Network


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

Hotspot


Dive into the research topics where Nicole Eyet is active.

Publication


Featured researches published by Nicole Eyet.


Journal of the American Chemical Society | 2009

Reactions of α-Nucleophiles with Alkyl Chlorides: Competition between SN2 and E2 Mechanisms and the Gas-Phase α-Effect

Stephanie M. Villano; Nicole Eyet; W. Carl Lineberger; Veronica M. Bierbaum

Reaction rate constants and deuterium kinetic isotope effects for the reactions of BrO(-) with RCl (R = methyl, ethyl, isopropyl, and tert-butyl) were measured using a tandem flowing afterglow-selected ion flow tube instrument. These results provide qualitative insight into the competition between two classical organic mechanisms, nucleophilic substitution (S(N)2) and base-induced elimination (E2). As the extent of substitution in the neutral reactants increases, the kinetic isotope effects become increasingly more normal, consistent with the gradual onset of the E2 channel. These results are in excellent agreement with previously reported trends for the analogous reactions of ClO(-) with RCl. [Villano et al. J. Am. Chem. Soc. 2006, 128, 728.] However, the reactions of BrO(-) and ClO(-) with methyl chloride, ethyl chloride, and isopropyl chloride were found to occur by an additional reaction pathway, which has not previously been reported. This reaction likely proceeds initially through a traditional S(N)2 transition state, followed by an elimination step in the S(N)2 product ion-dipole complex. Furthermore, the controversial alpha-nucleophilic character of these two anions and of the HO(2)(-) anion is examined. No enhanced reactivity is displayed. These results suggest that the alpha-effect is not due to an intrinsic property of the anion but instead due to a solvent effect.


Journal of the American Chemical Society | 2008

Gas-Phase Carbene Radical Anions: New Mechanistic Insights

Stephanie M. Villano; Nicole Eyet; W. C. Lineberger; Veronica M. Bierbaum

The gas-phase reactivity of the CHCl*- anion has been investigated with a series of halomethanes (CCl4, CHCl3, CH2Cl2, and CH3Cl) using a FA-SIFT instrument. Results show that this anion primarily reacts via substitution and by proton transfer. In addition, the reactions of CHCl*- with CHCl3 and CH2Cl2 form minor amounts of Cl2*- and Cl-. The isotopic distribution of these two products is consistent with an insertion-elimination mechanism, where the anion inserts into a C-Cl bond to form an unstable intermediate, which eliminates either Cl2*- or Cl- and Cl*. Neutral and cationic carbenes are known to insert into single bonds; however, this is the first observation of such reactivity for carbene anions.


Journal of Physical Chemistry A | 2011

The importance of NO(+)(H(2)O)(4) in the conversion of NO(+)(H(2)O)(n) to H(3)O(+)(H(2)O)(n): I. Kinetics measurements and statistical rate modeling.

Nicole Eyet; Nicholas S. Shuman; Albert A. Viggiano; J. Troe; Rachael A. Relph; Ryan P. Steele; Mark A. Johnson

The kinetics for conversion of NO(+)(H(2)O)(n) to H(3)O(+)(H(2)O)(n) has been investigated as a function of temperature from 150 to 400 K. In contrast to previous studies, which show that the conversion goes completely through a reaction of NO(+)(H(2)O)(3), the present results show that NO(+)(H(2)O)(4) plays an increasing role in the conversion as the temperature is lowered. Rate constants are derived for the clustering of H(2)O to NO(+)(H(2)O)(1-3) and the reactions of NO(+)(H(2)O)(3,4) with H(2)O to form H(3)O(+)(H(2)O)(2,3), respectively. In addition, thermal dissociation of NO(+)(H(2)O)(4) to lose HNO(2) was also found to be important. The rate constants for the clustering increase substantially with the lowering of the temperature. Flux calculations show that NO(+)(H(2)O)(4) accounts for over 99% of the conversion at 150 K and even 20% at 300 K, although it is too small to be detectable. The experimental data are complimented by modeling of the falloff curves for the clustering reactions. The modeling shows that, for many of the conditions, the data correspond to the falloff regime of third body association.


Journal of Physical Chemistry Letters | 2011

Reactions of Ions with Ionic Liquid Vapors by Selected-Ion Flow Tube Mass Spectrometry.

Steven D. Chambreau; Jerry A. Boatz; Ghanshyam L. Vaghjiani; Jeffrey F. Friedman; Nicole Eyet; A. A. Viggiano

Room-temperature ionic liquids exert vanishingly small vapor pressures under ambient conditions. Under reduced pressure, certain ionic liquids have demonstrated volatility, and they are thought to vaporize as intact cation-anion ion pairs. However, ion pair vapors are difficult to detect because their concentration is extremely low under these conditions. In this Letter, we report the products of reacting ions such as NO(+), NH4(+), NO3(-), and O2(-) with vaporized aprotic ionic liquids in their intact ion pair form. Ion pair fragmentation to the cation or anion as well as ion exchange and ion addition processes are observed by selected-ion flow tube mass spectrometry. Free energies of the reactions involving 1-ethyl-3-methylimidazolium bis-trifluoromethylsulfonylimide determined by ab initio quantum mechanical calculations indicate that ion exchange or ion addition are energetically more favorable than charge-transfer processes, whereas charge-transfer processes can be important in reactions involving 1-butyl-3-methylimidazolium dicyanamide.


Journal of Physical Chemistry A | 2010

Survey of the Reactivity of O2(a 1Δg) with Negative Ions

Nicole Eyet; Anthony J. Midey; Veronica M. Bierbaum; A. A. Viggiano

The reactivity of O(2)(a (1)Delta(g)) was studied with a series of anions, including (-)CH(2)CN, (-)CH(2)NO(2), (-)CH(2)C(O)H, CH(3)C(O)CH(2)(-), C(2)H(5)O(-), (CH(3))(2)CHO(-), CF(3)CH(2)O(-), CF(3)(-), HC(2)(-), HCCO(-), HC(O)O(-), CH(3)C(O)O(-), CH(3)OC(O)CH(2)(-), and HS(-). Reaction rate constants and product ion branching ratios were measured. All of the carbanions react through a common pathway to produce their major products. O(2)(a) adds across a bond at the site of the negative charge, resulting in the cleavage of this bond and the O=O bond. Oxyanions react through a hydride transfer to produce their major products. Proton transfer within these product ion-dipole complexes can occur, where the final branching ratios reflect the basicity of the resulting anions. Several of these anions (CF(3)(-), HC(2)(-), CH(3)OC(O)CH(2)(-)) were also found to undergo several sequential reactions within a single encounter. These three basic types of mechanisms are supported by calculations; a potential energy diagram for each type of reaction has been calculated. Additionally, six of these reactions had been qualitatively studied before; our results are in agreement with previous data.


Journal of Chemical Physics | 2009

Dissociative excitation transfer in the reaction of O-2(a(1)Delta(g)) with OH-(H2O)(1,2) clusters.

Albert A. Viggiano; Anthony J. Midey; Nicole Eyet; Veronica M. Bierbaum; J. Troe

Rate constants for the dissociation of OH(-)(H(2)O) and OH(-)(H(2)O)(2) by transfer of electronic energy from O(2)(a(1)Delta(g)) were measured. Values of 1.8x10(-11) and 2.2x10(-11) cm(3) molecule(-1) s(-1), respectively, at 300 K were derived and temperature dependences were obtained from 300 to 500 K for OH(-)(H(2)O) and from 300 to 400 K for OH(-)(H(2)O)(2). Dissociative excitation transfer with OH(-)(H(2)O) is slightly endothermic and the reaction appears to have a positive temperature dependence, but barely outside the uncertainty range. In contrast, the reaction of OH(-)(H(2)O)(2) is exothermic and appears to have a negative temperature dependence. The rate constants are analyzed in terms of unimolecular rate theory, which suggests that the dissociation is prompt and is not affected by collisions with the helium buffer gas.


The Astrophysical Journal | 2008

Gas Phase Study of C+ Reactions of Interstellar Relevance

Oscar Martinez; Nicholas B. Betts; Stephanie M. Villano; Nicole Eyet; Theodore P. Snow; Veronica M. Bierbaum

The current uncertainty in many reaction rate constants causes difficulties in providing satisfactory models of interstellar chemistry. Here we present new measurements of the rate constants and product branching ratios for the gas phase reactions of C+ with NH3, CH4, O2, H2O, and C2H2, using the flowing afterglow-selected ion flow tube (FASIFT) technique. Results were obtained using two instruments that were separately calibrated and optimized; in addition, low ionization energies were used to ensure formation of ground-state C+, the purities of the neutral reactants were verified, and mass discrimination was minimized.


Journal of Physical Chemistry A | 2013

Gas-Phase Reactions of Microsolvated Fluoride Ions: An Investigation of Different Solvents

Nicole Eyet; Stephanie M. Villano; Veronica M. Bierbaum

The gas-phase reactions of F(-)(DMSO), F(-)(CH(3)CN), and F(-)(C(6)H(6)) with t-butyl halides were investigated. Reaction rate constants, kinetic isotope effects, and product ion branching ratios were measured using the flowing afterglow selected ion flow tube technique (FA-SIFT). Additionally, the structure of F(-)(DMSO) was investigated both computationally and experimentally, and two stable isomers were identified. The reactions generally proceed by elimination mechanisms; however, the reaction of F(-)(C(6)H(6)) with t-butyl chloride occurs by a switching mechanism. These reactions are compared to previous studies of microsolvated reactions of t-butyl halides where the solvent molecules were polar, protic molecules.


Journal of Physical Chemistry A | 2010

Reexamination of the quenching of NO(+) vibrations by O(2)(a (1)Delta(g)).

Nicole Eyet; Albert A. Viggiano

The quenching of vibrationally excited NO(+) by O(2)(a (1)Delta(g)) has been examined using the monitor ion technique and chemical generation of O(2)(a (1)Delta(g)). In contrast to previous results which showed that the rate constant was much larger than for ground state O(2), this study finds that the rate constant for quenching is below the detection limit (<10(-11) cm(3) s(-1)) of this experiment. The previous experiments produced O(2)(a (1)Delta(g)) in a discharge, which would also produces O atoms. We found that the monitor ion CH(3)I(+) reacts with O atoms to produce CHIOH(+). This is the likely cause of error in the previous experiments.


Journal of Physical Chemistry A | 2010

Reactions of Negative Ions with ClN3 at 300 K

Nicole Eyet; Keith Freel; Michael C. Heaven; Albert A. Viggiano

The reactivity of ClN(3) with 17 negative ions has been investigated at 300 K. The electron affinity (EA) of ClN(3) was bracketed to be between that of NO(2) and N(3), giving EA(ClN(3)) = 2.48 +/- 0.20 eV, in agreement with an electronic structure calculation. Reaction rate constants and product ion branching ratios were measured. In nearly all cases the major product of the reaction was chloride ions. Charge transfer, N(3)(-) production, and O atom incorporation is also observed. DFT calculations of stable complexes and transition states are presented for two typical ions. Mechanistic details are discussed in terms of reaction coordinate diagrams.

Collaboration


Dive into the Nicole Eyet's collaboration.

Top Co-Authors

Avatar

Veronica M. Bierbaum

University of Colorado Boulder

View shared research outputs
Top Co-Authors

Avatar

Stephanie M. Villano

University of Colorado Boulder

View shared research outputs
Top Co-Authors

Avatar

Albert A. Viggiano

Air Force Research Laboratory

View shared research outputs
Top Co-Authors

Avatar

Keith Freel

Air Force Research Laboratory

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

W. Carl Lineberger

University of Colorado Boulder

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Anthony J. Midey

Air Force Research Laboratory

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

John M. Garver

University of Colorado Boulder

View shared research outputs
Researchain Logo
Decentralizing Knowledge