Abstract
Hydride transfer catalysis is shown to be enabled by the nonspectator reactivity of a transition metal-bound low-symmetry tricoordinate phosphorus ligand. Complex 1•[Ru]+, comprising a nontrigonal phosphorus chelate (1, P(N(o-N(2-pyridyl)C6H4)2) and an inert metal fragment ([Ru] = (Me5C5)Ru), reacts with NaBH4 to give a metallohydridophosphorane (1H•[Ru]) by P–H bond formation. Complex 1H•[Ru] is revealed to be a potent hydride donor (ΔG°H–,exp < 41 kcal/mol, ΔG°H–,calc = 38 ± 2 kcal/mol in MeCN). Taken together, the reactivity of the 1•[Ru]+/1H•[Ru] pair comprise a catalytic couple, enabling catalytic hydrodechlorination in which phosphorus is the sole reactive site of hydride transfer.
Graphical Abstract

The roles that ligands play in modulating the properties of organotransition metal complexes vary, especially as a function of their direct involvement in reactivity. Descriptors spanning a broad spectrum of reactivity from ancillary/spectator/innocent (A),1–2,3,4 to electronically-coupled/noninnocent (B),5–6,7,8 and participatory/cooperative (C)9–10,11 are well-known (Figure 1). On the far extreme of this continuum, a ‘functional’ class of ligands (D) —in which bond making and breaking is localized on the ligand itself in conjunction with a stabilizing ‘spectator metal’—are comparatively less common.12–13,14,15,16,17
Figure 1.

Ligand role in transition metal catalysis as a function of participation in reactivity. L = ligand, M = metal, X,Y= reactive substituents.
Although tricoordinate phosphorus (σ3-P) ligands are largely seen as archetypal spectator ligands within the foregoing continuum, the literature collects sporadic reference to reactions invoking their conversion to higher P-coordinate intermediates.18 In this vein, we have recently established that nontrigonal σ3-P compounds19–20,21,22 exhibit enhanced electrophilicity at P that enables access to stable higher coordinate (σ4-P)–M complexes by ‘nonspectator’ reactivity.23–24,25,26 Based on these findings, we hypothesized that a new opportunity might be realized for useful catalysis based on a “functional” σ3-P ligand if paired with an inert, coordinatively- and electronically-saturated metal guest. We report here the realization of this vision using a nontrigonal σ3-P-containing chelate in combination with a spectator (Me5C5)Ru+ fragment to drive hydride transfer reductions of organic acceptors and chloroalkanes without chemical or redox reactions at Ru. Instead, reactivity is solely localized at P. These results: (1) establish new precedent for σ3-P moieties as functional sites for ligand-catalyzed reactions, (2) validate higher-coordinate (σ4-P)–M metallophosphoranes as meritorious intermediates in catalysis, and (3) portend new opportunities for catalyst design by a role reversal of the metal-ligand construct, where reactive and functional ligands are tuned by decoration with spectator metal fragments.
Organometallic hydride transfer reactions usually access reactive metal sites for hydride formation,27,28 but on the basis of the well-known kinetic inertness of coordinatively-saturated Ru(II) complexes, the (Me5C5)Ru+ fragment was selected as an intentionally passive chelating template for a nontrigonal σ3-P ligand. Specifically, addition of the tridentate nontrigonal σ3-P ligand 1 to [(Me5C5)Ru(MeCN)3](PF6) provided in high yield the orange N,P,N-chelated complex 1•[Ru]+ ([Ru]+= (Me5C5)Ru+) as its hexafluorophosphate salt (Scheme 1). Compared to free 1 (δ 141.7 ppm, Figure 2a), complexation results in a downfield shift of the 31P{1H} NMR resonance for 1•[Ru]+ (δ 206.4 ppm, Figure 2b). In the solution phase, 1H NMR spectra for complex 1•[Ru]+ display an apparent mirror symmetry for the two phenylene and two pyridyl moieties, respectively, along with splitting of the Cp* methyl peak as a doublet (4JPH = 3.6 Hz), in line with the expected tridentate coordination of 1 to ruthenium. In the solid state, single-crystal X-ray diffraction confirms the local bonding environment about the coordinatively saturated, 18-electron Ru(II) center (Figure 3, left). As a function of the chelating structure, a short d(Ru–P) bond distance (2.1561(5) Å, Table 1) is found. With respect to the local geometry about phosphorus, coordination to Ru modestly perturbs the local nontrigonal structure, associated with an increase in the angle ∠N2–P–N4 to 121.6(5)° (Δ∠N2–P–N4 = +12.9° compared to 1). Given the ambient fluxional behavior of a related nontrigonal P(III) triamide by flexion of the ∠N2–P–N4 angle,29 this deformation is not expected to be energetically costly.
Scheme 1.

Synthesis of 1•[Ru]+ and 1H•[Ru].
Conditions: (i) [(Me5C5)Ru(MeCN)3][PF6], CH2Cl2, 25 °C, 16 h; 91% yield. (ii) NaBH4, MeCN, 40 °C, 1 h; 65% yield.
Figure 2.

31P NMR Spectra for (a) Compound 1 in C6D6. (b) Compound 1•[Ru]+ in CD3CN. (c) Compound 1H•[Ru] in C6D6.
Figure 3.

Structural representations of 1•[Ru]+ (left) and 1H•[Ru] (right) from single crystal X-ray diffraction. Ellipsoids are set at 50% occupancy. All hydrogen atoms (except those attached to P), counter anions, and additional molecules in the asymmetric unit are omitted for clarity. See the SI Section VIII for crystallographic details.
Table 1.
Selected bond distances (Å), angles (deg), and 31P NMR chemical shifts (ppm) for 1•[Ru]+ and 1H•[Ru].
| metric | 1 | 1•[Ru]+ | 1H•[Ru] |
|---|---|---|---|
| Ru–P (Å) | -- | 2.1561(5) | 2.229(3) |
| Ru–N1 (Å) | -- | 2.106(7) | 2.154(9) |
| P–N3 (Å) | 1.7485(8) | 1.7071(19) | 1.753(9) |
| P–N2 (Å) | 1.7341(8) | 1.741(7) | 1.882(9) |
| ∠N1–Ru–N5 (°) | -- | 93.0(5) | 103.7(3) |
| ∠N1–Ru–P (°) | -- | 80.9(2) | 76.6(3) |
| ∠N2–P–N4 (°) | 108.67(4) | 121.6(5) | 168.5(4) |
| 31P NMR δ (ppm) | 141.7 | 206.4 | 19.9 |
Despite the coordinative and electronic saturation of the ruthenium center in 1•[Ru]+, an orange acetonitrile solution of 1•[Ru]+ readily reacts with NaBH4 to give a red suspension within 1 h at 40 °C. 31P{1H} NMR spectroscopy of the isolated red solid revealed the absence of downfield resonances in the vicinity of the starting material, while a single upfield peak (δ 19.9 ppm) that evolves into a doublet in the proton-coupled 31P NMR spectrum (J = 469 Hz, Figure 2C) was observed. The corresponding coupling partner is identified in the 1H NMR spectrum as a doublet (J = 469 Hz) centered at δ 8.38 ppm. Both the chemical shift and coupling magnitude are inconsistent with metal-based hydride.30,31 Instead, these data indicate the formation of the metallohydridophosphorane complex 1H•[Ru], containing a higher-coordinate phosphorus moiety resulting from P–H bond formation.
Indeed, X-ray diffraction on crystalline 1H•[Ru] confirmed the presence of a 5-coordinate phosphorus center of intermediate square pyramidal/trigonal bipyramidal geometry (τ5 = 0.53)32 bearing a hydride with anti configuration with respect to the Me5C5 fragment about the Ru–P vector (Figure 3, right). The addition of the hydride results in significant local geometric changes at P for 1H•[Ru] as compared to 1•[Ru]+. First, the Ru–P bond distance is significantly elongated (d(Ru–P) = 2.229(3) Å, Table 1), in line with previously observed trends in metal-phosphorus bond distances between (σ3-P)–M and (σ4-P)–M complexes.33,34 Second, the angle ∠N2–P–N4 for 1H•[Ru] is increased significantly to 168.5(4)° (Δ∠N2–P–N4 = +46.9° compared to 1•[Ru]+) to accommodate the hydride substituent. A further consequence of this distortion is a substantial differential in the P–N distances, where d(P–N2)= 1.882(9) Å but d(P–N3) = 1.753(9) Å. This effect can be rationalized by reference to the Pimentel–Rundle model of 3c-4e bonding35 for the nearly linear N2–P–N4 triad.
DFT models at the M06/def2-TZVP level of theory for 1•[Ru]+ and 1H•[Ru] reproduce well the experimental structures. Bonding analysis within the NBO formalism36 reveals changes to the nature of the Ru–P σ-interactions as a function of hydride binding. A natural localized molecular orbital (NLMO) corresponding to a dative covalent P→Ru σ-interaction for 1•[Ru]+ shows modest polarization toward the phosphorus (P 62.0%/Ru 34.5%); the NLMO corresponding to the P–Ru bonding interaction in 1H•[Ru] indicates an increased degree of covalency (P 55.9%/Ru 41.8%). Further analysis of the electronic models also provide a rationale for the observed reactivity. Inspection of the Kohn-Sham orbitals of 1•[Ru]+ (Figure S52) reveals a low-lying unfilled orbital strongly localized on phosphorus and projecting into the concave space within the chelate that serves as the binding site for hydride to form 1H•[Ru]. Once formed, the P–H bond is described by a NLMO with leading contribution from H (P 47.2%/H 51.6%). Indeed, by natural population analysis (NPA), the P-H hydrogen is assigned more negative natural charge (−0.09) than any other hydrogen in the complex (+0.20), indicating hydridic character.
To test the hydridic reactivity28,37–38 of the metallophosphorane P–H bond, complex 1H•[Ru] was first treated with 1,3-dimethylbenzimidazolium (2+), a known hydride acceptor (ΔG°H– = 45 kcal/mol39 for conjugate 2H, Scheme 2A). Over the course of the reaction, complete consumption of 1H•[Ru] was observed with concomitant formation of 1•[Ru]+ and dihydrobenzimidazole 2H. Even the reaction of 1H•[Ru] with 1,3-dimethylpyridinum (3+, ΔG°H– = 41 kcal/mol40)—among the weakest tabulated organic hydride acceptors—proceeded cleanly and completely to give 1•[Ru]+ and dihydropyridine 3H (Scheme 2A). The failure to identify a hydride transfer partner that would permit measurement of a dynamic equilibrium establishes 1H•[Ru] as a potent hydride donor with a limiting thermodynamic hydricity of ΔG°H– ≤ 41 kcal/mol (Scheme 2B).41
Scheme 2.

(A) Hydride transfer reactions of 1H•[Ru] with organic acceptors. (B) Experimental and computed thermodynamic hydricity of 1H•[Ru].
To more accurately measure the hydricity of 1H•[Ru], hydride transfer equilibria with transition-metal based acceptors was also explored. However, these reactions were prone to competing side reactions over the course of of the long requisite reaction times (see SI Section V). Similarly, our attempts to measure hydricity via H2 heterolysis or pKa-potential methods were hindered by either by a lack of reactivity with H2 or irreversible reduction features observed in cyclic voltammograms of 1•[Ru]+.28 Therefore, to more precisely quantify the hydricity of 1H•[Ru], hydride transfer reactions between 1H•[Ru] and 2+ or 3+ were calculated by DFT at the M06/def2-TZVP level (Scheme 2). The computed hydricity value for 1H•[Ru] inferred from these calculations (ΔG°H–,calc = 38 ± 2 kcal/mol) is fully consistent with the experimental observations and marks 1H•[Ru] as the strongest ‘ligand-centered’ hydride donor yet quantified by almost 20 kcal/mol.42–43,44,45,46,47,48 Distortion-interaction analysis provides a partial rationalization for the strongly hydridic nature of 1H•[Ru] relative to prior ligand-centered hydrides; 31 kcal/mol of strain release is computed for relaxation to 1•[Ru]+ following a non-adiabatic hydride transfer from 1H•[Ru].
The strongly hydridic nature of 1H•[Ru] enables hydride transfer reactions to other weak electrophiles including chloroalkanes (Figure 4A). Representatively, dissolution of 1H•[Ru] in CDCl3 in a Teflon-sealed NMR tube results in a red-to-orange color change within minutes, after which complete consumption of 1H•[Ru] was observed. Both 1•[Ru]+ and CDHCl2 (δ 5.28 ppm, Figure S41) were identified as the sole products. Similarly, mild heating (50 °C, 16 h) of a solution of 1H•[Ru] in CD2Cl2 led to formation of 1•[Ru]+ along with CD2HCl (δ 2.98 ppm, Figure S43). In both instances, coordination of the displaced Cl− ion to the conjugate hydride acceptor 1•[Ru]+ is not observed, in contrast to both transition metal and diazaphospholene hydrodechlorination reactions.49–50,51,52 Indeed, dissolution of 1•[Ru]+ in a saturated solution of [nBu4N][Cl] in CD3CN (~35 equiv, Figures S45–46) showed no evidence of Cl− binding by NMR spectroscopy.
Figure 4.

(A) Stoichiometric hydrodechlorination of chloroalkanes using 1H•[Ru]. (B) Catalytic hydrodechlorination of CDCl3 using 1•[Ru]+ and [nBu4][BH4]. py = pyridine.
The combined reactivities of 1•[Ru]+ as a hydride acceptor from borohydride and 1H•[Ru] as a hydride donor to organic substrates suggested the possibility of employing this complex as a catalyst for hydride transfer (Figure 4B). Whereas the direct reaction of the soluble hydride donor [nBu4N][BH4] with bulk CDCl3 for 48 h at 25 °C proceeds only to minimal conversion (<5% yield, see SI Table S1), the addition of 1 mol% of 1•[Ru]+ to a CDCl3 solution containing [nBu4][BH4] (0.27 mM) and pyridine (0.54 mM, as a BH3 scavenger) gave unambiguous evidence for formation of CDHCl2 (1H NMR δ 5.11 ppm, J=1 Hz) with a turnover number of 31 ± 1 (average of two runs). Consistent with the qualitative rates of formation and consumption of 1•[Ru]+ and 1H•[Ru] from stoichiometric studies, in situ 31P NMR spectra implicate 1•[Ru]+ as the resting state of the catalytic cycle, and imply that future designs that increase the propensity of 1•[Ru]+ to accept hydride at P might further enhance the rate of catalysis.
These results demonstrate the capacity for nontrigonal σ3-P ligands to support hydride transfer catalysis via a functional mode of reactivity that transits through a higher-coordinate (σ4-P)–M metallohydridophosphorane. In these complexes, the metal is relegated to a peripheral position with respect to bond making and breaking. Indeed, since the reaction chemistry occurs exclusively at P and not Ru, it might be more apt to consider 1•[Ru]+ an electrophilic phosphonium ion53 bearing an ancillary (Me5C5)Ru substituent. From this point of view, the apparent reversal in roles for a metal-phosphorus complex demonstrated by the present results suggests fresh opportunities for the deliberate design of novel main group catalysts that incorporate spectator metal fragments.
Supplementary Material
ACKNOWLEDGMENT
This work was supported by the NIH (R21 GM134240). Q.J.B. acknowledges NIH NRSA postdoctoral fellowship support (F32 GM143865). A.T. acknowledges graduate fellowship support from the Honjo International Scholarship Foundation.
Footnotes
Supporting Information.
Experimental details, characterization data, and Cartesian Coordinates (PDF)
Crystallographic data (CIF)
This material is available free of charge via the Internet at http://pubs.acs.org.”
REFERENCES
- 1.Tolman CA Electron Donor-Acceptor Properties of Phosphorus Ligands. Substituent Additivity. J. Am. Chem. Soc 1970, 92, 2953–2956. [Google Scholar]
- 2.Tolman CA Steric Effects of Phosphorus Ligands in Organometallic Chemistry and Homogeneous Catalysis. Chem. Rev 1977, 77, 313–348. [Google Scholar]
- 3.Fryzuk MD; Haddad TS; Berg DJ; Rettig SJ Phosphine Complexes of the Early Metals and the Lanthanoids. Pure App. Chem 1991, 63, 845–850. [Google Scholar]
- 4.Crabtree RH NHC Ligands versus Cyclopentadienyls and Phosphines as Spectator Ligands in Organometallic Catalysis. J. Organomet. Chem 2005, 690, 5451–5457. [Google Scholar]
- 5.Chirik PJ Preface: Forum on Redox-Active Ligands. Inorg. Chem 2011, 50, 9737–9740. [DOI] [PubMed] [Google Scholar]
- 6.Eisenberg R; Gray HB Noninnocence in Metal Complexes: A Dithiolene Dawn Inorg. Chem 2011, 50, 9741–9751. [DOI] [PubMed] [Google Scholar]
- 7.Lyaskovskyy V; De Bruin B Redox Non-Innocent Ligands: Versatile New Tools to Control Catalytic Reactions. ACS Catal. 2012, 2, 270–279. [Google Scholar]
- 8.Broere DLJ; Plessius R; Van Der Vlugt JI New Avenues for Ligand-Mediated Processes – Expanding Metal Reactivity by the Use of Redox-Active Catechol, o-Aminophenol and o-Phenylenediamine Ligands. Chem. Soc. Rev 2015, 44, 6886–6915. [DOI] [PubMed] [Google Scholar]
- 9.Grützmacher H Cooperating Ligands in Catalysis. Angew. Chem. Int. Ed 2008, 47, 1814–1818. [DOI] [PubMed] [Google Scholar]
- 10.Khusnutdinova JR; Milstein D Metal-Ligand Cooperation. Angew. Chem. Int. Ed 2015, 54, 12236–12273. [DOI] [PubMed] [Google Scholar]
- 11.Whited MT Pincer-Supported Metal/Main-Group Bonds as Platforms for Cooperative Transformations. Dalton Trans. 2021, 50, 16443–16450. [DOI] [PubMed] [Google Scholar]
- 12.Annibale VT; Dalessandro DA; Song D Tuning the Reactivity of an Actor Ligand for Tandem CO2 and C–H Activations: From Spectator Metals to Metal-Free. J. Am. Chem. Soc 2013, 135, 16175–16183. [DOI] [PubMed] [Google Scholar]
- 13.Lee K; Donahue CM; Daly SR Triaminoborane-Bridged Diphosphine Complexes with Ni and Pd: Coordination Chemistry, Structures, and Ligand-Centered Reactivity. Dalton Trans. 2017, 46, 9394–9406. [DOI] [PubMed] [Google Scholar]
- 14.Luo GG; Zhang HL; Tao YW; Wu QY; Tian D; Zhang Q Recent Progress in Ligand-Centered Homogeneous Electrocatalysts for Hydrogen Evolution Reaction. Inorg. Chem. Front 2019, 6, 343–354. [Google Scholar]
- 15.Wegener AR; Kabes CQ; Gladysz JA Launching Werner Complexes into the Modern Era of Catalytic Enantioselective Organic Synthesis. Acc. Chem. Res 2020, 53, 2299–2313. [DOI] [PubMed] [Google Scholar]
- 16.Larionov VA; Feringa BL; Belokon YN Enantioselective “Organocatalysis in Disguise” by the Ligand Sphere of Chiral Metal-Templated Complexes. Chem. Soc. Rev 2021, 50, 9715–9740. [DOI] [PubMed] [Google Scholar]
- 17.Skaria M; Culpepper JD; Daly SR Leveraging Metal and Ligand Reactive Sites for One Pot Reactions: Ligand-Centered Borenium Ions for Tandem Catalysis with Palladium. Chem. Eur. J 2022, in press. DOI: 10.1002/chem.202201791 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Goodman J; Macgregor SA Metallophosphoranes: The Hidden Face of Transition Metal-Phosphine Complexes. Coord. Chem. Rev 2010, 254, 1295–1306. [Google Scholar]
- 19.Arduengo AJ; Stewart CA Low Coordinate Hypervalent Phosphorus. Chem. Rev 1994, 94, 1215–1237. [Google Scholar]
- 20.Brand A; Uhl W Sterically Constrained Bicyclic Phosphines: A Class of Fascinating Compounds Suitable for Application in Small Molecule Activation and Coordination Chemistry. Chem. Eur. J 2019, 25, 1391–1404. [DOI] [PubMed] [Google Scholar]
- 21.Lee K; Blake AV; Tanushi A; McCarthy SM; Kim D; Loria SM; Donahue CM; Spielvogel KD; Keith JM; Daly SR; Radosevich AT Validating the Biphilic Hypothesis of Nontrigonal Phosphorus(III) Compounds. Angew. Chem 2019, 131, 7067–7072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Abbenseth J; Goicoechea JM Recent Developments in the Chemistry of Non-Trigonal Pnictogen Pincer Compounds: From Bonding to Catalysis. Chem. Sci 2020, 11, 9728–9740. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Tanushi A; Radosevich AT Insertion of a Nontrigonal Phosphorus Ligand into a Transition Metal-Hydride: Direct Access to a Metallohydrophosphorane. J. Am. Chem. Soc 2018, 140, 8114–8118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Cleveland GT; Radosevich AT A Nontrigonal Tricoordinate Phosphorus Ligand Exhibiting Reversible “Nonspectator” L/X‐Switching. Angew. Chem 2019, 131, 15147–15151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Hwang SJ; Tanushi A; Radosevich AT Enthalpy-Controlled Insertion of a “Nonspectator” Tricoordinate Phosphorus Ligand into Group 10 Transition Metal-Carbon Bonds. J. Am. Chem. Soc 2020, 142, 21285–21291. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.For representative examples of analogous σ2-P/σ3-P interconversions please see:; A) Béguin A; Böttcher H-C; Süss-Fink G; Walther B Interplay of Phosphido and Phosphine Ligands in Dinuclear Ruthenium Complexes: Implication in the Catalytic Hydroformylation of Ethylene. J. Chem. Soc. Dalt. Trans 1992, 2133–2134. [Google Scholar]; B) Rosenberg L Metal Complexes of Planar PR2 Ligands: Examining the Carbene Analogy. Coord. Chem. Rev 2012, 256, 606–626. [Google Scholar]; C) Lee K; Thomas CM Nickel-Templated Replacement of Phosphine Substituents in a Tetradentate Bis(Amido)Bis(Phosphine) Ligand. Inorg. Chem 2021, 60, 17348–17356. [DOI] [PubMed] [Google Scholar]; D) Belli RG; Pantazis DA; McDonald R; Rosenberg L Reversible Silylium Transfer between P-H and Si-H Donors. Angew. Chem. Int. Ed 2021, 60, 2379–2384. [DOI] [PubMed] [Google Scholar]; E) Abhyankar P; MacMillan SN; Lacy DC Activation of H2 with Dinuclear Manganese(I)-Phosphido Complexes. Organometallics 2022, 41, 60–66. [Google Scholar]; F) Belli RG; Yang J; Bahena EN; McDonald R; Rosenberg L Mechanism and Catalyst Design in Ru-Catalyzed Alkene Hydrophosphination. ACS Catal. 2022, 12, 5247–5262. [Google Scholar]
- 27.Perutz RN; Sabo-Etienne S The σ-CAM Mechanism: σ Complexes as the Basis of σ-Bond Metathesis at Late-Transition-Metal Centers. Angew. Chem. Int. Ed 2007, 46, 2578–2592. [DOI] [PubMed] [Google Scholar]
- 28.Wiedner ES; Chambers MB; Pitman CL; Bullock RM; Miller AJM; Appel AM Thermodynamic Hydricity of Transition Metal Hydrides. Chem. Rev 2016, 116, 8655–8692. [DOI] [PubMed] [Google Scholar]
- 29.Zhao W; McCarthy SM; Lai TY; Yennawar HP; Radosevich AT “Reversible Intermolecular E–H Oxidative Addition to a Geometrically Deformed and Structurally Dynamic Phosphorous Triamide.” J. Am. Chem. Soc 2014, 136, 17634–17644. [DOI] [PubMed] [Google Scholar]
- 30.Creutz C; Chou MH Hydricities of d 6 Metal Hydride Complexes in Water. J. Am. Chem. Soc 2009, 131, 2794–2795. [DOI] [PubMed] [Google Scholar]
- 31.Häller LJL; Mas-Marzá E; Cybulski MK; Sanguramath RA; Macgregor SA; Mahon MF; Raynaud C; Russell CA; Whittlesey MK Computation Provides Chemical Insight into the Diverse Hydride NMR Chemical Shifts of [Ru(NHC)4(L)H]0/+ Species (NHC = N-Heterocyclic Carbene; L = Vacant, H2, N2, CO, MeCN, O2, P4, SO2, H−, F− and Cl−) and Their [Ru(R2PCH2CH2PR2)2(L)H]+ Congeners. Dalton Trans. 2017, 46, 2861–2873. [DOI] [PubMed] [Google Scholar]
- 32.Addison AW; Rao TN; Reedijk J; van Rijn J; Verschoor GC Synthesis, Structure, and Spectroscopic Properties of Copper(II) Compounds Containing Nitrogen–Sulphur Donor Ligands; the Crystal and Molecular Structure of Aqua[1,7-Bis(N-Methylbenzimidazol-2′-Yl)-2,6-Dithiaheptane]Copper(II) Perchlorate. J. Chem. Soc. Dalt. Trans 1984, 1349–1356. [Google Scholar]
- 33.Kubo K; Nakazawa H; Mizuta T; Miyoshi K Novel Synthesis, X-Ray Crystal Structures, and Spectroscopic Properties of Metalated Hypervalent Phosphorus Compounds, Cp(CO)LFe{P(OC6H4Y)OC6H4Z} (L = CO, P(OPh)3, P(OMe)3, PMe3; Y, Z = NH, NMe, O). Organometallics 1998, 17, 3522–3531. [Google Scholar]
- 34.Nakazawa H; Kawamura K; Kubo K; Miyoshi K Syntheses, Structures, and Berry Pseudorotation of Ruthenium-Phosphorane Complexes. Organometallics 1999, 18, 2961–2969. [Google Scholar]
- 35.Green MLH; Parkin G The Classification and Representation of Main Group Element Compounds That Feature Three-Center Four-Electron Interactions. Dalton Trans. 2016, 45, 18784–18795. [DOI] [PubMed] [Google Scholar]
- 36.Glendening ED; Landis CR; Weinhold F NBO 6.0: Natural Bond Orbital Analysis Program. J. Comput. Chem 2013, 34, 1429–1437. [DOI] [PubMed] [Google Scholar]
- 37.Ilic S; Alherz A; Musgrave CB; Glusac KD Thermodynamic and Kinetic Hydricities of Metal-Free Hydrides. Chem. Soc. Rev 2018, 47, 2809–2836. [DOI] [PubMed] [Google Scholar]
- 38.Brereton KR; Smith NE; Hazari N; Miller AJM Thermodynamic and Kinetic Hydricity of Transition Metal Hydrides. Chem. Soc. Rev 2020, 49, 7929–7948. [DOI] [PubMed] [Google Scholar]
- 39.Zhu X-Q; Zhang M-T; Yu A; Wang C-H; Cheng J-P Hydride, Hydrogen Atom, Proton, and Electron Transfer Driving Forces of Various Five-Membered Heterocyclic Organic Hydrides and Their Reaction Intermediates in Acetonitrile. J. Am. Chem. Soc 2008, 130, 2501–2516. [DOI] [PubMed] [Google Scholar]
- 40.Matsubara Y; Fujita E; Doherty MD; Muckerman JT; Creutz C Thermodynamic and Kinetic Hydricity of Ruthenium(II) Hydride Complexes. J. Am. Chem. Soc 2012, 134, 15743–15757. [DOI] [PubMed] [Google Scholar]
- 41.Here, ‘hydride transfer’ refers to the overall transformation with no assumptions made about microscopic mechanism, which could proceed ‘hydride ion transfer’ (the concerted transfer of the H- anion) or a stepwise ET/HAT pathway. In view of the relative reduction potentials of 1H•[Ru] and the substrates tested herein, we prefer a mechanism proceeding by concerted hydride ion transfer, although a two-step ET/HAT pathway with an initial endothermic ET cannot be rigorously excluded at this time.
- 42.Wu A; Masland J; Swartz RD; Kaminsky W; Mayer JM Synthesis and Characterization of Ruthenium Bis(β-Diketonato) Pyridine-Imidazole Complexes for Hydrogen Atom Transfer. Inorg. Chem 2007, 46, 11190–11201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Manner VW; Mayer JM Concerted Proton-Electron Transfer in a Ruthenium Terpyridyl-Benzoate System with a Large Separation between the Redox and Basic Sites. J. Am. Chem. Soc 2009, 131, 9874–9875. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.McLoughlin EA; Waldie KM; Ramakrishnan S; Waymouth RM Protonation of a Cobalt Phenylazopyridine Complex at the Ligand Yields a Proton, Hydride, and Hydrogen Atom Transfer Reagent. J. Am. Chem. Soc 2018, 140, 13233–13241. [DOI] [PubMed] [Google Scholar]
- 45.Chang MC; McNeece AJ; Hill EA; Filatov AS; Anderson JS Ligand-Based Storage of Protons and Electrons in Dihydrazonopyrrole Complexes of Nickel. Chem. Eur. J 2018, 24, 8001–8008. [DOI] [PubMed] [Google Scholar]
- 46.Chalkley MJ; Garrido-Barros P; Peters JC A Molecular Mediator for Reductive Concerted Proton-Electron Transfers via Electrocatalysis. Science 2020, 369, 850–854. [DOI] [PubMed] [Google Scholar]
- 47.Charette BJ; Ziller JW; Heyduk AF Exploring Ligand-Centered Hydride and H-Atom Transfer. Inorg. Chem 2021, 60, 5367–5375. [DOI] [PubMed] [Google Scholar]
- 48.Hua SA; Paul LA; Oelschlegel M; Dechert S; Meyer F; Siewert I A Bioinspired Disulfide/Dithiol Redox Switch in a Rhenium Complex as Proton, H Atom, and Hydride Transfer Reagent. J. Am. Chem. Soc 2021, 143, 6238–6247. [DOI] [PubMed] [Google Scholar]
- 49.Alonso F; Beletskaya IP; Yus M Metal-Mediated Reductive Hydrodehalogenation of Organic Halides. Chem. Rev 2002, 102, 4009–4091. [DOI] [PubMed] [Google Scholar]
- 50.Burck S; Gudat D; Nieger M; Du Mont WW P-Hydrogen-Substituted 1,3,2-Diazaphospholenes: Molecular Hydrides. J. Am. Chem. Soc 2006, 128, 3946–3955. [DOI] [PubMed] [Google Scholar]
- 51.Zhang J; Yang JD; Cheng JP Diazaphosphinanes as Hydride, Hydrogen Atom, Proton or Electron Donors under Transition-Metal-Free Conditions: Thermodynamics, Kinetics, and Synthetic Applications. Chem. Sci 2020, 11, 3672–3679. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Zhang J; Yang JD; Cheng JP Exploiting the Radical Reactivity of Diazaphosphinanes in Hydrodehalogenations and Cascade Cyclizations. Chem. Sci 2020, 11, 4786–4790 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Bayne JM; Stephan DW Phosphorus Lewis acids: emerging reactivity and applications in catalysis. Chem. Soc. Rev 2016, 45, 765–774 [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
