI] [Google Scholar]; b Hush N. S. Intervalence-Transfer Absorption. Part 2. Theoretical Considerations and Spectroscopic Data. Prog. Inorg. Chem. 1967, 8, 391–444. 10.1002/9780470166093.ch7. [DOI] [Google Scholar]; c D’Alessandro D. M.; Keene F. R. Current trends and future challenges in the experimental, theoretical and computational analysis of intervalence charge transfer (IVCT) transitions. Chem. Soc. Rev. 2006, 35, 424–440. 10.1039/b514590m. [DOI] [PubMed] [Google Scholar]
  14. Winter R. F. Half-Wave Potential Splittings ΔE1/2 as a Measure of Electronic Coupling in Mixed-Valent Systems: Triumphs and Defeats. Organometallics 2014, 33, 4517–4536. 10.1021/om500029x. [DOI] [Google Scholar]
  15. Cancela L.; Esteruelas M. A.; López A. M.; Oliván M.; Oñate E.; San-Torcuato A.; Vélez A. Osmium- and Iridium-Promoted C-H Bond Activation of 2,2’-Bipyridines and Related Heterocycles: Kinetic and Thermodynamic Preferences. Organometallics 2020, 39, 2102–2115. 10.1021/acs.organomet.0c00156. [DOI] [Google Scholar]
  16. See, for example:; a Bolaño T.; Esteruelas M. A.; Gay M. P.; Oñate E.; Pastor I. M.; Yus M. An Acyl-NHC Osmium Cooperative System: Coordination of Small Molecules and Heterolytic B-H and O-H Bond Activation. Organometallics 2015, 34, 3902–3908. 10.1021/acs.organomet.5b00418. [DOI] [Google Scholar]; b Eguillor B.; Esteruelas M. A.; Lezáun V.; Oliván M.; Oñate E.; Tsai J.-Y.; Xia C. A Capped Octahedral MHC6 Compound of a Platinum Group Metal. Chem. - Eur. J. 2016, 22, 9106–9110. 10.1002/chem.201601729. [DOI] [PubMed] [Google Scholar]; c Eguillor B.; Esteruelas M. A.; Lezáun V.; Oliván M.; Oñate E. Elongated Dihydrogen versus Compressed Dihydride in Osmium Complexes. Chem. - Eur. J. 2017, 23, 1526–1530. 10.1002/chem.201605843. [DOI] [PubMed] [Google Scholar]; d Babón J. C.; Esteruelas M. A.; Fernández I.; López A. M.; Oñate E. Redox-Assisted Osmium-Promoted C-C Bond Activation of Alkylnitriles. Organometallics 2018, 37, 2014–2017. 10.1021/acs.organomet.8b00326. [DOI] [Google Scholar]; e Babón J. C.; Esteruelas M. A.; Fernández I.; López A. M.; Oñate E. Evidence for a Bis(Elongated σ)-Dihydrideborate Coordinated to Osmium. Inorg. Chem. 2018, 57, 4482–4491. 10.1021/acs.inorgchem.8b00155. [DOI] [PubMed] [Google Scholar]; f Valencia M.; Merinero A. D.; Lorenzo-Aparicio C.; Gómez-Gallego M.; Sierra M. A.; Eguillor B.; Esteruelas M. A.; Oliván M.; Oñate E. Osmium-Promoted σ-Bond Activation Reactions on Nucleosides. Organometallics 2020, 39, 312–323. 10.1021/acs.organomet.9b00693. [DOI] [Google Scholar]
  17. a Skapski A. C.; Sutcliffe V. F.; Young G. B. Roll-over” 3-Metallation of Co-Ordinated 2,2′-Bipyridyl in the Thermal Rearrangement of Diary(Bipyridyl)Platinum(II) Complexes: Molecular Structure of (μ-Bidyl)[PtPh(Butpy)]2. J. Chem. Soc., Chem. Commun. 1985, 0, 609–611. 10.1039/C39850000609. [DOI] [Google Scholar]; b Zucca A.; Doppiu A.; Cinellu M. A.; Stoccoro S.; Minghetti G.; Manassero M. Multiple C–H Bond Activation. Threefold-Deprotonated 6-Phenyl-2,2’-Bipyridine as a Bridging Ligand in Dinuclear Platinum(II) Derivatives. Organometallics 2002, 21, 783–785. 10.1021/om010913h. [DOI] [Google Scholar]; c Petretto G. L.; Rourke J. P.; Maidich L.; Stoccoro S.; Cinellu M. A.; Minghetti G.; Clarkson G. J.; Zucca A. Heterobimetallic Rollover Derivatives. Organometallics 2012, 31, 2971–2977. 10.1021/om200660a. [DOI] [Google Scholar]; d Paziresh S.; Babadi Aghakhanpour R.; Rashidi M.; Nabavizadeh S. M. Simple tuning of the luminescence properties of the double rollover cycloplatinated(II) structure by halide ligands. New J. Chem. 2018, 42, 1337–1346. 