To see the other types of publications on this topic, follow the link: Vibrational energy transfer rates.

Journal articles on the topic 'Vibrational energy transfer rates'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Vibrational energy transfer rates.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Pavlov, A. V. "New electron energy transfer and cooling rates by excitation of O2." Annales Geophysicae 16, no. 8 (August 31, 1998): 1007–13. http://dx.doi.org/10.1007/s00585-998-1007-8.

Full text
Abstract:
Abstract. In this work I present the results of a study of the electron cooling rate, the production rates of vibrationally excited O2, and the production frequency of the O2 vibrational quanta arising from the collisions of electrons with O2 molecules as functions of the electron temperature. The electron energy transfer and cooling rates by vibrational excitation of O2 have been calculated and fitted to analytical expressions by use of the revised vibrationally excited O2 cross sections. These new analytical expressions are available to the researcher for quick reference and accurate computer modeling with a minimum of calculations. It is also shown that the currently accepted rate of electron energy loss associated with rotational transitions in O2 must be decreased by a factor of 13.
APA, Harvard, Vancouver, ISO, and other styles
2

Essafi, Stéphanie, and Jeremy N. Harvey. "Rates of Molecular Vibrational Energy Transfer in Organic Solutions." Journal of Physical Chemistry A 122, no. 14 (March 16, 2018): 3535–40. http://dx.doi.org/10.1021/acs.jpca.7b12563.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Adamovich, Igor V., Sergey O. Macheret, J. William Rich, and Charles E. Treanor. "Vibrational Energy Transfer Rates Using a Forced Harmonic Oscillator Model." Journal of Thermophysics and Heat Transfer 12, no. 1 (January 1998): 57–65. http://dx.doi.org/10.2514/2.6302.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Jones, D. B., L. Campbell, M. J. Bottema, and M. J. Brunger. "New electron-energy transfer rates for vibrational excitation of O2." New Journal of Physics 5 (September 25, 2003): 114. http://dx.doi.org/10.1088/1367-2630/5/1/114.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Pavlov, A. V. "New electron energy transfer rates for vibrational excitation of N." Annales Geophysicae 16, no. 2 (1998): 176. http://dx.doi.org/10.1007/s005850050591.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Pavlov, A. V. "New electron energy transfer rates for vibrational excitation of N<sub>2</sub>." Annales Geophysicae 16, no. 2 (February 28, 1998): 176–82. http://dx.doi.org/10.1007/s00585-998-0176-9.

Full text
Abstract:
Abstract. In this paper we present the results of a study of the electron cooling rate, the production rates of vibrationally excited N2(v), and the production frequency of the N2 vibrational quanta arising from the collisions of electrons with unexcited N2(0) and vibrationally excited N2(1) molecules as functions of the electron temperature. The electron energy transfer rates for vibrational excitation of N2 have been calculated and fit to analytical expressions by use of the revised vibrationally excited N2 cross sections. These new analytical expressions are available to the researcher for quick reference and accurate computer modeling with a minimum of calculations.Key words. Atmospheric composition and structure · Thermosphere · Ionosphere · Modeling and forecasting
APA, Harvard, Vancouver, ISO, and other styles
7

Bakos, J. S., P. N. Ign�cz, A. L�rincz, Zs S�rlei, and J. Szigeti. "Measurement of the vibrational energy-transfer rates in mixtures of polyatomic molecules." Applied Physics B Photophysics and Laser Chemistry 57, no. 2 (August 1993): 89–93. http://dx.doi.org/10.1007/bf00425989.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Skinner, J. L., and Kisam Park. "Calculating Vibrational Energy Relaxation Rates from Classical Molecular Dynamics Simulations: Quantum Correction Factors for Processes Involving Vibration−Vibration Energy Transfer†." Journal of Physical Chemistry B 105, no. 28 (July 2001): 6716–21. http://dx.doi.org/10.1021/jp010602k.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Fernández, José A., Alfredo Ortiz de Zárate, María N. Sánchez Rayo, and Fernando Castaño. "Direct Measurements of Removal Rates of CHF(X˜1A′(0,1,0)) by Simple Alkenes." Laser Chemistry 12, no. 1-2 (January 1, 1992): 43–52. http://dx.doi.org/10.1155/lc.12.43.

