Network


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

Hotspot


Dive into the research topics where Nicholas S. Shuman is active.

Publication


Featured researches published by Nicholas S. Shuman.


Science | 2008

Bond-selective control of a heterogeneously catalyzed reaction.

Daniel R. Killelea; Victoria L. Campbell; Nicholas S. Shuman; A. L. Utz

Energy redistribution, including the many phonon-assisted and electronically assisted energy-exchange processes at a gas-metal interface, can hamper vibrationally mediated selectivity in chemical reactions. We establish that these limitations do not prevent bond-selective control of a heterogeneously catalyzed reaction. State-resolved gas-surface scattering measurements show that the ν1 C-H stretch vibration in trideuteromethane (CHD3) selectively activates C-H bond cleavage on a Ni(111) surface. Isotope-resolved detection reveals a CD3:CHD2 product ratio > 30:1, which contrasts with the 1:3 ratio for an isoenergetic ensemble of CHD3 whose vibrations are statistically populated. Recent studies of vibrational energy redistribution in the gas and condensed phases suggest that other gas-surface reactions with similar vibrational energy flow dynamics might also be candidates for such bond-selective control.


Journal of Physical Chemistry A | 2009

Specific Rate Constants k(E) of the Dissociation of the Halobenzene Ions: Analysis by Statistical Unimolecular Rate Theories

William R. Stevens; Bálint Sztáray; Nicholas S. Shuman; Tomas Baer; Jürgen Troe

Specific rate constants k(E) of the dissociation of the halobenzene ions C6H5X+ --> C6H5+ + X* (X* = Cl, Br, and I) were measured over a range of 10(3)-10(7) s-1 by threshold photoelectron-photoion coincidence (TPEPICO) spectroscopy. The experimental data were analyzed by various statistical unimolecular rate theories in order to derive the threshold energies E0. Although rigid activated complex RRKM theory fits the data in the experimentally measured energy range, it significantly underestimates E0 for chloro- and bromobenzene. Phase space theory (PST) does not fit the experimentally measured rates. A parametrized version of the variational transition state theory (VTST) as well as a simplified version of the statistical adiabatic channel model (SSACM) incorporating an energy dependent rigidity factor provide excellent fits to the experimental data and predict the correct dissociation energies. Although both approaches have just two adjustable parameters, one of which is E0, SSACM is effective and particularly simple to apply.


Advances in Atomic Molecular and Optical Physics | 2012

Teaching an Old Dog New Tricks: Using the Flowing Afterglow to Measure Kinetics of Electron Attachment to Radicals, Ion–Ion Mutual Neutralization, and Electron Catalyzed Mutual Neutralization

Nicholas S. Shuman; Thomas M. Miller; Albert A. Viggiano; Jürgen Troe

Abstract Flowing tube apparatuses have been used to study a variety of electron and ion processes at thermal energies. We describe a new technique, variable electron and neutral density attachment mass spectrometry (VENDAMS), which has enabled measurement of rate coefficients for electron attachment to unstable radical molecules and other transient species and for mutual neutralization reactions of anions with noble gas cations, along with, in favorable cases, determination of neutral products of mutual neutralization. Additionally, the method has yielded evidence of a new ternary reaction in which an electron third body enhances anion–cation neutralization at high plasma densities.


Journal of Chemical Physics | 2011

Pressure and temperature dependence of dissociative and non-dissociative electron attachment to CF3: Experiments and kinetic modeling

Nicholas S. Shuman; Thomas M. Miller; Jeffrey F. Friedman; Albert A. Viggiano; A. I. Maergoiz; Juergen Troe

The kinetics of electron attachment to CF(3) as a function of temperature (300-600 K) and pressure (0.75-2.5 Torr) were studied by variable electron and neutral density attachment mass spectrometry exploiting dissociative electron attachment to CF(3)Br as a radical source. Attachment occurs through competing dissociative (CF(3) + e(-) → CF(2) + F(-)) and non-dissociative channels (CF(3) + e(-) → CF(3)(-)). The rate constant of the dissociative channel increases strongly with temperature, while that of the non-dissociative channel decreases. The rate constant of the non-dissociative channel increases strongly with pressure, while that of the dissociative channel shows little dependence. The total rate constant of electron attachment increases with temperature and with pressure. The system is analyzed by kinetic modeling in terms of statistical theory in order to understand its properties and to extrapolate to conditions beyond those accessible in the experiment.


Journal of Chemical Physics | 2015

Evaluation of the exothermicity of the chemi-ionization reaction Sm + O → SmO+ + e−

Richard M Cox; JungSoo Kim; Joshua H. Bartlett; Robert A. VanGundy; Michael C. Heaven; Shaun G. Ard; Joshua J. Melko; Nicholas S. Shuman; Albert A. Viggiano

The exothermicity of the chemi-ionization reaction Sm + O → SmO(+) + e(-) has been re-evaluated through the combination of several experimental methods. The thermal reactivity (300-650 K) of Sm(+) and SmO(+) with a range of species measured using a selected ion flow tube-mass spectrometer apparatus is reported and provides limits for the bond strength of SmO(+), 5.661 eV ≤ D0(Sm(+)-O) ≤ 6.500 eV. A more precise value is measured to be 5.725 ± 0.07 eV, bracketed by the observed reactivity of Sm(+) and SmO(+) with several species using a guided ion beam tandem mass spectrometer (GIBMS). Combined with the established Sm ionization energy (IE), this value indicates an exothermicity of the title reaction of 0.08 ± 0.07 eV, ∼0.2 eV smaller than previous determinations. In addition, the ionization energy of SmO has been measured by resonantly enhanced two-photon ionization and pulsed-field ionization zero kinetic energy photoelectron spectroscopy to be 5.7427 ± 0.0006 eV, significantly higher than the literature value. Combined with literature bond energies of SmO, this value indicates an exothermicity of the title reaction of 0.14 ± 0.17 eV, independent from and in agreement with the GIBMS result presented here. The evaluated thermochemistry also suggests that D0(SmO) = 5.83 ± 0.07 eV, consistent with but more precise than the literature values. Implications of these results for interpretation of chemical release experiments in the thermosphere are discussed.


Journal of Physical Chemistry A | 2014

Activation of Methane by FeO+: Determining Reaction Pathways through Temperature-Dependent Kinetics and Statistical Modeling

Shaun G. Ard; Joshua J. Melko; V. G. Ushakov; Ryan Johnson; Joseph A. Fournier; Nicholas S. Shuman; Hua Guo; Juergen Troe; Albert A. Viggiano

The temperature dependences of the rate constants and product branching ratios for the reactions of FeO(+) with CH4 and CD4 have been measured from 123 to 700 K. The 300 K rate constants are 9.5 × 10(-11) and 5.1 × 10(-11) cm(3) s(-1) for the CH4 and CD4 reactions, respectively. At low temperatures, the Fe(+) + CH3OH/CD3OD product channel dominates, while at higher temperatures, FeOH(+)/FeOD(+) + CH3/CD3 becomes the majority channel. The data were found to connect well with previous experiments at higher translational energies. The kinetics were simulated using a statistical adiabatic channel model (vibrations are adiabatic during approach of the reactants), which reproduced the experimental data of both reactions well over the extended temperature and energy ranges. Stationary point energies along the reaction pathway determined by ab initio calculations seemed to be only approximate and were allowed to vary in the statistical model. The model shows a crossing from the ground-state sextet surface to the excited quartet surface with large efficiency, indicating that both states are involved. The reaction bottleneck for the reaction is found to be the quartet barrier, for CH4 modeled as -22 kJ mol(-1) relative to the sextet reactants. Contrary to previous rationalizations, neither less favorable spin-crossing at increased energies nor the opening of additional reaction channels is needed to explain the temperature dependence of the product branching fractions. It is found that a proper treatment of state-specific rotations is crucial. The modeled energy for the FeOH(+) + CH3 channel (-1 kJ mol(-1)) agrees with the experimental thermochemical value, while the modeled energy of the Fe(+) + CH3OH channel (-10 kJ mol(-1)) corresponds to the quartet iron product, provided that spin-switching near the products is inefficient. Alternative possibilities for spin switching during the reaction are considered. The modeling provides unique insight into the reaction mechanisms as well as energetic benchmarks for the reaction surface.


Journal of Chemical Physics | 2011

Electron attachment to POCl3. III. Measurement and kinetic modeling of branching fractions

Nicholas S. Shuman; Thomas M. Miller; Albert A. Viggiano; Jürgen Troe

Electron attachment to POCl(3) was studied in the bath gas He over the pressure range 0.4-3.1 Torr and the temperature range 300-1210 K. Branching fractions of POCl(3)(-), POCl(2)(-), Cl(-), and Cl(2)(-) were measured. The results are analyzed by kinetic modeling, using electron attachment theory for the characterization of the nonthermal energy distribution of the excited POCl(3)(-∗) anions formed and chemical activation-type unimolecular rate theory for the subsequent competition between collisional stabilization of POCl(3)(-∗) and its dissociation to various dissociation products. Primary and secondary dissociations and∕or thermal dissociations of the anions are identified. The measured branching fractions are found to be consistent with the modeling results based on molecular parameters obtained from quantum-chemical calculations.


Journal of Physical Chemistry A | 2013

Temperature dependence of the OH- + CH3I reaction kinetics. experimental and simulation studies and atomic-level dynamics

Jing Xie; Swapnil C. Kohale; William L. Hase; Shaun G. Ard; Joshua J. Melko; Nicholas S. Shuman; Albert A. Viggiano

Direct dynamics simulations and selected ion flow tube (SIFT) experiments were performed to study the kinetics and dynamics of the OH(-) + CH3I reaction versus temperature. This work complements previous direct dynamics simulation and molecular beam ion imaging experiments of this reaction versus reaction collision energy (Xie et al. J. Phys. Chem. A 2013, 117, 7162). The simulations and experiments are in quite good agreement. Both identify the SN2, OH(-) + CH3I → CH3OH + I(-), and proton transfer, OH(-) + CH3I → CH2I(-) + H2O, reactions as having nearly equal importance. In the experiments, the SN2 pathway constitutes 0.64 ± 0.05, 0.56 ± 0.05, 0.51 ± 0.05, and 0.46 ± 0.05 of the total reaction at 210, 300, 400, and 500 K, respectively. For the simulations this fraction is 0.56 ± 0.06, 0.55 ± 0.04, and 0.50 ± 0.05 at 300, 400, and 500 K, respectively. The experimental total reaction rate constant is (2.3 ± 0.6) × 10(-9), (1.7 ± 0.4) × 10(-9), (1.9 ± 0.5) × 10(-9), and (1.8 ± 0.5) × 10(-9) cm(3) s(-1) at 210, 300, 400, and 500 K, respectively, which is approximately 25% smaller than the collision capture value. The simulation values for this rate constant are (1.7 ± 0.2) × 10(-9), (1.8 ± 0.1) × 10(-9), and (1.6 ± 0.1) × 10(-9) cm(3)s(-1) at 300, 400, and 500 K. From the simulations, direct rebound and stripping mechanisms as well as multiple indirect mechanisms are identified as the atomic-level reaction mechanisms for both the SN2 and proton-transfer pathways. For the SN2 reaction the direct and indirect mechanisms have nearly equal probabilities; the direct mechanisms are slightly more probable, and direct rebound is more important than direct stripping. For the proton-transfer pathway the indirect mechanisms are more important than the direct mechanisms, and stripping is significantly more important than rebound for the latter. Calculations were performed with the OH(-) quantum number J equal to 0, 3, and 6 to investigate the effect of OH(-) rotational excitation on the OH(-) + CH3I reaction dynamics. The overall reaction probability and the probabilities for the SN2 and proton-transfer pathways have little dependence on J. Possible effects on the atomistic mechanisms were investigated for the SN2 pathway and the probability of the direct rebound mechanism increased with J. However, the other atomistic mechanisms were not appreciably affected by J.


Journal of Physical Chemistry A | 2010

Variable Electron and Neutral Density Attachment Mass Spectrometry: Temperature-Dependent Kinetics of Electron Attachment to PSCl3 and PSCl2 and Mutual Neutralization of PSCl2− and PSCl− with Ar+†

Nicholas S. Shuman; Thomas M. Miller; Connor M. Caples; A. A. Viggiano

We describe the VENDAMS (variable electron and neutral density attachment mass spectrometry) technique to measure the rate constants of various processes occurring as primary, secondary, and higher order chemistry in a flowing afterglow at high charge densities over a temperature range of 300 to 550 K. In particular, we report measurements of rate constants of ion-ion mutual neutralization and electron attachment to radical species, processes which have proven difficult to study through other means. The product negative ion abundances from the addition of PSCl(3) to an Ar(+)/e(-) plasma have been measured as a function of initial electron densities between 1 × 10(8) and 4 × 10(10) cm(-3). Data at lower electron densities yield branching ratios of the primary electron attachment to PSCl(3); determination of the reactions and rate constants occurring at low electron densities then allows for determination of the greater number of reactions and rate constants contributing at higher electron densities. Reaction rate constants and branching ratios of electron attachment to PSCl(2) are reported; this is the first measurement of electron attachment to a radical as a function of temperature. The data show an unusual negative temperature dependence; however, a zero or even slightly positive dependence is within the uncertainty. Measured electron attachment rate constants are 1.4 × 10(-7), 1.1 × 10(-7), and 9.1 × 10(-8) ± 40% cm(3) s(-1) at 300, 400, and 550 K, respectively; the dominant product channel is PSCl + Cl(-) (95, 87, and 77% at 300, 400, and 550 K), and the minor channel is PSCl(-) + Cl. Ion-ion mutual neutralization rate constants of both PSCl(-) and PSCl(2)(-) with Ar(+) are reported over the investigated temperature range; rate constants at 300 K are 4.9 × 10(-8) ± 20% cm(3) s(-1) and 4.5 × 10(-8) ± 15% cm(3) s(-1) and show temperature dependences of T(-0.5±0.3) and T(-0.9±0.3), respectively.


Journal of Physical Chemistry A | 2010

Tunneling in a Simple Bond Scission: The Surprising Barrier in the H Loss from HCOOH+

Nicholas S. Shuman; Melanie Johnson; William R. Stevens; Michael E. Harding; John F. Stanton; Tomas Baer

The dissociation dynamics of gas phase formic acid ions (HCOOH(+), DCOOD(+), HCOOD(+), DCOOH(+)) are investigated by threshold photoelectron-photoion coincidence (TPEPICO) spectroscopy and high level ab initio calculations. The slow rate constants for this seemingly simple H loss reaction and the large onset energy shifts due to isotopic substitution point to a substantial exit barrier through which the H or D atoms tunnel. Modeling of the HCOOH(+) experimental data using RRKM theory with tunneling through an Eckart potential are best fitted with a barrier of about 17 kJ mol(-1). High level ab initio calculations support the experimental findings with a computed barrier of 15.9 kJ mol(-1), which is associated with the substantial geometry change between the product HOCO(+) cation and the corresponding HCOOH(+) molecular ion. Because of this exit channel barrier, the formic acid ion dissociation does not provide a route for determination of the HOCO(+) heat of formation. Rather, the most accurate value comes from the calculations employing the high accuracy extrapolated ab initio thermochemistry (HEAT) scheme, which yields a Δ(f)H(o)(0K)[HOCO(+)] = 600.3 ± 1.0 kJ mol(-1) (Δ(f)H(o)(298K)[HOCO(+)] = 597.3 ± 1.0 kJ mol(-1)). The calculated proton affinity of CO(2) is thus 534.7 ± 1.0 kJ mol(-1) at 0 K and 539.3 ± 1.0 kJ mol(-1) at 298.15 K.

Collaboration


Dive into the Nicholas S. Shuman's collaboration.

Top Co-Authors

Avatar

Albert A. Viggiano

Air Force Research Laboratory

View shared research outputs
Top Co-Authors

Avatar

Shaun G. Ard

University of Tennessee

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Joshua J. Melko

Pennsylvania State University

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Justin P. Wiens

University of Wisconsin-Madison

View shared research outputs
Top Co-Authors

Avatar

Hua Guo

University of New Mexico

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Tomas Baer

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge