You are on page 1of 335

Physical Chemistry 3:

— Chemical Kinetics —
F. Temps
Institute of Physical Chemistry
Christian-Albrechts-University Kiel
Olshausenstr. 40
D-24098 Kiel, Germany
Email: temps@phc.uni.kiel.de
URL: http://www.uni-kiel.de/phc/temps
Summer Term 2019

This scriptum contains notes for the lecture “Physical Chemistry 3:


Chemical Kinetics” (MNF-chem0405) in the summer term 2019 at the
Institute of Physical Chemistry of Christian-Albrechts-University (CAU)
Kiel. This module is the last of the three semester lecture course in
Physical Chemistry for B.Sc. students of Chemistry and Chemistry and
Economics at CAU Kiel. It includes 3 hours weekly (3 SWS) for lectures
and 1 hour weekly (1 SWS) for an exercise class.

c F. Temps (2004 — 2019)


° Last update: June 26, 2019
ii

Contents

Preface ix

Disclaimer xi

Organisational matters xii

1 Introduction 1

1.1 Types of chemical reactions 1

1.2 Historical events 2

1.3 Time scales for chemical reactions 4

1.4 References 7

2 Formal kinetics 8

2.1 Definitions and conventions 8

2.1.1 The rate of a chemical reaction 8

2.1.2 Rate laws and rate coefficients 9

2.1.3 The order of a reaction 10

2.1.4 The molecularity of a reaction 12

2.1.5 Mechanisms of complex reactions 12

2.2 Kinetics of irreversible first-order reactions 16

2.3 Kinetics of reversible first-order reactions (relaxation processes) 20

2.4 Kinetics of second-order reactions 24

2.4.1 A + A → products 24

2.4.2 A + B → products 25

2.4.3 Pseudo first-order reactions 26

2.4.4 Comparison of first-order and second-order reactions 27

2.5 Kinetics of third-order reactions 28

2.6 Kinetics of simple composite reactions 29

2.6.1 Consecutive first-order reactions 29

2.6.2 Reactions with pre-equilibrium 32

2.6.3 Parallel reactions 33

2.6.4 Simultaneous first- and second-order reactions* 33

2.6.5 Competing second-order reactions* 34


iii

2.7 Temperature dependence of rate coefficients 35

2.7.1 Arrhenius equation 35

2.7.2 Deviations from Arrhenius behavior 37

2.7.3 Examples for non-Arrhenius behavior 38

2.8 References 40

3 Kinetics of complex reaction systems 41

3.1 Determination of the order of a reaction 41

3.2 Application of the steady-state assumption 43

3.2.1 The reaction H2 + Br2 → 2 HBr 43

3.2.2 Further examples 44

3.3 Microscopic reversibility and detailed balancing 46

3.4 Generalized first-order kinetics* 47

3.4.1 Matrix method* 47

3.4.2 Laplace transforms* 54

3.5 Numerical integration 57

3.5.1 Taylor series expansions 58

3.5.2 Euler method 58

3.5.3 Runge-Kutta methods 59

3.5.4 Implicit methods (Gear): 61

3.5.5 Extrapolation methods (Bulirsch-Stoer) 62

3.6 Oscillating reactions* 64

3.6.1 Autocatalysis 64

3.6.2 Chemical oscillations 65

3.7 References 73

4 Experimental methods 74

4.1 End product analysis 74

4.2 Modulation techniques 75

4.3 The discharge flow (DF) technique 76

4.4 Flash photolysis (FP) 80

4.5 Cavity Ringdown Spectroscopy (CRDS) 81

4.6 Shock tube studies of high temperature kinetics 84

4.7 Stopped flow studies 85


iv

4.8 NMR spectroscopy 86

4.9 Relaxation methods 87

4.10 Femtosecond spectroscopy 88

4.11 References 93

5 Collision theory 94

5.1 Hard sphere collision theory 95

5.1.1 The gas kinetic collision frequency 95

5.1.2 Allowance for a threshold energy for reaction 99

5.1.3 Allowance for steric hindrance 100

5.2 Kinetic gas theory 101

5.2.1 The Maxwell-Boltzmann speed distributions of gas molecules in 1D and 3D 101

5.2.2 The rate of wall collisions and the ideal gas law (ideal gas pressure) 106

5.3 Transport processes in gases 111

5.3.1 The general transport equation 111

5.3.2 Diffusion 113

5.3.3 Heat conduction 114

5.3.4 Viscosity 115

5.4 Advanced collision theory 116

5.4.1 Fundamental equation for reactive scattering in crossed molecular beams 117

5.4.2 Crossed molecular beam experiments 118

5.4.3 From differential cross sections  (  ) to reaction rate constants  ( ) 121

5.4.4 Laplace transforms in chemistry 122

5.4.5 Bimolecular rate constants from collision theory 124

5.4.6 Long-range intermolecular interactions 133

6 Potential energy surfaces for chemical reactions 141

6.1 Diatomic molecules 141

6.2 Polyatomic molecules 142

6.3 Trajectory Calculations 146

7 Transition state theory 150

7.1 Foundations of transition state theory 150

7.1.1 Basic assumptions of TST 151


v

7.1.2 Formal kinetic model 153

7.1.3 The fundamental equation for  ( ) 154

7.1.4 Tolman’s interpretation of the Arrhenius activation energy 157

7.2 Applications of transition state theory 159

7.2.1 The molecular partition functions 159

7.2.2 Example 1: Reactions of two atoms A + B → A· · · B‡ → AB 161

7.2.3 Example 2: The reaction F + H2 → F· · · H· · · H‡ → FH + H 162

7.2.4 Example 3: Deviations from Arrhenius behavior (T dependence of  ( ))* 164

7.3 Thermodynamic interpretation of transition state theory 165

7.3.1 Fundamental equation of thermodynamic transition state theory 165

7.3.2 Applications of thermodynamic transition state theory 169

7.3.3 Kinetic isotope effects 171

7.3.4 Gibbs free enthalpy correlations 173

7.3.5 Pressure dependence of  173

7.4 Transition state spectroscopy 175

7.5 References 179

8 Unimolecular reactions 180

8.1 Experimental observations 180

8.2 Lindemann mechanism 183

8.3 Generalized Lindemann-Hinshelwood mechanism 187

8.3.1 Master equation 187

8.3.2 Equilibrium state populations 189

8.3.3 The density of vibrational states  () 190

8.3.4 Unimolecular reaction rate constant in the low pressure regime 193

8.3.5 Unimolecular reaction rate constant in the high pressure regime 195

8.4 The specific unimolecular reaction rate constants  () 196

8.4.1 Rice-Ramsperger-Kassel (RRK) theory 196

8.4.2 Rice-Ramsperger-Kassel-Marcus (RRKM) theory 198

8.4.3 The SACM and VRRKM model 203

8.4.4 Experimental results 204

8.5 Collisional energy transfer* 205

8.6 Recombination reactions 206


vi

8.7 References 207

9 Explosions 208

9.1 Chain reactions and chain explosions 208

9.1.1 Chain reactions without branching 208

9.1.2 Chain reactions with branching 209

9.2 Heat explosions 212

9.2.1 Semenov theory for “stirred reactors” ( = const) 212

9.2.2 Frank—Kamenetskii theory for “unstirred reactors” 216

9.2.3 Explosion limits 217

9.2.4 Growth curves and catastrophy theory 217

10 Catalysis 219

10.1 Kinetics of enzyme catalyzed reactions (homogeneous catalysis) 219

10.1.1 The Michaelis-Menten mechanism 220

10.1.2 Inhibition of enzymes 222

10.2 Kinetics of heterogeneous reactions (surface reactions) 224

10.2.1 Reaction steps in heterogeneous catalysis 224

10.2.2 Physisorption and chemisorption 224

10.2.3 The Langmuir adsorption isotherm 226

10.2.4 Surface Diffusion 228

10.2.5 Types of heterogeneous reactions 229

11 Reactions in solution 233

11.1 Qualitative model of liquid phase reactions 233

11.2 Diffusion-controlled reactions 236

11.2.1 Experimental observations 236

11.2.2 Derivation of  236

11.3 Activation controlled reactions 240

11.4 Electron transfer reactions (Marcus theory) 241

11.5 Reactions of ions in solutions 245

11.5.1 Effect of the ionic strength of the solution 245

11.5.2 Kinetic salt effect 245

11.6 References 247


vii

12 Photochemical kinetics 248

12.1 Radiative processes in a two-level system; Einstein coefficients 248

12.2 Light amplification by stimulated emission of radiation (LASER) 249

12.3 Fluorescence quenching (Stern-Volmer equation) 251

12.4 The radiative lifetime  0 and the fluorescence quantum yield Φ 252

12.5 Radiationless processes in photoexcited molecules 253

12.5.1 The Born-Oppenheimer approximation and its breakdown 254

12.5.2 Avoided crossings of two potential curves of a diatomic molecule 257

12.5.3 Quantum mechanical treatment of a coupled 2×2 system 259

12.5.4 Radiationless processes in polyatomic molecules: Conical intersections 261

12.5.5 Femtosecond time-resolved experiments 264

13 Atmospheric chemistry* 268

14 Combustion chemistry* 269

15 Astrochemistry* 270

16 Energy transfer processes* 271

17 Intramolecular dynamics & vibrational spectroscopy* 272

18 Reaction dynamics* 273

19 Advanced photochemical dynamics* 274

20 Electrochemical kinetics* 275

A Useful physicochemical constants 276

B The Marquardt-Levenberg non-linear least-squares fitting algorithm 277

C Solution of inhomogeneous differential equations 281

C.1 General method 281

C.2 Application to two consecutive first-order reactions 282

C.3 Application to parallel and consecutive first-order reactions 284

D Matrix methods 287

D.1 Definition 287


viii

D.2 Special matrices and matrix operations 288

D.3 Determinants 291

D.4 Coordinate transformations 292

D.5 Systems of linear equations 294

D.6 Eigenvalue equations 296

E Laplace transforms 299

F The Gamma function Γ () 301

G Ergebnisse der statistischen Thermodynamik 303

G.1 Boltzmann-Verteilung 304

G.2 Mittelwerte 306

G.3 Anwendung auf ein Zweiniveau-System 308

G.4 Mikrokanonische und makrokanonische Ensembles 311

G.5 Statistische Interpretation der thermodynamischen Zustandsgrößen 313

G.6 Chemisches Gleichgewicht 314

G.7 Zustandssumme für elektronische Zustände 316

G.8 Zustandssumme für die Schwingungsbewegung 317

G.9 Zustandssumme für die Rotationsbewegung 319

G.10 Zustandssumme für die Translationsbewegung 320


ix

Preface

This scriptum contains lecture notes for the module “Physical Chemistry 3: Chemical
Kinetics” (chem0405), the last part of the 3 semester course in Physical Chemistry for
B.Sc. students of Chemistry and Business Chemistry at CAU Kiel. It is assumed (but
not formally required) that students have previously taken courses on chemical thermo-
dynamics (“Physical Chemistry 1: Chemical Equilibrium” or equivalent) and elementary
quantum mechanics (“Physical Chemistry 2: Structure of Matter” or equivalent) as
well as “Mathematics for Chemistry” (parts 1 and 2, or equivalent). Module chem0405
includes 3 hours weekly (3 SWS) for lectures and 1 SWS of exercises (56 contact hours
combined). Students who started their studies at CAU Kiel should normally be in their
4th semester.
The scriptum gives a summary of the material covered in the lecture classes to allow
students to repeat the material more economically. It covers basic material that all
chemistry students should learn irrespective of their possible inclination towards inor-
ganic, organic or physical chemistry, but goes beyond the standard Physical Chemistry
textbooks used in the PC-1 and PC-2 courses. This is done in recognition of the estab-
lished research focus at the Institute of Physical Chemistry at CAU to enable students
to pursue B.Sc. thesis projects in Physical Chemistry. Some more specialized sections
have been marked by asterisks and may be omitted on first reading towards the B.Sc.
degree. Useful additional reference material is given in the Appendix.
A brief lecture scriptum can never replace a textbook. For additional review, students
are strongly encouraged (and at places required) to consult the recommended books
on Chemical Kinetics beyond the standard Physical Chemistry textbooks. The book
by Logan1 is a comprehensive overview of the field of chemical kinetics and a nice
introduction; it also has the advantage that it is written in German. The book by Pilling
and Seakins2 gives a good introduction especially to gas phase kinetics. The book by
Houston3 includes more information on the dynamics of chemical reactions. The book
by Barrante4 is highly recommended for everyone who wants to brush-up some math
skills. A highly recommended web site for looking up mathematical definitions and
recipies is the MathWorld online encyclopedia (http://mathworld.wolfram.com).
The computer has become indispensable in modern research. This applies especially
to Chemical Kinetics, where only computers can solve (by numerical integration) the
coupled nonlinear partial differential equation systems that describe important com-
plex reaction systems (e.g. reaction systems describing combustion, the atmosphere or
climate change). The integration of differential equations, a major task in chemical ki-
netics, can be performed using Wolfram Alpha (http://www.wolframalpha.com). More
complicated problems can be solved with versatile computer algebra systems, e.g. (in or-
der of increasing power) Derive,5 MuPad, MathCad,6 Maple, Mathematica7 . A graphics
program (e.g. Origin,8 or QtiPlot) may be needed for data evaluation and presentation.9

1
S. R. Logan, Grundlagen der Chemischen Kinetik, Wiley-VCH, Weinheim, 1996.
2
M. J. Pilling, P. W. Seakins, Reaction Kinetics, Oxford University Press, Oxford, 1995.
3
P. L. Houston, Chemical Kinetics and Reaction Dynamics, McGraw-Hill, 2001.
4
J. R. Barrante, Applied Mathematics for Physical Chemistry, Prentice Hall, 2003 ($ 47.11).
5
Derive is available as shareware without charge.
6
MathCad is installed in the PC lab.
7
Mathematica has been used by the author for some of the lecture material.
8
Student versions of Origin are available for ≈ 35 − 120  (depending on run time).
9
Excel also allows you to plot simple functions and data, but is not sufficient for serious scientific
x

As practically all original scientific literature is published in English (British and Amer-
ican; in fact many chemists at some time in their professional lifes have to master the
subtle differences between both), this scriptum is written in English to introduce stu-
dents to the predominating language of science at an early stage. Students should also
note that the working language of some of the research groups (including that of the
author) in the Chemistry Department at CAU is English.
Finally, I would like to encourage all students to ask questions when they appear, whether
during the lecture or in the exercise classes!

Kiel, June 26, 2019, F. Temps

applications.
xi

Disclaimer

Use of this text is entirely at the reader’s risk. Some sections are incomplete, some
equations still have to be checked, some figures need to be redrawn, references are still
missing, etc. No warranties in any form whatever are given by the author.
xii

Organisational matters

• List of participants
• Lecture schedule
• Exams and grades

— Homework assignments
— Short questions
— Written final exam (“Klausur”)

I Table 1: List of participants.

Chemistry (B.Sc.) 44
Business Chemistry (B.Sc.) 24
Biochemistry (B.Sc.)
2x-Chemistry (B.Sc.)
Physics (M.Sc.)
Other)
Total 68
)
Not registered for exercise classes and final written exam (“Klausur”).

I Figure 1: Physical chemistry courses at CAU Kiel.


xiii

I Figure 2: Organisational matters.

I Figure 3: Requirements and grading.


xiv

I Figure 4: Lecture contents.

I Figure 5: Table of contents (continued).a

a
Topics marked with asterisks will be omitted in this course for lack of time. They are the subject of
specialized lectures for advanced students.
xv

I Figure 6: Recommended textbooks on chemical kinetics.

References to original publications and other recommended literature for further reading
are given at the end of each Chapter.
1

1. Introduction

1.1 Types of chemical reactions

• Homogeneous reactions:

— Reactions in the gas phase, e.g.

H2 + Cl2 → 2 HCl

— Reactions in the liquid phase, e.g. H+ -transfer, reduction/oxidation reac-


tions, hydrolysis, . . . (most of organic and inorganic chemistry)
— Reactions in the solid phase (usually very slow, except for special systems
and special conditions; not covered in this lecture).

• Heterogeneous reactions:

— Reactions at the gas-solid interface, e.g. combustion of solids

C() + 12 O2 () → CO() 

— Heterogeneous catalysis (NH3 synthesis by Haber-Bosch process, catalytic


hydrogenation rxns, exhaust gas treatment from combustion, . . . ),
— Reactions at the gas-liquid interface, e.g. heterogeneous catalysis, reactions
on aerosols, e.g. Cl2 release from polar stratospheric clouds in antarctic
spring via the rxn

ClONO2 () + HCl() → Cl2 () + HNO3 ()

— Reactions at the liquid-solid boundary, e.g. solvation, heterogeneous cataly-


sis, electrochemical reactions (electrode kinetics), . . .

• Irreversible reactions, e.g.

H2 + 12 O2 → H2 O

• Reversible reactions, e.g.


H2 + I2 À 2 HI

• Composite reactions, e.g.

CH4 + 2 O2 → CO2 + 2 H2 O

• Elementary reactions, e.g.

OH + H2 → H2 O + H

CH3 NC + M → CH3 CN + M
1.2 Historical events 2

1.2 Historical events

I Sucrose inversion reaction (Wilhelmy, 1850): Kinetic investigation of the H+ cat-


alyzed hydrolysis of sucrose by monitoring the change as function of time of the optical
activity of the solution from reactant (sucrose) to products (glucose and fructose).10

I Table 1.1: Optical activity parameters for the sucrose inversion reaction.

C12 H22 O11 + H2 O −→ C6 H12 O6 + C6 H12 O6


sucrose glucose fructose
 (1 M)) +665
P

0◦ ◦
+520 P −920◦
 (1 M)) = +665 ◦
= −400◦

)
 (1 M) = specific angle of rotation of a 1 M solution at (Na-D).

I Figure 1.1: Observed concentration-time profile of sucrose.

10
This experiment is carried out in the Physical Chemistry Lab Course 1 at CAU (“Grundpraktikum”).
1.2 Historical events 3

I Empirical rate law:



ln = −  (1.1)
0

Differentiation gives
 ln   ln 0 
− = − ( ) = − (1.2)
 | {z } 
=0
or
1  
= − (1.3)
  

y Empirical rate law:


 
= −  (1.4)

Reaction rate :
 
≡− (1.5)

Notice that even though there is one molecule of H2 O taking part in the reaction, H2 O
does not appear in the rate law, because it is practically constant in the dilute solution.

I Figure 1.2: Milestones in chemical kinetics.


1.3 Time scales for chemical reactions 4

1.3 Time scales for chemical reactions

A chemical reaction requires some “contact” between the respective reaction partners.
The “time between contacts” thus determines the upper limit for the rate of a reaction.

I Reactions in the gas phase:

• Free mean pathlength  À molecular dimension (diameter) .


• Reaction rate is determined by time between collisions.
• Time between two collisions @   :

 ≈ 500 m s (1.6)

 ≈ 10−7 m (1.7)
y
∆ ≈  ≈ 2 × 10−10 s (1.8)
y
• Gas phase reactions take place on the nanosecond time scale:

I 1
@  ≤ = 5 × 109 s−1 (1.9)
∆

I Reactions in the liquid phase:

• Molecules are permanently in contact with each other.


• Typical distance between neighboring molecules (upper limit allowing for solvation
shell):
 ≈ 2 × 10−8    10−7 cm (1.10)

• Maximal reaction rate is determined by the rate of diffusion of the reacting mole-
cules in and out of the solvent cage.

I Reactions in the solid phase:

• Atoms and molecules are localized at specific lattice positions.


• Solid state reactions are usually very slow.
• Reaction rates in solids are determined by rate of diffusion (“hopping”) of the
atoms and molecules via vacancies (unoccupied lattice positions, “Fehlstellen”)
or interstitial sites (“Zwischengitterplätze”).
• Concentration of vacancies and interstitial sites controlled by thermodynamics.
• Hopping from one lattice position to another usually requires a corresponding high
activation energy.
1.3 Time scales for chemical reactions 5

I Time scale of molecular vibrations:

• The frequencies of the vibrational motions of the atoms in a molecule with respect
to each other determine an upper limit for the rate of a unimolecular reaction.
• Uncertainty principle (Fourier relation between frequency and time domain):
∆ ∆ ≥ ~ (1.11)
y
1
∆ ∆ ≥ (1.12)
2
With  =  =  ̃:
1
∆ ≈ (1.13)
2 ∆̃
• Application to a C-C stretching vibration:
̃(C—C) ≈ 1000 cm−1 (1.14)
y
 =  =  ̃ = 3 × 1013 s−1 (1.15)
y
 =  −1 = 333 × 10−14 s ≈ 33 fs (1.16)
• Application to slow torsional motions of large molecular groups in solution:
̃ ≈ 100 cm−1 (1.17)
y
∆ ≈ 330 fs (1.18)

I Femtochemistry: We can observe ultrafast chemical reactions directly nowadays by


following the motions of atoms or groups of atoms in the course of a reaction in real
time by femtosecond spectroscopy and femtosecond diffraction techniques. This new
area of “femtochemistry” has grown enormously in importance in the last decade, since
the award of the Nobel prize in chemistry to A.H. Zewail in 1999.
1 femtosecond = 1 fs = 10−15 s (1.19)

I Time scale of e− -jump processes: Only electron jumps are faster than vibrational
motions (by a factor of the order of   ).

• Estimated time scale for − -jump to excited orbital due to absorption of a photon:
1 
 ≈ × ≈ 5 × 10−18 s = 5 as (1.20)
 ̃ (H−H) 

• However, “chemical” − -transfer reactions (intra- or intermolecular − -transfer,


i.e. all red-ox reactions) are also limited by nuclear motions, because they usually
require substantial molecular rearrangements due to the structural differences be-
tween the reactants and products. The net reaction rate is thus again limited by
the time scale of vibrational motion (⇒ Marcus theory, Section 11.4).
1.3 Time scales for chemical reactions 6

I Attosecond spectroscopy: Since ≈ 2010, “attosecond spectroscopy” has taken off


with speed. By attosecond spectroscopy experiments, we can now observe the motion
of electrons in photoexcitation and photoionization experiments on the attosecond time
scale (1 as = 10−18 s).

• Since the periodic oscillation of a light wave at visible wavelengths takes much
longer than attoseconds, attosecond spectroscopy experiments require special
techniques, e.g. high harmonic generation (HHG) to reach soft x-ray wavelengths,
and ultraprecise ( as) interferometry.
• The group of F. Krausz (MPI for Quantum Optics, Munich) used “attosecond
metrology” based on an interferometric method to measure a delay of 21 ± 5 as in
the photo-induced emission of electrons from the 2 orbitals of neon atoms with
respect to emission from the 2 orbital by a 100 eV light pulse (Schultze et al.,
Science 328, 1658 (2010)).
• If true, these results indicate that the motion of an electron is affected by electron
correlation, because the photo-excited electron feels the effective electric field of
all electrons of an atom or molecule (⇒ configuration interaction, CI, in quan-
tum chemistry). There are indications that electron correlation can slow electron
motion from the 10−18 s time scale to several 10−17 s.
• Other fast electron transfer processes:

— Autoionization of a doubly excited state, by one electron falling to a lower


orbital, while the other is excited into the ionization continuum (⇒ Auger
spectroscopy).

I Conclusions:

• The time scales of chemical reactions span the enormous range of some 30 orders
of magnitude, from 1 fs = 10−15 s up to 109 y = 1016 s!
• In view of this huge span, a factor-to-two experimental uncertainty in an exper-
imental measurement or a theoretical prediction of a rate constant really means
rather little.
• The enormous range of reaction rates, their complex dependencies on temper-
ature, pressure, and species concentrations, and the question of the underlying
reaction mechanisms that determine the reaction rates, make the field of chemical
reaction kinetics and dynamics very attractive and rewarding!
1.3 Time scales for chemical reactions 7

1.4 References

Arrhenius 1889 S. A. Arrhenius, Über die Reaktionsgeschwindigkeit bei der Inversion


von Rohrzucker durch Säuren, Z. Physik. Chem. 4, 226 (1889).
Bunker 1966 D. L. Bunker, Theory of Elementary Gas Reaction Rates, Pergamon,
Oxford, 1966.
Glasstone 1941 S. Glasstone, K. Laidler, H. Eyring, The Theory of Rate Processes,
Mc Graw-Hill, New York, 1941.
Homann 1975 K. H. Homann, Reaktionskinetik, Grundzüge der Physikalischen
Chemie Bd. IV, R. Haase (Ed.), Steinkopf, Darmstadt, 1975.
Houston 1996 P. L. Houston, Chemical Kinetics and Reaction Dynamics, McGraw-
Hill, Boston, 2001.
Johnston 1966 H. S. Johnston, Gas Phase Reaction Rate Theory, Ronald, New York,
1966.
Kassel 1932 L. S. Kassel, The Kinetics of Homogeneous Gas Reactions, Chem. Cata-
log, New York, 1932.
Logan 1996 S. R. Logan, Grundlagen der Chemischen Kinetik, Wiley-VCH, Weinheim,
1996.
Pilling 1995 M. J. Pilling, P. W. Seakins, Reaction Kinetics, Oxford University Press,
Oxford, 1995.
Smith 1980 I. W. M. Smith, Kinetics and Dynamics of Elementary Gas Reactions,
Butterworths, London, 1980.
Steinfeld 1989 J. I. Steinfeld, J. S. Francisco, W. L. Haase, Chemical Kinetics and
Dynamics, Prentice Hall, Englewood Cliffs, 1989.
1.4 References 8

2. Formal kinetics

2.1 Definitions and conventions

2.1.1 The rate of a chemical reaction

I Rate of a simple reaction of type A + B → C + D:


 A  B  C  D
 ≡ − =− =+ =+ (2.1)
   
 [A]  [B]  [C]  [D]
= − =− =+ =+ (2.2)
   

We will usually write the concentrations of a species A by using square brackets, i.e.
[A]  and understand that [A] is time dependent, i.e. [A ()].

I Definition of the rate of reaction:11 Using the standard convention for the signs
of the stoichiometric coefficients of the reactants (−1) and products (+1), we write a
generic chemical reaction as

 1 B1 +  2 B2 +    →   B +  +1 B+1 +    (2.3)

In Chemical Thermodynamics (PC-1), we have described the progress of the reaction


from reactants to products using the reaction progress coordinate :
 
 = (2.4)

In Chemical Kinetics, we study the time dependence of , which leads directly to the
definition of the reaction rate.
I Definition 2.1: The reaction rate  is defined as
1 
≡ (2.5)
 

For the above reaction, therefore,

1  1 1  2 1   1  +1
=− =− =  = + =+ =  (2.6)
| 1 |   | 2 |   |  |   | +1 |  

11
Reaction rate = Reaktionsgeschwindigkeit. Von anderen wird der Begriff “Reaktionsrate” verwen-
det, wir wollen jedoch nach Bodenstein grundsätzlich den Ausdruck “Reaktionsgeschwindigkeit”
verwenden.
2.1 Definitions and conventions 9

I Units:12
• SI unit for the reaction rate:
[] = mol m3 s (2.7)
• Convenient units for gas phase reactions:
— molar units:
[] = mol cm3 s (2.8)
— molecular units:13
[] = molecules cm3 s ⇒ cm−3 s−1 (2.9)
— We prefer to use molar units for rate constants, because rate constants are
inherently quantities that are “averages” of some other quantity, e.g. a mole-
cular cross section weighted by the relevant statistical distribution function
(usually the Boltzmann distribution). Rate constants are thus properties of
an ensemble of molecules, they are not original molecular quantities.
• Convenient units for liquid phase reactions:
[] = mol l s = M s (2.10)

2.1.2 Rate laws and rate coefficients

I The rate equation (rate law): Experiments show that the reaction rate  generally
depends on the concentrations of the reactants. The functional dependence is expressed
by the rate equation (also called rate law),14 e.g.
1  [B1 ]
 = − =  (2.11)
| 1 |  
=  ([B1 ]  [B2 ]      [B ]  [B+1 ]    ) (2.12)
for reaction 2.3.
The rate laws of complex reactions (i.e. the  ([B1 ] , [B2 ] ,    , [B ] , [B+1 ] ,   ))
are often rather complicated; only rarely (as in the case of elementary reactions) are
they simple and intuitive.
Often (but not always),  ([B1 ] , [B2 ] ,    , [B ] , [B+1 ] ,   ) can be written as a
product of a rate coefficient  and some power series of the concentrations, for
example
1  [B1 ]
=− =  [B1 ] [B2 ] . . . (2.13)
| 1 |  
Note that the exponents  . . . are not usually equal to  1   2 . . .

I The rate constant (rate coefficient): The rate constant (also called rate coeffi-
cient)15  usually depends on temperature ( = ( )) and sometimes also on pressure

12
For general information on the SI system see (Mills 1988) or (Homann 1995).
13
IUPAC has decided that “molecules” is not a unit; see (Mills 1989) or (Homann 1995).
14
Rate law = Geschwindigkeitsgesetz oder Zeitgesetz.
15
Rate constant = Geschwindigkeitskonstante. Es wird manchmal auch die Übersetzung “Ratenkon-
stante” verwendet (aber bitte nicht “Ratekonstante”).
2.1 Definitions and conventions 10

( = ( )), but not on the concentrations ( 6=  ( )).


I Question: How can we determine reaction rates  and rate coefficients ?
I Answer:

• The most direct method — though not at all the most elegant method — is the
tangent method (see Fig. 2.1).
• We’ll learn about better methods for determining  for elementary reactions below.
Especially, we will consider first-order reactions in Section 2.2 and second-order
reactions in Section 2.4, and there will be homework assignments.

I Figure 2.1: A straightforward — and always applicable — method for determining the
reaction rate  and rate coefficient  is the tangent method. To determine , you
simply take the slope of a plot of () at time ; the rate coefficient  then follows via
Eq. 2.13.

2.1.3 The order of a reaction

I Rate laws for first-order and second-order reactions: Simple rate laws describe
reactions of first-order and second-order, e.g.

• Example of a first-order reaction: CH3 NC → CH3 CN


 [CH3 NC]  [CH3 CN]
=− =+ =  [CH3 NC] (2.14)
 
2.1 Definitions and conventions 11

• Example of a second-order reaction: C4 H9 + C4 H9 → C8 H18

1  [C4 H9 ]  [C8 H18 ]


=− =+ =  [C4 H9 ]2 (2.15)
2  

I Definition 2.2: The order of a reaction is defined as the sum of the exponents in the
rate law (Eq. 2.13).

I Example: Considering a generic rate law, like

 [A]
− = |  |  [A] [B]     (2.16)

the total order of the reaction is  =  + . We also say, the reaction is -th order in
[A] and -th order in B, . . .
As mentioned, the order of a reaction may be determined by the tangent method (cf.
Fig. 2.1): Its effects on the reaction rate  are observable by increasing (e.g. doubling)
the concentrations of the reaction partners one by one. Other methods: ⇒ homework
assignments.

I Caveats:

• The order of a reaction is, in general, an empirical quantity; it has to be determined


by an experimental measurement.
• The order of a complex reaction cannot be determined by simply looking at the
overall reaction.
• There are reactions with integer order, like the first-order or second-order reactions
above (Eqs. 2.14 and 2.15). However, the order can also be fractional, as for the
following reactions:

(1) CH3 CHO → CH4 + CO:

 [CH4 ]
=  [CH3 CHO]32 (2.17)


(2) H2 + Br2 → 2 HBr:

 [HBr] 0 [H2 ] [Br2 ]12


= (2.18)
 1 +  00 [HBr]  [Br2 ]

• A negative order in a concentration means that the respective species acts as an


inhibitor.
• A positive order in a product concentration means that the reaction rate is en-
hanced by autocatalysis.
2.1 Definitions and conventions 12

2.1.4 The molecularity of a reaction

I Definition 2.3: The molecularity of a reaction is the number of molecules in an ele-


mentary reaction step at the microscopic (molecular) level.

I There are only three possible cases for the molecularity (one example for each
case):

• monomolecular (= unimolecular) reactions:

CH3 NC → CH3 CN (2.19)

• bimolecular reactions:
O + H2 → OH + H (2.20)

• termolecular reactions:

H + O2 + M → HO2 + M (2.21)

• Events where more than three species come together in the right configuration
and react are extremely unlikely and do not play a role.

I Caveat: The order of a reaction can be integer or non-integer. However, the molec-
ularity can be only 1 or 2, or at the most 3.

2.1.5 Mechanisms of complex reactions

A stoichiometric formula rarely describes the reactive events happening at the micro-
scopic, molecular level. It is, for instance, completely irrational that the combustion of
1 -heptane molecule with 11 molecules of O2

-C7 H16 + 11 O2 → 7 CO2 + 8 H2 O (2.22)

in an internal combustion engine occurs in a single collision and gives 15 product mole-
cules. Instead, the net reaction is the result of a complex sequence of a very large
number (thousands, in this case) of unimolecular, bimolecular, and termolecular ele-
mentary reactions. Only these simple elementary reactions proceed in “single collision”
events (though may still be ruled by rather complex molecular dynamics).
The identification of the individual elementary reactions, the determination of their
rates as a function of  and , the understanding of the underlying molecular dynamical
processes, and the verification of the ensuing reaction mechanism are the major tasks
in the field of chemical kinetics.
We consider a few examples for complex reaction systems:

• H2 /O2 system (Fig. 2.2),


• combustion of CH4 (Fig. 2.3),
2.1 Definitions and conventions 13

• NO formation and removal in flames (Fig. 2.4).

I Figure 2.2: Reactions in the H2 /O2 system.

I Rate equations for the H2 /O2 system: Assuming that we know the reaction mech-
anism (i.e. all contributing elementary reactions), we can write down the rate equations
for all species. Doing so, we obtain a set of coupled non-linear differential equations
(DE’s). Usually, these coupled non-linear DE systems can be solved only numerically
(by more or less sophisticated numerical integration methods) in order to describe the
net reaction:
 [H]
= −1 [H] [O2 ] + 2 [O] [H2 ] + 3 [OH] [H2 ] − 4 [H] [O2 ] [M] (2.23)

−5 [H] [HO2 ] − 6 [H] [HO2 ] +   
 [OH]
= 1 [H] [O2 ] + 2 [O] [H2 ] − 3 [OH] [H2 ] + 2 6 [H] [HO2 ] (2.24)

−7 [OH] [HO2 ] − 2 8 [OH]2 [M]
+ (2.25)
 [O]
=  (2.26)

2.1 Definitions and conventions 14

I Figure 2.3: Mechanism of CH4 combustion (Warnatz 1996).

I Figure 2.4: NO in combustion processes.


2.1 Definitions and conventions 15

I How we will proceed:

• In the remainder of this Chapter, we shall learn how to analytically solve the rate
equations for elementary reactions and for simple composite reactions.
• In Chapter 3, we shall then consider different methods for solving coupled non-
linear differential equations for complex reaction systems.

— These methods are well established for stationary reaction systems.


— However, the modeling of instationary reaction systems (such as ignition
processes, explosions and detonations, 3D or 4D models of atmospheric
chemistry and climate evolution, . . . ) continues to be a tremendously chal-
lenging task.
2.2 Kinetics of irreversible first-order reactions 16

2.2 Kinetics of irreversible first-order reactions

A→B (2.27)

I Solution of the rate equations for A and B:

 [A]  [B]
− =+ =  [A] (2.28)
 

Mass balance:
 [A]  [B]
+ =0 (2.29)
 

Integration of the rate equation (Eq. 2.28):

 [A]
− =   (2.30)
[A]
Z Z Z
 ln [A] = −  = −  (2.31)
y
ln [A] = −  +  (2.32)

Initial value condition at  = 0:

[A ( = 0)] = [A]0 (2.33)

y
 = ln [A]0 (2.34)

Solution for [A ()]:


ln [A] = −  + ln [A]0 (2.35)
y
[A ()] = [A]0 −  (2.36)

Solution for [B ()]:


[B ()] = [A]0 − [A ()] (2.37)
y ¡ ¢
[B ()] = [A]0 1 − −  (2.38)
2.2 Kinetics of irreversible first-order reactions 17

I Figure 2.5: Kinetics of first-order reactions.

I Time constant: The rate coefficient  has the dimension of an inverse time. The
inverse of , the time constant
1
= (2.39)

1
thus has dimension of time and is the time after which [A] has dropped to (= 368

%) of its initial value:
[A]0 [A]0
[A]= = [A]0 − = ≈ (2.40)
 2718

1
I Half lifetime: The half lifetime 12 is the time after which [A] has dropped to of
2
its initial value (Fig. 2.5):
£ ¡ ¢¤ [A]0
A  = 12 = [A]0 − 12 = (2.41)
2
y
1
− 12 = (2.42)
2
 12 = ln 2 (2.43)
y
ln 2
12 = (2.44)

2.2 Kinetics of irreversible first-order reactions 18

I Conclusion:
The half lifetime for a first-order reactions is independent of the initial concentration!

I Experimental determination of k:

(1) Plot of ln [A] vs.  gives a straight line with slope  = − (see Fig. 2.5):
ln ([A]  [A]0 ) = −  (2.45)

Advantage: We do not need to know the absolute concentration of [A].


(2) Numerical fit of  to the exponential decay curve using the Marquardt-Levenberg
non-linear least-squares fitting algorithm (Fig. 2.6; see Appendix B for details).
This method is useful in particular for analyzing experimental decay curves with
a constant background term. Again, we do not need to know the absolute con-
centration of [A].
Advantages of non-linear fitting: Can be applied in presence of a constant back-
ground signal (noise, weak DC signal). Can also be applied for analyzing data
from pump-probe experiments, when the duration of the pump is not short to
the decay. In this case, the observed time profile is given by a convolution of
the exponential molecular decay signal with the pump pulse time profile, or more
generally with the time profile of the appropriate instrument response function
(IRF).

I Figure 2.6: Exponential decay curve of a particular vibration-rotation state of


the CH3 O radical resulting from its unimolecular dissociation reaction according to
CH3 O → H + H2 CO (Dertinger 1995). The small box is the output box from a fit
using the software package ORIGIN.
2.2 Kinetics of irreversible first-order reactions 19

I Figure 2.7: Excited-state relaxation dynamics of the adenine DNA dinucleotide after
UV excitation.

I Warning: With three parameters, you can draw an elephant; with a fourth parameter,
you can make him walk!
Non-linear least squares fitting requires a very cautious and critical analysis of the ob-
tained parameters! Careful error analysis and an investigation of the correlation between
the obtained fit parameters are mandatory. Anyone who wants to perform least-squares
fitting is urged to consult appropriate literature before doing so. The book by Bev-
ington16 is required reading material for all BSc, MSc or PhD students in Physical
Chemistry.

16
P. R. Bevington, D. K. Robinson, Data Reduction and Error Analysis for the Physical Sciences,
McGraw-Hill, Boston, 1992.
2.3 Kinetics of reversible first-order reactions (relaxation processes) 20

2.3 Kinetics of reversible first-order reactions (relaxation processes)

1
A À B (2.46)
−1

I Rate equation:
 [A]
= −1 [A] + −1 [B] (2.47)


I Equilibrium at t → ∞:

 [A]  [B]
=− = −1 [A] + −1 [B] = 0 (2.48)
 
y
[B]∞ 1
= =  (2.49)
[A]∞ −1

• If we have measured 1 and know  , the rate constant for the reverse reaction
−1 can be calculated.
• Or the equilibrium constant  can be determined from measurements of the
forward and reverse rate constants.
1
• From  = , we can then determine important thermochemical quantities
−1
that are difficult to obtain by other means. Note, however, that towards these
ends we need to use  instead of  :
∙ ¸
(B ª ) ∆ ª
 = = exp − (2.50)
(A ª ) 

I Important note:

• In the frequently encountered case that the number of species on the reactant
and product side of a reaction differ, one has to be very careful with the units of
 and  . Consider, for instance, a gas phase reaction of the type

A→B+C (2.51)

In this case,  has the dimension of mol cm3 :


B C
1 [B] [C] 1 B C 0
 = = =  A = = (2.52)
−1 [A]   A 

2.3 Kinetics of reversible first-order reactions (relaxation processes) 21

0 has the dimension of bar:


B C
0 = =  ×  (2.53)
A
But the thermodynamic equilibrium constant  is dimensionless:
∙ ¸
(B ª ) (C ª ) 0 ∆ ª
 = = ª = exp − (2.54)
(A ª )  

• General relation between   0 , and  :


µ ¶Σ  µ ¶Σ 
0 1 
 =  × =  × (2.55)
ª ª

I Solution of the rate equations for [A] and [B]:


Returning to the time dependence of the reaction
1
A À B (2.56)
−1

we now solve the rate equation


 [A]
= −1 [A] + −1 [B] (2.57)

• Mass balance:
[B] = [A]0 − [A] (2.58)
• Integration:
 [A]
= −1 [A] + −1 [B] (2.59)

= −1 [A] + −1 ([A]0 − [A]) (2.60)
= − (1 + −1 ) [A] + −1 [A]0 (2.61)
y
 [A]
= −  (2.62)
(1 + −1 ) [A] − −1 [A]0
• Solution by substitution:
 = (1 + −1 ) [A] − −1 [A]0 (2.63)

= (1 + −1 ) (2.64)
 [A]

 [A] = (2.65)
(1 + −1 )
y

= −  (2.66)
(1 + −1 ) 

= − (1 + −1 )   (2.67)

y
ln  = − (1 + −1 )  +  (2.68)
2.3 Kinetics of reversible first-order reactions (relaxation processes) 22

• Initial value condition at  = 0:


 = 0 (2.69)
y
 = ln 0 (2.70)
y 
= exp [− (1 + −1 ) ] (2.71)
0

I Solution for [A (t)]:

(1 + −1 ) [A] − −1 [A]0


= exp [− (1 + −1 ) ] (2.72)
(1 + −1 ) [A]0 − −1 [A]0
or
−1
[A] − [A]0
1 + −1
= exp [− (1 + −1 ) ] (2.73)
−1
[A]0 − [A]0
1 + −1

This expression is not easy to memorize, but we may recast it by using

1 [B]∞ [A]0 − [A]∞


= =  (2.74)
−1 [A]∞ [A]∞

which gives
−1
[A]∞ = [A]0  (2.75)
1 + −1
so that
[A ()] − [A]∞
= exp [− (1 + −1 ) ] (2.76)
[A]0 − [A]∞

With ∆ [A] = [A ()] − [A]∞ and ∆ [A]0 = [A]0 − [A]∞ , we write this result in compact
form as (see Fig. 2.8)
∆ [A]
= exp [− (1 + −1 ) ] (2.77)
∆ [A]0
2.3 Kinetics of reversible first-order reactions (relaxation processes) 23

I Figure 2.8: Kinetics of reversible first-order reactions.

I Relaxation processes:

• The transition from a deviation from equilibrium into the equilibrium is called
relaxation.
• (1 + −1 )−1 has the dimension of time and is called the relaxation time,

 = (1 + −1 )−1 (2.78)

• The temporal evolution of relaxation processes from a nonequilibrium to the equi-


libtium state is determined by the sum of the respective forward and reverse rate
constants!
2.4 Kinetics of second-order reactions 24

2.4 Kinetics of second-order reactions

2.4.1 A + A → products

A+A→P (2.79)

I Solution of the rate equation:

1  [A]
− =  [A]2 (2.80)
2 
Integration:
 [A]
= −2   (2.81)
[A]2
y
1
− = −2 +  (2.82)
[A]
Initial value condition at  = 0:

[A ( = 0)] = [A]0 (2.83)

y
1
=− (2.84)
[A]0
Solution for [A ()] (Fig. 2.15):

1 1
= + 2  (2.85)
[A] [A]0

I Experimental determination of k:

(1) Graphical method:


1 1
= + 2  (2.86)
[A] [A]0

⇒ A plot of [A]−1 vs.  gives a straight line with slope  = 2 (Fig. 2.9).
⇒ For second-order reactions, we do need to know the absolute concen-
tration of [A] to determine !
(2) If necessary, we can again use a non-linear least squares fitting method (Appendix
B).
2.4 Kinetics of second-order reactions 25

I Figure 2.9: Kinetics of second-order reactions.

I Half lifetime:
£ ¡ ¢¤ [A]0
A 12 = (2.87)
2
y
2 1
= + 2 12 (2.88)
[A]0 [A]0
y
1
12 = (2.89)
2 [A]0
y

The half lifetime for a second-order reactions depends on the initial concentration!

2.4.2 A + B → products

A+B→P (2.90)
2.4 Kinetics of second-order reactions 26

I Solution of the rate equation:

 [A]  [B]
− =− =  [A] [B] (2.91)
 

(1) Substitution:
 = ([A]0 − [A] ) = ([B]0 − [B] ) (2.92)
[A] = [A]0 −  (2.93)
y

=  ([A]0 − ) ([B]0 − ) (2.94)

(2) Integration using partial fractions (skipped here):



(3) General solution:


[B]0 [A]
= exp ([A]0 − [B]0 )   (2.95)
[A]0 [B]
This solution is rarely used in practice, however, compared to the approach below.

2.4.3 Pseudo first-order reactions

For experimental studies of second-order reactions of the type A + B → P, we like to


use a high excess of one of the reactions partners (e.g. [B] À [A]) so that [B] ≈ const
Under this condition, the reaction is said to be of pseudo first-order, because

 [B] = 0 (2.96)

so that the solution for [A ()] is simply


0
[A] = [A]0 −  = [A]0 −[B] (2.97)

I Experimental investigation of pseudo first-order reactions:

(1) Measurements of the pseudo first-order decay of [A] vs.  at different concentra-
tions of [B]: Plot of ln [A] vs.  give pseudo first-order rate constant 0 =  [B]
for each value of [B].
(2) Plot of 0 =  [B] vs. [B] gives straight line with slope .

⇒ We do not need the absolute concentration of [A], only the absolute concentration
of [B] is needed. This is a huge advantage if A is a highly reactive radical for which the
absolute concentration is hard to calibrate.
2.4 Kinetics of second-order reactions 27

I Figure 2.10: Kinetics of pseudo first-order reactions.

2.4.4 Comparison of first-order and second-order reactions

Consider a first-order and a second-order reaction with the same initial concentrations
and the same initial rate at  = 0. As shown in Fig. 2.11 below, the rate of the
second-order reaction rapidly slows down as time increases.

I Figure 2.11: Comparison of first-order and second-order reactions.


2.5 Kinetics of third-order reactions 28

2.5 Kinetics of third-order reactions

A+B+C→D (2.98)
y
 [A]
− =  [A] [B] [C] (2.99)


I Pseudo second-order combination reactions: In combination reactions of two


species A + A → A2 or A + B → AB (for example, the recombination of two free
radicals), one of the species (bath gas M) is usually present in large excess ([M] À [A],
[B]), so that the kinetics becomes pseudo second-order, giving the two cases:


A + A + M → A2 + M (2.100)


A + B + M → AB + M (2.101)

M plays the role of an inert collision partner which removes the excess energy from the
newly formed A2 or AB molecules.
2.6 Kinetics of simple composite reactions 29

2.6 Kinetics of simple composite reactions

2.6.1 Consecutive first-order reactions

1
A −→ B (2.102)
2
B −→ C (2.103)

I Coupled differential equations:

 [A]
= −1 [A] (2.104)

 [B]
= +1 [A] − 2 [B] (2.105)

 [C]
= +2 [B] (2.106)

In this case, since the DE’s are linear, there is an exact solution which we will examine
first. We will also look at an approximate solution which can be obtained with the
quasi steady-state approximation and at methods to obtain numerical solutions, which
we have to use for complex non-linear inhomogeneous DE systems.

I Mass balance:
[A] + [B] + [C] = [A]0 (2.107)

I Initial conditions:

[A ( = 0)] = [A]0 (2.108)


[B ( = 0)] = 0 (2.109)
[C ( = 0)] = 0 (2.110)

I Exact solution:

• First-order decay of [A]:


 [A]
= −1 [A] (2.111)

y
[A] = [A]0 −1  (2.112)

• Inhomogeneous DE for [B]:

 [B]
= +1 [A] − 2 [B] (2.113)

2.6 Kinetics of simple composite reactions 30

• Solution for [B ()]17 :



⎪ 1 [A]0 ¡ −1  ¢
⎨  − −2  if 1 =6 2
[B ()] = 2 − 1 (2.114)


1 [A]0  −1  if 1 = 2

• Reaction products by mass balance:

[C ()] = [A]0 − [A ()] − [B ()] =    (2.115)

I Figure 2.12: Kinetics of consecutive first-order reactions (case I).

17
The solution is detailed in Appendix C.
2.6 Kinetics of simple composite reactions 31

I Figure 2.13: Kinetics of consecutive first-order reactions (case II).

I Caveat: As seen, the general solution for [B] is the difference of two exponentials,

1 [A]0 ¡ −1  ¢
[B] =  − −2  , (2.116)
2 − 1
One exponential describes the rise, the other the fall of [B]. It is sometimes implicitly,
but wrongly assumed that the rise time corresponds to 1 and the decay time to 2 .
The truth is that we cannot tell just from the shape of the concentration-time profile
whether 1 or 2 correspond to the rise or to the fall of [B].

I Quasi steady-state approximation: Under the condition that

2 À 1 (2.117)

we can find a simple solution for the DE’s. Under this condition, after a short initial
induction time  (see Fig. 2.13), the change of [B] is very small compared to that of
[A]. Therefore, we have approximately

 [B]
≈0 (2.118)


This important approximation is known as the (quasi)steady-state approximation.18

18
Deutsch: Quasistationaritätsannahme.
2.6 Kinetics of simple composite reactions 32

I Quasi steady-state solution for [B (t)]: For    we have

 [B]
= 1 [A] − 2 [B] ≈ 0 (2.119)

y
1 [A] ≈ 2 [B] (2.120)
y
1 1
[B]ss = [A] = [A]0 −1  (2.121)
2 2

For    , [C ()] is therefore given by

 [C]
= 2 [B] ≈ 1 [A] (2.122)

y ¡ ¢
[C] = [A]0 1 − −1  (2.123)

2.6.2 Reactions with pre-equilibrium

1
A À B (2.124)
−1
2
B → C (2.125)

I Equilibrium between [A] and [B]:

1  −1 À 2 (2.126)

y
[B] 1
= = eq (2.127)
[A] −1
[B] = eq [A] (2.128)

I Time dependence of [C]:

 [C]
= 2 [B] = 2 eq [A] (2.129)

2.6 Kinetics of simple composite reactions 33

2.6.3 Parallel reactions

1
A −→ B (2.130)
2
A −→ C (2.131)
3
A −→ D (2.132)

y
 [A]
= − (1 + 2 + 3 ) [A] (2.133)

y
[A] = [A]0 −(1 +2 +3 ) (2.134)
and
1 [A]0 ¡ ¢
[B] = 1 − −(1 +2 +3 ) (2.135)
(1 + 2 + 3 )
2 [A]0 ¡ ¢
[C] = 1 − −(1 +2 +3 ) (2.136)
(1 + 2 + 3 )
3 [A]0 ¡ ¢
[D] = 1 − −(1 +2 +3 ) (2.137)
(1 + 2 + 3 )

Note that the decay of [A] is determined by the total removal rate constant  =
1 + 2 + 3

2.6.4 Simultaneous first- and second-order reactions*

1
A −→ C (2.138)
2
A + A −→ D (2.139)

y
 [A]
= −1 [A] − 22 [A]2 (2.140)

y µ ¶
1 22 22 1
=− + + exp (−1 ) (2.141)
[A] 1 1 [A]0

I Approximation for k1 t ¿ 1: Power series expansion of the exponential gives


µ ¶
1 1 1
≈ + + 22 ×  (2.142)
[A] [A]0 [A]0
2.6 Kinetics of simple composite reactions 34

2.6.5 Competing second-order reactions*

1
A + A −→ A2 (2.143)
2
A + B −→ P1 (2.144)
3
B −→ P2 (2.145)

I Solution for the case [A] À [B]:

(1)
 [A]
= −21 [A]2 − 2 [A] [B] (2.146)

≈ −21 [A]2 (2.147)
y
1 1
− = 21  (2.148)
[A] [A]0
or
¡ ¢−1
[A] = [A]0 −1 + 21  (2.149)
[A]0
= (2.150)
1 + 21 [A]0 

(2)
 [B]
= −2 [A] [B] − 3 [B] (2.151)

This is another inhomogeneous first-order DE, for which the solution can be found
as outlined in Appendix C, giving

[B] = [B]0 × (1 + 21 [A]0 )−2 21 × exp (−3 ) (2.152)

I Example:
1
CH3 + CH3 −→ C2 H6 (2.153)
2
CH3 + OH −→ CH2 + H2 O (2.154)
3
OH −→ P2 (2.155)

In the experimental study of reaction 2.154 (Deters 1998), the rate coefficient for the
reaction (2 according to Eq. 2.152) was fitted to measured OH concentration-time
profiles in the presence of high concentrations of CH3 using the Marquardt-Levenberg
nonlinear least squares fitting algorithm (Press 1992). The rate coefficients 1 and
3 were determined by independent measurements. The absolute CH3 concentrations,
which are needed for the evaluation of 2 according to Eq. 2.152, were obtained using
a quantitative titration reaction.
2.7 Temperature dependence of rate coefficients 35

2.7 Temperature dependence of rate coefficients

2.7.1 Arrhenius equation

I Arrhenius:19 Svante Arrhenius (Arrhenius 1889) found the following empirical relation
describing the temperature dependence of the reaction rate constant:

 ln  
=− (2.156)
(1 ) 

I Definition 2.4:  is called the Arrhenius activation energy of the reaction:

 ln 
 = − (2.157)
(1 )

I It is very important to keep in mind the following points:

• Eq. 2.156 is nothing else but a simple and very convenient way to describe the
empirical temperature dependence of reaction rate constants!
• Eq. 2.156 says that, within experimental error, a plot of ln  vs. 1 should give
a straight line. We will see later, however, that this is rarely exactly true. The
study and the understanding of non-Arrhenius behavior is in fact an important
topic in modern kinetics.
•  as determined by Eqs. 2.156 and 2.157 is therefore a purely empirical quantity!
• We will use Eq. 2.157 simply as the definition and recipy for determining  from
experimental data and to relate theoretical models to the experimental data!

I The rationale behind the Arrhenius equation (2.156): The rationale for the
Arrhenius equation is that molecules need additional energy so that chemical bonds can
be broken, atoms can rearrange, and new bonds can be formed (Fig. 2.14).
Van’t Hoff’s equation:
− ←
→ −
2  ln  2  ln

∆  =  =  (2.158)
 

− ←−
2  ln  2  ln 
=  −  (2.159)
 

− ←

=  −  (2.160)

Thus we can write


 ln  
= (2.161)
  2

19
Arrhenius is generally regarded as one of the founders of Physical Chemistry (together with Faraday,
van’t Hoff, and Ostwald).
2.7 Temperature dependence of rate coefficients 36

or20
 ln  
=− (2.164)
 (1 ) 

I Figure 2.14: Derivation of the Arrhenius equation.

I Experimental determination of E (see Fig. 2.14): A plot of ln  vs. 1 gives


a straight line with slope − 

I Integration of Eq. 2.156: Assuming that  is independent of  , Eq. 2.156 can be


integrated:

 ln  =  (2.165)
 2
y

ln  = − + const (2.166)

y
 ( ) =  −  (2.167)

Equation 2.167 is used to represent  ( ) over a limited  -range using the two para-
meters  and :

•  = Arrhenius activation energy (measure for the energy barrier of the reaction)
•  = pre-exponential factor (frequency factor, collision frequency).

20
(1 ) 1
=− 2 (2.162)
 
y
  = − 2  (1 ) (2.163)
2.7 Temperature dependence of rate coefficients 37

I It is important to realize, however, that the value of E is not equal to the


true height of the potential energy barrier for the reaction!
As said above,  is an empirical quantity! Its relation to the true threshold energy 0 ,
which is the minimum energy above which reaction may occur, will be a main subject
in the theory of bimolecular and unimolecular reactions (Sections 5 — 8). There is also
a statistical interpretation of  (see Section 8).

I Figure 2.15: The reaction HCO → H + CO (G. Friedrichs).

2.7.2 Deviations from Arrhenius behavior

In general,  is dependent on temperature,


√ i.e.  = ( ). For gas phase reactions,
for example, there is at least the  dependence of the hard sphere collision frequency.
The  -dependence of  is often small compared to  so that within experimental
error the dependence of  ( ) is determined by  .
If the  -dependence of  is not small compared to  , we will have to use a more
general expression for  ( ), as explained in the following.

I Generalized three-parameter expression for k (T ): To account for deviations


from Arrhenius because of  ( ), one conveniently employs the three parameter ex-
pression
( ) =    −0  (2.168)

with  and 0 now being  -independent quantities.


Range of  values for specific bimolecular reactions (justified by theory; see later):
− 15 ≤  ≤ +3 (2.169)
2.7 Temperature dependence of rate coefficients 38

I Question: What is the relation of Eq. 2.168 with the Arrhenius equation?
I Answer: We have to evaluate the Arrhenius parameters from Eq. 2.168 according to
the definitions of  and :

(1)  :

 ln 
 = + 2 (2.170)
 µ ¶
2  0
= + ln  +  ln  − (2.171)
 
y
 = 0 +  (2.172)
Only in the case that  ¿ 0 can we neglect the  term compared to 0 !
(2) : From the expression for  , we have

0 =  −  (2.173)

 ( ) =    −( − ) (2.174)


=     −  (2.175)
=  −  (2.176)

Comparing coefficients, we find

 =     (2.177)

2.7.3 Examples for non-Arrhenius behavior

Complex forming bimolecular reactions may show extreme non-Arrhenius behavior. An


extreme case is the reaction OH + CO → H + CO2 , which incidentally is the single
most important reaction in hydrocarbon combustion next to H + O2 → OH + O.
2.7 Temperature dependence of rate coefficients 39

I Figure 2.16: The reaction OH + CO → H + CO2 (Fulle 1996, Kudla 1992).

Another case is encountered if tunneling plays a role (for example, H+ transfer reactions
at low temperatures). In this case,  ( ) tends to a plateau value as  → 0 (Eisenberger
1991).
2.7 Temperature dependence of rate coefficients 40

2.8 References

Arrhenius 1889 S. A. Arrhenius, Über die Reaktionsgeschwindigkeit bei der Inversion


von Rohrzucker durch Säuren, Z. Physik. Chem. 4, 226 (1889).
Bevington 1992 P. R. Bevington, D. K. Robinson, Data Reduction and Error Analysis
for the Physical Sciences, McGraw-Hill, Boston, 1992.
Dertinger 1995 S. Dertinger, A. Geers, J. Kappert, F. Temps, J. W. Wiebrecht,
Rotation-Vibration State Resolved Unimolecular Dynamics of Highly Vibrationally
Excited CH3 O (2 ): III. State Specific Dissociation Rates from Spectroscopic Line
Profiles and Time Resolved Measurements, Faraday Discuss. Roy. Soc. 102, 31
(1995).
Deters 1998 R. Deters, M. Otting, H. Gg. Wagner, F. Temps, B. László, S. Dóbé,
T. Bérces, A Direct Investigation of the Reaction CH3 + OH at  = 298 K:
Overall Rate Coefficient and CH2 Formation, Ber. Bunsenges. Phys. Chem. 102,
58 (1998).
Eisenberger 1991 H. Eisenberger, B. Nickel, A. A. Ruth, W. Al-Soufi, K. H. Grell-
mann, M. Novo, Keto-Enol Tautomerization of 2-(20 -Hydroxyphenyl)benzoxazole
and 2-(20 -Hydroxy-40 -methylphenyl)benzoxazole in the Triplet State: Hydrogen
Tunneling and Isotope Effects. 2. Dual Phosphorescence Studies, J. Phys. Chem.
95, 10509 (1991).
Fulle 1996 D. Fulle, H. F. Hamann, H. Hippler, J. Troe, High Pressure Range of Addi-
tion Reactions of HO. II. Temperature and Pressure Dependence of the Reaction
HO + CO À HOCO → H + CO2 , J. Chem. Phys. 105, 983 (1996).
Homann 1995 K. H. Homann, SI-Einheiten, VCH, Weinheim, 1995.
Kudla 1992 K. Kudla, A. G. Koures, L. B. Harding, G. C. Schatz, A quasiclassical
trajectory study of OH rotational excitation in OH + CO collisions using ab initio
potential surfaces, J. Chem. Phys. 96, 7465 (1992).
Mills 1988 I. M. Mills et al., Units and Symbols in Physical Chemistry, Blackwell,
Oxford, 1988.
Press 1992 W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P. Flannery, Numerical
Recipes in Fortran, Cambridge University Press, Cambridge, 1992. Versions are
also available for C and Pascal.
Warnatz 1996 J. Warnatz, U. Maas, R. W. Dibble, Combustion. Physical and Chem-
ical Fundamentals, Modeling and Simulation, Experiments, Pollutant Formation,
Springer, Heidelberg, 1996.
2.8 References 41

3. Kinetics of complex reaction systems

Real reaction systems are usually much more complicated:

• Net reaction mechanisms may consist of ≈ 100 - 10 000 elementary reactions.


• No analytical solutions for the rate laws.
• In favorable cases, one may be able to use certain approximations to simplify the
reaction systems, like

— apply steady-state approximation where possible,


— take into account reactions in equilibrium,
— take into account microscopic reversibility (detailed balancing).

• For more acurate descriptions, however, one usually needs numerical solutions
for the rate equations. Using modern computers, this is not a problem even for
 10 000 reactions, as long as the transport processes are not too complicated
(as, for example, 1D reaction systems such premixed flames or 1D models of
atmospheric chemistry).
• However, numerical solutions are still extremely challenging for instationary 3D
reaction systems, such as

— 3D models of atmospheric chemistry including daily/annual variations and


couplings to the ocean,
— ignition processes (internal combustion engines).

3.1 Determination of the order of a reaction

Any reaction mechanism which we may postulate has to explain the observed reaction
order. Thus, the determination of the reaction order is a central topic.
We explore some general methods for determining the reaction order by considering an
example:
 [A]
=− =  [A] (3.1)


I First-order and second-order reactions: Test whether plots of ln  vs.  or 1 vs.
, give the straight lines expected for first-order or second-order reactions, respectively.

I Log-log plot of reaction rate vs. concentration: We see from Eq. 3.1 that a
log-log plot of  vs. [A] gives a straight line with slope :

lg  ∝  lg [A] (3.2)
3.1 Determination of the order of a reaction 42

I Half lifetime method (for n 6= 1):21

 [A]
− =  [A] (3.3)

y Z Z
 [A]
− =   (3.4)
[A]
y
1 1 1 1
− −1 = = + (3.5)
− + 1 [A]  − 1 [A]−1

Initial value condition at  = 0:

[A ( = 0)] = [A]0 (3.6)

y
1 1
= (3.7)
 − 1 [A]−1
0
y
1 1
−1 − = ( − 1)   (3.8)
[A] [A]−1
0

Half lifetime:
£ ¡ ¢¤ 1
A 12 = [A]0 (3.9)
2
y
2−1 1
−1 − = ( − 1)  12 (3.10)
[A]0 [A]−1
0
y
2−1 − 1 1
12 = (3.11)
( − 1) [A]−1
0
y
lg 12 ∝ −( − 1) lg [A]0 = (1 − ) lg [A]0 (3.12)

y A log-log plot of 12 vs. [A]0 gives a straight line with slope 1 − .

21
We leave aside a usually occuring stoichiometric factor ν here, since we can always define νk = k0
as rate constant k0 .
3.2 Application of the steady-state assumption 43

3.2 Application of the steady-state assumption

The steady state assumption allows us to reduce a large reaction mechanism to a smaller
one by eliminating the respective intermediate species.

3.2.1 The reaction H2 + Br2 → 2 HBr

The reaction between H2 and Br2 (Bodenstein et al.; Christiansen, Polanyi & Herzfeld)
runs as a thermal reaction or as a photochemically induced reaction (initiated by light).
In the latter case, it is a chain reaction with a quantum yield of
Φ ≈ 106 (3.13)

I Definition 3.1: Quantum yield Φ:


number of product molecules
Φ= (3.14)
number of absorbed photons

I Reaction mechanism (elementary reactions) describing the H2 -Br2 system:


Br2 → Br + Br chain initiation by thermal dissociation (1)
Br2 +  → Br + Br chain initiation by photolysis () (10 )
Br + H2 → HBr + H chain propagation (2)
H + Br2 → HBr + Br chain propagation (3)
H + HBr → H2 + Br inhibition (4) (3.15)
Br + HBr 6→ Br2 + H endothermic (∆  ª = +170 kJ mol) (6)
y reaction (6) can be neglected
+M
Br + Br → Br2 chain termination (5)

I Rate equations: With the steady state approximations for [Br] and [H], we obtain22
 [H]
= 2 [Br] [H2 ] − 3 [H] [Br2 ] − 4 [H] [HBr] ≈ 0 (3.16)

y 2 [Br] [H2 ] = 3 [H] [Br2 ] + 4 [H] [HBr] (3.17)
2 [Br] [H2 ]
y [H]ss = (3.18)
3 [Br2 ] + 4 [HBr]

 [Br]
= 21 [Br2 ] −2 [Br] [H2 ] + 3 [H] [Br2 ] + 4 [H] [HBr] −25 [Br]2 (3.19)
 | {z }
=−[H] =0
2
= 21 [Br2 ] − 25 [Br] ≈ 0 (3.20)
µ ¶12
1
y [Br]ss = [Br2 ] (3.21)
5
µ ¶12
1 2 [H2 ]
y [H]ss = [Br2 ] × (3.22)
5 3 [Br2 ] + 4 [HBr]

22
For the photolysis induced reaction, 1 has to be replaced by 10  , where  is the light intensity and
10 the corresponding photolysis rate constants (related to the absorption coefficient).
3.2 Application of the steady-state assumption 44

 [HBr]
= 2 [Br] [H2 ] + 3 [H] [Br2 ] − 4 [H] [HBr]


• Since Eq. 3.16 can be rewritten as

2 [Br] [H2 ] − 4 [H] [HBr] = 3 [H] [Br2 ] , (3.23)

 [HBr]   simplifies to
 [HBr]
= 2 3 [H] [Br2 ] . (3.24)


• Substituting [H] by [H]ss from Eq. 3.18, we obtain


µ ¶12
 [HBr] 1 2 [H2 ]
= 2 3 [Br2 ] × [Br2 ] × (3.25)
 5 3 [Br2 ] + 4 [HBr]
2 2 (1 5 )12 [H2 ] [Br2 ]12
= (3.26)
1 + (4 3 ) [HBr]  [Br2 ]

I Result:
 [HBr] 0 [H2 ] [Br2 ]12
= (3.27)
 1 + 00 [HBr]  [Br2 ]
with
0 = 2 2 (1 5 )12 and 00 = 4 3 (3.28)

I Conclusion: We see that HBr acts as an inhibitor of the total reaction.

3.2.2 Further examples

I Exercise 3.1: The thermal decomposition of acetaldehyde according to the overall


reaction
CH3 CHO → CH4 + CO (3.29)
is known to proceed via the following elementary steps (Rice-Herzfeld mechanism):

CH3 CHO → CH3 + HCO (1)


CH3 + CH3 CHO → CH4 + CH3 CO (2)
(3.30)
CH3 CO → CH3 + CO (3)
CH3 + CH3 → C2 H6 (4)

Derive an expression for the production rate of [CH4 ]. (Hint: Use the steady-state
approximations for [CH3 CO] and [CH3 ] and neglect [HCO] for the derivation of the
expression.) ¤
A more complete mechanism takes into account the thermal decomposition of the HCO
radicals according to HCO → H + CO (5) followed by the reaction H + CH3 CHO →
H2 + CH3 CO (6).
3.2 Application of the steady-state assumption 45

I Solution 3.1: With the steady state approximations for [CH3 CO] and [CH3 ], we obtain

 [CH3 CO]
= 2 [CH3 ] [CH3 CHO] − 3 [CH3 CO] ≈ 0 (3.31)

2 [CH3 ] [CH3 CHO]
y [CH3 CO]ss = (3.32)
3

 [CH3 ]
= 1 [CH3 CHO] − 2 [CH3 ] [CH3 CHO] (3.33)

+3 [CH3 CO] − 24 [CH3 ]2 (3.34)
= 1 [CH3 CHO] − 2 [CH3 ] [CH3 CHO] (3.35)
3 2 [CH3 ] [CH3 CHO]
+ − 24 [CH3 ]2 (3.36)
3
= 1 [CH3 CHO] − 24 [CH3 ]2 ≈ 0 (3.37)
µ ¶12
1 [CH3 CHO]
y [CH3 ]ss = (3.38)
24

and
 [CH4 ]
= 2 [CH3 ] [CH3 CHO]

y µ ¶
 [CH4 ] 1
= 2 [CH3 CHO]32 (3.39)
 24
¥
I Exercise 3.2: The thermal decomposition of ethane gives mainly H2 + C2 H4 . In
addition, small amounts of CH4 are formed. The decay of the C2 H6 concentration can
be described by the reaction steps

C2 H6 → CH3 + CH3 (1)


CH3 + C2 H6 → CH4 + C2 H5 (2)
C2 H5 → C2 H4 + H (3) (3.40)
H + C2 H6 → C2 H5 + H2 (4)
H + C2 H5 → C2 H6 (5)

Using the steady-state approximations for the radicals [CH3 ], [C2 H5 ], and [H], derive a
rate law describing the overall decay of [C2 H6 ]. Note: From the C-H bond dissociation
energies in C2 H6 and C2 H5 , it is known that 3 À 1 . ¤
I Solution 3.2:
" µ 2 ¶12 #
 [C2 H6 ] 3 1 1 3 4
− = 1 + + [C2 H6 ] (3.41)
 2 4 5

¥
3.3 Microscopic reversibility and detailed balancing 46

3.3 Microscopic reversibility and detailed balancing

The principle of microscopic reversibility is a consequence of the time reversal


symmetry of classical and quantum mechanics: If all atomic momenta of a molecule
that has reached some product state from an initial state are reversed, the molecule will
return to its initial state via the same path.
A macroscopic manifestation of the principle of microscopic reversibility is detailed
balancing which states that

(1) in equilibrium, the forward and reverse rates of a process are equal, i.e. for a
1
reaction A1 À A2 ,
−1
1 [A1 ] = −1 [A2 ] (3.42)
and
(2) if molecules A1 and A2 are in equilibrium and A2 and A3 are in equilibrium, both
with respect to each other, there is also equilibrium between A1 and A3 :
1 2
A1 À A2 À A3 (3.43)
−1 −2

y
3
A1 À A3 (3.44)
−3

Thus, we can write


[A2 ]eq 1
= = 12 (3.45)
[A1 ]eq −1
[A3 ]eq 2
= = 23 (3.46)
[A2 ]eq −2
[A3 ]eq 3
= = 13 (3.47)
[A1 ]eq −3
1 2
= 12 23 = (3.48)
−1 −2

Detailed balancing allows us to simplify complex reaction systems by eliminating  − 1


out of a sequence of  reversible reactions that are in equilibrium.
3.4 Generalized first-order kinetics* 47

3.4 Generalized first-order kinetics*

3.4.1 Matrix method*

For coupled complex reaction systems with only first-order reactions, the solution of
the rate equations can be reduced to an eigenvalue problem (Jost 1974). The solution
of the rate equations, i.e. the calculation of the time-dependent concentration profiles,
requires finding the eigenvalues and eigenvectors of the rate constant matrix of the
system.23
Eigenvalue problems are readily solved in practice using numerical methods (Press 1992).
We consider an example which we solved using conventional methods in Section 2.3.

I Example:
1
A1 À A2 (3.49)
2

The rate equations


 [A1 ]
= −1 [A1 ] + 2 [A2 ] (3.50)

 [A2 ]
= +1 [A1 ] − 2 [A2 ] (3.51)

can be written in a more compact form as a matrix equation:
⎛ · ⎞
µ ¶µ ¶
[A
⎝ · ⎠=
1 ] − 1 + 2 [A 1 ]
(3.52)
+1 −2 [A2 ]
[A2 ]

With the concentration vector


µ ¶
[A1 ]
A= (3.53)
[A2 ]
and the rate constant matrix
µ ¶
−1 +2
K= (3.54)
+1 −2
this becomes ·
A=K·A (3.55)

I Generalized matrix rate equation for coupled first-order reaction systems:


·
A=K·A (3.56)
To solve this matrix equation, as we shall see, we need the eigenvalues and the
eigenvectors of the K matrix. Both are readily obtained using a linear algebra com-
puter program.

23
A brief overview on matrices and determinants is given in Appendix D.
3.4 Generalized first-order kinetics* 48

I Ansatz: The ansatz for solving Eq. 3.56 assumes that we can express A in terms of
another vector B,24
A=P·B  (3.57)
where P · B means the dot product, using some rotation matrix P.

I Condition for P: P is required to diagonalize K via the eigenvalue equation

K · P = P · Λ. (3.58)

By multiplication from the left with P−1 , this condition yields

P−1 · K · P = Λ (3.59)

The eigenvalue matrix Λ is a diagonal matrix of the form


⎛ ⎞
1 0 0
Λ = ⎝ 0 2 0 ⎠  (3.60)
.
0 0 ..

As defined by Eq. 3.59, its elements on the diagonal {1  2    } are the eigenvalues of
the rate constant matrix. The matrix P is called the eigenvector matrix associated
with K. The columns of P are the respective eigenvectors p associated with the
respective eigenvalues  .

I Solution: Inserting the ansatz


A=P·B (3.61)
into our matrix equation25
·
A=K·A (3.62)
we have
 (P · B)
=K·P·B (3.63)

or ·
P·B=K·P·B (3.64)

Muliplication of this equation from the left by P−1 gives


·
P−1 · P · B = P−1 · K · P · B (3.65)

or ·
B=Λ·B (3.66)

This DE can be immediately integrated since Λ is diagonal. The solution is

B () = Λ · B0 (3.67)

24
The components of B will be seen to be linear combinations of the components of the concentration
vector A. B is immediately obtained once we have the matrix of eigenvectors P (see below).
25
The dot above a concentration vector indicates differentiation by .
3.4 Generalized first-order kinetics* 49

© ª
where Λ is a diagonal matrix with elements 1   2      , and Λ B0 means
element-wise multiplication. The result gives the time dependence of B, i.e. B(),
starting from the initial values B0 .
From B(), we obtain A () by transforming back into the original concentration basis
using
B = P−1 · A (3.68)
and
B0 = P−1 · A0 (3.69)
Inserting these expressions into into our solution for B () (Eq. 3.67), we obtain
P−1 · A = Λ · P−1 · A0 (3.70)
The final step is now a multiplication from the left with P which yields the solution for
Eq. 3.56
A = P · Λ · P−1 · A0 (3.71)

I Note: This may look complicated to someone who is not used to it. However, once
we have set up K, which is straightforward, everything is done by the computer: The
computer computes the

• eigenvalues of the rate constant matrix K,


• the eigenvector matrix P of the rate constant matrix K,
• all necessary inverse matrices, etc.,
• and does all the matrix multiplications.

All you usually have to do is to program the K matrix.

I Application to our example (solution “by hand”):

• Evaluation of the eigenvalue matrix Λ by solution of the eigenvalue equation


det (K − Λ · I) = 0 (3.72)
with I as unity matrix:
¯ ¯
¯ −1 −  +2 ¯
¯ ¯ (3.73)
¯ +1 −2 −  ¯ = 0
y
(−1 − ) · (−2 − ) − (1 · 2 ) = 0 (3.74)
1 2 + 1 + 2 + 2 − 1 2 = 0 (3.75)
 ( + (1 + 2 )) = 0 (3.76)
y
1 = − (1 + 2 ) (3.77)
2 = 0 (3.78)
y µ ¶
− (1 + 2 ) 0
Λ= (3.79)
0 0
3.4 Generalized first-order kinetics* 50

• Conclusions:

(1) The largest (negative) eigenvalue (1 in this case) determines the shortest
time scale and thus the short-time evolution of the reactant concentrations!
(2) The smallest (negative) eigenvalue (2 in this case) determines the longest
time scale of the reaction, i.e. the evolution of the reactant concentrations
at long times. The smallest (negative) eigenvalue thus determines the net
reaction rate!
(3) Note that in the example the concentrations at long times are constant
(equilibrium!) and therefore 2 = 0.

• Evaluation of the eigenvectors p1 und p2 by inserting the eigenvalues {1  2    }


one by one into
K·P=Λ·P (3.80)
yields, for each  and the respective p , a linear equation of the type

K · p =  · p (3.81)

y
(K −  ) · p = 0 (3.82)
µ ¶
1 
Naming the elements of each eigenvector for simplicity as p = , the
2 
linear equations are, explicitly,

(−1 −  ) 1 + 2 2 = 0 (3.83)


1 1 + (−2 −  ) 2 = 0 (3.84)

• Inserting 1 = − (1 + 2 ):

(−1 + 1 + 2 ) 11 + 2 21 = 0 (3.85)


1 11 + (−2 + 1 + 2 ) 21 = 0 (3.86)

2 11 + 2 21 = 0 (3.87)


1 11 + 1 21 = 0 (3.88)

y
11 = −21 (3.89)

This means that the elements 11  21 are determined only to a constant factor
:

11 =  (3.90)
21 = − (3.91)

y µ ¶

p1 = (3.92)
−
3.4 Generalized first-order kinetics* 51

• Inserting 2 = 0:

− 1 12 + 2 22 = 0 (3.93)


1 12 − 2 22 = 0 (3.94)

y
1 12 = 2 22 (3.95)
y
2 2
12 = · 22 = · (3.96)
1 1
22 =  (3.97)

where 22 has been arbitrarily set to a constant value .


⎛ ⎞
2
·⎠
p2 = ⎝ 1 (3.98)

y
• Eigenvector matrix (= rotation matrix P):
⎛ ⎞
2
 ·⎠
P=⎝ 1 (3.99)
− 

• The arbitrary factor  is determined by the initial conditions to be  = 1, so that


⎛ ⎞
2
1
P=⎝ 1 ⎠ (3.100)
−1 1
µ ¶
10
• Solution for A () with initial condition A0 = :
0

A = P · Λ · P−1 · A0 (3.101)

y
⎛ ⎞ ⎛ ⎞−1
2 µ −(1 +2 ) ¶ 2 µ ¶
1  0 1 
A = ⎝ 1 ⎠ 0
⎝ 1 ⎠
10
(3.102)
0  0
−1 1 −1 1
⎛µ ¶ ⎞ ⎛ ⎞
2 −(1 +2 ) 2 + 1 −(1 +2 )
⎜ 1 + 2 + 1 1 + 2 10 ⎟ ⎜ 1 + 2
10 ⎟
= ⎜

µ ¶ ⎟ = ⎜
⎠ ⎝ 1 − −(1 +2 )


1 −(1 +2 )
− 1 10 1 10
1 + 2 1 + 2 1 + 2
(3.103)

We shall verify this solution using computer algebra in the following paragraph.
3.4 Generalized first-order kinetics* 52

I Solution using symbolic algebra programs:

• rate constant matrix K: µ ¶


−1 +2
K= (3.104)
+1 −2

• eigenvalues(K) = {−1 − 2  0}:


µ ¶
− (1 + 2 ) 0
Λ= (3.105)
0 0

½µ ¶¾ (à 1 !)
−1 2
• eigenvectors(K) = ↔ −1 − 2  1 ↔ 0:
1 1
⎛ ⎞
2
−1
P=⎝ 1 ⎠ (3.106)
+1 1

• inverse(P): ⎛ ⎞
1 2

⎜ ⎟
P−1 = ⎝ 1+ 2 1 + 2 ⎠ (3.107)
1 1
1 + 2 1 + 2
• B0 = P−1 A0 :
⎛ ⎞
µ ¶ 1
−10
10 ⎜ 1 + 2 ⎟
B0 = P−1 · A0 = P−1 · =⎝ 1 ⎠ (3.108)
0
10
1 + 2

• Solution for B ():

B = Λ · B0 (3.109)
⎛ ⎞
µ −( + ) ¶ 1
− 10
 1 2 0 ⎜ 1 + 2 ⎟
= ·⎝ 1 ⎠ (3.110)
0 1
10
1 + 2
⎛ ⎞
1 −(1 +2 )
−10 ×
⎜ 1 + 2 ⎟
= ⎝ 1 ⎠ (3.111)
10 ×1
1 + 2
3.4 Generalized first-order kinetics* 53

• Solution for A ():


µ ¶
Λ −1 Λ −1 10
A = P ·  · P · A0 = P ·  · P · (3.112)
0
⎛ ⎞ ⎛ ⎞
µ −( + ) ¶ 1 2 µ ¶
2 −
−1  1 2
0 ⎜  +   +  ⎟ 
= ⎝ 1 ⎠ · ·⎝ 1
1
2 1 2
1 ⎠ ·
10
0 1 0
+1 1
1 + 2 1 + 2
(3.113)
⎛ ⎞ ⎛ ⎞
µ −( + ) ¶ 1
2 −10
−1  1 2
0 ⎜ 1 + 2 ⎟
= ⎝ 1 ⎠ · ·⎝ 1 ⎠ (3.114)
+1 1 0 1
10
1 + 2
⎛ µ ¶⎞
⎛ ⎞ −( + ) 1
2 ⎜
1 2
−10
⎝ −1 ⎠ ⎜ µ + 2 ⎟
1 ¶ ⎟
= 1 · ⎝ ⎠ (3.115)
1
+1 1 1 10
1 + 2
⎛ ⎞ ⎛ ⎞
2 −(1 +2 ) 2 + 1 −(1 +2 )
⎜ 10  +  + 10 1  +  ⎟ ⎜ 1 + 2
10 ⎟
= ⎜

1 2 1 2 ⎟ = ⎜ ⎟
1 −(1 +2 ) ⎠ ⎝ 1 − −(1 +2 ) ⎠
10 − 10 1 1 10
1 + 2 1 + 2 1 + 2
(3.116)

• This result is identical with the result for [A ()] obtained in Section 2.3, where
we had (after mapping the respective symbols)
2
A1 − A10
µ 1 + 2¶ = exp [− (1 + 2 ) ] (3.117)
2
1− A10
1 + 2
To show this, we rewrite the solution from Section 2.3 as
µ ¶
2 2
A1 = A10 + A10 1 − −(1 +2 ) (3.118)
1 + 2 1 + 2
µ µ ¶ ¶
2 2 −(1 +2 )
= + 1−  A10 (3.119)
1 + 2 1 + 2
µ ¶
2 + 1 −(1 +2 )
= A10 (3.120)
1 + 2
(3.121)
A2 = A10 − A1 (3.122)
µ ¶
2 + 1 −(1 +2 )
= A10 − A10 (3.123)
1 + 2
µ µ ¶¶
2 + 1 −(1 +2 )
= 1− A10 (3.124)
1 + 2
µ ¶
1 − −(1 +2 )
= 1 A10 (3.125)
1 + 2
3.4 Generalized first-order kinetics* 54

3.4.2 Laplace transforms*

The concept of Laplace and inverse Laplace transforms is extremely useful in chemical
kinetics because

(1) Laplace transforms provide a convenient method for solving differential equations,
(2) Laplace transforms make the connection between microscopic molecular properties
(i.e. microcanonical functions of energy, ) and statistically averaged quantities
(i.e. Boltzmann-averaged functions of temperature, / ).

I Definition 3.2: The Laplace transform L [ ()] of a function  () is defined as the
integral
Z∞
 () = L [ ()] =  () −   (3.126)
0

where

•  is a real variable,
•  () is a real function of the variable  with the property  () = 0 for   0,
•  is a complex variable,
•  () = L [ ()] is a function of the variable .

I Definition 3.3: The inverse Laplace transform L−1 [ ()] of the function  () is
defined as the integral
Z
+∞
1
 () = L−1 [ ()] =  ()   (3.127)
2
−∞

where

•  is an arbitrary real constant.

I Relation between f (t) and F (p): Since L−1 [ ()] recovers the original function
 (), the pair of functions  () and  () is said to form a Laplace pair:

L−1 [ ()] = L−1 [L [ ()]] =  () (3.128)

and £ ¤
L [ ()] = L L−1 [ ()] =  () (3.129)

A list of Laplace and inverse Laplace transforms of some functions can be found in
Appendix E (Table E.1).
3.4 Generalized first-order kinetics* 55

I Solution of differential equations using Laplace transforms: Consider the


Laplace transform of the derivative   ()   of a function  ():26
Z∞
  () −
L [  ()  ] =   (3.130)

0
Z∞
¯∞ ¡ ¢
=  () − ¯0 −  ()  − (3.131)
0
Z∞
= − ( = 0) +   () −  (3.132)
0
= −0 +   () (3.133)
We can solve this equation for  () via
L [  ()  ] + 0
 () = (3.134)

and recover the solution  () by inverse Laplace transformation,
∙ ¸
−1 L [  ()  ] + 0
 () = L (3.135)

I Example: Application to two consecutive first-order reactions. As an example,


we consider the DE’s for two consecutive first-order reactions
 
A →1 B →2 C (3.136)
We already solved this problem previously (in Section 2.6.1) using conventional methods,
and repeat the solution now using the Laplace transform method.
Towards these ends, in this example, we shall denote the concentrations [A ()] and
[B ()] as  () and  () and their Laplace transforms as  () and  (). The respective
initial concentrations shall be [A]0 = 0 and [B]0 = 0 = 0. Thus we have:

• DE for the concentration  ():


 = 1 0 −1  − 2  (3.137)

• Laplace transform:  () can be found using Eq. 3.133:


L [] = −0 +   () (3.138)
Taking the Laplace transform of the r.h.s. of Eq. 3.137:27
£ ¤
L [] = L 1 0 −1  − L [2  ()] (3.139)
£ ¤ 1 0
= 1 0 L −1  − 2 L [ ()] = − 2  () (3.140)
 + 1
= −0 +   () (Eq. 3.138) (3.141)

26
We used in line 2: integration by parts; in line 3: the derivative  (− )   = − − ; in line 4:
the definition  () = L [ ()] and the abbreviation  ( = 0) = 0 .
27 1
We used L [ ] = from Table E.1 and L [ ()] =  ().
−
3.4 Generalized first-order kinetics* 56

Since we have 0 = 0 from the initial condition, this becomes


1 0
  () + 2  () = (3.142)
 + 1

• Solution for  ():


1 1
 () = 1 0 (3.143)
 + 1  + 2

• Inverse Laplace transform:28

1 0 ¡ −1  ¢
 () = L [ ()] =  − −2  (3.144)
2 − 1

which is the same as obtained in Section 2.6.1.

∙ ¸
28 −1 1 1 ¡  ¢
From Table E.1, we see that L =  −  .
( − ) ( − ) ( − )
3.5 Numerical integration 57

3.5 Numerical integration

I Initial value problems: For complex reaction systems, we have to integrate the cou-
pled nonlinear DE systems numerically using the computer. We start with the known
·
initial values  of the concentrations at time 0 and the associated DE’s (  ) and want
to compute the values of  at time   0 . In numerics, these problems are known as
initial value problems. With additional conditions, for example for the spatial distrib-
utions of the concentration (fixed values at (0  0  0  1    )), they become boundary
value problems.
In solving these problems, we generally have to deal with systems of stiff nonlinear
coupled (ordinary) differential equations (DE systems are said to be stiff, if the
rate constants (more precisely: the eigenvalues) vary by many orders of magnitude;
under these conditons, the changes of the dependable variables differ by many orders of
magnitude).
Thus we have to solve

•  equations (kinetic equations + transport processes), with


•  variables (concentrations), and the given
• initial values and boundary values.

I Applications of numerical integration methods:

• Combustion chemistry:

— Stationary combustion (for example diffusion flames),


— Instationary combustion (ignition processes, turbulent combustion) - first
successful attempts have been made,
— Problem: For all but the simplest fuels, we have to deal with hundreds of
species and thousands of elementary reactions.

• Atmospheric chemistry:

— 2D “box” models ( ( )) are straightforward.


— 4D models are required (time, latitude, longitude, height)!
— Huge problems, when it comes to
∗ feedback with oceans, etc.?
∗ do we know all reactions?
∗ heterogeneous reactions?

• Physical organic chemistry:

— Elucidation of reaction mechanisms.

• Macromolecular chemistry (including industrial macromolecular synthesis):

— Prediction of chain length and other properties of polymers and co-polymers.


3.5 Numerical integration 58

I Survey of practical numerical integration methods:

• explicit methods (e.g. Runge-Kutta),


• implicit methods (e.g. Gear),
• extrapolation methods (e.g. Richardson, Bulirsch-Stoer, Deuflhard).

In the following, we consider the basics of the most convenient methods.

3.5.1 Taylor series expansions

The Taylor expansion is the basis for all numerical integration methods. We start from
the initial value of  (0 ) at some time 0

 (0 ) = 0 (3.145)

and want to determine the value  (0 + ) at time 0 + . Using the DE for  ,
·
 () =  (  ) , (3.146)

we expand  in a Taylor series around 0 :

· 2 ··
 (0 + ) =  (0 ) +  ·   (0  0 ) + ·  (0  0 ) +    (3.147)
2! 
2 ·
=  (0 ) +  · (0  0 ) + ·  (0  0 ) +    (3.148)
2!

Recursion formula:

+1 =  +  (3.149)
2
 ·
+1 =  +  ·  (   ) + ·  (   ) +    (3.150)
2!

The first-order term (   ) is given by the rate equation at  , the second-order
·
term (   ) (which is taken into account for example by the 2 order Runge-Kutta
method) and higher order terms (⇒ higher order Runge-Kutta methods) have to be
evaluated numerically by finite differences.

3.5.2 Euler method

The Euler method is based on a 1st order Taylor expansion. If the step size  is small
enough, the higher order terms can be neglected:
·
 (0 + ) =  (0 ) +  ·  (0 ) + O(2 ) (3.151)

Initial value problem:


·
 =  (  ) (3.152)
3.5 Numerical integration 59

with
 (0 ) = 0 (3.153)
·
Result by substituting  (0 ) =  (0   ):
¡ ¢
 (0 + ) =  (0 ) +   (0   ) + O 2 (3.154)

Recursion formula (see Fig. 3.1):

+1 =  +  (3.155)
+1 =  +   (   ) (3.156)

Problem: Large errors for practical step sizes.

I Figure 3.1: Euler method.

3.5.3 Runge-Kutta methods

I 2nd order Runge-Kutta method:


Initial value problem:
·
 =  (  ) (3.157)
with
 (0 ) = 0 (3.158)
3.5 Numerical integration 60

2nd order Runge-Kutta ansatz = Taylor expansion to 2nd order:

· 2 ··
 (0 + ) =  (0 ) +   (0 ) +  (0 ) + O(3 ) (3.159)
2! Ã !
· ·
· 2   (0 + ) −   (0 )
≈  (0 ) +   (0 ) + + O(3 ) (3.160)
2 
÷ ·
!
(0 + ) + (0 )
= (0 ) +  + O(3 ) (3.161)
2
| {z }
= slope of  () at midpoint

Recursion formula:

+1 =  +  (3.162)
+1 =  + 2 + O(3 ) (3.163)
µ ¶
 1
2 =    +   + (3.164)
2 2
1 =   (   ) (3.165)

I Figure 3.2: Illustration of the 2nd order Runge-Kutta method.


3.5 Numerical integration 61

I 4th order Runge-Kutta method: In practice, one frequently uses the 4th order
Runge-Kutta method.

I Figure 3.3: Illustration of the 4th order Runge-Kutta method.

3.5.4 Implicit methods (Gear):

The Runge-Kutta method becomes unstable for stiff DE systems. In this case, it is
preferable to employ implicit (also called iterative or backward) integration methods.
The starting point is the implicit form of the Euler method,
·
 (0 + ) =  (0 ) +   (0 + ) + O(2 ) (3.166)
·
For the initial value problem  =  (  ) with  (0 ) = 0 this leads to the recursion
formula:
+1 =  +  (3.167)
+1 =  +   (+1  +1 ) (3.168)

Using the so-called predictor-corrector methods, +1 is predicted in a first step by


starting from  using one of the explicit methods described above. The obtained trial
(0)
value +1 is then corrected by using a backward integration scheme (Gear 1971a, Gear
(1) (1)
1971a, Press 1992) to obtain a better estimate +1 . Then,  and +1 are used for
(2)
a new prediction-corrrection cycle, giving +1 . The procedure is repeated until the
difference between successive iterations falls below some tolerance limit .
The predictor-corrector method of Gear has been the method of choice for a long time
(Gear 1971a, Gear 1971a).
3.5 Numerical integration 62

3.5.5 Extrapolation methods (Bulirsch-Stoer)

The idea behind extrapolation methods is that the final answer for +1 is approximated
by an analytical function in the step size  (Richardson extrapolation). The best strategy
is to use a rational extrapolating function as the ansatz, and not a polynomial function
(Bulirsch-Stoer). This function is determined and fitted to the problem of interest by
performing calculations with various values of . The extrapolating function is then
used for extrapolating from  to +1 .
The optimal strategy for stiff DE systems is to follow the Bulirsch-Stoer method (Stoer
1980) in the form implemented by Deuflhard (Bader 1983, Deuflhard 1983, Deuflhard
1985). For Fortran subroutines, see (Press 1992).

I Figure 3.4: The principle of extrapolation methods.

I Exercise 3.3: Solve the coupled DEs for two consecutive first-order reactions29
1 2 
A −→ B −→ C (3.169)

(a) by exact integration using symbolic algebra software, and


(b) by numerical integration using suitable software.
(c) Plot your results using specific values for [A1 ]0 , 1 and 2 , e.g. [A1 ]0 = 1, 1 = 1
and 2 = 10 and compare them with plots of the exact solutions from Section 2.6.1.

29
To solve this problem, you need algebra software, e.g. Derive, MuPad, MathCad, Maple or Mathe-
matica. The exact and the numerical solutions are identical within numeric precision.
3.5 Numerical integration 63

(d) Repeat the exercise by numerical integration with an additional second-order reaction
included, e.g.
1 2 
A −→ B −→ C (3.170)
3
A + A −→ P (3.171)

with, e.g., [A1 ]0 = 1, 1 = 1 and 2 = 10 3 = 1 ¤

I Figure 3.5: Comparison of the exact solution (solid lines) and numerical solutions
(dashed blue) for simple consecutive reactions.
3.6 Oscillating reactions* 64

3.6 Oscillating reactions*

The reactant and product concentrations normally approach equilibrium in a smooth


(often exponenential) fashion. Under special circumstances, however, the concentra-
tions may also show oscillations and/or bistability as they approach equilibrium. Such
behavior can occur only if there is a chemical feedback by autocatalysis.

3.6.1 Autocatalysis

I Example for autocatalysis:



A+B→B+B (3.172)

I Rate law:
 [A]
− =  [A] [B] (3.173)


To solve this DE, we substitute for [B ()] using the mass balance relation

[A]0 − [A ()] = [B ()] − [B]0 (3.174)

and introduce  = [A ()] as reaction progress variable and write

[B ()] = [A]0 + [B]0 − [A ()] (3.175)


= [A]0 + [B]0 −  (3.176)

This gives the DE



− =   ([A]0 + [B]0 − ) (3.177)

y
Z Z

= −  (3.178)
([A]0 + [B]0 )  − 2
0 0

I Solution (see Fig. 3.6):


µ ¶
1 [A]0 [B ()]
ln = −  (3.179)
[A]0 + [B]0 [B]0 [A ()]
or
[A]0 + [B]0
[B ()] = (3.180)
1 + ([A]0  [B]0 ) exp [− ([A]0 + [B]0 ) ]
3.6 Oscillating reactions* 65

I Conclusions:

• As can be seen from Eq. 3.180, and illustrated in Fig. 3.6, the concentration-time
profile shows the transition from an initial state at  = 0, where [B (0)] = [B]0 ,
to a final state at  → ∞, where [B (∞)] = [A]0 + [B]0 .
• The rate of increase of [B] increases up to the inflection point ∗ , then slows down
again as [B (∞)] is approached.
• The shape of this sigmoid (-shaped) curve is typical for a population increase
from a low plateau to a new (higher) plateau after a favorable change of the
environment due to, e.g. the availability of more food, a technical revolution, a
reduction of mortality by a medical breakthrough, . . .

Many interesting games which can be played based on these ideas are described in (Eigen
1975).

I Figure 3.6: Transition from an initial to a final steady state in an autocatalytic reaction.

3.6.2 Chemical oscillations

The presence of one or more autocatalytic steps in a reaction mechanism can lead to
chemical oscillations.
3.6 Oscillating reactions* 66

a) The Lotka mechanism


The origin of chemical oscillations can be understood by considering the reaction mech-
anism proposed by Lotka in 1910 (see Lotka 1920):

A + X →1 X + X (1)

X + Y →2 Y + Y (2) (3.181)

Y →3 Z (3)

• Both reactions (1) and (2) are autocatalytic.


• The Lotka mechanism illustrates the principle, but it does not correspond to an
existing chemical reaction system.
• A real oscillating reaction is the Belousov-Zhabotinsky reaction; this reaction may
be described using a more complicated mechanism (see below).

I Boundary condition for [A] in a flow reactor: In order to model the resulting
chemical oscillations, we assume that the reaction takes place in a flow reactor, which
is constantly supplied with new A such that the concentration of A stays constant
([A] = [A]0 ), while the product Z is constantly removed.

I Rate equations and steady-state solutions:


 [X]
= +1 [A]0 [X] − 2 [X] [Y] (3.182)

 [Y]
= +2 [X] [Y] − 3 [Y] (3.183)


Steady-state solutions:
 [X]
= +1 [A]0 [X] − 2 [X] [Y] = 0 (3.184)

 [Y]
= +2 [X] [Y] − 3 [Y] = 0 (3.185)


Dividing these equations by [X] and [Y], respectively, we find


2 [Y]ss = 1 [A]0 (3.186)
2 [X]ss = 3 (3.187)
Since [A] = [A]0 , the steady state solutions for [X] and [Y] are independent of time!

I Displacements from steady-state solutions: What happens if the concentrations


are displaced from the steady-state values by small amounts  and ?

(1) Ansatz of small displacements of the concentrations:


[X] = [X]ss +  (3.188)
[Y] = [Y]ss +  (3.189)
3.6 Oscillating reactions* 67

(2) Rate equations:


= +1 [A]0 ([X]ss + ) − 2 ([X]ss + ) ([Y]ss + ) (3.190)

= +1 [A]0 [X]ss + 1 [A]0  − 2 [X]ss [Y]ss
| {z }
=1 [A]0 [X]ss

−2 [X]ss  − 2 [Y]ss  − 2  


| {z }
=1 [A]0 

= −2 [X]ss  − 2   (3.191)


= +2 ([X]ss + ) ([Y]ss + ) − 3 ([Y]ss + ) (3.192)

= 2 [X]ss [Y]ss + 2 [X]ss  + 2 [Y]ss 
+2   − 3 [Y]ss − 3 
|{z} |{z}
=2 [X]ss =2 [X]ss

= +2 [Y]ss  + 2   (3.193)

(3) Simplified DE’s for small displacements: We may neglect the terms contain-
ing  . Thus, we obtain the two coupled first-order DE’s
·
 = −2 [X]ss   (3.194)
·
 = +2 [Y]ss   (3.195)

or, with  = 2 [X]ss and  = 2 [Y]ss ,


·
 = −  (3.196)
·
 = +  (3.197)

(4) Conversion of the first-order DE’s (Eqs. 3.196 and 3.197) to a single
second-order DE:

a) From Eq. 3.196, we have


1· · 1 ··
=−  y =−  (3.198)
 
·
b) Substitution for  into Eq. 3.197:

· 1 ··
 = +  y −  = +  (3.199)

c) Resulting second-order DE:

··
 = −  (3.200)
3.6 Oscillating reactions* 68

(5) Solution of the second-order DE by the ansatz

 =   (3.201)
·
y  =    (3.202)
··
y  =  2  , (3.203)

y
 2  = −   (3.204)
y
2 = − (3.205)
y

 = ± ()12 (3.206)
¡ ¢12
= ± 22 [X]ss [Y]ss (3.207)
= ± (1 3 [A]0 )12 (3.208)
= ±  (3.209)

with
 = (1 3 [A]0 )12 (3.210)

Note that, since 1 3 [A]0 is positive, the solutions for  are purely imaginary!
(6) Result:
 () =   +  − = 0 cos  (3.211)

(7) Oscillating [X(t)] and [Y(t)]: Since  () = [X ()] − [X]ss , we see that

[X ()] = [X]ss + 0 cos  (3.212)

(and likewise, but 90◦ out of phase, [Y ()]) oscillate with frequency

 = (1 3 [A]0 )12 . (3.213)

This means that if the system is displaced from its steady-state by some amount
, the species concentrations [X ()] and [Y ()] will start to oscillate.
The displacements may even grow in magnitude until the amplitude reaches the
limiting value 0 (see limit cycle diagram in Fig. 3.7).
3.6 Oscillating reactions* 69

I Figure 3.7: Limit cycle diagram for chemical oscillations according to the Lotka mech-
anism.

b) The Brusselator30
Reaction scheme:

A →1 X (1)

B+X →2 R+Y (2)
 (3.214)
Y+2 X →3 3X (3)

X →4 S (4)
Here, A and B are reactants, R and S are final products, and X and Y are intermediates.
The autocatalytic reaction is reaction (3).
Rate equations (using “dimensionless time”  = 4 ) and steady-state concentrations:
 [X] · 1 [A] − 2 [B] [X] + 3 [X]2 [Y] − 4 [X]
= [X] = (3.215)
 4
· 2
 [X] 2 [B] [X] − 3 [X] [Y]
= [Y] = (3.216)
 4
Steady-state concentrations:
1 [A]
[X]ss = (3.217)
4
2 4 [B]
[Y]ss = (3.218)
1 3 [A]

30
The name ‘Brusselator’ stems from the workplace of its inventor Prigogine (Brussels). The model
does not correspond to a real chemical system.
3.6 Oscillating reactions* 70

We now investigate the effect of small displacements from steady-state by  [X] and
 [Y]. The reaction rates respond to these displacements according to
⎛ · ⎞ ⎛ · ⎞
⎝  [X] ⎠  [X] and ⎝  [Y] ⎠  [Y] (3.219)
 [X]  [Y]
ss ss
y
• The steady-state is stable if
⎛ ·
⎞ ⎛ ·

⎝  [X] ⎠ + ⎝  [Y] ⎠  0 (3.220)
 [X]  [Y]
ss ss

• The steady-state is unstable if


⎛ · ⎞ ⎛ · ⎞
⎝  [X] ⎠ + ⎝  [Y] ⎠ ≥ 0 (3.221)
 [X]  [Y]
ss ss

• Stability criterion:
⎛ · ⎞ ⎛ · ⎞
2 2
⎝  [X] ⎠ + ⎝  [Y] ⎠ = 2 [] − 1 3 [] − 1 (3.222)
 [X]  [Y] 4 42
ss ss

c) Belousov-Zhabotinsky reaction
Oxidation of malonic acid to CO2 by bromate, catalyzed by cerium ions in acidic solution:
2BrO− +
3 + 3CH2 (COOH)2 + 2H −→ 2BrCH(COOH)2 + 3CO2 + 4H2 O (3.223)
The reaction is probed by addition of the redox indicator ferroin/ferriin: The color varies
between red and blue depending on the concentrations of Ce4+ and Ce3+ 

I Figure 3.8: Mechanism of the Belousov-Zhabotinsky reaction.


3.6 Oscillating reactions* 71

I Figure 3.9: Phases I - III of the Belousov-Zhabotinsky reaction.

I Figure 3.10: Key steps of the Belousov-Zhabotinsky reaction.


3.6 Oscillating reactions* 72

d) The Oregonator31
The Oregonator model can be used to describe the Belousov-Zhabotinsky reaction.

I Schematic model and kinetic equations:



X+R
A+Y →1
(1)

2R
X+Y →2
(2)

A+X →3
2 X + 2 Z (3) (3.224)

A+R
2X →4
(4)
5 1
B+Z→ Y (5)
2
with A = BrO− 3  B = CH2 (COOH)2 + BrCH(COOH)2  R = HOBr X = HBrO2 
Y = Br−  Z = Ce4+ .
Using dimensionless time ( = 5 [B] ), we obtain
 [X] 1 [A] [Y] − 2 [X] [Y] + 3 [A] [X] − 24 [X2 ]
= (3.225)
 5 [B]
1
 [Y] − 1 [A] [Y] − 2 [X] [Y] + 5 [B] [Z]
= 2 (3.226)
 5 [B]
 [Z] 23 [A] [X] − 5 [B] [Z]
= (3.227)
 5 [B]

I Figure 3.11: Numerical integration of the Oregonator model (rate constants in


l mol−1 s−1 : 1 = 0005, 2 = 10, 3 = 10, 4 = 10, 5 = 0004; initial con-
centrations in mol l−1 : [A] = 001, [B] = 10, [X]0 = 00, [Y]0 = 00, [Z]0 = 000025.)

31
The name ‘Oregonator’ stems from the workplace of its inventor Noyes (University of Oregon).
3.6 Oscillating reactions* 73

3.7 References

Bader 1983 G. Bader, P. Deuflhard, Numerische Mathematik 41, 373 (1985)


Deuflhard 1983 P. Deuflhard, Numerische Mathematik 41, 399 (1985).
Deuflhard 1985 P. Deuflhard, SIAM Review 27, 505 (1985).
Eigen 1975 M. Eigen, R. Winkler, Das Spiel, Piper, München, 1975.
Gear 1971a C. W. Gear, Commun. Am. Comput. Mach. 14, 1976 (1971).
Gear 1971b C. W. Gear, Numerical Initial Value Problems in Ordinary Differential
Equations, Prentice-Hall, Englewood Cliffs, 1971.
Houston 1996 P. L. Houston, Chemical Kinetics and Reaction Dynamics, McGraw-
Hill, Boston, 2001.
Jost 1974 W. Jost, in Physical Chemistry: An Advanced Treatise, Ed.: W. Jost, Vol.
VIa, Academic Press, New York, 1974.
Lotka 1920 A. J. Lotka, Undamped Oscillations Derived from the Law of Mass Action,
J. Am. Chem. Soc. 44, 1595 (1920).
Press 1992 W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P. Flannery, Numerical
Recipes in Fortran, Cambridge University Press, Cambridge, 1992.
Scott 1994 S. K. Scott, Oscillations, Waves, and Chaos in Chemical Kinetics, Oxford
Science Publications, Oxford, 1994.
Stoer 1980 J. Stoer, R. Bulirsch, Introduction to Numerical Analysis, Springer, New
York, 1980.
3.7 References 74

4. Experimental methods

In this section we consider experimental techniques for determining rate coefficients.


We’ll get to faster reactions as we go down the list.
Special techniques used in quantum state-resolved reaction dynamics, photodissociation,
or transition state spectroscopy (molecular beams, state specific detection, photofrag-
ment translational spectroscopy, photofragment imaging, other pump-probe techniques,
fluorescence up-conversion, fs-CARS, fs-DFWM) will be considered later (in advanced
courses), but we’ll briefly touch photochemistry and “femtochemistry”.

4.1 End product analysis

• first step for setting up reaction mechanisms,


• all convenient analytical techniques (HPLC, GC, GC-MS, MS, UV, FTIR, . . . ),
• Problem: No “direct” kinetic information because of lack of time resolution.

I Figure 4.1: FTIR spectrometer for kinetic studies of photolysis-induced reactions.


4.2 Modulation techniques 75

4.2 Modulation techniques

• especially for photo-induced reactions.


• photolysis light has to be modulated at a frequency that is comparable to that of
the reaction y time scale of 10’s of ms and longer.
• kinetic information is derived from the “in-phase” and the “in-quadrature” com-
ponents of the reactant and product concentrations as function of modulation
frequency.

I Figure 4.2:
4.3 The discharge flow (DF) technique 76

4.3 The discharge flow (DF) technique

DF technique is used for gas phase reactions on ms timescale:

• laminar flow along tubular reactor:


∆
∆ = (4.1)

with  = mean flow speed, ∆ = distance along the flow tube.
• (pseudo)-first order reaction gives exponential decay of concentration along ∆.
• in reality ⇒ Hagen-Poisseuille flow (parabolic radial flow profile),
• but fast radial diffusion of reactants (at pressures of 1 to 10 mbar) gives “plug
shaped” radial concentration profiles despite parabolic flow profile.
• more complicated, when “back diffusion” and radial diffusion have to be taken
into account (at higher pressures).
• problem: heterogeneous reactions lead to first-order decay of reactive radicals
even in the absence of a reactant.

Combinations with numerous highly sensitive detection techniques, for instance:

• mass spectrometry (MS): universal, but often low sensitivity; problems with frag-
mentation of molecules in the ion source,
• laser induced fluorescence (LIF): excitation to electronically excited state followed
by relaxation by spontaneous emission. very sensitive (e.g. for OH (2 Σ ← 2 Π)
detection limit below 106 molecules cm3 ),
• electron spin resonance (ESR): for paramagnetic atoms by exploiting the Zeeman
effect depending on magnetic field 
 = 0 +     (4.2)
∆ = 1 y ∆ =    (4.3)
ESR transitions are in the microwave region (X band: 9 GHz).
• laser magnetic resonance (LMR): very sensitive; based on observation of rota-
tional transitions between Zeeman-shifted rotational levels of paramagnetic radi-
cals (rot = rotational constant).
 00 = rot  ( + 1) + 00  00  (4.4)
 0 = rot  ( + 1) + 0  0  (4.5)

∆ = ±1 ∆ = 0 ±1 : (4.6)


0 00 00 0 0
y  = rot ( + 1) ( + 2) +     (4.7)
00 0 0 00 00
y ∆ =   = 2rot ( + 1) +   (  −   ) (4.8)
LMR transitions are in the FIR region (100 m ≤  ≤ 2000 m).
4.3 The discharge flow (DF) technique 77

I Figure 4.3: Schematic diagram of a DF/MS combination.

I Figure 4.4: Schematic diagram of a DF/LIF combination.


4.3 The discharge flow (DF) technique 78

I Figure 4.5: Principles of Electron Spin Resonance (ESR) and Laser Magnetic Reso-
nance (LMR).

I Figure 4.6: Laser Magnetic Resonance (LMR).


4.3 The discharge flow (DF) technique 79

I Figure 4.7: Schematic diagram of a DF/LMR combination

I Figure 4.8: Flow reactor with laser flash photolysis and time-resolved mass spectro-
metric detection of transient reactants and products for “real-time” kinetics.
4.4 Flash photolysis (FP) 80

4.4 Flash photolysis (FP)

• “pump-probe” technique:

— generation of transient species by flash photolysis using fast flash lamp or


pulsed lasers (“pump”),
— fast detection of transient species by UV absorption spectrometry, laser in-
duced fluorescence (LIF), cavity ringdown spectroscopy (CRDS) (”probe”).

• for gas and liquid phase reactions on nanosecond timescale (Norrish & Porter).

I Figure 4.9: Typical “pump-probe” set-up for kinetics studies by pulsed laser photolysis
and detection by laser-induced fluorescence.
4.5 Cavity Ringdown Spectroscopy (CRDS) 81

4.5 Cavity Ringdown Spectroscopy (CRDS)

• CRDS is an ideal method for absorption spectroscopy with pulsed lasers despite
of their large (±5 %) amplitude fluctuations.
• The absorption measurement is transformed into the time domain:

— time dependence of signal transmitted through a resonantor:

 = 0 exp (− ) (4.9)

— decay of intensity for pass through cavity (length , mirror reflectivity ,


absorption path length ):
µ ¶
 1− 
=−  −  (4.10)
  | 
{z }
| {z }
∆ due to reflection losses at mirrors ∆ due to absorption

— with
 =   (4.11)
y
 
= − [(1 − ) + ]  (4.12)
 
— empty cavity ( = 0):
 (1 − )
 −1
0 = (4.13)

— with absorber ( 6= 0):

 [(1 − ) + ]
 −1 = (4.14)

— The difference of the ringdown times measures the absorption coefficient


and is related to absorber concentration [A] by

 −1 −  −1
0 =  =  [A]  (4.15)

• Performance of a CRD spectrometer:

— mirrors:
 = 99999% (4.16)
— measured empty cavity decay time:

 0 ≈ 300 s (4.17)

— effective absorption pathlength (1 m cell):

eff = 90 km (4.18)

• applications to chemical kinetics:


4.5 Cavity Ringdown Spectroscopy (CRDS) 82

— slow reactions: pump-probe scheme with variable time delay.


first-order reaction with rate constant 1 :

[A] = [A]0 −1  (4.19)

y ¡ ¢
ln  −1 −  −1
0 =  − 1  (4.20)
— fast reactions: kinetics and ringdown occur on the same time scale. y
(extended) Kinetics and Ringdown model ((e)SKaR); see Brown2000 and
Guo2003 for details.

I Figure 4.10:
4.5 Cavity Ringdown Spectroscopy (CRDS) 83

I Figure 4.11:
4.6 Shock tube studies of high temperature kinetics 84

4.6 Shock tube studies of high temperature kinetics

• Detection of atoms by atomic resonance absorption spectrometry (ARAS),


• Detection of small radicals by frequency modulation absorption spectrometriy
(FM-AS)

I Figure 4.12:

I Figure 4.13:
4.7 Stopped flow studies 85

4.7 Stopped flow studies

• the analogue of the flow tube technique for liquid phase reactions.
• applications mainly kinetics of enzyme reactions, studies of protein folding dy-
namics

I Figure 4.14:
4.8 NMR spectroscopy 86

4.8 NMR spectroscopy

• line coalescence
• good for interconversion of two conformers (“isomerization”)

I Figure 4.15:
4.9 Relaxation methods 87

4.9 Relaxation methods

• application to fast reactions in the liquid phase (Eigen)


•  -jump, -jump, -jump, . . . :

(1) ⇒ displacement from equilibrium


(2) ⇒ measurements of relaxation into new equilibrium state by transient ab-
sorption spectroscopy, conductivity, . . .
4.10 Femtosecond spectroscopy 88

4.10 Femtosecond spectroscopy

I Figure 4.16: Femtosecond pump-probe spectroscopy for the investigation of ultrafast


reactions (time resolution down to the order of 10 fs).

I Figure 4.17: Signals in broadband femtosecond transient absorption spectroscopy.


4.10 Femtosecond spectroscopy 89

I Figure 4.18: Broadband femtosecond transient absorption spectroscopy of pho-


tochromic molecular switches in SFB 677.

I Analysis of femtosecond time-resolved data: In order to determine the fastest


time scales of chemical reactions, the analysis of the temporal signals measured in a
pump-probe experiment needs to take into account the experimental time resolution.
The latter is usually limited by the duration of the pump and probe laser pulses and by
other experimental factors. Under such conditions, the experimentally observable time
profile  () is given by the convolution of the “true” molecular signal  () with the
so-called instrument response function (IRF, here  ()):
Z ∞
 () =  ( )  ( −  )  = [ ∗ ] () (4.21)
−∞

In general, such convolution integrals must be evaluated numerically.


If the IRF can be described by a (normalized) Gaussian centered at 0 with width  (1
standard deviation) µ ¶
1 ( − 0 )
 () = √ exp − (4.22)
2 2 2
and the molecular signal by an exponential with decay time  and amplitude  (or a
sum of such exponentials with parameters {    }), the convolution integral becomes
µ 2 ¶µ µ ¶¶
   − 0 0 −  + ( 2  )
 () = exp − 1 − erf √ (4.23)
2 2 2  2
(or a corresponding sum of such terms). In practice, one uses a non-linear least-squares
fitting routine to fit the parameters  , , 0 , and  to the measured experimental data
following Eq. 4.23.
4.10 Femtosecond spectroscopy 90

I Figure 4.19: Broadband femtosecond transient absorption spectroscopy of Disperse


Red 1.

I Figure 4.20: Broadband femtosecond transient absorption spectroscopy of Furyl


Fulgides.
4.10 Femtosecond spectroscopy 91

I Figure 4.21: Ultrafast Electronic Deactivation in DNA Building Blocks.

I Figure 4.22: Femtosecond fluorescence up-conversion data for DNA strands.


4.10 Femtosecond spectroscopy 92

I Figure 4.23: Ultrafast Electronic Deactivation in DNA Building Blocks: d(pApA) as


Model System.
4.10 Femtosecond spectroscopy 93

4.11 References

Busch 1999 K. W. Busch, M. A. Busch, Cavity-Ringdown Spectroscopy, ACS Sym-


posium Series, Vol 720, American Chemical Society, Washington, 1999.
Brown 2000 S. S. Brown, A. R. Ravishankara, H. Stark, J. Phys. Chem. A 104, 7044
(2000).
Friebolin 1999 H.Friebolin, Ein- und zweidimensionale NMR-Spektroskopie, Wiley-
VCH, Weinheim, 1999.
Guo 2003 Y. Guo, M. Fikri, G. Friedrichs, F. Temps, An Extended Simultaneous Ki-
netics and Ringdown Model: Determination of the Rate Constant for the Reaction
SiH2 + O2 . Phys. Chem. Chem. Phys. 5, 4622 (2003).
OKeefe 1988 A. O’Keefe, D. A. G. Deacon, Rev. Sci. Instrum. 59, 2544 (1988).
4.11 References 94

5. Collision theory

So far, we have considered more or less complex chemical reactions following more or
less complicated rate laws, but we have only considered the rate coefficients appearing
in the rate laws as empirical quantities, which had to be determined experimentally.
We shall now ask the question, whether we can rationalize and predict those rate coef-
ficients using theoretical models. Towards these ends, we have to look at the reactions
at a microscopic, molecular level. In particular, we have to

• consider the relation between the microscopic (individual molecule) properties and
the macroscopic (thermally averaged) rate quantities (rate constants, product
branching ratios)
• develop models for the microscopic (molecular level) progress of individual ele-
mentary reactions, including the different molecular degrees of freedom (quantum
states).
• develop sufficiently simplified models averaging over the molecular degrees of
freedom which allow us to make practically useful, relevant, and still sufficiently
accurate predictions.

We shall do so in the next four chapters.

I Collision theory: To begin with, this chapter considers bimolecular reactions in the
gas phase. These obviously require a collision between two molecules as condition for a
reaction to take place.

• We start with the hard sphere collision model of structureless particles as the
simplest model for bimolecular reactions. Although this model has two major
flaws, it leads to the right result because of a compensation of errors.
• We then consider results of kinetic gas theory, which allow us to properly take
into account the molecular speed distributions.
• We apply the hard sphere collision model to understand transport processes in the
gas phase (diffusion, heat conductivity, viscosity).
• We then proceed with a correct derivation of advanced collision theory models
starting from differential cross sections (which can be measured in crossed molec-
ular beam experiments). Advanced collision theory is the method of choice for fast
gas phase reactions which can be studied in detail by molecular beam experiments
and handled by quantum theory.
5.1 Hard sphere collision theory 95

5.1 Hard sphere collision theory

5.1.1 The gas kinetic collision frequency

I The gas kinetic collision frequency as upper limit for rate constants of bimole-
cular reactions: The gas kinetic collision frequency constitues an upper limit for the
rate constant of bimolecular gas phase reactions. We assume that the molecules A and
B are structureless hard spheres (radii  ,  ) moving around with fixed mean speeds
 and  and look at the number of collisions between A and B per unit time.

I Figure 5.1: Derivation of the hard sphere gas kinetic collision frequency.

I Discussion of the molecular quantities affecting the rate constant:

• collision parameter (= projected distance between centers of mass):

 (5.1)

• maximal collision parameter = max  value leading to a collision max :

 ≤ max =  +  =  (5.2)

• collision cross section  = 2 (= reactive cross section)32 :

  = 2 =  ( +  )2 (5.3)

32
Deutsch: reaktiver Wirkungsquerschnitt.
5.1 Hard sphere collision theory 96

• mean speed of A and B:


r r
8  8
 = = (5.4)
 
r r
8  8
 = = (5.5)
 

• mean relative speed (2 different molecules, i.e. A 6= B, A—B collisions):33


r
8 
 = (5.9)


with the reduced mass


 
= (5.10)
 + 

• mean relative speed (single type of molecules, i.e. A = B, A—A collisions):


  1
= =  (5.11)
 +  2
y r
√ 8 
 = 2× (5.12)


I Hard sphere gas kinetic collision frequency:

• We want to determine the number of collisions  of a single molecule A with


molecules B in time  . This is given by the product of the (volume that A has
swept in  ) × (number density of B):

— volume of cylinder that A has swept in time  :


volume = cross section ×   (5.13)
y
2 ×   (5.14)

33
The mean relative speed of two molecules with velocities u and u is found by looking at the
square of the difference speed

2 = |u − u |2 = | |2 + | |2 − 2 | | | | cos  (5.6)

The third term averaged over cos  is 0, therefore

3  3  3  ( +  ) 3 


2 = | |2 + | |2 = + = = (5.7)
    
³ ´12
Since  and 2 differ only by a constant factor, we also find that
s
8 
 = (5.8)

5.1 Hard sphere collision theory 97

— number density of B
 (5.15)
— Number of collisions of a single A in time  : ⇒ number of molecules B in
that volume
 = 2 ×  ×  (5.16)

• Number of all collisions between A and B:

 = 2 ×  ×   (5.17)

• Modification for collisions between identical molecules (collisions must not be


counted twice):
1
 = × 2 ×  ×   (5.18)
2√
2
= × 2 ×  ×   (5.19)
2

• Comparison with rate expression:


 
− =− =  ×   =  (5.20)
 
y
• Expression for rate constant for collisions A + B:

max = 2 ×  =   ×  (5.21)

y
µ ¶12
 2 8 
max =  × (5.22)


• Modification for collisions between identical molecules (A + A):



 1 2 2 2
max =  ×  =  ×  (5.23)
2 2

These expressions give the upper limits for the rate constants of bimolecular
reactions. The upper limits are reached only if

• there are no energetic restrictions (no energy barriers),


• all collisions lead to reaction (no steric constraints),
• molecules can be approximated as hard spheres (unrealistic, usually one has   
  ),
• all molecules collide with the same relative speed (in reality there is a distribution
of speeds).
5.1 Hard sphere collision theory 98

I Mean free pathlength:

• The mean free pathlength is the distance that a molecule flies between two suc-
cessive collisions:
distance distance/time 
= = = (5.24)
# of collisions # of collisions/time 

• For collisions between identical molecules (A + A):



= √ (5.25)
2 2 
y
1
= √ (5.26)
2
2 × 

I Time between two collisions:


1 1
= 2 (5.27)
   × 

I Example 5.1: Collision frequency for O2 - N2 @  = 298 K  = 1000 mbar:


Molecular parameters:34

2 = 22 = 3467 Å (5.28)


2 = 22 = 3798 Å (5.29)
y  = 36 Å = 36 × 10−10 m (5.30)
2
  = 2 = 41 Å (5.31)
28 × 32 1
 = × 10−3 kg mol × (5.32)
28 + 32 
s
8 
 = = 650 m s (5.33)

cf. :  (O2 ) = 444 m s (5.34)

Rate constant:
µ ¶12
 2 8 
max =  × (5.35)

= 26 × 10 cm3  s
−10
(5.36)
= 16 × 1014 cm3  mol s (5.37)

34
Note that we write the Loschmidt number  = 6 022 1 × 1023 mol−1 for the Avogadro number
 in this lecture to avoid confusion with the number density of A molecules  .
5.1 Hard sphere collision theory 99

Mean free pathlength:


 = 1000 mbar (5.38)
 = 40 × 10−5 mol cm3 (5.39)
= 25 × 1025  m3 (5.40)
 = 36 × 10−10 m (5.41)
 = 1 × 10−7 m (5.42)

Rule-of-thumb:
 = 03 mbar y  = 03 mm (5.43)

Time between collisions:


 = 1000 mbar (5.44)
1 1
= 2
(5.45)
   × 
= 15 × 10−10 s (5.46)
≈ 01 ns (5.47)
¤

5.1.2 Allowance for a threshold energy for reaction

I The threshold energy E0 : The majority of reactions exhibit a threshold energy, be-
low which the reaction does not occur (quantum mechanical tunneling is excluded here).
The molecules must have a minimum energy min to overcome this reaction threshold en-
ergy 0 . For our structureless spherical molecules, we consider only translational energy
 :35
  0 : no reaction (5.48)
  0 : reaction (5.49)

I Boltzmann distribution: Fraction of molecules with translational energy   0 :


µ ¶ µ ¶
(  0 ) 0 0
= exp −  = exp −  (5.50)
   

I Resulting rate constant:


µ ¶12
 2 8 
max =  × × −0   (5.51)

or µ ¶12
 2 8 
max =  × × −0  (5.52)


35
We use  and 0 to indicate the energies in molecular units, and  and 0 for the respective
values in molar units.
5.1 Hard sphere collision theory 100

I Arrhenius activation energy: Applying the definition of the apparent (Arrhenius)


activation energy  , we find that

 ln  1
 =  2 = 0 +  (5.53)
 2


As we see,  is not constant but (slightly)  -dependent: The slope of a plot of ln max
−1 −1
vs.  increases as  → ∞ (or  → 0). The Arrhenius preexponential factor 
would not be constant as well, it is also slightly  -dependent.

5.1.3 Allowance for steric hindrance

I The sterical factor p: In addition, even if   0 , not every collision will lead to
a reaction, because real molecules are not structureless: A reaction requires a specific
orientation of the molecules. This is taken into account by introducing a steric factor
:
µ ¶12
 2 8 
max =  ×  × × −0  (5.54)

5.2 Kinetic gas theory 101

5.2 Kinetic gas theory

5.2.1 The Maxwell-Boltzmann speed distributions of gas molecules in 1D and 3D

I Figure 5.2: Experimental measurement of molecular speed distribution using a rotating


disk velocity selector and an effusive molecular beam (orrifice diameter ¿ ).

I Rotating disk velocity selector:

• Flight time of molecules with velocity  between two rotating disks (distance ):

1 = (5.55)


• Time delay of second slit with respect to first slit ( = angle,  = angular velocity):

2 = (5.56)

• Transmission only if
1 = 2 (5.57)
y 
 =  × (5.58)

In practice, one uses a number of velocity selectors in series. A measurement of the 1D


( )
speed distribution ( )  =  using this method gives

5.2 Kinetic gas theory 102

I The 1D Maxwell-Boltzmann speed distribution (Fig. 5.3):


r µ ¶
 2
 ( )  = × exp −  (5.59)
2  2 

Z∞
Note that this distribution is already normalized for  ( )  = 1 We check this
−∞
by substituting
r r r
  2 
=  y  =  y  =  (5.60)
2  2  
y
Z∞ r Z∞ r Z∞
 2  −2 1 2 1 √
 ( )  =   = √ −  = √  = 1 (5.61)
2    
−∞ −∞ −∞

I Figure 5.3: 1D Maxwell-Boltzmann speed distribution.


5.2 Kinetic gas theory 103

I Velocity distribution in 3 dimensions: Since the velocities in    direction are


independent, the probability that a molecule has a velocity between  and  +  ,
 and  +  , and  and  +  is given be the product of  ( )  , ( )  ,
and ( )  :

(     )
   = ( ) ( )  ( )   
 Ã
µ ¶32 ¡ 2 2 2
¢!
   +   + 
= × exp −    (5.62)
2  2 

We are, however, not interested in the distribution of velocities pointing into a specific
direction (     ), but in the probability that a molecule has a certain speed || in
any direction, i.e. that the velocity vector ends somewhere on a sphere with radius
q
|| = 2 + 2 + 2 (5.63)

To obtain this distribution, we must integrate over all polar angles  and . This
corresponds to replacing the volume element    with the volume 42  of
a spherical shell with radius  and thickness :

   → 42 

I The 3D Maxwell-Boltzmann speed distribution (Fig. 5.4):


µ ¶32 µ ¶
 2 2
 ()  = × 4 × exp −  (5.64)
2  2 

Z∞
Note that this distribution is also normalized for  ()  = 1 We check this by
0
substituting
r r µ ¶12
  2 
 =  y  =  y  =  (5.65)
2  2  
2  2
2 =  (5.66)

y
Z∞ µ ¶32 Z∞ µ ¶12
 2  2 −2 2 
 ()  = × 4    (5.67)
2   
0 0
µ ¶32 Z∞ µ ¶32
1 2 −2 1 1√
= 4 ×    = 4 × ×  = 1 (5.68)
  4
0
5.2 Kinetic gas theory 104

I Figure 5.4: 3D Maxwell-Boltzmann speed distribution.

I Notice the differences between the following quantities:


• Value at maximum of  ():
r
2 
max = (5.69)

• Mean speed:
r
8 
= (5.70)

• Root mean square value:
³ ´12 r 3 

2 = (5.71)

I Table 5.1: Mean molecular speed values for some gases @  = 298 K

gas  [ m s]
H 2511
H2 1776
He 1255
O2 444
CO2 379
Hg 178
238
U 163
5.2 Kinetic gas theory 105

I Kinetic energy distribution function: We may substitute  for  in Eq. 5.64:


1 2
 =  (5.72)
2
µ ¶12
2
y = (5.73)

µ ¶12 µ ¶12
 2 1 −12 1
y = × × ( ) = ( )−12 (5.74)
  2 2
µ ¶12
1
y  = ( )−12  (5.75)
2

Insertion into Eq. 5.64 gives (see Fig. 5.5)


µ ¶32 µ ¶ µ ¶12 µ ¶12
 2 −   1 1
 ( )  = × 4 × ×  (5.76)
2   2 
y
µ ¶32
1
 ( )  = 2 ( )12 −    (5.77)
 

I Figure 5.5: Maxwell-Boltzmann distribution function for the kinetic energy.

• The fraction of molecules with   0 depends on temperature!


5.2 Kinetic gas theory 106

5.2.2 The rate of wall collisions and the ideal gas law (ideal gas pressure)

Earlier,36 we derived the ideal gas law by calculating the rate of wall collisions of the
molecules of a gas with mass  and number density  in a container of volume  and
1
the resulting gas pressure  using a simplified model assuming that of all molecules
r 6
8 
have a fixed speed  = in + direction. In this simple picture, we neglected

the 3D molecular velocity distribution. We obtained the correct result only due to a
1 1 2
cancelation of two errors (assumption of + =  instead of + =  and kin =
6 4 2
2
instead of kin = ). In the following,
2

(1) we briefly review the simplified picture,


(2) we then proceed to derive the correct results by appropriate averaging over the
velocity/speed distribution.

I Simplified model for the calculation of the pressure of an ideal gas (Fig. 5.6):

I Figure 5.6: Simplified model for the calculation of the pressure of an ideal gas.

36
See lecture notes for PC-1.
5.2 Kinetic gas theory 107

• In the time , all gas molecules (mass ) in the volume  +  hit the wall
area . We assume (first mistake) that
1
+ =  (5.78)
6
• Number of wall collisions  in time  (with number density ):
1
 = ×  ×  ×   (5.79)
6
• Momentum change  due to wall collision, force on wall, and pressure:
1
 = 2  ×    (5.80)
6
• Gas pressure:
 1  1 1
 = = = × 2  ×  (5.81)
    6
y
1
 =   2 (5.82)
3

• Result for 1 mol, with  = ( = 6022 1 × 1023 mol−1 ,  = molar volume)

2
and (second mistake) kin = :
2
1 2 2 3
  =   2 =  kin =  ×   =  (5.83)
3 3 3 2

I Correct derivation of the gas kinetic wall collision frequency:

I Figure 5.7: On the derivation of the gas kinetic wall collision frequency.
5.2 Kinetic gas theory 108

• Number
p of molecules with velocities      in a volume of  +  with
 = 2 + 2 + 2 (cf. Fig. 5.7) that will hit the area  in the time :
 (+ )  ( )  ( )
 (+     ) =   +     (5.84)
  
• To find the number of all molecules hitting  in , we have to integrate over all
positive  and over all positive and negative  and  , using the 1D-Maxwell-
 ( )
Boltzmann distributions  ( )  =  , etc.:

µ ¶32

 =    ×
2 
Z∞ Z∞ Z∞ µ ¶ µ ¶ µ ¶
2 2 2
 exp − exp − exp −    (5.85)
2  2  2 
0 −∞−∞

We know that the distributions over  and  are normalized to 1. Hence, we


only have to solve the integral over  , which we do by the substitution
µ ¶12 µ ¶12 µ ¶12
  2 
=  y  =  y  =  (5.86)
2  2  
y
Z∞ µ ¶ Z∞ µ ¶12 µ ¶12
2 2  −2 2 
 exp −  =   (5.87)
2   
0 0
Z∞
2  2 2  1  
=  −  = × = (5.88)
  2 
0

• Result:37
µ ¶12 µ ¶12
    
 =    × =    × (5.89)
2   2
y µ ¶12
  
wall = =× (5.90)
  2
This can be rewritten using the mean molecular speed , because
µ ¶12 µ ¶12
1 16  1 8  1
wall =  × × =× × =  × ×  (5.91)
4 2 4  4
with µ ¶12
8 
= (5.92)


µ ¶12
37 
Note that two of the three factors in Eq. 5.85 disappear by the integrations over 
2 
µ ¶12

and  , only one factor remains.
2 
5.2 Kinetic gas theory 109

• We thus obtain the final result for the gas kinetic collision frequency
1
wall =  (5.93)
4
1 1
with the infamous factor of rather than .
4 6

I Correct derivation of the ideal gas pressure and ideal gas law:

• Likewise, we take the differential number of wall collisions for specific speeds (Eq.
5.84) to compute the total momentum change  (+     ) on wall collisions
in time 
 (+     ) = 2 + ×   +  ×  (+ ) +  ( )   ( ) 
(5.94)
• Inserting again the 1D-Maxwell-Boltzmann distributions and integrating over all
positive  and all    , we find the total momentum change in time :
µ ¶32

 =    × 2×
2 
Z∞ Z∞ Z∞ µ ¶ µ ¶ µ ¶
2 2 2 2
 exp − exp − exp −   
2  2  2 
0 −∞−∞
(5.95)


• Pressure  for  = :

µ ¶12 Z∞ µ ¶
   2 2
 = = × 2  exp −  (5.96)
   2  2 
0

Using again the substitution


µ ¶12 µ ¶12 µ ¶12
  2 
=  y  =  y  =  ,
2  2  
(5.97)
we obtain
µ ¶12 Z∞ µ ¶12
  2  2 −2 2 
 = × 2    (5.98)
 2   
0
Z∞
 −12 2
=  × 4  × 2 −  (5.99)

0
 −12 1
=  × 4  ×  12 (5.100)
 4

=   (5.101)

y
 =  (5.102)
5.2 Kinetic gas theory 110

Thus, we have fixed our earlier simple derivation of the ideal gas law by appropriately
integrating over the Maxwell-Boltzmann velocity distribution of the molecules.
5.3 Transport processes in gases 111

5.3 Transport processes in gases

5.3.1 The general transport equation



I Definition of the flux J of a quantity: We consider a generic transport property

Γ (5.103)

which differs in magnitude along the  direction, i.e. has the gradient

grad Γ (5.104)

The flux of Γ through an area  in time  is defined as



→ Γ
Γ= (5.105)
 
and can be written as −

 Γ = − grad Γ (5.106)
Kinetic gas theory allows us to derive a general equation for the proportionality constant
.

I Figure 5.8: Derivation of the general transport equation.


5.3 Transport processes in gases 112

I The general transport equation: We consider the net flux of Γ through an area 
at  = 0 that is transported by gas molecules coming from an area above at  = 0 + 
and from below from an area above at  = 0 − (where  is the mean free pathlength).
Denoting the respective transport quantity per molecule as Γ, these partial fluxes are
given by the number of gas kinetic collisions hitting on  ( = 0 ) times the Γ carried
by each gas molecule (number density ):

1
  (Eq. 5.93), we obtain
• Using the gas kinetic wall collision frequency wall =
4
µ µ ¶ ¶
1 Γ
Γ+ =   ×  × Γ0 +   (5.107)
4 
and µ µ ¶ ¶
1 Γ
Γ− =   ×  × Γ0 −   (5.108)
4 
• Net effect: µ ¶
1 Γ
Γ = Γ+ − Γ− =   ×  ×   (5.109)
2 

• Net flux (with − sign to account for the fact that the flux is directed along the
downhill gradient of Γ):
µ ¶

→ Γ 1 Γ
Γ= = −  (5.110)
  2 

with  given by (see the derivation of wall above)


µ ¶12
8 
= (5.111)


and from Eq. 5.26, for only one type of gas molecules A (i.e. A - A collisons),

1
= √ (5.112)
22 × 

• If  also varies along  (diffusion), Eq. 5.110 may be recast by defining the overall
0
transport quantity Γ from the transport quantity per molecule Γ as
0
Γ =Γ (5.113)

into the form à !


0

→ 1 Γ
 Γ = −  (5.114)
2 

In the following,
³ we shall´ apply Eqs. 5.110 or 5.114 to determine the self-diffusion coef-
0 ¡ ¢ ¡ ¢
ficient Γ = 1 Γ =  , the heat conductivity Γ =  , and the viscosity Γ = 
of gases.
5.3 Transport processes in gases 113

5.3.2 Diffusion

I Figure 5.9: The diffusion in a mixture A + B of one sort of molecules A against a



time-independent concentration gradient is described by Fick’s first law.


• Fick’s first law: µ ¶



→  
= = − (5.115)
  

• In order to derive an expression for the self-diffusion coefficient  (diffusion of an


isotope of A in normal A), we make the ansatz
0
Γ = 1 Γ =  (5.116)
y µ ¶

→ 1 
  = −  (5.117)
2 
• Result:
1
=  (5.118)
2
• Fick’s second law (time-dependent concentration gradient):
  2 
= (5.119)
  2
• Einstein’s diffusion law in 1D:
∆ 2 = 2 (5.120)
• Einstein’s diffusion law in 3D:
∆2 = 6 (5.121)
5.3 Transport processes in gases 114

5.3.3 Heat conduction

I Figure 5.10: Heat conduction is described by Fourier’s law.

• Fourier’s law (for constant temperature gradient):


µ ¶

→  
= = − (5.122)
  

• Ansatz:
Γ =  = internal energy per molecule (5.123)
y
µ ¶

→ 1 
  = −  (5.124)
2 
µ ¶µ ¶
1  
= −  (5.125)
2  
µ ¶µ ¶
1   
= −  (5.126)
2  
µ ¶
1 
= −     (5.127)
2 

• Result:
1
 =     (5.128)
2
with  = number density and  = specific heat per molecule.
5.3 Transport processes in gases 115

5.3.4 Viscosity

I Figure 5.11: Newton’s law describes the inner friction in laminar flow.

• Newton’s law: µ ¶
 
= − (5.129)
 

• Ansatz:

Γ =  = momentum per molecule in  direction (5.130)

y
µ ¶

→ 1  ( )
  = −   
2 
1  ( ) 1  
= = =
    

• Result:
1
 =  (5.131)
2
with  = number density and  = mass of molecule.
5.4 Advanced collision theory 116

5.4 Advanced collision theory

The hard sphere model is a very primitive model. In reality, we have to account for the
following:
• The speed distribution follows the Maxwell-Boltzmann distribution, i.e.
 →  ()  →  ( )  (5.132)
• The reaction probability depends on exact details of the collision, i.e.
hard sphere model →  (     ) (5.133)
• Molecules are not hard spheres with   = 2 but “soft”, i.e. the reactive cross
section depends on the relative speed (or kinetic energy)
  → () → ( ) (5.134)
• Molecules in different quantum states  may behave differently, i.e.
( ) → ( ;  ) (5.135)

I Examples: The complexity of the problem is illustrated in Fig. 5.12 by two examples:
D(D ) + H2 (H2  H2  H2 ) −→ HD(HD  HD  HD ) + H(H ) (5.136)

OH(OH  OH  Ω OH ) + H2 (H2  H2  H2 )


→ H2 O(H2 O  1  2  3    ) + H(H ) (5.137)
As can be seen, the reaction rate is expected to depend on  (speed, translational
energy),  (vibrational quantum numbers ( 1 = symmetric stretch,  2 = bend,  3 =
antisymmetric stretch)),      (rotational quantum numbers), Ω (fine structure
state, electronic state). In addition, the products are scattered into the spatial angles
 .

I Figure 5.12: Quantum state-specific elementary reactions.


5.4 Advanced collision theory 117

We shall now derive the fundamental equation of collision theory that allows us to
account for those effects.

5.4.1 Fundamental equation for reactive scattering in crossed molecular beams

I Figure 5.13: Derivation of the fundamental equation of collision theory.

• Consider the following reactive scattering process (Fig. 5.13):

A+B→C (5.138)

• Intensity of beam of A:
 =  ×  (5.139)
Is this reasonable? ⇒ Look at the physical dimensions of  ×  :
£ ¤ £ ¤
[ cm s] × molecules/ cm3 = molecules cm2 s = [intensity] (5.140)

• Depletion of intensity of A due to scattering by B:

 = −  ×  () ×  ×   (5.141)


↑ ↑ ↑
     

y rate equation:

= −()    (5.142)

5.4 Advanced collision theory 118

• Rate coefficient for a given value of :

() = ()  (5.143)

• Averaging over the speed distribution  () leads to the fundamental equation
of all improved collision theories:

 ( ) = h  ()i (5.144)

or
R∞
 ( ) =   ()  (  )  (5.145)
0

• Frequently used approximation: Since the explicit form of  () is usually not
known, we often write Eqs. 5.144 and 5.145 in simplified, separated form as

( ) = hi h()i (5.146)

with
µ ¶12
8 
hi = (5.147)

and
h()i = -averaged (reactive) cross section (5.148)

5.4.2 Crossed molecular beam experiments

I Differential and integral reactive cross sections: Experiments with crossed mole-
cular beams allow us to measure differential and integral reactive cross sections
(Herschbach 1987, Lee 1987).
We consider the reaction

 (   ) +  (   ) −→  (   ) +  (   ) (5.149)

where  denotes the speeds of the molecules and  stands for the quantum numbers
of their energy states (electronic, vibrational, rotational, nuclear spin). In an ideal
experiment (which has yet to be performed), all  of the reactants are defined by, e.g.,
using some specific laser excitation scheme, and the  of the products are determined
by some probe laser. The  can be determined in a molecular beam experiment using
a rotating disk velocity selector (see Fig. 5.14).
5.4 Advanced collision theory 119

I Figure 5.14: Measurement of differential cross sections in crossed molecular beam


experiments.

I Differential and integral cross sections from crossed molecular beam experi-
ments:

• We define the differential cross section  (         ) for scattering


of products C into the solid angle element Ω at   such that

 (         ) Ω


=  (         )

×  ×  ×  (   ) ×  ×  (   ) ×  Ω (5.150)

• Corresponding integral cross sections are obtain by integration over  and .


• Depending on the scattering process of interest, we distinguish between

— elastic cross sections ( elast. ): measure of the size of the molecules,


— inelastic cross sections ( inelast. ): measure of the efficiency of energy transfer
processes,
— reactive cross sections ( react. ): measure of reactivity.
5.4 Advanced collision theory 120

I Figure 5.15: Measurement of differential cross sections in crossed molecular beam


experiments.
5.4 Advanced collision theory 121

5.4.3 From differential cross sections σ (. . .) to reaction rate constants k (T )

I Figure 5.16: The strategy.

I 1 → 2: Integration over angles:


Z2 Z
0 0
  ( ) = (  ) (   ) sin    (5.151)
0 0

I 2 → 3: Summation over all product quantum states α0 :


X 0
  ( ) =   ( ) (5.152)
0

I 3 → 4: Thermal averaging over initial quantum states α:


X
 ( ) =  ×   ( ) (5.153)

with the Boltzmann factor  and partition function 


 −    −  
 = P = (5.154)
 −   
X
=  −   (5.155)

y
X  −  
 ( ) =   ( ) × (5.156)


5.4 Advanced collision theory 122

I 5 → 6: Thermal averaging over all initial quantum states α:


X
 ( ) =  ×  ( )

X  −  
= ×  ( )


Z
 −  
= ×  ( ) 


I 3 → 5 and 4 → 6: Laplace Transformation:


Z∞
 ( ) ∝  ×  ( ) × −    (5.157)
0

This expression follows from the fundamental equation (Eq. 5.145) by substituting 
for . The exact transformation (giving the correct factors before the integral) will be
carried out below (see Eq. 5.169).

5.4.4 Laplace transforms in chemistry

Formally, the transformation (Eq. 5.157) is a Laplace transformation (see Appendix E).
A Laplace transform generally describes the transformation from a microscopic (“mi-
crocanonical”) quantity such as  ( ) to a macroscopic (“macrocanonical”, i.e. tem-
perature dependent) quantity such as  ( ).

I Definition of the Laplace Transformation: The Laplace transform L [ ()] of a


function  () is defined as the integral
Z∞
 () = L [ ()] =  () −  (5.158)
0

I Inverse Laplace Transformation:

Z
+∞
1
 () = L−1 [ ()] =  ()   (5.159)
2
−∞

with an arbitrary real number .


5.4 Advanced collision theory 123

I Notes:

(1) The so far best known Laplace transform to us is the partition function
X Z ∞
− 
 () =   =  () −   (5.160)
 0

with
1
= (5.161)
 
Thus, for continuous state densities,  ( ) is normally written as

Z∞
P
 ( ) =   −   =  () −   . (5.162)
0

(2) While the pathway from  ( ) to  ( ) by forward Laplace transformation is


straightforward, it is not generally straightforward to derive  ( ) by inverse
Laplace transformation from  ( ), because  ( ) would be needed from  = 0
to  = ∞.
5.4 Advanced collision theory 124

5.4.5 Bimolecular rate constants from collision theory

I Transformation from velocity to energy space: Taking our results from kinetic
gas theory, we are now in the position to apply the fundamental equation of collision
theory (Eq. 5.145)
Z∞
( ) =  ()  (  )  (5.163)
0

using various functional forms for . Towards these ends, we first insert  () from Eq.
5.64, with the reduced mass  instead of , to obtain
µ ¶32 Z∞ µ ¶
 3 2
( ) = 4  () exp −  (5.164)
2  2 
0

We then transform from  to  by substituting for  according to


1 2
 =  (5.165)
2
µ ¶12
2
y = (5.166)

µ ¶12 µ ¶12 µ ¶12 µ ¶12
 2 1 1 1 1
y = = (5.167)
  2  2 
µ ¶12 µ ¶12
1 1
y  =   (5.168)
2 
which gives
µ ¶12 µ ¶32 Z∞ µ ¶
1 2 
( ) =   ( ) exp − 
    
0 (5.169)

The integration from 0 to ∞ takes into account that a reaction can occur only if
  0 .
This expression is interpreted as follows:

• The function  −   describes the fraction of molecules with energy  .
• The function  ( ) in the integrand is called the excitation function38 . This
can, in general, be a rather complicated (and not necessarily smooth) function
(see Fig. 5.17).
• The complete integrand   ( ) −   is called the reaction function39 ,
because it describes the reactive fraction of the molecules.
• The rate constant  ( ) corresponds to the area under the reaction function in
Fig. 5.18.

38
“Anregungsfunktion”
39
“Reaktionsfunktion”
5.4 Advanced collision theory 125

I Figure 5.17: Advanced collision theory: The excitation function.

I Figure 5.18: Advanced collision theory: The reaction function.


5.4 Advanced collision theory 126

I Non-equilibrium energy distribution:* The above expression for  ( ) is valid if


the reacting molecules are in internal equilibrium, i.e. the energy states are populated
according to the Boltzmann distribution.
Deviations from the Boltzmann distributions will occur in the case of very fast reactions,
if the reaction is faster than the relaxation between the energy states. In this case,  ( )
decreases, because the mean energy of the reacting molecule decreases.
Mean energy of the reacting molecules (see Fig. 5.19):

X Z∞
=   = ()  (5.170)
 0

(1) For thermal equilibrium:


() = () (5.171)
with () being the Boltzmann distribution.
(2) Non-equilibrium:
()  () (5.172)

y
non-eq.  eq. (5.173)

I Figure 5.19: Mean energy of the reacting molecules in thermal equilibrium and for a
non-equilibrium situation.

I Conclusion: The non-equilibrium distribution leads to a reduced rate coefficient.


5.4 Advanced collision theory 127

a) Advanced collision theory for hard spheres (Fig. 5.20)

I Figure 5.20: The hard sphere model.

(1) Ansatz for the reactive cross section:


½
2   0
 ( ) = (5.174)
0   0

(2) Inserting this ansatz into Eq. 5.169:


µ ¶12 µ ¶32 Z∞ µ ¶
HS 1 2 2 
 ( ) =   exp −  (5.175)
    
0

(3) Result for HS ( ): With


Z

   = ( − 1) (5.176)
2
we obtain
µ ¶12 µ ¶32
HS 1 2
 ( ) = 2
  
µ ¶µ ¶¯∞
  ¯
2
× (  ) × exp − − − 1 ¯¯ (5.177)
    0
y
µ ¶12 µ ¶ µ ¶
HS 2 8  0 0
 ( ) =  1+ exp − (5.178)
    
5.4 Advanced collision theory 128

(4) Arrhenius activation energy:

 ln 
 =  2 (5.179)

µ ¶
1 0 0
ln  ∝ ln  + ln 1 + − (5.180)
2  
µ ¶ µ ¶
 ln  11 1 0 0
= + − + (5.181)
 2 1 + 0   2  2
y
1 0
 = 0 +  − (5.182)
2 1 + 0 

b) Line-of-Centers model (Figs. 5.21 - 5.22)


Even for hard spheres, because of angular momentum conservation (see below), not the
entire collision energy, but only the fraction “directed” along the line of centers ()
is effective.
In particular, we have

• for a collision with  = 0:


LC =  (5.183)
y central collisions are most effective,
• for a collision with  = :
LC = 0 (5.184)
y glancing collisions are ineffective because LC = 0,
• general expression (see below):
µ ¶
2
LC =  1 − 2 (5.185)

y LC decreases with increasing .


5.4 Advanced collision theory 129

I Figure 5.21: Collision energy along line-of-centers

Functional expression for  ( ):

(1) We assume that a reaction can occur only if LC ≥ 0 . Thus, we first determine
LC (see Fig. 5.21):

LC =  cos  (5.186)


2LC = 2 cos2  (5.187)
¡ ¢
= 2 1 − sin2  (5.188)
µ ¶
2 2
=  1− 2 (5.189)

y µ ¶
2
LC =  1 − 2 (5.190)

(2) We see from this expression that LC decreases with increasing . Thus, for a
given  , we have a maximal value of  (i.e. a value max ), for which LC ≥ 0 .
We determine max from the condition that
µ ¶
2
LC =  1 − 2 ≥ 0 (5.191)

y µ ¶
2 2 0
 ≤ 1− = 2max (5.192)

y µ ¶
LC 2 2 0
 ( ) = max =  1 − (5.193)

5.4 Advanced collision theory 130

y µ⎧ ¶
⎨ 0 2
 1 −   0
 LC ( ) =  (5.194)

0   0

 LC ( ) is shown as function of  in Fig. 5.22. Crossed molecular beam experi-


ments have indeed shown similar dependencies for a number of reactions.

I Figure 5.22: Reactive cross section for the line-of-centers model.

(3) Result for  ( ) from Eq. 5.169:


µ ¶12 µ ¶32 Z∞ µ ¶ µ ¶
LC 1 2 2 0 
 ( ) = ×  1 −  exp − 
     
0
(5.195)
µ ¶12 µ ¶32 Z∞ µ ¶
1 2 2 
= ×  ( − 0 ) exp − 
    
0
(5.196)

y
µ ¶12 µ ¶
LC 8  2 0
 ( ) =  exp − (5.197)
  
5.4 Advanced collision theory 131

This result40 is identical to that of our primitive “hard sphere-fixed mean relative
speed” model from Section 5.1!
(4) Arrhenius activation energy:

 ln 
 =  2 (5.200)

y
1
 = 0 +  (5.201)
2

c) Allowance for “steric effects”

I Steric factor: In order to account for the “structures” of the molecules which can
undergo a reaction only if the reaction partners have a specific relative orientation, we
introduce a (generally  -dependent) steric factor  ( ):
µ ¶12 µ ¶
8  2 0
( ) = ( )  exp − (5.202)
  

In ( ), we include all sorts of factors leading to deviations from the predicitions by
simple collision theory.
A certain form for ( ) can, for example, be rationalized by considering reaction prob-
abilities (so-called “opacity functions”) as shown in Fig. 5.23:

40
We used the integrals
Z
1 
  =  (5.198)

Z

   = 2 ( − 1) (5.199)

5.4 Advanced collision theory 132

I Figure 5.23: Allowance for steric effects in collision theory.

I Reaction probability (opacity function; Fig. 5.23): In favorable cases,  ( ) can


be predicted by theory. Towards these ends, we define a reaction probability  (  )
(also called the opacity function, and calculated by theory) and integrate according to
Z0max

 ( ) =  (  ) 2  (5.203)


0

I Angle-dependent reaction probability (Fig. 5.23): For some reactions, it may be


convenient to use an angle-dependent reaction threshold energy ∗0 :

∗0 = 0 + 0 (1 − cos ) (5.204)

y
∗0 = 0 for  = 0
(5.205)
∗0  0 for  6= 0

Collision parameter-dependent and angle-dependent reaction probabilities are helpful for


explaining steric factors in the range of 1 ≤  ≤ 001.
5.4 Advanced collision theory 133

I Table 5.2: Comparison of predictions by collision theory with experimental data.

I Harpooning: The value of ≥ 1 for the steric factor of the K + Br2 reaction can
be understood by the “harpooning mechanism”: An electron jumps from the K to the
Br2 at a long intermolecular distance. The subsequent collision and reaction at closer
distance then occurs between a K+ and Br−2 . The collision between these two is strongly
affected by the Coulomb attraction.

5.4.6 Long-range intermolecular interactions

To this point, we have neglected any interactions between the reacting molecules other
than the the collision itself. Thus, only straight line trajectories were allowed. In reality,
however, there will be attractive or repulsive interactions between the colliding molecules
(see Fig. 5.24). These will inevitably affect the collision dynamics. An extreme example
is the K + Br2 reaction mentioned above.
We will investigate the potential energy hypersurfaces governing the reactions in Section
6. Here, we will consider only long-range intermolecular interactions which can be easily
implemented in collision theory.
5.4 Advanced collision theory 134

I Figure 5.24: Examples of attractive and repulsive trajectories for the collision between
two molecules.

a) Types of long-range interactions


The long-range interactions between molecules can be described using multipole expan-
sions of the potentials (charge: 1; dipole: 2; induced dipole: 3; quadrupole: 3 . . . In
general, the interaction potential between two multipoles of ranks  and 0 is proportional
to − with  =  + 0 − 1:
µ ¶

 () = − (5.206)

I Table 5.3: Types of long-range intermolecular interactions.

system  ()
1 2 2
charge-charge (Coulomb)  () =
40 

−  cos 
charge-dipole  ( ) = − → · =−
40 2
3 (μ1 · r) (μ2 · r)
μ1 · μ2 −
dipole-dipole  ( 1  2 ) = 2
40 3
charge-quadrupole Θ ∝ −3
dipole-quadrupole Θ ∝ −4
charge-induced dipole  ∝ −4
quadrupole-quadrupole ΘΘ ∝ −5
induced dipole-induced dipole   ∝ −6
5.4 Advanced collision theory 135

3
I Table 5.4: Long-range intermolecular interactions ( = 15 D,  = 3 Å ,  = 1,
 = 0; energies in kJ mol; for comparison:  = 25 kJ mol @  = 298 K).
³ ´ ³ ´
system   = 5 Å   = 10 Å
ion-ion 278 139
ion-dipole 17.4 4.35
ion-induced dipole 3.34 0.21
dipole-dipole 2.17 0.27

b) Lennard-Jones (6-12) potential for nonpolar molecules


A realistic intermolecular potential for non-polar molecules is the Lennard-Jones (6-12)
potential
∙³ ¸
LJ ´12 ³  LJ ´6
LJ () = 4LJ − (5.207)
 

LJ is usually given in reduced form in units of K as


LJ
∗LJ = (5.208)


I Figure 5.25: Schematic plot of the Lennard-Jones potential.


5.4 Advanced collision theory 136

I Estimation of LJ parameters: Values for  LJ and LJ are typically derived from
transport properties (e.g. viscosity), thermodynamic data (vdW parameters), or — if
there is no better way — from empirical correlations (see, e.g. Svehla1963) with critical
data, boiling temperatures, . . .

I Table 5.5: Lennard-Jones for some gases (Mourits1977).

 LJ  ( pm) (LJ  )  ( K)  LJ  ( pm) (LJ  )  ( K)


He 2560 102 N2 3738 820
Ne 2822 320 O2 3480 1026
Ar 3465 1135 CH4 3790 1421
Kr 3662 1780 CF4 4486 1673
Xe 4050 2302 CCl4 5611 4155
H2 2930 370 CO2 3943 2009
SF6 5199 2120

I Figure 5.26: Lennard-Jones interaction potentials for He, Ne, Ar, Kr, and Xe.
5.4 Advanced collision theory 137

I Modified gas kinetic collision frequency with LJ potential: To account for at-
tractive/repulsive intermolecular interactions during molecular collisions, we have to
modify the standard “hard sphere” expression for the gas kinetic collision frequency:
p
LJ =  2LJ × 8  × Ω(22)∗ (5.209)

• The reduced collision integral Ω(22)∗ averages over the the detailed trajectories in
the presence of attractive or repulsive LJ interactions between “real molecules”
in comparision to hard sphere collisions. Its value depends on the reduced tem-
perature
 
∗ = (5.210)
LJ

• Values of Ω(22)∗ ( ∗ ) are tabulated (see, e.g., Hirschfelder1967, Reid1987). For


LJ interactions, in the range 09   ∗  5, Ω(22)∗ can be written as (Kim1967)
µ ¶
(22)∗ ∗ 109
Ω ( ) = 07616 1 + ∗ (5.211)

with an error of  01%.


• For binary systems (A-B pairs), one usually takes the arithmetic resp. geometric
mean values:
1
 LJ (AB) = [ LJ (AA) +  LJ (BB)] (5.212)
2
LJ (AB) = [LJ (AA) LJ (BB)]12 (5.213)

• For polar molecules, people have developed various modified potential forms (e.g.,
the Stockmayer potential given above) and also evaluated corresponding reduced
collision integrals (see Hirschfelder1967).

c) Orbital angular momentum and centrifugal barrier*


In the case of a non-central collision, part of the translational energy is converted
to “rotational energy” (orbital angular momentum; see Fig. 5.27). This is seen by
partitioning the velocity  vector between the  direction ( ) and the radial
direction ( ). The component  corresponds to a rotational motion.
5.4 Advanced collision theory 138

I Figure 5.27: Orbital angular momentum in binary collisions.

• Radial velocity component:


 
cos  = = (5.214)
 
y

 =  (5.215)

• Kinetic energy:
1 2 1 2
 =  +  (5.216)
|2 {z } |2 {z }
translation via LC 1
= 2 2 (rotation)
2

• Angular velocity:
     
= = = = = 2 (5.217)
      
• Kinetic energy:
1 1
 =  ( )2 +  ( )2 (5.218)
2 2
1 · 2 1 2 2
=  +  2 (5.219)
2 2 
1 · 2 1 2
=  +  2 (5.220)
2 2 
1 · 2 1 2 2 2
=  + (5.221)
2 2 2
5.4 Advanced collision theory 139

• Orbital angular momentum:  is the orbital angular momentum 



 = 2 =  =  (5.222)

y
1 ·2 2
 =  + (5.223)
2 22

From quantum mechanics, 2 = ~2  ( + 1) with  as orbital angular momentum


quantum number. y

1 · 2 ~2  ( + 1)
 =  + (5.224)
2 22
1 ·2
=  + 2  ( + 1) (5.225)
2
~2
where 2 = is formally an -dependent rotational constant.
22
~2  ( + 1)
• Centrifugal barrier: The centrifugal kinetic energy term acts as a
22
centrifugal barrier
1 2 ~2  ( + 1)
centrifugal =  2 = (5.226)
2  22
¡ ¢
• With a long-range attractive term of the form −  , the intermolecular po-
tential thus has the form
µ ¶
 2
eff () = − +  2 (5.227)
 
µ ¶
 ~2  ( + 1)
= − + (5.228)
 22

i.e.
eff () ∝ −− + −2 (5.229)

— The −2 term is important in practice if   3. In effect, it leads to an


effective energy barrier (“centrifigual” barrier).
— The larger  for a given  , the larger the orbital angular momentum .
5.4 Advanced collision theory 140

I Figure 5.28: Potential energy curves with centrifugal barriers for a diatomic system.

d) Langevin “capture” rate constant for ion-molecule reactions*


An nice illustration of the above results is the rate constant for ion-molecule reactions.
The intermolecular interaction is governed by the point charge-induced dipole interac-
tion.41 In addition, the centrifugal barrier has to be taken into account, so that
µ ¶2
 ()2 
 () = − 2 +  (5.230)
(40 ) 24 

The resulting expression for the rate constant is


à !12 µ ¶
2
L 4 () 1
 ( ) = ×Γ (5.231)
(40 )2  2

Typical values of L are 100 × HS 

41
 () needs to be checked again!!
5.4 Advanced collision theory 141

6. Potential energy surfaces for chemical reactions

6.1 Diatomic molecules

I Morse potential (anharmonic oscillator):

 () =  [1 − exp (−  ( −  ))]2 (6.1)

I Vibrational states of a harmonic oscillator:


µ ¶ µ ¶
1 1
 = ~   + =    + (6.2)
2 2

I Vibrational states of an anharmonic oscillator:


µ ¶ µ ¶2
1 1
 = ~  + − ~    + +  (6.3)
2 2

I Dissociation energies D and D0 :


1 1
 = 0 + ~  − ~   +    (6.4)
2 4

I Figure 6.1: Potential energy curve of diatomic molecules.


6.2 Polyatomic molecules 142

6.2 Polyatomic molecules

The potential energy becomes dependent of the different internal coordinates.


I Question: How many internal coordinates (dimensions) do we need?
I Answer:

• Total number of degrees of freedom of an -atomic molecule:

3 (6.5)

• for translational motion of center of mass:

3 (6.6)

• for rotational motion around center of mass

2 for linear molecules (6.7)


3 for nonlinear molecules (6.8)

• resulting number of vibrational coordinates:

— linear molecules:
3 − 5 (6.9)
— nonlinear molecules:
3 − 6 (6.10)

I 3N − 6 dimensional potential energy hypersurface:

• The internal coordinates are described using the normal coordinates  (vibra-
tional displacements).
• The potential energy function

 (1  2      3 −6 ) (6.11)

is a hypersurface in a 3 − 6 (3 − 5) dimensional hyperspace. But we can plot


 only in 2-D or 3-D space.
6.2 Polyatomic molecules 143

I Figure 6.2: Potential energy hypersurface for a simple atom transfer reaction (plot for
fixed angle, e.g. collinear collision and other fixed coordinates depending on dimension-
ality of the system).

I Figure 6.3: Schematic potential energy hypersurface of a triatomic molecule.


6.2 Polyatomic molecules 144

I Figure 6.4: Potential energy hypersurface of HCO.

I Figure 6.5: Potential energy hypersurface of HCO.


6.2 Polyatomic molecules 145

I Models of potential energy surfaces:

(1) Empirical (e.g. LEPS) or semiempirical models,


(2) Single points  ( ) are nowadays easily obtained by ab initio quantum chemistry
calculations.
(3) Problems and challenges:

a) High dimensionality for polyatomic molecules,


b) Fitting a global hypersurface to the ab initio points.

I Kinetics and Dynamics:

(1) TST = Transition State Theory (statistical theory; ⇒ Chapter 7),


(2) Classical trajectory calculations (by solving Newton’s equations; ⇒ Section 6.3),
(3) Fully quantum mechanical wave packet calculations (by solving the time-
dependent Schrödinger equation; ⇒ Theoretical Chemistry).
6.3 Trajectory Calculations 146

6.3 Trajectory Calculations

The “exact” dynamics of the molecules on a potential energy hypersurface can be


followed by classical trajectory calculations.

I Phase Space: 6-dimensional space spanned by the spatial coordinates ( ) and
the conjugated momenta ( ):
 ↔  (6.12)
where
 ·
 =  =   (6.13)


I Hamilton function:
2
 = + = +  () (6.14)
2
Note that  depends only on  and  depends only on .

I Equations of motion:

• for a 1-particle system:

(1)
  ()
 =  = =− (6.15)
 
y
·  
 =− =− (6.16)
 
since  depends only on  , but not .
(2)
2
= + (6.17)
2
y µ ¶
 1 1 · ·
= × = ×  =  (6.18)
  
y
 ·
= (6.19)

since  depends only on , but not on  .

• for an  atomic system:

X
3

 = + = 2 +  (1  2      3−6 ) (6.20)
=1
2

y
· 
 = − (6.21)

6.3 Trajectory Calculations 147

and
· 
 = (6.22)


Thus, we obtain a set of coupled differential equations which can be solved using
the Runge-Kutta or Gear methods.
• Thermal rate constants  ( ) follow by suitable averaging over the initial condi-
tions ( vibrational states, rotational states, collision angles, . . . ).
• Major problems: Zero-point energy? Tunneling? Surface crossings?

I Figure 6.6: Trajectory calculations.


6.3 Trajectory Calculations 148

I Figure 6.7: Trajectory calculations.

I Figure 6.8: The reaction H + Cl2 is a reaction with an early TS. The reaction is known
to yield HCl with an inverted vibrational state distribution (⇒ HCl-Laser).
6.3 Trajectory Calculations 149

I Figure 6.9: Bimolecular reaction with a late TS.


6.3 Trajectory Calculations 150

7. Transition state theory

7.1 Foundations of transition state theory

The concept of transition state theory (TST) or activated complex theory (ACT) goes
back to H. Eyring and M. Polanyi (1935). TST is a statistical theory; it does not give
information on the exact dynamics of a reaction. It does, however, give (statistical)
information on product state distributions.
There are two forms of TST:

• Macrocanonical TST ⇒  ( ) for a macrocanonical ensemble of molecules with


a constant  (an ensemble of molecules with energy states populated according
to the Boltzmann distribution - each state being equally likely.)
• Microcanonical TST (TST) ⇒  () for a microcanonical ensemble of molecules
with specific energy .

We will consider only the first version.


The detailed implementations of TST also depend on whether or not the reaction of
interest has a distinctive potential energy barrier (⇒ conventional TST or variational
TST).

I Figure 7.1: Transition State Theory (TST): Fundamental equations and applications.
7.1 Foundations of transition state theory 151

7.1.1 Basic assumptions of TST

(1) The Born-Oppenheimer approximation (separation of electronic and nuclear mo-


tion) is a valid approximation. Thus we can describe the nuclear motion during
the reaction on a single electronic potential energy surface (PES; see Fig. 7.2).42
(2) The reactant molecules are in internal equilibrium (the quantum states are popu-
lated according to the Boltzmann distribution).
(3) “Transition state” and “dividing surface“:

a) Conventional TST (valid for so-called “type I” PES’s = PES’s with a dis-
tinctive energy barrier along the reaction coordinate (RC)): We can localize
a distinctive “transition state” (TS) for the reaction at the saddle point on
the PES that separates the reactant and product “valleys”. The TS is lo-
cated at the maximum of the PES along the minimum energy path (MEP;
see Fig. 7.3).
b) Variational TST (version that can be used for so-called “type II” PES’s =
PES’s without a clear energy barrier along the RC (we can’t localize a simple
TS)): We can localize a 3 − 7-dimensional “dividing surface” (DS) on the
3 − 6-dimensional PES that separates the reactant and product regions
on the PES in such a way that the “reactive flux” through the DS becomes
minimal (Fig. 7.2). The TS is then placed at the point where the MEP
intersects the DS. In this case, the position of the TS depends strongly on
the finer details such as angular momentum (because of centrifugal barriers
which move inwards with increasing  or increasing ).

(4) The motion along the minimum energy path across the TS (through the DS) can
be separated from other degrees of freedom and treated as translational motion.
At the TS, we therefore have 1 translational degree of freedom  ‡ (the RC) and
3 − 7 (vibrational) degrees of freedom   orthogonal to the RC.
(5) Reactant molecules passing across the TS/DS never come back (no recrossing).
(6) The states of the TS are in quasi-equilibrium with the states of the reactant
molecule, even if we have no equilibrium between reactants and products (the
individual molecules don’t know that this equilibrium is not present - this quasi-
equilibrium assumption can be dropped in dynamical derivations of TST; see
below).

I Conditions for TS at the saddle point:

(1)  has maximum along ‡ :


 2
= 0 and 0 (7.1)
‡ ‡2
y 1 imaginary frequency (criterion for TS in quantum chemical calculations)!

42
TST cannot be used at least in the simple form considered in this lecture when two potential surfaces
are involved that are energetically close to each other (or even become degenerate) and the reaction
involves a “non-adiabatic transition” between these two potentials.
7.1 Foundations of transition state theory 152

(2)  has minimum along all 3 − 7 other coordinates:


 2
= 0 and 0 (7.2)
   2
y 3 − 7 real other frequencies.

I Figure 7.2: Transition state theory: The dividing surface with minimal reactive flux.

I Figure 7.3: Minimum energy path (MEP) and transition state (TS) of a simple atom
transfer reaction.
7.1 Foundations of transition state theory 153

7.1.2 Formal kinetic model

In order to derive the TST expression for  ( ), we first look at a simple formal kinetic
model. This formal kinetic model has its roots in the original idea that the TS can be
considered as some sort of an “activated complex” with a local energy minimum on
the PES. It is not a good derivation of the TST expression for  ( ), but leads to the
correct expression.

I Activated complex model of TST:


1 3
A + BC À ABC‡ → P (7.3)
2

y
£ ¤ 1 [A] [BC]
ABC‡ ss = (7.4)
2 + 3

I Assumption: 3 ¿ 2 y
£ ¤ 1
ABC‡ ss = [A] [BC] = 12 [A] [BC] (7.5)
2

y Equilibrium between the reactants (A + BC) and the TS (ABC‡ ):

 [P] £ ¤
= 3 ABC‡ ss (7.6)

= 3 12 [A] [BC] (7.7)
| {z }
=

I ACT expression for rate constant:

 = 3 12 (7.8)

Below, we will find that


 
3 =  ‡ = (7.9)

and µ ¶
∆‡
12 ∝ exp − (7.10)

7.1 Foundations of transition state theory 154

7.1.3 The fundamental equation for k (T )

The above derivation of the TST rate constant was based on an oversimplified picture,
since the TS is not at a local minimum, but at a saddle point on the PES. In this section,
we derive the TST fundamental equation more carefully by bringing in some dynamical
aspects. In (1) - (4), we explore the implications of the quasi-equilibrium assumption,
and in (5) - (9) we develop a dynamical semiclassical model:43,44

(1) We start by assuming equilibrium between the reactants and products:


1 3
A + BC À ABC‡ À P (7.11)
2 4

(2) We consider two dividing surfaces at the TS separated by a distance of ≤ .


 is the “length” of a “phase space cell” in semiclassical theory which is determined
by Heisenberg’s uncertainty principle:
∆ ‡ × ∆‡ =  × ∆‡ =  (7.12)

(3) We now consider the number densities of molecules passing through these surfaces

→ ←−
from left to right ( ‡ ) and from right to left ( ‡ ). The total number density  ‡
at the TS in equilibrium is their sum, i.e.

→ ← −
 ‡ =  ‡ +  ‡ = ‡ A BC (7.13)
In equilibrium, we must have

→ ← −
‡ = ‡ (7.14)
Thus,
−→‡ ← −‡ 1 ‡ ‡
 = =  = A BC (7.15)
2 2
←−
(4) We now suddenly remove all the products. Then  ‡ = 0 However, there is no
−→
reason for  ‡ to change, because the molecules on the reactant side remain in
internal equilibrium (this is the quasi-equilibrium assumption).45 y

→ ‡ ‡
‡ = =  A BC (7.16)
2 2


Thus, we can still express  ‡ using the equilibrium constant ‡ , even if there is
no equilibrium between reactants and products.

43
Before continuing with the derivation of the fundamental equation of TST, the novice should review
some important results from statistical thermodynamics (see Appendix G) which we will use.
44
The model is called semiclassical, because the translation is handled classically, while all other degrees
of freedom are described quantum mechanically.
45
−→
The quasi-equilibrium assumption for  ‡ is made in canonical TST. In a microcanonical derivation
· ·
−→‡ ←−‡
of the TST expression for  (), we consider the fluxes  and  through the dividing surface
at the TS. Then, the individual molecules “dont’t know” that the products have been removed and
· ·
←−‡ −→‡
 = 0. And since they don’t know about this, the flux  has to remain unchanged.
7.1 Foundations of transition state theory 155

(5) We now turn to the reaction rate, which is given by the number density of mole-


cules  ‡ in the time interval  that are crossing the DS (i.e. the TS) in the
forward direction. We can write this as
−→ →

 ‡ − →‡ ‡ ‡ ‡
=  =  (7.17)
 
where


•  ‡ is the decay rate of the TS in the forward direction (or the frequency
factor for the molecules crossing the DS in the forward direction),

− →

• ‡ =   ‡ is the mean speed of the molecules crossing the DS in the forward
direction.

(6) We now use the equilibrium constant to determine  ‡ via



ABC
‡ = (7.18)
A BC

y  ‡ = ABC = ‡ A BC (7.19)

y −→ − →‡
 ‡ 
= ‡ A BC =  A BC (7.20)
 
with −‡


 = ‡ (7.21)

(7) From statistical thermodynamics (see Appendix G), we have the following ex-
pression for the equilibrium constant ‡ ( ) in terms of the molecular partition
functions: µ ¶
‡ ∗ABC ABC ∆0
 ( ) = = exp − (7.22)
A BC A BC  

• The ’s are the respective molecular partition functions (per unit volume).
• ∗ABC is written with respect to the zero-point level of the reactants.
• The exp (−∆0   ) factor arises subsequently, because we now express
the value of partition function for the TS relative to the zero-point level of
the TS (∗ABC = ABC exp (−∆0   )).

(8) Separating the motion along the reaction coordinate ‡ from the other degrees of
freedom, we write
ABC = ‡ ‡ABC (7.23)
where ‡ is the 1-D (translational) partition function describing the translational
motion of the molecules along  ‡ across the TS and ‡ABC is the partition function
for all 3 − 7 other (rotational-vibrational-electronic) degrees of freedom.
Here,
7.1 Foundations of transition state theory 156

• the 1-D translational partition function is


µ ¶12
2 
‡ = × (7.24)
2

• the mean speed along  ‡ for crossing the DS in the forward direction is given
by R ∞ −2 2  µ ¶12
→‡
− 0
    
 = R ∞ −2 2  = (7.25)
−∞
   2

(9) Collecting and inserting the results into


−‡


 ( ) = ‡ ( ) (7.26)

we obtain
µ ¶12 µ ¶12 µ ¶
1   2  ‡ABC ∆0
 ( ) = × × exp − (7.27)
 2 2 A BC  
y µ ¶
  ‡ABC ∆0
 ( ) = exp − (7.28)
 A BC  

I Result: In order to account () for some recrossing of the TS and () for quantum
mechanical tunneling, we add an additional factor  called the transmission coefficient.
For well-behaved reactions,  ≈ 1. Thus, we end up with the fundamental equation
of TST in the form:
µ ¶
  ‡ABC ∆0
 ( ) =  exp − (7.29)
 A BC  

Note that ‡ABC is the partition function of the TS excluding the RC.
Equation 7.29 allows us to calculate  ( ) using the following data:

(1) The threshold energy for the reaction ∆0 (or, per mol, ∆0 ) and
(2) structural information on the reactants and the TS (i.e. bond lengths and angles,
vibrational frequencies).

∆0 can be obtained from ab initio quantum chemistry calculations, from thermody-
namic data (e.g. for bond dissociation reactions), for  = 0 K of course, or from the
experimentally measured activation energy, corrected for the difference between  and
∆0 using the TST expression.
7.1 Foundations of transition state theory 157

I Figure 7.4: Illustration of ∆ ‡ , ∆0 , and  .

∆ ‡ : Born-Oppenheimer potential energy barrier (7.30)


∆0 : zero-point corrected threshold energy for the reaction (7.31)
 : thermal Arrhenius activation energy (7.32)
∆0 is given with respect to the zero-point energies of the reactants and the TS,
1X
 =   (7.33)
2

7.1.4 Tolman’s interpretation of the Arrhenius activation energy

I We now understand the difference between E and ∆E0 :


 ln  ( )
 =  2 (7.34)
 Ã !
µ ¶
2    ‡ABC ∆0
=  ln  exp − (7.35)
  A BC 
‡ µ ¶
2  ln ABC 2  ln A 2  ln BC
= ∆0 +  +  −  +  (7.36)
| {z  } |  
{z }
=hreacting molecules i =hall molecules i

• ∆0 is difference of the zero-point levels of the reactants and the TS,
• the  term is the mean translational energy in the RC (from the    factor),
 ln 
• the  2 terms are the internal (vibration-rotation) energies of the TS and

the reactants, respectively. y
7.1 Foundations of transition state theory 158

I Tolman’s interpretation of E (see Fig. 7.4):

 = hreacting molecules i − hall molecules i (7.37)


7.2 Applications of transition state theory 159

7.2 Applications of transition state theory

µ ¶
  ‡ABC ∆0
 ( ) =  exp − (7.38)
 A BC  

TST has gained enormous importance in all areas of physical chemistry because it
provides the basis for understanding – and/or calculating –

• quantitative values of preexponential factors,


• ∆0 from measured  values,
•  -dependence of  and deviations from Arrhenius behavior,
• gas phase and liquid phase reactions (proton transfer, electron transfer, organic
reactions, reactions of ions in solutions),
• pressure dependence of reaction rate constants (activation volumes),
• kinetic isotope effects,
• structure-reactivity relations (e.g. Hammett relations in organic chemistry),
• features of biological reactions (enzyme reactions),
• heterogeneous reactions (heterogeneous catalysis, electrode kinetics), . . .

7.2.1 The molecular partition functions

In order to evaluate Eq. 7.29 for a given reaction, we need, besides ∆0 , the molecular
partition functions for the reactants and the TS. We summarize only the main points
here; further information is given in Appendix G.

I Definition:
P
=   −   (7.39)

I Physical interpretation:

• The molecular partition function is a number which describes how many states
are available to the molecule at a given temperature  .
• The ratio
‡ABC
(7.40)
A BC
in Eq. 7.29 is thus the ratio of the number of states available to the TS and
‡ABC
the reactant molecules. Since each state is equally likely, is the relative
A BC
probability of the TS vs. the probability of the reactants.

Note again that ‡ABC is the partition function of the TS excluding the translational
motion along the RC.
7.2 Applications of transition state theory 160

I Separation of degrees of freedom: For separable degrees of freedom, we can write


the molecular partition function  as

 = trans × rot × vib × el (7.41)

Nuclear spin (ns ) may have to be taken into account in special cases as well.

I Table 7.1: Molecular partition functions.

type of motion energy partition function magnitude


µ ¶12
2 2 2 
translation (1-D)  108 × 
8 2 2
µ ¶ µ ¶32
2 2 2 2 2 
translation (3-D) + +  1024 × 
8 2 2 2 2
µ ¶12
~2 2  12 2  12
rotation (1-D)ab)  2
×  10
2 µ ~¶
~2 1 2 
rotation (2-D)c)  ( + 1) × 102
2  ~2
µ ¶32
 12 2 
rotation (3-D)d)  or   2
× (   )12 103    104
 ~
vibratione)    [1
P − exp (−    )]−1 1    10
electronic    exp (−   ) 1

a
For a free internal rotor with reduced moment of inertia −1 = 1−1 + 2−1 .
b
For a hindered rotor (torsional vibrational mode), the partition function has to be determined by an
explicit calculation over the hindered rotor quantum states.
c
Linear molecule.
d
Nonlinear molecule.
e
For each harmonic oscillator degree of freedom measured from zero-point level.
7.2 Applications of transition state theory 161

7.2.2 Example 1: Reactions of two atoms A + B → A· · · B‡ → AB

This example is somewhat artifical, unless there is some mechanism to remove the bond
energy (for example, in associative ionization, A + B → A· · · B‡ → AB+ + − ).
µ ¶
  ‡AB ∆0
 ( ) = exp − (7.42)
 A B 
à ! à ! µ ¶
‡ ‡
  AB  AB ∆0
= exp − (7.43)
 (A  ) (B  ) 1 
trans rot
µ ¶32 µ 2 ‡ ¶
  2 (A + B )   2 2 8   
=
 2 2A   2B   2
µ ¶
∆0
× exp − (7.44)

µ ¶32 Ã 2 !
  2 A + B 8   A B (A + B )2
=
 2  A B 2 A + B
µ ¶
∆0
× exp − (7.45)

y µ ¶12 µ ¶
8  2 ∆0
 ( ) =  (A + B ) exp − (7.46)
 
This result is identical to that obtained by the LC model in collision theory!
7.2 Applications of transition state theory 162

7.2.3 Example 2: The reaction F + H2 → F· · · H· · · H‡ → FH + H

I Exercise 7.1: Evaluate the rate constant for the reaction

F + H2 → F · · · H · · · H‡ → FH + H (7.47)

using the data in Table 7.2 at  = 200, 300, 500, 1000 and 2000 K, determine the
Arrhenius parameters, and compare the results with the experimental value of

 = 2 × 1014 exp (−800 ) cm3  mol s (7.48)

I Table 7.2: Properties of the reactants and transition state of the reaction F + H2 .

parameter F · · · H · · · H‡ a F H2
2 (F · · · H) ( Å) 1602
b
1 (H · · · H) ( Å) 0756 07417
 1  ( cm−1 ) 40076 43952
 2  ( cm−1 ) 3979
 3  ( cm−1 ) 3979
 4  ( cm−1 ) 3108  c
³( u) ´ 21014 189984 2016
2
 u Å 7433 0277
 1 2
 d ¡ ¢ 4 4 1
0  kJ mol−1 657 000

a
The transition state FHH‡ is assumed to be linear.
b
= (H-H).
c
One imaginary frequency describes the TS along the RC.
d
The electronic ground state of F is 2 32 . We neglect the 2 12 spin-orbit component at (2 12 ) =
404 cm−1 because it does not correlate with the products H + HF. The electronic state of linear
FHH‡ is 2 Π. The ground state of H2 is 1 Σ+ .

I Solution 7.1:
µ ¶
  ‡FHH ∆0
 ( ) = exp − (7.49)
 F H2 
à ! à ! à ! à !
  ‡FHH  ‡FHH ‡FHH ‡FHH
=
 (F  ) (H2  ) F H2 F H2 F H2
trans rot vib el
µ ¶
∆0
× exp − (7.50)

7.2 Applications of transition state theory 163

The different factors can be evaluated separately:


à ! µ ¶32 µ ¶32
‡FHH  F + H2 2
= (7.51)
(F  ) (H2  ) F H2 2 
trans
à ! µ ¶
‡FHH 8 2    ‡ 2 ‡
= =  H2 (7.52)
F H2 8 2   H2  H2 2 H2
rot
à !
‡FHH 1 − exp (− H2   )
= 3 (7.53)
F H2 Yh ³ ´i
vib
1 − exp − ‡  
=1
à !
‡FHH 4
= (7.54)
F H2 4×1
el

Note: The difference between the TST result and experiment can be ascribed (a)
to tunneling and (b) deficiencies of the ab initio PES. Indeed, according to newer ab
initio calculations, the TS is actually slightly bent and the TS parameters differ slightly,
yielding better agreement with experiment. ¥
7.2 Applications of transition state theory 164

7.2.4 Example 3: Deviations from Arrhenius behavior (T dependence of k (T ))*

I Exercise 7.2: The temperature dependence of bimolecular gas phase reactions of the
type A + B → products can usually be expressed in the form
 ( ) =  ×   × exp (−∆0  ) (7.55)
Using the TST expression for  ( )
µ ¶
  ‡AB ∆0
 ( ) = exp − , (7.56)
 A B 
determine the values of  for the types of reactions in Table 7.3, assuming that the
vibrational degrees of freedom are not excited. ¤

I Table 7.3:
 
A +B → TS‡ transl. rot. 

 32 1
atom atom linear TS 1 +05
 32  32 1
1
atom linear molecule linear TS  1  −32 −05
 132

atom linear molecule nonlin. TS 1  −32 1
00
32

atom nonlin. molecule nonlin. TS 1  −32 −05
 32
1
linear molecule linear molecule linear TS 1  −32 −15
 132
1

linear molecule linear molecule nonlin. TS 1  −32 −10
 1 32
1

linear molecule nonlin. molecule nonlin. TS 1  −32 −15
 1 32
32

nonlin. molecule nonlin. molecule nonlin. TS 1  −32 −20
 32  32

I Solution 7.2: See Table 7.3. ¥


Using the above results, we can recast  ( ) and determine the following expression for
the Arrhenius activation energy, with  as above:
‡ X  ln reactants
2  ln vib 2 vib
 = ∆0 +  +  −  (7.57)
 reactants

We see that the value for  depends on the vibrational frequencies of the TS and the
reactants. Often, the frequencies of the reactants are very high so that reactant
vib = 1.

However, the TS may have soft vibrations so that vib may approach the classical value
 
vib = [1 − exp (−  )]−1 → for  → ∞ (7.58)

Thus, each “soft” vibrational mode in the TS increases  by 1, each “soft” vibrational
mode in the reactants decreases  by 1.
7.3 Thermodynamic interpretation of transition state theory 165

7.3 Thermodynamic interpretation of transition state theory

7.3.1 Fundamental equation of thermodynamic transition state theory

As shown above,
  ‡
 ( ) =  (7.59)

‡ can often be interpreted using thermodynamic arguments, i.e. ∆‡ , ∆ ‡ , and ∆ ‡ .
For liquid phase reactions, this is straightforward. For gas phase reactions, we have to
take into account the difference between ‡ and ‡ .

a) Thermodynamic transition state theory for unimolecular gas phase reactions


For unimolecular isomerization and dissociation reactions of the type
ABC → ABC‡ → ACB‡ (7.60)
or
ABC → ABC‡ → A + BC‡ , (7.61)
we have £ ¤

ABC  ‡
‡ ( ) = = ABC = ‡ ( ) (7.62)
[ABC] ABC
‡ ( ) = ‡ ( ) =  ‡ ( ) is dimensionless (∆ ‡ = 0) and we do not have to worry
about a difference between ‡ ( ) = ‡ ( ) and about units or different standard
states.

I Results: Eq. 7.59 thus gives the following results:


• Rate constant:
µ ¶
  ‡   ∆‡
 ( ) =  ( ) = exp − (7.63)
  
y µ ¶ µ ¶
  ∆ ‡ ∆ ‡
 ( ) = exp exp − (7.64)
  
• Arrhenius activation energy:
∙ µ ¶ µ ¶¸
2  ln   2   ∆ ‡ ∆ ‡
 =  =  ln exp exp − (7.65)
    
Thus, if ∆ ‡ , ∆ ‡ 6=  ( ) in the temperature range of interest, we obtain
 = ∆ ‡ +  (7.66)
• Arrhenius preexponential factor:
∆ ‡ =  −  (7.67)
y µ ¶ µ ¶
  ∆ ‡ 
 ( ) =  exp exp − (7.68)
  
y µ ¶
  ∆ ‡
= exp (7.69)
 
7.3 Thermodynamic interpretation of transition state theory 166

b) Thermodynamic transition state theory for bimolecular gas phase reactions

I Conversion from K‡ (T ) to K‡ (T ): For bimolecular reactions

A + BC → ABC‡ , (7.70)
we have to account for the different numbers of species ∆ ‡ =  ‡ −  reactants :
∆ ‡ = −1 . (7.71)
We therefore have to convert from ‡ ( ) to ‡ ( ) via

£ ¤ ABC‡ ABC‡ ª



ABC 1
‡ ( ) = = A
 = 
ª ª × ª (7.72)
[A] [BC] BC A  BC  
   
µ ¶−∆ ‡ µ ¶
‡  ‡ 
=  ( ) × =  ( ) × . (7.73)
ª ª

Here, ‡ ( ) is conveniently given in units of l mol−1 .46 However, ‡ , ∆‡ , ∆ ‡ and
∆ ‡ are related to the standard state at ª = 1 bar.47 Therefore, to keep track of units
l bar
in the conversion, it is useful to apply  in by writing
mol K
J m3 Pa l bar
 = 83145 = 83145 = 0083145 = 0 , (7.75)
mol K mol K mol K
so that
ABC‡ ª µ 0 ¶
‡  1 ‡ 
 ( ) = ª ª × ª =  ( ) × . (7.76)
A  BC   ª
 
Thus, via Eq. 7.59, we obtain
µ ¶
  0  ‡   0  ∆ª‡
 ( ) =  ( ) = exp − , (7.77)
 ª   ª 
in the desired units of
µ ¶ µ ¶
  0  l
dim ( ( )) = dim × dim = (7.78)
 ª mol s
and with the standard state ª = 1 bar for the thermodynamic quantities indicated by
the ª symbol.

¡ ¢∆
46
With ‡ ( ) in units of mol l−1 , we get as close as possible to the standard state ∗ =
−1
1 mol kg for reactions in aqueous solution without having to invoke a suitable equation of state.
47
We recall that the thermodynamic equilibrium constant  ( ) = exp (−∆ ª  ) referred to
the standard state at ª = 1 bar is dimensionless. Instead of the dimensionless ‡ ( ), one may
also decide to use the equilibrium constant in pressure units ‡ 0 :
h i
ABC‡
ABC‡
‡ ( ) = = A ‡0 0
BC =  ( ) ×   (7.74)
[A] [BC]
 
7.3 Thermodynamic interpretation of transition state theory 167

I Results:

• Rate constant:
µ ¶ µ ¶
  0  ∆ ª‡ ∆ ª‡
 ( ) = exp exp − (7.79)
 ª  

• Arrhenius activation energy:


 ln 
 =  2 (7.80)


If ∆ ª‡ 6=  ( ) in the temperature range of interest:

 = ∆ ª‡ + 2 (7.81)

• Arrhenius preexponential factor:

∆ ª‡ =  − 2 (7.82)

y
µ ¶ µ ¶
  0  ∆ ª‡  − 2
 ( ) = exp exp − (7.83)
 ª  
0
µ ª‡
¶ µ ¶
2     ∆ 
=  exp exp − (7.84)
 ª  

Comparison with the Arrhenius expression  =  exp (−  ) thus gives
µ ¶
  0 
2 ∆ ª‡
= exp (7.85)
 ª 

• Standard state: Note again that all thermodynamic quantities above apply to
the standard state ª = 1 bar.

I Note of caution on the choice of units and standard states:*  and  above
are given in units of l mol−1 s−1 (we may indicate this by writing them as  and
 ), but ∆ ª‡ and ∆ ª‡ used above are referred to the standard state for gases of
ª = 1 bar. However, we may want to relate ∆ ª‡ resp. ∆ ª‡ and ∆ ª‡ to the
respective quantities ∆ ⊕‡ and ∆ ⊕‡ at ⊕ = ⊕  ⊕ = 1 mol l:

• For ideal gases, ∆ ‡ and ∆ ‡ are related via

∆ ª‡ = ∆ ª‡ + ∆ ( ) = ∆ ª‡ + ∆ ‡  = ∆ ª‡ −  = ∆ ⊕‡ − 
(7.86)
ª ⊕ ‡
where we have used ∆ = ∆ for ∆ and ∆ = −1.
7.3 Thermodynamic interpretation of transition state theory 168

• Likewise, for a mole change of ∆ = ∆ ‡ , ∆ ª‡ and ∆ ⊕‡ are related via


Z ª Z ª Z ª Z ª
ª‡ ⊕‡    ∆   ‡
∆ − ∆ = = = = ∆   ln 
⊕  ⊕  ⊕  ⊕
(7.87)
ª ⊕ 0 ª ⊕ 0
      
= ∆ ‡  ln ⊕ = ∆ ‡  ln ⊕
= ∆ ‡  ln
  ª
(7.88)

y
⊕ 0 
∆ ª‡ = ∆ ⊕‡ −  ln . (7.89)
ª

c) Thermodynamic transition state theory for reactions in the liquid phase


For reactions in condensed phases, the standard state may be set to ⊕ = 1 mol l−1 .
Furthermore,
£ ¤ µ ¶ µ ¶
‡ ABC‡ ¡ ⊕ ¢∆ ‡ ∆ ⊕‡ ∆ ⊕‡
 ( ) = =  exp − exp − (7.90)
[A] [BC]  
µ ¶ µ ¶
¡ ⊕ ¢∆ ‡ ∆ ⊕‡ ∆ ⊕‡
≈  exp − exp − , (7.91)
 
since
∆ ≈ ∆  (7.92)
y

• for unimolecular reactions:


µ ¶ µ ¶
  ∆ ⊕‡ ∆ ⊕‡
 ( ) = exp − exp − (7.93)
  

• for bimolecular reactions:


µ ¶ µ ¶
  ∆ ⊕‡ ∆ ⊕‡
 ( ) = ⊕ exp − exp − . (7.94)
   

With ⊕ = 1 mol l−1 ,  ( ) again has units of


µ ¶
  l
dim ( ( )) = dim ⊕
= (7.95)
  mol s

Note: The conversion between values for ∆ ⊕‡ and ∆ ª‡ and, likewise, between
values for ∆ ⊕‡ and ∆ ª‡ for a given reaction in the liquid and gas phase requires
some caution.48

48
See D. M. Golden, J. Chem. Ed. 48, 235 (1971).
7.3 Thermodynamic interpretation of transition state theory 169

7.3.2 Applications of thermodynamic transition state theory

a) Comparison of ∆S ‡ -values for different types of transition states

I Bimolecular reactions:
A + BC → ABC‡ (7.96)
y
∆ ‡  0 (7.97)
or even
∆ ‡ ¿ 0 (7.98)
Furthermore, for chemically otherwise similar molecules

 (nonlinear molecule)   (linear molecule) À  (cyclic molecule) (7.99)

Thus,
¯ ‡ ¯ ¯ ¯ ¯ ¯
¯∆ (nonlinear TS)¯  ¯∆ ‡ (linear TS)¯ ¿ ¯∆ ‡ (cyclic TS)¯ (7.100)

For example:
¯ ⎛ ⎞¯ ¯ ¯ ⎛ ⎞¯
¯ ‡ ¯ ⎛ ‡ ⎞¯¯ ¯ B ¯¯

¯ A ··· B ¯ ¯¯ ¯ ¯ A ···
¯ ‡⎜ ... ⎟¯ ¯ ⎜ . .. ⎟¯
⎠¯  ¯¯∆ ⎝ A · · · B · · · C ⎠¯¯ ¿ ¯∆ ⎝ ..
‡ ‡
¯∆ ⎝ . ⎠¯¯
¯ ¯ ¯
¯ C ¯ ¯ ¯ ¯ C ··· D ¯
(7.101)
Thus, since all ∆ ‡ values are negative,

cyclic ¿ linear  nonlinear (7.102)

I Unimolecular reactions:
ABC → ABC‡ (7.103)

• For isomerization reactions with a tight transition state:


⎛ ‡⎞
B
∆ ‡ ⎝ Á Â ⎠.0 (7.104)
A – C

• For bond fission reactions with a loose transition state:


¡ ¢
∆ ‡ AB · · · C‡  0 (7.105)

y
loose  tight (7.106)
7.3 Thermodynamic interpretation of transition state theory 170

b) Thermochemical estimation of A-factors*


Equations 7.69 and 7.85 allow us to estimate the -factors of chemical reaction rate
constants by the following procedure (Benson1976):

(1) We setup a model for the TS, with a certain structure and vibrational freuqencies.
(2) We take a reference molecule that is similar to the TS and use the  ref of the
reference molecule for a zero-order estimation of ∆ ‡ according to
X
∆ ‡ =  ref − ( reactants ) (7.107)
reactants

(3) We apply corrections for the differences between  ref and  ‡ based on the statis-
tical thermodynamics result of
 ln 
 =  ln  +  (7.108)

y µ ‡¶
cor ‡ ref ‡  
∆ =  −  =  ln ref +  ln (7.109)
  ref
Usually, the correction term accounts for differences of the entropies for transla-
tion, rotation, internal rotations, vibrations, symmetry, electron spin, plus some-
times other terms (e.g. optical isomers).

I Example 7.1: Estimation of the -factor for the reaction O + CH4 →


(O · · · H · · · CH3 )‡ → OH + CH3 .

∆ ‡ =  (O · · · H · · · CH3 )‡ −  (O) −  (CH4 ) (7.110)


 ref =  (CH3 F) (7.111)
∆ cor thus depends on a comparison of CH3 F and OHCH‡3 (see Table ). ¤

I Table 7.4: Estimation of the -factor for the reaction O + CH4 →


(O · · · H · · · CH3 )‡ → OH + CH3 .

Contribution ∆ ‡ ( J mol−1 K−1 )


ª (CH3 F) − ª (O) − ª (CH4 ) −1243
Spin: + ln 3 + 91
1 (1 2 3 )‡
external rotation:  ln + 79
3 (1 2 3 )ref
translation − 04
 CF = 1000 cm−1 →  − 04
 OH = 2000 cm−1 + 00
−1
2 OHC ()
³ = 600 cm ´ + 42
∆ ‡ = ª O · · · H · · · CH‡3 − ª (O) − ª (CH4 ) −1039

Correction ∆‡ → ∆‡ : + ln (0  ) y ∆ ‡ = −198 J mol−1 K−1 .


7.3 Thermodynamic interpretation of transition state theory 171

Estimated -value:
µ ¶
2   ∆ ‡
 =  exp (7.112)
 
µ ¶
−197 J mol−1 K−1
= 739 × 625 × 10 × exp12
cm3 mol−1 s−1 (7.113)
831 J mol−1 K−1
= 43 × 1012 cm3 mol−1 s−1 (7.114)

Experimental -value:

expt = 19 × 1012 cm3 mol−1 s−1 (7.115)

The difference is probably due to a poor estimate of the TS structure — one can do
much better, I just didn’t try hard enough here.49

7.3.3 Kinetic isotope effects

I Figure 7.5: The kinetic isotope effect due to the difference in the zero-point energies
of the reactants and the respective TS structures.

49
For H abstraction reactions from a series of hydrocarbons by 3 CH2 , the agreement that was achieved
was better than ±30 %.
7.3 Thermodynamic interpretation of transition state theory 172

Kinetic isotope effects occur in particular in the case of H/D atom transfer reactions.
The reason is that the threshold energy ∆0‡ depends on the vibrational zero-point
energies.
We consider the reactions (with A being an atom)

A + H—R → A—H—R‡ → products (7.116)


A + D—R → A—D—R‡ → products (7.117)

where

∆0‡ (H—R) = ∆ ‡ + ‡ (A—H—R) −  (H—R) (7.118)


∆0‡ (D—R) = ∆ + ‡
‡ (A—D—R) −  (D—R) (7.119)

with
1X
 =    (7.120)
2 
y
³ ´

∆ ∆0(DH) = ∆0‡ (D—R) − ∆0‡ (H—R) (7.121)
= +‡ (A—D—R) − ‡ (A—H—R) − ( (D—R) −  (H—R))
(7.122)

y ³⎛ ´⎞

 (D—R) ∆ ∆0(D/H)
= exp ⎝− ⎠ (7.123)
 (H—R) 

I Example 7.2: R-H/R-D substitution. The vibrational frequencies depend on the force
constants  and reduced masses :
µ ¶12
1 
= (7.124)
2 

For H/D substitution:


 (HR) 1
≈ (7.125)
 (DR) 2
y µ ¶12
 (HR)  (HR) √
= ≈ 2 (7.126)
 (DR)  (DR)
With  (HR) = 3000 cm−1 and  (DR) = 2100 cm−1 , we find ∆ =
1
(3000 − 2100) cm−1 = 450 cm−1 , by which the DR reactant is stabilized compared
2
to the HR reactant. For  = 300 K:
µ ¶
 (DR) 450
= exp − ≈ −225 ≈ 01 (7.127)
 (HR) 200
¤
For more realistic estimates, one has to take into account the change of all vibrational
frequencies in the reacting molecules and the TS.
7.3 Thermodynamic interpretation of transition state theory 173

7.3.4 Gibbs free enthalpy correlations

I Figure 7.6: Linear free enthalpy correlations.

Application of the TST expression


µ ¶
  ∆‡
= exp − (7.128)
 
to a series of related reactions (1) and (2) often gives a relation of the type50
¡ ¢
∆‡1 − ∆‡2 =  ∆ª 1 − ∆2
ª
(7.129)
y
1 1
ln =  ln (7.130)
2 2

7.3.5 Pressure dependence of k

Liquid phase reactions often show a pressure dependence of the rate constant. Using
the relation µ ¶

= (7.131)
 
we can predict this pressure dependence. Starting with
µ ¶
  ∆‡
= exp − (7.132)
 
we find µ ¶
 ln  ∆ ‡
=− (7.133)
  

50
A well-known example is the Hammett correlation in organic chemistry.
7.3 Thermodynamic interpretation of transition state theory 174

I Activation volume:
∆ ‡ (7.134)

I Example 7.3: Comparison of the pressure dependencies of the rate constants for  1
and  2 reactions:

(CH3 )3 CCl + H2 O → (CH3 )3 C+ + Cl− + H2 O →    ∆ ‡  0

y ↑ y  ↓ )  1

⎡ ⎤−
|
HO− + CH3 I → ⎣ HO · · · C · · · I ⎦ →  ∆ ‡  0
ÁÂ

y ↑ y  ↑ )  2

)
Note the effect of the −-sign in Eq. 7.133. ¤
7.4 Transition state spectroscopy 175

7.4 Transition state spectroscopy

Until a few years ago, the “transition state” was a rather “elusive and strange” species.
This situation has completely changed (a) because of the rapidly improving quantum
chemical methods for computing TS properties and (b) because of transition state
spectroscopy, in the frequency domain (J. Polanyi, Kinsey), and in the time domain
(Zewail).

I Figure 7.7: Transition state spectroscopy: The reaction K + NaCl.

I Figure 7.8: Transition state spectroscopy: Raman emission of dissociating molecules.


7.4 Transition state spectroscopy 176

I Figure 7.9: Transition state spectroscopy: The reaction F + H2 .

I Figure 7.10: Transition state spectroscopy: NaI Photodissociation.


7.4 Transition state spectroscopy 177

I Figure 7.11: Transition state spectroscopy: NaI Photodissociation.

I Figure 7.12: Transition state spectroscopy: Reaction a in van-der-Waals Complex.


7.4 Transition state spectroscopy 178

I Figure 7.13: Transition state spectroscopy: Stimulated Emission Pumping.


7.4 Transition state spectroscopy 179

7.5 References

Benson 1976 S. W. Benson, Thermochemical Kinetics, Methods for the Estimation


of Thermochemical Data and Rate Parameters, 2nd Ed., Wiley, New York, 1976.
Evans 1935 M. G. Evans, M. Polanyi, Trans. Faraday Soc. 31, 875 (1935).
Eyring 1935 H. Eyring, J. Chem. Phys. 3, 107 (1935).
Herschbach 1987 D. R. Herschbach, Molekulardynamik einfacher chemischer Reak-
tionen, Angew. Chem. 99, 1251 (1987).
Lee 1987 Y. T. Lee, Molekularstrahluntersuchungen chemischer Elementarprozesse,
Angew. Chem. 99, 967 (1987).
Polanyi 1987 J. C. Polanyi, Konzepte der Reaktionsdynamik, Angew. Chem. 99, 981
(1987).
Wigner 1938 E. Wigner, Trans. Faraday Soc. 34, 29 (1938).
7.5 References 180

8. Unimolecular reactions

Unimolecular reactions are reactions in which only one species experiences chemical
change.

I Figure 8.1: Examples of unimolecular reactions.

8.1 Experimental observations

(1) There are many reactions governed by a first-order rate law,


 [A]
= −∞ [A] , (8.1)

for example isomerization and dissociation reactions like
Cyclopropane → Propene (8.2)
-Butene → -Butene (8.3)
C2 H6 → 2 CH3 (8.4)
Cycloheptatriene → Toluene (8.5)

(2) In these reactions, no other species than the reactants experience any chemical
change.
(3) These reactions have the same rates in the gas phase and in solution,
gas phase ≈ solution phase , (8.6)
i.e. the environment has no effect on the reaction rates!
8.1 Experimental observations 181

(4) There are many similar other reactions, however, for which the rate law is second-
order, i.e.
 [A]
= − [A] [M] (8.7)

Examples are

Br2 → Br + Br (8.8)
H2 O2 → OH + OH (8.9)

(5) More detailed measurements show that at a given temperature the order of the
reaction may depend on pressure (see Fig. 8.2), for instance for the reaction

CH3 NC → CH3 CN (8.10)

one has found that

a) for  → 0:
 [CH3 NC]
= − [CH3 NC] [M] (8.11)

This is called the “low pressure range” for the reaction.
b) for  → ∞:
 [CH3 NC]
= −∞ [CH3 NC] (8.12)

This is called the “high pressure range” for the reaction”.

(6) The transition from the low to the high pressure range depends

a) on the size of the molecule (see Fig. 8.3),


b) for a given molecule, it depends on temperature.

(7) The activation energy in the low pressure regime is much smaller than in the high
pressure regime:
 ln 
 =  2 (8.13)

where
0  ∞ (8.14)
and
∞ ≈ 0 (8.15)

(8) The preexponential factors in the low pressure regime are much higher than the
collision frequencies, i.e.
0   (8.16)

(9) For all these reactions, the molecules must somehow be “energized”. There
are other reactions, in which the molecules are energized chemically (“chemically
activated reactions”) or by photons. At the microscopic, i.e. molecular level, these
reactions should have related mechanisms.
8.1 Experimental observations 182

I Questions:

(1) What are the differences of the reactions under (1), (4) and (5)?
(2) How can we rationalize their reaction mechanisms?

I Figure 8.2: Schematic fall-off curve of the unimolecular rate constant.

I Figure 8.3: Experimentally observed fall-off curves of unimolecular reactions.


8.2 Lindemann mechanism 183

8.2 Lindemann mechanism

First attempts to explain the pressure dependence of unimolecular reaction rate con-
stants due to radiation and the dissociation dynamics due to centrifugal forces (rotational
excitation) proved to be wrong. In addition, the historic examples for unimolecular re-
actions, for instance the N2 O5 decomposition, were later proven to have more complex
reaction mechanisms. The first “realistic” model that explained the pressure dependence
of  was that proposed by Lindemann (1922).

I Figure 8.4: Lindemann mechanism.

I Figure 8.5: Lindemann mechanism: High pressure regime.


8.2 Lindemann mechanism 184

I Figure 8.6: Lindemann mechanism: Low pressure regime.

I Figure 8.7: Fall-off curve: The proper way to plot log  is vs. log [M].ab

a
Since 0 ∝ [M], we can also plot log( ∞ ) vs. log 0 . This gives the so-called doubly reduced
fall-off curves.
b
In reality, the fall-off regime extends over several orders of magnitude in [M].
8.2 Lindemann mechanism 185

I Half-pressure: We have found the following results:

2 1 [M]
 = (8.17)
−1 [M] + 2

1
∞ = 2 (8.18)
−1
0 = 1 [M] (8.19)
Thus, if −1 [M] = 2 , we have

2 1 [M] 2 1 [M] 1 2 1 1
 = = = = ∞ (8.20)
−1 [M] + 2 2−1 [M] 2 −1 2

Half “pressure”:
2
[M]12 = (8.21)
−1

I Lifetimes of energized molecules: It is instructive to look at the reciprocal values:

1 1
= (8.22)
−1 [M] 2

1
• is the time between two collisions. This is, in fact, the lifetime of the
−1 [M]
energized molecules with respect to collisional deactivation.
1
• is the lifetime of the excited molecules with respect to reaction.
2

Thus, at the half pressure, we have

 Collisional Deactivation =  Unimolecular Decay (8.23)

I Determination of k∞ :
2 1 [M]
 = (8.24)
−1 [M] + 2
y
1 −1 1 1
= + (8.25)
 1 2 1 [M]
1 1 1
= + (8.26)
∞ 1 [M]
1 1
= + (8.27)
∞ 0

y Plot of  −1 vs. [M]−1 should give a straight line with slope 1 −1 and intercept
∞ −1 .
8.2 Lindemann mechanism 186

I Deficiencies of the Lindemann model:

I Figure 8.8:

• Activation energies:
0  ∞ (8.28)
In particular:
∞ ≈ 0 (8.29)
 0 ≈ 0 − ( − 15)  (8.30)
8.3 Generalized Lindemann-Hinshelwood mechanism 187

8.3 Generalized Lindemann-Hinshelwood mechanism

8.3.1 Master equation

I Figure 8.9: Strong collision model.

I Figure 8.10: Generalized Lindemann-Hinshelwood weak collision model.


8.3 Generalized Lindemann-Hinshelwood mechanism 188

I Figure 8.11: Generalized Lindemann-Hinshelwood weak collision model.

I Figure 8.12: Generalized Lindemann-Hinshelwood model: Rate constants in the high


and low pressure (strong and weak collision) limits.
8.3 Generalized Lindemann-Hinshelwood mechanism 189

8.3.2 Equilibrium state populations

As seen from the master equation analysis, the unimolecular reaction rate constant
in the high and in the low pressure regimes depends on the equilibrium (Boltzmann)
state distributions  (for discrete states) or  () (continuous state distributions). In
a polyatomic molecule, due to the large number of vibrational degrees of freedom, the
number of vibrational states increases very rapidly with increasing vibrational excitation
energy. The quanta of vibrational excitation can be distributed over the oscillators. In
order to calculate the number of combinations, we start with a single oscillator and
expand our scope first to two oscillators and then to  = 3 − 6 oscillators.

a) Equilibrium state population for a single oscillator:

I Boltzmann distribution:
 −  
 = (8.31)

with the energy quanta
 =  ×  ;  = 0 1 2    (8.32)
the degeneracy factor
 (8.33)
and the partition function
1
 = (8.34)
1 − − 

I Limiting case for high T (transition to classical mechanics):  ¿   y


 
classical
 = (8.35)

since, with  =   ,
− ≈ 1 −  for  → 0 (8.36)
Note that this classical description of vibrations is not so bad for unimolecular reactions,
where we are interested in the kinetic behavior at high temperatures.

b) Equilibrium state population for s oscillators:

 () →  ()  (8.37)


y
 () −  
 ()  = (8.38)

with  () being the density of vibrational states at the energy  and
Y

1
 = (8.39)
=1
1− −   

and
 = 3 − 6 (resp.  = 3 − 5) (8.40)
8.3 Generalized Lindemann-Hinshelwood mechanism 190

8.3.3 The density of vibrational states ρ (E)

The density of vibrational states is defined as


number of states in energy interval 
 () = (8.41)
energy interval 
i.e.
 ()
 () = (8.42)


In a polyatomic molecule, as we shall see in the following,  () increases with  very
rapidly owing to the large number of overtone and combination states, and reaches truly
astronomical values:

a) s = 1:

1
 () = (8.43)


b) s = 2:
We consider the ways, in which we can distribute two identical energy quanta  between
the two oscillators:
 osc.1 osc.2 osc.1 osc.2 osc.1 osc.2 osc.1 osc.2 
0 - - 1
 • - - • 2
2 •• - • • - •• 3
3 ••• - •• • • •• - ••• 4 (8.44)
4 5
5 6
.. .. .. ..
. . . .

c) s oscillators:

• For simplicity, we consider a molecule with  identical harmonic oscillators.


• This molecule shall be excited with  quanta, each of size , i.e. the molecule
has the excitation energy  =  

I Question: How can we distribute these  quanta over the  oscillators?


8.3 Generalized Lindemann-Hinshelwood mechanism 191

I Answer: We are asking for the number of distinguishable possibilities to distribute 


quanta between  oscillators,  (). This is identical to the problem of distributing 
balls between  boxes, for example for  = 11 and  = 8:
osc.1 osc.2 osc.3 osc.4 osc.5 osc.6 osc.7 osc.8
(8.45)
••• • • • •• • ••
That number is identical to the number of distinguishable permutations of  balls ( • )
and  − 1 walls ( | ),
( +  − 1)!
 () = (8.46)
! ( − 1)!
where ( +  − 1)! is the total number of permutations, which is corrected by dividing
through ! ( − 1)! to account for the number of indistinguishable permutations among
the balls (!) and walls (( − 1)!) among each other (which give indistinguishable re-
sults).
I Question: How can we compute  ()?
I Answer:
 ()
 () = (8.47)


(1) We consider the energy interval


∆ =  (8.48)
at the excitation energy
 =  ×  (8.49)
In this energy interval, we have the number of states
∆() =  () (8.50)
y
 () ∆ () 1 ( +  − 1)!
 () = ≈ = (8.51)
 ∆  ! ( − 1)!

(2) For  À , we can use the approximation51


( +  − 1)!
≈  −1 (8.54)
!
to obtain
1  −1
 () = (8.55)
 ( − 1)!

51
The approximation
( +  − 1)!
≈  −1 (8.52)
!
can be derived using Stirling’s formula (ln  ! =  ln  −  ) and the power series expansion of

ln (1 − ) ≈  (8.53)

for || ¿ 1 (see homework assignment).


8.3 Generalized Lindemann-Hinshelwood mechanism 192

(3) Inserting  =  into  (), we obtain

1 ()−1
 () = (8.56)
 ( − 1)!

(4) Result:
 −1
 () = (8.57)
( − 1)! ()

d) Notes:*

(1) The above derivation may seem artificial because in a real molecule not all oscil-
lators are identical. However, the derivation can be easily generalized.
(2) For instance, we can start from the number of states of the first oscillator 1 (1 )
and convolute this with the density of states 2 () of the second oscillator
1
2 (2 ) = 2 ( − 1 ) = (8.58)

and compute the total (combined) density of states for two oscillators by convo-
lution according to
Z 
 () = 1 (1 ) 2 ( − 1 ) 1 (8.59)
0

etc. up to  oscillators. The result is

 −1
 () = Q (8.60)
( − 1)! =1  

(3) We can also use inverse Laplace transforms: Since  is the Laplace transform
of  () according to
Z ∞
 =  () −   = L [ ()] (8.61)
0

we can determine  () from  (which we know) by inverse Laplace transfor-
mation.
(4) With corrections for zero-point energy:

( +  )−1
 () = Q (8.62)
( − 1)! =1  
8.3 Generalized Lindemann-Hinshelwood mechanism 193

(5) With (empirical) Whitten-Rabinovitch corrections:

( +  ()  )−1
 () = Q (8.63)
( − 1)! =1  

where  () is a correction factor that is of the order of 1 except at very low
energies.
(6) Exact values of  () for harmonic oscillators can be obtained by direct state
counting algorithms.
(7) Corrections for anharmonicity can be applied using different means.

8.3.4 Unimolecular reaction rate constant in the low pressure regime

From the master equation analysis above, we obtained the thermal unimolecular reaction
rate constant in the low pressure regime as
XX 
0 = −1() [M]  (8.64)
 


The reduction of the excited state populations compared to the equilibrium (Boltzmann)
distribution is shown in Fig. 8.13. Two cases can be distinguished:

(1) With the strong collision assumption (h∆ i À  ), we can apply the so-called
equilibrium theories which assume that  =  for  ≤ 0 . Thus, the state
population below 0 remains as in the high pressure regime, whereas it is reduced
above 0 .Then,
XX  XX X
0 = −1() [M]  = −1() [M]  =  [M]  (8.65)
 
   

y
Z∞
0 =  [M]  ()  (8.66)
0

Using the expressions for  ()   () and  in the classical limit ( ¿   )
derived above and doing the integration by parts, we obain the Hinshelwood
equation for 0 :
µ ¶−1
0 1
0 =  [M] −0   (8.67)
  ( − 1)!

This equation is usually used to describe experimental data with  as a fit para-
meter (usually about half the real ).
Since  ∝  05 and  =  2  ln  , the low pressure activation energy
becomes
0 = 0 − ( − 15)  (8.68)
Thus,  may be significantly smaller than 0 . This can be understood as
illustrated in Fig. 8.14.
8.3 Generalized Lindemann-Hinshelwood mechanism 194

(2) With the weak collision assumption (h∆ i ≤  ) made by the so-called non-
equilibrium theories, the bottleneck in the state population falls below the equi-
librium distribution already below 0 (see Fig. 8.14). Using the weak collision
assumption, Troe derived the expression

   (0 ) −0  


0 =    [M]  (8.69)


where  is the Lennard-Jones collision frequency and   is the so-called colli-
sion efficiency (  ≤ 1) which can be expressed in terms of the average energy
transferred per collision h∆ i.  becomes even smaller than the Hinshelwood
 above. A simple estimate of the weak collision effect based on Fig. 8.14 may
lead to
0 ≈ 0 − ( − 05)  (8.70)

I Figure 8.13: State populations.


8.3 Generalized Lindemann-Hinshelwood mechanism 195

I Figure 8.14: Activation energies for unimolecular reactions in the high pressure, the
low pressure strong collision, and the low pressure weak collision limits.

8.3.5 Unimolecular reaction rate constant in the high pressure regime

From the master equation analysis above, the thermal unimolecular reaction rate con-
stant in the high pressure regime is (see Fig. 8.12)

Z∞
∞ =  ()  ()  (8.71)
0

with the Boltzmann distribution


 () −  
 ()  = (8.72)

as shown in Fig. 8.13.
Averaging over the Boltzmann distribution gives the TST expression

  ‡ −0  
 ( ) =  (8.73)
 
8.4 The specific unimolecular reaction rate constants  () 196

8.4 The specific unimolecular reaction rate constants k (E)

We may intuitively expect that the specific uinimolecular reaction rate constants  ()
depend on the energy content (i.e. the excitation energy ) of the energetically acti-
vated molecules molecules A∗ . More precisely, we will also have to take into account
the dependence of the specific rate constants on other conserved degrees of freedom,
in particular total angular momentum  (⇒  ( )), and perhaps other quantities
(symmetry, components of  if -mixing is weak, (⇒  ( ; ;  ))).

8.4.1 Rice-Ramsperger-Kassel (RRK) theory

I RRK model: In order to derive an expression for the specific rate constants  (),
we realize that it is not enough that the molecules are energized to above 0 . Indeed,
for the reaction to take place, the energy also has to be in the right oscillator, namely
in the RC. In particular, for the reaction to take place, of the total excitation energy ,
the amount  † has to be concentrated in the RC such that  † ≥ 0 . Thus, we can
write the reaction scheme as
†
∗ → † →   (8.74)

where † denotes those excited molecules that have enough energy in the RC to over-
come 0 . We assume that once this critical configuration († ) is reached, the molecules
will immediately cross the TS and go to the product side.

I Probabilities of A† and A∗ : The probability of † compared to ∗ can be derived


using statistical arguments by considering the ratio of the density of states. We consider
a molecule with

•  oscillators,
•  =  quanta which have to be distributed over the  oscillators,
• a threshold energy of 0 such that a minimum of  = 0  quanta have to be
concentrated in the RC and only  −  quanta can be distributed freely,
•  À  and  −  À .

The probability of † relative to ∗ is given by


¡ ¢ ¡ ¢
 †  †
= (8.75)
 (∗ )  (∗ )
where
1 ( +  − 1)!
 (∗ ) ∝ ∗ () = (8.76)
 ! ( − 1)!
and, since in the critical configuration only  −  quanta can be distributed freely,
¡ ¢ 1 ( −  +  − 1)!
 † ∝ † () = (8.77)
 ( − )! ( − 1)!
8.4 The specific unimolecular reaction rate constants  () 197

y
¡ ¢
 † ( −  +  − 1)! ! ( − 1)!

= × (8.78)
 ( ) ( − )! ( − 1)! ( +  − 1)!
( −  +  − 1)! !
= × (8.79)
( − )! ( +  − 1)!

I Specific rate constant: The critical configuration † is reached with the rate coeffi-
cient (≡ frequency factor)  † (see the reaction scheme above). Since the energy is now
in place (in the RC), † will cross the TS to products.
The specific rate constant  () is therefore simply
¡ †¢
 
 () =  † × (8.80)
 (∗ )
y
( −  +  − 1)! !
 () =  † × × (8.81)
( − )! ( +  − 1)!

I Kassel formula: With the approximations according to Eq. 8.54),

( +  − 1)!
≈  −1 for  À  (8.82)
!
and
( −  +  − 1)!
≈ ( − )−1 for  −  À  , (8.83)
( − )!
we obtain ¡ ¢ µ ¶−1
 † ( − )−1 −
= = (8.84)
 (∗ )  −1 
y ¡ ¢ µ ¶−1
 †  − 0
= (8.85)
 (∗ ) 
so that the specific rate constant becomes
µ ¶−1
†  − 0
 () =  (8.86)

This expression is known as the Kassel formula for  ().

I Notes on the Kassel formula:

• Owing to the approximations  À  and  −  À , the Kassel formula is valid


only at  À 0 .
8.4 The specific unimolecular reaction rate constants  () 198

• The Kassel formula describes experimental data qualitatively well. However, in-
stead of the full value of  = 3 − 6, one usually has to use an effective number
eff . As suggested by Troe, eff can be estimated from the specific heat of the
reactant molecules via
µ ¶ µ ¶
  2  ln 
 = =    ≈ eff  (8.87)
   
using the known expression for  ,
Y

1
 = (8.88)
=1
1 − exp (−    )

• Often, eff is equal to about one half of the number of oscillators,


(3 − 6)
eff ≈ (8.89)
2

I Figure 8.15: Specific rate constants: RRK formula.

8.4.2 Rice-Ramsperger-Kassel-Marcus (RRKM) theory

RRK theory fails in the following points:

• It can be used only with  as an effective parameter (eff ).


• It leads to the wrong results close to threshold, where  → 0 (see the conditions
above).
• It is difficult to derive for oscillators with different frequencies.
• RRK theory is essentially a classical theory. It does not know about zero-point
energy.
8.4 The specific unimolecular reaction rate constants  () 199

I RRKM expression: These problems are essentially solved by RRKM theory which
gives the expression
 ‡ ( − 0 )
 () = (8.90)
 ()
where

•  ‡ ( − 0 ) is the number of open channels (essentially the number of states at


the transition state configuration exluding the RC), counted from the zero-point
level of the TS, at the energy  − 0 ,
•  () is the density of states of the reacting molecules at the energy .

These quantities are the two critical quantities in microcanonical transition state theory.

I Figure 8.16: RRKM expression for ().


8.4 The specific unimolecular reaction rate constants  () 200

I Figure 8.17: RRKM expression for ( ) with angular momentum conservation.

I Figure 8.18: The sum and the density of states.


8.4 The specific unimolecular reaction rate constants  () 201

I RRKM expression with angular momentum conservation: Eq. 8.90 can be gen-
eralized to include other conserved quantities, e.g. total angular momentum,

 ‡ ( − 0  )
 ( ) = (8.91)
 ( )

I Relation between RRKM and RRK theory: To elucidate the relation between Eqs.
8.86 and 8.90, we look at  () and  ( − 0 ) more closely.

(1) Density of vibrational states  ():

number of states in energy interval  ()


 () = = (8.92)
energy interval 

a) RRK result for the number of indistinguishable permutations of  quanta



over  oscillators;  = À :

1 ( +  − 1)! 1  −1 1 ()−1
 () = ≈ = (8.93)
 ! ( − 1)!  ( − 1)!  ( − 1)!
y
 −1
 () = (8.94)
( − 1)! ()

b) It is relatively easy to show that for oscillators with different frequencies, the
expression for  () becomes

 −1
 () = Q (8.95)
( − 1)! =1  

c) With corrections for zero-point energy:

( +  )−1
 () = Q (8.96)
( − 1)! =1  

d) With (empirical) Whitten-Rabinovitch correction:

( +  ()  )−1
 () = Q (8.97)
( − 1)! =1  

where  () is a correction factor that is of the order of 1 except at very


low energies.
e) Exact values of  () for harmonic oscillators can be obtained by direct
state counting algorithms. Corrections for anharmonicity can be applied
using different means.
8.4 The specific unimolecular reaction rate constants  () 202

(2) Number of states  () and  ‡ ( − 0 ): Since  () =  () , the
number of states from  = 0 to  can be obtained by integration,

Z
 () =  ()  (8.98)
0

a) RRK expression:

Z Z Z
 −1 1 1
 () =  ()  =   =  −1 
( − 1)! () ( − 1)! ()
0 0 0
(8.99)
y

 () = (8.100)
! ()

b) Allowing for different frequencies, we modify this expression to


 () = Q (8.101)
! =1  

c) As above, a correction for zero-point energy gives

( +  )
 () = Q (8.102)
! =1  

d) With the Whitten-Rabinovitch correction, this becomes

( +  ()  )
 () = Q (8.103)
! =1  

e) Correction for  → 0: At  = 0, we must have at least one state, i.e.

( +  ()  ) ( (0)  )
 ‡ () = 1 + Q − Q (8.104)
! =1   ! =1  

(3) Application to unimolecular reactions:

a) At the TS, we have only  − 1 oscillators since the RC has to be excluded.


Furthermore, the states range from the zero-point energy 0 of the TS to
, while we count  from the zero-point of the reactants. y
³ ´−1 ³ ´−1
 − 0 + ‡ ( − 0 ) ‡ ‡ (0) ‡
 ‡ ( − 0 ) = 1 + Q − Q
( − 1)! −1
=1  ‡
 ( − 1)! −1 ‡
=1  
(8.105)

This expression gives the correct value of  ( − 0 ) = 1 at  = 0 .
8.4 The specific unimolecular reaction rate constants  () 203

b) Density of states:
( +  ()  )−1
 () = Q (8.106)
( − 1)! =1  

c) Specific rate constant:


³ ´−1 ³ ´−1
‡ ‡ ‡ ‡
1  − 0 +  ( − 0 )   (0) 
 () = + Q−1 ‡ − Q
 ()  () ( − 1)! =1    () ( − 1)! −1 ‡
=1  
(8.107)

(4) For energies sufficiently above 0 , we can use the expressions


³ ´−1
 − 0 + ‡
 ‡ ( − 0 ) = Q (8.108)
( − 1)! −1 ‡
=1  

and
( +  )−1
 () = Q (8.109)
( − 1)! =1  
to obtain
³ ´−1

Q  −  +  ‡
 ( − 0 ) =1  
0 
 () = = Q−1 (8.110)
 () ‡
=1    ( +  )−1
y à !−1
Q ‡
=1    − 0 + 
 () = Q−1 ‡
(8.111)

=1 
 + 

The resemblence to the Kassel formula is obvious.

8.4.3 The SACM and VRRKM model

More advanced models allow for the tightening of the TS with increasing  (variational
RRKM theory = VRRKM theory) or even for individual adiabatic channel potential
curves (statistical adiabatic channel model = SACM). The latter model has to be used
for reactions with loose transition states.
8.4 The specific unimolecular reaction rate constants  () 204

8.4.4 Experimental results

I Figure 8.19: Experimental methods for the preparation of highly vibrationally excited
molecules with defined excitation energy.

I Figure 8.20: Comparison of theoretically predicted specific rate constants  () for uni-
molecular isomerization of cycloheptatrienes with directly measured experimental data.
8.5 Collisional energy transfer* 205

8.5 Collisional energy transfer*

Skipped for lack of time.


8.6 Recombination reactions 206

8.6 Recombination reactions

See homework assignment.


8.6 Recombination reactions 207

8.7 References

Lindemann 1922 F. A. Lindemann, Trans. Faraday Soc. 17, 598 (1922).


Hinshelwood 1927 C. N. Hinshelwood, Proc. Roy. Soc. A113, 230 (1927).
Kassel 1928a J. Phys. Chem. 32, 225 (1928).
Kassel1928b J. Phys. Chem. 32, 1065 (1928).
Kassel 1932 L. S. Kassel, The Kinetics of Homogeneous Gas Reactions, Chem. Cata-
log, New York, 1932.
Rice 1927 O. K. Rice, H. C. Ramsperger, J. Am. Chem. Soc. 49, 1617 (1927).
Rice 1928 O. K. Rice, H. C. Ramsperger, J. Am. Chem. Soc. 50, 617 (1928).
Troe1979 J. Troe, J. Phys. Chem. 83, 114 (1979).
Troe1994 J. Chem. Soc. Faraday Trans. 90, 2303 (1994).
8.7 References 208

9. Explosions

9.1 Chain reactions and chain explosions

9.1.1 Chain reactions without branching

I Reaction scheme:

(1) A →X chain initiation


(2) X + B → P + X chain propagation
(4) X →A chain termination

I Rate laws:
 [X]
= 1 [A] − 4 [X] (9.1)

 [P]
= 2 [B] [X] (9.2)

As we see, reaction (2) does not appear in the rate law because  →  !

I Solutions of Eq. 9.1 for two limiting cases (cf. Fig. 9.1):

(1) Solution for  → 0 and [A] ≈ const by substitution, using

 = 1 [A] − 4 [X]

y
 
= −4 y  [X] = −
 [X] 4
y
1  
− =y = −4 
4  
y
 =  −4  = 1 [A] − 4 [X] (9.3)
y
1
[X] = [A] −  0 −4  (9.4)
4
@ = 0:
1
[X] = 0 y  0 = [] (9.5)
4
y
1 [A] ³ ´
[X] = 1 − −4  (9.6)
4
9.1 Chain reactions and chain explosions 209

(2) Solution with steady state assumption for [X] at  À  (i.e, after the induction
time  ):
 [X]
≈0 (9.7)

y
1
[X]ss = [A] (9.8)
4
y Considering the free radical concentration profile [X ()], no desasters are hap-
pening.

I Figure 9.1: Evolution of the free radical concentration in a normal chain reaction.

9.1.2 Chain reactions with branching

I Reaction scheme:

(1) A →X chain initiation


(3) X + B → P +  X chain branching
 = branching factor
(4) X →A chain termination

I Most important example for a chain branching reaction:


H + O2 → OH + O (9.9)
This reaction is responsible for flame propagation in all hydrocarbon flames: Since H is
small, it has the highest diffusion coefficient and diffuses rapidly into the unburnt gases.
9.1 Chain reactions and chain explosions 210

I Rate laws:
 [X]
= 1 [A] + ( − 1) 3 [X] [B] − 4 [X] (9.10)

 [P]
= 3 [B] [X] (9.11)


I Solutions for three limiting cases:

(1)  = 1 or 4  ( − 1)3 [B]:


 [X]
≈0 (9.12)

y
1 [A]
[X] = (9.13)
4 − ( − 1)3 [B]
y
 [P]
= 3 [B] [X] =    (9.14)

The overall reaction is stable, because the radical concentration is stable. The
product formation rate is final (constant, well behaved).
(2) 4 = ( − 1)3 [B]: At short times, [A]  [B] ≈ const y

 [X]
= 1 [A] = const (9.15)

y
[X] → ∞ (9.16)
The overall reaction turns unstable, because the radical concentration goes to
infinity and thus the product formation rate too, leading to chain explosion!
(3) 4  3 ( − 1) [B]: At short times, [A]  [B] ≈ const y

 [X]
+ 4 [X] − ( − 1)3 [X] [B] = 1 [A] (9.17)

 [X]
+ [X] (4 − ( − 1)3 [B]) = 1 [A] (9.18)

y ³ ´
1 [A]
[X] = × exp [( − 1) 3 [B] ] − 1 (9.19)
( − 1) 3 [B] − 4
The overall reaction turns unstable, because the radical concentration increases
exponentially, leading to an exponentially growing product formation rate, and
therefore chain explosion!
9.1 Chain reactions and chain explosions 211

I Figure 9.2: Evolution of the free radical concentrations in a chain reaction with branch-
ing (three limiting cases).
9.2 Heat explosions 212

9.2 Heat explosions

We consider an exothermic bimolecular reaction in a closed reactor

A+B→C+D ∆  ª  0 (9.20)

and look at the radial temperature profile in the reactor.

I Figure 9.3: Temperature profiles in a stirred and in an unstirred reactor.

9.2.1 Semenov theory for “stirred reactors” (T = const)

I Heat production in an exothermic reaction (∆ H ª < 0):

• Energy released per volume:


£ ¤
∆ ∆  ª kJ cm3 (9.21)

• Reaction rate:

= − [] [] (9.22)

• heat release:
 
 =  ∆  ª (9.23)

ª
=  ∆  0 −    (9.24)
| {z } | {z }
Arrhenius rate constant in pressure units
9.2 Heat explosions 213

• positive feedback effect:


reaction rate increases
y temperature  ↑
y reaction rate ↑
y heat explosion !!

I Figure 9.4: Heat explosions.

I Heat removal: At constant temperature in the stirred reactor, we use Newton’s law

 =   ( − 0 ) (9.25)
where
 : heat conduction coefficient (depends on material) (9.26)
 : reactor surface area
 : temperature
0 : wall temperature
As we see, the heat removal increases (becomes steeper), when we increase the reactor
surface area (at constant reactor volume).

I Stability limits:
(1) 0 = 1 : Heat removal wins over heat production y negative feedback y stable.
(2) 0 = 2 : Heat removal wins over heat production y negative feedback y stable.
(3) 0 = 3 : Heat production wins over heat removal y positive feedback y heat
explosion, but the system should never reach point [3].
(4) 0 = 4 : tangent y stability limit, the system runs away upon the smallest
perturbation!
(5) 0 = 5 : always unstable!
9.2 Heat explosions 214

I Conclusions:

• The quantity that determines whether the system is stable or not is the wall
temperature 0 , because 0 determines the initial point of the heat removal line.
• Smaller surface-to-volume ratio is dangerous, because smaller surface at a given
volume leads to a smaller slope of the heat removal line y instability!
• Security measure: internal cooling system (cold water pipes, heat exchanger).
• Emergency measure: when cooling system fails, the reaction mixture should be
strongly diluted by addition of solvent immediately!

I Quantitative estimation of the stability limits:* At point (4) in Fig. 9.4, we have:
 
 =  (9.27)
y   ( − 0 ) =  ∆  ª 0 −    (9.28)
 
   
= (9.29)
 

y   =  ∆  ª 0 −    (9.30)
 2
2
Eq  
y = ( − 0 ) (9.31)
Eq  
 
y 2 − + 0 = 0 (9.32)
 s
µ ¶2
  
y  = ± − 0 (9.33)
2 2 

Solution is: ⎧ ³ ´⎫
⎪ 1 p ⎪
⎨ =  + (2 − 40  ) ⎬
2 ³ p ´ (9.34)
⎪ 1
⎩ =  − ( − 40  ) ⎪
2 ⎭
2
This can be simplified:

(1) For practically important reactions, we have:



 À 0 y 0 ¿ (9.35)

(2) We also can exclude unreasonably hight temperatures y only the −-sign before
√ counts: y
sµ ¶
2
  
∗ = − − 0 (9.36)
2 2 
r
  40
= − 1− (9.37)
2 2 
à r !
 40
= 1− 1− (9.38)
2 
9.2 Heat explosions 215

(3) We can expand the √ :

40
 = ; 1; (9.39)

√ 1 1 2
y 1−=1− −  −  (9.40)
µ 2 2·4 ¶
∗  20 16 · 2 02
y  = 1−1+ + +  (9.41)
2  2 · 4 · ( )2
02
y  ∗ = 0 + (9.42)

02
y  ∗ − 0 = (9.43)


(4) Insert into in (I) and power expansion:


⎡ ⎤
µ ¶
02 
· · =  · ∆  ª · 0 · exp ⎣− ³  ´ ⎦ ·  · (9.44)
 0 1 +  0

∙ ¸

=  · ∆  ª · 0 ·  · exp − · 2 ·  ·  (9.45)
0

(5) Explosion limit:


∙ ¸
2  · ·
2
· exp − = ª = const (9.46)
0 0  · ∆  ·  ·  · 0 ·  · 

y
 ∝  ·   (9.47)

Power expansion of the exponent in step (4):


⎡ ⎤
∙ ¸
⎣  ⎦ 1 1 0
exp − ³ ´ = exp −   = ¿ 1 (9.48)
0 1 +  0 1+ 

∙ ¸
1¡ 2
¢
= exp − 1 −  +     (9.49)

∙ ¸
1
≈ exp − (1 − ) (9.50)

∙ ¸
1
= exp 1 − (9.51)

∙ ¸
1 1
=  · exp − (9.52)

∙ ¸

=  · exp − (9.53)
0
9.2 Heat explosions 216

I Figure 9.5:

9.2.2 Frank—Kamenetskii theory for “unstirred reactors”

I Figure 9.6: Temperature profiles in an unstirred reactor.

Boundary effects are neglected in the Figure. In reality, there is a smooth transition in
the boundary layer.
9.2 Heat explosions 217

9.2.3 Explosion limits

I Figure 9.7: Explosion limits in the H2 /O2 system.

9.2.4 Growth curves and catastrophy theory

I Exponential growth:

 ∝ · y =  (9.54)


I Hyperbolic growth:
1 
∝ y =  1+ mit 1 +   1 (9.55)
( − ) 

I Heat explosion:

 ∝    quadratic, i.e. we get hyperbolic growth ! (9.56)
9.2 Heat explosions 218

I Figure 9.8: Growth curves.


9.2 Heat explosions 219

10. Catalysis

The net rate of a chemical reaction may be affected by the presence of a third species
even though that species is neither consumed nor produced by the reaction. If the rate
increases, we speak of catalysis. If the rate decreases, we speak of inhibition.
In general, we have to distinguish

• homogeneous catalysis,
• heterogeoues catalysis.

10.1 Kinetics of enzyme catalyzed reactions (homogeneous catalysis)

The mechanism describing the homogeneous catalytical activity of enzymes has first
been elucidated by Michaelis and Menten (1914).

I Experimental observations on enzyme catalysis:

(1) For constant substrate concentration [S]0 , the reaction rate  is proportional to
the concentration of the enzyme [E]0
 ∝ [E]0 for [S]0 = const (10.1)

(2) At an applied enzyme concentration [E]0 , the reaction rate  first increases with
increasing substrate concentrations [S]0 and then reaches saturation, as is shown
in Fig. 10.1.

I Figure 10.1: Enzyme catalysis: Michelis-Menten kinetics.


10.1 Kinetics of enzyme catalyzed reactions (homogeneous catalysis) 220

10.1.1 The Michaelis-Menten mechanism

Basic reaction scheme:


1
E + S À ES (10.2)
−1


ES →2 E + P (10.3)

Typical values of 2 :

2 ≈ 102 103 s−1 (in some cases up to 106 s−1 ) (10.4)

Reaction rate:
 [P]  [S]
=− = 2 [ES] (10.5)
 

 [ES]
= 1 [E] [S] − −1 [ES] − 2 [ES] (10.6)

≈ 0 (10.7)

Mass balance:

[E]0 = [E] + [ES] (10.8)


y [E] = [E]0 − [ES] (10.9)
[S]0 = [S] + [ES] + [P] (10.10)
y [S] = [S]0 − [ES] − [P] (10.11)

Steady state ES concentration:

 [ES]
= 1 ([E]0 − [ES]) ([S]0 − [ES] − [P]) − (−1 + 2 ) [ES] (10.12)

≈ 0 (10.13)

Approximations:

[E]0 ¿ [S]0 (10.14)


[ES] ¿ [S]0 (10.15)
[P] ≈ 0 for early times (10.16)

y
 [ES]
= 1 [E]0 [S]0 − 1 [S]0 [ES] − (−1 + 2 ) [ES] (10.17)

≈ 0 (10.18)

Steady state ES concentration:

1 [E]0 [S]0
[ES]ss = (10.19)
1 [S]0 + −1 + 2
10.1 Kinetics of enzyme catalyzed reactions (homogeneous catalysis) 221

Resulting rate expression:

 [P]
 = = 2 [ES] (10.20)

1 2 [E]0 [S]0
= (10.21)
1 [S]0 + −1 + 2
2 [E]0
= (10.22)
−1 + 2
1+
1 [S]0
y
2 [E]0
= (10.23)
1 +   [S]0
with the Michaelis-Menten constant

−1 + 2
 = (10.24)
1

Limiting cases:


(1) [S]0 → ∞ y ¿ 1: y
[S]0

 [P]
= = 2 [E]0 (10.25)



(2) [S]0 → 0 y À 1: y
[S]0

 [P] 2 [E]0 [S]0


= = (10.26)
 

Determination of 2 and  from a Lineweaver-Burk plot (Fig. 10.2):

1 1  1
= + (10.27)
 2 [E]0 2 [E]0 [S]0

• ⇒ A plot of −1 vs. [S]0 −1 should give a straight line with intercept (2 [E]0 )−1
and slope  2 [E]0 . y

slope −1 + 2
=  = (10.28)
intercept 1

• ⇒ If [E]0 can be reliably determined (which is often difficult in practice), 2 is


found from the intercept.
10.1 Kinetics of enzyme catalyzed reactions (homogeneous catalysis) 222

I Figure 10.2: Lineweaver-Burk plot.

10.1.2 Inhibition of enzymes

I Mechanisms for enzyme inhibition:

• Competitive inhibition: The inhibitor I binds to the active site of the enzyme.
• Noncompetitive inhibition: The inhibitor I binds to the enzyme (though not at
its active site) in a way that it still inhibits product formation.

I Extended reaction mechanism for competitive inhibition:


1
E + S À ES (10.29)
−1


ES →2 E + P (10.30)
3
E + I À EI (10.31)
−3

The EI complex is catalytically inactive, since I blocks the active site of E.


Inhibition constant:
[EI] 3
 = = (10.32)
[E] [I] −3
10.1 Kinetics of enzyme catalyzed reactions (homogeneous catalysis) 223

Reaction rate (Fig. 10.2):


 [ES]
=  ≈ 0 (10.33)

y
1 1 (1 +  [I])  1
= + (10.34)
 2 [E]0 2 [E]0 [S]0
y The slope of the plot of −1 vs. [S]0 −1
goes up with increasing [I] (see Fig. 10.2).
I Exercise 10.1: Derive Eq. 10.34 with a reasonable assumption for the concentration
of EI. ¤

I Extended reaction mechanism for noncompetitive inhibition:


1
E + S À ES (10.35)
−1


ES →2 E + P (10.36)
3
E + I À EI (10.37)
−3

4
ES + I À ESI (10.38)
−4

ESI 6→ EI + P (10.39)
EI may still bind S, but the complex ESI can no longer produce any P, because the
inhibitor inactivates E, e.g. changes the structure of E, such that E becomes inactive.
Inhibition constant:
[EI] 3
 = = (10.40)
[E] [I] −3
[ESI] 4
= = (10.41)
[ES] [I] −4

Reaction rate:
 [ES]
=  ≈ 0 (10.42)

y ∙ ¸
1 1  1
= + [1 +  [I]] (10.43)
 2 [E]0 2 [E]0 [S]0
1
y Both the slope of the plot of −1 vs. [S]0 −1
and the intercept with the -axis

increase with increasing [I].
I Exercise 10.2: Derive Eq. 10.43 with reasonable assumptions for the concentrations
of EI and ESI. ¤
10.2 Kinetics of heterogeneous reactions (surface reactions) 224

10.2 Kinetics of heterogeneous reactions (surface reactions)

I Examples for heterogeneous reactions (gas-solid interface): Many reactions,


which are very slow in the gas phase, take place as heterogeneous reactions in the
presence of a heterogeneous catalyst:

• catalytic hydrogenation
• NH3 synthesis (Haber-Bosch)
• catalytic exhaust gas treatment (internal combustion engines)
• catalytic DeNO processes
• CO oxidation
• oxidative coupling reactions of CH4 (oxide catalysts)
• atmospheric reactions (at gas-solid and gas-liquid interface)

10.2.1 Reaction steps in heterogeneous catalysis

Heterogeneous reactions take place via a series of steps:

• Diffusion in gas (or liquid) phase to catalyst surface.


• Adsorption:

— “physical” adsorption (“physisorption”),


— “chemical” adsorption (“chemisorption”).

• Diffusion on the surface.


• Chemical reaction on the surface.
• Desorption.
• Diffusion away from surface.

10.2.2 Physisorption and chemisorption

We distinguish between “physical” adsorption (“physisorption”) and “chemical” adsorp-


tion (“chemisorption”) on a surface based upon the adsorption energies (surface binding
energies). ⇒ Fig. 10.3.

• Physisorption:

— van-der-Waals type interaction,


— typical bond energy: ∆ ≤ −40 kJ mol−1 ,
10.2 Kinetics of heterogeneous reactions (surface reactions) 225

— example: Ar on a metal or oxide surface.

• Chemisorption:

— formation of a chemical bond between adsorbate and surface.


— typical bond energy: −300 kJ mol−1 ≤ ∆ ≤ −100,
— example 1: CO on Pd:
O
||
C (10.44)
|
− Pd − Pd −
∗ chemical bond by overlap of  orbital of the C atom with free bands of
Pd,
∗ back donation from occupied bands into  ∗ orbital of the CO substrate.
∗ experiment shows that (Pd − C = O) ≈ 236 cm−1 .
— example 2: thiols on Au(111):
∗ S is strongly bound to one Au atom.
∗ thiol molecules can move around on the Au surface with the Au atom,
because the S—Au bond is stronger than the Au-Au metallic bond.

I Figure 10.3: Potential curves for physisorption and chemisorption.


10.2 Kinetics of heterogeneous reactions (surface reactions) 226

10.2.3 The Langmuir adsorption isotherm

I Reaction scheme:
A
| 1 |
A() + − S − À −S− (10.45)
−1

|
− S− = “active surface site” (10.46)

I Fraction of occupied active surface sites Θ:

• Adsorption rate:
1 × (1 − Θ) ×  (10.47)

• Desorption rate:
−1 × Θ (10.48)

• Steady state:
1 (1 − Θ)  = −1 Θ (10.49)
y
1  = (−1 + 1 ) Θ (10.50)
y

I Langmuir isotherm (Fig. 10.4):

1   
Θ= = (10.51)
−1 + 1  1 +  

with
1
 = (10.52)
−1
10.2 Kinetics of heterogeneous reactions (surface reactions) 227

I Figure 10.4: Langmuir adsorption isotherm.

I Determination of Θ and the adsorption enthalpy ∆H : At a series of tem-


peratures  , we need to measure
 (  )
Θ (  ) = ∞
(10.53)
 ( )
as function of  by pressure/volume measurements in an ultrahigh vacuum apparatus.

• From Eq. 10.51:


1 ∞ 1
=  = 1 + (10.54)
Θ   

• Multiplication of l.h.s. and r.h.s. with ∞ :

 1 
= ∞
+ ∞ =  +  (10.55)
   

• A plot of vs.  gives a straight line (see the inset in Fig. 10.4) with slope

∞ −1
 = ( ) (10.56)
and intercept
∞ −1
 = (  ) (10.57)

slope 
 = = (10.58)
intercept 
• van’t Hoff plot:
 ln  ∆
=− (10.59)
 (1 ) 
10.2 Kinetics of heterogeneous reactions (surface reactions) 228

I Desorption rate: If ∆0  0 we can describe the desorption rate by TST (see Figs.
10.3 und 10.6). µ ¶
  ‡ ∆0
−1 = exp − (10.60)
    
The ’s are partition functions per unit area.

10.2.4 Surface Diffusion

I Figure 10.5: Surface Diffusion.

I Diffusion is an activated process:


µ ¶
∆
 = 0 exp − (10.61)

∆ = energy barrier of jump of adsorbate from one site to another.

I Fick’s 2nd law:


 ( )  2  ( )
=× (10.62)
 2

I Einstein’s diffusion law:

• 1D:
∆2 = 2 (10.63)
• 2D:
∆2 = 4 (10.64)
10.2 Kinetics of heterogeneous reactions (surface reactions) 229

10.2.5 Types of heterogeneous reactions

a) Dissociative Adsorption

I Scheme:
A A
| | 2 | |
A2 () + − S − S − À −S − S− (10.65)
−2

I Adsorption isotherm:

• Adsorption rate:
2 2 (1 − Θ)2 (10.66)

• Desorption rate:
−2 Θ2 (10.67)

y
( 2 )12
Θ= (10.68)
1 + ( 2 )12
with
2
 = (10.69)
−2

I Figure 10.6: Potential energy curves for dissociative adsorption.


10.2 Kinetics of heterogeneous reactions (surface reactions) 230

b) Surface catalyzed unimolecular dissociation

I Scheme:
A
| 3 |

A() + − S − À − S− →4 Products (10.70)
−3

I Example: Decomposition of PH3 on a W metal surface.

I Decomposition rate and two limiting cases:


 [P]  A
= 4 Θ = 4 × (10.71)
 1 +  A

(1)  A ¿ 1: The reaction becomes first-order in A .


 [P]
= 4  A (10.72)

(2)  A À 1: The reaction becomes zeroth order, i.e. independent of A !
 [P]
= 4 (10.73)


c) Bimolecular reactions

I Scheme:
A
|  |
A() + − S − À −S− (10.74)
−

B
|  |
B() + − S − À −S− (10.75)
−

A B A−B
| | |
5 
− S −+− S − → − S−S − →6 Products (10.76)

I Reaction rate: Assuming that A and B compete for the same active sites, we have
  (1 − Θ − Θ ) = − Θ (10.77)
  (1 − Θ − Θ ) = − Θ (10.78)
y
 [P] 5    
= 5 Θ Θ = (10.79)
 (1 +   +   )2
10.2 Kinetics of heterogeneous reactions (surface reactions) 231

I Limiting cases: It is often the case that one species is adsorbed much more strongly
than the other.

• Assume, for instance, that   ¿   . Then,


 [P] 5    
= (10.80)
 (1 +   )2

(1)   À 1:
 [P] 5  
= (10.81)
  
The reaction is inhibited by B, which is strongly adsorbed: At high  , all
active sites are blocked by B.
(2)   ¿ 1:
 [P]
= 5     (10.82)

The reaction is truly second order.

 [P]
⇒ We can see from these two cases that, as  increases, the rate first

increases, reaches a maximum, and then decreases again.

d) Example 1: Heterogeneous oxidation of CO

I Figure 10.7: Heterogeneous oxidation of CO on Pt. (a) Langmuir-Hinshelwood and


Eley-Rideal mechanisms, (b) potential energy diagram (values in kJ mol−1 ).
10.2 Kinetics of heterogeneous reactions (surface reactions) 232

e) Example 2: Haber-Bosch NH3 synthesis

I Figure 10.8: Potential energy diagram for the catalytic reaction N2 + 3H2 → 2 NH3
on a Fe3 O4 catalyst according to Ertl and co-workers (1982). Energies in kJ mol−1 .

f) How can we control and accelerate heterogeneous reactions?


The answer depends on which step is rate determining:

• If diffusion to/from the surface is the rate determining step, we can accelerate
the reaction by stirring
• If the adsorption step is rate determining (because of an activation energy), we
can accelerate the reaction by increasing  and  and, in particular, by increasing
surface area.
10.2 Kinetics of heterogeneous reactions (surface reactions) 233

11. Reactions in solution

“Everyone has problems - chemists have solutions.”

I Comparison of gas and liquid phase reactions (@ T = 298 K and p = 1 bar):

• Molecules in the gas phase occupy ≈ 02 % of the total volume, while molecules
in the liquid phase occupy ≈ 50 % of the total volume.
• Molecules in liquid are in permanent contact with each other, and undergo per-
manent collisions with each other.
• The crossing of a potential barrier in a reaction is not continuous as in gas phase
but affected by multiple collisions during the crossing.
• As a result of the permanent collisions, we can expect the reactant molecules to
be in thermodynamic equilibrium even above 0 . Thus TST should be applicable.
• However, it is very difficult to evaluate the partition functions in liquids. Hence,
one usually employs the thermodynamic version TST.
• Solvent molecules act as an efficient heat bath.

I Observations on liquid phase reactions: Experience has shown that reactions in


liquids have similar rates as in the gas phase, except if

• the reaction occurs with the solvent,


• there are strong interactions of the reactant molecules with the solvent (e.g. ionic
reactions; ⇒ solvent shells),
• the reactions are diffusion controlled.

11.1 Qualitative model of liquid phase reactions

In contrast to gas phase reactions for which we have advanced theories (⇒ transition
state theory, unimolecular rate theory), the theory for liquid phase reactions is much
less well developed. However, we can understand the required basic reaction steps as
sketched in Fig. 11.1 and a number of limiting cases resulting from these steps.
11.1 Qualitative model of liquid phase reactions 234

I Figure 11.1: Structure and dynamics of solutions.

I Formal kinetics: We distinguish between the formation of the encounter complex in


the solvent cage (by diffusion) and the actual reaction:
 
+ À {} →  (11.1)
−

y
 [{}]
=  [] [] − − [{}] −  [{}] ≈ 0 (11.2)


[{}] = [] [] (11.3)
− + 
y
 [ ]  
=  [{}] = [] [] =  [] [] (11.4)
 − + 
with
 
= (11.5)
− + 

I Rate constant for liquid phase reactions:

 
= (11.6)
− + 
11.1 Qualitative model of liquid phase reactions 235

I Two limiting cases:

(1) Diffusion control:  À − y

 =  (11.7)

This case is observed when  = 0 and the diffusion coefficient  is small.
We shall find below that
 = 4  (11.8)

(2) Reaction (or activation) control:  ¿ − y


=  = {}  (11.9)
−
In this case, the reaction is said to be activation controlled because usually it has
a sizable  .
 can be described by TST:
  ∆ +  −∆ + 
 =   (11.10)

11.2 Diffusion-controlled reactions 236

11.2 Diffusion-controlled reactions

11.2.1 Experimental observations

The transition to diffusion control can be observed in the gas phase by studying the
kinetics in highly compressed gases (up to  = 1000 bar) or, even better, in supercritical
media. At  ≈ 1000 bar, the gas density reaches the density of the liquid phase.

I Smoluchowsky-Kramers limit: At high densities, the rate constant becomes

1
∝∝ (11.11)

I Figure 11.2: Diffusion control of a unimolecular reaction: At very high pressures, the
fragments recombine before they become separated by diffusion y the reaction becomes
diffusion controlled.

11.2.2 Derivation of k

In order to derive a theoretical expression for  , we consider the distribution of the


molecules  around a molecule  in the center of a sphere as shown in Fig. 11.3.
11.2 Diffusion-controlled reactions 237

I Figure 11.3: Radial distribution of reactants in a diffusion-controlled reaction.

(1) We start by considering the boundary conditions for the concentration [] (cf.
Fig. 11.3):

• At a distance  → ∞, the concentration of  is identical to the mean


concentration in the solution, i.e.

 → ∞ y [] = []=∞ = [] (11.12)

• At the distance  =   for a diffusion controlled reaction, the concentra-


tion of  vanishes because of its fast reaction with , i.e.

 → 0 y []=0 = 0 (11.13)

 []
(2) The result is a concentration gradient around , which gives rise to a net

− 
flux  by diffusion as described by Fick’s first law:
 []
 =  ×  × (11.14)

• We define this flux, measured in units of molecules/s, in the direction from
 = ∞ (high concentration) to  =  (low concentration); thus there is
no − sign in the usually written form of Fick’s law.
•  is the mean diffusion coefficient, measured in m2  s,

 =  +  (11.15)
11.2 Diffusion-controlled reactions 238

(3) With  = 42 as the surface of the sphere with radius  around , we obtain
 []
 =  × 42 × (11.16)

(4) To determine the concentration profile [] , we integrate over  [] :
Z [] Z ∞

 [] = 2
 (11.17)
[]  4 
y ¯∞
 ¯
[]|∞ =− ¯ (11.18)

4  ¯
y

[] − [] = −0 + (11.19)
4 
(5)  is found from the boundary conditions that [] = 0 at  =  :
 = 4  × [] (11.20)

(6) The concentration gradient of [] is obtained by inserting the above result for 
into Eq. 11.19:
 []
[] − [] = (11.21)

y ³  ´
[] = [] 1 − (11.22)

This expression is plotted in Fig. 11.3.
(7) To determine  , we write the rate of the diffusion controlled reaction as
 [ ]
=  [] (11.23)

Inserting the expression for  from Eq. 11.20, we obtain
 [ ]
= 4  × [] [] (11.24)

The diffusion-limited rate constant is thus (in molecular units)
 = 4  (11.25)

(8) Nernst-Einstein relation between diffusion constant and viscosity for spherical par-
ticles with radius  :
 
= (11.26)
6
y
 
 ∝ (11.27)

(9) Keep in mind that diffusion is an activated process:


 = 0 −  (11.28)
Typical value for H2 O:  ≈ 15 kJ mol−1 .
11.2 Diffusion-controlled reactions 239

I Typical values:
 ≈ 2 × 10−5 cm2 s−1 (11.29)
 ≈ 5 × 10−8 cm (11.30)
y Typical magnitude of  :
 ≈ 8 × 109 l mol−1 s−1 (11.31)

I Table 11.1: Diffusion constants of ions in aqueous solution at  = 298 K.

ion  ( cm2 s−1 ) ion  ( cm2 s−1 )


H+ 91 × 10−5 OH− 52 × 10−5
Li+ 10 × 10−5 Cl− 20 × 10−5
Na+ 13 × 10−5 Br− 21 × 10−5

I Examples of diffusion controlled reaction in solution:


• radical recombination reactions, e.g.
 +  → 2 :  = 13 × 1010 l mol−1 s−1 , (11.32)
• electronic deactivation by quenching with third partner [M] in solution,
• H+ transfer reactions
 + +  − → 2  :  = 14 × 1011 l mol−1 s−1 . (11.33)
This reaction is ten times faster than a typical diffusion controlled reaction, be-
cause H+ transfer is a concerted process (Grotthus mechanism, see Fig. 11.4)!

I Figure 11.4: Structures involved in fast H+ transfer processes in liquid water.


11.3 Activation controlled reactions 240

11.3 Activation controlled reactions

The rate constant for an activation controlled reaction is

 =  {} (11.34)

I Estimation of K{} : We can estimate the equilibrium constant {} using simple
arguments based on the coordination number. Accordingly, the concentration of the
encounter complex is

[{}] = [] × (probability of finding  next to ) (11.35)


 []
= [] × (11.36)
[]
where [] is the solvent concentration and  is the coordination number. y
[{}] 
{} = = (11.37)
[] [] []

For H2 O at  = 298 K, [] = 555 mol l−1 . Thus, taking for example  = 8, we find
that
{} = 014 l mol−1 (11.38)

I Application of thermodynamic TST: TST can be applied in two ways:

(1) Considering the overall reaction, we can write


  ¡ ¢
= exp −∆‡  (11.39)

where ∆‡ is the overall Gibbs free enthalphy for the reaction.
(2) We can separate the effects from  and {} by writing
³ ´
ª
{} = exp −∆{}  (11.40)

and
  ¡ ¢
 =  exp −∆+  (11.41)

∆ª{} is the Gibbs free enthalpy change for the formation of the encounter
complex, and ∆+ is the Gibbs free enthalpy of activation from the encounter
complex. Thus,
  ³ ³ ´ ´
= exp − ∆ª {} + ∆ +
 (11.42)

I Pressure dependence of the rate constant: ⇒ See section 7.3.5.


11.4 Electron transfer reactions (Marcus theory) 241

11.4 Electron transfer reactions (Marcus theory)

Electron transfer is what happens in oxidation-reduction reactions. Hence, electron


transfer reactions are extremely important. They are strongly affected by the solvent
because the change of the charge distribution results in a large reorganization energy.

I Figure 11.5:

I Marcus theory: For a simplified derivation of  ( )  we refer to Fig. 11.5.

(1) The energy of the donor and acceptor molecules will generally depend on all 3 −6
internal coordinates. We consider the dependence on an effective coordinate ,
which describes the minimum energy path from reactant to product (Fig. 11.5).
(2) The exergonicity of the ET reaction ∆ª is negative for an exergonic reaction
(convention; cf. Fig. 11.5).
(3) For the donor (D) and the acceptor (A), the Gibbs energy profiles  () and
 () shall be of parabolic form, which we may write as
 () = 2 (11.43)
and
 () = ( −  )2 + ∆ª . (11.44)
The origin  = 0 has been placed at the minimum of the donor; the acceptor has
its minimum at  =  .52

52
We assume that  ∝  2 , taking any proportionality constant into account by proper rescaling of the
abscissa.
11.4 Electron transfer reactions (Marcus theory) 242

(4) The ET rate constant is described by TST as


  ¡ ¢
 = exp −∆+  (11.45)

The Gibbs free energy of activation

∆+ = 
2
(11.46)

is the difference from the donor minimum to the TS at the donor/acceptor curve
crossing at point  .
(5) In order to relate ∆+ to the  potential curves, we define the “reorganiza-
tion energy”  as the energy that the acceptor would have at the equilibrium
geometry of the donor. This is the energy of the acceptor parabola  () at
 = 0 relative to the minimum of  () at  =  , i.e.

 = 2 (11.47)

(6) We now determine the Gibbs free energy at the intersection point  of the two
parabolas from the condition

 ( ) =  ( ) . (11.48)

Inserting  into Eqs. 11.43 and 11.44 yields, and using Eq. 11.47, yields
2
 = ( −  )2 + ∆ª (11.49)
2
=  − 2   + 2 + ∆ª (11.50)

y
2   = 2 + ∆ª (11.51)
y
2 + ∆ª
 = (11.52)
2 
(7) By inserting Eq. 11.52 for  into Eq. 11.46 and replacing 2 by  (Eq. 11.47),
we obtain the following expressions for ∆+ and  ( ):
2
(2 + ∆ª ) ( + ∆ª )2
∆+ = = (11.53)
4 2 4 
y à !
  ( + ∆ª )2
 ( ) =  exp − (11.54)
 4 

(8) Limiting cases: We can distinguish between

• −∆ª   : Regular region – the reaction becomes faster, the more


exergonic the reaction becomes (see Figs. 11.6A and B).
• −∆ª =  : turning point (Fig. 11.6C).
• −∆ª   : Inverted region – the reaction becomes slower, the more
exergonic the reaction becomes (Fig. 11.6D).
11.4 Electron transfer reactions (Marcus theory) 243

I Conclusion: Equation 11.54 predicts the ET rate constant to increase with increasing
exoergicity (−∆ª ) of the reaction. However, once −∆ª   , the rate constant
is predicted to decrease again (see Figs. 11.6D and 11.7). The prediction of this
unexpected inverted Marcus regime before any experimental data were available is
the hallmark of Marcus theory.

I Figure 11.6: Gibbs free energy parabolas for (A) an endergonic ET reaction, (B) a
“regular” ET reaction, (C) an ET reaction with zero activation energy ( = −∆ª ),
and (D) an ET reaction in the Marcus inverted region.
11.4 Electron transfer reactions (Marcus theory) 244

I Figure 11.7: Intramolecular electron transfer from a biphenyl anion as donor (bottom
right) to a series of acceptors A connected to the donor via a rigid spacer as function
of −∆ª . The regular ET region can be seen on the left, the Marcus inverted region
on the right.
11.5 Reactions of ions in solutions 245

11.5 Reactions of ions in solutions

Reactions of ionic species in liquids are one exception to the rule that reactions in the
liquid phase usually have similar rates as in the gas phase (the other exception being
electron transfer reactions). The reason is that the charged species are very sensitive
to their environment, in particular solvation. Changes of the charge distribution are
accompanied by correspondingly large solvation shell rearrangements.

11.5.1 Effect of the ionic strength of the solution

The main effect arises from the stabilization of shielding every ion in solution by the op-
positely charged ”ionic atmosphere” or “ion cloud” (⇒ Debye-Hückel theory, Appendix
??).

h i
I

Equilibrium constant for AB :
£ ¤
‡ ‡  ‡  ‡
 = = (11.55)
      [] []

I Activity coefficient from Debye-Hückel theory:

log   = − 2 12 (11.56)

with (for water at  = 298 K, according to Debye-Hückel theory)

 = 0509 l12 mol−12 (11.57)

I Ionic strength:
1X
 =  2 (11.58)
2 

11.5.2 Kinetic salt effect

I TST rate constant:


 [ ] £ ¤  
=  ‡  ‡ =  ‡ ‡   ‡ [] [] (11.59)
  

Defining 0 as the rate constant for the ideal solution, where all   = 1, this becomes
  
 = 0 (11.60)
 ‡
11.5 Reactions of ions in solutions 246

I Dependence of the rate constant on the ionic strength:


³ ´
‡ 2
log  = log 0 −  2 + 2 −  12 (11.61)


with  =  +  .

Inserting  =  +  , we obtain
¡ ¢
log  = log 0 −  2 + 2 − ( +  ) 2 12 (11.62)

y
 12
log = +2    (11.63)
0

I Conclusions:

• If  and  have the same sign, then  will increase with increasing , because
the repulsion is reduced by the shielding effect of the ion cloud.
• If  and  have opposite sign, then  will decrease with increasing , because
the attraction is reduced by the shielding effect of the ion cloud.

I Figure 11.8: Kinetic salt effect for reactions of ions in solution.


11.5 Reactions of ions in solutions 247

11.6 References

Eigen 1954a M. Eigen, Disc. Faraday Soc. 17, 194 (1954).


Eigen 1954b M. Eigen Z. Physik. Chem. NF 1, 176 (1954).
11.6 References 248

12. Photochemical kinetics

12.1 Radiative processes in a two-level system; Einstein coefficients

I Three fundamental radiative processes (Einstein):

• absorption,
• spontaneous emission,
• stimulated emission:

I Rate equation using Einstein coefficients:


2 1
=− = 12  () 1 − 21  () 2 − 21 2 (12.1)
 
with
8 3
21 = 21 (12.2)
3
and (for a true two-level system, with 1 = 2 )

12 = 21 (12.3)

I Figure 12.1: Absorption, stimulated emission, and spontaneous emission.


12.2 Light amplification by stimulated emission of radiation (LASER) 249

12.2 Light amplification by stimulated emission of radiation (LASER)

I Requirements for a laser: To build a laser, we want stimulated emission in Eq. 12.1
to predominate. This requires a population inversion. However, a population inversion
cannot be achieved in a two-level system in thermal equilibrium. Even at  → ∞, we
can only achieve equal populations. A working laser needs, at minimum, three states
(cf. 12.2).

I Components for a laser:

• optical medium with population inversion,


• pump source to achieve population inversion,
• optical resonator for feedback.

I Figure 12.2: Three-level and four-level laser.


12.2 Light amplification by stimulated emission of radiation (LASER) 250

I Figure 12.3: Light amplification by stimulated emission of radiation (LASER).


12.3 Fluorescence quenching (Stern-Volmer equation) 251

12.3 Fluorescence quenching (Stern-Volmer equation)

I Kinetic model:
A +  → A∗  1
A∗ → A +   (12.4)

A +Q →A +Q 

If any other radiationless processes can be neglected, we find

 1 [A]
[A∗ ]ss = (12.5)
 +  [Q]

I Fluorescence intensity I0 in the absence of the quencher Q:

0 ∝  [A∗ ]ss =  1 [A] (12.6)

I Fluorescence intensity I in the presence of the quencher Q:

 1 [A]
 ∝  [A∗ ]ss =  (12.7)
 +  [Q]

I Stern-Volmer equation:

0  +  [Q]  +  [Q]


=  1 [A] × = (12.8)
   1 [A] 
y
0
= 1 +   0 [Q]  (12.9)


I Radiative lifetime:
1
0 = (12.10)


I Figure 12.4: A plot of 0  vs. [Q] should give a straight line with intercept 1 and
slope   0 [Q]. Such a plot is called a Stern-Volmer plot.
12.4 The radiative lifetime  0 and the fluorescence quantum yield Φ 252

12.4 The radiative lifetime τ 0 and the fluorescence quantum yield Φ

I Determination of the radiative lifetime τ 0 : Assuming that there are no competing


nonradiative processes, we can determine the radiative lifetime  0 by

(1) time-resolved measurements of the (exponential) fluorescence decay   () after


pulsed excitation according to

2 () 20 = exp(−21 ) = exp(− 0 )  (12.11)

(2) high-resolution measurement of the natural (Lorentzian) linewidth according to

∆ ∆ = ~ (12.12)

y
1
0 = (12.13)
2 ∆ 0

(3) evaluation from the molar absorption coefficient (Lambert-Beer) using the rela-
tionship between the Einstein coefficients 21 and 21 ,53

8 3
21 = 21 (12.14)
3

(4) calculation from first principles of the value of


¯ ¯2
¯ ¯ ∝ 21 ∝ 21 ∝ 12 ∝ . (12.15)

I Fluorescence quantum yield: Apart from quenching, the fluorescence is reduced by


other radiationless processes (IC, ISC, R, . . . ). The experimentally measured fluores-
cence lifetime  is therefore smaller (often much smaller) than the radiative lifetime  0 .
The ratio is the fluorescence quantum yield:
 
Φ = = (12.16)
 +  +  +  +  [Q] +     0

53
The actual procedure, which involves integration over the absorption band, corrects for the influence
of the refraction index of the solvent, etc., is called the Strickler-Berg analysis.
12.5 Radiationless processes in photoexcited molecules 253

12.5 Radiationless processes in photoexcited molecules

I Figure 12.5: Jablonski Diagram.

I Primary photochemical processes:


¯ ¯2
• Absorption of a photon (depends on transition dipole moment ¯ ¯ ),
• Intramolecular processes:

— fluorescence (≡ spontaneous emission of a photon, FLUO),


— stimulated emission (SE), (⇒ Einstein coefficients)
— intramolecular vibrational (energy) redistribution (IVR),
— internal conversion (IC),
— chemical reaction in the excited state (RXN),
— intersystem crossing (ISC),
— phosphorescence (PHOS).

• Intermolecular (i.e. collisional) processes:

— electronic deactivation (quenching, Q),


∗ may affect fluorescence as well as phosphorescence,
— vibrational energy transfer (VET),
— rotational energy transfer (RET),
— collision-induced intersystem crossing (CIISC).
12.5 Radiationless processes in photoexcited molecules 254

12.5.1 The Born-Oppenheimer approximation and its breakdown

a) The full Hamiltonian


We consider a non-moving, non-rotating molecule in the laboratory framework. The
molecule is described by the Schrödinger equation (SE)

̂ (r R) =  (r R) (12.17)

or ³ ´
̂ −   (r R) = 0 (12.18)

with  being the energy eigenvalue associated with the wavefunction  (r R) as function
of the electronic (e) coordinates r and the nuclear (N) coordinates R.
The Hamilton operator ̂ appearing in the SE can be written as

̂ = ̂ + ̂ = ̂ (r) + ̂ (R) +  (r R) (12.19)

where

~2 X  
̂ = − ∇2 (12.20)
2 =1 

~2 X
̂ = − ∇2 (12.21)
2 =1 

and
 (r R) =  (R) +  (r R) +  (r) (12.22)
with

2 X X  0
  = + (12.23)
40 0 0 =1 0
 X
2 X


 = − (12.24)
40 =1 =1 

2 X X

1
 = + (12.25)
40 0 0 =1 0

Electronic spin, spin-orbit interaction, and nuclear spins are neglected here. Further,
we have left aside the so-called electronic mass and electronic nuclear cross polarization
terms, which appear in the case of a moving molecule as result of the separation of the
center-of-mass.

b) Adiabatic separation of nuclear and electronic motion


We now want to separate the slow motion of the nuclei from the the fast motion of
the electrons. This separation can be made because the fast electrons can be assumed
12.5 Radiationless processes in photoexcited molecules 255

to very rapidly follow the slow nuclei. The kinetic energy of the nuclei is assumed to be
small compared to the electronic energy.
Thus, for every nuclear configuration R, there is an electronic wavefunction el (r R)
belonging to the electronic state |i specified by quantum number . This el (r R)
depends on the nuclear coordinates R, but only very little on the nuclear velocities.
We say that the electrons adiabatically follow the periodic vibrational motion of the
nuclei. This approximation is therefore called adiabatic approximation.

I Ansatz for Ĥ:


̂ = ̂0 + ̂ 0 (12.26)
with

~2 X

̂0 = ̂ +  (r R) = − ∇2 +  (r R) (12.27)
2 =1 

~2 X
0
̂ = ̂ = − ∇2 (12.28)
2 =1

I The zero-order Hamiltonian Ĥ0 :

• The zero-order Hamiltonian ̂0 depends only on the electron kinetic energy op-
erator  and the potential energy  and describes the rigid molecule (all R are
fixed!) via the unperturbed SE

̂0 el (r; R) =  (0) (R) el (r; R) (12.29)

• The zero-order electronic wave functions el (r; R) describe the electrons and
¯ ¯2
their distribution ¯el (r; R)¯ for the case of fixed nuclei in the electronic state
|i with the electronic energy el .
• The el (r; R) depend on R only as a parameter, not as a variable (we do not
differentiate or integrate with respect to the  in the zero-order SE).
• The el (r; R) form a complete orthonormal system.
• The eigenvalues  (0) (R) of the electronic SE are what we have come to know as
the molecular potential energy functions (or hypersurfaces)  (R) of the molecule
with all nuclei at rest.

I The perturbation Ĥ 0 :

• The perturbation ̂ 0 is the nuclear kinetic energy operator ̂ .


12.5 Radiationless processes in photoexcited molecules 256

c) Solution of the complete SE

• To solve the complete SE


³ ´
̂0 + ̂ 0 −   (r R) = 0 (12.30)

we use the ansatz X


 (r R) =  (R) el (r; R) (12.31)

with the expansion coefficients  =  (R) depending only on R but not on r.


• Inserting the ansatz 12.31 into the complete SE,
³ ´X
0 0
̂ + ̂ −   (R) el (r; R) = 0 (12.32)

and doing some algebraic manipulation, we arrive at a


• system of coupled equations for the electronic wavefunctions el (r; R) for the
electronic state el (r; R) and the nuclear wavefunctions  (R):

̂0 el (r; R) = (0) (R) el (r; R) (12.33)

and
X ¡ ¢
̂ 0  (R) +  (R)  (R) =  − (0) (R)  (R) (12.34)

with the coupling coefficients


Z
 (R) = ∗ (r; R) ̂ 0  (r; R) r
r
Z X 1  
−~ 2
∗ (r; R)  (r; R) r (12.35)
   
r

• We interprete these Eqs. as follows:


(0)
(1)  (R) can be interpreted as the zero-order potential functions of the th
zero-order electronic state that is given by the solution of the electronic SE
(Eq. 12.29) for fixed nuclei R.
(2) Without the  (R), Eq. 12.34 describes the kinetic energy of the nuclei
(0)
in the potential  (R).54 Eq. 12.34 is thus simply the vibrational SE for
(0)
the electronic potential  (R).
(3) The coefficients  (R) defined by Eq. 12.35 are coupling terms that con-
nect (i.e. mix) the electronic states  and  . They describe how the
different electronic states  and  are coupled by the nuclear motion (the
two terms on the RHS resulting from the nuclear kinetic energy operator
̂ ; see Eq. 12.35).

54
We usually denote these potential energy functions as  (R).
12.5 Radiationless processes in photoexcited molecules 257

d) Born-Oppenheimer approximation
The Born-Oppenheimer (BO) approximation completely neglects the coupling coeffi-
cients  . The complete SE is therefore reduced to two uncoupled equations, which
we denote as electronic SE
(0)
̂0 el (r) =  (R) el (r) (12.36)
and nuclear (vibrational) SE
h i
(0)
̂ +   (R) =   (R) (12.37)

which describe, respectively, the electronic wavefunction el for fixed nuclear coordinates
R in the electronic state el and the set of nuclear wavefunctions  (R) for the energy
state  of the nuclei in the electronic state el .

e) Breakdown of the Born-Oppenheimer approximnation


When two (or more) electronic states el (R), el (R) become nearly isoenergetic or,
at a crossing, even degenerate, the coupling matrix element  (R) may no longer be
neglected, because through the  (R), the nuclear motion leads to a mixing between
the different electronic BO-states.

I This breakdown of the BO approximnation leads to non-adiabatic coupling of


electronic states:

• For a diatomic molecule, the coupling leads to an avoided crossing of the two
electronic potential curves.
• For polyatomic molecules, the coupling can lead to a conical intersection (CI)
between the two electronic potential energy hypersurfaces.

As will be outlined in the following, the existence of conical intersections between dif-
ferent excited electronic potential energy hypersurfaces and between an excited and
the ground electronic potential energy hypersurface has dramatic consequences for the
dynamics of photochemical reactions.55

12.5.2 Avoided crossings of two potential curves of a diatomic molecule

I Illustration of an avoided crossing using two diabatic model potential functions


+ coupling term:
1 () = 1 [1 − exp [− 1 ( − 1 )]]2 (12.38)
2 () = 2 +  exp [− 2 ( − 2 )] (12.39)
12 () = 1 [1 + tanh [−2 ( − 12 )]] (12.40)
12 () is just a convenient model function with the physically correct behavior
12 () → 0 for  → ∞.

55
The recognition since about the years 2000 - 2005 that conical intersections in polyatomic molecules
are the rule rather than the exception has indeed revolutionized our understanding of photochemical
reactions.
12.5 Radiationless processes in photoexcited molecules 258

I Figure 12.6: Diabatic model potential curves for a diatomic molecule.

I Figure 12.7: Avoided crossing of two potential curves of a diatomic molecule.


12.5 Radiationless processes in photoexcited molecules 259

I Figure 12.8: Conical intersection of potential energy surfaces of a polyatomic molecule.

12.5.3 Quantum mechanical treatment of a coupled 2×2 system

We can easily verify the “avoided crossing” of two coupled potential energy curves as
follows:

I Schrödinger equation:
( − ) |i = 0 (12.41)

I Hamilton-Operator:
 =  (0) +  (1) (12.42)

I Ansatz:
|i = 1 |1 i + 2 |2 i (12.43)
(0)
with orthonormal basis vectors that are eigenvectors of  , i.e.
½
­ ® 1 for  = 
 | =   = (12.44)
0 for  6= 

and the zero-order energies


(0)
 (0) |1 i = 1 |1 i (12.45)
(0)
 (0) |2 i = 2 |2 i (12.46)
12.5 Radiationless processes in photoexcited molecules 260

I Solution of the Schrödinger equation: Insertion of the ansatz (Eq. 12.43) into the
SE and multiplication from the left by either h1 | or by h2 | gives the two equations:

1 h1 | ( − ) |1 i + 2 h1 | ( − ) |2 i = 0 (12.47)


1 h2 | ( − ) |1 i + 2 h2 | ( − ) |2 i = 0 (12.48)

This is a system of coupled linear equations (“secular equations”):

1 (11 − ) + 2 12 = 0 (12.49)


1 21 + 2 (22 − ) = 0 (12.50)

with the “matrix elements” (integrals over all r resp. R):

 = h |  | i (12.51)


= h |  (0) +  (1) | i (12.52)
= h |  (0) | i + h |  (1) | i (12.53)

Specifically, we have
(0)
11 = h1 |  (0) |1 i = 1 (12.54)
(0) (0)
22 = h2 |  |2 i = 2 (12.55)

and
12 = h1 |  (1) |2 i = h2 |  (1) |1 i = 21 (12.56)

Therefore, we obtain the secular equations as follows:

I Secular equations:
³ ´
(0)
1 1 −  + 2 12 = 0 (12.57)
³ ´
(0)
1 21 + 2 2 −  = 0 (12.58)

I Secular determinant: A non-trivial solution for the secular equations requires that
¯ ¯
¯ (0) ¯
¯ 1 −  12 ¯
¯ ¯=0 (12.59)
¯ 21 2(0) −  ¯

I Eigenvalues:
³ ´³ ´
(0) (0)
0 = 1 −  2 −  − |12 |2 (12.60)
³ ´
(0) (0) (0) (0)
= 1 2 −  1 + 2 + 2 − |12 |2 (12.61)

with
|12 |2 = 12 21 (12.62)
(1)
because  is Hermitian.
12.5 Radiationless processes in photoexcited molecules 261

With some rewriting56 , we obtain

1 ³ (0) ´ 1 r³ ´2
(0) (0) (0) (0) (0)
± = 1 + 2 ± 1 + 2 − 41 2 + 4 |12 |2 (12.63)
2 2
y r³
1 ³ (0) (0)
´ 1
(0) (0)
´2
± = 1 + 2 ± 1 − 2 + 4 |12 |2 (12.64)
2 2

I Shorthand form: Using the abbreviations

± (R) = ± (12.65)
(0)
1 (R) = 1 (12.66)
(0)
2 (R) = 2 (12.67)
 2 (R) = |12 |2 (12.68)

1
Σ (R) = (1 (R) + 2 (R)) (12.69)
2
1
∆ (R) = (1 (R) − 2 (R)) (12.70)
2
we can rewrite the solution for the eigenvalues + and − in a compact form as
q
± (R) = Σ (R) ± (∆ (R))2 + ( (R))2 (12.71)

12.5.4 Radiationless processes in polyatomic molecules: Conical intersections

I “Traditional” view on the cis-trans isomerization of C2 H4 :

• Reaction coordinate can be approximated by torsion angle  (0 ≤  ≤ 180◦ ).


• Reaction proceeds via a perpendicular transition state ( = 90◦ ), in which the C
atoms are connected by a single bond.
• Correlation diagram shows that the  orbital of the trans-isomer transforms into
the  ∗ orbital of the cis-isomer and vice versa.
• Crossing of the two diabatic states corresponds to an avoided crossing of the
adiabatic states.

56

³ ´2 ³ ´2 ³ ´2
(0) (0) (0) (0) (0) (0) (0) (0) (0) (0)
1 + 2 − 41 2 = 1 + 2 + 21 2 − 41 2
³ ´2 ³ ´2 ³ ´2
(0) (0) (0) (0) (0) (0)
= 1 + 2 − 21 2 = 1 − 2
12.5 Radiationless processes in photoexcited molecules 262

• The reaction is thus thermally forbidden because of the high energy barrier at the
avoided crossing between the trans and cis isomers on the left and right.
• The reaction is photochemically allowed because of the barrierless adiabatic cor-
relation between the ∗ states on the left and right.

I Figure 12.9: Cis-trans isomerization of C2 H4 : HMO picture with transformation of 


and  ∗ orbitals and conical intersection.

I Figure 12.10: Internal coordinates relevant for the cis-trans isomerization of C2 H4 .


The out-of-plane bending angle  plays a role as coupling mode leading to a CI; the
angles  and  play a role in the isomerization to H3 CCH (also by a CI).
12.5 Radiationless processes in photoexcited molecules 263

I Figure 12.11: Diabatic model PES’s for C2 H4 .

I Figure 12.12: Adiabatic model PES’s for C2 H4 with the conical intersection at  = 90◦ .
12.5 Radiationless processes in photoexcited molecules 264

I Figure 12.13: Wave packet motion in a complex photo-induced reaction.

12.5.5 Femtosecond time-resolved experiments

I Figure 12.14:
12.5 Radiationless processes in photoexcited molecules 265

I Figure 12.15:

I Figure 12.16:
12.5 Radiationless processes in photoexcited molecules 266

I Figure 12.17:

I Figure 12.18:
12.5 Radiationless processes in photoexcited molecules 267

I Figure 12.19:
12.5 Radiationless processes in photoexcited molecules 268

13. Atmospheric chemistry*

(skipped for lack of time in the semester)


13 Atmospheric chemistry* 269

14. Combustion chemistry*

(skipped for lack of time in the semester)


14 Combustion chemistry* 270

15. Astrochemistry*

(skipped for lack of time in the semester)


15 Astrochemistry* 271

16. Energy transfer processes*

(See lecture in M.Sc. course.)


16 Energy transfer processes* 272

17. Intramolecular dynamics & vibrational spectroscopy*

(See lecture in M.Sc. course.)


17 Intramolecular dynamics & vibrational spectroscopy* 273

18. Reaction dynamics*

(See lecture in M.Sc. course.)


18 Reaction dynamics* 274

19. Advanced photochemical dynamics*

(See lecture in M.Sc. course.)


19 Advanced photochemical dynamics* 275

20. Electrochemical kinetics*

(skipped for lack of time in the semester)


Appendix B 276

Appendix A: Useful physicochemical constants

I Table A.1: List of useful physical constants (Cohen 1986).

physical constant symbol value


Loschmidt (or Avogadro) number   = 60221367 × 1023 mol−1
gas constant   = 8314511 J mol−1 K−1
Boltzmann constant   = 1380658 × 10−23 J K−1
electron charge   = 160217733 × 10−19 C
Faraday constant   = 96485309 C mol−1
speed of light   = 299792458 × 108 m s−1
Planck’s constant   = 66260755 × 10−34 J s
~ ~ = 10547266 × 10−34 J s
electron mass   = 91093897 × 10−31 kg
proton mass   = 16726231 × 10−27 kg
neutron mass   = 16749286 × 10−27 kg
atomic mass unit   = 16605402 × 10−27 kg
Rydberg constant e
 e = 10973731534 × 107 m−1

Bohr radius 0 0 = 529177249 × 10−11 m

electric field constant 0 0 = 1(0 2 )


0 = 8854187817 × 10−12 F m−1
magnetic field constant 0 0 = 4 × 10−7 N A−2
0 = 125664 × 10−6 N A−2

I References:

Cohen 1986 E. R. Cohen and B. N. Taylor, The 1986 Adjustment of the Fundamental
Constants, CODATA Bull. 63, 1 (1986). An updated list is contained in every
August issue of Physics Today.
Appendix B 277

Appendix B: The Marquardt-Levenberg non-linear least-squares


fitting algorithm

With the omnipresence of the computer, the Marquardt-Levenberg algorithm (Bevington


1992, Press 1992) has become an extremely important data analysis tool. It is generally
the method of choice in cases where it is not possible to make simple linear plots to
determine rate constants.

I Problem: Experimentally, we measure the values of some quantity  at the points  ,


 ,   . Suppose we take  measurements. Now suppose we would like to describe our
 data points  by some model. Towards these ends, we take a physically reasonable
model function
 =  (     ; (1       )) (B.1)
which we would like to fit to our data in such a way that  (  ) describes the data
points  in “the best possible way”. The model function  depends directly on the
variables     , and it also depends parametrically on  nonlinear parameters  . In
kinetics, for example,  is usually some concentration, e.g.  , which depends on time
 i.e.  =  (). The parameters  may be the rate constants  (or decay times  )
and the initial concentration 0 . Our job is to adjust and optimize these  parameters
by somehow “fitting” them. The method of choice is, of course, “least-squares fitting”:
We start by defining the so-called figure-of-merit or goodness-of-fit function57
" #
X
[ −  (      )]2
2 = 2
(B.2)
=1
 

Here, the  and the independent variables  ,      are the measured quantities, the
  are the experimental uncertainties of the  , and  (      ) is the calculated value
of the model fit function  at the point {      } (the parameters 1       are
omitted here in  ). What we have to do is to adjust the  parameters  in the model
function such that 2 reaches a minimum, i.e. we have to search the -dimensional
parameter space for the (global) minimum of 2 .

I Solution: The Marquardt-Levenberg algorithm is a method to determine this mini-


mum of 2 as function of the fit parameters  in the model fit function  .
The minimum of 2 as function of the fit parameters a is defined by the conditions
" #
 X [ −  (      )]2

 2
= =0 (B.3)
    =1 2
X ∙ ¸
[ −  (   )]   (      )
= −2 (B.4)
=1
2  

Thus, taking the partial derivatives of 2 with respect to each of the  parameters
 , we obtain a set of  coupled equations in the  unknown parameters  . These
equations are, in general, nonlinear in the  .

57
Deutsch: Fehlerquadratsumme.
Appendix B 278

The Marquardt-Levenberg algorithm gives a recipy for finding the minimum of 2 using
a combination of the method of steepest descent (following the gradient of the hyper-
surface defined by 2 as a function of the parameters 1       and, near the minimum
of 2 , a parabolic Taylor series expansion of the fitting function  around the minimum).
At the end of a fit, we shall usually report the  (± 2 standard deviations), the
standard deviation of the  ’s, and the so-called reduced-2 , which is 2 divided by the
number of degrees of freedom,  =  − , i.e.
" #
1 X
[ −  (      )]2
2 = (B.5)
 −  =1 2

I Example 1: Exponential decay. Suppose we measure the time dependent concen-


tration  () of a molecule in some quantum state, which is populated at some time 0
by a laser pulse and then decays monoexponentially. As a model fit function, we use
the expression
 () = 1 exp [−2 ( − 3 )] + 4 + 5  (B.6)
The parameters we are interested in are 2 (the decay rate coefficient) and 1 (the
initial amplitude or initial concentration). The parameter 3 just describes the trigger
delay time (= 0 ). We have also added the parameters 4 and 5 to describe a constant
background term (due to some offset of our experimental signal) and a linear drift in
this background term (perhaps due to some experimental instability, such as a gradual
temperature increase in the lab). Our figure of merit is then
" #
X 
[ −  ()]2
2 = (B.7)
=1
 2

The  (in general nonlinear) conditions for the minimum of 2 , from which we determine
the parameters  , are
" #  ∙ ¸
 X [ −  ()]2 X

 2 [ −  ()]   ()
= = −2 =0 (B.8)
 1  1 =1 2 =1
 
2   1

 2
=  (B.9)
 2
 (B.10)

Figure B.1 below shows a simple two-parameter fit (parameters 1 and 1) to such a
signal.
Appendix B 279

I Figure B.1: Exponential decay curve of a particular vibration-rotation state of the


CH3 O radical resulting from the unimolecular dissociation reaction of the radical ac-
cording to CH3 O → H + H2 CO (Dertinger 1995). The small box is the output box
from a fit using the ORIGIN program.

I Example 2: Multiexponential decay. In femtosecond spectroscopy, we often ob-


serve ultrafast multiexponential decays of laser-excited molecules. The laser prepares
an excited wavepacket, which usually does not decay single-exponentially. Further, we
need to take into account the final duration of the pump laser pulse (by deconvolution,
or by forward convolution).

• Model function to be fitted to measured data:


X
 () =  exp (−  ) (B.11)

with adjustable parameters  and   .


• Instrument response function (IRF): Often represented by a Gaussian centered at
time 0 Ã !
1 ( − 0 )2
 () = √ exp − (B.12)
 IRF 2 2 2IRF
with width parameter  IRF related to the full width at half maximum (FWHM)
of the IRF by √
FWHM = 8 ln 2 ≈ 2355IRF (B.13)
Appendix C 280

• Convolution of the molecular intensity  () and  () gives the signal function
Z+∞ Z+∞
0 0 0
 () =  ( )  ( −  )  =  ( − 0 )  (0 ) 0 =  () ⊗  () (B.14)
−∞ −∞

where ⊗ denotes the convolution.


• Resulting model function to be fitted to the data:
∙ 2 ¸∙ µ ¶¸
1X 1  IRF ( − 0 ) ( − 0 )   −  2IRF
 () =  exp − 1 + erf √ +
2  2  2  2 IRF  
(B.15)
where erf () is the error function and  is a simple constant background term (can
be replaced by background + drift  + ).

I Figure B.2: Excited-state relaxation dynamics of the adenine dinucleotide after UV


photoexcitation.

I References:
Bevington 1992 P. R. Bevington, D. K. Robinson, Data Reduction and Error Analysis
for the Physical Sciences, McGraw-Hill, Boston, 1992.
Dertinger 1995 S. Dertinger, A. Geers, J. Kappert, F. Temps, J. W. Wiebrecht,
Rotation-Vibration State Resolved Unimolecular Dynamics of Highly Vibrationally
Excited CH3 O (2 ): III. State Specific Dissociation Rates from Spectroscopic Line
Profiles and Time Resolved Measurements, Faraday Discuss. Roy. Soc. 102, 31
(1995).
Press 1992 W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P. Flannery, Numerical
Recipes in Fortran, Cambridge University Press, Cambridge, 1992. Versions are
also available for C and Pascal.
Appendix C 281

Appendix C: Solution of inhomogeneous differential equations

C.1 General method

An inhomogeneous DE is a DE of the type

 0 +  () ×  =  () (C.1)

In order to solve Eq. C.1, we first find a solution  of the homogeneous DE

 0 +  () ×  = 0 (C.2)

and then determine a particular solution  of the inhomogeneous DE:

• Solution of the homogeneous DE by separation of variables:

 0 +  () ×  = 0 (C.3)

 0 = − () ×  (C.4)
 ln 
= − ()   (C.5)

y 
 =  × −  () 
(C.6)

• Determination of a particular solution of the inhomogeneoues DE by variation of


constant:
 →  () (C.7)
 0 () =    (C.8)

Insertion of  () and  0 () into Eq. C.1 and integration gives

 =    (C.9)

The general solution of Eq. C.1 is then given by

 =  +  (C.10)
Appendix C 282

C.2 Application to two consecutive first-order reactions

Consider the reaction system


1
A −→ B (C.11)
2
B −→ C (C.12)

which is described by the following rate equations

 [A]
= −1 [A] (C.13)

 [B]
= +1 [A] − 2 [B] (C.14)

 [C]
= +2 [B] (C.15)


The solution for Eq. C.13 is


[A] = [A]0 −1  (C.16)

Thus we have to find the solution of the inhomogeneous DE for [B] (Eq. C.14)

 [B]
+ 2 [B] = 1 [A]0 −1  (C.17)


• Solution of the homogeneous DE by separation of variables:

 [B]
= −2 [B] (C.18)

y
[B] =  × −2  (C.19)

• Determination of a particular solution by variation of constant:

 =  () (C.20)

y
[B] =  () × −2  (C.21)
 [B]   ()
= × −2  −  () × 2 −2  (C.22)
 
 [B]
Insertion of [B] and into Eq. C.17 gives

  ()
× −2  −  () × 2 −2  +  () × 2 −2  = 1 [A]0 −1  (C.23)

which simplifies to
  ()
= 1 [A]0 (2 −1 ) (C.24)

This DE is easily integrated to obtain  ():
Appendix C 283

— If 1 6= 2 we obtain
1 [A]0 (2 −1 )
 () =  (C.25)
2 − 1
and thus

[B] =  () −2  (C.26)


1 [A]0 (2 −1 ) −2 
=   (C.27)
2 − 1
1 [A]0 −1 
=  (C.28)
2 − 1

— If 1 = 2 we obtain
  ()
= 1 [A]0 (2 −1 ) (C.29)

= 1 [A]0 (C.30)

y
 () = 1 [A]0  (C.31)
y

[B] =  () −2  (C.32)


= 1 [A]0  −2  (C.33)

• General solution:
[B] = [B] + [B] (C.34)
y ⎧
⎪ 1 [A]0 −1 
⎨  × −2  +  if 1 =
6 2
[B] = 2 − 1 (C.35)


 × −2  + 1 [A]0  −2  if 1 = 2

Initial value condition at  = 0:

[B ( = 0)] = 0 (C.36)

y ⎧
⎪  [A]
⎨ − 1 0 if 1 = 6 2
= 2 − 1 (C.37)


0 if 1 = 2

Final solutions for [B ()]:



⎪ 1 [A]0 ¡ −1  ¢
⎨  − −2  if 1 =6 2
[B ()] = 2 − 1 (C.38)


1 [A]0  −2  if 1 = 2
Appendix C 284

C.3 Application to parallel and consecutive first-order reactions

Consider the reaction system



1
A −→ P (C.39)

1 2 
A −→ B −→ P (C.40)

which is described by the following rate equations

 [A]
= − (1 + 1 ) [A] = −1 [A] (C.41)

 [B]
= +1 [A] − 2 [B] (C.42)

 [P]
= +1 [A] + 2 [B] (C.43)

with 1 = (1 + 1 ).58
The solution for Eq. C.41 is

[A] = [A]0 −(1 +1 )  = [A]0 −1  (C.44)

Thus we have to find the solution of the inhomogeneous DE for [B] (Eq. C.42)

 [B]
+ 2 [B] = +1 [A]0 −1  (C.45)

using the above standard procedure:

• Solution of the homogeneous DE by separation of variables:

 [B]
= −2 [B] (C.46)

y
[B] =  × −2  (C.47)

• Determination of a particular solution by variation of constant:

 =  () (C.48)

y
[B] =  () × −2  (C.49)
 [B]   ()
= × −2  −  () × 2 −2  (C.50)
 

58
An example is the radiationless deactivation of a  ∗ electronically excited molecule A directly to
the ground state P or via an intermediatate optically dark ∗ state B, which decays to the ground
state more slowly.
Appendix C 285

 [B]
Insertion of [B] and into Eq. C.45 gives

  ()
× −2  −  () 2 −2  + 2  () −2  = 1 [A]0 −1  (C.51)

The terms with ± () 2 −2  cancel, so that

  ()
× −2  = 1 [A]0 −1  (C.52)

which we rewrite in order to solve  () as

  ()
= 1 [A]0 −1  × +2  (C.53)

or
  ()
= 1 [A]0 (2 −1 )  (C.54)

This DE is easily integrated to obtain  ():

— If 2 6= 1 = (1 + 1 ) we obtain

1 [A]0 (2 −1 )


 () =  (C.55)
2 − 1
and thus

[B] =  () −2  (C.56)


1 [A]0 (2 −1 ) −2 
=   (C.57)
2 − 1
1 [A]0 −1 
=  (C.58)
2 − 1

— The case 2 = 1 does not interest us here, as the  ∗ state B should be a


longer-lived one.

• General solution for 2 6= 1 :

[B] = [B] + [B] (C.59)

y
1 [A]0 −1 
[B] =  × −2  +  (C.60)
2 − 1
Initial value condition at  = 0:

[B ( = 0)] = 0 (C.61)

y
1 [A]0
=− (C.62)
2 − 1
Appendix D 286

Final solutions for [B ()]:

1 [A]0 1 [A]0 −1 


[B ()] = − × −2  +  (C.63)
2 − 1 2 − 1
1 [A]0 −1  1 [A]0
=  − × −2  (C.64)
2 − 1 2 − 1
i.e.
1 [A]0 ¡ −2  ¢
[B ()] = ×  − −1  (C.65)
1 − 2

• In the time-resolved fluorescence measurements, we observe

 () ∝ [A ()] +  [B ()] (C.66)

with some unknown factor  accounting for the (lower) transition probability of
state B, i.e. the fluorescence-time profile  () is

1 [A]0 ¡ −2  ¢


 () = [A]0 −1  +  ×  − −1  (C.67)
1 − 2
1 [A]0 1 [A]0
= [A]0 −1  +  × −2  −  × −1 
1 − 2 1 − 2
(C.68)
µ ¶
1 [A]0 −2  1 [A]0
=  × + [A]0 −  × −1 
1 − 2 1 − 2
(C.69)
−2  −1 
= 2 ×  + 1 ×  (C.70)

This is the same result as would be obtained by a fit to a sum of two expo-
nentials with the rate constants 1 = 1 + 1 and 2 , i.e. time constants
 1 = (1 + 1 )−1 and  2 = 2 .
Appendix D 287

Appendix D: Matrix methods

This appendix has been taken from the “Introduction to Molecular Spectroscopy” script,
not all subsections apply to chemical kinetics.

D.1 Definition

I Matrices: A matrix is a rectangular,  ×  dimensional array of numbers  :


⎛ ⎞
11 12 13    1
⎜ 21 22 23    2 ⎟
A=⎜
⎝ ... .. .. . ⎟ (D.1)
. .    .. ⎠
1 2 3    

The  are called the matrix elements.


We will need to consider only  ×  dimensional square matrices, which we can think
of as a collection of column vectors a :
⎛ ⎞
11 12    1
⎜ 21 22    2 ⎟
A=⎜ ⎝ ... .. . ⎟ (D.2)
.    .. ⎠
1 2    

I Notation: Matrices will be denoted by bold symbols:

A (D.3)

I Matrix manipulation by computer programs: Matrix manipulations are carried


out efficiently using computer programs (see, e.g. Press1992) or with symbolic algebra
programs.59

59
The most commmon symbolic math programs are MathCad, MuPad, Maple, and Mathematica.
MathCad and Mathematica are available in the PC lab.
Appendix D 288

D.2 Special matrices and matrix operations

I Identity matrix: I = 1

⎛ ⎞
1 0  0
⎜0 1  0 ⎟
I=1=⎜
⎝ ... .. .⎟ (D.4)
.    .. ⎠
0 0  1

We frequently use the Kronecker :

 =   (D.5)

where ½
1 for  = 
  = (D.6)
0 for  =
6 

I Diagonal matrices: ⎛ ⎞
11     0
⎜ 0 22  0 ⎟
⎜ . .. .. .. ⎟ (D.7)
⎝ .. . . . ⎠
0 0    

I Block diagonal matrices:


⎛ ⎞
11 12  0
⎜ 21 22  0 ⎟
⎜ . .. .. .. ⎟ (D.8)
⎝ .. . . . ⎠
0 0    

I Complex conjugate (c.c.) of a matrix: A∗


To obtain the complex conjugate A∗ of the matrix A, change  to −:

(∗ ) = ( )∗ (D.9)

I Transpose of a matrix: 
To obtain the transpose A of the matrix A, interchange rows and colums:
¡ ¢
  = ( ) (D.10)
Appendix D 289

I Hermitian conjugate (“transpose conjugate”) of a matrix: A†


To obtain the Hermitian conjugate (or “transpose conjugate” or “adjoint”) A† of the
matrix A, take the complex conjugate and transpose:
¡ ¢ ¡ ¢∗
† = ∗ =  (D.11)

where ¡ †¢
  = ( )∗ (D.12)

I Hermitian matrices: A matrix A is Hermitian, if

A† = (A∗ ) = A (D.13)

i.e. ¡ †¢
  = ( )∗ =  (D.14)

I Inverse of a matrix: A−1


A matrix A−1 is the inverse of the original matrix A, if

AA−1 = A−1 A = I (D.15)

I Orthogonal matrices: A matrix A is orthogonal, if the inverse of A equals its trans-


pose:
A−1 = A (D.16)

I Unitary matrices: A matrix U is unitary, if the inverse of U equals its transpose


conjugate:
U−1 = U† = (U∗ ) (D.17)
y
U† U = I (D.18)

I Trace of a matrix:
X

tr (A) =  (D.19)
=1

I Matrix addition:
C=A+B (D.20)
with
 =  +  (D.21)
Appendix D 290

I Multiplication by a scalar:
C = A (D.22)
with
 =  (D.23)

I Matrix multiplication:
C = A · B = AB (D.24)
with
X

 =   (D.25)
=1

Example:
µ ¶µ ¶ µ ¶µ ¶
1 2 5 6 1×5+2×7 1×6+2×8 19 22
= (D.26)
3 4 7 8 3×5+4×7 3×6+4×8 43 50

Notes:

(1) Matrix multiplication is defined only if the number of columns of the first matrix
is identical to the number of rows of the second matrix.
(2)
A B 6= B A (D.27)

I Singular matrices: A matrix A is called singular, if its determinant vanishes:

det (A) = |A| = 0 (D.28)


Appendix D 291

D.3 Determinants

I Determinant of a matrix: A determinant of a matrix is, in general, a number which


is evaluated according to the rule
X

det (A) = |A| =   (D.29)
=1

The  are called the co-factors which are obtained by omitting the ’th row and ’th
column (marked below in red) and multiplying by (−1)+ :
¯ ¯
¯ 11    1    1 ¯
¯ . .. .. ¯¯
¯ .
¯ .  .  . ¯
¯ ¯
 = (−1)+ ¯ 1         ¯ (D.30)
¯ . . . ¯
¯ ..    ..    .. ¯
¯ ¯
¯ 1         ¯

I Evaluation of simple determinants:


¯ ¯
¯ ¯
¯ ¯ (D.31)
¯   ¯ =  − 

¯ ¯
¯ 11 12 13 ¯
¯ ¯
¯ 21 22 23 ¯ = (D.32)
¯ ¯
¯ 31 32 33 ¯
= 11 22 33 + 12 23 31 + 13 21 32
− (13 22 31 + 11 23 32 + 12 21 33 )

where we “extended” the original determinant according to the scheme


¯ ¯
¯ 11 12 13 ¯ 11 12
¯ ¯
¯ 21 22 23 ¯ 21 22 (D.33)
¯ ¯
¯ 31 32 33 ¯ 31 32
Appendix D 292

D.4 Coordinate transformations

I Multiplication of a vector by a matrix: x0 = A x


The multiplication of an -dimensional column vector x by an  ×  matrix A gives a
new vector x0 with components corresponding to those of x in a transformed coordinate
system.

I Example: Multiplication of a 2-dim vector


µ ¶
1
x= (D.34)
2

by a matrix µ ¶
cos  − sin 
A= (D.35)
sin  cos 
gives
µ ¶µ ¶
cos  − sin  1
Ax = (D.36)
sin  cos  2
µ ¶
1 cos  − 2 sin 
= (D.37)
1 sin  + 2 cos 
= x0 (D.38)

I Figure D.1: Rotation of a coordinate system.


Appendix D 293

I The result is a rotation:

• In the original coordinate system, the new vector x0 corresponds to the result of
an anti-clockwise rotation of x around .
• Alternatively, we can say that x0 is given in a new coordinate system that is
obtained by a clockwise rotation of the old coordinate system around .

I Rotation matrices and rotation operators: A is called a rotation matrix. We shall


also call A a rotation operator.
I Exercise D.1: Show that Eqs. D.36 - D.38 does describe an anti-clockwise rotation
of the vector x to the new vector x0 and determine its coordinates using trigonometric
arguments. ¤
Appendix D 294

D.5 Systems of linear equations

I Solutions of linear equations (I): Cramer’s rule. Matrices and determinants can
be generally used for solving systems of linear equations of the type

11 1 + 12 2 + 13 3 +    1  = 1


21 1 + 22 2 + 23 3 +    2  = 2
 =  (D.39)
1 1 + 2 2 + 3 3 +      = 

These equations can be written in matrix form:


⎛ ⎞⎛ ⎞ ⎛ ⎞
11 12    1 1 1
⎜ 21 22    2 ⎟ ⎜ 2 ⎟ ⎜ 2 ⎟
⎜ . .. .. ⎟ ⎜ . ⎟=⎜ . ⎟ (D.40)
⎝ .. . . ⎠ ⎝ .. ⎠ ⎝ .. ⎠
1 2      
or
Ax = c (D.41)

Simple algebraic manipulations (see any math textbook) gives the expressions

|A|  = |A | (D.42)

where |A| is the determinant of the coefficient matrix A and |A | is the respective
determinant of |A| in which the th column is replaced by the 1 2  3    , e.g.
⎛ ⎞
11 1    1
⎜ 21 2    2 ⎟
|A2 | = ⎜
⎝ ... ... .. ⎟ (D.43)
. ⎠
1     

Solutions for the  : From Eq. D.42 we obtain the non-trivial solutions for the 
according to
|A |
 = (D.44)
|A|
under the condition that the A matrix is not singular, i.e. the determinant of coefficients
does not vanish
|A| 6= 0 (D.45)
(The trivial and uninteresting solutions are 1  2  3     = 0).

I Solutions of linear equations (II): In the following, we shall only consider the special
linear equations of the type

11 1 + 12 2 + 13 3 +    1  = 0


21 1 + 22 2 + 23 3 +    2  = 0
.
   = .. (D.46)
1 1 + 2 2 + 3 3 +      = 0
Appendix D 295

or
Ax = 0 (D.47)

The trivial and uninteresting solutions of this equation are 1  2  3     = 0.


The interesting solutions require that the matrix of coefficients A is singular, i.e. the
determinant of coefficients has to vanish

|A| = 0 (D.48)

This type of equations is encountered in eigenvalue problems.


Appendix D 296

D.6 Eigenvalue equations

Consider a set of linear equations as considered in the previous section which can be
written in the form
A x = x (D.49)
where

(1) A is an  × -dimensional matrix,


(2)  is a scalar constant, called an eigenvalue of A ( is any one of the set of 
eigenvalues of A), and
(3) x is an -dimensional column vector (which can in general be complex), called
the eigenvector of A belonging to the particular eigenvalue.

Eq. D.49 is called an eigenvalue equation. It has the following property: The multipli-
cation of x by  (and hence that of x by the matrix A) changes the length of x (by
the factor ), but not the direction.

I Eigenvalues and eigenvectors of the molecular Hamiltonian: The determina-


tion of eigenvalues and eigenvectors of the molecular Hamiltonian H, i.e. the “energy
matrix”, or the matrix of the Hamilton operator , b is the most important problem in
spectroscopy (in general, it is the key problem!!).

I Secular equation and eigenvalues λ : Eq. D.49 can be rewritten as

(A − I) x = 0 (D.50)

In order not to obtain the trivial solutions 1 = 2 = 3 =    = 0, det (A − I) has to


vanish, i.e. det (A − I) = 0. This is the so-called characteristic equation or secular
equation:
¯ ¯
¯ 11 −  12    1 ¯
¯ ¯
¯ 21 22 −     2 ¯
|A − I| = ¯ ¯ .. .. .. ¯=0 (D.51)
. .    . ¯
¯ ¯
¯     − ¯
1 2 

The secular equation has  roots

{1  2       } (D.52)

which are called the eigenvalues.

I Eigenvectors x : For each eigenvalue  , we can write an eigenvalue equation

A x1 = 1 x1 (D.53)
A x2 = 2 x2 (D.54)
..
. (D.55)
A x =  x (D.56)
Appendix D 297

In matrix notation, this becomes

AX=XΛ (D.57)

with the diagonal eigenvalue matrix


⎛ ⎞
1 0  0
⎜ 0 2  0 ⎟
Λ=⎜
⎝ ... .. . ⎟ (D.58)
.    .. ⎠
0 0    

The different eigenvectors x are usually normalized to unity, i.e.


¡ ¢12
|x| = x† x =1 (D.59)

I Matrix diagonalization: The matrix of the eigenvectors X can be determined by


multiplying the equation
AX=XΛ (D.60)
from the left with X−1 , which gives

X−1 A X = X−1 X Λ (D.61)

i.e.
X−1 A X = Λ (D.62)
y
The determination of the eigenvector matrix is equivalent to finding a matrix X that
transforms the matrix A into diagonal form (Λ).

I Example: µ ¶
2 3
A= (D.63)
3 10

Eigenvalue equations:
A x = x (D.64)

or in matrix form:
AX=XΛ (D.65)

Secular equation:
¯ ¯
¯2− 3 ¯
det (A − I) = ¯¯ ¯ (D.66)
3 10 −  ¯
= (2 − ) (10 − ) − 9 = 0 (D.67)

y
0 = 20 − 2 − 10 + 2 − 9 = 2 − 12 + 11 (D.68)
Appendix E 298

sµ ¶2
12 12 √
12 = + ± − 11 = 6 ± 36 − 11 (D.69)
2 2
= 6±5 (D.70)
y
Eigenvalues:
1 = 1 (D.71)
2 = 11 (D.72)

Eigenvectors: Inserting the eigenvalues one by one into the eigenvalue equation yields:

(1) µ ¶µ ¶ µ ¶ µ ¶
2 3 11 11 11
= 1 = 11 (D.73)
3 10 21 21 21
y µ ¶
−3
(D.74)
1

(2) µ ¶µ ¶ µ ¶ µ ¶
2 3 12 12 12
= 2 =1 (D.75)
3 10 22 22 22
y µ ¶
1
(D.76)
3

Unnormalized eigenvalue matrix:


µ ¶
−3 1
(D.77)
1 3

Normalized eigenvalue matrix:


⎛ −3 1 ⎞
√ √
⎜ ⎟
X = ⎝ 110 310 ⎠ (D.78)
√ √
10 10

I Final note: In practice, matrix manipulations such as matrix inversion, the evaluation
of determinants, the determination of eigenvalues and eigenvectors, etc., are carried out
numerically by using computers.

I References

Press 1992 W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P. Flannery, Numerical


Recipes in Fortran, Cambridge University Press, Cambridge, 1992. Versions are
also available for C and Pascal.
Appendix E 299

Appendix E: Laplace transforms

The concept of Laplace and inverse Laplace transforms is extremely useful in chemical
kinetics for two reasons:

(1) They connect microscopic molecular properties and statistically averaged quanti-
ties:

a)  () ↔  ( ): Collision cross section vs. thermal rate constant for bimole-
cular reactions (section ??),
b)  () ↔  ( ): Specific rate constant vs. thermal rate constant for uni-
molecular reactions (section 8),
c)  () ↔  ( ): Density of states vs. partition function in statistical rate
theories 8).

(2) They provide a convenient method for solving of differential equations (section
3.4.2).

I Definition E.1: The Laplace transform L [ ()] of a function  () is defined as the
integral
Z∞
 () = L [ ()] =  () −  (E.1)
0

where

•  is a real variable,
•  () is a real function of the variable  with the property  () = 0 for   0,
•  is a complex variable,
•  () = L [ ()] is a function of the variable .

I Definition E.2: The inverse Laplace transform L−1 [ ()] of the function  () is
defined as the integral

Z
+∞
−1 1
 () = L [ ()] =  ()   (E.2)
2
−∞

where

•  is an arbitrary real constant.


Appendix F 300

I Notes:

• Since L−1 [ ()] recovers the original function  (), the pair of functions  ()
and  () is said to form a Laplace pair.
• The properties of the Laplace and inverse Laplace transforms can be derived from
those of Fourier transforms (Zachmann 1972).
• Laplace transforms can be computed using symbolic algebra computer programs
(Maple, Mathematica, etc.). In favorable cases, it is also possible to obtain inverse
Laplace transforms by computer programs. However, it is often more convenient
to take the Laplace and inverse Laplace transforms from corresponding Tables.60

I Table E.1: Table of Laplace transforms.

 ()  () = L [ ()]  ()  () = L [ ()]


R∞
 ()  () = 0
 () − 

  () 1
  () −  ( = 0)  
 −
1 1
1  
 ( − )2

1 
 sin 
2 2 + 2

2 1 
cos 
2 3 2 + 2

−1 1 1 ¡  ¢ 1
 − 
Γ ()  ( − ) ( − ) ( − )

Γ () is the Gamma function. For integer , Γ () = ( − 1)! (see Appendix F).

I References:

Houston 1996 P. L. Houston, Chemical Kinetics and Reaction Dynamics, McGraw-


Hill, Boston, 2001.
Steinfeld 1989 J. I. Steinfeld, J. S. Francisco, W. L. Haase, Chemical Kinetics and
Dynamics, Prentice Hall, Englewood Cliffs, 1989.
Zachmann 1972 H. G. Zachmann, Mathematik für Chemiker, Verlag Chemie, Wein-
heim, 1972.

60
Extensive Tables of Laplace transforms are given by (Steinfeld 1989).
Appendix F 301

Appendix F: The Gamma function Γ (x)

• If  is a positive integer, the Gamma function is

Γ () = ( − 1)! (F.1)

• The Gamma function smoothly interpolates between all points ( ) given by
 = ( − 1)! for positive non-integer  (cf. Fig. F.1).
• The Gamma function is defined for all complex numbers  with a positive real
part as Z ∞
Γ () = −1 −  (F.2)
0
This integral converges for complex numbers with a positive real part. It has poles
for complex numbers with a negative real part (cf. Fig. F.2).

I Figure F.1: The Gamma function.


Appendix G 302

I Figure F.2: For negative arguments the Gamma function has poles (at all negative
integers).
Appendix G 303

Appendix G: Ergebnisse der statistischen Thermodynamik

In diesem Kapitel sollen wichtige Ergebnisse der statistischen Thermodynamik zusam-


mengefasst werden. Im Vordergrund steht die Berechnung der thermodynamischen Zu-
standsfunktionen und die Berechnung von chemischen Gleichgewichten aus molekularen
Eigenschaften. Wir benutzen die Ergebnisse der Quantenmechanik:

• Ein Molekül kann nur bestimmte Energiezustände  einnehmen.


• Diese Energiezustände sind nicht kontinuierlich, sondern diskret verteilt.
• Jeder Energiezustand hat ein bestimmtes statistisches Gewicht  . Das statistische
Gewicht gibt an, wie oft ein Zustand bei einer bestimmten Energie vorkommt.
Beispiele:

(G.1)
 = 1 eindimensionaler harmonischer Oszillator
 = 2 isotroper zweidimensionaler harmonischer Oszillator
(z.B. Biegeschwingung von CO2 )
 = 2 + 1 Rotation eines linearen Moleküls

• Die Ergebnisse der statischen Thermodynamik unterscheiden sich, je nachdem


ob das betrachtete System aus unterscheidbaren oder aus ununterscheidbaren
Teilchen besteht:

Beispiele für unterscheidbare Teilchen: Atome im Kristallgitter


Beispiele für ununterscheidbare Teilchen: 1.) Gasmoleküle
2.) Elektronen (Fermi-Statistik)
(G.2)
• Die anzuwendende Statistik hängt vom Spin der Teilchen ab:

(G.3)
Spin gerade: Bose-Einstein-Statistik
Spin ungerade: Fermi-Dirac-Statistik

Auf die Unterschiede zwischen der Bose-Einstein- und der Fermi-Dirac-Statistik


wird an anderer Stelle eingegangen (⇒ Vorlesung “Statistische Thermody-
namik”).
• Im Grenzfall   À  gilt die Boltzmann-Statistik.
• Im Folgenden werden behandelt: Elektronische Anregung, Schwingungsanregung,
Rotationsanregung, Translationsanregung.
Die Translation erfordert abhängig vom Problem manchmal eine etwas andere
Behandlung (klassische statistische Thermodynamik bzw. Semiklassik). Quanten-
mechanik und klassische Mechanik stimmen im Ergebnis für   À  überein.
Appendix G 304

G.1 Boltzmann-Verteilung

Die Besetzungswahrscheinlichkeit eines Zustands  wird durch die Boltzmann-Verteilung


angegeben. Diese ergibt sich als die wahrscheinlichste Verteilung aus dem Boltzmann-
Ansatz für die Entropie
 =  ln  (G.4)
 wird als die thermodynamische Wahrscheinlichkeit bezeichnet.
Die Boltzmann-Verteilung kann hergeleitet werden, wenn man entsprechend dem 2.
Hauptsatz den Maximalwert für  sucht (Methode der Lagrange’schen Multiplikatoren).
Hier setzen wir die Boltzmann-Verteilung als gegeben voraus (z.B. als empirisches, ex-
perimentelles Ergebnis):

I Boltzmann-Verteilung: Für diskrete Energiezustände  ist die Beset-


zungswahrscheinlichkeit eines Zustands proportional zu −  :

 = ∝ −  (G.5)

Bei Berücksichtigung einer möglichen Entartung von Zuständen (Entartungsfaktor bzw.
statistisches Gewicht  ) und mit einer Proportionalitätskonstante  0 zur Normierung
gilt

 = =  0 ×  × −  (G.6)

Dabei ist  die Zahl der Teilchen im Energiezustand  (mit dem statistischen Gewicht
 ) und  die Gesamtzahl der Teilchen.
Die Funktion  gibt die Wahrscheinlichkeit der Besetzung des Zustands  an. Die
Normierungskonstante  0 ergibt sich aus der Normierungsbedingung, dass die Gesamt-
wahrscheinlichkeit zu 1 herauskommt:
P X X
 
=  =  0 ×  −   = 1 (G.7)
  
1
y 0 = P −  
. (G.8)
  

I Normierte Boltzmann-Verteilungsfunktion:

  −  
 = =P −  
(G.9)
   

P
I Molekulare Zustandssumme: Die Summe   −   im Nenner der o.a. Glei-
chung wird als molekulare Zustandssumme bezeichnet:
P
=   −   (G.10)

Die Zustandssumme  ist die zentrale Grösse der statistischen Thermodynamik.  ist
von der Temperatur  abhängig.
Appendix G 305

I Anschauliche Bedeutung der molekularen Zustandssumme:


 = Zahl der bei der Temperatur  im Mittel besetzten Energiezustände eines Molekül.

I Boltzmann-Verteilungsfunktion für kontinuierlich verteilte Energiezustände:


Für kontinuierlich verteilte Energiezustände ersetzt man  durch () und erhält
analog eine kontinuierliche Besetzungswahrscheinlichkeit:

() () × −   


() = = (G.11)
 

I Zustandsdichte: () = Zustandsdichte (=Anzahl der Zustände pro Energie-


Intervall):
()
() = (G.12)

Appendix G 306

G.2 Mittelwerte

I Mittlere Energie ²: In der Thermodynamik interessieren Mittelwerte, z.B. die mit-


tlere Energie . Für Größen, die durch eine Verteilungsfunktion beschrieben werden,
erhält man den Mittelwert nach
P P P P
       X  
= P = P = =   = P (G.13)
    

 P
mit  = und  = 1.

Übergang zu einer kontinuierlichen Verteilungsfunktion liefert
Z∞
 ()
 = 0Z∞ (G.14)
()
0

Diese Ausdrücke vereinfachen sich, wenn () auf 1 normiert ist.

I Allgemeine Berechnung der Mittelwerts y einer Größe y(²):


X
 

= X (G.15)


bzw.
Z∞
()()
0
= Z∞ (G.16)
 ()
0

I Example G.1: Berechnung der mittleren Energie für ein 3-Niveausystem:

0 = 0 1 = 1 2 = 21 (G.17)
0 = 3 2 = 2 2 = 1
X
=  = 6 (G.18)
Appendix G 307

Ergebnis:
1 X X  X
 =   =  =  (G.19)
  
 
1
= (30 + 21 + 12 ) (G.20)
6
1
= (30 + 21 + 21 ) (G.21)
6
1
= (0 + 41 ) (G.22)
6
4
= 1 (G.23)
6
2
 = 1 (G.24)
3
¤

I Mittlere Energie eines Moleküls: Bei Gültigkeit der Boltzmann-Verteilung kann die
mittlere Energie eines Moleküle aus der molekularen Zustandssumme berechnet werden:
1 X X  X
 =   =  =   (G.25)
  
 
P
  −  
= P   −   (G.26)
   
P
  −     2
= |× (G.27)
   2
  2 X  −  
=   (G.28)
 
  2
  2 X  ¡ −   ¢
=   |(durch differenzieren zeigen) (G.29)
 

  2  X −  
=   (G.30)
  
  2 
= (G.31)
 
y
 ln 
 =   2 (G.32)

Appendix G 308

G.3 Anwendung auf ein Zweiniveau-System

I Chemisch relevante Beispiele für Zweiniveau-Systeme:

• Paramagnetische Atome u. Moleküle


• NMR-Spektroskopie: Atomkerne mit Kernspin. Typischer Frequenzbereich:
100 MHz.
• Molekül mit cis-trans-Isomeren oder 2 Konformeren

I Figure G.1: Zwei-Niveau-System: Energiezustände

The Two Level System: Energy Levels

Zustandssumme:
 Q   gi e i / kT
 g
i

 1  e 0 / kT  1  e 1 / kT
 1  e 1 / kT
Besetzungsverhältnis:
N 0 g0e 0 / kT

N Q
N1 g1e 1 / kT

N Q
N1 g1e 1 / kT
  g    e 1 / kT
N 0 g0e 0 / kT

Physical Chemistry III: Chemical Kinetics • © F. Temps, IPC Kiel 173

I Besetzungsverhältnis:
1 1 −1 
= (G.33)
 
mit

 = 0 −0  + 1 −1  (G.34)


= 1 + −1  (G.35)

(1) Grenzfall für  → 0:


1
→ ∞ y −1  → −∞ = 0 (G.36)

y Bei  → 0 ist nur der tiefste Energiezustand besetzt.
Appendix G 309

(2) Grenzfall für  → ∞:


1
→ 0 y −1  → −0 = 1 (G.37)

y Bei  → ∞ ergibt sich eine Gleichbesetzung der Energiezustände. Dies ist die
maximale Besetzung im oberen Zustand!

• Eine Besetzungsinversion (1  0 ) würde formal eine negative Temperatur


erfordern (→ Laser).

I Figure G.2: Zwei-Niveau-System: Boltzmann-Besetzung der Zustände (rot: unterer


Zustand, blau: oberer Zustand).

The Two Level System: Boltzmann Distribution

Ni
1
 = 1000 J/mol
N

0.75

0.5

0.25

0
0 250 500 750 1000
T / (K)

Physical Chemistry III: Chemical Kinetics • © F. Temps, IPC Kiel 174

I Mittlere thermische Energie:

(1) 1 = 125 kJ mol (entsprechend  = 1000 cm−1 ; z.B. C-C-Schwingung im IR


oder Energieabstand zwischen cis-trans-Isomeren):
½
1 −1   8 × 10−3 @  = 300 K
= = (G.38)
0 023 @  = 1000 K

(2) 1 = 004 kJ mol (entsprechend Energieabstand zwischen Kernspinzuständen bei


100 MHz NMR):
1
= −1   = 0999984 @  = 300 K (G.39)
0
Reihenentwicklung:
1 1
+  ≈ 1 − −1  = 1 −
(G.40)
 
y sehr kleiner Besetzungsunterschied y Wunsch der NMR-Spektroskopiker zu
immer größeren Frequenzen (Rekord heute 950 MHz).
Appendix G 310

I Figure G.3: Zwei-Niveau-System: Mittlere thermische Energie.

The Two Level System: Energy

E 1000
 = 1000 J/mol

750

500

250

0
0 250 500 750 1000
T / (K)

Physical Chemistry III: Chemical Kinetics • © F. Temps, IPC Kiel 175

I Spezifische Wärme:

 = (G.41)


I Figure G.4: Zwei-Niveau-System: Spezifische Wärme.

The Two Level System: Specific Heat

cv
5
 = 1000 J/mol

0
0 250 500 750 1000
T / (K)

Physical Chemistry III: Chemical Kinetics • © F. Temps, IPC Kiel 176


Appendix G 311

G.4 Mikrokanonische und makrokanonische Ensembles

Die statistische Thermodynamik macht Aussagen über statistische Ensembles


(Ansammlungen von Systemen, Molekülen). Wir betrachten verschiedene Ensembles
von miteinander wechselwirkenden Molekülen.

I Der Begriff des Ensembles: Wir betrachten ein abgeschlossenes System bei kon-
stantem Volumen, konstanter Zusammensetzung und konstanter Temperatur (bzw. bes-
timmter Energie (s.u.)). Ein Ensemble erhalten wir, wenn wir uns N Replikationen dieser
Systeme denken, die wir unter Einhaltung bestimmter Regeln erhalten (Regel = Canon).
Alle Systeme in dem Ensemble sollen im thermischen Gleichgewicht miteinander stehen
(sie dürfen zwischeneinander Energie austauschen).
Es werden verschiedene Sorten von Ensembles unterschieden:

• Mikrokanonisches Ensemble: Ensemble von  Systemen (z.B. Atomen,


Molekülen) mit konstanter Teilchenzahl, konstantem Volumen, konstanter En-
ergie.
• Makrokanonisches Ensemble (kurz kanonisches Ensemble genannt): Ensemble
von  Systemen (z.B. Atomen, Molekülen) mit konstanter Teilchenzahl, konstan-
tem Volumen, konstanter Temperatur.
• Großkanonisches Ensemble: Ensemble von  offenen Systemen (z.B. Teilchen,
Molekülen) mit konstantem Volumen, konstanter Temperatur, die zwischeneinan-
der Teilchen austauschen können.

I Systemzustandssumme (= kanonische Zustandssumme): Molekulare Zu-


standssumme: P
 = alle Molekülzustände  −  (G.42)

Systemzustandssumme (= kanonische Zustandssumme):


P
= alle Systemzustände  −  (G.43)

Innere Energie des Systems:

 ln 
 =  2 (G.44)


Bei mehreren Teilchensorten (A, B, C, . . . ): Zustandsummen sind multiplikativ,


da die Energien additiv sind, y

 =  ×  ×  ×    (G.45)
Appendix G 312

• Systemzustandssumme für  unterscheidbare Teilchen der Sorte  (z.B.


Kristall):
Y
=  =  

(G.46)

y für 1 mol:
 ln   ln 
 =  2 =  2 (G.47)
 

• Systemzustandssumme für  nicht unterscheidbare Teilchen der Sorte  (z.B.


Gas):
 

=  (G.48)
 !
y für 1 mol ( =  ):

 ln   
 =  2 =  2 ln (G.49)
   !
Appendix G 313

G.5 Statistische Interpretation der thermodynamischen Zustandsgrößen

I Innere Energie:
 ln 
 =  2 (G.50)


I Helmholtz-Energie:
 
= − (G.51)
 
µ ¶

µ ¶  − µ ¶
    
= − (G.52)
   2  
  −     ln 
= − = − = − (G.53)
2  2 
y

= − ln  (G.54)

y
 = − ln  (G.55)

I Entropie:
µ ¶
  −
=− = =  (G.56)
  

I Druck: µ ¶ µ ¶
  ln 
 =− =  =  (G.57)
   
Appendix G 314

G.6 Chemisches Gleichgewicht

+ À (G.58)
y µ ¶
∆ ª
 = exp − (G.59)

bzw.
∆ ª = − ln  (G.60)

I Gleichgewichtskonstante: Wir berechnen zunächst


 = − ln  , (G.61)
 

indem wir  =  etc.. . . einsetzen und die Stirling’sche Näherung ln ! =


 !
 ln  −  anwenden:
 = − [ln  + ln  + ln  ] (G.62)
" #

 

= − ln  + ln  + ln  (G.63)
 !  !  !
= − [( ln  − ln  !)
+ ( ln  − ln  !)
+ ( ln  − ln  !)] (G.64)
= − [( ln  −  ln  +  )
+ ( ln  −  ln  +  )
+ ( ln  −  ln  +  )] (G.65)

I Gleichgewichtsbedingung: Im chemischen Gleichgewicht muss gelten


µ ¶

=0 (G.66)
 
y

0 = [( ln  −  ln  +  )

+ ( ln  −  ln  +  )
+ ( ln  −  ln  +  )] (G.67)
µ ¶
1
= ln  −  − ln  + 1

µ ¶
1
+ ln  −  − ln  + 1

µ ¶
1
− ln  −  − ln  + 1 (G.68)

= (ln  − ln  ) + (ln  − ln  ) − (ln  − ln  ) (G.69)
  
= ln + ln (G.70)
  
Appendix G 315

y
 
= (G.71)
   

Übergang auf Teilchendichten und volumenbezogenen Zustandssummen:

(  ) (  )
= (G.72)
(  ) (  ) (  ) (  )
y
(  ) ∗
 = = ∗ ∗ (G.73)
(  ) (  )  

Bezug auf die getrennten molekularen Nullpunktsenergien:

0
 = 0 0 × −∆0   (G.74)
 

mit 0 = die auf das Volumen bezogene Molekülzustandssumme:


0 = ×  ×  ×  (G.75)

I Endergebnis: Die Indices ∗ und 0 wurden hier zur Klarheit benutzt. Der Einfachheit
halber werden diese üblicherweise weggelassen.
y

 = × −∆0   (G.76)
 
Appendix G 316

G.7 Zustandssumme für elektronische Zustände

Die elektronische Zustandssumme muss unter Einschluss aller elektronischen Entartun-


gen explizit berechnet werden:
X
 =  −  (G.77)

Speziell sind zu berücksichtigen:

• Spinentartung (Spinmultiplizität ):

 = 2 + 1 (G.78)

• elektronischer Bahndrehimpuls ( (Atome) bzw. Λ (lin. Moleküle) bzw. Symme-


trierasse  oder  für nichtlin. Moleküle):
• elektronische Feinstrukturkomponenten (Index  im Atom-Termsymbol):

 = 2 + 1 (G.79)

Nur niedrig liegende Elektronenzustände liefern einen Beitrag. Die Zustandsumme ist
explizit auszurechnen.
Appendix G 317

G.8 Zustandssumme für die Schwingungsbewegung

I Harmonischer Oszillator:

 =    = 0 1 2    (G.80)

I Schwingungszustandssumme:
X
∞ X


 = −  = −· mit  = (G.81)

©=0 − −2
=0
ª
= 1+ + + −3 +    mit  = −  1 (G.82)
2 3
= 1 +  +  +  +  (G.83)
1
= für   1 (G.84)
1−
y
∙ µ ¶¸−1
1 
 = = 1 − exp − (G.85)
1 − − 

I s Oszillatoren:
Y ∙ µ ¶¸−1
 
 = 1 − exp − (G.86)
=1


I typische Werte bei Zimmertemperatur:

 ≈ 1    10 (G.87)

I Grenzwert
µ von ¶
Q für T → ∞ bzw. für ν → 0: Für  → ∞ bzw. für  → 0

kann exp − entwickelt werden:

µ ¶
 
exp − ≈1− +  (G.88)
 
y
1 1
 = µ ¶= µ ¶ (G.89)
 
1 − exp − 1− 1− + 
 
y

 ≈ (G.90)


Für  klassische Oszillatoren:


Y

 ≈ (G.91)
=1
 
Appendix G 318

I Abweichungen vom harmonischen Oszillator:

• Für anharmonische Oszillatoren


Pist die−Zustandssumme am besten durch explizite
numerische Berechnung nach ∞ =0   
zu ermitteln.
• Torsionsschwingungen können bei tiefen Temperaturen als harmonische Oszilla-
toren beschrieben werden, bei hohen Temperaturen als freie 1D-Rotatoren. Im
Zwischenbereich muss die Zustandssumme ebenfalls explizit berechnet werden
(wobei die Energiezustände für den gehinderten 1D-Rotator zugrundegelegt wer-
den müssen).
Appendix G 319

G.9 Zustandssumme für die Rotationsbewegung

I lineare Moleküle:
~2
 = ( + 1)  = 0 1 2    (G.92)
2
~2
= (G.93)
2

I Rotationszustandssumme:

1X

 (2) =  −  (G.94)
 =0
1X

2
= (2 + 1) −~ (+1)2 (G.95)
 =0
Z∞
1 2
= (2 + 1) −~ (+1)2  (G.96)

=0

y
2  
 (2) = = (G.97)
~2 

I Symmetriezahl:
 (G.98)

I Rotationszustandssumme für nichtlineare Moleküle:  ist nur für symmetrische


Kreisel analytisch darstellbar, nicht für asymmetrische Kreisel. Für nicht zu tiefe Tem-
peraturen ist
µ ¶32
12 2  12
 (3) = (   )12 = ( )32 (  )−12 (G.99)
 ~2 

I typische Werte bei Zimmertemperatur:

 ≈ 100    1000 (G.100)


Appendix G 320

G.10 Zustandssumme für die Translationsbewegung

I Teilchen im 1D-Kasten:
2 ³  ´2
 = ×   = 1 2 3    (G.101)
8 

I 1D-Translationszustandssumme:

X
∞ Z∞
2 2 2
 = −  = −  8   (G.102)
 1
µ ¶12
2 ¡ ¢
= × in cm−1 (G.103)
2
y
µ ¶12
 2
= (G.104)
 2

I Teilchen im 3D-Kasten:
"µ ¶2 µ ¶2 µ ¶2 #
2   
 = × + + (G.105)
8   

I 3DTranslationszustandssumme:
µ ¶32
2 ¡ ¢
 = × in cm−3 (G.106)
2
y
µ ¶32
 2
= (G.107)
 2

I typische Werte:
 ≈ 1024 ×  (G.108)
y

≈ 1024 cm−3 (G.109)

You might also like