Full text
Abstract:
Direct measurements of the rate constants for collisional removal of CHF(X˜1A′(0,1,0)) by argon (Ar), ethylene (C2H4), propene (C3H6), 1-butene (1-C4H8), iso-butene (i-C4H8), 1,3-butadiene (C4H6), are reported. CHF(X˜1A′(0,1,0)) was prepared by IRMPD of precursor CH2F2. Fluence independence removal rate constants by Ar, ethylene and propene are: 0.0045, 2.70 and 2.46 × 10−11 cm3 molecule−1 s−1, faster than the rates for the vibrational ground state. 1,3-butadiene was measured at a fluence of 42Jcm−2, yielding a value of 0.81 × 10−11 cm3 molecule−1 s−1. Iso-butene has fluence dependence removal rates, given 1.28 and 0.92 × 10−11 cm3 molecule−1 s−1 for fluences of 50 and 32Jcm−2 respectively. For 1-butene a non-linear Stern-Volmer behaviour has been observed and attributed to fast resonant vibrational-to-vibrational (V-to-V) intermolecular energy transfer. The influence of this (V-to-V) energy transfer on the removal rates is discussed.
APA, Harvard, Vancouver, ISO, and other styles
10

Reid, Korey M., Takahisa Yamato, and David M. Leitner. "Scaling of Rates of Vibrational Energy Transfer in Proteins with Equilibrium Dynamics and Entropy." Journal of Physical Chemistry B 122, no. 40 (September 17, 2018): 9331–39. http://dx.doi.org/10.1021/acs.jpcb.8b07552.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Wang, L., J. R. Xu, and W. E. Jones. "A BOXCARS investigation of vibrational relaxation in highly excited 1, 2-trans-dichloroethene." Canadian Journal of Physics 71, no. 11-12 (November 1, 1993): 547–51. http://dx.doi.org/10.1139/p93-083.

Full text
Abstract:
The BOXCARS technique has been used to study the collisional vibrational energy transfer from 1, 2-trans-dichloroethene excited into a quasicontinuum by a pulsed CO2 laser. The temporal evolution behaviour for vibrational energies in different modes was obtained. It has been shown that both the rate and maximum energy transferral to the ν4 mode are slightly larger than rates and energy transferral to the ν1 and ν2 modes and that this specificity declines with increase in excitation energy. The mechanism for this specificity is discussed.
APA, Harvard, Vancouver, ISO, and other styles
12

Kato, Shuji, Michael J. Frost, Veronica M. Bierbaum, and Stephen R. Leone. "Vibrational specificity for charge transfer versus deactivation in N2+(υ = 0, 1, 2) + Ar and O2 reactions." Canadian Journal of Chemistry 72, no. 3 (March 1, 1994): 625–36. http://dx.doi.org/10.1139/v94-087.

Full text
Abstract:
Rate constants for charge transfer and vibrational deactivation in the N2+(X2Σg+, υ = 0, 1,2) + Ar and O2 reactions are directly measured by a state-resolved optical detection method. The novel, selected-ion flow tube, laser-induced fluorescence (SIFT–LIF) technique is used to study the vibrationally specific reactions at near-thermal collision energy. The total rate constant for N2+(υ = 1,2) + Ar increases by more than a factor of 40 relative to N2+(υ = 0). This enhancement is due exclusively to an increase in the charge transfer channel. The charge transfer rate constants for the N2+(υ) + Ar reaction are found to be almost identical for υ = 1 and υ = 2; this differs slightly from previous results at higher collision energies. The vibrational deactivation rate constant for the N2+(υ = 1) + Ar reaction is measured for the first time; the upper limit for the branching fraction is ≈3%, confirming that this reaction is a useful monitor for N2+(υ > 0). The total rate constant for N2+(υ = 1, 2) + O2 increases by factors of 2.6 and 3.3, respectively, relative to N2+(υ = 0). In contrast to the N2+ + Ar reaction, this enhancement is largely due to the occurrence of vibrational deactivation, which is found to be slightly faster for υ = 2 than for υ = 1. For N2+(υ = 2) + O2, the υ = 2 → 1 and υ = 2 → 0 vibrational deactivation channels are found to occur with comparable rates. The lack of substantial enhancement in the charge transfer channel in the N2+(υ) + O2 reaction by vibrational excitation (up to υ = 2) is in contrast to the observed translational enhancement, which opens a higher lying, endothermic O2+(a4Πu) product channel. These results are consistent with a short-range, curve-crossing mechanism that efficiently channels energy into the O2+(a4Πu) state.
APA, Harvard, Vancouver, ISO, and other styles
13

