Download as pdf or txt
Download as pdf or txt
You are on page 1of 41

Unimolecular kinetics.

Parts 2 and 3,
Collisional energy transfer and the
master equation - eBook PDF
Visit to download the full and correct content document:
https://ebooksecure.com/download/unimolecular-kinetics-parts-2-and-3-collisional-en
ergy-transfer-and-the-master-equation-ebook-pdf/
Elsevier
Radarweg 29, PO Box 211, 1000 AE Amsterdam, Netherlands
The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom
50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States

© 2019 Elsevier B.V. All rights reserved.

No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical,
including photocopying, recording, or any information storage and retrieval system, without permission in writing
from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies
and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency,
can be found at our website: www.elsevier.com/permissions.
This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as
may be noted herein).

Notices
Knowledge and best practice in this field are constantly changing. As new research and experience broaden our
understanding, changes in research methods, professional practices, or medical treatment may become necessary.
Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any
information, methods, compounds, or experiments described herein. In using such information or methods they
should be mindful of their own safety and the safety of others, including parties for whom they have a professional
responsibility.
To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for
any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any
use or operation of any methods, products, instructions, or ideas contained in the material herein.

Library of Congress Cataloging-in-Publication Data


A catalog record for this book is available from the Library of Congress

British Library Cataloguing-in-Publication Data


A catalogue record for this book is available from the British Library
ISBN: 978-0-444-64207-3
ISSN: 0069-8040

For information on all Elsevier publications visit our


website at https://www.elsevier.com/books-and-journals

Publisher: Susan Dennis


Acquisition Editor: Anneka Hess
Editorial Project Manager: Katerina Zaliva
Production Project Manager: Omer Mukthar
Cover Designer: Alan Studholme
Typeset by SPi Global, India
CONTENTS

COfltnbulorl ,
,

�l'f{.I(� •

Part 2. Collision Energy Transfer 1

1. Experiments on collisional energy transfer 3

Keith 0. King. John R. Barker

I. ln"l'C)(f,i("hOn 4
2. """'' 5
3. �".:t!O• l;"J:petiment C�l('C'JOl'iCS 12
4. Some COl)(lu� dr3Vl'fl from direct e�men1s 52
s. Conclvdl.•..g fenlatks 56
Actrlo\vledgn'lef'"I'> S7
Rt:feri>ores SI

2. Quantum scattering theory for collisional energy transfer 63


G)l'Orgy lendvay, Geo19e C. Schi112

I. !n�rod.x:oon 64
2. Wicthods 65
). Cc'llcu!.,tiOllS of tilt- s t<1
�e resolvtdl"\(.'f<JY 11,,nsfCT prcperties 74
4. Cakv!<tbOl'ls in ttie higt-.-eneigy 1egime %

s. Con(lu!>IOn •04

�O()\V!«jgmet·b 'OS

Rtferenct:s '00

3. Cla$sical trajectory studies of collisional energy trans.fer 109


Gy()rgy Len<tv-'Y

1. ln1rocloc:t()n 10
2. Method! ll
l. Rc'SultS o l tr.ij('(lfJIY '>l00ics of�oc;llviciMI t'X<.il«i' nlOk.'CVto'. CoU1dcr P-'�'> '9

4. Cner9y lr.>ntr
i picb.lbihty diltribvtiOM 2J7
s. AAonylxxfy >f'lul.)tJOns <I.eri�mblc �cl.a:<dt10n 247
6. Ef'll"IJY lt.>n�t." In Colli)i()n-ioduct.'d dis'SOCidtiOn 250
7. The mechanism of ene1gy uansfe1 2S•

8. 51.lmrn.)I)' (Ind W'!lool\ 261

Ac'i;l'lO\\'ledgment:s 264
Ref�e•'ICe� 7(>4

v
CONTRIBUTORS

John R. Barker
Department of Climate and Space Sciences & Engineering, University of Michigan, Ann Arbor,
MI, United States of America
Terry J. Frankcombe
School of Physical, Environmental and Mathematical Sciences, University of New South Wales,
Canberra, ACT, Australia
Nicholas J.B. Green
Physical and Theoretical Chemistry Laboratory, Oxford University, Oxford, United Kingdom
Keith D. King
School of Chemical Engineering, University of Adelaide, Adelaide, SA, Australia
Gy€ orgy Lendvay
Institute of Materials and Environmental Chemistry, Research Centre for Natural Sciences,
Hungarian Academy of Sciences, Budapest, Hungary
Struan H. Robertson
Dassault Systèmes Ltd., Cambridge Science Park, Cambridge, United Kingdom
George C. Schatz
Department of Chemistry, Northwestern University, Evanston, IL, United States of America
Sean C. Smith
Research School of Physics and Engineering, The Australian National University, Canberra,
ACT, Australia

ix
PART 2

Collision Energy Transfer


CHAPTER 1

Experiments on collisional energy


transfer
Keith
a
D. Kinga, John R. Barkerb
School of Chemical Engineering, University of Adelaide, Adelaide, SA, Australia
b
Department of Climate and Space Sciences & Engineering, University of Michigan, Ann Arbor, MI, United States of America

Contents
1. Introduction 4
2. Theory 5
2.1 Master equation 5
2.2 Energy transfer quantities 6
2.3 Steady-state unimolecular reactions 7
2.4 Time-dependent systems 9
2.4.1 Shock-wave excitation and reaction 9
2.4.2 Photoactivation 12
3. Major experiment categories 12
3.1 Thermal activation 12
3.1.1 Steady-state methods 13
3.1.2 Shock wave methods 17
3.1.3 Energy transfer in shocks 19
3.2 Chemical activation 21
3.2.1 Chemical mechanisms 22
3.2.2 Energy distribution 23
3.2.3 Measured quantities and energy transfer 24
3.3 Photoactivation 26
3.3.1 Infrared multiphoton excitation 26
3.3.2 Photophysical radiationless electronic transitions 31
3.4 Time-resolved infrared fluorescence 33
3.4.1 Theory and calibration 33
3.4.2 Experimental methods 34
3.4.3 Vibrational deactivation 35
3.5 Ultraviolet absorption 38
3.5.1 Principles and calibration 38
3.5.2 Experimental methods 41
3.5.3 Vibrational deactivation 41
3.6 Kinetically controlled selective ionization 41
3.6.1 Principles and experimental techniques 41
3.6.2 Experimental details 45
3.6.3 Extraction of vibrational energy transfer parameters 46

Comprehensive Chemical Kinetics, Volume 43 © 2019 Elsevier B.V.


ISSN 0069-8040, https://doi.org/10.1016/B978-0-444-64207-3.00001-9 All rights reserved. 3
4 Unimolecular Kinetics

3.7 Time-resolved tunable diode laser absorption spectroscopy 47


3.7.1 Principles and experimental techniques 48
3.7.2 Energy transfer categories and results 50
3.8 Crossed molecular beams 51
4. Some conclusions drawn from direct experiments 52
4.1 Collision frequency 52
4.2 Energy transfer mechanisms 53
4.3 Collision step-size distribution, P(E,E0 ) 54
5. Concluding remarks 56
Acknowledgments 57
References 57

1. Introduction
Unimolecular reactions are particularly fascinating because their rate constants depend on
pressure. The pressure dependence comes about because of collisional energy transfer,
which is required for activation and deactivation of vibrationally excited reactants. The
transition from the low-pressure limit, where energy transfer is rate limiting, to the
high-pressure limit, where energy transfer is only implicit, introduces a complexity not
found in simple bimolecular reactions. It also makes predictions much more challenging.
A goal of many researchers is to make accurate theoretical predictions of reaction
rates. Quantum chemical methods are now capable of predicting reaction thermo-
chemistry and molecular structures that can be used with Transition State Theory
(Chapter 2 of Part 1) or the Adiabatic Channel Model to make reasonably accurate pre-
dictions of k(E), the specific rate constant for dissociation as a function of internal
energy (E). The weak link in making routine predictions of unimolecular reaction rate
constants is the current inability to predict energy transfer properties accurately [1].
Theoretical methods for calculating energy transfer are described in Chapters 2 and
3 of this volume, but all of the methods are cumbersome and impractical for routine
predictions. The semiempirical models described in Chapter 4 are much less cumber-
some to use, but they rely on experimental data. When empirical energy transfer models
are combined with quantum chemistry calculations and theoretical predictions for k(E),
it is possible to make quite accurate predictions of unimolecular reaction rate constants
[2], as demonstrated recently by the research groups of Green [3,4] and Lin [5–10]. One
must conclude that experimental information on energy transfer will be required for
some time to come.
Historically, energy transfer has been associated with unimolecular reaction systems
ever since Lindemann [11] and Christiansen [12] realized that the activation process could
be considered separately from the dissociation process. Several major reviews have
appeared during the past three decades [13–16]. The early experiments employed thermally
Experiments on collisional energy transfer 5

activated unimolecular reactions, from which relative collision efficiencies were derived.
The advent of the chemical activation method constituted a major advance, because it
enabled the activation process to be divorced from the collider bath temperature and it
results in much narrower initial energy distribution functions. Photoactivation carries with
it the same advantages. All of the chemical methods are subject to complications due to
secondary reactions and complex chemical mechanisms. The next major advance arrived
in the early 1980s when nonchemical means such as spontaneous infrared fluorescence and
light absorption from lamps or lasers were used to monitor vibrational energy losses in
highly excited molecules. The most current technique is a pump-probe method that
exploits multiphoton ionization. Each of these advances has produced important improve-
ments in experimental measurements of collisional energy transfer involving highly excited
molecules.
In this chapter, the principal experimental methods for determining energy transfer
parameters appropriate for unimolecular reactions are reviewed. The older methods are
included for completeness and because they are the basis for subsequent developments.
There are many variations and combinations of the basic methods reviewed here.
Omitted from the review are various novel methods that have been described in the
literature, but have not been widely used. The chapter is organized roughly historically,
and it is clear that even in the earliest days, a great deal of ingenuity has been brought to
bear on the enduring problem of energy transfer in unimolecular reactions.

2. Theory
2.1 Master equation
The master equation can be written as follows:
ð∞
dyðE 0 , tÞ 0
dE ¼ f ðE0 , tÞdE 0 + ω dE ½P ðE 0 , E ÞdE0 yðE, tÞ
dt 0
(1.1)
X
channels
0 0 0 0 0
 ωyðE , tÞdE  km ðE ÞyðE , tÞdE
m¼1

where y(E0 ,t)dE0 is the concentration of species with vibrational energy in the range E0 to
E0 +dE0 , ω is the collision rate for inelastic vibrational energy transfer (VET), P(E0 ,E) is the
probability for a collisional transition from energy range E0 to E0 + dE0 to the new energy
range E + dE, f(E0 ,t)dE0 is a source term (e.g., chemical or photoactivation), and km(E0 ) is
the unimolecular reaction rate constant for molecules at energy E0 reacting via the mth
channel. Terms involving radiative emission and absorption [17] have been neglected.
The step-size distributions for activating and deactivating collisions are related via
detailed balance [13,14,16]:
6 Unimolecular Kinetics

 
P ðE, E 0 Þ ρðEÞ E  E0
¼ exp  , (1.2)
P ðE0 , EÞ ρðE 0 Þ kB T
where ρ(E) is the density of states at energy E and kB is the Boltzmann constant. Because
of this relationship, it is convenient and efficient to specify collision models only in terms
of the deactivating collisions.

2.2 Energy transfer quantities


The step-size distribution, P(E,E0 ), contains all of the information about energy transfer.
Although several simple theories have been proposed, there is no complete theory for
describing P(E,E0 ) and, hence, P(E,E0 ) has been described by empirical models that
are defined in terms of deactivating collisions (i.e., for E0 > E); for activating collisions,
the distribution is found by application of detailed balance Eq. (1.2). The strong collision
model has been very important in the development of unimolecular reaction rate theory
[18–20]. It consists of assuming that every collision results in the establishment of a
Boltzmann vibrational distribution:
 
0 1 E
Psc ðE, E Þ ¼ ρðEÞ exp , (1.3)
N ðE 0 Þ kB T
The two other models that are most popular are the exponential and bi-exponential models:
 
1 E0  E
Pe ðE, E0 Þ ¼ exp  (1.4)
N ðE 0 Þ αðE 0 Þ
    
1 E0  E E0  E
Pbe ðE, E 0 Þ ¼ a1 exp  + ð1  a 1 Þexp  (1.5)
N ðE 0 Þ α1 ðE 0 Þ α2 ðE0 Þ
A generalized exponential model (similar to one employed by King, Gilbert, and coworkers
[21]) provides the best current description of energy transfer data [22,23]:
  0 
1 E  Eγ
0
Pge ðE, E Þ ¼ exp   (1.6)
N ðE 0 Þ αðE 0 Þ 
In all of these expressions, the distribution function is written for deactivating collisions,
N(E0 ) is the normalization factor, αι(E0 ) is a polynomial function of vibrational energy,
and γ is a parameter that ranges from 0.5 to 1.5. When γ < 1, the “wings” of the step-
size distribution have enhanced relative probabilities that qualitatively resemble a
bi-exponential distribution. The venerable exponential model is obtained from the
above expression when γ ¼ 1. The exponential model, which is mathematically very
convenient, has been used for many years and is still a very good choice when making
estimates.
Experiments on collisional energy transfer 7

In the exponential model, the parameter α is closely related to the average energy
transferred in deactivating collisions, hΔ E(E0 )idown, which is convenient for summarizing
the energy transfer effectiveness:
ð E0
ðE0  E ÞP ðE, E0 ÞdE
0
hΔEðE Þidown ¼ 0
ð E0 (1.7)
0
P ðE, E ÞdE
0
0 0
Note that hΔE(E )idown differs from hΔE(E )iall, a second measure of energy transfer
effectiveness, in that for the latter quantity, the sign of the expression is reversed and
the upper integration limits are extended to ∞.
ð∞
ðE  E0 ÞP ðE, E0 ÞdE
hΔEðE0 Þiall ¼ 0 ð ∞ (1.8)
P ðE, E0 ÞdE
0
0
Note that hΔE(E )iall includes the temperature-dependence associated with activation
collisions (see Eq. 1.2). In comparisons with experimental data, hΔE(E0 )iall is often more
useful than hΔE(E0 )idown, but the latter is more useful when comparing data obtained at
different temperatures; the latter is also more useful in master equation models.
For the exponential model, the relationship [24] between the energy transfer param-
eter α (which may be energy-dependent) and hΔEiall is obtained with good accuracy
by using the Whitten–Rabinovitch approximation [25,26] for densities of states. The
resulting expression is
1
hΔEiall ¼ α (1.9)
C
where
1 1 s + r=2  1
C¼ +  (1.10)
α kB T E + aðE ÞEz
In these expressions, s and r are the numbers of vibrational and rotational degrees of free-
dom, respectively, Ez is the zero point energy, E is the rovibrational energy in excess of
Ez, and a(E) is the Whitten–Rabinovitch parameter [25,26].

2.3 Steady-state unimolecular reactions


In both the thermal and the chemical activation systems, reaction rates are most sensitive
to energy transfer at low pressures, where the greatest deviations from the Boltzmann
distribution are achieved. Consider the following simplified Lindemann-like mechanism
8 Unimolecular Kinetics

involving molecule A, which can be collisionally activated to energy E in a thermal


system by collisions with M:
ωBðEÞ
A + M ! AðEÞ + M (1.11)
ω
AðE Þ + M ! A + M (1.12)
kðEÞ
AðE Þ ! products (1.13)
For strong collisions and steady state, the relative population distribution F(E) is propor-
tional to [A(E)]ss, the steady-state concentration of excited molecules:
½AðEÞss ωBðEÞ
F ðE Þ ¼ ¼ (1.14)
½A ω + kðE Þ
where ω is the collision frequency, B(E) is the Boltzmann distribution, and k(E) is the
energy-dependent specific unimolecular rate constant. At the high-pressure limit
(ω ¼ ∞), the population distribution above the reaction threshold is Boltzmann, while
at low pressures (ω ≪ k(E)) it is depleted by an amount that depends on k(E), which is
a strong near-exponential function of vibrational energy.
The enhanced reaction yield found in a low-pressure chemical activation system is
evident from the following simplified mechanism:
kca
C + D ! AðEÞ (1.15)
kðE Þ
AðEÞ ! G (1.16)
ω
AðE Þ ! A (1.17)
where C and D are initial reactants (e.g., a free radical and a molecule) that react with
bimolecular rate constant kca, G is a product, and the other quantities are as defined above.
For this system, the steady-state population distribution Fca(E) is proportional to the
steady-state concentration of excited molecules:
kca ½C ½D
Fca ðEÞ∝ ½AðE Þss ¼ (1.18)
kðE Þ + ω
It is clear from this expression and from the mechanism that the steady-state population at
high energy and the resulting yield of product G are larger at lower pressures.
Eqs. (1.14) and (1.18) are for strong collisions, which by definition are perfectly efficient
in thermalizing excited species. For weak collisions, which are not perfectly efficient, the
collision frequency can be modified by the collision efficiency βc to give simple expressions.
This is the approach taken in conventional RRKM theory [18–20,27], according to which
the pressure-dependent unimolecular reaction rate constant kuni can be written as
Experiments on collisional energy transfer 9

ð∞
kB T βc ωkðEÞBðE Þ
kuni ¼ dE (1.19)
hQA E0 β c ω + kðE Þ

where QA is the partition function of species A and the other symbols are defined above.
Multiplying the collision frequency ω by βc amounts to scaling the pressure axis of a
fall-off plot by βc. At the high-pressure limit, there is no dependence on βc, but at the
low-pressure limit, the unimolecular rate constant is directly proportional to βc:
ð
kB T ∞
klow ¼ βc ω BðE ÞdE (1.20)
hQA E0
The parameter βc can be found by fitting kuni as a function of collider gas pressure or by
comparing the measured value of klow to the one calculated from Eq. (1.20) with βc set
equal to unity [28].
For the exponential model with parameter α independent of energy, the master equa-
tion can be (approximately) solved to obtain an analytical expression for the collision
efficiency:
 2
α
βc  (1.21)
α + FE k B T
where FE is a factor related to the equilibrium energy distribution [28]. Thus, the collision
efficiency can be related to the more fundamental parameter α and to the average collision
step sizes hΔEidown and hΔEiall.
Although the determination of collision efficiencies has been very important histori-
cally, steady-state master equation simulations that lead to values of hΔEi are far more
accurate. A numerical solution of the master equation is discussed elsewhere in this volume.
Modern master equation computer codes that are freely available can be used readily for
fitting experimental data and for simulations [29–31].

2.4 Time-dependent systems


In steady-state descriptions such as those described above, the competition between
collision frequency and unimolecular reaction results in the fall-off curve. By analyzing
the fall-off curve, one can determine the average collision step size. In time-dependent sys-
tems, additional phenomena are observable and the theoretical analysis is more demanding.
In return, more detailed information about energy transfer can be extracted.

2.4.1 Shock-wave excitation and reaction


Prior to the passage of a shock wave, the vibrational energy is described by a Boltzmann
distribution at ambient temperature. The passage of the shock wave produces a near-
instantaneous jump in translational temperature due to adiabatic compression of the
gas. Translations and rotations relax much more rapidly than vibrations, and therefore
10 Unimolecular Kinetics

the vibrational (V) temperature remains near ambient while the translational/rotational
(T/R) temperature equilibrates within a few collisions. Collisional vibrational activation
now occurs, and eventually the vibrational energy distribution reaches a new steady state
that depends on the translational temperature and the energy-dependent rate of unim-
olecular reaction [32–34].
The vibrational activation process requires many collisions, since the average amount
of energy transferred per collision is much smaller than the threshold energy for reaction.
As the activation takes place, the vibrational energy distribution is characterized by an
increasing average energy and width, as illustrated in Fig. 1.1 for norbornene [34]. Follow-
ing the passage of a shock wave, the average vibrational energy (Fig. 1.2) increases smoothly
and monotonically until a new Boltzmann distribution is established at the new temper-
ature. The exact details of the energy relaxation depend on the energy transfer model.
When a fraction of the reactant is activated to energies above the reaction threshold,
unimolecular reaction occurs. The “incubation time” (τinc) is the delay between the passage
of the shock wave and the onset of unimolecular reaction. Shortly after onset of this reac-
tion, the population distribution achieves steady state and the unimolecular reaction rate
constant corresponds to the steady state kuni discussed above. The onset of unimolecular
reaction, the incubation time, and steady-state unimolecular reaction are shown in Fig. 1.3.
The “vibrational relaxation time” (τvib) characterizes the transition from the initial to
the final vibrational energy distribution. In the absence of unimolecular reaction, the final
vibrational energy distribution is thermal and at the same temperature as the bath. When
unimolecular reaction is significant relative to collisional activation, the final steady-state
vibrational energy distribution is depleted relative to the thermal distribution, as discussed
above, resulting in unimolecular reaction rate coefficient fall-off.

7 ´ 104
0 µs Norbornene C2H4 + c-C5H6
Population (ppm) / Bin (250 cm–1)

6 ´ 104
1.2 µs
2
(T,E) = 40 + 5.8 ´ 10–6 T E
5 ´ 104

2.4 µs Shock #30


4 ´ 104
4.8 µs
9.6 µs
3 ´ 104

2 ´ 104

1 ´ 104

0 ´ 100
0 5000 10,000 15,000 20,000
–1
Energy (cm )
Fig. 1.1 Evolution of the population distribution (Shock #30 from [35]).
Experiments on collisional energy transfer 11

15,000 = 40 + 5.8 ´ 10–6 T E


2(E,T)

Vibrational energy (cm–1)


3(E,T) = 40 + 0.0063 E
10,000

1(E,T) = 10 + 1.1 ´ 10–5 T E


5000
Shock #76

Norbornene (+Kr) C2H4 + c-C5H6 (+Kr)


0
0 ´ 100 5 ´ 10–6 1 ´ 10–5 1.5 ´ 10–5 2 ´ 10–5 2.5 ´ 10–5
Time (s)
Fig. 1.2 Relaxation of average energy as calculated with three energy transfer models for Shock #76
from KS [35].

inc
1

1(E,T ) = 10 + 1.1 ´ 10–5 T E


[NB] / [NB]0

2(E,T ) = 40 + 5.8 ´ 10–6 T E

3(E,T ) = 40 + 0.0063 E

Shock #76

Norbornene (+Kr) C2H4 + c-C5H6 (+Kr)


0.1
0 ´ 100 1 ´ 10–5 2 ´ 10–5 3 ´ 10–5 4 ´ 10–5
Time (s)
Fig. 1.3 Incubation and unimolecular reaction in Shock #76 [35] calculated with three models. The
fluctuations are due to the stochastic solution of the master equation.

The vibrational relaxation is assumed to be driven by the difference in energy between


the vibrational energy E and the final steady-state vibrational energy E∞ [35]:
dE ðE  E∞ Þ
¼ (1.22)
dt τvib
In general, the phenomenological τvib is a function of time because the relaxation involves
many energy levels and the vibrational energy is the sum of the level energies weighted by
the level populations. Eq. (1.22) can be integrated to give
12 Unimolecular Kinetics

Evib ðtÞ ¼ E∞ + ðEi  E∞ ÞeW ðtÞ , (1.23)


where Ei is the initial vibrational energy. The function W(t) depends on the time depen-
dence of τvib. In simulations of norbornene excited by a shock wave, the expression
τvib ¼ (a + bt)  1 provides an excellent description of the phenomenological vibrational
relaxation time [34].

2.4.2 Photoactivation
Pulsed photoactivation experiments have been used extensively for so-called “direct” mea-
surements of energy transfer in nonreacting systems as well as for unimolecular reaction
studies [13,14,16,36,37]. Single-photon photoactivation produces an initial vibrational
energy distribution that is very narrow: it is thought to be a thermal distribution (at ambient
temperature) that has been shifted upward in energy by hν, the absorbed photon energy
[38]. In multiple photon excitation using high-power infrared lasers, the excited molecules
absorb photons sequentially to produce a rather broad distribution function [39–41]. The
details of the initial distribution depend on the combined effects of stimulated emission and
the variation of the absorption coefficient with vibrational energy. The average vibrational
energy corresponds to the average absorbed laser energy, which can be measured using
optoacoustic techniques or by measuring the attenuation of the laser energy [42–44].
For energy transfer studies, knowledge of the initial average vibrational energy is sufficient
for analyzing the experiments to obtain the average energy transferred per collision.
Subsequent to excitation, the excited molecules undergo collisional energy transfer,
which results in changes in the vibrational energy distribution. The rates of deactivation
and broadening of the vibrational energy distribution depend on the collision step-size
distribution. Analysis is best carried out by using master equation modeling, but simpler
phenomenological methods are also useful. These methods are described in the sections
describing the various experimental techniques.

3. Major experiment categories


3.1 Thermal activation
A quantitative understanding of collisional energy transfer is essential for the interpreta-
tion and prediction of rate coefficients for thermal unimolecular and recombination
reactions over wide ranges of temperature and pressure. These fundamental reactions
occur in such complex chemical systems as pyrolysis, combustion, and atmospheric
chemical processes. Traditionally, collisional energy transfer has been investigated in
unimolecular reaction systems using thermal, chemical, or photoactivation where a com-
petition is established between chemical reaction and collisional energy transfer. Because
the energy transfer information has to be deduced from the pressure-dependent reaction
rate coefficients or pressure-dependent product yields [13,14,16,45], these are “indirect”
Another random document with
no related content on Scribd:
back
back
back
back
back
back
back
back
back
back
back
back
back
back
back

You might also like