Network


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

Hotspot


Dive into the research topics where John A. S. Roberts is active.

Publication


Featured researches published by John A. S. Roberts.


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.


Inorganic Chemistry | 2013

Direct determination of equilibrium potentials for hydrogen oxidation/production by open circuit potential measurements in acetonitrile.

John A. S. Roberts; R. Morris Bullock

Open circuit potentials were measured for acetonitrile solutions of a variety of acids and their conjugate bases under 1 atm H2. Acids examined were triethylammonium, dimethylformamidium, 2,6-dichloroanilinium, 4-cyanoanilinium, 4-bromoanilinium, and 4-anisidinium salts. These potentials, along with the pKa values of the acids, establish the value of the standard hydrogen electrode (SHE) potential in acetonitrile as -0.028(4) V vs the ferrocenium/ferrocene couple. Dimethylformamidium forms homoconjugates and other aggregates with dimethylformamide; open circuit potentials (OCPs) were used to quantify the extent of these reactions. Overpotentials for electrocatalytic hydrogen production and oxidation were determined from open circuit potentials and voltammograms of acidic or basic catalyst solutions under H2. For these solutions, agreement between OCP values and potentials calculated using the Nernst equation is within 12 mV. Use of the measured equilibrium potential allows direct comparison of catalytic systems in different media; it requires neither pKa values, homoconjugation constants, nor the SHE potential.


Journal of the American Chemical Society | 2013

Minimal Proton Channel Enables H2 Oxidation and Production with a Water-Soluble Nickel-Based Catalyst

Arnab Dutta; Sheri Lense; Jianbo Hou; Mark H. Engelhard; John A. S. Roberts; Wendy J. Shaw

Hydrogenase enzymes use first-row transition metals to interconvert H2 with protons and electrons, reactions that are important for the storage and recovery of energy from intermittent sources such as solar, hydroelectric, and wind. Here we present Ni(P(Cy)2N(Gly)2)2, a water-soluble molecular electrocatalyst with the amino acid glycine built into the diphosphine ligand framework. Proton transfer between the outer coordination sphere carboxylates and the second coordination sphere pendant amines is rapid, as observed by cyclic voltammetry and FTIR spectroscopy, indicating that the carboxylate groups may participate in proton transfer during catalysis. This complex oxidizes H2 (1-33 s(-1)) at low overpotentials (150-365 mV) over a range of pH values (0.1-9.0) and produces H2 under identical solution conditions (>2400 s(-1) at pH 0.5). Enzymes employ proton channels for the controlled movement of protons over long distances-the results presented here demonstrate the effects of a simple two-component proton channel in a synthetic molecular electrocatalyst.


Proceedings of the National Academy of Sciences of the United States of America | 2012

Acidic ionic liquid/water solution as both medium and proton source for electrocatalytic H2 evolution by [Ni(P2N2)2]2+ complexes

Douglas H. Pool; Michael P. Stewart; Molly O’Hagan; Wendy J. Shaw; John A. S. Roberts; R. Morris Bullock; Daniel L. DuBois

The electrocatalytic reduction of protons to H2 by (where in the highly acidic ionic liquid dibutylformamidium bis(trifluoromethanesulfonyl)amide shows a strong dependence on added water. A turnover frequency of 43,000–53,000 s-1 has been measured for hydrogen production at 25 °C when the mole fraction of water (χH2O) is 0.72. The same catalyst in acetonitrile with added dimethylformamidium trifluoromethanesulfonate and water has a turnover frequency of 720 s-1. Thus, the use of an ionic liquid/aqueous solution enhances the observed catalytic rate by more than a factor of 50, compared to a similar acid in a traditional organic solvent. Complexes (X = H, OMe,CH2P(O)(OEt)2, Br) are also catalysts in the ionic liquid/water mixture, and the observed catalytic rates correlate with the hydrophobicity of X.


Chemical Communications | 2013

Production of H2 at fast rates using a nickel electrocatalyst in water–acetonitrile solutions

Wesley A. Hoffert; John A. S. Roberts; R. Morris Bullock; Monte L. Helm

We report a synthetic nickel complex containing proton relays, [Ni(P(Ph)2N(C6H4OH)2)2](BF4)2 (P(Ph)2N(C6H4OH)2 = 1,5-bis(p-hydroxyphenyl)-3,7-diphenyl-1,5-diaza-3,7-diphosphacyclo-octane), that catalyzes the production of H2 in aqueous acetonitrile with turnover frequencies of 750-170,000 s(-1) at experimentally determined overpotentials of 310-470 mV.


Proceedings of the National Academy of Sciences of the United States of America | 2014

Amino acid modified Ni catalyst exhibits reversible H2 oxidation/production over a broad pH range at elevated temperatures

Arnab Dutta; Daniel L. DuBois; John A. S. Roberts; Wendy J. Shaw

Significance Enzymes achieve rapid and reversible H2 oxidation catalysis by cooperative behavior between the active site and the protein scaffold. To better understand the role of the enzyme scaffold, we have attached amino acids (glycine, arginine, and arginine methyl ester) to an active functional mimic of hydrogenase to give [Ni(P2CyN2Amino acid)2]2+. The resulting complexes are fully reversible catalysts with the arginine complex exhibiting high activity for both H2 oxidation/production, functionality achieved by the addition of an outer coordination sphere. Hydrogenases interconvert H2 and protons at high rates and with high energy efficiencies, providing inspiration for the development of molecular catalysts. Studies designed to determine how the protein scaffold can influence a catalytically active site have led to the synthesis of amino acid derivatives of [Ni(P2RN2R′)2]2+ complexes, [Ni(P2CyN2Amino acid)2]2+ (CyAA). It is shown that these CyAA derivatives can catalyze fully reversible H2 production/oxidation at rates approaching those of hydrogenase enzymes. The reversibility is achieved in acidic aqueous solutions (pH = 0–6), 1 atm 25% H2/Ar, and elevated temperatures (tested from 298 to 348 K) for the glycine (CyGly), arginine (CyArg), and arginine methyl ester (CyArgOMe) derivatives. As expected for a reversible process, the catalytic activity is dependent upon H2 and proton concentrations. CyArg is significantly faster in both directions (∼300 s−1 H2 production and 20 s−1 H2 oxidation; pH = 1, 348 K, 1 atm 25% H2/Ar) than the other two derivatives. The slower turnover frequencies for CyArgOMe (35 s−1 production and 7 s−1 oxidation under the same conditions) compared with CyArg suggests an important role for the COOH group during catalysis. That CyArg is faster than CyGly (3 s−1 production and 4 s−1 oxidation) suggests that the additional structural features imparted by the guanidinium groups facilitate fast and reversible H2 addition/release. These observations demonstrate that outer coordination sphere amino acids work in synergy with the active site and can play an important role for synthetic molecular electrocatalysts, as has been observed for the protein scaffold of redox active enzymes.


Angewandte Chemie | 2014

Arginine‐Containing Ligands Enhance H2 Oxidation Catalyst Performance

Arnab Dutta; John A. S. Roberts; Wendy J. Shaw

Hydrogenase enzymes use Ni and Fe to oxidize H2 at high turnover frequencies (TOF) (up to 10,000 s(-1)) and low overpotentials (<100 mV). In comparison, the fastest reported synthetic electrocatalyst, [Ni(II)(P(Cy)2N(tBu)2)2](2+), oxidizes H2 at 60 s(-1) in MeCN under 1 atm H2 with an unoptimized overpotential of ca. 500 mV using triethylamine as a base. Here we show that a structured outer coordination sphere in a Ni electrocatalyst enhances H2 oxidation activity: [Ni(II)(P(Cy)2N(Arg)2)2](8+) (Arg=arginine) has a TOF of 210 s(-1) in water with high energy efficiency (180 mV overpotential) under 1 atm H2 , and 144,000 s(-1) (460 mV overpotential) under 133 atm H2. The complex is active from pH 0-14 and is faster at low pH, the most relevant condition for fuel cells. The arginine substituents increase TOF and may engage in an intramolecular guanidinium interaction that assists in H2 activation, while the COOH groups facilitate rapid proton movement. These results emphasize the critical role of features beyond the active site in achieving fast, efficient catalysis.


Inorganic Chemistry | 2015

Standard Reduction Potentials for Oxygen and Carbon Dioxide Couples in Acetonitrile and N,N-Dimethylformamide

Michael L. Pegis; John A. S. Roberts; Derek J. Wasylenko; Elizabeth A. Mader; Aaron M. Appel; James M. Mayer

A variety of next-generation energy processes utilize the electrochemical interconversions of dioxygen and water as the oxygen reduction reaction (ORR) and the oxygen evolution reaction (OER). Reported here are the first estimates of the standard reduction potential of the O2 + 4e(-) + 4H(+) ⇋ 2H2O couple in organic solvents. The values are +1.21 V in acetonitrile (MeCN) and +0.60 V in N,N-dimethylformamide (DMF), each versus the ferrocenium/ferrocene couple (Fc(+/0)) in the respective solvent (as are all of the potentials reported here). The potentials have been determined using a thermochemical cycle that combines the free energy for transferring water from aqueous solution to organic solvent, -0.43 kcal mol(-1) for MeCN and -1.47 kcal mol(-1) for DMF, and the potential of the H(+)/H2 couple, - 0.028 V in MeCN and -0.662 V in DMF. The H(+)/H2 couple in DMF has been directly measured electrochemically using the previously reported procedure for the MeCN value. The thermochemical approach used for the O2/H2O couple has been extended to the CO2/CO and CO2/CH4 couples to give values of -0.12 and +0.15 V in MeCN and -0.73 and -0.48 V in DMF, respectively. Extensions to other reduction potentials are discussed. Additionally, the free energy for transfer of protons from water to organic solvent is estimated as +14 kcal mol(-1) for acetonitrile and +0.6 kcal mol(-1) for DMF.


Inorganic Chemistry | 2013

Synthesis and electrochemical studies of cobalt(III) monohydride complexes containing pendant amines.

Eric S. Wiedner; John A. S. Roberts; William G. Dougherty; W. Scott Kassel; Daniel L. DuBois; R. Morris Bullock

Two new tetraphosphine ligands, P(nC-PPh2)2N(Ph)2 (1,5-diphenyl-3,7-bis((diphenylphosphino)alkyl)-1,5-diaza-3,7-diphosphacyclooctane; alkyl = (CH2)2, n = 2 (L2); (CH2)3, n = 3 (L3)), have been synthesized. Coordination of these ligands to cobalt affords the complexes [Co(II)(L2)(CH3CN)](2+) and [Co(II)(L3)(CH3CN)](2+), which are reduced by KC8 to afford [Co(I)(L2)(CH3CN)](+) and [Co(I)(L3)(CH3CN)](+). Protonation of the Co(I) complexes affords [HCo(III)(L2)(CH3CN)](2+) and [HCo(III)(L3)(CH3CN)](2+). The cyclic voltammetry of [HCo(III)(L2)(CH3CN)](2+), analyzed using digital simulation, is consistent with an ErCrEr reduction mechanism involving reversible acetonitrile dissociation from [HCo(II)(L2)(CH3CN)](+) and resulting in formation of HCo(I)(L2). Reduction of HCo(III) also results in cleavage of the H-Co bond from HCo(II) or HCo(I), leading to formation of the Co(I) complex [Co(I)(L2)(CH3CN)](+). Under voltammetric conditions, the reduced cobalt hydride reacts with a protic solvent impurity to generate H2 in a monometallic process involving two electrons per cobalt. In contrast, under bulk electrolysis conditions, H2 formation requires only one reducing equivalent per [HCo(III)(L2)(CH3CN)](2+), indicating a bimetallic route wherein two cobalt hydride complexes react to form 2 equiv of [Co(I)(L2)(CH3CN)](+) and 1 equiv of H2. These results indicate that both HCo(II) and HCo(I) can be formed under electrocatalytic conditions and should be considered as potential catalytic intermediates.


Energy and Environmental Science | 2014

Electrocatalytic H2 production with a turnover frequency >107 s−1: the medium provides an increase in rate but not overpotential

Jianbo Hou; Ming Fang; Allan Jay P. Cardenas; Wendy J. Shaw; Monte L. Helm; R. Morris Bullock; John A. S. Roberts; Molly O'Hagan

Rapid proton movement results in exceptionally fast electrocatalytic H2 production (up to 3 × 107 s−1) at overpotentials of ∼400 mV when catalysed by [Ni(PPh2NC6H4x2)2]2+ complexes in an acidic ionic liquid–water medium ([(DMF)H]NTf2–H2O, χH2O = 0.71).

Collaboration


Dive into the John A. S. Roberts's collaboration.

Top Co-Authors

Avatar

R. Morris Bullock

Pacific Northwest National Laboratory

View shared research outputs
Top Co-Authors

Avatar

Wendy J. Shaw

Pacific Northwest National Laboratory

View shared research outputs
Top Co-Authors

Avatar

Daniel L. DuBois

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

Aaron M. Appel

Pacific Northwest National Laboratory

View shared research outputs
Top Co-Authors

Avatar

Arnab Dutta

Pacific Northwest National Laboratory

View shared research outputs
Top Co-Authors

Avatar

Sheri Lense

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

Mark H. Engelhard

Environmental Molecular Sciences Laboratory

View shared research outputs
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge