Network


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

Hotspot


Dive into the research topics where Alexander J. M. Miller is active.

Publication


Featured researches published by Alexander J. M. Miller.


Journal of the American Chemical Society | 2010

E-Type Delayed Fluorescence of a Phosphine-Supported Cu2(μ-NAr2)2 Diamond Core: Harvesting Singlet and Triplet Excitons in OLEDs∥

Joseph C. Deaton; Steven C. Switalski; Denis Y. Kondakov; Ralph H. Young; Thomas D. Pawlik; David J. Giesen; Seth B. Harkins; Alexander J. M. Miller; Seth F. Mickenberg; Jonas C. Peters

A highly emissive bis(phosphine)diarylamido dinuclear copper(I) complex (quantum yield = 57%) was shown to exhibit E-type delayed fluorescence by variable temperature emission spectroscopy and photoluminescence decay measurement of doped vapor-deposited films. The lowest energy singlet and triplet excited states were assigned as charge transfer states on the basis of theoretical calculations and the small observed S(1)-T(1) energy gap. Vapor-deposited OLEDs doped with the complex in the emissive layer gave a maximum external quantum efficiency of 16.1%, demonstrating that triplet excitons can be harvested very efficiently through the delayed fluorescence channel. The function of the emissive dopant in OLEDs was further probed by several physical methods, including electrically detected EPR, cyclic voltammetry, and photoluminescence in the presence of applied current.


Journal of the American Chemical Society | 2011

A two-coordinate nickel imido complex that effects C-H amination

Carl A. Laskowski; Alexander J. M. Miller; Gregory L. Hillhouse; Thomas R. Cundari

An exceptionally low coordinate nickel imido complex, (IPr*)Ni═N(dmp) (2) (dmp = 2,6-dimesitylphenyl), has been prepared by the elimination of N2 from a bulky aryl azide in its reaction with (IPr*)Ni(η6-C7H8) (1). The solid-state structure of 2 features two-coordinate nickel with a linear C−Ni−N core and a short Ni−N distance, both indicative of multiple-bond character. Computational studies using density functional theory showed a Ni═N bond dominated by Ni(dπ)−N(pπ) interactions, resulting in two nearly degenerate singly occupied molecular orbitals (SOMOs) that are Ni−N π* in character. Reaction of 2 with CO resulted in nitrene-group transfer to form (dmp)NCO and (IPr*)Ni(CO)3 (3). Net C−H insertion was observed in the reaction of 2 with ethene, forming the vinylamine (dmp)NH(CH═CH2) (5) via an azanickelacyclobutane intermediate, (IPr*)Ni{N,C:κ2-N(dmp)CH2CH2} (4).


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

Kinetic and structural studies, origins of selectivity, and interfacial charge transfer in the artificial photosynthesis of CO

Jonathan M. Smieja; Eric E. Benson; Bhupendra Kumar; Kyle A. Grice; Candace S. Seu; Alexander J. M. Miller; James M. Mayer; Clifford P. Kubiak

The effective design of an artificial photosynthetic system entails the optimization of several important interactions. Herein we report stopped-flow UV-visible (UV-vis) spectroscopy, X-ray crystallographic, density functional theory (DFT), and electrochemical kinetic studies of the Re(bipy-tBu)(CO)3(L) catalyst for the reduction of CO2 to CO. A remarkable selectivity for CO2 over H+ was observed by stopped-flow UV-vis spectroscopy of [Re(bipy-tBu)(CO)3]-1. The reaction with CO2 is about 25 times faster than the reaction with water or methanol at the same concentrations. X-ray crystallography and DFT studies of the doubly reduced anionic species suggest that the highest occupied molecular orbital (HOMO) has mixed metal-ligand character rather than being purely doubly occupied , which is believed to determine selectivity by favoring CO2 (σ + π) over H+ (σ only) binding. Electrocatalytic studies performed with the addition of Brönsted acids reveal a primary H/D kinetic isotope effect, indicating that transfer of protons to Re -CO2 is involved in the rate limiting step. Lastly, the effects of electrode surface modification on interfacial electron transfer between a semiconductor and catalyst were investigated and found to affect the observed current densities for catalysis more than threefold, indicating that the properties of the electrode surface need to be addressed when developing a homogeneous artificial photosynthetic system.


Journal of the American Chemical Society | 2008

Reductive Coupling of Carbon Monoxide in a Rhenium Carbonyl Complex with Pendant Lewis Acids

Alexander J. M. Miller; Jay A. Labinger; John E. Bercaw

Phosphinoborane ligands impart unique reactivity to a rhenium carbonyl cation relative to simple phosphine complexes. Addition of either triethylborohydride or a platinum hydride (that can be formed from H2) forms a rhenium boroxycarbene. This carbene, which crystallizes as a dimer, disproportionates over a period of days to afford the starting cation and a structurally unprecedented boroxy(boroxymethyl)carbene, in which a new C-C bond has been formed between two reduced CO ligands. This product of C-C bond formation can be independently synthesized by addition of 2 equiv of hydride to the rhenium carbonyl cation.


Chemical Communications | 2010

Dehydrogenation of amine–boranes with a frustrated Lewis pair

Alexander J. M. Miller; John E. Bercaw

Bulky tertiary phosphine/borane Lewis pairs P(t)Bu(3)/B(C(6)F(5))(3) react with amine-boranes to afford dehydrocoupling products and phosphonium borohydride salts.


Journal of the American Chemical Society | 2011

Synthesis and Characterization of Three-Coordinate Ni(III)-Imide Complexes

Vlad M. Iluc; Alexander J. M. Miller; John S. Anderson; Marisa J. Monreal; Mark P. Mehn; Gregory L. Hillhouse

A new family of low-coordinate nickel imides supported by 1,2-bis(di-tert-butylphosphino)ethane was synthesized. Oxidation of nickel(II) complexes led to the formation of both aryl- and alkyl-substituted nickel(III)-imides, and examples of both types have been isolated and fully characterized. The aryl substituent that proved most useful in stabilizing the Ni(III)-imide moiety was the bulky 2,6-dimesitylphenyl. The two Ni(III)-imide compounds showed different variable-temperature magnetic properties but analogous EPR spectra at low temperatures. To account for this discrepancy, a low-spin/high-spin equilibrium was proposed to take place for the alkyl-substituted Ni(III)-imide complex. This proposal was supported by DFT calculations. DFT calculations also indicated that the unpaired electron is mostly localized on the imide nitrogen for the Ni(III) complexes. The results of reactions carried out in the presence of hydrogen donors supported the findings from DFT calculations that the adamantyl substituent was a significantly more reactive hydrogen-atom abstractor. Interestingly, the steric properties of the 2,6-dimesitylphenyl substituent are important not only in protecting the Ni═N core but also in favoring one rotamer of the resulting Ni(III)-imide, by locking the phenyl ring in a perpendicular orientation with respect to the NiPP plane.


Chemical Reviews | 2016

Thermodynamic Hydricity of Transition Metal Hydrides

Eric S. Wiedner; Matthew B. Chambers; Catherine L. Pitman; R. Morris Bullock; Alexander J. M. Miller; Aaron M. Appel

Transition metal hydrides play a critical role in stoichiometric and catalytic transformations. Knowledge of free energies for cleaving metal hydride bonds enables the prediction of chemical reactivity, such as for the bond-forming and bond-breaking events that occur in a catalytic reaction. Thermodynamic hydricity is the free energy required to cleave an M-H bond to generate a hydride ion (H(-)). Three primary methods have been developed for hydricity determination: the hydride transfer method establishes hydride transfer equilibrium with a hydride donor/acceptor pair of known hydricity, the H2 heterolysis method involves measuring the equilibrium of heterolytic cleavage of H2 in the presence of a base, and the potential-pKa method considers stepwise transfer of a proton and two electrons to give a net hydride transfer. Using these methods, over 100 thermodynamic hydricity values for transition metal hydrides have been determined in acetonitrile or water. In acetonitrile, the hydricity of metal hydrides spans a range of more than 50 kcal/mol. Methods for using hydricity values to predict chemical reactivity are also discussed, including organic transformations, the reduction of CO2, and the production and oxidation of hydrogen.


Journal of the American Chemical Society | 2008

Probing the Electronic Structures of [Cu2(μ-XR2)]n+ Diamond Cores as a Function of the Bridging X Atom (X = N or P) and Charge (n = 0, 1, 2)

Seth B. Harkins; Neal P. Mankad; Alexander J. M. Miller; Robert K. Szilagyi; Jonas C. Peters

A series of dicopper diamond core complexes that can be isolated in three different oxidation states ([Cu2(mu-XR2)]n+, where n = 0, 1, 2 and X = N or P) is described. Of particular interest is the relative degree of oxidation of the respective copper centers and the bridging XR2 units, upon successive oxidations. These dicopper complexes feature terminal phosphine and either bridging amido or phosphido donors, and as such their metal-ligand bonds are highly covalent. Cu K-edge, Cu L-edge, and P K-edge spectroscopies, in combination with solid-state X-ray structures and DFT calculations, provides a complementary electronic structure picture for the entire set of complexes that tracks the involvement of a majority of ligand-based redox chemistry. The electronic structure picture that emerges for these inorganic dicopper diamond cores shares similarities with the Cu2(mu-SR)2 CuA sites of cytochrome c oxidases and nitrous oxide reductases.


Journal of the American Chemical Society | 2014

Cation-Modulated Reactivity of Iridium Hydride Pincer-Crown Ether Complexes

Matthew R. Kita; Alexander J. M. Miller

Complexes of a new multidentate ligand combining a rigid, strongly donating pincer scaffold with a flexible, weakly donating aza-crown ether moiety are reported. The pincer-crown ether ligand exhibits tridentate, tetradentate, and pentadentate coordination modes. The coordination mode can be changed by Lewis base displacement of the chelating ethers, with binding equilibria dramatically altered through lithium and sodium cation-macrocycle interactions. Cation-promoted hydrogen activation was accomplished by an iridium monohydride cation ligated in a pentadentate fashion by the pincer-crown ether ligand. The rate can be controlled on the basis of the choice of cation (with lithium-containing reactions proceeding about 10 times faster than sodium-containing reactions) or on the basis of the concentration of the cation. Up to 250-fold rate enhancements in H/D exchange rates are observed when catalytic amounts of Li(+) are added.


Inorganic Chemistry | 2010

Thermodynamic Studies of [H2Rh(diphosphine)2]+ and [HRh(diphosphine)2(CH3CN)]2+ Complexes in Acetonitrile

Aaron D. Wilson; Alexander J. M. Miller; Daniel L. DuBois; Jay A. Labinger; John E. Bercaw

Thermodynamic studies of a series of [H(2)Rh(PP)(2)](+) and [HRh(PP)(2)(CH(3)CN)](2+) complexes have been carried out in acetonitrile. Seven different diphosphine (PP) ligands were selected to allow variation of the electronic properties of the ligand substituents, the cone angles, and the natural bite angles (NBAs). Oxidative addition of H(2) to [Rh(PP)(2)](+) complexes is favored by diphosphine ligands with large NBAs, small cone angles, and electron donating substituents, with the NBA being the dominant factor. Large pK(a) values for [HRh(PP)(2)(CH(3)CN)](2+) complexes are favored by small ligand cone angles, small NBAs, and electron donating substituents with the cone angles playing a major role. The hydride donor abilities of [H(2)Rh(PP)(2)](+) complexes increase as the NBAs decrease, the cone angles decrease, and the electron donor abilities of the substituents increase. These results indicate that if solvent coordination is involved in hydride transfer or proton transfer reactions, the observed trends can be understood in terms of a combination of two different steric effects, NBAs and cone angles, and electron-donor effects of the ligand substituents.

Collaboration


Dive into the Alexander J. M. Miller's collaboration.

Top Co-Authors

Avatar

Catherine L. Pitman

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar

Jay A. Labinger

California Institute of Technology

View shared research outputs
Top Co-Authors

Avatar

John E. Bercaw

California Institute of Technology

View shared research outputs
Top Co-Authors

Avatar

Jonas C. Peters

Massachusetts Institute of Technology

View shared research outputs
Top Co-Authors

Avatar

Peter S. White

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Javier Grajeda

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Kelsey R. Brereton

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar

Matthew R. Kita

University of North Carolina at Chapel Hill

View shared research outputs
Researchain Logo
Decentralizing Knowledge