10.1039/C7NJ03817H. [DOI] [Google Scholar]; e Leist M.; Kerner C.; Ghoochany L. T.; Farsadpour S.; Fizia A.; Neu J. P.; Schön F.; Sun Y.; Oelkers B.; Lang J.; Menges F.; Niedner-Schatteburg G.; Salih K. S. M.; Thiel W. R. Roll-over cyclometalation: A versatile tool to enhance the catalytic activity of transition metal complexes. J. Organomet. Chem. 2018, 863, 30–43. 10.1016/j.jorganchem.2018.03.022. [DOI] [Google Scholar]; f Paziresh S.; Babadi Aghakhanpour R.; Fuertes S.; Sicilia V.; Niroomand Hosseini F.; Nabavizadeh S. M. A double rollover cycloplatinated(II) skeleton: a versatile platform for tuning emission by chelating and non-chelating ancillary ligand systems. Dalton Trans. 2019, 48, 5713–5724. 10.1039/C9DT00807A. [DOI] [PubMed] [Google Scholar]
  18. See, for example:; a Esteruelas M. A.; García-Raboso J.; Oliván M.; Oñate E. N-H and N-C Bond Activation of Pyrimidinic Nucleobases and Nucleosides Promoted by an Osmium Polyhydride. Inorg. Chem. 2012, 51, 5975–5984. 10.1021/ic300639j. [DOI] [PubMed] [Google Scholar]; b Casarrubios L.; Esteruelas M. A.; Larramona C.; Muntaner J. G.; Oliván M.; Oñate E.; Sierra M. A. Chelated Assisted Metal-Mediated N-H Bond Activation of β-Lactams: Preparation of Irida-, Rhoda-, Osma-, and Ruthenatrinems. Organometallics 2014, 33, 1820–1833. 10.1021/om500162m. [DOI] [Google Scholar]; c Casarrubios L.; Esteruelas M. A.; Larramona C.; Lledós A.; Muntaner J. G.; Oñate E.; Ortuño M. A.; Sierra M. A. Mechanistic Insight into the Facilitation of β-Lactam Fragmentation through Metal Assistance. Chem. - Eur. J. 2015, 21, 16781–16785. 10.1002/chem.201503595. [DOI] [PubMed] [Google Scholar]; d Esteruelas M. A.; Larramona C.; Oñate E. Osmium-Mediated Direct C-H Bond Activation at the 8-Position of Quinolines. Organometallics 2016, 35, 1597–1600. 10.1021/acs.organomet.6b00264. [DOI] [Google Scholar]; e Esteruelas M. A.; Lezáun V.; Martínez A.; Oliván M.; Oñate E. Osmium Hydride Acetylacetonate Complexes and Their Application in Acceptorless Dehydrogenative Coupling of Alcohols and Amines and for the Dehydrogenation of Cyclic Amines. Organometallics 2017, 36, 2996–3004. 10.1021/acs.organomet.7b00521. [DOI] [Google Scholar]; f Buil M. L.; Esteruelas M. A.; Gay M. P.; Gómez-Gallego M.; Nicasio A. I.; Oñate E.; Santiago A.; Sierra M. A. Osmium Catalysts for Acceptorless and Base-Free Dehydrogenation of Alcohols and Amines: Unusual Coordination Modes of a BPI Anion. Organometallics 2018, 37, 603–617. 10.1021/acs.organomet.7b00906. [DOI] [Google Scholar]; g Babón J. C.; Esteruelas M. A.; Fernández I.; López A. M.; Oñate E. Reduction of Benzonitriles via Osmium-Azavinylidene Intermediates Bearing Nucleophilic and Electrophilic Centers. Inorg. Chem. 2019, 58, 8673–8684. 10.1021/acs.inorgchem.9b01013. [DOI] [PubMed] [Google Scholar]
  19. Alabau R. G.; Esteruelas M. A.; Oliván M.; Oñate E. Preparation of Phosphorescent Osmium(IV) Complexes with N,N’,C- and C,N,C’- Pincer Ligands. Organometallics 2017, 36, 1848–1859. 10.1021/acs.organomet.7b00193. [DOI] [Google Scholar]
  20. See, for example:; a Chou P.-T.; Chi Y. Osmium- and Ruthenium-Based Phosphorescent Materials: Design, Photophysics, and Utilization in OLED Fabrication. Eur. J. Inorg. Chem. 2006, 2006, 3319–3332. 10.1002/ejic.200600364. [DOI] [Google Scholar]; b Chi Y.; Chou P.-T. Contemporary progresses on neutral, highly emissive Os(II) and Ru(II) complexes. Chem. Soc. Rev. 2007, 36, 1421–1431. 10.1039/b608951h. [DOI] [PubMed] [Google Scholar]; c Alabau R. G.; Eguillor B.; Esler J.; Esteruelas M. A.; Oliván M.; Oñate E.; Tsai J.-Y.; Xia C. CCC-Pincer-NHC Osmium Complexes: New Types of Blue-Green Emissive Neutral Compounds for Organic Light-Emitting Devices (OLEDs). Organometallics 2014, 33, 5582–5596. 10.1021/om500905t. [DOI] [Google Scholar]; d Alabau R. G.; Esteruelas M. A.; Oliván M.; Oñate E.; Palacios A. U.; Tsai J.-Y.; Xia Osmium(II) Complexes Containing a Dianionic CCCC-Donor Tetradentate Ligand. Organometallics 2016, 35, 3981–3995. 10.1021/acs.organomet.6b00776. [DOI] [Google Scholar]