Wang, Lixin, and W. E. Jones. "A BOXCARS investigation of the interspecies V–V energy transfer from highly excited SF6 to CH4." Canadian Journal of Physics 74, no. 1-2 (January 1, 1996): 34–38. http://dx.doi.org/10.1139/p96-006.

Full text
Abstract:
The BOXCARS technique was used to investigate the V–V energy transfer between highly excited SF6 and CH4. The rates and the amounts of energy transferred to both the ν1 and ν3 modes depend strongly on excitation intensity and partial pressure of SF6 and CH4, and within experimental error, the variation of these quantities in both modes is identical, which is contrary to the situation in other polyatomic molecules. The results indicate that V–T energy transfer in CH4 plays an important role in the relaxation of the excess vibrational energy transferred from SF6 to CH4, and that the intermode V–V energy transfer between the ν1 and ν3 modes of CH4 is much faster than the interspecies V–V energy transfer between highly excited SF6 and CH4.
APA, Harvard, Vancouver, ISO, and other styles
14

Xu, J. R., L. Wang, and W. E. Jones. "A BOXCARS investigation of the interspecies V-V energy transfer from highly excited SF6 to N2O: excitation of ν1 and ν3 modes of N2O." Canadian Journal of Physics 72, no. 3-4 (March 1, 1994): 113–19. http://dx.doi.org/10.1139/p94-018.

Full text
Abstract:
The BOXCARS technique has been used to investigate the vibrational energy transfer between highly excited SF6 and N2O. It was found that the apparent rate and the amount of energy transferred to the ν1 (1285.0 cm−1) and ν3 (2223.5 cm−1) modes of N2O strongly depend on the excitation energy. The amount of energy transferred to the ν1 mode is slightly larger than that transferred to the ν3 mode, while the rate of energy transfer to the ν1 mode is slightly less than that to the ν3 mode. The rates and the amounts of energy transferred to both the ν1 and ν3 modes show greater dependence on the partial pressure of SF6 than on the partial pressure of N2O. A model has been proposed to explain the observed results.
APA, Harvard, Vancouver, ISO, and other styles
15

Simserides, Constantinos, Andreas Morphis, and Konstantinos Lambropoulos. "Hole Transfer in Open Carbynes." Materials 13, no. 18 (September 8, 2020): 3979. http://dx.doi.org/10.3390/ma13183979.

Full text
Abstract:
We investigate hole transfer in open carbynes, i.e., carbon atomic nanowires, using Real-Time Time-Dependent Density Functional Theory (RT-TDDFT). The nanowire is made of N carbon atoms. We use the functional B3LYP and the basis sets 3-21G, 6-31G*, cc-pVDZ, cc-pVTZ, cc-pVQZ. We also utilize a few Tight-Binding (TB) wire models, a very simple model with all sites equivalent and transfer integrals given by the Harrison ppπ expression (TBI) as well as a model with modified initial and final sites (TBImod) to take into account the presence of one or two or three hydrogen atoms at the edge sites. To achieve similar site occupations in cumulenes with those obtained by converged RT-TDDFT, TBImod is sufficient. However, to achieve similar frequency content of charge and dipole moment oscillations and similar coherent transfer rates, the TBImod transfer integrals have to be multiplied by a factor of four (TBImodt4times). An explanation for this is given. Full geometry optimization at the B3LYP/6-31G* level of theory shows that in cumulenes bond length alternation (BLA) is not strictly zero and is not constant, although it is symmetrical relative to the molecule center. BLA in cumulenic cases is much smaller than in polyynic cases, so, although not strictly, the separation to cumulenes and polyynes, approximately, holds. Vibrational analysis confirms that for N even all cumulenes with coplanar methylene end groups are stable, for N odd all cumulenes with perpendicular methylene end groups are stable, and the number of hydrogen atoms at the end groups is clearly seen in all cumulenic and polyynic cases. We calculate and discuss the Density Functional Theory (DFT) ground state energy of neutral molecules, the CDFT (Constrained DFT) “ground state energy” of molecules with a hole at one end group, energy spectra, density of states, energy gap, charge and dipole moment oscillations, mean over time probabilities to find the hole at each site, coherent transfer rates, and frequency content, in general. We also compare RT-TDDFT with TB results.
APA, Harvard, Vancouver, ISO, and other styles
16

GE, SU-HONG, GUANG-XING DONG, XIN-LU CHENG, and GUI-HUA SUN. "DENSITY FUNCTIONAL THEORY STUDY OF THE ENERGY TRANSFER RATES, MOLECULAR SIZE, AND ATOMIZATION ENERGIES OF SOME SECONDARY EXPLOSIVE MOLECULES." Journal of Theoretical and Computational Chemistry 07, no. 01 (February 2008): 81–90. http://dx.doi.org/10.1142/s0219633608003617.

Full text
Abstract:
In this paper, we suggested a theoretical relationship between the property of molecular atomization energy and energy transfer rate in explosive detonation. According to the theory of Dlott and Fayer (J Chem Phys92(6):3798, 1990) some explosives are stable molecules with large energy barriers to chemical reaction in shock or impact initiation, so, a sizable amount of phonon energy must be converted to the molecular internal higher vibrations by multiphonon up pumping. To investigate the relationship between atomization energies and energy transfer rate, the number of doorway modes of explosives is estimated by their theory in which the rate is proportional to the number of normal mode vibrations. We evaluated frequencies of normal mode vibrations of TNB, TNAP, TNA, DATB, TATB, 2,4,6-trinitro-benzylalcohol ( C 7 H 5 N 3 O 7), and TNR by means of density functional theory (DFT) at the B3P86/6-31G(d, p) level. It is found that the number of doorway modes shows a linearly correlation to the atomization energies also calculated by means of DFT at the B3P86/6-31G(d, p) level. Besides, we studied the relation between the number of atoms and atomization energies for these molecules, and confirmed that for those secondary explosives molecules with similar molecular structure and similar molecular weight, the correlation between the atomization energy and the number of doorway modes is higher. This relationship is beneficial to the understanding of the property of explosive in detonation.
APA, Harvard, Vancouver, ISO, and other styles
17

Holmström, E., P. Spijker, and A. S. Foster. "The interface of SrTiO 3 and H 2 O from density functional theory molecular dynamics." Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 472, no. 2193 (September 2016): 20160293. http://dx.doi.org/10.1098/rspa.2016.0293.

Full text
Abstract:
We use dispersion-corrected density functional theory molecular dynamics simulations to predict the ionic, electronic and vibrational properties of the SrTiO 3 /H 2 O solid–liquid interface. Approximately 50% of surface oxygens on the planar SrO termination are hydroxylated at all studied levels of water coverage, the corresponding number being 15% for the planar TiO 2 termination and 5% on the stepped TiO 2 -terminated surface. The lateral ordering of the hydration structure is largely controlled by covalent-like surface cation to H 2 O bonding and surface corrugation. We find a featureless electronic density of states in and around the band gap energy region at the solid–liquid interface. The vibrational spectrum indicates redshifting of the O–H stretching band due to surface-to-liquid hydrogen bonding and blueshifting due to high-frequency stretching vibrations of OH fragments within the liquid, as well as strong suppression of the OH stretching band on the stepped surface. We find highly varying rates of proton transfer above different SrTiO 3 surfaces, owing to differences in hydrogen bond strength and the degree of dissociation of incident water. Trends in proton dynamics and the mode of H 2 O adsorption among studied surfaces can be explained by the differential ionicity of the Ti–O and Sr–O bonds in the SrTiO 3 crystal.
APA, Harvard, Vancouver, ISO, and other styles
18

Stone, John Alfred, Xiaoping Li, and Patricia Anne Turner. "A high pressure mass spectrometric study of proton transfer between tri-, tetra-, penta-, and hexamethylbenzene." Canadian Journal of Chemistry 64, no. 10 (October 1, 1986): 2021–30. http://dx.doi.org/10.1139/v86-334.

Full text
Abstract:
The proton affinities (PA) of some methylaromatic compounds and the rates of proton transfer reactions have been measured using high pressure mass spectrometry. The equilibria studied were of the form [Formula: see text]. Van't Hoff plots yielded the following PA values (kcal mol−1) relative to PA(ethylacetate) = 200.7 as standard: mesitylene 201.0, 1,2,3,5-tetramethylbenzene 203.2, pentamethylbenzene 204.4, hexamethylbenzene 206.6. ΔS0 values for the proton transfer equilibria are not fully determined by changes in rotational symmetry numbers and it is suggested that vibrational and torsional changes must be considered. Proton transfer is slow in both the forward (exothermic) and reverse (endothermic) directions and, in addition, for all except the most endothermic reaction studied (1,2,3,5-tetramethylbenzene + protonated hexamethylbenzene) the rate constants for proton transfer increase with decreasing temperature. Such behaviour can be associated with the potential energy profile along the reaction coordinate for the reaction. A limited number of charge transfer experiments involving molecular ions [Formula: see text] showed that reaction occurs at every collision in the forward (exothermic) direction and the reaction efficiency in the reverse (endothermic) direction is small but increases with increasing temperature.
APA, Harvard, Vancouver, ISO, and other styles
19

NAKANE, Hideaki, and Soji TSUCHIYA. "Modulation technique for determination of vbrational relaxation rates of polyatomic molecules - Intra- and intermode vibrational energy transfers in methyl fluoride." NIPPON KAGAKU KAISHI, no. 8 (1989): 1195–203. http://dx.doi.org/10.1246/nikkashi.1989.1195.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Glover, Starla D., Benjamin J. Lear, J. Catherine Salsman, Casey H. Londergan, and Clifford P. Kubiak. "Electron transfer at the class II/III borderline of mixed valency: dependence of rates on solvent dynamics and observation of a localized-to-delocalized transition in freezing solvents." Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 366, no. 1862 (September 7, 2007): 177–85. http://dx.doi.org/10.1098/rsta.2007.2149.

Full text
Abstract:
The dependence of the rates of intramolecular electron transfer (ET) of mixed-valence complexes of the type {[Ru 3 O(OAc) 6 (CO)(L)] 2 -BL} −1 , where L is the pyridyl ligand and BL is the pyrazine on solvent type and temperature is described. Complexes were reduced chemically to obtain the mixed-valence anions in acetonitrile (CH 3 CN) and methylene chloride (CH 2 Cl 2 ). Rate constants for intramolecular ET were estimated by simulating the observed degree of ν (CO) infrared (IR) bandshape coalescence in the mixed-valence state. In the strongly coupled mixed-valence states of these complexes, the electronic coupling, H AB , approaches λ /2, where λ is the total reorganization energy. The activation energy is thus nearly zero, and rate constants are in the ‘ultrafast’ regime where they depend on the pre-exponential terms within the frequency factor, ν N . The frequency factor contains both external (solvent dynamics) and internal (molecular vibrations) contributions. In general, external solvent motions are slower than internal vibrations, and therefore control ET rates in fluid solution. A profound increase in the degree of ν (CO) IR bandshape coalescence is observed as the temperature approaches the freezing points of the solvents methylene chloride (f.p. −92°C) and acetonitrile (f.p. −44°C). Decoupling the slower solvent motions involved in the frequency factor ν N for ET by freezing the solvent causes a transition from solvent dynamics to internal vibration-limited rates. The solvent phase transition causes a localized-to-delocalized transition in the mixed-valence ions that accelerates the rate of ET.
APA, Harvard, Vancouver, ISO, and other styles
21

Cooper, Bridgette, Maria Tudorovskaya, Sebastian Mohr, Aran O’Hare, Martin Hanicinec, Anna Dzarasova, Jimena Gorfinkiel, et al. "Quantemol Electron Collisions (QEC): An Enhanced Expert System for Performing Electron Molecule Collision Calculations Using the R-Matrix Method." Atoms 7, no. 4 (October 17, 2019): 97. http://dx.doi.org/10.3390/atoms7040097.

Full text
Abstract:
Collisions of low energy electrons with molecules are important for understanding many aspects of the environment and technologies. Understanding the processes that occur in these types of collisions can give insights into plasma etching processes, edge effects in fusion plasmas, radiation damage to biological tissues and more. A radical update of the previous expert system for computing observables relevant to these processes, Quantemol-N, is presented. The new Quantemol Electron Collision (QEC) expert system simplifyies the user experience, improving reliability and implements new features. The QEC graphical user interface (GUI) interfaces the Molpro quantum chemistry package for molecular target setups, and the sophisticated UKRmol+ codes to generate accurate and reliable cross-sections. These include elastic cross-sections, super elastic cross-sections between excited states, electron impact dissociation, scattering reaction rates, dissociative electron attachment, differential cross-sections, momentum transfer cross-sections, ionization cross sections, and high energy electron scattering cross-sections. With this new interface we will be implementing dissociative recombination estimations, vibrational excitations for neutrals and ions, and effective core potentials in the near future.
APA, Harvard, Vancouver, ISO, and other styles
22

Gorski, Alexandr, Barbara Leśniewska, Grażyna Orzanowska, and Jacek Waluk. "Influence of alkyl substituents in corrphycene on geometry, electronic structure, hydrogen bonding, and tautomerization." Journal of Porphyrins and Phthalocyanines 20, no. 01n04 (January 2016): 367–77. http://dx.doi.org/10.1142/s1088424616500140.

Full text
Abstract:
Geometry, electronic and vibrational structure, and relative energies of different tautomeric forms have been calculated for free base corrphycene and its five tetra-, octa-, and dodeca-alkyl-substituted derivatives, of which only one has been synthesized so far. The results demonstrate that the lowest energy structure always corresponds to the transtautomeric form. Comparison with the experimental IR and X-ray data available for 2,3,6,7,11,12,17,18-octaethylcorrphycene shows, in contrast to previous suggestions, that th. transspecies is dominant both in solution and crystalline phase. Based on structural, spectral, and computational data, rates of double hydrogen transfer can be predicted. Tautomerization is expected to occur in a nanosecond or subnanosecond range. The rate should strongly decrease upon substituting parent corrphycene with alkyl groups at the pyrrolic [Formula: see text] positions, whereas the opposite effect is expected for mesosubstitution. Depending on the position of substituent, planar or nonplanar geometries of the corrphycene chromophore are predicted. The nonplanarity leads to a substantial increase of the intensity of the low energy electronic transitions (Q-bands).
APA, Harvard, Vancouver, ISO, and other styles
23

Flynn, George W., Charles S. Parmenter, and Alec M. Wodtke. "Vibrational Energy Transfer." Journal of Physical Chemistry 100, no. 31 (January 1996): 12817–38. http://dx.doi.org/10.1021/jp953735c.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Adelman, S. A., R. Muralidhar, and R. H. Stote. "Theory of vibrational energy relaxation in liquids: Vibrational–vibrational energy transfer." Journal of Chemical Physics 99, no. 2 (July 15, 1993): 1333–39. http://dx.doi.org/10.1063/1.465377.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Miller, David W., and Steven A. Adelman. "Vibrational energy transfer in fluids." International Reviews in Physical Chemistry 13, no. 2 (September 1994): 359–86. http://dx.doi.org/10.1080/01442359409353300.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Spears, Kenneth, Xiaoning Wen, and Steven Arrivo. "Electron Transfer Rates with Vibrational State Resolution." Journal of Physical Chemistry 98, no. 39 (September 1994): 9693–96. http://dx.doi.org/10.1021/j100090a601.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Spears, Kenneth G., Xiaoning Wen, and Ruihua Zhang. "Electron Transfer Rates from Vibrational Quantum States." Journal of Physical Chemistry 100, no. 24 (January 1996): 10206–9. http://dx.doi.org/10.1021/jp960444a.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Howard, Stephen L. "Vibrational energy transfer in symmetric N2 charge transfer." Canadian Journal of Physics 69, no. 5 (May 1, 1991): 584–87. http://dx.doi.org/10.1139/p91-097.

Full text
Abstract:
Results of the crossed-beam investigation of the symmetric charge-transfer reaction of N2+ (X2Σg, ν = 0) with N2 (X1Σg, ν = 0) near 10 eV collision energy showed a symmetrically resonant channel with Δν = 0 as well as a series of inelastically scattered channels. Upon deconvolution to remove the Δν = 0 contribution, the inelastic Δν = 1, 2, and 4 channels were readily observed.
APA, Harvard, Vancouver, ISO, and other styles
29

Egorov, Vladimir V. "Dynamic Symmetry in Dozy-Chaos Mechanics." Symmetry 12, no. 11 (November 11, 2020): 1856. http://dx.doi.org/10.3390/sym12111856.

Full text
Abstract:
All kinds of dynamic symmetries in dozy-chaos (quantum-classical) mechanics (Egorov, V.V. Challenges 2020, 11, 16; Egorov, V.V. Heliyon Physics 2019, 5, e02579), which takes into account the chaotic dynamics of the joint electron-nuclear motion in the transient state of molecular “quantum” transitions, are discussed. The reason for the emergence of chaotic dynamics is associated with a certain new property of electrons, consisting in the provocation of chaos (dozy chaos) in a transient state, which appears in them as a result of the binding of atoms by electrons into molecules and condensed matter and which provides the possibility of reorganizing a very heavy nuclear subsystem as a result of transitions of light electrons. Formally, dozy chaos is introduced into the theory of molecular “quantum” transitions to eliminate the significant singularity in the transition rates, which is present in the theory when it goes beyond the Born–Oppenheimer adiabatic approximation and the Franck–Condon principle. Dozy chaos is introduced by replacing the infinitesimal imaginary addition in the energy denominator of the full Green’s function of the electron-nuclear system with a finite value, which is called the dozy-chaos energy γ. The result for the transition-rate constant does not change when the sign of γ is changed. Other dynamic symmetries appearing in theory are associated with the emergence of dynamic organization in electronic-vibrational transitions, in particular with the emergence of an electron-nuclear-reorganization resonance (the so-called Egorov resonance) and its antisymmetric (chaotic) “twin”, with direct and reverse transitions, as well as with different values of the electron–phonon interaction in the initial and final states of the system. All these dynamic symmetries are investigated using the simplest example of quantum-classical mechanics, namely, the example of quantum-classical mechanics of elementary electron-charge transfers in condensed media.
APA, Harvard, Vancouver, ISO, and other styles
30

Brugmans, Marco J. P., Huib J. Bakker, and Ad Lagendijk. "Direct vibrational energy transfer in zeolites." Journal of Chemical Physics 104, no. 1 (January 1996): 64–84. http://dx.doi.org/10.1063/1.470876.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Uzer, T., and W. H. Miller. "Theories of intramolecular vibrational energy transfer." Physics Reports 199, no. 2 (January 1991): 73–146. http://dx.doi.org/10.1016/0370-1573(91)90140-h.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Gortel, Z. W., P. Piercy, R. Teshima, and H. J. Kreuzer. "Lateral vibrational energy transfer in photodesorption." Surface Science Letters 179, no. 1 (January 1987): A7. http://dx.doi.org/10.1016/0167-2584(87)90252-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Gortel, Z. W., P. Piercy, R. Teshima, and H. J. Kreuzer. "Lateral vibrational energy transfer in photodesorption." Surface Science 179, no. 1 (January 1987): 176–86. http://dx.doi.org/10.1016/0039-6028(87)90128-2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

DeLeon, Robert L., and J. William Rich. "Vibrational energy exchange rates in carbon monoxide." Chemical Physics 107, no. 2-3 (September 1986): 283–92. http://dx.doi.org/10.1016/0301-0104(86)85008-x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Moritsugu, K., O. Miyashita, and A. Kidera. "Vibrational Energy Transfer in a Protein Molecule." Seibutsu Butsuri 41, supplement (2001): S173. http://dx.doi.org/10.2142/biophys.41.s173_4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Cook, J. C., and E. M. McCash. "Vibrational energy-transfer processes in the system." Surface Science 371, no. 2-3 (February 1997): 213–22. http://dx.doi.org/10.1016/s0039-6028(96)01095-3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Coletti, Cecilia, and Gert D. Billing. "Vibrational energy transfer in molecular oxygen collisions." Chemical Physics Letters 356, no. 1-2 (April 2002): 14–22. http://dx.doi.org/10.1016/s0009-2614(02)00279-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Tanaka, Shigenori, Satoshi Itoh, and Noriyuki Kurita. "Vibrational energy transfer in azodendrimer and azobenzene." Chemical Physics 272, no. 2-3 (October 2001): 171–84. http://dx.doi.org/10.1016/s0301-0104(01)00475-x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Barker, John R., and Keith D. King. "Vibrational energy transfer in shock‐heated norbornene." Journal of Chemical Physics 103, no. 12 (September 22, 1995): 4953–66. http://dx.doi.org/10.1063/1.470581.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Moritsugu, Kei, Osamu Miyashita, and Akinori Kidera. "Vibrational Energy Transfer in a Protein Molecule." Physical Review Letters 85, no. 18 (October 30, 2000): 3970–73. http://dx.doi.org/10.1103/physrevlett.85.3970.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Shi, L., F. Li, and J. L. Skinner. "Resonant vibrational energy transfer in ice Ih." Journal of Chemical Physics 140, no. 24 (June 28, 2014): 244503. http://dx.doi.org/10.1063/1.4883913.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Hooper, Joe. "Vibrational energy transfer in shocked molecular crystals." Journal of Chemical Physics 132, no. 1 (January 7, 2010): 014507. http://dx.doi.org/10.1063/1.3273212.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Ree, Jongbaik, Yoo Hang Kim, and Hyung Kyu Shin. "Vibrational Energy Transfer in a Water Chain." Bulletin of the Korean Chemical Society 41, no. 1 (January 2020): 38–47. http://dx.doi.org/10.1002/bkcs.11915.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Breshears, W. D., and I. W. M. Smith. "Vibrational energy transfer in HBr–C2D2 mixtures." Journal of Chemical Physics 83, no. 1 (July 1985): 140–44. http://dx.doi.org/10.1063/1.449806.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Cederbaum, Lorenz S. "Efficient non-resonant intermolecular vibrational energy transfer." Molecular Physics 117, no. 15-16 (May 18, 2018): 1950–55. http://dx.doi.org/10.1080/00268976.2018.1473654.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Schwartz, Steven D. "Vibrational energy transfer from resummed evolution operators." Journal of Chemical Physics 101, no. 12 (December 15, 1994): 10436–41. http://dx.doi.org/10.1063/1.467861.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Kustova, E., and M. Mekhonoshina. "Multi-temperature vibrational energy relaxation rates in CO2." Physics of Fluids 32, no. 9 (September 1, 2020): 096101. http://dx.doi.org/10.1063/5.0021654.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Liu Chuan-Ping, Wang Li, and Zhang Fu-Weng. "Energy transfer and dissipation in vibrational granular bed." Acta Physica Sinica 63, no. 4 (2014): 044502. http://dx.doi.org/10.7498/aps.63.044502.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Chen, Sheah, Xiaoyu Hong, Jeffrey R. Hill, and Dana D. Dlott. "Ultrafast Energy Transfer in High Explosives: Vibrational Cooling." Journal of Physical Chemistry 99, no. 13 (March 1995): 4525–30. http://dx.doi.org/10.1021/j100013a023.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Zenevich, Vladimir A., Gert D. Billing, and Georges Jolicard. "Vibrational–rotational energy transfer in H2–H2 collisions." Chemical Physics Letters 312, no. 5-6 (October 1999): 530–35. http://dx.doi.org/10.1016/s0009-2614(99)00975-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography