Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 18 de 18
Filter
Add more filters










Publication year range
1.
J Phys Chem A ; 127(34): 7221-7227, 2023 Aug 31.
Article in English | MEDLINE | ID: mdl-37584597

ABSTRACT

Rate constants and product branching fractions were measured for reactions of Ar+, O2+, and NO+ with isoprene (2-methyl-1,3-butadiene C5H8) as a function of temperature. The rate constants are large (∼2 × 10-9 cm3 s-1) and increase with temperature, exceeding the ion-dipole/induced dipole capture rate. Adding a hard sphere term to the collision rate provides a more useful upper limit and predicts the positive temperature dependences. Previous kinetic energy-dependent rate constants show a similar trend. NO+ reacts only by non-dissociative charge transfer. The more energetic O2+ reaction has products formed through both non-dissociative and dissociative charge transfer, or possibly through an H atom transfer. The very energetic Ar+ has essentially only dissociative products; assumption of statistical behavior in the dissociation reasonably reproduces the product branching fractions.

2.
Mass Spectrom Rev ; 2023 Jul 02.
Article in English | MEDLINE | ID: mdl-37394838

ABSTRACT

Starting in the 1960s, flow tube apparatuses have played a central role in the study of ion-molecule kinetics, allowing for immense chemical diversity of cationic, anionic, and neutral reactants. Here, we review studies of oxygen allotropes, excluding ground state O2 ( X 3 ∑ g - ${X}^{3}{\sum }_{g}^{-}$ ), and focusing instead on reactions of cations, anions, and metal chemi-ionization reactions with ground state atomic oxygen (O 3 P), vibrationally excited molecular oxygen (O2 (v)), electronically excited molecular oxygen (O2 ( a 1 Δ g ${a}^{1}{{\rm{\Delta }}}_{g}$ )), and ozone (O3 ). Historical outlines of work over several decades are given along with a focus on more recent work by our group at the Air Force Research Laboratory.

3.
J Phys Chem A ; 126(40): 7202-7209, 2022 Oct 13.
Article in English | MEDLINE | ID: mdl-36169997

ABSTRACT

We report kinetics studies of H3O+(H2O)n=0-3 with isoprene (2-methyl-1,3-butadiene, C5H8) as a function of temperature (300-500 K) measured using a flowing afterglow-selected ion flow tube. Results are supported by density functional (DFT) calculations at the B3LYP/def2-TZVP level. H3O+ (n = 0) reacts with isoprene near the collision limit exclusively via proton transfer to form C5H9+. The first hydrate (n = 1) also reacts at the collision limit and only the proton transfer product is observed, although hydrated protonated isoprene may have been produced and dissociated thermally. Addition of a second water (n = 2) lowers the rate constant by about a factor of 10. The proton transfer of H3O+(H2O)2 to isoprene is endothermic, but transfer of the water ligands lowers the thermicity and the likely process occurring is H3O+(H2O)2 + C5H8 → C5H9+(H2O)2 + H2O, followed by thermal dissociation of C5H9+(H2O)2. Statistical modeling indicates the amount of reactivity is consistent with the process being slightly endothermic, as is indicated by the DFT calculations. This reactivity was obscured in past experiments due to the presence of water in the reaction zone. The third hydrate is observed not to react and helps explain the past results for n = 2, as n = 2 and 3 were in equilibrium in that flow tube experiment. Very little dependence on temperature was found for the three species that did react. Finally, the C5H9+ proton transfer product further reacted with isoprene to produce mainly C6H9+ along with a small amount of clustering.

4.
J Phys Chem A ; 123(13): 2586-2591, 2019 Apr 04.
Article in English | MEDLINE | ID: mdl-30848902

ABSTRACT

Rate constants and product ions were determined for a series of anions reacting with singlet molecular oxygen O2 (a 1Δg) at thermal energy using an electrospray ionization-selected ion flow tube. The 20 naturally occurring amino acids were used to produce corresponding deprotonated anions; only [Cys-H]- and [Pro-H]- were found to be reactive with O2 (a 1Δg), generating OSCH2CH(NH2)CO2- + HO and C5H6NO2- + H2O2, respectively. The reaction of O2 (a 1Δg) with [Cys-H]- has a rate constant more than ten times larger than the reaction of O2 (a 1Δg) with [Pro-H]-. Furthermore, reactions of O2 (a 1Δg) with carboxylic acid and thiol anions were carried out to elucidate the reactivity of the sulfur-containing functional groups. Potential energy surfaces and overall reaction exothermicities were calculated for representative reactions using density functional theory. Reactions in which attack occurs at the sulfur produce HCSO- as an ionic product. Reactions of several carboxylic acid anions likely proceed through a hydroperoxide intermediate that is analogous to that calculated for reactions with amino acid anions at a higher collision energy. Overall, rate constants for reactions of carboxylic acid anions RC(O)O- were found to be smaller for larger R groups.

5.
J Phys Chem A ; 117(6): 1136-43, 2013 Feb 14.
Article in English | MEDLINE | ID: mdl-23145533

ABSTRACT

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.


Subject(s)
Acetonitriles/chemistry , Benzene Derivatives/chemistry , Dimethyl Sulfoxide/chemistry , Fluorides/chemistry , Gases/chemistry , Ions/chemistry , Molecular Structure , Quantum Theory , Solvents/chemistry
6.
J Phys Chem A ; 115(26): 7582-90, 2011 Jul 07.
Article in English | MEDLINE | ID: mdl-21591810

ABSTRACT

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.


Subject(s)
Models, Statistical , Nitric Oxide/chemistry , Water/chemistry , Kinetics
7.
J Phys Chem Lett ; 2(8): 874-9, 2011 Apr 21.
Article in English | MEDLINE | ID: mdl-26295621

ABSTRACT

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.

8.
J Phys Chem A ; 114(28): 7506-8, 2010 Jul 22.
Article in English | MEDLINE | ID: mdl-20578766

ABSTRACT

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.

9.
J Phys Chem A ; 114(25): 6832-6, 2010 Jul 01.
Article in English | MEDLINE | ID: mdl-20524675

ABSTRACT

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.

10.
Eur J Mass Spectrom (Chichester) ; 16(3): 255-68, 2010.
Article in English | MEDLINE | ID: mdl-20530835

ABSTRACT

The 351.1 nm photoelectron spectrum of the peroxyacetate anion, (CH(3)C(O)OO(-)) was measured. Analysis of the spectrum shows that the observed spectral features arise almost exclusively from transitions between the trans-conformer of the anion and the X(2)A'' and A (2)A' states of the corresponding radical. The electron affinity of trans-CH(3)C(O)OO is 2.381+/- 0.007 eV and the term energy splitting of the A (2)A' state is 0.691 +/- 0.009 eV, in excellent agreement with two prior values [Zalyubovsky et al. J. Phys. Chem. A 107, 7704 (2003); Hu et al. J. Phys. Chem. 124, 114305/1 (2006); Hu et al. J. Phys. Chem. 110, 2629 (2006)]. The gas-phase acidity of trans-peroxyacetic acid was bracketed between the acidity of acetic acid and tert-butylthiol at Delta(a)G(298)(trans-CH(3)C(O)OOH)=1439 +/- 14 kJ mol(-1) and Delta(a)H(298)(trans-CH(3)C(O)OOH)=1467+/-14 kJ mol(-1). The acidity of cis-CH(3)C(O)OOH was found by adding a calculated energy correction to the acidity of the trans-conformer; Delta(a)G(298)[cis-CH(3)C(O)OOH] = 1461 +/- 14 kJ mol(-1) and Delta(a)H(298)[cis- CH(3)C(O)OOH]=1490+/-14 kJ mol(-1). The O-H bond dissociation energies for both conformers were determined using a negative ion thermodynamic cycle to be D(0)[trans- CH(3)C(O)OOH]=381+/-14 kJ mol(-1) and D(0)[cis- CH(3)C(O)OOH]=403+/-14 kJ mol(-1). The atmospheric implications of these results and relations to the thermochemistry of peroxyacetyl nitrate are discussed briefly.

11.
J Am Chem Soc ; 132(11): 3808-14, 2010 Mar 24.
Article in English | MEDLINE | ID: mdl-20187638

ABSTRACT

Direct comparisons of the reactivity and mechanistic pathways for anionic systems in the gas phase and in solution are presented. Rate constants and kinetic isotope effects for the reactions of methyl, ethyl, isopropyl, and tert-butyl iodide with cyanide ion in the gas phase, as well as for the reactions of methyl and ethyl iodide with cyanide ion in several solvents, are reported. In addition to measuring the perdeutero kinetic isotope effect (KIE) for each reaction, the secondary alpha- and beta-deuterium KIEs were determined for the ethyl iodide reaction. Comparisons of experimental results with computational transition states, KIEs, and branching fractions are explored to determine how solvent affects these reactions. The KIEs show that the transition state does not change significantly when the solvent is changed from dimethyl sulfoxide/methanol (a protic solvent) to dimethyl sulfoxide (a strongly polar aprotic solvent) to tetrahydrofuran (a slightly polar aprotic solvent) in the ethyl iodide-cyanide ion S(N)2 reaction in solution, as the "Solvation Rule for S(N)2 Reactions" predicts. However, the Solvation Rule fails the ultimate test of predicting gas phase results, where significantly smaller (more inverse) KIEs indicate the existence of a tighter transition state. This result is primarily attributed to the greater electrostatic forces between the partial negative charges on the iodide and cyanide ions and the partial positive charge on the alpha carbon in the gas phase transition state. Nevertheless, in evaluating the competition between S(N)2 and E2 processes, the mechanistic results for the solution and gas phase reactions are strikingly similar. The reaction of cyanide ion with ethyl iodide occurs exclusively by an S(N)2 mechanism in solution and primarily by an S(N)2 mechanism in the gas phase; only approximately 1% of the gas phase reaction is ascribed to an elimination process.

12.
J Phys Chem A ; 114(3): 1270-6, 2010 Jan 28.
Article in English | MEDLINE | ID: mdl-19639975

ABSTRACT

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.


Subject(s)
Oxygen/chemistry , Anions/chemistry , Computer Simulation , Models, Chemical
13.
J Phys Chem A ; 114(1): 191-200, 2010 Jan 14.
Article in English | MEDLINE | ID: mdl-19827803

ABSTRACT

The 351.1 nm photoelectron spectrum of the peroxyformate anion has been measured. The photoelectron spectrum displays vibronic features in both the 2A'' ground and 2A' first excited states of the corresponding radical. Franck-Condon simulations of the spectrum show that the ion is formed exclusively in the trans-conformation. The electron affinity (EA) of the peroxyformyl radical was determined to be 2.493 +/- 0.006 eV, while the term energy splitting was found to be 0.783-0.020+0.060 eV. Extended progressions in the C-OO (973 +/- 20 cm-1) and O-O (1098 +/- 20 cm-1) stretching modes are observed in the ground state of the radical. The fundamental frequency of the in-plane OCO bend was found to be 574 +/- 35 cm-1. The gas-phase acidity of peroxyformic acid has been determined using an ion-molecule bracketing technique. On the basis of the size of the trans- to cis- isomerization barrier, the measured acidity was assigned to the higher energy trans-conformer of the acid. The gas-phase acidity of the lower energy cis-conformer of peroxyformic acid was found from the measured acidity for the trans-form and a calculated energy correction: DeltaaG298(cis-peroxyformic acid) = 346.8 +/- 3.3 kcal mol-1 and DeltaaH298(cis-peroxyformic acid) = 354.4 +/- 3.3 kcal mol-1. Using a negative ion EA/acidity thermochemical cycle, the O-H bond dissociation energy (D0) values of the trans- and cis-conformers of peroxyformic acid to form the trans-radical were determined to be 94.0 +/- 3.3 and 97.1 +/- 3.3 kcal mol-1, respectively. The heat of formation (DeltafH298) of the trans-peroxyformyl radical was found to be -22.8 +/- 3.5 kcal mol-1.


Subject(s)
Formates/chemistry , Thermodynamics , Anions/chemistry , Photoelectron Spectroscopy
14.
J Chem Phys ; 131(9): 094303, 2009 Sep 07.
Article in English | MEDLINE | ID: mdl-19739854

ABSTRACT

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.


Subject(s)
Carbon Dioxide/chemistry , Iron/chemistry , Nitrogen/chemistry , Nitrous Oxide/chemistry , Oxygen/chemistry , Water/chemistry , Carbon Dioxide/metabolism , Computer Simulation , Iron/metabolism , Nitrogen/metabolism , Nitrous Oxide/metabolism , Oxygen/metabolism , Quantum Theory , Water/metabolism
15.
J Am Chem Soc ; 131(23): 8227-33, 2009 Jun 17.
Article in English | MEDLINE | ID: mdl-19456156

ABSTRACT

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.

16.
J Am Chem Soc ; 130(23): 7214-5, 2008 Jun 11.
Article in English | MEDLINE | ID: mdl-18481846

ABSTRACT

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.


Subject(s)
Hydrocarbons, Chlorinated/chemistry , Methane/analogs & derivatives , Anions/chemistry , Gases , Kinetics , Methane/chemistry , Thermodynamics
17.
J Phys Chem A ; 111(35): 8579-87, 2007 Sep 06.
Article in English | MEDLINE | ID: mdl-17696508

ABSTRACT

The 351.1 nm photoelectron spectra of the N-methyl-5-pyrazolide anion and the N-methyl-5-imidazolide anion are reported. The photoelectron spectra of both isomers display extended vibrational progressions in the X2A' ground states of the corresponding radicals that are well reproduced by Franck-Condon simulations, based on the results of B3LYP/6-311++G(d,p) calculations. The electron affinities of the N-methyl-5-pyrazolyl radical and the N-methyl-5-imidazolyl radical are 2.054 +/- 0.006 eV and 1.987 +/- 0.008 eV, respectively. Broad vibronic features of the A(2)A' ' states are also observed in the spectra. The gas-phase acidities of N-methylpyrazole and N-methylimidazole are determined from measurements of proton-transfer rate constants using a flowing afterglow-selected ion flow tube instrument. The acidity of N-methylpyrazole is measured to be Delta(acid)G(298) = 376.9 +/- 0.7 kcal mol(-1) and Delta(acid)H(298) = 384.0 +/- 0.7 kcal mol(-1), whereas the acidity of N-methylimidazole is determined to be Delta(acid)G(298) = 380.2 +/- 1.0 kcal mol(-1) and Delta(acid)H(298)= 388.1 +/- 1.0 kcal mol(-1). The gas-phase acidities are combined with the electron affinities in a negative ion thermochemical cycle to determine the C5-H bond dissociation energies, D(0)(C5-H, N-methylpyrazole) = 116.4 +/- 0.7 kcal mol(-1) and D(0)(C5-H, N-methylimidazole) = 119.0 +/- 1.0 kcal mol(-1). The bond strengths reported here are consistent with previously reported bond strengths of pyrazole and imidazole; however, the error bars are significantly reduced.

18.
J Am Soc Mass Spectrom ; 18(6): 1046-51, 2007 Jun.
Article in English | MEDLINE | ID: mdl-17448673

ABSTRACT

This work describes the first experimental studies of deuterium kinetic isotope effects (KIEs) for the gas-phase E2 reactions of microsolvated systems. The reactions of F(-)(H(2)O)(n) and OH(-)(H(2)O)(n), where n = 0, 1, with (CH(3))(3)CX (X = Cl, Br), as well as the deuterated analogs of the ionic and neutral reactants, were studied utilizing the flowing afterglow-selected ion flow tube technique. The E2 reactivity is found to decrease with solvation. Small, normal kinetic isotope effects are observed for the deuteration of the alkyl halide, while moderately inverse kinetic isotope effects are observed for the deuteration of the solvent. Minimal clustering of the product ions is observed, but there are intriguing differences in the nature and extent of the clustering process. Electronic structure calculations of the transition states provide qualitative insight into these microsolvated E2 reactions.


Subject(s)
Deuterium/chemistry , Gases/chemistry , Solvents/chemistry , Spectrometry, Mass, Electrospray Ionization/methods , Isotope Labeling/methods , Kinetics , Phase Transition
SELECTION OF CITATIONS
SEARCH DETAIL
...