  21. See, for example:; a Henwood A. F.; Zysman-Colman E. Lessons learned in tuning the optoelectronic properties of phosphorescent iridium(III) complexes. Chem. Commun. 2017, 53, 807–826. 10.1039/C6CC06729H. [DOI] [PubMed] [Google Scholar]; b Li T.-Y.; Wu J.; Wu Z.-G.; Zheng Y.-X.; Zuo J.-L.; Pan Y. Rational design of phosphorescent iridium(III) complexes for emission color tunability and their applications in OLEDs. Coord. Chem. Rev. 2018, 374, 55–92. 10.1016/j.ccr.2018.06.014. [DOI] [Google Scholar]; c Lee S.; Han W.-S. Cyclometalated Ir(III) complexes towards blue-emissive dopant for organic light-emitting diodes: fundamentals of photophysics and designing strategies. Inorg. Chem. Front. 2020, 7, 2396–2422. 10.1039/D0QI00001A. [DOI] [Google Scholar]; d Bonfiglio A.; Mauro M. Phosphorescent Tris-Bidentate IrIII Complexes with N-Heterocyclic Carbene Scaffolds: Structural Diversity and Optical Properties. Eur. J. Inorg. Chem. 2020, 2020, 3427–3442. 10.1002/ejic.202000509. [DOI] [Google Scholar]
  22. See, for example:; a McGuire R.; McGuire M. C.; McMillin D. R. Platinum(II) polypyridines: A tale of two axes. Coord. Chem. Rev. 2010, 254, 2574–2583. 10.1016/j.ccr.2010.04.013. [DOI] [Google Scholar]; b Li K.; Ming Tong G. S.; Wan Q.; Cheng G.; Tong W. Y.; Ang W. H.; Kwong W. L.; Che C. M. Highly phosphorescent platinum(II) emitters: photophysics, materials and biological applications. Chem. Sci. 2016, 7, 1653–1673. 10.1039/C5SC03766B. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Cebrián C.; Mauro M. Recent Advances in Phosphorescent Platinum Complexes for Organic Light-emitting Diodes. Beilstein J. Org. Chem. 2018, 14, 1459–1481. 10.3762/bjoc.14.124. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Haque A.; Xu L.; Al-Balushi R. A.; Al-Suti M. K.; Ilmi R.; Guo Z.; Khan M. S.; Wong W.-Y.; Raithby P. R. Cyclometallated tridentate platinum(II) arylacetylide complexes: old wine in new bottles. Chem. Soc. Rev. 2019, 48, 5547–5563. 10.1039/C8CS00620B. [DOI] [PubMed] [Google Scholar]
  23. a Crespo O.; Eguillor B.; Esteruelas M. A.; Fernández I.; García-Raboso J.; Gómez-Gallego M.; Martín-Ortiz M.; Oliván M.; Sierra M. A. Synthesis and characterisation of [6]-azaosmahelicenes: the first d4-heterometallahelicenes. Chem. Commun. 2012, 48, 5328–5330. 10.1039/c2cc30356f. [DOI] [PubMed] [Google Scholar]; b Eguillor B.; Esteruelas M. A.; Fernández I.; Gómez-Gallego M.; Lledós A.; Martín-Ortiz M.; Oliván M.; Oñate E.; Sierra M. A. Azole Assisted C-H Bond Activation Promoted by an Osmium-Polyhydride: Discerning between N and NH. Organometallics 2015, 34, 1898–1910. 10.1021/acs.organomet.5b00174. [DOI] [Google Scholar]; c Castro-Rodrigo R.; Esteruelas M. A.; Gómez-Bautista D.; Lezáun V.; López A. M.; Oliván M.; Oñate E. Influence of the Bite Angle of Dianionic C,N,C-Pincer Ligands on the Chemical and Photophysical Properties of Iridium(III) and Osmium(IV) Hydride Complexes. Organometallics 2019, 38, 3707–3718. 10.1021/acs.organomet.9b00466. [DOI] [Google Scholar]
  24. D’Alessandro D. M.; Keene F. R. A cautionary warning on the use of electrochemical measurements to calculate comproportionation constants for mixed-valence compounds. Dalton. Trans. 2004, 3950–3954. 10.1039/b413980a. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

ic0c03680_si_001.pdf (12.2MB, pdf)
ic0c03680_si_002.xyz (317.9KB, xyz)

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

RESOURCES