Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2022 Aug 31;41(17):2513–2524. doi: 10.1021/acs.organomet.2c00338

Reactions of an Osmium–Hexahydride Complex with 2-Butyne and 3-Hexyne and Their Performance in the Migratory Hydroboration of Aliphatic Internal Alkynes

Juan C Babón 1, Miguel A Esteruelas 1,*, Ana M López 1, Enrique Oñate 1
PMCID: PMC9969483  PMID: 36864948

Abstract

graphic file with name om2c00338_0017.jpg

Reactions of the hexahydride OsH6(PiPr3)2 (1) with 2-butyne and 3-hexyne and the behavior of the resulting species toward pinacolborane (pinBH) have been investigated in the search for new hydroboration processes. Complex 1 reacts with 2-butyne to give 1-butene and the osmacyclopropene OsH22-C2Me2)(PiPr3)2 (2). In toluene, at 80 °C, the coordinated hydrocarbon isomerizes into a η4-butenediyl form to afford OsH24-CH2CHCHCH2)(PiPr3)2 (3). Isotopic labeling experiments indicate that the isomerization involves Me-to-COs hydrogen 1,2-shifts, which take place through the metal. The reaction of 1 with 3-hexyne gives 1-hexene and OsH22-C2Et2)(PiPr3)2 (4). Similarly to 2, complex 4 evolves to η4-butenediyl derivatives OsH24-CH2CHCHCHEt)(PiPr3)2 (5) and OsH24-MeCHCHCHCHMe)(PiPr3)2 (6). In the presence of pinBH, complex 2 generates 2-pinacolboryl-1-butene and OsH{κ2-H,H-(H2Bpin)}(η2-HBpin)(PiPr3)2 (7). According to the formation of the borylated olefin, complex 2 is a catalyst precursor for the migratory hydroboration of 2-butyne and 3-hexyne to 2-pinacolboryl-1-butene and 4-pinacolboryl-1-hexene. During the hydroboration, complex 7 is the main osmium species. The hexahydride 1 also acts as a catalyst precursor, but it requires an induction period that causes the loss of 2 equiv of alkyne per equiv of osmium.

Introduction

Understanding the behavior of transition-metal polyhydride complexes is one of the challenges in chemistry. The presence of metal–hydrogen bonds of different nature offers interesting chemical opportunities,1 as is evident in their ability to activate one of the widest ranges of σ-bonds.2 This property allows them to be relevant in a variety of fields including materials science,3 energy and environment,4 or organic synthesis based on metal-mediated catalysis.5 Thus, for example, this type of compounds are catalysts or catalyst precursors for some 40 different classes of organic reactions; the vast majority of them involve some σ-bond activation elemental step, in accordance with their tendency to activate σ-bonds.

The main fact explaining such versatility is probably the variety of roles that the coordinate hydrogen atoms can play during the catalysis. They can be transferred as a proton6 or hydride,7 undergo reductive elimination with other co-ligand to generate coordination vacancies,8 insert unsaturated organic molecules to afford organic ligands with a rich reactivity,9 promote C–H bond heterolytic activation acting as an internal Brønsted base,2 or even cooperate with the metal in the coordination of acidic ligands such as boranes.10 The latter are the main reagents in hydroboration and borylation reactions, processes of great current relevance because of the synthetic importance of organoboranes in organic chemistry.11

Several polyhydride complexes of rhenium and metals of the iron triad promote the hydroboration of olefins,12 alkynes,13 nitriles,14 N-heterocycles,15 and CO2.16 In addition, trihydride–iridium(III) complex IrH33-P,O,P-[xant(PiPr2)2]} (xant(PiPr2)2 = 9,9-dimethyl-4,5-bis(diisopropylphosphino)xanthene) catalyzes the borylation of arenes.17 Alkyne hydroboration is of special interest as is the most straightforward procedure to prepare useful alkenylborane synthetic intermediates. Polyhydrides for this catalysis have focused on the use of dihydride-(Kubas type-dihydrogen)-iron(II) and -ruthenium(II) complexes for the Z-hydroboration of terminal alkynes with pinacolborane (pinBH), while avoiding internal alkynes although the hydroboration of these substrates is attracting great attention with other families of catalysts.18

Metal-catalyzed hydroboration of internal alkynes can produce four different alkenylborane products, resulting from syn- and anti-BH-addition (a19 and b20 in Scheme 1). The formation of one or the other depends upon the nature of the catalyst (Scheme 2). The syn-products are the most frequent and are particularly favored when the catalyst carries a hydride ligand, which provides metal-(E-alkenyl) intermediates. The reaction of the latter with the borane completes the catalysis (a in Scheme 2). In contrast, the anti-products are common for catalysts with boryl ligands. Insertion of the alkyne in the metal–boron bond affords metal-(E-borylalkenyl) intermediates. A boryl group at the Cβ atom of the C–C double bond favors the E-to-Z isomerization of the alkenyl moiety,21 which is the key to the appearance of the anti-products.22 The boryl group lowers the activation energy to the formation of the metalacyclopropene responsible for the isomerization.21 Once Z-stereochemistry is achieved at the C–C double bond, reductive elimination involving the hydride and borylalkenyl ligands leads to the anti-products (b in Scheme 2). In addition, it has been recently observed that 3-hexyne and 4-octyne undergo rhodium-mediated dehydrogenative borylation–hydroboration with B2pin2, to give equimolecular mixtures of conjugated boryldienes and borylolefins (c in Scheme 1).23 The mixtures, which also contain the respective E and Z isomers, result from the addition of the B–B bond of the diborane to different molecules of alkyne and the hydride transfer from one to the other. Complexes Rh(Bpin){κ3-P,O,P-[xant(PiPr2)2]} and RhH{κ3-P,O,P-[xant(PiPr2)2]} collaborate to perform both borylations in a sequential and cyclic manner (c in Scheme 2). The rhodium(I)-boryl derivative promotes the stoichiometric dehydrogenative borylation to afford mixtures of E and Z boryldiene isomers and the rhodium(I)–hydride complex. The latter is responsible for the stoichiometric hydroboration, which yields the E and Z borylolefins and regenerates the rhodium–boryl compound to initiate the cycle again.

Scheme 1. Different Products Resulting from Hydroboration and Dehydrogenative Borylation of Internal Alkynes.

Scheme 1

Scheme 2. Mechanisms for the Hydroboration and Dehydrogenative Borylation of Internal Alkynes.

Scheme 2

Hexahydride complex OsH6(PiPr3)2 is a prominent member within the family of polyhydride derivatives,24 with a rich stoichiometric and catalytic reactivity and use as starting material to prepare compounds of interest in material science. It activates a vast range of σ-bonds,2,25 coordinates boranes stabilizing different bonding modes,26 promotes uncommon reactions as the metathesis between E–C(spn) and H–C(sp3) σ-bonds (E = Si, Ge; n = 2, 3)27 is a precursor of osmium(II) and osmium(IV) phosphorescent emitters,28 and catalyzes a variety of organic transformations such as hydrogenation of nitriles to symmetrical and asymmetrical secondary amines,29 hydrogen transfer from 2-propanol to unsaturated organic substrates,24 dihydroboration of nitriles,14 deuteration of pyridines,30 hydration of nitriles,31 Tishchenko dimerization of cyclohexanecarboxaldehyde and benzaldehyde, and aldol-Tishchenko trimerization of isobutyraldehyde.32 Our interest in taking a further step in understanding the behavior of this fascinating polyhydride prompted us to study its reactivity toward internal alkynes such as 2-butyne and 3-hexyne, in the search of new hydroboration reactions. This paper describes the research steps taken to discover the migratory hydroboration of both alkynes (d in Scheme 1).

Results and Discussion

Metalacyclopropene and Butenediyl Compounds

Treatment of polyhydride complex OsH6(PiPr3)2 (1) solutions, in toluene, with 2 equiv of 2-butyne, at 50 °C, for 18 h produces the release of 1 equiv of H2, the migratory hydrogenation33 of 1 equiv of alkyne to 1-butene, and the quantitative formation (according to the 31P{1H} NMR spectrum of the reaction crude) of the dihydride derivative OsH22-C2Me2)(PiPr3)2 (2), which was isolated as an orange solid in 78% yield (Scheme 3) and characterized by X-ray diffraction analysis.

Scheme 3. Reaction of 1 with 2-Butyne.

Scheme 3

The structure (Figure 1) resembles that of the known complex OsH2Cl2(PiPr3)224 with the CMe units of the hydrocarbon occupying the positions of the chloride ligands. Six-coordinate structures that deviate significantly from the typical octahedron are characteristic for complexes of d4-ions bearing a π-donor ligand, in particular for osmium(IV) hydride derivatives.34 The distortion destabilizes a half-occupied orbital of the t2g set but stabilizes the other one, resulting in a diamagnetic species. The process produces a partial cancelation of the electron deficiency at the metal center, which receives additional electron density from the π-donor ligand via the corresponding π-bond. Such disposition in the case of 2 indicates that the hydrocarbon, which acts as a 4e donor ligand,35 undergoes the oxidative addition of one of its π-bonds to the metal center, whereas the other works to cancel the metal electron deficiency. In agreement with the oxidative addition of one of the π-bonds of the alkyne, the C(1)–C(1) distance of 1.308(7) Å and the C(1)–C(1)–C(2) angle of 134.4(2)° (1.34 Å and 136.3° in the density functional theory (DFT)-optimized structure) are consistent with the formation of a metalacyclopropene. The coordinated carbon atoms give rise to a triplet (2JC–P = 5.9 Hz) at 168.7 ppm in the 13C{1H} NMR spectrum, in toluene-d8, at room temperature. The most notable resonance in the 1H NMR spectrum is a triplet at −19.79 ppm, with an H–P coupling constant of 33.1 Hz, corresponding to the hydride ligands. The 31P{1H} NMR spectrum contains a singlet at 48.9 ppm.

Figure 1.

Figure 1

Molecular structure of 2 with ellipsoids at the 50% probability level. Hydrogen atoms are omitted for clarity (except for hydride ligands). Selected bond distances (Å) and angles (°) for the X-ray structure and DFT-optimized (in square brackets): Os–C(1) = 1.985(3) [2.01], Os–P(1) = 2.3155(8) [2.37], C(1)–C(1) = 1.308(8) [1.34], C(1)–C(2) = 1.501(5) [1.49]; P(1)–Os–P(1) = 123.99(4) [119.8], H(01)–Os–H(01) = 106(3) [112.4], C(1)–C(1)–C(2) = 134.1(2) [136.3].

Complex 2 is quantitatively transformed into the butenediyl isomer OsH24-CH2CHCHCH2)(PiPr3)2 (3) after 24 h, at 80 °C, in toluene. The isomerization involves osmium-mediated 1,2-hydrogen shifts from the methyl substituents to the coordinated atoms of the metalacyclopropene unit. In agreement with the participation of the metal in the process, a complete deuterium distribution between the metal center and the butenediyl positions was observed in the deuterated 3d2 species, when the dideuteride derivative OsD22-C2Me2)(PiPr3)2 (2d2) was used as starting point (Scheme 4).

Scheme 4. Isomerization of 2 into 3.

Scheme 4

The isomerization can be rationalized according to Scheme 5. The initial migration of one of the hydride ligands to a coordinated carbon atom should give the E-alkenyl intermediate a, which could undergo E-to-Z isomerization of the alkenyl group to afford b. The subsequent C–H bond activation of the methyl substituent at the Cβ-atom of the C–C double bond disposed syn to the metal center should lead to the metalacycle c, which would evolve into d by reductive migration of one of the hydride ligands to the C(sp2)-metalated carbon atom. Thus, the equilibrium between allyl species could convert d into f via e. A β-hydrogen elimination on the methyl substituent at the σ-allyl ligand should finally yield 3.

Scheme 5. Mechanism for the Isomerization of 2 into 3.

Scheme 5

Complex 3 was isolated as a white solid in 83% yield and characterized by X-ray diffraction analysis. The structure (Figure 2) proves the isomerization of the hydrocarbon. The arrangement of ligands around the metal center can be described as a four-legged piano stool, with the butenediyl at the seat, whereas the hydride and phosphine ligands occupy the legs in an alternated manner. This disposition is usual in cations of the class [Os(η5-C5R5)H2(PR3)2]+36 and suggests a +4 oxidation state for the metal center. In accordance with these species, the P(1)–Os–P(2) and H(01)–Os–H(02) angles are 119.22(7) and 110(6)° (119.2 and 120° in the DFT-optimized structure), respectively. The butenediyl coordinates with Os–C distances in the range 2.192(12)–2.235(11) Å (2.21–2.27 Å in the optimized structure). The butenediyl C–C distances of 1.43(2)–1.46(2) Å (1.43–1.44 in the DFT-optimized structure) are essentially the same and point out a very light partial double-bond character of the bonds between the carbons. In this context, it should be noted that the delocalized electron density between the four carbon atoms is that corresponding to one bond (2e). In toluene-d8, at room temperature, the butenediyl rotates around an osmium-butenediyl axis as revealed by the 31P{1H} NMR spectrum, which displays a singlet at 37.9 ppm for the inequivalent phosphines. In the 1H NMR spectrum, the resonance corresponding to the hydride ligands appears at −13.65 ppm as a triplet with an H–P coupling constant of 32.0 Hz, which is consistent with the relative cisoid disposition of the hydride and phosphine ligands in the four legs of the stool. Characteristic resonances of the butenediyl in this spectrum are three signals at 4.53 (CH), 2.07, and −0.52 (CH2) ppm, which fit with resonances at 68.4 (CH) and 25.5 (CH2) in the 13C{1H} NMR spectrum.

Figure 2.

Figure 2

Molecular structure of 3 with ellipsoids at the 50% probability level. Hydrogen atoms are omitted for clarity (except for hydride ligands). Selected bond distances (Å) and angles (°) for the X-ray structure and DFT-optimized (in square brackets): Os–C(1) = 2.235(11) [2.27], Os–C(2) = 2.197(12) [2.21], Os–C(3) = 2.192(12) [2.21], Os–C(4) = 2.235(11) [2.27], Os–P(1) = 2.2535(15) [2.33], Os–P(2) = 2.3984(15) [2.37], Os–H(01) = 1.590(10) [1.64], Os–H(02) = 1.592(10) [1.64], C(1)–C(2) = 1.43(2) [1.43], C(2)–C(3) = 1.43(2) [1.43], C(3)–C(4) = 1.46(2) [1.44]; P(1)–Os–P(2) = 119.22(7) [119.2], H(01)–Os–H(02) = 110(6) [120.0], C(1)–C(2)–C(3) = 116.2(14) [116.9], C(2)–C(3)–C(4) = 118(2) [117.4].

3-Hexyne displays similar behavior to 2-butyne; it undergoes migratory hydrogenation to 1-hexene and coordinates to the resulting metal center to afford the counterpart metalacyclopropene complex OsH22-C2Et2)(PiPr3)2 (4), although the hydrocarbon of the latter appears to have a smaller activation energy for the isomerization into butenediyl than in 2. The formation rate of 4 and the rate of isomerization are comparable. As a consequence, the isomerization starts before the complete formation of 4. Furthermore, the isomerization affords two products, the ethylbutenediyl derivative OsH24-CH2CHCHCHEt)(PiPr3)2 (5) and the dimethylbutenediyl species OsH24-MeCHCHCHCHMe)(PiPr3)2 (6), the first of them being the main one in a 9:1 molar ratio (Scheme 6). The formation of both butenediyls takes place by metal-mediated ethyl-to-COs 1,3-hydrogen shifts. Complex 5 results from migrations from the CH2 and CH3 units of an ethyl substituent, while the generation of the minor isomer 6 involves the CH2 groups of both ethyl substituents.

Scheme 6. Reaction of 1 with 3-Hexyne.

Scheme 6

Complexes 4–6 were fully characterized by NMR spectroscopy. The spectra of 4 (Figures S14–S17) agree well with those of 2; the hydride resonance in the 1H spectrum appears at −18.14 (2JH–P = 33.1 Hz) ppm, whereas the 13C{1H} spectrum shows the signal due to the coordinated C atoms at 176.4 (2JC–P = 6.3 Hz) ppm. The 31P{1H} spectrum contains a singlet at 48.2 ppm. Spectra of the butenediyl derivative 5 (Figures S18–S25) are consistent with the asymmetry introduced in the molecule by the ethyl substituent of the butenediyl, which converts both hydrides and phosphines into inequivalent ligands. Furthermore, the substituent makes the rotation of the butenediyl around the osmium-butenediyl axis difficult. Thus, 1H and 31P{1H} spectra are temperature-dependent. At 233 K, the hydride ligands generate two doublets of doublets (2JH–P = 36.6 and 27.3 Hz) at −13.25 and −14.40 ppm, in the 1H, whereas an AB spin system (JAB = 86 Hz, Δν = 979 Hz) centered at 28.7 ppm is observed in the 31P{1H}. In contrast to 5, only the phosphine ligands are inequivalent in 6, whereas the presence of two substituents in the butenediyl prevents its rotation, even at room temperature. Thus, the 1H shows a doublet of doublets (2JH–P = 36.8 and 28.2 Hz) at −14.11 ppm, due to the equivalent hydrides (Figure S20), whereas the inequivalent phosphines give rise to an AB spin system (JAB = 89 Hz, Δν = 1145 Hz) at 27.0 ppm in the 31P{1H} (Figure S22).

Reaction of the Metalacyclopropene Compound 2 with Pinacolborane

Having studied the access of the alkynes to the metal center and analyzed the alkyne–osmium interaction, we decided to investigate the entry of the other component of the hydroboration process, pinBH. Thus, we carried out the reaction of the metalacyclopropene complex 2 with the borane. Treatment of solutions of this complex, in toluene, with 5 equiv of pinBH, at room temperature for 3 h leads to the dihydrideborate-osmium(II)-(elongated σ-borane) derivative OsH{κ2-H,H-(H2Bpin)}(η2-HBpin)(PiPr3)2 (7) and 2-pinacolboryl-1-butene, as a result of the migratory hydroboration of the coordinated hydrocarbon and the coordination of two borane molecules to the resulting dihydride fragment (Scheme 7). Complex 7 was isolated as a white solid in 71% and characterized by X-ray diffraction analysis.

Scheme 7. Reaction of 2 with Pinacolborane.

Scheme 7

The structure (Figure 3) proves the presence of two borane molecules at the metal center, coordinated in a different manner: dihydrideborate (B(1)) and elongated σ-borane (B(2)). The dihydrideborate ligand acts as κ2-H,H-chelate with a bite angle of 73(5)°; the separation between its boron and the osmium atom of 2.222(10) (2.24 Å in the DFT-optimized structure) compares well with those found in other crystallographically characterized related osmium complexes.37 The coordination of the B(2)–H(02) bond of the second borane molecule gives rise to a metal–(σ-bond) interaction weaker than those found in other elongated σ-borane complexes as proved by the B(2)–H(02) distance of 1.44(10) Å (1.48 Å in the DFT-optimized structure), which is between 0.1 and 0.2 Å shorter than the B–H distances reported for complexes OsHCl(η2-HBR2){κ3-P,O,P-[xant(PiPr2)2]} (1.60–1.69 Å)38 and Rh(η5-C5Me5)(Bpin)22-HBpin) (1.53(2) and 1.69(3) Å).39 The B(2)–H(02) bond length compares well with the B–H distance found in the iridium complex Ir{κ3-P,C,P-[C6H3-1,3-OPtBu2]}(η2-HBpin) (1.47(6) Å).40 The B–H bonds of the dihydrideborate and elongated σ-borane groups are located, together with a hydride ligand, in a perpendicular plane to an ideal direction defined by the mutually trans-disposed phosphine ligands and the osmium atom (P(1)–Os–P(2) = 160.69(6)°), in the expected octahedron for a six-coordinate d6-species.

Figure 3.

Figure 3

Molecular structure of 7 with ellipsoids at the 50% probability level. Hydrogen atoms are omitted for clarity (except for hydrides). Selected bond distances (Å) and angles (°) for the X-ray and DFT-optimized (in square brackets) structures: Os–B(1) = 2.222(10) [2.24], Os–B(2) = 2.132(9) [2.14], Os–H(01) = 1.587(10) [1.76], Os–H(02) = 1.585(10) [1.69], Os–H(03) = 1.585(10) [1.74], Os–H(04) = 1.587(10) [1.66], Os–P(1) = 2.3522(18) [2.36], Os–P(2) = 2.3624(18) [2.36], B(1)–H(01) = 1.28(9) [1.41], B(1)–H(03) = 1.39(10) [1.44], B(2)–H(02) = 1.44(10) [1.48]; P(1)–Os–P(2) = 160.69(6) [162.1], H(01)–Os–H(04) = 167(5) [172.6], H(01)–Os–H(03) = 73(5) [78.9], H(02)–Os–H(04) = 75(5) [72.6], H(03)–Os–H(04) = 95(5) [93.9], H(01)–Os–B(2) = 75(3) [72.3], H(02)–Os–B(2) = 42(4) [43.7].

Atoms in Molecules (AIM) analysis of 7 confirmed its dihydrideborate-osmium(II)-(elongated σ-borane) nature (Figure 4). In accordance with other dihydrideborate complexes,37c the four-membered cycle formed by the osmium atom and the H2B moiety displays two Os–H and two B–H bond critical points associated with one OsHHB ring critical point. The metal–(σ-bond) interaction shows the characteristic triangular topology of elongated σ-bonds acting as 2e donor ligands,38 which is defined by bond critical points located between the involving atoms, associated with bond paths running between them, and all complemented by a ring critical point. The coordination mode of the dihydrideborate group in 7 contrasts with that found in the dihydridecatecholborate counterpart derivative OsH(η3-H2Bcat)(η2-HBcat)(PiPr3)2, which has been described as bis(elongated σ) on the basis an AIM analysis and the energy decomposition analysis (EDA) method.26

Figure 4.

Figure 4

Contour line diagram ∇2ρ(r) for complex 7 in the OsH4B2 plane. The black lines connecting the atomic nuclei are the bond paths, while the small green and red spheres designate the corresponding bond and ring critical points, respectively.

The coordinated hydrogen atoms of the dihydrideborate and borane groups and the hydride ligand exchange their positions in solution. Thus, the four inequivalent nuclei give rise to only one resonance, at −10.65 ppm, in the 1H NMR spectrum, in toluene-d8, at room temperature. The exchange is thermally activated. As a consequence, the spectrum is temperature-dependent. About 233 K, the resonance decoalesces, and between 213 and 183 K, three signals at −9.44, −9.64, and −11.88 ppm are clearly observed. The exchange is also supported by the 11B spectrum at room temperature, which shows a broad resonance at 37.9 ppm for both borane and dihydrideborate. In contrast to the 1H spectrum, the 31P{1H} is temperature-invariant, displaying a singlet at 33.7 ppm in agreement with the equivalence of the phosphines.

The interaction between the σ-bond of the borane and the metal center is certainly weak. Thus, in spite of the saturated character of 7, it reacts with a weak Lewis base such as molecular hydrogen. The H2 molecule displaces the elongated σ-borane ligand from the metal coordination sphere to give the previously reported trihydride-osmium(IV)-dihydrideborate derivative OsH32-H,H-(H2Bpin)}(PiPr3)2 (8).26 This finding explains why complex 7 is inaccessible by direct reaction between the hexahydride 1 and pinBH and why complex 8 is the isolated species. The lability of the elongated σ-borane ligand was certainly a promising feature of 7, given the presence of an additional molecule of borane coordinated in the dihydrideborate form. Such lability stimulated our interest to know the potential performance of 7 in the alkyne hydroboration and prompted us to study the reaction between this bis-borylated complex and alkyne in excess. The addition of 5 equiv of 2-butyne to 7, in tolune-d8, at room temperature gives 2 equiv of 2-pinacolboryl-1-butene and regenerates 2 (Scheme 8).

Scheme 8. Reactions of 7 with Hydrogen or 2-Butyne.

Scheme 8

Catalytic Hydroboration of 2-Butyne and 3-Hexyne

The reaction of 2 with pinBH to give 2-pinacolboryl-1-butene and 7 (Scheme 7) and the reaction of the latter with 2-butyne to afford 2-pinacolboryl-1-butene again and regenerate 2 (Scheme 8) form a cycle. Accordingly, complex 2 catalyzes the migratory hydroboration of 2-butyne to 2-pinacolboryl-1-butene. The reaction was performed in toluene, at 60 °C, using a 5 mol % of catalyst and an alkyne/borane molar ratio of 1:1.5. The borylated olefin was selectively formed after 2 h (91%) and isolated in 82% yield (d in Scheme 1). Under the same conditions, 3-hexyne was transformed into a mixture of the migratory hydroboration product 4-pinacolboryl-1-hexene (85%, d in Scheme 1) and the syn-hydroboration product 3-pinacolboryl-3-hexene (14%, a in Scheme 1) after 3 h. The former was separated from the mixture by flash chromatography in silica gel, using hexane as eluent, and isolated pure in 77% yield.

Migratory hydrofunctionalization of internal alkynes catalyzed by transition-metal complexes is a scarcely explored class of reactions. It has been mainly focused on hydrosilylation and hydrogermylation processes promoted by cobalt-hydride compounds. Mechanistic details are even scarcer, with the reaction mechanisms being true black boxes.41

The 1H and 31P{1H} NMR spectra of the catalytic solutions revealed that complex 2 is rapidly and quantitatively transformed into 7, which is the main metallic species while alkyne and borane are present in solution. Its incorporation into the catalytic cycle takes place by means of the dissociation of a borane molecule. In agreement with this, the hydroboration rate decreases by increasing the borane concentration; the catalysis is inhibited for borane:alkyne molar ratios >10. Previous DFT calculations about the dissociation of molecular hydrogen from 8 revealed that the interaction hydride-borane affording the dihydridoborate group is broken in the resulting unsaturated five-coordinate intermediate, which was described as the dihydride-osmium(II)-(elongated σ-borane) species OsH22-HBpin)(PiPr3)2 (g).14 Therefore, this is the species formed as a result of the dissociation of a borane molecule from 7 (Scheme 9). The coordination of the alkyne to intermediate g should lead to h. Related compounds with a carbonyl group instead of an alkyne ligand have been experimentally observed in equilibrium with dihydrideborate-osmium(II)-hydride derivatives.37b Dihydride-osmium(II)-(elongated σ-borane) species like h, with a coordinated nitrile instead of a carbonyl group or an alkyne ligand, have been also proposed as key intermediates for the hydroboration of aliphatic nitriles, on the basis of DFT calculations.14 Intermediate h could be also formed by the addition of the borane to 2 and, since it contains both components of the hydroboration products, it should be the species responsible for the catalysis and thus the species connecting 2 and 7 (Scheme 9).

Scheme 9. Formation of Catalytic Active Species.

Scheme 9

The catalysis can be rationalized according to Scheme 10. The insertion of the coordinated alkyne into the Os–B bond of h should give metal-(E-borylalkenyl) intermediates i, which would evolve in a different manner depending upon the size of the substituents at the C–C double bond. When the substituent is methyl (the 2-butyne case), the E-to-Z isomerization of the borylalkenyl group takes place. Such isomerization should lead to j. The unsaturated character of the metal center of the latter and its syn-disposition to the β-methyl group would favor the C–H bond activation of this substituent to give k. Intermediate k could generate the σ-allyl derivative l by reductive elimination of the alkenyl moiety. Once formed l, it should be transformed into the σ-allyl counterpart n, via the π-allyl m. Thus, in the presence of the borane, the reductive elimination of the σ-allyl could yield 2-pinacolboryl-1-butene and regenerate g. The bigger steric hindrance of ethyl with respect to methyl appears to prevent the E-to-Z isomerization of the borylalkenyl group of the corresponding intermediate i. Thus, for the 3-hexyne case, the C–H bond activation of the methyl group of the ethyl substituent at the Cα atom of the borylalkenyl group would lead to o, which could evolve into p by reductive elimination of the borylalkenyl moiety. A subsequent β-hydrogen elimination should give the unsaturated tetrahydride q and 4-pinocolboryl-1,3-hexadiene. The reduction of one of the C–C double bonds of the diene by the tetrahydride complex would afford the hydroboration products and a dihydride-metal fragment that could regenerate g by coordination of the borane. The reduction of the internal olefinic bond should yield the main product, 4-pinacolboryl-1-hexene, the terminal olefin resulting from the migratory hydroboration, while the reduction of the terminal C–C double bond should lead to 3-pinacolboryl-3-hexene. The latter could be also formed by reductive elimination from i.

Scheme 10. Proposed Catalytic Cycles.

Scheme 10

In short, the formation of α-vinylborane or homoallylborane products depends on whether or not the borylalkenyl intermediate undergoes E-to-Z isomerization.

Having achieved the migratory hydroboration of 2-butyne and 3-hexyne and analyzed the formation of the products, we decided to address the direct use of the hexahydride complex 1 as a catalyst precursor. Indeed, this polyhydride is also an efficient catalyst precursor for the migratory hydroboration of both alkynes, the main difference with regard to the reactions performed with the cyclopropane complex 2 is the additional formation of the terminal olefin, resulting from the migratory hydrogenation of the alkyne. The formed amount of this product is about twice the percentage of moles of the catalytic precursor used in the reaction (about 10% of alkyne). Figure 5a shows the formation of the product of the migratory hydroboration of 3-hexyne, 4-pinacolboryl-1-hexene, as a function of time in the presence of 1 and 2, respectively. The formation of the borylated olefin in the presence of the hexahydride 1 displays an induction period, which is not observed when the precursor is the metalacyclopropene complex 2. However, the slope of the line defining the reaction course is almost identical in both cases after the induction time. This finding suggests that both precursors give rise to the same active species. To confirm this conclusion and to understand the induction period, we analyzed the 1H NMR spectra of the catalytic solutions generated from 1 and 2, in the high-field region, as a function of time (Figure 5b,c, respectively). As expected, the main osmium species is complex 7 in both cases. However, it is generated in a different way. Complex 2 is rapidly transformed into 7, according to Scheme 9, via the spectroscopically undetected intermediates g and h (Figure 5c); it is not necessary for any induction period. In contrast, hexahydride complex 1 is first converted into 8, which is transformed into 7 when the excess of hydrogen provided by 1 is consumed (Figure 5b). So, an induction period is necessary to eliminate two hydrogen molecules from the reaction medium.

Figure 5.

Figure 5

(a) Formation of 4-pinacolboryl-1-hexene by hydroboration of 3-hexyne (0.15 M) with pinBH (0.23 M) in the presence of 1 (blue •) and 2 (red •) (7.5 × 10–3 M), respectively, in toluene-d8 at 60 °C. (b) High-field region of the 1H NMR spectrum of the catalytic solutions using 1 as a catalyst precursor as a function of time. (c) High-field region of the 1H NMR spectrum of the catalytic solutions using 2 as catalyst precursor as a function of time.

In conclusion, hexahydride complex 1 is also an efficient catalyst precursor for the migratory hydroboration of internal alkynes but causes the loss of 2 equiv of alkyne per equiv of catalyst precursor, generating as a byproduct the olefin resulting from a migratory hydrogenation.

Concluding Remarks

This study has revealed that the hexahydride complex OsH6(PiPr3)2 induces the stoichiometric migratory hydrogenation of 2-butyne and 3-hexyne to afford the corresponding terminal olefins and a dihydride-osmium(II) fragment, which oxidatively adds one of the three bonds of the triple bond of a new alkyne molecule, to form formally unsaturated d4-dihydride-osmacyclopropene complexes. These compounds isomerize into (η4-butenediyl)-osmium(IV)-dihydride derivatives by hydrogen shifts from the substituents of the three-member ring to the carbon atoms of the ring. Isotopic labeling experiments indicate that the hydrogen shifts take place through the metal center and involve C–H bond activation processes.

Complexes dihydride-osmacyclopropene react with pinacolborane to produce α-vinylborane (2-butyne) or homoallylborane (3-hexyne) derivatives and the dihydrideborate-osmium(II)-(elongated σ-borane) compound OsH{κ2-H,H-(H2Bpin)}(η2-H-Bpin)(PiPr3)2. In agreement with the formation of the borylated olefins, these dihydride–osmacylopropene complexes promote the migratory hydroboration of 2-butyne to 2-pinacolboryl-1-butene and 3-hexyne to 4-pinacolboryl-1-hexene. During the hydroboration process, the dihydrideborate-osmium(II)-(elongated σ-borane) compound is the main osmium species. The hexahydride complex is also a catalyst precursor for these reactions, but it needs an induction period to reach the maximum activity that causes the loss of 2 equiv of alkyne per equiv of osmium.

In summary transition-metal polyhydride complexes are promising catalyst precursors for migratory hydrofunctionalization reactions of internal aliphatic alkynes due to their ability to activate σ-bonds, in particular C–H.

Experimental Section

General Information

All reactions were carried out with exclusion of air at an argon/vacuum manifold using standard Schlenk tube or glovebox techniques. Complexes OsH6(PiPr3)2 (1)24 and OsD6(PiPr3)2 (1d6)25i were prepared according to the published methods. Instrumental methods used for characterization, X-ray information, and computational details are given in the Supporting Information. Chemical shifts (in ppm) are referenced to residual solvent peaks (1H, 13C{1H}), external H3PO4 (31P{1H}), or BF3·OEt2 (11B). Coupling constants are given in hertz.

Preparation of OsH22-C2Me2)(PiPr3)2 (2)

2-Butyne (152 μL, 1.93 mmol) was added over a solution of complex 1 (500 mg, 0.96 mmol) in toluene (2 mL). The mixture was heated at 50 °C for 18 h. After this time, the 1H NMR spectrum of reaction crude in toluene-d8 showed the formation of 2 along with 1-butene. Then, it was concentrated to dryness to give an orange oil. The addition of pentane (2 mL) at −78 °C afforded an orange solid that was washed with cold pentane (2 × 2 mL). Orange single crystals suitable for X-ray diffraction analysis were obtained from a saturated solution of 2 in pentane at −30 °C. Yield: 425 mg (78%). Anal. Calcd. for C22H50OsP2: C, 46.62; H, 8.89. Found: C, 46.27; H, 9.39. HR-MS (electrospray): m/z calcd for C22H49OsP2 [M – H]+ 567.2919; found 567.2903. IR (ATR, cm–1): ν(Os–H) 2136. 1H NMR (300.13 MHz, C7D8, 298 K): δ 2.46 (s, 6H, CMe), 2.25 (m, 6H, CH iPr), 1.16 (dvt, 3JH–H = 7.1, N = 12.6, 36H, CH3iPr3), −19.79 (t, 2JH–P = 33.1, 2H, OsH2). 31P{1H} NMR (121.4 MHz, C7D8, 298 K): δ 48.9 (s; t under off-resonance conditions, 2JC–P = 32.8). 13C{1H} APT NMR (75.48 MHz, C7D8, 298 K): δ 168.7 (t, 2JC–P = 5.9, OsC), 30.3 (vt, N = 24.4, CH iPr), 20.8 (s, CH3iPr), 20.1 (CMe). T1(min) (ms, OsH, 300.13 MHz, C7D8, 203 K): 254 ± 5 (−19.79 ppm).

Preparation of OsH24-H2CCHCHCH2)(PiPr3)2 (3)

A solution of complex 2 (50 mg, 0.088 mmol) in toluene (1 mL) was heated at 80 °C for 24 h. Then, it was concentrated to dryness giving a white solid. Yield: 41.5 mg (83%). Colorless single crystals suitable for X-ray diffraction analysis were obtained from a saturated solution of 3 in pentane at −30 °C. Anal. Calcd. for C22H50OsP2: C, 46.62; H, 8.89. Found: C, 46.93; H, 9.21. HR-MS (electrospray): m/z calcd for C22H49OsP2 [M – H]+ 567.2919; found 567.2935. IR (ATR, cm–1): ν(Os–H) 2033. 1H NMR (300.13 MHz, C7D8, 298 K): δ 4.53 (br, 2H, HCCH), 2.07 (br, 2H, CH2), 1.96 (m, 6H, CH iPr), 1.09 (dvt, 3JH–H = 5.6, N = 12.7, 36H, CH3iPr3), −0.52 (br, 2H, CH2), −13.65 (t, 2JH–P = 32.0, 2H, OsH2). 31P{1H} NMR (121.4 MHz, C7D8, 298 K): δ 37.9 (s). 13C{1H} APT NMR (75.48 MHz, C7D8, 298 K): δ 68.4 (HCCH), 30.7 (vt, N = 26.2, CH iPr), 25.5 (CH2), 20.0 (s, CH3iPr). T1(min) (ms, OsH, 400.13 MHz, C7D8, 203 K): 356 ± 5 (−13.65 ppm).

Preparation of 3d2

2-Butyne (6.0 μL, 0.077 mmol) was added to two NMR tubes containing deuterated complex 1d6 (20 mg, 0.039 mmol) in toluene-d8 (0.5 mL). The tubes were heated at 50 °C for 18 h. After this time, their 1H and 31P{1H} NMR spectra showed the quantitative transformation of 1d6 into OsD22-C2Me2)(PiPr3)2 (2d2). Then, they were concentrated to dryness to give orange oils. The 1H NMR (300.13 MHz, toluene-d8, 298 K) data were identical to that reported for 2 except for the almost total disappearance of the signal at δ −19.79 (OsH). The two NMR tubes (A and B) with a solution of deuterated complex 2d2 were heated at 80 °C for 24 h. Tetrachloroethane (8.2 μL, 0.078 mmol) was added to tube A as the internal standard. The 1H NMR (300.13 MHz, toluene-d8, 298 K) data were identical to those of 3 except for a decrease in the intensity of the signals δ 4.53 (77%, HCCH), −0.52 (76%, CH2), −13.65 (79%, OsH2). The percentages indicate the amount of hydrogen atoms in those positions. Tube B was concentrated to dryness to give a colorless oil, which was checked in nondeuterated toluene. 2H NMR (61.42 MHz, toluene, 298 K): δ 4.47 (br, DCCD), 2.02 (br, CD2), −0.57 (br, CD2), −13.66 (br, OsD2).

Reaction of 1 with 3-Hexyne: Formation of OsH22-C2Et2)(PiPr3)2 (4)

3-Hexyne (43 μL, 0.38 mmol) was added over a solution of complex 1 (100 mg, 0.19 mmol) in toluene (2 mL). The mixture was heated at 50 °C for 18 h. After this time, the 1H and 31P{1H} NMR spectra of the reaction crude shows a mixture of complexes 4 and 5 in an 88:22 ratio along with 1-hexene and 3-hexene. The mixture was concentrated to dryness giving an orange oil, which was washed with cold methanol (3 × 2 mL) and dried under vacuum. Anal. Calcd. for C24H54OsP2: C, 48.46; H, 9.15. Found: C, 48.10; H, 9.01. HR-MS (electrospray): m/z calcd for C24H53OsP2 [M – H]+ 595.3233; found 595.3255. 1H NMR (300.13 MHz, C7D8, 298 K): δ 3.08 (q, 3JH–H = 7.4, 4H, CH2), 2.14 (m, 6H, CH iPr), 1.30 (t, 3JH–H = 7.4, 6H, CH3 Et), 1.16 (dvt, 3JH–H = 6.3, N = 12.6, 36H, CH3iPr3), −18.14 (t, 2JH–P = 33.1, 2H, OsH2). 31P{1H} NMR (121.4 MHz, C7D8, 298 K): δ 48.2 (s; t under off-resonance conditions, 2JH–P = 33.1). 13C{1H} APT NMR (75.48 MHz, C7D8, 298 K): δ 176.4 (t, 2JH–C = 6.3, OsC), 30.2 (s, CH2), 29.8 (vt, N = 23.6, CH iPr), 20.1 (s, CH3iPr), 13.36 (s, CH3 Et).

Isomerization of 4: Formation of OsH24-H2CCHCHCHEt)(PiPr3)2 (5) and OsH24-MeHCCHCHCHMe)(PiPr3)2 (6)

A mixture of complexes 4 and 5 in 88:22 molar ratio (30 mg, 0.05 mmol) was heated in toluene (0.5 mL) at 80 °C for 24 h. The mixture was concentrated to dryness obtaining a light-yellow oil. The oil was washed with cold pentane (2 × 1 mL) and dried under vacuo. 1H NMR spectra show a mixture of complexes 5 and 6 in a 9:1 molar ratio. Anal. Calcd. for C24H54OsP2: C, 48.46; H, 9.15. Found: C, 48.12; H, 8.88. HR-MS (electrospray): m/z calcd for C24H53OsP2 [M – H]+ 595.3233; found 595.3251. Spectroscopic data for 5: 1H NMR (300.13 MHz, C6D6, 293 K): δ 4.42 (m, 1H, CHCHEt), 4.40 (m, 1H, CH2CH), 2.31 (m, 1H, CH2 Et), 2.12–1.95 (br, 7H, 1H CH2CH and 6H CH iPr3), 1.68 (m, 1H, CH2 Et), 1.26 (t, 3JH–H = 7.4, 3H, CH3 Et), 1.12 (br, 36H, CH3iPr3), 0.23 (br, 1H, CHEt), −0.32 (br, 1H, CH2CH), −13.46 (br, 1H, OsH), −14.14 (br, 1H, OsH). 1H NMR (300.13 MHz, C7D8, 253 K): δ 4.35 (m, 1H, CHCHEt), 4.33 (m, 1H, CH2CH), 2.35 (m, 1H, CH2 Et), 2.14 (br, 3H, CH iPr3), 1.83 (br, 4H, 1H CH2CH and 3 CH iPr3), 1.64 (m, 1H, CH2 Et), 1.29 (t, 3JH–H = 7.3, 3H, CH3 Et), 1.20 (m, 18H, CH3iPr3), 1.01 (m, 18H, CH3iPr3), 0.14 (br, 1H, CHEt), −0.40 (br, 1H, CH2CH), −13.25 (dd, 2JH–P = 27.3, 36.6, 1H, OsH), −14.40 (dd, 2JH–P = 27.3, 36.6, 1H, OsH). 31P{1H} NMR (121.4 MHz, C7D8, 298 K): 28.7 (br AB system). 31P{1H} NMR (121.4 MHz, C7D8, 253 K): δ 28.7 (AB system, JAB = 86.5 Hz; Δυ = 979 Hz). 13C{1H} APT NMR (75.48 MHz, C6D6, 298 K): δ 73.2 (s, CHCHEt), 66.0 (s, CH2CH), 49.4 (s, CHEt), 31.5 (s, CH2 Et), 30.8 (br, CH iPr), 30.5 (br, CH iPr), 25.3 (s, CH2CH), 21.0 (s, CH3 Et), 20.2 (s, CH3iPr), 20.1 (s, CH3iPr). T1(min) (ms, OsH, 400.13 MHz, C7D8, 213 K): 288 ± 5 (−18.36 ppm), 288 ± 5 (−19.50). Selected spectroscopic data for 6: 1H NMR (300.13 MHz, C6D6, 298 K): δ 4.28 (m, 2H, CHCHMe), 1.90 (3H, CHMe overlapped with CH PiPr3), 0.23 (inferred from COSY spectrum, 2H, CHMe), −14.11 (dd, 2JH–P = 36.8, 28.2, 2H, OsH2). 31P{1H} NMR (121.4 MHz, C7D8, 298 K): δ 27.0 (AB system, JAB = 89, Hz; Δυ = 1145 Hz). 13C{1H} APT NMR (75.48 MHz, C6D6, 298 K): δ 71.0 (t, 2JC–P = 3.4, CHCHMe), 39.3 (dd, 2JC–P = 8.5, 6.1, CHMe), 21.7 (s, CHMe). T1(min) (ms, OsH, 400.13 MHz, C7D8, 213 K): 309 ± 5 (−19.15 ppm).

Preparation of OsH{κ2-H,H-(H2Bpin)}(η2-HBpin)(PiPr3)2 (7)

A solution of 2 (20 mg, 0.035 mmol) in toluene (0.5 mL) was treated with pinacolborane (25.4 μL, 0.175 mmol). The mixture was stirred at room temperature for 3 h (or 1 h at 60 °C), during which time the solution turned colorless and the 1H NMR spectrum of reaction crude in toluene-d8 showed the formation of 2 along with 1-butene. The reaction crude was concentrated to dryness affording a white solid. The solid was washed with cold pentane (0.5 mL, −72 °C) and dried under vacuum. Yield: 19 mg (71%). Colorless single crystals suitable for X-ray diffraction analysis were obtained by removing the solvent of a saturated solution of 7 in toluene at room temperature under vacuum. Anal. Calcd. for C30H70B2O4OsP2: C, 46.88; H, 9.18. Found: C, 47.27; H, 8.84. 1H NMR (300.13 MHz, C7D8, 298 K): δ 2.43 (m, 6H, CH iPr3), 1.28 (dvt, 3JH–H = 5.7, N = 12.8, 36H, CH3iPr3), 1.14 (s, 24H, CH3 Bpin), −10.65 (br, 4H, OsH4). 1H NMR (300.13 MHz, C7D8, 193 K): δ 2.43 (br, 6H, CH iPr3), 1.28 (br, 36H, CH3iPr3), 1.14 (s, 24H, CH3 Bpin), −9.44 (br, 1H, OsH), −9.64 (br t, 1H, OsH), −11.88 (br, 2H, OsH2).31P{1H} NMR (121.4 MHz, C7D8, 298 K): δ 33.7. 11B{1H} NMR (94.29 MHz, C7D8, 298 K): δ 37.9. T1(min) (ms, OsH, 400.13 MHz, C7D8, 273 K): 313 ± 5 (OsH4, −10.72 ppm).

Reaction of 7 with H2

A solution of complex 7 (10 mg, 0.013 mmol) in toluene-d8 (0.5 mL) was kept for 3 h under H2 atmosphere at room temperature. After that time, the 1H and 31P NMR spectra of the reaction mixture showed the quantitative transformation of 7 into OsH32-H,H-H2Bpin)(PiPr3)2 (8).

Reaction of 7 with 2-Butyne

2-Butyne (7 μL, 0.195 mmol) was added to a solution of 7 (30 mg, 0.039 mmol) in 0.5 mL of toluene-d8 at room temperature, and the mixture was stirred for 30 min. After this time, the 31P NMR showed the quantitative transformation of 7 into 2, whereas the 1H NMR showed also the formation of 2-pinacolboryl-1-butene.

Hydroboration of 2-Butyne and 3-Hexyne

The corresponding alkyne (0.18 mmol) was added over a solution of complex 2 (5 mg, 9.0 × 10–3 mmol) or complex 1 (4.7 mg, 9.0 × 10–3 mmol), pinacolborane (39 μL, 0.27 mmol), and mesytilene (25 μL, 0.18 mmol) as internal standard in toluene-d8 (0.5 mL). The mixture was heated at 60 °C and followed by 1H NMR spectra until the complete disappearance of the alkyne.

Isolation of Borylated Olefins

The corresponding alkyne (0.72 mmol) was added over a solution of complex 2 (20 mg, 0.036 mmol) and pinacolborane (156 μL, 1.08 mmol) in toluene (2 mL). The mixture was heated at 60 °C, 2 h for 2-butyne and 3 h for 3-hexyne. The solvent was evaporated to dryness under vacuum giving orange oils, which were purified by silica-gel flash chromatography using hexane as eluent. 2-Pinacolboryl-1-butene was isolated as a colorless oil in 82% yield, and 2-(hex-5-en-3-yl)-pinacolborane was isolated as a colorless oil in 77% yield.

Spectroscopic Data for 2-Pinacolboryl-1-butene

1H NMR (300.13 MHz, CDCl3, 298 K): δ 5.76 (d, 1H, 3JH–H = 3.4, =CH2), 5.62 (br, 1H, =CH2), 2.18 (q, 2H, 3JH–H = 7.5, CH2), 1.28 (s, 3H, CH3 Bpin), 1.03 (t, 3H, 3JH–H = 7.5, CH3). 13C{1H} APT NMR (75.48 MHz, CDCl3, 298 K): δ127.8 (=CH2), 83.3 (Cq Bpin), 28.2 (CH2), 24.7 (CH3 Bpin), 13.6 (CH3), (CBpin, not observed). 11B NMR (96.29 MHz, CDCl3, 298 K): δ 30.0 (Bpin). These spectroscopic data agree with the reported data.42

Spectroscopic Data for 4-Pinacolboryl-1-hexene

1H NMR (300.13 MHz, CDCl3, 298 K): δ 5.74 (ddt, 3JH–H = 17.0, 3JH–H = 10.1, 3JH–H = 6.9, 1H, CH2=CH), 4.94 (ddt, 3JH–H = 17.0, 3JH–H = 2.1, 4JH–H = 1.5, 1H, CH=CH2), 4.86 (ddt, 3JH–H = 10.1, 3JH–H = 2.1, 4JH–H = 1.0, 1H, CH=CH2), 2.07 (m, 3H, =CHCH2), 1.34 (m, 3H, CHBpinCH2CH3), 1.17 (s, 12H, CH3 Bpin), 0.84 (t, 3JH–H = 6.9, 3H, CH2CH3). 13C{1H} APT NMR (75.48 MHz, CDCl3, 298 K): δ 138.7 (−CH=CH2), 114.6 (−CH=CH2), 82.8 (Cq Bpin), 35.3 (=CHCH2), 24.8 (CH3 Bpin), 23.8 (CH2CH3), 13.5 (CH2CH3). 11B NMR (96.29 MHz, CDCl3, 298 K): δ 34.4 (Bpin). These spectroscopic data are in agreement with those reported for the closely related 4-pinacolboryl-1-nonene and 4-pinacolboryl-1-pentene.43

Acknowledgments

Financial support from the MICIN/AEI/10.13039/501100011033 of Spain (Projects PID2020-115286GB-I00 and RED2018-102387-T), Gobierno de Aragón (Group E06_20R and LMP23_21), 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.organomet.2c00338.

  • General information, crystallographic data, computational details, and NMR spectra (PDF)

  • Cartesian coordinates of calculated structures (XYZ)

The authors declare no competing financial interest.

Supplementary Material

om2c00338_si_001.pdf (1.5MB, pdf)
om2c00338_si_002.xyz (15.2KB, xyz)

References

  1. Norton J. R.; Sowa J. Introduction: Metal Hydrides. Chem. Rev. 2016, 116, 8315–8317. 10.1021/acs.chemrev.6b00441. [DOI] [PubMed] [Google Scholar]
  2. 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]
  3. Buil M. L.; Esteruelas M. A.; López A. M. Recent Advances in Synthesis of Molecular Heteroleptic Osmium and Iridium Phosphorescent Emitters. Eur. J. Inorg. Chem. 2021, 2021, 4731–4761. 10.1002/ejic.202100663. [DOI] [Google Scholar]
  4. a Rossin A.; Peruzzini M. Ammonia-Borane and Amine-Borane Dehydrogenation Mediated by Complex Metal Hydrides. Chem. Rev. 2016, 116, 8848–8872. 10.1021/acs.chemrev.6b00043. [DOI] [PubMed] [Google Scholar]; b Gianotti E.; Taillades-Jacquin M.; Rozière J.; Jones D. J. High-Purity Hydrogen Generation via Dehydrogenation of Organic Carriers: A Review on the Catalytic Process. ACS Catal. 2018, 8, 4660–4680. 10.1021/acscatal.7b04278. [DOI] [Google Scholar]; c Shimbayashi T.; Fujita K.-i. Metal-Catalyzed Hydrogenation and Dehydrogenation Reactions for Efficient Hydrogen Storage. Tetrahedron 2020, 76, 130946 10.1016/j.tet.2020.130946. [DOI] [Google Scholar]
  5. Babón J. C.; Esteruelas M. A.; López A. M.. Homogeneous Catalysis by Transition Metal Polyhydride Complexes Chem. Soc. Rev. (submitted). [DOI] [PubMed] [Google Scholar]
  6. Morris R. H. Brønsted–Lowry Acid Strength of Metal Hydride and Dihydrogen Complexes. Chem. Rev. 2016, 116, 8588–8654. 10.1021/acs.chemrev.5b00695. [DOI] [PubMed] [Google Scholar]
  7. Wiedner E. S.; Chambers M. B.; Pitman C. L.; Bullock R. M.; Miller A. J. M.; Appel A. M. Thermodynamic Hydricity of Transition Metal Hydrides. Chem. Rev. 2016, 116, 8655–8692. 10.1021/acs.chemrev.6b00168. [DOI] [PubMed] [Google Scholar]
  8. Jordan A. J.; Lalic G.; Sadighi J. P. Coinage Metal Hydrides: Synthesis, Characterization, and Reactivity. Chem. Rev. 2016, 116, 8318–8372. 10.1021/acs.chemrev.6b00366. [DOI] [PubMed] [Google Scholar]
  9. Eberhardt N. A.; Guan H. Nickel Hydride Complexes. Chem. Rev. 2016, 116, 8373–8426. 10.1021/acs.chemrev.6b00259. [DOI] [PubMed] [Google Scholar]
  10. Maity A.; Teets T. Main Group Lewis Acid-Mediated Transformations of Transition Metal Hydride Complexes. Chem. Rev. 2016, 116, 8873–8911. 10.1021/acs.chemrev.6b00034. [DOI] [PubMed] [Google Scholar]
  11. a Miyaura N. Metal-Catalyzed Reactions of Organoboronic Acids and Esters. Bull. Chem. Soc. Jpn. 2008, 81, 1535–1553. 10.1246/bcsj.81.1535. [DOI] [Google Scholar]; b Carreras J.; Caballero A.; Pérez P. Alkenyl Boronates: Synthesis and Applications. Chem. Asian J. 2019, 14, 329–343. 10.1002/asia.201801559. [DOI] [PubMed] [Google Scholar]; c Bontemps S.; Vendier L.; Sabo-Etienne S. Ruthenium-Catalyzed Reduction of Carbon Dioxide to Formaldehyde. J. Am. Chem. Soc. 2014, 136, 4419–4425. 10.1021/ja500708w. [DOI] [PubMed] [Google Scholar]
  12. Caballero A.; Sabo-Etienne S. Ruthenium-Catalyzed Hydroboration and Dehydrogenative Borylation of Linear and Cyclic Alkenes with Pinacolborane. Organometallics 2007, 26, 1191–1195. 10.1021/om0610851. [DOI] [Google Scholar]
  13. a Gunanathan C.; Hölscher M.; Pan F.; Leitner W. Ruthenium Catalyzed Hydroboration of Terminal Alkynes to Z-Vinylboronates. J. Am. Chem. Soc. 2012, 134, 14349–14352. 10.1021/ja307233p. [DOI] [PubMed] [Google Scholar]; b Gorgas N.; Alves L. G.; Stöger B.; Martins A. M.; Veiros L. F.; Kirchner K. Stable, Yet Highly Reactive Nonclassical Iron(II) Polyhydride Pincer Complexes: Z-Selective Dimerization and Hydroboration of Terminal Alkynes. J. Am. Chem. Soc. 2017, 139, 8130–8133. 10.1021/jacs.7b05051. [DOI] [PubMed] [Google Scholar]; c Gorgas N.; Stöger B.; Veiros L. F.; Kirchner K. Iron(II) Bis(Acetylide) Complexes as Key Intermediates in the Catalytic Hydrofunctionalization of Terminal Alkynes. ACS Catal. 2018, 8, 7973–7982. 10.1021/acscatal.8b01942. [DOI] [Google Scholar]
  14. Babón J. C.; Esteruelas M. A.; Fernández I.; López A. M.; Oñate E. Dihydroboration of Alkyl Nitriles Catalyzed by an Osmium-Polyhydride: Scope, Kinetics, and Mechanism. Organometallics 2020, 39, 3864–3872. 10.1021/acs.organomet.0c00582. [DOI] [Google Scholar]
  15. Donnelly L. J.; Parsons S.; Morrison C. A.; Thomas S. P.; Love J. B. Synthesis and Structures of Anionic Rhenium Polyhydride Complexes of Boron-Hydride Ligands and Their Application in Catalysis. Chem. Sci. 2020, 11, 9994–9999. 10.1039/D0SC03458D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. a Bontemps S.; Vendier L.; Sabo-Etienne S. Borane-Mediated Carbon Dioxide Reduction at Ruthenium: Formation of C1 and C2 Compounds. Angew. Chem., Int. Ed. 2012, 51, 1671–1674. 10.1002/anie.201107352. [DOI] [PubMed] [Google Scholar]; b Bontemps S.; Sabo-Etienne S. Trapping Formaldehyde in the Homogeneous Catalytic Reduction of Carbon Dioxide. Angew. Chem., Int. Ed. 2013, 52, 10253–10255. 10.1002/anie.201304025. [DOI] [PubMed] [Google Scholar]; c Bontemps S.; Vendier L.; Sabo-Etienne S. Ruthenium-Catalyzed Reduction of Carbon Dioxide to Formaldehyde. J. Am. Chem. Soc. 2014, 136, 4419–4425. 10.1021/ja500708w. [DOI] [PubMed] [Google Scholar]
  17. Esteruelas M. A.; Martínez A.; Oliván M.; Oñate E. Direct C–H Borylation of Arenes Catalyzed by Saturated Hydride-Boryl-Iridium-POP Complexes: Kinetic Analysis of the Elemental Steps. Chem. - Eur. J. 2020, 12632–12644. 10.1002/chem.202001838. [DOI] [PubMed] [Google Scholar]
  18. a Yoshida H. Borylation of Alkynes under Base/Coinage Metal Catalysis: Some Recent Developments. ACS Catal. 2016, 6, 1799–1811. 10.1021/acscatal.5b02973. [DOI] [Google Scholar]; b Wang X.; Wang Y.; Huang W.; Xia C.; Wu L. Direct Synthesis of Multi(boronate) Esters from Alkenes and Alkynes via Hydroboration and Boration Reactions. ACS Catal. 2021, 11, 1–18. 10.1021/acscatal.0c03418. [DOI] [Google Scholar]; c Rej S.; Das A.; Panda T. K. Overview of Regioselective and Stereoselective Catalytic Hydroboration of Alkynes. Adv. Synth. Catal. 2021, 363, 4818–4840. 10.1002/adsc.202100950. [DOI] [Google Scholar]
  19. a Hall J. W.; Unson D. M. L.; Brunel P.; Collins L. R.; Cybulski M. K.; Mahon M. F.; Whittlesey M. K. Copper-NHC-Mediated Semihydrogenation and Hydroboration of Alkynes: Enhanced Catalytic Activity Using Ring-Expanded Carbenes. Organometallics 2018, 37, 3102–3110. 10.1021/acs.organomet.8b00467. [DOI] [Google Scholar]; b Rami F.; Bächtle F.; Plietker B. Hydroboration of internal alkynes catalyzed by FeH(CO)(NO)(PPh3)2: a case of boron-source controlled regiosesectivity. Catal. Sci. Technol. 2020, 10, 1492–1497. 10.1039/C9CY02461A. [DOI] [Google Scholar]; c González M. J.; Bauer F.; Breit B. Cobalt-Catalyzed Hydroboration of Terminal and Internal Alkynes. Org. Lett. 2021, 23, 8199–8203. 10.1021/acs.orglett.1c02854. [DOI] [PubMed] [Google Scholar]
  20. a Sundararaju B.; Fürstner A. A trans-Selective Hydroboration of Internal Alkynes. Angew. Chem., Int. Ed. 2013, 52, 14050–14054. 10.1002/anie.201307584. [DOI] [PubMed] [Google Scholar]; b Longobardi L. E.; Fürstner A. trans-Hydroboration of Propargyl Alcohol Derivatives and Related Substrates. Chem. – Eur. J. 2019, 25, 10063–10068. 10.1002/chem.201902228. [DOI] [PubMed] [Google Scholar]
  21. Curto S. G.; Esteruelas M. A.; Oliván M.; Oñate E. Insertion of Diphenylacetylene into Rh–Hydride and Rh–Boryl Bonds: Influence of the Boryl on the Behavior of the β-Borylalkenyl Ligand. Organometallics 2019, 38, 4183–4192. 10.1021/acs.organomet.9b00513. [DOI] [Google Scholar]
  22. Curto S. G.; Esteruelas M. A.; Oliván M.; Oñate E.; Vélez A. β-Borylalkenyl ZE Isomerization in Rhodium-Mediated Diboration of Nonfunctionalized Internal Alkynes. Organometallics 2018, 37, 1970–1978. 10.1021/acs.organomet.8b00259. [DOI] [Google Scholar]
  23. Curto S. G.; Esteruelas M. A.; Oliván M.; Oñate E. Rhodium-Mediated Dehydrogenative Borylation–Hydroborylation of Bis(alkyl)alkynes: Intermediates and Mechanism. Organometallics 2019, 38, 2062–2074. 10.1021/acs.organomet.9b00104. [DOI] [Google Scholar]
  24. Aracama M.; Esteruelas M. A.; Lahoz F. J.; Lopez J. A.; Meyer U.; Oro L. A.; Werner H. Synthesis, Reactivity, Molecular Structure, and Catalytic Activity of the Novel Dichlorodihydridoosmium(IV) Complexes OsH2Cl2(PR3)2 (PR3 =P-i-Pr3, PMe-t-Bu2). Inorg. Chem. 1991, 30, 288–293. 10.1021/ic00002a025. [DOI] [Google Scholar]
  25. a 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]; 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 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]; e 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]; 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]; h Buil M. L.; Esteruelas M. A.; Izquierdo S.; Nicasio A. I.; Oñate E. N–H and C–H Bond Activations of an Isoindoline Promoted by Iridium- and Osmium-Polyhydride Complexes: A Noninnocent Bridge Ligand for Acceptorless and Base-Free Dehydrogenation of Secondary Alcohols. Organometallics 2020, 39, 2719–2731. 10.1021/acs.organomet.0c00316. [DOI] [Google Scholar]; i 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]; j 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]; k Babón J. C.; Esteruelas M. A.; Fernández I.; López A. M.; Oñate E. Assembly of a Dihydrideborate and Two Aryl Nitriles to Form a C,N,N′-Pincer Ligand Coordinated to Osmium. Organometallics 2021, 40, 635–642. 10.1021/acs.organomet.0c00690. [DOI] [PMC free article] [PubMed] [Google Scholar]; l Buil M. L.; Cabeza J. A.; Esteruelas M. A.; Izquierdo S.; Laglera-Gándara C. J.; Nicasio A. I.; Oñate E. Alternative Conceptual Approach to the Design of Bifunctional Catalysts: An Osmium Germylene System for the Dehydrogenation of Formic Acid. Inorg. Chem. 2021, 60, 16860–16870. 10.1021/acs.inorgchem.1c02893. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. 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]
  27. Esteruelas M. A.; López A. M.; Oñate E.; Raga E. Metathesis between E–C(spn) and H–C(sp3) σ-Bonds (E = Si, Ge; n = 2, 3) on an Osmium-Polyhydride. Angew. Chem. Int. Ed. 2022, 61, e202204081 10.1002/anie.202204081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. 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 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]; c 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]; d Casarrubios L.; Esteruelas M. A.; Larramona C.; Muntaner J. G.; Oñate E.; Sierra M. A. 2-Azetidinones as Precursors of Pincer Ligands: Preparation, Structure, and Spectroscopic Properties of CC′N-Osmium Complexes. Inorg. Chem. 2015, 54, 10998–11006. 10.1021/acs.inorgchem.5b02138. [DOI] [PubMed] [Google Scholar]; e Alabau R. G.; Esteruelas M. A.; Oliván M.; Oñate E.; Palacios A. U.; Tsai J.-Y.; Xia C. Osmium(II) Complexes Containing a Dianionic CCCC-Donor Tetradentate Ligand. Organometallics 2016, 35, 3981–3995. 10.1021/acs.organomet.6b00776. [DOI] [Google Scholar]; f 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]; g 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]; h Cancela L.; Esteruelas M. A.; Galbán J.; Oliván M.; Oñate E.; Vélez A.; Vidal J. C. Electronic Communication in Binuclear Osmium- and Iridium-Polyhydrides. Inorg. Chem. 2021, 60, 2783–2796. 10.1021/acs.inorgchem.0c03680. [DOI] [PMC free article] [PubMed] [Google Scholar]; i Cancela L.; Esteruelas M. A.; Oliván M.; Oñate E. Azolium Control of the Osmium-Promoted Aromatic C–H Bond Activation in 1,3-Disubstituted Substrates. Organometallics 2021, 40, 3979–3991. 10.1021/acs.organomet.1c00565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Babón J. C.; Esteruelas M. A.; López A. M.; Oñate E. Osmium-Promoted Transformation of Alkyl Nitriles to Secondary Aliphatic Amines: Scope and Mechanism. Organometallics 2020, 39, 2177–2188. 10.1021/acs.organomet.0c00236. [DOI] [Google Scholar]
  30. Eguillor B.; Esteruelas M. A.; García-Raboso J.; Oliván M.; Oñate E. Stoichiometric and Catalytic Deuteration of Pyridine and Methylpyridines by H/D Exchange with Benzene-d6 Promoted by an Unsaturated Osmium Tetrahydride Species. Organometallics 2009, 28, 3700–3709. 10.1021/om900335b. [DOI] [Google Scholar]
  31. Babón J. C.; Esteruelas M. A.; López A. M.; Oñate E. Hydration of Aliphatic Nitriles Catalyzed by an Osmium Polyhydride: Evidence for an Alternative Mechanism. Inorg. Chem. 2021, 60, 7284–7296. 10.1021/acs.inorgchem.1c00380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Barrio P.; Esteruelas M. A.; Oñate E. Reactions of a Hexahydride-Osmium Complex with Aldehydes: Double C-Hα Activation-Decarbonylation and Single C-Hα Activation-Hydroxylation Tandem Processes and Catalytic Tishchenko Reactions. Organometallics 2004, 23, 1340–1348. 10.1021/om034389l. [DOI] [Google Scholar]
  33. Liu X.; Liu B.; Liu Q. Migratory Hydrogenation of Terminal Alkynes by Base/Cobalt Relay Catalysis. Angew. Chem., Int. Ed. 2020, 59, 6750–6755. 10.1002/anie.201916014. [DOI] [PubMed] [Google Scholar]
  34. Esteruelas M. A.; Oliván M.; Oñate E.. Sigma-bond Activation Reactions Induced by Unsaturated Os(IV)-Hydride Complexes. In Advances in Organometallic Chemistry; Elsevier, 2020; Vol. 74, pp 53–104. [Google Scholar]
  35. a Carbó J. J.; Crochet P.; Esteruelas M. A.; Jean Y.; Lledós A.; López A. M.; Oñate E. Two- and Four-Electron Alkyne Ligands in Osmium-Cyclopentadienyl Chemistry: Consequences of the π→M Interaction. Organometallics 2002, 21, 305–314. 10.1021/om010645n. [DOI] [Google Scholar]; b Barrio P.; Esteruelas M. A.; Oñate E. Reactions of Elongated Dihydrogen-Osmium Complexes Containing Orthometalated Ketones with Alkynes: Hydride-Vinylidene-π-Alkyne versus Hydride-Osmacyclopropene. Organometallics 2003, 22, 2472–2485. 10.1021/om021043m. [DOI] [Google Scholar]; c Bolaño T.; Castarlenas R.; Esteruelas M. A.; Oñate E. Assembly of an Allenylidene Ligand, a Terminal Alkyne, and an Acetonitrile Molecule: Formation of Osmacyclopentapyrrole Derivatives. J. Am. Chem. Soc. 2006, 128, 3965–3973. 10.1021/ja058355b. [DOI] [PubMed] [Google Scholar]
  36. See for example:; a Esteruelas M. A.; Gutiérrez-Puebla E.; López A. M.; Oñate E.; Tolosa J. I. Reactions of Os(η5-C5H5)Cl(PiPr3)2 with NH=CPh2 and PPh3: The Unit Os(η5-C5H5)Cl(PiPr3) as Support for the Study of the Competitive Alkane-Arene Intramolecular C–H Activation. Organometallics 2000, 19, 275–284. 10.1021/om990538h. [DOI] [Google Scholar]; b Esteruelas M. A.; López A. M.; Oñate E.; Royo E. Synthesis and Reactivity of Osmium Complexes Containing a Cyclopentadienyl Ligand with a Pendant Phosphine Donor Group. Organometallics 2004, 23, 3021–3030. 10.1021/om049904+. [DOI] [Google Scholar]; c Esteruelas M. A.; López A. M.; Oñate E.; Royo E. Influence of the Solvent in the Synthesis of Osmium Complexes Containing Cyclopentadienyl Ligands with a Pendant Donor Group. Organometallics 2004, 23, 5633–5636. 10.1021/om049514n. [DOI] [Google Scholar]
  37. a Frost P. W.; Howard J. A. K.; Spencer J. L. An Osmium Tetrahydroborate Complex with Unusual Dynamic Behaviour: X-Ray Crystal Structure of [Os(BH4)H3{P(c-C5H9)3}2] (c-C5H9 = cyclo-C5H9). J. Chem. Soc., Chem. Commun. 1984, 1362–1363. 10.1039/c39840001362. [DOI] [Google Scholar]; b Esteruelas M. A.; López A. M.; Mora M.; Oñate E. Boryl-Dihydrideborate Osmium Complexes: Preparation, Structure, and Dynamic Behavior in Solution. Organometallics 2015, 34, 941–946. 10.1021/om501309g. [DOI] [Google Scholar]; c Buil M. L.; Cardo J. J. F.; Esteruelas M. A.; Fernández I.; Oñate E. An Entry to Stable Mixed Phosphine–Osmium–NHC Polyhydrides. Inorg. Chem. 2016, 55, 5062–5070. 10.1021/acs.inorgchem.6b00658. [DOI] [PubMed] [Google Scholar]
  38. Esteruelas M. A.; Fernández I.; García-Yebra C.; Martín J.; Oñate E. Elongated σ-Borane versus σ-Borane in Pincer–POP–Osmium Complexes. Organometallics 2017, 36, 2298–2307. 10.1021/acs.organomet.7b00234. [DOI] [Google Scholar]
  39. Hartwig J. F.; Cook K. S.; Hapke M.; Incarvito C. D.; Fan Y.; Webster C. E.; Hall M. B. Rhodium Boryl Complexes in the Catalytic, Terminal Functionalization of Alkanes. J. Am. Chem. Soc. 2005, 127, 2538–2552. 10.1021/ja045090c. [DOI] [PubMed] [Google Scholar]
  40. Hebden T. J.; Denney M. C.; Pons V.; Piccoli P. M. B.; Koetzle T. F.; Schultz A. J.; Kaminsky W.; Goldberg K. I.; Heinekey D. M. σ-Borane Complexes of Iridium: Synthesis and Structural Characterization. J. Am. Chem. Soc. 2008, 130, 10812–10820. 10.1021/ja801898m. [DOI] [PubMed] [Google Scholar]
  41. a Kim Y. B.; Kim D.; Dighe S. U.; Chang S.; Park J.-W. Cobalt-Hydride-Catalyzed Hydrosilylation of 3-Alkynes. ACS Catal. 2021, 11, 1548–1553. 10.1021/acscatal.0c05424. [DOI] [Google Scholar]; b Sahoo M. K.; Kim D.; Chang D. S.; Park J.-W. Regioselective Access to α-Vinylsilanes and α-Vinylgermanes by Cobalt-Catalyzed Migratory Hydrofunctionalization of 2-Alkynes. ACS Catal. 2021, 11, 12777–12784. 10.1021/acscatal.1c03769. [DOI] [Google Scholar]
  42. Xu S.; Geng P.; Li Y.; Liu G.; Zhang L.; Guo Y.; Huang Z. Pincer Iron Hydride Complexes for Alkene Isomerization: Catalytic Approach to Trisubstituted (Z)-Alkenyl Boronates. ACS Catal. 2021, 11, 10138–10147. 10.1021/acscatal.1c02432. [DOI] [Google Scholar]
  43. Yun W. J.; Han J. T.; Yun J. NHC-Copper-Catalyzed Tandem Hydrocupration and Allylation of Alkenyl Boronates. Synthesis 2017, 49, 4753–4758. 10.1055/s-0036-1588811. [DOI] [Google Scholar]

Associated Data

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

Supplementary Materials

om2c00338_si_001.pdf (1.5MB, pdf)
om2c00338_si_002.xyz (15.2KB, xyz)

Articles from Organometallics are provided here courtesy of American Chemical Society

RESOURCES