Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where R. Morris Bullock is active.

Publication


Featured researches published by R. Morris Bullock.


Science | 2011

A Synthetic Nickel Electrocatalyst with a Turnover Frequency Above 100,000 s−1 for H2 Production

Monte L. Helm; Michael P. Stewart; R. Morris Bullock; M. Rakowski DuBois; Daniel L. DuBois

Precisely shaped basic ligands protect highly reactive protons and electrons to help accelerate catalytic hydrogen formation. Reduction of acids to molecular hydrogen as a means of storing energy is catalyzed by platinum, but its low abundance and high cost are problematic. Precisely controlled delivery of protons is critical in hydrogenase enzymes in nature that catalyze hydrogen (H2) production using earth-abundant metals (iron and nickel). Here, we report that a synthetic nickel complex, [Ni(PPh2NPh)2](BF4)2, (PPh2NPh = 1,3,6-triphenyl-1-aza-3,6-diphosphacycloheptane), catalyzes the production of H2 using protonated dimethylformamide as the proton source, with turnover frequencies of 33,000 per second (s−1) in dry acetonitrile and 106,000 s−1 in the presence of 1.2 M of water, at a potential of –1.13 volt (versus the ferrocenium/ferrocene couple). The mechanistic implications of these remarkably fast catalysts point to a key role of pendant amines that function as proton relays.


Journal of the American Chemical Society | 2011

[Ni(PPh2NC6H4X2)2]2+ Complexes as Electrocatalysts for H2 Production: Effect of Substituents, Acids, and Water on Catalytic Rates

Uriah J. Kilgore; John A. S. Roberts; Douglas H. Pool; Aaron M. Appel; Michael P. Stewart; M. Rakowski DuBois; William G. Dougherty; W. Scott Kassel; R. Morris Bullock; Daniel L. DuBois

A series of mononuclear nickel(II) bis(diphosphine) complexes [Ni(P(Ph)(2)N(C6H4X)(2))(2)](BF(4))(2) (P(Ph)(2)N(C6H4X)(2) = 1,5-di(para-X-phenyl)-3,7-diphenyl-1,5-diaza-3,7-diphosphacyclooctane; X = OMe, Me, CH(2)P(O)(OEt)(2), Br, and CF(3)) have been synthesized and characterized. X-ray diffraction studies reveal that [Ni(P(Ph)(2)N(C6H4Me)(2))(2)](BF(4))(2) and [Ni(P(Ph)(2)N(C6H4OMe)(2))(2)](BF(4))(2) are tetracoordinate with distorted square planar geometries. The Ni(II/I) and Ni(I/0) redox couples of each complex are electrochemically reversible in acetonitrile with potentials that are increasingly cathodic as the electron-donating character of X is increased. Each of these complexes is an efficient electrocatalyst for hydrogen production at the potential of the Ni(II/I) couple. The catalytic rates generally increase as the electron-donating character of X is decreased, and this electronic effect results in the favorable but unusual situation of obtaining higher catalytic rates as overpotentials are decreased. Catalytic studies using acids with a range of pK(a) values reveal that turnover frequencies do not correlate with substrate acid pK(a) values but are highly dependent on the acid structure, with this effect being related to substrate size. Addition of water is shown to dramatically increase catalytic rates for all catalysts. With [Ni(P(Ph)(2)N(C6H4CH2P(O)(OEt)2)(2))(2)](BF(4))(2) using [(DMF)H](+)OTf(-) as the acid and with added water, a turnover frequency of 1850 s(-1) was obtained.


Nature | 2003

A recyclable catalyst that precipitates at the end of the reaction

Vladimir K. Dioumaev; R. Morris Bullock

Homogeneous catalysts—which exist in the same (usually liquid) phase as reactants and products—are usually more selective than heterogeneous catalysts and far less affected by limitations due to slow transport of reactants and products, but their separation from reaction products can be costly and inefficient. This has stimulated the development of strategies that facilitate the recycling of homogeneous catalysts. Some of these methods exploit the preference of a catalyst for one of two solvents with thermoregulated miscibility; others exploit a dramatic decrease in catalyst solubility as one reagent is consumed or temperature changed after completion of the reaction. Here we describe a tungsten catalyst for the solvent-free hydrosilylation of ketones that retains its activity until essentially all of the liquid substrate is converted to liquid products, which we can then simply decant to separate the catalyst that precipitates from the products of the reaction. We attribute the ability of the catalyst to retain its solubility and hence activity until completion of the reaction to the transient formation of liquid clathrate that contains a few molecules of the substrate per molecule of the otherwise solid catalyst. Insights into the fundamental processes controlling the formation of this liquid clathrate might help to tailor other catalysts and substrates, so as to develop efficient and solvent-free schemes for reactions of practical interest.


Science | 2013

Abundant Metals Give Precious Hydrogenation Performance

R. Morris Bullock

Iron- or cobalt-based catalysts have outperformed traditional precious metal catalysts in several hydrogenation reactions. [Also see Reports by Jagadeesh et al., Friedfeld et al., and Zuo et al.] Homogeneous catalysts based on precious metals have enabled highly selective synthesis of organic compounds. Precious metal catalysts including ruthenium (Ru), rhodium (Rh), and platinum (Pt) have been superstars of both homogeneous and heterogeneous catalysis. In recent years, increasing efforts have been devoted to the design and discovery of homogeneous catalysts that incorporate base metals, such as iron (Fe) and cobalt (Co) (1–3) (see the figure). Catalytic hydrogenations are one of the extraordinary success stories of homogeneous catalysis, and three reports in this issue describe remarkable progress in the use of Earth-abundant metals as catalysts for hydrogenations. Hydrogenations are conceptually simple—H2 is added across a C=C or C=O double bond—but mechanistic studies have revealed considerable complexity with respect to how the metal catalyzes hydrogenations (4). On page 1080, Zuo et al. (5) report iron catalysts for asymmetric hydrogenation of C=O bonds. On page 1076, cobalt catalysts for asymmetric hydrogenation of C=C bonds are described by Friedfeld et al. (6). On page 1073, Jagadeesh et al. (7) report on nanoscale iron catalysts for synthesis of functionalized anilines through hydrogenation of nitroarenes.


Journal of the American Chemical Society | 2011

Moving Protons with Pendant Amines: Proton Mobility in a Nickel Catalyst for Oxidation of Hydrogen

Molly O’Hagan; Wendy J. Shaw; Simone Raugei; Shentan Chen; Jenny Y. Yang; Uriah J. Kilgore; Daniel L. DuBois; R. Morris Bullock

Proton transport is ubiquitous in chemical and biological processes, including the reduction of dioxygen to water, the reduction of CO(2) to formate, and the production/oxidation of hydrogen. In this work we describe intramolecular proton transfer between Ni and positioned pendant amines for the hydrogen oxidation electrocatalyst [Ni(P(Cy)(2)N(Bn)(2)H)(2)](2+) (P(Cy)(2)N(Bn)(2) = 1,5-dibenzyl-3,7-dicyclohexyl-1,5-diaza-3,7-diphosphacyclooctane). Rate constants are determined by variable-temperature one-dimensional NMR techniques and two-dimensional EXSY experiments. Computational studies provide insight into the details of the proton movement and energetics of these complexes. Intramolecular proton exchange processes are observed for two of the three experimentally observable isomers of the doubly protonated Ni(0) complex, [Ni(P(Cy)(2)N(Bn)(2)H)(2)](2+), which have N-H bonds but no Ni-H bonds. For these two isomers, with pendant amines positioned endo to the Ni, the rate constants for proton exchange range from 10(4) to 10(5) s(-1) at 25 °C, depending on isomer and solvent. No exchange is observed for protons on pendant amines positioned exo to the Ni. Analysis of the exchange as a function of temperature provides a barrier for proton exchange of ΔG(‡) = 11-12 kcal/mol for both isomers, with little dependence on solvent. Density functional theory calculations and molecular dynamics simulations support the experimental observations, suggesting metal-mediated intramolecular proton transfers between nitrogen atoms, with chair-to-boat isomerizations as the rate-limiting steps. Because of the fast rate of proton movement, this catalyst may be considered a metal center surrounded by a cloud of exchanging protons. The high intramolecular proton mobility provides information directly pertinent to the ability of pendant amines to accelerate proton transfers during catalysis of hydrogen oxidation. These results may also have broader implications for proton movement in homogeneous catalysts and enzymes in general, with specific implications for the proton channel in the Ni-Fe hydrogenase enzyme.


Journal of the American Chemical Society | 2009

Mechanistic Insights into Catalytic H2 Oxidation by Ni Complexes Containing a Diphosphine Ligand with a Positioned Amine Base

Jenny Y. Yang; R. Morris Bullock; Wendy J. Shaw; Brendan Twamley; Kendra Fraze; M. Rakowski DuBois; Daniel L. DuBois

The mixed-ligand complex [Ni(dppp)(P(Ph)(2)N(Bz)(2))](BF(4))(2), 3, (where P(Ph)(2)N(Bz)(2) is 1,5-dibenzyl-3,7-diphenyl-1,5-diaza-3,7-diphosphacyclooctane and dppp is 1,3-bis(diphenylphosphino)propane) has been synthesized. Treatment of this complex with H(2) and triethylamine results in the formation of the Ni(0) complex, Ni(dppp)(P(Ph)(2)N(Bz)(2)), 4, whose structure has been determined by a single-crystal X-ray diffraction study. Heterolytic cleavage of H(2) by 3 at room temperature forms [HNi(dppp)(P(Ph)(2)N(Bz)(mu-H)N(Bz))](BF(4))(2), 5a, in which one proton interacts with two nitrogen atoms of the cyclic diphosphine ligand and a hydride ligand is bound to nickel. Two intermediates are observed for this reaction using low-temperature NMR spectroscopy. One species is a dihydride, [(H)(2)Ni(dppp)(P(Ph)(2)N(Bz)(2))](BF(4))(2), 5b, and the other is [Ni(dppp)(P(Ph)(2)N(Bz)(2)H(2))](BF(4))(2), 5c, in which both protons are bound to the N atoms in an endo geometry with respect to nickel. These two species interconvert via a rapid and reversible intramolecular proton exchange between nickel and the nitrogen atoms of the diphosphine ligand. Complex 3 is a catalyst for the electrochemical oxidation of H(2) in the presence of base, and new insights into the mechanism derived from low-temperature NMR and thermodynamic studies are presented. A comparison of the rate and thermodynamics of H(2) addition for this complex to related catalysts studied previously indicates that for Ni(II) complexes containing two diphosphine ligands, the activation of H(2) is favored by the presence of two positioned pendant bases.


Inorganic Chemistry | 2011

Studies of a Series of [Ni(PR2NPh2)2(CH3CN)]2+ Complexes as Electrocatalysts for H2 Production: Substituent Variation at the Phosphorus Atom of the P2N2 Ligand

Uriah J. Kilgore; Michael P. Stewart; Monte L. Helm; William G. Dougherty; W. Scott Kassel; Mary Rakowski DuBois; Daniel L. DuBois; R. Morris Bullock

A series of [Ni(P(R)(2)N(Ph)(2))(2)(CH(3)CN)](BF(4))(2) complexes containing the cyclic diphosphine ligands [P(R)(2)N(Ph)(2) = 1,5-diaza-3,7-diphosphacyclooctane; R = benzyl (Bn), n-butyl (n-Bu), 2-phenylethyl (PE), 2,4,4-trimethylpentyl (TP), and cyclohexyl (Cy)] have been synthesized and characterized. X-ray diffraction studies reveal that the cations of [Ni(P(Bn)(2)N(Ph)(2))(2)(CH(3)CN)](BF(4))(2) and [Ni(P(n-Bu)(2)N(Ph)(2))(2)(CH(3)CN)](BF(4))(2) have distorted trigonal bipyramidal geometries. The Ni(0) complex [Ni(P(Bn)(2)N(Ph)(2))(2)] was also synthesized and characterized by X-ray diffraction studies and shown to have a distorted tetrahedral structure. These complexes, with the exception of [Ni(P(Cy)(2)N(Ph)(2))(2)(CH(3)CN)](BF(4))(2), all exhibit reversible electron transfer processes for both the Ni(II/I) and Ni(I/0) couples and are electrocatalysts for the production of H(2) in acidic acetonitrile solutions. The heterolytic cleavage of H(2) by [Ni(P(R)(2)N(Ph)(2))(2)(CH(3)CN)](BF(4))(2) complexes in the presence of p-anisidine or p-bromoaniline was used to determine the hydride donor abilities of the corresponding [HNi(P(R)(2)N(Ph)(2))(2)](BF(4)) complexes. However, for the catalysts with the most bulky R groups, the turnover frequencies do not parallel the driving force for elimination of H(2), suggesting that steric interactions between the alkyl substituents on phosphorus and the nitrogen atom of the pendant amines play an important role in determining the overall catalytic rate.


Angewandte Chemie | 2012

Reversible Electrocatalytic Production and Oxidation of Hydrogen at Low Overpotentials by a Functional Hydrogenase Mimic

Stuart E. Smith; Jenny Y. Yang; Daniel L. DuBois; R. Morris Bullock

An efficient ligand combination: A new bis(diphosphine) nickel(II) complex (see picture) is described. A ΔG° value of 0.84 kcal mol^(−1) for hydrogen addition for this complex was calculated from the experimentally determined equilibrium constant. This complex displayed reversible electrocatalytic activity for hydrogen production and oxidation at low overpotentials, which are characteristic for hydrogenase enzymes.


Chemical Communications | 2010

Hydrogen oxidation catalysis by a nickel diphosphine complex with pendant tert-butyl amines

Jenny Y. Yang; Shentan Chen; William G. Dougherty; W. Scott Kassel; R. Morris Bullock; Daniel L. DuBois; Simone Raugei; Roger Rousseau; Michel Dupuis; M. Rakowski DuBois

A bis-diphosphine nickel complex with tert-butyl functionalized pendant amines [Ni(P(Cy)(2)N(t-Bu)(2))(2)](2+) has been synthesized. It is a highly active electrocatalyst for the oxidation of hydrogen in the presence of base. The turnover rate of 50 s(-1) under 1.0 atm H(2) at a potential of -0.77 V vs. the ferrocene couple is 5 times faster than the rate reported heretofore for any other synthetic molecular H(2) oxidation catalyst.


Journal of the American Chemical Society | 2013

High catalytic rates for hydrogen production using nickel electrocatalysts with seven-membered cyclic diphosphine ligands containing one pendant amine

Michael P. Stewart; Ming Hsun Ho; Stefan Wiese; Mary Lou Lindstrom; Colleen E. Thogerson; Simone Raugei; R. Morris Bullock; Monte L. Helm

A series of Ni-based electrocatalysts, [Ni(7P(Ph)2N(C6H4X))2](BF4)2, featuring seven-membered cyclic diphosphine ligands incorporating a single amine base, 1-para-X-phenyl-3,6-triphenyl-1-aza-3,6-diphosphacycloheptane (7P(Ph)2N(C6H4X), where X = OMe, Me, Br, Cl, or CF3), have been synthesized and characterized. X-ray diffraction studies have established that the [Ni(7P(Ph)2N(C6H4X))2](2+) complexes have a square planar geometry, with bonds to four phosphorus atoms of the two bidentate diphosphine ligands. Each of the complexes is an efficient electrocatalyst for hydrogen production at the potential of the Ni(II/I) couple, with turnover frequencies ranging from 2400 to 27,000 s(-1) with [(DMF)H](+) in acetonitrile. Addition of water (up to 1.0 M) accelerates the catalysis, giving turnover frequencies ranging from 4100 to 96,000 s(-1). Computational studies carried out on the [Ni(7P(Ph)2N(C6H4X))2](2+) family indicate the catalytic rates reach a maximum when the electron-donating character of X results in the pKa of the Ni(I) protonated pendant amine matching that of the acid used for proton delivery. Additionally, the fast catalytic rates for hydrogen production by the [Ni(7P(Ph)2N(C6H4X))2](2+) family relative to the analogous [Ni(P(Ph)2N(C6H4X)2)2](2+) family are attributed to preferred formation of endo protonated isomers with respect to the metal center in the former, which is essential to attain suitable proximity to the reduced metal center to generate H2. The results of this work highlight the importance of precise pKa matching with the acid for proton delivery to obtain optimal rates of catalysis.

Collaboration


Dive into the R. Morris Bullock's collaboration.

Top Co-Authors

Avatar

Daniel L. DuBois

Pacific Northwest National Laboratory

View shared research outputs
Top Co-Authors

Avatar

Simone Raugei

Pacific Northwest National Laboratory

View shared research outputs
Top Co-Authors

Avatar

Monte L. Helm

Pacific Northwest National Laboratory

View shared research outputs
Top Co-Authors

Avatar

David J. Szalda

City University of New York

View shared research outputs
Top Co-Authors

Avatar

Shentan Chen

Pacific Northwest National Laboratory

View shared research outputs
Top Co-Authors

Avatar

Aaron M. Appel

Pacific Northwest National Laboratory

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Eric S. Wiedner

Pacific Northwest National Laboratory

View shared research outputs
Top Co-Authors

Avatar

Jenny Y. Yang

University of California

View shared research outputs
Top Co-Authors

Avatar

Roger Rousseau

Pacific Northwest National Laboratory

View shared research outputs
Researchain Logo
Decentralizing Knowledge